timestamp
timestamp[s]
yymm
stringlengths
4
4
arxiv_id
stringlengths
16
16
language
stringclasses
2 values
url
stringlengths
38
38
categories
sequence
id
stringlengths
16
16
submitter
stringlengths
2
27
authors
stringlengths
5
193
title
stringlengths
14
125
comments
stringlengths
5
508
journal-ref
stringlengths
23
180
doi
stringlengths
16
31
report-no
stringlengths
5
90
license
stringclasses
2 values
abstract
stringlengths
73
1.64k
versions
list
update_date
timestamp[s]
authors_parsed
sequence
primary_category
stringclasses
1 value
text
stringlengths
1
472k
1995-12-05T06:20:27
9512
alg-geom/9512003
en
https://arxiv.org/abs/alg-geom/9512003
[ "alg-geom", "math.AG" ]
alg-geom/9512003
Bumsig Kim
Bumsig Kim
Quot schemes for flags and Gromov invariants for flag varieties
16 pages, wrtten by latex 209, compile twice
null
null
null
null
Using Quot schemes and a localization theorem we study Gromov-Witten invariants for partial flag varieties. The strategy is to extend A. Bertram's result of Gromov-Witten invariants for special Schubert varieties of Grassmannians to the case of partial flag varieties. To do so a Grothendieck's Quot scheme is generalized for flags and proven to be an irreducible, rational, smooth, projective variety following Str\o mme \cite{St}.
[ { "version": "v1", "created": "Mon, 4 Dec 1995 21:58:00 GMT" } ]
2015-06-30T00:00:00
[ [ "Kim", "Bumsig", "" ] ]
alg-geom
\section{Introduction} To define Gromov-Witten invariants arise in mirror symmetry, there are two general rigorous methods so far \cite{RT}\cite{Ko}. In particular Kontsevich introduced the notion of stable maps for a compactification of moduli spaces. For Grassmannians, however, there is a natural compactification of the space $% Mor_d({\PP}^1,Gr(n,r))$ of all holomorphic maps from ${\PP}^1$ to Grassmannians with a given degree $d$ where $Gr(n,r)$ is the Grassmannian of all rank $r$ quotient vector spaces of ${\CC}^n.$ We may see $Mor_d({\PP}% ^1,Gr(n,r))$ as the set of all rank $r,$ degree $d,$ quotient bundles of $% {\CC}^n\otimes {\cal O}_{{\PP}^1}.$ It is not a compact space. Hence we come to a Grothendieck's Quot scheme $Quot_d(Gr(n,r)),$ the set of all rank $r,$ degree $d,$ quotient sheaves of ${\CC}^n\otimes {\cal O}_{{\PP}^1}$ \cite {Gr}$.$ It is proven to be a smooth projective variety by S. A. Str\o mme \cite{St}. Bertram and Franco-Reina used Grothendieck's Quot schemes for Gromov-Witten invariants and quantum cohomology of Grassmannians respectively \cite{Be}\cite{FR}. In this paper using Quot schemes and a localization theorem we study Gromov-Witten invariants for partial flag varieties. The strategy is the following. We extend A. Bertram's result of Gromov-Witten invariants for special Schubert varieties of Grassmannians to the case of partial flag varieties. To do so a Grothendieck's Quot scheme is generalized for flags and proven to be an irreducible, rational, smooth, projective variety following Str\o mme \cite{St}. The Hilbert schemes for flags have already studied \cite{Se}. On a partial flag manifold there is an action by a special linear group. It induces an action on the Quot scheme for flags. There is another action on it by the multiplicative group ${\CC}^{\times }.$ It is induced from the action $% \CC ^{\times }$ on ${\PP}^1.$ The analogous action by ${\CC}^{\times }$ does not exist on the Kontsevich moduli space. These two actions are commutative. These together give isolated fixed points. Using a localization by action an explicit formula of Gromov invariant for special Schubert classes with a certain condition is given. Note that Kontsevich also uses torus actions on his moduli space of stable maps \cite{Ko}. In his case he has to deal with summation over trees since the fixed subsets are rather complicated. For projective spaces the formula derived in the sequel is shown to agree to the residue formula \cite {Ki}. The author does not know how to directly relate with the result of Givental and Astashkevich-Sadov's computation \cite{GK}\cite{AS}. Now we state our main results. \bigskip For given integers $s_0=0<s_1=n-r_1<\cdots <s_l=n-r_l<n=s_{l+1},$ a flag variety $Fl:=F(s_1,s_2,...,s_l;n)$ is, by definition, the set of all flags of complex subspaces $V_1\subseteq V_2\subseteq \cdots \subseteq V={\Bbb C}^n, \;\dim V_i=s_i$. There are universal vector bundles $S_i$ and universal quotient bundles $Q_i$ over $Fl$ with fibers ${\Bbb C}^{s_i}$ and $\CC ^n/\CC ^{s_i}$ respectively. We are interested in a moduli space, the set $Mor_d({\Bbb P}^1,Fl)$ of all morphisms $% \varphi $ from ${\Bbb P}^1$ to $Fl$ with $<{\Bbb P}^1,c_1(\varphi ^{*}Q_k)>=d_k$, $d=(d_1,d_2,...,d_l)$. Since $Fl$ is the fine moduli space such that the associated flag functor is equivalent to $Mor(\cdot ,Fl)$ where the image of the functor at a scheme $S$ is the set of all flag quotient bundles of $V\otimes\cal{O}_S$ with ranks $r_i$. From the point of view as above, $Mor_d(\PP ^1 ,Fl)$ is realized as the set of all flag quotient bundles $(F_1,...,F_l)$ of $V\otimes\cal{O}_{\PP ^1}$ with rank $r_i$ and degrees $d_i$ and hence it can be compactifyed by collecting flag quotient sheaves. More precisely, \bigskip \begin{theorem}\label{thm1} There is a smooth compactification $fQuot_d(Fl)$ of $Mor_d({\Bbb P}^1,Fl)$. The underlying set of $fQuot_d(Fl)$ is the set of all flag quotient sheaves $(F_1,...,F_l)$ of $V\bigotimes {\cal O}_{{\Bbb P}^1}$ over ${\Bbb P}^1$ where rank of $F_i$ is $r_i$ and its degree is $d_i.$ Over the irreducible, rational, projective variety $fQuot_d(Fl)\times \PP ^1$ there are tautological bundles $\cal {E}_i$ and sheaves $\cal {Q}_i$, $i=1,...,l$. They form exact sequences $$ 0\rightarrow \cal {E}_i\rightarrow V\otimes O_{\PP ^1\times fQuot_d(Fl)} \rightarrow \cal {Q}_i\rightarrow 0.$$ The induced sheaf morphisms $\cal {E}_i\rightarrow \cal{Q}_{i+1} $ are identically zero. \end{theorem} This fine moduli space $fQuot_d(Fl)$ will give the Gromov-Witten invariants defined in \cite{KM}\cite{Ko}. \begin{theorem}\label{thm2} Let $p_1,...,p_N$ be fixed N distinct points in $\PP ^1$. For i=1,...,N, let $\alpha _i$ be integers in $\{s_1,...,s_{l+1}\}$ and let $\beta _i$ be positive integers less than $s_{\alpha _i+1}-s_{\alpha _i-1}$. Then the number of morphism $\varphi $ from $\PP ^1$ to $Fl$ such that each $\varphi (p_i)$ is in each the Poinc\'{a}re dual Schubert subvariety to the classes $c_{\beta _i}(S_{\alpha _i})$ and $<\PP ^1 ,c_1(\varphi ^*Q_k)>=d_k,$ $% k=1,...,l,$ is well-defined and it is \[ \int _{fQuot_d(Fl)}\wedge _i c_{\beta _i} (\cal {E}_{p_i}^i). \] The integration is not depend on the choices of the point $p_1,...,p_N$ in $\PP ^1$. \end{theorem} By the torus action on $fQuot_d(fl)$ induced from the standard $\CC ^{\times }$-action on $\PP ^1$ and the standard $(\CC ^{\times})^n$ on $Fl$ one can apply Bott's residue formula to the above integration to get \begin{theorem}\label{thm3} The integration in the theorem \ref{thm2} is \[ \sum_{\text{all integers as in (\ref{integer})}}\frac{\prod _i (\sigma _{\alpha _i}^{\beta _i}) \prod (\text{characters as in (\ref{tang1}))}} {\prod (\text{characters as in (\ref{tang2}))}}. \] The notations will be explained as follows. \end{theorem} Consider a sequence of data by nonnegative integers $d_{i,j}$ and $a_{i,j}$: \begin{equation} (d_{1,1},a_{1,1};...;d_{1,s_1},a_{1,s_1})\cdots (d_{l,1},a_{l,1};...;d_{l,s_l},a_{l,s_l}) \label{integer} \end{equation} such that $d_{i,j}-a_{i,j}\geq d_{i+1,j}-a_{i+1,j}\geq 0,\;a_{i,j}\geq a_{i+1,j}$ and $\sum_{j=1}^{r_i}d_{i,j}=d_i.$ Set $b_{i,j}:=d_{i,j}-a_{i,j}.$ Then we consider \begin{eqnarray} (p-a_{i,j})\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }% 0\leq p\leq a_{i,j^{\prime }}-1,\;1\leq j,j^{\prime }\leq s_i, \nonumber \label{tang1} \\ (b_{i,j}-p)\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }% 0\leq p\leq b_{i,j^{\prime }}-1,\;1\leq j,j^{\prime }\leq s_i, \label{tang1} \\ (p-a_{i,j})\hbar +\lambda _{m}-\lambda _{j}.\;\text{for }0\leq p\leq d_{i,j},\;1\leq j\leq s_i,\;s_i+1\leq m\leq n, \nonumber \end{eqnarray} and \begin{eqnarray} \label{tang2} \\ (p-a_{i,j})\hbar +\lambda _{j^{\prime }}-\lambda _{j},\; \text{for }0 &\leq &p\leq a_{i+1,j^{\prime }}-1,\;1\leq j,\leq s_i,\;1\leq j^{\prime }\leq s_{I+1}, \nonumber \\ (b_{i,j}-p)\hbar +\lambda _{j^{\prime }}-\lambda _{j},\; \text{for }0 &\leq &p\leq b_{i+1,j^{\prime }}-1,\;1\leq j,\leq s_i,\;1\leq j^{\prime }\leq s_{i+1}, \nonumber \\ (p-a_{i,j})\hbar +\lambda _{m}-\lambda _{j},\; \text{for }0 &\leq &p\leq d_{i,j},\;1\leq j\leq s_i,\;\;s_{i+1}+1\leq m\leq n. \nonumber \end{eqnarray} Finally let $\sigma ^k_i$ be the $k$-th elementary symmetry function of $a_{i,j}\hbar +\lambda _{j}$, $j=1,...,s_i$. \bigskip {\bf Acknowledgments and remarks: }My special thanks goes to my advisor A. Givental for his wonderful guide. I would also like to thank A. Bertram, R. Hartshorne, M. Kontsevich, and S. A. Str\o mme for answering my questions. After finishing a preliminary version of the paper I learned there are more advanced results by A. Bertram \cite{Be1} and I. Ciocan-Fontanine \cite{CF} in some cases. But works seem complementary. I would like to express my thanks to I. Ciocan-Fontanine for pointing out an error in the preliminary version of the paper. By \cite{Be1}\cite{CF} the condition $\beta _i < s_{\alpha _i+1}-s_{\alpha _i-1}$ could be omitted. \section{flag-Quot schemes} All schemes will be assumed to be algebraic schemes over an algebraically closed field ${\bf k}$ of characteristic $0\;$and all sheaves will be quasi-coherent. The space of all rank $r,$ degree $d,$ quotients of trivial sheaf $V\otimes {\cal O}_{{\Bbb P}^1}$---or equivalently, all subsheaves of rank $s=n-r$, degree $-d$---is a smooth, rational, irreducible, projective variety $% Qout(V\otimes {\cal O}_{{\Bbb P}^1}$, Hilbert polynomial $rx+r+d)$ \cite{Gr}% \cite{St}. Let us denote by $Quot_d(s,V)$ the Quot scheme. It could be considered as a compactification of the space $Mor_d({\Bbb P}^1,Gr(n,r))$ of all degree $d$ holomorphic maps from the projective line ${\Bbb P}^1$ to the Grassmannian $Gr(n,r)$ which is the set of all rank $r$ quotient spaces of $V$ \cite{FR}. It is equipped with a universal locally free sheaf ${\cal E}$ and a universal quotient sheaf ${\cal Q}$ over ${\Bbb P}^1\times Quot$ with an exact sequence $0\rightarrow {\cal E}\rightarrow V\bigotimes {\cal O}_{{\Bbb P% }^1\times Quot}\rightarrow {\cal Q}\rightarrow 0.$ They are flat over $Quot.$ For the special Schubert varieties, Gromov invariants can be defined via Quot schemes as enumerative invariants \cite{Be}\cite{FR}. To extend their results to flag varieties (not necessary complete flag varieties), we will construct Quot schemes for flags, a compactification of $Mor_d(\PP ^1, Fl)$. \bigskip Let $s=(s_1,...,s_l),\;s_{i+1}>s_i>0,$ and $d=(d_{1,}...,d_l)$ be multi-indices of nonnegative integers. For this moduli problem, first we introduce a moduli functor $F^s_d$. A contravariant functor $F_d^s$ from the category of schemes to the category of sets is defined to be: for a scheme $S$, $F_d^s(S):=$the set of all flag subsheaves $(E_1,E_2,...,E_l,E% _n=V\bigotimes {\cal O}_{{\Bbb P}^1\times S})$ over ${\Bbb P}^1\times S$ where sheaf ${\cal E}_i$ is subsheaf of ${\cal E}_j$ if $i<j,$ sheaves are flat over $S$, and the rank of ${\cal E}_i$ is $s_i$ and its degree is $-d\;$over ${\Bbb P% }^1$. The functor $F_d^s\,$can be defined by quotients in a more transparent way, for different data $0\rightarrow E\rightarrow V\otimes\cal{O}_{\PP ^1\times S}$ could give the same data $V\otimes \cal{O}_{\PP ^1\times S}\rightarrow F\rightarrow 0$ \cite{FR}. Now we will show the functor is representable. First some notations; for a sheaf $\cal F$ over $\PP ^1 \times S$, denote by $\cal {F}(m)$ $\cal{F}\otimes \pi ^*\cal {O}(m)$, where $\pi$ is the projection from $\PP ^1\times S$ to $\PP ^1$. For the second projection from $\PP ^1\times S$ to $S$ we will use $\pi _S$. Taking advantage of the existence of Quot schemes, the obvious candidate for the scheme representing the functor $F^s_d$ is the appropriate subscheme of $Quot_{d_1}(s_1,n)\times ...\times Quot_{d_l}(s_l,n)$. Over the product of Quot schemes there are universal subsheaves $\cal {E}_i$ and quotient sheaves $\cal {Q}_i$ induced each from $Quot_{d_i}(s_i,n)$. Define $fQuot_d(Fl) $ as the degeneracy loci of \[ \bigoplus_{i=1}^{l-1} \left( (\pi _{\Pi Quot_{d_i}(s_i,n)}){*}{\cal E}_i(m)\rightarrow V/% (\pi _{\Pi Quot_{d_i}(s_i,n)})_{*}({\cal E}_{i+1}(m))\right) \] for any $m\geq \max_i\{d_i\}-1$. Note that $(\pi _{\Pi Quot_{d_i}(s_i,n)})_{*}{\cal E}_i(m)$ and $% V/(\pi _{\Pi Quot_{d_i}(s_i,n)})_{*}({\cal E}_{i+1}(m))$ are locally free so that there is no problem giving a scheme structure on $fQuot_d(Fl)$. For shorthand, we will write $Quot_{d_i}(s_i,n)$ by $Quot_i$ and $% fQuot_d(Fl)$ by $fQuot$, then now we are ready for \begin{proposition} The functor $F_d^s$ is representable by the (unique) projective scheme, $% fQuot_d(Fl)$. \end{proposition} \proof The statement means that for any scheme $S$, $F_d^s(S)=Mor(S,fQuot)$ in a functorial way. Let $(E_1,...,E_l)\in F_d^s(S).$ Then we have the morphism $g:S\rightarrow Quot_1\times ...\times Quot_l$ from the fine moduli property of the Quot schemes. We see that $g^{*}(\pi _{\Pi Quot_i})_{*}{\cal E}_i(m) \cong (\pi _S)_{*}E_i(m)$ naturally \cite{Mu}. The fact that $E_i\subset E_{i+1}$ implies that $(\pi _S)_{*}E_i(m)\subset (\pi _S)_{*}E_{i+1}(m).$ Hence $g^{*}(\pi _{\Pi Quot_i})_{*}{\cal E}_i(m)\subset g^{*} (\pi _{\Pi Quot_i})_{*}% {\cal E}_{i+1}(m)$ and $g$ factor through $fQuot$. $\Box $ \bigskip Over $fQuot$ there are exact sequences of universal sheaves \[ 0\rightarrow {\cal E}_i\rightarrow V_{{\Bbb P}^1\times fQuot}\rightarrow {\cal Q}_i\rightarrow 0 \] and surjections ${\cal Q}_i\rightarrow {\cal Q}_{i+1}.$ Each ${\cal E}_i$ are locally free since it is a subsheaf of a locally free sheaf over ${\Bbb P}% ^1\times fQuot$ and it is flat over $fQuot.$ For given $f\in Mor(S,fQuot),$ the corresponding quotient sheaves over ${\Bbb P}^1\times S$ is just the pull back $(id\times f)^{*}({\cal Q}_i).$ \subsection{Irreducibility and Smoothness} To show the Quot scheme for flags is irreducible and smooth, one can simply adapt Str\o mme's proof \cite{St}. \begin{theorem}\label{thm4} $fQuot$ is an irreducible, rational, nonsingular, projective variety. \end{theorem} \proof We will work by the language of quotients rather than subsheaves. For $m=0,-1$ and $i=1,...,l,$ let ${\cal Q}_m^i$ be $(\pi _{fQuot})_{*}{\cal Q}_i(m)$, a locally free sheaf over $% fQuot$ of rank $(m+1)r_i+d_i$, and let $X_m^i\rightarrow fQuot$ be the associated principal $GL((m+1)r_i+d_i)$-bundle. One has a smooth morphism $\rho :$% \[ \begin{tabular}{l} $\prod_{fQuot,i=1}^l(X_{-1}^i\times _{fQuot}X_0^i)=:Y$ \\ $\;\;\;\;\;\;\;\;\downarrow \rho $ \\ $\;\;\;\;\;\;\;fQuot.$% \end{tabular} \] We will show that $Y$ is an irreducible and smooth variety after finding an isomorphism to a smooth irreducible affine quasi-variety. Since $\rho $ is smooth, we conclude $fQuot$ is a smooth, irreducible, projective variety. Let $% N_m^i:=V^{(m+1)r_i+d_i}$ for $m=0,-1$ and$\;i=1,...,l.$ Here $V^r$ is, by definition, the $r$-dimensional vector space over the ground field $k.$ Let $% W:=Hom(V,V^{r_1+d_1})\times Hom_{{\Bbb P}^1}(\pi ^{*}V^{d_1}(-1),\pi ^{*}V^{r_1+d_1})$ $\times Hom(V^{d_1},V^{d_2})\times Hom(V^{r_1},V^{r_2+d_2})\times Hom_{{\Bbb P% }^1}(\pi ^{*}V^{d_2}(-1),\pi ^{*}V^{r_2+d_2})$ $\times \cdots \times Hom(V^{d_{l-1}},V^{d_l})\times Hom(V^{r_l},V^{r_l+d_l})\times Hom_{{\Bbb P}^1}(\pi ^{*}V^{d_l}(-1),\pi ^{*}V^{r_l+d_l}).$ Let $\overline{X}:=$associated affine space. On ${\Bbb P}^1% \times \overline{X},$ there is a tautological diagram: \begin{eqnarray*} && \begin{array}{ccc} \pi _{\overline{X}}^{*}V^{d_l}(-1) & {\rightarrow }_{v_1} & \pi _{% \overline{X}}^{*}V^{r_l+d_l} \\ & & \uparrow \mu _l \end{array} \\ &&\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,% \,\,\,\,\,\,\,\,\vdots \\ && \begin{array}{ccc} \pi _{\bar{X}}^{*}V^{d_2}(-1) & {\rightarrow }_{v_2} & \pi _{\bar{X}% }^{*}V^{r_2+d_2} \\ & & \uparrow \mu _2 \end{array} \\ && \begin{array}{ccc} \pi _{\overline{X}}^{*}V^{d_1}(-1) & {\rightarrow }_{v_1} & \pi _{% \overline{X}}^{*}V^{r_1+d_1} \\ & & \uparrow \mu _1 \\ & & \pi _{\overline{X}}^{*}V. \end{array} \end{eqnarray*} \ Let $Z\subset \overline{X}$ be the nonsingular irreducible quasi-variety defined by the conditions i) $v_i$ is injective on each fiber over $X$, ii) the induced map $\overline{\mu }_i:\pi _Z^{*}V\rightarrow $coker ($% v_i)$ is surjective, and iii) coker$(v_i)$ is flat over $Z$ with rank $r_i$ and degree $d_i$ on the fibers of $\pi _Z.$ The above conditions are open conditions in algebraic geometry. Let $% W^{\prime }:=Hom(V^{d_1},V^{r_2+d_2})\times \cdots \times Hom(V^{d_{l-1}},V^{r_l+d_l}).$ Let $X$ be a closed subvariety of $Z$ $\,$ of codimension $\sum_{i=1}^ld_i(r_{i+1}+d_{i+1})$ defined by the inverse image of a morphism \begin{equation} W\rightarrow W^{\prime } \label{ww} \end{equation} measuring commutativity of the diagram \[ \begin{array}{ccc} \pi ^{*}_ZV^{d_{i+1}}(-1) & \rightarrow _d & \pi ^{*}_ZV^{r_{i+1}+d_{i+1}} \\ \uparrow a & & \uparrow b \\ \pi ^{*}_ZV^{d_i}(-1) & \rightarrow _c & \pi ^{*}_ZV^{r_i+d_i} \end{array} \] by $d\circ a-b\circ c.$ $X$ is a product of hypersurfaces defined by irreducible quadratic polynomials. It is smooth away where the right vertical arrow is zero map. And so $X$ is smooth and irreducible. There is a natural morphism $% g:X\rightarrow fQuot$ by the construction of $X.$ By the construction, $g^{*}{\cal Q}_m^i=(N_m^i)\otimes \cal {O}_X=((m+1)r_i+d_i){\cal O}_X.$ Therefore we have a morphism $s$ in the diagram \[ \begin{tabular}{lll} $X\ $ \ & ${\rightarrow }_s$ & ${% \prod_{fQuot,i}(X_{-1}^i}\times _{fQuot}X_0^i)=:Y$ \\ $g\searrow $ & & $\swarrow \rho $ \\ & $fQuot$ &. \end{tabular} \] We will show $X$ and $Y$ are isomorphic finding the inverse of $s$ which complete the proof. We are given isomorphisms on $Y$ \[ \lambda _m:\rho ^{*}{\cal Q}_m^i\rightarrow (N_m^i)_Y. \] By a proposition (1,1) in \cite{St}, we get a diagram on ${\Bbb P}^1\times Y$ except $\Uparrow$ \[ \begin{array}{ccc} 0\rightarrow \pi _Y^{*}N_{-1}^i(-1)\rightarrow & \pi _Y^{*}N_0^i & \rightarrow (1\times \rho )^{*}{\cal Q}_i\rightarrow 0 \\ & \Uparrow & \uparrow \\ 0\rightarrow \pi _Y^{*}N_{-1}^{i-1}(-1)\rightarrow & \pi _Y^{*}N_0^{i-1} & \rightarrow (1\times \rho )^{*}{\cal Q}_{i-1}\rightarrow 0. \end{array} \] Note here that$\;\pi _Y^{*}(\pi _Y)_{*}(1\times \rho )^{*}{\cal Q}_{i-1}=\pi _Y^{*}$ $\rho ^{*}{\cal Q}_0^{i-1}\cong \pi _Y^{*}(N_0^{i-1})_Y.$ Since a morphism from free sheaf is determined and can be defined by a morphism between the space of global sections, there exists a unique lifting as indicated by the vertical arrow $\Uparrow $. By the defining property of $% X$, there is an induced morphism which is the inverse of $s$. Since $X$ is an irreducible smooth affine quasi-variety and $g$ is smooth of relative dimension $\sum_{i=1}^l(d_i^2+(r_i+d_i)^2),$ $R$ is smooth and irreducible. It's dimensions is $d_1(n-r_2)+d_2(r_1-r_3)+\cdots +d_{l-1}(r_{l-2}-r_l)+d_lr_{l-1}$ $+nr_1+r_1r_2+r_2r_3+\cdots +r_{l-1}r_l-r_1^2-r_2^2-\cdots -r_l^2.$ The rationality will be from Bialynski-Birula's theorem after considering an action \cite{St}. The action will be studied in the following section 3. $\Box $ \subsection{Gromov-Witten invariants and flag-Quot schemes} From now on we will work over the complex number field ${\Bbb C}$ to consider complex manifolds. We shall recall the definition of Gromov-Witten invariants for homogeneous projective variety $X$. The variety is always smooth. Denote by $% \overline{{\cal M}}_N(X,d)$ the moduli stack of stable maps of degree $d$ and genus $0$. The stack is represented by a smooth compact oriented orbifold. The same notations will be taken for the stack and the coarse moduli orbifold. It has morphisms, contraction $\pi ^X$ and evaluations $ev_i$ at the $i$-th marked point: \[ \begin{array}{ccc} \overline{{\cal M}}_N(X,d) & {\rightarrow }_{ev_i}X & \\ \downarrow _{\pi ^X} & & \\ \overline{{\cal M}}_N & & \end{array} \] where $\overline{{\cal M}}_N$ is the coarse moduli space of stable $n$ marked points of genus zero. $\overline{{\cal M}}_N$ is a smooth compact oriented manifold. The (tree level) Gromov-Witten classes $I_{N,d}^X:H^{*}(X, \QQ)^{\otimes N}\rightarrow H^{*}(\overline{{\cal M}}_N,\QQ )$ are defined as follows: \[ I_{N,d}^X(a_1\otimes \cdots \otimes a_N):=(\pi ^X)_{!}(ev_1^{*}(a_1)\otimes \cdots \otimes ev_N^{*}(a_N)). \] In the sequel we are interested in $I_{N,d}^X(a_1\otimes \cdots \otimes a_N)[p] \in \QQ$ where $[p]$ is the homology class defined by a point $p$ in $\overline{{\cal M}}_N$. If we choose any $N$ ordered distinct points $p_i$ in $\PP ^1$, we can naturally embed $Mor_d(\PP ^1, X)$ into $(\pi ^X)^{-1}(p)$ for any generic point $p$. The boundary $(\pi ^X)^{-1}(p)\smallsetminus Mor_d(\PP ^1, X)$ does not matter much, namely \begin{proposition} Let $Y_i$ be Schubert subvarieties of $X.$ Then $% \bigcap_{i=1}^Nev_i^{-1}(g_iY_i)=\bigcap_{i=1}^Nev_i^{-1}(g_iY_i)\cap Mor_d(% {\Bbb P}^1,X)$ for generic $g_i.$ \end{proposition} \proof The proof follows from the following general setting.% $\Box$ \begin{proposition} $M$ be open subvariety of a variety $\bar{M}.$ Suppose a connected algebraic group $G$ acts transitively on another variety $X$. Given a morphism $f:\bar{M}\rightarrow X$ subvarieties $Y_i$ of pure dimension, for generic $g_i\in G,$ \[ \bigcap_if^{-1}(g_iY_i)=(\bigcap_if^{-1}(g_iY_i))\cap M \] provided dimensional condition $\sum_i\codim Y_i=\dim\bar{M}$. \end{proposition} \proof Apply Kleiman's theorem in Fulton's Book \cite{Fu}. For generic $g_i$, $(\bar{M}\backslash M)\cap \bigcap_if^{-1}(g_iY_i)=\emptyset $ and for generic $g_i,$ $% (\bigcap_if^{-1}(g_iY_i))\cap M$ is proper. Hence for generic $g_i,$ $% (\bigcap_if^{-1}(g_iY_i))\cap M$ is proper and ($\bar{M}\backslash M)\cap \bigcap_if^{-1}(g_iY_i)=\emptyset $. $\Box$ Since $Mor_d(\PP ^1, X) $ is a nonsingular quasi-projective variety, we can consider its Chow group $A_*(Mor_d(\PP ^1,X))$ with products. Let us use the same notation $ev_i$ for the restriction of the evaluation map to $Mor_d(\PP ^1 ,X)$. For $[ev^{-1}_1(Y_1)]\cdot ...\cdot [ev^{-1}_1(Y_1)] \in A_0(Mor_d(\PP ^1 ,X)$, in $\ZZ $ is \[ \int _{Mor_d(\PP ^1 ,X)} [ev^{-1}_1(Y_1)]\cdot ...\cdot [ev^{-1}_1(Y_1)] \] after summing up the coefficients of cycles of points in $Mor_d(\PP ^1 ,X)$, which is well-defined in these intersections. It is equal to $I_{N,d}^X(a_1\otimes \cdots \otimes a_N)$ for the Poincare dual classes $a_i$ of $Y_i$ because $(\pi ^X)^{-1}(p)$ is a projective variety and has a resolution of singularities to avoid the intersection theory of algebraic (smooth) stacks. We would like to do a similar thing in $Quot$ schemes following Bertram \cite{Be}. \begin{proposition} (c.f. Bertram) Suppose $Y\subset X\,$ is an irreducible subvariety of codimension $c$ and suppose $Z\subset Mor_d({\Bbb P}^1,X)$ is an irreducible subvariety. Then for any $p\in {\Bbb P}^1$ and a generic translate $g,$ the intersection $Z\cap ev_p^{-1}(gY)$ is either empty or has codimension $c$ in $Z$ where $ev_p$ denoted the evaluation map at $p$. \end{proposition} \proof Apply Kleiman's theorem in Fulton's Book \cite{Fu}. $\Box$ \begin{corollary} \label{indexco}Let $c_i=$co$\dim _XY_i$ in the setting of the above definition. If $\sum_{i=1}^Nc_i>\dim (Mor_d(C,X)),$ then, for generic elements $g_1,...,g_N$ , $\bigcap_{i=1}^Nev_{p_i}^{-1}(g_iY_i)=\emptyset .$ If $\sum_{i=1}^Nc_i=\dim (Mor_d(C,X)),$ then, for generic elements $% g_1,...,g_N$ , $\bigcap_{i=1}^Nev_{p_i}^{-1}(g_iY_i)$ is isolated or empty. \end{corollary} The points $p_i$ in the above could not be distinct. \bigskip Let $({\cal E}_i)_p$ be the restriction of the sheaf ${\cal E}_i$ at $p$ in $% {\Bbb P}^1.$ Consider a commutative diagram \[ \begin{array}{ccc} Mor({\Bbb P}^1{\bf ,}Fl) & \rightarrow & Mor({\Bbb P}^1,Gr(n,r_i)) \\ \downarrow ev_p & & \downarrow ev_p \\ Fl & \rightarrow & Gr(n,r_i) \end{array} \] where $ev_p$ is the evaluation map at $p\in {\Bbb P}^1.$ Let $W$ be the subspace of $V^{*}$ used for defining $Z,$ i.e., the special Schubert varieties associated to $W$. Let $V_d(p,Z)$ be the degenerate locus of the sheaf homomorphism $W\bigotimes {\cal O}_{Quot}\rightarrow ({\cal E}_i)_p^{*}.$ In this setting we have \begin{proposition} $V_d(p,Z)$ represents the ($s_i+1-\dim (W))$-th Chern class of $({\cal E}_i)_p^*$ over $fQuot$. \end{proposition} \proof When the flag variety $Fl$ is a Grassmannian, it is proven by A. Bertram \cite{Be}. For the general case, just consider the morphism $fQuot\rightarrow Quot_i$ from the embedding $fQuot\rightarrow \prod Quot_i$ followed by the projection $\prod Quot_i\rightarrow Quot_i$. It is smooth since both schemes are smooth and the induced homomorphism between tangent spaces is surjective after looking at \ref{ww} in the proof of the theorem \ref{thm4}. This implies the degeneracy locus has the expected dimension and $[V_d(p,Z)]$ in the Chow ring of the smooth projective variety $fQuot$ is the $(\codim V_d(p,Z)$)-th Chern class of $({\cal E}_i)_p^*$ which complete the proof. $\Box$ Using a Pl\"{u}cker embedding and a stratification $Quot_d({\Bbb P}^1,{\Bbb P}% ^n)=\coprod_{0\leq m\leq d}C_m\times Mor_{d-m}({\Bbb P}^1,{\Bbb P}^n)$ (by locally closed schemes) where $C_m$ is the $m$-th symmetric product of ${\Bbb % P}^1,$ one can extend Bertram's result for flag varieties. \begin{proposition} \label{Ber}Let $Z_i$ be a special Schubert variety with $c_i$ codimension $% \leq s_{k_{i+1}}-s_{k_{i-1}}-1$ representing a Chern class of $S_{k_i}^{*}.$ Suppose $\sum_{i=1}^Nc_i\geq \dim (fQuot),$ then $% \bigcap_{i=1}^Nev_{d,p_i}^{-1}(g_iZ_i)=\bigcap_{i=1}^NV_d(p_i,g_iZ_i)$ for distinct points $p_i\in {\Bbb P}^1$. \end{proposition} \proof We will use induction on the total degree $% |d|=d_1+\cdots +d_l.$ When $|d|=0,$ it can be done by Kleiman's theorem. Let $\widetilde{C}_m=\prod_{i=1}^lC_{m_i},$ where $m=(m_1,...,m_l)$ is a multi-index$.$ Using the Pl\"{u}cker morphisms on Quot schemes, let us consider the morphism \begin{eqnarray*} J &:&fQuot\hookrightarrow \prod_{i=1}^lQuot_{d_i}(s_i,n)\rightarrow \prod_{i=1}^lQuot_{d_i}({\Bbb P}^1{\Bbb ,P}^{M_i}) \\ &=&Mor_d({\Bbb P}^1 ,\prod_{i=1}^l{\Bbb P}^{M_i})\cup \bigsqcup_{|m|=1}^{|d|}\widetilde{C}_m\times Mor_{d-m}({\Bbb P}^1 ,% \prod_{i=1}^l{\Bbb P}^{M_i}). \end{eqnarray*} We would like to show that \[ \bigcap_{i=1}^NV_d(p_i,g_iZ_i)\cap J^{-1}\left( \bigsqcup_{|m|=1}^{|d|}% \widetilde{C}_m\times Mor_{d-m}({\Bbb P}^1 ,\prod_{i=1}^l{\Bbb P}% ^{M_i})\right) =\emptyset . \] Then we are done. To do so one has to show, for each $m$, $|m|>0$, \begin{equation} \emptyset =\bigcap_{i=1}^NV_d(p_i,g_iZ_i)\cap J^{-1}\left( \widetilde{C}% _m\times Mor_{d-m}({\Bbb P}^1 ,\prod_{i=1}^l{\Bbb P}^{M_i})\right). \label{inter} \end{equation} For any subset $P$ of $\{p_1,...,p_N\}$ let \begin{eqnarray*} A_P &=&\{\text{quotient sheaves }Q=(Q_1,...,Q_l)\in \text{LHS of (\ref{inter}) } \\ |\text{ }\dim _{k(p_i)}(Q_{k_i})_{p_i}\otimes k(p_i) &<&r_{k_i}\text{ iff }% p_i\in P\} \end{eqnarray*} \[ A_P\subset \bigcap_{p_i\notin P}J^{-1}(\widetilde{C}_m\times ev_{d-m,p_i}^{-1}(g_iZ_i)). \] But for generic $g_i,\;\bigcap_{p_i\notin P}J^{-1}(\widetilde{C}_m\times ev_{d-m,p_i}^{-1}(g_iZ_i))=\emptyset $ since $\bigcap_{p_i\notin P}ev_{d-m,p_i}^{-1}(g_iZ_i)=\emptyset $ by dimension counting in $% fQuot(d-m;s_1,...,s_l;n)$: \begin{eqnarray*} \sum_{p_i\notin P}\codim (ev_{d-m,p_i}^{-1}(g_iZ_i)) &\geq &\dim (fQuot_d(Fl)) \\ &&-\sum_{p_i\in P}(s_{k_{i+1}}-s_{k_{i-1}}-1) \\ &&\dim (fQuot_d(Fl)) \\ &&-\sum_{p_i\in P}(s_{k_{i+1}}-s_{k_{i-1}}). \\ &\geq &\dim (fQuot_{d-m}(Fl)). \end{eqnarray*} Since $p_i$ are distinct, we conclude the last inequality above. $\Box$ The proof of the theorem \ref{thm2} follows from what are done. \section{A Formula by Localization} \subsection{Equivariant action on ${\cal E}_i\rightarrow fQuot\times {\PP}% ^1$} By the standard action of $SL(n)\times PGL(2)$ on $V\times \PP ^1$, the group acts on the space of stalks of $V\otimes \cal {O}_{fQuot\times \PP ^1}$ and hence on the subsheaves $\cal {E}_i$ and $fQuot$. The action on the sheaves is equivariant. In particular the maximal complex torus action of $T\times \CC ^{\times}$ will formulate integrations of wedges products of Chern classes of $(\cal {E}_i)_p$ as certain finite sums of characters using the localization theorem \cite{AB}. For simplicity of notations let us do it for the Quot schemes $Quot$. Consider the action by $T\times\CC ^{\times}$ on $Quot\times {\PP}^1$, then the action has a lift on the total space of the vector bundle ${\cal E}$. Let ${\cal E}% _p$ be the restriction of the sheaf ${\cal E}$ at $p$ in ${\PP}^1$. The action has the lifting to vector bundles $\cal{E}_0$ and $\cal{E}_\infty$. It means $\cal{E}_{0(\infty )}$ is an equivariant vector bundle and its equivariant Chern classes can be considered. For other points, say $p$, transitive $% PSL(2)$-action on ${\PP}^1$ will show $\cal{E}_0$, $\cal{E}_\infty$, and $\cal{E}_p$ are isotropic: \[ \begin{array}{ccc} {\cal E} & & {\cal E} \\ \downarrow & & \downarrow \\ Quot\times {\PP}^1 & \rightarrow _g & Quot\times {\PP}^1 \end{array} \begin{array}{ccc} {\cal E}_0 & & {\cal E}_p \\ \downarrow & & \downarrow \\ Quot & \rightarrow _g & Quot \end{array} \] where $g\cdot 0=p.$ In particular the Chern classes of ${\cal E}_p\,$ are independent to $p$ since the map induced by $g$ is homotopic to identity. Let $\frac 12\hbar \;($resp. $\lambda _i)$ is (are) the ${\CC}^{\times }\;( $resp. $T)\;$characteristic classes. Then, \begin{eqnarray} \int_{Quot}\phi (c_{i_1}({\cal E}_{p_1}),...,c_{i_m}({\cal E}_{p_m})) &=&\int_{Quot}\phi (c_{i_1}({\cal E}_0),...,c_{i_m}({\cal E}_0)) \nonumber \\ &=&[\text{push forward of }\phi \text{ of equivariant classes } \nonumber \\ &&\text{of }(c_1,...,c_r)\text{ at }{\cal E}_0]_{\hbar =\lambda _i=0} \nonumber \\ &=&[\text{localization into components $P$ of the fixed subset}, \nonumber \label{local} \\ &&\text{ i.e.,}\;\sum_P\text {push forward} \frac{i_{*}^P\phi }{E(v_P)}]_{\hbar =\lambda _i=0} \label{local}, \end{eqnarray} where $i^P$ is the inclusion $P\subset Quot$ and $E(v_P) $ is the equivariant Euler class of the normal bundle of P in $Quot$. The last expression is independent to $\hbar $ and $\lambda _i$, without letting them zeros, if the quasi-homogeneous degree of $\phi $ given by degrees of the Chern classes agrees the dimension of $Quot$. The complete analog hold for $fQuot$. It is easy to see that the fixed subset consists of finite points. Therefore $E(v_P)$ in (\ref{local}) is the equivariant Euler class of the normal space over the point $P\in Quot.$ It is the product of the complex characters of the representation of $T\times\CC ^{\times}$ in the irreducible complex one dimensional subspaces of the tangent space of $% Quot$ at $p.$ In the following subsection, we devote ourselves to spell out the all fixed points and all characteristics of the representation to finish the proof of the theorem \ref{thm3}. \subsection{Computation} Let us use the standard maximal torus $T\times \CC ^{\times}$ in the picture of flag manifolds $Fl$ and the projective line ${\PP}^1.$ Fix a sequence of $(e_{k_1},e_{k_2},...,e_{k_l})$ where $\{e_i\}_{i=1}^n$ is the standard basis of $V.$ Then, for data (\ref{integer}) in introduction, one may associate a flag of subsheaves \[ {\cal O}(-d_{i,j})\longrightarrow {\cal O}(-d_{i+1,j}) \] by the global section \[ x^{(a_{i,j}-a_{i+1,j})}y^{(b_{i,j}-b_{i+1,j})}. \] It is a fixed point by the action. For such a (\ref{integer}) and a sequence, we can associate any fixed point in $fQuot$. We have found all fixed points. Note that the tangent space at $x$ of a scheme $X$ is the first order infinitesimal deformation $Mor_x(D,X),$ the set of all morphisms sending the closed point of Spectrum of the ring $D$ of dual numbers to $x$. Therefore, at a subsheaf ${\cal S}$ over ${\PP}$ of $% V\otimes\cal{O}_{{\PP}^1},$ the tangent space of Quot schemes is the set of flat families of quotient sheaves over the Spec$D$ whose fiber over the closed point of Spec$D$ is ${\cal S}$. It is $Hom({\cal S},V\otimes\cal{O}_{{\PP}^1}/{\cal S}).$ For the flag-Quot scheme consider the following equivariant short exact sequence at a fixed point ${\cal S}$ of $fQuot$ \begin{eqnarray*} 0 &\rightarrow &T_{{\cal S}}fQuot\left( d_1,...,d_l;s_1,...,s_l;n\right) \\ &\rightarrow &T_{{\cal S}}\{Quot_{d_1}(s_1,n)\times \cdots \times Quot_{d_l}(s_l,n)\}\rightarrow \prod_{i=1}^lHom({\cal S}_i,{\cal Q}% _{i+1})\rightarrow 0. \end{eqnarray*} At the fixed point associated to \ref{tang1} we find all characters of irreducible subspace of $T_{{\cal S}}Quot$ by the torus action. They are, for all $1\leq i\leq l,$ \begin{eqnarray*} (p-a_{i,j})\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }0 &\leq &p\leq a_{i,j^{\prime }}-1,\;1\leq j,j^{\prime }\leq s_i, \\ (b_{i,j}-p)\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }0 &\leq &p\leq b_{i,j^{\prime }}-1,\;1\leq j,j^{\prime }\leq s_i, \\ (p-a_{i,j})\hbar +\lambda _{m}-\lambda _{j}.\;\text{for }0 &\leq &p\leq d_{i,j},\;1\leq j\leq s_i,\;s_i+1\leq m\leq n, \end{eqnarray*} from \begin{eqnarray*} &&\bigoplus_{i=1}^lHom({\cal S}_i,{\cal Q}_i) \\ &=&\bigoplus_{i=1}^l[\bigoplus_{j=1,j^{\prime }=1}^{s_i,s_i}Hom(.x^{a_{i,j}}y^{b_{i,j}}{\cal O}_{j}(-d_{i,j}),{\cal O}% _{j^{\prime }}/x^{a_{i,j^{\prime }}}y^{b_{i,j^{\prime }}}\cal{O}_j) \\ &&\bigoplus \bigoplus_{j=1,m=s_i+1}^{s_j,n}Hom(x^{a_{i,j}}y^{b_{i,j}}{\cal O% }_{j}(-d_{i,j}),{\cal O}_{m})]. \end{eqnarray*} Characters from $\prod_{i=1}^{l-1}Hom({\cal S}_i,{\cal Q}_{i+1})$ are, for $% 1\leq i\leq l-1,$% \begin{eqnarray*} (p-a_{i,j})\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }0 &\leq &p\leq a_{i+1,j^{\prime }}-1,\;1\leq j,\leq s_i,\;1\leq j^{\prime }\leq s_{i+1}, \\ (b_{i,j}-p)\hbar +\lambda _{j^{\prime }}-\lambda _{j},\;\text{for }0 &\leq &p\leq b_{i+1,j^{\prime }}-1,\;1\leq j,\leq s_i,\;1\leq j^{\prime }\leq s_{i+1}, \\ (p-a_{i,j})\hbar +\lambda _{m}-\lambda _{j},\;\text{for }0 &\leq &p\leq d_{i,j},\;1\leq j\leq s_i,\;\;s_{i+1}+1\leq m\leq n. \end{eqnarray*} The fiber space of ${\cal S}_i$ at the point $0$ has characters \[ a_{i,j}\hbar +\lambda _{j} \] for $0\leq j\leq s_i.$ Therefore the $k$-th Chern character is the $k$-th symmetric function in those characters. Let us denote it by $\sigma _i^k$. The proof of theorem \ref{thm3} follows from the proposition 5. \subsection{Projective spaces} In this section we will relate the our result to the residue formula of intersection pairing in \cite{Ki} for projective spaces. The author does not know for the other cases. Let $x$ be the Chern class of ${\cal O}_{{\PP}^n}(-1).\,$Then the Gromov-Witten invariant $I^{\PP ^n}_{N,d}(x^{\otimes (n+1)d+n})$ is \begin{equation} \sum\Sb 0\leq i\leq n \\ 0\leq k\leq d \endSb \frac{(\lambda _i+k\hbar )^{(n+1)d+n}}{\prod\Sb 0\leq p\leq d \\ p\ne k \endSb ((p-k)\hbar )\prod\Sb 0\leq q\leq d \\ 0\leq j\ne i\leq n \endSb ((q-k)\hbar +\lambda _j-\lambda _i)} \label{proj} \end{equation} \begin{proposition} $\sum_{N=0}^{N=\infty } \frac {1}{N!}q^d I^{\PP ^n}_{N.d}(x^{\otimes N})$ is a global residue \[ \frac 1{2\pi }\oint \frac{f(x)dx}{x^{n+1}-q} \] where $q$ is a formal variable. \end{proposition} \proof The identity \begin{eqnarray*} \frac 1{2\pi }\oint \frac{x^{(n+1)d+n}dx}{x^{(n+1)(d+1)}} &=&\left[ \frac 1{2\pi }\oint \frac{x^{(n+1)d+n}}{\prod\Sb 0\leq i\leq n \\ 0\leq k\leq d \endSb (x-\lambda _i-k\hbar )}\right] _{\lambda _i=\hbar =0} \\ &=&(\text{\ref{proj}}) \end{eqnarray*} implies the proof. $\Box$
1996-02-05T06:20:18
9512
alg-geom/9512016
en
https://arxiv.org/abs/alg-geom/9512016
[ "alg-geom", "math.AG" ]
alg-geom/9512016
Goncharov
Alexander Goncharov
Deninger's conjecture on $L$-function of elliptic curves at $s=3$
LaTeX
null
null
null
null
I compute explicitly the regulator map on $K_4(X)$ for an arbitrary curve $X$ over a number field. Using this and Beilinson's theorem about regulators for modular curves ([B2]) I prove a formula expressing the value of the $L$-function $L(E,s)$ of a modular elliptic curve $E$ over $\Bbb Q$ at $s=3$ by the double Eisenstein-Kronecker series.
[ { "version": "v1", "created": "Mon, 25 Dec 1995 20:28:39 GMT" }, { "version": "v2", "created": "Fri, 2 Feb 1996 18:04:47 GMT" } ]
2008-02-03T00:00:00
[ [ "Goncharov", "Alexander", "" ] ]
alg-geom
\section{Appendix} {\bf 1. Proof of theorem \ref{z2}b)}. Let me remind the formulation of this theorem {\bf Theorem \ref{z2}} a) $f_{4}(3)$ and $f_{5}(3)$ {\it do not depend on the choice of $\omega$. b) The homomorphisms $f_*(3)$ provide a morphism of complexes. } Proof. a) See the proof of similar results in chapter 3 of [G2]. b) We have to prove that $f_4(3) \circ d = \delta \circ f_5(3)$ and $f_5(3) \circ d = \delta \circ f_6(3)$. For the first result see chapter 3 in [G2]. The second one is much more subtle. As pointed out H.Gangl, the geometric proof given in [G2] (see theorem 3.10 there) has some errors. Namely, in lemma 3.8 $r = -r_3$ but not $r=r_3$ as clamed, and as a result the proof of theorem 3.10 become more involved; further, the correct statement in theorem 3.10 is $f_5(3) \circ d = \delta \circ 1/15 \cdot f_6(3)$ (the coefficient $1/15$ in the definition of $f_6(3)$ was missed). Another proof was given in [G1]. It was actually the first proof of the statement b). However in this proof we used a different definition for homomorphism $f_6(3)$ (the map $M_3$ in [G1]). Moreover the proof was rather complicated and the relation between the homomorphisms $f_6(3)$ and $M_3$ not easy to see. Therefore I will present in detail a completely different proof togerther with some corrections to chapter 3 in [G2]. Let us suppose that in a three dimensional vector space $V_3$ we choose a volume form $\omega$. Then for any two vectors $a,b$ one can define the cross product $a \times b \in V_3^*$ as follows: $<a \times b, c>: = \Delta(a,b,c)$. The volume form $\omega$ defines the dual volume form in $V_3^*$, so we can define $\Delta(x,y,z)$ for any three vectors in $V_3^*$. \begin{lemma} \label{gz2} For any $6$ vectors in generic position $a_1,a_2,a_3,b_1,b_2,b_3$ in $V_3$ $$ \Delta(a_1,a_2,b_1) \cdot \Delta(a_2,a_3,b_2) \cdot \Delta(a_3,a_1,b_3) - \Delta(a_1,a_2,b_2) \cdot \Delta(a_2,a_3,b_3) \cdot \Delta(a_3,a_1,b_1) = $$ $$ \Delta(a_1,a_2,a_3) \cdot \Delta(a_1 \times b_1,a_2 \times b_2,a_3 \times b_3) $$ \end{lemma} {\bf Proof}. The left hand side is zero if the vectors $a_1,a_2,a_3$ are linearly dependent. So $\Delta(a_1,a_2,a_3)$ divides it. Similarly the left hand side is zero if $a_i$ is collinear to $b_i$ or $\alpha_1 a_1 + \beta_1 b_1 = \alpha_2 a_2 + \beta_2 b_2= \alpha_3 a_3 + \beta_3 b_3$ for some numbers $\alpha_k, \beta_k$. This implies that $\Delta(a_1 \times b_1,a_2 \times b_2,a_3 \times b_3)$ also divides the left hand side. It is easy to deduce the formula from this. However it perhaps easier to check the formula directly. Consider the following special configuration of vectors: $$ \begin{array} {cccccc} a_1&a_2&a_3&b_1&b_2&b_3\\ -&-&-&-&-&-\\ 1&0&0&x_1&y_1&z_1\\ 0&1&0&x_2&y_2&z_2\\ 0&0&1&x_3&y_3&z_3 \end{array} $$ Then the left hand side is equal to $x_3 y_1 z_2 - y_3 z_1 x_2$, and the computation of the right hand side gives the same result. The lemma is proved. {\bf Remark}. Let $a_1,...,a_n,b_1,...,b_n$ be a configuration of $2n$ vectors in an $n$-dimensional vector space $V_n$. Set $\Delta({\hat a}_n,b_1):= \Delta(a_1,...,a_{n-1},b_1)$ and so on. Then $$ \Delta({\hat a}_1,b_1) \cdot ... \cdot \Delta({\hat a}_{n},b_n) - \Delta({\hat a}_1,b_2) \cdot ... \cdot \Delta({\hat a}_n,b_1) = $$ $$ \Delta(a_1,...,a_n) \cdot \Delta(a_1 \times ... \times a_{n-2} \times b_n, ... ,a_n \times ... \times a_{n-3} \times b_{n-1}) $$ Notice that $f_5(3) \circ d - \delta \circ f_6(3) \in B_2(F) \otimes F^*$. There is a homomorphism $$ \delta \otimes id: B_2(F) \otimes F^* \longrightarrow \wedge^2F^*\otimes F^*, \qquad \{x\}_2 \otimes y \longmapsto (1-x) \wedge x \otimes y $$ The crucial step of the proof is the following \begin{proposition} \label{gz3} $$ (\delta \otimes id) \circ \Bigl (f_5(3) \circ d - \delta \circ f_6(3)\Bigr)(v_1,...,v_6) =0 \quad \mbox{in} \quad \wedge^2 F^* \otimes F^* $$ \end{proposition} {\bf Proof}. We will use notation $\Delta ( i,j,k) $ for $\Delta ( v_i, v_j, v_k)$. According to lemma (\ref{gz2}) $$ 1 - \frac{\Delta ( 1, 2, 4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)} = \frac{\Delta (1,2,3) \Delta (v_1 \times v_4, v_2\times v_5, v_3\times v_6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)} $$ Using the cyclic permutation $1-> 2->3 ->1, 4->5->6->4$ we see that one has to calculate the element $$ 3 \cdot {\rm Alt}_{6}\left\{ \frac{\Delta (1,2,4) \Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)} \wedge \frac{\Delta (1,2,3) \Delta (v_1 \times v_4, v_2\times v_5, v_3\times v_6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)} \otimes \frac{\Delta(1,2,4)}{\Delta(1,2,5)}\right\} $$ in $\wedge^2 F^* \otimes F^*$. Let us do this. We will compute first the contribution of the factor $\otimes \Delta(1,2,4)$. What we need to find is $$ {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2, 5)\Delta (2,3,6)\Delta (3,1,4)} \wedge \frac{\Delta (1,2,3) \Delta (v_1 \times v_4, v_2\times v_5, v_3\times v_6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)}\right\} $$ in $\wedge^2 F^*$. Here ${\rm Alt}_{(1,2,4);(3,5,6)}$ is the skewsymmetrization with respect to the group $S_3 \times S_3$ which permutes the indices $(1,2,4)$ and $(3,5,6)$. i) Consider $$ {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2,5 )\Delta (2,3,6)\Delta (3,1,4)} \wedge \Delta (v_1 \times v_4,v_2\times v_5, v_3\times v_6)\right\} $$ Using the skewsymmetry with respect to the permutation exchanging $1$ with $3$ as well as $4$ with $ 6$ (notation: $: (13)(46)$) we see that this expression is zero. ii)Look at $$ - {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3,5) \Delta(3,1,6)} {\Delta(1,2,5)\Delta (2,3,6)\Delta (3,1,4)} \wedge \Delta (2,3,6) \otimes \Delta(1,2,4)\right\} $$ The skewsymmetry with respect to $(14)$ or with respect to $(36)$ imply that it is also zero. iii) Consider $$ - {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2, 5)\Delta (2,3,6)\Delta (3,1,4)} \wedge \Delta (1,2,5) \otimes \Delta(1,2,4)\right\} $$ The skewsymmetry with respect to the permutations $(12)$ as well as $(36)$ leads to $$ -{\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta(2,3,5)}{\Delta(1,3,4)}\wedge \Delta(1,2,5) \otimes \Delta(1 ,2,4)\right\} $$ iv) Look at the term with $\Delta(3,1,4)$: $$ - {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2, 5)\Delta (2,3,6)\Delta (3,1,4)} \wedge \Delta (3,1,4) \otimes \Delta(1,2,4)\right\} $$ Using the permutation $(14)$ we get $$ -{\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta(3,1,6)}{\Delta(1,2,5)}\wedge \Delta(1,3,4) \otimes \Delta(1 ,2,4)\right\} $$ v) Finally, using $(12)$ and $(56)$ we see that $$ {\rm Alt}_{(1,2,4);(3,5,6)}\left\{\frac{\Delta (1,2,4)\Delta (2,3, 5) \Delta(3,1,6)} {\Delta(1,2, 5)\Delta (2,3,6)\Delta (3,1,4)} \wedge \Delta (1,2,3) \otimes \Delta(1,2,4)\right\} = $$ $$ - 3 \cdot {\rm Alt}_{(1,2,4);(3,5,6)} \left\{\Delta(1,3,5) \wedge \Delta(1,2,3) \otimes \Delta(1 ,2,4)\right\} $$ Therefore we get $$ {\rm Alt}_{(1,2,4);(3,5,6)}\Bigl( \Delta(1,2,5) \wedge \frac{\Delta(2,3,5)}{\Delta(1,3,4)} + \Delta(1,3,4) \wedge \frac{\Delta(1,3,6)}{\Delta(1,2,5)} + $$ $$ 3 \cdot \Delta(1,2,3) \wedge \Delta(1,3,5) \Bigr)\otimes \Delta(1,2,4) = $$ $$ {\rm Alt}_{(1,2,4);(3,5,6)} \Bigl(\Delta(1,2,5) \wedge \Delta(2,3,5) + \Delta(1,3,4) \wedge \Delta(1,3,6) + $$ $$ 3 \cdot \Delta(1,2,3) \wedge \Delta(1,3,5) \Bigr)\otimes \Delta(1,2,4) = $$ $$ 5 \cdot {\rm Alt}_{(1,2,4);(3,5,6)} 1759 \Delta(1,2,3) \wedge \Delta(1,3,5) \otimes \Delta(1,2,4) $$ The computation of the contribution of $\Delta(1,2,5)$ goes similarly and gives the same answer. {\it So the total result of our computation is} \begin{equation} \label{su2} -30 \cdot {\rm Alt}_{6} \left\{ \Delta(1,2,4) \wedge \Delta(1,4,5) \otimes \Delta(1,2,3) \right\} \end{equation} Here we get the coefficient $-30$ taking into account the action of the cyclic group of order $3$ generated by $1->2->3->1, 4->5->6->3$. Now let us compute $f_5(3) \circ d(v_1,...,v_6)$. We will use the formula \begin{equation} \label{su1} \delta \{r(v_1,v_2,v_3,v_4)\}_2 = 1/2\cdot {\rm Alt}_{4}\left\{\Delta(v_1,v_2) \wedge \Delta(v_1,v_3)\right\} \end{equation} Neglecting for a moment the constant $c, c'$ we get $$ (\delta \otimes id) \Bigl(f_5(3) \circ d(v_1,...,v_6)\Bigr) = c \cdot {\rm Alt}_{6} \{r(v_1|v_2,v_3,v_4,v_5\}_2 \otimes \Delta(1,2,3) = $$ $$ c' \cdot {\rm Alt}_{6} \Delta(1,2,4) \wedge \Delta(1,4,5)\otimes \Delta(1,2,3) $$ To justify this we used here formula (\ref{su1}) and the symmetry considerations for transpositions $i<->j$ where $1 \leq i<j\leq 3$. More careful consideration shows $c' =-2$. It remains to compare it with (\ref{su2}). That's why we need in the definition of $f_6(3)$ the coefficient $1/15$. We have proved that $$ (f_5(3) \circ d - \delta \circ f_6(3)\Bigr)(v_1,...,v_6) = \sum_{1 \leq i<j<k \leq 6 } \gamma_{i,j,k} \otimes \Delta(i,j,k) $$ where $\gamma_{i,j,k} \in B_2(F)$ and moreover $\delta(\gamma_{i,j,k}) =0$ in $\wedge^2F^*$. According to [S2] \begin{equation} \label{susl} Ker\Bigl( B_2(F) \stackrel{\delta}{\longrightarrow} \wedge^2F^*\Bigr)\otimes \Bbb Q = K_3^{ind}(F)\otimes \Bbb Q \end{equation} One knows that $K_3^{ind}(F(t)) \otimes \Bbb Q = K_3^{ind}(F)\otimes \Bbb Q$. Therefore the left hand side of (\ref{susl}) is rationaly invariant. On the other hand one can connect by a rational curve the configurations $(v_1,v_2,...,v_6)$ and $(v_2,v_1,...,v_6)$ (interchanging $v_1$ with $v_2$) in the space of all generic configurations. This implies that $\gamma(1,2,3) = \gamma(2,1,3)$ modulo torsion. But $\gamma(1,2,3) = - \gamma(2,1,3)$ modulo torsion by the skewsymmetry. So $\gamma(1,2,3) =0$ modulo torsion, and the same conclusion is valid for $\gamma(i,j,k)$. With more work one can show that $f_5(3) \circ d - \delta \circ f_6(3) = 0$ at least modulo 6-torsion, but we do not need this. Theorem is proved. {\bf 2. The geometrical definition of the homomorphism $f_6(3)$} Let $(a_{1},a_{2},a_{3},b_{1},b_{2},b_{3})$ be a configuration of 6 distinct points in $P^{2}$ as on fig.\ 1. Let $P^{2}=P(V_{3})$. Choose vectors in $V_{3}$ such that they are projected to points $a_{i},b_{i}$. We denote them by the same letters. Choose $f_{i} \in V_{3}^{\ast}$ such that $f_{i}(a_{i}) = f_{i}(a_{i+1}) = 0$. Put \begin{equation} r'_{3}(a_{1},a_{2},a_{3},b_{1},b_{2},b_{3}) = \frac{f_{1}(b_{2}) \cdot f_{2}(b_{3})\cdot f_{3}(b_{1})} {f_{1}(b_{3})\cdot f_{2}(b_{1})\cdot f_{3}(b_{2})}\; . \end{equation} The right-hand side of (3.10) does not depend on the choice of vectors $f_{i},b_{j}$. \begin{center} \begin{picture}(100,80) \put(47,78){$a_{2}$} \put(47,68){$\bullet$} \put(50,70){\vector(2,-3){40}} \put(50,70){\vector(-2,-3){40}} \put(23,31){$\bullet$}% \put(12,33){$b_{1}$} \put(91,10){\vector(-1,0){80}} \put(60,47){$\bullet$} % \put(66,50){$b_{2}$} \put(8,7){$\bullet$} % \put(87,7){$\bullet$}% \put(93,0){$a_{3}$} \put(0,0){$a_{1}$} \put(36,7){$\bullet$} \put(33,0){$c_3$} \put(63,7){$\bullet$} \put(60,0){$b_{3}$} \put(33,-20){(fig. 1)} \end{picture} \end{center} \vskip 1cm \vskip 3mm \noindent {\bf Lemma 3.8} $-r(b_{1}\vert a_{2},a_{3},b_{2},b_{3}) = r'_{3}(a_{1},a_{2},a_{3},b_{1},b_{2},b_{3})$. \vskip 3mm \noindent {\bf Proof.} The same as the one of lemma 3.8 in [G2] Now let $\hat b_3$ be the of the line $b_1b_2$ with the line $a_1a_3$. Further, let $x$ be the intersection point of the lines $a_1b_2$ and $a_3b_1$. Let us denote by $c_3$ the intersection point of the line $a_2 x$ with the line $a_1a_3$. Then \begin{equation} r'_{3}(a_{1},a_{2},a_{3},b_{1},b_{2},b_{3}) = r(a_1,a_3,c_3,b_3) \end{equation} Indeed, by the well known theorem $r(a_1,a_3,\hat b_3,b_3) =-1$. Now returning to a configuration $(v_1,...,v_6)$ (see fig 2) \begin{center} \begin{picture}(100,80) \put(47,78){$a_{2}$} \put(47,68){$\bullet$} \put(50,70){\vector(2,-3){40}} \put(50,70){\vector(-2,-3){40}} \put(23,31){$\bullet$} \put(12,33){$v_{1}$} \put(71,31){$\bullet$} \put(80,33){$v_{5}$} \put(91,10){\vector(-1,0){80}} \put(60,47){$\bullet$} \put(66,50){$v_{2}$} \put(25,50){$v_{4}$} \put(35,47){$\bullet$} \put(8,7){$\bullet$} \put(87,7){$\bullet$} \put(93,0){$a_{3}$} \put(0,0){$a_{1}$} \put(36,7){$\bullet$} \put(33,0){$v_{6}$} \put(63,7){$\bullet$} \put(60,0){$v_{3}$} \put(33,-20){(fig. 2)} \end{picture} \end{center} \vskip 1cm we see that one has proceed as follows: Put $b_1:=v_1,b_2:=v_2,b_3:=v_3$ and apply the given above definition to the configuration $(a_1,a_2,a_3,b_1,b_2,b_3)$ and then alternate. Notice that the configuration $(a_1,a_2,a_3,b_1,b_2,b_3)$ is defined by three flags $(v_1,v_1v_4),(v_2,v_2v_5),(v_3,v_3v_6)$. \vskip 3mm \noindent {\bf REFERENCES} \begin{itemize} \item[{[B1]}] Beilinson A.A.: {\it Higher regulators and values of $L$-functions}, VINITI, 24 (1984), 181--238 (in Russian); English translation: J. Soviet Math. 30 (1985), 2036--2070. \item[{[B2]}] Beilinson A.A.: {\it Higher regulators for modular curves} Contemporary Mathematics, vol. 55, 1987, 1-35. \item[{[BL]}] Beilinson A.A., Levin A.M.: {\it Elliptic polylogarithms}. Symposium in pure mathematics, 1994, vol 55, part 2, 101-156. \item[{[Bl1]}] Bloch S.: {\it Higher regulators, algebraic $K$- theory and zeta functions of elliptic curves}, Lect. Notes U.C. Irvine, 1977. \item[{[Bl2]}] Bloch S.: {\it 2 letters to Deninger regarding [D1]} Fall 1990. \item[{[Bl3]}] Bloch S.: {\it Lectures on algebraic cycles}, Duke Math. Lect. Series, 1980. \item[{[BMS]}] Beilinson A.A., MacPherson R.D. Schechtman V.V: {\it Notes on motivic cohomology}. Duke Math. J., 1987 vol 55 p. 679-710 \item[{[GGL]}] Gabrielov A.M., Gelfand I.M., Losic M.V.: {\it Combinatorial computation of characteristic classes}, Funct.\ Analysis and its Applications V. 9 No. 2 (1975) p. 103--115 and No. 3 (1975) p. 5--26 (in Russian). \item[{[G1]}] Goncharov A.B.:{\it Geometry of configurations, polylogarithms and motivic cohomology}. Advances in Mathematics, 1995. 197 - 318. \item[{[G2]}] Goncharov A.B., {\it Polylogarithms and motivic Galois groups}, Symposium in pure mathematics, 1994, vol 55, part 2, p. 43 - 96. \item[{[G3]}] Goncharov A.B., {\it Explicit construction of characteristic classes}. Advances in Soviet mathematics, 1993, vol 16, p. 169-210 (Special issue dedicated to I.M.Gelfand 80-th birthday) \item[{[G4]}] Goncharov A.B., {\it Special values of Hasse-Weil $L$-functions and generalized Eisenstein-Kronecker series}. To appear. \item[{[GL]}] Goncharov A.B., Levin A.M. {\it Zagier's conjecture on $L(E,2)$}. Preprint IHES 1995. \item[{[Del]}] Deligne P.: {\it Symbole modere} Publ. Math. IHES 1992 \item[{[D1]}] Deninger C.: {\it Higher order operations in Deligne cohomology}. Inventiones Math. 122 N1 (1995). \item[{[D2]}] Deninger C.: {\it Higher regulators and Hecke L-series of imaginary quadratic fields I} Invent. Math. 96 (1989), 1-69. \item[{[J]}] De Jeu R. {\it On $K_4^{(3)}$ of curves over number fields} Preprint 1995. \item[{[S1]}] Suslin A.A.: {\it Homology of $GL_{n}$, characteristic classes and Milnor's $K$-theory}. Springer Lecture Notes in Math. 1046 (1989), 357--375. \item[{[S2]}] Suslin A.A.: {\it $K_{3}$ of a field and Bloch's group}, Proceedings of the Steklov Institute of Mathematics 1991, Issue 4. \item[{[Z1]}] Zagier D.:{\it Polylogarithms, Dedekind zeta functions and the algebraic $K$-theory of fields}. Arithmetic Algebraic Geometry (G.v.d.Geer, F.Oort, J.Steenbrink, eds.), Prog. Math., Vol 89, Birkhauser, Boston, 1991, pp. 391--430. \item[{[Z2]}] Zagier D.:{\it The Bloch - Wigner -Ramakrishnan polylogarithm function} Math. Ann. 286 (1990), 613-624 \end{itemize} \end{document} {\bf Proof}. Let us use induction on $n-k$. One has $$ \sum_i \int_{X(\Bbb C)}d{\cal L}_{2}(f_i)\log^{n-3}|f_i|\log|g_i|\wedge \omega = -i \sum_i \int_{X(\Bbb C)}\alpha(1-f_i, f_i)\log^{n-3}|f_i|\log|g_i|\wedge \omega $$ Further, $$ \sum_i \int_{X(\Bbb C)}d{\cal L}_{3}(f_i)\log^{n-4}|f_i|\log|g_i|\wedge \omega = $$ $$ - \sum_i \int_{X(\Bbb C)}{\cal L}_{2}(f_i)d\arg|f_i|\log^{n-4}|f_i|\log|g_i|\wedge \omega +\frac{1}{3}\sum_i \int_{X(\Bbb C)}\alpha(1-f_i,f_i)\log^{n-3}|f_i|\log|g_i|\wedge \omega $$ The second term is already in the desired form. The first one can be written as follows: $$ -i\cdot \sum_i \int_{X(\Bbb C)}{\cal L}_{2}(f_i)d\log|f_i|\log^{n-4}|f_i|\log|g_i|\wedge \omega = $$ $$ -\frac{i}{n-2}\cdot \sum_i \int_{X(\Bbb C)}{\cal L}_{2}(f_i)d(\log^{n-3}|f_i|\log|g_i|)\wedge \omega = $$ $$ \frac{i}{n-2}\cdot \sum_i \int_{X(\Bbb C)}d{\cal L}_{2}(f_i)\log^{n-3}|f_i|\log|g_i|\wedge \omega $$ After this we apply tinduction. The general case is in complete analogy with this one: we use formula (\ref{ss}) for $d{\cal L}_{n-k}(f_i)$ to rewrite $$ \sum_i \int_{X(\Bbb C)} d{\cal L}_{n-k}(f_i)\log^{k-1}|f_i|\log|g_i|\wedge \omega $$ and then (\ref{3.11}), (\ref{masha}) and induction. Proposition is proved. {\bf Example 2: n=4}. Set \begin{eqnarray*} & & r_{4}(1): \{ f\}_{4}\mapsto {\cal L}_{4}(f )\\ & &r_{4}(2) : \{ f \}_{3} \otimes g \mapsto {\cal L}_{3} (f ) d \arg g -\frac{1}{3} {\cal L}_{2} (f )\log \vert g \vert \cdot d \log \vert f\vert\\ & & r_{4}(3) : \{ f \}_{2} \otimes g_1 \wedge g_2 \mapsto - {\cal L}_{2} (f ) d \arg g_1 \wedge d \arg g_2 + \frac{1}{3} \alpha(1-f,f) \cdot \\ & &\Bigl(\log \vert g_1 \vert d \arg \vert g_2\vert - \log \vert g_2 \vert d \arg \vert g_1\vert\Bigr) \quad + \frac{1}{3} {\cal L}_{2} (f ) d \log \vert g_1 \vert \wedge d \log \vert g_2 \vert\\ & &r_{4}(4) : g_{1}\wedge ... \wedge g_{4}\mapsto {\rm Alt} ( \frac{1}{6} \cdot \log \vert g_{1} \vert d \arg g_{2}\wedge d \arg g_{3}\wedge d \arg g_{4} - \\ & & \qquad -\frac{1}{6} \log \vert g_{1} \vert d\log \vert g_{2}\vert d\log \vert g_3 \vert d\arg g_4 )\in {\cal A}^{2}_{X};\\ & & d\log^{\wedge^{4}} : g_{1}\wedge ... \wedge g_{4}\mapsto d\log g_{1} \wedge ... \wedge d\log g_{4}\in \Omega^{4}_{X} \end{eqnarray*}
1995-12-14T06:20:18
9512
alg-geom/9512009
en
https://arxiv.org/abs/alg-geom/9512009
[ "alg-geom", "math.AC", "math.AG" ]
alg-geom/9512009
null
Takesi Kawasaki
On Macaulayfication of certain quasi-projective schemes
AMSLaTeX v 1.1 with amsart.sty, amscd.sty, 25 pages
null
null
null
null
The notion of Macaulayfication, which is analogous of the desingularization, was introduced by Faltings in 1978 and he construct a Macaulayfication of quasi-projective scheme whose non-Cohen-Macaulay locus is of dimension 0 or 1 by a characteristic free method. In this paper, we gave a Macaulayfication of a quasi-projective scheme whose non-Cohen-Macaulay locus is of dimension 2. Of course out method is independent of the characteristic.
[ { "version": "v1", "created": "Thu, 14 Dec 1995 04:45:18 GMT" } ]
2015-06-30T00:00:00
[ [ "Kawasaki", "Takesi", "" ] ]
alg-geom
\section{Introduction} Let $X$ be a Noetherian scheme. A birational proper morphism~$Y \rightarrow X$ of schemes is said to be a {\em Macaulayfication\/} of~$X$ if $Y$ is a Cohen-Macaulay scheme. This notion was introduced by Faltings~\cite{Faltings:78:Macaulay} and he established that there exists a Macaulayfication of a quasi-projective scheme over a Noetherian ring possessing a dualizing complex if its non-Cohen-Macaulay locus is of dimension $0$~or~$1$. Of course, a desingularization is a Macaulayfication and Hironaka gave a desingularization of arbitrary algebraic variety over a field of characteristic~$0$. But Faltings' method to construct a Macaulayfication is independent of the characteristic of a scheme. Furthermore, several authors are interested in a Macaulayfication. For example, Goto and Schenzel independently showed the converse of Faltings' result in a sense. Let $A$ be a Noetherian local ring possessing a dualizing complex, hence its non-Cohen-Macaulay locus is closed, and assume that $\dim A / {\frak p} = \dim A$ for any associated prime ideal~${\frak p}$ of~$A$. Then the non-Cohen-Macaulay locus of~$A$ consists of only the maximal ideal if and only if $A$ is a generalized Cohen-Macaulay ring but not a Cohen-Macaulay ring~\cite{Schenzel:75:einige}. When this is the case, Faltings~\cite[Satz 2]{Faltings:78:Macaulay} showed that there exists a parameter ideal~${\frak q}$ of~$A$ such that the blowing-up~$\operatorname{Proj} A[{\frak q} t]$ of~$\operatorname{Spec} A$ with center~${\frak q}$ is Cohen-Macaulay, where $t$ denotes an indeterminate. Conversely, Goto~\cite{Goto:82:blowing} proved that if there is a parameter ideal~${\frak q}$ of~$A$ such that $\operatorname{Proj} A[{\frak q} t]$ is Cohen-Macaulay, then $A$ is a generalized Cohen-Macaulay ring. Moreover, he showed that $A$ is Buchsbaum if and only if $\operatorname{Proj} A[{\frak q} t]$ is Cohen-Macaulay for every parameter ideal~${\frak q}$ of~$A$: see also \cite{Schenzel:83:standard}. Brodmann~\cite{Brodmann:83:local} also studied the blowing-up of a generalized Cohen-Macaulay ring with center a parameter ideal. Furthermore, he constructed Macaulayfications in a quite different way from Faltings. Let $A$ be a Noetherian local ring possessing a dualizing complex. We let $d= \dim A$ and $s$ be the dimension of its non-Cohen-Macaulay locus. If $s=0$, then Brodmann~\cite[Proposition 2.13]{Brodmann:83:two} gave an ideal~${\frak b}$ of height~$d-1$ such that $\operatorname{Proj} A[{\frak b} t]$ is Cohen-Macaulay. If $s=1$, then Faltings' Macaulayfication~\cite[Satz 3]{Faltings:78:Macaulay} of~$\operatorname{Spec} A$ consists of two consecutive blowing-ups $Y \rightarrow X \rightarrow \operatorname{Spec} A$ where the center of the first blowing-up is an ideal of height~$d-1$. In this case, Brodmann gave two other Macaulayfications of~$\operatorname{Spec} A$: the first one~\cite{Brodmann:80:Kohomologische} is the composite of a blowing-up $X \rightarrow \operatorname{Spec} A$ with center an ideal of height~$d-1$ and a finite morphism~$Y \rightarrow X$; the second one~\cite[Corollary 3.11]{Brodmann:83:two} consists of two consecutive blowing-ups $Y \rightarrow X \rightarrow \operatorname{Spec} A$ where the center of the first blowing-up is an ideal of height~$d-2$. In this article, we are interested in a Macaulayfication of the Noetherian scheme whose non-Cohen-Macaulay locus is of dimension~$2$. Let $A$ be a Noetherian ring possessing a dualizing complex and $X$ a quasi-projective scheme over~$A$. Then $X$ has a dualizing complex with codimension function~$v$. Furthermore the non-Cohen-Macaulay locus~$V$ of~$X$ is closed. We define a function $u \colon X \to {\Bbb Z}$ to be $u(p) = v(p) + \dim \overline{\{p\}}$. We will establish the following theorem: \begin{thm} \label{mthm} If $\dim V \leq 2$ and $u$ is locally constant on~$V$, then $X$ has a Macaulayfication. \end{thm} If $\dim V \leq 1$, then $u$ is always locally constant on~$V$. Therefore, this theorem contains Faltings' result. Furthermore, we note if $X$ is a projective scheme over a Gorenstein local ring, then $u$ is constant on~$X$. We agree that $A$ denotes a Noetherian local ring with maximal ideal~${\frak m}$ except for Section~\ref{sec:6}. Assume that $d = \dim A > 0$. We refer the reader to~% \cite{Hartshorne:66:residue,% Hartshorne:77:algebraic,% Matsumura:89:commutative,% Stuckrad-Vogel:86:Buchsbaum} for unexplained terminology. \section{Preliminaries} In this section, we state some definitions and properties of a local cohomology and an ideal transform. Let ${\frak b}$ be an ideal of~$A$. \begin{dfn} The local cohomology functor~$H_{\frak b}^p(-)$ and the ideal transform functor~$D_{\frak b}^p(-)$ with respect to~${\frak b}$ are defined to be $$ H_{\frak b}^p(-) = \mathop{\varinjlim}_m \operatorname{Ext}_A^p(A/{\frak b}^m, -) \quad \text{and} \quad D_{\frak b}^p(-) = \mathop{\varinjlim}_m \operatorname{Ext}_A^p({\frak b}^m, -), $$ respectively. \end{dfn} For an $A$-module $M$, there exists an exact sequence \begin{equation} \label{eqn:2.1.1} 0 @>>> H_{\frak b}^0(M) @>>> M @>\iota>> D_{\frak b}^0(M) @>>> H_{\frak b}^1(M) @>>> 0 \end{equation} and isomorphisms $$ D_{\frak b}^p(M) \cong H_{\frak b}^{p+1}(M) \quad \text{for all $p>0$}. $$ They induces that \begin{equation} \label{eqn:2.1.2} H_{\frak b}^p D_{\frak b}^0(M) = \begin{cases} 0, & p=0, 1; \\ H_{\frak b}^p(M), & \text{otherwise}. \end{cases} \end{equation} If ${\frak b}$ contains an $M$-regular element~$a$, then we can regard $D_{\frak b}^0(M)$ as a submodule of the localization~$M_a$ with respect to~$a$ and $\iota$ is the inclusion. It is well-known that $H_{\frak b}^p(-)$ is naturally isomorphic to the direct limit of Koszul cohomology. In particular, let ${\frak b} = (f_1, \dots, f_h)$ and $M$ be an $A$-module. Then $$ H_{\frak b}^h(M) = \mathop{\varinjlim}_m M/(f_1^m, \dots, f_h^m)M \quad \text{and} \quad H_{\frak b}^0(M) = \bigcap_{i=1}^h 0 \qtn_M \angled{f_i}, $$ where $0 \qtn \angled{f_i}$ denotes $\bigcup_{m = 1}^\infty 0 \qtn f_i^m$. Furthermore, let $A \rightarrow B$ be a ring homomorphism. Then there exists a natural isomorphism $H_{\frak b}^p(M) \cong H_{{\frak b} B}^p(M)$ for a $B$-module~$M$. The following lemma is frequently used in this article. \begin{lem}[Brodmann \cite{Brodmann:83:Einige}] \label{lem:2.2} Let ${\frak b} = (f_1, \dots, f_h)$ and ${\frak c} = (f_1, \dots, f_{h-1})$ be two ideals. Then there exists a natural long exact sequence $$ \cdots @>>> [H_{\frak c}^{p-1}(-)]_{f_h} @>>> H_{\frak b}^p(-) @>>> H_{\frak c}^p(-) @>>> [H_{\frak c}^p(-)]_{f_h} @>>> \cdots. $$ \end{lem} Next we state on the annihilator of local cohomology modules. \begin{dfn} For any finitely generated $A$-module~$M$, we define an ideal~${\frak a}_A(M)$ to be $$ {\frak a}_A(M) = \prod_{p=0}^{\dim M -1} \operatorname{ann} H_{\frak m}^p(M). $$ \end{dfn} We note that a finitely generated $A$-module~$M$ is Cohen-Macaulay if and only if ${\frak a}_A(M) = A$, and that $M$ is generalized Cohen-Macaulay if and only if ${\frak a}_A(M)$ is an ${\frak m}$-primary ideal. The notion of~${\frak a}_A(-)$ plays a key role in this article. In fact, Schenzel~\cite{Schenzel:79:dualizing} showed that $V({\frak a}_A(A))$ coincides with the non-Cohen-Macaulay locus of~$A$ if it possesses a dualizing complex and is equidimensional. He also gave the following lemma~% \cite{Schenzel:79:dualizing,Schenzel:82:cohomological}: \begin{lem} \label{lem:2.4} Let $M$ be a finitely generated $A$-module and $x_1$,~\dots, $x_n$ a system of parameters for~$M$. Then $ (x_1, \dots, x_{i-1})M \qtn x_i \subseteq (x_1, \dots, x_{i-1})M \qtn {\frak a}_A(M) $ for any $1 \leq i \leq n$. In particular, if $x_i \in {\frak a}_A(M)$, then the equality holds. \end{lem} Let $R = \bigoplus_{n \geq 0} R_n$ be a Noetherian graded ring where $R_0 = A$. A graded module~$M = \bigoplus_n M_n$ is said to be {\em finitely graded\/} if $M_n=0$ for all but finitely many~$n$. The following lemma is an easy consequence of~\cite{Faltings:78:uber}. \begin{lem} \label{lem:2.5} Let ${\frak b}$ be a homogeneous ideal of~$R$ containing~$R_+ = \bigoplus_{n>0} R_n$ and $M$ a finitely generated graded $R$-module. We assume that $A$ possesses a dualizing complex. Let $p$ be the largest integer such that, for all~$q \leq p$, $H_{\frak b}^q(M)$ is finitely graded. Then $\operatorname{depth} M\hlz{\frak p} \geq p$ for any closed point~${\frak p}$ of~$\operatorname{Proj} R$, that is, ${\frak p}$ is a homogeneous prime ideal such that $\dim R/ {\frak p} = 1$ and $R_+ \not\subseteq {\frak p}$. \end{lem} \section{A Rees algebra obtained by an ideal transform} \label{sec:3} \begin{dfn} A sequence $f_1$,~\dots, $f_h$ of elements of~$A$ is said to be a d-sequence on an $A$-module~$M$ if $ (f_1, \dots, f_{i-1})M \qtn f_i f_j = (f_1, \dots, f_{i-1})M \qtn f_j $ for any $1 \leq i \leq j \leq h$. We shall say that $f_1$,~\dots, $f_h$ is an unconditioned strong d-sequence (for short, {\em u.s.d-sequence\/}) on~$M$ if $f_1^{n_1}$,~\dots, $f_h^{n_h}$ is a d-sequence on~$M$ in any order and for arbitrary positive integers $n_1$,~\dots, $n_h$. \end{dfn} The notion of u.s.d-sequences was introduced by Goto and Yamagishi~\cite{Goto-Yamagishi::theory} to refine arguments on Buchsbaum rings and generalized Cohen-Macaulay rings. Their theory contains Brodmann's study on the Rees algebra with respect to an ideal generated by a pS-sequences~\cite{Brodmann:83:local}. But Brodmann~\cite{Brodmann:84:local} also studied the ideal transform of such a Rees algebra. The purpose of this section is to study an ideal transform of the Rees algebra with respect to an ideal generated by a u.s.d-sequence. Let $f_0$,~\dots, $f_h$ be a sequence of elements of~$A$ where $h \geq 1$ and ${\frak q} = (f_1, \dots, f_h)$. \begin{lem} \label{lem:3.2} If $f_1$,~\dots, $f_h$ be a d-sequence on~$A/ f_0A$, then $$ [(f_1, \dots, f_k) {\frak q}^n] \qtn f_0 = (f_1, \dots, f_k) [{\frak q}^n \qtn f_0] + 0 \qtn f_0 $$ for any $1 \leq k \leq h$ and $n>0$. \end{lem} \begin{pf} It is obvious that the left hand side contains the right one. We shall prove the inverse inclusion by induction on~$k$. Let $a$ be an element of the left hand side. When $k=1$, we put $f_0 a = f_1 b$ where $b \in {\frak q}^n$. By using \cite[Theorem 1.3]{Goto-Yamagishi::theory}, we obtain $b \in (f_0) \qtn f_1 \cap {\frak q}^n \subseteq (f_0)$. If we put $b = f_0 a'$, then $a' \in {\frak q}^n \qtn f_0$ and $f_0 (a - f_1 a') =0$. Thus we get $a \in f_1 [{\frak q}^n \qtn f_0] + 0 \qtn f_0$. When $k > 1$, we put $f_0 a = b + f_k c$ where $ b \in (f_1, \dots, f_{k-1}) {\frak q}^n $ and $c \in {\frak q}^n$. Then we obtain \begin{align*} c & \in (f_0, \dots, f_{k-1}) \qtn f_k \cap {\frak q}^n \\ & \subseteq (f_0) + (f_1, \dots, f_{k-1}) {\frak q}^{n-1} \end{align*} by using~\cite[Theorem 1.3]{Goto-Yamagishi::theory} again. If we put $c = f_0 a' + b'$ where $$ b' \in (f_1, \dots, f_{k-1}) {\frak q}^{n-1}, $$ then $a' \in {\frak q}^n \qtn f_0$. Thus we get \begin{align*} a - f_k a' & \in [(f_1, \dots, f_{k-1}) {\frak q}^n] \qtn f_0 \\ & = (f_1, \dots, f_{k-1}) [{\frak q}^n \qtn f_0] + 0 \qtn f_0 \end{align*} by induction hypothesis. The proof is completed. \end{pf} Let $\trans{\frak q} = {\frak q} \qtn \angled{f_0}$. If $f_0$ is $A$-regular and $f_1$,~\dots, $f_h$ is a d-sequence on~$A/ f_0^lA$ for all~$l > 0$, then Lemma~\ref{lem:3.2} assures us that \begin{equation} \label{eqn:3.2.1} {\frak q}^{n-1} \trans{\frak q} = \trans{\frak q}^n = {\frak q}^n \qtn \angled{f_0} \quad \text{for all $n>0$}. \end{equation} Therefore the Rees algebra $\trans R = A[\trans{\frak q} t]$ is finitely generated over $R = A [{\frak q} t]$. The following is an analogue of~\cite[Lemma 3.4]{Goto:82:blowing}. \begin{thm} \label{thm:3.3} Let $B = A[\trans{\frak q}/f_h] = \trans R\hlz{f_ht}$. If $f_0$ is $A$-regular and $f_1$,~\dots, $f_h$ is a d-sequence on~$A/ f_0^lA$ for all~$l > 0$, then $f_h$, $f_1/ f_h$,~\dots, $f_{h-1}/f_h$, $f_0$ is a regular sequence on~$B$. \end{thm} \begin{pf} First we note that $f_1$,~\dots, $f_h$ is a d-sequence on~$A$. In fact, by using Krull's intersection theorem, we obtain \begin{align*} (f_1, \dots, f_{i-1}) \qtn f_i f_j & = \bigcap_{l =1}^\infty (f_0^l, f_1, \dots, f_{i-1}) \qtn f_i f_j \\ & = \bigcap_{l =1}^\infty (f_0^l, f_1, \dots, f_{i-1}) \qtn f_j \\ & = (f_1, \dots, f_{i-1}) \qtn f_j \end{align*} for any $1 \leq i \leq j \leq h$. Next we show that \begin{equation} \label{eqn:3.3.1} (f_1, \dots, f_{k-1}) \qtn f_k \cap \trans {\frak q}^n = (f_1, \dots, f_{k-1}) \trans {\frak q}^{n-1}, \end{equation} for any $1 \leq k \leq h+1$ and $n>1$, where $f_{h+1} = 1$. If $a$ is an element of the left hand side, then $f_0^l a \in {\frak q}^n$ for a sufficiently large~$l$. By~\cite[Theorem 1.3]{Goto-Yamagishi::theory}, we have \begin{align*} f_0^l a & \in (f_1, \dots, f_{k-1}) \qtn f_k \cap {\frak q}^n \\ & = (f_1, \dots, f_{k-1}) {\frak q}^{n-1}. \end{align*} Lemma~\ref{lem:3.2} says $$ a \in [(f_1, \dots, f_{k-1}) {\frak q}^{n-1}] \qtn \angled{f_0} = (f_1, \dots, f_{k-1}) \trans {\frak q}^{n-1}. $$ The inverse inclusion is clear. By~\eqref{eqn:3.3.1} and \cite[Theorem 1.7]{Goto-Yamagishi::theory}, we obtain that $$ f_h, \frac{f_1}{f_h}, \dots, \frac{f_{h-1}}{f_h} $$ is a regular sequence on~$B$. Finally we shall show that $f_0$ is regular on $B/(f_h, f_1/f_h, \dots, f_{h-1}/f_h)B$. Let $\alpha \in (f_h, f_1/f_h, \dots, f_{h-1}/f_h)B \qtn f_0$. For a sufficiently large~$n>1$, we may assume $\alpha = a_0/f_h^n$ and $$ f_0 \frac{a_0}{f_h^n} = f_h \frac{a_h}{f_h^n} + \frac{f_1}{f_h} \frac{a_1}{f_h^n} + \dots + \frac{f_{h-1}}{f_h} \frac{a_{h-1}}{f_h^n} $$ where $a_0$,~\dots, $a_h \in \trans{\frak q}^n$. Therefore $$ f_h^{m+1} f_0 a_0 = f_h^m(f_h^2 a_h + f_1 a_1 + \dots + f_{h-1} a_{h-1}) $$ in~$A$ for some~$m>0$. Take an integer~$l$ such that $f_0^l a_h \in {\frak q}^n$. Then \begin{align*} f_h^{m+2} f_0^l a_h & \in (f_0^{l+1}, f_1, \dots, f_{h-1}) \cap {\frak q}^{n+m+2} \\ & = (f_0^{l+1}) \cap {\frak q}^{n+m+2} + (f_1, \dots, f_{h-1}) {\frak q}^{n+m+1} \\ & \subseteq f_0^{l+1} \trans{\frak q}^{n+m+2} + (f_1, \dots, f_{h-1}) {\frak q}^{n+m+1}. \end{align*} If we put $$ f_h^{m+2} f_0^l a_h = f_0^{l+1} b_0 + f_1 b_1 + \dots + f_{h-1} b_{h-1} $$ where $b_0 \in \trans{\frak q}^{n+m+2}$ and $b_1$,~\dots, $b_{h-1} \in {\frak q}^{n+m+1}$, then \begin{align*} f_h^{m+2} a_h - f_0 b_0 & \in [(f_1, \dots, f_{h-1}) {\frak q}^{n+m+1}] \qtn \angled{f_0} \\ & = (f_1, \dots, f_{h-1}) \trans{\frak q}^{n+m+1}. \end{align*} Let $$ f_h^{m+2} a_h - f_0 b_0 = f_1 c_1 + \dots + f_{h-1} c_{h-1} $$ where $c_1$,~\dots, $c_{h-1} \in \trans{\frak q}^{n+m+1}$. Then $$ f_0(f_h^{m+1} a_0 - b_0) \in (f_1, \dots, f_{h-1}) {\frak q}^{n+m}. $$ Therefore $$ f_h^{m+1} a_0 - b_0 \in (f_1, \dots, f_{h-1}) \trans{\frak q}^{n+m}, $$ that is, $$ \alpha - f_h \frac{b_0}{f_h^{n+m+2}} \in \left( \frac{f_1}{f_h}, \dots, \frac{f_{h-1}}{f_h} \right)B. $$ The proof is completed. \end{pf} In the rest of this section, we assume that $f_0$ is $A$-regular and that $f_1$,~\dots, $f_h$ is a u.s.d-sequence on~$A/ f_0^l A$ for all~$l>0$. Let $G = \bigoplus_{n \geq 0} {\frak q}^n/ {\frak q}^{n+1}$ and $ \trans G = \bigoplus_{n \geq 0} \trans{\frak q}^n/ \trans{\frak q}^{n+1} $ be associated graded rings with respect to~${\frak q}$ and $\trans{\frak q}$, respectively. We shall compute local cohomology modules of~$\trans G$ and~$\trans R$ with respect to~${\frak N} = (f_0, \dots, f_h)R + R_+$. \begin{thm} If $p < h+1$, then $$ [H_{\frak N}^p(\trans G)]_n = 0 \quad \text{for $n \ne 1-p$}. $$ Furthermore $$ [H_{\frak N}^{h+1}(\trans G)]_n = 0 \quad \text{for $n > -h$}. $$ \end{thm} \begin{pf} We shall prove that \begin{equation} \label{eqn:3.4.1} [H_{(f_0, f_1t, \dots, f_kt)}^p(\trans G)]_n = 0 \quad \text{for $n \ne 1-p$} \end{equation} if $p < k+1$ by induction on~$k$. It is obvious that $f_0$ is $\trans G$-regular. Therefore $H_{(f_0)}^0(\trans G) = 0$. Suppose $k > 0$. Then $H_{(f_0, f_1t, \dots, f_{k-1}t)}^p(\trans G)_{f_kt} = 0$ for $p < k$ by induction hypothesis. By Lemma~\ref{lem:2.2}, we obtain isomorphisms $$ H_{(f_0, f_1t, \dots, f_kt)}^p(\trans G) \cong H_{(f_0, f_1t, \dots, f_{k-1}t)}^p (\trans G) \quad \text{for $p<k$}. $$ Therefore \eqref{eqn:3.4.1} is proved if $p < k$. We also obtain an exact sequence $$ 0 @>>> H_{(f_0, f_1t, \dots, f_kt)}^k(\trans G) @>>> H_{(f_0, f_1t, \dots, f_{k-1}t)}^k(\trans G) @>>> H_{(f_0, f_1t, \dots, f_{k-1}t)}^k(\trans G)_{f_kt} $$ from Lemma~\ref{lem:2.2}. Hence $H_{(f_0, f_1t, \dots, f_kt)}^k(\trans G)$ is the limit of the direct system~$\{K_m\}_{m>0}$ such that $$ K_m = \frac {(f_0^m, (f_1t)^m, \dots, (f_{k-1}t)^m) \trans G \qtn \angled{f_kt}} {(f_0^m, (f_1t)^m, \dots, (f_{k-1}t)^m) \trans G} \, (m(k-1)) \quad \text{for $m > 0$} $$ and the homomorphism~$K_m \to K_{m'}$ is induced from the multiplication of $(f_0 \cdot f_1t \cdots f_{k-1}t)^{m' - m}$ for any $m' > m$. We shall show that it is the zero map except for degree~$1-k$ if $m'$ is sufficiently larger than~$m$. Let $\alpha$ be a homogeneous element of~$K_m$ of degree~$n$ and $a$ its representative. That is, $a \in \trans{\frak q}^{n+m(k-1)}$ and $$ f_k^l a \in f_0^m \trans{\frak q}^{n+m(k-1)+l} + (f_1^m, \dots, f_{k-1}^m) \trans{\frak q}^{n+m(k-2)+l} + \trans{\frak q}^{n+m(k-1)+l+1} $$ for some~$l>0$. Take an integer~$m' > m$ such that $f_0^{m' - m}\trans {\frak q} \subseteq {\frak q}$. Then $f_0^{m' - m}\trans {\frak q}^n \subseteq {\frak q}^n$ for any~$n>0$ by~\eqref{eqn:3.2.1}. By replacing $\alpha$ by its image in~$K_{m'}$, we may assume that $a \in {\frak q}^{n+m(k-1)}$ and $$ f_k^l a \in f_0^m \trans{\frak q}^{n+m(k-1)+l} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2) + l} + {\frak q}^{n+m(k-1) + l+1}. $$ We put $f_k^l a = b + c$ where $ b \in f_0^m \trans{\frak q}^{n+m(k-1)+l} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2) + l} $ and $c \in {\frak q}^{n+m(k-1)+l+1}$. Then, by using~\cite[Theorem 2.6]{Goto-Yamagishi::theory}, we obtain \begin{align*} c & \in (f_0^m, \dots, f_{k-1}^m, f_k^l) \cap {\frak q}^{n+m(k-1) + l+1} \\ & \subseteq f_0^m \trans{\frak q}^{n+m(k-1) + l+1} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2) + l+1} + f_k^l {\frak q}^{n+m(k-1) + 1}. \end{align*} If we put $c = b' + f_k^l a'$ where $ b' \in f_0^m \trans{\frak q}^{n+m(k-1) + l+1} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2) + l+1} $ and $a' \in {\frak q}^{n+m(k-1)+1}$, then $a-a'$ is also a representative of~$\alpha$. Therefore we may assume that $c=0$. By using~\cite[Theorem 2.8]{Goto-Yamagishi::theory}, we obtain \begin{align*} a & \in (f_0^m, \dots, f_{k-1}^m) \qtn f_k \cap {\frak q}^{n+m(k-1)} \\ & = (f_0^m) \cap {\frak q}^{n+m(k-1)} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2)} \\ & \quad + \sum \begin{Sb} I \subseteq \{1, \dots, k-1\} \\ \sharp I \cdot (m-1) \geq n + m(k-1) \end{Sb} \left\{ \prod_{i \in I} f_i^{m-1} \right\} \{ [(f_0^m) + (f_i \mid i \in I)] \qtn f_k \} \\ & \subseteq f_0^m \trans{\frak q}^{n+m(k-1)} + (f_1^m, \dots, f_{k-1}^m) {\frak q}^{n+m(k-2)} + {\frak q}^{n+m(k-1) + 1} \\ & \quad + \sum \begin{Sb} I \subseteq \{1, \dots, k-1\} \\ \sharp I \cdot (m-1) = n + m(k-1) \end{Sb} \left\{ \prod_{i \in I} f_i^{m-1} \right\} \{ [(f_0^m) + (f_i \mid i \in I)] \qtn f_k \} \end{align*} Here $\sharp I$ denotes the number of elements in~$I$. If $n > 1-k$, then there is no subset~$I$ of~$\{1, \dots, k-1\}$ such that $\sharp I \cdot (m-1) = n+m(k-1)$. If $n < 1-k$, then such $I$ is a proper subset. Let $j \in \{1, \dots, k-1\} \setminus I$ and $$ d \in [(f_0^m) + (f_i \mid i \in I)] \qtn f_k = [(f_0^m) + (f_i \mid i \in I)] \qtn f_j. $$ Then $$ (f_0 \cdots f_{k-1}) \left\{ \prod_{i \in I} f_i^{m-1} \right\} d \in f_0^{m+1} \trans{\frak q}^{n+(m+1)(k-1)} + (f_1^{m+1}, \dots, f_{k-1}^{m+1}) {\frak q}^{n+(m+1)(k-2)}. $$ In fact, if we put $f_j d = f_0^m e + g$ where $g \in (f_i \mid i \in I)$, then $e \in \trans{\frak q}$. Thus the image of~$\alpha$ in~$K_{m+1}$ is zero if $n \ne 1-k$. Put $k=h$. Then $$ [H_{\frak N}^p(\trans G)]_n = [H_{(f_0, f_1t, \dots, f_ht)}^p(\trans G)]_n = 0 \quad \text{for $n \ne 1 - p$} $$ if $p < h+1$. The first assertion is proved. Next we compute $H_{(f_0, f_1t, \dots, f_ht)}^{h+1}(\trans G)$. It is the limit of the direct system~$\{K'_m\}_{m>0}$ such that $$ K'_m = \trans G/ (f_0^m, (f_1t)^m, \dots, (f_ht)^m) \trans G \, (mh) \quad \text{for $m>0$} $$ and the homomorphism $K'_m \to K'_{m'}$ is induced from the multiplication of $(f_0 \cdot f_1t \cdots f_ht)^{m' - m}$ for any $m' > m$. We shall show that it is the zero map for degree $n > -h$ if $m'$ is sufficiently larger than~$m$. Let $\alpha$ be a homogeneous element of~$K'_m$ of degree~$n$ and $a$ its representative. That is, $a \in \trans{\frak q}^{n+mh}$. If $n > -h$, then $$ (f_0 \cdots f_h)^{m' - m} a \in {\frak q}^{n+m'h} = (f_1^{m'}, \dots, f_h^{m'}) {\frak q}^{n+m'(h-1)} $$ for a sufficiently larger~$m'$ than~$m$. Thus the image of~$\alpha$ in~$K'_{m'}$ is zero if $n> -h$. Therefore $[H_{\frak N}^{h+1}(\trans G)]_n = 0$ for $n > -h$. \end{pf} By this theorem, we can compute local cohomology of~$\trans R$. \begin{cor} \label{cor:3.5} If $h=1$, $2$, then $$ H_{\frak N}^p(\trans R) = 0 \quad \text{for $p\ne 1$, $h+2$} $$ and $ H_{\frak N}^1(\trans R) = [H_{\frak N}^1(\trans R)]_0 = H_{(f_0, \dots, f_h)}^1(A) $. If $h \geq 3$, then $$ H_{\frak N}^p(\trans R) = 0 \quad \text{for $p=0$, $2$, $3$} $$ and $ H_{\frak N}^1(\trans R) = [H_{\frak N}^1(\trans R)]_0 = H_{(f_0, \dots, f_h)}^1(A) $. Furthermore, if $ 4 \leq p \leq h+1$, then $$ [H_{\frak N}^p(\trans R)]_n = \begin{cases} H_{(f_0, \dots, f_h)}^{p-1}(A), & \text{for $-1 \geq n \geq 3-p$}; \\ 0, & \text{otherwise}. \end{cases} $$ \end{cor} \begin{pf} Passing through the completion, we may assume that $A$ possesses a dualizing complex. Since $H_{\frak N}^p(\trans G)$ is finitely graded for~$p < h+1$, $H_{\frak N}^p(\trans R)$ is finitely graded for~$p \leq h+1$ \cite[Proposition~3]{Marley:94:finitely}. Considering the following two exact sequences $$ 0 @>>> \trans R_+ @>>> \trans R @>>> A @>>> 0 \quad \text{and} \quad 0 @>>> \trans R_+(1) @>>> \trans R @>>> \trans G @>>> 0, $$ we obtain the assertion: see the proof of~\cite[Theorem 4.1]{Brodmann:84:local}. \end{pf} Let $S = \trans R/ R$, that is, $S = \bigoplus_{n>0} \trans{\frak q}^n / {\frak q}^n$. The following proposition shall play an important role in the next section. \begin{prop} \label{prop:3.6} If $p < h$, then $$ [H_{\frak N}^p(S)]_n =0 \quad \text{for $n \ne 1-p$.} $$ Moreover, $$ [H_{\frak N}^h(S)]_n = 0 \quad \text{for $n > 1-h$}. $$ \end{prop} \begin{pf} In the same way as the proof of Theorem~\ref{thm:3.3}, we find that $f_1$,~\dots, $f_h$ is a u.s.d-sequence on~$A$. Hence, by using \cite[Theorem 4.2]{Goto-Yamagishi::theory}, $$ [H_{(f_1t, \dots, f_ht)}^p(G)]_n = 0 \quad \text{for $n \ne -p$} $$ if $p<h$. Furthermore, $$ [H_{(f_1t, \dots, f_ht)}^h(G)]_n = 0 \quad \text{for $n> -h$}. $$ By using Lemma~\ref{lem:2.2}, we obtain $$ [H_{\frak N}^p(G)]_n = 0 \quad \text{for $n \ne 1-p$, $-p$} $$ if $p < h$ and $$ [H_{\frak N}^p(G)]_n = 0 \quad \text{for $n>1-p$} $$ if $p = h$, $h+1$. Since $ \trans{\frak q}^2 = {\frak q} \trans{\frak q} $, there exists an exact sequence $$ 0 @>>> S(1) @>>> G @>\phi>> \trans G @>>> S @>>> 0. $$ Let $T$ be the image of~$\phi$. We shall show $$ [H_{\frak N}^p(S)]_n = [H_{\frak N}^p(T)]_n = 0 \quad \text{for $n> 1-p$} $$ by induction on~$h-p$. If $p > h+1$, then the assertion is obvious. Let $p \leq h+1$. Then following two exact sequences \begin{gather*} H_{\frak N}^p(\trans G) @>>> H_{\frak N}^p(S) @>>> H_{\frak N}^{p+1}(T) @>>> H_{\frak N}^{p+1}(\trans G), \\ H_{\frak N}^p(G) @>>> H_{\frak N}^p(T) @>>> H_{\frak N}^{p+1}(S)(1) @>>> H_{\frak N}^{p+1}(G) \end{gather*} and the induction hypothesis imply $$ [H_{\frak N}^p(S)]_n = [H_{\frak N}^p(T)]_n = 0 \quad \text{for $n>1-p$}. $$ In the same way, we can prove that $$ [H_{\frak N}^p(S)]_n = [H_{\frak N}^p(T)]_n = 0 \quad \text{for $n< 1-p$} $$ if $p< h$ by induction on~$p$. \end{pf} Finally we show that $\trans R$ is an ideal transform of~$R$ in a sense. \begin{prop} $\trans R_+ = D_{(f_0, \dots, f_h)}^0(R_+)$. \end{prop} \begin{pf} We first show that $f_0$, $f_1$ is a regular sequence on~$\trans R_+$. Let $n>0$. Since $f_0$ is $A$-regular, it is also $\trans{\frak q}^n$-regular. Let $ a \in [f_0 \trans{\frak q}^n] \qtn f_1 \cap \trans {\frak q}^n $. Then $f_0^l a \in {\frak q}^n$ for a sufficiently large~$l$. Since $f_1 a \in (f_0)$, we have $ f_0^l a \in (f_0^{l+1}) \qtn f_1 \cap {\frak q}^n \subseteq f_0^{l+1} \trans{\frak q}^n $, that is, $a \in f_0 \trans{\frak q}^n$. Thus we have shown that $f_1$ is $\trans R_+ / f_0 \trans R_+$-regular. By this and \eqref{eqn:2.1.1}, we obtain \begin{equation} \label{eqn:3.7.1} D_{(f_0, \dots, f_h)}^0(R_+) \subseteq D_{(f_0, \dots, f_h)}^0(\trans R_+) = \trans R_+. \end{equation} Since $ \trans{\frak q}^n = {\frak q}^{n-1} \trans{\frak q} $ for $n \geq 2$, $(f_0^l, f_1, \dots, f_h) \trans R_+ \subseteq R_+$ for a sufficiently large~$l$. Hence, we obtain the inverse inclusion of~\eqref{eqn:3.7.1}. The proof is completed. \end{pf} \setcounter{equation}{0} \section{% A blowing-up with respect to a certain subsystem of parameters} \label{sec:4} In this section, we assume that $A$ possesses a dualizing complex. We fix an integer $s \geq \dim A/ {\frak a}_A(A)$. Since $\dim A/ {\frak a}_A(M) < \dim M$ for any finitely generated $A$-module~$M$ \cite[Korollar 2.2.4]{Schenzel:82:dualisierende}, there exists a system of parameters $x_1$,~\dots, $x_d$ for~$A$ such that \begin{equation} \label{eqn:4.0.1} \begin{cases} x_{s+1}, \dots, x_d \in {\frak a}_A(A); \\ x_i \in {\frak a}_A(A/(x_{i+1}, \dots, x_d)), & \text{for $i \leq s$}. \end{cases} \end{equation} This notion is a slight improvement of a p-standard system of parameters, which was introduced by Cuong~\cite{Cuong:91:dimension}. He also gave the statement~(1) of Theorem~\ref{thm:4.2}. The author was personally taught it by him. \begin{lem} \label{lem:4.1} Let $n_1$,~\dots, $n_i$ be arbitrary positive integers. Then \begin{multline*} (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) \qtn x_i^{n_i} \cap (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_k, \dots, x_d) \\ = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) \end{multline*} for any $1 \leq i \leq k \leq d$. \end{lem} \begin{pf} It is obvious that the left hand side contains the right one. Let $a$ be an element of the left hand side and $a = b + x_k c$ where $ b \in (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) $. Then \begin{align*} c & \in (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) \qtn x_i^{n_i} x_k \\ & = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) \qtn x_k \end{align*} by Lemma~\ref{lem:2.4}. Therefore $ x_k c, a \in (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{k+1}, \dots, x_d) $. The proof is completed. \end{pf} Let ${\frak q} = (x_{s+1}, \dots, x_d)$. Lemma~\ref{lem:2.2} assures us that $x_{s+1}$, \dots, $x_d$ is a u.s.d-sequence on~$A$. Furthermore, we have the following theorem: \begin{thm} \label{thm:4.2} \rom{(1)} The sequences $x_1^{n_1}$,~\dots, $x_s^{n_s}$, $x_{\sigma(s+1)}^{n_{s+1}}$,~\dots, $x_{\sigma(d)}^{n_d}$ is a d-sequence on~$A$ for any positive integers $n_1$,~\dots, $n_d$ and for any permutation~$\sigma$ on $s+1$,~\dots, $d$. \rom{(2)} If $s>0$, then $x_1^{n_1}$,~\dots, $x_s^{n_s}$ is a d-sequence on~$A/ {\frak q}^n$ for any positive integers $n_1$,~\dots, $n_s$ and~$n$. \end{thm} \begin{pf} (1):~Let $1 \leq i \leq j \leq d$. We have only to prove that $$ (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn x_i^{n_i} x_j^{n_j} = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn x_j^{n_j} $$ for any positive integers $n_1$,~\dots, $n_d$. If $j>s$, then the both sides are equal to $(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn {\frak a}_A(A)$. Assume that $j \leq s$ and take an element~$a$ of the left hand side. By using Lemma~\ref{lem:2.4}, we get \begin{align*} a & \in (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{j+1}, \dots, x_d) \qtn x_i^{n_i} x_j^{n_j} \\ & = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{j+1}, \dots, x_d) \qtn x_j^{n_j}. \end{align*} Hence we have \begin{align*} x_j^{n_j} a & \in (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn x_i^{n_i} \cap (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{j+1}, \dots, x_d) \\ & = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \end{align*} by repeating to use Lemma~\ref{lem:4.1}. (2):~If $n=1$, then the assertion is proved in the same way as above. Let $1 \leq i \leq j \leq s$ and $n>1$. Then $x_{s+1}$,~\dots, $x_d$ is a d-sequence on~$A/(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_i^{n_i} x_j^{n_j})$. By using Lemma~\ref{lem:3.2}, we obtain \begin{align*} \lefteqn{[(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) + {\frak q}^n] \qtn x_i^{n_i} x_j^{n_j}} \qquad \\ & = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn x_i^{n_i} x_j^{n_j} + {\frak q}^{n-1} [(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{s+1}, \dots, x_d) \qtn x_i^{n_i} x_j^{n_j}] \\ & = (x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) \qtn x_j^{n_j} + {\frak q}^{n-1} [(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}, x_{s+1}, \dots, x_d) \qtn x_j^{n_j}] \\ & \subseteq [(x_1^{n_1}, \dots, x_{i-1}^{n_{i-1}}) + {\frak q}^n] \qtn x_j^{n_j}. \end{align*} Here the second equality follows from the case of~$n=1$. Thus the proof is completed. \end{pf} In the same way as the proof of Theorem~\ref{thm:3.3}, we find that any subsequence of $x_1^{n_1}$,~\dots, $x_d^{n_d}$ is a d-sequence on~$A$ and any subsequence of $x_1^{n_1}$,~\dots, $x_s^{n_s}$ is a d-sequence on~$A/ {\frak q}^n$ for arbitrary positive integers $n_1$,~\dots, $n_d$ and~$n$. \begin{cor} Fix an integer~$k$ such that $1 \leq k \leq d$. Then $$ H_{(x_k, \dots, x_d)}^p(A) = \mathop{\varinjlim}_m \frac{(x_k^m, \dots, x_{k+p-1}^m) \qtn x_{k+p}} {(x_k^m, \dots, x_{k+p-1}^m)} \quad \text{for $p< d-k+1$}. $$ \end{cor} \begin{pf} We shall prove that $$ H_{(x_k, \dots, x_l)}^p(A) = \mathop{\varinjlim}_m \frac{(x_k^m, \dots, x_{k+p-1}^m) \qtn x_{k+p}} {(x_k^m, \dots, x_{k+p-1}^m)} \quad \text{for $p< l-k+1$} $$ by induction on~$l \geq k$. If $l=k$, then $H_{(x_k)}^0(A) = 0 \qtn_A x_k$. Suppose $l > k$. Then $x_k$,~\dots, $x_{l-1}$ is a regular sequence on~$A_{x_l}$ because $x_k$,~\dots, $x_l$ is a d-sequence on~$A$. Hence we obtain isomorphisms $$ H_{(x_k, \dots, x_l)}^p(A) \cong H_{(x_k, \dots, x_{l-1})}^p(A) \quad \text{for all $p<l-k$} $$ and an exact sequence $$ 0 @>>> H_{(x_k, \dots, x_l)}^{l-k}(A) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k}(A) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k}(A)_{x_l} $$ by Lemma~\ref{lem:2.2}. This exact sequence is the direct limit of the exact sequence $$ 0 @>>> \frac{(x_k^m, \dots, x_{l-1}^m) \qtn x_l} {(x_k^m, \dots, x_{l-1}^m)} @>>> A/(x_k^m, \dots, x_{l-1}^m) @>>> [A/(x_k^m, \dots, x_{l-1}^m)]_{x_l} $$ Thus the proof is completed. \end{pf} If $s=0$, then $\operatorname{Proj} A[{\frak q} t] \rightarrow \operatorname{Spec} A$ is a Macaulayfication of~$\operatorname{Spec} A$: see Theorem~\ref{thm:5.1} for details. In the rest of this section, we shall observe $\operatorname{Proj} A[{\frak q} t]$ when $s>0$. Assume that $s>0$ and fix an integer~$k$ such that $1 \leq k \leq s$. We shall compute local cohomology modules of $R = A[{\frak q} t]$ with respect to~$(x_k, \dots, x_{s+1})$. Let ${\frak M} = {\frak m} R + R_+$. \begin{thm} $H_{(x_k, \dots, x_{s+1})}^0(R) = 0 \qtn_A x_k$. \end{thm} \begin{pf} Since $x_k$, $x_{s+1}$,~\dots, $x_d$ is a d-sequence on~$A$, $0 \qtn_A x_k \cap {\frak q}^n = 0$ for~$n>0$ by~\cite[Theorem 1.3]{Goto-Yamagishi::theory}. That is, $$ H_{(x_k, \dots, x_{s+1})}^0({\frak q}^n) = \begin{cases} 0 \qtn_A x_k, & \text{if $n=0$}; \\ 0, & \text{otherwise}. \end{cases} $$ Therefore, $ H_{(x_k, \dots, x_{s+1})}^0(R) = \bigoplus_{n \geq 0} H_{(x_k, \dots, x_{s+1})}^0({\frak q}^n) = 0 \qtn_A x_k $. \end{pf} Let $C = A[t]/R$, that is, $C = \bigoplus_{n>0} A/ {\frak q}^n$. \begin{lem} For $k \leq l \leq s+1$ and $p \leq l-k$, the natural homomorphism $$ \alpha_l^p \colon H_{(x_k, \dots, x_l)}^p(A[t]) @>>> H_{(x_k, \dots, x_l)}^p(C) $$ is a monomorphism except for degree~$0$. \end{lem} \begin{pf} We shall work by induction on~$l$. If $l=k$, then $0 \qtn_A x_k \cap {\frak q}^n = 0$ for $n>0$. Therefore $$ \alpha_k^0 \colon 0 \qtn_{A[t]} x_k @>>> \bigoplus_{n>0} {\frak q}^n \qtn x_k / {\frak q}^n $$ is a monomorphism except for degree~$0$. Let $k < l \leq s$. Then $x_k$,~\dots, $x_{l-1}$ is a regular sequence on~$A_{x_l}$ and on~$C_{x_l}$ by Theorem~\ref{thm:4.2}. By using Lemma~\ref{lem:2.2}, we obtain commutative diagrams $$ \begin{CD} H_{(x_k, \dots, x_l)}^p(A[t]) @>\sim>> H_{(x_k, \dots, x_{l-1})}^p(A[t]) \\ @V{\alpha_l^p}VV @V{\alpha_{l-1}^p}VV \\ H_{(x_k, \dots, x_l)}^p(C) @>\sim>> H_{(x_k, \dots, x_{l-1})}^p(C) \end{CD} \quad \text{for $p<l-k$} $$ and $$ \begin{CD} 0 @>>> H_{(x_k, \dots, x_l)}^{l-k} (A[t]) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k} (A[t]) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k} (A[t])_{x_l} \\ @. @V{\alpha_l^{l-k}}VV @VVV @VVV \\ 0 @>>> H_{(x_k, \dots, x_l)}^{l-k} (C) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k} (C) @>>> H_{(x_k, \dots, x_{l-1})}^{l-k} (C)_{x_l} \end{CD} $$ whose rows are exact. Therefore the assertion is true for $p<l-k$ and we find that $\alpha_l^{l-k}$ is the direct limit of $$ \alpha_{l,m} \colon \frac{(x_k^m, \dots, x_{l-1}^m) A[t] \qtn x_l} {(x_k^m, \dots, x_{l-1}^m) A[t]} @>>> \bigoplus_{n>0} \frac{[(x_k^m, \dots, x_{l-1}^m) + {\frak q}^n] \qtn x_l} {(x_k^m, \dots, x_{l-1}^m) + {\frak q}^n}. $$ Since $x_l$, $x_{s+1}$,~\dots, $x_d$ is a d-sequence on~$A/(x_k^m, \dots, x_{l-1}^m)$, $$ (x_k^m, \dots, x_{l-1}^m) \qtn x_l \cap [(x_k^m, \dots, x_{l-1}^m) + {\frak q}^n] = (x_k^m, \dots, x_{l-1}^m) \quad \text{for $n>0$}. $$ Therefore $\alpha_{l,m}$ is a monomorphism except for degree~$0$ and $\alpha_l^{l-k}$ is also. If $l=s+1$, then $x_k$,~\dots, $x_s$ is a regular sequence on~$A_{x_{s+1}}$ and $C_{x_{s+1}} = 0$. The assertion is proved in the same way as above. \end{pf} Of course, $\alpha_{s+1}^p$ is the zero map in degree~$0$. Therefore there exists an exact sequence \begin{equation} \label{eqn:4.5.1} 0 @>>> \operatorname{Coker} \alpha_{s+1}^{p-1} @>>> H_{(x_k, \dots, x_{s+1})}^p(R) @>>> H_{(x_k, \dots, x_{s+1})}^p(A) @>>> 0 \end{equation} for $0 < p \leq s-k+1$. \begin{thm} \label{thm:4.6} Let $0 \leq q \leq s-k$. Then $$ (x_{k+q}, \dots, x_d) \operatorname{Coker} \alpha^q_{s+1} = 0 $$ and $H_{\frak M}^p(\operatorname{Coker} \alpha^q_{s+1})$ is finitely graded for~$p< d-s$. \end{thm} \begin{pf} We know that $ \operatorname{Coker} \alpha_{s+1}^q = \operatorname{Coker} \alpha_{k+q}^q = \mathop{\varinjlim}_m \operatorname{Coker} \alpha_{k+q,m} $ and $$ \operatorname{Coker} \alpha_{k+q,m} = \bigoplus_{n>0} \frac{[(x_k^m, \dots, x_{k+q-1}^m) + {\frak q}^n] \qtn x_{k+q}} {(x_k^m, \dots, x_{k+q-1}^m) \qtn x_{k+q} + {\frak q}^n}. $$ By using~Theorem~\ref{thm:4.2} and Lemma~\ref{lem:3.2}, we obtain \begin{multline} \label{eqn:4.6.1} [(x_k^m, \dots, x_{k+q-1}^m) + {\frak q}^n] \qtn x_{k+q} \\ = (x_k^m, \dots, x_{k+q-1}^m) \qtn x_{k+q} + {\frak q}^{n-1} [(x_k^m, \dots, x_{k+q-1}^m, x_{s+1}, \dots, x_d) \qtn x_{k+q}]. \end{multline} Therefore $\operatorname{Coker} \alpha_{k+q,m}$ is annihilated by~$(x_{k+q}, \dots, x_d)$ and $\operatorname{Coker} \alpha_{s+1}^q$ is also. Next we compute local cohomology modules of~$\operatorname{Coker} \alpha_{s+1}^q$. We note that $x_{k+q}$ is a regular element on~$A/(x_k^m, \dots, x_{k+q-1}^m) \qtn x_{k+q}$ and that $x_{s+1}$,~\dots, $x_d$ is a u.s.d-sequence on~$A/ (x_k^m, \dots, x_{k+q-1}^m) \qtn x_{k+q} + (x_{k+q}^l)$ for any $l>0$: see \cite[Proposition 2.2]{Huneke:82:theory}. Therefore, by Proposition~\ref{prop:3.6}, \begin{equation} \label{eqn:4.6.2} \text{ $ H_{(x_{k+q}, x_{s+1}, \dots, x_d)R + R_+}^p (\operatorname{Coker} \alpha_{k+q, m}) $ is concentrated in degree~$1-p$} \end{equation} if $p<d-s$. Hence $ H_{(x_{k+q}, x_{s+1}, \dots, x_d)R + R_+}^p (\operatorname{Coker} \alpha_{s+1}^q) $ is also. By the spectral sequence $ E_2^{pq} = H_{\frak M}^p H_{(x_{k+q}, x_{s+1}, \dots, x_d)R + R_+}^q (-) \Rightarrow H_{\frak M}^{p+q}(-) $, we obtain the second assertion. \end{pf} Next we compute $H_{(x_k, \dots, x_{s+1})}^{s-k+2}(R)$. \begin{thm} Let $A_m = A/(x_k^m, \dots, x_s^m)$ and ${\frak q}_m = {\frak q} A_m$ for any positive integer~$m$. Then $$ H_{(x_k, \dots, x_{s+1})}^{s-k+2}(R) = \mathop{\varinjlim}_{m,l} A_m[{\frak q}_m t]/ x_{s+1}^l A_m[{\frak q}_m t]. $$ In particular, $H_{\frak M}^p H_{(x_k, \dots, x_{s+1})}^{s-k+2}(R)$ is finitely graded for~$p<d-s$. \end{thm} \begin{pf} We consider the exact sequence \begin{multline*} H_{(x_k, \dots, x_s)}^{s-k}(A[t]) @>\alpha_s^{s-k}>> H_{(x_k, \dots, x_s)}^{s-k}(C) @>>> H_{(x_k, \dots, x_s)}^{s-k+1}(R) @>>> \\ @>>> H_{(x_k, \dots, x_s)}^{s-k+1}(A[t]) @>\beta>> H_{(x_k, \dots, x_s)}^{s-k+1}(C). \end{multline*} Since $\beta$ is the direct limit of $$ A[t]/(x_k^m, \dots, x_s^m) A[t] @>>> C/ (x_k^m, \dots, x_s^m) C, $$ we have $\ker \beta = \mathop{\varinjlim}_m A_m [{\frak q}_m t]$. Taking local cohomology modules of a short exact sequence $$ 0 @>>> \operatorname{Coker} \alpha_s^{s-k} @>>> H_{(x_k, \dots, x_s)}^{s-k+1}(R) @>>> \ker \beta @>>> 0 $$ with respect to $(x_{s+1})$, we obtain \begin{equation} \label{eqn:4.7.1} H_{(x_{s+1})}^1 H_{(x_k, \dots, x_s)}^{s-k+1}(R) = H_{(x_{s+1})}^1 (\ker \beta), \end{equation} because $\operatorname{Coker} \alpha_s^{s-k} = \operatorname{Coker} \alpha_{s+1}^{s-k}$ is annihilated by~$x_{s+1}$. The left hand side of~\eqref{eqn:4.7.1} coincides with~$H_{(x_k, \dots, x_{s+1})}^{s-k+2}(R)$ by Lemma~\ref{lem:2.2}. Thus the first assertion is proved. Since $x_{s+1}$,~\dots, $x_d$ is a u.s.d-sequence on~$A_m$, $H_{(x_{s+1}, \dots, x_d)R + R_+}^p(A_m [{\frak q}_m t])$ is concentrated in degree $0 \geq n \geq s-d+2$ if $p \leq d-s$: see \cite[Theorem 4.1]{Goto-Yamagishi::theory}. From the exact sequence $$ 0 @>>> 0 \qtn_{A_m} x_{s+1} @>>> A_m [{\frak q}_m t] @>x_{s+1}^l>> A_m [{\frak q}_m t] @>>> A_m[{\frak q}_m t] / x_{s+1}^l A_m [{\frak q}_m t] @>>> 0 $$ and the spectral sequence $ E_2^{pq} = H_{\frak M}^p H_{(x_{s+1}, \dots, x_d)R + R_+}^q(-) \Rightarrow H_{\frak M}^{p+q}(-) $, we find that \begin{equation} \label{eqn:4.7.2} \text{ $H_{\frak M}^p(A_m[{\frak q}_m t]/ x_{s+1}^l A_m [{\frak q}_m t])$ is concentrated in degree $0 \geq n \geq s-d+2$} \end{equation} if $p< d-s$. Taking the direct limit of it, we obtain the second assertion. \end{pf} Finally we compute local cohomology modules of $B = A[{\frak q}/ x_{s+1}] = R\hlz{x_{s+1}t}$. \begin{thm} \label{thm:4.8} Let ${\frak n}$ be a maximal ideal of~$B$. Then $$ H_{\frak n}^p H_{(x_k, \dots, x_{s+1})}^q(B) = 0 \quad \text{if $q=0$ or $p<d-s-1$.} $$ Furthermore $(x_{k+q-1}, \dots, x_{s+1}) H_{(x_k, \dots, x_{s+1})}^q(B) =0$ for~$q < s-k+2$. \end{thm} \begin{pf} Since the blowing-up $\operatorname{Proj} R \to \operatorname{Spec} A$ is a closed map, there exists a homogeneous prime ideal~${\frak p}$ of~$R$ such that $x_{s+1}t \notin {\frak p}$, $\dim R/ {\frak p} = 1$ and ${\frak n} = [{\frak p} R_{x_{s+1}t}]_0$. Since $x_{s+1}$ is $B$-regular, $H_{(x_k, \dots, x_{s+1})}^0(B) =0$. Let $1 \leq q \leq s-k+1$. By applying Lemma~\ref{lem:2.5} to~\eqref{eqn:4.6.2}, we obtain $$ H_{\frak n}^p((\operatorname{Coker} \alpha_{k+q-1,m})\hlz{x_{s+1}t}) =0 \quad \text{for $p<d-s-1$}. $$ By taking the direct limit of it and using~\eqref{eqn:4.5.1}, we have $$ H_{\frak n}^p H_{(x_k, \dots, x_{s+1})}^q(B) = 0 \quad \text{for $p<d-s-1$}. $$ Moreover Theorem~\ref{thm:4.6} also assures us $ (x_{k+q-1}, \dots, x_{s+1}) H_{(x_k, \dots, x_{s+1})}^q(B) =0 $. Next we consider $H_{(x_k, \dots, x_{s+1})}^{s-k+2}(B)$. By applying Lemma~\ref{lem:2.5} to~\eqref{eqn:4.7.2} and by taking direct limit, we have $$ H_{\frak n}^p H_{(x_k, \dots, x_{s+1})}^{s-k+2}(B) = 0 \quad \text{for $p<d-s-1$}. $$ Thus the proof is completed. \end{pf} \section{Macaulayfications of local rings} \label{sec:5} In this section, we shall construct a Macaulayfication of the affine scheme~$\operatorname{Spec} A$ if its non-Cohen-Macaulay locus is of dimension~$2$. Assume that $A$ possesses a dualizing complex and $\dim A/ {\frak p} = d$ for any associated prime ideal~${\frak p}$ of~$A$. Then $V({\frak a}_A(A))$ coincides with the non-Cohen-Macaulay locus of~$A$. We fix an integer $s \geq \dim A/ {\frak a}_A(A)$ and let $x_1$,~\dots, $x_d$ be a system of parameters for~$A$ satisfying~\eqref{eqn:4.0.1}. First we refine Faltings' results~% \cite[Satz 2, 3]{Faltings:78:Macaulay}. Let ${\frak q} = (x_{s+1}, \dots, x_d)$, $R = A[{\frak q} t]$ and $X = \operatorname{Proj} R$. \begin{thm} \label{thm:5.1} With notation as above, $$ \operatorname{depth} {\cal O}_{X,p} \geq d-s \quad \text{for any closed point~$p$ of~$X$.} $$ If $s=0$ or $A/ {\frak q}$ is Cohen-Macaulay, then $X$ is Cohen-Macaulay. \end{thm} \begin{pf} Since $x_{s+1}$,~\dots, $x_d$ is a u.s.d-sequence on~$A$, $H_{(x_{s+1}, \dots, x_d)R + R_+}^p(R)$ is finitely graded for $p \leq d-s$: see \cite[Theorem 4.1]{Goto-Yamagishi::theory}. By using Lemma~\ref{lem:2.5}, we obtain the first assertion. Furthermore since $\dim {\cal O}_{X,p} =d$ for any closed point~$p$ of~$X$, $X$ is Cohen-Macaulay if $s=0$. Assume that $s>0$ and $A / {\frak q}$ is Cohen-Macaulay. Then $x_1$,~\dots, $x_s$ is a regular sequence on~$A/ {\frak q}$. We use theorems in Section~\ref{sec:4} as $k=1$. From~\eqref{eqn:4.6.1}, we find that $\operatorname{Coker} \alpha_{s+1}^q =0$ for all~$q \leq s-1$. That is, $H_{\frak M}^p H_{(x_1, \dots, x_{s+1})}^q(R)$ is finitely graded if $p<d-s$ or $q<s+1$. By the spectral sequence~% $ E_2^{pq} = H_{\frak M}^p H_{(x_1, \dots, x_{s+1})}^q(-) \Rightarrow H_{\frak M}^{p+q} (-) $, we find that $H_{\frak M}^p(R)$ is finitely graded for $p<d+1$. Lemma~\ref{lem:2.5} assures us $$ \operatorname{depth} {\cal O}_{X,p} \geq d \quad \text{for any closed point~$p$ of~$X$}. $$ The proof is completed. \end{pf} From now on, we assume that $s > 0$. Since $x_s$ is $A$-regular, ${\frak q}$ is a reduction of~$\trans{\frak q} = {\frak q} \qtn x_s$ by~\eqref{eqn:3.2.1}. We put $\trans R = A[\trans{\frak q} t]$ and $\trans X = \operatorname{Proj} \trans R$. Then $\trans X \rightarrow X$ is a finite morphism. \begin{thm} With notation as above, $$ \operatorname{depth} {\cal O}_{\trans X, \trans p} \geq d-s+1 \quad \text{for any closed point~$\trans p$ of~$\trans X$} $$ In particular, if $s=1$, then $\trans X$ is Cohen-Macaulay. \end{thm} \begin{pf} By Corollary~\ref{cor:3.5}, $H_{(x_s, \dots, x_d)R + R_+}^p(\trans R)$ is finitely graded for $p \leq d-s+1$. By using Lemma~\ref{lem:2.5}, we obtain the assertion. \end{pf} Next we consider an ideal~% $ {\frak b} = {\frak q}^2 + x_s {\frak q} = (x_s, \dots, x_d) {\frak q} $. We put $S = A[{\frak b} t]$ and $Y = \operatorname{Proj} S$. Then $Y$ is the blowing-up of~$X$ with center~$(x_s, \dots, x_d) {\cal O}_X$. \begin{thm} \label{thm:5.3} With notation as above, $$ \operatorname{depth} {\cal O}_{Y,q} \geq d-s+1 \quad \text{for any closed point~$q$ of~$Y$}. $$ Furthermore, if $s=1$ or $A$ is Cohen-Macaulay, then $Y$ is Cohen-Macaulay. \end{thm} \begin{pf} Since $ (x_s x_{s+1}, \dots, x_s x_d, x_{s+1}^2, \dots, x_d^2) {\frak b}^{d-s-1} = {\frak b}^{d-s} $, we have only to compute the depth of $C_0 = A[{\frak b}/ x_s x_{s+1}]$ and $C_1 = A[{\frak b}/ x_{s+1}^2]$. If we put $B = A[{\frak q}/ x_{s+1}]$, then \begin{align*} C_0 & = B[x_{s+1}/ x_s] \cong B[T]/(x_sT - x_{s+1}) \qtn \angled{x_s}, \\ C_1 & = B[x_s/ x_{s+1}] \cong B[T]/(x_{s+1}T - x_s) \qtn \angled{x_{s+1}}, \end{align*} where $T$ denotes an indeterminate. We note that $B$, $C_0$, $C_1$ are subrings of the total quotient ring of~$A$ because $x_1$,~\dots, $x_d$ are $A$-regular elements. First we consider $C_0$. We regard it as a homomorphic image of~$B[T]$. Let ${\frak l}_0$ be a maximal ideal of~$C_0$ and ${\frak n} = {\frak l}_0 \cap B$. Then ${\frak n}$ is a maximal ideal of~$B$ because $\operatorname{Spec} C_0 \cup \operatorname{Spec} C_1 \rightarrow \operatorname{Spec} B$ is a blowing-up with center $(x_s, x_{s+1})B$, hence a closed map. There exists a polynomial~$f$ over~$B$ such that ${\frak l}_0 = {\frak n} C_0 + f C_0$ and the leading coefficient of~$f$ is not contained in~${\frak n}$. By Lemma~\ref{lem:2.2} and Theorem~\ref{thm:4.8}, we have, for any $1 \leq k \leq s$, \begin{equation} \label{eqn:5.3.1} H_{{\frak n} B[T] + fB[T]}^p H_{(x_k, \dots, x_{s+1})}^q(B[T]) = 0 \quad \text{if $p<d-s$ or $q=0$.} \end{equation} In fact, the leading coefficient of~$f$ is a regular element on $H_{\frak n}^{d-s} H_{(x_k, \dots, x_{s+1})}^q(B[T])$ because it acts on the injective envelope of~$B/ {\frak n}$ as isomorphism. Taking the local cohomology of a short exact sequence $$ 0 @>>> B[T] @>x_sT - x_{s+1}>> B[T] @>>> B[T]/ (x_sT - x_{s+1}) @>>> 0 $$ with respect to~% $ (x_k, \dots, x_{s+1}) = (x_k, \dots, x_s, x_sT - x_{s+1}) $, we obtain an exact sequence \begin{multline*} 0 @>>> H_{(x_k, \dots, x_{s+1})}^{s-k+1}(B[T]) @>>> H_{(x_k, \dots, x_{s+1})}^{s-k+1}(B[T]/(x_sT - x_{s+1})) @>>> \\ @>>> H_{(x_k, \dots, x_{s+1})}^{s-k+2}(B[T]) @>>> H_{(x_k, \dots, x_{s+1})}^{s-k+2}(B[T]) @>>> 0, \end{multline*} because $(x_s, x_{s+1}) H_{(x_k, \dots, x_{s+1})}^{s-k+1}(B)=0$ by Theorem~\ref{thm:4.8}. This and \eqref{eqn:5.3.1} show that $$ H_{{\frak n} B[T] + fB[T]}^p H_{(x_k, \dots, x_{s+1})}^{s-k+1} (B[T]/ (x_s T - x_{s+1})) = 0 \quad \text{for $p<d-s$}. $$ Taking the local cohomology of an exact sequence $$ 0 @>>> \frac{(x_sT - x_{s+1}) \qtn \angled{x_s}}{(x_s T - x_{s+1})} @>>> B[T]/ (x_sT - x_{s+1}) @>>> C_0 @>>> 0 $$ with respect to $(x_k, \dots, x_{s+1})$, we obtain $$ H_{(x_k, \dots, x_{s+1})}^{s-k+1}(C_0) = H_{(x_k, \dots, x_{s+1})}^{s-k+1}(B[T]/(x_s T - x_{s+1})), $$ that is, \begin{equation} \label{eqn:5.3.2} H_{{\frak l}_0}^p H_{(x_k, \dots, x_{s+1})}^{s-k+1}(C_0) = 0 \quad \text{for $p<d-s$}. \end{equation} We note that $x_s$ is $C_0$-regular. Put $k=s$. Then we have $$ H_{{\frak l}_0}^p H_{(x_s, x_{s+1})}^q(C_0) = 0 \quad \text{if $p<d-s$ or $q<1$}. $$ By the spectral sequence $ E_2^{pq} = H_{{\frak l}_0}^p H_{(x_s, x_{s+1})}^q(-) \Rightarrow H_{{\frak l}_0}^{p+q}(-), $ we obtain \begin{equation} \label{eqn:5.3.3} H_{{\frak l}_0}^p(C_0) = 0 \quad \text{for $p<d-s+1$,} \end{equation} that is, $\operatorname{depth} (C_0)_{{\frak l}_0} \geq d-s+1$. In the same way, we can show that $\operatorname{depth} (C_1)_{{\frak l}_1} \geq d-s+1$ for any maximal ideal~${\frak l}_1$ of~$C_1$. Thus the first assertion is proved. In particular, $Y$ is Cohen-Macaulay if $s=1$. Assume that $A$ is Cohen-Macaulay. Using \cite[Lemma 1]{Faltings:78:Macaulay} twice, we find that $$ x_{s+1} T_{s+2} - x_{s+2}, \dots, x_{s+1} T_d - x_d, x_s T_{s+1} - x_{s+1} $$ is a regular sequence on~$A[T_{s+1}, \dots, T_d]$. Therefore $$ C_0 \cong A[T_{s+1}, \dots, T_d] / (x_{s+1} T_{s+2} - x_{s+2}, \dots, x_{s+1} T_d - x_d, x_s T_{s+1} - x_{s+1}) $$ is Cohen-Macaulay. In the same way, we can show that $C_1$ is Cohen-Macaulay. The proof is completed. \end{pf} In the rest of this section, we assume that $s \geq 2$ and let $\trans {\frak b} = {\frak b} \qtn \angled{x_{s-1}}$. \begin{lem} \label{lem:5.4} For any positive integer~$n$, $$ \trans {\frak b}^n = {\frak b}^n \qtn \angled{x_{s-1}} = {\frak q} {\frak b}^{n-1} [(x_s, \dots, x_d) \qtn x_{s-1}] + x_s^n {\frak q}^{n-1} [{\frak q} \qtn x_{s-1}]. $$ In particular, $\trans {\frak b}^2 = {\frak b} \trans {\frak b}$. \end{lem} \begin{pf} It is sufficient to prove $$ {\frak b}^n \qtn \angled{x_{s-1}} \subseteq {\frak q} {\frak b}^{n-1} [(x_s, \dots, x_d) \qtn x_{s-1}] + x_s^n {\frak q}^{n-1} [{\frak q} \qtn x_{s-1}]. $$ Take $a \in {\frak b}^n \qtn \angled{x_{s-1}}$. Then, by Lemma~\ref{lem:2.4}, Lemma~\ref{lem:3.2} and Theorem~\ref{thm:4.2}, we have \begin{align*} a & \in (x_s, \dots, x_d)^{2n} \qtn \angled{x_{s-1}} \\ & = (x_s, \dots, x_d)^{2n-1} [(x_s, \dots, x_d) \qtn x_{s-1}] \\ & = [{\frak q}^{2n-1} + x_s {\frak q}^{2n-2} + \dots + (x_s^{2n-1})] [(x_s, \dots, x_d) \qtn x_{s-1}] \\ & \subseteq {\frak q} {\frak b}^{n-1} [(x_s, \dots, x_d) \qtn x_{s-1}] + (x_s^n). \end{align*} If we put $a = b + x_s^n a'$ where $b \in {\frak q} {\frak b}^{n-1} [(x_s, \dots ,x_d) \qtn x_{s-1}]$, then $x_s^n a' \in {\frak b}^n \qtn \angled{x_{s-1}}$. Since $x_{s-1}^l x_s^n a' \in {\frak b}^n$ for a sufficiently large~$l$, we can put $x_{s-1}^l x_s^n a' = c + x_s^n d$ where $c \in {\frak q}^{2n} + \dots + x_s^{n-1} {\frak q}^{n+1}$ and $d \in {\frak q}^n$. Then $ x_{s-1}^l a' - d \in {\frak q}^{n+1} \qtn \angled{x_s} = {\frak q}^n [{\frak q} \qtn x_s] $. Hence, $x_{s-1}^l a' \in {\frak q}^n$ and $ a' \in {\frak q}^n \qtn \angled{x_{s-1}} = {\frak q}^{n-1} [{\frak q} \qtn x_{s-1}] $. The proof is completed. \end{pf} Therefore the Rees algebra~$\trans S = A[\trans {\frak b} t]$ is finitely generated over~$S$. Let $\trans Y = \operatorname{Proj} \trans S$. \begin{prop} \label{prop:5.5} $D_{(x_{s-1}, x_s, x_{s+1})}^0(S_+) = \trans S_+$. \end{prop} \begin{pf} First show that $x_{s-1}$, $x_s$ is an $\trans S_+$-regular sequence. Let $n > 0$. It is clear that $x_{s-1}$ is $\trans {\frak b}^n$-regular because it is $A$-regular. Let $a \in (x_{s-1} \trans {\frak b}^n \qtn x_s) \cap \trans {\frak b}^n$. Then $x_{s-1}^l a \in {\frak b}^n$ for a sufficiently large~$l$. Since $x_s a \in (x_{s-1})$ and $x_s$,~\dots, $x_d$ is a d-sequence on~$A/ x_{s-1}^{l+1}A$, \begin{align*} x_{s-1}^l a & \in (x_{s-1}^{l+1}) \qtn x_s \cap {\frak b}^n \\ & \subseteq (x_{s-1}^{l+1}) \qtn x_s \cap (x_{s-1}^{l+1}, x_s, \dots, x_d) \\ & = (x_{s-1}^{l+1}). \end{align*} Hence $a \in (x_{s-1})$. If we put $a = x_{s-1} a'$, then $ a' \in {\frak b}^n \qtn x_{s-1}^{l+1} \subseteq \trans {\frak b}^n $, that is, $a \in x_{s+1}\trans {\frak b}^n$. Thus we have proved that $x_s$ is $\trans S_+ / x_{s-1} \trans S_+$-regular. By~\eqref{eqn:2.1.1}, we have \begin{equation} \label{eqn:5.5.1} D_{(x_{s-1}, x_s, x_{s+1})}^0(S_+) \subseteq D_{(x_{s-1}, x_s, x_{s+1})}^0(\trans S_+) = \trans S_+. \end{equation} Since ${\frak q} \qtn x_{s-1} \subseteq {\frak q} \qtn x_s$ by Theorem~\ref{thm:4.2}, $ (x_{s-1}, \dots, x_d) \trans {\frak b}^n \subseteq {\frak b}^n $ for all~$n>0$ by Lemma~\ref{lem:5.4}, that is, $(x_{s-1}, \dots, x_d) \trans S_+ \subseteq S_+$. We have shown the inverse inclusion of~\eqref{eqn:5.5.1}. \end{pf} The following theorem is one of main aims of this section. \begin{thm} With notation as above, $$ \operatorname{depth} {\cal O}_{\trans Y, \trans q} \geq d-s+2 \quad \text{for any closed point~$\trans q$ of~$\trans Y$}. $$ In particular, if $s=2$, then $\trans Y$ is Cohen-Macaulay. \end{thm} \begin{pf} We have only to compute the depth of $$ \trans C_0 = A[\trans {\frak b}/x_s x_{s+1}] \quad \text{and} \quad \trans C_1 = A[\trans {\frak b}/ x_{s+1}^2]. $$ Proposition~\ref{prop:5.5} says that $\trans C_i = D_{(x_{s-1}, x_s, x_{s+1})}^0(C_i)$ and it is a finitely generated $C_i$-module for $i=0$,~$1$. Let $\trans{\frak l}_i$ be a maximal ideal of~$\trans C_i$ and ${\frak l}_i = \trans{\frak l}_i \cap C_i$. Then ${\frak l}_i$ is a maximal ideal of~$C_i$ because $\trans C_i$ is integral over~$C_i$. We use \eqref{eqn:5.3.2} as $k=s-1$, that is, \begin{equation} \label{eqn:5.6.1} H_{{\frak l}_i}^p H_{(x_{s-1}, x_s, x_{s+1})}^2(C_i) =0 \quad \text{for $p<d-s$}. \end{equation} By using~\eqref{eqn:2.1.2}, we obtain $$ H_{{\frak l}_i}^p H_{(x_{s-1}, x_s, x_{s+1})}^q(\trans C_i) = 0 \quad \text{if $p<d-s$ or $q<2$}. $$ By the spectral sequence $ E_2^{pq} = H_{{\frak l}_i}^p H_{(x_{s-1}, x_s, x_{s+1})}^q(-) \Rightarrow H_{{\frak l}_i}^{p+q}(-) $, we find \begin{equation} \label{eqn:5.6.2} H_{{\frak l}_i}^p(\trans C_i) = 0 \quad \text{for $p<d-s+2$,} \end{equation} that is, $\operatorname{depth} (\trans C_i)_{\trans{\frak l}_i} \geq d-s+2$. Thus the proof is completed. \end{pf} The following corollary shall be used in the next section. \begin{cor} \label{cor:5.7} If $A/(x_s, \dots, x_d)$ is Cohen-Macaulay, then $$ \operatorname{depth} {\cal O}_{Y,q} \geq d-s+2 \quad \text{for any closed point~$q$ of~$Y$.} $$ \end{cor} \begin{pf} It is sufficient to prove $\trans {\frak b}={\frak b}$. Let $a \in \trans {\frak b}$ and $l$ be an integer such that $x_{s-1}^l a \in {\frak b}$. Then we have \begin{align*} a & \in (x_s, \dots, x_d)^2 \qtn x_{s-1}^l \\ & = (x_s, \dots, x_d) [(x_s, \dots, x_d) \qtn x_{s-1}^l] \\ & = (x_s, \dots, x_d)^2 = {\frak b} + (x_s^2) \end{align*} by Lemma~\ref{lem:3.2}. Hence, we may assume that $a \in (x_s^2)$. Let $a = x_s^2 a'$. Since $x_{s-1}^l a \in {\frak b} \subseteq {\frak q}$, $a' \in {\frak q} \qtn x_{s-1}^l x_s^2 = {\frak q} \qtn x_s$ by Theorem~\ref{thm:4.2}. Hence $a = x_s^2 a' \in x_s {\frak q} \subset {\frak b}$. \end{pf} We shall give another Macaulayfication of~$\operatorname{Spec} A$ by considering an ideal ${\frak c} = (x_{s-1}, \dots, x_d) {\frak b}$. Let $Z = \operatorname{Proj} A[{\frak c} t]$, which is the blowing-up of~$Y$ with center $(x_{s-1}, \dots, x_d) {\cal O}_Y$. \begin{thm} \label{thm:5.8} With notation as above, $$ \operatorname{depth} {\cal O}_{Z,r} \geq d-s+2 \quad \text{for any closed point~$r$ of~$Z$}. $$ Furthermore, if $s=2$ or $A$ is Cohen-Macaulay, then $Z$ is Cohen-Macaulay. \end{thm} \begin{pf} Since $ (x_{s-1}x_s, x_s^2) {\frak q} + x_{s-1} (x_{s+1}^2, \dots, x_d^2) + (x_{s+1}^3, \dots, x_d^3) $ is a reduction of~${\frak c}$, we have only to compute the depth of \begin{align*} D_0 & = A[{\frak c}/ x_{s-1} x_s x_{s+1}] = C_0[x_s/ x_{s-1}], \\ D_1 & = A[{\frak c}/ x_s^2 x_{s+1}] = C_0[x_{s-1}/ x_s], \\ D_2 & = A[{\frak c}/ x_{s-1} x_{s+1}^2] = C_1[x_{s+1}/ x_{s-1}], \\ \intertext{and} D_3 & = A[{\frak c}/ x_{s+1}^3] = C_1[x_{s-1}/ x_{s+1}]. \end{align*} For $i=0$ or~$1$, let ${\frak l}_i$ be a maximal ideal of~$C_i$. By~\eqref{eqn:2.1.1}, there exists an exact sequence $$ 0 @>>> C_i @>>> \trans C_i @>>> H_{(x_{s-1}, x_s, x_{s+1})}^1(C_i) @>>> 0. $$ By using~\eqref{eqn:5.3.3} and \eqref{eqn:5.6.2}, we obtain $$ H_{{\frak l}_i}^p H_{(x_{s-1}, x_s, x_{s+1})}^1(C_i) = 0 \quad \text{for all $p<d-s$.} $$ Furthermore, $(x_{s-1}, \dots, x_d) \trans C_i \subseteq C_i$: see the proof of Proposition~\ref{prop:5.5}. Therefore, by~\eqref{eqn:5.6.1}, we have \begin{equation} \label{eqn:5.8.1} H_{{\frak l}_i}^p H_{(x_{s-1}, x_s, x_{s+1})}^q(C_i) = 0 \quad \text{if $p<d-s$ or $q = 0$} \end{equation} and \begin{equation} \label{eqn:5.8.2} (x_{s-1}, \dots, x_d) H_{(x_{s-1}, x_s, x_{s+1})}^1(C_i) =0. \end{equation} Therefore we can prove $$ \operatorname{depth} (D_i)_{{\frak r}_i} \geq d-s+2 $$ for any maximal ideal~${\frak r}_i$ of~$D_i$ and $i=0$,~\dots, $3$ in the same way as Theorem~\ref{thm:5.3}. To make sure, we compute the depth of~% $ D_0 \cong C_0[T]/(x_{s-1}T- x_s) \qtn \angled{x_{s-1}} $. First we note that $x_{s+1} \in x_s C_0$ and $x_{s+1} \in x_s D_0$. Let ${\frak r}_0$ be a maximal ideal of~$D_0$ and ${\frak l}_0 = {\frak r}_0 \cap C_0$. Then ${\frak l}_0$ is a maximal ideal of~$C_0$ and there exists a polynomial~$f$ over~$C_0$ such that ${\frak r}_0 = {\frak l}_0 D_0 + f D_0$ and the leading coefficient of~$f$ is not contained in~${\frak l}_0$. We obtain $$ H_{{\frak l}_0 C_0[T] + f C_0[T]}^p H_{(x_{s-1}, x_s)}^q (C_0[T]) = 0 \quad \text{if $p<d-s+1$ or $q = 0$} $$ from~\eqref{eqn:5.8.1}. Taking the local cohomology of an exact sequence $$ 0 @>>> C_0[T] @>x_{s-1}T - x_s>> C_0[T] @>>> C_0[T]/(x_{s-1}T - x_s) @>>> 0, $$ we have an exact sequence \begin{multline*} 0 @>>> H_{(x_{s-1}, x_s)}^1(C_0[T]) @>>> H_{(x_{s-1}, x_s)}^1(C_0[T]/ (x_{s-1}T - x_s)) @>>> \\ @>>> H_{(x_{s-1}, x_s)}^2(C_0[T]) @>>> H_{(x_{s-1}, x_s)}^2(C_0[T]) @>>> 0 \end{multline*} because of~\eqref{eqn:5.8.2}. This says that $$ H_{{\frak l}_0 C_0[T] + f C_0[T]}^p H_{(x_{s-1}, x_s)}^1 (C_0[T]/ (x_{s-1}T - x_s)) = 0 \quad \text{for $p<d-s+1$}. $$ Taking the local cohomology of an exact sequence $$ 0 @>>> \frac{(x_{s-1}T-x_s) \qtn \angled{x_{s-1}}} {(x_{s-1}T - x_s)} @>>> C_0[T]/(x_{s-1}T - x_s) @>>> D_0 @>>> 0 $$ with respect to~$(x_{s-1}, x_s)$, we obtain $$ H_{{\frak r}_0}^p H_{(x_{s-1}, x_s)}^1(D_0) = 0 \quad \text{for $p<d-s+1$}. $$ Of course, $H_{(x_{s-1}, x_s)}^0(D_0) = 0$. By the spectral sequence $$ E_2^{pq} = H_{{\frak r}_0}^p H_{(x_{s-1}, x_s)}^q(-) \Rightarrow H_{{\frak r}_0}^{p+q}(-), $$ we get $H_{{\frak r}_0}^p(D_0) = 0$ for any $p<d-s+2$. That is, $\operatorname{depth} (D_0)_{{\frak r}_0} \geq d-s+2$. The last assertion is also proved in the same way as Theorem~\ref{thm:5.3}. \end{pf} \setcounter{equation}{0} \section{The proof of Theorem~\ref{mthm}} \label{sec:6} This section is devoted to the proof of Theorem~\ref{mthm}. Let $A$ be a Noetherian ring possessing a dualizing complex and $X$ a quasi-projective scheme over~$A$. That is, $X$ is a dense open subscheme of $X\closure = \operatorname{Proj} R$ where $R = \bigoplus_{n \geq 0} R_n$ is a Noetherian graded ring such that $R_0$ is a homomorphic image of~$A$ and $R$ is generated by~$R_1$ as an $R_0$-algebra. Let $V\closure$ be the non-Cohen-Macaulay locus of~$X\closure$ and $U\closure = X\closure \setminus V\closure$. Of course $V = V\closure \cap X$ is the non-Cohen-Macaulay locus of~$X$. Let $\dc$ be a dualizing complex of~$R$ with codimension function~$v$. Assume that $X$ satisfies the assumption of Theorem~\ref{mthm}. Without loss of generality, we may assume that \begin{equation} \label{eqn:6.0.1} v({\frak p}) = 0 \quad \text{for all associated prime ideal~${\frak p}$ of~$R$:} \end{equation} see \cite[p.~191]{Faltings:78:Macaulay}. Then the local ring~${\cal O}_{X,p}$ of~$p \in X$ satisfies the assumption of Section~\ref{sec:5}, that is, $\dim {\cal O}_{X,p}/ {\frak p} = \dim {\cal O}_{X,p}$ for any associated prime ideal~${\frak p}$ of~${\cal O}_{X,p}$. For the sake of completeness, we sketch out the proof. Let ${\frak a}$ be a homogeneous ideal of~$R$ such that $V\closure = V({\frak a})$. Then the closed immersion $\operatorname{Proj} R/H_{\frak a}^0(R) \to X\closure$ is birational as follows. For any minimal prime ideal~${\frak p}$ of~$R$, ${\frak a} \not\subset {\frak p}$ and $H_{\frak a}^0(R) \subseteq {\frak p}$ because $R_{\frak p}$ is Cohen-Macaulay. Hence the underlying set of~$\operatorname{Proj} R/H_{\frak a}^0(R)$ coincides with the one of~$X\closure$. Furthermore, $f^{-1}(U\closure) \to U\closure$ is an isomorphism and $U\closure$ is dense in~$X\closure$. By replacing $R$ by~$R/H_{\frak a}^0(R)$, we may assume that \begin{equation} \label{eqn:6.0.2} \text{ every associated prime ideal of~$R$ is minimal. } \end{equation} Next we fix a primary decomposition of~$(0)$ in~$R$. For all integer~$i$, let ${\frak q}_i$ be the intersection of all primary component~${\frak q}$ of~$(0)$ such that $v(\sqrt{\frak q}) = i$. Then $g \colon \coprod_i \operatorname{Proj} R/{\frak q}_i \to X\closure$ is a finite morphism and $g^{-1}(U\closure) \to U\closure$ is an isomorphism as follows. Note that ${\frak q}_i = R$ for all but finitely many~$i$. Furthermore, for any ${\frak p} \in U\closure$, ${\frak p} \supseteq {\frak q}_i$ if and only if $v({\frak p}) - \dim R_{\frak p} = i$ because $R_{\frak p}$ is Cohen-Macaulay, hence equidimensional. Therefore $U\closure$ is the disjoint union of $U\closure \cap V({\frak q}_i)$. Moreover $R\hlz{\frak p} = [R/{\frak q}_i]\hlz{\frak p}$ if ${\frak p} \in U\closure \cap V({\frak q}_i)$. Because of~\eqref{eqn:6.0.2}, $g^{-1}(U\closure)$ and $U\closure$ are dense in $\operatorname{Proj} R/{\frak q}_i$ and $X\closure$, respectively. Thus $g^{-1}(X) \to X$ is birational proper and the connected components of~$g^{-1}(X)$ satisfy the assumption of Theorem~\ref{mthm}. Since $u$ is locally constant, $V_i = u^{-1}(i) \cap V$ is closed for any positive integer~$i$. We put $s_i = \dim V_i$. By~\eqref{eqn:6.0.1}, we find that $V_1 = \emptyset$, $s_2 \leq 0$ and $s_3 \leq 1$. Let $d$ be the largest integer such that $V_d \ne \emptyset$ and $s = s_d$. We shall give a closed subscheme~$W$ of~$X$ such that $V_d = V \cap W$ and ${\cal O}_{Y,q}$ is Cohen-Macaulay for all $q \in \pi^{-1}(W)$ where $\pi \colon Y \to X$ is the blowing-up of~$X$ with center~$W$. Let ${\frak a} = \prod_{i>0} \operatorname{ann} H^i(\dc)$, which is finite product. Then it is obvious that $V\closure = V({\frak a})$. Fix a primary decomposition of~${\frak a}$ and let ${\frak a}_d$ be the intersection of all primary component~${\frak q}$ of~${\frak a}$ such that $\sqrt{\frak q} \in V_d$. Then we can take homogeneous elements $z_1$,~\dots, $z_d \in R$ such that \begin{gather} \label{eqn:6.0.3} V_i \cap V((z_{d-s_i}, \dots, z_d)) = \emptyset \quad \text{for $i < d$} \\ d({\frak p}) = d \quad \text{for all minimal prime ideal~${\frak p}$ of $R/(z_1, \dots, z_d) \qtn \angled{R_+}$} \\ \begin{cases} z_{s+1}, \dots, z_d \in {\frak a}_d; \\ \label{eqn:6.0.5} z_i \in \prod_{j>d-i} \operatorname{ann} H^j(\hom(R/(z_{i+1}, \dots, z_d), \dc)), & \text{for $i \leq s$} \end{cases} \end{gather} in the same way as Section~\ref{sec:4}. We put $$ {\frak b} = \begin{cases} (z_1, \dots, z_d), & \text{if $s=0$}; \\ (z_1, \dots, z_d) (z_2, \dots, z_d), & \text{if $s=1$}; \\ (z_1, \dots, z_d) (z_2, \dots, z_d) (z_3, \dots, z_d), & \text{if $s=2$} \end{cases} $$ and prove that $W = V({\frak b}) \cap X$ satisfies the required properties. Because of~\eqref{eqn:6.0.3}, $V_i \cap W = \emptyset$ for~$i < d$. Let $\pi \colon Y \to X$ be the blowing-up of~$X$ with center~$W$, $q$ a closed point of~$\pi^{-1}(W)$ and ${\frak p} \subseteq R$ the image of~$q$. Take an element $y \in R_1 \setminus {\frak p}$ and put $x_i = z_i / y^{\deg z_i}$ for all~$i$. Since $(\dc)\hlz{\frak p}$ is a dualizing complex of~$R\hlz{\frak p}$, we obtain $$ \begin{cases} x_{s+1}, \dots, x_d \in {\frak a}_{R\hlz{\frak p}}(R\hlz{\frak p}); \\ x_i \in {\frak a}_{R\hlz{\frak p}}( R\hlz{\frak p}/(x_{i+1}, \dots, x_d)), & \text{for $i \leq s$}. \end{cases} $$ from~\eqref{eqn:6.0.5}. When $s=2$, there exist three cases: If $z_1$, $z_2 \in {\frak p}$, then $x_1$,~\dots, $x_d$ is a system of parameters for~$R\hlz{\frak p}$ satisfying~\eqref{eqn:4.0.1} or a regular sequence on the Cohen-Macaulay ring~$R\hlz{\frak p}$. Since $ {\frak b}\hlz{\frak p} = (x_1, \dots, x_d) (x_2, \dots, x_d) (x_3, \dots, x_d) $, ${\cal O}_{Y,q}$ is Cohen-Macaulay by Theorem~\ref{thm:5.8}. If $z_2 \in {\frak p}$ but $z_1 \notin {\frak p}$, then $x_2$,~\dots, $x_d$ is a subsystem of parameters for~$R\hlz{\frak p}$ satisfying~\eqref{eqn:4.0.1} or a regular sequence on the Cohen-Macaulay ring~$R\hlz{\frak p}$. Furthermore ${\frak b}\hlz{\frak p} = (x_2, \dots, x_d)(x_2, \dots, x_d)$ and $R\hlz{\frak p}/(x_2, \dots, x_d)$ is Cohen-Macaulay because $x_1 \in {\frak a}_{R\hlz{\frak p}}(R\hlz{\frak p}/(x_2, \dots, x_d))$ is a unit. Hence ${\cal O}_{Y,q}$ is Cohen-Macaulay by Corollary~\ref{cor:5.7}. If $z_1$, $z_2 \notin {\frak p}$, then $x_3$,~\dots, $x_d \in {\frak a}_{R\hlz{\frak p}}(R\hlz{\frak p})$ is a subsystem of parameters for~$R\hlz{\frak p}$ and $R\hlz{\frak p}/(x_3, \dots, x_d)$ is Cohen-Macaulay. Since ${\frak b}\hlz{\frak p} = (x_3, \dots, x_d)$, ${\cal O}_{Y,q}$ is Cohen-Macaulay by Theorem~\ref{thm:5.1}. When $s=0$ or $1$, we can prove the assertion in the same way as above. By repeating this procedure, we obtain a Macaulayfication of~$X$. We complete the proof of Theorem~\ref{mthm}. \subsection*{Acknowledgment} The author is grateful to Professor S.~Goto for his helpful discussions and to Professor K.~Kurano for careful reading of the draft. \ifx\undefined\bysame \newcommand{\bysame}{% \leavevmode\hbox to3em{\hrulefill}\,} \fi
1996-07-19T01:09:21
9512
alg-geom/9512001
en
https://arxiv.org/abs/alg-geom/9512001
[ "alg-geom", "math.AG" ]
alg-geom/9512001
Richard Hain
Richard Hain
Infinitesimal presentations of the Torelli groups
55 pages, LaTeX2e, amsart.cls. Author supplied dvi file available from http://www.math.duke.edu/faculty/hain/
null
null
null
null
This is a significant revision of the early version of this paper which was posted last December. The speculative section has been removed in light of some recent results of Morita and Kawazumi. Numerous typos have been fixed. The companion paper "The Hodge de Rham Theory of Relative Malcev Completion" has just been posted.
[ { "version": "v1", "created": "Fri, 1 Dec 1995 21:49:54 GMT" }, { "version": "v2", "created": "Thu, 18 Jul 1996 23:03:13 GMT" } ]
2008-02-03T00:00:00
[ [ "Hain", "Richard", "" ] ]
alg-geom
\section{Introduction} By a theorem of Malcev \cite{malcev}, every torsion free nilpotent group can be imbedded canonically as a discrete, cocompact subgroup of a real nilpotent Lie group. One can therefore associate to a finitely generated group $\pi$ a tower of nilpotent Lie groups \begin{equation}\label{tower} \cdots \to G_3 \to G_2 \to G_1 = H_1(\pi,{\mathbb R}) \end{equation} by taking $G_k$ to be the nilpotent Lie group associated to the maximal torsion free quotient of $\pi$ of length $k$. Since each nilpotent Lie group is simply connected, the tower (\ref{tower}) is determined by the corresponding tower $$ \cdots \to {\mathfrak g}_3 \to {\mathfrak g}_2 \to {\mathfrak g}_1 = H_1(\pi,{\mathbb R}) $$ of nilpotent Lie algebras. The inverse limit ${\mathfrak g}$ of this tower is a pronilpotent Lie algebra, called the {\it Malcev Lie algebra associated to $\pi$.} This Lie algebra has the property that the graded Lie algebra $\Gr {\mathfrak g}$ associated with its lower central series is isomorphic to $\left(\Gr \pi\right) \otimes {\mathbb R}$, where $\Gr \pi$ is the graded ${\mathbb Z}$-Lie algebra associated to the filtration of $\pi$ by its lower central series. In this paper we give a presentation of the Malcev Lie algebra ${\mathfrak t}_{g,r}^n$ associated to the Torelli group $T_{g,r}^n$ for all $g\ge 6$. Since each Torelli group injects into its unipotent completion (at least when $n+r>0$), the corresponding Malcev Lie algebra should contain significant information about the group. Recall that the mapping class group $\Gamma_{g,r}^n$ \label{group_def} is defined as follows. Fix a compact orientable surface $S$ of genus $g$, together with $n+r$ distinct points \begin{equation}\label{points} x_1,\dots,x_n; y_1,\dots,y_r \end{equation} of $S$ and $r$ non-zero tangent vectors $v_1,\dots,v_r$, where $v_j$ is tangent to $S$ at $y_j$. The group $\Gamma_{g,r}^n$ is the group of isotopy classes of orientation preserving diffeomorphisms of $S$ that fix each of the points (\ref{points}) and each of the tangent vectors $v_j$.% \footnote{One can replace each tangent vector in the definition by a boundary component --- with this change, the diffeomorphisms are required to be the identity on each boundary component.} The Torelli group $T_{g,r}^n$ \label{torelli_def}is defined to be the kernel of the natural homomorphism \begin{equation}\label{rho} \Gamma_{g,r}^n \to \Aut H_1(S,{\mathbb Z}). \end{equation} Observe that the classical pure braid group $P_n$ is $T_{0,1}^n$. Our presentation of ${\mathfrak t}_{g,r}^n$ generalizes the well-known presentation of ${\mathfrak p}_n$, the Malcev Lie algebra of the pure braid group $P_n$, which is of importance in the theory of Vassiliev invariants (cf.\ \cite{kontsevich:vass,cartier:knots,bar-natan}) and was first written down by Kohno \cite{kohno:braids}. Denote the free Lie algebra generated by indeterminates $X_1,\dots, X_m$ by ${\mathbb L}(X_1,\dots,X_m)$, and that generated by a vector space $V$ by \label{lie_def} ${\mathbb L}(V)$. Then ${\mathfrak p}_n$ is the completion of the graded Lie algebra $$ {\mathbb L}(X_{ij} : ij\text{ is a two element subset of }\{1,\dots,n\})/R $$ where $R$ is the ideal generated by the quadratic relations \begin{align*}\label{braid_relns} [X_{ij},X_{kl}]&\text{ when $i,j,k$ and $l$ are distinct;}\cr [X_{ij},X_{ik} + X_{jk}]& \text{ when $i,j$ and $k$ are distinct.} \end{align*} The property that ${\mathfrak p}_n$ is the completion of the associated graded Lie algebra $\Gr {\mathfrak p}_n$ does not hold for the generic group, but does hold for all Torelli groups as we shall see. It is easiest to first state the result for the absolute Torelli group, $T_g := T_{g,0}^0$. It follows from Dennis Johnson's computation of the first homology of $T_g$ \cite{johnson:h1} that each graded quotient of the lower central series of ${\mathfrak t}_g$ is a representation of the algebraic group $Sp_g$. We will give a presentation of $\Gr {\mathfrak t}_g$ in the category of representations of $Sp_g$. Chose a set $\lambda_1,\dots, \lambda_g$ of fundamental weights of $Sp_g$. Denote the representation of $Sp_g$ with highest weight $\lambda = \sum n_j \lambda_j$ by $V(\lambda)$. Johnson's fundamental computation is that there is a natural $Sp_g({\mathbb Z})$ equivariant isomorphism between $H_1(T_g,{\mathbb Q})$ and $V(\lambda_3)$. For all $g \ge 3$, the representation $\Lambda^2 V(\lambda_3)$ contains a unique copy of $V(2\lambda_2) + V(0)$. Denote the $Sp_g$ invariant complement of this by $R_g$. Since the quadratic part of the free Lie algebra ${\mathbb L}(V)$ is $\Lambda^2 V$, we can view $R_g$ as being a subspace of the quadratic elements of ${\mathbb L}(V(\lambda_3))$. \begin{theorem} For all $g\neq 2$, ${\mathfrak t}_g$ is the completion of its associated graded $\Gr{\mathfrak t}_g$. When $g\ge 6$, this has presentation $$ \Gr {\mathfrak t}_g = {\mathbb L}(V(\lambda_3))/(R_g), $$ where $R_g$ is the set of quadratic relations defined above. When $3 \le g < 6$, the relations in $\Gr {\mathfrak t}_g$ are generated by the quadratic relations $R_g$, and possibly some cubic relations. \end{theorem} In fact, we will show that in genus 3 there are no quadratic relations and the cubic relations contain a copy of $V(\lambda_3)$. Dennis Johnson has proved that $T_g$ is finitely generated for all $g\ge 3$, but it is not known for any $g\ge 3$ whether or not $T_g$ is finitely presented. Geoff Mess \cite{mess} proved that $T_2$ is a countably generated free group. (Note that when $g = 0,1$, $T_g$ is trivial.) \begin{corollary} For all $g\neq 2$, and for all $r,n \ge 0$, $t_{g,r}^n$ is finitely presented. \qed \end{corollary} In the decorated case, we have the extension \begin{equation}\label{extension} 1 \to \pi_{g,r}^n \to T_{g,r}^n \to T_g \to 1, \end{equation} where $\pi_{g,r}^n$ denotes the fundamental group of the configuration space of $n$ points and $r$ tangent vectors in $S$. After applying the Malcev Lie algebra functor, we obtain an extension $$ 0 \to {\mathfrak p}_{g,r}^n \to {\mathfrak t}_{g,r}^n \to {\mathfrak t}_g \to 0, $$ where ${\mathfrak p}_{g,r}^n$ denotes the Malcev Lie algebra of $\pi_{g,r}^n$. \begin{theorem} For all $g \ge 0$, and all $r,n \ge 0$, the Lie algebra ${\mathfrak p}_{g,r}^n$ is the completion of its associated graded $\Gr {\mathfrak p}_{g,r}^n$. The associated graded has a presentation with only quadratic relations. \end{theorem} The explicit presentation is given in Section \ref{braids2}. In order to give the presentation for ${\mathfrak t}_{g,r}^n$, we prove that (\ref{extension}) remains exact after taking graded quotients. Thus, in order to give a presentation of ${\mathfrak t}_{g,r}^n$, it suffices to determine the map $$ [\phantom{x},\phantom{x}] : \left(\Gr^1 {\mathfrak t}_g \otimes \Gr^1 {\mathfrak p}_{g,r}^n\right) \oplus \left(\Gr^1 {\mathfrak t}_g \otimes \Gr^1 {\mathfrak t}_g\right) \to \Gr^2 {\mathfrak t}_{g,r}^n. $$ determined by the bracket. We do this in Section \ref{decorated} to obtain the presentation of ${\mathfrak t}_{g,r}^n$ in general. Our results complement, and sometimes overlap with, the beautiful work \cite{morita:casson,morita:cocycles,morita:trace,morita:conj} of Shigeyuki Morita who began the study of the ``higher Johnson homomorphisms'' studied in this paper. Our main theorem allows us to answer several questions about Torelli groups, and to prove a conjecture of Morita. These and other applications are discussed in Section~\ref{applications}. Another feature of the classical case is the existence of a canonical universal integrable connection. Denote the classifying space $$ {\mathbb C}^n - \left\{(z_1,\dots,z_n) : \text{ the $z_i$ are not distinct}\right\} $$ of $P_n$ by $X_n$. Denote the complex of global meromorphic $k$-forms on a complex manifold $Y$ by $\Omega^k(Y)$. The universal integrable connection on $X_n$ is given by the ${\mathfrak p}_g$ valued 1-form $$ \sum_{ij} d\log(z_i - z_j)\, X_{ij} \in \Omega^1(X_n) \otimes {\mathfrak p}_n. $$ It plays a central role in the theory of Vassiliev invariants (cf. \cite{kohno:KZ}, \cite{cartier:knots}, \cite{kassel}.) We are able to prove that there is a canonical universal connection form with ``scalar curvature'' for each $T_{g,r}^n$, provided $g\neq 2$, although, to date, we have not been able to give an explicit formula for it. The universal connection is discussed in Section \ref{applications}. The basic approach in this paper is to use Hodge theory. The main technical theorem of the paper is: \begin{theorem} Suppose that $g\neq 2$ and that $r,n \ge 0$. For each choice of a complex structure on the decorated reference surface $$ \left(S;x_1,\dots,x_n;y_1,\dots,y_r;v_1,\dots,v_r\right) $$ there is a mixed Hodge structure on ${\mathfrak t}_{g,r}^n$ for which the bracket is a morphism of mixed Hodge structures. \end{theorem} This mixed Hodge structure is canonical once one fixes an isomorphism of $\Gamma_{g,r}^n$ with the (orbifold) fundamental group of the moduli space of smooth projective curves of genus $g$ with $n$ marked points, and $r$ non-zero tangent vectors. The theorem is proved using the mixed Hodge structure on the completion of the mapping class group $\Gamma_{g,r}^n$ relative to the homomorphism $\Gamma_{g,r}^n \to Sp_g({\mathbb Q})$ induced by (\ref{rho}), the existence of which follows from \cite{hain:derham}. The mixed Hodge structure on the Lie algebra ${\mathfrak u}_{g,r}^n$ of the prounipotent radical of the relative completion is lifted to ${\mathfrak t}_{g,r}^n$ using two results from \cite{hain:comp}. The first states that we have a central extension $$ 0 \to \Ga \to {\mathfrak t}_{g,r}^n \to {\mathfrak u}_{g,r}^n \to 0 $$ when $g\ge 3$. The second gives an explicit relationship between this extension and the algebraic 1-cycle $C - C^-$ in the jacobian of an algebraic curve $C$. The theory of relative completion is reviewed in Section \ref{rel_comp}. When $g \ge 3$, the weight filtration of ${\mathfrak t}_{g,r}^n$ is its lower central series. The fact that the weight graded functors are exact on the category of mixed Hodge structures then allows the reduction to the associated graded with impunity when studying ${\mathfrak t}_{g,r}^n$, ${\mathfrak u}_{g,r}^n$, ${\mathfrak p}_{g,r}^n$ and maps between them. In order to bound the degrees of relations in ${\mathfrak t}_g$ by $N$, we need to know that if ${\mathbb V}$ is a variation of Hodge structure of weight $n$ over ${\mathcal M}_g$ (the moduli space of curves) that comes from a rational representation of $Sp_g$, then the weights on $$ H^2({\mathcal M}_g,{\mathbb V}) $$ are bounded between $2+n$ and $n+N$ --- see Section \ref{cts_coho_tor}. There is no {\it a priori} uniform bound on the weights of $H^k(X,{\mathbb V})$, where $X$ is a smooth variety and ${\mathbb V}$ is a variation of Hodge structure over $X$ of weight $l$, as there is in the case of ${\mathbb Q}$ coefficients where the weights are bounded between $k$ and $2k$. For example, if $\Gamma$ is a finite index subgroup of $SL_2({\mathbb Z})$ and $X$ the quotient of the upper half plane by $\Gamma$, then the non-trivial weights on $H^1(X,S^n{\mathbb V})$ are $n+1$ and $2n+2$ for infinitely many $n$, as can be seen from results in \cite{zucker}. Here ${\mathbb V}$ denotes the fundamental representation of $SL_2$ viewed as a variation of Hodge structure over $X$ of weight 1 and $S^n{\mathbb V}$ its $n$th symmetric power. Thus, one of the main technical ingredients in the paper is the result of Kabanov \cite{kabanov} (see also \cite{kabanov:purity} which states that one can take $N$ to be 2 when $g\ge 6$, and 3 when $3 \le g < 6$. The existence of the mixed Hodge structure on the Malcev Lie algebra associated to the Torelli group was obtained several years ago. The quadratic relations (proved in Section \ref{quadratic_relns}) were derived in \cite{hain:letter}. Subsequently Morita (unpublished) proved that when the genus is sufficiently large there are no cubic or quartic relations in ${\mathfrak t}_g$. Kabanov's purity theorem allows us to avoid Morita's involved computations and to show there are no higher order relations. \medskip \noindent{\it Acknowledgements.} It is a pleasure to thank all those with whom I have had useful discussions, especially A.~Borel, P.~Deligne, Alexander Kabanov, Eduard Looijenga, Shigeyuki Morita, and Steven Zucker. I would also like to thank Hiroaki Nakamura for his numerous comments on the manuscript. I would like to thank the Institute for Advanced Study, the Institut des Hautes \'Etudes Scientifiques, the Institut Henri Poincar\'e, and the Universit\"at Essen, each of which supported me during my sabbatical during which this paper was written. \section{Braid Groups in Positive Genus} \label{braids1} Throughout this section $g$ will be positive. Suppose that $S$ is a compact oriented surface of genus $g$, and that $r$ and $n$ are integers $\ge 0$. The configuration space of $m\ge 1$ points on $S$ is $$ F^m(S) = S^m - \Delta, $$ where $\Delta$ is the union of the various diagonals $x_i = x_j$. Denote the tangent bundle of $S$ by $TS$, and the bundle of non-% zero tangent vectors by $V$. The pullback of $V$ to $F^m(S)$ along the $j$th projection $p_j : F^m(S) \to S$ will be denoted by $V_j$. For a subset $A$ of $\{1,\dots,m\}$ denote the fibered product of the $V_j$, where $j \in A$, by $V_A$. The {\it configuration space \label{config_def} $F_{g,r}^n$ of $n$ points and $r$ non-zero tangent vectors of $S$} is defined to be the total space of the bundle $$ V_A \to F^{r+n}(S) $$ where $A = \{n+1,\dots,n+r\}$. Fix a base point $f_o$ of $F_{g,r}^n$. Define \label{fund_def} $$ \pi_{g,r}^n = \pi_1(F_{g,r}^n,f_o). $$ When $r=0$ this is just the group of pure braids with $n$ strings on the surface $S$. In general, this group can be thought of as the group of pure braids on $S$ with $r+n$ strings where $r$ of the strings are framed. It is a standard fact that the space $F_{g,r}^n$ is an Eilenberg-MacLane space of type $K(\pi,1)$ \cite[\S 1.2]{birman}. In contrast with the genus 0 case, we have: \begin{proposition}\label{h1_braid} For each $g\ge 0$, there is a short exact sequence $$ 0 \to \left({\mathbb Z}/(2g-2){\mathbb Z}\right)^r \to H_1(\pi_{g,r}^n,{\mathbb Z}) \stackrel{p}{\to} H_1(S^{n+r},{\mathbb Z}) \to 0, $$ where $p$ is induced by the natural map $F_{g,r}^n \to S^{n+r}$. \end{proposition} \begin{proof} We first consider the case when $r=0$. In this case $F_{g,r}^n$ is $S^n - \Delta$. The divisor $\Delta$ is the union of the diagonals $\Delta_{ij}$ where the $i$th and $j$th point of $S^n$ are equal. We therefore have a Gysin sequence $$ \cdots \to H_2(S^n) \stackrel{\gamma}{\to} \bigoplus_{i<j}{\mathbb Z} \stackrel{t}{\to} H_1(S^n - \Delta) \to H_1(S^n) \to 0. $$ The map $\gamma$ takes a cycle $z$ to the element of $\oplus_{i<j}{\mathbb Z}$ whose $ij$th term is the intersection number $z\cdot\Delta_{ij}$. The map $t$ takes the generator of the $ij$th factor to the homology class of a small circle which winds about about $\Delta_{ij}$ in the positive direction. Fix a base point $x_o$ of $S$. For $u\in H_k(S)$ and $i\in \{1,\dots,n\}$, denote by $u^i$ the element $$ x_o \times \dots \times x_o \times \stackrel{i}{u} \times x_o \times \dots \times x_o $$ of $H_k(S^n,{\mathbb Z})$, where $u$ is placed in the $i$th factor. For elements $u$ and $v$ of $H_1(S)$ and $i,j\in\{1,\dots,n\}$, denote the element $$ x_o \times \dots \times x_o \times \stackrel{i}{u} \times x_o \times \dots \times x_o \times \stackrel{j}{v} \times x_o \times \dots \times x_o $$ of $H_2(S,{\mathbb Z})$ by $u^i \times v^j$, where $u$ is placed in the $i$th factor and $v$ in the $j$th. By choosing representatives of $u$ and $v$ which do not pass through the base point, one sees immediately that $$ \gamma : u^i \times u^j \mapsto (u\cdot v)\Delta_{ij} $$ from which it follows that $\gamma$ is surjective and that $t$ is trivial. This proves the result when $r=0$. Observe that $$ \gamma : S^i \mapsto \sum_{j\neq i} \Delta_{ij}. $$ If $a$ and $b$ are elements of $H_1(S,{\mathbb Z})$ with intersection number 1, then $$ S^i - \sum_{j\neq i} a^i \times b^j $$ is in the kernel of $\gamma$, and therefore lifts to an element $\sigma_i$ of $H_2(S^n-\Delta,{\mathbb Z})$. We prove the general case by induction on $r$. Our inductive hypothesis is that the result has been proven for $F_{g,s}^m$ when $s<r$, and that there are classes $$ \sigma_1,\dots, \sigma_m \in H_2(F_{g,s}^m,{\mathbb Z}) $$ whose images under the the maps $$ {p_j}_\ast : H_2(F_{g,s}^m,{\mathbb Z}) \to H_2(S,{\mathbb Z}) $$ induced by the various projections $p_j : F_{g,s}^m \to S$ satisfy $p_j(\sigma_i) = \delta_{ij}[S]$. We have proved this when $r=0$. Suppose that $r>0$. We have the projection $$ q: F_{g,r}^n \to F_{g,r-1}^{n+1}, $$ which replaces the first tangent vector by its anchor point. This is a principal ${\mathbb C}^\ast$ bundle. It fits into a commutative square $$ \begin{CD} F_{g,r}^n @>>> V \cr @VqVV @VVV \cr F_{g,r-1}^{n+1} @>>> S \cr \end{CD} $$ One therefore has a map $$ \begin{CD} H_2(F_{g,r-1}^{n+1}) @>\gamma_F>> H_0(F_{g,r}^n) @>>> H_1(F_{g,r}^n) @>>> H_1(F_{g,r-1}^{n+1}) @>>> 0 \cr @VVV @VVV @VVV @VVV \cr H_2(S) @>\gamma_S>> H_0(S) @>>> H_1(V) @>>> H_1(S) @>>> 0 \cr \end{CD} $$ of Gysin sequences. The map $\gamma_S$ is simply multiplication by the euler characteristic. Since $\sigma_{n+1} \mapsto [S]$, we see that $\gamma_F$ takes $\sigma_{n+1}$ to $2 - 2g$. The result follows by induction. \end{proof} \begin{remark}\label{error} This result (with $r=0$) can also be proved by considering the natural fibrations $F_g^{n+1} \to F_g^n$ obtained by forgetting the last point. The fiber is an $n$ punctured copy of $S$, and its homology therefore fits into an exact sequence $$ 0 \to {\mathbb Z}^n/\text{diagonal} \to H_1(S - n\text{ points}) \to H_1(S) \to 0. $$ One has to be a little careful as the monodromy action is not trivial. A simple geometric argument shows that the monodromy acts trivially on the kernel and quotient in the sequence above, and therefore is given by a homomorphism $$ \pi_g^n \to \Hom(H_1(S),{\mathbb Z}^n/\text{diagonal}) \cong H^1(S^n)/\text{diagonal}. $$ By induction on $n$, $H_1(\pi_g^n)$ is isomorphic to $H_1(S^n)$. Since the monodromy is abelian, it factors through the quotient map $$ \pi_g^n \to H_1(S)^{\oplus n}. $$ A straightforward geometric argument shows that the action of the latter is given by the map $$ H_1(S)^{\oplus n} \stackrel{PD^{\oplus n}}{\longrightarrow} H^1(S)^{\oplus n}/\text{diagonal}, $$ where $PD$ denotes Poincar\'e duality. The coinvariants are therefore given by $$ H_0(F_g^n,H_1(\text{fiber})) = H_1(S). $$ An elementary spectral sequence argument completes the inductive step. Kohno and Oda \cite[p.~208]{kohno-oda} use this method, but their result contradicts ours as they mistakenly assume that the monodromy representation is trivial. \end{remark} \section{Relative Completion of Mapping Class Groups} \label{rel_comp} In this section we recall the main theorem of \cite{hain:comp} which makes precise the relationship between the Malcev completion of $T_{g,r}^n$ and the unipotent radical of the relative Malcev completion of $\Gamma_{g,r}^n$. We first recall the definition of relative Malcev completion, which is due to Deligne. A reference for this material is \cite[\S\S 2--4]{hain:comp}. Suppose that $\Gamma$ is a discrete group, $S$ a reductive linear algebraic group over a field $F$ of characteristic zero, and that $\rho : \Gamma \to S(F)$ is a representation whose image is Zariski dense. The {\it Malcev completion of $\Gamma$ over $F$ relative to $\rho$} is a homomorphism $\tilde{\rho} : \Gamma \to {\mathcal G}$ of $\Gamma$ into a proalgebraic group ${\mathcal G}$, defined over $F$, which is an extension $$ 1 \to {\mathcal U} \to {\mathcal G} \stackrel{p}{\to} S \to 1 $$ of $S$ by a prounipotent group ${\mathcal U}$ such that the diagram $$ \begin{CD} \Gamma @>{\rho}>> S \cr @V{\tilde{\rho}}VV @| \cr {\mathcal G} @>>p> S \end{CD} $$ commutes. It is characterized by a universal mapping property: If $G$ is a linear (pro)algebraic group over $F$ which is an extension $$ 1 \to U \to G \to S \to 1 $$ of $S$ by a (pro)unipotent group, and if $\tau : \Gamma \to G$ is a homomorphism whose composition with $G \to S$ is $\rho$, then there is a unique homomorphism ${\mathcal G} \to G$ such that the diagram $$ \begin{CD} \Gamma @>{\tilde{\rho}}>> {\mathcal G} \cr @V{\tau}VV @VpVV\cr G @>>> S \end{CD} $$ commutes. When $S$ is the trivial group, the relative completion of $\Gamma$ coincides with the classical Malcev (or unipotent) completion of $\Gamma$. Suppose that $K/F$ is an extension of fields of characteristic zero. When $S$ is defined over $F$ and $\rho : \Gamma \to S(F)$, one can ask if the $K$-form of the completion of $\Gamma$ relative to $\rho$ is obtained from the $F$-form by extension of scalars. If this is the case for all such field extensions, we will say that the relative completion of $\Gamma$ relative to $\rho$ can be defined over $F$. The action of the mapping class group on $S$ preserves the intersection pairing $q : H_1(S,{\mathbb Z})^{\otimes 2} \to Z$. We therefore have a homomorphism \begin{equation}\label{map} \rho : \Gamma_{g,r}^n \to \Aut (H_1(S,{\mathbb Z}),q) \cong Sp_g({\mathbb Z}). \end{equation} For a positive integer $l$, we define the {\it level $l$ subgroup} $\Gamma_{g,r}^n[l]$ \label{level_def} to be the kernel of the induced map $$ \Gamma_{g,r}^n \to \Aut (H_1(S,{\mathbb Z}/l{\mathbb Z}),q) \cong Sp_g({\mathbb Z}/l{\mathbb Z}). $$ Here we interpret $Sp_g({\mathbb Z}/l{\mathbb Z})$ as the trivial group when $l=1$. \begin{theorem}\label{rational} For all $g\ge 3$ and all $l \ge 1$, the completion of the mapping class group $\Gamma_{g,r}^n[l]$ relative to the homomorphism $\rho : \Gamma_{g,r}^n[l] \to Sp_g({\mathbb Q})$ induced by (\ref{map}) is defined over ${\mathbb Q}$. \qed \end{theorem} This result was proved in \cite[(4.14)]{hain:comp} under the assumption that $g \ge 8$ and that $l=1$. That the stronger result is true follows from the strengthening \cite{borel:improved} of Borel's stability theorem \cite{borel:triv,borel:twisted} for the symplectic group, stated below, which ensures that the hypothesis \cite[(4.10)]{hain:comp} is satisfied when $l\ge 1$ and $g \ge 3$. \begin{theorem}\label{imp_borel} Suppose that $V$ is an irreducible rational representation of the algebraic group $Sp_g$ and that $\Gamma$ is a finite index subgroup of $Sp_g({\mathbb Z})$. If $k < g$, then $H^k(\Gamma,V)$ vanishes when $V$ is non-trivial, and agrees with the stable cohomology of $Sp_g({\mathbb Z})$ when $V$ is the trivial representation. \qed \end{theorem} Denote the completion of $\Gamma_{g,r}^n$ relative to $\rho$ by $\tilde{\rho} : \Gamma_{g,r}^n \to {\mathcal G}_{g,r}^n$. \label{comp_def} Denote the prounipotent radical of ${\mathcal G}_{g,r}^n$ by \label{ugp_def} ${\mathcal U}_{g,r}^n$, and its Lie algebra by \label{ulie_def} ${\mathfrak u}_{g,r}^n$. \begin{proposition}\label{level} If $g\ge 3$, then for all $l\ge 1$, the composite $$ \Gamma_{g,r}^n[l] \hookrightarrow \Gamma_{g,r}^n \to {\mathcal G}_{g,r}^n $$ is the completion of $\Gamma_{g,r}^n[l]$ relative to the restriction of $\rho$ to $\Gamma_{g,r}^n[l]$. \end{proposition} \begin{proof} This follows directly from results in \cite[\S4]{hain:comp} as we shall explain. Denote the relative completion of $\Gamma_{g,r}^n[l]$ by ${\mathcal G}_{g,r}^n[l]$ and its prounipotent radical by ${\mathcal U}_{g,r}^n[l]$. There is a natural map ${\mathcal U}_{g,r}^n[l] \to {\mathcal U}_{g,r}^n$, the surjectivity of which follows from (\ref{imp_borel}) and \cite[(4.6)]{hain:comp}. Injectivity follows directly from (\ref{imp_borel}) and \cite[(4.13)]{hain:comp}. \end{proof} We have an extension $$ 1 \to {\mathcal U}_{g,r}^n \to {\mathcal G}_{g,r}^n \to Sp_g \to 1 $$ of proalgebraic groups over ${\mathbb Q}$. The homomorphism $\tilde{\rho}$ induces a map $T_{g,r}^n \to {\mathcal U}_{g,r}^n$. Denote the classical Malcev completion of \label{comptor_def} $T_{g,r}^n$ by ${\mathcal T}_{g,r}^n$, and its Lie algebra by \label{lietor_def} ${\mathfrak t}_{g,r}^n$. Since ${\mathcal U}_{g,r}^n$ is prounipotent, $\tilde{\rho}$ induces a homomorphism $$ \theta : {\mathcal T}_{g,r}^n \to {\mathcal U}_{g,r}^n $$ of prounipotent groups. The following theorem is the main result of \cite{hain:comp}.% \footnote{There is a minor error in proof of the case ``$A_{g,r}^n$ implies $A_{h,r}^n$'' of the proof of \cite[(7.4)]{hain:comp}. It is easily fixed.} There it is proved for all $g\ge 8$, but in view of (\ref{imp_borel}), it holds for all $g\ge 3$. (Cf.\ the third footnote on page~76 of \cite{hain:comp}.) \begin{theorem}\label{central_ext} For all $g\ge 3$, the homomorphism $\theta$ is surjective and has a one dimensional kernel isomorphic to $\Ga$ which is central in ${\mathcal T}_{g,r}^n$ and is trivial as an $Sp_g({\mathbb Z})$ module. Moreover, the extensions are all pulled back from that of ${\mathcal T}_g$; that is, the diagram $$ \begin{CD} 0 @>>> \Ga @>>> {\mathcal T}_{g,r}^n @>>> {\mathcal U}_{g,r}^n @>>> 1 \cr @. @| @VVV @VVV \cr 0 @>>> \Ga @>>> {\mathcal T}_g @>>> {\mathcal U}_g @>>> 1 \end{CD} $$ commutes. \qed \end{theorem} It is a standard fact that the sequence $$ 1 \to \pi_{g,r}^n \to \Gamma_{g,r}^n \to \Gamma_g \to 1 $$ is exact. Restricting to $T_g$, we obtain an extension \begin{equation} \label{exten} 1 \to \pi_{g,r}^n \to T_{g,r}^n \to T_g \to 1. \end{equation} \begin{proposition}\label{h1_tor} If $g\ge 3$, then the extension (\ref{exten}) induces an exact sequence $$ 0 \to H_1(\pi_{g,r}^n,{\mathbb Q}) \to H_1(T_{g,r}^n,{\mathbb Q}) \to H_1(T_g,{\mathbb Q}) \to 0. $$ \end{proposition} \begin{proof} It follows from (\ref{h1_braid}) that the natural map $\pi_{g,r}^n \to \pi_g^{n+r}$ induces an isomorphism on $H_1$ with rational coefficients. The corresponding surjection $T_{g,r}^n \to T_g^{n+r}$ induces a map $$ \begin{CD} H_1(\pi_{g,r}^n,{\mathbb Q}) @>>> H_1(T_{g,r}^n,{\mathbb Q}) @>>> H_1(T_g,{\mathbb Q}) @>>> 0\cr @VVV @VVV @| \cr H_1(\pi_g^{n+r},{\mathbb Q}) @>>> H_1(T_g^{n+r},{\mathbb Q}) @>>> H_1(T_g,{\mathbb Q}) @>>> 0\cr \end{CD} $$ of exact sequences. Since the middle vertical map is a surjection, and the two outside maps are isomorphisms, it follows that the middle map is an isomorphism. It therefore suffices to prove the result when $r=0$. To prove this, we need to prove that the map \begin{equation}\label{red} H_1(\pi_g^n,{\mathbb Q}) \to H_1(T_g^n,{\mathbb Q}) \end{equation} is injective. We first remark that this is easily proved when $n=1$ using the Johnson homomorphism $$ \tau_g^1 : H_1(T_g^1) \to H_3(\Jac C). $$ (Cf.\ \cite{johnson:def} and \cite[\S 3]{hain:normal}.) The composition of $\tau_g^1$ with the map $$ H_1(\pi_g^1,{\mathbb Q}) \to H_3(\Jac C,{\mathbb Q}) $$ is easily seen to be the map $H_1(C) \to H_3(\Jac C)$ which takes a class in $H_1$ to its Pontrjagin product with the class of $C$ in $H_2(\Jac C)$. Since this map is injective, it follows that (\ref{red}) is injective when $n=1$. The general case follows from this by considering the maps $p_j : H_1(T_g^n) \to H_1(T_g^1)$ induced by the $n$ forgetful maps $T_g^n \to T_g^1$. \end{proof} Define ${\mathcal P}_{g,r}^n$ \label{comppi_def} to be the Malcev completion of $\pi_{g,r}^n$ and \label{liepi_def} ${\mathfrak p}_{g,r}^n$ to be the corresponding Malcev Lie algebra. Applying Malcev completion to (\ref{exten}) we obtain a sequence $$ {\mathfrak p}_{g,r}^n \to {\mathfrak t}_{g,r}^n \to {\mathfrak t}_g $$ of Malcev Lie algebras. \begin{proposition}\label{seq} If $g\neq 2$, then the sequence $$ 0 \to {\mathfrak p}_{g,r}^n \to {\mathfrak t}_{g,r}^n \to {\mathfrak t}_g \to 0 $$ associated to (\ref{exten}) is exact. \end{proposition} \begin{proof} By \cite[(5.6)]{hain:cycles} it suffices to verify two conditions. First, that $T_g$ acts unipotently on $H^1(\pi_{g,r}^n,{\mathbb Q})$; this follows from (\ref{h1_braid}). The second condition there is satisfied if, for example, the extension $$ 0 \to H_1(\pi_{g,r}^n,{\mathbb Q}) \to G \to T_g \to 1 $$ obtained by pushing (\ref{exten}) out along $\pi_{g,r}^n \to H_1(\pi_{g,r}^n,{\mathbb Q})$ is split. In our case this follows from (\ref{h1_tor}) as the extension above can be pulled back from the extension $$ 0 \to H_1(\pi_{g,r}^n,{\mathbb Q}) \to H_1(T_{g,r}^n,{\mathbb Q}) \to H_1(T_g,{\mathbb Q}) \to 0 $$ which is split for trivial reasons. \end{proof} The standard homomorphism $\Gamma_{g,r}^n \to \Gamma_g$ induces a homomorphism ${\mathcal G}_{g,r}^n \to {\mathcal G}_g$ of relative completions. The inclusion ${\mathcal P}_{g,r}^n \to {\mathcal T}_{g,r}^n$ induces a homomorphism ${\mathcal P}_{g,r}^n \to {\mathcal U}_{g,r}^n$. We therefore have a sequence $$ {\mathcal P}_{g,r}^n \to {\mathcal G}_{g,r}^n \to {\mathcal G}_g \to 1 $$ of proalgebraic groups. \begin{lemma}\label{exactness} If $g\ge 3$, then the sequence $$ 1 \to {\mathcal P}_{g,r}^n \to {\mathcal G}_{g,r}^n \to {\mathcal G}_g \to 1 $$ is exact. In particular, the sequence $$ 0 \to {\mathfrak p}_{g,r}^n \to {\mathfrak u}_{g,r}^n \to {\mathfrak u}_g \to 0 $$ is exact. \end{lemma} \begin{proof} To prove the result, it suffices to prove that the sequence $$ 1 \to {\mathcal P}_{g,r}^n \to {\mathcal U}_{g,r}^n \to {\mathcal U}_g \to 1 $$ is exact. But this follows immediately from (\ref{central_ext}) and (\ref{seq}). \end{proof} Suppose that ${\mathfrak g}$ is a finitely generated pronilpotent Lie algebra. Denote the group of automorphisms of ${\mathfrak g}$ by $\Aut{\mathfrak g}$. Denote the subgroup of $\Aut{\mathfrak g}$ consisting of the elements that act trivially on $H_1({\mathfrak g})$ by $L^1\Aut {\mathfrak g}$. Since the action of an automorphism on the graded quotients of the lower central series is determined by the action on the first graded quotient, $\Aut {\mathfrak g}$ is a proalgebraic group which is an extension $$ 1 \to L^1\Aut {\mathfrak g} \to \Aut {\mathfrak g} \to S \to 1 $$ of a closed subgroup $S$ of $\Aut H_1({\mathfrak g})$ by the prounipotent group consisting of those automorphisms of ${\mathfrak g}$ that act trivially on the graded quotients of the lower central series. Its Lie algebra is the Lie algebra $\Der {\mathfrak g}$ of derivations of ${\mathfrak g}$. This is an extension $$ 0 \to L^1\Der {\mathfrak g} \to \Der {\mathfrak g} \to {\mathfrak s} \to 0 $$ of the Lie algebra ${\mathfrak s}$ of $S$ by the pronilpotent Lie algebra of derivations of ${\mathfrak g}$ that act trivially on the graded quotients of the lower central series of ${\mathfrak g}$. If ${\mathcal G}$ is the prounipotent group corresponding to ${\mathfrak g}$, then $\Aut {\mathcal G}$ and $\Aut {\mathfrak g}$ are isomorphic, as can be seen using the Baker-Campbell-Hausdorff formula. \begin{lemma}\label{rep} For all $g\ge 0$ the natural action of $\Gamma_{g,r}^n$ on $\pi_{g,r}^n$ induces a representation $$ {\mathcal G}_{g,r}^n \to \Aut {\mathfrak p}_{g,r}^n. $$ \end{lemma} \begin{proof} Suppose that $g \ge 0$. The mapping class group $\Gamma_{g,r}^n$ acts on ${\mathfrak p}_{g,r}^n$. We therefore have a homomorphism \begin{equation}\label{act} \Gamma_{g,r}^n \to \Aut {\mathfrak p}_{g,r}^n. \end{equation} By (\ref{h1_braid}) we know that $$ \Aut H_1({\mathfrak p}_{g,r}^n) = \Aut H_1(S)^{\oplus(n+r)}. $$ There is a diagonal copy of $Sp_g$ contained in this group, and it is easy to see that this is the Zariski closure of the image of $\Gamma_{g,r}^n$ in $\Aut H_1({\mathfrak p}_{g,r}^n)$. It follows that the Zariski closure of the image of (\ref{act}) is an extension of this diagonal $Sp_g$ by a prounipotent group. Since the homomorphism from $\Gamma_{g,r}^n$ to this copy of $Sp_g$ is the standard representation, the universal mapping property of the relative completion implies that (\ref{act}) induces a homomorphism ${\mathcal G}_{g,r}^n \to \Aut{\mathfrak p}_{g,r}^n$. \end{proof} \begin{remark}\label{sl2} When $g=1$, the results (\ref{level}) and (\ref{central_ext}) are false. That (\ref{level}) and (\ref{central_ext}) fail can be deduced from \cite[(10.3)]{hain:derham}, a special case of which states that there is a natural isomorphism $$ H^1({\mathcal M}_1[l],S^n{\mathbb V}) \cong \left(H^1({\mathcal U}_{1}[l])\otimes S^n{\mathbb V}\right)^{SL_2}. $$ Here ${\mathcal M}_1[l]$ denotes the moduli space of elliptic curves with a level $l$ structure, ${\mathbb V}$ denotes the variation of Hodge structure over ${\mathcal M}_1[l]$ of weight 1 corresponding to $H^1$ of the universal elliptic curve, and $S^n {\mathbb V}$ denotes its $n$th symmetric power. Since the level $l$ congruence subgroup of $SL_2({\mathbb Z})$ is free when $g\ge 4$, it follows by an Euler characteristic argument that $H^1({\mathcal M}_1[l],S^k{\mathbb V})$ is non-zero whenever $g\ge 4$. Since ${\mathcal T}_1$ is trivial, it cannot surject onto ${\mathcal U}_1[l]$. Since the rank $r_l$ of the level $l$ subgroup of $SL_2({\mathbb Z})$ depends on $l$, and since $$ \dim H^1({\mathcal M}_1[l],S^n{\mathbb V}) = (r_l-1)\dim S^k{\mathbb V}, $$ it follows that the rank of the $S^n{\mathbb V}$ isotypical part of $H_1({\mathcal U}_1[l])$ depends on $l$. So (\ref{level}) does not hold. \end{remark} \section{Mixed Hodge Structures on Torelli Groups} Denote by ${\mathcal M}_{g,r}^n[l]$ \label{mod_def} the moduli space of ordered $(n+r+1)$-tuples $$ (C;x_1,\dots,x_n;v_1,\dots,v_r) $$ where $C$ is a smooth complex projective curve with a level $l$ structure, the $x_j$ are distinct points of $C$, and the $v_j$ are non-zero holomorphic tangent vectors of $C$ which are anchored at $r$ distinct points of $C$ which are also distinct from the $x_j$. We shall omit the $l$ when it is 1, and $r$ and $n$ when they are zero. So, for example, ${\mathcal M}_g$ denotes the moduli space of smooth projective curves of genus $g$. For each point $x$ of ${\mathcal M}_{g,r}^n[l]$, there is a natural isomorphism of $\Gamma_{g,r}^n[l]$ with the (orbifold) fundamental group $\pi_1({\mathcal M}_{g,r}^n[l],x)$ of ${\mathcal M}_{g,r}^n[l]$. We will denote the latter by $\Gamma_{g,r}^n[l](x)$. We shall denote the subgroup of $\Gamma_{g,r}^n(x)$ corresponding to $T_{g,r}^n$ by $T_{g,r}^n(x)$. Denote the relative Malcev completion of $\Gamma_{g,r}^n(x)$ by ${\mathcal G}_{g,r}^n(x)$, its prounipotent radical by ${\mathcal U}_{g,r}^n(x)$, etc. The Lie algebras corresponding to $T_{g,r}^n(x)$ and $U_{g,r}^n(x)$ will be denoted by ${\mathfrak t}_{g,r}^n(x)$ and ${\mathfrak u}_{g,r}^n(x)$, respectively. In this section we prove that for each choice of a point $x$ in ${\mathcal M}_{g,r}^n$, there is a canonical ${\mathbb Q}$ mixed Hodge structure (MHS) on ${\mathfrak t}_{g,r}^n(x)$. The first ingredient in the construction of this MHS is the following theorem, which is proved in \cite[(13.1)]{hain:derham}. \begin{theorem}\label{mhs_gen} Suppose that $X$ is a smooth quasi-projective algebraic variety and $({\mathbb V},\langle\phantom{x},\phantom{x}\rangle)$ is a polarized variation of Hodge structure over $X$ of geometric origin whose monodromy representation $$ \rho : \pi_1(X,x_o) \to \Aut_{\mathbb R}(V_o,\langle\phantom{x},\phantom{x}\rangle) $$ has Zariski dense image. Then the coordinate ring of the completion of $\pi_1(X,x_o)$ relative to $\rho$ and its unipotent radical both have natural real MHSs such that the product, coproduct, and antipode of each are morphisms of MHSs. \qed \end{theorem} We will say that a homomorphism ${\mathcal G} \to{\mathcal H}$ between proalgebraic groups, each of whose coordinate rings is a Hopf algebra in the category of mixed Hodge structures, is a morphism of MHSs if the corresponding map on coordinate rings is. Since $\Gamma_{g,r}^n(x)$ is the orbifold fundamental group of $({\mathcal M}_{g,r}^n,x)$, the following result is not unexpected. \begin{theorem} For all $g,r,n \ge 0$, and for each choice of a point $$ x = [C;x_1,\dots,x_n;v_1,\dots,v_r] $$ of ${\mathcal M}_{g,r}^n$, there is a canonical real MHS on the coordinate ring of ${\mathcal G}_{g,r}^n(x)$ for which the product, coproduct and antipode are morphisms of MHS. Moreover, the homomorphisms ${\mathcal G}_{g,r}^n(x) \to {\mathcal G}_{g,r}^{n-1}(x')$ and ${\mathcal G}_{g,r}^n(x) \to {\mathcal G}_{g,r-1}^{n+1}(x'')$, induced by forgetting a point or by replacing a tangent vector by its anchor point, are morphisms of mixed Hodge structure. \end{theorem} \begin{proof} Since the mapping class group is not, in general, the fundamental group of ${\mathcal M}_{g,r}^n$, we need to pass to a level. Choose an integer $l$ such that $\Gamma_{g,r}^n[l]$ is torsion free. In this case, the moduli space ${\mathcal M}_{g,r}^n[l]$ is smooth and has fundamental group isomorphic to $\Gamma_{g,r}^n[l]$. Since $\Gamma_{g,r}^n[l]$ is torsion free, there is a universal curve $$ \pi : {\mathcal C} \to {\mathcal M}_{g,r}^n[l]. $$ Take ${\mathbb V}$ to be the dual of the local system $R^1\pi_\ast {\mathbb Z}$. This is a polarized variation of Hodge structure of weight $-1$ and is clearly of geometric origin. Its monodromy representation is $$ \rho : \Gamma_{g,r}^n[l] \to Sp_g({\mathbb Z}). $$ So by (\ref{level}) and (\ref{mhs_gen}), there is a canonical real MHS on the coordinate ring of the relative completion ${\mathcal G}_{g,r}^n(x)$ of $\Gamma_{g,r}^n(x)$ for each choice of a point of ${\mathcal M}_{g,r}^n[l]$ that lies over $x$. Denote the projection ${\mathcal M}_{g,r}^n[l] \to {\mathcal M}_{g,r}^n$ by $p$. The set of lifts of a point $x$ of ${\mathcal M}_{g,r}^n$ to ${\mathcal M}_{g,r}^n[l]$ is permuted transitively by the Galois group $Sp_g({\mathbb Z}/l{\mathbb Z})$. It follows from the naturality of the MHS on the relative completion that the MHSs on ${\mathcal G}_{g,r}^n(x)$ with respect any two points of $p^{-1}(x)$ are canonically isomorphic. The MHS on $\Gamma_{g,r}^n(x)$ is therefore indepenent of the choice of a point of $p^{-1}(x)$, and is therefore canonical. To show that the MHS on $\Gamma_{g,r}^n(x)$ constructed above is independent of the choice of the level $l$, suppose that $l_1$ and $l_2$ are two levels for which the mapping class group is torsion free. One can then compare the corresponding MHSs by passing to the level corresponding to the least common multiple of $l_1$ and $l_2$. The naturality statement follows directly from \cite[(13.12)]{hain:derham}. \end{proof} \begin{corollary} For all $g\ge 0$, and for each choice of a point $$ x = [C;x_1,\dots,x_n;v_1,\dots,v_r] $$ of ${\mathcal M}_{g,r}^n$, the pronilpotent Lie algebra ${\mathfrak u}_{g,r}^n(x)$ of the prounipotent radical of ${\mathcal G}_{g,r}^n(x)$ has a canonical real MHS for which the bracket is a morphism of MHS. Moreover, the morphisms ${\mathfrak u}_{g,r}^n(x) \to {\mathfrak u}_{g,r}^{n-1}(x')$ and ${\mathfrak u}_{g,r}^n(x) \to {\mathfrak u}_{g,r-1}^{n+1}(x")$, obtained by forgetting a point or replacing a tangent vector by its anchor point, are morphisms of mixed Hodge structure. \qed \end{corollary} Given a point $x$ of ${\mathcal M}_{g,r}^n$, there are two {\it a priori} different MHSs on ${\mathfrak p}_{g,r}^n(x)$. The first is the one obtained from the construction given in \cite{hain:dht}. The second arises as ${\mathfrak p}_{g,r}^n(x)$ is the kernel of the natural surjection ${\mathfrak u}_{g,r}^n(x) \to {\mathfrak u}_g(x)$. The following assertion follows directly from the naturality properties \cite[(13.12)]{hain:derham} of the mixed Hodge structure relative completions. \begin{proposition} These two MHSs are identical.\qed \end{proposition} Fix a point $x$ of ${\mathcal M}_{g,r}^n$. Then both of ${\mathcal G}_{g,r}^n(x)$ and ${\mathfrak p}_{g,r}^n(x)$ have canonical MHSs. It is natural to expect that the natural action \begin{equation}\label{action} {\mathcal G}_{g,r}^n(x) \to \Aut {\mathfrak p}_{g,r}^n(x) \end{equation} constructed in (\ref{rep}) is compatible with these. \begin{lemma}\label{action_morph} The action (\ref{action}) is a morphism of MHS. Consequently, the morphism $$ {\mathfrak u}_{g,r}^n(x) \to \Der {\mathfrak p}_{g,r}^n(x) $$ is also morphism of MHS with respect to the canonical MHSs determined by $x\in {\mathcal M}_{g,r}^n$. \end{lemma} \begin{proof} It follows immediately from (\ref{exactness}) that ${\mathcal P}_{g,r}^n(x)$ is a normal subgroup of ${\mathcal G}_{g,r}^n(x)$. Since the coordinate ring of ${\mathcal G}_{g,r}^n(x)$ has a natural mixed Hodge structure compatible with its operations, the action of ${\mathcal G}_{g,r}^n(x)$ on ${\mathcal P}_{g,r}^n(x)$ via conjugation is a morphism of MHS. But this action is easily seen to coincide with the canonical action of ${\mathcal G}_{g,r}^n(x)$ on ${\mathcal P}_{g,r}^n(x)$. \end{proof} For a curve $C$ of genus $g\ge 3$, denote by $PH_3(\Jac C,{\mathbb Q})$ the {\it primitive three dimensional homology} of its jacobian $\Jac C$ --- that is, the subspace of $H_3(\Jac C,{\mathbb Q})$ corresponding to $PH^{2g-3}(\Jac C,{\mathbb Q})$ under Poincar\'e duality. It has a natural Hodge structure of weight $-3$. \begin{proposition}\label{purity} If $g \ge 3$, then for each $x = [C;x_1,\dots,x_n;v_1,\dots,v_r] \in {\mathcal M}_{g,r}^n$ the canonical real mixed Hodge structure on $H_1({\mathfrak u}_{g,r}^n(x))$ is of weight $-1$ and is canonically isomorphic to $$ PH_3(\Jac C,{\mathbb R}(-1)) \oplus H_1(C,{\mathbb R})^{\oplus(r+n)}. $$ \end{proposition} \begin{proof} As in the proof of (\ref{h1_tor}), we reduce to proof to showing that it is true for ${\mathfrak u}_g^1$. Then, by (\ref{action_morph}), the composite \begin{equation}\label{comp} H_1({\mathfrak u}_g^1(x)) \to W_{-1}H_1(\Der {\mathfrak p}_g^1(x)) \to \Gr^W_{-1}\Der{\mathfrak p}_g^1(x) \end{equation} is a morphism of MHS. Observe that $$ \Gr^W_{-1}\Der{\mathfrak p}_g^1(x) \subset \Hom(\Gr^W_{-1}{\mathfrak p}_g^1(x),\Gr^W_{-2}{\mathfrak p}_g^1(x)). $$ {}From the work of Johnson \cite{johnson:def} (see also \cite[\S4]{hain:normal}), it follows that (\ref{comp}) is injective, from which the result follows for ${\mathfrak u}_g^1$. \end{proof} The fact that $H_1({\mathfrak u}_{g,r}^n)$ is pure of weight $-1$ allows us to conclude that the weight filtration of ${\mathfrak u}_{g,r}^n$ is essentially its lower central series. This follows from the following general fact. \begin{lemma}\label{wt=lcs} Suppose that ${\mathfrak g}$ is a pronilpotent Lie algebra in the category of mixed Hodge structures with finite dimensional $H_1$. If the induced MHS on $H_1({\mathfrak g})$ is pure of weight $-1$, then $W_{-l}{\mathfrak g}$ is the $l$th term of the lower central series of ${\mathfrak g}$. \end{lemma} \begin{proof} Denote the $l$th term of the lower central series of ${\mathfrak g}$ by ${\mathfrak g}^{(l)}$. Since $H_1({\mathfrak g})$ is pure of weight $-1$, since ${\mathfrak g}$ is pronilpotent, and since the bracket is a morphism of MHS, it follows that ${\mathfrak g} = W_{-1}{\mathfrak g}$. An elementary argument using the Jacobi identity shows that the bracket \begin{equation}\label{bra} {\mathfrak g} \otimes {\mathfrak g}^{(l)} \to {\mathfrak g}^{(l+1)} \end{equation} is surjective. Since the bracket is a morphism of MHS, it follows that ${\mathfrak g}^{(l)} \subseteq W_{-l}{\mathfrak g}$. The fact that $H_1({\mathfrak g})$ is pure of weight $-1$, forces ${\mathfrak g}^{(2)} = W_{-2}\,{\mathfrak g}$. The result now follows by an induction argument (induct on $l$) using the fact that (\ref{bra}), being a morphism of MHS, is strict with respect to the weight filtration. \end{proof} \begin{corollary}\label{lcs} The $l$th term of the lower central series of ${\mathfrak u}_{g,r}^n$ is $W_{-l}{\mathfrak u}_{g,r}^n$. \qed \end{corollary} This result implies that the weight filtration on ${\mathcal G}_{g,r}^n$ is defined over ${\mathbb Q}$ which implies that this MHS is really defined over ${\mathbb Q}$. \begin{corollary} The weight filtration of the canonical MHSs on ${\mathcal G}_{g,r}^n(x)$ and ${\mathfrak u}_{g,r}^n(x)$ associated to a point of ${\mathcal M}_{g,r}^n$ are topologically determined and therefore defined over ${\mathbb Q}$. Consequently, the MHSs on ${\mathcal G}_{g,r}^n(x)$ and ${\mathfrak u}_{g,r}^n(x)$ each have a canonical lift to ${\mathbb Q}$-MHSs. \qed \end{corollary} We are now ready to lift the MHS from ${\mathfrak u}_{g,r}^n(x)$ to ${\mathfrak t}_{g,r}^n(x)$: \begin{theorem}\label{mhs_torelli} Suppose that $g\neq 2$ and that $r,n \ge 0$. For each choice of a base point $$ x = \left[C;x_1,\dots,x_n;v_1,\dots,v_r\right] $$ of ${\mathcal M}_{g,r}^n$ there is a canonical ${\mathbb Q}$-MHS on ${\mathfrak t}_{g,r}^n(x)$ for which the bracket and the quotient map ${\mathfrak t}_{g,r}^n(x) \to u_{g,r}^n(x)$ are morphisms of MHS. Moreover, $W_{-l}{\mathfrak t}_{g,r}^n(x)$ is the $l$th term of the lower central series of ${\mathfrak t}_{g,r}^n(x)$ and the central $\Ga$ is isomorphic to ${\mathbb Q}(1)$. \end{theorem} \begin{proof} For all $g \ge 0$, we have the exact sequence $$ 1 \to \pi_{g,r}^n \to T_{g,r}^n \to T_g \to 1. $$ When $g =0,1$, $T_g$ is the trivial group, so that $T_{g,r}^n$ is isomorphic to $\pi_{g,r}^n$. It follows that in these cases ${\mathfrak t}_{g,r}^n$ is isomorphic to the Malcev Lie algebra ${\mathfrak p}_{g,r}^n$ associated to $\pi_{g,r}^n$. The choice of the base point of ${\mathcal M}_{g,r}^n$ gives the configuration space $(F_{g,r}^n,f_o)$ the structure of a pointed smooth complex algebraic variety. Since $\pi_{g,r}^n$ is the fundamental group of $(F_{g,r}^n,f_o)$, the existence of the MHS on ${\mathfrak t}_{g,r}^n(x)$ when $g=0,1$ follows from \cite[(6.3.1)]{hain:dht}. Now suppose that $g\ge 3$. To construct a MHS on ${\mathfrak t}_{g,r}^n(x)$, it suffices to show that ${\mathfrak t}_g(x)$ has a MHS such that ${\mathfrak t}_g(x) \to {\mathfrak u}_g(x)$ is a morphism as it follows from (\ref{central_ext}), (\ref{seq}) and (\ref{exactness}) that the diagram $$ \begin{CD} {\mathfrak t}_{g,r}^n(x) @>>> {\mathfrak t}_g(x) \cr @VVV @VVV \cr {\mathfrak u}_{g,r}^n(x) @>>> {\mathfrak u}_g(x) \cr \end{CD} $$ is a pullback square in the category of pronilpotent Lie algebras. It is useful to begin by explaining the philosophy behind the proof. The essential point is that $\Gamma_g(x)$ acts on ${\mathfrak t}_g(x)$ and on ${\mathfrak u}_g(x)$ --- the action is induced by the action of $\Gamma_g(x)$ on $T_g(x)$ by conjugation. The central extension \begin{equation}\label{extn} 0 \to \Ga \to {\mathfrak t}_g(x) \to {\mathfrak u}_g(x) \to 0. \end{equation} given by (\ref{central_ext}) can be viewed as an extension of local systems over ${\mathcal M}_g$ (in the orbifold sense, of course) where $\Ga$ is a trivial local system. Although we have not proved it yet, ${\mathfrak u}_g(x)$ should be a variation of MHS over ${\mathcal M}_g$. So we should try to construct the MHS on ${\mathfrak t}_g(x)$ so that (\ref{extn}) is both an extension of local systems and an extension of mixed Hodge structures. This, and the fact that the bracket has to be a morphism of MHS, gives us no choice. We now carry this out this program. The first point is that we know that, since $H_1({\mathfrak t}_g)$ does not contain any copies of the trivial representation, the central $\Ga$ is contained in $[{\mathfrak t}_g,{\mathfrak t}_g]$. The second is that by the computations in \S8 of \cite{hain:comp} we know that the central $\Ga$ lies in the image of the map $$ \Lambda^2 H_1({\mathfrak t}_g) \to \Gr^{\mathrm{lcs}}_{-2} {\mathfrak t}_g $$ induced by the bracket. So in order that the bracket be a morphism of MHS, the central $\Ga$ must be of weight $-2$. Since it is one dimensional, it has to be isomorphic to ${\mathbb Q}(1)$. Now fix a base point $x$ of ${\mathcal M}_g$. There is a corresponding ${\mathbb Q}$-MHS on ${\mathfrak u}_g(x)$. To lift this MHS to ${\mathfrak t}_g(x)$, we have to give an element of $$ \Ext^1_{\mathcal H}({\mathfrak u}_g(x),{\mathbb Q}(1)), $$ where $\Ext_{\mathcal H}$ denote the Ext group in the category ${\mathcal H}$ of ${\mathbb Q}$ mixed Hodge structures. Applying the functor $\Ext_{\mathcal H}^{\bullet}$ to the sequence $$ 0 \to W_{-2} {\mathfrak u}_g(x) \to {\mathfrak u}_g(x) \to H_1({\mathfrak u}_g(x)) \to 0, $$ we see that the natural map $$ \Ext^1_{\mathcal H}(H_1({\mathfrak u}_g(x)),{\mathbb Q}(1)) \to \Ext^1_{\mathcal H}({\mathfrak u}_g(x),{\mathbb Q}(1)) $$ is an isomorphism. Now let the base point vary. By (\ref{purity}), $H_1({\mathfrak u}_g)$ is a variation of ${\mathbb Q}$-MHS over ${\mathcal M}_g$ (in the orbifold sense) of weight $-1$ --- cf.\ \cite[(9.1)]{hain:normal}. Denote the category of admissible variations of ${\mathbb Q}$-MHS over a smooth variety $X$ by ${\mathcal H}(X)$ and the category of ${\mathbb Q}$ local systems over $X$ by ${\mathcal L}(X)$. Then, by \cite[(8.4)]{hain:normal}, the the forgetful map \begin{equation}\label{ext_iso} \Ext^1_{{\mathcal H}({\mathcal M}_g)}(H_1({\mathfrak u}_g),{\mathbb Q}(1)) \to \Ext^1_{{\mathcal L}({\mathcal M}_g)}(H_1({\mathfrak u}_g),{\mathbb Q}(1)) \cong H^1(\Gamma_g,H^1({\mathfrak u}_g)) \end{equation} is an isomorphism.\footnote{This group is one dimensional and generated by the Johnson homomorphism, although we do not really need to know this here --- see \cite[(5.2)]{hain:normal}.} We will lift the MHS on ${\mathfrak u}_g(x)$ to ${\mathfrak t}_g(x)$ using an element of the right hand group.\footnote{The class we seek, not surprisingly, is the one corresponding to the Johnson homomorphism, and is half the class associated to the algebraic cycle $C-C^-$ in $\Jac C$ --- see \cite[\S8]{hain:normal}.} We do this by producing an element of the right hand group which corresponds to the central extension (\ref{extn}). Denote the $k$th term of the lower central series of ${\mathfrak t}_g(x)$ by ${\mathfrak t}_g(x)^{(k)}$. We can form the extension \begin{equation}\label{extn2} 0 \to {\mathfrak t}_g(x)^{(2)}/{\mathfrak t}_g(x)^{(3)} \to {\mathfrak t}_g(x)/{\mathfrak t}_g(x)^{(3)} \to H_1({\mathfrak t}_g(x)) \to 0 \end{equation} As has been pointed out above, the image of ${\mathfrak t}_g(x)^{(2)}/{\mathfrak t}_g(x)^{(3)}$ in the central $\Ga$ in ${\mathcal T}_{g,r}^n(x)$ is non-trivial. The kernel of the extension (\ref{extn2}) is a rational representation of $Sp_g$. Since $H_1({\mathfrak t}_g(x))$ is irreducible, its second exterior power contains exactly one copy of the trivial representation. There is therefore a unique non-zero $Sp_g$-invariant projection $$ {\mathfrak t}_g(x)^{(2)}/{\mathfrak t}_g(x)^{(3)} \to \Ga. $$ If we push the extension (\ref{extn2}) out along this map, we obtain a canonical element of $$ \Ext^1_{{\mathcal L}({\mathcal M}_g)}(H_1({\mathfrak u}_g),{\mathbb Q}(1)). $$ By the isomorphism (\ref{ext_iso}), we obtain an element of $$ \Ext^1_{{\mathcal H}({\mathcal M}_g)}(H_1({\mathfrak u}_g),{\mathbb Q}(1)) $$ which allows us, for each point $x$ of ${\mathcal M}_g$, to lift the MHS on ${\mathfrak u}_g(x)$ associated to $x$ to a MHS on ${\mathfrak t}_g(x)$. Our last task is to show that the bracket is a morphism of MHS. We only need show that the bracket preserves the Hodge and weight filtrations. First observe that since $\Ga$ is central and contained in $W_{-2}$, the bracket preserves the weight filtration, and its restriction to $W_{-2}{\mathfrak t}_g(x)$ is a morphism of MHS. It remains to show that the the bracket preserves the Hodge filtration. In view of these facts, it suffices to prove that $$ \left[F^p{\mathfrak t}_g(x),F^q{\mathfrak t}_g(x)\right] = 0 $$ when $p+q > -1$. This is easily deduced from the the fact that the map $$ [\phantom{x},\phantom{x}] : \Lambda^2 H_1({\mathfrak t}_g) \to \Gr^W_{-2} {\mathfrak t}_g(x) \stackrel{\text{proj}}{\to} \Ga\cong {\mathbb Q}(1) $$ induced by the bracket is a polarization of $H_1({\mathfrak t}_g(x))$ as it is $Sp_g$ equivariant and non-zero by results in \cite[\S7]{hain:comp}. \end{proof} \begin{remark} It follows immediately from (\ref{action_morph}) and (\ref{mhs_torelli}) that for each choice of base point in ${\mathcal M}_{g,r}^n$, the canonical morphism ${\mathfrak t}_{g,r}^n(x) \to \Der {\mathfrak p}_{g,r}^n(x)$ is a morphism of MHS. \end{remark} \section{Review of Continuous Cohomology} \label{cts_coho} In this section, we briefly review the theory of continuous cohomology of discrete groups, which is mainly developed in \cite{hain:cycles}. It will be our principal tool in proving that Torelli has a presentation with only quadratic relations. As a warm up, we show how it can be used to give a new and simpler proof of Morgan's theorem that the complex form of the Lie algebra associated to the fundamental group of a smooth variety has a weighted homogenous presentation with generators of weights equal to those occurring in $H_1(X)$, and relations of weight contained in those of $H_2(X)$. Define the continuous cohomology of a discrete group $\pi$ to be the direct limit of the rational cohomology of its finitely generated nilpotent quotients: $$ H_{\mathrm{cts}}^{\bullet}(\pi,{\mathbb Q}) := \lim_\to H^{\bullet}(N,{\mathbb Q}) \label{ctsgp_def} $$ where $N$ ranges over the finitely generated nilpotent quotients of $\pi$. There is an obvious natural homomorphism \begin{equation}\label{natural_homom} H_{\mathrm{cts}}^{\bullet}(\pi,{\mathbb Q}) \to H^{\bullet}(\pi,{\mathbb Q}). \end{equation} If $X$ is a topological space with fundamental group $\pi$, then we also have a natural homomorphism $$ H_{\mathrm{cts}}^{\bullet}(\pi,{\mathbb Q}) \to H^{\bullet}(X,{\mathbb Q}) $$ as there is a canonical map $H^{\bullet}(\pi) \to H^{\bullet}(X)$. \begin{proposition}\label{cts_ord} If $H_1(\pi,{\mathbb Q})$ is finite dimensional, then the natural homomorphism (\ref{natural_homom}) is an isomorphism in degree 1 and injective in degree 2. \qed \end{proposition} This is really a restatement of the result of Dennis Sullivan which asserts that the Lie algebra of the 1-minimal model of a space is the Malcev Lie algebra of the fundamental group. A more direct proof can be found, for example, in \cite[(5.1)]{hain:cycles}. We present a new proof because it is elementary. \begin{proof} The group $H^2(G,{\mathbb Q})$ parameterizes central extensions of $G$ by ${\mathbb Q}$. Suppose that $\alpha \in H_{\mathrm{cts}}^2(\pi,{\mathbb Q})$ whose image in $H^2(\pi,{\mathbb Q})$ is trivial. Then there is a nilpotent quotient $N$ of $\pi$ and an element $\tilde{\alpha}$ of $H^2(N,{\mathbb Q})$ that is a lift of $\alpha$. There is a central extension \begin{equation}\label{alpha} 1 \to {\mathbb Q} \to E \to N \to 1 \end{equation} corresponding to by $\tilde{\alpha}$. The key point to note is that $E$ is nilpotent. To say that the image of $\alpha$ is trivial in $H^2(\pi,{\mathbb Q})$ is to say that the pullback of the extension (\ref{alpha}) to $\pi$ is split. Composing a splitting of this projection with the projection of $\pi \to N$ gives a homomorphism $\pi \to E$ which lifts $\pi \to N$. Denote the image of $\pi$ in $E$ by $\tilde{N}$. It is easy to see that the pullback of the extension (\ref{alpha}) to $\tilde{N}$ splits. Since $\tilde{N}$ is a nilpotent quotient of $\pi$, the class $\alpha$ vanishes. \end{proof} Similarly, we can define the continuous cohomology of a pronilpotent Lie algebra ${\mathfrak g}$ to be the direct limit of the cohomology of its finite dimensional nilpotent quotients: $$ H_{\mathrm{cts}}^{\bullet}({\mathfrak g}) := \lim_\to H^{\bullet}({\mathfrak n}) \label{ctslie_def} $$ where ${\mathfrak n}$ ranges over the finite dimensional nilpotent quotients of ${\mathfrak g}$. A mild generalization of a theorem of Nomizu \cite{nomizu} states that for each finitely generated nilpotent group $N$ there is a natural isomorphism $$ H^{\bullet}({\mathfrak n}) \cong H^{\bullet}(N,{\mathbb Q}) $$ where ${\mathfrak n}$ is the Lie algebra of the Malcev completion of $N$. It follows immediately from the definitions that if $\pi$ is a finitely generated group and ${\mathfrak p}$ the associated Malcev Lie algebra, then there is a natural isomorphism $$ H_{\mathrm{cts}}^{\bullet}({\mathfrak p}) \cong H_{\mathrm{cts}}^{\bullet}(\pi,{\mathbb Q}). $$ The continuous cohomology of a pronilpotent Lie algebra ${\mathfrak g}$ can be computed using the standard complex ${\mathcal C}^{\bullet}({\mathfrak g})$ of continuous cochains of ${\mathfrak g}$. This is defined to be the direct limit of the Chevalley-Eilenberg cochains of the finite dimensional nilpotent quotients of ${\mathfrak g}$. Denote the continuous dual of ${\mathfrak g}$ by ${\mathfrak g}^\ast$. Then we have a d.g.a.\ isomorphism $$ {\mathcal C}^{\bullet}({\mathfrak g}) = \Lambda^{\bullet} \left({\mathfrak g}^\ast[-1]\right); $$ the differential is derivation of degree 1 whose restriction to ${\mathfrak g}^\ast$ is minus the dual of the bracket. The definition of continuous cohomology can be extended to the case where the coefficients are ${\mathbb Q}$ modules on which ${\mathfrak g}$ acts via a representation of one of its nilpotent quotients --- cf.\ \cite{hain:cycles}. Suppose now that $H_1({\mathfrak g})$ is finite dimensional. If ${\mathfrak g}$ has a MHS, then, by linear algebra, so do ${\mathcal C}^{\bullet}({\mathfrak g})$ and $H^{\bullet}({\mathfrak g})$. We will call such a pronilpotent Lie algebra a {\it Hodge Lie algebra}. It follows that if $X$ is an algebraic variety, $x\in X$, then $H_{\mathrm{cts}}^{\bullet}(\pi_1(X,x),{\mathbb Q})$ has a canonical MHS. One can show that this MHS does not depend on the base point $x$ of $X$ --- \cite{hain:cycles}. Since the weight filtration of a MHS splits canonically over ${\mathbb C}$, each finite dimensional Hodge Lie algebra ${\mathfrak g}$ is canonically isomorphic to the graded Lie algebra $\Gr^W_{\bullet}{\mathfrak g}$ after tensoring with ${\mathbb C}$. The following result therefore follows by taking inverse limits. \begin{proposition}\label{canon_split} If ${\mathfrak g}$ is a Hodge Lie algebra, all of whose weights are negative, then there is a canonical Lie algebra isomorphism $$ {\mathfrak g}_{\mathbb C} \cong \prod_{l\ge 1} \Gr^W_{-l}{\mathfrak g}_{\mathbb C}. \qed $$ \end{proposition} Since each choice of a base point of ${\mathcal M}_{g,r}^n$ determines a canonical MHS on ${\mathfrak t}_{g,r}^n$, we have: \begin{corollary} For each choice of a base point of ${\mathcal M}_{g,r}^n$, there is a canonical isomorphism $$ {\mathfrak t}_{g,r}^n\otimes {\mathbb C} \cong \prod_{l\ge 1}\left(\Gr^W_{-l} {\mathfrak t}_{g,r}^n \otimes {\mathbb C}\right). \qed $$ \end{corollary} The following result is proved, for example, in \cite[(11.7)]{carlson-hain}. In Section \ref{cts_coho_tor} we will prove a generalization needed for studying the relations in ${\mathfrak t}_g$. \begin{theorem}\label{cts_morph} If $X$ is a smooth algebraic variety, then the natural homomorphism $$ H_{\mathrm{cts}}^{\bullet}(\pi_1(X),{\mathbb Q}) \to H^{\bullet}(X,{\mathbb Q}) $$ is a morphism of mixed Hodge structures. \end{theorem} The final two results in this section together will allow us to use continuous cohomology as an effective tool for studying relations in Hodge Lie algebras in general, and ${\mathfrak t}_g$ in particular. The cochains, and therefore the cohomology, of a graded Lie algebra both have an extra grading, and are therefore bigraded algebras. If ${\mathfrak g}$ is a Hodge Lie algebra, then $\Gr^W_{\bullet}{\mathfrak g}$ has an extra grading is by weight. Since the functor $\Gr^W$ is exact on the category of MHS, we have: \begin{proposition} If ${\mathfrak g}$ is a Hodge Lie algebra, then there is a canonical bigraded algebra isomorphism $$ \Gr^W_{\bullet} H_{\mathrm{cts}}^{\bullet}({\mathfrak g}) \cong H^{\bullet}(\Gr^W_{\bullet} {\mathfrak g}). \qed $$ \end{proposition} If ${\mathfrak g}$ is a graded Lie algebra with negative weights, then we can write ${\mathfrak g}$ as a quotient of the free graded Lie algebra ${\mathfrak f}$ generated by $H_1({\mathfrak g})$ modulo a homogeneous ideal ${\mathfrak r}$. Note that we are not assuming that $H_1({\mathfrak g})$ is pure --- in general it will be graded. The group $$ H_0({\mathfrak f}/{\mathfrak r}) = {\mathfrak r}/[{\mathfrak f},{\mathfrak r}] $$ is graded. One can obtain a minimal set of relations of ${\mathfrak g}$ by taking the image of any splitting of the projection $$ {\mathfrak r} \to H_0({\mathfrak f}/{\mathfrak r}). $$ The following result is an analogue of Hopf's description of the second homology of a group in terms of a presentation. \begin{proposition} If ${\mathfrak g}$ is a graded Lie algebra with negative weights, then there is a canonical isomorphism of graded vector spaces $$ H_0({\mathfrak f}/{\mathfrak r}) \cong H_2({\mathfrak g}). $$ \end{proposition} \begin{proof} There are several ways to see this. One is look closely at the Chevalley-Eilenberg cochains of ${\mathfrak g}$. The second is the use the fact that a sub-Lie algebra of a free Lie algebra is free \cite% [(2.5)]{reutenauer} to deduce that, as a Lie algebra, ${\mathfrak r}$ is free. Then apply the Lie algebra analogue of the Hochschild-Serre spectral sequence to the extension $$ 0 \to {\mathfrak r} \to {\mathfrak f} \to {\mathfrak g} \to 0. $$ The details are standard and are omitted. \end{proof} \begin{corollary}\label{gr_presentn} If ${\mathfrak g}$ is a graded Lie algebra with negative weights, then there is an injective linear map $$ \delta : H_2({\mathfrak g}) \hookrightarrow {\mathbb L}(H_1({\mathfrak g})) $$ of graded vector spaces such that ${\mathfrak g}$ has presentation $$ {\mathbb L}(H_1({\mathfrak g}))/(\im \delta) $$ in the category of graded Lie algebras. \end{corollary} Combining (\ref{canon_split}), (\ref{gr_presentn}) and the existence of a canonical MHS on the Malcev Lie algebra ${\mathfrak g}(X,x)$ associated to a pointed variety, we obtain one of Morgan's theorems \cite[(10.3)]{morgan}. \begin{theorem}\label{morgan} If $X$ is a smooth complex algebraic variety and $x\in X$, then then the complex Malcev Lie algebra ${\mathfrak g}(X,x)_{\mathbb C}$ associated to $\pi_1(X,x)$ has the property that $$ {\mathfrak g}(X,x)_{\mathbb C} \cong \prod_{l \ge 1}\Gr^W_{-l}{\mathfrak g}_C $$ and there is a homomorphism of graded vector spaces $$ \delta : H_2(X,{\mathbb C}) \to {\mathbb L}(\Gr^W_{\bullet} H_1(X)) $$ such that $$ \Gr^W_{\bullet} {\mathfrak g}_{\mathbb C} \cong {\mathbb L}(\Gr^W_{\bullet} H_1(X,{\mathbb C}))/(\delta(\Gr^W_{\bullet} H_2(X,{\mathbb C}))) $$ in the category of graded Lie algebras. \qed \end{theorem} \section{Remarks on the Representations of $\sp_g$} \label{reps} In this section we review some basic facts from the representation theory that we shall need in subsequent sections. A good reference is \cite{fulton-harris}. Denote the Lie algebra of $Sp_g$ by \label{symp_def} $\sp_g$. The representation theory of the group and the Lie algebra are the same. Denote their common representation ring by \label{rep_def} $R(\sp_g)$. Choose a symplectic basis $a_1,\dots,a_g,b_1,\dots,b_g$ of the fundamental representation of $\sp_g$. Denote by ${\mathfrak h}$ the torus in $\sp_g$ consisting of matrices that are diagonal with respect to this basis. Choose coordinates $t=(t_1,\ldots,t_g)$ on ${\mathfrak h}$ so that $$ t\cdot a_i = t_ia_i \text{ and } t\cdot b_i = - t_i b_i. $$ The subalgebra of positive nilpotents ${\mathfrak n}$ has basis the elements $S_{i,j}, (i<j)$, $T_i$, and $F_{i,j}, (i\neq j)$ of $\sp_g$, where $$ S_{i,j} (a_j) = a_i, \quad S_{i,j}(b_i) = - b_j, \quad S_{i,j}(\text{other basis vectors}) = 0, $$ $$ T_i(b_i) = a_i, \quad T_i(\text{other basis vectors}) = 0, $$ $$ F_{i,j}(b_i) = a_j, \quad F_{i,j} (b_j) = a_i, \quad F_{i,j}(\text{other basis vectors}) = 0. $$ A fundamental set of weights of $\sp_g$ is $\lambda_j:{\mathfrak h} \to {\mathbb R}$, $1\le j \le g$, where $\lambda_j$ is defined by $$ \lambda_j(t) = t_1 + t_2 + \cdots + t_j. \label{wt_def} $$ The irreducible representations of $\sp_g$ correspond to positive integral linear combinations $\lambda$ of the $\lambda_j$. Denote the irreducible representation of $\sp_g$ with highest weight $\lambda$ by \label{module_def} $V(\lambda)$. The irreducible representations of $\sp_g$ can also be indexed by partitions $\alpha$ of an integer $n$ into $\le g$ parts: $$ n = \alpha_1 + \alpha_2 + \dots + \alpha_g $$ where $$ \alpha_1 \ge \alpha_2 \ge \dots \ge \alpha_g \ge 0. $$ The irreducible representation corresponding to $\alpha$ has highest weight $$ t \mapsto \sum_j \alpha_j t_j. $$ We shall denote the integer $$ \sum_{k=1}^g \alpha_k = \sum_{k=1}^g k\, n_k \label{size_def} $$ by $|\alpha|$ or by $|\lambda|$ according to whether we are using partitions or highest weights. This can be considered as a measure of the size of the corresponding irreducible representation; it is the smallest positive integer $d$ such that $V(\lambda)$ occurs in the $d$th tensor power of the fundamental representation. There is a notion of stability of the decomposition of tensor products and Schur functors of representations of symplectic groups. In order to state the result, we need to first define the {\it depth}, $\delta(V)$, of a representation $V$ of $\sp_g$. If the module is irreducible with highest weight $\sum n_k\,\lambda_k$, define $\delta(V)$ to be the largest $d$ such that $n_d\neq 0$ --- or equivalently, it is the number of rows in the corresponding Young diagram. Define the depth of an arbitrary representation to be the maximum of the depths of its irreducible components. In order to discuss stability, we will need a stabilization map. When $h \ge g$, define a group homomorphism $$ R(\sp_g) \hookrightarrow R(\sp_h) $$ by taking the irreducible representation of $\sp_g$ corresponding to the partition $\alpha$ to the representation of $\sp_h$ corresponding to the same partition. Equivalently, take the representation of $\sp_g$ with highest weight $\sum n_k\, \lambda_k$ to the representation of $\sp_h$ with the same highest weight decomposition. Recall that to each partition $\beta$ of a positive integer $n$, one has a Schur functor $\Schur_\beta$ defined on the category of representations of each group. For example, if $\beta = [n]$, then $\Schur_\beta$ is the $n$th symmetric power, if $\beta = [1^n]$, then $\Schur_\beta$ is the $n$th exterior power. We shall denote the integer $n$ by $|\beta|$. The second assertion of the following stability result appears to be folklore --- the only proof I know of is in Kabanov's thesis. \begin{theorem}\label{kab_stab} \begin{enumerate} \item (\cite[p.~424]{fulton-harris}) If $V$ and $W$ are representations of $\sp_g$ and $\delta(V) + \delta(W) \le g$, then the irreducible representations and their multiplicities occurring in the decomposition of $V\otimes W$ is independent of $g$. \item (\cite{kabanov} --- see also \cite{kabanov:stab}) If $V$ is a representation of $\sp_g$ and $\beta$ is a partition with $|\beta|\delta(V) \le g$, then the decomposition of $\Schur_\beta V$ into irreducible components is independent of $g$. \end{enumerate} \qed \end{theorem} \begin{remark}\label{method_comp} Some of the computations of highest weight decompositions in this paper have been made using the computer program {\textsf{LiE}}\ from the University of Amsterdam. The computations were performed for a particular $g$ in the stable range. The stability theorem was then used to deduce the decomposition for all $g$ in the stable range. Note that all such computations were checked using several values of $g$ in the stable range. In addition, many of the unstable computations were done using {\textsf{LiE}}. \end{remark} By composition with the canonical homomorphism $$ \Gamma_{g,r}^n \to Sp_g({\mathbb Q}) $$ we see that each representation $V$ of $\sp_g$ gives rise to a local system over ${\mathcal M}_{g,r}^n$, at least in the orbifold sense. It is a standard fact that each such local system arising from an irreducible representation of $\sp_g$ is an admissible variation of Hodge structure over ${\mathcal M}_{g,r}^n$ in a unique way up to Tate twist --- cf.\ \cite[(9.1)]{hain:normal}. It can always be realized as a variation of weight $|\lambda|$, and we shall take this as the default weight. We would like to discuss the cohomology of ${\mathcal M}_{g,r}^n$ with coefficients in such a local system. To do this, first choose a level $l$ such that $\Gamma_{g,r}^n$ is torsion free. Then ${\mathcal M}_{g,r}^n[l]$ is smooth and the variation of Hodge structure ${\mathbb V}$ corresponding to an irreducible representation $V$ of $Sp_g$ is defined over ${\mathcal M}_{g,r}^n[l]$ and has a natural $Sp_g({\mathbb Z}/l{\mathbb Z})$ action. From the work of M.~Saito \cite{saito}, we know that $H^k({\mathcal M}_{g,r}^n[l],{\mathbb V})$ has a canonical mixed Hodge structure with weights $\ge k + \text{weight}(V)$. The action of $Sp_g({\mathbb Z}/l{\mathbb Z})$ preserves this MHS. So we can define $$ H^{\bullet}({\mathcal M}_{g,r}^n,{\mathbb V}) = H^0(Sp_g({\mathbb Z}/l{\mathbb Z}),H^{\bullet}({\mathcal M}_{g,r}^n[l],{\mathbb V})). $$ as a MHS. Note that the underlying group is canonically isomorphic to $H^{\bullet}(\Gamma_{g,r}^n,V)$. \section{Continuous Cohomology of Torelli Groups} \label{cts_coho_tor} The next step in finding a presentation of ${\mathfrak t}_{g,r}^n$ is to determine the relations in $\Gr^W_{\bullet} {\mathfrak t}_{g,r}^n$. Since this is a graded Lie algebra generated in degree $-1$, the generators of the ideal of relations is homogeneous. In this section we will use a result of Kabanov \cite{kabanov} (see also \cite{kabanov:purity}) about the second cohomology of $\Gamma_{g,r}^n$ to show that the ideal of relations in $\Gr{\mathfrak t}_{g,r}^n$ is generated by quadratic and cubic generators when $g\ge 3$, and quadratic relations alone when $g\ge 6$. Our principal tool will be the continuous cohomology defined in Section \ref{cts_coho}. First some notation. Take $X$ and ${\mathbb V}$ as in the statement of (\ref{mhs_gen}). Denote the Lie algebra associated to the prounipotent radical of the completion of $\pi_1(X,x)$ relative to $\rho$ by ${\mathfrak u}(x)$. This is a Hodge Lie algebra. The next result is a generalization of (\ref{cts_morph}). \begin{proposition}\label{morphism} Suppose that ${\mathbb W}$ is an admissible variation of Hodge structure over $X$ which is a subquotient of a tensor power of ${\mathbb V}$. Then for all $k\ge 0$ and each $x\in X$, there is a natural homomorphism $$ H^0(X,H_{\mathrm{cts}}^k({\mathfrak u}(x))\otimes W_x) \to H^k(X,{\mathbb W}) $$ which is a morphism of MHS. It is an isomorphism when $k=1$ and injective when $k=2$. \end{proposition} \begin{proof} The case $k=1$ is proved in \cite[(10.3),(13.8)]{hain:derham}. We will prove the result when $k>1$ by induction. The most important case for us is when $k=2$, so we will give that argument in more detail and briefly sketch the remaining cases. We will assume throughout that the reader is familiar with \cite{hain:derham}. A convenient auxiliary reference for rational homotopy theory is \cite[\S2]{hain:dht}. We begin by recalling some well known facts from rational homotopy theory. The base point $x\in X$ determines an augmentation $$ \epsilon_x : \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathbb R}, $$ where $P$ is the principal bundle defined in \cite[\S 4]{hain:derham}. We shall write ${\mathcal O}$ instead of ${\mathcal O}(P)$. We can form the bar construction $$ B(\Efin^{\bullet}(X,{\mathcal O})) := B({\mathbb R},\Efin^{\bullet}(X,{\mathcal O}),{\mathbb R}) $$ where both copies of ${\mathbb R}$ are regarded as $\Efin^{\bullet}(X,{\mathcal O})$ modules via $\epsilon_x$. The Lie algebra ${\mathfrak u}(x)$ is determined by $B(\Efin^{\bullet}(X,{\mathcal O}))$ as follows: the dual $$ H^0(B(\Efin^{\bullet}(X,{\mathcal O}))) $$ is a complete Hopf algebra, ${\mathfrak u}(x)$ is its set of primitive elements. (See, for example, \cite[(2.4.5) and \S 2.6]{hain:dht}.) There is an augmentation preserving d.g.a.\ homomorphism \begin{equation}\label{min_model} {\mathcal C}^{\bullet}({\mathfrak u}(x)) \to \Efin^{\bullet}(X,{\mathcal O}), \end{equation} unique up to homotopy, which induces the map $$ \theta : H_{\mathrm{cts}}^{\bullet}({\mathfrak u}(x)) \to H^{\bullet}(\Efin^{\bullet}(X,{\mathcal O})) $$ on homology. The map $\theta$ is an isomorphism in degree 1 and injective in degree 2.\footnote{In the language of Sullivan \cite{sullivan}, the map (\ref{min_model}) is the 1-minimal model of $\Efin^{\bullet}(X,{\mathcal O})$.} There is a canonical isomorphism $$ H^k(X,{\mathbb W}) \cong H^0(X,H^k(\Efin^{\bullet}(X,{\mathcal O}))\otimes{\mathbb W}) $$ of MHSs for each VHS ${\mathbb W}$ over $X$ whose monodromy representation is the pullback of a rational representation of $\Aut(V_o,q)$ via the representation $\rho$. Since ${\mathfrak u}(x)$ has a canonical MHS, and since $\Efin^{\bullet}(X,{\mathcal O})$ is a mixed Hodge complex, each of the domain and target of $\theta$ have a canonical MHS. To prove the result, we need only prove that $\theta$ is a morphism of MHS. First we give an intuitive proof. The image of the map $$ \theta^2 : H_{\mathrm{cts}}^2({\mathfrak u}) \to H^2(\Efin^{\bullet}(X,{\mathcal O})) $$ is the subspace of the right hand side generated by the cup product $H^1\otimes H^1 \to H^2$, all Massey triple products of 1-forms, all Massey quadruple products of 1-forms, etc. Since the cup product and all Massey $k$-fold products have domain which is a sub-MHS of $\otimes^k H^1$ and are themselves morphisms, it follows that the image of $\theta^2$ is a MHS. That $\theta^2$ is a morphism follows as $\theta^1$ is an isomorphism of MHS. One can continue in an analogous fashion to prove that each $\theta^k$ is a morphism. We now make this argument precise. The spectral sequence associated to the standard filtration of the bar construction is called the Eilenberg-Moore spectral sequence (EMss): for an augmented d.g.a.\ $A^{\bullet}$ with connected homology, it takes the form $$ E_1^{-s,t} = \left[\otimes^s H^+(A)\right]^t \implies H^{t-s}(B(A)). $$ Denote the EMss associated to ${\mathcal C}({\mathfrak u})^{\bullet}$ by $\{E_r({\mathfrak u})\}$ and the EMss associated to $\Efin^{\bullet}(X,{\mathcal O})$ by $\{E_r(X)\}$. The map (\ref{min_model}) induces a morphism of Eilenberg-Moore spectral sequences. Each of these is a spectral sequence of MHSs as both the domain and target of (\ref{min_model}) are mixed Hodge complexes, but we have to prove that the map between them is a morphism of MHSs. This is the case in total degree 0 as $E_1^{-s,s}$ is $\otimes^s H^1$ and $\theta^1$ is an isomorphism of MHS. It is a standard fact that $$ H^k(B({\mathcal C}^{\bullet}({\mathfrak u}))) = 0 $$ when $k>0$; cf.\ \cite[(2.6.2)]{hain:dht} and \cite{bloch-kriz}. Therefore, the $E^{-1,2}_\infty$ term of the associated EMss vanishes. (This is a precise way to say that $H_{\mathrm{cts}}^2({\mathfrak u})$ is generated by Massey products.) The edge homomorphisms $$ H_{\mathrm{cts}}^k({\mathfrak u}) = E_1^{-1,k}({\mathfrak u}) \to E_r^{-1,k}({\mathfrak u}) $$ are all surjective. Let $M_r^k$ be the inverse image in $H_{\mathrm{cts}}^k({\mathfrak u})$ of the image of $$ d_{r-1} : E_{r-1}^{-r,k+r-2}({\mathfrak u}) \to E_{r-1}^{-1,k}({\mathfrak u}). $$ Then the fact that the higher cohomology of $B({\mathcal C}^{\bullet}({\mathfrak u}))$ vanishes implies that whenever $k\ge 2$ $$ H_{\mathrm{cts}}^k({\mathfrak u}) = \bigcup_r M_r^k. $$ Since the spectral sequence is a spectral sequence of MHS, each $M_r^k$ is a sub-MHS of $H_{\mathrm{cts}}^k({\mathfrak u})$. Since both spectral sequences are spectral sequences of MHSs, it follows that the image of $$ H_{\mathrm{cts}}^2({\mathfrak u})=E_1^{-1,2}({\mathfrak u}) \to E_1^{-1,2}(X) = H^2(\Efin^{\bullet}(X,{\mathcal O})) $$ is a sub-MHS and that $\theta^2$ is a morphism of MHS. If $k>2$, one can assume by induction that $\theta^m$ is a morphism whenever $m<k$. It follows easily that the natural map $$ E_1^{-s,t}({\mathfrak u}) \to E_1^{-s,t}(X) $$ is a morphism of MHS whenever $-s+t < k-1$, and therefore that its image is a sub-MHS of $E_1^{-s+t}(X)$. But since these spectral sequences are spectral sequences in the category of MHSs, and since $E_\infty^{-1,k}({\mathfrak u})$ vanishes, it follows that $\theta^k$ is a morphism. \end{proof} \begin{remark} This is a continuation of Remark~\ref{sl2}. The previous result implies that ${\mathcal U}_1[l]$ is a free pronilpotent group as $SL_2({\mathbb Z})[l]$ has a free subgroup of finite index, which implies that $H^2({\mathcal M}_1[l],S^n V)$ vanishes for all $n$. It follows that $H^2({\mathfrak u}_1[l])$ vanishes and from (\ref{gr_presentn}) that ${\mathfrak u}_1[l]$ is free. \end{remark} We thus have the following version of (\ref{morphism}) for moduli spaces of curves. \begin{proposition}\label{morph_u} If $g\ge 3$ and ${\mathbb V}$ is a variation of Hodge structure over ${\mathcal M}_{g,r}^n$ whose monodromy representation comes from a rational representation of $Sp_g$, then for all $k$, there is a natural map $$ H^0(Sp_g,H_{\mathrm{cts}}^k({\mathfrak u}_{g,r}^n(x))\otimes V_x) \to H^k({\mathcal M}_{g,r}^n,{\mathbb V}) $$ which is a morphism of MHS. Here $V_x$ denotes the fiber of ${\mathbb V}$ over $x$. \qed \end{proposition} This yields the following useful result about differentials in the Hochschild-Serre spectral sequence associated to the group extension \begin{equation}\label{std} 1 \to T_{g,r}^n \to \Gamma_{g,r}^n \to Sp_g({\mathbb Z}) \to 1. \end{equation} If we take coefficients in the irreducible representation $V(\lambda)$ of $Sp_g$, this spectral sequence takes the form $$ E_2^{s,t} = H^0(Sp_g({\mathbb Z}),H^t(T_{g,r}^n)\otimes V(\lambda)) \implies H^{s+t}(\Gamma_{g,r}^n,V(\lambda)). $$ \begin{corollary}\label{vanishing} For each $\lambda$, the image of the composite $$ H^0(Sp_g,H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n)\otimes V(\lambda)) \to H^0(Sp_g({\mathbb Z}),H^2(T_{g,r}^n)\otimes V(\lambda)) = E_2^{0,2} $$ is contained in $E_\infty^{0,2}$. \end{corollary} \begin{proof} The result follows immediately from the fact that the diagram $$ \begin{CD} H^0(Sp_g,H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n)\otimes V(\lambda)) @>>> H^0(Sp_g,H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n)\otimes V(\lambda)) \cr @VVV @VVV \cr H^2(\Gamma_{g,r}^n,V(\lambda)) @>>> H^0(Sp_g({\mathbb Z}),H^2(T_{g,r}^n)\otimes V(\lambda)) \end{CD} $$ commutes, where the top map is the surjection induced by the projection of $t_{g,r}^n$ onto ${\mathfrak u}_{g,r}^n$, the right hand vertical map by the canonical map $$ H_{\mathrm{cts}}^{\bullet}({\mathfrak t}_{g,r}^n) \to H^{\bullet}(T_{g,r}^n) $$ described in Section \ref{cts_coho}, the bottom map is the canonical restriction map, and the left hand vertical map is the one given by (\ref{morph_u}). This assertion can be proved using the constructions in \cite[\S4]{hain:derham} by restricting to a leaf in the principal bundle $P \to {\mathcal M}_{g,r}^n$ associated to the representation (\ref{map}). In this case, each leaf is a copy of the classifying space of $T_{g,r}^n$. \end{proof} Actually, we have proved a stronger statement than asserted. The stronger claim will be stated in \S\ref{applications}. Denote the fiber over $x\in {\mathcal M}_{g,r}^n$ of the variation of Hodge structure ${\mathbb V}(\lambda)$ corresponding to the irreducible representation $V(\lambda)$ of $Sp_g$ by $V(\lambda)_x$. \begin{corollary} If $g\ge 3$, then for each irreducible representation $V(\lambda)$ of $Sp_g$, there is a canonical monomorphism of MHS $$ H^0(Sp_g,H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n(x))\otimes V(\lambda)_x) \hookrightarrow H^2({\mathcal M}_{g,r}^n,{\mathbb V}(\lambda)). $$ \end{corollary} \begin{proof} We first prove the existence of the homomorphism. Injectivity will then follow directly from (\ref{morphism}). It follows from (\ref{central_ext}) and \cite[(5.5)]{hain:cycles} that the sequence of $Sp_g$ modules $$ 0 \to {\mathbb Q}(1) \to H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n(x)) \to H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n(x)) \to 0 $$ is an exact sequence of MHSs. It follows that the natural map $H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n(x)) \to H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n(x))$ is a surjective morphism of MHS. According to (\ref{morph_u}), the map $$ \left[H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n(x))\otimes V(\lambda)_x\right]^{Sp_g} \to H^2({\mathcal M}_{g,r}^n,{\mathbb V}(\lambda)) $$ is a morphism of MHS. So to construct the homomorphism, it suffices to show that this map factors through the quotient map $$ H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n) \to H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n). $$ We first assume that $\lambda \neq 0,\lambda_1,\lambda_3$. Consider the Hochschild-Serre spectral sequence $$ E_2^{s,t} = H^s(Sp_g({\mathbb Z}),H^t(T_{g,r}^n)\otimes V(\lambda)) \implies H^{s+t}(\Gamma_{g,r}^n,V(\lambda)). $$ By (\ref{imp_borel}), $E_2^{s,t}$ vanishes when $s\le 1$ and $t\le 2$ provided $g \ge 3$ (cf.\ \cite[(5.2)]{hain:normal}). It follows that $$ H^2(\Gamma_{g,r}^n,{\mathbb V}(\lambda)) = H^0(Sp_g({\mathbb Z}),H^2(T_{g,r}^n)\otimes V(\lambda)). $$ The result now follows from (\ref{cts_ord}). When $\lambda=\lambda_3$, we have $E_2^{2,1} \cong {\mathbb Q}$ (cf.\ \cite[(5.2)]{hain:normal}), so there is a possibility of having a non-trivial differential $d_2 : E_2^{0,2} \to E_2^{2,1}$. But by (\ref{vanishing}) this cannot occur. The argument is completed as in the previous case. The case of $\lambda_1$ is proved in the same way. Finally, we consider the case of the trivial representation. In this case, we have $E_2^{2,1}=E_2^{3,0}=0$, but $E_2^{2,0}={\mathbb Q}$. It follows that we have an exact sequence $$ 0 \to {\mathbb Q} \to H^2(\Gamma_{g,r}^n,{\mathbb Q}) \to H^0(Sp_g({\mathbb Z}),H^2(T_{g,r}^n)) \to 0. $$ The result in this case now follows using the exact sequence in the first paragraph of this proof. \end{proof} Denote the $\lambda$ isotypical part of an $Sp_g$ module $V$ by \label{iso_def} $V_\lambda$. \begin{corollary} If $g\ge 3$ and $\lambda$ is a dominant integral weight of $Sp_g$, then $$ \dim \Gr^W_l H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n)_\lambda \le \dim Gr^W_{l+|\lambda|} H^2({\mathcal M}_{g,r}^n,{\mathbb V}(\lambda)). \qed $$ \end{corollary} So, in order to bound the degrees of the relations in ${\mathfrak t}_{g,r}^n$, it suffices to give a bound the weights on $H^2({\mathcal M}_{g,r}^n,{\mathbb V}(\lambda))$. In the absolute case we have the following result of Kabanov \cite{kabanov,kabanov:purity} which is proved using intersection homology. \begin{theorem}[Kabanov] For each irreducible rational representation $V(\lambda)$ of $Sp_g$, we have $$ \Gr^W_{k + |\lambda|}H^2({\mathcal M}_g,{\mathbb V}(\lambda)) = 0 $$ when $$ \begin{cases} k \neq 2 & \text{ when $g\ge 6$;}\cr k \neq 2,3 & \text{ when $3 \le g < 6$.} \end{cases} $$ \end{theorem} Combining Kabanov's result with the previous results, we obtain: \begin{corollary} If $g\ge 3$, then $\Gr^W_{\bullet}{\mathfrak t}_g$ has a presentation with only quadratic and cubic relations, and only quadratic relations when $g\ge 6$. \qed \end{corollary} It is now an easy matter to insert the decorations: \begin{corollary} If $g\ge 3$, then $\Gr^W_{\bullet}{\mathfrak t}_{g,r}^n$ has a presentation with only quadratic and cubic relations, and only quadratic relations when $g\ge 6$. \qed \end{corollary} \section{The Lower Central Series Quotients of a Surface Group} In this section we gather some information about $\Gr{\mathfrak p}_g^1$ that will be useful when computing relations in $\Gr{\mathfrak t}_g$ and $\Gr{\mathfrak t}_g^1$. Our basic tool, once again, is continuous cohomology. A group is called {\it pseudo-nilpotent} if $\theta$ is an isomorphism. A proof of the following result is sketched by Kohno and Oda in \cite[(4.1)]{kohno-oda}. \begin{theorem}\label{curve} If $g \ge 1$, then $\pi_g^1$ is pseudo-nilpotent. \qed \end{theorem} Even though we will not be needing it, we record the following result which is stated by Kohno and Oda \cite[(4.1)]{kohno-oda}. Their proof is incorrect --- cf.\ (\ref{error}). Nonetheless, the result follows directly from (\ref{curve}) and \cite[(5.7)]{hain:cycles}. \begin{corollary}\label{kohno-oda} If $g=0$ and $r\ge 1$, or if $g \ge 1$, then, for all $n\ge 0$, each of the decorated pure braid groups $F_{g,r}^n$ is pseudo-nilpotent. \qed \end{corollary} Since $H_1({\mathfrak p}_g^1)$ is the fundamental representation of $\sp_g$, $\Gr^W_{\bullet} {\mathfrak p}_g^1$ is a graded Lie algebra in $R(\sp_g)$, and its complex of chains $\Lambda^{\bullet} \Gr^W_{\bullet}{\mathfrak p}_g^1$ is a complex in $R(\sp_g)$. We shall write ${\mathfrak p}_g$ \label{p_def} instead of ${\mathfrak p}_g^1$, and $\pi_g$ \label{pi_def} instead of $\pi_g^1$. We shall denote the $l$th weight graded quotient of a Hodge Lie algebra ${\mathfrak g}$ by \label{gr_def} ${\mathfrak g}(l)$. In particular, we shall denote $\Gr^W_{-l}{\mathfrak p}_g^1$ by \label{pgr_def} ${\mathfrak p}_g(l)$. \begin{corollary}\label{complex} If $g\ge 1$, then, for each $l\ge 3$, the complex $$ \Gr^W_{-l} \Lambda^{\bullet}\Gr^W_{\bullet}{\mathfrak p}_g $$ is an acyclic complex of $\sp_g$ modules. When $l=2$, we have an exact sequence $$ 0 \to {\mathbb Q}(1) \to \Lambda^2 {\mathfrak p}_g(1) \to {\mathfrak p}_g(2) \to 0 $$ of $\sp_g$ modules. \qed \end{corollary} This result allows us to compute the ${\mathfrak p}_g(l)$ inductively as elements of $R(\sp_g)$. As before, we fix a set $\lambda_1,\dots,\lambda_g$ of fundamental weights of $\sp_g$. \begin{proposition}\label{lcs_quots} For all $g\ge 3$, the highest weight decomposition of ${\mathfrak p}_g(l)$ when $1\le l \le 4$ is given by $$ {\mathfrak p}_g(l) = \begin{cases} V(\lambda_1) & \text{ when $l=1$}; \cr V(\lambda_2) & \text{ when $l=2$}; \cr V(\lambda_1 + \lambda_2) & \text{ when $l=3$}; \cr V(2\lambda_1) + V(2\lambda_1+\lambda_2) + V(\lambda_1 + \lambda_3) & \text{ when $l=4$}. \end{cases} $$ \end{proposition} \begin{proof} This is a straightforward consequence of (\ref{complex}). To show how this works, we prove the case where $l=3$. In this case, we have the exact sequence $$ 0 \to \Lambda^3 {\mathfrak p}_g(1) \to {\mathfrak p}_g(1)\otimes {\mathfrak p}_g(2) \to {\mathfrak p}_g(3) \to 0 $$ in $R(\sp_g)$. Taking euler characteristics and applying the result for $l=2$ and $g\ge 3$, we see that $$ {\mathfrak p}_g(3) = V(\lambda_1)\otimes V(\lambda_2) - \Lambda^3 V(\lambda_1) = V(\lambda_1 + \lambda_2). $$ \end{proof} Since the $k$th exterior power is the Schur functor corresponding to the Young diagram with $k$ rows and one box in each row, and since ${\mathfrak p}_g(1) = H_1(\pi_g)$ is the fundamental representation of $\sp_g$, we obtain the following stability result for the graded quotients of the lower central series of $\pi_g$. \begin{corollary} The highest weight decomposition of ${\mathfrak p}_g(l)$ is independent of $g$ when $l\ge g$. \qed \end{corollary} \section{The Action of ${\mathfrak t}_g^1$ on ${\mathfrak p}_g$} \label{inf_action} In this section we obtain a lower bound for the size of $\Gr^W_l{\mathfrak t}_g$ when $l=2,3$ and $g\ge 3$ by studying the action of ${\mathfrak t}_g^1$ on ${\mathfrak p}_g$. This will provide an upper bound on the size of the quadratic and cubic relations of $\Gr^W_{\bullet}{\mathfrak t}_g$. Related results have been obtained Asada-Kaneko \cite{japanese}, Morita \cite{morita:trace} and Asada-Nakamura \cite{asada-nakamura}. We also use a result \cite{asada-nakamura} of Asada and Nakamura to prove that ${\mathfrak t}_g$ is infinite dimensional. (This fact also follows from a recent result of Oda \cite{oda}.) First, some notation. Denote the pronilpotent Lie algebra $W_{-1}\Der {\mathfrak p}_g$ by \label{der_def} ${\mathfrak d}_g$, and the quotient of this by inner automorphisms by \label{out_def} ${\mathfrak o}_g$. Once a base point $x$ of ${\mathcal M}_g^1$ has been chosen, each of these acquires the structure of a Hodge Lie algebra. We have natural representations $$ {\mathfrak u}_g^1 \to {\mathfrak d}_g \text{ and } {\mathfrak u}_g \to {\mathfrak o}_g. $$ These induce homomorphisms of their associated graded Lie algebras. It is clear that there is an injective homomorphism $$ {\mathfrak d}_g(l) \hookrightarrow \Hom({\mathfrak p}_g(1),{\mathfrak p}_g(l+1)). $$ Each element $\delta : {\mathfrak p}_g(1) \to {\mathfrak p}_g(l+1)$ determines a derivation $\tilde{\delta}$ of the free Lie algebra ${\mathbb L}({\mathfrak p}_g(1))$, the second graded quotient of which is isomorphic to ${\mathfrak p}_g(2) \oplus {\mathbb Q}\omega$, where ${\mathbb Q} \omega$ is the unique copy of the trivial representation in $\Lambda^2 {\mathfrak p}_g(1)$. By taking the image of $\tilde{\delta}(\omega)$ under the projection $$ {\mathbb L}({\mathfrak p}_g(1)) \to \Gr^W_{\bullet} {\mathfrak p}_g, $$ we obtain an element $\sigma_g(\delta)$ of ${\mathfrak p}_g(l+1)$. Observe that $\delta$ induces a derivation of $\Gr{\mathfrak p}_g$ if and only if $\sigma_g(\delta)$ vanishes. We therefore have a surjection $$ {\mathfrak d}_g \to \ker\left\{\Hom({\mathfrak p}_g(1),{\mathfrak p}_g(l+1)) \stackrel{\sigma}{\to} {\mathfrak p}_g(l+2)\right\}. $$ \begin{proposition}\label{gr_der} The map $\sigma$ is surjective. Consequently, $$ {\mathfrak d}_g(l) = {\mathfrak p}_g(1)\otimes{\mathfrak p}_g(l+1) - {\mathfrak p}_g(l+2) $$ in $R(\sp_g)$. \end{proposition} \begin{proof} Consider the diagram $$ \begin{CD} \Hom({\mathfrak p}_g(1),{\mathfrak p}_g(l+1)) @>{\sigma_g}>> {\mathfrak p}_g(l+2)\cr @VVV @| \cr {\mathfrak p}_g(1)\otimes {\mathfrak p}_g(l+1) @>{[\phantom{x},\phantom{x}]}>> {\mathfrak p}_g(l+2) \end{CD} $$ where the left hand vertical map is induced by the quadratic form $\omega = \sum a_i\wedge b_i$. This diagram commutes as the left hand map satisfies $$ \delta \mapsto \sum_{i=1^g} a_i\otimes\delta(b_i) - b_i\otimes\delta(a_i), $$ which goes to $$ \sigma_g(q) = \sum_{i=1}^g\left([\delta(a_i),b_i] + [a_i,\delta(b_i)]\right) $$ under the bracket. Since the bottom map is surjective and all maps are $\sp_g$ equivariant, the result follows. \end{proof} Combining this with the computation of the first few graded quotients of ${\mathfrak p}_g$ given in (\ref{lcs_quots}), we obtain the following result. \begin{corollary}\label{der_quots} For all $g\ge 3$, we have $$ {\mathfrak d}_g(l) = \begin{cases} V(\lambda_3)+V(\lambda_1) &\text{ when }l=1;\cr V(2\lambda_2)+V(\lambda_2)&\text{ when }l=2;\cr V(2\lambda_1+\lambda_3)+V(\lambda_1+\lambda_2)+V(3\lambda_1)& \text{ when } l=3. \end{cases} $$ \qed \end{corollary} The computation of ${\mathfrak d}_g(1)$ is simply another formulation of the Johnson homomorphism. It is proven in \cite[$A^\prime$, p.~149]{japanese} that the center of $\Gr{\mathfrak p}_g$ is trivial, so that the inclusion ${\mathfrak p}_g \to {\mathfrak d}_g$ of the inner automorphisms is injective. \begin{proposition} For all $g\ge 3$ and all $l\ge 1$, ${\mathfrak o}_g(l) = {\mathfrak d}_g(l) - {\mathfrak p}_g(l)$. \qed \end{proposition} Combining (\ref{lcs_quots}) and (\ref{der_quots}), we obtain the following computation. \begin{corollary}\label{out_quots} For all $g\ge 3$, we have $$ {\mathfrak o}_g(l) = \begin{cases} V(\lambda_3) &\text{ when }l=1;\cr V(2\lambda_2)&\text{ when }l=2;\cr V(2\lambda_1+\lambda_3)+V(3\lambda_1)& \text{ when } l=3. \end{cases} $$ \qed \end{corollary} It does not seem obvious {\it a priori}, that ${\mathfrak t}_g$ is infinite dimensional.\footnote{This result also follows quite directly from a result of Oda \cite{oda}.} \begin{proposition}\label{quotients} For all $g\ge 3$, the image of ${\mathfrak t}_g$ in ${\mathfrak o}_g$ is infinite dimensional. \end{proposition} \begin{proof} Since ${\mathfrak t}_g \to {\mathfrak o}_g$ is a morphism of MHS, the image ${\mathfrak g}$ has a MHS. Since $\Gr^W_{\bullet}$ is an exact functor, $$ \Gr^W_{\bullet}{\mathfrak g} = \text{ image of }\{\Gr^W_{\bullet}{\mathfrak t}_g \to \Gr^W_{\bullet}{\mathfrak o}_g\}. $$ So it suffices to show that each graded quotient of ${\mathfrak g}$ is non-trivial. It follows from the result Asada and Nakamura \cite[Theorem~B]{asada-nakamura} that the image of $$ {\mathfrak t}_g^1(l) \to {\mathfrak d}_g(l) $$ contains the representation $V(2m\lambda_1 + \lambda_3)$ when $l = 2m+1$, and $V(2m\lambda_1 + 2\lambda_2)$ when $l = 2m+2$. These representations both have the maximal possible depth, $l+2$. But the inner automorphisms ${\mathfrak p}_g(l)$ in ${\mathfrak d}_g(l)$ have depth at most $l$. The result follows. \end{proof} We can now bound below the low degree relations in ${\mathfrak t}_g$. \begin{proposition}\label{upper} For all $g\ge 3$, the image of ${\mathfrak u}_g(l)$ in ${\mathfrak o}_g(l)$ is $$ \begin{cases} V(\lambda_3) & \text{ when } l=1;\cr V(2\lambda_2) & \text{ when } l=2;\cr V(2\lambda_1+\lambda_3) & \text{ when } l=3. \end{cases} $$ \end{proposition} \begin{proof} It follows from (\ref{quotients}) that when $g\ge 3$, the image of ${\mathfrak u}_g(l) \to {\mathfrak o}_g(l)$ is non-trivial for all $l$. Since this map is $\sp_g$ equivariant, the image of ${\mathfrak u}_g(2)$ must be all of ${\mathfrak o}_g(2)$. Since $\Gr^W_{\bullet} {\mathfrak u}_g$ is generated by ${\mathfrak u}_g(1)$, and since $V(3\lambda_1)$ does not appear in ${\mathfrak u}_g(1)\otimes {\mathfrak u}_g(2)$, the assertion for $l=3$ follows. \end{proof} Note that the copy of $V(3\lambda_1)$ is detected by Morita's trace \cite{morita:trace}. \section{Quadratic Relations} \label{quadratic_relns} In this section, we find some obvious quadratic relations in ${\mathfrak t}_g$ for each $g\ge 4$. These give a lower bound for the relations in ${\mathfrak t}_g$. Serendipitously, this coincides with the upper bound (\ref{upper}) derived in the previous section, thus yielding all the quadratic relations. \begin{theorem}\label{lower} For all $g \ge 3$, we have $$ \Gr^W_{-2} {\mathfrak t}_g = \Gr^W_{-2} {\mathfrak u}_g = V(2\lambda_2) + V(0). $$ \end{theorem} The proof occupies the rest of this section. We prove the result by finding a pair of commuting elements $\phi$ and $\psi$ of the $T_g$ whose class $$ \tau(\phi)\wedge\tau(\phi) \in \Lambda^2 V(\lambda_3) $$ generates the $Sp_g$ complement of $V(2\lambda_2) + V(0)$ for all $g \ge 3$. Since we know, by (\ref{upper}), that the quadratic relations are contained in the complement of $V(2\lambda_2) + V(0)$, we have found all quadratic relations. \begin{lemma}\label{computation} If $g\ge 3$, then \begin{multline*} \Lambda^2 V(\lambda_3) = \\ \begin{cases} V(\lambda_6) + V(\lambda_4) + V(\lambda_2) + V(\lambda_2 + \lambda_4) + V(2\lambda_2) + V(0) & \text{when $g\ge 6$;}\cr V(\lambda_4) + V(\lambda_2) + V(\lambda_2 + \lambda_4) + V(2\lambda_2) + V(0) & \text{when $g=5$;}\cr V(\lambda_2) + V(\lambda_2 + \lambda_4) + V(2\lambda_2) + V(0) & \text{when $g = 4$;}\cr V(2\lambda_2) + V(0) & \text{when $g=3$.} \end{cases} \end{multline*} \qed \end{lemma} {}From (\ref{mhs_torelli}), we know that $V(0)$ occurs in ${\mathfrak t}_g(2)$. By (\ref{upper}) and the previous proposition, there is nothing to prove when $g=3$. So we suppose that $g\ge 4$. We use the notation introduced in Section \ref{reps}. Set $$ \omega = a_1\wedge b_1 + \cdots + a_g\wedge b_g. $$ \begin{proposition}\label{construction} When $g\ge 3$, there are elements $\phi_{i,j}$, $1\le i < j \le g$ of the Torelli group whose image under the Johnson homomorphism $$ \tau_g : H_1(T_g) \to V(\lambda_3) $$ is given by $$ (g-1)\tau_g(\phi_{i,j}) = (g-1) a_i\wedge a_j \wedge b_j - a_i \wedge \omega $$ Here we are viewing $V(\lambda_3)$ as a submodule of $\Lambda^3 V(\lambda_1)$. Moreover, we can choose them such that $\phi_{1,2}$ and $\phi_{3,4}$ commute when $g\ge 4$. \end{proposition} \begin{proof} For $1 \le i < j \le g$ is easy to construct elements $\phi_{i,j}$ of the pointed Torelli group $T^1_g$ with $$ \tau_g^1(\phi_{i,j}) = a_i\wedge a_j \wedge b_j \in \Lambda^3 V(\lambda_1). $$ (To compute $\tau_g^1: H_1(T_g^1)\to \Lambda^3V(\lambda_1)$, use Johnson's original definition in terms of the action of $T_g^1$ on $\pi_g$.) \vspace*{1.5in}\\ It is also easy to arrange for $\phi_{1,2}$ and $\phi_{3,4}$ to have disjoint supports, and therefore commute. To compute $\tau_g(\phi_{i,j}) \in V(\lambda_3)$, we just use the fact that the maps $$ \underline{\phantom{x}}\wedge\omega : V(\lambda_1) \to \Lambda^3 V(\lambda_1) $$ and $$ p: \Lambda^3 V(\lambda_1) \to V(\lambda_1) $$ defined by $p(x\wedge y\wedge z) = q(x,y)z + q(y,z) x + q(z,x) y$ are $\sp_g$-equivariant and satisfy $$ p\circ (\underline{\phantom{x}}\wedge\omega) = (g-1)\id. $$ It follows that $V(\lambda_3)$ is the kernel of $p$ and that $$ (g-1) \tau_g(\phi_{i,j}) = (g-1) a_i\wedge a_j\wedge b_j - a_i \wedge \omega. $$ \end{proof} Take $\phi = \phi_{1,2}$ and $\psi = \phi_{3,4}$. Since these commute, $$ v :=\tau_g(\phi)\wedge\tau_g(\psi) \in \Lambda^2 V(\lambda_3) $$ will lie inside the $\sp_g$ module of quadratic relations. Denote the $\sp_g$ submodule of $\Lambda^2 V(\lambda_3)$ generated by $v$ by $V$. By (\ref{construction}), $$ v = [(g-1)\, a_1\wedge a_2 \wedge b_2 - a_1\wedge \omega] \wedge [(g-1)\, a_3\wedge a_4 \wedge b_4 - a_3\wedge \omega]. $$ Recall that elements of $\sp_g$ act on exterior powers as derivations. Note also that for all $X \in \sp_g$, $X\omega = 0$. We now compute the highest weight decomposition of $V$. \smallskip \noindent{$\mathbf \lambda_2 + \lambda_4$:} Apply $F_{2,3}$, then $F_{1,4}$, then $T_{2,3}$ to $v$ to get $$ (g-1)^2 [a_1\wedge a_2 \wedge a_3 ] \wedge [a_1\wedge a_2 \wedge a_4] \in \Lambda^2\Lambda^3 V(\lambda_1) $$ which is a highest weight vector on which ${\mathfrak h}$ acts via the character $$ \lambda_2 + \lambda_4 = (t_1 + t_2) + ( t_1 + t_2 + t_3 + t_4). $$ To decompose the rest of $V$, consider the $\sp_g$-equivariant map $$ \Lambda^2 V(\lambda_3) \hookrightarrow \Lambda^2\Lambda^3 V(\lambda_1) \to \Lambda^6 V(\lambda_1). $$ Denote the image of $V$ under this map by $W$. It is spanned by the image of $v$ in $\Lambda^6 V(\lambda_1)$. For the time being, we suppose that $g\ge 6$. \smallskip \noindent{$\mathbf \lambda_6$:} In this case, the image of $v$ in $W$ is $$ w := ((g-1)\, a_1\wedge a_2 \wedge b_2 - a_1\wedge \omega) \wedge ((g-1)\, a_3\wedge a_4 \wedge b_4 - a_3\wedge \omega). $$ To find a highest weight vector for the representation it generates, first apply $F_{2,5}$, then $F_{4,6}$ to this vector to get the highest weight vector $$ (g-1)^2 a_1\wedge a_2 \wedge a_3 \wedge a_4 \wedge a_5 \wedge a_6 $$ of $W$ on which ${\mathfrak h}$ acts via the character $$ \lambda_6 = t_1 + t_2 + t_3 + t_4 + t_5 + t_ 6. $$ To show that the weights $\lambda_2$ and $\lambda_4$ occur in $V$, it is useful to recall that for all $k\ge 2$, there is an $\sp_g$ equivariant projection \begin{equation}\label{contraction} \theta_k : \Lambda^k V(\lambda_1) \to \Lambda^{k-2} V(\lambda_1) \end{equation} which is defined by $$ x_1 \wedge \ldots \wedge x_k \mapsto \sum_{1 \le i < j \le k} (-1)^{i+j+1} q(x_i,x_j)\, x_1 \wedge \ldots \wedge \hat{x_i}\wedge \ldots \wedge \hat{x_j} \wedge \ldots \wedge x_k. $$ \noindent{$\mathbf \lambda_4$: } The image of $V$ in $\Lambda^4 V(\lambda_1)$ is generated by $\theta_6(w)$ which is $$ (g-1)^2 a_1 \wedge a_3 \wedge (a_2 \wedge b_2 + a_4 \wedge b_4) - (g-1)(g-3)\, a_1 \wedge a_3 \wedge(a_2 \wedge b_2 + a_4 \wedge b_4) $$ $$ - 2(g-1)\, a_1 \wedge a_3 \wedge \omega - 2(g-2)\, a_1 \wedge a_3 \wedge \omega $$ $$ = 2(g-1)\, a_1 \wedge a_2 \wedge (a_3 \wedge b_3 + a_4 \wedge b_4) -2\, a_1 \wedge a_2 \wedge \omega. $$ Applying $F_{3,6}$, then $S_{4,6}$, one gets the highest weight vector $$ 2(g-1)\, a_1\wedge a_2\wedge a_3\wedge a_4 $$ on which ${\mathfrak h}$ acts via the character $\lambda_4 = t_1 + t_2 + t_3 + t_4$. \smallskip \noindent{ $\mathbf \lambda_2$:} The image of $V$ in $\Lambda^2 V(\lambda_1)$ is generated by the image under $\theta_4$ of $\theta_6(w)$. This is $$ 4(g-1)\, a_1 \wedge a_3 - 2(g-2)\, a_1 \wedge a_3 = 2g\, a_1 \wedge a_3. $$ Apply $S_{2,3}$ to this to get $2g\, a_1 \wedge a_2$ upon which ${\mathfrak h}$ acts with highest weight $\lambda_2 = t_1 + t_2$. We sketch the remaining cases $g=4,5$. When $g=5$, $W$ is generated by the vector $$ (g-1)\left(a_1\wedge a_2\wedge a_3\wedge b_2 - a_1\wedge a_3\wedge a_4\wedge b_4\right)\wedge \omega. $$ By contracting with $q$ as above, it is easy to see that this vector generates a submodule of $$ \omega \wedge \Lambda^4 V(\lambda_1) \cong \Lambda^6 V(\lambda_1) $$ isomorphic to $V(\lambda_4) + V(\lambda_2)$. When $g=4$, $W$ is generated by $a_1\wedge a_3 \wedge \omega^2$. Again, by contracting with $q$, it is easy to see that this vector generates a submodule of $$ w^2\wedge \Lambda^2 V(\lambda_1) \subset \Lambda^6 V(\lambda_1) $$ isomorphic to $V(\lambda_2)$. \begin{remark} Note that we have determined the quadratic relations for all $g\ge 3$. One should be able to determine the cubic relations when $3 \le g \le 5$ by applying similar methods and the fact that the Dehn twist about the separating curve $C$ below commutes with the bounding pair map associated to the curves $C'$ and $C''$. The Dehn twist about $C$ is in the kernel of the Johnson homomorphism, but has non-trivial image in the second graded quotient of ${\mathfrak t}_g$. \vspace*{2in} Note that there have to be cubic relations in genus 3 as there are no quadratic relations, and there has to be one copy of $V(\lambda_3)$ in the cubic relations to ensure the existence of the central $\Ga$. \end{remark} \section{Presentations of ${\mathfrak t}_g$, $t_{g,1}$ and ${\mathfrak t}_g^1$} \label{special} Recall that ${\mathbb L}(V)$ denotes the free Lie algebra generated by the vector space $V$. In this section we shall give presentations of ${\mathfrak t}_g$, ${\mathfrak t}_g^1$ and ${\mathfrak t}_{g,1}$ when $g\ge 6$. First ${\mathfrak t}_g$ --- combining (\ref{upper}), (\ref{lower}) and (\ref{construction}), we have: \begin{theorem} For all $g\ge 6$, $\Gr^W_{\bullet} {\mathfrak t}_g$ is isomorphic to $$ {\mathbb L}(V(\lambda_3))/R_g $$ as a graded Lie algebra in $R(\sp_g)$, where $R_g$ is the ideal generated by the quadratic relations $$ V(\lambda_6) + V(\lambda_4) + V(\lambda_2) + V(\lambda_2 +\lambda_4) \subseteq \Lambda^2 V(\lambda_3). \qed $$ \end{theorem} Since ${\mathfrak t}_g\otimes{\mathbb C} \cong \prod_l \Gr^W_l {\mathfrak t}_g\otimes{\mathbb C}$, this gives the desired presentation of ${\mathfrak t}_g$ for $g\ge 6$. We next consider ${\mathfrak t}_g^1$. Fix a point $[C;x]$ of ${\mathcal M}_g^1$ so that ${\mathfrak t}_g^1$ and ${\mathfrak p}_g$ have canonical MHSs. Then the sequence $$ 0 \to {\mathfrak p}_g^1 \to {\mathfrak t}_g^1 \to {\mathfrak t}_g \to 0 $$ is an exact sequence of MHSs. Since $\Gr^W_{\bullet}$ is an exact functor, the sequence $$ 0 \to \Gr^W_{\bullet} {\mathfrak p}_g^1 \to \Gr^W_{\bullet} {\mathfrak t}_g^1 \to \Gr^W_{\bullet} {\mathfrak t}_g \to 0 $$ is exact in $R(\sp_g)$. Since the sequence of $H_1$'s is canonically split in $R(\sp_g)$, there is a canonical lift $$ {\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet}{\mathfrak t}_g^1 $$ of the natural homomorphism ${\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet}{\mathfrak t}_g$. Since ${\mathfrak p}_g(2)$ is isomorphic to $V(\lambda_2)$, and since this representation does not occur in ${\mathfrak t}_g(2)$, we can take the $\lambda_2$ component of the bracket $$ \Lambda^2 H_1({\mathfrak t}_g) \hookrightarrow \Lambda^2 H_1(t_g^1) \to {\mathfrak t}_g^1(2) $$ to obtain an $\sp_g$ module map \begin{equation}\label{bracket} c : \Lambda^2 H_1({\mathfrak t}_g) \to {\mathfrak p}_g(2) \cong V(\lambda_2). \end{equation} This and the map \begin{equation}\label{action2} H_1({\mathfrak t}_g) \otimes H_1({\mathfrak p}_g) \to {\mathfrak p}_g(2) \end{equation} induced by the bracket completely determine $\Gr^W_{\bullet} {\mathfrak t}_g^1$ given $\Gr^W_{\bullet} {\mathfrak t}_g$ and $\Gr^W_{\bullet} {\mathfrak p}_g$. The map (\ref{action2}) is simply the adjoint of the Johnson homomorphism $$ \tau_g^1 : H_1({\mathfrak t}_g) \to \Lambda^3 V \subset \Hom(H_1({\mathfrak p}_g),{\mathfrak p}_g(2)). $$ So, to give a presentation of ${\mathfrak t}_g^1$, we have to determine the map (\ref{bracket}). We do this by studying the action of ${\mathbb L}(V(\lambda_3))$ on ${\mathbb L}(V(\lambda_1))$. Set $V=V(\lambda_1)$. We identify $V$ with $H_1(C)$ and $H_1(T_g^1)$ with $\Lambda^3 V$ via the Johnson homomorphism. Recall that $V(\lambda_3)$ can be realized as the kernel of the map $p:\Lambda^3 V \to V$ defined by \begin{equation}\label{projn} p : v_1\wedge v_2 \wedge v_3 \mapsto (v_1\cdot v_2)v_3 + (v_2\cdot v_3) v_1 + (v_3\cdot v_1) v_2. \end{equation} We identify $H_1(T_g) \cong \Lambda^3 V/V$ with $V(\lambda_3)$ via the map $$ V(\lambda_3) = \ker p \hookrightarrow \Lambda^3 V \to \Lambda^3 V/V. $$ The natural action of $\Lambda^3 V$ on ${\mathbb L}(V)$ is defined by \begin{multline*} e_1\wedge e_2 \wedge e_3 \mapsto -\left\{v \mapsto (e_1\cdot v) [e_2,e_3] + (e_2\cdot v)[e_3,e_1] + (e_3\cdot v)[e_1,e_2]\right\} \cr \in \Hom(V,\Lambda^2 V) \subseteq \Der {\mathbb L}(V). \end{multline*} (With this choice of sign, $\sum x\wedge a_j\wedge b_j\mapsto\ad(x)$.) It follows from the definition of the Johnson homomorphism that the composite $$ H_1(T_{g,1}) \stackrel{\tau_{g,1}}{\longrightarrow} \Lambda^3 V \hookrightarrow \Hom(V,\Lambda^2 V) $$ is the map induced by the action of $T_{g,1}$ on $\pi_{g,1}$. The action descends to the action of $\Gr^W_{\bullet}{\mathfrak t}_g^1$ on $\Gr^W_{\bullet} {\mathfrak p}_g$. Define a projection $r : \Lambda^2 V(\lambda_3) \to V(\lambda_2)$ to be the composite $$ \Lambda^2 V(\lambda_3) \hookrightarrow \Lambda^2 \Lambda^3 V \stackrel{\text{mult}}{\longrightarrow} \Lambda^6 V \stackrel{\theta_4\theta_6}{\longrightarrow} \Lambda^2 V \to V(\lambda_2) $$ where $\theta_k$ is the contraction (\ref{contraction}) defined in the previous section, and the last map is the standard projection $$ u\wedge v \mapsto u\wedge v - (u\cdot v)\,\omega/(g-1). $$ Since there is only one copy of $V(\lambda_2)$ in $\Lambda^2 V(\lambda_3)$, this projection is unique up to a scalar. \begin{proposition}\label{bra_const} The map (\ref{bracket}) is given by $$ c[u,v] = -\frac{1}{2g+2} \ad(r(u\wedge v)) \in \Hom(H_1(p_g),{\mathfrak p}_g(3)). $$ In particular, this map is non-zero, and the extensions $$ 0 \to {\mathfrak p}_g \to {\mathfrak t}_g^1 \to {\mathfrak t}_g \to 0 \text{ and } 0 \to {\mathfrak p}_g \to {\mathfrak u}_g^1 \to {\mathfrak u}_g \to 0 $$ are not split. \end{proposition} We now sketch the proof. Denote the degree $k$ part of the free Lie algebra ${\mathbb L}(V)$ by ${\mathbb L}(V)(k)$. Recall that there is a standard exact sequence $$ 0 \to \Lambda^3 V \stackrel{j}{\to} V\otimes \Lambda^2 V \stackrel{b}{\to} {\mathbb L}(V)(3) \to 0. $$ The first map is the ``Jacobi identity'' map $$ j : v_1\wedge v_2 \wedge v_3 \mapsto v_1\otimes v_2\wedge v_3 + v_2 \otimes v_3\wedge v_1 + v_3 \otimes v_1 \wedge v_2, $$ and the second map is the bracket. (We are identifying ${\mathbb L}(V)(2)$ with $\Lambda^2 V$ in the standard way.) \begin{lemma}\label{compn} The bracket $[e_1\wedge e_2 \wedge e_3, f_1\wedge f_2 \wedge f_3]$ of two elements of $\Lambda^3 V$ as derivations of ${\mathbb L}(V)$ is obtained by summing the expression \begin{multline*} (e_1\cdot f_1) \big( e_2\otimes [e_3,[f_2,f_3]] - e_3\otimes [e_2,[f_2,f_3]] + f_2\otimes [f_3,[e_2,e_3]] - f_3\otimes [f_2,[e_2,e_3]] \big) \cr \quad \in V\otimes {\mathbb L}(V)(3) \cong \Hom(V,{\mathbb L}(V)(3)) \end{multline*} cyclically in $(e_1,e_2,e_3)$ and in $(f_1,f_2,f_3)$. \qed \end{lemma} We shall view this expression as an element of $\left(V\otimes V \otimes \Lambda^2 V\right) /\left(V \otimes \Lambda^3 V\right)$. The next step is to write down the projections of this group onto $V(\lambda_2)$. There are four copies of $V(\lambda_2)$ in $V\otimes V \otimes \Lambda^2 V$. These are detected by the following four projections onto $\Lambda^2 V$: \begin{align*} p_1 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto (u_1\cdot u_2)u_3\wedge u_4 \cr p_2 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto (u_3\cdot u_4) u_1\wedge u_2 \cr p_3 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto \big((u_1\cdot u_4)u_2\wedge u_3 - (u_1\cdot u_3)u_2\wedge u_4\big)/2\cr p_4 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto \big((u_2\cdot u_3)u_1\wedge u_4 - (u_2\cdot u_4)u_1\wedge u_3\big)/2. \end{align*} One can easily check that there are two copies of $V(\lambda_2)$ in $V\otimes \Lambda^3 V$ and that the projections $p_1 - p_3$ and $p_2 - p_4$ vanish on these. This leaves two copies of $V(\lambda_2)$ in $\left(V\otimes V \otimes \Lambda^2 V\right) /\left(V \otimes \Lambda^3 V\right)$. One of these vanishes in $\Hom(V,{\mathfrak p}_g(3))$ as $V\otimes V \otimes \omega$ projects to zero there. We are now ready to compute. Since $$ u_j = a_j\wedge a_3 \wedge b_3 - a_j\wedge a_4 \wedge b_4 $$ lies in the kernel of the projection $p$ above when $j=1,2$, $u_1\wedge u_2$ is an element of $\Lambda^2 V(\lambda_3)$. The projection $r$ takes $u_1\wedge u_2$ to $-4\, a_1\wedge a_2$. On the other hand, by straightforward computations using (\ref{compn}), we have $$ p_1([u_1,u_2]) = p_2([u_1,u_2]) = 0,\text{ and } p_3([u_1,u_2]) = p_4([u_1,u_2]) = -4\, a_1\wedge a_2. $$ Consequently, $$ (p_1 - p_3)([u_1,u_2]) = (p_2 - p_4)([u_1,u_2]) = 4\, a_1\wedge a_2. $$ Next observe that $\ad [a_1,a_2]$ corresponds to the element $$ - \sum_{j=1}^g \left( a_j\otimes b_j \otimes a_1\wedge a_2 - b_j \otimes a_j \otimes a_1 \wedge a_2\right). $$ of $\left(V\otimes V \otimes \Lambda^2 V\right) /\left(V \otimes \Lambda^3 V\right)$. Since $\sum\, [a_j,b_j] = 0$, $\ad [a_1,a_2]$ is also represented by $$ z := - \sum_{j=1}^g \left( a_j\otimes b_j \otimes a_1\wedge a_2 - b_j \otimes a_j \otimes a_1 \wedge a_2\right) - 2 \sum_{j=1}^g a_1\otimes a_2\otimes a_j\wedge b_j. $$ By direct computation, we have $$ (p_1-p_3)(z) = (p_2-p_4)(z) = -(2g+2)\, a_1\wedge a_2. $$\, This concludes the proof of Proposition \ref{bra_const}. \qed Next we consider the case of ${\mathfrak t}_{g,1}$. Fix a point $(C;x,v)$ of ${\mathcal M}_{g,1}$ so that ${\mathfrak t}_{g,1}$, ${\mathfrak p}_{g,1}$, etc.\ all have compatible MHSs. By strictness, the sequence $$ 0 \to \Gr^W_{\bullet} {\mathfrak p}_{g,1} \to \Gr^W_{\bullet} {\mathfrak t}_{g,1} \to \Gr^W_{\bullet} {\mathfrak t}_g \to 0 $$ is exact in $R(\sp_g)$. Since the sequence $$ 0 \to {\mathbb Q}(1) \to {\mathfrak p}_{g,1} \to {\mathfrak p}_g \to 0 $$ is exact, it follows that ${\mathfrak p}_{g,1}(2)$ is isomorphic to $\Lambda^2 V$ via the bracket. As in the case of ${\mathfrak t}_g^1$, there is a canonical lifting ${\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet}{\mathfrak t}_{g,1}$ of the natural surjection ${\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet} {\mathfrak t}_g$. It follows that to give a presentation of $\Gr^W_{\bullet} {\mathfrak t}_{g,1}$ given presentations of ${\mathfrak t}_g$ and ${\mathfrak p}_{g,1}$, it suffices to give the map \begin{equation}\label{first} H_1({\mathfrak t}_g)\otimes H_1({\mathfrak p}_{g,1}) \longrightarrow {\mathfrak p}_{g,1}(2) \end{equation} induced by the bracket, together with the $\lambda_2$ component \begin{equation}\label{second} \Lambda^2 H_1({\mathfrak t}_g) \longrightarrow V(\lambda_2) \subset {\mathfrak p}_{g,2} \end{equation} and the invariant part \begin{equation}\label{third} c_0 : \Lambda^2 H_1({\mathfrak t}_g) \longrightarrow {\mathfrak t}_{g,1}(2)^{Sp_g} \cong {\mathbb Q}(1)^2 \end{equation} of the bracket. As in the case of ${\mathfrak t}_g^1$, the first map (\ref{first}) is the adjoint of the Johnson homomorphism and the second (\ref{second}), by naturality with respect to the projection ${\mathfrak t}_{g,1} \to {\mathfrak t}_g^1$, is the map $c$ determined in (\ref{bra_const}). It remains to determine the map (\ref{third}). Observe that the sequence $$ 0 \to {\mathfrak p}_{g,1}(2)^{Sp_g} \to {\mathfrak t}_{g,1}(2)^{Sp_g} \to {\mathfrak t}_g(2)^{Sp_g} \to 0 $$ splits canonically as the canonical central $\Ga$ in ${\mathfrak t}_{g,1}$ projects to the canonical central $\Ga$ in ${\mathfrak t}_g$ by (\ref{central_ext}), and because $\Ga = {\mathfrak t}_g(2)^{Sp_g}$. As a generator of ${\mathfrak p}_{g,1}(2)^{Sp_g}$ we take $\sum\, [a_j,b_j]$. Fix an invariant bilinear form $\bil \phantom{x} \phantom{x}$ on $V(\lambda_3)$ by insisting that $$ \bil {a_1\wedge a_2 \wedge a_3} {b_1 \wedge b_2 \wedge b_3} = 1. $$ We can therefore choose a generator $\gamma$ of $\Ga$ such that if $u,v\in H_1({\mathfrak t}_g)$, then the invariant component of $[u,v]$ in ${\mathfrak t}_g(2)$ is $\bil u v \, \gamma$. \begin{proposition}\label{triv_cpt} If $u,v \in H_1({\mathfrak t}_g)$, then $$ c_0[u,v] = \bil u v \,\gamma - \frac{6\bil u v}{g(2g+1)}\sum_{j=1}^g\, [a_j,b_j]. $$ \end{proposition} As in the previous case, we determine the coefficient by studying the action of ${\mathbb L}(V(\lambda_3))$ on ${\mathbb L}(V)$. Note that $\Gamma_{g,1}$ acts on the free group $\pi_1(C - \{x\},v)$.% \footnote{This notation denotes Deligne's fundamental group of $C - \{x\}$ with base point the tangent vector $v \in T_x C$.} We therefore have a representation ${\mathfrak t}_{g,1} \to {\mathfrak p}(C-\{x\},v) \cong {\mathbb L}(V)$.\footnote{If we put the limit MHS on $\pi_1(C - \{x\},v)$ associated with the tangent vector $v$, then this action is a morphism of MHS.} We continue with the notation in the proof of (\ref{bra_const}). There are two copies of the trivial representation in $V\otimes V \otimes \Lambda^2 V$. The corresponding projections to ${\mathbb Q}$ are: \begin{align*} q_1 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto (u_1\cdot u_2)(u_3\cdot u_4) \cr q_2 : u_1\otimes u_2 \otimes u_3 \wedge u_4 &\mapsto \bigl((u_1\cdot u_4)(u_2\cdot u_3) - (u_1\cdot u_3) (u_2\cdot u_4)\bigr)/2. \end{align*} There is one copy of the trivial representation in $V\otimes \Lambda^3 V$ and $q_1 - q_2$ vanishes on it. The vectors $$ u_1 = a_1\wedge a_2 \wedge a_3 \text{ and } u_2 = b_1\wedge b_2 \wedge b_3 $$ both lie in $V(\lambda_3)$ and $\bil {u_1} {u_2} = 1$. It follows from the formula (\ref{bracket}) that $[u_1,u_2]$ is obtained by summing the expression $$ a_2 \otimes [a_3,[b_2,b_3]] - a_3\otimes [a_2,[b_2,b_3]] + b_2 \otimes [b_3,[a_2,a_3]] - b_3 \otimes [b_2,[a_2,a_3]] $$ over the cyclic group generated by $(1,2,3)$. We have $(q_1 - q_2) ([u_1,u_2]) = 6$. On the other hand, $\ad \sum[a_j,b_j]$ is represented by $$ \sum_{j=1}^g \sum_{k=1}^g (b_j\otimes a_j - a_j \otimes b_j)\otimes a_k\wedge b_k $$ The projection $q_1 - q_2$ takes the value $-g(2g +1)$ on this. The result follows. \begin{remark} The formulas in (\ref{bra_const}) and (\ref{triv_cpt}) are closely related to those in Theorem~3.1 of Morita's paper \cite{morita:cocycles}. \end{remark} \section{A Presentation of ${\mathfrak p}_{g,r}^n$} \label{braids2} In this section we give an explicit quadratic presentation of the pure braid Lie algebras ${\mathfrak p}_{g,r}^n$ for all $g > 0$. We continue with the notation of Section~\ref{braids1}. We fix a complex structure on and a base point of $F_{g,r}^n$ by choosing a point $$ [C;x_1,\dots,x_n;v_1,\dots,v_r] $$ of ${\mathcal M}_{g,r}^n$. In Section~\ref{braids1} we showed that $H^1(F_{g,r}^n(C))$ is pure of weight 1. We will show that $H^2(F_{g,r}^n(C))$ is pure of weight 2, from which the existence of a quadratic presentation of ${\mathfrak p}_{g,r}^n$ will follow via Morgan's Theorem (\ref{morgan}). First, some notation. Denote the projection of $C^n$ onto its $i$th factor by $p_i$. Denote the image of the inclusion $$ p_i^\ast : H^{\bullet}(C) \hookrightarrow H^{\bullet}(C^n) $$ by $H^{\bullet}(C_i)$. For $x\in H^{\bullet}(C)$, denote $p_i^\ast x$ by $x\sup{i}$. Denote the component of $\Delta$ where the $i$th and $j$th coordinates are equal by $\Delta_{ij}$. Fix a symplectic basis $a_1, \dots, a_g,b_1,\dots, b_g$ of $H_1(C)$, and let $\alpha_1,\dots, \alpha_g,\beta_1,\dots,\beta_g$ be the dual basis of $H^1(C)$. Denote the positive integral generator of $H^2(C)$ by $\zeta$, and the intersection form $$ \sum_{r=1}^g \alpha_r \wedge \beta_r $$ by $q$. When $i\neq j$, set $$ q_{ij} = \sum_{r=1}^g \alpha_r\sup{i} \wedge \beta_r\sup{j} + \alpha_r\sup{j} \wedge \beta_r\sup{i}. $$ \begin{lemma} The Poincar\'e dual $PD(\Delta_{ij})$ of $\Delta_{ij}$ is $\zeta\sup{i} + \zeta\sup{j} - q_{ij}$. \qed \end{lemma} This is elementary. Another elementary fact we shall need is the following statement. It is easily proved using a Mayer-Vietoris argument. \begin{lemma}\label{isom} The natural map $$ \bigoplus_{i<j} H_{2n-3}(\Delta_{ij}) \to H_{2n-3}(\Delta) $$ is an isomorphism. \qed \end{lemma} We can therefore write the Gysin map $\gamma : H_{2n-3}(\Delta) \to H^3(C^n)$ as the sum of the Gysin maps $\gamma_{ij} : H^1(\Delta_{ij}) \to H^3(C^n)$; the map $\gamma_{ij}$ being given by cup product with $PD(\Delta_{ij})$. \begin{lemma}\label{formula} The composite $H^1(C) \stackrel{p_k^\ast}{\longrightarrow} H^1(\Delta_{ij}) \stackrel{\gamma_{ij}}{\longrightarrow} H^3(C^n)$ is given by $$ x \mapsto \begin{cases} \zeta\sup{i}\wedge x\sup{j} + \zeta\sup{j}\wedge x\sup{i} & \text{ if $k \in \{i,j\}$;}\cr \zeta\sup{i}\wedge x\sup{k} + \zeta\sup{j}\wedge x\sup{k} - q_{ij}\wedge x\sup{k} & \text{ if $k\not\in \{i,j\}$. \qed} \end{cases} $$ \end{lemma} It follows from (\ref{h1_braid}) that the part $$ 0 \to \bigoplus_{i<j} {\mathbb Z} \to H^2(C^n) \to H^2(C^n - \Delta) \to H_{2n-3}(\Delta) \to H^3(C^n) $$ of the Gysin sequence is exact. Purity of $H^2(F_g^n(C))$ therefore follows from the following proposition. \begin{proposition} The Gysin map $\gamma : H_{2n-3}(\Delta) \to H^3(C^n)$ is injective. \end{proposition} \begin{proof} The Gysin sequence can be viewed as the fiber over $[C] \in {\mathcal M}_g$ of an exact sequence of (orbifold) local systems over ${\mathcal M}_g$. It follows from (\ref{isom}) that the the monodromy actions of the last two terms of the Gysin sequence above factor through the symplectic group. It is convenient, though not necessary, to decompose these groups under the action of $Sp_g$. First note that $H_{2n-3}(\Delta_{ij})$ is isomorphic to $H^1(\Delta_{ij})$, which is isomorphic to $n-1$ copies of the fundamental representation $V$. Next, $$ H^3(C^n) = \bigoplus_{i\neq j} \left(H^2(C_i)\otimes H^1(C_j)\right) \oplus \bigoplus_{i<j<k} H^1(C_i)\otimes H^1(C_j)\otimes H^1(C_k). $$ Each of the terms $H^2(C_i)\otimes H^1(C_j)$ is a copy of the fundamental representation that we shall denote by $V^i_j$. The term $H^1(C_i)\otimes H^1(C_j)\otimes H^1(C_k)$ is isomorphic to $V^{\otimes 3}$. It contains 3 copies of $V$. If, for $i,j,k$ distinct, we set $$ V_{ij}^k = \text{ the image of } \left\{H^1(C_k) \stackrel{\wedge q_{ij}}{\to} H^3(C^n)\right\}, $$ then $$ \left[H^1(C_i)\otimes H^1(C_j)\otimes H^1(C_k)\right]_{\lambda_1} = V_{ij}^k \oplus V_{jk}^i \oplus V_{ki}^j. $$ It is now easy to see that $\gamma$ is injective. Indeed, by (\ref{formula}), we see that the images of the maps $$ H^1(C) \stackrel{p_i^\ast}{\to} H^1(\Delta_{ij}) \stackrel{\gamma_{ij}}{\to} H^3(C^n) $$ are independent copies of $V$, and also, when $k \not\in \{i,j\}$, that the image of $$ H^1(C) \stackrel{p_k^\ast}{\to} H^1(\Delta_{ij}) \stackrel{\gamma_{ij}}{\to} H^3(C^n) $$ is congruent to $V_{ij}^k$ modulo the sum of the $V_b^a$. \end{proof} Similarly, one can show that the $r$ Chern classes of the central extensions $$ 0 \to {\mathbb Z}^r \to \pi_{g,r}^n \to \pi_g^{r+n} \to 1 $$ are linearly independent in $H^2(\pi_g^{r+n})$ as they correspond to independent copies of the trivial representation in $H^2(F_g^{r+n})$. It follows that $H^2(F_{g,r}^n)$ is also pure of weight 2. Assembling all this, we obtain: \begin{proposition} For each choice of a base point $[C]$ of ${\mathcal M}_g$ and for all $g\ge 1$ and $n,r \ge 0$, the natural MHS on $H^1(F_{g,r}^n)$ is pure of weight 1 and that on $H^2(F_{g,r}^n)$ is pure of weight 2. In addition, the cup product $$ \Lambda^2 H^1(F_{g,r}^n,{\mathbb Q}) \to H^2(F_{g,r}^n) $$ is surjective. \qed \end{proposition} {}From Morgan's Theorem we deduce that ${\mathfrak p}_{g,r}^n$ has a quadratic presentation for all non-negative $g$, $r$ and $n$.\footnote{In the genus zero case, it is well known that $H^1$ has weight 2 and $H^2$ weight 4 as the corresponding classifying spaces are complements of hyperplanes in affine space.} Our final task is to determine the relations explicitly. First some notation. The Lie algebra ${\mathfrak p}_{g,r}^n$ is a quotient of the free Lie algebra generated by $$ H_1({\mathfrak p}_{g,r}^n) \cong H_1(C^{n+r}) \cong \bigoplus_{i=1}^{n+r} H_1(C_i). $$ We shall think of elements of $H_1(C^{n+r})$ as indeterminates, and write them as upper case letters. If $X\in H_1(C)$, we shall denote the corresponding element of $H_1(C_i)$ by $X\sup{i}$. Fix a symplectic basis $A_1,\dots,A_g,B_1,\dots, B_g$ of $H_1(C)$. Denote the intersection number of $X$ and $Y \in H_1(C)$ by $(X\cdot Y)$. \begin{theorem} For all $g\ge 1$ and all $r,n\ge 0$, $$ \Gr^W_{\bullet} {\mathfrak p}_{g,r}^n \cong {\mathbb L}(H_1(C)^{\oplus(n+r)})/R $$ where $R$ is the ideal generated by the relations \begin{xalignat*}{2} [X\sup{i},Y\sup{j}] & = [X\sup{j},Y\sup{i}] & \text{all $i$ and $j$;} \cr [X\sup{i},Y\sup{j}] & = \frac{(X\cdot Y)}{g} \sum_{k=1}^g\, [{\mathcal A}{i},\B{j}] & \text{all $i$ and $j$;} \cr \sum_{k=1}^g\, [{\mathcal A}{i},\B{i}] & = % \frac{1}{g} \sum_{j\neq i} \sum_{k=1}^g\, [{\mathcal A}{i},\B{j}] & 1 \le i \le n. \end{xalignat*} where $X$ and $Y$ are arbitrary elements of $H_1(C)$. \end{theorem} Note that the last relation holds only for those factors corresponding to a marked point, and not those corresponding to a marked tangent vector. \begin{proof} If ${\mathfrak g}$ is a graded Lie algebra generated in weight $-1$ and $H_2({\mathfrak g})$ of weight 2, then we have an exact sequence $$ 0 \to H_2({\mathfrak g}) \stackrel{\text{cup}^\ast}{\longrightarrow} \Lambda^2 H_1({\mathfrak g}) \stackrel{\text{bracket}}{\longrightarrow} Gr^W_{-2}{\mathfrak g} \to 0, $$ where the first map is the dual of the cup product.\footnote{There are many ways to see this --- the easiest being from the standard complex of Lie algebra cochains. However, the statement holds in greater generality --- cf.\ \cite{sullivan:les}.} In our case, the natural injection $$ H^2({\mathfrak p}_{g,r}^n) \to H^2(F_{g,r}^n,{\mathbb Q}) $$ is an isomorphism because the cup product is surjective. The coproduct is the obvious inclusion of $H_2(F_{g,r}^n,{\mathbb Q})$ into $\Lambda^2 H_1(C^n,{\mathbb Q})$, and the sequence is a sequence of $Sp_g$ modules: $$ 0 \to H_2(F_{g,r}^n,{\mathbb Q}) \to \Lambda^2 H_1(C^n,{\mathbb Q}) \to \Gr^W_{-2} {\mathfrak p}_{g,r}^n \to 0. $$ We will consider one weight at a time. Note that the three weights occurring in $\Lambda^2 H_1(C^n)$ are 0, $\lambda_2$ and $2\lambda_1$ --- the last being the symmetric square of $H_1(C)$ and second being the primitive part of $H_2(\Jac C)$. We also have the exact sequence $$ 0 \to H_2(F_{g,r}^n) \to H_2(C^n) \stackrel{\gamma^\ast}{\to} \bigoplus_{i<j}{\mathbb Q} \to 0 $$ of $Sp_g$ modules. The last map is the dual of the Gysin map. It follows that $$ H_2(F_{g,r}^n,{\mathbb Q})_{\lambda} = H_2(C^n,{\mathbb Q})_{\lambda} $$ when $\lambda$ is $2\lambda_1$ or $\lambda_2$. The $2\lambda_1$ isotypical component is spanned by elements of the form $$ X\sup{i}\times Y\sup{j} + Y\sup{i}\times X\sup{j}. $$ This gives the first relation: \begin{equation}\label{comm} [X\sup{i},Y\sup{j}] = [X\sup{j},Y\sup{i}]. \end{equation} Since $V(\lambda_2)$ is the kernel of the symplectic form $\Lambda^2 V(\lambda_1) \to {\mathbb Q}$, the $\lambda_2$ isotypical component of $H_2(C^n)$ is spanned by vectors of the form $$ X\sup{i}\times Y\sup{j} - Y\sup{i}\times X\sup{j} - \frac{(X\cdot Y)}{g} \sum_{k=1}^g \left( {\mathcal A}{i}\times \B{j} - \B{i}\times {\mathcal A}{j} \right) . $$ This gives relations of the form $$ [X\sup{i}, Y\sup{j}] + [X\sup{j}, Y\sup{i}] = \frac{(X\cdot Y)}{g} \sum_{k=1}^g \left([{\mathcal A}{i}, \B{j}] + [{\mathcal A}{j}, \B{i}] \right) $$ which simplifies to the second relation after applying (\ref{comm}). For the time being, we assume that $r=0$. The trivial isotypical component lies in an exact sequence $$ 0 \to H_2(F_{g,r}^n)^{Sp_g} \to H_2(C^n)^{Sp_g} \stackrel{\gamma^\ast}{\to} H^{2g-2}(\Delta) \to 0. $$ The map $\gamma^\ast$ takes $W\in H_2(C^n)$ to the functional $$ \{\Delta_{ij} \mapsto W\cdot\Delta_{ij}\}. $$ Note that $$ H_2(C^n)^{Sp_g} = \bigoplus_{i=1}^n H_2(C_i) \oplus \bigoplus_{i<j} \left[H_1(C_i)\otimes H_1(C_j)\right]^{Sp_g}. $$ The first terms has basis the $Z\sup{i}$, where $Z$ denotes the integral generator of $H_2(C)$. The second term has basis consisting of the $$ Q_{ij} := \sum_{k=1}^n \left({\mathcal A}{i}\times \B{j} - \B{i}\times {\mathcal A}{j}\right). $$ We next determine a basis of $\ker \gamma^\ast$. Choose $n$ distinct points $u_1,\dots,u_n$ of $C$. For $i<j$, let $C_{ij}$ be the image of the map $C\hookrightarrow C^n$ defined by $$ x \mapsto (u_1,\dots, u_{i-1},x,u_{i+1},\dots,u_{j-1},x,u_{j+1},\dots,u_n). $$ Denote its homology class by $Z_{ij}$. It is easily seen that $$ Z_{ij} = Z_i + Z_j + Q_{ij}. $$ Observe that $$ Z_{ij}\cdot \Delta_{kl} = \begin{cases} 0 & \text{$i$, $j$, $k$, $l$ distinct};\cr 1 & \#\{i,j,k,l\} = 3;\cr 2 - 2g & ij = kl. \end{cases} $$ The first two assertions are clear, the second follows from the projection formula applied to the projection of $C^n$ onto $C^2$ along the $i$th and $j$th factors and the fact that the self intersection of the diagonal in $C^2$ is the Euler number of $C$. It follows immediately that a basis of $$ H_2(F_g^n)^{Sp_g} = \ker\{H_2(C^n)^{Sp_g} \to H^2(\Delta)\} $$ consists of the $$ Z_i - \frac{1}{2g}\sum_{j\neq i} Q_{ij}. $$ These give the relations $$ \sum_{k=1}^g\, [{\mathcal A}{i},\B{i}] = \frac{1}{2g} \sum_{j\neq i} \sum_{k=1}^g\, \left([{\mathcal A}{i},\B{j}] + [{\mathcal A}{j},\B{i}]\right) $$ which becomes the third relation after an application of (\ref{comm}). Finally, the third relation is dual to the first Chern class of the pullback of the tangent bundle of $C$ along $p_i : C^n \to C$. It follows that in the general case, we do not get any relations coming from the trivial representation associated to an index corresponding to a tangent vector. \end{proof} We conclude this section with a computation of the generating function of the lower central series of $\pi_{g,r}^n$. This corrects the formula in \cite{kohno-oda}. (The galois analogue of this corrected formula is also stated in \cite[(2.14)]{nak-tak-ueno}.) \begin{theorem} For all $g\ge 1$ and all $n\ge 1$, we have $$ \prod_{k=1}^n \left(1 - 2g\,t -(k-2)t^2\right) = \prod_{l=1}^\infty\left(1 - t^l\right)^{r_l} $$ where $r_l$ is the rank of the $l$th graded quotient of the lower central series of $\pi_g^n$. \end{theorem} \begin{proof} Since $F_g^n$ is a smooth variety and a rational $K(\pi,1)$ by (\ref{kohno-oda}), we can apply (\ref{wt=lcs}) and the formula \cite[(9.7)]{hain:cycles} to deduce that $$ W_{F_g^n}(t) = \prod_{l=1}^\infty\left(1 - t^l\right)^{r_l}, $$ where, for a graded (variation of) MHS $H$ $$ W_H(t) = \sum_{k\ge 0} \chi(\Gr^W_k H)\, t^k $$ and, for an algebraic variety $X$, $W_X(t) = W_{H^{\bullet}(X)}(t)$. (Here $\chi$ denotes Euler characteristic.) The result now follows by induction on $n$ using the fact that $$ W_{C - \{x_1,\dots,x_n\}}(t) = 1 - 2g\, t + (n-1)t^2 $$ and the following lemma which is proved by induction on the length of the weight filtration of $V$. \end{proof} \begin{lemma} If ${\mathbb V}$ is an admissible unipotent variation of MHS over a smooth variety $X$, then $$ W_{H^{\bullet}(X,{\mathbb V})}(t) = W_X(t)W_{\mathbb V}(t). \qed $$ \end{lemma} \section{The General Case} \label{decorated} In this section we assemble results from Sections \ref{special} and \ref{braids2} to obtain a presentation of ${\mathfrak t}_{g,r}^n$ for all $g\ge 6$ and all $r$ and $n \ge 0$. Fix a base point $$ [C;x_1,\dots,x_n;v_1,\dots,v_n] $$ of ${\mathcal M}_{g,r}^n$ so that ${\mathfrak t}_{g,r}^n$ and ${\mathfrak p}_{g,r}^n$, etc.\ all have mixed Hodge structures. The sequence of Lie algebras $$ 0 \to {\mathfrak p}_{g,r}^n \to {\mathfrak t}_{g,r}^n \to {\mathfrak t}_g \to 0 $$ is exact in the category of MHSs, and therefore remains exact after applying $Gr^W_{\bullet}$: $$ 0 \to \Gr^W_{\bullet} {\mathfrak p}_{g,r}^n \to \Gr^W_{\bullet}{\mathfrak t}_{g,r}^n \to \Gr^W_{\bullet}{\mathfrak t}_g \to 0. $$ By (\ref{canon_split}), ${\mathfrak t}_{g,r}^n \otimes {\mathbb C}$ is isomorphic to the completion of $\Gr^W_{\bullet}{\mathfrak t}_{g,r}^n\otimes {\mathbb C}$. So, to find a presentation of ${\mathfrak t}_{g,r}^n$, it suffices to find a presentation of its associated graded. As in \S\ref{special}, there is a lift of the canonical homomorphism ${\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet} {\mathfrak t}_g$ to a homomorphism ${\mathbb L}(H_1({\mathfrak t}_g)) \to \Gr^W_{\bullet} {\mathfrak t}_{g,r}^n$. Given presentations of $\Gr^W_{\bullet} {\mathfrak t}_g$ and $\Gr^W_{\bullet} {\mathfrak p}_{g,r}^n$, a presentation of $\Gr^W_{\bullet} {\mathfrak t}_{g,r}^n$ is determined by maps \begin{gather*} a : H_1({\mathfrak t}_g)\otimes H_1({\mathfrak p}_{g,r}^n) \to {\mathfrak p}_{g,r}^n(2)\\ c : \Lambda^2 H_1({\mathfrak t}_g) \to {\mathfrak p}_{g,r}^n(2)_{\lambda_2}\\ c_0 : \Lambda^2 H_1({\mathfrak t}_g) \to {\mathfrak t}_{g,r}^n(2)^{Sp_g} \end{gather*} induced by the bracket. Observe that the homomorphism $$ {\mathfrak p}_{g,r}^n \to {\mathfrak p}_g^{\oplus(n+r)} $$ induced by the inclusion $F_{g,r}^n(C) \hookrightarrow C^{n+r}$ induces isomorphisms $$ H_1({\mathfrak p}_{g,r}^n) \cong H_1({\mathfrak p}_g)^{\oplus(n+r)}\text{ and } {\mathfrak p}_{g,r}^n(2)_{\lambda_2} \to {\mathfrak p}_g(2)^{\oplus(n+r)}. $$ By a naturality argument, the map $a$ is easily seen to be the adjoint of the map $$ H_1({\mathfrak t}_g) \to \Hom\left(\bigoplus_{j=1}^{r+n} H_1({\mathfrak p}_g), \bigoplus_{j=1}^{r+n} {\mathfrak p}_g(2)_{\lambda_2}\right) $$ which is the direct sum of $n+r$ copies of the Johnson homomorphism. The map $c$ is simply the sum over all the marked points and tangent vectors $$ \Lambda^2 H_2({\mathfrak t}_g) \to \bigoplus_{j=1}^{r+n}{\mathfrak p}_g(2) \cong {\mathfrak p}_{g,r}^n(2)_{\lambda_2}. $$ of the maps (\ref{bracket}) which is determined in (\ref{bra_const}). In remains to determine $c_0$. As in the case of ${\mathfrak t}_{g,1}$ considered in Section~\ref{special}, there is a canonical splitting of the sequence $$ 0 \to {\mathfrak p}_{g,r}^n(2)^{Sp_g} \to {\mathfrak t}_{g,r}^n(2)^{Sp_g} \to {\mathfrak t}_g^{Sp_g}(2) \to 0 $$ from which we obtain a canonical decomposition $$ {\mathfrak t}_{g,r}^n(2)^{Sp_g} = {\mathfrak p}_{g,r}^n(2)^{Sp_g} \oplus \Ga. $$ As in \S\ref{special}, we identify $H_1({\mathfrak t}_g)$ with the subspace of $\Lambda^3 V$ which is the kernel of the projection (\ref{projn}), denote by $\bil \phantom{x} \phantom{x}$ the unique $Sp_g$ invariant bilinear form on $H_1({\mathfrak t}_g)$ such that $$ \bil {a_1\wedge a_2 \wedge a_3} {b_1\wedge b_2 \wedge b_3} = 1, $$ and choose a generator $\gamma$ of $\Ga$ such that if $u,v\in H_1({\mathfrak t}_g)$, then the invariant part of $[u,v]$ in ${\mathfrak t}_g(2)$ is $\bil u v \,\gamma$. Observe that there is an exact sequence $$ 0 \to \bigoplus_{1\le i<j\le r+n} {\mathbb Q}(1) \oplus \bigoplus_{i=1}^r {\mathbb Q}(1) \to {\mathfrak p}_{g,r}^n(2) \to \bigoplus_{j=1}^{r+n} {\mathfrak p}_g(2) \to 0 $$ of $Sp_g$ modules. It follows that $$ {\mathfrak p}_{g,r}^n(2)^{Sp_g} \cong \bigoplus_{1\le i<j\le r+n} {\mathbb Q}(1) \oplus \bigoplus_{i=1}^r {\mathbb Q}(1). $$ The terms indexed by $1\le i \le r$ correspond to the $r$ marked tangent vectors; those indexed by $1\le i < j \le r+n$ to the diagonals $\Delta_{ij}$. It is easy to see that the composition $$ \Lambda^2 H_1({\mathfrak t}_g) \to \bigoplus_{i=1}^r {\mathbb Q}(1) $$ of $c_0$ with the projection $$ {\mathfrak t}_{g,r}^n(2)^{Sp_g} \to \bigoplus_{i=1}^r {\mathbb Q}(1) $$ is the sum of the maps $c_0$ associated to ${\mathfrak t}_{g,1}$ computed in (\ref{triv_cpt}). So it remains to determine the composition $$ \Lambda^2 H_1({\mathfrak t}_g) \to \bigoplus_{1\le i<j\le r+n} {\mathbb Q}(1) $$ of $c_0$ with the projection $$ e^{r+n} : {\mathfrak t}_{g,r}^n(2)^{Sp_g} \to \bigoplus_{1\le i<j\le r+n} {\mathbb Q}(1). $$ To do this, it suffices to compute $e^2$ in the case of ${\mathfrak t}_g^2$, for then the map $e^{r+n}$ is simply the sum of the $e^2$s over all diagonals. In order to compute $$ e^2 : {\mathfrak t}_{g,r}^n(2)^{Sp_g} \to {\mathbb Q}(1) $$ we use the fact that a punctured tubular neighbourhood of the diagonal $\Delta$ in $C\times C$ is homeomorphic to the frame bundle of the tangent bundle of $C$. In this way, we obtain homomorphisms $$ {\mathfrak t}_{g,1} \to {\mathfrak t}_g^2 \text{ and } {\mathfrak p}_{g,1} \to {\mathfrak p}_g^2. $$ In particular, we have a map \begin{equation}\label{rest} {\mathfrak p}_{g,1}(2)^{Sp_g} \to {\mathfrak p}_g^2(2)^{Sp_g}. \end{equation} Using the fact that $\Gr^W_{\bullet} {\mathfrak p}_{g,1} \to \Gr^W_{\bullet} {\mathfrak p}_g^2$ is a homomorphism and that on $H_1$, it is the diagonal map $V\to V\oplus V$, we see that the map (\ref{rest}) takes the generator $\sum_k [A_k,B_k]$ of ${\mathfrak p}_{g,1}(2)^{Sp_g}$ to \begin{multline*} \sum_{k=1}^g\, \left( [{\mathcal A} 1, \B 1] + [{\mathcal A} 1, \B 2] + [{\mathcal A} 2, \B 1] + [{\mathcal A} 2, \B 2]\right) \\ = 2 \sum_{k=1}^g\,\left( [{\mathcal A} 1, \B 2] + [{\mathcal A} 2, \B 1]\right). \end{multline*} It follows that in ${\mathfrak t}_g^2$, the map $c_0$ is given by $$ c_0[u,v] = \bil u v \,\gamma - \frac{12\bil u v}{g(2g+1)} \sum_{k=1}^g\, \left([{\mathcal A} 1, \B 2] + [{\mathcal A} 2, \B 1]\right). $$ This completes the determination of $c_0$ in general and, with it, the descriptions of the ${\mathfrak t}_{g,r}^n$. \section{Applications} \label{applications} \subsection{Cup products and Massey products} \label{cup} We have shown that for all $g \ge 6$ and all $r,n\ge 0$, the Lie algebra $\Gr^W_{\bullet} {\mathfrak t}_{g,r}^n$ has a presentation with only quadratic relations. This implies, using the short exact sequence in \cite{sullivan:les} for example, that the cup product $$ \Lambda^2 \Gr^W_{\bullet} H_{\mathrm{cts}}^1({\mathfrak t}_{g,r}^n) \to \Gr^W_{\bullet} H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n) $$ is surjective. It follows that the cup product $$ \Lambda^2 H_{\mathrm{cts}}^1({\mathfrak t}_{g,r}^n) \to H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n) $$ is also surjective. Recall that the $l$-fold Massey products constructed from $H^1(A^{\bullet})$, where $A^{\bullet}$ is a d.g.a., are defined on a subspace $D_l$ of $H^1(A^{\bullet})^{\otimes l}$ and take values in $H^2(A^{\bullet})/I_{l-1}$, where $I_{l-1}$ denotes the lift to $H^2(A^{\bullet})$ of the image of the Massey products of order $< l$. It follows that all Massey products in $H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n)$ of order $\ge 3$ vanish when $g\ge 6$ as the cup product (Massey products of order 2) map is surjective. Since the natural map $H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n) \to H^2(T_{g,r}^n,{\mathbb Q})$ is injective (cf.\ (\ref{cts_ord})) and preserves Massey products, we have: \begin{theorem} For all $g \ge 6$, all Massey products of order $\ge 3$ in $H^2(T_{g,r},{\mathbb Q})$ vanish. \qed. \end{theorem} \begin{remark} It follows from the fact that there are non-trivial cubic relations and no quadratic relations in a minimal presentation of ${\mathfrak t}_3$ that the cup product $$ \Lambda^2 H^1(T_3,{\mathbb Q}) \to H^2(T_3,{\mathbb Q}) $$ vanishes, and that the Massey triple product map $$ H^1(T_3,{\mathbb Q})^{\otimes 3} \to H^2(T_3,{\mathbb Q}) $$ is non-trivial. \end{remark} It follows from (\ref{lower}) and (\ref{computation}) that for all $g \ge 6$, $H_{\mathrm{cts}}^2({\mathfrak t}_g)$ has highest weight decomposition $$ H_{\mathrm{cts}}^2({\mathfrak t}_g) \cong V(\lambda_6) + V(\lambda_4) + V(\lambda_2) + V(\lambda_2 + \lambda_4). $$ \begin{theorem} For all $g \ge 3$, there is an (unnatural) isomorphism $$ H_{\mathrm{cts}}^2({\mathfrak t}_{g,r}^n) \cong H_{\mathrm{cts}}^2({\mathfrak t}_g) \oplus \left(H_{\mathrm{cts}}^1({\mathfrak p}_{g,r}^n)\otimes H_{\mathrm{cts}}^1({\mathfrak t}_g)\right) \oplus H_{\mathrm{cts}}^2({\mathfrak p}_{g,r}^n) $$ of $Sp_g$ modules. \end{theorem} \begin{proof} Chose a base point of ${\mathcal M}_{g,r}^n$. Then $$ 0 \to {\mathfrak p}_{g,r}^n \to {\mathfrak t}_{g,r}^n \to {\mathfrak t}_g \to 0 $$ is an exact sequence of mixed Hodge structures, and the corresponding spectral sequence $$ E_2^{s,t} = H_{\mathrm{cts}}^s({\mathfrak t}_g,H_{\mathrm{cts}}^t({\mathfrak p}_{g,r}^n)) \implies H_{\mathrm{cts}}^{s+t}({\mathfrak t}_{g,r}^n) $$ is a spectral sequence in the category of mixed Hodge structures. Since ${\mathfrak t}_g$ has negative weights, the weights on $H_{\mathrm{cts}}^k({\mathfrak t}_g)$ are $\ge k$. This and the fact that $H^k({\mathfrak p}_{g,r}^n)$ is a trivial ${\mathfrak t}_g$ module when $k = 1$ and 2 imply that $E_\infty^{s,t} = E_2^{s,t}$ when $s+t=2$. The result follows. \end{proof} \subsection{Period space is not contractible when $g\ge 4$} Denote by ${\mathfrak h}_g$ the Siegel upper half space; that is, the space of symmetric $g\times g$ complex matrices with positive definite imaginary part. Denote the image of the period map $$ \text{Teichm\"uller space } \to {\mathfrak h}_g $$ by ${\mathcal J}_g$, and its closure by $\overline{{\mathcal J}}_g$. When $g \le 3$, $\overline{\J}_g={\mathfrak h}_g$, so that $\overline{\J}_g$ is contractible in these cases. \begin{theorem} For each $g\ge 4$, $H^2(\overline{\J}_g,{\mathbb Q})$ is non-trivial. Consequently, $\overline{\J}_g$ is not contractible. \end{theorem} The proof proceeds in two steps. We begin by making a definition. \begin{definition} The {\it extended Torelli group} $\widehat{T}_g$ is the subgroup of $\Gamma_g$ consisting of those mapping classes which act as $\pm$ the identity on the first homology of the reference surface. \end{definition} We have group extensions \begin{equation}\label{extensions} 1 \to T_g \to \widehat{T}_g \to {\mathbb Z}/2{\mathbb Z} \to 0\text{ and } 1 \to \widehat{T}_g \to \Gamma_g \to PSp_g({\mathbb Z}) \to 1 \end{equation} where $PSp_g({\mathbb Z})$ denotes the quotient of $Sp_g({\mathbb Z})$ by $\pm I$. Note that the first sequence gives rise to a natural action of ${\mathbb Z}/2{\mathbb Z}$ on $H^{\bullet}(T_g)$. The first step is: \begin{proposition} For all $g \ge 3$, there are natural isomorphisms $$ H^{\bullet}({\mathcal J}_g,{\mathbb Q}) \cong H^{\bullet}(\widehat{T}_g,{\mathbb Q}) \cong H^{\bullet}(T_g,{\mathbb Q})^{{\mathbb Z}/2{\mathbb Z}}. $$ Moreover, when $g\ge 4$, $H^2(\widehat{T}_g,{\mathbb Q})$ is non-trivial. \end{proposition} \begin{proof} Since $g\ge 3$, ${\mathcal J}_g$ is the quotient of Teichm\"uller space by $\widehat{T}_g$. Since the mapping class group acts on Teichm\"uller space virtually freely, this implies (via standard arguments) that there is a natural isomorphism $$ H^{\bullet}(\widehat{T},{\mathbb Q}) \cong H^{\bullet}({\mathcal J}_g,{\mathbb Q}). $$ Applying the Hochschild-Serre spectral sequence to the first of the extensions (\ref{extensions}) above, we see that $$ H^k(\widehat{T}_g,{\mathbb Q}) \cong H^k(T_g,{\mathbb Q})^{{\mathbb Z}/2}. $$ But $-I \in Sp_g({\mathbb Z})$ acts trivially on $H_{\mathrm{cts}}^2({\mathfrak t}_g)$, which implies that $$ H_{\mathrm{cts}}^2({\mathfrak t}_g) \subseteq H^2(\widehat{T}_g,{\mathbb Q}). $$ The result follows. \end{proof} \begin{remark} This argument also shows that the image of the cup product $$ \Lambda^2 H^1(T_3,{\mathbb Z}) \to H^2(T_3,{\mathbb Z}) $$ is 2-torsion. \end{remark} To complete the proof of the theorem, note that ${\mathcal J}_g = \overline{\J}_g - {\mathcal R}$ where ${\mathcal R}$ is the locus of reducible jacobians. By standard arguments, each component of ${\mathcal R}$ has complex codimension $\ge 2$ in $\overline{\J}_g$. Combining Lefschetz duality and the Gysin sequence, we obtain an exact sequence $$ H^{BM}_{6g-k-6}({\mathcal R}) \to H^k(\overline{\J}_g) \to H^k({\mathcal J}_g) \to H^{BM}_{6g-k-7}({\mathcal R}), $$ where $H_{\bullet}^{BM}$ denotes Borel-Moore homology. Since ${\mathcal R}$ has real codimension 4, it follows that $H^2(\overline{\J}_g) \cong H^2({\mathcal J}_g)$. The theorem follows as $H^2({\mathcal J}_g)$ is non-trivial. \subsection{Johnson's conjecture} In \cite{johnson:survey}, Johnson constructed maps $$ \phi_k : H_k(T_g) \to H_{k+2}(\Jac S)/[S]\times H_k(\Jac S), $$ which generalize the classical Johnson homomorphism, which is the case $k=1$. He conjectured that these homomorphisms are isomorphisms for all $k$ and sufficiently large $g$. The following result is an improvement of some unpublished computations of Morita (cf.\ \cite[\S4]{morita:jap_acad}). \begin{theorem} For all $g \ge 3$, the map $\phi_2$ is not injective. \end{theorem} \begin{proof} It is not difficult to see that each $\phi_k$ is $Sp_g({\mathbb Z})$ equivariant. Consider its adjoint $$ \phi_k^t : H^{k+2}(\Jac S,{\mathbb Q})/\omega\wedge H^k(\Jac S,{\mathbb Q}) \to H^k(T_g,{\mathbb Q}). $$ This is also $Sp_g({\mathbb Z})$ equivariant. The domain of $\phi_2^t$ is the primitive cohomology group $PH^4(\Jac S,{\mathbb Q})$. This is the restriction to $Sp_g({\mathbb Z})$ of the rational representation of $Sp_g$ with highest weight $\lambda_4$. Since this is an irreducible $Sp_g({\mathbb Z})$ module, the image of $\phi_2^t$ is either isomorphic to $V(\lambda_4)$ or trivial. But $H^2(T_g,{\mathbb Q})$ contains the rational representation $H_{\mathrm{cts}}^2({\mathfrak t}_g)$. It follows from the results in \S\ref{cup} that $$ H_{\mathrm{cts}}^2({\mathfrak t}_g)/\im \phi_2^t \cap H_{\mathrm{cts}}^2({\mathfrak t}_g) $$ is non-trivial as it contains $V(\lambda_6) + V(\lambda_2 + \lambda_4)$ when $g\ge 6$; $V(\lambda_2 + \lambda_4)$ when $g = 4,5$; and $V(\lambda_3)$ when $g = 3$. The result follows. \end{proof} \subsection{Filtrations of $T_g^1$}\label{filtn} Define a filtration \begin{equation}\label{filtration} T_g = L^1 T_g^1 \supseteq L^2 T_g^1 \supseteq L^3 T_g^1 \supseteq \cdots \end{equation} of $T_g^1$ by $$ L^k T_g^1 = \left\{\phi \in T_g^1 : \phi_\ast : \pi_1(S,x) \to \pi_1(S,x) \text{ is congruent to the identity mod } \Gamma^{k+1}\right\}. $$ It is quite common in the literature for this filtration to be called the {\it relative weight filtration}, as it is in \cite{asada-nakamura} and \cite{oda}. In view of (\ref{mhs_torelli}) and (\ref{unequal}), I feel that this terminology is likely to result in confusion. \begin{proposition} This filtration is a descending central series of $T_g^1$ with torsion free quotients and has the property that $$ \bigcap_{k=1}^\infty L^k T_g^1 $$ is trivial. \end{proposition} \begin{proof} This follows directly from the fact that the fundamental group of a compact Riemann surface is residually nilpotent \cite{baumslag}, and the fact that the graded quotients of the lower central series of a surface group are torsion free \cite{labute}. \end{proof} The most rapidly descending series with torsion free quotients of a group $G$ is the series $$ G = D^1 G \supseteq D^2 G \supseteq D^3 G \supseteq \cdots $$ where $$ D^k G = \{g\in G : \text{ there is an integer $n>0$ such that } g^n \in \Gamma^k G\}. $$ This filtration has the property that $D^kG/D^{k+1}$ is the $k$th term of the lower central series of $G$ mod torsion. Proofs of these assertions can be found in \cite{passman}. In the current situation, we have $$ D^k T_g^1 \subseteq L^k T_g^1. $$ Johnson's Theorem \cite{johnson:h1} implies that $D^2 T_g^1 = L^2 T_g^1$ when $g \ge 3$. The computations (\ref{upper}) and (\ref{lower}) imply that the kernel of $D^2 T_g^1 / D^3 \to L^2 T_g^1 /L^3$ is isomorphic to ${\mathbb Z}$. Morita was aware of the fact that the kernel was at least this big --- cf.\ his work on the Casson invariant \cite{morita:casson}, and asked whether there is a $k$ such that $D^3 T_g^1 \supseteq L^k T_g^1$. That is, whether the kernel of $D^2 T_g^1 / D^3 \to L^2 T_g^1 /L^3$ can be detected by the action of $T_g^1$ on the quotients of $\pi_g$ by the terms of its lower central series. More generally, one can ask if the topologies on $T_g^1$ determined by the filtrations $D^{\bullet}$ and $L^{\bullet}$ are equivalent. (Both are separated.) That is, for each $k\in {\mathbb N}$, can one find a positive integer $n(k)$ such that $L^{n(k)} T_g^1 \subseteq D^k T_g^1$~? Since the groups $T_g^1/D^k$ and $T_g^1/L^k$ are torsion free nilpotent, they imbed as a Zariski dense subgroup of a unipotent group defined over ${\mathbb Q}$. One obtains two inverse systems of unipotent groups. It is clear that the first prounipotent group is the Malcev completion ${\mathcal T}_g^1$ of $T_g^1$, and the second is the prounipotent group associated to the pronilpotent Lie algebra ${\mathfrak h}_g:=\im\{{\mathfrak t}_g^1 \to {\mathfrak d}_g\}$, where ${\mathfrak d}_g$ is the pronilpotent Lie algebra defined in \S \ref{inf_action}. The two topologies on $T_g^1$ are equivalent if and only if the natural map ${\mathfrak t}_g^1 \to {\mathfrak h}_g$ is an isomorphism. Equivalently, they are equivalent if and only if ${\mathfrak t}_g^1 \to {\mathfrak d}_g$ is injective. It is also clear that the filtration $L^{\bullet}$ of ${\mathfrak t}_g^1$ induced from that of $T_g^1$ is the pullback of the weight filtration of ${\mathfrak h}_g^1$, so that $$ \left(L^kT_g^1/L^{k+1}\right) \otimes{\mathbb Q} \cong \Gr^k_L {\mathfrak t}_g^1 \cong \Gr^W_{-k} {\mathfrak h}_g^1 $$ and that $L^k{\mathfrak t}_g^1 \supseteq \Ga$ for all $k\ge 1$ --- cf.\ (\ref{central_ext}). \begin{theorem}\label{unequal} For all $g \ge 3$, and all $k\ge 1$, $L^k{\mathfrak t}_g^1 \supseteq \Ga$ so that the natural representation ${\mathfrak t}_g^1 \to {\mathfrak d}_g$ is not injective as its kernel contains $\Ga$. In particular, there is no $k\ge 1$ such that $W_{-3}{\mathfrak t}_g^1 \supseteq L^k{\mathfrak t}_g^1$. \qed \end{theorem} One can define a filtration $L^{\bullet}$ of $T_g$ by defining $L^k T_g$ to be the image of $L^k T_g^1$. Using similar arguments, one can prove that the filtrations $L^{\bullet}$ and $D^{\bullet}$ of $T_g$ do not define equivalent topologies. \subsection{A question of Asada and Nakamura} There is an issue raised by Asada and Nakamura in \cite[(4.5)]{asada-nakamura} which is closely related to Morita's question. Denote by $\pi_{g,1}$ the fundamental group $\pi_1(S,v)$ of $S$ with respect to the tangent vector $v$. It is naturally isomorphic to the fundamental group of the punctured surface $S$ minus the anchor point $x$ of $v$. Note that $T_{g,1}$ acts on $\pi_{g,1}$. They define a filtration $M^{\bullet}$ of $T_g^1$ as follows: First define a filtration $L^{\bullet}$ of $T_{g,1}$ as in the previous section: $\phi$ is in $L^kT_{g,1}$ if and only if $\phi$ induces the identity on $\pi_{g,1}$ modulo the $(k+1)$st term of its lower central series. Define $M^kT_g^1$ to be the image of $L^kT_{g,1}$ in $T_g^1$. They then ask whether, after tensoring with ${\mathbb Q}$, the sequence $$ 0 \to \Gr^W_{\bullet} \pi_g \to \Gr^M_{\bullet} T_g^1 \to \Gr^L_{\bullet} T_g \to 0 $$ is exact. (Recall from (\ref{wt=lcs}) that the lower central series of ${\mathfrak p}_g$ agrees with its weight filtration.) We now give a proof that this is indeed the case. We continue with the notation of the previous section. The filtration $M^{\bullet}$ induces a filtration of ${\mathfrak t}_g^1$. Their question then beocmes: is the sequence $$ 0 \to \Gr^W_{\bullet} {\mathfrak p}_g \to \Gr^M_{\bullet} {\mathfrak t}_g^1 \to \Gr^L_{\bullet} {\mathfrak t}_g \to 0 $$ exact? Fix a base point of ${\mathcal M}_{g,1}$ so that ${\mathfrak t}_{g,1}$, ${\mathfrak t}_g^1$, $\pi_{g,1}$, ${\mathfrak d}_g$, ${\mathfrak p}_g$, etc.\ all have compatible MHSs; the MHS on ${\mathfrak p}_{g,1}$ is the limit MHS on $\pi_1(S-\{x\},x_o)$ associated to the ``degeneration'' where $x_o$ approaches $x$ from the direction of $v$. Denote the image of ${\mathfrak t}_g^1$ in ${\mathfrak d}_g$ by ${\mathfrak h}_g^1$, and the image of ${\mathfrak t}_g$ in ${\mathfrak o}_g$ by ${\mathfrak h}_g$. These have canonical mixed Hodge structures determined by the choice of the base point. Since the diagram $$ \begin{CD} 0 @>>> {\mathfrak p}_g @>>> {\mathfrak t}_g^1 @>>> {\mathfrak t}_g @>>> 0\\ @. @| @VVV @VVV \\ 0 @>>> {\mathfrak p}_g @>>> {\mathfrak h}_g^1 @>>> {\mathfrak h}_g @>>> 0\\ \end{CD} $$ commutes and since the top row is exact, it follows that the bottom row is exact. Since $\Gr^W_{\bullet}$ is an exact functor, and since $\Gr^W_k {\mathfrak h}_g^n \cong \Gr^L_k{\mathfrak t}_g^n$ when $n = 0,1$, this implies that the sequence $$ 0 \to \Gr^W_{\bullet} {\mathfrak p}_g \to \Gr^L_{\bullet} {\mathfrak t}_g^1 \to \Gr^L_{\bullet} {\mathfrak t}_g \to 0 $$ is exact. To complete the proof, we show that the filtrations $L^{\bullet}$ and $M^{\bullet}$ of ${\mathfrak t}_g^1$ are equal. Denote by $b_o$ the element of $\pi_{g,1}$ that corresponds to rotating the tangent vector once about $x$ --- this is a ``Dehn twist about the boundary of $S-\{x\}$.'' The action of $T_{g,1}$ on $\pi_{g,1}$ fixes $b_o$, and therefore induces a homomorphism $T_{g,1} \to \Aut(\pi_{g,1},b_o)$ into the automorphisms of $\pi_{g,1}$ that fix $b_o$. Set $w_o = \log b_o$. This we interpret as an element of ${\mathfrak p}_{g,1}$. The homomorphism above induces a homomorphism $T_{g,1} \to \Aut({\mathfrak p}_{g,1},w_o)$, and therefore a Lie algebra homomorphism $$ {\mathfrak t}_{g,1} \to \Der({\mathfrak p}_{g,1},w_o) $$ into the derivations of ${\mathfrak p}_{g,1}$ that annhilate $w_o$. It follows from standard properties of limit MHSs that $w_o$ spans a copy of ${\mathbb Q}(1)$ in $\Der{\mathfrak p}_{g,1}$. But $\Der({\mathfrak p}_{g,1},w_o)$ is the kernel of the map $\Der {\mathfrak p}_{g,1} \to {\mathfrak p}_{g,1}$ that takes $\phi$ to $\phi(w_o) - w_o$. Since $w_o$ is a Hodge class, this is a morphism of MHS. It follows that $\Der({\mathfrak p}_{g,1},w_o)$ has a natural MHS. Since ${\mathfrak p}_g$ is the quotient of ${\mathfrak p}_{g,1}$ by the ideal generated by $w_o$ as MHS, the homomorphism $$ \Der({\mathfrak p}_{g,1},w_o) \to \Der {\mathfrak p}_g $$ is a morphism of MHS. The filtration $M^{\bullet}$ of ${\mathfrak t}_{g,1}$ is the inverse image of the weight filtration under the homomorphism ${\mathfrak t}_{g,1}\to \Der({\mathfrak p}_{g,1},w_o)$. The equality of the filtrations $L^{\bullet}$ and $M^{\bullet}$ of ${\mathfrak t}_g^1$ now follows from the strictness properties of the weight filtration as the diagram $$ \begin{CD} {\mathfrak t}_{g,1} @>>> {\mathfrak t}_g^1 \cr @VVV @VVV \cr \Der({\mathfrak p}_{g,1},w_o) @>>> \Der {\mathfrak p}_g \cr \end{CD} $$ commutes and all arrows are morphisms of MHS. \subsection{Cohomology of ${\mathfrak t}_g$ and vanishing differentials} Since ${\mathfrak t}_{g,r}^n = W_{-1}{\mathfrak t}_{g,r}^n$, it follows that $$ W_{k-1}H_{\mathrm{cts}}^k({\mathfrak t}_{g,r}^n) = 0 $$ for all $k\ge 0$. The {\it lowest weight subring of} $H_{\mathrm{cts}}^{\bullet}({\mathfrak t}_{g,r}^n)$ is defined to be the subring $$ \bigoplus_{k\ge 0} W_kH_{\mathrm{cts}}^k({\mathfrak t}_{g,r}^n). $$ By \cite[(9.2)]{hain:cycles}, this is a quadratic algebra generated by $H_{\mathrm{cts}}^1({\mathfrak t}_{g,r}^n)$ and where the relations are dual to the second graded quotient of the lower central series of $T_{g,r}^n$. The following result is a refinement of (\ref{vanishing}). \begin{theorem}\label{van_diffls} For each irreducible representation $V(\lambda)$ of $Sp_g$, the image of the natural homomorphism $$ \left[W_kH_{\mathrm{cts}}^k({\mathfrak t}_{g,r}^n) \otimes V(\lambda)\right]^{Sp_g} \to H^0(Sp_g({\mathbb Z}),H^k(T_{g,r}^n)\otimes V(\lambda)) = E_2^{0,k} $$ is contained in $$ E_\infty^{0,k} = \im\left\{ H^k(\Gamma_{g,r}^n,V(\lambda)) \to H^k(T_{g,r}^n)\otimes V(\lambda) \right\}. $$ \end{theorem} \begin{proof} Fix a base point of ${\mathcal M}_{g,r}^n$ so that ${\mathfrak t}_{g,r}^n$, ${\mathfrak u}_{g,r}^n$, etc.\ all have compatible MHSs. Since the extension $$ 0 \to \Ga \to {\mathfrak t}_{g,r}^n \to {\mathfrak u}_{g,r}^n \to 0 $$ is central with kernel isomorphic to ${\mathbb Q}(1)$, it follows from the Gysin sequence that the induced map $$ \bigoplus_{k\ge 0} W_kH_{\mathrm{cts}}^k({\mathfrak u}_{g,r}^n) \to \bigoplus_{k\ge 0} W_kH_{\mathrm{cts}}^k({\mathfrak t}_{g,r}^n). $$ is surjective, with kernel the ideal generated by the cohomology class in $W_2H_{\mathrm{cts}}^2({\mathfrak u}_{g,r}^n)$ corresponding to the extension above. By (\ref{morph_u}), there is a canonical map $$ \left[H_{\mathrm{cts}}^k({\mathfrak u}_{g,r}^n)\otimes V(\lambda)\right]^{Sp_g} \longrightarrow H^k(\Gamma_{g,r}^n,{\mathbb V}(\lambda)). $$ The result follows because the diagram $$ \begin{CD} \left[W_kH_{\mathrm{cts}}^k({\mathfrak u}_{g,r}^n)\otimes V(\lambda)\right]^{Sp_g} @>>> H^{\bullet}(\Gamma_{g,r}^n,{\mathbb V}(\lambda)) \\ @VVV @VVV \\ \left[W_kH_{\mathrm{cts}}^k({\mathfrak t}_{g,r}^n)\otimes V(\lambda)\right]^{Sp_g} @>>> H^0(Sp_g({\mathbb Z}),H^k(T_{g,r}^n)\otimes V(\lambda)) \end{CD} $$ commutes, and because the left hand vertical map is surjective. \end{proof} \subsection{Morita's Conjecture} We now prove a result which is, in some sense, an affirmation of Morita's conjecture \cite[2.7]{morita:conj}. Our result is an analogue of his theorem \cite[6.2]{morita:conj} which is a solution to the conjecture in the first non-trivial case. He also informs me that he has proved the second non-trivial case of the conjecture over ${\frac{1}{24}}{\mathbb Z}$. Suppose that $g\ge 3$. Denote the $k$th term of the lower central series of $\pi_g$ by $\pi^{(k)}$. Set $$ \pi_{(k)} = \pi_g/\pi^{(k+1)}. $$ We know from Labute's theorem \cite{labute} that this is a torsion free nilpotent group. For each $k\ge 1$, there is a natural representation $$ \rho_k : \Gamma_g^1 \to \Aut \pi_{(k)}. $$ The first is simply the standard representation $\Gamma_g^1 \to Sp_g({\mathbb Z})$. Denote the ${\mathbb Q}$ form of the unipotent completion of $\pi_{(k)}$ by ${\mathcal P}_{(k)}$. Since $\pi_{(k)}$ is torsion free, the canonical map $\pi_{(k)} \to {\mathcal P}_{(k)}$ is injective. By the universal mapping property of unipotent completion, we see that each $\rho_k$ extends to a homomorphism $$ \tilde{\rho}_k : \Gamma_g^1 \to \Aut {\mathcal P}_{(k)}. $$ Denote the Lie algebra of ${\mathcal P}_{(k)}$ by ${\mathfrak p}_{(k)}$. Then $\Aut {\mathcal P}_{(k)} \cong \Aut {\mathfrak p}_{(k)}$. It follows that $\Aut {\mathcal P}_{(k)}$ is a linear algebraic group. Denote the Zariski closure of the image of $\tilde{\rho}_k$ in this by $G_k$. It is easy to see that $G_k$ is an extension of $Sp_g({\mathbb Q})$ by a unipotent group: $$ 1 \to U_k \to G_k \to Sp_g({\mathbb Q}) \to 1 $$ This extension is split exact, so that $$ G_k \cong Sp_g({\mathbb Q}) \ltimes U_k. $$ By the universal mapping property of the relative completion of $\Gamma_g^1$, there is a homomorphism ${\mathcal G}_g^1 \to G_k$ which commutes with the projections to $Sp_g$. The following result follows directly from the fact (\ref{central_ext}) that the natural map ${\mathcal T}_g^1 \to {\mathcal U}_g^1$ is surjective.\footnote{A direct proof of the lemma can be given --- cf.\ the proof of \cite[(4.6)]{hain:comp}.} \begin{lemma} For each $k\ge 2$, $\tilde{\rho}(T_g^1)$ is Zariski dense in $U_k$. \qed \end{lemma} \begin{proposition} For each $k \ge 2$, the image of $\tilde{\rho}_k$ is a discrete subgroup of $G_k({\mathbb R})$, and the quotient $\im\rho_k\backslash G_k({\mathbb R})$ has finite volume with respect to any left invariant metric on $G_k({\mathbb R})$. \end{proposition} \begin{proof} Since every finitely generated subgroup of the ${\mathbb Q}$ points of a unipotent group $U$ is discrete in $U({\mathbb R})$, it follows that $\tilde{\rho}_k(T_g^1)$ is a discrete subgroup of $U_k({\mathbb R})$. Since it is also Zariski dense, it is cocompact. The result now follows as the image of $\Gamma_g^1$ in $Sp_g({\mathbb R})$ is $Sp_g({\mathbb Z})$, which is discrete and of finite covolume. \end{proof} We should note that Morita works with $\Gamma_{g,1}$ rather than with $\Gamma_g^1$ as we do. Our arguments work equally well in his case; we chose to work with $\Gamma_g^1$ as it seems more natural. In conclusion, we remark that the Lie algebra of $U_k$ is simply the image ${\mathfrak h}_g^1/W_{-k-1}$ of ${\mathfrak u}_g^1$ in $\Der {\mathfrak p}_{(k)}$. It follows that the Lie algebra of $U_k$ has a MHS, and is therefore isomorphic to its associated graded after tensoring with ${\mathbb C}$. \section{The Universal Connection} \label{connection} In this section we construct a universal connection form $$ \widetilde{\omega} \in E^{\bullet}(\text{Torelli space})\comptensor \Gr^{\bullet} {\mathfrak t}_{g,r}^n $$ with ``scalar curvature'' on Torelli space when $g\ge 3$. Here $E^{\bullet}(X)$ denotes the $C^\infty$ de~Rham complex of a smooth manifold $X$, and $\comptensor$ the completed tensor product.% \footnote{The completed tensor product $E^{\bullet}(X)\comptensor\Gr^{\bullet}{\mathfrak g}$ is defined to be the inverse limit $$ \lim_\leftarrow E^{\bullet}(X)\otimes \Gr^{\bullet}{\mathfrak g} /\oplus_{l\ge m} \Gr^l{\mathfrak g}. $$ } This is the analogue of the universal connection $$ \sum_{ij} d\log(z_i - z_j)\, X_{ij} \in E^{\bullet}(X_n) \otimes \Gr^{\bullet}{\mathfrak p}_n $$ for the braid group $P_n$. Here $X_n$ denotes the classifying space $$ {\mathbb C}^n - \left\{\text{fat diagonal}\right\} $$ of the pure braid group, $(z_1,\dots,z_n)$ its coordinates, and ${\mathfrak p}_n$ the Malcev Lie algebra associated to $P_n$. A reasonably precise dictionary between the case of braid groups and the absolute mapping class groups $\Gamma_g$ is given in the table. \begin{table} {\footnotesize \begin{tabular}{c|c|p{1.2in}} \hline Braid Groups & Mapping Class Groups & \quad Comments \\ \hline &&\\ $B_n$ & $\Gamma_g$ & {\raggedright the group of interest}\\ &&\\ $\Sigma_n$ & $Sp_g({\mathbb C})$ & {\raggedright a semi-simple algebraic group $G$}\\ &&\\ $\rho:B_n \to \Sigma_n$ & $\rho:\Gamma_g \to Sp_g({\mathbb C})$ & {\raggedright homomorphism to $G$ with dense image}\\ &&\\ $\Sigma_n$ & $Sp_g({\mathbb Z})$ & {\raggedright the image of $\rho$, an arithmetic group}\\ &&\\ $P_n$ & $T_n$ & {\raggedright the kernel of $\rho$, a residually torsion free nilpotent group}\\ &&\\ ${\mathcal P}_n$ & ${\mathcal U}_g$ & {\raggedright prounipotent radical of the relative completion}\\ &&\\ $B_n \to \Sigma_n \ltimes {\mathcal P}_n$ & $\Gamma_g \to {\mathcal G}_g \cong Sp_g({\mathbb C})\ltimes {\mathcal U}_g$ & {\raggedright the relative completion}\\ &&\\ ${\mathcal P}_n$ & ${\mathcal T}_g$ & {\raggedright the unipotent completion of the kernel of $\rho$}\\ &&\\ $\id: {\mathcal P}_n \to {\mathcal P}_n$ & ${\mathcal T}_g \to {\mathcal U}_g$ & {\raggedright the homomorphism to the prounipotent radical}\\ &&\\ ${\mathfrak p}_n$ & ${\mathfrak t}_g$ & {\raggedright the pronilpotent Lie algebra corresponding to $\ker\rho$}\\ &&\\ $\Gr^{\bullet} {\mathfrak p}_n = {\mathbb L}(H_1(P_n))/R$ & $\Gr^{\bullet} {\mathfrak t}_g = {\mathbb L}(H_1(T_g))/R$ & {\raggedright quadratic presentations as graded Lie algebras in the category of representations of $G$}\\ &&\\ $\left\{[X_{ij},X_{kl}],[X_{ij},X_{ik}+X_{jk}]\right\}$ & $V(\lambda_6), V(\lambda_4), V(\lambda_2), V(\lambda_2 + \lambda_4)$ & {\raggedright the quadratic relations}\\ &&\\ $X_n := {\mathbb C}^n - \{\text{fat diagonal}\}$ & ${\mathcal H}_g := \text{ Torelli space}$ & {\raggedright the classifying space of the kernel of $\ker\rho$}\\ &&\\ $Y_n := \Sigma_n\backslash X_n$ & ${\mathcal M}_g = Sp_g({\mathbb Z})\backslash {\mathcal H}_g$ & {\raggedright the classifying space of the group of interest}\\ &&\\ $\sum_{ij} w_{ij}\, X_{ij} \in E^{\bullet}(X_n)\otimes \Gr^{\bullet}{\mathfrak p}_n$ & $\omega \in E^{\bullet}({\mathcal H}_g)\comptensor \Gr^{\bullet}{\mathfrak t}_g$ & {\raggedright the ``universal (projectively) flat connection'' on the classifying space of $\ker\rho$} \end{tabular} } \end{table} \begin{question} The Lie algebra $\Gr^{\bullet} {\mathfrak p}_n$ has interesting finite dimensional representations; namely those associated to Hecke algebras. Are there analogous representations of $\Gr^{\bullet} {\mathfrak t}_{g,r}^n$ where the canonical central $\Ga$ acts via scalar transformations? These should lead to interesting projective representations of $\Gamma_{g,r}^n$. \end{question} We now give the construction of the connection. First recall that Torelli space ${\mathcal H}_{g,r}^n$ is the quotient of the Teichm\"uller space associated to $\Gamma_{g,r}^n$ by $T_{g,r}^n$. It is the moduli space of isomorphism classes of $(n+r+2g+1)$-tuples $$ (C;x_1,\dots,x_n;v_1,\dots,v_r;a_1,\dots,a_g,b_1,\dots,b_g) $$ where $C$ is a compact Riemann surface of genus $g$, $x_1,\dots,x_n$ are $n$ marked points, $v_1,\dots,v_r$ are $r$ marked tangent vectors, and $a_1,\dots, b_g$ is a symplectic basis of $H_1(C,{\mathbb Z})$. Since $T_{g,r}^n$ is torsion free and Teichm\"uller space is contractible, ${\mathcal H}_{g,r}^n$ is the classifying space of $T_{g,r}^n$. The bulk of the work needed for the construction of the connection has already been done in \cite[\S14]{hain:derham}. Fix a point $x \in {\mathcal M}_{g,r}^n$. It follows from \cite[\S14.2]{hain:derham} that there is a 1-form $$ \omega \in E^{\bullet}({\mathcal H}_{g,r}^n)\comptensor \Gr^W_{\bullet} {\mathfrak u}_{g,r}^n $$ which is integrable and is $Sp_g({\mathbb Z})$ invariant in that \begin{equation}\label{invariance} s^\ast \omega = Ad(s)\, \omega \end{equation} for all $s\in Sp_g({\mathbb Z})$. That is, if $$ \omega = \sum_I w_I X_I, \text{ where } w_I \in E^1({\mathcal H}_{g,r}^n) \text{ and } X_I \in \Gr^W_{\bullet} {\mathfrak u}_{g,r}^n, $$ then for all $s\in Sp_g({\mathbb Z})$, $$ \sum_I (s^\ast w_I) X_I = \sum_I w_I (X_I\cdot s^{-1}) $$ where $X\cdot s$ denotes the canonical action of $s\in S$ on $X\in {\mathfrak u}_{g,r}^n$. This should be compared with the case of braids where the corresponding formula is easily verified --- cf. \cite[(14.6)]{hain:derham}. Since there is a canonical isomorphism $$ {\mathfrak u}_{g,r}^n(x) \cong \prod_{l\ge 1}\Gr^W_{-l}{\mathfrak u}_{g,r}^n(x) $$ this form gives rise to a flat connection on the trivial right ${\mathcal U}_{g,r}^n$ principal bundle $$ {\mathcal H}_{g,r}^n \times {\mathcal U}_{g,r}^n \to {\mathcal H}_{g,r}^n. $$ Note that $Sp_g({\mathbb Z})$ acts on this bundle via the diagonal action. The composite $$ {\mathcal U}_{g,r}^n \to {\mathcal H}_{g,r}^n \to {\mathcal M}_{g,r}^n $$ is a left principal $Sp_g({\mathbb Z})\ltimes {\mathcal U}_{g,r}^n$ bundle (in the orbifold sense.) The invariance condition (\ref{invariance}) means that the connection defined by $\omega$ is invariant under the $Sp_g({\mathbb Z})\ltimes {\mathcal U}_{g,r}^n$ action. The monodromy yields a representation $$ \Gamma_{g,r}^n \to Sp_g({\mathbb C})\ltimes {\mathcal U}_{g,r}^n(x) $$ As proved in \cite[\S14.2]{hain:derham}, this is the ${\mathbb C}$ form of the completion of $\Gamma_{g,r}^n$ with respect to the canonical homomorphism $\Gamma_{g,r}^n \to Sp_g({\mathbb C})$. Since the sequence $$ 0 \to \Ga \to {\mathfrak t}_{g,r}^n \to {\mathfrak u}_{g,r}^n \to 0 $$ splits canonically over ${\mathbb C}$ (given the choice of the base point $x$), $\omega$ has a canonical lift $$ \widetilde{\omega} \in E^{\bullet}({\mathcal H}_{g,r}^n)\comptensor \Gr^W_{\bullet} {\mathfrak t}_{g,r}^n $$ to $\Gr^W_{\bullet}{\mathfrak t}_{g,r}^n$. This form is not integrable, but since $\omega$ is integrable, the curvature of $\widetilde{\omega}$ takes values in the central $\Ga$. It also has the invariance property (\ref{invariance}). We will say that a representation $\phi : \Gr^W_{\bullet}{\mathfrak t}_{g,r}^n \to \End(V)$ is {\it projective} if the image of $\Ga$ consists of scalar matrices. If $V$ is an $Sp_g$ module and $\phi$ is $Sp_g$ equivariant, then $\phi$ should integrate to a homomorphism $$ \Gamma_{g,r}^n \to PGL(V), $$ at least when $\phi$ is ``sufficiently small,'' since, in this case, the composite $$ \omega_\phi \in E^{\bullet}(X_n)\otimes \End(V)/\text{scalars}, $$ an infinite sum, should converge to an integrable 1-form. The equivariance of $\phi$ implies that $\omega_\phi$ has the invariance property (\ref{invariance}), leading to a projectively flat bundle over ${\mathcal M}_{g,r}^n$ with fiber $V$ over the base point $x$.
1995-12-20T06:20:20
9512
alg-geom/9512012
en
https://arxiv.org/abs/alg-geom/9512012
[ "alg-geom", "math.AG" ]
alg-geom/9512012
Fernando Torres
Fernando Torres
Remarks on numerical semigroups
Latex2e, ICTP preprint
null
null
null
null
We extend results on Weierstrass semigroups at ramified points of double covering of curves to any numerical semigroup whose genus is large enough. As an application we strengthen the properties concerning Weierstrass weights in \cited{To}.
[ { "version": "v1", "created": "Tue, 19 Dec 1995 06:21:54 GMT" } ]
2008-02-03T00:00:00
[ [ "Torres", "Fernando", "" ] ]
alg-geom
\section{Introduction} Let $H$ be a numerical semigroup, that is, a subsemigroup of $(\mathbb N, +)$ whose complement is finite. Examples of such semigroups are the Weierstrass semigroups at non-singular points of algebraic curves. In this paper we deal with the following type of semigroups: \begin{definition}\label{def} Let $\gamma\ge 0$ an integer. $H$ is called $\gamma$-hyperelliptic if the following conditions hold: \begin{itemize} \item[($E_1$)] $H$ has $\gamma$ even elements in $[2,4\gamma]$. \item[($E_2$)] The $(\gamma+1)$th positive element of $H$ is $4\gamma+2$. \end{itemize} A 0-hyperelliptic semigroup is usually called hyperelliptic. \end{definition} The motivation for study of such semigroups comes from the study of Weierstrass semigroups at ramified points of double coverings of curves. Let $\pi: X\to \tilde X$ be a double covering of projective, irreducible, non-singular algebraic curves over an algebraically closed field $k$. Let $g$ and $\gamma$ be the genus of $X$ and $\tilde X$ respectively. Assume that there exists $P\in X$ which is ramified for $\pi$, and denote by $m_i$ the $i$th non-gap at $P$. Then the Weierstrass semigroup $H(P)$ at $P$ satisfies the following properties (cf. \cite{To}, \cite[Lemma 3.4]{To1}): \begin{itemize} \item[($P_1$)] $H(P)$ is $\gamma$-hyperelliptic, provided $g\ge 4\gamma+1$ if ${\rm char}(k)\neq 2$, and $g\ge 6\gamma-3$ otherwise. \item[($P_2$)] $m_{2\gamma+1}=6\gamma+2$, provided $g\ge 5\gamma+1$. \item[($P_3$)] $m_{\frac{g}{2}-\gamma-1}=g-2$ or $m_{\frac{g-1}{2}-\gamma}=g-1$, provided $g\ge 4\gamma+2$. \item[($P_4$)] The weight $w(P)$ of $H(P)$ satisfies $$ \binom{g-2\gamma}{2}\le w(P)< \binom{g-2\gamma+2}{2}. $$ \end{itemize} Conversely if $g$ is large enough and if any of the above properties is satisfied, then $X$ is a double covering of a curve of genus $\gamma$. Aposteriori the four above properties become equivalent whenever $g$ is large enough. The goal of this paper is to extend these results for any semigroup $H$ such that $g(H):= \#(\mathbb N\setminus H)$ is large enough. We remark that there exist semigroups of genus large enough that cannot be realized as Weierstrass semigroups (see \cite{Buch1}, \cite[Scholium 3.5]{To}). The key tool used to prove these equivalences is Theorem 1.10 in Freiman's book \cite{Fre} which have to do with addition of finite sets. From this theorem we deduce Corollary \ref{cor-cast} which can be considered as analogous to Castelnuovo's genus bound for curves in projective spaces (\cite{C}, \cite[p.116]{ACGH}, \cite[Corollary 2.8]{R}). Castelnuovo's result is the key tool to deal with Weierstrass semigroups. This Corollary can also be proved by means of properties of addition of residue classes (see Remark \ref{cauchy}). In \S2 we prove the equivalences $(P_1)\Leftrightarrow (P_2)\Leftrightarrow (P_3)$. The equivalence $(P_1)\Leftrightarrow (P_2)$ is proved under the hypothesis $g(H)\ge 6\gamma+4$, while $(P_1)\Leftrightarrow (P_3)$ is proved under $g(H)=6\gamma+5$ or $g(H)\ge 6\gamma+8$. In both cases the bounds on $g(H)$ are sharp (Remark \ref{sharp}). We mention that the cases $\gamma\in\{1,2\}$ of $(P_1)\Leftrightarrow (P_3)$ were fixed by Kato \cite[Lemmas 4,5,6,7]{K2}. In \S3 we deal with the equivalence $(P_1)\Leftrightarrow (P_4)$. To this purpose we determine bounds for the weight $w(H)$ of the semigroup $H$, which is defined by $$ w(H):= \sum_{i=1}^{g}(\ell_i -i), $$ where $g:=g(H)$ and $\mathbb N\setminus H = \{\ell_1,\ldots,\ell_g\}$. It is well known that $0\le w(H)\le \binom{g}{2}$; clearly $w(H)=0\Leftrightarrow H=\{g+i:i\in \mathbb Z^+\}$, and one can show that $w(H)=\binom{g}{2} \Leftrightarrow H$ is $\mathbb N$, or $g(H)\ge 1$ and $H$ is hyperelliptic (see e.g. \cite[Corollary III.5.7]{F-K}). Associated to $H$ we have the number \begin{equation}\label{even-gap} \rho=\rho(H):=\{\ell \in \mathbb N\setminus H: \ell\ {\rm even}\}. \end{equation} In \cite[Lemma 2.3]{To} it has been shown that $\rho(H)$ is the unique number $\gamma$ satisfying $(E_1)$ of Definition \ref{def}, and \begin{itemize} \item[($E_2'$)] $4\gamma+2 \in H$. \end{itemize} Thus we observe the following: \begin{lemma}\label{feto0} Let $H$ be a $\gamma$-hyperelliptic semigroup. Then $$ \rho(H)=\gamma. $$ \end{lemma} We also observe that if $g(H)\ge 1$, then $H$ is hyperelliptic if and only if $\rho(H)=0$. In general, $\rho(H)$ affects the values of $w(H)$. Let us assume that $\rho(H)\ge 1$ (hence $w(H)<\binom{g}{2}$); then we find $$ \binom{g-2\rho}{2}\le w(H)\le \left\{ \begin{array}{ll} \binom{g-2\rho}{2}+2\rho^2 & {\rm if\ } g\ge 6\rho+5 \\ \frac{g(g-1)}{3} & {\rm otherwise} \end{array} \right. $$ (see Lemmas \ref{bo-weight} and \ref{opt-weight}). These bounds strengthen results of Kato \cite[Thm.1]{K1} and Oliveira \cite[p.435]{Oliv} (see Remark \ref{oliv}). From this result we prove $(P_1)\Leftrightarrow (P_4)$ (Theorem \ref{char-weight1}) under the hypothesis \begin{equation}\label{bound-g} g(H)\ge \left\{ \begin{array}{ll} {\rm max}\{12\gamma-1,1\} & {\rm if\ } \gamma\in\{0,1,2\}, \\ 11\gamma+1 & {\rm if\ } \gamma \in\{3,5\}, \\ \frac{21(\gamma-4)+88}{2} & {\rm if\ }\gamma \in \{4,6\}, \\ \gamma^2+4\gamma+3 & {\rm if\ } \gamma\ge 7. \end{array} \right. \end{equation} The cases $\gamma \in \{1,2\}$ of that equivalence was fixed by Garcia (see \cite{G}). In this section we use ideas from Garcia's \cite[Proof of Lemma 8]{G} and Kato's \cite[p. 144]{K1}. In \S1 we recollet some arithmetical properties of numerical semigroups. We mainly remark the influence of $\rho(H)$ on $H$. It is a pleasure to thank Pablo Azcue and Gustavo T. de A. Moreira for discussions about \S2. \section{Preliminaries} Throughout this paper we use the following notation \begin{itemize} \itemsep=0.5pt \item semigroup:\quad numerical semigroup. \item Let $H$ be a semigroup. The {\it genus} of $H$ is the number $g(H):= \#(\mathbb N\setminus H)$, which throughout this article will be assumed bigger than 0. The positive elements of $H$ will be called the {\it non-gaps} of $H$, and those of $G(H):= \mathbb N\setminus H$ will be called the {\it gaps} of $H$. We denote by $m_i(H)$ the $i$th non-gap of $H$. If a semigroup is generated by $m,n,\ldots $ we denote $H=\langle m,n,\ldots \rangle$. \item $[x]$ stands for the integer part of $x\in \mathbb R$. \end{itemize} In this section we recall some arithmetical properties of semigroups. Let $H$ be a semigroup of genus $g$. Set $m_j:= m_j(H)$ for each $j$. If $m_1=2$ then $m_i=2i$ for $i=1,\ldots,g$. Let $m_i\ge 3$. By the semigroup property of $H$ the first $g$ non-gaps satisfy the following inequalities: \begin{equation}\label{prop-sem} m_i\ge 2i+1\ \ {\rm for}\ i=1,\ldots,g-2,\ \ m_{g-1}\ge 2g-2,\ \ m_g=2g \end{equation} (see \cite{Buch}, \cite[Thm.1.1]{Oliv}). \medskip Let $\rho$ be as in (\ref{even-gap}). From \cite[Lemma 2.3]{To1} we have that \begin{equation}\label{feto} \{4\rho+2i: i\in \mathbb N\} \subseteq H. \end{equation} {}From the definition of $\rho$, $H$ has $\rho$ odd non-gaps in $[1,2g-1]$. Let denote by $$ u_{\rho} <\ldots < u_1 $$ such non-gaps. \begin{lemma}\label{feto1} Let $H$ be a semigroup of genus $g$, and $\rho$ the number of even gaps of $H$. Then $$ 2g \ge 3\rho. $$ \end{lemma} \begin{proof} Let us assume that $g\le 2\rho -1$. From $u_1\le 2g-1$ we have $u_{2\rho-g+1}\le 4g-4\rho-1$. Let $\ell$ be the biggest even gap of $H$. Then $\ell \le 4g-4\rho$. For suppose that $\ell \ge 4g-4\rho+2$. Thus $\ell-u_{2\rho-g+1}\ge 3$, and then $H$ would has $g-\rho+1$ odd gaps, namely $1, \ell-u_{2\rho-g+1},\ldots, \ell-u_{\rho}$, a contradicition. Now since in $[2,4g-4\rho]$ there are $2g-2\rho$ even numbers such that $\rho$ of them are gaps, the lemma follows. \end{proof} Denote by $f_i:=f_i(H)$ the $i$th even non-gap of $H$. Hence by (\ref{feto}) we have \begin{equation}\label{gene} H=\langle f_1,\ldots, f_{\rho},4\rho+2,u_{\rho},\ldots,u_1\rangle. \end{equation} Observe that $f_{\rho}=4\rho$, and \begin{equation}\label{even} f_{g-\rho}=2g. \end{equation} By \cite[Lemma 2.1]{To1} and since $g\ge 1$ we have \begin{equation}\label{des-odd-1} u_{\rho} \ge\ {\rm max}\{2g-4\rho +1, 3\}. \end{equation} In particular, if $g\ge 4\rho$ we obtain \begin{equation}\label{first=even} m_1=f_1, \ldots, m_{\rho}=f_{\rho}. \end{equation} Note that (\ref{des-odd-1}) is only meanful for $g\ge 2\rho$. For $g\le2\rho-1$ we have: \begin{lemma}\label{feto2} Let $H$ be a semigroup of genus $g$, and $\rho$ the number of even gaps of $H$. If $g\le 2\rho-1$, then $$ u_{\rho}\ge 4\rho -2g+1. $$ \end{lemma} \begin{proof} {}From the proof of Lemma \ref{feto1} we have that $H$ has $2g-3\rho$ even non-gaps in $[2,4g-4\rho]$. Consider the following sequence of even non-gaps: $$ 2u_{\rho}<\ldots< u_{\rho}+u_{4\rho-2g}. $$ Since in this sequence we have $2g-3\rho+1$ even non-gaps, then $$ u_{\rho}+u_{4\rho-2g}\ge 4g-4\rho+2. $$ Now, since $u_{4\rho-2g}\le 6g-8\rho+1$ the proof follows. \end{proof} \section{$\gamma$-hyperelliptic semigroups} In this section we deal with properties $(P_1)$, $(P_2)$ and $(P_3)$ stated in \S0. For $i\in \mathbb Z^+$ set $$ d_i(H):= \gcd(m_1(H),\ldots,m_i(H)). $$ \begin{theorem}\label{char1} Let $\gamma \in \mathbb N$, $H$ a semigroup of genus $g \ge 6\gamma +4$ if $\gamma\ge 1$. Then the following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $\gamma$-hyperelliptic. \item[(ii)] $m_{2\gamma+1}(H)= 6\gamma +2$. \end{itemize} \end{theorem} \begin{theorem}\label{char2} Let $\gamma$ and $H$ be as in Theorem \ref{char1}, and assume that $g\ge 1$ if $\gamma=0$. Then the following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $t$-hyperelliptic for some $t\in \{0,\ldots,\gamma\}$. \item[(ii)] $m_{2\gamma+1}(H)\le 6\gamma+2$. \item[(iii)] $\rho(H)\le \gamma$. \end{itemize} \end{theorem} \begin{theorem}\label{char3} Let $\gamma \in \mathbb N$, $H$ a semigroup of genus $g=6\gamma+5$ or $g\ge 6\gamma+7$. Set $r:= \[x]-\gamma-1$. Then the following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $\gamma$-hyperelliptic. \item[(ii)] $m_r(H)=g-2$ if $g$ is even; $m_r(H)=g-1$ if $g$ is odd. \item[(iii)] $m_r(H)\le g-1 < m_{r+1}(H)$. \end{itemize} \end{theorem} \begin{theorem}\label{char4} Let $\gamma$, $H$ and $r$ be as in Theorem \ref{char3}. Then the following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $t$-hyperelliptic for some $t\in\{0,\ldots,\gamma\}$. \item[(ii)] $m_r(H)\le g-2$ if $g$ is even; $m_r(H)\le g-1$ if $g$ is odd. \item[(iii)] $m_r(H)\le g-1$. \item[(iv)] $\rho(H)\le \gamma$. \end{itemize} \end{theorem} To prove these results we need a particular case of the result below. For $K$ a subset of a group we set $2K:=\{a+b: a, b \in K\}$. \begin{lemma}[[Fre, Thm. 1.10 {]}]\label{thm-fre} Let $K=\{0<m_1<\ldots<m_i\} \subseteq \mathbb Z$ be such that $\gcd(m_1,\ldots,m_i)=1$. If $m_i\ge i+1+b$, where $b$ is an integer satisfying $0\le b<i-1$, then $$ \# 2K \ge 2i+2+b. $$ \hfill $\Box$ \end{lemma} \begin{corollary}\label{cor-cast} Let $H$ be a semigroup of genus $g$, and $i\in \mathbb Z^+$. If $$ d_i(H)= 1\qquad{\rm and}\qquad i\le g+1, $$ then we have $$ 2m_i(H) \ge m_{3i-1}(H). $$ \end{corollary} \begin{proof} Let $K:=\{0,m_1(H),\ldots,m_i(H)\}$. Then by (\ref{prop-sem}), we can apply Lemma \ref{thm-fre} to $K$ with $b=i-2$. \end{proof} \begin{remark}\label{rem-cast} Both the hypothesis $d_i(H)=1$ and $i\le g+1$ of the corollary above are necessaries. Moreover the conclusion of that result is sharp: \begin{itemize} \item[(i)] Let $i=g+h$, $h\ge 2$. Then $2m_{g+h}=m_{3i-h}$. \item[(ii)] Let $m_1=4$, $m_2=6$ and $m_3=8$. Then $d_3=2$ and $2m_3=m_7$. \item[(iii)] Let $m_1=5$, $m_2=10$, $m_3=15$, $m_4=18$, $m_5=20$. Then $2m_6=m_{14}$. \end{itemize} \end{remark} \begin{remark}\label{cauchy} (i) The Corollary above can also be proved by using results on the addition of residue classes: let $H$ and $i$ be as in \ref{cor-cast}; assume further that $2\le i\le g-2$ (the remaining cases are easy to prove), and consider $\tilde K:=\{m_1,\ldots,m_i\}\subseteq \mathbb Z_{m_i}$ (i.e. a subset of the integers modulus $m_i$). Let $N:= \# 2\tilde K$. Then it is easy to see that $$ 2m_i\ge m_{i+N}. $$ Consequently we have a proof of the above Corollary provided $N\ge 2i-1\ (*)$. Since $m_i\ge 2i+1$ (see (\ref{prop-sem})), we get $(*)$ provided $m_i$ prime (Cauchy \cite{Dav1}, Davenport \cite{Dav}, \cite[Corollary 1.2.3]{M}), or provided $\gcd(m_j,m_i)=1$ for $j=1,\ldots,i-1$ (Chowla \cite[Satz 114]{Lan}, \cite[Corollary 1.2.4]{M}). In general by using the hypothesis $d_i(H)=1$ we can reduce the proof of the Corollary to the case $\gcd(m_{i-1},m_i)=1$. Then we apply Pillai's \cite[Thm 1]{Pi} generalization of Davenport and Chowla results (or Mann's result \cite[Corollary 1.2.2]{M}). (ii) Let $H$ and $i$ be as above and assume that $2m_i=m_{3i-1}$. Then from (i) we have $N=\# 2\tilde K=2i-1$. Thus by Kemperman \cite[Thm 2.1]{Kem} (or by \cite[Thm. 1.11]{Fre}) $2\tilde K$ satisfies one of the following conditions: (1) there exist $m, d\in \mathbb Z_{m_i}$, such that $2\tilde K=\{m+dj:j=0,1,\ldots,N-1\}$, or (2) there exists a subgroup $F$ of $\mathbb Z_{m_i}$ of order $\ge 2$, such that $2\tilde K$ is the disjoint union of a non-empty set $C$ satisfying $C+F=C$, and a set $C'$ satisfying $C'\subseteq c+F$ for some $c\in C'$. For instance example (iii) of \ref{rem-cast} satisfies property (2). \end{remark} Set $m_j:= m_j(H)$ for each $j$. \begin{proof} {\it (Theorem \ref{char1}).} By definition $H$ is hyperelliptic if and only if $m_1=2$. So let us assume that $\gamma\ge 1$. (i) $\Rightarrow$ (ii): From Lemma \ref{feto0} and (\ref{des-odd-1}) we find that $u_{\gamma}\ge 6\gamma+3$ if $g\ge 5\gamma+1$. Then (ii) follows from (\ref{first=even}) and (\ref{feto}). (ii) $\Rightarrow$ (i): We claim that $d_{2\gamma+1}(H)=2$. For suppose that $d_{2\gamma+1}(H) \ge 3$. Then $6\gamma+2= m_{2\gamma+1} \ge m_1+ 6\gamma$ and so $m_1\le 2$, a contradiction. Hence $d_{2\gamma+1}(H)\le 2$. Now suppose that $d_{2\gamma+1}(H) =1$. Then Corollary \ref{cor-cast} implies $$ 2(6\gamma+2) = 2m_{2\gamma+1} \ge m_{6\gamma +2}. $$ But, since $g-2 \ge 6\gamma +2$, by (\ref{prop-sem}) we would have $$ m_{6\gamma +2} \ge 2(6\gamma+2) +1 $$ which leads to a contradiction. This shows that $d_{2\gamma+1}(H)=2$. Now since $m_{2\gamma+1}=6\gamma+2$ we have that $m_\gamma \le 4\gamma$. Moreover, there exist $\gamma$ even gaps of $H$ in $[2, 6\gamma+2]$. Let $\ell$ be an even gap of $H$. The proof follows from the following claim: \begin{claim*} $\ell <m_\gamma$. \end{claim*} \begin{proof} {\it (Claim).} Suppose that there exists an even gap $\ell$ such that $\ell>m_\gamma$. Take the smallest $\ell$ with such a property; then the following $\gamma$ even gaps: $\ell-m_\gamma< \ldots, \ell -m_1$ belong to $[2,m_\gamma]$. Thus, since $m_1>2$, we must have $\ell-m_\gamma=2$. This implies that $H$ has $\gamma+1$ even non-gaps in $[2,6\gamma+2]$, namely $\ell-m_\gamma,\ldots,\ell-m_1,\ell$, a contradiction. \end{proof} This finish the proof of Theorem \ref{char1}. \end{proof} \begin{proof} {\it (Theorem \ref{char2}).} The case $\gamma=0$ is trivial; so let assume $\gamma\ge 1$. (i) $\Rightarrow$ (ii): Since $g\ge 5\gamma+1\ge 5t+1$ by Theorem \ref{char1} we have $m_{2t+1}=6t+2$. Thus (ii) follows from Lemma \ref{feto0} and (\ref{feto}). (ii) $\Rightarrow$ (iii): From the proof of (ii) $\Rightarrow$ (i) of Theorem \ref{char1} it follows that $d_{2\gamma+1}(H)=2$. Consequently by using the hypothesis on $m_{2\gamma+1}$, and again from the mentioned proof we have that all the gaps of $H$ belong to $[2,m_\gamma]$. Since $m_\gamma\le 4\gamma$ then we have $\rho(H)\le \gamma$ (iii) $\Rightarrow$ (i) Since $g\ge 4\gamma+1\ge 4\rho(H)+1$, the proof follows from $(E_1)$ and $(E_2')$ (see \S0). \end{proof} \begin{proof} {\it (Theorem \ref{char3}).} (i) $\Rightarrow$ (ii): Similar to the proof of (i) $\Rightarrow$ (ii) of Theorem \ref{char1} (here we need $g\ge 4\gamma+3$ (resp. $g\ge 4\gamma+4$) if $g$ is odd (resp. even)). \smallskip Before proving the other implications we remark that $g\le 3r-1$: in fact, if $g\ge 3r$ we would have $g\le 6\gamma+6$ (resp. $g\ge 6\gamma+3$) provided $g$ even (resp. odd) - a contradiction. \smallskip (ii) $\Rightarrow$ (iii): Let $g$ even and suppose that $m_{r+1}=g-1$. Then by Corollary \ref{cor-cast} we would have $2g-2=2m_{r+1}\ge m_{3r+2}$ and hence $g-1\ge 3r+2$. This contradicts the previous remark. (iii) $\Rightarrow$ (i): We claim that $d_r(H)= 2$. Suppose that $d_r(H)\ge 3$. Then we would have $g-1\ge m_r \ge m_1+3(r-1) \ge 3r-1$, which contradicts the previous remark. Now suppose that $d_r(H)=1$. Then by Corollary \ref{cor-cast} we would have $$ 2g-2\ge m_r \ge m_{3r-1}, $$ which again contradicts the previous remark. Thus the number of even gaps of $H$ in $[2,g-1]$ is $\gamma$, and $m_\gamma\le 4\gamma$. Let $\ell$ be an even gap of $H$. As in the proof of the Claim in Theorem \ref{char1} here we also have that $\ell<m_\gamma$. Now the proof follows. \end{proof} \begin{proof} {\it (Theorem \ref{char4}).} (i) $\Rightarrow$ (ii): By Theorem \ref{char3} and since $t\le \gamma$ we have $g-2 =m_{g/2-t-1}$ or $g-1=m_{(g-1)/2-t}$. This implies (ii). The implication (ii) $\Rightarrow$ (iii) is trivial. (iii) $\Rightarrow$ (iv): As in the proof of Theorem \ref{char3} we obtain $d_r(H)=2$. Then the number of even gaps of $H$ in $[2,g-1]$ is at most $\gamma$. We claim that all the even gaps of $H$ belong to that interval. For suppose there exists an even gap $\ell>g-1$. Choose $\ell$ the smallest one and consider the even gaps $\ell-m_1<\ldots<\ell-m_r\le g-1$. Then $r\le \gamma$ which yields to $g\le 4\gamma+2$, a contradiction. Consequently $\rho(H)\le \gamma$. The implication (iv) $\Rightarrow$ (i) follows from Theorem \ref{char2}. \end{proof} \begin{remark}\label{sharp} The hypothesis on the genus in the above theorems is sharp. To see this let $\gamma\ge 0$ an integer, and let $X$ be the curve defined by the equation $$ y^4=\mathop{\prod}\limits^{I}_{j=1}(x-a_j), $$ where the $a_j's$ are pairwise distinct elements of a field $k$, $I=4\gamma+3$ if $\gamma$ is odd; $I=4\gamma+5$ otherwise. Let $P$ be the unique point over $x=\infty$. Then $H(P)=\langle 4, I\rangle$ and so $g(H(P))=6\gamma+3$ (resp. $6\gamma+6$), $m_{2\gamma+1}(H(P))= 6\gamma+2$ (resp. $m_{2\gamma+2}(H(P))=6\gamma+5$), and $\rho(H(P))=2\gamma+1$ (resp. $\rho(H(P))=2\gamma+2$) provided $\gamma$ odd (resp. $\gamma$ even). \end{remark} \section{Weight of semigroups} \subsection{Bounding the weight.} Let $H$ be a semigroup of genus $g$. Set $m_j=m_j(H)$ for each $j$ and $\rho=\rho(H)$ (see (\ref{even-gap}). Due to $m_g=2g$ (see (\ref{prop-sem})), the weight $w(H)$ of $H$ can be computed by \begin{equation}\label{weight} w(H)=\frac{3g^2+g}{2}-\mathop{\sum}\limits^{g}_{j=1} m_j. \end{equation} Consequently the problem of bounding $w(H)$ is equivalent to the problem of bounding $$ S(H):= \sum_{j=1}^{g} m_g. $$ If $\rho=0$, then we have $m_i=2i$ for each $i=1,\ldots,g$. In particular we have $w(H)=\binom{g}{2}$. Let $\rho\ge 1$ (or equivalently $f_1\ge 4$). Then by (\ref{gene}) we have \begin{equation}\label{weight1} S(H)=\sum_{f\in \tilde H,\ f\le g} 2f + \sum_{i=1}^{\rho} u_i, \end{equation} where $$ \tilde H:= \{f/2 : f\in H,\ f\ {\rm even}\}. $$ \begin{lemma'}\label{bounds} With the notation of \S1 we have: \begin{itemize} \item[(i)] If $f_1=4$, then $f_i=4i$ for $i=1,\ldots,\rho$. \item[(ii)] If $f_1\ge 6$, then $$ f_i\ge 4i+2\ \ {\rm for}\ i=1, \ldots,\rho-2,\ \ f_{\rho-1}\ge 4\rho-4,\ \ f_{\rho}=4\rho. $$ \item[(iii)] $f_i\le 2\rho+2i$ for each $i$. \item[(iv)] $2g-4j+1 \le u_j \le 2g-2j+1$, for $j=1,\ldots,\rho$. \end{itemize} \end{lemma'} \begin{proof} By (\ref{feto}), we have $$ \tilde H= \{\frac{f_1}{2},\ldots, \frac{f_\rho}{2}\}\cup \{4\rho +i: i\in \mathbb N \}. $$ Thus $\tilde H$ is a semigroup of genus $\rho$. Then (i) is due to the fact that $f_1/2=2$ and (ii) follows from (\ref{prop-sem}). Statement (iii) follows from (\ref{even}). \smallskip (iv) The upper bound follows from $u_1\le 2g-1$. To prove the lower bound we procced by induction on $i$. The case $i=\rho$ follows from (\ref{des-odd-1}). Suppose that $u_i\ge 2g-4i+1$ but $u_{i-1} < 2g-4(i-1)+1$, for $1<i\le \rho$. Then $u_i=2g-4i+1$, $u_{i-1}= 2g-4i+3$, and there exists an odd gap $\ell$ of $H$ such that $\ell>u_{i-1}$. Take the smallest $\ell$ with such a property. Set $I:=[\ell-u_{i-1}, \ell-u_\rho]\subseteq [2,4\rho-2]$ and let $t$ be the number of non-gaps of $H$ belonging to $I$. By the election of $\ell$ we have that $\ell-u_{i-1}<f_1$. Now, since $\ell-u_j \in I$ for $j=i-1,\ldots,\rho$ we also have that $$ \frac{u_{i-1}-u_\rho}{2}+1 \ge t + \rho -(i-1)+1. $$ Thus $u_\rho \le 2g-2\rho-2i-2t+1$. Now, since $u_\rho+f_{t+1}>u_{i-1}$ and since by statement (iii) $f_{t+i-1}\le 2\rho + 2t +2i-2$, we have that the odd non-gaps $u_\rho +f_{t+1}, \ldots, u_\rho +f_{t+i-1}$ belong to $[\ell+2,2g-1]$. This is a contradiction because $H$ would have $(\rho -i+2)+(i-1) = \rho +1$ odd non-gaps. \end{proof} \begin{lemma'}\label{bo-weight} Let $H$ be a semigroup of genus $g$. With notation as in \S1, we have \begin{itemize} \item[(i)] $w(H)\ge \binom{g-2\rho}{2}$. Equality holds if and only if $f_1=2\rho+2$ and $u_{\rho}=2g-2\rho+1$. \item[(ii)] If $g\ge 2\rho$, then $w(H)\le \binom{g-2\rho}{ 2}+2\rho^2 $. Equality holds if and only if $H=\langle 4, 4\rho,2g-4\rho+1\rangle$. \item[(iii)] If $g\le 2\rho-1$, then $w(H)\le \binom{g+2\rho}{ 2}-4g-6\rho^2+8\rho $. \end{itemize} \end{lemma'} \begin{proof} (i) By (\ref{weight}) we have to show that $$ S(H) \le g^2+(2\rho+1)g-2\rho^2-\rho, $$ and that the equality holds if and only if $f_1=2\rho+2$ and $u_{\rho}=2g-2\rho+1$. Both the above statements follow from Lemma \ref{bounds} (i), (iv). \smallskip (ii) Here we have to show that \begin{equation*} S(H)\ge g^2 + (2\rho+1)g-4\rho^2-\rho,\tag{$\dag$} \end{equation*} and that equality holds if and only $H=\langle 4,4\rho+2,2g-4\rho+1\rangle$. Since $g\ge 2\rho$ by (\ref{feto}) we obtain \begin{equation}\label{sum1} S(H)=\sum_{i=1}^{\rho}(f_i(H)+u_i(H))+ g^2+g-4\rho^2-2\rho. \end{equation} Thus we obtain $(\dag)$ by means of Lemma \ref{bounds} (ii), (iii) and (iv). Moreover the equality in $(\dag)$ holds if and only if $\sum_{i=1}^{\rho}(f_i+u_i)=2\rho g +\rho$. Then the second part of (ii) also follows from the above mentioned results. \smallskip (iii) In this case, due to the proof of Lemma \ref{feto1}, instead of equation (\ref{sum1}) we have \begin{equation}\label{sum2} S(H) = \sum_{i=1}^{2g-3\rho}f_i + \sum_{i=2g-3\rho+1}^{g-\rho}(2i+2\rho) + \sum_{i=1}^{\rho}u_i. \end{equation} We will see in the next remark that in this case we have $f_1\ge 6$. Thus by using Lemmas \ref{feto2} and \ref{bounds} (iii), (iv) we obtain $$ S(H)\ge 4\rho^2-(2g+7)\rho +g^2 +5g, $$ from where it follows the proof. \end{proof} \begin{remark'}\label{remark3} (i) If $f_1=4$, then $g\ge 2\rho$. This follows from the fact the biggest even gap of $H$ is $4(\rho-1)+2$. Moreover, one can determinate $u_{\rho},\ldots,u_1$ as follows: let $J\in \mathbb N$ satisfying the inequalities below $$ {\rm max}\{1, \frac{3\rho+2-g}{2}\}\le J\le {\rm min}\{\rho+1,\frac{g-\rho+3}{2}\}, $$ provided $g$ even, otherwise replace $J$ by $\rho-J+2$; then \begin{eqnarray*} \{u_{\rho},\ldots,u_1\} & = & \{ 2g-4\rho+4J-7+4i: i=1,\ldots,\rho-J+1\}\\ & &\mbox{}\cup\{2g-4J+3+4i: i=1,\ldots,J-1\}. \end{eqnarray*} (see \cite[\S3]{Ko}, \cite[Remarks 2.5]{To}). Consequently from (\ref{sum1}) and (\ref{weight}) we obtain $$ w(H)=\binom{g-2\rho}{2}+ 2\rho^2+4\rho+6+4J^2-(4\rho+10)J. $$ In particular we have $$ \binom{g-2\rho}{2}+\rho^2-\rho\le w(H)\le \binom{g-2\rho}{2}+2\rho^2. $$ Let $C$ be an integer such that $0\le 2C\le \rho^2+\rho$. Then $w(H)=\binom{g-2\rho}{2}+\rho^2-\rho +2C$ if and only if $4+32C$ is a square. The lower bound is attained if and only if $H=\langle 4,4\rho+2, 2g-2\rho+1,2g-2\rho+3\rangle$. \smallskip \noindent (ii) Let $u_{\rho}=3$. Them from (\ref{des-odd-1}) and Lemma \ref{feto2} we find that $g\in \{2\rho-1,2\rho,2\rho+1\}$. Moreover, in this case one can also obtain a explicit formula for $w(H)$ (\cite[Lemma 6]{K1}). Let $g\equiv r \pmod{3}$, $r=0,1,2$ and let $s$ be an integer such that $0\le s\le (g-r)/3$. If $r=0,1$ (resp. $r=2$), then \begin{align*} w(H) & =\frac{g(g-1)}{3}+3s^2-gs-s\le \frac{g(g-1)}{3} \\ \intertext{resp.} w(H) & =\frac{g(g-2)}{3}+3s^2-gs+s\le \frac{g(g-2)}{3}. \end{align*} If $r=0,1$ (resp. $r=2$), equality occurs if and only if $H=\langle 3, g+1\rangle$ (resp. $H=\langle 3, g+2,2g+1\rangle$). \end{remark'} Let $g\le 2\rho-1$. The way how we bound from below equation (\ref{sum2}) is far away from being sharp. We do not know an analogous to the lower bound of Lemma \ref{bounds} (iv) in this case. However, for certain range of $g$ the bounds in \ref{remark3} (ii) are the best possible: \begin{lemma'}\label{opt-weight} Let $H$ be a semigroup of genus $g\ge 11$, $r\in \{1,2,3,4,5,6\}$ such that $g\equiv r \pmod{6}$. Let $\rho$ be the number of even gaps of $H$. If $$ \rho>\left\{ \begin{array}{ll} \frac{g-5}{6} & {\rm if\ } r=5 \\ \frac{g-r}{6}-1 & {\rm if\ } r\neq 5, \end{array} \right. $$ then $$ w(H)\le \left\{ \begin{array}{ll} \frac{g(g-2)}{3} & {\rm if\ } r= 2,5 \\ \frac{g(g-1)}{3} & {\rm if\ } r=1,3,4,6. \end{array} \right. $$ If $r=2,5$ (resp. $r\not\in\{2,5\}$) equality above holds if and only if $H=\langle 3,g+2,2g+1\rangle$ (resp. $H=\langle 3,g+1\rangle$). \end{lemma'} \begin{proof} We assume $g\equiv 5 \pmod{6}$; the other cases can be proven in a similar way. By Remark \ref{remark3} (ii) we can assume $u_1>3$, and then by (\ref{weight}) we have to prove that \begin{equation*} S(H) > \frac{7g^2+7g}{6}.\tag{$*$} \end{equation*} Now, since $\rho>(g-5)/6$, by Theorem \ref{char4} and Lemma \ref{feto0} we must have $$ m_{\frac{g+1}{3}}=m_{\frac{g-1}{2}-\frac{g-5}{6}}\ge g. $$ (A) Let $S':= \sum_{i} m_i$, $(g+1)/3\le i \le g$:\quad Define $$ F:= \{ i\in \mathbb N: \frac{g+1}{3}\le i\le g,\ m_i\le 2i+\frac{g-5}{3}\}, $$ and let $f:= {\rm min}(F)$. Then $f\ge (g+4)/3$, $m_f=2f+\frac{g-5}{3}, m_{f-1}= 2f+\frac{g-8}{3}$. Thus for $g\ge i\ge f$, $d_i=1$ and hence by Corollary \ref{cor-cast}, $2m_i\ge m_{3i-1}=g+3i-1$. In particular, $f\ge (g+7)/3$. Now we bound $S'$ in three steps: \smallskip \noindent Step (i). $(g+1)/3\le i\le f-1$: By definiton of $f$ we have that $m_i\ge 2i+\frac{g-2}{3}$ and hence \begin{equation}\label{aux0} \sum_{i} m_i \ge f^2+\frac{g-5}{3}f -\frac{2g^2-2g-4}{9}. \end{equation} \noindent Step (ii). $f\le i\le (6f-g-7)/3$: Here we have that $m_i\ge m_f+i-f=i+f+\frac{g-5}{3}$. Hence $$ \sum_{i} m_i \ge \frac{5}{2}f^2-\frac{4g+37}{6}f -\frac{g^2-13g-68}{18}. $$ \noindent Step (iii). $(6f-g-4)/3\le i\le g$: Here we have $m_i+m_{i+1}\ge g+3i+1$ for $i$ odd, $6f-g-4\le i\le g-2$. Since $m_g=2g$ then we have $$ \sum_{i} m_i \ge -3f^2+6f+\frac{4g^2+2g-8}{3}. $$ (B) Let $S'':= \sum_{i} m_i$, $1\le i\le (g-2)/3$:\quad By Theorem \ref{char2} and Lemma \ref{feto0} we have that $m_i\ge 3i$ for $i$ odd, $i=3,\ldots, (g-2)/3$. First we notice that for $i$ odd and $3\le i\le (g-8)/3$ we must have $m_{i+1}\ge 3i+3$. Otherwise we would have $d_{i+1}=1$ and hence by Corollary \ref{cor-cast} and (\ref{prop-sem}) we would have $2m_{i+1}\ge m_{3i+2}\ge 6i+5$, a contradiction. \begin{claim*} Let $i$ odd and $3\le i\le (g-8)/3$. If $m_i=3i$ or $m_{i+1}=3i+3$, then $m_1=3$. \end{claim*} \begin{proof} {\it (Claim).} It is enough to show that $d_i=3$ or $d_{i+1}=3$. Suppose that $m_i=3i$. Since $i$ is odd, $d_i$ is one or three. Suppose $d_i=1$. Then by Corollary \ref{cor-cast} we have $6i=2m_i\ge m_{3i-1}$ and hence $6i=2m_i=m_{3i-1}$. Let $\ell \in G(H)$. Then $\ell \ge m_{3i-1}+3$. In fact if $\ell>m_{3i-1}+3$, by choosing the smallest $\ell$ with such a property we would have $3i+2$ gaps in $[1, 6i]$ namely, $1,2,3,\ell-m_{3i-1},\ldots,\ell-m_1$, a contradiction. Then it follows that $g\le 3i+1+3=3i+4$ or $g+2\le 3i+4$. \smallskip Now suppose that $m_{i+1}=3i+3$; as in the previous proof here we also have that $d_{i+1}>1$. Suppose that $d_{i+1}=2$. Then $m_1>3$ and hence $m_i=3i+1$. Since we know that $m_{i+2}\ge 3i+6$, then the even number $\ell=3i+5$ is a gap of $H$. Then we would fine $2i+2$ even numbers in $[2,3i+3]$, namely $m_1,\ldots,m_{i+1}$, and $\ell-m_{i+1},\ell-m_1$, a contradiction. Hence $d_{i+1}=3$ and then $m_1=3$. \end{proof} Then, since we assume $u_1>3$, we have $m_i+m_{i+1}\ge 6i+5$ for $i$ odd $3\le i\le (g-8)/3$, $m_{\frac{g-2}{3}}\ge g-2$, and so \begin{equation}\label{aux} \begin{split} \sum_{i=1}^{(g-2)/3} m_i & \ge \sum_{j=1}^{(g-11)/6} (12j+10) +m_1+m_2+m_{\frac{g-2}{3}} \\ & \ge \frac{g^2+g-78}{6} + m_1+m_2.\\ \end{split} \end{equation} Summing up (i), (ii), (iii) and (B) we get $$ S(H)\ge \frac{3f^2-(2g+11)f}{6}+\frac{22g^2+32g-206}{18}+m_1+m_2. $$ The function $\Gamma(x):= 3x^2-(2g+11)x$ attains its minimum for $x=(2g+11)/6<(g+7)/3\le f$. Suppose that $f\ge (g+13)/3$. Then we find $$ S(H)\ge \frac{7g^2+7g-60}{6}+m_1+m_2. $$ We claim that $m_1+m_2>11$. Otherwise we would have $m_3=m_1+m_2=10$ which is impossible. From the claim we get $(*)$. In all the computations below we use the fact that $2g\le (m-1)(n-1)$ whenever $m,n \in H$ with $\gcd(m,n)=1$ (see e.g. Jenkins \cite{J}). Now suppose that $f=(g+10)/3$. Here we find $$ S(H)\ge \frac{7g^2+7g-72}{6}+m_1+m_2. $$ Suppose that $m_1+m_2\le 12$ (otherwise the above computation imply $(*)$.). If $g>11$, then $m_4\ge 13$ and so $m_3=m_1+m_2\in \{9,11,12\}$. If $m_1+m_2=9$, then $g\le 6$; if $m_1+m_2=11$ then $g\le 10$; if $m_1+m_2=12$ then $g\le 11$ or $m_1=4$, $m_2=8$. Let $s$ denote the first odd non-gap of $H$. Then $2g\le 3(s-1)$ and so $s>(2g+2)/3$. In the interval $[4,(2g+2)/3]$ does not exist $h\in H$ such that $h\equiv 2 \pmod{4}$: In fact if such a $h$ exists then we would have $4\rho+2\le (2g-4)/3$ or $ \rho\le (g-5)/6$. Consequently $m_3=12,\ldots,m_{(g+1)/6}=(2g+2)/3$. Thus we can improve the computation in (\ref{aux}) by summing it up $\sum_{i=1}^{j}(4i+1)$, where $j=(g-5)/12$ or $j=(g-11)/12$. Then we get $(*)$. If $g=11$, the first seven non-gaps are $\{4,8,10,12,14,15,16\}$ or $\{5,7,10,12,14,15,16\}$. In both cases the computation in (\ref{aux0}) increases at least by one, and so we obtain $(*)$. Finally let $f=(g+7)/3$. Here we find $$ S(H)\ge \frac{7g^2+7g-78}{6}+m_1+m_2, $$ and we have to analize the cases $m_1+m_2=\{9, 11, 12, 13\}$. This can be done as in the previous case. This finish the proof of Lemma \ref{opt-weight}. \end{proof} \subsection{The equivalence $(P_1)\Leftrightarrow (P_4)$.} We are going to characterize $\gamma$-hyperelliptic semigroups by means of weights of semigroups. We begin with the following result, which has been proved by Garcia for $\gamma\in\{1,2\}$ \cite[Lemmas 8 and 10]{G}. \begin{theorem'}\label{char-weight} Let $\gamma\in \mathbb N$ and $H$ a semigroup whose genus $g$ satisfies (\ref{bound-g}). Then the following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $t$-hyperelliptic for some $t\in \{0,\ldots,\gamma\}$. \item[(ii)] $w(H)\ge \binom{g-2\gamma}{2}$. \end{itemize} \end{theorem'} \begin{theorem'}\label{char-weight1} Let $\gamma$, $H$ and $g$ be as in Theorem \ref{char-weight}. The following statements are equivalent: \begin{itemize} \item[(i)] $H$ is $\gamma$-hyperelliptic. \item[(ii)] $\binom{g-2\gamma}{2}\le w(H)\le \binom{g-2\gamma}{2}+2\gamma^2$. \item[(iii)] $\binom{g-2\gamma}{2}\le w(H)<\binom{g-2\gamma+2}{2}$. \end{itemize} \end{theorem'} \begin{proof} {\it (Theorem \ref{char-weight}).} (i) $\Rightarrow$ (ii): By Lemma \ref{feto0} and Lemma \ref{bo-weight} (i) we have $w(H)\ge \binom{g-2t}{2}$. This implies (ii). (ii) $\Rightarrow$ (i): Suppose that $H$ is not $t$-hyperelliptic for any $t\in \{0,\ldots\gamma\}$. We are going to prove that $w(H)<\binom{g-2\gamma}{2}$, which by (\ref{weight}) is equivalent to prove that: \medskip $(*)$\hfill $\sum_{i=1}^{g} m_i > g^2+(2\gamma+1)g-2\gamma^2-\gamma.$\hfill \medskip \noindent We notice that by Lemma \ref{feto0} we must have $\rho\ge \gamma+1$. \smallskip \noindent Case 1: $g$ satisfies the hypothesis of Lemma \ref{opt-weight}.\quad From that lemma we have $ S(H)\ge (7g^2+5g)/3$ and then we get $(*)$ provided $$ g>\bar\gamma:= 12\gamma+1+\sqrt{96\gamma^2+1}. $$ We notice that $\gamma^2+4\gamma+3\ge \bar\gamma$ if $\gamma\ge 7$. For $\gamma=1,4,6$ we need respectively $g>11$, $g>44$ and $g>65$. By noticing that 11, 44 and 65 are congruent to 2 modulus 3, we can use $g=11$, $g=44$ and $g=65$ because in these cases $S(H)\ge (7g^2+7g)/3$. For the other values of $\gamma$ we obtain the bounds of (\ref{bound-g}). \smallskip \noindent Case 2: $g$ does not satisfy the hypothesis of Lemma \ref{opt-weight}.\quad Here we have $g\ge 6\rho+5$. From (\ref{sum1}) and Lemma \ref{bounds} we have $S(H)\ge g^2+(2\rho+1)g-4\rho^2-\rho$. The function $\Gamma(\rho):= (2g-1)\rho-4\rho^2$ satisfies $$ \Gamma(\rho)\ge \Gamma(\gamma+1)=(2\gamma+2)g-4\gamma^2-9\gamma-5, $$ for $\gamma+1 \le \rho\le [(2g-1)/4]-\gamma-1$. Thus we obtain condition $(*)$ provided $g\ge \gamma^2 + 4\gamma+3$. \end{proof} \begin{remark'}\label{oliv} Let $H$ be a semigroup of genus $g$, $r$ the number defined in Lemma \ref{opt-weight}. Put $c:= (g-5)/6$ if $r=5$, and $c:= (g-r)/6-1$ otherwise. {}From the proof of Case 2 of the above result we see that $S(H)\ge g^2+3g-5$ whenever $1\le \rho(H)\le (g-3)/2$. Hence this result and Lemma \ref{opt-weight} imply $$ w(H)\le\left\{ \begin{array}{ll} (g^2-5g+10/2 & {\rm if\ } \rho(H)\le c\\ {\rm min}\{(g^2-5g+10)/2, (g-1)g/3\} & {\rm if\ } c<\rho(H)\le (g-3)/2\\ (g-1)g/3 & {\rm if\ } \rho(H)>(g-3)/2. \end{array} \right. $$ \end{remark'} \begin{proof} {\it (Theorem \ref{char-weight1}).} (i) $\Rightarrow$ (ii) follows from Lemma \ref{bo-weight}. (ii) $\Rightarrow$ (iii) follows from the hypothesis on $g$. (iii) $\Rightarrow$ (i): By Theorem \ref{char-weight} we have that $H(P)$ is $t$-hyperelliptic for some $t\in \{0,\ldots,\gamma\}$. Then by Lemma \ref{bo-weight} and hypothesis we have $$ \binom{g-2\gamma+2}{2}>w(H)\ge \binom{g-2t}{2}, $$ from where it follows that $t=\gamma$. \end{proof} \begin{remark'}\label{sharp-weight} The hypothesis on $g$ in the above two theorems is sharp: \smallskip (i) Let $\gamma\ge 7$ and considerer $H:=\langle 4, 4(\gamma+1), 2g-4(\gamma+1)+1\rangle$ where $g$ is an integer satisfying ${\rm max}\{4\gamma+4, \frac{\gamma^2+6\gamma-3}{2}\}<g\le \gamma^2+4\gamma+2$. Then $H$ has genus $g$ and $\rho(H)=\gamma+1$. In particular $H$ is not $\gamma$-hyperelliptic. By Lemma \ref{bo-weight} (ii) we have $w(H)=\binom{g-2(\gamma+1)}{2}+2(\gamma+1)^2$. Now it is easy to check that $w(H)$ satisfies Theorem \ref{char-weight} (ii) and Theorem \ref{char-weight1} (iii). \smallskip (ii) Let $\gamma\le 6$ and consider $H=\langle 3,g+1\rangle$, where $g=10, 22, 33, 43, 55, 64$ if $\gamma=1,2,3,4,5,6$ respectively. $H$ has genus $g$ and it can be easily checked that $H$ is not $\gamma$-hyperelliptic by means of inequality (\ref{des-odd-1}) and Lemma \ref{feto1}. Moreover $w(H)=g(g-1)/3$ (see Remark \ref{remark3} (ii)). Now it is easy to check that $w(H)$ satisfies Theorem \ref{char-weight} (ii) and Theorem \ref{char-weight} (iii). \smallskip (iii) The semigroups considered in (i) and (ii) are also Weierstrass semigroups (see Komeda \cite{Ko}, Maclachlan \cite[Thm. 4]{Mac}). \end{remark'} \subsection{Weierstrass weights} In this section we apply Theorem \ref{char-weight1} in order to characterize double coverings of curves by means of Weierstrass weights. Specifically we strengthen \cite[Theorem B]{To} and hence all its corollaries. The basic references for the discussion below are Farkas-Kra \cite[III.5]{F-K} and St\"ohr-Voloch \cite[\S1]{S-V}. Let $X$ be a non-singular, irreducible and projective algebraic curve of genus $g$ over an algebraically closed field $k$ of characteristic $p$. Let $\pi:X\to \mathbb P^{g-1}$ be the morphism induced by the canonical linear system on $X$. To any $P\in X$ we can associate the sequence $j_i(P)$ ($i=0,\ldots,g-1$) of intersection multiplicities at $\pi(P)$ of $\pi(X)$ with hyperplanes of $\mathbb P^{g-1}$. This sequence is the same for all but finitely many points (the so called Weierstrass points of $X$). These points are supported by a divisor $\mathcal{W}$ in such a way that the Weierstrass weight at $P$, $v_P(\mathcal{W})$, satisfies $$ v_P({\mathcal{W}})\ge w(P):= \sum_{i=1}^{g-1}(j_i(P)-\epsilon_i), $$ where $0=\epsilon_0<\ldots<\epsilon_{g-1}$ is the sequence at a generic point. One has $j_i(P)\ge \epsilon_i$ for each $i$, and from the Riemann-Roch theorem follows that $G(P):=\{j_i(P)+1:i=0,\ldots,g-1\}$ is the set of gaps of a semigroup $H(P)$ of genus $g$ (the so called Weierstrass semigroup at $P$). $X$ is called {\it classical} if $\epsilon_i=i$ for each $i$ (e.g. if $p=0$ or $p>2g-2$). In this case we have $v_P({\mathcal{W}})=w_P(R)$ for each $P$, and the number $w(P)$ is just the weight of the semigroup $H(P)$ defined in \S0. The following result strengthen \cite[Thm.B]{To}. The proof follows from \cite[Thm.A]{To}, \cite[Thm.A]{To1}, and Theorem \ref{char-weight1}. \begin{theorem'} Let $X$ be a classical curve, and assume that $g$ satisfies (\ref{bound-g}). Then the following statements are equivalent: \begin{itemize} \item[(i)] $X$ is a double covering of a curve of genus $\gamma$. \item[(ii)] There exists $P\in X$ such that $$ \binom{g-2\gamma}{2}\le w(P)\le \binom{g-2\gamma}{2}+2\gamma^2. $$ \item[(iii)] There exists $P\in X$ such that $$ \binom{g-2\gamma}{2}\le w(P)< \binom{g-2\gamma+2}{2}. $$ \end{itemize} \end{theorem'} Remark \ref{sharp-weight} says that the bound for $g$ above is the best possible. Further applications of \S3.1 and \S3.2 will be published elsewhere \cite{To2}.
1995-12-07T06:20:24
9512
alg-geom/9512004
en
https://arxiv.org/abs/alg-geom/9512004
[ "alg-geom", "math.AG" ]
alg-geom/9512004
Torsten Ekedahl
Torsten Ekedahl
Varieties of CM-type
plain TeX
null
null
null
null
We introduce the notion of a variety (or more generally a motive) of CM-type which generalises the well known notion of abelian variety of CM-type. Just as in that particular case it will turn out that the cohomology of the variety is determined by purely combinatorial data; the type of the variety. As applications we will show that the \l-adic representations are given by algebraic Hecke characters whose algebraic parts are determined by the type and give a method for computing the discriminant of the N\'eron-Severi group of super-singular Fermat surfaces.
[ { "version": "v1", "created": "Tue, 5 Dec 1995 13:08:10 GMT" }, { "version": "v2", "created": "Wed, 6 Dec 1995 08:42:58 GMT" } ]
2008-02-03T00:00:00
[ [ "Ekedahl", "Torsten", "" ] ]
alg-geom
\part{\ifnum\p@rtno>0\endgraf\fi\advance\p@rtno1 \def\br@mp[##1]{\e@\global\e@\edef\csname p@rt##1\endcsname {\romannumeral\p@rtno)}\romannumeral\p@rtno) \def\tr@mp{\if\t@mp\space\e@\e@tspace\fi}\futurelet\t@mp\tr@mp}% \def\t@mp{\if\tr@mp[\e@\br@mp\fi}% \futurelet\tr@mp\t@mp }} \def\pr@{\par \ifdim\lastskip<\medskipamount \removelastskip\penalty55\medskip\fi} \def\r@fnumb@ring{\sectname.\number\lop} \def\tr@mp#1.\@@#2{\def\@@{}\global\edef\t@mp{#1}}% \declareunit{procl} \declareparent partial\of procl \unitaction{procl}#1 {% \global\ch@pqtrue \global\advance\lop by 1% \ifempty{#1}% \edef\t@mp{\r@fnumb@ring.}% \else {\tr@mp#1\@@.\@@\@@}% \e@\global\e@\edef \csname l@p\t@mp\endcsname{\ifchap\the\chapno @\fi \iffinal\r@fnumb@ring\else \t@mp\fi}% \fi \medbreak\noindent{\bf \csname\presentunit name\endcsname\ \iffinal \r@fnumb@ring.% \else {\rm\ignorespaces\t@mp} \fi}% \enspace\sl} \declareunit{stmt} \declareparent procl\of stmt \langdef\pr@@f\rum\@a\rum english={PROOF},\rum\@a\rum swedish={BEVIS},\rum\@a\rum french={PREUVE}. \addeveryunit{stmt}{\def\pro{\egroup\let\ifpr@\iffalse \pr@{\bf \sevenrm\pr@@f}: }\bgroup\let\ifpr@\iftrue} \afterunit{stmt}{\ifpr@ \errhelp={You seem to have forgotten to to finish the statement and (at the same time) begin the proof with a \pro.}% \errmessage{No proof in statement.}\egroup \fi \slut} \declareunit{theorem} \declareparent stmt\of theorem \namedef theorem\rum\@a\rum english={Theorem},\rum\@a\rum swedish={Sats},% \rum\@a\rum french={Th{\'e}or{\`e}me}. \declareunit{proposition} \declareparent stmt\of proposition \namedef proposition\rum\@a\rum english={Proposition},\rum\@a\rum swedish={Proposition},% \rum\@a\rum french={Proposition}. \declareunit{lemma} \declareparent stmt\of lemma \namedef lemma\rum\@a\rum english={Lemma},\rum\@a\rum swedish={Lemma},\rum\@a\rum french={Lemme}. \declareunit{definition-lemma} \declareparent stmt\of definition-lemma \namedef definition-lemma\rum\@a\rum english={Definition-Lemma},% \rum\@a\rum swedish={Definition-Lemma},\rum\@a\rum french={Definition-Lemme}. \declareunit{corollary} \declareparent stmt\of corollary \namedef corollary\rum\@a\rum english={Corollary},\rum\@a\rum swedish={F{\"o}ljdsats},% \rum\@a\rum french={Lemme}. \declareunit{def} \declareparent procl\of def \afterunit{def}{\pr@} \declareunit{definition} \declareparent def\of definition \namedef definition\rum\@a\rum english={Definition},\rum\@a\rum swedish={Definition},% \rum\@a\rum french={Definition}. \declareunit{namedpar} \declareparent partial\of namedpar \everyunitadd{namedpar}{\smallskip\noindent {\bf\csname\presentunit name\endcsname}\rm:\hskip 0.5em plus2pt minus2pt\ignorespaces} \afterunit{namedpar}{\par\smallskip} \declareunit{remark} \declareparent namedpar\of remark \namedef remark\rum\@a\rum english={Remark},\rum\@a\rum swedish={Anm{\"a}rkning},% \rum\@a\rum french={R{\'e}marque}. \declareunit{example} \declareparent namedpar\of example \namedef example\rum\@a\rum english={Example},\rum\@a\rum swedish={Exempel},% \rum\@a\rum french={Exemple}. \ovnnr=1 \declareunit{exercise} \declareparent namedpar\of exercise \namedef exercise\rum\@a\rum english={Exercise \the\ovnnr},% \rum\@a\rum swedish={{\"O}vning \the\ovnnr},\rum\@a\rum french={Exercise \the\ovnnr}. \addafterunit{exercise}{\global\advance\ovnnr by 1} \forbidinself{exercise} \def\markexercise#1{\ifdef{ex@r@f#1} \errhelp{#1 has already been used to mark an exercise.}% \errmessage{Reuse of exercise mark.}% \else \e@\xdef\csname ex@r@f#1\endcsname{\the\ovnnr}% \fi} \def\exerciseref#1{\ifdef{ex@r@f#1}\csname ex@r@f#1\endcsname \else \errhelp{#1 has not been used to mark any exercise.}% \errmessage{Reference to unmarked exercise.}% \fi} \declareunit{solution} \namedef solution\rum\@a\rum english={Solution \the\ovnnr},% \rum\@a\rum swedish={L{\"o}sning \the\ovnnr},\rum\@a\rum french={Solution \the\ovnnr}. \requireunit solution\inside{exercise} \forbidinself{solution} \declareunit{hint} \namedef hint\rum\@a\rum english={Hint \the\ovnnr},% \rum\@a\rum swedish={Ledning \the\ovnnr},\rum\@a\rum french={Suggestion \the\ovnnr}. \requireunit hint\inside{exercise} \forbidinself{hint} \def\exerciselast{ \startout{exercise}\writeandput \startout{hint}\writeonly \startout{solution}\writeonly \afterunitadd{document}{\ovnnr0 \unitaction{exercise}{} \everyunitadd{exercise}{\ignorein{hint}}% \everyunitadd{exercise}{\ignorein{solution}}% \def\markexercise##1{} \vfill\eject\centerline{\bf {\"O}vningar} \vskip1cm \csname input\endcsname\t@stpt\jobname..\relax.exercise\ \declareparent namedpar\of hint \norequireds{hint} \vfill\eject\centerline{\bf Ledningar} \vskip1cm \csname input\endcsname\t@stpt\jobname..\relax.hint\ \declareparent namedpar\of solution \norequireds{solution} \vfill\eject\centerline{\bf L{\"o}sningar} \vskip1cm \csname input\endcsname\t@stpt\jobname..\relax.solution\ }} \def\hglue\fontdimen2\the\font \relax{\hglue\fontdimen2\the\font \relax} {\catcode`\^^M=\active\catcode`\ =\active \gdef\v@rb{\let^^M\ \let \ } \gdef\v@rbdis{\let^^M\par\let \hglue\fontdimen2\the\font \relax} \global\def\setesc@pe#1{\if#1 \catcode`@=0\else\catcode`#1=0\fi{}} } \def\showcs#1{{\tt\char`\\#1}} \declareunit{verb} \everyunit{verb}{\bgroup\def\do#1{\catcode`#1=12}\dospecials \def\egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{\egroup\egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt}\catcode`\^^M=\active\catcode`\ =\active\tt} \unitaction{verb}{\setesc@pe} \declareunit{verbatim} \declareparent verb\of verbatim \everyunitadd{verbatim}{\v@rb} \forbidinself{verbatim} \declareunit{verbatimdisplay} \declareparent verb\of verbatimdisplay \addeveryunit{verbatimdisplay}{\vskip-\lastskip\smallskip} \everyunitadd{verbatimdisplay}{\v@rbdis\baselineskip10pt\parindent0pt \leftskip2cm} \afterunit{verbatimdisplay}{\vskip-\lastskip\smallskip\noindent} \forbidinself{verbatimdisplay} \declareunit{condition} \unitaction{condition}#1{\medskip\setbox0=\hbox{#1}% \setbox1=\vbox\bgroup\advance\hsize-1.5cm\advance\hsize-\wd0\noindent} \afterunit{condition}{\egroup\dimen1=\ht1 \hbox to\hsize{\hfill\vbox to\dimen1{\vfill\box0\vfill}\hfill \box1\hskip.75cm}\medskip} \declareunit{bibliography} \everyunit{bibliography}{\bibl} \def\finaltrue\hfuzz=2pt\errorstopmode{\finaltrue\hfuzz=2pt\errorstopmode} \def\pr@l@m{\special{" gsave 55 rotate /Times-Roman findfont 110 scalefont setfont 0.9 setgray -50 -70 translate (PRELIMINARY) stringwidth -1 mul exch -1 mul exch moveto (PRELIMINARY) show grestore }} \def\draft{\finalfalse\hfuzz=2pt\errorstopmode \global\headline={% \ifnum\pageno=1\hfill\pr@l@m \else\tenrm\ifsvensk\idag\else\today\fi \hfill PRELIMINARY VERSION -- No preprint\hfill\t@stpt\jobname..\relax\quad\fi}} \def\t@stpt#1.#2.{#1\e@tif.} \def\t@stpt\jobname..\relax{\t@stpt\jobname..\relax} \def\head#1\par{% \global\headline={% \iffinal {\ifnum\pageno>1\hfill\sevenrm \uppercase{#1}\hfill\else\hfill\fi}% \else {\tenrm\ifsvensk\idag\else\today\fi \hfill\t@stpt\jobname..\relax\quad}% \fi}} \chapno=0 \def\ch@pter#1\par{ \ifch@pq \ifchap \else \errhelp={You must either use or not use chapters in a document. In references chapter numbers will now be ignored.}% \errmessage{Chapter introduced in the middle of document.}% \fi \else \global\chaptrue \fi \global\advance\chapno by1\sect=0\lop=0 \vfill\eject \noindent \centerline{\bf\ifnum0=\chapno0\else \uppercase\e@{\romannumeral\chapno}\fi: #1}% \nobreak\vskip1cm\nobreak\noindent} \def\noreferrtrue{\noreferrtrue} \def\fe@#1@{\ifnum#1=\chapno \else\ifnum0=#1 0\else \uppercase\e@{\romannumeral #1}\fi:\fi} \makesplit\spl@tco: \def\isd@fin@d#1{% \ifdef{l@p#1}% \@strue \else \ifnoreferr \else \errhelp=\e@{#1 does not refer to anything.}% \errmessage{Reference to nonexistent result.}% \fi \@sfalse \fi} \def\refpart#1{\csname p@rt#1\endcsname} \def\(#1){% (% \spl@tco{#1}% \edef\r@st{\csname l@p\firstsplit\endcsname\ \secondsplit\unskip)}% \isd@fin@d\firstsplit \ifchap \tr@th \e@\fe@\r@st \else #1)% \fi \else \r@st \fi } \def\ref#1{% \isd@fin@d{#1}% \ifchap \tr@th \e@\e@\e@ \fe@\csname l@p#1\endcsname \else(#1)% \fi \else \csname l@p#1\endcsname \fi } \def\slut{\nobreak\hfill \hbox{% \vrule height 5pt \kern-.4pt \vbox{ \hrule width 5pt depth0pt height.4pt \kern4.6pt \hrule } \kern-3.75pt \vrule height 5pt}\kern1pt \par} \def\t@sta#1 #2{\t@{#1}\e@t} \def\tag#1${% \global\ch@pqtrue \global\advance\lop by 1% \t@sta#1 @ \unskip \global% \e@\e@ \e@\edef \e@\csname \e@ l\e@ @\e@ p\the\t@\endcsname \ifchap \the\chapno \fi \iffinal \r@fnumb@ring \else \the\t@% \fi \leqno\hbox{(% \iffinal \r@fnumb@ring \else \the\t@% \fi)% $% \def\leqno{\leqno} \def\leqalignno{\leqalignno} \def\mkref#1{% \global\advance\lop by 1% \e@\global\e@\edef \csname l@p#1\endcsname{\ifchap\the\chapno @\fi \iffinal\r@fnumb@ring\else #1\fi} \iffinal (\r@fnumb@ring)% \else {\rm\ignorespaces(#1)} \fi } \def\x@struf{\dp\strutbox} \def\x@notnu{\strut\vadjust{\kern-\x@struf\vtop to\x@struf{% \baselineskip\x@struf\vss\llap{\sevenrm\the\x@ant\quad}\null}}} \x@ant=0 \x@sid={. s. } \x@note={} \long\def\note#1{% \iffinal \else \global\advance\x@ant1 \iffinal \else \x@notnu \t@ \e@\e@ \e@ {\e@\the\e@\x@ant\e@\the\e@\x@sid \the\pageno: #1}% \global\x@note=\e@{\the\e@\x@note \e@\par\the\t@}% \fi \fi} \x@bib={} \long\def\x@ny#1{\t@={#1}% \global\x@bib=\e@{\the\e@\x@bib\e@\par\the\t@} \def\[#1]{[#1]\iffinal\else\x@ny{#1}\fi} \namedef bibliography\rum\@a\rum english={Bibliography},\rum\@a\rum swedish={Bibliogafi},% \rum\@a\rum french={Bibliographie}. \bibliographytrue \def\bibl{ \ifbibliography \medbreak \centerline{\bf\csname\presentunit name\endcsname} \nobreak \medskip \nobreak \fi \ninerm \frenchspacing \def\egroup\setbox1=\hbox\bgroup\smc{\egroup\setbox1=\hbox\bgroup\smc} \def\egroup\setbox2=\hbox\bgroup{\egroup\setbox2=\hbox\bgroup} \def\egroup\prep@true\setbox2=\hbox\bgroup\sl{\egroup\prep@true\setbox2=\hbox\bgroup\sl} \def\egroup\jour@true\setbox3=\hbox\bgroup\sl{\egroup\egroup\jour@true\setbox3=\hbox\bgroup\sl@true\setbox3=\hbox\bgroup\sl} \def\egroup\setbox4=\hbox\bgroup\bf{\egroup\setbox4=\hbox\bgroup\bf} \def\egroup\setbox5=\hbox\bgroup{\egroup\setbox5=\hbox\bgroup} \def\egroup\pages@true\setbox6=\hbox\bgroup{\egroup\egroup\pages@true\setbox6=\hbox\bgroup@true\setbox6=\hbox\bgroup} \def\egroup\page@true\setbox6=\hbox\bgroup p. {\egroup\egroup\page@true\setbox6=\hbox\bgroup p. @true\setbox6=\hbox\bgroup p. } \def\egroup\book@true\setbox2=\hbox\bgroup\sl{\egroup\egroup\book@true\setbox2=\hbox\bgroup\sl@true\setbox2=\hbox\bgroup\sl} \def\egroup\lect@true\setbox7=\hbox\bgroup{\egroup\lect@true\setbox7=\hbox\bgroup} \def\egroup\setbox3=\hbox\bgroup{\egroup\setbox3=\hbox\bgroup} \def\egroup\setbox4=\hbox\bgroup{\egroup\setbox4=\hbox\bgroup} \def\egroup\setbox4=\hbox\bgroup{\egroup\setbox4=\hbox\bgroup} \def\egroup\inbook@true\setbox7=\hbox\bgroup{\egroup\egroup\inbook@true\setbox7=\hbox\bgroup@true\setbox7=\hbox\bgroup} \def\egroup\sln@true\setbox4=\hbox\bgroup{\egroup\egroup\sln@true\setbox4=\hbox\bgroup@true\setbox4=\hbox\bgroup} \def\egroup\spec@true\setbox8=\hbox\bgroup{\egroup\egroup\spec@true\setbox8=\hbox\bgroup@true\setbox8=\hbox\bgroup} \def\[##1:##2\par{\egroup\jour@true\setbox3=\hbox\bgroup\sl@false\egroup\book@true\setbox2=\hbox\bgroup\sl@false\egroup\inbook@true\setbox7=\hbox\bgroup@false\egroup\pages@true\setbox6=\hbox\bgroup@false \egroup\sln@true\setbox4=\hbox\bgroup@false\egroup\page@true\setbox6=\hbox\bgroup p. @false\prep@false\egroup\spec@true\setbox8=\hbox\bgroup@false\lect@false \noindent\hangindent2.3\parindent\hangafter1\hbox to2.3\parindent{[##1:\hfill}\bgroup##2\egroup\unskip \unhbox1\unskip, \ifjour@ \unhbox2\unskip, \unhbox3\unskip\ \unhbox4\unskip\ (\unhbox5\unskip), \unhbox6\unskip.\fi \ifbook@ \unhbox2\unskip, \unhbox4\unskip, \unhbox3\unskip, \unhbox5\unskip \ifpages@\unskip, pp. \unhbox6\unskip.\else\ifpage@\unskip ,\unhbox6\unskip.\else\unskip.\fi\fi\fi \iflect@ \unhbox2\unskip, \unhbox7\unskip\ \unhbox4\unskip, \unhbox3\unskip, \unhbox5\unskip \ifpages@\unskip , pp. \unhbox6\unskip.\else\ifpage@\unskip ,\unhbox6\unskip.\else\unskip.\fi\fi\fi \ifinbook@ \unhbox2\unskip, in \unhbox7\unskip, \unhbox4\unskip, \unhbox3\unskip, \unhbox5\unskip, \ifpages@ pp. \fi \unhbox6\unskip.\fi \ifsln@ \unhbox2\unskip, SLN \unhbox4\unskip, Springer-Verlag \ifpages@\unskip, pp. \unhbox6\unskip.\else\ifpage@\unskip, \unhbox6\unskip.\else\unskip.\fi\fi\fi \ifprep@ \unhbox2\unskip, (preprint) \unhbox4\unskip, \unhbox5\unskip.\fi \ifspec@\unhbox8\fi \par} } \def\sc@le#1#2{\divide#1 by 1000\multiply#1#2} \def\p@llbox#1#2#3{\hbox{% \getsc@le{#1}% #2% \ifx\scalep@rt\relax\def\scalep@rt{1000}\else \sc@le\h@ight\scalep@rt \sc@le\w@dth\scalep@rt\fi \vbox to \h@ight{\hrule width \w@dth height 0pt depth 0pt #3{#1}% \vfill\special{\t@mp}}}} \def\eatsc@led scaled{} \def\insc@le#1scaled#2scaled{\global\def\namep@rt{#1 } \ifempty{#2}\global\let\scalep@rt\relax\else \global\def\scalep@rt{#2}\e@\eatsc@led\fi} \def\getsc@le#1{\insc@le#1scaledscaled} \newread\@psffile \dimendef\h@ight=0 \dimendef\w@dth=1 \def\pictg@tdim#1#2{\h@ight=#2\unskip\w@dth=#1\unskip} \def\pictt@mpl#1{\def\t@mp{picture #1}} \def\pictbox#1 by #2 (#3){\p@llbox{#3}{\pictg@tdim{#1}{#2}\unskip}\pictt@mpl} \def\picture #1 by #2 (#3){\vskip 6pt\centerline{% \pictbox #1 by #2 (#3)}\vskip 6pt\noindent} { \catcode`\%=11 \catcode`\|=14 \global\def\illg@tdim{| \openin\@psffile=\namep@rt \unskip {| \catcode`\%=11| \catcode`\_=11| \catcode`\&=11| \def\h@ndle##1 ##2 ##3 ##4:{{\globaldefs1\def\l@ft{##1 bp}| \def\b@t{##2 bp}\def\r@ght{##3 bp}\def\t@p{##4 bp}}}| \def\t@est \def\e@tspace##1{\if##1 \else##1\fi}| \def\BBt@st##1:## \edef\s@tt{\e@tspace##2}| \e@\h@ndle\s@tt\unskip\let\n@xt\relax\fi}| \read\@psffile to \h@p \def\n@xt{| \read\@psffile to \h@p \e@\BBt@st\h@p \n@xt}| \n@xt }| \closein\@psffile \e@\h@ight\t@p \e@\advance\e@\h@ight\e@-\b@t \e@\w@dth\r@ght \e@\advance\e@\w@dth\e@-\l@ft} } \def\illt@mpl#1{\def\t@mp{illustration #1}} \def\illbox(#1){\p@llbox{#1}\illg@tdim\illt@mpl} \def\illustration(#1){\vskip 6pt plus 2pt minus 2pt\centerline{\illbox(#1)}% \vskip 6pt plus 2pt minus 2pt\noindent} \let\psprol\relax \let\psend\relax \def\pst@mpl#1{\e@\w@dth\l@ft \e@\h@ight\b@t \sc@le\w@dth\scalep@rt \sc@le\h@ight\scalep@rt \divide\w@dth65536 \divide\h@ight65536 {\count255=\scalep@rt \divide\count255 10 \count254\w@dth \count253\h@ight \xdef\t@mp{psfile=\namep@rt hoffset=-\number\count254 \space\space voffset=-\number\count253\space\space vscale=\number\count255\space\space hscale=\number\count255}% }} \def\psbox(#1){\psprol\p@llbox{#1}\illg@tdim\pst@mpl\psend} \def\psillustration(#1){\vskip 6pt plus 2pt minus 2pt\centerline{\psbox(#1)}% \vskip 6pt plus 2pt minus 2pt\noindent} \ifmac\else \let\illbox\psbox \let\illustration\psillustration \fi \langdef\yrs\rum\@a\rum english={Yours\iffriendly\else sincerely\fi},% \rum\@a\rum swedish={\iffriendly H\else V{\"a}nliga h\fi {\"a}lsningar},% \rum\@a\rum french={Bien {\'a} \iffriendly toi\else vous\fi}. \def\Torsten{\vskip1cm\leftline{\hskip10cm \yrs}\nobreak \vskip.75cm\nobreak \leftline{\hskip10cm Torsten \iffriendly\else Ekedahl\fi}} \def\letter{\footline={\hss\ifnum\pageno=1\else\tenrm\folio\fi\hss}% \headline={\ifnum\pageno>1T.~Ekedahl\hss\today\else\hfil\fi}% \rightline{Stockholm \today\qquad}\vskip3cm} \let\friendly\friendlytrue \langdef\dear\rum\@a\rum english={Dear},\rum\@a\rum swedish={},\rum\@a\rum french={Cher}. \def\beginletter#1\par{\letter \dear #1\unskip,\par\qquad} \def\endletter#1\par{\if\relax#1\relax \else\friendlytrue\fi\nobreak\Torsten\vfill\eject\@nd} \def\hsize=14cm \voffset=0pt \hoffset=.2cm{\hsize=14cm \voffset=0pt \hoffset=.2cm} \def\Zblatt{\finaltrue\hfuzz=2pt\errorstopmode\nopagenumbers\voffset=3.75cm\hsize=14cm \hoffset=2.5cm\vsize=18cm} \catcode`\@=12 \ifdump \let\dmp\dump \else \let\dmp\relax \fi \dmp \finaltrue\hfuzz=2pt\errorstopmode \def\symb{Tr}{\symb{Tr}} \def\symb{Irr}{\symb{Irr}} \def\symb{Frac}{\symb{Frac}} \mdef\coh{H^*(X,r)} \mdef{\hodg#1.#2.}{H^{#1}(X,{\gOm}^{#2}_{X/k})} \candef W \head varieties of CM-type \begin{document} \begin{start} Varieties of CM-type \egroup\setbox1=\hbox\bgroup\smc Torsten Ekedahl \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{start} \begin{introduction} We will introduce the notion of a variety (or more generally a motive) of CM-type which generalises the well known notion of abelian variety of CM-type. Just as in that particular case it will turn out that the cohomology of the variety is determined by purely combinatorial data; the type of the variety. As applications we will show that the \l-adic representations are given by algebraic Hecke characters whose algebraic parts are determined by the type and give a method for computing the discriminant of the N{\'e}ron-Severi group of super-singular Fermat surfaces. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{introduction} \begin{section}Preliminaries. To begin with let us recall the following facts from category theory. If \Cal A is an additive category all of whose idempotents have kernels and $R$ is a ring, then for a finitely right resp.~left projective $R$-module $P$ resp.~$Q$ and a left $R$-object $M$ in \Cal A we can define objects $P\bigotimess _RM$ resp.~$Hom_R(Q,M)$ of \Cal A characterised by $$ \leqalignno{Hom_{\Cal A}(P\bigotimess _RM,N)&=Hom_R(P,Hom_{\Cal A}(M,N))\cr \noalign{\leftline{resp. }} Hom_{\cal A}(N,Hom_R(Q,M))&=Hom_R(Q,Hom_{\cal A}(N,M)).\cr} $$ We always have a natural $R$-morphism $ev\co P\to Hom_{End_R(P)-\Cal A}(Hom_R(P,M),M)$, the evaluation map, obtained by interpreting an element $p\rum\@a\rum P$ as an $R$-morphism $R\to P$ and using $M=Hom_R(R,M)$. If R=$\bigoplus P_i^{n_i}$ and $Hom_R(P_ i,P_j)=0$ for $i\ne j$ then for any $R$-object $M$ in \Cal A we have $$ M=\bigoplus P_i{\textstyle\bigotimess }_{S_i}Hom_R(P_i,M),\tag 1.1 $$ where $S_i:=End_R(P_i)$ and the map is defined using the evaluation maps. To see this we first note that $P_i=Hom_R(R,P_i)\cong(S_i)^{n_i}$ so that $P_i$ is $S_i$-projective and then the desired equivalence follows by decomposing the two $R$-factors of $M=R\bigotimess _RHom_R(R,M)$. The following setup will be with us during the rest of the paper: We let \k\ be a perfect field of characteristic $p\ge 0$, $X$ a proper, smooth variety over \k\ (alternatively a motive) and $S$ a set of \k-correspondences of $X$. Furthermore, \coh, $r$ prime, will denote the \l-adic cohomology of $X_{\k}$, where \bk\ is a fixed algebraic closure of \k, when $r\ne p$ and the crystalline cohomology of $X/\k$ when $r=p$. Recall that when $r\ne p$ \coh\ is a graded $\Z_r$-algebra, finitely generated as $\Z_r$-module, having a continuous action of $Gal(\bk/\k)$ and that when $r=p$, \coh\ is a graded $\W(\k)$-algebra, finitely generated as $\W(\k)$-module having a {\gsi}-linear endomorphism $F$. Here $\W(\k)$ is the ring of Witt vectors of \k\ and {\gsi} sends a Witt vector $(x_i)$ to $(x^p_i)$. We let $L_r$ be an algebraically closed field containing $\Z_r$ resp.~$\W(\k)$. Furthermore we will denote by $K$ the fraction field of $\W(\k)$. Finally, if $p>0$ we will have need of the following technical condition. There is a scheme $T$ of finite type over $\F_p$, a smooth and proper morphism $\Cal X\to T$ and a cartesian diagram $$ \diagram{ X&\mapright{}&\cal {X}\cr \mapdown{}&&\mapdown{}\cr \Sp\;k&\mapright{}&T\cr} $$ such that for every closed point $t\rum\@a\rum T$ the eigenvalues of the Frobenius with respect to $\k(t)$ on $H^i(X_t,p)$ are algebraic integers all of whose archimedean absolute values are $|k(t)|^{i/2}$. This condition is fulfilled when $X$ (possibly over \bk) is the image of a smooth and projective variety (\[K-M]) and that this is always the case has recently been verified by J.~de Jong (\[Jo]). For a field $L$ and a set $R$ let $K(R,L)$ be the Grothendieck group of the category of finite dimensional representations (i.e.~maps of $R$ into the set of endomorphisms) of $R$. Then $K(R,L)$ is a functor in $R$ and $L$; contravariant in $R$ and covariant in $L$. If $M(R)$ is the free monoid generated by $R$ then the trace map gives an additive map $\symb{Tr}: K(R,L)\to L^{M(R)}$. \begin{lemma}1.2. Let $L$ be algebraically closed of characteristic 0. i) $\symb{Tr}: K(R,L)\to L^{M(R)}$ is injective. ii) Let $L'\subseteq L$ be a subfield of $L$ and $N$ a semi-simple $L$-representation of $R$ s.t.~for all $r\rum\@a\rum M(R)$ $\symb{Tr}_N(r)\rum\@a\rum L'$. If $L'\langle R\rangle$ is the free associative $L'$-algebra on $R$ then $I :=\ker (L'\langle R\rangle\to End_L(N))$ depends only on the function $\symb{Tr}_N: M(R)\to L'$ and $L'\langle R\rangle/I\bigotimess _{L'}L\to End_L(N)$ is an injection. In particular, if $L'$ is algebraically closed $N$ is isomorphic to the scalar extension of some $L'$-representation of $R$. iii) If $L'$ is an algebraically closed field and $L'\to L$ a field homomorphism, then the following diagram $$ \diagram{ K(R,L')&\mapright{}&K(R,L)\cr \mapdown{}&&\mapdown{}\cr L'{}^{M(R)}&\mapright{}&L^{M(R)}\cr} $$ is cartesian. \pro Let us begin with ii). Note first that as $\mathop{\rm im}\nolimits(L\langle R\rangle\to End_L(N))$ is semi-simple and that for a finite dimensional semi-simple L-algebra $M$ and a faithful finite dimensional $L$-representation $V$, the linear form $$ \eqalign{ M \times\;M&\to L\cr (m,m')\mapsto Tr_V(mm')\cr} $$ is non-degenerate. Hence $t\rum\@a\rum L\langle R\rangle$ acts as zero on $N$ iff $\symb{Tr}_N(rt)=0$ for all $r\rum\@a\rum M(R)$. If $t=\sum _{r\rum\@a\rum M(R)}{\gla}_rr$ then this is a set of linear conditions on the ${\gla}_r$ with coefficients in $L'$ depending only on $\symb{Tr}_N: M(R)\to L'$. Furthermore, $L'\langle R\rangle/I\bigotimess _{L'}L\to End_L(N)$ is injective iff whenever there is an $L$-linear relation in $End_L(N)$ between elements in $L'\langle R\rangle$ there is also an $L'$-linear relation. This also follows from the fact that the above conditions have $L'$-coefficients. Now iii) follows immediately from i) and ii) whereas i) is well known (cf.~\[C-R:Thm. 30.12]). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{lemma} If $L$ is a field of characteristic zero and $L'$ an algebraic closure of $L$, we put $\ovl K(R,L):=K(R,L')\bigcap L^{M(R)}$. This is clearly independent of the choice of $L'$ and we have $K(R,L)\subseteq \ovl K(R,L)$. If $N$ is a representative over $L'$ of an $n\rum\@a\rum \ovl K(R,L)$ then we can construct the $L\langle R\rangle/I$ of lemma \ref{1.1}, which depends only on $n$. It is a semi-simple $L$-algebra as its scalar extension to $L'$ is, and we will denote it $A_L(n)$. In case $L=\Q$ then we put $A(n):=\mathop{\rm im}\nolimits(\Z\langle R\rangle\to A_\Q(n)$, where $\Z\langle R\rangle$ denotes the free associative algebra on $R$. If $M$ is an over-field of $L$ then we say that $n$ is realisable over $M$ if the induced element in $\ovl K(R,M)$ belongs to $K(R,M)$. This is equivalent to $N$ being realisable by an $A_L(n)\bigotimess _LM$-representation. For any $n\rum\@a\rum K(R,L')$ we let $\symb{Irr}(n)\subset K(R,L')$ be the set of irreducible constituents of n. If now $n\rum\@a\rum\ovl K(R,L)$ then $\symb{Irr}(n)$ is a finite set stable under the action of $Gal(\bar\Q/\Q)$. Under the correspondence of Galois theory, $\symb{Irr}(n)$ then corresponds the {\'e}tale $L$-algebra $Z(A_L(n))$. If we return to the situation at hand we have elements $[H^i(X,r)\bigotimess _{\Z_r}L_r]$ (resp.~$[H^i(X,p)\bigotimess _{\W(k)}L_p]$) in $K(S,L_r)$. Let $\bar \Q$ be an algebraic closure of \Q. \begin{lemma}1.3. Let $K_r$ denote $\Q_r$ when $r\ne p$ and $K_r$ when it isn't. There exists a unique element $[H^i(X)]\rum\@a\rum \ovl K(S,\Q)$ whose image in $K(S,K_r)$ coincides with $[H^i(X,r)]$. Furthermore, $A([H_i(X)])$ is finitely generated as \Z-module and $A\bigotimess K_r$ equals $A_r/rad(A_r)$ where $$ A_r:=Im(\Q_r\langle S\rangle\to End_{\Q_r}(H^i(X,r))\;\;\;\;\;\;(resp.\;\ldots). $$ \pro I first claim that, for every $s\rum\@a\rum M(S)$, $Tr(s,H_i(X,r))$ is a rational number independent of $r$. By standard specialisation arguments we reduce to \k\ being a finite field where it is \[K-M:Thm 2] (supplemented by \[Gr] for the definition of the cycle map in crystalline cohomology). Note that if $p=0$, using Chow's lemma and resolution of singularities we can get a reduction for which our technical condition is fulfilled. In this case, a transcendental argument can also be used. This already, using \(1.2:ii), gives the existence of $[H_i(X)]$ and that $A\bigotimess L_r = A_r/rad(A_r)$. Hence $A\bigotimess \Q$ is a finite dimensional semi-simple \Q-algebra. For any $t\rum\@a\rum \Z\langle S\rangle$ the characteristic polynomial of $t$ on $H_i(X,r)\bigotimess L_r$ is independent of $r$ and has rational coefficients by [loc.~cit.]. As $t$ stabilises a $\Z_{\ell}$- (resp.~$\W(\k)$-) lattice in $H_i(X,r)\bigotimess L_r$, those coefficients are $r$-integral for all $r$ and so integral. By the Cayley-Hamilton theorem the image of $t$ in $A\bigotimess \Q$ is integral over \Z\ and so $A$, being equal to $\mathop{\rm im}\nolimits:\Z\langle S\rangle\to A\bigotimes \Q$, is finitely generated as it is contained in the different ideal of any order containing it. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{lemma} If still $L$ is algebraically closed of characteristic 0 and $R$ and $T$ are two sets, then we let $K(R,T,L)$ denote the Grothendieck group of the category of finite dimensional $L$-representations of $R \disjunion T$ such that every element of $R$ commutes with every element of $T$. It is easy to see that every simple object of this category is a tensor product of an irreducible representation of $R$ and one of $T$ and that the two factors are well-determined up to isomorphism. Hence $K(R,T,L)=K(R,L)\bigotimess K(T,L)$. Furthermore, as $K(R,L)$ has a canonical base consisting of irreducible representations we get a canonical pairing $$ K(R,L)\bigotimess K(R,L)\to \Z $$ where the the canonical base is orthonormal. Using this we get a mapping $$ K(R,L)\bigotimess K(R,T,L)=K(R,L)\bigotimess K(R,L)\bigotimess K(T,L)\to K(T,L) $$ and so for each $N\rum\@a\rum K(R,T,L)$ a mapping $$ N\cap \co K(R,L)\to K(T,L).\tag1.4 $$ This mapping is compatible, in the obvious way, with the mappings obtained from homomorphisms $L\to L'$ of algebraically closed fields. Hence we get \begin{corollary}1.5. Let $S'$ be a set of \k-correspondences of $X$ and suppose that every element of $S$ commutes up to homological equivalence with $S'$. Then \(1.4) gives a $Gal(\bQ/\Q)$-equivariant homomorphism $$ [H^i(X)]\cap : K(S,\bQ)\to K(S',\bQ)\tag(1.6) $$ which equals, for each $r$, the restriction of $[H^i(X,r)\bigotimess L_r]\cap $ to K(S,$\bQ$). \pro \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{corollary} \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{section} \begin{section}Varieties of CM-type. Let us fix an $n\rum\@a\rum \N$ and assume, for simplicity, that $$b_n(X) =\sum _{i+j=n}\dim_kH^i(X,{\gOm}^j_{X/k}),$$ where $b_n(X) := \dim_{ L{_r}} H^n(X,r)$ for any $r$. (This is of course always true when $p=0$.) If $p>0$ this implies (cf.~\[Ek: IV, 1.2] or \[B-Og:\S8]) that $H^n_{DR}(X/k)=H^n(X,p)/pH^n(X,p)$ and that if $$ M^i:= \mathop{\rm im}\nolimits(F^{-1}p^iH^n(X,p)\to H^n(X,p)/pH^n(X,p)) $$ then $$ M^i/M^{i+1}=H^{n-i}(X,{\gOm}^i_{X/k}). $$ Hence, no matter the value of $p$, $H^n_{DR}(X/k)$ has a Hodge filtration with the $H^{n-i}(X,{\gOm}^i_{X/k})$ as successive quotients. \begin{definition}2.1. $(X,S)$ is said to be of {\deffont separable CM-type in degree $n$} if the $A_i$ of \(1.3) has the property that $A\bigotimess \Z_{(p)}$ is a separable (cf.~\[D-I:II,1]) $\Z_{(p)}$-algebra and for every $0\le i\ne j\le n$, $H^{n-i}(X,{\gOm}^i_{X/k})$ and $H^{n-j}(X,{\gOm}^j_{X/k})$ are disjoint $S$-modules (i.e.~they have no common composition factors). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition} \begin{remark} As the quotient of a separable algebra is separable it suffices to verify that $\Z_{(p)}\langle S\rangle$ factored by some known relations is separable. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{remark} \begin{example}i) If $X$ is an abelian variety of CM-type in the usual sense and $End_k(X)$ is separable at $p$, which is always true if $p=0$, then $(X,End_k(X))$ is of CM-type in degree 1 (and in fact in all other degrees). ii) Kummer surfaces associated to abelian surfaces of CM-type are of CM-type in degree 2. Hence, by \[S-I], K3-surfaces in characteristic 0 for which the rank of the N{\'e}ron-Severi group is 20 are of CM-type in degree 2. iii) (Fermat hyper-surfaces, diagonal automorphisms). This is well known (cf.~e.g.~\[Ka:Sect.~6]). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{example} \begin{lemma}2.2. Suppose $(X,S)$ is of separable CM-type. If $p=0$ then $[H^n(X)]$ is realisable over \Q\ and if $p>0$ then $A^i\bigotimess _\Z\Z_p$ is unramified (i.e.~a product of matrix algebras over unramified extensions of $\Z_p$) and in particular $[H^n(X)]$ is realisable over $\Q_p$. \pro The case $p=0$ follows by transcendental methods, in fact $[H_n(X)]$ is realised by singular cohomology, and the $p>0$ is well-known (use the fact that the Brauer group of a finite field is trivial and lift an idempotent). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{lemma} \begin{remark} Is it possible to give an algebraic proof of the first part of the lemma? The existence of \l-adic cohomology implies that it suffices to prove realisability over \R. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{remark} Suppose now that $(X,S)$ is of separable CM-type and let $M$ be an irreducible component of $[H_n(X)]$. If $p=0$ there is then an irreducible $S$-module $N$ such that $ M\bigotimess _\bQ\bk$ is a factor of $M\bigotimess _\k \bk$ for an embedding of $\bQ$ in \bk\ and $N$ is a sub-quotient of $H^n_{DR}(X/k)$. (Note that the base extension of $[H^n(X)]$ to \bk\ equals the extension of $[H^n_{DR}(X/k)]\rum\@a\rum K(S,\k)$ to \bk, which is seen by either using a constructibility argument to reduce to \(1.3) or a transcendental argument.) By assumption there is a unique $i$, $0\le i\le n$, such that $N$ occurs as a sub-quotient of \hodg n-i.i. and this $i$ depends only on $M$. If $p>0$ we get in the same way an irreducible $K\langle S\rangle$-module $N$, $K:=\W(\k)\bigotimess \Q$, such that $M\bigotimess _\bQ\bar K$ occurs in $N\bigotimess _{K}\bar K$ for an algebraic closure $\bar K$ of $K$ and an embedding of \bQ\ in $\bar K$ and $N$ occurs in $H^ n(X,p)\bigotimess _{\W(k)}K$. As $A\bigotimess \W(\k)$ is separable, there is a unique, up to isomorphism, $A\bigotimess \W(\k)$-lattice $N'$ with $N'\bigotimess _\W k=N$ (this follows from \(2.2)) and $N'\bigotimess \k$ is an irreducible $A\bigotimess \k$-module and we see that $N'\bigotimess \k$ is an irreducible sub-quotient of of $H^n_{DR}(X/\k)$. By assumption there is then a unique $i$, $0\le i\le n$, such that $N\bigotimess \k$ occurs in a \hodg n-i.i. and this $i$ depends only on $M$. In both cases we put ${\gta}(M) := i$. In conclusion we have obtained a mapping $$ {\gta}: \symb{Irr}([H^n(X)])\to n+1\;(:=\{0,1,\relax\ifmmode \ldots\else \dots\fi ,n\}). $$ Note that the action of $Gal$(\bk/\k) on $\ovl \symb{Irr}([H^n(X)]$) obtained through the action of $Gal(\bQ/\Q)$ on it and the induced map $Gal(\bk/\k)\to Gal(\bar\Q/\Q)$ (resp.~$Gal(\bk/k)\to Gal(\W(\bk)/\W(k))$) preserves the fibers of {\gta} by construction. \begin{definition}2.3. Under the assumption of \(2.1) the type of $(X,S)$ is the pair ($[H^n(X)]$,{\gta}). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition} Finally, when $p>0$ the condition that $A$ is separable at $p$ implies that the action of $Gal(\bar \Q/\Q)$ on $\symb{Irr}([H^n(X)])$ is unramified at $p$ so that we may unambiguously speak about the action of the Frobenius morphism on $\symb{Irr}([H^n(X)])$ having once and for all chosen an embedding of \bQ\ in $L_p$. This permutation of $Irr([H^n(X)])$ we will denote {\gsi}. We have now come to the main result of the present paper. Before we formulate it we will need to introduce some constructions. Let $T$ be a set, $M\rum\@a\rum \ovl K(T,Q)$ and ${\gta}\co \symb{Irr}(M)\to n+1$ a function. Let us choose an embedding of \bQ\ into \C\ and let ${\gio}\rum\@a\rum Gal(\bQ/\Q)$ be the element corresponding through this embedding to complex conjugation. Suppose that for every ${\grh}\rum\@a\rum \symb{Irr}(M)$, ${\gta}({\gio}({\grh}))=n-{\gta}({\grh})$ and also that $M$ is realisable over \Q\ by a module $V$. For each simple factor $A_r$ of $A_\Q$ we let $V_r$ be an irreducible $A_r$-module. We then put a rational Hodge structure on $V_r$, of weight $n$, as follows: For ${\grh}\rum\@a\rum \symb{Irr}(M)\bigcap \symb{Irr}(V_r\bigotimess \C)$ we let $V_{r,{\grh}}$ be the {\grh}-isotypical component of $V_r\bigotimess \C$ and then we put $(V_r\bigotimess \C)^{i,n-i}:=\sum _{{\gta}({\grh})=i}V_{r,{\grh}}$. We then put a Hodge structure on $V$ by forcing $V=\bigoplus V_i\bigotimess _{End(V_i)}Hom(V_i,V)$ to be an isomorphism of Hodge structures. By construction $T$ acts as morphisms of Hodge structures. However, the Hodge structure itself depends only on the action of $Z(A_\Q(M))$ on $V$ so that an alternative method of construction is to start with the set $\symb{Irr}(M)$ with its action of $Gal(\bar\Q/\Q)$, let $K$ be the associated {\'e}tale \Q-algebra, let $V$ be the $K$-module of dimension specified by $M$ and then let $(V\bigotimess \C)^{i,n-i}:=\sum _{{\gta}({\grh})=i}V_{\grh}$, where {\grh} runs over the \Q-algebra homomorphisms $K\to \C$. In this way it is seen that $V$ as a rational Hodge structure depends only on the $Gal(\bQ/\Q)$-set $\symb{Irr}(M)$ and two functions ${\gta}\co \symb{Irr}(M)\to n+1$ and $\dim\co \symb{Irr}(M)\to \N$, where $\dim$ is defined by $\dim(n)=dim_{Z(A_\C(n))}N$ for a representative of $N$ (\C\ can of course be replaced by any algebraically closed field). If \k\ is a subfield of \C\ such that the action of $Gal(\bk/\k)$ on $\symb{Irr}(M)$ preserves the fibers of {\gta} then for a choice of descent of the Hodge filtration on each $V_r\bigotimess \C$ to $V_r\bigotimess k$ for each we get a descent of the Hodge filtration on $V\bigotimess \C$ again by forcing the isomorphism above to preserve the descent. If $p$ is a prime such that $A(M)\bigotimess \Z_{(p)}$ is finitely generated as $\Z_{(p)}$-module and separable we associate, in a similar way, an $F$-crystal to $(M,{\gta})$: Suppose that the action of $Gal(\bk/\k)$ on $\symb{Irr}(M)$ preserves the fibers of {\gta}. We know that $M$ is realisable over $\Q_p$ and there is, up to isomorphism, a unique $A(M)\bigotimess \Z_p$-lattice $V$ such that $V\bigotimess \Q$ is such a realisation. We also get analogous $V_r$. Further, $V_r\bigotimess _{\Z_p}\W(k)$ is the sum of its isotypical components $(V_r\bigotimess _{\Z_p}\W( k))_{\grh}$. The {\gsi}-linear isomorphism $1\bigotimess {\gsi}$ takes $(V_r\bigotimess _{\Z_p}\W(k))_{\grh}$ to $(V_r\bigotimess _{\Z_p}\W( k))_{{\gsi}({\grh})}$ and we define the structure of an F-crystal on $V_r\bigotimess _{\Z_p}\W(k)$ by $F=p^{{\gta}(t)}(1\bigotimess {\gsi})\co (V_r\bigotimess _{\Z_p}\W(k))_{\grh}\to (V_r\bigotimess _{\Z_p}\W(k))_{{\gsi}({\grh})}$. The $F$-crystal structure on $V\bigotimess \W(k)$ is constructed as before. Again $T$ acts by endomorphisms and there is an alternative way of constructing the $F$-crystal if one is prepared to forget the $T$-action. Indeed, consider the set $\symb{Irr}(M)$ with its action of {\gsi} and the two functions {\gta} and $\dim$. We let $R$ be the set containing for each $n\rum\@a\rum \symb{Irr}(M)\dim(n)$ copies of $n$ with {\gsi} and {\gta} extended in the obvious way. We then consider $\W(\bk)[R]$, the free $\W(\bk)$-module on $R$, and define the Frobenius map by $F{\ninerm[} r{\ninerm]}=p^{{\gta}(r)}{\ninerm[} {\gsi}(r){\ninerm]}$. Also if we have for each $V_i$ an $End(V_i)$-representation {\grh} of $Gal(\bk/\k)$ on $V_i$, then we can twist by this by letting $F$ act by $p^{{\gta}(t)}({\grh}\bigotimess {\gsi})$. Finally, we would like to associate to $(M,{\gta})$ the \l-adic analogue of this, that is a Hecke character. We will be able to associate to our data the algebraic part of a Hecke character but a problem arises as there is no canonical choice for a Hecke character with a given algebraic part, indeed such a character may exist only after an extension of the coefficient field. This will have as a consequence that our description of the \l-adic cohomology will not be as satisfactory as the description of the Hodge structure or $F$-crystal of a variety of CM-type. In case $[H_n(X)]$ is multiplicity free we will be able to do better however. In any case the algebraic part (cf.~\[De:5.3]) can be associated to our data as follows. Assume that for any embedding of \bQ\ in \C\ we have ${\gta}({\gio}({\grh}))=n-{\gta}({\grh})$ as above for the corresponding {\gio}. Assume also that \k\ is a number field for which the action of $Gal(\bQ/K)$ on $\symb{Irr}(M)$ stabilises the fibers of {\gta}. If again $K$ is the \Q-algebra corresponding to $\symb{Irr}(M)$, then we can define a multiplicative map $$ \eqalign{ k^\times &\;\longrightarrow\;\; K^\times \cr {\gla}&\mapsto \prod _{{\grh}\rum\@a\rum \symb{Irr}(M)}{\grh}(N_{k/\Q}({\gla})) ^{{\gta}({\grh})}.\cr} $$ By the assumption on \k\ this is well-defined and by the assumption on {\gta} the projection onto each simple factor of $K$ fulfills the conditions for being the algebraic part of a Hecke character of weight $n$ so we obtain in this way a set of algebraic parts of Hecke characters. \begin{theorem}2.4. Let $(X,S)$ be of separable CM-type in degree $n$. i) If $k\subseteq \C$ then $H^n_{sing}(X(k),\Q)$ is isomorphic as a Hodge structure with $S$-action to the one associated to the type of $(X,S)$ with a descent of the Hodge filtration to \k\ of the sort described. ii) If $p>0$ then $H^n(X,p)$ is isomorphic as $F$-crystal to the $F$-crystal associated to the type of $(X,S)$ and a representation of $Gal(\bk/\k)$. iii) After a finite extension of \k\ the $Gal(\bk/\k)$-representation on $H_n(X,r)$, $(r\ne p)$ factors through the Galois group of the algebraic closure $K$ of the prime field in (the finite extension of) \k. If $p=0$ this representation is given, on the Galois group of a finite extension of $K$, by a direct sum of algebraic Hecke characters whose algebraic parts are the ones associated to the type of $(X,S)$ and with multiplicities given by $\dim$. iv) If $\symb{Irr}(M)$ is multiplicity free (i.e.~every irreducible representation of $S$ occurs at most once in $\symb{Irr}(M)$) and \k\ is a number field then the $Gal$(\bk/\k)-representation on $H_n(X,r)$ is given by a direct sum of algebraic Hecke characters with values in the simple components of $Z(A_\Q(\symb{Irr}(M)))$. \pro To begin with let $R$ be \Q, $\Z_p$ resp.~$\Q_r$. Then $S$ generates an $R$-subalgebra $B$ of the algebra of endomorphisms of $H^n_{sing}(X(\k),\Q)$, $H^n(X,p)$ resp.~$H^n(X,r)\bigotimess _{\Z_r}\Q_r$ such that $$ B/(\hbox{\rm maximal nilpotent ideal})=A^n\bigotimess _\Z R $$ (the $A^n$ being that of \(2.1)). By \[C-R:Thm. 72.19] $B\to A^n\bigotimess R$ splits as an algebra map and so we can assume that $A\bigotimess R$ acts on $H^n_{sing}(X(\k),\Q)$, $H^n(X,p)$ resp. $H^n(X,r)\bigotimess _{\Z_r}\Q_r$. Using \(1.1) we reduce to the case when $A\bigotimess R$ is a division algebra or, in the case of ii), isomorphic to $W(\F)$ for a finite field \F. The proof of i) is then easy: By the comparison theorem $H^n_{sing}(X(\k),\Q)$ is a representative of $[H^n(X)]$ and as the Hodge decomposition on $H^n_{sing}(X(\k),\Q)\bigotimess _\Q\C$ is stable under $A\bigotimess \C$, the assumption of CM-type forces the Hodge decomposition to be obtained as the lumping together of isotypical components. As for ii), the fact that $A\bigotimess \W(k)$ is separable implies that we have a unique isotypical decomposition $H^n(X,p)=\bigoplus M_{\grh}$, where {\grh}\ runs over the irreducible $A\bigotimess K$-modules ($K:=\W(\k)\bigotimess _\Z\Q$) occurring in $H^n(X,p)\bigotimess \Q$. As $F$ is \gSi-linear and commutes with $A$ it maps $M_{\grh}$ to $M_{\gSi({\grh})}$. If $(p^{n_1},p^{n_2},\ldots,p^{n_k})$ are the elementary divisors of the linear mapping $F\co M_{\grh}\to\gSi_* M_{\gSi({\grh}))}$, the characterisation of the Hodge filtration recalled at the beginning of this section shows that $M_{\grh}/pM_{\grh}$ is non-disjoint from \hod X{n-i}ik exactly when $i$ equals some $n_j$. Hence, by assumption, all the $n_j$ equal ${\gta}({\gLa})$ where {\gLa} is any component of $[H_n(X)]$ which occurs in ${\grh}\bigotimess _K\bar K$. We will denote, by abuse, this common value ${\gta}({\grh})$. Thus $F\co M_{\grh}\to\gSi_* M_{\gSi({\grh})}$ is p$^{{\gta}({\gLa})}$ times an isomorphism. Hence if we define $F'\co H^n(X,p)\to H^n(X,p)$ as being $p^{-{\gta}({\grh})}F$ on $M_{\grh}$, $H_n(X,p)$ becomes a unit root crystal and is hence described by the $Gal$(\bk/\k)-representation on the fixed points of $F$ (over \bk). This action commutes with $A\bigotimess _\Z{\Z_p}$ and so gives the desired description. Let us now turn to iii) and let us begin with the case $p=0$. We may assume that \k\ is a finitely generated field. We have a representation ${\gph}\co Gal(\bar k/k)\to Aut_{Z\bigotimess \Q_r}(H^n(X,r)\bigotimess \Q)$ and after possibly enlarging $Z$ ($:=Z(\symb{Irr}(M))$) and \k\ we may assume that there exists an algebraic Hecke character $I_m(k')\to Z^\times $, where $k'$ is the algebraic closure of \Q\ in \k, whose algebraic part is the one coming from the type of $(X,S)$ (using that the condition on {\gta} is fulfilled by the transcendental theory). Twist {\gph} by this character, considered as a character of $Gal(\bk/\k)$ through the morphism $Gal(\bk/\k)\to Gal(\bk'/\k')$ and (\[Se:II,2.7]). What we now need to prove is that this twist ${\gph}'$ has finite image (cf.~\[De:Thm.~5.10]). Let us first show that if $\gSi\rum\@a\rum Gal(\bar k/k)$ then ${\gph}'(\gSi)$ is quasi-unipotent. The possible orders for the eigenvalues of a quasi-unipotent matrix over $\Q_r$ of given order is bounded as the degrees of the extensions of $\Q_r$ obtained by adjoining an $m$th root of unity goes to infinity with $m$. It is therefore enough to verify the quasi-unipotence on a dense set of Frobenius elements. By the \v Ceboratev density theorem it suffices to check quasi-unipotence for the Frobenius elements corresponding to maximal ideals for some thickening of \Sp\k\ over which $A_n$ is separable, $X$ is smooth and $$ b_n(X)=\sum _{i+j=n}dim_h\hod Xijh, $$ where $h$ is the residue field. If $F_m$ is the Frobenius element of $Gal(\bk/\k)$ we then want to show that all the eigenvalues of ${\gph}'(F_m)$ are roots of unity or, as they are all algebraic numbers, that all their absolute values are equal to 1. For the infinite primes we use the Riemann hypothesis for $X$. At finite places away from $q:=\symb{char} h$ there is no problem. Let us therefore consider the places over $q$. Pick a place $v$ of $Z$ lying over $q$ normalised so that $v(|h|)=1$. By definition $v({\gph}(F_ m))$ equals the average of {\gta} over the orbit of $v$ of the action of \gSi\ on $\symb{Irr}([H^n(X)])$, where $v$ is seen as a homomorphism $Z\to L_q$ and thus giving an element of $\symb{Irr}([H^n(X)])$ (recall that $L_q$ is an algebraic closure of $\Q_q$). Hence we want to show that any eigenvalue of the action of $F_m$ on the $v$-isotypical part of $H^n(X,q)$ has the same valuation. By construction $(X_h,S)$ is of separable CM-type in degree $n$ and the eigenvalues of $F_m$ are of course the same for $X_k$ and $X_h$ so we may replace $k$ by $h$. Applying \(1.5) to $S$ and $\{F_h\}$ we see, as $F_m=F^*_h$ on $H^n(X,r)$, that the eigenvalues of $F_m$ on the $v$-isotypical part of of $H^n(X,r)$ are the same as the the eigenvalues of $F^*_h$ on $H^n(X,q)$. By ii), if $u$ is the length of the \gSi-orbit of $v$ then $F^u$ is divisible exactly by $p^t$ on the $v$-isotypical component of $H^n(X,q)$, where $t$ is the sum of the values of {\gta} over the \gSi-orbit of $v$. As $F^*_h=F^r$, where $|h|=p^r$, we immediately get what we want. Now again as the orders of the eigenvalues of the elements of $Gal$(\bk/\k) are bounded after replacing \k\ by a finite extension we may assume that the image of ${\gph}'$ consists entirely of unipotent matrices and so by Engel's theorem ${\gph}'$ is a unipotent representation. We aim to show that it is in fact trivial. As $G:={\gph}'(Gal(\bar k/k))$ is a compact $r$-adic Lie group the closed subgroup of $G$ generated by $r$th powers is of finite index in $G$ and by the Frattini lemma any closed subgroup mapping surjectively onto the quotient of $G$ by this subgroup equals all of G. We may then apply the Hilbert irreducibility theorem to get a number field specialisation $\k''$ of \k\ such that $X$ has good reduction at $\k''$ and that the composed map $Gal(\bk''/\k'')\to Gal(\bk/\k)\to G$ is surjective. Hence we may assume that \k\ is a number field. I claim that for each prime of \k\ over $r$, the inertia group of that prime has finite image in $G$. Indeed, ${\gph}'$ is Hodge-Tate as an algebraic Hecke character is always Hodge-Tate and by \[Fa]. The finiteness then follows from (\[Se1:1.4,Cor.~3]) as the unipotence implies that the Hodge-Tate weight is zero. As ${\gph}'$ is unipotent this implies that ${\gph}'$ is unramified over $r$. The other monodromy groups automatically have finite, and therefore trivial, images. The triviality of $G$ then follows from the finiteness of the Hilbert class field of \k. We have therefore proved iii) when $p=0$. The case $p>0$ is similar up to the point where we have arrived at a unipotent representation. Any homomorphism $Gal(S,\bar s)\to \Z_r$ for a finitely generated $\F_p$-scheme $S$ is geometrically trivial, by \[K-L:Thm~1], so by thickening \k\ we finish. As for iv) we start as above so that we have an action of $A:=A^n\bigotimess \Q$ on $H^n(X)$ by correspondences. Note that the assumption of multiplicity freeness implies that the commutant of $A$ in $End(H^n(X,r))$ equals $Z\bigotimess \Q_r$. Let $v$ be a place of \k\ at which $X$ has good reduction with fiber $X_v$ over the residue field $\F_v$. Apply the construction of a semi-simple algebra of correspondences to all correspondences so as to get $B$. Then $B$ contains the Frobenius correspondence in its center as well as the subalgebra $A$. Let $C$ be the commutant of $A$ in $B$ so that $B=AC$. By the observation just made $C\bigotimess \Q_r=Z\bigotimess \Q_r\subseteq B\bigotimess \Q_r$ and so $C\subseteq B$ and therefore $A=B$. Hence the Frobenius correspondence $F_v$ lies in $Z$. Therefore we have associated to every place $v$ of \k\ outside a finite set an element $F_v$ of $Z$. Extending by multiplicativity we get a homomorphism $I_m(k)\to Z^\times$ for a suitable $m$. As in the proof of iii) we show that the projections onto the simple factors of $Z$ are algebraic Hecke characters with algebraic parts given by the type of $(X,S)$. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{theorem} \begin{remark} i) It is probably true that in ii) we also get the conclusion that $[H^n(X,p)]$ is geometrically constant. This would follow from a good theory of over-convergent $F$-crystals. ii) Can one find a good extension of iii) that would contain iv) as a special case? iii) An example showing that there are problems in the \l-adic case is obtained as follows. Pick an imaginary quadratic field $K$ with class number greater than one. There is an elliptic curve $E$ with complex multiplication by the ring of integers $R$ of $K$ defined over the Hilbert class field $H$ of K. The pair ($R_{H/K}E$,R) is of CM-type in degree 1 over $K$ yet there is no algebraic Hecke character whose algebraic part is that obtained from the type of ($R_{H/K}E$,R). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{remark} As will come as no particular surprise, for abelian varieties our notion coincides with the traditional one. \begin{proposition}2.4. Suppose $(X,S)$ is of separable CM-type in degree 1. Then its Albanese variety is of CM-type in the usual sense possibly after a finite extension of \k. \pro This follows from \(2.4) and the Tate conjecture for homomorphisms between abelian varieties. Another proof is for $p=0$ to note that \(2.3:i) says that the degree 1 Hodge structure of $X$ is visibly of CM-type and for $p>0$ that $(Alb X,S)$ is rigid as by definition $Hom_S(\hod X01k,\hod X10k)$ is equal to 0 and so after a finite extension of \k, $Alb\, X$ can be defined over a finite field and is hence of CM-type. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{proposition} \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{section} \begin{section}Hereditary CM-type \mdef\End{\symb{End}\,} Theorem 2.3 suffers somewhat on the $p$-adic side as the very natural example of $(E,\End E)$ where $E$ is a supersingular elliptic curve is not of separable CM-type; $\End E$ is not separable at $p$. It is possible to give a result which in that case specialises to a satisfactory answer. In this section we will give a generalisation of the previous results that will cover this case. The maximal possible generality would seem to be to assume that $A^n\bigotimess \Z_{(p)}$ should be a {\deffont hereditary} order which means that any $A^n$-splitting of crystalline cohomology tensored with \Q\ comes from an $A^n$-splitting of crystalline cohomology itself. Let us recall that an order is hereditary if each lattice over it is projective. \begin{remark} The meaning of the term differs somewhat in various areas of the literature as hereditary sometimes means just that a submodule of a projective module is projective. The definition used here means that the base extension of the order to the fraction field of its base ring is semi-simple together with the fact that every sub-module of a projective module is projective. We will want this extra condition and hence adopt the current definition (which is to be found for instance in \[Re]). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{remark} On the other hand, the example of an automorphism of order $p$ acting (non-trivially) on a curve of genus $(p-1)/2$ shows that the condition that different Hodge pieces be disjoint is not reasonable as the cyclic group of order $p$ has only one irreducible representation mod $p$. The situation will no longer be as simple as in the separable case. It is still true that one to any irreducible $A^n\bigotimess \k$-module can associate an irreducible $A^n\bigotimess K$-module but this map is no longer injective (though surjective). We will use \[Re:Ch.~9] as a general reference to the theory of hereditary orders. For the reader's convenience we repeat the salient facts in the following proposition as well as adding a result -- a weak version of the elementary divisor theorem -- which is not to be found in \[loc.~cit.] (but no doubt is not new). \begin{proposition}eldiv Let $A$ be a hereditary order over a henselian discrete valuation ring $R$ with fraction field $K$. i) Any submodule of an $A$-lattice is a submodule of finite colength of a direct factor of the lattice. ii) In every indecomposable $A$-lattice there is exactly one submodule of a given colength. iii) If $M$ is an indecomposable $A$-lattice and $M\hookrightarrow N_i$ two inclusions of finite colength. Then one of these inclusions is contained in the other. iv) Every indecomposable finitely generated torsion $A$-module is a quotient of an indecomposable $A$-lattice. v) Let $M$ be an $A$-lattice and $N$ a sub-lattice of it. Then there is a decomposition of $M$ as a direct sum of indecomposable submodules whose intersection with $N$ also gives a decomposition of $N$ into a direct sum of indecomposable submodules. \pro For i) we take the saturation of the submodule. The quotient of the lattice by that saturation is torsion-free and hence projective and the saturation is therefore a direct factor. For ii) we notice that by i) any submodule of the lattice is also indecomposable so we may assume by induction that the given colength is 1. However, the lattice being projective is the projective hull of its co-socle (the maximal semi-simple quotient) and so being indecomposable the co-socle is simple which means that there is a unique submodule of colength 1; the radical. For iii) we note that $M\hookrightarrow N_i$ are included in a common inclusion of finite colength (being of finite colength). We then apply ii). As for iv) we use induction on the length of the module $M$. We therefore find a simple quotient $S$ of $M$ and apply the induction hypothesis to the kernel $M'$ of this map. We will temporarily (and improperly) call a torsion quotient of an indecomposable lattice a {\it cyclic} module. Thus we may assume that $M$ is an extension of a sum of cyclic modules by the simple module $S$. This extension is the sum, as extension, of the extension of the cyclic summands by $S$. Let us first study the latter extensions and let us denote by $P$ the projective hull of $S$, by $Q$ its radical, by $V$ the cyclic summand and by $R$ its projective hull. Then every extension of $V$ by $S$ comes from pushout by a map from $Q$ to $V$, the same extensions being obtained if the difference of two morphism extends to a map from $P$ to $S$. Now, I claim that all non-surjective maps $Q\to S$ so extend. In fact the map lifts to a map $Q\to R$ which necessarily is injective as $Q$ is indecomposable. If the map $Q\to S$ is not surjective then the map $Q\to R$ is neither. By applying iii) we see that $P$ must be isomorphic to the unique submodule of $R$ containing $Q$ as a submodule of colength 1 and thus the original map lifts to $P$. This result shows that if $V$ is not a quotient of $Q$ then any extension of $V$ by $S$ is trivial and if it is, then the group of extensions can be identified with maps from the co-socle of $Q$ to the co-socle of $V$ which are isomorphic simple modules. We will now show that, after possibly changing the direct sum decomposition of $M'$ we may assume that all but one of the extension classes of direct summands by $S$ are trivial. This will clearly show iv). For this we may immediately discard summands of $M'$ which are not quotients of $Q$ as their extension classes have just been shown to be trivial. Furthermore, we may use induction on the number of non-trivial extension classes. Note now that if $V_1$ and $V_2$ are summands then for any map \pil\phi{V_1}{V_2} we may consider the automorphism of $V$ which maps $v\rum\@a\rum V_1$ to $v+\phi v$ and acts as the identity on all other factors. If $e_i$ are the extension classes then all of them but $e_2$ are unchanged and $e_2$ is changed into $e_2+\phi_*e_1$. If we identify extension classes of $V_i$ with homomorphisms from the co-socle of $Q$ to that of $V_i$ then $\phi_*$ is just composition by the map on co-socles induced by $\phi$. As $V_1$ and $V_2$ are both both quotients of $Q$, by ii) on is a quotient by the other and we may assume that $V_2$ is a quotient of $V_1$. In that case, any endomorphism of $Q$ induces a morphism $V_1\to V_2$ and by the projetivity of $Q$, any map from the co-socle of $V_1$ to that of $V_2$ is induced by an endomorphism of $Q$. Putting this together we see that any extension class is of the form $\phi_*e_1$ so that the we may choose $\phi$ so that $\phi_*e_1=-e_2$ which allows us to decrease the number of non-zero extension classes. To finally prove v) we consider the module $M/N$. This is a direct sum of a lattice and a torsion module and the lattice may be split off from $M$ without changing $N$. Thus we may assume that $M/N$ is torsion. We then use iv) to write that quotient as a direct sum of quotients of indecomposable projective modules. The sum of the projective hulls of each summand is a projective hull of the sum. That projective hull is a direct summand of the map $M\to M/N$. This immediately gives the pair $(M,N)$ as a direct sum of of pairs $(P_i,P'_i)$, where $P_i$ is indecomposable and a factor $(M',M')$. As $M'$ is a sum of indecomposables, v) follows. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{proposition} We will need a definition which is very special to the situation at hand. \begin{definition} Let $A$ be a hereditary order over a henselian mixed characteristic discrete valuation ring $R$ with positive residue field characteristic $p$ and let $M$ be a finitely generated torsion module killed by $p$. By the \definition{complementary module} to $M$ we mean the torsion module (defined up to isomorphism) obtained as $P/pP'$, where $P$ is a projective hull of $M$ and $P'$ is the kernel of the natural map $P\to M$. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition} As there can be, as opposed to the separable case, non-trivial extensions of modules we also will need to recall the definition of block, well-known in the theory of general orders, \begin{definition} Let $A$ be a hereditary order over a henselian discrete valuation ring $R$ with fraction field $K$. Two indecomposable (finitely generated) $A$-modules belong to the same block if there is a non-zero morphism from the projective hull of one to the other. This is equivalent to the two hulls tensored with $K$ being isomorphic. If $M$ is a finitely generated $A$-module then the \definition{$B$-component}of $M$ is the sum of all indecomposable factors belonging to the block $B$. (It is clear that any finitely generated $A$-module is the direct sum of its components associated to different blocks.) \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition} What is different with the hereditary case as opposed to the separable case is that we may have non-semisimple (f.g.) modules killed by $p$. This will imply that to define CM-type it is not enough to look at what simple modules occur in which Hodge piece; the more precise module structure needs to be taken into account. As we will see this forces certain relations between Hodge pieces. Our results will be purely algebraic so we will, rather than sticking to the notation of this article as a whole, use the following notation: $A$ will be a hereditary $\Z_p$-order and $M$ will be an $F$-crystal with an action of $A$. We define the Hodge filtration on $M/pM$ by $M^i:=F^{-1}p^iM/pM$ and the Hodge modules $H^i:=M^i/M^{i+1}$ (which may be considered as $A\bigotimes \W$-modules). \begin{definition-lemma} We define the \definition{$A$-primitive part} of $H^i$ as the direct factor (defined up to isomorphism only) by induction on $i$. For $i=0$ we let the primitive part be all of $H^i$. For $i>0$ the complement of the primitive part of $H^{i-1}$ is a direct factor of $H^i$ and we let the primitive part be a complementary factor of it. \pro What is to be proven is the statement about the complement of the primitive part being a direct summand. We will give another description of the primitive part which will make this obvious. Consider therefore the Frobenius map as a $\W$-linear map $\sigma^*M\to M$, which then also is a $A\bigotimess \W$-linear. This is a hereditary order so we may by \(eldiv:v) split this map up in indecomposable factors. Using Mazur-Ogus' characterisation (\[B-Og]) of the Hodge filtration and the fact that submodules are linearly ordered we immediately see that each indecomposable factor will contribute a cyclic module to one Hodge piece and its complement to the next. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition-lemma} We are now ready to define what we mean by CM-type in the context of actions of hereditary orders. \begin{definition} The pair $(M,A)$ is of \definition{hereditary CM-type} if $A\bigotimess \Z_p$ is a hereditary order and for each block $B$, the $B$-component of the primitive part of $H^i$ is non-zero for at most one $i$ and all indecomposable factors of that $B$-component have the same length. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{definition} We have now set up our definitions so that we may carry through the same analysis as in the separable case (it should be noted that in the case that $A^n\bigotimess \Z_{(p)}$ is actually separable this definition coincides with the previous one). \begin{lemma}galois Let $B$ be a $\W$-algebra, finitely generated and free as a $\W$-module. Suppose $n$ is a positive integer and $T$ a $\sigma^n$-linear automorphism of $B$. Using $T$ to get a \Z-action on $B$ we have that $H^1(\Z,B^\times)=*$. \pro If $B'$ is the $\Z_p$-algebra of $T$-fixpoints then we have that $B=B'\bigotimes \W$ and an element of $H^1(\Z,B^*)$ is given by an automorphism class of a finitely generated right $B'$-module whose extension of scalars to $\W$ is isomorphic as $B$-module to $B$ itself. As the extension $\Z_p\to \W$ is faithfully flat this means that such a $B'$-module is projective. Hence it is determined up to isomorphism by its co-socle and to prove the lemma it is enough to show that if we have two semi-simple $B'$-modules which become isomorphic under extension of scalars to $\W$ are isomorphic. This however is obvious (using for instance the independence of central characters of a semi-simple algebra). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{lemma} \begin{theorem} Suppose $(M,A)$ is of hereditary CM-type and that {\k} is algebraically closed. Then it is determined up to isomorphism by which blocks appear in the primitive part of which Hodge-modules and the common length of indecomposable factors of each such block. \pro Note first that we can make {\gsi} act on the blocks of $A\bigotimes \W$ by the condition that $N$ belongs to the block $B$ iff $\sigma^*N$ belongs to $\sigma^*B$. If we now split up $M$ in blocks, $M=\bigoplus_B M_B$, then it is clear that $F$, considered as a map $\sigma^*M\to M$ is a sum of maps $\sigma^*M_B\to M_{\sigma^*B}$. Now, for an indecomposable $B\bigotimes W$-lattice $N$ the length of $N/pN$ only depends on which block $N$ belongs to. Indeed, any two indecomposable $B\bigotimes W$-lattices in the same block are contained in each other with quotient of finite length, the kernel and cokernel of the map induced by reduction modulo $p$ then has the same length. We now consider the component of $F$, $\sigma^*M_B\to M_{\sigma^*B}$, as a $W$-linear map and split it up into indecomposable pieces according to lemma \ref{eldiv}. Looking at each indecomposable piece we see that if $B$ appears in the $i$'th Hodge piece of $M$ then $F$ maps $\sigma^*M_B$ into $p^iM_{\sigma^*B}$, the image contains $p^{i+1}M_{\sigma^*B}$ and the length of each indecomposable factor of $p^iM_B/M'$, $M'$ being the image, has the same length (as each such length added to the common length of the indecomposables of the primitive $B$-part of the Hodge piece adds up to the common length of an indecomposable lattice in $B$ modulo $p$). This means that any indecomposable factor of $M_{\sigma^*B}/M'$ has the same length, which is the same as saying that $M'=\symb{rad}^mM_{\sigma^*B}$ for a suitable $m$, where $\symb{rad}(-)$ is the radical functor. We may thus use $F$ to identify $\sigma^*M_B$ with $\symb{rad}M_{\sigma^*B}$, where $m$ is determined by $i$ and the common length of indecomposables of the primitive part belonging to the block $B$. If $n$ is the smallest positive integer for which $\sigma^{n*}B=B$, then $F^n$ maps $M_B$ onto $\symb{rad}^kM_B$ for a suitable $k$. It is then enough to show that all such maps are conjugate under automorphisms of $M_B$. Fix one such map {\gph}. Now, I claim that the relation $\phi\circ f^\sigma=g\circ\phi$ defines an automorphism $f\mapsto g$ of $\End(M_{\sigma^*B})$. Indeed, for any $g$ there is an $f$ fulfilling that relation as the image of {\gph} is equal to $\symb{rad}^mM_{\sigma^*B}$. Conversely, the inverse image of $M_{\sigma^*B}$ in $\sigma^*M_B\bigotimes K$ under {\gph} is equal to its largest sub-lattice for which the quotient by $\sigma^*M_B$ has all its indecomposable components of length less than or equal to $m$ which shows that to any $f$ there is a $g$. As any map fulfilling the conditions imposed on {\gph} differs from it by an automorphism of $\sigma^*M_B$ we can apply lemma \ref{galois} to include that there is, up to isomorphism, only one $F$. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{theorem} I would also like to record that, just as in the separable case, multiplicity freeness implies CM-type. \begin{proposition} Suppose that $(M,A)$ is multiplicity free in the sense that an irreducible $A\bigotimes K$-module appears at most once in $M\bigotimes K$. Then $(M,A)$ is of hereditary CM-type. \pro The condition implies that for any block, the component of $M$ in that block is indecomposable. That immediately implies that a given block appears in the primitive part of just a single Hodge piece and that part is indecomposable so the condition on length is fulfilled. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{proposition} We finish this section with some examples. \begin{example} i) Consider a supersingular elliptic curve $E$ and its ring of endomorphisms $A$, which is a maximal order in a division ring and hence hereditary. The action of $A$ on the first crystalline cohomology group is multiplicity free and hence of hereditary CM-type. More precisely, $H^1(E,p)$ is an indecomposable $A\bigotimes W$-lattice and $\sigma^*H^1(E,p)$ is the other indecomposable $A\bigotimes W$-lattice -- both are in the same block. Hence, $H^0(E,\Omega^1)$ is one irreducible $A\bigotimes W$-module and $H^1(E,\Cal O_E)$ the other. The image of $\sigma^*H^1(E,p)$ under $F$ is the maximal proper submodule of $H^1(E,p)$. ii) Let $C$ be the projective, smooth completion of the curve $y^p-y=x^2$, $p\ne2$, and consider the action of $\Z/p$ given by $y\mapsto y+\alpha$. The action of the group algebra of $\Z/p$ on $H^1(C,p)$ factors through the quotient $A$ that is the ring of $p$'th roots of unity. $H^1(C,p)$ is then a free $A\bigotimes \W$-module of rank 1 and hence $(C,A)$ is of hereditary CM-type. This time the situation is simpler as the ring is commutative and to prove the classification theorem we could simply have divided $F$ by $(\zeta-1)^{(p-1)/2}$ to obtain a unit root crystal. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{example} \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{section} \begin{section}Applications to the N{\'e}ron-Severi group. In this section we will suppose that $X$ is a surface and that $(X,S)$ is of separable type in degree 2. Then $S$ acts on the N{\'e}ron-Severi group $NS$ of $X_{\bar k}$. If $p=0$, \(2.4) and the Lefschetz theorem on $(1,1)$-classes show that $[NS\bigotimess \bar\Q]\rum\@a\rum K(S,\bQ)$ equals the sum of all irreducible {\grh}\ in $[H^2(X)]$ such that ${\gta}({\gsi}({\grh}))=1$ for all ${\gsi}\rum\@a\rum Gal(\bar\Q/\Q)$ and the Tate conjectures implies this in all characteristics. However, \(2.3) can be used to obtain further information on $NS$. To illustrate this let us suppose that $p>0$ and that $\mathop{\rm rk}\nolimits NS=b_2(X)$. By possibly extending \k\ we may assume that $NS$ is defined over \k. As $c_1\co NS\bigotimess \Z_\ell\to H^2(X_{\bk},\Z_\ell)$ ($\ell\ne p$) has torsion free cokernel (cf.~\[Gro:8.7]) and the two modules have the same rank, $c_1$ is an isomorphism. By Poincar{\'e} duality the intersection pairing is perfect at \l. By \[Ill:II,5.8.5,5.20] the image of $c_1\co NS\bigotimess \W(k)\to H^2(X,p)$ is the largest sub-$F$-crystal in which $F$ is divisible by $p$ and, again by Poincar{\'e} duality, if ${\gsi}_0$ is the $\W(\k)$-length of the cokernel, then $p^{2{\gsi}_0}$ is the exact power of $p$ dividing $disc(NS)$. Hence by the Hodge index theorem $disc(NS)=(-1)^{b_2-1}p^{2{\gsi}_0}$. As the whole $H^2(X,p)$ is determined by the type of $(X,S)$, ${\gsi}_0$ is as well and we will now see how this can be done explicitly. By \(2.3) $M:=H^2(X,p)=\bigoplus _{{\grh}\rum\@a\rum \symb{Irr}([H^2(X)])}M_{\grh}$ and $F\co M_{\grh}\to M_{{\gsi}({\grh})}$ is $p^{{\gta}({\grh})}$ times an isomorphism. Let $N\subseteq M$ be the maximal sub-$F$-crystal on which $F$ is divisible by $p$. Consider $T:=\symb{Irr}([H^2(X)])$ with the functions {\gta} and $\dim$ and the action of {\gsi}. We shall now describe an algorithm for computing ${\gsi}_0$. To do this we start by by considering $M$ with $F':=p^{-1}F$ as a virtual $F$-crystal i.e.~$p^{-1}F$ takes $M$ into $M\bigotimess \Q$ rather than into $M$ itself. Now $N$ can then be characterised as the maximal sub-$F$-virtual-crystal which is actually a crystal. As it is unique it is a sub-representation and so it is the direct sum of the $N_{\grh}$. We will now concentrate on one specific {\gsi}-orbit on $T$ and assume that $M$ is in fact the $F$-crystal associated to it. Pick one {\grh} in this orbit. As $M_{\grh}$ is of rank 1 $N_{\grh}$ is equal to $p_nM_{\grh}$ for some $n$. All powers of $F'$ must take $N_{\grh}$ to $M$ which means that $n+\sum_{j=0}^k({\gta}({\gsi}^j{\grh})-1)$ is greater than or equal to 0 for all $k$. Hence if we put $n$ equal to $-\min_k\sum _{j=0}^k({\gta}({\gsi}^j{\grh})-1)$ and define $N'$ as $\bigoplus p^{n_{{\grh}'}}M$, where $m_{{\grh}'}:=n+\sum _{j=0}^k({\gta}({\gsi}^j{\grh})-1)$ with ${\grh}'={\gsi}^k{\grh}$ we have a sub-$F$-virtual-crystal of $M$ which clearly is an actual $F$-crystal (here we use the fact that the sum of ${\gta}-1$ over the orbit is 0). We have also seen that $N_{\grh}\subseteq N'_{\grh}$ and as $N$ is the maximal sub-$F$-crystal we have equality. Finally, again using that the sum of ${\gta}-1$ over the orbit is 0 it is immediately realised that $N'$ is independent of the choice {\grh} and so has to be equal to $N$. In particular we see that the contribution of this orbit to ${\gsi}_0$ equals the multiplicity of the orbit times the sum of the $m_{{\grh}'}$. \begin{example} We consider one orbit for {\gsi} and describe such an orbit by $({\gta}(t),{\gta}({\gsi}(t)),\ldots,{\gta}({\gsi}^{h-1}(t)))$, where $h$ is the length of the orbit. We also assume that the starting point {\grh} is the first element of this list. \noindent i)\ (0,2) gives partial sums $(-1,0)$ and so $n=1$ and the list of the $m_{{\grh}'}$ is $(0,1)$ and finally the contribution to ${\gsi}_0$ is 1. \noindent ii)\ $(0,2,1)$ gives partial sums $(-1,0,0)$, $m$s $(0,1,1)$ and ${\gsi}_0=2$. \noindent iii)\ $(2,1,1,1,1,0)$ gives partial sums $(1,1,1,1,1,0)$, $m$s $(1,1,1,1,1,0)$ and so ${\gsi}_0=5$. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{example} As a geometric example let us first consider the Fermat surface $X_m= \{X^m_0+X^m_1+X^m_2+X^m_3=0\}$ and the group of diagonal automorphisms $G_m=\mu^4_m/(scalars)$. The irreducible representations of this group are the elements of the dual group $$ \check G_m:=\{(b_0,b_1,b_2,b_3)\rum\@a\rum(\Z/m)^4:\sum _{i=0}^3b_i=0\} $$ and it is well known (cf.~\[Ka:Sect.~6]) that each character occurs at most once in $H^2(X_m)$ and those that occur are exactly those in the set $T:=\{(b_0,b_1,b_2,b_3)\rum\@a\rum G:\forall i:i\ne0\}\cup\{(0,0,0,0)\}$. Furthermore, if we for $b\rum\@a\rum\Z/m$ let $\langle b\rangle$ be the unique integer s.t.~$\langle b\rangle\rum\@a\rum b$ and $0\le \langle b\rangle< m$ then ${\gta}((\b b))=1/m\sum _{i=0}^3\langle b_i\rangle$ if $(\undl b)\ne (\undl 0)$ and 1 if not. Finally, the action by $Gal$(\bQ/\Q) is given by $F_p((\undl b))=(p\undl b)$ for a prime $p\not|m$. \begin{remark} The proof of this in \[Ka:Sect.~6] uses transcendental methods. A purely algebraic proof can be given by tracing the action of $G$ through the calculations of \[SGA7:Exp.~XI]. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{remark} The Fermat surfaces verify the Tate conjecture (\[S-K]) so $\mathop{\rm rk}\nolimits\,NS=b_2$ over a field of positive characteristic iff the average of {\gta} over any {\gsi}-orbit equals 1. Now complex conjugations in $Gal(\bQ/\Q)$ exchanges the values 0 and 2 and fixes 1 so we see that if the subgroup generated by {\gsi} contains complex conjugation this is always the case. Hence if $-1\rum\@a\rum\langle p\rangle\subseteq(\Z/m)^\times$ then $rk~NS(X_m)=b_2(X_m)$ in characteristic p. \begin{example} i) $p\equiv -1 \pmod m$. Then all the orbits are of type $(1,1,\ldots,1)$ or (0,2) giving a contribution of 0 resp.~1 to ${\gsi}_0$. It is a general fact (true whenever $rk\,NS=b_2$ and $b_2=\sum _{i+j=2}dim\;\hod Xijk$) that $p^{2p_g}|disc\ NS$ as the morphism $H^2(X,p)\to \hod X20k$ is surjective and vanishes on $NS\bigotimess \W$. ii) m=5, $p\equiv 2,3 \pmod5$. Then there are four orbits of type $(0,1,2,1)$ and the rest are of type $(1,1,\ldots,1)$. Now the algorithm applied to $(0,1,2,1)$ gives partial sums $(-1,-1,0,0)$ and a contribution of $2$ to ${\gsi}_0$ for each copy of this orbit and hence $disc\,NS=p^{16}$, whereas $p_g=4$ so that we get a higher power of $p$ than is guaranteed by $p^{2p_g}|disc\ NS$. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{example} \mdef\tx{\tilde X} \def\tilde X{\tilde X} \begin{proposition}3.1. Suppose that $X$ is a smooth surface over \k\ and that $p>0$. Suppose that $G$ is a finite group of order prime to $p$ acting on $X$. Let \tx be a minimal resolution of $X/G$. Then $H^2(\tilde X,p)=H^2(X,p)^G\perp E$, where $E$ is the \W-module spanned by the Chern classes of the exceptional curves of $\tilde X\to X/G$ and orthogonality is wrt the cup product. Furthermore, the cup product pairing restricted to $E$ is perfect. \pro Let ${\gpi}\co X'\to X$ be a $G$-equivariant blowing up of $X$ such that we have a map ${\grh}\co X'\to\tilde X$ covering the quotient map $X\to X/G$. The cup product pairing on $E$ is perfect because the cokernel of $E\to \check E$ equals \W tensored with the sum of the local Picard groups of the singularities of $X/G$ (cf.~\[Li:14.4]) and these are killed by the order of $G$ by the existence of a norm map. Hence we may write $H^2(\tilde X,p)=V\perp E$. Now ${\grh}_*{\grh}^*=|G|$ so ${\grh}^*$ is injective on $H^2(\tx,p)$ and the image is contained in the $G$-invariants and is a direct factor. Furthermore, by the projection formula, ${\grh}^* V$ is orthogonal to the submodule of $H^2(X',p)$ spanned by the curves exceptional for {\gpi}. Therefore ${\grh}^* V\subseteq \pi^* H^2(X,p)^G$ and we are finished if we can show that this is an equality. First, we show this for the $p$-torsion. Indeed, consider the slope spectral sequence for $X'$ and \tx\ (cf.~\[Ill:II,3]). By duality (cf.~\[Ek1]) ${\grh}_*$ is defined as a map of spectral sequences and we still have ${\grh}_*{\grh}^*=|G|$. Furthermore, ${\grh}^*$ is an isomorphism on $H^*(\tx,\W\ko \tx)\to H^*(X',\W\ko {X'})^G$ as $H^*(X',\W\ko {X'})^G=H^*(X,\W\ko X)^G= H^*(X^G,\W\ko {X^G}) =H^*(\tx,\W\ko \tx)$ the last as the singularities are rational. Hence as $H^0(-,\W{\gOm}_X^2)$ is torsion free as $W{\gOm}_X^2$ is we see that we have equality on torsion groups if we have equality for the torsion of $H^1(-,\W{\gOm}^1_-)$. The nilpotent torsion (cf.~\[Ek1:IV,3.3.13]) of it is dual to the nilpotent torsion of $H^2(-,\W\ko -)$ (loc.~cit.) and is hence taken care of, whereas the semi-simple torsion (loc.~cit.:IV,3.4) comes from $H^2(-,\Z_p(1))$ which in turn comes from the N{\'e}ron-Severi group \[Ill:II,5.8.5] which is taken care of by noting that $\undl {Pic}^{\gta}(\tilde X)=\undl {Pic}^{\gta}(X/G)$ as the singularities are rational and $\undl {Pic}^{\gta}(X/G)=\undl {Pic}^{\gta}(X)^G$ outside of the order of $G$. Hence, as ${\grh}^* V$ is a direct factor of $\pi^* H^2(X,p)^G$, it suffices to show that that they have the same rank. As the rank of $V$ is the rank of $H^2(\tx,p)$ minus the number of exceptional curves we may replace $p$ by \l\ and then $H^2(X,\ell)^G=H^2(X/G,\ell)$ and the latter space is isomorphic to the orthogonal complement of the exceptional curves of $\tilde X\to X/G$ by the Leray spectral sequence. \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{proposition} Using the proposition we get a description of the crystalline cohomology of the minimal resolution of the quotient of $X_m$ by any subgroup of $G$. We can also compute other invariants of Fermat surfaces and their quotients. Consider for instance the formal Brauer group of a surface $X$ which is Mazur-Ogus (cf.~\[Ek:IV,1.1]) (in positive characteristic) or rather $H^2(X,\W\ko X)$ the knowledge of which is equivalent to knowing the formal Brauer group. It follows from \[loc.~cit.:III,Thm~4.3] that $H^2(X,W\ko X)$ is the quotient of $H^2_{cris}(X/W)\bigotimess _{W[F]}D$, where $D$ is the Dieudonn{\'e}-ring (with power series in $V$), by the submodule generated by $m\bigotimess 1-V(n\bigotimess 1)$ for all $m,n\rum\@a\rum H^2_{cris}(X/W)$ for which $Fm=pn$. Hence if $X$ is of CM-type in degree 2 we get a description of $H^2(X,W\ko X)$. \begin{example} i) $(0,1,0,2,1,2)$ gives a $D$-module with generators $a$ and $b$ and relations $Fa=Vb$ and $Fb=0$. This is the Dieudonn{\'e}-module of a 2-dimensional formal group isogenous but not isomorphic to $W_2$. For $p\equiv3\pmod7$ this appears in the cohomology of the Fermat surface of degree 7 (the orbit of $(1,1,1,4)\rum\@a\rum (\Z/7)^4$). ii) $(0,0,2,2)$ gives a $D$-module with generator a and relation $F^2a=0$. This is the Dieudonn{\'e}-module of $W_2$. For $p\equiv5\pmod7$ this appears in the cohomology of the Fermat surface of degree 13 (the orbit of $(3,3,3,4)\rum\@a\rum (\Z/7)^4$). \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{example} \begin{bibliography} \[B-Og]:\egroup\setbox1=\hbox\bgroup\smc P. Berthelot, A. Ogus \egroup\book@true\setbox2=\hbox\bgroup\sl Notes on crystalline cohomology\egroup\setbox3=\hbox\bgroup Princeton Univ. Press \egroup\setbox4=\hbox\bgroup Princeton\yr1978 \[C-R]: \egroup\setbox1=\hbox\bgroup\smc C. W. Curtis, I. Reiner\egroup\book@true\setbox2=\hbox\bgroup\sl Representation theory of finite groups \yr1962\egroup\setbox3=\hbox\bgroup Interscience publishers\egroup\setbox4=\hbox\bgroup New York \[De]:\egroup\setbox1=\hbox\bgroup\smc P. Deligne\egroup\setbox2=\hbox\bgroup Applications de la formule des traces aux sommes trigonom{\'e}triques \egroup\sln@true\setbox4=\hbox\bgroup 569 \egroup\pages@true\setbox6=\hbox\bgroup 168--232 \[D-I]: \egroup\setbox1=\hbox\bgroup\smc F. DeMeyer, E. Ingraham \egroup\setbox2=\hbox\bgroup Separable algebras over commutative rings\sln181 \[Ek]:\egroup\setbox1=\hbox\bgroup\smc T. Ekedahl\egroup\book@true\setbox2=\hbox\bgroup\sl Diagonal complexes and F-gauge structures\egroup\setbox3=\hbox\bgroup Hermann \egroup\setbox4=\hbox\bgroup Paris \yr1986 \[Ek1]:\egroup\setbox1=\hbox\bgroup\smc T. Ekedahl \egroup\setbox2=\hbox\bgroup On the multiplicative properties of the de Rham-Witt complex I\egroup\jour@true\setbox3=\hbox\bgroup\sl Arkiv f{\"o}r mate\-ma\-tik \vol22 \yr1984\pages185--239 \[Fa]:\egroup\setbox1=\hbox\bgroup\smc G. Faltings \egroup\setbox2=\hbox\bgroup p-adic Hodge theory \yr1988 \egroup\jour@true\setbox3=\hbox\bgroup\sl Journal of the AMS \vol1\pages255--299 \[Gr]: \egroup\setbox1=\hbox\bgroup\smc M. Gros \egroup\setbox2=\hbox\bgroup Classes de Chern et classes de cycles en cohomologie de Hodge-Witt logarithmique\egroup\jour@true\setbox3=\hbox\bgroup\sl M{\'e}m. de la Soc. Math. de France\vol21\pages1--87\yr1985 \[Gro]:\egroup\setbox1=\hbox\bgroup\smc A. Grothendieck \egroup\setbox2=\hbox\bgroup Le groupe de Brauer III \egroup\inbook@true\setbox7=\hbox\bgroup Dix expos{\'e}s sur la cohomologie des sch{\'e}mas \egroup\setbox3=\hbox\bgroup North-Holland \egroup\setbox4=\hbox\bgroup Amsterdam \yr1968 \pages88--188 \[Ill]:\egroup\setbox1=\hbox\bgroup\smc L. Illusie \egroup\setbox2=\hbox\bgroup Complexe de de Rham-Witt et cohomologie cristalline \egroup\jour@true\setbox3=\hbox\bgroup\sl Ann. scient. {\'E}c. Norm. Sup \egroup\setbox4=\hbox\bgroup\bf 12 \yr1979 \pages501--661 \[Jo]:\egroup\setbox1=\hbox\bgroup\smc J. de Jong\egroup\spec@true\setbox8=\hbox\bgroup Article to appear \[Ka]:\egroup\setbox1=\hbox\bgroup\smc N. M. Katz\egroup\setbox2=\hbox\bgroup On the intersection matrix of a hypersurface \egroup\jour@true\setbox3=\hbox\bgroup\sl Ann. scient. {\'E}c. Norm. Sup\vol2\yr1969\egroup\pages@true\setbox6=\hbox\bgroup 583--589 \[K-L]:\egroup\setbox1=\hbox\bgroup\smc N. M. Katz, S. Lang \egroup\setbox2=\hbox\bgroup Finiteness theorems in geometric class field theory \egroup\jour@true\setbox3=\hbox\bgroup\sl L'Ens. Math \vol27 \yr1981\pages286--314 \[K-M]: \egroup\setbox1=\hbox\bgroup\smc N. M. Katz, W. Messing \egroup\setbox2=\hbox\bgroup Some consequences of the Riemann hypothesis for varieties over finite fields \egroup\jour@true\setbox3=\hbox\bgroup\sl Invent. Math. \yr1974 \vol23\pages73--77 \[Li]:\egroup\setbox1=\hbox\bgroup\smc J. Lipman\egroup\setbox2=\hbox\bgroup Rational singularities with applications to algebraic surfaces and unique factorization \egroup\jour@true\setbox3=\hbox\bgroup\sl Publ. IHES \vol36\yr1969\pages195--280 \[Re]:\egroup\setbox1=\hbox\bgroup\smc I. Reiner\egroup\book@true\setbox2=\hbox\bgroup\sl Maximal orders\egroup\setbox3=\hbox\bgroup Academic Press\egroup\setbox4=\hbox\bgroup London \egroup\setbox5=\hbox\bgroup 1975 \[Se]:\egroup\setbox1=\hbox\bgroup\smc J.-P. Serre \egroup\book@true\setbox2=\hbox\bgroup\sl Abelian \l-adic representations and elliptic curves \egroup\setbox3=\hbox\bgroup W. A. Benjamin\egroup\setbox4=\hbox\bgroup New York\yr1968 \[Se1]:\egroup\setbox1=\hbox\bgroup\smc J.-P. Serre \egroup\setbox2=\hbox\bgroup Groupes alg{\'e}briques associ{\'e}s aux modules de Hodge-Tate \egroup\jour@true\setbox3=\hbox\bgroup\sl Ast{\'e}rix 65 \egroup\setbox5=\hbox\bgroup 1979\pages155--188 \[SGA7]: \egroup\setbox1=\hbox\bgroup\smc P. Deligne, N. M. Katz \egroup\setbox2=\hbox\bgroup Groupes de monodromie en g{\'e}om{\'e}trie alg{\'e}brique \sln340 \[S-K]:\egroup\jour@true\setbox3=\hbox\bgroup\sl Toh{\^o}ku Math. Jour. \egroup\setbox1=\hbox\bgroup\smc T. Shioda, T. Katsura \egroup\setbox5=\hbox\bgroup 1979 \vol31\egroup\setbox2=\hbox\bgroup On Fermat varieties \egroup\pages@true\setbox6=\hbox\bgroup 97-115 \[S-I]: \egroup\setbox1=\hbox\bgroup\smc T. Shioda, H. Inose \egroup\setbox2=\hbox\bgroup On singular K3-surfaces \egroup\inbook@true\setbox7=\hbox\bgroup Complex analysis \& algebraic geometry \egroup\setbox3=\hbox\bgroup Cambridge Univ. Press \egroup\setbox4=\hbox\bgroup Cambridge \yr1977\egroup\pages@true\setbox6=\hbox\bgroup 117--136 \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{bibliography} \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{section} \egroup\end}\catcode`\^^M=\active\catcode`\ =\active\tt{document} \vfill \supereject \@nd
1996-03-05T06:22:21
9512
alg-geom/9512013
fr
https://arxiv.org/abs/alg-geom/9512013
[ "alg-geom", "math.AG" ]
alg-geom/9512013
Laurent Bonavero
Laurent Bonavero
In\'egalit\'es de Morse et vari\'et\'es de Moishezon
Latex (plain). PhD-thesis of the author, defended in Grenoble (France) 18-12-1995, hard copies available on request at [email protected] http://www-fourier.ujf-grenoble.fr/THESE/
null
null
null
null
The central topic of this thesis is the study of some properties of a class of complex compact manifolds~: Moishezon manifolds. In the first part, we generalize J.-P. Demailly's holomorphic Morse inequalities to the case of a line bundle equipped with a metric with analytic singularities on an arbitrary compact complex manifold. Our inequalities give an estimate of the cohomology groups with values in the line bundle tensor powers twisted by the corresponding sequence of multiplier ideal sheaves introduced by Nadel. As a consequence, we obtain a necessary and sufficient analytic condition, invariant by bimeromorphism, for a manifold to be Moishezon. In the second part, we use Mori theory to analyze the structure of Moishezon manifolds with infinite cyclic Picard group, with big canonical bundle, and which become projective after one single blow-up with smooth center. We study the dimension and the structure of the center of the blow-up. In dimension four, we show that this locus is always a surface, and when the canonical bundle
[ { "version": "v1", "created": "Wed, 20 Dec 1995 10:49:04 GMT" } ]
2008-02-03T00:00:00
[ [ "Bonavero", "Laurent", "" ] ]
alg-geom
\chapter*{Introduction} Le sujet central de cette th\`ese est l'\'etude de certaines propri\'et\'es d'une classe de vari\'et\'es analytiques complexes compactes~: les vari\'et\'es de Moishezon. Ces derni\`eres sont particuli\`erement int\'eressantes car elles forment la plus petite classe de vari\'et\'es complexes stable par application bim\'eromorphe et contenant les vari\'et\'es projectives. Il est bien connu depuis K.\ Kodaira que les vari\'et\'es projectives sont caract\'eris\'ees par l'existence d'un fibr\'e en droites ample ou de fa\c con \'equivalente d'un fibr\'e en droites muni d'une m\'etrique hermitienne \`a courbure strictement positive. Le fil conducteur de cette th\`ese est l'\'etude de l'existence ou de l'inexistence de fibr\'es en droites v\'erifiant des propri\'et\'es de positivit\'e faible sur les vari\'et\'es de Moishezon. L'\'etude que nous avons faite est divis\'ee en deux parties~: un point de vue analytique suivant et g\'en\'eralisant une d\'emarche pr\'esente dans certains travaux de J.-P.\ Demailly et Y.-T.\ Siu, et un point de vue plus alg\'ebrique reposant sur l'utilisation de la th\'eorie de Mori, d\'emarche pr\'esente dans certains travaux de J.\ Koll\'ar et T.\ Peternell. \bigskip \bigskip \noindent{\bf \'Etude analytique} \medskip Cette \'etude a d\'emarr\'e avec les travaux de J.-P.\ Demailly et Y.-T.\ Siu qui, r\'epondant \`a une conjecture de H.\ Grauert et O.\ Riemenschneider, ont donn\'e ind\'ependamment des conditions analytiques suffisantes (existence de fibr\'es en droites \`a courbure semi-positive et g\'en\'eriquement positive) pour qu'une vari\'et\'e complexe compacte soit de Moishezon. Une de nos motivations vient du fait qu'aucune de ces conditions n'est n\'ecessaire, comme le montre l'\'etude de constructions r\'ecentes. L'un des premiers r\'esultats de cette th\`ese est de donner une caract\'erisation analytique des vari\'et\'es de Moishezon. Pour cela, nous montrons, et c'est le th\'eor\`eme principal de la premi\`ere partie de notre travail, que les in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly se g\'en\'eralisent au cas d'un fibr\'e en droites $E$ muni d'une m\'etrique singuli\`ere $h$ au dessus d'une vari\'et\'e complexe compacte $X$. Nos in\'egalit\'es donnent une estimation asymptotique de la dimension des groupes de cohomologie \`a valeurs dans les puissances tensorielles $E^{\otimes k}$, tordues par une suite de faisceaux d'id\'eaux $ {\cal I}_{k} (h)$ naturellement associ\'ee aux singularit\'es de la m\'etrique $h$~: la suite des faisceaux d'id\'eaux multiplicateurs de Nadel. La pr\'esence de ces faisceaux d'id\'eaux constitue le ph\'enom\`ene nouveau par rapport au cas o\`u la m\'etrique est lisse. Comme dans ce dernier cas, l'estimation fait intervenir des int\'egrales de la courbure $\Theta(E)$. Notre r\'esultat est le suivant~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Si la m\'etrique $h$ a des singularit\'es analytiques, alors pour tout fibr\'e $F$ de rang $r$ et pour tout $q$ compris entre $0$ et $n = \dim (X)$, on a~: $$ \sum _{j=0}^{q}(-1)^{q-j} \dim H^{j}(X,{\cal O}(E^{k}\otimes F) \otimes {\cal I}_{k}(h)) \leq r\frac{k^{n}}{n!} \int _{X(\leq q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}) $$ (avec \'egalit\'e si $q=n$), o\`u $X(\leq q,E)$ d\'esigne l'ouvert de $X$ des points lisses de la m\'etrique d'indice inf\'erieur \`a $q$. } \bigskip Ce r\'esultat est \`a mettre en parall\`ele avec la g\'en\'eralisation du th\'eor\`eme de Ka\-wa\-ma\-ta-Viehweg donn\'ee par A.\ Nadel. Nous montrons ensuite, g\'en\'eralisant un r\'esultat de S.\ Ji et B.\ Shiffman obtenu ind\'ependamment et simultan\'ement au n\^otre, que les crit\`eres de J.-P.\ Demailly et Y.-T.\ Siu deviennent, dans ce cadre plus souple, n\'ecessaires et suffisants. Donnons par exemple le~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Une vari\'et\'e compacte $X$ de dimension $n$ est de Moishezon si et seulement s'il existe sur $X$ un courant ferm\'e $T$ de bi-degr\'e $(1,1)$ tel que~: (i) $ \displaystyle{\{ T \} \in H^{2}(X,{\Bbb Z}) }$, (ii) $\displaystyle{ T= \frac{i}{\pi} \partial \overline{\partial} \varphi + \alpha }$, o\`u $\varphi$ est une fonction r\'eelle \`a singularit\'es analytiques et o\`u $\alpha$ est un repr\'esentant $ {\cal C}^{\infty}$ de $\{ T \}$, (iii) $\displaystyle{\int_{X(\leq 1,T)} T^{n} > 0}$ o\`u l'int\'egrale est prise sur les points lisses du courant $T$. } \bigskip Comme nous l'avons d\'ej\`a mentionn\'e, ce type de crit\`ere a la propri\'et\'e d'\^etre invariant par morphisme bim\'eromorphe. \bigskip \bigskip \noindent{\bf \'Etude alg\'ebrique} \medskip La deuxi\`eme partie de cette th\`ese consiste \`a \'etudier en d\'etail la classe des vari\'et\'es de Moishezon dont le groupe de Picard est ${\Bbb Z}$, et dont le fibr\'e canonique $K_X$ est gros (``big"). Une de nos motivations provient d'un r\'esultat de J.\ Koll\'ar affirmant qu'en dimension $3$, et sous les hypoth\`eses pr\'ec\'edentes, le fibr\'e canonique est alors num\'eriquement effectif (nef). Il n'est donc pas possible de trouver dans cette classe des vari\'et\'es de Moishezon de dimension $3$ ne v\'erifiant pas les crit\`eres de J.-P.\ Demailly et Y.-T.\ Siu. Nous montrons que ceci n'est plus vrai en dimension sup\'erieure ou \'egale \`a $4$ en construisant explicitement des exemples~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Pour tout entier $n$ sup\'erieur ou \'egal \`a $4$, il existe des vari\'et\'es de Moishezon $X$, non projectives, de dimension $n$ v\'erifiant~: (i) $\operatorname{Pic} (X) = {\Bbb Z}$, (ii) $K_X$ est gros, (iii) $K_X$ n'est pas nef.} \bigskip La construction donnant ce r\'esultat montre que les vari\'et\'es $X$ obtenues deviennent projectives apr\`es un \'eclatement le long d'une sous-vari\'et\'e isomorphe \`a $\displaystyle{{\Bbb P} ^{n-2}}$. Plus g\'en\'eralement, un r\'esultat fondamental de B.\ Moishezon affirme qu'une vari\'et\'e de Moishezon peut \^etre rendue projective apr\`es une succession finie d'\'eclatements le long de sous-vari\'et\'es lisses. Ce r\'esultat difficile ne donne malheureusement aucune indication quant au choix explicite des sous-vari\'et\'es en question. Gr\^ace \`a l'utilisation de la c\'el\`ebre th\'eorie de Mori sur un mod\`ele projectif, nous avons \'etudi\'e le centre de l'\'eclatement en toutes dimensions~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Soit $X$ une vari\'et\'e de Moishezon (non projective) de dimension $n$, avec $\operatorname{Pic} (X) = {\Bbb Z}$ et $K_X$ gros. Supposons de plus que $X$ est rendue projective apr\`es \'eclatement le long d'une sous-vari\'et\'e $Y$ lisse. \noindent Alors, si $K_X$ n'est pas nef, on a $\displaystyle{ \dim Y > \frac{n-1}{2}.}$ } \bigskip En dimension $4$, ce r\'esultat peut \^etre pr\'ecis\'e, y compris dans le cas o\`u le fibr\'e canonique est nef~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Soit $X$ une vari\'et\'e de Moishezon (non projective) de dimension $4$, avec $\operatorname{Pic} (X) = {\Bbb Z}$ et $K_X$ gros. Supposons de plus que $X$ est rendue projective apr\`es \'eclatement le long d'une sous-vari\'et\'e $Y$ lisse. \noindent Alors $Y$ est n\'ecessairement une surface. Autrement dit, et dans cette situation particuli\`ere, il ne suffit pas d'\'eclater une courbe pour ``rentrer dans le monde projectif".} \bigskip Nous avons vu pr\'ec\'edemment que $K_X$ n'est pas n\'ecessairement nef \`a partir de la dimension $4$. Le r\'esultat suivant montre que l'exemple que nous avons construit est, en un sens, le seul possible en dimension $4$ dans le cas o\`u $K_X$ n'est pas nef. \bigskip \noindent {\bf Th\'eor\`eme } {\em Sous les hypoth\`eses pr\'ec\'edentes et si $K_{X}$ n'est pas nef, alors le couple $\displaystyle{(Y,N_{Y/X})}$ est isomorphe \`a $\displaystyle{({\Bbb P} ^2, {\cal O}_{{\Bbb P} ^{2}}(-1)^{\oplus 2})}$.} \bigskip Ces r\'esultats pr\'ecis sont accessibles en dimension $4$ car les contractions de Mori ont \'et\'e \'etudi\'ees par T.\ Ando, Y.\ Kawamata et M.\ Beltrametti. Nos r\'esultats peuvent \^etre vus comme un premier pas vers une analyse du caract\`ere non projectif des vari\'et\'es de Moishezon de dimension sup\'erieure ou \'egale \`a $4$~; la situation en dimension $3$ \'etant maintenant assez bien comprise suite aux travaux de J.\ Koll\'ar et T.\ Peternell. \bigskip \bigskip \noindent{\bf Plan du texte} \medskip - le chapitre 1 est un chapitre de pr\'eliminaires~; il contient une description pr\'ecise des motivations et des objets utilis\'es dans le reste de la th\`ese. Nous y d\'etaillons en particulier une d\'emonstration du th\'eor\`eme de Siegel et donnons les constructions de I.\ Nakamura et K.\ Oguiso montrant que les crit\`eres analytiques de J.-P.\ Demailly et Y.-T.\ Siu ne sont pas n\'ecessaires en g\'en\'eral. \medskip - le chapitre 2 est consacr\'e \`a l'\'etude analytique. Nous rappelons les premiers r\'esultats li\'es aux m\'etriques singuli\`eres, \'enon\c cons et d\'emontrons notre version des in\'egalit\'es de Morse holomorphes. On en d\'eduit les caract\'erisations analytiques des vari\'et\'es de Moishezon. Enfin, nous donnons une version alg\'ebrique singuli\`ere des in\'egalit\'es de Morse. \medskip - le chapitre 3 est consacr\'e \`a l'\'etude alg\'ebrique. Nous commen\c cons par rappeler le r\'esultat de J.\ Koll\'ar en dimension $3$, puis nous faisons une \'etude des vari\'et\'es de Moishezon \`a groupe de Picard ${\Bbb Z}$, \`a fibr\'e canonique gros et devenant projectives apr\`es un seul \'eclatement de centre lisse et projectif. Nous obtenons une restriction sur la dimension du centre de l'\'eclatement. En dimension $4$, cette restriction implique que ce dernier est n\'ecessairement une surface. Nous d\'ecrivons alors notre exemple et montrons qu'en dimension $4$, cette construction est essentiellement la seule dans le cas o\`u le fibr\'e canonique n'est pas nef. \medskip Mentionnons que les chapitres 2 et 3 sont dans une large mesure ind\'ependants et peuvent \^etre lus dans un ordre quelconque. \chapter{Pr\'eliminaires} Ce chapitre a pour but d'introduire les principales notions utilis\'ees par la suite, de pr\'esenter les premi\`eres motivations en d\'etail et de rappeler un certain nombre de r\'esultats auxquels nous nous r\'ef\'erons dans les chapitres suivants. \section{\! Quelques rappels de g\'eom\'etrie analytique complexe} \subsection{Vari\'et\'es, fibr\'es vectoriels} Pr\'ecisons tout d'abord que dans toute cette th\`ese, et sauf mention explicite du contraire, le mot {\bf vari\'et\'e} sera utilis\'e pour d\'esigner une {\bf vari\'et\'e analytique complexe {\em non singuli\`ere}} suppos\'ee de plus connexe. Pour toutes les notions introduites ici, nous renvoyons de fa\c con g\'en\'erale \`a \cite{G-H78}. \medskip Un {\bf fibr\'e vectoriel} complexe $F$ au dessus d'une vari\'et\'e $X$ est la donn\'ee d'une vari\'et\'e $F$ et d'une application $\displaystyle{\pi : F \to X}$ de sorte qu'il existe un recouvrement de $X$ par des ouverts trivialisants $U_{\alpha}$ et des isomorphismes (appel\'es trivialisations) $$\theta _{\alpha} : \pi ^{-1}(U_{\alpha}) \to U_{\alpha} \times {\Bbb C} ^r$$ respectant la structure d'espace vectoriel des fibres, i.e $$\theta _{\alpha \beta}(x,\xi) := \theta _{\alpha} \circ \theta _{\beta} ^{-1} (x,\xi) = (x, g_{\alpha \beta}(x)\xi )$$ o\`u $g_{\alpha \beta}$ est une application holomorphe sur $U_{\alpha} \cap U_{\beta}$ \`a valeurs dans le groupe des matrices complexes inversibles de taille $r$. Nous notons $\xi _x$ un point de $F$ au dessus du point $x$ de $X$ (i.e tel que $\pi (\xi _x) =x$). L'entier $r$ est le {\bf rang} du fibr\'e $F$. Si $r = 1$, on parle de {\bf fibr\'e en droites}. Un exemple important de fibr\'e en droites est le fibr\'e ${\cal O}(D)$ associ\'e \`a un diviseur $D$ sur $X$~: si $D$ est un diviseur irr\'eductible donn\'e sur $U_{\alpha}$ par l'\'equation $f_{\alpha}=0$, le fibr\'e ${\cal O}(D)$ est le fibr\'e associ\'e au cocycle $$g_{\alpha \beta} = \frac{f_{\alpha}}{f_{\beta}} \ .$$ \medskip Toutes les constructions d'alg\`ebre lin\'eaire s'\'etendent aux fibr\'es~: dual, produit tensoriel, produit ext\'erieur. Ainsi, un exemple fondamental de fibr\'e en droites sur une vari\'et\'e $X$ de dimension $n$ est le {\bf fibr\'e canonique} d\'efini par $$K_X := \det (T^{\ast}X) = \bigwedge ^n T^{\ast}X,$$ o\`u $T^{\ast}X$ d\'esigne le fibr\'e cotangent, dual du fibr\'e tangent holomorphe $TX$ de $X$. Un autre exemple important de fibr\'e en droites est le fibr\'e ${\cal O}_{{\Bbb P} ^n}(1)$ sur l'espace projectif ${\Bbb P} ^n$~: en associant \`a un point $[x]$ de ${\Bbb P} ^n$ la droite ${\Bbb C} x$, on construit un sous-fibr\'e en droites du fibr\'e trivial ${\Bbb P} ^n \times {\Bbb C} ^{n+1}$~; le dual de ce fibr\'e en droites est par d\'efinition le fibr\'e ${\cal O}_{{\Bbb P} ^n}(1)$. L'ensemble des fibr\'es en droites, modulo isomorphisme, sur une vari\'et\'e $X$ est naturellement muni d'une structure de groupe pour le produit tensoriel~: on l'appelle {\bf groupe de Picard} de $X$ et on le note $\operatorname{Pic} (X)$. Mentionnons ici que nous identifions suivant l'usage un fibr\'e en droites $E$ au faisceau inversible ${\cal O}(E)$ des germes de sections holomorphes de $E$. La $k$-i\`eme puissance tensorielle d'un fibr\'e en droites $E$ sera not\'ee indiff\'eremment $E^{\otimes k}$, $E^k$,${\cal O}(kE)$ ou m\^eme $kE$. \medskip Un fibr\'e vectoriel $E$ peut \^etre muni d'une m\'etrique hermitienne ${\cal C}^{\infty}$, on parle alors de fibr\'e vectoriel {\bf hermitien} et on note g\'en\'eralement $h$ une telle m\'etrique~: elle correspond \`a la donn\'ee d'une forme hermitienne sur chaque fibre $E_x$ de $E$, d\'ependant de fa\c con ${\cal C}^{\infty}$ de $x$. Dans le cas particulier d'un fibr\'e en droites, une m\'etrique hermitienne est donn\'ee localement sur un ouvert trivialisant $U_{\alpha}$ par $$h(\xi _{x}) = ||\xi _x||_h := |\xi|\exp(-\varphi _{\alpha} (x))$$ (la fonction $\exp(-\varphi _{\alpha})$ est appel\'ee {\bf poids} de la m\'etrique $h$ dans la trivialisation $\theta _{\alpha}$) o\`u la fonction r\'eelle $\varphi _{\alpha}$ est de classe ${\cal C}^{\infty}$ sur $U_{\alpha}$. Lorsque le fibr\'e tangent $TX$ est muni d'une m\'etrique hermitienne, on dit que la vari\'et\'e $X$ est hermitienne. Comme il est d'usage, nous identifions toujours la donn\'ee d'une m\'etrique hermitienne sur une vari\'et\'e $X$ \`a celle de la $(1,1)$ forme r\'eelle, g\'en\'eralement not\'ee $\omega$, qui lui est naturellement associ\'ee ($\omega$ est \`a un facteur $-2$ pr\`es la partie imaginaire de la m\'etrique). Ainsi, une vari\'et\'e est {\bf k\"ahl\'erienne} si elle poss\`ede une m\'etrique hermitienne pour laquelle $\omega$ est une forme ferm\'ee. Une vari\'et\'e {\bf projective} est une vari\'et\'e isomorphe \`a une sous-vari\'et\'e ferm\'ee d'un espace projectif ${\Bbb P} ^N$. \medskip Pour un fibr\'e en droites hermitien $(E,h)$, on note $\Theta (E)$ la $(1,1)$ {\bf forme de courbure} de $(E,h)$~: c'est la forme r\'eelle d\'efinie globalement sur $X$ et donn\'ee localement par $$\displaystyle{ \Theta (E)= \frac{i}{\pi} \partial \overline{\partial} \varphi _{\alpha}};$$ c'est aussi la courbure de la connexion de Chern du fibr\'e hermitien $E$. La classe de cohomologie de $\Theta (E)$ appartient \`a $H^2(X,{\Bbb Z})$ et ne d\'epend pas de la m\'etrique $h$~; c'est la premi\`ere classe de Chern de $E$ et elle est not\'ee $c_1(E)$. Remarquons qu'il n'y a pas de sens \`a parler des valeurs propres de la forme de courbure, mais que la {\bf signature} de la courbure (i.e le nombre de ``valeurs propres" nulles, strictement positives et strictement n\'egatives) est une notion bien d\'efinie sans donn\'ee suppl\'ementaire. Par exemple, le fibr\'e ${\cal O}_{{\Bbb P} ^n}(1)$ muni de la m\'etrique induite de celle de ${\Bbb C} ^{n+1}$ est \`a courbure strictement positive~: la forme de courbure est la m\'etrique de Fubini-Study de ${\Bbb P} ^n$. \subsection{Th\'eor\`eme de Kodaira} Nous sommes en mesure d'\'enoncer maintenant le c\'el\`ebre th\'eor\`eme de plongement de Kodaira \cite{Kod54}~: \bigskip \noindent{\bf Th\'eor\`eme (K.\ Kodaira, 1954)} {\em Une vari\'et\'e compacte $X$ est projective si et seulement si elle poss\`ede un fibr\'e en droites hermitien $E$ \`a courbure strictement positive.} \bigskip Signalons \'evidemment qu'un sens est ais\'e~: si une vari\'et\'e est projective, la restriction \`a $X$ du fibr\'e $\displaystyle{ {\cal O}_{{\Bbb P} ^N} (1) }$ muni de sa m\'etrique naturelle ayant pour courbure la forme de Fubini-Study de ${\Bbb P} ^N$ donne le fibr\'e souhait\'e. L'autre sens consiste \`a montrer que pour $k$ entier assez grand, il est possible de plonger $X$ dans l'espace projectif des hyperplans de $H^0(X,E^{\otimes k})$, o\`u $H^0(X,E^{\otimes k})$ d\'esigne l'espace vectoriel des sections holomorphes globales de $E^{\otimes k}$. Rappelons ici qu'un fibr\'e en droites pouvant \^etre muni d'une m\'etrique \`a courbure strictement positive est dit {\bf ample}. C'est le th\'eor\`eme de Kodaira que H.\ Grauert et O.\ Riemenschneider \cite{GrR70} se proposaient de g\'en\'eraliser aux vari\'et\'es de Moishezon, vari\'et\'es que nous introduisons dans le paragraphe suivant. \section{Quelques rappels sur les vari\'et\'es de Moishezon} Les vari\'et\'es de Moishezon sont, parmi les vari\'et\'es compactes, celles qui poss\`edent le ``plus" de fonctions m\'eromorphes alg\'ebriquement ind\'ependantes. Cette d\'efinition heuristique est justifi\'ee par le th\'eor\`eme de Siegel que nous rappelons maintenant. \subsection{Th\'eor\`eme de Siegel} En 1955, C.L.\ Siegel d\'emontre le r\'esultat suivant \cite{Sie55}~: \bigskip \noindent{\bf Th\'eor\`eme (C.L.\ Siegel, 1955) } {\em Si $X$ est une vari\'et\'e compacte de dimension $n$, alors $X$ poss\`ede au plus $n$ fonctions m\'eromorphes alg\'ebriquement ind\'ependantes.} \bigskip La d\'emonstration originale de C.L.\ Siegel repose sur une application \'el\'ementaire du lemme de Schwarz. Il existe maintenant plusieurs d\'emonstrations diff\'erentes, certaines g\'en\'eralisant cet \'enonc\'e aux espaces complexes compacts. Nous en donnons ici une preuve ``moderne" dans le cas non singulier. Pour cela, nous utilisons un r\'esultat de P.\ Gauduchon \cite{Gau77}~: {\em toute vari\'et\'e analytique complexe compacte de dimension $n$ poss\`ede une m\'etrique hermitienne $\omega$ de classe ${\cal C}^{\infty}$ et d'excentricit\'e nulle, i.e telle que $\displaystyle{ \partial \overline{\partial} (\omega ^{n-1}) =0}$.} Commen\c cons par montrer le lemme suivant~: \medskip \noindent{\bf Lemme} {\em Soit $X$ une vari\'et\'e compacte que l'on munit d'une m\'etrique de Gauduchon $\omega$ et soit $x_0$ un point de $X$. Alors, il existe une constante $C := C(X,x_0,\omega )$ telle que pour tout fibr\'e en droites hermitien $(E,h)$ au dessus de $X$ et pour toute section holomorphe $s$ de $E$ non identiquement nulle, on ait~: $$ \operatorname{mult} \, (s,x_0) \leq C \int _X \omega ^{n-1} \wedge \Theta (E).$$} \medskip \noindent {\bf D\'emonstration} Soit $r$ un r\'eel strictement positif fix\'e ``petit" (de sorte qu'il existe une carte centr\'ee en $x_0$ et contenant la boule $B(x_o,r)$). Alors, la multiplicit\'e de $s$ en $x_0$ v\'erifie~: $$ \operatorname{mult} \, (s,x_0) \leq \frac{ \operatorname{vol} _{n-1}(Z_s \cap B(x_o,r))}{\operatorname{vol} _{n-1}(B_{n-1}(x_o,r))} +o(r) $$ o\`u $\operatorname{vol} _{n-1}$ est le volume $(n-1)$-dimensionnel mesur\'e avec la m\'etrique $\omega$, et o\`u $Z_s$ d\'esigne le lieu des z\'eros de $s$. En effet, la multiplicit\'e de $s$ en $x_0$ est en fait \'egale \`a la limite d\'ecroissante lorsque $r$ tend vers $0$ de la quantit\'e du membre de droite de l'in\'egalit\'e. Comme $X$ est compacte, on a {\em a fortiori}~: $$ \operatorname{mult} \, (s,x_0) \leq C \operatorname{vol} _{n-1}(Z_s) = C \int_{Z_s} \omega ^{n-1}.$$ Mais l'\'equation de Lelong-Poincar\'e affirme que~: $$ \frac{i}{\pi}\partial \overline{\partial}\log ||s|| = [ Z_s ] - \Theta (E),$$ o\`u $[ Z_s ]$ d\'esigne le courant d'int\'egration sur l'ensemble $Z_s$. Comme $\omega ^{n-1}$ est $\partial \overline{\partial}$-ferm\'ee, la formule de Stokes donne de suite~: $$ \int_{Z_s} \omega ^{n-1} = \int _X \omega ^{n-1} \wedge \Theta (E),$$ ce qui prouve le lemme.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Ce lemme implique le r\'esultat suivant~: \medskip \noindent{\bf Proposition} {\em Soient $X$ une vari\'et\'e compacte de dimension $n$, $\omega$ une m\'etrique de Gauduchon sur $X$ et $E$ un fibr\'e en droites hermitien au dessus de $X$. Alors~: (i) $\dim H^0(X,E) \leq {n+p \choose n}$ o\`u $p$ est la partie enti\`ere de $\displaystyle{C \int _X \omega ^{n-1} \wedge \Theta (E)}$ et o\`u $C$ est la constante du lemme pr\'ec\'edent, (ii) $\displaystyle{\dim H^0(X,E^{\otimes k}) \leq A \left(\int _X \omega ^{n-1} \wedge \Theta (E)\right)^n k^n + o(k^n)}$, o\`u $A$ est une constante ind\'ependante de $E$ et $k$. } \medskip \noindent {\bf D\'emonstration} Il suffit de remarquer que ${n+p \choose n}$ est la dimension de l'espace vectoriel des polyn\^omes de $n$ variables et de degr\'e inf\'erieur ou \'egal \`a $p$~: le lemme implique en effet que l'application lin\'eaire qui \`a une section holomorphe de $E$ associe son $p$-i\`eme jet en $x_0$ est injective, d'o\`u le point (i). Le point (ii) est cons\'equence du fait que la forme de courbure du fibr\'e $(E^{\otimes k},h^k)$ est donn\'ee par $\displaystyle{ \Theta (E^{\otimes k}) = k \Theta (E)}$. On applique alors (i) au fibr\'e $(E^{\otimes k},h^k)$.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Remarquons que la proposition pr\'ec\'edente affirme que la dimension de l'espace vectoriel des sections holomorphes des puissances $E^{\otimes k}$ d'un fibr\'e en droites sur une vari\'et\'e compacte de dimension $n$ cro\^\i t au plus comme $k^n$. Ce fait est bien classique et nous avons estim\'e la dimension en fonction d'int\'egrales de courbure. Des estimations bien plus pr\'ecises, valables pour la dimension de tous les groupes de cohomologie seront donn\'ees par les in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly dans le paragraphe suivant. \medskip \noindent {\bf D\'emonstration du th\'eor\`eme de Siegel} L'argument est standard~: soient $\displaystyle{(f_{i})_{1\leq i\leq N}}$ $N$ fonctions m\'eromorphes alg\'ebriquement ind\'ependantes sur $X$. Notons $D$ la somme des diviseurs des p\^oles des $f_{i}$ et ${\cal O}(D)$ le fibr\'e en droites associ\'e. Rappelons que $H^{0}(X,{\cal O}(D))$ est isomorphe \`a l'espace vectoriel des fonctions m\'eromorphes sur $X$ v\'erifiant $\operatorname{div} (f) + D \geq 0$. Alors, si $P$ est un polyn\^ome \`a coefficients complexes en $N$ variables, de degr\'e total inf\'erieur ou \'egal \`a $k$, la fonction m\'eromorphe $P(f_{1},\ldots,f_{N})$ est une section holomorphe de ${\cal O}(kD)$, et comme les $f_{i}$ sont alg\'ebriquement ind\'ependantes, on a ${N+k \choose N}$ telles sections lin\'eairement ind\'ependantes. De l\`a~: $$ \dim H^{0}(X,{\cal O}(D)^{\otimes k}) \geq {N+k \choose N} \sim_{k \to +\infty} \frac{k^N}{N!}.$$ Avec la proposition, il vient $N \leq n$.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{\'Eclatements et vari\'et\'es de Moishezon} Nous commen\c cons ce paragraphe par quelques rappels sur les \'eclatements. Ces derniers jouent un r\^ole important dans la th\'eorie des vari\'et\'es de Moishezon et la construction d'exemples explicites. \subsub{\'Eclatements} Une r\'ef\'erence standard est \`a nouveau \cite{G-H78}. \noindent Si $X$ est une vari\'et\'e, et $Y$ une sous-vari\'et\'e de $X$ de codimension sup\'erieure ou \'egale \`a $2$, on construit une vari\'et\'e $\tilde{X}$ appel\'ee {\bf \'eclatement de $X$ le long de $Y$} en rempla\c cant les points de $Y$ par l'espace des directions normales \`a $Y$ dans $X$. On note g\'en\'eralement $\pi : \tilde{X} \to X$ l'\'eclatement et $E := \pi ^{-1}(Y)$ le {\bf diviseur exceptionnel}. La sous-vari\'et\'e $Y$ est appel\'ee {\bf centre de l'\'eclatement}. Par construction, $\pi$ induit un isomorphisme $$\pi _{| \tilde{X} \backslash E } : \tilde{X} \backslash E \to X \backslash Y.$$ \noindent Signalons que le centre de l'\'eclatement $Y$ peut \^etre r\'eduit \`a un point. La restriction de $\pi$ au diviseur exceptionnel $E$ munit $E$ d'une structure de fibr\'e en espaces projectifs au dessus de $Y$~: plus pr\'ecis\'ement, $E$ est isomorphe \`a $\displaystyle{{\Bbb P} (N^{\ast}_{Y/X})}$ (projectivis\'e en droites du fibr\'e normal $N_{Y/X}$ suivant la convention de Grothendieck). De plus, le fibr\'e normal $\displaystyle{N_{E/\tilde{X}} = {\cal O}(E)_{| E}}$ est isomorphe au fibr\'e ${\cal O}_{{\Bbb P} (N_{Y/X}^{\ast})}(-1)$. Mentionnons aussi que le groupe de Picard de $\tilde{X}$ est \'egal \`a $\pi ^*\operatorname{Pic} (X) \oplus {\Bbb Z} \cdot {\cal O}(E)$. Par exemple, le fibr\'e canonique est donn\'e par $$K_{\tilde{X}} = \pi ^{\ast}K_X + (r-1){\cal O}(E)$$ o\`u $r$ est la codimension du centre de l'\'eclatement. Enfin, si $Z$ est une sous-vari\'et\'e de $X$, non incluse dans le centre de l'\'eclatement $Y$, alors l'adh\'erence de $\pi ^{-1}(Z \cap (X \backslash Y))$ dans $\tilde{X}$ est appel\'ee {\bf transform\'ee stricte} de $Z$. Si $Z'$ est la transform\'ee stricte de $Z$, alors $$\pi _{|Z'}~: Z' \to Z$$ \noindent est l'\'eclatement de $Z \cap Y$ dans $Y$. \medskip Le r\'esultat suivant, d\^u \`a A.\ Fujiki et S.\ Nakano \cite{FuN72} donne un crit\`ere pour qu'un diviseur soit le diviseur exceptionnel d'un \'eclatement~; ce crit\`ere est une extension du crit\`ere de Castelnuovo sur les surfaces. Nous l'utiliserons souvent lors de la construction d'exemples explicites de vari\'et\'es de Moishezon non projectives. \bigskip {\bf Th\'eor\`eme (A.\ Fujiki, S.\ Nakano, 1972) } {\em Soit $Z$ une vari\'et\'e et $D$ une sous-vari\'et\'e de $Z$ de codimension $1$. On suppose que $D$ est isomorphe \`a ${\Bbb P} (G)$ o\`u $G$ est un fibr\'e vectoriel sur une vari\'et\'e $Y$ ; on note $p : {\Bbb P} (G) \to Y$ la projection. On suppose enfin que $N_{D/Z}$ est isomorphe \`a ${\cal O}_{{\Bbb P} (G)}(-1)$. Alors, il existe une vari\'et\'e $Z'$ contenant $Y$ comme sous-vari\'et\'e et une application $\pi : Z \to Z'$ de sorte que $\pi$ soit l'\'eclatement de $Z'$ le long de $Y$ et que la restriction de $\pi$ \`a $D$ soit \'egale \`a $p$.} \bigskip Notons qu'au vu de ce qui pr\'ec\`ede, les hypoth\`eses faites sur le diviseur $D$ dans cet \'enonc\'e sont \'evidemment n\'ecessaires~: ce crit\`ere remarquable montre qu'elles sont suffisantes. \subsub{Vari\'et\'es de Moishezon} Le th\'eor\`eme de Siegel motive les d\'efinitions suivantes~: \medskip \noindent {\bf D\'efinition } {\em Un fibr\'e en droites sur une vari\'et\'e compacte de dimension $n$ est dit {\bf gros} ({\bf big} en anglais) si la dimension de l'espace vectoriel des sections holomorphes globales de ses puissances $E^{\otimes k}$ cro\^\i t exactement comme $k^n$.} \medskip \noindent {\bf Remarque } De fa\c con g\'en\'erale, pour un fibr\'e en droites $E$ sur une vari\'et\'e compacte $X$, il existe un entier $\kappa (E)$ tel que la dimension de $H^0(X,E^k)$ cro\^\i t comme $\displaystyle{k^{ \kappa (E)}}$. Cet entier (\'egal \`a $-\infty$ si tous les $H^0(X,E^k)$ sont nuls) est appel\'e {\bf dimension de Kodaira-Iitaka}. Cet entier est inf\'erieur ou \'egal \`a la dimension de $X$, et selon ce qui pr\'ec\`ede, le fibr\'e $E$ est gros si et seulement si $\kappa (E) = \dim X$. \medskip \noindent {\bf D\'efinition } {\em Une vari\'et\'e compacte de dimension $n$ est {\bf de Moishezon} si elle poss\`ede exactement $n$ fonctions m\'eromorphes alg\'ebriquement ind\'ependantes ou, de fa\c con \'equivalente, si elle poss\`ede un fibr\'e en droites gros.} \medskip Les premiers exemples de vari\'et\'es de Moishezon sont les vari\'et\'es projectives. En particulier, toutes les courbes sont de Moishezon. En fait, le th\'eor\`eme suivant, difficile et fondamental, montre qu'une vari\'et\'e de Moishezon n'est pas tr\`es loin d'\^etre projective. Ce r\'esultat est d\^u \`a B.\ Moishezon \cite{Moi67}~: \bigskip \noindent{\bf Th\'eor\`eme (B.\ Moishezon, 1967)} {\em Une vari\'et\'e compacte est de Moishezon si et seulement si elle peut \^etre rendue projective apr\`es un nombre fini d'\'eclatements de centres lisses. On peut m\^eme choisir les centres projectifs.} \bigskip \noindent {\bf Remarque } Mentionnons d\`es \`a pr\'esent que ce th\'eor\`eme ne donne aucune m\'ethode pour choisir les sous-vari\'et\'es le long desquelles il faut \'eclater. Ce probl\`eme figure dans une liste de 100 probl\`emes ouverts en g\'eom\'etrie \'etablie par S.\ T.\ Yau \cite{Yau93}. Nous donnerons (modestement) quelques r\'eponses dans cette direction dans la deuxi\`eme partie de cette th\`ese. \medskip Une cons\'equence du th\'eor\`eme pr\'ec\'edent est le th\'eor\`eme de Chow-Kodaira~: {\em une surface complexe compacte lisse est de Moishezon si et seulement si elle est projective}. En effet, on ne peut qu'\'eclater des points en dimension $2$. Or, de fa\c con g\'en\'erale, si $X$ est une vari\'et\'e compacte et si $\tilde{X}$ est la vari\'et\'e $X$ \'eclat\'ee au point $x$, alors $X$ est projective si et seulement si $\tilde{X}$ l'est (voir par exemple \cite{Kle66}). Ceci explique le fait que tous les exemples de vari\'et\'es de Moishezon qui figurent dans notre travail sont de dimension sup\'erieure ou \'egale \`a $3$. Le premier exemple de vari\'et\'e de Moishezon non projective a \'et\'e construit par H.\ Hironaka (voir par exemple \cite{Har77}). Nous donnons deux constructions dues \`a I.\ Nakamura et K.\ Oguiso \`a la fin de ces pr\'eliminaires. \section{Les in\'egalit\'es de Morse de J.-P.\ Demailly} Nous rappelons ici les in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly~: on renvoie \`a \cite{Dem85} pour la d\'emonstration de ce r\'esultat. Dans ce qui suit, $X$ d\'esigne une vari\'et\'e compacte de dimension $n$ et $(E,h)$ un fibr\'e en droites hermitien sur $X$. Au couple $(E,h)$, on associe pour tout entier $q$ compris entre $0$ et $n$ l'ouvert $X(q,E)$ form\'e des points $x$ de $X$ pour lesquels $\Theta (E)(x)$ poss\`ede exactement $q$ valeurs propres strictement n\'egatives et $n-q$ valeurs propres strictement positives ; finalement on pose $X(\leq q,E) = X(0,E)\cup \cdots \cup X(q,E)$. Les in\'egalit\'es de Morse holomorphes donnent une estimation des groupes de cohomologie \`a valeurs dans les puissances tensorielles $E^{\otimes k}$ en fonction d'int\'egrales de courbure sur $X$. L'\'enonc\'e pr\'ecis est le suivant~: \bigskip \noindent{\bf Th\'eor\`eme (J.-P.\ Demailly, 1985)} {\em Pour tout $q$ compris entre $0$ et $n$, et si $F$ est un fibr\'e vectoriel holomorphe de rang $r$ sur $X$, on a~: (i) $\displaystyle{ \sum _{j=0}^{q}(-1)^{q-j} \dim H^{j}(X,E^{k}\otimes F) \leq r\frac{k^{n}}{n!} \int _{X(\leq q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n})}$ (avec \'e\-ga\-li\-t\'e si $q=n$), (ii) $ \displaystyle{ \dim H^{q}(X,E^{k}\otimes F) \leq r\frac{k^{n}}{n!} \int _{X(q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}).} $} \bigskip Le point (i) est d\'esign\'e sous le nom plus pr\'ecis d'in\'egalit\'es de Morse fortes, alors que le point (ii), qui est une cons\'equence imm\'ediate de (i), est d\'esign\'e sous le nom d'in\'egalit\'es de Morse faibles. Ces in\'egalit\'es ont de nombreuses applications dont certaines tr\`es r\'ecentes~: elles sont utilis\'ees dans les travaux de J.-P.\ Demailly et Y.-T.\ Siu en direction de la conjecture de Fujita \cite{Dem94}, \cite{Siu94}. En g\'en\'eral, ces in\'egalit\'es peuvent se substituer aux th\'eor\`emes d'annulation lorsque la signature de la forme de courbure n'est pas constante. En particulier, J.-P.\ Demailly les a utilis\'ees initialement pour renforcer un r\'esultat de Y.T.\ Siu donnant un crit\`ere analytique suffisant pour qu'une vari\'et\'e soit de Moishezon. Ces r\'esultats fournissent la solution \`a la conjecture de H.\ Grauert et O.\ Riemenschneider \cite{GrR70}~: \bigskip \noindent{\bf Th\'eor\`eme (J.-P.\ Demailly, Y.T.\ Siu, 1985)} {\em Une vari\'et\'e compacte $X$ est de Moishezon d\`es que $X$ poss\`ede un fibr\'e $E$ en droites muni d'une m\'etrique hermitienne lisse dont la forme de courbure $\Theta(E)$ v\'erifie l'une des conditions suivantes : (i) $\Theta(E)$ est partout semi-positive et d\'efinie positive en au moins un point (``crit\`ere de Siu"), (ii) $\displaystyle{ \int_{X(\leq 1,E)} \Theta(E)^{n} > 0}$ (``crit\`ere de Demailly"). } \bigskip Les deux \'enonc\'es d\'ecoulent des in\'egalit\'es de Morse~: elles impliquent dans les deux cas que le fibr\'e $E$ est gros. Avant de commenter ce r\'esultat, rappelons qu'un fibr\'e en droites $E$ sur une vari\'et\'e compacte $X$ est dit {\bf num\'eriquement effectif} (en abr\'eg\'e {\bf nef}) si pour toute m\'etrique hermitienne $\omega$ sur $X$ et pour tout $\varepsilon > 0$, le fibr\'e $E$ poss\`ede une m\'etrique lisse $h_{\varepsilon}$ telle que $\Theta_{h_{\varepsilon}}(E) \geq -\varepsilon\omega$. Cette notion a \'et\'e introduite par J.-P.\ Demailly, T.\ Peternell et M.\ Schneider \cite{DPS94} et admet une formulation \'equivalente sur les vari\'et\'es projectives~: sur une vari\'et\'e projective $X$, un fibr\'e en droites est nef si et seulement si son intersection avec toute courbe de $X$ est semi-positive. Sur une vari\'et\'e quelconque, il peut ne pas y avoir de courbes, cependant il y en a ``suffisamment" sur une vari\'et\'e de Moishezon et Mihai Paun a \'etendu le r\'esultat pr\'ec\'edent \`a ces derni\`eres~: {\em sur une vari\'et\'e de Moishezon, un fibr\'e en droites est nef si et seulement si son intersection avec toute courbe est semi-positive} \cite{Pau95}. \medskip \noindent {\bf Remarque-exemple } Un fibr\'e en droites satisfaisant au crit\`ere de Siu est simultan\'ement gros et nef. Par ailleurs, des exemples de vari\'et\'es de Moishezon satisfaisant les crit\`eres (i) et (ii) sont donn\'es par les vari\'et\'es de Moishezon qui admettent un morphisme g\'en\'eriquement fini vers une vari\'et\'e projective. \section{Exemples explicites} Nous donnons dans ce paragraphe deux constructions, l'une utilis\'ee par I.\ Nakamura \cite{Nak87} et J.\ Koll\'ar \cite{Kol91} et l'autre due \`a K.\ Oguiso. Nous \'etudions en d\'etail la premi\`ere et expliquons plus bri\`evement celle de K.\ Oguiso. Ces constructions nous permettent de montrer qu'aucun des deux crit\`eres analytiques pr\'ec\'edents pour qu'une vari\'et\'e soit de Moishezon n'est n\'ecessaire~: ces \'enonc\'es n'admettent donc pas de r\'eciproque dans le cadre des fibr\'es hermitiens \`a m\'etrique lisse. \subsection{La premi\`ere construction} La construction qui suit exhibe une famille de vari\'et\'es de dimension $3$ complexe d\'epen\-dant d'un param\`etre entier $m$. L'origine de cette construction n'est pas tr\`es claire~; elle est utilis\'ee dans \cite{Nak87} et mentionn\'ee dans \cite{Kol91} \S5. Une des motivations de J.\ Koll\'ar, lorsqu'il mentionne cet exemple, est de construire une vari\'et\'e de Moishezon, dont le groupe de Picard est ${\Bbb Z}$ et dont le g\'en\'erateur gros du groupe de Picard est d'auto-intersection n\'egative. En particulier, ce g\'en\'erateur n'est pas nef. \subsub{Construction explicite} La construction est tr\`es simple~: elle consiste \`a \'eclater ${\Bbb P} ^3$ le long d'une courbe contenue dans une quadrique, et \`a contracter lorsque ceci est possible la transform\'ee stricte de la quadrique sur une courbe rationnelle lisse. Soit donc ${\cal Q} \subset {\Bbb P} ^{3}$ une quadrique lisse, donn\'ee par exemple par l'\'equation homog\`ene $xy=zt$, o\`u $\lbrack x:y:z:t \rbrack$ sont les coordonn\'ees homog\`enes sur ${\Bbb P} ^{3}$. La quadrique ${\cal Q}$ est isomorphe \`a ${\Bbb P}^{1} \times {\Bbb P}^{1}$ et nous notons $$L_{1} = \{ \ast \} \times {\Bbb P} ^{1} \ \mbox{et} \ L_{2} = {\Bbb P} ^{1} \times \{ \ast \}$$ les g\'en\'erateurs de $H_2(\cal Q,{\Bbb Z}) \simeq {\Bbb Z} ^2$. On a \'evidemment $$L_{i} \cdot L_{i}=0 \ \mbox{et} \ L_{1} \cdot L_{2}=1.$$ De plus, tout diviseur $D$ de ${\cal Q}$ est num\'eriquement caract\'eris\'e par un couple d'entiers $(a,b)$ donn\'e par l'intersection de $D$ avec $L_{1}$ et $L_{2}$~: $$(a,b) =(D \cdot L_{1},D \cdot L_{2}) \in {\Bbb Z}^{2}.$$ Ce couple est appel\'e le type de $D$. Par exemple, le diviseur canonique $K_{{\cal Q}}$ est de type $(-2,-2)$. \medskip \noindent {\bf Affirmation } {\em Pour tous $n$ et $m$ entiers positifs, il existe une courbe lisse $C_{n,m}$ incluse dans ${\cal Q}$ et de type $(n,m)$. Une telle courbe est de genre $g_{n,m}=(n-1)(m-1)$ et de degr\'e $n+m$.} \medskip Soit $C_{n,m}$ une telle courbe. Nous \'eclatons alors ${\Bbb P} ^{3}$ le long de $C_{n,m}$~: on obtient une vari\'et\'e projective $\tilde{X}$, et un morphisme $$\pi_{1} : \tilde{X} \to {\Bbb P} ^{3}.$$ Notons $E_{n,m}$ le diviseur exceptionnel de l'\'eclatement~; il est isomorphe \`a ${\Bbb P} (N^{\ast}_{C_{n,m}/{\Bbb P} ^{3}})$. Comme le groupe de Picard de ${\Bbb P} ^3$ est ${\Bbb Z}$, celui de $\tilde{X}$ est ${\Bbb Z} ^{2}$. Si $\tilde{{\cal Q}}$ d\'esigne la transform\'ee stricte de ${\cal Q}$ et $\tilde{L_{i}}$ celle de $L_{i}$, alors $\tilde{{\cal Q}}$ et $\tilde{L_{i}}$ sont respectivement isomorphes \`a ${\cal Q}$ et $L_{i}$ car $C_{n,m}$ est incluse dans ${\cal Q}$. De plus, le type du fibr\'e normal $N_{\tilde{{\cal Q}}/\tilde{X}}$ de $\tilde{{\cal Q}}$ dans $\tilde{X}$ est donn\'e par l'affirmation suivante que nous d\'emontrons plus loin~: \medskip \noindent {\bf Affirmation } {\em On a $N_{\tilde{{\cal Q}}/\tilde{X}} \cdot \tilde{L_{1}} = 2-n$ et $N_{\tilde{{\cal Q}}/\tilde{X}} \cdot \tilde{L_{2}} = 2-m$. } \medskip Comme cas particulier de l'affirmation pr\'ec\'edente, consid\'erons le cas o\`u $n=3$. La restriction \`a $\tilde{L_{1}}$ du fibr\'e $N_{\tilde{{\cal Q}}/\tilde{X}}$ est alors isomorphe au fibr\'e ${\cal O}_{{\Bbb P}^{1}}(-1)$. Par le crit\`ere de contraction de Fujiki-Nakano, il existe donc une vari\'et\'e $X_{m}$ et une application $$\pi_{2} : \tilde{X} \to X_{m}$$ de sorte que $\pi_{2}$ soit l'\'eclatement d'une courbe lisse rationnelle $C_{m}$, de fibr\'e normal projectivement trivial (\'egal \`a ${\cal O}_{{\Bbb P} ^1}(-m)^{\oplus 2}$) tel que le diviseur exceptionnel de $\pi_{2}$ est exactement $\tilde{{\cal Q}}$. Evidemment, $X_{m}$ est bim\'eromorphiquement \'equivalente \`a ${\Bbb P}^{3}$ donc est de Moishezon. De plus, le groupe de Picard de $X_{m}$ est ${\Bbb Z}$. \bigskip \vspace{+3mm} \setlength{\unitlength}{0.0100in} \begin{picture}(470,544)(0,-10) \put(373,117){\ellipse{194}{182}} \put(271,429){\ellipse{212}{200}} \path(157,300)(127,255) \path(129.774,262.766)(127.000,255.000)(133.102,260.547) \path(322,300)(352,255) \path(345.898,260.547)(352.000,255.000)(349.226,262.766) \path(307,87) (308.728,91.454) (310.345,95.593) (311.857,99.431) (313.272,102.984) (314.595,106.268) (315.833,109.298) (318.081,114.656) (320.069,119.183) (321.850,123.000) (325.000,129.000) \path(325,129) (326.547,131.899) (328.570,135.473) (334.000,141.000) \path(334,141) (338.893,141.000) (344.322,138.895) (349.840,137.093) (355.000,138.000) \path(355,138) (360.838,143.296) (362.959,146.772) (364.596,150.952) (365.784,155.951) (366.557,161.882) (366.799,165.233) (366.950,168.861) (367.016,172.778) (367.000,177.000) \path(367,177) (371.390,176.923) (375.473,176.836) (379.264,176.739) (382.779,176.630) (386.032,176.508) (389.041,176.372) (394.382,176.053) (398.925,175.666) (402.794,175.200) (409.000,174.000) \path(409,174) (415.135,172.017) (418.821,170.528) (423.071,168.637) (427.999,166.291) (433.721,163.438) (436.915,161.805) (440.350,160.026) (444.040,158.093) (448.000,156.000) \path(367,162) (371.390,161.923) (375.473,161.836) (379.264,161.739) (382.779,161.630) (386.032,161.508) (389.041,161.372) (394.382,161.054) (398.925,160.666) (402.794,160.200) (409.000,159.000) \path(409,159) (415.135,157.017) (418.821,155.528) (423.071,153.637) (427.999,151.291) (433.721,148.439) (436.915,146.805) (440.350,145.026) (444.040,143.093) (448.000,141.000) \path(307,87) (311.390,86.923) (315.473,86.836) (319.264,86.739) (322.779,86.630) (326.032,86.508) (329.041,86.372) (334.382,86.054) (338.925,85.666) (342.794,85.200) (349.000,84.000) \path(349,84) (355.135,82.017) (358.821,80.528) (363.071,78.637) (367.999,76.291) (373.721,73.439) (376.915,71.805) (380.350,70.026) (384.040,68.093) (388.000,66.000) \path(388,66) (389.728,70.454) (391.345,74.593) (392.857,78.431) (394.272,81.984) (395.595,85.268) (396.833,88.298) (399.081,93.656) (401.069,98.183) (402.850,102.000) (406.000,108.000) \path(406,108) (407.547,110.899) (409.570,114.473) (415.000,120.000) \path(415,120) (419.893,120.000) (425.322,117.895) (430.840,116.093) (436.000,117.000) \path(436,117) (441.838,122.296) (443.959,125.772) (445.596,129.952) (446.784,134.951) (447.557,140.882) (447.799,144.233) (447.950,147.861) (448.016,151.778) (448.000,156.000) \path(184,426) (189.926,423.721) (195.420,421.547) (200.501,419.467) (205.189,417.471) (209.504,415.546) (213.466,413.683) (217.094,411.870) (220.409,410.096) (226.175,406.622) (230.924,403.174) (234.813,399.662) (238.000,396.000) \path(238,396) (240.821,391.325) (243.149,385.576) (244.147,382.408) (245.043,379.101) (245.845,375.700) (246.560,372.248) (247.195,368.788) (247.757,365.364) (248.253,362.019) (248.691,358.796) (249.419,352.892) (250.000,348.000) \path(250,348) (250.000,345.000) \path(250,345) (254.247,349.854) (257.908,354.037) (261.043,357.620) (263.715,360.674) (267.915,365.474) (271.000,369.000) \path(271,369) (275.498,374.379) (278.244,377.726) (281.157,381.243) (284.112,384.730) (286.987,387.986) (292.000,393.000) \path(292,393) (297.433,397.137) (300.943,399.495) (305.134,402.141) (310.121,405.144) (316.018,408.574) (319.344,410.472) (322.940,412.503) (326.821,414.676) (331.000,417.000) \path(199,420) (203.247,424.854) (206.908,429.037) (210.043,432.620) (212.715,435.674) (216.915,440.474) (220.000,444.000) \path(220,444) (224.498,449.379) (227.244,452.726) (230.157,456.243) (233.112,459.730) (235.987,462.986) (241.000,468.000) \path(241,468) (246.433,472.137) (249.943,474.495) (254.134,477.141) (259.121,480.144) (265.018,483.574) (268.344,485.472) (271.940,487.503) (275.821,489.676) (280.000,492.000) \path(184,426) (188.247,430.854) (191.908,435.037) (195.043,438.620) (197.715,441.674) (201.915,446.474) (205.000,450.000) \path(205,450) (209.498,455.379) (212.244,458.726) (215.157,462.243) (218.112,465.730) (220.987,468.986) (226.000,474.000) \path(226,474) (231.433,478.137) (234.943,480.495) (239.134,483.141) (244.121,486.144) (250.018,489.574) (253.344,491.472) (256.940,493.503) (260.821,495.676) (265.000,498.000) \path(292,393) (293.728,397.454) (295.345,401.593) (296.857,405.431) (298.272,408.984) (299.595,412.268) (300.833,415.298) (303.081,420.656) (305.069,425.183) (306.850,429.000) (310.000,435.000) \path(310,435) (311.547,437.899) (313.570,441.473) (319.000,447.000) \path(319,447) (323.893,447.000) (329.322,444.895) (334.840,443.093) (340.000,444.000) \path(340,444) (345.838,449.296) (347.959,452.772) (349.596,456.952) (350.784,461.951) (351.557,467.882) (351.799,471.233) (351.950,474.861) (352.016,478.778) (352.000,483.000) \path(202,384) (205.676,385.460) (209.100,386.805) (215.246,389.167) (220.544,391.121) (225.099,392.702) (229.017,393.945) (232.403,394.885) (238.000,396.000) \path(238,396) (241.793,396.383) (246.115,396.567) (251.123,396.543) (256.976,396.303) (260.268,396.099) (263.832,395.837) (267.685,395.517) (271.849,395.137) (276.343,394.697) (281.186,394.195) (286.398,393.629) (292.000,393.000) \path(13,132) (18.926,129.721) (24.420,127.547) (29.501,125.467) (34.189,123.471) (38.504,121.546) (42.466,119.683) (46.094,117.870) (49.409,116.096) (55.175,112.622) (59.924,109.174) (63.813,105.662) (67.000,102.000) \path(67,102) (69.821,97.325) (72.149,91.576) (73.147,88.408) (74.043,85.101) (74.845,81.700) (75.560,78.248) (76.195,74.788) (76.757,71.364) (77.253,68.019) (77.691,64.796) (78.419,58.892) (79.000,54.000) \path(79,54) (79.000,51.000) \path(79,51) (83.247,55.854) (86.908,60.037) (90.043,63.620) (92.715,66.674) (96.915,71.474) (100.000,75.000) \path(100,75) (104.498,80.379) (107.244,83.726) (110.157,87.243) (113.112,90.730) (115.987,93.986) (121.000,99.000) \path(121,99) (126.433,103.137) (129.943,105.495) (134.134,108.141) (139.121,111.144) (145.018,114.574) (148.344,116.472) (151.940,118.503) (155.821,120.676) (160.000,123.000) \path(94,204) (99.926,201.721) (105.420,199.547) (110.501,197.467) (115.189,195.471) (119.504,193.546) (123.466,191.683) (127.094,189.870) (130.409,188.096) (136.175,184.622) (140.924,181.174) (144.813,177.662) (148.000,174.000) \path(148,174) (150.821,169.325) (153.149,163.576) (154.147,160.408) (155.043,157.101) (155.845,153.700) (156.560,150.248) (157.195,146.788) (157.757,143.364) (158.253,140.019) (158.691,136.796) (159.419,130.892) (160.000,126.000) \path(160,126) (160.000,123.000) \path(28,126) (32.247,130.854) (35.908,135.037) (39.043,138.620) (41.715,141.674) (45.915,146.474) (49.000,150.000) \path(49,150) (53.498,155.379) (56.244,158.726) (59.157,162.243) (62.112,165.730) (64.987,168.986) (70.000,174.000) \path(70,174) (75.433,178.137) (78.943,180.495) (83.134,183.141) (88.121,186.144) (94.018,189.575) (97.344,191.472) (100.940,193.503) (104.821,195.676) (109.000,198.000) \path(13,132) (17.247,136.854) (20.908,141.037) (24.043,144.620) (26.715,147.674) (30.915,152.474) (34.000,156.000) \path(34,156) (38.498,161.379) (41.244,164.726) (44.157,168.243) (47.112,171.730) (49.987,174.986) (55.000,180.000) \path(55,180) (60.433,184.137) (63.943,186.495) (68.134,189.141) (73.121,192.144) (79.018,195.575) (82.344,197.472) (85.940,199.503) (89.821,201.676) (94.000,204.000) \path(67,99) (68.728,103.454) (70.345,107.593) (71.857,111.431) (73.272,114.984) (74.595,118.268) (75.833,121.298) (78.081,126.656) (80.069,131.183) (81.850,135.000) (85.000,141.000) \path(85,141) (86.547,143.899) (88.570,147.473) (94.000,153.000) \path(94,153) (98.893,153.000) (104.322,150.895) (109.840,149.093) (115.000,150.000) \path(115,150) (120.838,155.296) (122.959,158.772) (124.596,162.952) (125.784,167.951) (126.557,173.882) (126.799,177.233) (126.950,180.861) (127.016,184.778) (127.000,189.000) \path(28,147) (33.926,144.721) (39.420,142.547) (44.501,140.467) (49.189,138.471) (53.504,136.546) (57.466,134.683) (61.094,132.870) (64.409,131.096) (70.175,127.622) (74.924,124.174) (78.813,120.662) (82.000,117.000) \path(82,117) (84.821,112.325) (87.149,106.576) (88.147,103.408) (89.043,100.101) (89.845,96.700) (90.560,93.248) (91.195,89.788) (91.757,86.364) (92.253,83.019) (92.691,79.796) (93.419,73.892) (94.000,69.000) \path(94,69) (94.000,66.000) \put(97,123){\ellipse{194}{182}} \path(334,156) (339.926,153.721) (345.420,151.547) (350.501,149.467) (355.189,147.471) (359.504,145.546) (363.466,143.683) (367.094,141.870) (370.409,140.096) (376.175,136.622) (380.924,133.174) (384.813,129.662) (388.000,126.000) \path(388,126) (390.821,121.325) (393.149,115.576) (394.147,112.408) (395.043,109.101) (395.845,105.700) (396.560,102.248) (397.195,98.788) (397.757,95.364) (398.253,92.019) (398.691,88.796) (399.419,82.892) (400.000,78.000) \path(400,78) (400.000,75.000) \put(61,6){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\egtrm ${\Bbb P}^{3}$}}}}} \path(199,444) (204.926,441.721) (210.420,439.547) (215.501,437.467) (220.189,435.471) (224.504,433.546) (228.466,431.683) (232.094,429.870) (235.409,428.096) (241.175,424.622) (245.924,421.174) (249.813,417.662) (253.000,414.000) \path(253,414) (255.821,409.325) (258.149,403.576) (259.147,400.408) (260.043,397.101) (260.845,393.700) (261.560,390.248) (262.195,386.788) (262.757,383.364) (263.253,380.019) (263.691,376.796) (264.419,370.892) (265.000,366.000) \path(265,366) (265.000,363.000) \path(265,498) (270.926,495.721) (276.420,493.547) (281.501,491.467) (286.189,489.471) (290.504,487.546) (294.466,485.683) (298.094,483.870) (301.409,482.096) (307.175,478.622) (311.924,475.174) (315.813,471.662) (319.000,468.000) \path(319,468) (321.821,463.325) (324.149,457.576) (325.147,454.408) (326.043,451.101) (326.845,447.700) (327.560,444.248) (328.195,440.788) (328.757,437.364) (329.253,434.019) (329.691,430.796) (330.419,424.892) (331.000,420.000) \path(331,420) (331.000,417.000) \path(202,384) (203.728,388.454) (205.345,392.593) (206.857,396.431) (208.272,399.984) (209.595,403.268) (210.833,406.298) (213.081,411.656) (215.069,416.183) (216.850,420.000) (220.000,426.000) \path(220,426) (221.547,428.899) (223.570,432.473) (229.000,438.000) \path(229,438) (233.893,438.000) (239.322,435.895) (244.840,434.093) (250.000,435.000) \path(250,435) (255.838,440.296) (257.959,443.772) (259.596,447.952) (260.784,452.951) (261.557,458.882) (261.799,462.233) (261.950,465.861) (262.016,469.778) (262.000,474.000) \path(262,474) (265.676,475.460) (269.100,476.805) (275.246,479.167) (280.544,481.121) (285.099,482.702) (289.017,483.945) (292.403,484.885) (298.000,486.000) \path(298,486) (301.793,486.383) (306.115,486.567) (311.123,486.543) (316.976,486.303) (320.268,486.099) (323.832,485.837) (327.685,485.517) (331.849,485.137) (336.343,484.697) (341.186,484.195) (346.398,483.629) (352.000,483.000) \path(262,456) (265.676,457.460) (269.100,458.805) (275.246,461.167) (280.544,463.121) (285.099,464.702) (289.017,465.945) (292.403,466.885) (298.000,468.000) \path(298,468) (301.793,468.383) (306.115,468.567) (311.123,468.543) (316.976,468.303) (320.268,468.099) (323.832,467.837) (327.685,467.517) (331.849,467.137) (336.343,466.697) (341.186,466.195) (346.398,465.629) (352.000,465.000) \path(238,396) (239.728,400.454) (241.345,404.593) (242.857,408.431) (244.272,411.984) (245.595,415.268) (246.833,418.298) (249.081,423.656) (251.069,428.183) (252.850,432.000) (256.000,438.000) \path(256,438) (257.547,440.899) (259.570,444.473) (265.000,450.000) \path(265,450) (269.893,450.000) (275.322,447.895) (280.840,446.093) (286.000,447.000) \path(286,447) (291.838,452.296) (293.959,455.772) (295.596,459.952) (296.784,464.951) (297.557,470.882) (297.799,474.233) (297.950,477.861) (298.016,481.778) (298.000,486.000) \put(331,165){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $C_{m}$}}}}} \put(334,0){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $X_{m}$}}}}} \put(121,285){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\pi_{1}$}}}}} \put(346,288){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\pi_{2}$}}}}} \put(82,174){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $L_{1}$}}}}} \put(100,135){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $C_{3,m}$}}}}} \put(139,96){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm ${\cal Q}$}}}}} \put(94,105){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $L_{2}$}}}}} \put(232,306){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\tilde{X}$}}}}} \put(310,390){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\tilde{{\cal Q}}$}}}}} \put(328,471){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\tilde{F}$}}}}} \put(277,501){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\tilde{L_{1}}$}}}}} \put(178,447){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $\tilde{L_{2}}$}}}}} \put(208,372){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $E_{3,m}$}}}}} \put(409,87){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\tenrm $E_{3,m}$}}}}} \end{picture} \medskip Avant d'\'etudier plus en d\'etail la vari\'et\'e $X_m$, d\'emontrons les deux affirmations n\'ecessaires \`a sa construction. \medskip \noindent {\bf D\'emonstration des affirmations} L'existence de $C_{n,m}$ r\'esulte du fait que $\displaystyle{{\cal O}(n,m)= \mathop{\rm pr}\nolimits_{1}^{\ast}{\cal O}(n)\otimes \mathop{\rm pr}\nolimits_{2}^{\ast}{\cal O}(m)}$ est tr\`es ample sur ${\Bbb P} ^{1} \times {\Bbb P} ^{1}$. Enfin, le calcul du genre est donn\'e par la formule classique $$\displaystyle{2g_{n,m}-2 = C_{n,m} \cdot (C_{n,m}+K_{{\cal Q}})}.$$ Ici, $$C_{n,m} \cdot C_{n,m}=2nm \ \mbox{et} \ C_{n,m} \cdot K_{{\cal Q}}= -2(n+m).$$ Ceci d\'emontre la premi\`ere affirmation.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Pour la deuxi\`eme, la suite exacte $$ 0 \rightarrow T \tilde{{\cal Q}} \to T\tilde{X}_{| \tilde{{\cal Q}}} \to N_{\tilde{{\cal Q}}/\tilde{X}} \to 0$$ donne $N_{\tilde{{\cal Q}}/\tilde{X}}=K_{\tilde{{\cal Q}}}- K_{\tilde{X}|\tilde{{\cal Q}}}$ o\`u $K_{\tilde{X}}=\pi_{1}^{\ast}K_{{\Bbb P}^{3}}+ {\cal O}(E_{n,m})$. De l\`a~: \vspace{-3mm} \begin{eqnarray*} N_{\tilde{{\cal Q}}/\tilde{X}} \cdot \tilde{L_{i}} & = & K_{\tilde{{\cal Q}}} \cdot \tilde{L_{i}}- \pi_{1}^{\ast}K_{{\Bbb P}^{3}} \cdot \tilde{L_{i}}- {\cal O}(E_{n,m}) \cdot \tilde{L_{i}}\\ & = & K_{{\cal Q}} \cdot L_{i}- K_{{\Bbb P}^{3}} \cdot L_{i}-C_{n,m} \cdot L_{i}. \end{eqnarray*} Or, $K_{{\cal Q}}={\cal O}_{{\Bbb P}^{3}}(-2)_{|{\cal Q}}$ et $K_{{\Bbb P}^{3}}= {\cal O}_{{\Bbb P}^{3}}(-4)$. Ceci conclut le calcul.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsub{Deux propri\'et\'es de $X_m$} Nous montrons ici que les crit\`eres de Demailly et Siu ne sont pas satisfaits pour la vari\'et\'e de Moishezon $X_m$. \bigskip \noindent {\bf Th\'eor\`eme A} {\em La vari\'et\'e $X_m$ v\'erifie les deux propri\'et\'es suivantes~: (i) si $m$ est strictement plus grand que $3$, $X_{m}$ ne poss\`ede pas de fibr\'e en droites \`a la fois gros et nef, et donc ne satisfait pas au crit\`ere de Siu, (ii) si $m$ est strictement plus grand que $5$, $X_{m}$ ne poss\`ede pas de fibr\'e en droites $E$ muni d'une m\'etrique hermitienne lisse $h$ telle que la forme de courbure $\Theta (E)$ v\'erifie : $$ \int_{X(\leq 1,E)}\Theta (E)^{3} > 0.$$ } \medskip L'affirmation (i) est due \`a J.\ Koll\'ar, nous en donnons une preuve \'el\'ementaire, tandis que (ii) est nouveau \`a notre connaissance. \medskip \noindent {\bf D\'emonstration de (i)} Soit $E$ un fibr\'e holomorphe de rang $1$ sur $X_{m}$, que l'on suppose non trivial (le fibr\'e trivial, bien que nef, n'est pas gros ! ). Il existe alors des entiers $k$ et $l$ tels que~: $$\pi_{2}^{\ast}E = \pi_{1}^{\ast}{\cal O}_{{\Bbb P}^{3}}(l)-{\cal O}(kE_{3,m}).$$ \noindent Comme $\tilde{L_{1}}$ est une fibre de $\pi_{2}$, on a~: $\pi_{2}^{*}E \cdot \tilde{L_{1}} = 0$. On en d\'eduit la relation $l=3k$ et donc $$\pi_{2}^{*}E = k(3\pi_{1}^{\ast}{\cal O}_{{\Bbb P}^{3}}(1)-{\cal O}(E_{3,m})),$$ o\`u $k$ est un entier non nul. En particulier, si $\tilde{F}$ est une fibre non triviale de $\pi_{1}$ dans $\tilde{X}$, on a les nombres d'intersection suivants~: $$\left\{ \begin{array}{l} \pi_{2}^{\ast}E \cdot \tilde{L_{2}}=k(3-m) \\ \pi_{2}^{\ast}E \cdot \tilde{F}=k \end{array} \right. $$ \noindent On en d\'eduit que pour $m>3$, le fibr\'e $\pi_{2}^{\ast}E$ n'est pas nef (sinon son intersection avec toute courbe serait positive ou nulle), et donc $E$ n'est pas nef.\hskip 3pt \vrule height6pt width6pt depth 0pt \noindent {\bf D\'emonstration de (ii)} Notons dans la suite ${\cal O}_{X_m}(1)$ le g\'en\'erateur du groupe de Picard de $X_{m}$ tel que $\pi_{2}^{\ast}{\cal O}_{X_m}(1) = \pi_{1}^{\ast}{\cal O}_{{\Bbb P}^{3}}(3)-{\cal O}(E_{3,m})$. \bigskip \noindent {\bf Affirmation } {\em Le fibr\'e canonique $K_{X_{m}}$ est \'egal \`a ${\cal O}_{X_m}(-2)$.} \medskip En effet, on a $$K_{\tilde{X}}= \pi_{2}^{\ast}K_{X_{m}}+{\cal O}(\tilde{{\cal Q}}) = \pi_1^*{\cal O}_{{\Bbb P}^{3}}(-4)+ {\cal O}(E_{3,m})$$ \noindent par construction. Or, $$ {\cal O}(\tilde{{\cal Q}}) = \pi_1^*{\cal O}_{{\Bbb P}^{3}}(2) - {\cal O}(E_{3,m}),$$ d'o\`u l'affirmation. \bigskip \noindent {\bf Affirmation } {\em Les espaces de sections holomorphes $H^{0}(X_m,{\cal O}_{X_m}(k))$ sont nuls pour tout entier $k < 0$ et le fibr\'e ${\cal O}_{X_m}(1)$ est gros.} \medskip Par invariance bim\'eromorphe des plurigenres, les groupes $H^{0}(X_m,{\cal O}_{X_m}(k))$ sont nuls pour tout entier $k$ strictement n\'egatif et pair. Comme $X_{m}$ est de Moishezon de groupe de Picard ${\Bbb Z}$, $X_{m}$ poss\`ede un fibr\'e gros qui n'est donc pas ${\cal O}_{X_m}(-1)$, c'est donc que ${\cal O}_{X_m}(1)$ est gros et que les $H^{0}(X_m,{\cal O}_{X_m}(k))$ sont nuls pour tout entier $k$ strictement n\'egatif. Ceci d\'emontre l'affirmation. \bigskip Par dualit\'e de Serre, on d\'eduit des deux affirmations pr\'ec\'edentes que {\em les groupes de cohomologie $H^{3}(X_{m},{\cal O}_{X_m}(k))$ sont nuls pour tout entier $k > -2$.} \medskip Nous sommes maintenant en mesure de passer \`a la preuve de (ii) proprement dit~: raisonnons par l'absurde et supposons que ${\cal O}_{X_m}(1)$ poss\`ede une telle m\'etrique. D'apr\`es les in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly appliqu\'ees pour $q=1$ au fibr\'e hermitien $({\cal O}_{X_m}(1),h)$, on a~: $$\dim H^{0}(X_{m},{\cal O}_{X_m}(k))- \dim H^{1}(X_{m},{\cal O}_{X_m}(k)) \geq \frac{1}{6} \left(\int_{X(\leq 1,E)}\Theta (E)^{3}\right)k^{3}+ o(k^{3}).$$ \noindent Comme les groupes de cohomologie $H^{3}(X_{m},{\cal O}_{X_m}(k))$ sont nuls pour tout entier $k > -2$, on a successivement~: \vspace{-3mm} \begin{eqnarray*} c_{1}({\cal O}_{X_m}(1))^{3}\frac{k^{3}}{6} + o(k^{3}) & = & \sum _{i=0}^3 (-1)^i \dim H^{i}(X_{m},{\cal O}_{X_m}(k)) \\ & = & \sum _{i=0}^2 (-1)^i \dim H^{i}(X_{m},{\cal O}_{X_m}(k)) \\ & \geq & \dim H^{0}(X_{m},{\cal O}_{X_m}(k))- \dim H^{1}(X_{m},{\cal O}_{X_m}(k)) \\ & \geq & \left(\int_{X(\leq 1,E)}\Theta (E)^{3}\right)\frac{k^{3}}{6}+ o(k^{3}). \end{eqnarray*} \noindent On en d\'eduit que $\displaystyle{c_{1}({\cal O}_{X_m}(1))^{3} \geq \int_{X(\leq 1,E)}\Theta (E)^{3} >0}$. Il suffit donc de montrer pour obtenir la contradiction cherch\'ee que pour $m > 5$, on a $c_{1}({\cal O}_{X_m}(1))^{3} \leq 0$. Or, cette derni\`ere quantit\'e est ais\'ement calculable~: \medskip \noindent {\bf Affirmation } {\em La quantit\'e $c_{1}({\cal O}_{X_m}(1))^{3}$ est \'egale \`a $6-m$.} \medskip En effet \vspace{-3mm} \begin{eqnarray*} c_{1}({\cal O}_{X_m}(1))^{3} & = & c_{1}(\pi_{1}^{*}{\cal O}_{{\Bbb P}^{3}}(3)-{\cal O}(E_{3,m}))^{3} \\ & = & c_{1}({\cal O}_{{\Bbb P}^{3}}(3))^{3} -3\int_{E_{3,m}} c_{1}(\pi_{1}^{*}{\cal O}_{{\Bbb P}^{3}}(3))^{2} \\ & & +3\int_{\tilde{X}}c_{1}(\pi_{1}^{*}{\cal O}_{{\Bbb P}^{3}}(3)) \wedge c_{1}({\cal O}(E_{3,m}))^{2} -E_{3,m}^{3}. \end{eqnarray*} \noindent On a \'evidemment $c_{1}({\cal O}_{{\Bbb P}^{3}}(3))^{3}=27$ et $\displaystyle{\int_{E_{3,m}} c_{1}(\pi_{1}^{*}{\cal O}_{{\Bbb P}^{3}}(3))^{2}= 0.}$ \noindent Pour les deux termes restants, on commence par remarquer que $\displaystyle{c_{1}({\cal O}(E_{3,m}))_{|E_{3,m}}}$ est \'egale \`a $-\xi$ o\`u $\displaystyle{\xi=c_{1}({\cal O}_{{\Bbb P} (N_{C_{3,m}/{\Bbb P} ^{3}}^{*})}(1))}$ d\'esigne la classe fondamentale de l'\'eclatement. On en d\'eduit que~: \vspace{-3mm} \begin{eqnarray*} \int_{\tilde{X}}c_{1}(\pi_{1}^{*}{\cal O}_{{\Bbb P}^{3}}(3)) \wedge c_{1}({\cal O}(E_{3,m}))^{2} & = & -\int_{E_{3,m}} \pi_{1}^{*}c_{1}({\cal O}_{{\Bbb P}^{3}}(3)) \wedge \xi \\ & = & -\int_{C_{3,m}}c_{1}({\cal O}_{{\Bbb P}^{3}}(3)) \\ & = & -3(3+m), \end{eqnarray*} \noindent la derni\`ere \'egalit\'e venant du fait que $C_{3,m}$ est de degr\'e $3+m$ dans ${\Bbb P} ^{3}$ (rappelons que $\displaystyle{{\cal O}_{{\Bbb P} ^{3}}(1)_{|Q}= {\cal O}(1,1)}$). Finalement, il reste \`a calculer $E_{3,m}^{3}$. Pour cela, rappelons que $\xi$ v\'erifie la formule fondamentale suivante~: $$ \xi^{2} - \pi_{1}^{*}c_{1}(N_{C_{3,m}/{\Bbb P}^{3}}^{*})\xi + \pi_{1}^{*}c_{2}(N_{C_{3,m}/{\Bbb P}^{3}}^{*})=0 $$ qui se r\'eduit ici \`a $\displaystyle{\xi^{2} - \pi_{1}^{*}c_{1}(N_{C_{3,m}/{\Bbb P}^{3}}^{*})\xi =0}$. \noindent Il vient alors $\displaystyle{E_{3,m}^{3}= \int_{E_{3,m}}\xi^{2}=\int_{E_{3,m}} \pi_{1}^{*}c_{1}(N_{C_{3,m}/{\Bbb P}^{3}}^{*})\xi = \int_{C_{3,m}}c_{1}(N_{C_{3,m}/{\Bbb P}^{3}}^{*}).}$ \noindent Or la suite exacte~: $$ 0 \rightarrow TC_{3,m} \rightarrow T{\Bbb P}^{3}_{| C_{3,m}} \rightarrow N_{C_{3,m}/{\Bbb P}^{3}} \rightarrow 0$$ donne de suite : $$ \int_{C_{3,m}}c_{1}(N_{C_{3,m}/{\Bbb P}^{3}}^{*})= \int_{C_{3,m}}c_{1}({\cal O}_{{\Bbb P}^{3}}(-4))-2g_{3,m}+2 =-6-8m .$$ \noindent Il reste finalement~: $$c_{1}({\cal O}_{X_m}(1))^{3} = 27 - 27 - 9m + 6 + 8m = 6-m.$$ \noindent Ceci ach\`eve la preuve de l'affirmation et par suite celle du th\'eor\`eme.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{La construction de K.\ Oguiso} Pour ce paragraphe, la r\'ef\'erence est \cite{Ogu94}. Dans cet article, K.\ Oguiso construit une vari\'et\'e de Moishezon non projective, de dimension $3$ et qui est de plus de Calabi-Yau~: ceci signifie que cette vari\'et\'e est simplement connexe et \`a fibr\'e canonique trivial. Mettons en garde le lecteur sur l'usage fait ici de l'expression Calabi-Yau. En effet, la vari\'et\'e en question n'est pas k\"ahl\'erienne. Rappelons plus g\'en\'eralement qu'un th\'eor\`eme de B.\ Moishezon affirme que {\em toute vari\'et\'e de Moishezon k\"ahl\'erienne est projective} (voir le ``survey" de T.\ Peternell \cite{Pet95} pour une d\'emonstration rapide de ce r\'esultat). Le r\'esultat de K.\ Oguiso est le suivant~: \bigskip \noindent {\bf Th\'eor\`eme (K.\ Oguiso, 1994)} {\em Il existe une vari\'et\'e de Moishezon $Y$, de dimension $3$ et de Calabi-Yau telle que $\displaystyle{H^2(Y,{\Bbb Z}) = \operatorname{Pic} (Y) = {\Bbb Z} \cdot L}$ o\`u $L$ satisfait $L^3 := c_1(L)^3 =0$. Autrement dit, la forme cubique d'intersection sur $H^2(Y,{\Bbb Z})$ est identiquement nulle.} \bigskip Nous donnons dans la derni\`ere partie de cette th\`ese une construction, diff\'erente de celle d'Oguiso, permettant de retrouver ce r\'esultat. Dans \cite{Ogu94}, K.\ Oguiso obtient ce th\'eor\`eme comme cons\'equence du r\'esultat suivant~: \bigskip \noindent {\bf Th\'eor\`eme (K.\ Oguiso, 1994)} {\em Soit $d$ un entier positif. Alors il existe une vari\'et\'e projective $X_d$ de dimension $3$, intersection compl\`ete d'une quadrique et d'une quartique dans ${\Bbb P} ^5$ et contenant une courbe rationnelle lisse $C_d$ de degr\'e $d$ dont le fibr\'e normal $N_{C_d/X_d}$ est ${\cal O}_{C_d}(-1)^{\oplus 2}$.} \bigskip Montrons que ce dernier r\'esultat implique le premier. \noindent Soit donc $X_d$ comme ci dessus et soit $\tilde{X}_d$ la vari\'et\'e projective obtenue en \'eclatant $X_d$ le long de la courbe $C_d$. Un argument identique \`a celui d\'evelopp\'e dans la construction pr\'ec\'edente assure que l'on peut contracter le diviseur exceptionnel $E \simeq C_d \times {\Bbb P} ^1 = {\Bbb P} ^1 \times {\Bbb P} ^1$ dans l'autre direction~: on note $Y_d$ la vari\'et\'e obtenue. Par construction, $Y_d$ est de Moishezon, de Calabi-Yau et son groupe de Picard est ${\Bbb Z}$. Enfin, si $\displaystyle{{\cal O}_{Y_d}(1)}$ est le g\'en\'erateur gros de $\displaystyle{\operatorname{Pic} (Y_d)}$, on montre comme pr\'ec\'edemment que $\displaystyle{c_1({\cal O}_{Y_d}(1))^3 = 8 - d^3}$. Le premier r\'esultat est d\'emontr\'e en choisissant $d=2$ i.e $Y := Y_2$. \noindent Comme pour les vari\'et\'es $X_m$ pr\'ec\'edemment construites, {\em les vari\'et\'es $Y_d$ pour $d$ sup\'erieur ou \'egal \`a $2$ ne satisfont pas les crit\`eres de Siu et Demailly.} \medskip \noindent {\bf Remarque } Les vari\'et\'es $Y_d$ ont \'et\'e obtenues apr\`es une transformation birationnelle classique en th\'eorie de Mori appel\'ee ``flop". Nous revenons sur ce type de construction dans la deuxi\`eme partie de cette th\`ese. \subsection{Quelques commentaires} Ces pr\'eliminaires illustrent une partie des motivations de cette th\`ese. En effet, alors que les deux constructions pr\'ec\'edentes montrent que les crit\`eres de J.-P.\ Demailly et Y.-T.\ Siu ne sont pas n\'ecessaires dans le cadre des fibr\'es hermitiens munis d'une m\'etrique ${\cal C}^{\infty}$, nous montrons dans le deuxi\`eme chapitre de cette th\`ese que ces crit\`eres deviennent n\'ecessaires et suffisants dans le cadre plus souple des m\'etriques singuli\`eres. Pour cela, nous \'etendons les in\'egalit\'es de Morse en autorisant un certain type de singularit\'es aux m\'etriques des fibr\'es consid\'er\'es. Remarquons ensuite que les constructions pr\'ec\'edentes donnent des exem\-ples de vari\'et\'es de Moishezon de dimension $3$, de groupe de Picard ${\Bbb Z}$ avec respectivement $-K_X$ gros et $K_X$ trivial. Une des motivations du dernier chapitre de cette th\`ese est de r\'epondre \`a la question suivante~: peut-on construire des exemples analogues avec $K_X$ gros ? Un r\'esultat de J.\ Koll\'ar montre que ce n'est pas possible en dimension $3$~: nous montrons en revanche que de tels exemples existent en dimension sup\'erieure. Ceci nous conduira naturellement \`a \'etudier la structure du centre de l'\'eclatement projectif donn\'e abstraitement par le th\'eor\`eme de Moishezon lorsque la vari\'et\'e devient projective apr\`es un \'eclatement seulement. \chapter{In\'egalit\'es de Morse singuli\`eres} Le but central de ce chapitre est d'\'etendre les in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly au cas d'un fibr\'e en droites $E$ muni d'une m\'etrique singuli\`ere au dessus d'une vari\'et\'e complexe compacte $X$. Ces in\'egalit\'es nous permettent ensuite de caract\'eriser analytiquement les vari\'et\'es de Moishezon. Enfin, nous donnons une version alg\'ebrique singuli\`ere de nos in\'egalit\'es de Morse. \section{M\'etriques singuli\`eres} La notion de m\'etrique singuli\`ere pour des fibr\'es en droites a \'et\'e introduite par J.-P.\ Demailly, A.\ Nadel et H.\ Tsuji. Nous commen\c cons ce paragraphe en rappelant les premi\`eres d\'efinitions et les exemples classiques. Une r\'ef\'erence est \cite{Dem90}. \subsection{Premi\`eres d\'efinitions} Soit $X$ une vari\'et\'e et $E$ un fibr\'e en droites sur $X$. Une m\'etrique hermitienne {\bf singuli\`ere} sur $E$ est donn\'ee localement sur un ouvert trivialisant $U_{\alpha}$ par $$h(\xi _{x}) = ||\xi _x||_h := |\xi|\exp(-\varphi _{\alpha} (x))$$ o\`u la fonction r\'eelle $\varphi _{\alpha}$ est seulement suppos\'ee {\bf localement int\'egrable}. Cette derni\`ere hypoth\`ese suffit \`a donner encore un sens \`a la notion de courbure d'un tel fibr\'e~: en effet, on pose toujours $\displaystyle{\Theta(E) := \frac{i}{\pi} \partial \overline{\partial} \varphi _{\alpha}}$ o\`u le $\partial \overline{\partial}$ est pris au sens des distributions. L'objet ainsi d\'efini n'est plus une forme ${\cal C}^{\infty}$ mais un {\bf courant} (appel\'e courant de courbure) de bi-degr\'e $(1,1)$. Le lemme de Dolbeault-Grothendieck \'etant vrai pour les courants, la cohomologie de De Rham est calculable aussi bien avec les courants qu'avec les formes et la classe de cohomologie du courant de courbure $\Theta(E)$ est \'egale comme dans le cas lisse \`a la premi\`ere classe de Chern du fibr\'e $E$. \medskip \noindent {\bf Exemples } Les deux exemples suivants jouent un r\^ole essentiel. (i) Si $\displaystyle{D = \sum \alpha _j D_j}$ est un diviseur de $X$ et si $g_j$ est l'\'equation locale de $D_j$ sur un ouvert $U_{\alpha}$, alors la fonction $\displaystyle{\varphi _{\alpha} = \sum \alpha _j \log |g_j|}$ d\'efinit une m\'etrique singuli\`ere naturelle sur le fibr\'e en droites ${\cal O}(D)$ associ\'e au diviseur $D$. \noindent Pour cette m\'etrique, l'\'equation de Lelong-Poincar\'e est $\displaystyle{\Theta({\cal O}(D)) = [D]}$ o\`u $[D]$ d\'esigne le courant d'int\'egration sur le diviseur $D$. \noindent Remarquons que dans le cas o\`u $X$ est une courbe complexe et $E$ le fibr\'e tangent de $X$, la construction pr\'ec\'edente donne une m\'etrique plate avec des masses de Dirac de courbure~: ce type de m\'etrique est bien connu en g\'eom\'etrie riemannienne sous le nom de {\bf m\'etrique plate \`a singularit\'es coniques}. (ii) Soit $E$ un fibr\'e en droites sur une vari\'et\'e $X$ et soient $s_1,\ldots \!,s_p$ des sections de $E^{\otimes k}$. Alors, $$||\xi_x||^2 := \left( \frac{|\theta _{\alpha}(\xi)|^2}{|\theta _{\alpha}(s_1(x))|^2+ \cdots+|\theta _{\alpha}(s_p(x))|^2} \right) ^{1/k}$$ o\`u $\theta _{\alpha}$ est une trivialisation locale de $E$ et $E^{\otimes k}$, d\'efinit une m\'etrique singuli\`ere sur le fibr\'e $E$. La fonction $\varphi _{\alpha}$ est ici $\displaystyle{\varphi _{\alpha}(x) = \frac{1}{2k} \log(|\theta _{\alpha}(s_1(x))|^2+ \cdots+|\theta _{\alpha}(s_p(x))|^2)}$. \medskip Comme dans le cas ${\cal C}^{\infty}$, nous d\'esirons pouvoir donner un sens \`a l'expression ``\^etre \`a courbure positive ou strictement positive" pour un fibr\'e en droites muni d'une m\'etrique singuli\`ere. Pour cela, rappelons qu'un courant $T$ de bi-degr\'e $(1,1)$ est {\bf positif} si pour toutes formes $\alpha_1,\ldots \!,\alpha_{n-1}$ de type $(1,0)$ et ${\cal C}^{\infty}$ \`a support compact, on a~: $$ < T , i \alpha_1 \wedge \overline{\alpha}_1\wedge \ldots \wedge i \alpha_{n-1} \wedge \overline{\alpha}_{n-1} > \ \ \geq \ \ 0,$$ o\`u $<\cdot, \cdot >$ est la dualit\'e entre les courants et les formes. On note alors $T \geq 0$. De m\^eme, on dira qu'un courant $T$ de bi-degr\'e $(1,1)$ est {\bf strictement positif} s'il existe une m\'etrique $\omega$ hermitienne ${\cal C}^{\infty}$ sur $X$ telle que $T - \omega$ est un courant positif. On note dans ce cas $T >0$. Nous introduisons alors la d\'efinition suivante~: \medskip \noindent {\bf D\'efinition } {\em Un fibr\'e en droites muni d'une m\'etrique singuli\`ere est positif (respectivement strictement positif) si le courant de courbure associ\'e est positif (respectivement strictement positif).} \medskip En reprenant les notations de l'exemple ci-dessus, le fibr\'e en droites ${\cal O}(D)$ muni de sa m\'etrique singuli\`ere est positif si et seulement si le diviseur $D$ est effectif (i.e tous les coefficients $\alpha _j$ sont positifs ou nuls). L'exemple (ii) d\'efinit toujours un courant de courbure positif car une fonction de la forme $\varphi = \log \sum |f_j|^2$ o\`u les $f_j$ sont des fonctions holomorphes est plurisousharmonique (en abr\'eg\'e {\bf psh}). \subsection{Faisceaux d'id\'eaux multiplicateurs de Nadel} L'\'etude d'un fibr\'e en droites muni d'une m\'etrique singuli\`ere est facilit\'ee par l'utilisation d'un outil pertinent associ\'e aux singularit\'es de la m\'etrique~: il s'agit d'un faisceau d'id\'eaux introduit par A.\ Nadel. \medskip \noindent {\bf D\'efinition } {\em Soit $(E,h)$ un fibr\'e en droites sur une vari\'et\'e $X$. On appelle ``faisceau multiplicateur de Nadel" le faisceau d'id\'eaux ${\cal I}(h)$ des germes de fonctions holomorphes $L^2$ par rapport au poids de la m\'etrique singuli\`ere, i.e l'ensemble des germes $f \in {\cal O}_{X,x}$ tels que $|f|^2 \exp (-2\varphi)$ est int\'egrable par rapport \`a la mesure de Lebesgue dans des coordonn\'ees locales au voisinage de $x$. Plus g\'en\'eralement, si $\varphi$ est une fonction r\'eelle sur un ouvert $\Omega$, nous notons aussi ${\cal I}(\varphi)$ le faisceau des germes $f \in {\cal O}_{\Omega}$ tels que $|f|^2 \exp (-2\varphi)$ est localement int\'egrable.} \medskip La propri\'et\'e essentielle satisfaite par ce faisceau d'id\'eaux, due \`a Nadel, est que {\em si $\varphi$ est une fonction plurisousharmonique, alors ${\cal I}(\varphi)$ est un faisceau coh\'erent.} \medskip Avant de donner des exemples, mentionnons le r\'esultat suivant qui g\'en\'eralise le th\'eor\`eme de Kawamata-Viehweg~: il s'agit du th\'eor\`eme d'annulation de Nadel \cite{Nad89} \cite{Dem89}. \bigskip \noindent {\bf Th\'eor\`eme d'annulation de Nadel} {\em Soit $X$ une vari\'et\'e k\"ahl\'erienne compacte, et $E$ un fibr\'e en droites muni d'une m\'etrique singuli\`ere \`a courbure strictement positive. Alors, pour tout $q \geq 1$, on a~: $$ H^q (X,{\cal O}(E+K_X) \otimes {\cal I}(h)) = 0.$$ } \medskip \vspace{-2mm} Ce r\'esultat montre que pour g\'en\'eraliser un th\'eor\`eme d'annulation dans le contexte des fibr\'es en droites munis de m\'etrique singuli\`ere, une bonne m\'ethode consiste \`a ``tordre" la cohomologie du fibr\'e par le faisceau de Nadel. \medskip Donnons quelques exemples de faisceaux multiplicateurs, qui bien que tr\`es simples jouent un r\^ole important. \medskip \noindent {\bf Exemples} (i) Soit $\varphi$ une fonction r\'eelle sur un ouvert $\Omega$ de ${\Bbb C} ^n$ contenant l'origine. Si $\varphi$ est minor\'ee au voisinage de l'origine, alors pour tout $x$ proche de $0$, on a clairement ${\cal I}(\varphi)_x = {\cal O}_{\Omega,x}$. En particulier, si $\varphi$ est continue sur $\Omega$, alors ${\cal I}(\varphi) = {\cal O}_{\Omega}$. (ii) Pla\c cons nous dans $\displaystyle{{\Bbb C} ^n}$ au voisinage de l'origine et, pour $\alpha_{1},\ldots \!,\alpha_{p}$ des r\'eels positifs et $k$ un entier naturel, posons~: $$ \varphi _k (z)= \frac{k}{2}\log(|z_{1}|^{2\alpha_{1}}+ \cdots +|z_{p}|^{2\alpha_{p}}).$$ Alors, ${\cal I}(\varphi _k)_{{\Bbb C} ^n,0}$ est ${\cal O}_{{\Bbb C} ^n,0}$-engendr\'e par les $\displaystyle{\prod _{j=1}^{p}z_{j}^{\beta_{j}}}$ tels que~: $$ \sum_{j=1}^{p}\frac{\beta_{j}+1}{\alpha_{j}} > k .$$ \indent En particulier, si tous les $\alpha_{i}$ sont \'egaux \`a $\alpha$, alors ${\cal I}(\varphi _k)_{{\Bbb C} ^n,0}$ est ${\cal O}_{{\Bbb C} ^n,0}$-engendr\'e par les $\displaystyle{\prod _{j=1}^{p}z_{j}^{\beta_{j}}}$ o\`u $$\displaystyle{\sum_{j=1}^{p}\beta_{j} \geq \lfloor k\alpha - p \rfloor +1}$$ ($\lfloor x \rfloor$ d\'esigne la partie enti\`ere du r\'eel $x$). Autrement dit, $\displaystyle{{\cal I}(\varphi _k) = {\cal I}_{Y}^{ \lfloor k\alpha \rfloor -p+1}}$ o\`u $Y$ est la sous-vari\'et\'e $\{ z_{1}= \cdots =z_{p}=0 \}$ et ${\cal I}_{Y}$ son id\'eal annulateur. (iii) Soit $\varphi$ une fonction de la forme $\sum \alpha _j \log |g_j|$ o\`u les $\alpha _j$ sont des r\'eels positifs et o\`u les fonctions holomorphes $g_j$ sont telles que les $D_j := g_j^{-1}(0)$ soient des diviseurs irr\'eductibles lisses se coupant transversalement (on dit alors que $\displaystyle{D = \sum \alpha _j D_j}$ est un diviseur {\bf lisse \`a croisements normaux}). Dans ce cas, le faisceau multiplicateur ${\cal I}(\varphi)$ s'identifie au faisceau inversible de rang un ${\cal O}(-\sum_{j} \lfloor \alpha _j \rfloor D_{j})$. \medskip \noindent {\bf D\'emonstration} Le point (i) est trivial. Pour (ii), il s'agit d'estimer l'int\'egrale suivante sur un voisinage de $0$~: $$ I := \int_{D(0,\varepsilon) ^n} \frac{|\sum a_{\beta}z_1 ^{\beta _1} \ldots z_n ^{\beta _n}|^2}{ (|z_{1}|^{2\alpha_{1}}+ \cdots +|z_{p}|^{2\alpha_{p}})^k} \ d\lambda(z).$$ Le passage en coordonn\'ees polaires $z_j = \rho _j e^{i\theta _j}$ donne $$ I = (2\pi)^n \sum |a_{\beta}|^2 \int_{[0,\varepsilon]^n} \frac{\rho_1^{2\beta_{1}+1}\cdots\rho_n^{2\beta_{n}+1}} {(|\rho_{1}|^{2\alpha_{1}}+ \cdots +|\rho_{p}|^{2\alpha_{p}})^k}\ d\rho_1 \ldots d\rho_n,$$ puis $$ I = (2\pi)^n \sum |a_{\beta}|^2 \frac{\varepsilon^{2\beta_{p+1}+2}\ldots \varepsilon^{2\beta_{n}+2}} {(2\beta_{p+1}+2)\ldots (2\beta_{n}+2)} \int_{[0,\varepsilon]^p} \frac{\rho_1^{2\beta_{1}+1}\cdots\rho_p^{2\beta_{p}+1}} {(\rho_{1}^{2\alpha_{1}}+ \cdots +\rho_{p}^{2\alpha_{p}})^k}\ d\rho_1 \ldots d\rho_p.$$ Par le changement de variables $u_j = \rho_j^{\alpha_j}$, l'int\'egrale ci-dessus est \'egale \`a $$ \int_{[0,\varepsilon]^p} \frac{u_1^{((2\beta _1 +2)/\alpha_1) -1}\ldots u_p^{((2\beta _p +2)/\alpha_p)-1}} {(u_1^2+ \cdots + u_p^2)^k} \ du_1 \ldots du_p,$$ Par homog\'en\'eit\'e, cette derni\`ere int\'egrale converge si et seulement si l'int\'egrale $$\displaystyle{ \int_0^{\varepsilon} \frac{ t^{ 2\sum_{j=1}^p ((\beta _j +1)/\alpha_j) - p}}{t^{2k}}t^{p-1} \ dt}$$ converge, soit $\displaystyle{2\sum_{j=1}^p \frac{\beta _1 +1}{\alpha_1} - p -2k + p - 1 > -1}$. Ceci donne bien la condition annonc\'ee. Pour le point (iii), il s'agit de d\'eterminer le crit\`ere pour qu'une fonction de la forme $$\displaystyle{ \frac{|f|^{2}}{|g_1|^{2\alpha_1}\ldots |g_n|^{2\alpha_n}}}$$ \noindent soit dans $L_{\mbox{\scriptsize loc}}^{1}$. Comme les $g_j$ fournissent des coordonn\'ees locales, et si $p_{j}$ d\'esigne l'ordre d'annulation de $f$ le long de $D_j = \{ g_{j}=0 \}$, la condition n\'ecessaire et suffisante est que $2p_{j}-2\alpha _j > -2$, soit $p_{j} \geq \lfloor \alpha _j \rfloor$. Comme les sections du fibr\'e ${\cal O}(-D_j)$ s'identifient aux fonctions holomorphes s'annulant \`a un ordre sup\'erieur ou \'egal \`a un le long de $D_j$, le r\'esultat en d\'ecoule.\hskip 3pt \vrule height6pt width6pt depth 0pt \section{In\'egalit\'es de Morse singuli\`eres} Dans cette partie, nous \'enon\c cons et d\'emontrons notre version singuli\`ere des in\'e\-galit\'es de Morse holomorphes. \subsection{\'Enonc\'e du r\'esultat principal} Dans tout ce qui suit, $X$ d\'esigne une vari\'et\'e compacte et $(E,h)$ un fibr\'e en droites sur $X$ muni d'une m\'etrique singuli\`ere. A $(E,h)$, on associe pour tout entier $k$ positif le faisceau d'id\'eaux de Nadel donn\'e par la m\'etrique singuli\`ere $h^k$ du fibr\'e $E^k$~; il s'agit ici du faisceau des germes de fonctions holomorphes $f$ telles que $|f|^{2}\exp(-2k\varphi)$ est $L_{\mbox{\scriptsize loc}}^{1}$ (o\`u $\exp (-\varphi)$ est le poids local de $h$). Nous notons ce faisceau ${\cal I}_{k} (h)$. Pour des raisons qui apparaitront plus loin, nous sommes contraints de n'accepter qu'un certain type de singularit\'es, que nous appelons {\bf ``singularit\'es analytiques"}. Plus pr\'ecis\'ement, nous faisons l'hypoth\`ese suivante sur $h$ (c'est-\`a-dire localement sur $\varphi$ )~: \bigskip \noindent {\bf Hypoth\`ese ({\cal S}) :} {\em la fonction $\varphi$ s'\'ecrit localement~: $$ \varphi=\frac{c}{2}\log(\sum \lambda_{j}|f_{j}|^{2})+\psi,$$ o\`u les $f_{j}$ sont holomorphes, les $\lambda_{j}$ sont des fonctions r\'eelles positives ${\cal C}^{\infty}$ sans z\'eros communs, $\psi$ est ${\cal C}^{\infty}$ et $c$ est un rationnel positif ou nul. Mentionnons que la somme $\sum \lambda_{j}|f_{j}|^{2}$ peut poss\'eder une infinit\'e de termes.} \medskip Cette hypoth\`ese implique que la fonction $\varphi$ est {\bf quasi plurisousharmonique} (ce qui signifie que son hessien complexe est minor\'e par une $(1,1)$-forme \`a coefficients continus) et donc en particulier que le faisceau $ {\cal I}_{k} (h)$ est un faisceau coh\'erent d'apr\`es le r\'esultat de A.\ Nadel pr\'ec\'edemment rencontr\'e. En terme de courbure, l'hypoth\`ese ({\cal S}) implique en particulier que le courant de courbure $\displaystyle{ \Theta (E)= \frac{i}{\pi} \partial \overline{\partial} \varphi}$ est quasi positif (c'est-\`a-dire minor\'e par une $(1,1)$-forme \`a coefficients continus). \bigskip \noindent {\bf Remarque } Signalons d\`es maintenant que cette hypoth\`ese est ``raisonnable" car c'est pr\'ecis\'ement par des fonctions ayant ce type de singularit\'es que J.-P.\ Demailly approche une fonction quasi plurisousharmonique quelconque \cite{Dem92}. Ceci aura une importance capitale dans les applications. Remarquons aussi que tous les exemples rencontr\'es jusqu'\`a pr\'esent satisfont l'hypoth\`ese ({\cal S}). Nous renvoyons ici au paragraphe 2.3.1 pour une discussion plus compl\`ete de l'origine de ce type de singularit\'es autoris\'ees. \medskip Dans notre contexte, nous introduisons les notations suivantes~: \medskip \noindent {\bf Notations } Le symbole $X(q,E)$ d\'esigne l'ouvert de $X$ form\'e des points $x$ au voisinage desquels $\varphi$ est born\'ee et $\Theta (E) (x)$ poss\`ede exactement $q$ valeurs propres strictement n\'egatives et $n-q$ valeurs propres strictement positives~; finalement et comme dans le cas lisse $X(\leq q,E) = X(0,E)\cup \cdots \cup X(q,E)$. \medskip Notre r\'esultat est le suivant~: \bigskip \noindent {\bf Th\'eor\`eme B } {\em Soit $(E,h)$ \`a singularit\'es analytiques sur une vari\'et\'e compacte $X$ de dimension $n$ et soit $F$ un fibr\'e de rang $r$ sur $X$. Alors pour tout $q$ compris entre $0$ et $n$~: \noindent \ (i) $\displaystyle{\ \sum _{j=0}^{q}(-1)^{q-j} \dim H^{j}(X,{\cal O}(E^{k}\otimes F) \otimes {\cal I}_{k}(h)) \leq r\frac{k^{n}}{n!} \int _{X(\leq q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}) }$,\\ \noindent (a\-vec \'egalit\'e si $q=n$), \noindent \ (ii) $\displaystyle{ \ \dim H^q(X,{\cal O}(E^{k}\otimes F) \otimes {\cal I}_{k}(h)) \leq r\frac{k^{n}}{n!} \int _{X(q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}).}$} \bigskip Comme dans le cas o\`u la m\'etrique est lisse, le point (ii) est cons\'equence imm\'ediate du point (i). \medskip Remarquons que comme dans le th\'eor\`eme d'annulation de Nadel, le ph\'enom\`ene nouveau par rapport au cas ${\cal C}^{\infty}$ est la pr\'esence des faisceaux d'id\'eaux~; les groupes de cohomologie que nous estimons sont les groupes de cohomologie \`a valeurs dans les grandes puissances de $E$, tordues par la suite des faisceaux multiplicateurs naturellement associ\'ee aux singularit\'es de la m\'etrique. Evidemment, si la m\'etrique est lisse, nous retrouvons les in\'egalit\'es de J.-P.\ Demailly car tous les ${\cal I}_{k}(h)$ sont \'egaux au faisceau structural ${\cal O}_X$. Cependant, mentionnons d\`es maintenant que notre d\'emonstration repose sur les in\'egalit\'es dans le cas ${\cal C}^{\infty}$ ! \subsection{D\'emonstration du th\'eor\`eme B} \subsub{Plan de la preuve} La d\'emarche suivie pour d\'emontrer nos in\'egalit\'es est la suivante~: a) apr\`es \'eclatement de $X$ le long de sous-vari\'et\'es d\'efinies par les singularit\'es de $h$, on se ram\`ene \`a un fibr\'e muni d'une m\'etrique lisse~; on peut alors appliquer les in\'egalit\'es de Morse holomorphes dans le cas $h$ lisse \`a la vari\'et\'e obtenue $\tilde{X}$, b) on relie les groupes de cohomologie sur $\tilde{X}$ \`a ceux de $X$. Pour cela, nous \'etudions le comportement des faisceaux multiplicateurs par rapport aux \'eclatements en montrant que les dimensions des groupes de cohomologie associ\'es sont asymptotiquement de m\^eme dimension. \subsub{R\'eduction au cas lisse} Commen\c cons par expliquer la premi\`ere partie de la preuve. Pour cela, il est bon de remarquer que l'hypoth\`ese ({\cal S}) implique que les singularit\'es de la m\'etrique $h$ sont localis\'ees le long d'un ensemble analytique, d\'efini localement par $ \{ x|\ \forall j,\ f_{j}(x)=0 \}$. Comme nous l'avons vu dans les exemples, cet ensemble analytique n'est pas n\'ecessairement irr\'eductible et ses composantes irr\'eductibles sont en g\'en\'eral de dimension quelconque. La r\'eduction au cas lisse consiste dans un premier temps \`a se ramener \`a une vari\'et\'e $\tilde{X}$, obtenue en \'eclatant $X$ le long de centres lisses $\pi : \tilde{X} \to X$ de telle sorte que la m\'etrique $\tilde{h}=\pi^{*}h$ sur le fibr\'e $\tilde{E}=\pi^{*}E$ n'ait ses singularit\'es qu'en codimension $1$ (ou, de fa\c con \'equivalente, de telle sorte que le faisceau $\tilde{{\cal I}}_{k} (\tilde{h})$ soit inversible). Dans un deuxi\`eme temps, nous appliquerons les in\'egalit\'es de Morse dans le cas lisse. \bigskip \noindent {\bf a) D\'esingularisation de ${\cal I}_{k} (h)$} \medskip Pour rendre les faisceaux ${\cal I}_{k} (h)$ localement libres, l'id\'ee (classique) est d'\'eclater l'id\'eal ``engendr\'e par les $f_j$". Cependant, une difficult\'e appara\^{\i}t ici car la donn\'ee des $f_j$ est locale sur $X$ et la notion d'id\'eal engendr\'e par ces fonctions n'a donc pas de sens {\em a priori}. Nous avons cependant la proposition suivante~: \medskip \noindent {\bf Proposition } {\em Il existe un faisceau d'id\'eaux global ${\cal J}$, qui co\"{\i}ncide avec la cl\^oture int\'egrale du faisceau d'id\'eaux engendr\'e par les $f_{j}$ sur chaque ouvert o\`u $\varphi$ s'\'ecrit comme dans l'hypoth\`ese ({\cal S}). } \medskip \noindent {\bf D\'emonstration} En effet, c'est un fait bien connu (r\'esultant par exemple du th\'eor\`eme de Brian\c con-Skoda \cite{BSk74}, voir aussi \cite{L-T74}) que la cl\^oture int\'egrale du faisceau d'id\'eaux engendr\'e par les $f_{j}$ est donn\'ee par~: $${\cal J}_{x} = \{ f \in {\cal O}_{X,x} \ ; \ \exists C >0 \ , \ |f(z)| \leq C\exp(\frac{1}{c}\varphi(z)) \ \mbox{au voisinage de}\ x \}.$$ Si maintenant $\varphi _{\alpha}$ et $\varphi _{\beta}$ d\'esignent les poids de la m\'etrique sur des ouverts trivialisants $U_{\alpha}$ et $U_{\beta}$, alors sur l'intersection $U_{\alpha} \cap U_{\beta}$, on a $\varphi _{\alpha} = \varphi _{\beta} + O(1)$. La caract\'erisation ci-dessus implique donc bien que ${\cal J}$ est d\'efini globalement sur $X$.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Nous sommes en mesure de d\'emontrer la proposition suivante~: \medskip \noindent {\bf Proposition } {\em Sous les hypoth\`eses pr\'ec\'edentes, il existe une vari\'et\'e $\tilde{X}$ et $\pi : \tilde{X} \rightarrow X$ une compos\'ee d'un nombre fini d'\'eclate\-ments de centres lisses tels que le fibr\'e $\tilde{E} :=\pi^{*}E$ muni de la m\'etrique singuli\`ere $\tilde{h}=\pi^{*}h$ de poids local $\exp(-\tilde{\varphi})$ v\'erifie la propri\'et\'e suivante : pour tout $x_{0} \in \tilde{X}$, il existe des coordonn\'ees holomorphes $w_{1},\ldots \!,w_{n}$ centr\'ees en $x_{0}$ et une fonction $\tilde{\psi}$ de classe ${\cal C}^{\infty}$ telles que~: $$ \tilde{\varphi} (w) = c \sum_{j}a_{j}\log|g_{j}(w)|+\tilde{\psi}(w)$$ o\`u les $a_{j}$ sont des entiers positifs ou nuls et o\`u les $g_{j}$ sont irr\'eductibles dans ${\cal O}_{\tilde{X},x_{0}}$ et d\'efinissent un diviseur lisse \`a croisements normaux.} \medskip \noindent {\bf D\'emonstration} \'Eclatons l'id\'eal ${\cal J}$ de sorte que l'image inverse $\pi^{-1}{\cal J}.{\cal O}_{X'}$ soit un faisceau inversible. Par le th\'eor\`eme d'aplatissement d'Hironaka \cite{Hir75}, on peut dominer cet \'eclatement par une vari\'et\'e $\tilde{X}$ obtenue par une suite d'\'eclatements de centres lisses dans $X$, et l'image inverse de ${\cal J}$ est toujours inversible ! \noindent Mais alors, la m\'etrique image r\'eciproque sur $\pi^{*}E$ est donn\'ee par \begin{eqnarray*} \ \tilde{\varphi} & = & \frac{c}{2}\log(\sum (\lambda_{j}\circ\pi)|f_{j}\circ\pi|^{2})+ \psi\circ\pi \\ & = & \frac{c}{2}\log(|g|^{2}) + \frac{c}{2}\log(\sum (\lambda_{j}\circ\pi)|h_{j}|^{2})+ \psi\circ\pi \\ & = & \frac{c}{2}\log(|g|^{2}) + \tilde{\psi}, \end{eqnarray*} \noindent o\`u on a not\'e $g$ le g\'en\'erateur local du faisceau d'id\'eaux engendr\'e par les $f_{j}\circ\pi$. Si la d\'ecomposition de $g$ en facteurs irr\'eductibles dans $\displaystyle{{\cal O}_{\tilde{X},x_{0}}}$ s'\'ecrit~: $\displaystyle{g = \prod_{j}g_{j}^{a_{j}}}$, on a le r\'esultat apr\`es application du th\'eor\`eme de d\'esingularisation d'Hironaka \cite{Hir64} pour rendre le diviseur d\'efini par les $g_{j}$ \`a croisements normaux.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Dans tout ce qui suit, les objets utilis\'es sont ceux obtenus apr\`es application de la proposition pr\'ec\'edente. Notons $\displaystyle{\tilde{{\cal I}}_{k} (\tilde{h})}$ le faisceau d'id\'eaux des germes de fonctions holomorphes sur $\tilde{X}$ telles que $\displaystyle{|f|^{2}e^{-2k\tilde{\varphi}}}$ est $L_{\mbox{\scriptsize loc}}^{1}$ (c'est le faisceau multiplicateur de Nadel associ\'e au fibr\'e $\pi^*E$). Si nous notons $b_{j,k} = \lfloor ca_{j}k \rfloor$ et $\tilde{D}_{j}$ le diviseur d\'efini localement par $g_{j}=0$, alors la d\'etermination du faisceau $\displaystyle{\tilde{{\cal I}}_{k} (\tilde{h})}$ rel\`eve des exemples pr\'ec\'edents si bien que le lemme suivant en d\'ecoule~: \medskip \noindent {\bf Lemme }{\em Sous les conditions pr\'ec\'edentes, le faisceau d'id\'eaux $\tilde{{\cal I}}_{k} (\tilde{h})$ s'identifie au faisceau inversible de rang un ${\cal O}(-\sum_{j}b_{j,k}\tilde{D}_{j})$.} \medskip \noindent {\bf b) Exemples } \medskip Illustrons ce qui pr\'ec\`ede en reprenant les notations de l'exemple (ii) du 2.1.2. Dans ces cas \'evidemment simples, il n'est nul besoin d'appliquer les th\'eor\`emes d'Hironaka~: on explicite directement le choix des \'eclatements ! \medskip Si on suppose que tous les $\alpha_{i}$ sont \'egaux \`a $\alpha$, nous avons vu alors que ${\cal I}(\varphi_{k} )$ est \'egal \`a ${\cal I}_{Y}^{ \lfloor k\alpha \rfloor -p+1}$ o\`u $Y$ est la sous-vari\'et\'e de codimension $p$ donn\'ee par $\{ z_{1}= \cdots =z_{p}=0 \}$. \noindent Si $p=1$, le faisceau d'id\'eaux est d\'ej\`a inversible, sinon \'eclatons ${\Bbb C} ^n$ le long de $Y$. L'expression de la nouvelle m\'etrique est donn\'ee dans la premi\`ere carte par $$\displaystyle{\tilde{\varphi}(w)= \alpha\log(|w_{1}|) + \frac{1}{2}\log(1+|w_{2}|^{2\alpha}+\cdots +|w_{p}|^{2\alpha})}$$ si bien que $\tilde{{\cal I}}(\varphi_{k}) = {\cal O}(- \lfloor k\alpha \rfloor D)$ o\`u $D$ est le diviseur exceptionnel de l'\'eclatement. Il suffit donc dans ce cas d'un \'eclatement en codimension $p$ pour obtenir le r\'esultat souhait\'e. \medskip De m\^eme, si $\alpha$ est un entier positif et si $\displaystyle{\varphi(z) = \frac{1}{2}\log(|z_{1}|^{2}+|z_{2}|^{2\alpha})}$ dans ${\Bbb C} ^n$, il faut cette fois $\alpha$ \'eclatements en codimension~2. \noindent En effet, \'eclatons ${\Bbb C} ^n$ le long de $\{ z_{1}=z_{2}=0 \}$. L'expression de la nouvelle m\'etrique est donn\'ee dans la premi\`ere carte par $$\displaystyle{\tilde{\varphi}(w)= \log(|w_{1}|) + \frac{1}{2}\log(1+|w_{1}w_{2}|^{2\alpha})}$$ qui est de la forme voulue alors qu'on obtient dans la deuxi\`eme carte $$\displaystyle{\tilde{\varphi}(w)= \log(|w_{2}|) + \frac{1}{2}\log(|w_{1}|^{2}+|w_{2}|^{2(\alpha-1)})}.$$ On \'eclate alors dans la deuxi\`eme carte le long de $\{ w_{1}=w_{2}=0 \}$. En r\'ep\'etant ce proc\'ed\'e $\alpha$ fois, on obtient une m\'etrique de la forme souhait\'ee en tout point. \noindent D\'ecrivons le faisceau d'id\'eaux obtenu~: pour tout $j$ compris entre $1$ et $\alpha$, notons $D_{j}$ la transform\'ee stricte dans $\tilde{X}$ du diviseur exceptionnel du $j$-i\`eme \'eclatement. Alors $D_{j}$ et $D_{j+1}$ se coupent transversalement et on a $\tilde{{\cal I}}_{k} (\tilde{\varphi}) = {\cal O}(-kD)$ o\`u $D$ d\'esigne le diviseur \`a croisements normaux $\displaystyle{D=\sum_{j=1}^{\alpha}jD_{j}}$. \medskip \noindent {\bf c) In\'egalit\'es de Morse sur $\tilde{X}$ } \medskip On montre maintenant comment appliquer les in\'egalit\'es de Morse holomorphes dans le cas ${\cal C}^{\infty}$ sur $\tilde{X}$ au fibr\'e en droites $\tilde{E} = \pi ^{\ast}E$ muni de la m\'etrique image r\'eciproque. Pour cela, nous devons montrer que $(\tilde{E})^k \otimes \tilde{{\cal I}}_{k} (\tilde{h})$ peut essentiellement s'\'ecrire comme la $k$-i\`eme puissance tensorielle d'un fibr\'e en droites hermitien fixe. Notons $\displaystyle{c=\frac{u}{v}}$ et supposons que $k = vk'$ est un multiple du d\'enomina\-teur de $c$. Comme $b_{j,k}=ca_{j}k$, le faisceau multiplicateur $\displaystyle{\tilde{{\cal I}}_k (\tilde{h})}$ est \'egal au faisceau inversible $\displaystyle{{\cal O}(-k'\tilde{D})}$ o\`u $\displaystyle{\tilde{D}=u\sum_{j}a_{j}\tilde{D}_{j}}$. Avec ces notations, on a la~: \medskip \noindent {\bf Proposition } {\em Notons $\hat{E}$ le fibr\'e $\tilde{E}^{v} \otimes {\cal O}(-\tilde{D})$. Alors~: (i) pour tout $k = vk'$, on a $(\hat{E})^{k'} = \tilde{E}^{k} \otimes \tilde{{\cal I}}_{k} (\tilde{h}),$ (ii) la m\'etrique hermitienne sur $\hat{E}$, produit de la m\'etrique $\tilde{h} ^v$ et de la m\'etrique singuli\`ere naturelle sur ${\cal O}(-\tilde{D})$, est une m\'etrique hermitienne ${\cal C}^{\infty}$. De plus, et en dehors des singularit\'es de la m\'etrique $\tilde{h}$, on a l'\'egalit\'e $\Theta(\hat{E}) = v\Theta(\tilde{E})=\pi^{\ast}\Theta(E)$. } \medskip \noindent {\bf D\'emonstration } Le point (i) est \'evident. Pour le point (ii), la m\'etrique produit naturelle est donn\'ee localement par le poids $\displaystyle{\tilde{\chi}(z)= v \tilde{\phi} (z) -\sum_{j=1}^{n}ua_{j}\log|g_{j}| = v \tilde{\psi}(z).}$ Ainsi, cette m\'etrique est lisse et l'\'egalit\'e de courbure d\'ecoule de suite du fait que $\displaystyle{\partial \overline{\partial} \log|g_j| = 0}$ l\`a o\`u $g_j$ ne s'annule pas. \hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Cette proposition nous permet d'estimer les groupes de cohomologie qui nous int\'eressent~: \medskip \noindent {\bf Proposition }{\em On a pour tout $k$ : $$\sum_{j=0}^{q}(-1)^{q-j}\dim H^{j}(\tilde{X},{\cal O}(\tilde{E}^{k} \otimes \tilde{F}) \otimes \tilde{{\cal I}}_{k} (\tilde{h})) \leq r\frac{k^{n}}{n!} \int _{X(\leq q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}) $$ o\`u l'int\'egrale est prise sur les points lisses de la m\'etrique de $E$. } \medskip \newpage \noindent {\bf D\'emonstration} D'apr\`es la proposition pr\'ec\'edente, on peut appliquer les in\'egalit\'es de Morse de Demailly au fibr\'e $\hat{E}$, si bien que pour $k=k'v$, on a~: $$\sum_{j=0}^{q}(-1)^{q-j}\dim H^{j}(\tilde{X},{\cal O}(\tilde{E}^{k} \otimes \tilde{F}) \otimes \tilde{{\cal I}}_{k} (\tilde{h})) \leq r\frac{k'{}^{n}}{n!} \int _{\tilde{X}(\leq q,\hat{E})} (-1)^{q} \Theta (\hat{E})^{n} + o(k'{}^{n}).$$ \noindent Relions alors l'int\'egrale de courbure sur $\tilde{X}$ \`a une int\'egrale de courbure sur $X$. \noindent Comme $\Theta(\hat{E})=v\Theta(\tilde{E})$ sur les points lisses de la m\'etrique de $\tilde{E}$ si $k=k'v$, on a $$k'{}^{n} \int _{\tilde{X}(\leq q,\hat{E})} \Theta (\hat{E})^{n} + o(k'{}^{n})= k^{n} \int _{\tilde{X}(\leq q,\tilde{E})} \Theta (\tilde{E})^{n} + o(k^{n}).$$ \noindent Notons alors $S$ la r\'eunion des diviseurs exceptionnels de $\pi : \tilde{X} \rightarrow X $~; $S$ est n\'egligeable pour la mesure de Lebesgue et on a \begin{eqnarray*} \int _{\tilde{X}(\leq q,\tilde{E})} \Theta (\tilde{E})^{n} & = & \int _{\tilde{X}(\leq q,\tilde{E}) \setminus S} \Theta (\tilde{E})^{n} = \int _{\tilde{X}(\leq q,\tilde{E}) \setminus S} \Theta (\pi ^{*}E)^{n} \\ & = & \int _{\tilde{X}(\leq q,\tilde{E}) \setminus S} \pi^{*}(\Theta (E)^{n}) = \int _{X(\leq q,E) \setminus \pi(S)} \Theta (E)^{n} \\ & = & \int _{X(\leq q,E)} \Theta (E)^{n}. \end{eqnarray*} \noindent On a donc pour les entiers $k$ multiples d'un entier fixe~: $$\sum_{j=0}^{q}(-1)^{q-j}\dim H^{j}(\tilde{X},{\cal O}(\tilde{E}^{k} \otimes \tilde{F}) \otimes \tilde{{\cal I}}_{k} (\tilde{h})) \leq r\frac{k^{n}}{n!} \int _{X(\leq q,E)} (-1)^{q} \Theta (E)^{n} + o(k^{n}) $$ o\`u l'int\'egrale est prise sur les points lisses de la m\'etrique de $E$. \medskip Pour terminer cette preuve, il suffit de montrer que l'estimation pr\'ec\'edente est valable sans restriction sur $k$. \noindent En reprenant les notations du lemme et en posant $\displaystyle{c=\frac{u}{v}}$, on a, pour $k=k'v+r$ $$ b_{j,k} = ua_{j}k'+r' $$ o\`u $r'$ ne prend qu'un nombre fini de valeurs enti\`eres. On en d\'eduit alors \begin{eqnarray*} \! \tilde{E} ^{k} \otimes \tilde{F} \otimes {\cal O}(-\sum_{j}b_{j,k}D_{j}) & = & (\tilde{E}^{v})^{k'}\otimes \tilde{E}^{r'} \otimes \tilde{F} \otimes {\cal O}(-k'u\sum_{j}a_{j}D_{j}) \otimes {\cal O}(-r'\sum_{j}D_{j})\\ & = & (\tilde{E}^{v})^{k'}\otimes {\cal O}(-k'u\sum_{j}a_{j}D_{j}) \otimes \hat{F}_{r'}. \end{eqnarray*} On raisonne alors comme pr\'ec\'edemment~: on munit $\hat{F}_{r'}$ d'une m\'etrique lisse quelconque tandis que $(\tilde{E}^{v})^{k'}\otimes {\cal O}(-k'u\sum_{j}a_{j}D_{j})$ est muni de la m\'etrique lisse naturelle donn\'ee localement par $v\tilde{\psi} (z)$. Ceci d\'emontre la proposition.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsub{Lien entre cohomologie sur $\tilde{X}$ et cohomologie sur $X$} Pour achever la preuve du th\'eor\`eme B, il reste \`a relier les groupes $$H^{q}(\tilde{X},{\cal O}(\tilde{E}^{k} \otimes \tilde{F}) \otimes \tilde{{\cal I}}_{k} (\tilde{h}))$$ de la proposition pr\'ec\'edente et les $$H^{q}(X,{\cal O}(E^{k}\otimes F) \otimes {\cal I}_{k}(h))$$ qui nous int\'eressent directement. Nous adoptons ici une d\'emarche un peu diff\'erente de celle de notre travail \cite{Bo93b}. Soient $X$ une vari\'et\'e compacte et $\mu : X' \to X$ l'\'eclatement de $X$ le long d'une sous-vari\'et\'e lisse $Y$. Soit $E$ un fibr\'e en droites sur $X$, muni d'une m\'etrique hermitienne singuli\`ere $h$ de poids local $\exp (-\varphi)$. On note ${\cal I}_k(\varphi)$ le faisceau multiplicateur associ\'e \`a la m\'etrique sur $E^k$. Soit enfin $F$ un fibr\'e vectoriel sur $X$. Nous montrons dans ce paragraphe la proposition suivante~: \bigskip \noindent {\bf Proposition } {\em Supposons qu'au voisinage de tout point du diviseur exceptionnel de $\mu$, la fonction $\varphi \circ \mu$ s'\'ecrive~: $$ \varphi \circ \mu = \alpha \log |f| + \psi,$$ o\`u $\alpha $ est strictement positif, $f$ d\'esigne une \'equation locale du diviseur exceptionnel de $\mu$ et $\psi$ est une fonction psh. \noindent Alors, si $k \alpha >1$, on a pour tout entier $q \geq 0$~: $$ H^q(X',K_{X'} \otimes \mu^*(E^k \otimes F) \otimes {\cal I}_k(\varphi \circ \mu)) \simeq H^q(X,K_X \otimes E^k \otimes F \otimes {\cal I}_k(\varphi)) .$$} \vspace*{-3mm} \noindent {\bf Commentaires } Nous ne supposons plus dans ce dernier r\'esultat que la m\'etrique est \`a singularit\'es analytiques, ni que les faisceaux ${\cal I}_k(\varphi \circ \mu)$ sont inversibles. Le point important est que l'hypoth\`ese faite sur l'\'ecriture de $\varphi \circ \mu$ est automatiquement satisfaite si $\varphi$ est \`a singularit\'es analytiques et le centre de l'\'eclatement $Y$ est inclus dans le lieu singulier de la m\'etrique. La fin de la d\'emonstration de nos in\'egalit\'es de Morse se conclut par application r\'ep\'et\'ee de cette proposition aux \'eclatements successifs dont $\pi : \tilde{X} \to X$ est la compos\'ee.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Avant de d\'emontrer la proposition, mentionnons le probl\`eme suivant~: \medskip \noindent {\bf Question } {\em Si les singularit\'es de $\varphi$ sont quelconques, a-t-on toujours $$ \dim H^q(X',K_{X'} \otimes \mu^*(E^k \otimes F) \otimes {\cal I}_k(\varphi \circ \mu)) = \dim H^q(X,K_X \otimes E^k \otimes F \otimes {\cal I}_k(\varphi)) + o(k^n) ?$$} \vspace*{-3mm} \noindent {\bf D\'emonstration de la proposition } Elle repose tout d'abord sur le fait que les faisceaux de Nadel se comportent bien par image directe. De fa\c con pr\'ecise, si $(E,h)$ est un fibr\'e en droites muni d'une m\'etrique singuli\`ere au dessus d'une vari\'et\'e $X$, et si $\mu : \tilde{X} \to X$ est une modification, alors \cite{Dem94} $$\mu _*(K_{\tilde{X}}\otimes {\cal I}(\mu^*h)) = K_X \otimes {\cal I}(h).$$ \noindent \'Evidemment, ceci ne suffit pas pour d\'emontrer qu'il y a isomorphisme en cohomologie~: l'obstruction est mesur\'ee par les images directes sup\'erieures. Ainsi, il suffit de montrer, gr\^ace au th\'eor\`eme de Leray, que pour tout entier $q \geq 1$ et pour tout $k$ assez grand, $$R^{q}\mu_{*}(K_{X'} \otimes {\cal I}_k(\varphi \circ \mu)) = 0.$$ \noindent On note $r$ la codimension de $Y$, centre de l'\'eclatement. \noindent Dans ce cas, le faisceau $q$-i\`eme image directe sup\'erieure $R^{q}\mu_{*}(K_{X'} \otimes {\cal I}_k(\varphi \circ \mu))$ est un faisceau \`a support dans $Y$, la fibre au dessus d'un point $y$ de $Y$ \'etant \'egale \`a~: $$ F_{k,y} = \limind_{y \in U} H^{q}(\mu^{-1}(U),(K_{X'} \otimes {\cal I}_k(\varphi \circ \mu))_{|\mu^{-1}(U)}),$$ o\`u la limite porte sur les voisinages $U$ de $y$ dans $X$. \noindent Soit donc $U$ un ouvert de Stein voisinage de $y$ et soit $\omega$ une m\'etrique hermitienne sur $X'$. Il s'agit de montrer que pour toute forme $u$ de type $(n,q)$ sur $\mu^{-1}(U)$, \`a coefficients localement $L^2$ satisfaisant~: (i) $\displaystyle{\overline{\partial}u = 0}$, (ii) $\displaystyle{ I := \int _{\mu^{-1}(U)}|u|^2 \exp(-2k \varphi \circ \mu) dV_{\omega} < +\infty ,}$ \noindent il existe une forme $v$ de type $(n,q-1)$ \`a coefficients localement $L^2$ satisfaisant~: (i)' $\displaystyle{\overline{\partial}v = u}$, (ii)' $\displaystyle{ \int _{\mu^{-1}(U)}|v|^2 \exp(-2k \varphi \circ \mu) dV_{\omega} < +\infty .}$ \noindent R\'esoudre un tel probl\`eme est en g\'en\'eral possible gr\^ace aux estimations $L^2$ de H\"ormander. La difficult\'e ici est que la fonction $k \varphi \circ \mu$ n'est pas strictement psh au voisinage du diviseur exceptionnel $D$ de l'\'eclatement. C'est exactement ici que nous utilisons l'hypoth\`ese faite sur $\varphi$. \noindent En effet, l'\'egalit\'e $$ \varphi \circ \mu = \alpha \log |f| + \psi$$ se traduit par~: $$ {\cal I}_k(\varphi \circ \mu) = {\cal O}(- \lfloor k\alpha \rfloor D) \otimes {\cal I} \left( k\psi + (k\alpha - \lfloor k\alpha \rfloor)\log |f|) \right).$$ \noindent Il s'agit alors de montrer l'annulation des groupes $$H^{n,q}\left (\mu^{-1}(U),{\cal O}(- \lfloor k\alpha \rfloor D) \otimes {\cal I}(k\psi + (k\alpha - \lfloor k\alpha \rfloor)\log |f|\right ),$$ autrement dit de r\'esoudre le probl\`eme du $\overline{\partial}$~: (i)' $\displaystyle{\overline{\partial}v = u}$, (ii)' $\displaystyle{ \int _{\mu^{-1}(U)}|v|^2 \exp(-2k \psi) dV_{\omega} < +\infty ,}$ \noindent pour des $(n,q)$ formes {\em \`a valeurs dans le fibr\'e ${\cal O}(- \lfloor k\alpha \rfloor D)$}. Or, le fibr\'e ${\cal O}(-D)_{|D} = {\cal O}_D (1)$ est strictement positif sur les fibres de l'\'eclatement. On peut donc munir ${\cal O}(-\lfloor k\alpha \rfloor D)$ d'une m\'etrique \`a courbure strictement positive au voisinage du diviseur exceptionnel, et comme $\psi$ peut \^etre suppos\'ee strictement psh en dehors de $D$, nous sommes maintenant dans les hypoth\`eses d'application des estimations $L^2$ de H\"ormander.\hskip 3pt \vrule height6pt width6pt depth 0pt \section{Caract\'erisations analytiques des vari\'et\'es \- de \- Moishezon} Dans la lign\'ee des conditions suffisantes donn\'ees par Y.-T.\ Siu et J.-P.\ Demailly pour caract\'eriser les vari\'et\'es de Moishezon, nous montrons la caract\'erisation suivante~: \bigskip \noindent {\bf Th\'eor\`eme C} {\em Une vari\'et\'e compacte $X$ de dimension $n$ est de Moishezon si et seulement si il existe sur $X$ un courant $T$ ferm\'e de bidegr\'e $(1,1)$ tel que~: (i) $ \displaystyle{\{ T \} \in H^{2}(X,{\Bbb Z}) }$, (ii) $\displaystyle{ T= \frac{i}{\pi} \partial \overline{\partial} \varphi + \alpha }$, o\`u $\varphi$ est une fonction r\'eelle \`a singularit\'es analytiques et o\`u $\alpha$ est un repr\'esentant $ {\cal C}^{\infty}$ de $\{ T \}$, (iii) $\displaystyle{\int_{X(\leq 1,T)} T^{n} > 0}$ o\`u l'int\'egrale est prise sur les points lisses du courant $T$. } \bigskip L'id\'ee de donner une caract\'erisation analytique des vari\'et\'es de Moishezon en terme de courant de courbure est aussi pr\'esente dans un travail de S.\ Ji et B.\ Shiffman \cite{JiS93} simultan\'e au n\^otre. Nos in\'egalit\'es permettent ainsi de d\'emontrer le r\'esultat suivant~: \bigskip \noindent {\bf Th\'eor\`eme D \cite{JiS93}, \cite{Bo93b}} {\em Une vari\'et\'e compacte $X$ de dimension $n$ est de Moishezon si et seulement si il existe sur $X$ un courant $T$ ferm\'e de bidegr\'e $(1,1)$ tel que~: (i) $ \displaystyle{\{ T \} \in H^{2}(X,{\Bbb Z}) }$, (ii) le courant $T$ est strictement positif (i.e minor\'e par une $(1,1)$-forme ${\cal C}^{\infty}$ hermitienne). } \bigskip Ces deux \'enonc\'es fournissent l'extension naturelle aux vari\'et\'es de Moishezon du th\'eor\`eme de plongement de Kodaira pour les vari\'et\'es projectives. Avant de d\'emontrer les r\'esultats ci-dessus, nous rappelons deux r\'esultats sur les courants. \subsection{Deux rappels} Il est bien connu \cite{G-H78} que la seule obstruction pour qu'une $(1,1)$-forme ferm\'ee r\'eelle ${\cal C}^{\infty}$ soit la forme de courbure d'un fibr\'e en droites hermitien est que sa classe de cohomologie soit enti\`ere, i.e appartienne \`a $H^2(X,{\Bbb Z})$. Nous montrons dans la proposition suivante que ce r\'esultat persiste pour les courants quasi positifs~: \medskip \noindent {\bf Proposition } {\em Soit $T$ un courant quasi-positif ferm\'e, de bi-degr\'e $(1,1)$ dont la classe de cohomologie $\{ T \}$ est dans $H^2(X,{\Bbb Z})$. Alors, il existe un fibr\'e en droites $E$ muni d'une m\'etrique singuli\`ere dont le courant de courbure est \'egal \`a $T$.} \medskip \noindent {\bf D\'emonstration } Elle suit la d\'emonstration du cas ${\cal C}^{\infty}$ \cite{S-S85}. On recouvre la vari\'et\'e $X$ par des ouverts de Stein contractiles dont les intersections mutuelles sont elles aussi contractiles. Sur chaque ouvert $U_{\alpha}$, on \'ecrit $\displaystyle{T = \frac{i}{\pi}\partial \overline{\partial} \varphi_{\alpha} }$. Comme $T$ est quasi positif, les fonctions $\varphi_{\alpha}$ sont quasi plurisousharmoniques, donc localement int\'egrables (c'est le seul point qui diff\`ere du cas ${\cal C}^{\infty}$). \noindent De l\`a, on \'ecrit successivement $\varphi_{\alpha \beta} := \varphi_{\beta}-\varphi_{\alpha}$ sur l'intersection $U_{\alpha}\cap U_{\beta}$, puis $$ i(\overline{\partial} - \partial)\varphi_{\alpha \beta} = 2\pi d u_{\alpha \beta}.$$ Les fonctions $c_{\alpha \beta \gamma} := u_{\beta \gamma} - u_{\alpha \gamma} + u_{\alpha \beta}$ sont constantes. Comme $\{ T \}$ est enti\`ere, la classe $\{ c_{\alpha \beta \gamma} \}$ l'est aussi. Il existe donc une $1$-c\^ochaine \`a coefficients r\'eels $\{ b_{\alpha \beta} \}$ telle que~: $$ c_{\alpha \beta \gamma} + b_{\beta \gamma} - b_{\alpha \gamma} + b_{\alpha \beta} = m_{\alpha \beta \gamma} \in {\Bbb Z}.$$ On pose alors~: $$ g_{\alpha \beta} := \exp (\varphi_{\alpha \beta} + 2i\pi(u_{\alpha \beta} + b_{\alpha \beta})).$$ Les $g_{\alpha \beta}$ sont holomorphes sans z\'ero et forment un cocycle~: on note $E$ le fibr\'e en droites associ\'e. Comme $$ |g_{\alpha \beta}|\exp(-\varphi_{\alpha})=\exp(-\varphi_{\beta}),$$ les poids $\exp(-\varphi_{\alpha})$ d\'efinissent une m\'etrique singuli\`ere sur $E$ dont le courant de courbure est $T$.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Comme nos in\'egalit\'es de Morse supposent que la m\'etrique est \`a singularit\'es analytiques, nous avons besoin d'un r\'esultat d'approximation permettant de s'y ramener. De fa\c con g\'en\'erale, \'etant donn\'e un courant ferm\'e $T$ sur une vari\'et\'e compacte, c'est un probl\`eme classique que de vouloir le r\'egulariser dans la m\^eme classe de cohomologie. Malheureusement, il n'est pas possible en g\'en\'eral de r\'egulariser ${\cal C}^{\infty}$ tout en perdant aussi peu de positivit\'e que souhait\'e. L'obstruction \`a le faire est mesur\'ee par les nombres de Lelong du courant. Pour ne pas perdre de positivit\'e, on r\'egularise en autorisant des singularit\'es analytiques. Le r\'esultat que nous utilisons est le th\'eor\`eme d'approximation des courants de J.-P.\ Demailly \cite{Dem92}~: \bigskip \noindent {\bf Th\'eor\`eme (J.-P.\ Demailly, 1992)} {\em Soit $T$ un courant ferm\'e de bi-degr\'e $(1,1)$ de sorte que $T \geq \alpha$ o\`u $\alpha$ est une $(1,1)$ forme ${\cal C}^{\infty}$. Alors pour toute m\'etrique hermitienne $\omega$ de classe ${\cal C}^{\infty}$ sur $X$, il existe une suite de courants $T_{\varepsilon}$ telle que~: (i) $\{ T_{\varepsilon} \} = \{ T \}$, (ii) $T_{\varepsilon}$ tend (faiblement) vers $T$ lorsque $\varepsilon$ tend vers $0$, (iii) $T_{\varepsilon} \geq \alpha - \varepsilon \omega$, (iv) $\displaystyle{ T_{\varepsilon} = i \partial \overline{\partial} \varphi _{\varepsilon} + \beta }$, o\`u $\varphi _{\varepsilon}$ est une fonction r\'eelle \`a singularit\'es analytiques et o\`u $\beta$ est un repr\'esentant $ {\cal C}^{\infty}$ de $\{ T \}$. } \bigskip Il n'est pas question ici de donner la d\'emonstration de ce r\'esultat, mais pour \'eclairer le lecteur, mentionnons la premi\`ere \'etape de la d\'emonstration, qui est une version locale du r\'esultat~: \medskip \noindent {\bf Proposition } {\em Soit $\varphi$ une fonction psh dans la boule unit\'e $B$ de ${\Bbb C} ^n$. Pour $k$ entier positif, notons ${\cal H}(k \varphi)$ l'espace de Hilbert d\'efini de la fa\c con suivante~: $$ {\cal H}(k \varphi) = \{ f \in {\cal O}(B) \ | \ \int _B |f|^2 \exp(-2k\varphi) d\lambda < +\infty \}.$$ Soit enfin $(\sigma_{j,k})_j$ une base orthonorm\'ee de ${\cal H}(k \varphi)$. Alors la suite de fonctions $$ \varphi _k := \frac{1}{2k} \log (\sum_j |\sigma_{j,k}|^2) $$ converge vers $\varphi$ simplement et dans $L_{\mbox{\scriptsize loc}}^1$ lorsque $k$ tend vers $+\infty$.} \medskip La difficult\'e du th\'eor\`eme r\'eside dans le recollement des diverses approximations locales donn\'ees par la proposition pr\'ec\'edente. Les fonctions $\lambda _j$ figurant dans la d\'efinition de l'hypoth\`ese {\cal S} proviennent essentiellement de partition de l'unit\'e. En particulier, elles ne s'annulent pas toutes simultan\'ement si bien que les singularit\'es sont vraiment donn\'ees par les z\'eros communs d'une famille de fonctions holomorphes. \subsection{D\'emonstration des th\'eor\`emes C et D} Nous d\'emontrons ici les deux caract\'erisations analytiques. Pour cela, on commence par remarquer que le sens faux dans le cadre des m\'etriques lisses est v\'erifi\'e de fa\c con presque imm\'ediate dans le cadre plus souple des vari\'et\'es de Moishezon. Soit en effet $X$ une vari\'et\'e de Moishezon et $\pi : \hat{X} \to X$ une modification projective. Si $\hat{\omega}$ est une $(1,1)$ forme ${\cal C}^{\infty}$ d\'efinie positive sur $\hat{X}$ telle que $\{ \hat{\omega} \} \in H^{2}(\hat{X},{\Bbb Z})$, et si $\omega$ est une $(1,1)$ forme ${\cal C}^{\infty}$ d\'efinie positive sur $X$, alors la forme $\pi^{*}\omega$ est ${\cal C}^{\infty}$ et semi-positive. Il existe donc une constante $A > 0$ telle que $\hat{\omega} \geq A\pi^{*}\omega$. Par cons\'equent, le courant $T=\pi_{*} \hat{\omega}$ v\'erifie $T \geq A\omega$~: en effet, si $\alpha$ est une $(n-1,n-1)$ forme positive ${\cal C}^{\infty}$ sur $X$, on a \vspace{-3mm} \begin{eqnarray*} \ <T,\alpha> & = & \int_{\hat{X}} \hat{\omega} \wedge \pi^{*} \alpha \\ & \geq & \int_{\hat{X}} A \pi^{*}\omega \wedge \pi^{*} \alpha \\ & = & A \int_{X} \omega \wedge \alpha \\ & = & <A\omega,\alpha>. \end{eqnarray*} Le courant $T$ satisfait les points (i) et (ii) du th\'eor\`eme D, et le th\'eor\`eme d'approximation des courants de Demailly rappel\'e pr\'ec\'edemment implique qu'il existe un courant $T' \in \{ T \}$ ayant localement les singularit\'es de l'hypoth\`ese {\cal S} et tel que $T' \geq (A/2) \omega$. Ainsi, $T'$ v\'erifie la conclusion du th\'eor\`eme C. \medskip Pour la r\'eciproque dans le th\'eor\`eme C, soit $T$ satisfaisant (i), (ii) et (iii). Alors, il existe un fibr\'e en droites sur $X$ muni d'une m\'etrique hermitienne singuli\`ere dont le courant de courbure est \'egal \`a $T$. Les in\'egalit\'es de Morse singuli\`eres impliquent que~: $$ \dim H^0(X,E^k \otimes {\cal I}_k(h)) - \dim H^1(X,E^k \otimes {\cal I}_k(h)) \sim_{k \to +\infty} k^n.$$ {\em A fortiori}, $$ \dim H^0(X,E^k) \sim_{k \to +\infty} k^n ,$$ i.e le fibr\'e $E$ est gros et $X$ est de Moishezon. Pour le th\'eor\`eme D, si $T$ est strictement positif, on peut supposer par le th\'eor\`eme d'approximation des courants que $T$ a localement les singularit\'es de l'hypoth\`ese {\cal S}. On conclut alors comme ci-dessus.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip \noindent {\bf Remarque } La preuve du th\'eor\`eme D donn\'ee par S.\ Ji et B.\ Shiffman suit la d\'emarche suivante~: on approche, comme pr\'ec\'edemment, le courant $T$ par un courant strictement positif et singulier sur un ensemble analytique $S$. Puis on montre directement gr\^ace aux estimations $L^2$ de H\"ormander appliqu\'ees \`a la vari\'et\'e k\"ahl\'erienne compl\`ete $X \backslash S$ que les grandes puissances de $E$ engendrent les $1$-jets en dehors de $S$. \subsection{Quelques commentaires} Les th\'eor\`emes C et D sont satisfaisants car ils donnent une caract\'erisation analytique des vari\'et\'es de Moishezon. Cependant, on souhaiterait pouvoir se dispenser de l'hypoth\`ese faite sur les singularit\'es dans le th\'eor\`eme C ou affaiblir l'hypoth\`ese de stricte positivit\'e du th\'eor\`eme D. On conjecture le r\'esultat suivant~: \medskip \noindent {\bf Conjecture } {\em Une vari\'et\'e compacte $X$ est de Moishezon si et seulement si il existe sur $X$ un courant $T$ positif ferm\'e de bi-degr\'e $(1,1)$ tel que~: (i) $ \displaystyle{\{ T \} \in H^{2}(X,{\Bbb Z}) }$, (ii) il existe un ouvert $U$ sur lequel $T$ est strictement positif.} \medskip La condition (ii) est une fa\c con d'assurer que le support de $T$ n'est pas n\'egligeable pour la mesure de Lebesgue. Evidemment, l'hypoth\`ese {\cal S} nous a permis de d\'efinir les int\'egrales de courbure en int\'egrant simplement sur la partie lisse de la m\'etrique. Dans le cas g\'en\'eral, d\'efinir un produit de courants $T^n$ est un probl\`eme de type {\bf Monge-Amp\`ere}. \medskip Le r\'esultat suivant va dans le sens de la conjecture~: \medskip \noindent {\bf Proposition } {\em La conjecture est vraie dans le cas des surfaces complexes (et dans ce cas, $X$ \'etant de Moishezon est donc projective).} \medskip \noindent {\bf D\'emonstration } Soit $\omega$ une m\'etrique hermitienne sur $X$, et pour $\varepsilon > 0$, soit $T_{\varepsilon}$ un courant donn\'e par le th\'eor\`eme d'approximation des courants v\'erifiant $T_{\varepsilon} \geq - \varepsilon \omega$. Par le th\'eor\`eme C, il suffit de montrer que pour $\varepsilon$ assez petit, on a $$ \displaystyle{\int_{X(\leq 1,T_{\varepsilon})} T_{\varepsilon}^{n} > 0}.$$ Comme il existe un ouvert $U$ sur lequel $T$ est strictement positif (disons sup\'erieur \`a $C_U \omega _{| U}$ o\`u $C_U$ est une constante strictement positive), il existe un ouvert plus petit $U'$ ind\'ependant de $\varepsilon$ sur lequel $T_{\varepsilon}$ est sup\'erieur \`a $(C_{U}/2) \omega _{| U'}$. On en d\'eduit qu'il existe une constante $C>0$, ind\'ependante de $\varepsilon$ telle que pour tout $\varepsilon$ petit, on ait $$\displaystyle{\int_{X(0,T_{\varepsilon})} T_{\varepsilon}^{n} \geq \int_{U'} T_{\varepsilon}^{n} > C}.$$ Il suffit donc de montrer que $$ \lim _{\varepsilon \to 0}\int_{X(1,T_{\varepsilon})} T_{\varepsilon}^{n} =0.$$ Or, sur l'ouvert $X(q,T_{\varepsilon})$, on a $$ 0 \leq (-1)^q T_{\varepsilon}^{n} \leq \frac{n!}{q! (n-q)!} \varepsilon ^q \omega ^q \wedge (T_{\varepsilon} + \varepsilon \omega )^{n-q}.$$ Il suffit donc de contr\^oler les produits de Monge-Amp\`ere et plus pr\'ecis\'ement de montrer que pour $q >0$ (et dans la perspective de la conjecture, $q=1$ suffit), on a~: $$ \lim _{\varepsilon \to 0} \varepsilon ^q \int_X \omega ^q \wedge (T_{\varepsilon} + \varepsilon \omega )^{n-q} =0.$$ C'est \'evidemment vrai pour $q=n$ et, si on suppose de plus que la m\'etrique $\omega$ est une m\'etrique de Gauduchon, alors la formule de Stokes donne $$ \int_X \omega ^{n-1} \wedge (T_{\varepsilon} + \varepsilon \omega ) = \mbox{Cste} + O(\varepsilon),$$ et donc c'est aussi vrai pour $q = n-1$. Dans le cas des surfaces, on a l'estimation souhait\'ee pour $q= n-1 =1$, donc $X$ est de Moishezon par le th\'eor\`eme C.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Comme le montre la preuve ci-dessus, le cas g\'en\'eral pourrait d\'ecouler du contr\^ole des masses de Monge-Amp\`ere des approximations $\varphi _{\varepsilon}$ utilis\'ees par J.-P.\ Demailly dans la d\'emonstration de son th\'eor\`eme d'approximation des courants. \section{Une version alg\'ebrique singuli\`ere des in\'egalit\'es de Morse} Dans ce paragraphe, nous donnons quelques exemples ``alg\'e\-bri\-ques" de faisceaux d'id\'eaux de Nadel. Leur origine en g\'eom\'etrie alg\'ebrique se situe dans la version du th\'eor\`eme de Kawamata-Viehweg pour les diviseurs \`a coefficients rationnels. Comme dans \cite{Dem90} ou \cite{EsV92}, ce faisceau d'id\'eaux sert de terme correctif dans le cas o\`u les diviseurs consid\'er\'es ne sont pas \`a croisements normaux. Par ailleurs, il existe une version alg\'ebrique des in\'egalit\'es de Morse holomorphes de Demailly \cite{Dem94} dans le cas d'un fibr\'e en droites diff\'erence de deux fibr\'es amples. Il est alors naturel de donner une version analogue dans le cadre singulier. \'Evidemment, il s'agit essentiellement d'une reformulation dans un cas particulier de notre theor\`eme~B. \subsection{Th\'eor\`eme de Kawamata-Viehweg} Soient $X$ une vari\'et\'e projective, et $M$ un diviseur rationnel effectif de $X$. On note $M = \sum a_i D_i$ o\`u les $a_i$ sont des rationnels positifs et les $D_i$ sont des diviseurs irr\'eductibles. Nous notons ${\cal I}(M)$ le faisceau d'id\'eaux de Nadel associ\'e \`a la m\'etrique singuli\`ere $\phi = \sum a_i \log|g_i|$ o\`u $g_i$ est un g\'en\'erateur local de $D_i$. On rappelle que si $M$ est \`a croisements normaux, alors ${\cal I}(M)$ n'est rien d'autre que le faisceau inversible ${\cal O}(- \lfloor M \rfloor)$ o\`u $\lfloor M \rfloor := \sum \lfloor a_i \rfloor D_i$. Avec ces notations, rappelons le th\'eor\`eme de Kawamata-Viehweg \cite{Kaw82}, \cite{Vie82}~: \medskip \noindent {\bf Th\'eor\`eme d'annulation de Kawamata-Viehweg } {\em Soient $X$ une vari\'et\'e projective et $L$ un fibr\'e en droites sur $X$. On suppose que $L = M + \sum_{j} \alpha _jE_j$ o\`u~: (i) $M$ est un ${\Bbb Q}$-diviseur effectif gros et nef, (ii) les $\alpha _j$ sont des r\'eels v\'erifiant $0 \leq \alpha _j <1$, (iii) le diviseur $\sum_{j} \alpha _jE_j$ est \`a croisements normaux. \noindent Alors $$ H^q(X,K_X + L)=0 \ \mbox{pour tout}\ q \geq 1 .$$} Dans le cas o\`u $\sum_{j} \alpha _jE_j$ n'est pas \`a croisements normaux dans l'\'enonc\'e du th\'eor\`eme de Kawamata-Viehweg, le faisceau multiplicateur de Nadel sert de terme correctif. Nous allons d\'etailler un exemple utilis\'e par L.\ Ein et R.\ Lazarsfeld dans l'\'etude des diviseurs \`a singularit\'e presque isol\'ee (voir le papier de R.\ Lazarsfeld \cite{Laz93} pour plus de d\'etails). Le contexte est le suivant : soient $X$ une vari\'et\'e projective, $A$ un fibr\'e gros et nef sur $X$ et $D$ un diviseur dans le syst\`eme lin\'eaire $|kA|$. Choisissons $\pi : \tilde{X} \rightarrow X$ une compos\'ee d'un nombre fini d'\'eclate\-ments de centres lisses de sorte que $\pi ^* D$ soit un diviseur \`a croisements normaux et soit $K_{\tilde{X}/X}$ la diff\'erence des fibr\'es canoniques~: $$K_{\tilde{X}/X} = K_{\tilde{X}} - \pi^*K_{X}.$$ Alors, pour $\lambda$ rationnel, $0\leq \lambda <1$, on pose~: $${\cal I}_{\lambda}= \pi_*\left(K_{\tilde{X}/X}- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right).$$ La proposition suivante donne les propri\'et\'es de ${\cal I}_{\lambda}$~: \medskip \noindent {\bf Proposition }{\em On a (i) le faisceau d'id\'eaux $\displaystyle{{\cal I}_{\lambda}}$ est \'egal au faisceau multiplicateur $\displaystyle{{\cal I}(\lambda \frac{D}{k})}$, en particulier ${\cal I}_{\lambda}$ est ind\'ependant de la r\'esolution $\pi$ choisie, (ii) les images directes sup\'erieures $\displaystyle{R^q \pi_*\left(K_{\tilde{X}/X}- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right)}$ sont nulles pour tout $q \geq 1$, (iii) les groupes de cohomologie $\displaystyle{H^q(X,(K_X + A)\otimes{\cal I}_{\lambda})}$ sont nuls pour tout $q \geq 1$. } \medskip \noindent {\bf D\'emonstration} Pour le point (i), nous utilisons \`a nouveau l'identit\'e $$\mu _*(K_{\tilde{X}}\otimes {\cal I}(\mu^*h)) = K_X \otimes {\cal I}(h).$$ Dans notre situation, on a \vspace{-2mm} $$ \pi_*\left(K_{\tilde{X}}\otimes {\cal I}(\pi^*(\lambda \frac{D}{k}))\right) = K_X \otimes {\cal I}(\lambda \frac{D}{k}).$$ \vspace{-3mm} Comme $\displaystyle{\pi^*(\lambda \frac{D}{k})}$ est \`a croisements normaux, on a \vspace{-2mm} $$ {\cal I}(\pi^*(\lambda \frac{D}{k}))= {\cal O}\left(- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right).$$ De l\`a, on d\'eduit successivement \vspace{-3mm} \begin{eqnarray*} {\cal I}_{\lambda} & = & \pi_*\left(K_{\tilde{X}/X}- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right)\\ & = & \pi_*\left(K_{\tilde{X}}-\pi^*K_X - \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right)\\ & = & -K_X \otimes K_X \otimes {\cal I}(\lambda \frac{D}{k}) \\ & = & {\cal I}(\lambda \frac{D}{k}). \end{eqnarray*} \noindent Ceci d\'emontre le point (i). Pour le point (ii), la d\'emonstration repose sur l'observation classique suivante (d\'ej\`a observ\'ee par H.\ Grauert et O.\ Riemenschneider \cite{GrR70})~: les faisceaux $R^q\pi_*{\cal F}$ sont nuls si et seulement si pour tout fibr\'e $L$ sur $X$ suffisamment ample et tout $q \geq 1$, on a $H^q(\tilde{X},{\cal F}\otimes \pi^*L) = 0$. Comme $$ R^q \pi_* \left(\pi^*(K_X + \pi^*A) + K_{\tilde{X}/X}- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right) =$$ $$(K_X + \pi^*A)\otimes R^q \pi_* \left( K_{\tilde{X}/X}- \lfloor \pi^*(\lambda \frac{D}{k}) \rfloor \right),$$ il suffit de montrer que pour tout fibr\'e $L$ sur $X$ suffisamment ample et tout $q \geq 1$, on a $$ H^q\left(\tilde{X},\pi^*L + \pi^*A + K_{\tilde{X}}- \lfloor \pi^*(\lambda \frac{D}{k})\rfloor \right) = 0. $$ Or ce dernier groupe est \'egal \`a $$ H^q\left(\tilde{X}, K_{\tilde{X}} + \pi^*L + \left(1-\lambda\right)\pi^*A + \left(\lambda\pi^*A - \lfloor \lambda\pi^*A\right)\rfloor \right)$$ qui est nul gr\^ace au th\'eor\`eme de Kawamata-Viehweg appliqu\'e au ${\Bbb Q}$-diviseur effectif gros et nef $$ M:= \pi^*L + (1-\lambda)\pi^*A .$$ Pour le point (iii), on a $$ H^q\left(X,\left(K_X + A\right)\otimes{\cal I}_{\lambda}\right)= H^q\left(\tilde{X}, K_{\tilde{X}} + \pi^*A-\lfloor \pi^*(\lambda \frac{D}{k})\rfloor \right).$$ Ce dernier groupe s'\'ecrit encore $\displaystyle{H^q\left(\tilde{X}, K_{\tilde{X}} +\left(1-\lambda\right)\pi^*A + \lambda\pi^*A - \lfloor \lambda\pi^*A \rfloor \right)}$ qui est nul \`a nouveau par le th\'eor\`eme de Kawamata-Viehweg.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{In\'egalit\'es de Morse alg\'ebriques singuli\`eres} En nous inspirant de l'exemple pr\'ec\'edent, nous sommes en mesure de donner une version alg\'ebrique des in\'egalit\'es de Morse holomorphes singuli\`eres. Pour cela, rappelons au pr\'ea\-la\-ble la version suivante des in\'egalit\'es de Morse holomorphes de J.-P.\ Demailly \cite{Dem94}~: \bigskip \noindent {\bf Th\'eor\`eme } {\em Soit $X$ une vari\'et\'e k\"ahl\'erienne de dimension $n$ et soient $F$ et $G$ deux fibr\'es en droites nef sur $X$. Alors, on a~: $$ \sum _{j=0}^{q}(-1)^{q-j} \dim H^{j}(X,k(F-G)) \leq \frac{k^{n}}{n!}\sum _{j=0}^{q}(-1)^{q-j}{n \choose j} F^{n-j}\cdot G^j + o(k^{n}).$$ } \bigskip Ce r\'esultat a \'et\'e obtenu dans un premier temps par J.-P.\ Demailly comme cons\'equence des in\'egalit\'es de Morse, et plus r\'ecemment, F.\ Angelini en a donn\'e une d\'emonstration purement alg\'ebrique \cite{Ang95}. Auparavant, S.\ Trapani \cite{Tra91} et Y.-T.\ Siu \cite{Siu93} avaient d\'emontr\'e le cas particulier $q = 1$ en vue d'obtenir des crit\`eres num\'eriques pour l'existence de sections. Le terme alg\'ebrique a ici une double origine~: les estimations font intervenir des nombres d'intersection \`a la place d'int\'egrales de courbure, et un cas particulier du th\'eor\`eme est celui d'une vari\'et\'e projective et d'un fibr\'e \'ecrit comme diff\'erence de deux fibr\'es amples. Nous montrons le r\'esultat suivant~: \bigskip \noindent {\bf Th\'eor\`eme E } {\em Soit $X$ une vari\'et\'e k\"ahl\'erienne de dimension $n$ et soient $F$ et $G$ deux fibr\'es en droites sur $X$. On suppose que $G$ est nef, et qu'il existe un entier positif $m$, un fibr\'e en droites nef $A$ et un diviseur effectif $D$ de sorte que~: $mF = A + D.$ Alors, on a~: $$ \sum _{j=0}^{q}(-1)^{q-j} \dim H^{j}((X,k(F-G)\otimes {\cal I}_{k}(m^{-1}D)) $$ $$ \leq \frac{k^{n}}{n!} \sum _{j=0}^{q}(-1)^{q-j}{n \choose j} m^{-n+j}A^{n-j}\cdot G^j + o(k^{n}).$$ } \bigskip \noindent {\bf D\'emonstration} Soit $\pi : \tilde{X} \rightarrow X$ une compos\'ee d'un nombre fini d'\'eclate\-ments de centres lisses de sorte que $\pi ^* D$ soit un diviseur \`a croisements normaux. \noindent Comme dans le cadre purement analytique de nos in\'egalit\'es de Morse singuli\`eres, on travaille sur $\tilde{X}$ o\`u l'on applique simplement l'\'enonc\'e pr\'ec\'edent. \noindent D\'etaillons bri\`evement~: le faisceau d'id\'eaux ${\cal I}_{k}(m^{-1}D)$ est \'egal \`a l'image directe $$\pi_*\left(K_{\tilde{X}/X}-\lfloor \pi^*(km^{-1}D)\rfloor \right),$$ si bien qu'il s'agit d'estimer les dimensions $$\dim H^{q}\left(\tilde{X},k\pi^*(F-G)-\lfloor \pi^*(km^{-1}D)\rfloor\right).$$ \noindent Or pour $k$ multiple de $m$, on a $$k\pi^*(F-G)-\lfloor \pi^*(km^{-1}D)\rfloor = k\pi^*(m^{-1}A-G),$$ \noindent et il suffit d'appliquer le th\'eor\`eme pr\'ec\'edent aux fibr\'es nef $\pi^*A$ et $\pi^*G$.\hskip 3pt \vrule height6pt width6pt depth 0pt \chapter{\'Etude de certaines vari\'et\'es de Moishezon dont le groupe de Picard est infini cyclique} Le th\`eme central de ce chapitre est l'\'etude d'une classe particuli\`ere de vari\'et\'es de Moishezon~: celles dont le groupe de Picard est ${\Bbb Z}$ et dont le fibr\'e canonique est gros. En faisant l'hypoth\`ese suppl\'ementaire que la vari\'et\'e $X$ devient projective apr\`es un seul \'eclatement de centre lisse et projectif, nous \'etudions ce centre {\em via} la th\'eorie de Mori sur le mod\`ele projectif. Nous obtenons alors une restriction sur la dimension du centre de l'\'eclatement dans le cas o\`u le fibr\'e canonique n'est pas nef. Apr\`es avoir donn\'e une nouvelle famille de vari\'et\'es de Moishezon ne poss\'edant pas de fibr\'e en droites gros et nef et s'inscrivant dans ce cadre d'\'etude, nous nous restreignons \`a la dimension $4$. Nous obtenons alors une description pr\'ecise du centre de l'\'eclatement et montrons que notre construction est essentiellement unique dans le cas o\`u le fibr\'e canonique n'est pas nef. Enfin, nous obtenons aussi des restrictions partielles en dimension $4$ dans le cas o\`u le fibr\'e canonique est nef. \section{Un th\'eor\`eme de J.\ Koll\'ar} \subsection{\'Enonc\'e du r\'esultat} Nous avons vu dans les pr\'eliminaires de cette th\`ese qu'il existe des vari\'et\'es de Moishezon ne poss\'edant pas de fibr\'e en droites simultan\'ement gros et num\'eriquement effectif. Plus pr\'ecis\'ement, nous avons rencontr\'e le r\'esultat suivant~: \bigskip \noindent {\bf Th\'eor\`eme (J.\ Koll\'ar, K.\ Oguiso) } {\em (i) Il existe des vari\'et\'es de Moishezon $X$ de dimension $3$ dont le groupe de Picard est \'egal \`a ${\Bbb Z}$, avec $-K_X$ gros et ne poss\'edant pas de fibr\'e gros et nef, (ii) il existe des vari\'et\'es de Moishezon $X$ de dimension $3$ dont le groupe de Picard est \'egal \`a ${\Bbb Z}$, dont le fibr\'e canonique est trivial et ne poss\'edant pas de fibr\'e gros et nef. } \bigskip Remarquons que pour une vari\'et\'e de Moishezon dont le groupe de Picard est ${\Bbb Z}$, un et un seul g\'en\'erateur de $\operatorname{Pic}(X)$ est gros. Suivant J.\ Koll\'ar \cite{Kol91}, nous notons ce g\'en\'erateur ${\cal O}_X(1)$ et nous \'ecrivons $\operatorname{Pic}(X) = {\Bbb Z} \cdot {\cal O}_X(1)$. Notons aussi $m_X$ l'entier d\'efini par la relation $K_X = {\cal O}_X(m_X)$. Remarquons ici que les trois cas $m_X < 0$ (respectivement $m_X=0$ et $m_X >0$) correspondent aux trois possibilit\'es $\kappa(X) = -\infty$ (respectivement $\kappa(X) = 0$ et $\kappa(X) = \dim X$), o\`u $\kappa(X)$ d\'esigne la dimension de Kodaira de $X$. \medskip Evidemment, il reste un cas non couvert par l'\'enonc\'e pr\'ec\'edent et la question suivante est naturelle~: \medskip \noindent {\bf Question } {\em Existe-t-il des vari\'et\'es de Moishezon $X$, avec $\operatorname{Pic}(X) = {\Bbb Z} \cdot {\cal O}_X(1)$ et $m_X > 0$ ne poss\'edant pas de fibr\'e gros et nef \ ? } \medskip En dimension $3$, la r\'eponse \`a cette question est n\'egative comme le montre le r\'esultat suivant~: \bigskip \noindent {\bf Th\'eor\`eme (J.\ Koll\'ar, 1991) } {\em Soit $X$ une vari\'et\'e de Moishezon de dimension $3$. On suppose que le groupe de Picard $\operatorname{Pic}(X)$ est ${\Bbb Z}$ et que le fibr\'e canonique $K_X$ est gros. Alors $K_X$ est nef.} \bigskip \noindent {\bf Remarque } Dans le cas o\`u le fibr\'e canonique est gros et nef, l'un de ses multiples est globalement engendr\'e par le ``base-point free theorem". Au vu du r\'esultat pr\'ec\'edent, il n'existe donc pas d'exemple de vari\'et\'e de Moishezon de dimension $3$, de groupe de Picard ${\Bbb Z}$ et \`a fibr\'e canonique gros ne satisfaisant pas aux crit\`eres de J.-P.\ Demailly et Y.-T.\ Siu. \medskip La suite de ce paragraphe consiste \`a rappeler la d\'emonstration de ce r\'esultat car les id\'ees qu'elle contient seront pr\'esentes dans tout le chapitre. \subsection{D\'emonstration} La r\'ef\'erence pr\'ecise est \cite{Kol91}, page 170 et suivantes. \medskip Le lemme suivant, bien qu'\'el\'ementaire est essentiel~: \medskip \noindent {\bf Lemme } {\em Soit $X$ une vari\'et\'e de Moishezon, de dimension quelconque, avec $\operatorname{Pic}(X) = {\Bbb Z} \cdot {\cal O}_X(1)$. Soit $\pi : \tilde{X} \rightarrow X$ une modification projective, de lieu exceptionnel $\tilde{S} \subset \tilde{X} \stackrel{\pi}{\rightarrow} S \subset X$. \noindent Alors pour toute courbe $C$ de $X$, non incluse dans $S$, on a ${\cal O}_X(1) \cdot C > 0$.} \bigskip \setlength{\unitlength}{0.0125in} \begin{picture}(395,148)(0,-10) \put(337,81){\ellipse{116}{104}} \path(215.000,76.000)(223.000,78.000)(215.000,80.000) \path(223,78)(148,78)(148,78) \spline(31,111) (49,60)(79,39) \spline(37,42) (43,75)(73,105) \spline(301,111) (346,78)(364,51) \spline(310,51) (322,93)(346,114) \put(55,75){\ellipse{110}{104}} \put(79,51){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{C}$}}}}} \put(325,3){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $X$}}}}} \put(58,108){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{S}$}}}}} \put(31,0){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{X}$}}}}} \put(361,60){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $C$}}}}} \put(355,102){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $S$}}}}} \put(175,87){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi$}}}}} \end{picture} \bigskip \noindent {\bf Corollaire } {\em Sous les hypoth\`eses du lemme, et si de plus $K_X$ est gros, alors pour toute courbe $C$ (respectivement $\tilde{C}$) de $X$ (respectivement de $\tilde{X}$) non incluse dans $S$ (respectivement $\tilde{S}$), on a $K_X\cdot C > 0$ (respectivement $K_{\tilde{X}} \cdot \tilde{C} > 0$).} \medskip \noindent {\bf D\'emonstration du lemme } Soit $\tilde{H}$ un diviseur ample dans $\tilde{X}$, et $H := \pi _{*}(\tilde{H})$. Alors $H$ est gros, donc s'\'ecrit ${\cal O}_X(p)$, o\`u $p$ est un entier strictement positif. Comme $${\cal O}_X(1) \cdot C = \frac{1}{p}H \cdot C ,$$ il suffit de montrer que pour toute courbe $C$ non incluse dans $S$, on a $H \cdot C > 0$. \noindent Comme $C$ n'est pas incluse dans $S$, si $\tilde{C}$ d\'esigne la transform\'ee stricte de $C$, l'\'egalit\'e suivante est v\'erifi\'ee~: $$ H \cdot C = \pi^{*}(H) \cdot \tilde{C} = (\tilde{H} + \sum a_iE_i) \cdot \tilde{C},$$ o\`u les $a_i$ sont des entiers positifs ou nuls, et les $E_i$ les composantes irr\'eductibles de $\tilde{S}$. De l\`a, $H \cdot C > 0$ car $\tilde{H} \cdot \tilde{C} > 0$ et pour tout $i$, on a $E_i \cdot \tilde{C} \geq 0$. \hskip 3pt \vrule height6pt width6pt depth 0pt \medskip \noindent {\bf Remarque } Ce lemme affirme en particulier que les courbes sur lesquelles ${\cal O}_X(1)$ est n\'egatif ou nul sont incluses dans un ensemble analytique de codimension sup\'erieure ou \'egale \`a $2$. \medskip \noindent {\bf D\'emonstration du th\'eor\`eme } En dimension $3$, on d\'eduit du lemme pr\'ec\'edent qu'il n'y a qu'un nombre fini de courbes sur lesquelles ${\cal O}_X(1)$ est n\'egatif. Si $K_X$ est gros, et si $C$ est une courbe telle que $K_X \cdot C < 0$, alors une telle courbe se d\'eforme dans $X$ car la formule de Riemann-Roch donne $$ \chi (N_{C/X}) = -K_X \cdot C + (n-3)(1-g) = -K_X \cdot C > 0,$$ o\`u $N_{C/X}$ est le fibr\'e normal de $C$ dans $X$. Mentionnons que si la courbe $C$ est singuli\`ere, on d\'efinit $N_{C/X}$ comme \'etant \'egal \`a $\displaystyle{ \nu ^{\ast}T X / T \tilde{C} }$ o\`u $\nu : \tilde{C} \to C$ est la normalis\'ee de la courbe $C$. Ceci donne bien la contradiction.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{Commentaires} La d\'emonstration ci-dessus repose sur un argument de d\'eformation. Comme nous utilisons dans la suite ce type d'argument, il est sans doute bon de faire ici un bref rappel. \'Etant donn\'ees une vari\'et\'e $X$ et une sous-vari\'et\'e $Y$ de $X$, c'est un probl\`eme classique et important de d\'eterminer les d\'eformations de $Y$ dans $X$. Dans le cadre analytique, ce probl\`eme a \'et\'e consid\'er\'e par K.\ Kodaira \cite{Kod62} et par A.\ Grothendieck et D.\ Mumford dans le cadre alg\'ebrique o\`u la notion de {\bf sch\'ema de Hilbert} joue un r\^ole essentiel~: une r\'ef\'erence importante est le travail r\'ecent de J.\ Koll\'ar \cite{Kol94}. La ``solution" au probl\`eme est donn\'ee par le~: \bigskip \noindent {\bf Th\'eor\`eme \cite{Gro62}, \cite{Kod62} }{\em Soit $Y$ une sous-vari\'et\'e d'une vari\'et\'e $X$. Alors le sch\'ema de Hilbert $\operatorname{Hilb} (X)$ des sous-ensembles analytiques de $X$ admet $H^0(Y,N_{Y/X})$ comme espace tangent de Zariski en $[Y]$. La dimension de $\operatorname{Hilb} (X)$ en $[Y]$ satisfait~: $$ \dim H^0(Y,N_{Y/X}) - \dim H^1(Y,N_{Y/X}) \leq \dim_{[Y]} \operatorname{Hilb} (X) \leq \dim H^0(Y,N_{Y/X}).$$ En particulier, si $H^1(Y,N_{Y/X}) =0$, alors $\operatorname{Hilb} (X)$ est lisse au voisinage de $[Y]$.} \bigskip Dans le cadre analytique, la construction de K.\ Kodaira consiste \`a trouver explicitement en coordonn\'ees locales les s\'eries enti\`eres d\'efinissant les sous-ensembles proches de $Y$, tandis que dans le cadre alg\'ebrique, l'id\'ee est qu'une vari\'et\'e projective $Z$ est d\'etermin\'ee par le sous-espace vectoriel des polyn\^omes de degr\'e suffisamment grand qui s'annulent sur $Z$. \section{Quelques rappels sur la th\'eorie de Mori} La th\'eorie de Mori, n\'ee dans les ann\'ees 80, consiste \`a \'etendre et approfondir en dimension sup\'erieure ou \'egale \`a $3$ la classification des surfaces complexes et l'\'etude des applications bim\'eromorphes entre surfaces complexes. Cependant, cette th\'eorie n'est valable que sur les vari\'et\'es projectives. Comme une vari\'et\'e de Moishezon est domin\'ee par une vari\'et\'e projective, une id\'ee naturelle est d'appliquer certains r\'esultats de la th\'eorie de Mori pour obtenir des renseignements concernant la structure des vari\'et\'es de Moishezon ou de leur caract\`ere non projectif. Nous rappelons dans ce paragraphe les r\'esultats essentiels utilis\'es dans ce chapitre. Deux excellentes r\'ef\'erences sont \cite{CKM88} et \cite{KMM87}. Mentionnons aussi qu'une des grandes id\'ees de S.\ Mori est, m\^eme pour l'\'etude des vari\'et\'es non singuli\`eres, de quitter le monde lisse pour autoriser certains types de singularit\'es~; cette analyse fut en particulier mise en oeuvre par M.\ Reid. Il est cependant bon de pr\'eciser ici que la plupart des \'enonc\'es que nous utilisons sont d'une difficult\'e moindre dans le cas lisse, cadre dans lequel nous les appliquons. \subsection{C\^one des courbes effectives} Dans tout ce paragraphe, $X$ est une vari\'et\'e projective. \subsub{Notations} Rappelons tout d'abord que la notation $N_1(X,{\Bbb R})$ d\'esigne l'espace vectoriel des combinaisons lin\'eaires finies (\`a coefficients r\'eels) de courbes (irr\'eductibles et \'eventuellement singuli\`eres) de $X$, modulo l'\'equivalence num\'erique~: deux courbes sont \'equivalentes si et seulement si leurs intersections avec tout diviseur sont \'egales. \noindent Pour une vari\'et\'e projective (et m\^eme de Moishezon), cet espace vectoriel est de dimension finie et est en dualit\'e naturelle ({\em via} la forme d'intersection) avec le groupe de N\'eron-Severi $\displaystyle{(\operatorname{Pic} (X)/ \operatorname{Pic} ^0(X))\otimes _{{\Bbb Z}} {\Bbb R}}$~; la dimension de $N_1(X,{\Bbb R})$ est appel\'ee {\bf nombre de Picard de $X$}. L'espace vectoriel $N_1(X,{\Bbb R})$ est naturellement un sous-espace vectoriel de $H_2(X,{\Bbb R})$. Enfin, nous notons suivant l'usage $\operatorname{NE} (X)$ le sous-c\^one convexe de $N_1(X,{\Bbb R})$ engendr\'e par les classes d'homologie des courbes effectives. L'adh\'erence de ce c\^one est not\'ee $\overline{\operatorname{NE} }(X)$. \subsub{Th\'eor\`eme du c\^one} L'un des premiers succ\`es de la th\'eorie de Mori est de montrer que {\em si le fibr\'e canonique $K_X$ n'est pas nef, alors il existe une courbe rationnelle sur laquelle il est strictement n\'egatif}. L'\'enonc\'e pr\'ecis est le suivant~: \medskip \noindent {\bf Th\'eor\`eme du c\^one } {\em Soit $X$ une vari\'et\'e projective. Alors il existe un ensemble minimal (fini ou d\'enombrable) de courbes rationnelles $C_i$ dans $X$ de sorte que~: \vspace{+1mm} (i) pour tout $i$, on a $\displaystyle{0 < -K_X \cdot C_i \leq \dim X +1}$, \vspace{+2mm} (ii) $\displaystyle{ \overline{\operatorname{NE} }(X) = \overline{\operatorname{NE} }(X)_{K_X \geq 0} + \sum_i {\Bbb R}_+ \ [C_i]}$, \noindent o\`u $\displaystyle{ \overline{\operatorname{NE} }(X)_{K_X \geq 0} := \{ [C] \in N_1(X,{\Bbb R}) \ | \ K_X \cdot C \geq 0 \} }$. } \medskip Les courbes rationnelles $C_i$ sont appel\'es {\bf courbes rationnelles extr\^emales} et les ${\Bbb R}_+ \ [C_i]$ sont appel\'ees {\bf rayons extr\^emaux}. On ne connait pas de version analogue de ce th\'eor\`eme pour les vari\'et\'es de Moishezon. \`A notre connaissance, la question suivante est ouverte~: \medskip \noindent {\bf Question } {\em Soit $X$ une vari\'et\'e de Moishezon, dont le fibr\'e canonique n'est pas nef. Existe-t-il alors une courbe rationnelle $C$ telle que $K_X \cdot C < 0$ \ ? } \subsection{Contraction de Mori} Les rayons extr\^emaux jouent le m\^eme r\^ole dans la th\'eorie de Mori que les courbes rationnelles lisses d'auto-intersection n\'egative dans la th\'eorie des surfaces complexes~: ils peuvent \^etre contract\'es. Le th\'eor\`eme suivant d\'ecrit les diff\'erents types de contraction obtenus en contractant un rayon extr\^emal. \bigskip \noindent {\bf Th\'eor\`eme de contraction \cite{CKM88} } {\em Soit $X$ une vari\'et\'e projective dont le fibr\'e canonique n'est pas nef. Soient $C$ une courbe rationnelle extr\^emale et $R := {\Bbb R}_+ \ [C]$ le rayon extr\^emal engendr\'e par $C$. Alors, il existe une vari\'et\'e (\'eventuellement singuli\`ere) projective, normale, et une application $$f : X \to Y ,$$ not\'ee aussi $\operatorname{cont}_R$, de sorte que~: (i) une courbe de $X$ est contract\'ee par $f$ si et seulement si sa classe d'homologie appartient au rayon $R$, (ii) le fibr\'e $-K_X$ est $f$-ample (i.e la restriction de $-K_X$ \`a toute fibre de $f$ est ample). \noindent De plus, on distingue trois types de contractions~: (a) $\dim X > \dim Y$ et $f$ est une fibration Fano (i.e la fibre g\'en\'erique de $f$ est une vari\'et\'e lisse dont le fibr\'e anti-canonique est ample), (b) $\dim X = \dim Y$ et $f$ est une contraction divisorielle (i.e $f$ est birationnelle et contracte un diviseur), (c) $\dim X = \dim Y$ et $f$ est une petite contraction (i.e $f$ est birationnelle et contracte un sous-ensemble alg\'ebrique de codimension sup\'erieure ou \'egale \`a $2$).} \medskip Les cas (a) et (b) sont les ``bons" cas~: dans le cas (a), on r\'eduit la compr\'ehension de la vari\'et\'e $X$ \`a celle d'une vari\'et\'e de dimension plus petite et \`a la structure des fibres qui sont des vari\'et\'es de Fano. Dans le cas (b), la vari\'et\'e singuli\`ere $Y$ est ${\Bbb Q}$-factorielle, \`a singularit\'es terminales (ce sont les singularit\'es qui permettent de donner encore un sens \`a l'expression ``$K_Y$ est ou n'est pas nef") et le nombre de Picard de $Y$ est strictement plus petit que celui de $X$. Le cas (c) est le ``mauvais" cas~: les singularit\'es de $Y$ sont telles que $Y$ ne poss\`ede pas de fibr\'e canonique et il n'est pas clair que $Y$ soit ``plus simple" que $X$. \subsection{Contractions divisorielles} Nous utilisons dans la suite un r\'esultat plus pr\'ecis que le th\'eor\`eme de contraction dans le cas d'une contraction divisorielle. Sous cette forme, il appara\^{\i}t dans les travaux de T.\ Ando \cite{And85} et M.\ Beltrametti \cite{Bel86}~: \bigskip \noindent {\bf Th\'eor\`eme (T.\ Ando, M.\ Beltrametti, 1985) } {\em Soient $X$ une vari\'et\'e projective et $f : X \to Y$ une contraction divisorielle d'un rayon extr\^emal. Soit enfin $F$ une fibre g\'en\'erale de $f_E : E \to f(E)$ o\`u $E$ est le diviseur exceptionnel de $f$. \noindent Alors il existe un fibr\'e en droites $L$ sur $X$ tel que~: (i) $\operatorname{Im} (\operatorname{Pic} (X) \to \operatorname{Pic} (F)) = {\Bbb Z} \cdot L_{|F}$ o\`u $L_{|F}$ est ample sur $F$, (ii) ${\cal O}_F(-K_X) \simeq {\cal O}_F(pL)$ et ${\cal O}_F(-E) \simeq {\cal O}_F(qL)$ o\`u $p$ et $q$ sont deux entiers positifs. \noindent Enfin, si $F$ est de dimension $2$, alors $F$ est isomorphe \`a ${\Bbb P} ^2$ ou \`a la quadrique ${\cal Q}_2$. } \medskip Nous utiliserons ce r\'esultat lorsque $X$ est de dimension $4$. \subsection{L'in\'egalit\'e de Wi\'sniewski} Avant d'\'enoncer cette in\'egalit\'e, nous avons besoin de deux notations~: soient $X$ une vari\'et\'e projective, $R$ un rayon extr\^emal de $\overline{\operatorname{NE} } (X)$ et $f$ la contraction de Mori associ\'ee. On note $\displaystyle{ l(R) = \min \{ -K_X \cdot C \ | \ C \mbox{ est une courbe rationnelle telle que } [C] \in R \} }.$ Le nombre $l(R)$ est la {\bf longueur} du rayon $R$. On note aussi $A(R)$ le lieu de $X$ couvert par les courbes dont la classe appartient \`a $R$. Dans \cite{Wis91}, J.\ Wi\'sniewski d\'emontre l'in\'egalit\'e fondamentale suivante~: \bigskip \noindent {\bf Th\'eor\`eme (J.\ Wi\'sniewski, 1991) } {\em Pour toute fibre non triviale $F$ de $f$, on a~: $$ \dim F + \dim A(R) \geq \dim X + l(R) - 1.$$} Cette in\'egalit\'e, dont une forme faible est due \`a P.\ Ionescu \cite{Ion86}, a de nombreuses cons\'equences dans la classification des vari\'et\'es projectives de dimension sup\'erieure ou \'egale \`a $3$. \section{Un premier r\'esultat} Rappelons pour commencer ce paragraphe que les exemples de J.\ Koll\'ar et K.\ O\-gui\-so d\'ej\`a cit\'es v\'erifient la propri\'et\'e suppl\'ementaire suivante~: les vari\'et\'es $X$ cons\-trui\-tes ne sont, bien s\^ur, pas projectives, mais le deviennent apr\`es exactement un \'eclatement le long d'une sous-vari\'et\'e $Y \subset X$. Dans toute cette partie, nous nous pla\c cons dans la situation analogue suivante~: \medskip \noindent {\bf Hypoth\`ese } {\em $X$ est une vari\'et\'e de Moishezon non projective de dimension $n$ dont le groupe de Picard est \'egal \`a ${\Bbb Z}$, dont le fibr\'e canonique est gros et telle qu'il existe une sous-vari\'et\'e $Y \subset X$ de sorte que l'\'eclatement $\pi : \tilde{X} \to X$ de $X$ le long de $Y$ d\'efinisse une vari\'et\'e projective $\tilde{X}$. On note $E$ le diviseur exceptionnel de l'\'eclatement. } \medskip La remarque suivante est tr\`es importante~: \medskip \noindent {\bf Remarque } D'apr\`es le corollaire 3.1.2, on sait alors que $K_X$ (respectivement $K_{\tilde{X}}$) est strictement positif sur les courbes non incluses dans $Y$ (respectivement dans $E$). La cons\'equence suivante sera utilis\'ee dans la suite~: si $C$ est une courbe de $Y$ sur laquelle $K_X$ est strictement n\'egatif, cette courbe ne peut pas se d\'eformer (dans $X$) hors de $Y$. \medskip La m\'ethode que nous adoptons pour \'etudier $X$ et $Y$ est d'appliquer la th\'eorie de Mori \`a la vari\'et\'e projective $\tilde{X}$. \subsection{C\^one des courbes sur $\tilde{X}$} Dans le cadre de notre \'etude, comme $\operatorname{Pic} (X) = {\Bbb Z}$, l'espace vectoriel $N_1(\tilde{X},{\Bbb R})$ est isomorphe \`a ${\Bbb R} ^2$. Dans la suite, nous repr\'esentons dans $N_1(\tilde{X},{\Bbb R}) \simeq {\Bbb R} ^2$ le c\^one ferm\'e $\overline{\operatorname{NE} }(\tilde{X})$~; dans les figures ci-dessous, ce dernier correspond \`a la partie hachur\'ee. Si $D$ est un \'el\'ement de $\operatorname{Pic} (\tilde{X})$, nous notons $D > 0$ (respectivement $D = 0$, respectivement $D < 0$) les ensembles $$\{ [C] \in \overline{\operatorname{NE} }(\tilde{X}) \ | \ D \cdot C > 0 \}$$ (respectivement $D \cdot C = 0$, respectivement $D \cdot C < 0$). Deux cas se pr\'esentent suivant que $K_X$ est nef ou non. Ces deux cas se distinguent naturellement~; ils correspondent, comme nous le verrons plus loin, au fait que $X$ admet ou non un morphisme vers une vari\'et\'e (\'eventuellement singuli\`ere) projective de m\^eme dimension, \medskip (i) soit $K_X$ est nef et le dessin est le suivant~: \setlength{\unitlength}{0.0125in} \begin{picture}(404,250)(0,-10) \path(0,144)(288,69) \path(165,153)(120,96) \path(225,183)(150,72) \path(45,162)(234,0) \path(171,54)(270,207) \path(0,72)(297,219) \put(309,219){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}=0$}}}}} \put(258,117){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}<0$}}}}} \put(300,66){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}=0$}}}}} \put(195,222){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}<0$}}}}} \put(255,159){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}>0$}}}}} \put(240,30){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}>0$}}}}} \end{picture} (ii) soit $K_X$ n'est pas nef et le dessin est le suivant~: \setlength{\unitlength}{0.0125in} \begin{picture}(404,333)(0,-10) \path(0,180)(288,105) \path(45,72)(192,318)(189,315) \path(141,243)(150,237) \path(120,195)(150,105) \path(141,237)(192,69) \path(174,285)(240,27) \path(144,246)(147,234) \path(0,108)(297,255) \path(36,204)(273,0) \put(108,249){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $[\tilde{C}]$}}}}} \put(195,273){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}<0$}}}}} \put(309,255){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}=0$}}}}} \put(231,195){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi^{*}K_{X}>0$}}}}} \put(300,102){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}=0$}}}}} \put(243,66){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}>0$}}}}} \put(246,150){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $K_{\tilde{X}}<0$}}}}} \end{picture} \noindent {\bf Quelques commentaires sur ces diagrammes} - dans les deux cas, le fait que la droite $\{ K_{\tilde{X}} = 0 \}$ coupe le c\^one effectif vient du fait qu'il y a \`a la fois des courbes sur lesquelles $ K_{\tilde{X}}$ est strictement positif (celles non contenues dans le diviseur exceptionnel) et des courbes sur lesquelles $ K_{\tilde{X}}$ est strictement n\'egatif (toute courbe incluse dans les fibres de l'\'eclatement), - dans le deuxi\`eme cas, la position relative de $\{ \pi ^{\ast} K_{X} = 0 \}$ est justifi\'ee par le fait qu'il y a des courbes sur lesquelles $\pi ^{\ast} K_{X}$ et $K_{\tilde{X}}$ sont strictement positifs (celles non contenues dans le diviseur exceptionnel d'apr\`es 3.1.2) et que $\pi ^{\ast} K_{X}$ est nul sur toute courbe incluse dans les fibres de l'\'eclatement. \subsub{Quelques cons\'equences de ces diagrammes} Nous regroupons ici les renseignements provenant directement de la description de $\overline{\operatorname{NE} }(\tilde{X})$. Pour cela, appliquons le th\'eor\`eme du c\^one \`a la vari\'et\'e projective $\tilde{X}$. On en d\'eduit que le rayon extr\^emal du c\^ot\'e $K_{\tilde{X}} < 0$ est engendr\'e par la classe d'une courbe rationnelle $\tilde{C}$ dans $\tilde{X}$. Alors, le th\'eor\`eme de contraction assure l'existence d'une vari\'et\'e (en g\'en\'eral singuli\`ere) projective $Z$ et d'un morphisme $f$ associ\'es \`a la courbe extr\^emale rationnelle $\tilde{C}$ de sorte que la situation suivante ait lieu~: \begin{center} \setlength{\unitlength}{0.0125in} \begin{picture}(216,110)(0,-10) \path(147,69)(147,18) \path(145.000,26.000)(147.000,18.000)(149.000,26.000) \path(99,81)(24,81) \path(32.000,83.000)(24.000,81.000)(32.000,79.000) \put(60,84){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $f$}}}}} \path(117,72)(117,15) \path(115.000,23.000)(117.000,15.000)(119.000,23.000) \put(108,75){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{X} \supset E$}}}}} \put(0,75){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $Z$}}}}} \put(126,39){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi$}}}}} \put(111,0){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $X \supset Y$}}}}} \end{picture} \end{center} \medskip \noindent (i) Cas o\`u $K_X$ est nef. Alors le rayon extr\^emal est engendr\'e par la classe d'une courbe rationnelle incluse dans une fibre non triviale de $\pi$. Toutes les fibres de $\pi$ sont donc contract\'ees par $f$ si bien que $f$ se factorise en une application $g : X \to Z$ $$ \tilde{X} \stackrel{\pi}{\to} X \stackrel{g}{\to} Z \ \mbox{et} \ \ f = g \circ \pi.$$ \medskip \noindent (ii) Cas o\`u $K_X$ n'est pas nef. Alors les fibres de $f$ et les fibres de $\pi$ ne se coupent que sur un nombre fini de points~: en effet, il n'existe pas de courbes simultan\'ement contract\'ees par $\pi$ et $f$ car les rayons engendr\'es par $[\tilde{C}]$ et la classe d'une courbe rationnelle incluse dans une fibre non triviale de $\pi$ sont distincts. \subsub{Une application imm\'ediate} Dans le cas o\`u $K_X$ n'est pas nef, la courbe rationnelle $\tilde{C}$ n'\'etant pas contract\'ee par $\pi$, la courbe rationnelle $C = \pi (\tilde{C})$ v\'erifie $K_X \cdot C < 0$. Le r\'esultat suivant en d\'ecoule~: \medskip \noindent {\bf Proposition }{\em Sous les hypoth\`eses pr\'ec\'edentes, si $K_X$ n'est pas nef, il existe une courbe rationnelle $C \subset Y $ sur laquelle $K_X$ est strictement n\'egatif. } \subsection{Contraction de Mori de $\tilde{X}$} Nous \'etudions ici plus en d\'etail la contraction de Mori $f$ associ\'ee \`a la courbe extr\^emale rationnelle $\tilde{C} \subset \tilde{X}$ obtenue pr\'ec\'edemment pour en d\'eduire une estimation de la dimension de $Y$ en toute dimension. \subsub{\'Enonc\'e du r\'esultat} Nous avons vu qu'il y a trois types de contractions extr\^emales. Le th\'eor\`eme suivant restreint les possibilit\'es dans notre situation~: \bigskip \noindent {\bf Th\'eor\`eme F } {\em Soit $X$ de Moishezon avec $\textstyle{\operatorname{Pic} (X)} = {\Bbb Z}$ et $K_X$ gros. Si $X$ est rendue projective apr\`es \'eclatement $\pi : \tilde{X} \to X$ le long de $Y$, alors~: (i) on a $\dim \tilde{X} = \dim Z$, autrement dit $f$ est une application birationnelle, (ii) si $f$ est une contraction divisorielle, son diviseur exceptionnel est \'egal \`a celui de $\pi$ (not\'e $E$ pr\'ec\'edemment)~; ce cas est le seul possible lorsque $K_X$ est nef, (iii) si $f$ est une contraction divisorielle et si $K_X$ n'est pas nef, les in\'egalit\'es suivantes sont satisfaites~: $$ \operatorname{codim} Y -1 \leq \dim f(E) < \dim Y \ \ \mbox{\rm et} \ \ \dim Y > \frac{n-1}{2}, $$ (iv) si $f$ est une petite contraction et si $K_X$ n'est pas nef, l'in\'egalit\'e suivante est satisfaite~: $$ \dim Y \geq \frac{n+1}{2}.$$ } \smallskip On peut reformuler ce r\'esultat sans faire intervenir la contraction de Mori~: \medskip \noindent {\bf Th\'eor\`eme F' } {\em Soit $X$ une vari\'et\'e de Moishezon de dimension $n$ avec $\textstyle{\operatorname{Pic} (X)} = {\Bbb Z}$ et $K_X$ gros. Supposons que $X$ est rendue projective apr\`es \'eclatement le long d'une sous-vari\'et\'e lisse $Y$. \noindent Alors, si $K_X$ n'est pas nef, on a $\displaystyle{ \dim Y > \frac{n-1}{2}}$.} \bigskip \noindent {\bf Remarque } Le fait que $K_X$ soit gros est ici essentiel. En effet, les constructions de J.\ Koll\'ar et K.\ Oguiso montrent que les in\'egalit\'es du point (iii) ne sont pas vraies en g\'en\'eral. La construction de J.\ Koll\'ar donne aussi un exemple o\`u les diviseurs exceptionnels de $\pi$ et $f$ ne sont pas \'egaux. \subsub{D\'emonstration du th\'eor\`eme F-F'} Les points (i) et (ii) du th\'eor\`eme~F sont faciles~: par hypoth\`ese, $f$ ne contracte que des courbes sur lesquelles $K_{\tilde{X}}$ est strictement n\'egatif, donc incluses dans $E$ d'apr\`es le corollaire 3.2.1~; en particulier le point (i) est d\'emontr\'e. \noindent Ceci montre aussi que si $f$ est une contraction divisorielle, son diviseur exceptionnel \'etant inclus dans $E$ est donc \'egal \`a $E$. R\'eciproquement, si $K_X$ est nef, $f$ se factorise \`a travers $\pi$ et donc est une contraction divisorielle. Le point (ii) est d\'emontr\'e. \bigskip Montrons le point (iii) du th\'eor\`eme F~: $f$ est une contraction divisorielle et $K_X$ n'est pas nef. La situation est r\'esum\'ee par le diagramme suivant~: \begin{center} \setlength{\unitlength}{0.0125in} \begin{picture}(207,130)(0,-10) \path(138,60)(138,15) \path(136.000,23.000)(138.000,15.000)(140.000,23.000) \path(168,90)(168,15) \path(166.000,23.000)(168.000,15.000)(170.000,23.000) \path(44.000,77.000)(36.000,75.000)(44.000,73.000) \path(36,75)(126,75) \path(147,87) (151.921,89.353) (155.375,90.591) (159.000,90.000) \path(159,90) (158.773,86.022) (156.650,82.657) (153.000,78.000) \put(147,45){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\pi$}}}}} \path(41.000,107.000)(33.000,105.000)(41.000,103.000) \path(33,105)(153,105) \put(132,0){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $X \supset Y $}}}}} \put(135,69){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{X}$}}}}} \put(0,102){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $f(E)$}}}}} \put(12,87){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\bigcap$}}}}} \put(165,99){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $E$}}}}} \put(9,69){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $Z$}}}}} \end{picture} \end{center} \noindent Dans cette situation, on \'ecrit $$ K_{\tilde{X}}=f^{\ast}K_Z + aE = \pi^{\ast}K_X + (r-1)E $$ o\`u $r= \operatorname{codim} Y$ et o\`u $a$ est un nombre rationnel. Cette \'egalit\'e est une \'egalit\'e de ${\Bbb Q}$-diviseurs de Cartier~: dans le cas d'une contraction divisorielle, le diviseur canonique de $Z$ n'est pas de Cartier en g\'en\'eral mais l'un de ses multiples entiers l'est. Rappelons que toutes les notions de positivit\'e (telle que par exemple \^etre gros, nef ou ample) s'\'etendent naturellement aux ${\Bbb Q}$-diviseurs de Cartier. \noindent Les in\'egalit\'es cherch\'ees d\'ecoulent imm\'ediatement du lemme suivant~: \medskip \noindent {\bf Lemme } {\em Si $K_X$ n'est pas nef, les nombres $a$ et $r$ v\'erifient les deux in\'egalit\'es suivantes~: (i) $a > r-1$, (ii) $ \operatorname{codim} f(E) + r \leq n + 1$. \noindent L'in\'egalit\'e suivante est vraie en toute g\'en\'eralit\'e pour une contraction divisorielle~: (iii) $a \leq \operatorname{codim} f(E) - 1$. } \bigskip \noindent {\bf D\'emonstration du lemme} \medskip \noindent {\bf In\'egalit\'e (i)~:} Comme $Z$ est projective avec $\operatorname{Pic} (Z) = {\Bbb Z}$ et $K_Z$ gros, on en d\'eduit que $K_Z$ est ample et donc que $f^{\ast}K_Z$ est nef, et strictement positif sur les courbes de $\tilde{X}$ non contract\'ees par $f$. Choisissons alors une courbe rationnelle $R$ incluse dans une fibre non triviale de $\pi$ (ces derni\`eres sont des ${\Bbb P} ^{r-1}$, on prend pour $R$ une droite ${\Bbb P} ^{1}$). \noindent L'\'egalit\'e $$f^{\ast}K_Z \cdot R + a E \cdot R = \pi^{\ast}K_X \cdot R + (r-1)E \cdot R$$ donne alors~: $$a-(r-1) = f^{\ast}K_Z \cdot R > 0$$ car, $K_X$ n'etant pas nef, $R$ n'est pas contract\'ee par $f$. \hskip 3pt \vrule height6pt width6pt depth 0pt \medskip \noindent {\bf In\'egalit\'e (ii)~:} Cette in\'egalit\'e d\'ecoule de suite du fait que les fibres de $f$ et $\pi$ dans $E$ ne peuvent se couper qu'en un nombre fini de points. De l\`a ~: $$ (n-1 - \dim f(E) ) + (r-1) \leq n-1. \hskip 3pt \vrule height6pt width6pt depth 0pt$$ \medskip \noindent {\bf In\'egalit\'e (iii)~:} Soit $F$ une fibre g\'en\'erique de la restriction de $f$ \`a $E$, et soit $\tilde{C}$ une courbe dans $F$. \noindent Alors, on a $$aE \cdot \tilde{C} = K_{\tilde{X}} \cdot \tilde{C}$$ et par la formule d'adjonction~: $$ K_{\tilde{X} | E} = K_E - E_{| E}.$$ \noindent De l\`a, on en d\'eduit~: $$a+1 = \frac{K_E \cdot \tilde{C}}{E \cdot \tilde{C}}.$$ \noindent Comme le fibr\'e canonique $K_F$ est simplement la restriction de $K_{E}$ \`a $F$, on obtient~: $$a = \frac{K_F \cdot \tilde{C}}{E \cdot \tilde{C}} - 1.$$ \noindent Or, la vari\'et\'e $F$ est Fano, et par le th\'eor\`eme du c\^one appliqu\'e \`a $F$, on peut supposer que $\tilde{C}$ est une courbe (rationnelle) satisfaisant~: $$ 0 < -K_F \cdot \tilde{C} \leq \dim F + 1 = n-1 -\dim f(E) +1 = \operatorname{codim} f(E) .$$ \noindent De l\`a, comme $E \cdot \tilde{C}$ est un entier strictement n\'egatif, il vient $a \leq \operatorname{codim} f(E) - 1$. Ceci termine la preuve du lemme.\hskip 3pt \vrule height6pt width6pt depth 0pt \medskip \noindent {\bf Remarque } L'in\'egalit\'e (iii) peut aussi se d\'eduire de l'in\'egalit\'e de Wi\'sniewski~: en effet $f$ \'etant divisorielle, on a $\dim A(R) = n-1$ et la dimension de la fibre g\'en\'erique non triviale est $n-1-\dim f(E)$. De l\`a, l'in\'egalit\'e de Wi\'sniewski donne $$\operatorname{codim} f(E) -1 \geq l(R).$$ Or, comme $K_{\tilde{X}} = f^*K_Z + a E$, on a $-K_{\tilde{X}}\cdot C \geq a$ pour toute courbe contract\'ee, d'o\`u $l(R) \geq a$ comme souhait\'e. \bigskip Montrons maintenant le point (iv) du th\'eor\`eme~F. Pour cela, on applique l'in\'egalit\'e de Wi\'sniewski~: comme $f$ est une petite contraction, on a \'evidemment $$\dim A(R) \leq n-2,$$ et si $F$ est une fibre non triviale de $f$ cette derni\`ere est incluse dans $E$ et ne coupe les fibres de $\pi$ que sur un ensemble fini. On en d\'eduit que $\dim F \leq \dim Y$ d'o\`u~: $$ n-2 + \dim Y \geq n + l(R) -1,$$ soit $$ \dim Y \geq l(R) +1.$$ Il suffit alors d'estimer $l(R)$. Or, on a $K_{\tilde{X}} = \pi^*K_X + (r-1) E$ et comme $K_X$ n'est pas nef, $ \pi^*K_X$ est strictement n\'egatif sur les courbes contract\'ees par $f$. On en d\'eduit que $$ -K_{\tilde{X}} \cdot C \geq r$$ pour toute courbe contract\'ee par $f$. De l\`a, $ l(R) \geq r$ et en reportant $$ 2 \dim Y \geq n+1$$ qui est l'in\'egalit\'e souhait\'ee.\hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{Application \`a la dimension $3$} On d\'eduit du th\'eor\`eme F un r\'esultat pr\'ecisant celui de J.\ Koll\'ar dans notre situation~: \medskip \noindent {\bf Corollaire } {\em Soit $X$ une vari\'et\'e de Moishezon non projective de dimension $3$, avec $\operatorname{Pic}(X) = {\Bbb Z}$ et $K_X$ gros. \noindent Si $X$ peut \^etre rendue projective apr\`es un \'eclatement seulement, alors $X$ est une petite modification d'une vari\'et\'e singuli\`ere projective ayant une unique singularit\'e nodale ordinaire (dont le mod\`ele local est $xy-zt = 0$ dans $({\Bbb C} ^4,0)$). En particulier, le fibr\'e canonique $K_X$ est nef.} \medskip \noindent {\bf D\'emonstration du corollaire} On note toujours $\pi$ l'\'eclatement rendant $X$ projective et $f$ la contraction de Mori d\'efinie sur la vari\'et\'e projective $\tilde{X}$. D'apr\`es le th\'eor\`eme F, $f$ est birationnelle, et comme il n'y a pas de petites contractions en dimension $3$ d'une vari\'et\'e non singuli\`ere, c'est que $f$ est une contraction divisorielle. De plus, les in\'egalit\'es (iii) du th\'eor\`eme F ne peuvent \^etre v\'erifi\'ees ici car elles impliquent $\operatorname{codim} Y = 1$. C'est donc que $K_X$ est nef (on retrouve ainsi le r\'esultat de J.\ Koll\'ar), et que le rayon extr\^emal du c\^ot\'e $K_{\tilde{X}} < 0$ est engendr\'e par la classe d'homologie des fibres de $\pi$. Il y a alors exactement deux possibilit\'es~: - les fibres de la contraction de Mori (restreinte au diviseur exceptionnel $E$) sont de dimension $1$ et alors cette derni\`ere co\"{\i}ncide avec $\pi$. Dans ce cas, $X$ est projective, ce que l'on a exclu, - la contraction de Mori contracte le diviseur exceptionnel $E$ sur un point. Dans ce cas, nous appliquons le r\'esultat de S.\ Mori \cite{Mor82} qui donne la liste de toutes les contractions extr\^emales d'une vari\'et\'e non singuli\`ere de dimension $3$. On en d\'eduit que le diviseur exceptionnel de $\pi$ (\'egal \`a celui de $f$) est isomorphe \`a ${\Bbb P} ^1 \times {\Bbb P} ^1$ et que $\displaystyle{{\cal O}_{E} (E) = N_{E/\tilde{X}}}$ est de type $(-1,-1)$. La situation est alors la suivante~: $$ \tilde{X} \stackrel{\pi}{\to} X \stackrel{g}{\to} Z \ \mbox{et} \ \ f = g \circ \pi,$$ o\`u $Z$ est une vari\'et\'e singuli\`ere projective ayant une unique singularit\'e nodale ordinaire (dont le mod\`ele local est $xy-zt = 0$ dans $({\Bbb C} ^4,0)$). Dans ce cas, la contraction de Mori est alors $g \circ \pi$ et correspond \`a l'\'eclatement du point singulier~: le centre $Y$ de l'\'eclatement $\pi$ est une courbe rationnelle lisse.\hskip 3pt \vrule height6pt width6pt depth 0pt \bigskip \noindent {\bf Exemple } La situation pr\'ec\'edente peut effectivement \^etre r\'ealis\'ee~: soit $Z$ une hypersurface de ${\Bbb P} ^4$ d'\'equation $$h_0x_0^2 + h_1x_1^2 + h_2x_2^2 + h_3x_3^2 = 0,$$ o\`u $[x_0 : \cdots : x_4]$ sont les coordonn\'ees homog\`enes dans ${\Bbb P}^4$ et o\`u les $h_i$ sont quatre polyn\^omes homog\`enes de degr\'e $d$ sup\'erieur ou \'egal \`a $4$ ne s'annulant pas en $[0:0:0:0:1]$ et g\'en\'eriques parmi les polyn\^omes ayant ces propri\'et\'es. Alors l'hypersurface $Z$ est lisse except\'e au point $[0:0:0:0:1]$ o\`u elle poss\`ede une singularit\'e nodale ordinaire. On obtient $X$ comme d\'ecrit pr\'ec\'edemment en r\'esolvant la singularit\'e puis en contractant dans une direction de la quadrique exceptionnelle. \medskip Nous donnons dans la suite d'autres applications du th\'eor\`eme F mais nous commen\c cons par montrer dans le paragraphe suivant que le r\'esultat de J.\ Koll\'ar ne s'\'etend pas en dimension sup\'erieure ou \'egale \`a $4$. \section{Une famille de vari\'et\'es de Moishezon} Le but de cette partie est de montrer le r\'esultat suivant~: \medskip \noindent {\bf Th\'eor\`eme G} {\em Pour tout entier $n$ sup\'erieur ou \'egal \`a $4$, il existe des vari\'et\'es de Moishezon $X$ de dimension $n$ v\'erifiant~: (i) $\operatorname{Pic}(X) = {\Bbb Z}$, (ii) $K_X$ est gros, (iii) $K_X$ n'est pas nef. } \medskip Ainsi, le r\'esultat de J.\ Koll\'ar est propre \`a la dimension $3$. \medskip \noindent {\bf Remarque } Les vari\'et\'es obtenues dans la construction qui suit rel\`event toutes du cas ``contraction divisorielle" \'evoqu\'e dans le paragraphe pr\'ec\'edent. Il serait bien s\^ur int\'eressant de cons\-truire de telles vari\'et\'es relevant du cas ``petite contraction". Cependant, nous verrons plus loin que ce cas ne peut pas se produire en dimension~$4$. \subsection{Un r\'esultat interm\'ediaire} La d\'emonstration du th\'eor\`eme G repose sur la proposition suivante, que nous prouvons plus loin. Mentionnons qu'il nous a \'et\'e signal\'e par un rapporteur anonyme que cette proposition se trouve dans \cite{BVV78}. Pour $n$ entier, nous notons $[x_0: \cdots :x_{n+1}]$ les coordonn\'ees homog\`enes dans ${\Bbb P} ^{n+1}$. On d\'esigne par ${\Bbb P} _{x_n} ^{1}$ la droite $\{ x_0 = \dots = x_{n-1} = 0 \}$. Choisissons alors $n$ polyn\^omes homog\`enes $h_0,\ldots \!,h_{n-1}$ de degr\'e $2n-2$ et consid\'erons l'hypersurface $Z$ de degr\'e $2n-1$ dans ${\Bbb P} ^{n+1}$ et d'\'equation $$\sigma = x_0h_0 +\cdots+ x_{n-1}h_{n-1} = 0.$$ Cette hypersurface contient ${\Bbb P} _{x_n} ^{1}$ et peut \^etre singuli\`ere. On a cependant le r\'esultat suivant~: \medskip \noindent {\bf Proposition }{\em Si $n$ est sup\'erieur ou \'egal \`a $3$ et si les $h_{i}$ sont choisis g\'en\'eriquement, alors~: \smallskip \!(i) l'hypersurface $Z$ est non singuli\`ere, \smallskip \!(ii) le fibr\'e normal $N_{{\Bbb P} ^{1}/ Z}$ est \'egal \`a ${\cal O}_{{\Bbb P} ^{1}}(-1)^{\oplus n-1}$, \smallskip \!(iii) $K_Z$ est \'egal \`a ${\cal O}_{{\Bbb P} ^{n+1}}(n-3)_{| Z}$, \smallskip \!(iv) $\operatorname{Pic} (Z) = {\Bbb Z}$. } \subsection{D\'emonstration du th\'eor\`eme G } La construction qui suit nous a \'evidemment \'et\'e inspir\'ee par l'analyse du paragraphe pr\'ec\'edent, dans le cas o\`u la contraction de Mori est une contraction divisorielle~: si une vari\'et\'e de dimension $4$ de Moishezon satisfait le point (iii) du th\'eor\`eme~F, c'est en \'eclatant une surface, puis en contractant sur une courbe rationnelle que l'on obtient un mod\`ele projectif. Nous donnons cependant la construction g\'en\'erale en toute dimension. \medskip \noindent {\bf Construction explicite : } On se fixe dor\'enavant une hypersurface $Z$ donn\'ee par la proposition pr\'ec\'edente. La vari\'et\'e $X$ cherch\'ee va \^etre obtenue en effectuant un ``flip" (plus exactement l'inverse d'un flip) \`a partir de $Z$. \medskip \begin{center} \setlength{\unitlength}{0.0125in} \begin{picture}(322,130)(0,-10) \path(129,60)(129,15) \path(127.000,23.000)(129.000,15.000)(131.000,23.000) \path(159,90)(159,15) \path(157.000,23.000)(159.000,15.000)(161.000,23.000) \path(35.000,77.000)(27.000,75.000)(35.000,73.000) \path(27,75)(117,75) \path(138,87) (142.921,89.353) (146.375,90.591) (150.000,90.000) \path(150,90) (149.773,86.022) (147.650,82.657) (144.000,78.000) \path(32.000,107.000)(24.000,105.000)(32.000,103.000) \path(24,105)(144,105) \put(3,87){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\bigcap$}}}}} \put(123,0){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $X \supset {\Bbb P}^{n-2} \supset {\Bbb P} ^1$}}}}} \put(0,69){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $Z$}}}}} \put(126,69){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $\tilde{X}$}}}}} \put(0,102){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm ${\Bbb P} ^{1}$}}}}} \put(156,99){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{\shortstack[l]{{\twlrm $E={\Bbb P}^{1} \times {\Bbb P} ^{n-2} $}}}}} \end{picture} \end{center} Suivant la figure ci-dessus, notons $\tilde{X}$ la vari\'et\'e projective obtenue en \'eclatant $Z$ le long de ${\Bbb P} ^{1}$. Le diviseur exceptionnel de l'\'eclatement est alors $E = {\Bbb P} ^{1} \times {\Bbb P} ^{n-2}$, et pour pouvoir contracter dans l'autre direction, il s'agit de montrer, d'apr\`es le crit\`ere de contraction de Fujiki-Nakano, que $\displaystyle{ {\cal O}(E)_{| {\Bbb P} ^1} = {\cal O}_{{\Bbb P} ^1}(-1) }$. Pour cela, les deux suites exactes suivantes~: $$ 0 \to N_{{\Bbb P}^1 /E} = {\cal O}_{{\Bbb P} ^1}^{\oplus n-2} \to N_{{\Bbb P}^1/\tilde{X}} \to N_{E/ \tilde{X} | {\Bbb P} ^1} = {\cal O}(E)_{| {\Bbb P} ^1} \to 0 ,$$ $$ 0 \to T{\Bbb P} ^1 \to T\tilde{X}_{| {\Bbb P} ^1} \to N_{{\Bbb P}^1/\tilde{X}} \to 0 $$ \noindent donnent successivement~: $$\deg (N_{{\Bbb P}^1/\tilde{X}})= \deg ({\cal O}(E)_{| {\Bbb P} ^1}) \ ,\ \deg (K_{\tilde{X} | {\Bbb P} ^1}) = -2 -\deg (N_{{\Bbb P}^1/\tilde{X}}).$$ \noindent Comme $K_{\tilde{X}} = f^{\ast}K_Z + (n-2){\cal O}(E)$ et $K_Z = {\cal O}_{{\Bbb P} ^{n+1}}(n-3)_{| Z}$, on en d\'eduit bien que ${\cal O}(E)_{| {\Bbb P} ^1} = {\cal O}_{{\Bbb P} ^1}(-1)$. \noindent La contraction de ${\Bbb P} ^{1}$ d\'efinit donc une vari\'et\'e de Moishezon, contenant un ${\Bbb P} ^{n-2}$ et telle que $\operatorname{Pic} (X) = {\Bbb Z}$. Montrons maintenant que $N_{{\Bbb P} ^{n-2} / \tilde{X}} = {\cal O}_{{\Bbb P} ^{n-2}}(-1) \oplus {\cal O}_{{\Bbb P} ^{n-2}}(-1)$. Comme $$E = {\Bbb P} ^{1} \times {\Bbb P} ^{n-2} = {\Bbb P}(N^{\ast}_{{\Bbb P} ^{n-2} / \tilde{X}}),$$ le fibr\'e normal $N_{{\Bbb P} ^{n-2} / \tilde{X}}$ est de la forme ${\cal O}_{{\Bbb P} ^{n-2}}(a) \oplus {\cal O}_{{\Bbb P} ^{n-2}}(a)$. Comme pr\'ec\'edemment, la suite exacte~: $$ 0 \to T{\Bbb P} ^{n-2} \to TX_{| {\Bbb P} ^{n-2}} \to N_{{\Bbb P}^{n-2}/X} \to 0 $$ \noindent donne $2a = -\deg (K_{X | {\Bbb P} ^{n-2}}) -n+1$, puis $$\deg (K_{X | {\Bbb P} ^{n-2}}) = \deg (K_{\tilde{X} | {\Bbb P} ^{n-2}} - {\cal O}(E)_{| {\Bbb P} ^{n-2}}) = -(n-2) + 1,$$ \noindent d'o\`u finalement $a=-1$. \noindent Par ailleurs, nous venons de montrer que $$ K_{X | {\Bbb P} ^{n-2}} = {\cal O}_{{\Bbb P} ^{n-2}}(3-n).$$ \noindent Ainsi, si $n$ est sup\'erieur ou \'egal \`a $4$, $-K_X$ est ample sur ${\Bbb P} ^{n-2}$. Finalement, le fibr\'e $K_X$ bien que gros n'est pas nef et le th\'eor\`eme est d\'emontr\'e. \hskip 3pt \vrule height6pt width6pt depth 0pt \bigskip \noindent {\bf Remarque } La construction pr\'ec\'edente, en dimension $3$, donne un nouvel exemple de vari\'et\'e de Moishezon, de ``Calabi-Yau" satisfaisant $\displaystyle{\operatorname{Pic} (X) = {\Bbb Z} \cdot {\cal O}_X (1) ,}$ (voir aussi \cite{Ogu94}). \subsection{D\'emonstration de la proposition} On d\'emontre (i) et (ii) simultan\'ement. Les points singuliers de $Z$ sont des z\'eros communs des \'equations $$ x_0h_0 +\cdots+ x_{n-1}h_{n-1} = 0$$ et $$x_0 \frac{\partial h_0}{\partial x_i} +\cdots+ x_{n-1} \frac{\partial h_{n-1}}{\partial x_i} + h_i = 0 \ , \ i=0,\ldots \!,n-1.$$ En particulier, $Z$ est lisse au voisinage de ${\Bbb P} ^{1} = \{ x_0 = \dots = x_{n-1} = 0 \}$ d\`es que les $h_i$ ne s'annulent pas simultan\'ement sur ${\Bbb P} ^{1}$. Ceci est vrai pour un choix g\'en\'erique des $h_i$ d\`es que $n$ est sup\'erieur ou \'egal \`a $2$. On d\'eduit alors du th\'eor\`eme de Bertini \cite{G-H78} que si les $h_i$ sont \`a nouveau g\'en\'eriques, l'hypersurface $Z$ est non singuli\`ere partout~: en effet, de fa\c{c}on g\'en\'erale, une relation $$\displaystyle{ \sum_i s_i f_i = 0 }$$ d\'efinit une vari\'et\'e non singuli\`ere en dehors des z\'eros communs des $s_i$ d\`es que les $f_i$ sont g\'en\'eriques dans l'espace des sections holomorphes d'un fibr\'e engendrant en tout point les jets d'ordre inf\'erieur ou \'egal \`a $1$. \medskip D\'eterminons ensuite le fibr\'e normal $N_{{\Bbb P} ^{1}/ Z}$, et pour cela, consid\'erons la suite exacte des fibr\'es normaux : $$ 0 \to N_{{\Bbb P} ^{1}/ Z} \to N_{{\Bbb P} ^{1}/ {\Bbb P} ^{n+1}} = {\cal O}_{{\Bbb P} ^{1}}(1)^{\oplus n} \stackrel{d \sigma}{\to} {\cal O}_{{\Bbb P} ^{n+1}}(2n-1)_{| {\Bbb P} ^1} \to 0 .$$ Il est alors clair que $N_{{\Bbb P} ^{1}/ Z}$ est de degr\'e $-(n-1)$. Pour montrer qu'il est \'egal \`a ${\cal O}_{{\Bbb P} ^{1}}(-1)^{\oplus n-1}$, il suffit donc de montrer qu'il n'a pas de sections (rappelons en effet qu'un th\'eor\`eme d'A.\ Grothendieck affirme que tout fibr\'e vectoriel sur ${\Bbb P} ^1$ est scind\'e). Par la suite exacte pr\'ec\'edente, une section de $N_{{\Bbb P} ^{1}/ Z}$ peut \^etre vue comme une section de ${\cal O}_{{\Bbb P} ^{1}}(1)^{\oplus n}$, annul\'ee par $d \sigma$. Une telle section correspond \`a la donn\'ee d'un $n$-uplet $(s_0,\ldots \!,s_{n-1})$ o\`u les $s_i$ sont des polyn\^omes homog\`enes de degr\'e $1$ en les variables $x_n , x_{n+1}$, que l'on \'ecrit $s_i (x) = s_{i,n}x_n + s_{i,n+1}x_{n+1}$. Dans $N_{{\Bbb P} ^{1}/ {\Bbb P} ^{n+1}}$, on a alors~: $$\displaystyle{s = \sum_{i=0}^{n-1} s_i \frac{\partial}{\partial x_i}}.$$ De m\^eme, notons $$\displaystyle{h_i(x) = \sum_{p=o}^{2n-2} h_{i,p}x_{n}^{p}x_{n+1}^{2n-2-p}}$$ la restriction de $h_i$ \`a ${\Bbb P} ^1$. La relation $d \sigma (s) = 0$ donne ici~: $$\displaystyle{\sum_{i=0}^{n-1} s_i h_i = 0}.$$ Comme $$\displaystyle{ d \sigma = \sum_{i=0}^{n-1} h_idx_i }$$ le long de ${\Bbb P} ^1$, cette relation se traduit par un syst\`eme lin\'eaire \`a $2n$ \'equations en les $2n$ inconnues $s_{i,n},s_{i,n+1}$. Il s'agit de montrer que pour un choix g\'en\'erique des $h_i$, le d\'eterminant de la matrice suivante~: $$ \left( \begin{matrix} h_{0,0}&h_{1,0}&\dots&h_{n-1,0}&0&0&\dots&0\\ h_{0,1}&h_{1,1}&\dots&h_{n-1,1}&h_{0,0}&h_{1,0}&\dots&h_{n-1,0}\\ \vdots&\vdots&&\vdots&\vdots&\vdots&&\vdots\\ h_{0,2n-2}&h_{1,2n-2}&\dots&h_{n-1,2n-2}&h_{0,2n-3}&h_{1,2n-3}&\dots&h_{n-1,2n-3}\\ 0&0&\dots&0&h_{0,2n-2}&h_{1,2n-2}&\dots&h_{n-1,2n-2} \end{matrix} \right) $$ n'est pas nul, ce qui est clair en prenant par exemple $$h_{0,0}= \lambda _0, \ldots \!, h_{n-1,n-1}= \lambda _{n-1}, h_{0,n-1}= \mu _{0}, \ldots \!, h_{n-1,2n-2}= \mu _{n-1}$$ avec $\lambda _i \neq 0$, $\mu_i \neq 0$. \noindent Ainsi, il existe un choix des $h_i$ de sorte que l'hypersurface $Z$ (\'eventuellement singuli\`ere) est lisse au voisinage de ${\Bbb P} ^1$, avec $N_{{\Bbb P} ^{1}/ Z} = {\cal O}_{{\Bbb P} ^{1}}(-1)^{\oplus n-1}$. \medskip Si maintenant les $h_i$ sont choisis de sorte que (i) et (ii) soient satisfaits, alors (iii) est imm\'ediat par adjonction et (iv) d\'ecoule du th\'eor\`eme de Lefschetz~: l'hypersurface $Z$ est le diviseur d'une section d'un fibr\'e ample et les fibr\'es en droites sur $Z$ sont restriction de fibr\'es en droites sur ${\Bbb P} ^{n+1}$.\hskip 3pt \vrule height6pt width6pt depth 0pt \bigskip \noindent {\bf Remarque } Le cas $n=3$ de la proposition pr\'ec\'edente correspond \`a celui des quintiques dans ${\Bbb P} ^4$. Il a \'et\'e consid\'er\'e par S.\ Katz dans \cite{Kat86}. Dans ce cadre, S.\ Katz d\'etermine le fibr\'e normal $N_{{\Bbb P} ^{1}/ Z}$ dans le cas o\`u les $h_i$ sont g\'en\'eriques, mais analyse aussi la situation non g\'en\'erique. Signalons aussi \cite{Cle83}, o\`u H.\ Clemens consid\`ere des questions analogues, toujours en dimension $3$. \medskip \noindent {\bf Remarque } On peut reprendre plus g\'en\'eralement la construction pr\'ec\'edente pour les hypersurfaces de ${\Bbb P} ^{n+1}$ de degr\'e $2n-2k+1$ passant par un ${\Bbb P} ^{k}$ lin\'eaire. On peut alors \`a nouveau montrer que (g\'en\'eriquement) $N_{{\Bbb P} ^{k}/ Z}$ (qui est de degr\'e $k-n$) n'a pas de sections. Cependant, le fibr\'e $N_{{\Bbb P} ^{k}/ Z}$ n'est pas scind\'e si $k \geq 2$. \medskip D\'emontrons ce dernier point en consid\'erant \`a nouveau la suite exacte des fibr\'es normaux~: $$ 0 \to N_{{\Bbb P} ^{k}/ Z} \to N_{{\Bbb P} ^{k}/ {\Bbb P} ^{n+1}} = {\cal O}_{{\Bbb P} ^{k}}(1)^{\oplus n+1-k} \stackrel{d \sigma}{\to} {\cal O}_{{\Bbb P} ^{n+1}}(2n-2k+1)_{| {\Bbb P} ^k} \to 0 .$$ En dualisant cette suite, il vient~: $$ 0 \to {\cal O}_{{\Bbb P} ^{n+1}}(-2n+2k-1)_{| {\Bbb P} ^k} \to {\cal O}_{{\Bbb P} ^{k}}(-1)^{\oplus n+1-k} \to N_{{\Bbb P} ^{k}/ Z}^{\ast} \to 0 .$$ Si $k \geq 2$, comme le fibr\'e ${\cal O}_{{\Bbb P} ^{n+1}}(-2n+2k-1)_{| {\Bbb P} ^k}$ est n\'egatif, la suite exacte longue de cohomologie donne $H^0 ({\Bbb P} ^k, N_{{\Bbb P} ^{k}/ Z}^{\ast}) = 0$. Ceci exclut de suite le fait que $N_{{\Bbb P} ^{k}/ Z}$ soit scind\'e car il serait alors \'egal \`a ${\cal O}_{{\Bbb P} ^{k}}(-1)^{\oplus n-k}$.~\hskip 3pt \vrule height6pt width6pt depth 0pt \section{Une classification en dimension $4$} Dans ce paragraphe, nous consid\'erons des vari\'et\'es de Moishezon (non projectives) $X$ de dimension $4$. Comme pr\'ec\'edemment, nous supposons que $\operatorname{Pic} (X) = {\Bbb Z}$, que $K_X$ est gros et que $X$ est rendue projective apr\`es \'eclatement le long d'une sous-vari\'et\'e lisse $Y$. \subsection{\'Enonc\'e des r\'esultats} Nous montrons les deux r\'esultats suivants~: \medskip \noindent {\bf Th\'eor\`eme H }{\em Sous les hypoth\`eses pr\'ec\'edentes, $Y$ est n\'ecessairement une surface. \noindent Autrement dit, et dans cette situation particuli\`ere, il ne suffit pas d'\'eclater une courbe pour rentrer dans le monde projectif. } \medskip Nous avons vu pr\'ec\'edemment que $K_X$ n'est pas n\'ecessairement nef \`a partir de la dimension $4$. Le r\'esultat suivant montre que l'exemple construit dans le paragraphe pr\'ec\'edent est le ``seul possible" dans le cas o\`u $K_X$ n'est pas nef. Nous reprenons les notations des paragraphes pr\'ec\'edents, $\displaystyle{ \pi : \tilde{X} \to X}$ d\'esigne l'\'eclatement de $X$ le long de $Y$ et $\displaystyle{ f : \tilde{X} \to Z}$ d\'esigne la contraction extr\^emale de Mori sur $\tilde{X}$. \medskip \noindent {\bf Th\'eor\`eme I }{\em Sous les hypoth\`eses pr\'ec\'edentes et si $K_{X}$ n'est pas nef, alors~: (i) le couple $(Y,N_{Y/X})$ est \'egal \`a $({\Bbb P} ^2, {\cal O}_{{\Bbb P} ^{2}}(-1)^{\oplus 2})$, (ii) $f$ contracte le diviseur exceptionnel de $\pi$ sur une courbe rationnelle lisse \`a fibr\'e normal ${\cal O}_{{\Bbb P} ^{1}}(-1)^{\oplus 3}$ dans une vari\'et\'e projective lisse $Z$. En particulier, $f$ est une contraction divisorielle.} \bigskip Ces r\'esultats sont accessibles en dimension $4$ car les contractions de Mori sont ``bien comprises" gr\^ace aux r\'esultats, rappel\'es pr\'ec\'edemment de T.\ Ando \cite{And85} et M.\ Beltrametti \cite{Bel86} pour les contractions divisorielles. \bigskip \subsection{D\'emonstration du th\'eor\`eme H} \medskip Nous traitons s\'epar\'ement les cas $K_X$ nef et $K_X$ non nef. \noindent (i) Le cas $K_X$ non nef~. Il d\'ecoule directement du th\'eor\`eme~F~: en effet, si la contraction de Mori $f$ est une contraction divisorielle, le point (iii) assure que $Y$ est une surface et que $f(E)$ est une courbe. Par ailleurs, le point (iv) exclut la possibilit\'e que $f$ soit une petite contraction (car sinon $\dim Y \geq 3$ !). Mentionnons qu'une premi\`ere version de ce travail \cite{Bo95b} excluait ce cas en utilisant le difficile th\'eor\`eme de structure des petites contractions de Kawamata \cite{Kaw89}. Ceci ach\`eve le cas $K_X$ non nef. \bigskip \noindent (ii) Le cas $K_X$ nef. Dans ce cas, rappelons que la contraction de Mori se factorise par $\pi$. Notons $\displaystyle{ g : X \to Z}$ de sorte que $f = g \circ \pi$. Raisonnons alors par l'absurde en supposant que $Y$ est une courbe. Dans ce cas, il est clair que $f(E)$ (\'egal \`a $g(Y)$) est un point. En effet, dans le cas contraire, $f(E)$ est une courbe et l'application $g$ est finie. Comme $Z$ est projective, on en d\'eduit que $X$ est projective, ce que l'on a rejet\'e. Ainsi $f(E)$ est un point et le diviseur $E$ est une vari\'et\'e de Fano. \noindent Nous allons montrer que $E$ est en fait isomorphe \`a la quadrique de dimension $3$, ce qui fournira la contradiction~; une quadrique de dimension $3$, dont le nombre de Picard est $1$, ne pouvant \^etre \'egale au projectivis\'e d'un fibr\'e de rang $3$ sur une courbe, pour lequel le nombre de Picard est $2$ ! \noindent Pour cela, remarquons que $Z$ \'etant ${\Bbb Q}$-factorielle \`a singularit\'es terminales, il existe un entier $m$ non nul tel que $mK_Z$ est de Cartier. Alors~: $$mK_X = g^{\ast}(mK_Z).$$ En particulier, la restriction de $K_X$ \`a $Y$ est triviale. Il en d\'ecoule que $K_Y = \det N_{Y/X}$, et par cons\'equent, $$K_E = \pi ^{\ast}(K_Y - \det N_{Y/X} ) + 3{\cal O}_{E}(-1) = 3{\cal O}_{E}(-1).$$ On en d\'eduit que ${\cal O}_{E}(1)$ est ample et que $E$ est une vari\'et\'e (de dimension $3$) d'indice $3$~; rappelons que l'indice d'une vari\'et\'e de Fano $V$ est le plus grand entier $r >0$ tel qu'il existe un fibr\'e en droites $L$ avec $-K_V = rL$. Or, le th\'eor\`eme de Kobayashi-Ochiai \cite{KoO73} affirme qu'{\em une vari\'et\'e de Fano de dimension $n$ et d'indice $n$ est isomorphe \`a la quadrique ${\cal Q}_n$}. On en d\'eduit ici que $E$ est la quadrique ${\cal Q}_3$ comme annonc\'e.\hskip 3pt \vrule height6pt width6pt depth 0pt \bigskip Dans la situation du th\'eor\`eme H et lorsque $K_X$ est nef, nous avons vu que la contraction de Mori $f$ sur $\tilde{X}$ se factorise en une application birationnelle $g : X \to Z$ qui contracte la surface $Y$. La proposition suivante pr\'ecise le cas o\`u $g(Y)$ est r\'eduit \`a un point~: \bigskip \noindent {\bf Proposition } {\em Si $K_X$ est nef et si $f(E)$ (\'egal \`a $g(Y)$) est un point, alors le couple $(Y,N_{Y/X})$ est \'egal \`a $({\Bbb P} ^2, T^{\ast}{\Bbb P} ^2)$, $({\Bbb P} ^2, {\cal O}_{{\Bbb P} ^2}(-1) \oplus {\cal O}_{{\Bbb P} ^2}(-2) )$ ou $({\cal Q}_2, {\cal O}_{{\cal Q}_2}(-1,-1)^{\oplus 2} )$.} \bigskip Nous ne connaissons pas d'exemples explicites o\`u ces possibilit\'es sont effectivement r\'ealis\'ees, mais nous pouvons remarquer qu'aucune n'est exclue {\em a priori} par les r\'esultats de T.\ Ando et M.\ Beltrametti. \bigskip La d\'emonstration de la proposition d\'ecoule directement du th\'eor\`eme suivant de T.\ Peternell \cite{Pet91}~: \medskip \noindent {\bf Th\'eor\`eme (T.\ Peternell, 1991) } {\em Soit $V$ une vari\'et\'e projective de dimension $n$ et soit $E$ un fibr\'e vectoriel de rang $n$ sur $V$ de sorte que $c_1(E) = c_1(X)$. Alors, le couple $(V,E)$ est \'egal \`a $({\Bbb P} ^n, {\cal O}_{{\Bbb P} ^n}(2) \oplus {\cal O}_{{\Bbb P} ^n}(1)^{\oplus n+1})$, $({\Bbb P} ^n, T {\Bbb P} ^n)$ ou $({\cal Q}_n, {\cal O}_{{\cal Q}_n}(1)^{\oplus n})$.} \bigskip \noindent {\bf D\'emonstration de la proposition} La d\'emonstration du th\'eor\`eme H montre que $$K_Y = \det N_{Y/X}$$ et que $${\cal O}_E(1) = {\cal O}_{{\Bbb P} (N_{Y/X}^{\ast})}(1)$$ est ample, donc que $N_{Y/X}^{\ast}$ est aussi ample. Le r\'esultat d\'ecoule du th\'eor\`eme de T.\ Peternell appliqu\'e au couple $(Y, N_{Y/X}^{\ast})$. \hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{D\'emonstration du th\'eor\`eme I} Notons $F$ la fibre g\'en\'erale de $f$ restreinte au diviseur exceptionnel $E$. Comme $f(E)$ est une courbe, $F$ est de dimension $2$. D'apr\`es le th\'eor\`eme de T.\ Ando et M.\ Beltrametti, $F$ est \'egal \`a ${\Bbb P} ^2$ ou \`a la quadrique ${\cal Q}_2$. De plus, il a \'et\'e vu pr\'ec\'edemment que $F$ coupe les fibres de $\pi$ sur des points. On en d\'eduit que $\pi _{| F} : F \to Y$ est une application surjective finie. Deux cas sont \`a distinguer~: \medskip - $F$ est \'egal \`a ${\Bbb P} ^2$. \noindent Dans ce cas, $Y$ est aussi \'egal \`a ${\Bbb P} ^2$. En effet, un r\'esultat de R.\ Lazarsfeld \cite{Laz84} affirme que si {\em $h : {\Bbb P} ^n \to V$ est une application holomorphe surjective finie sur une vari\'et\'e de dimension $n$, alors $V$ est isomorphe \`a ${\Bbb P} ^n$}~; en dimension $2$, on peut trouver une d\'emonstration \'el\'ementaire dans \cite{BPV84}. \noindent Montrons alors que $$\displaystyle{ \pi _{| F} : F \simeq {\Bbb P} ^2 \to Y \simeq {\Bbb P} ^2 }$$ est un isomorphisme. Pour cela, il suffit de montrer que $\pi _{| F}$ est un isomorphisme local, car alors $\pi _{| F}$ est un rev\^etement donc le rev\^etement trivial. Soient donc $x$ dans $F$ et $L$ un ${\Bbb P} ^1$ quelconque passant par $\pi (x)$. Sa pr\'e-image $\pi ^{-1}(L)$ est une surface d'Hirzebruch bi-r\'egl\'ee donc ${\Bbb P} ^1 \times {\Bbb P} ^1$. L'intersection $F \cap \pi ^{-1}(L)$ est alors une r\'eunion de ${\Bbb P} ^1$ ``horizontaux". La restriction de $\pi$ au ${\Bbb P} ^1$ horizontal passant par $x$ est donc un isomorphisme sur son image. Ceci \'etant vrai pour tout ${\Bbb P} ^1$ passant par $\pi (x)$, ceci montre bien que $d \pi_{| F}(x)$ est surjective, donc inversible, et que $\pi _{| F}$ est un isomorphisme local. Ainsi, $$\displaystyle{ \pi _{| F} : F \simeq {\Bbb P} ^2 \to Y \simeq {\Bbb P} ^2 }$$ est un isomorphisme. On en d\'eduit que le fibr\'e normal $N_{Y/X}$ est scind\'e~; on d\'efinit alors $a$ et $b$ en posant~: $$N_{Y/X} = {\cal O}_{{\Bbb P} ^2}(a) \oplus {\cal O}_{{\Bbb P} ^2}(b).$$ Le fait que $\pi ^{-1}(L) \simeq {\Bbb P} ^1 \times {\Bbb P} ^1$ montre m\^eme que $a=b$. \noindent Comme $K_X$ n'est pas nef, $K_X$ est n\'egatif sur $Y$. Il vient alors~: $$\displaystyle{\deg (K_{X | Y}) = -3-2a < 0}$$ d'o\`u $a \geq -1$. L'affirmation suivante permet de conclure~: \medskip \noindent {\bf Affirmation } {\em L'entier $a$ est strictement n\'egatif.} \medskip \noindent {\bf D\'emonstration } Par l'absurde, supposons que $a \geq 0$. Alors, si $C$ d\'esigne un ${\Bbb P} ^1$ de $Y = {\Bbb P} ^2$, on a~: $$ H^1(C,N_{C/X}) = H^1({\Bbb P}^1, {\cal O}_{{\Bbb P} ^1}(1) \oplus {\cal O}_{{\Bbb P} ^1}(a)^{\oplus 2}) = 0,$$ d'o\`u~: $$ \dim_{[C]} \operatorname{Hilb} (X) = \dim H^0({\Bbb P}^1, {\cal O}_{{\Bbb P} ^1}(1) \oplus {\cal O}_{{\Bbb P} ^1}(a)^{\oplus 2}) = 2a + 4.$$ Or, $$\dim_{[C]} \operatorname{Hilb} (Y) = \dim_{[{\Bbb P} ^1]} \operatorname{Hilb} ({\Bbb P} ^2) = 2.$$ Comme $a \geq 0$, on en d\'eduit que~: $$ \dim_{[C]} \operatorname{Hilb} (X) > \dim_{[C]} \operatorname{Hilb} (Y)$$ si bien que $C$ se d\'eforme dans $X$ hors de $Y$. Ceci n'est pas possible comme nous l'avons d\'ej\`a rencontr\'e car $K_X$ est positif sur les courbes non incluses dans $Y$. Ici, $K_X$, n'\'etant pas nef, est n\'egatif sur $C$. Contradiction ! \hskip 3pt \vrule height6pt width6pt depth 0pt \medskip Ainsi, $a=-1$ et $f$ contracte $E$ sur une courbe rationnelle lisse \`a fibr\'e normal ${\cal O}_{{\Bbb P} ^1}(-1)^{\oplus 3}$ dans la vari\'et\'e projective lisse $Z$. \medskip - $F$ est \'egal \`a la quadrique ${\cal Q}_2$. \noindent Nous montrons que ce cas ne peut pas arriver. En effet, $Y$ est alors isomorphe \`a ${\Bbb P} ^2$ ou ${\cal Q}_2$. Le cas $Y \simeq {\Bbb P} ^2$ s'exclut exactement comme pr\'ec\'edemment~: $\pi _{| F}$ r\'ealise un isomorphisme entre la quadrique et ${\Bbb P} ^2$ ! \noindent Si $Y \simeq {\cal Q}_2$, le raisonnement est plus simple et il est inutile de montrer que $$\pi _{| F} : F \simeq {\cal Q}_2 \to Y \simeq {\cal Q}_2$$ est un isomorphisme. Choisissons en effet un ${\Bbb P} ^1$ dans $Y$, \`a savoir un des g\'en\'erateurs de $H_2({\cal Q}_2,{\Bbb Z})$, sur lequel $K_X$ est strictement n\'egatif (il en existe car $K_X$ n'est pas nef). Alors $N_{Y/X}$ restreint \`a ${\Bbb P} ^1$ est de la forme $${\cal O}_{{\Bbb P} ^1}(a) \oplus {\cal O}_{{\Bbb P} ^1}(a)$$ (ceci comme pr\'ec\'edemment car $\pi ^{-1}({\Bbb P} ^1) \simeq {\Bbb P} ^1 \times {\Bbb P} ^1$). La suite exacte~: $$ 0 \to T {\Bbb P} ^1 \to TX_{| {\Bbb P} ^1} \to N_{{\Bbb P} ^1/X} \to 0,$$ et le fait que $N_{{\Bbb P} ^1/{\cal Q}_2}$ est trivial entrainent que $$\deg (-K_{X | {\Bbb P} ^1}) = 2 + 2a > 0$$ et donc que $a \geq 0$. Ceci est, comme dans le cas pr\'ec\'edent, absurde car ce ${\Bbb P} ^1$ se d\'eformerait alors dans $X$ hors de $Y$ ! \hskip 3pt \vrule height6pt width6pt depth 0pt \subsection{Quelques commentaires} Comme nous venons de le voir, la situation en dimension $4$ est tr\`es satisfaisante lorsque $K_X$ n'est pas nef. Dans le cas o\`u $K_X$ est nef, nous avons obtenu une restriction sur le centre de l'\'eclatement seulement lorsque le diviseur exceptionnel $E$ est contract\'e sur un point. Au moment o\`u nous finissions la r\'edaction de cette th\`ese, nous avons appris que M.\ Andreatta et J.A.\ Wi\'sniewski terminent la r\'edaction d'un travail consistant \`a classifier les contractions extr\^emales divisorielles en dimension $4$ sur une vari\'et\'e non-singuli\`ere, \'etendant ainsi les r\'esultats de M.\ Beltrametti au cas o\`u le diviseur est contract\'e sur une courbe ou sur une surface. Nous sommes en mesure d'appliquer leurs r\'esultats dans notre situation pour obtenir la proposition suivante. Pr\'ecisons cependant que nous n'avons pas encore une version \'ecrite du travail en question mais que notre seule r\'ef\'erence est une s\'erie de discussions informelles avec M.\ Andreatta, M.\ Mella et J.A.\ Wi\'sniewski. \medskip \noindent {\bf Proposition } {\em Soit $X$ comme dans le th\'eor\`eme H. On suppose que $K_X$ est nef. Si $f$ est la contraction de Mori d\'efinie sur $\tilde{X}$, alors~: (i) le diviseur exceptionnel $E$ est contract\'e sur une courbe ou un point. Autrement dit, $f(E)$ n'est pas une surface, (ii) si $f(E)$ est une courbe, cette derni\`ere est une courbe lisse de singularit\'es nodales ordinaires $3$-dimensionelles et le centre $Y$ de l'\'eclatement $\pi$ est une surface r\'egl\'ee dont les fibres ${\Bbb P} ^1$ ont pour fibr\'e normal ${\cal O}_{{\Bbb P} ^1}\oplus {\cal O}_{{\Bbb P} ^1}(-1)^{\oplus 2}$. Autrement dit, la situation est localement le produit d'une courbe par le mod\`ele analogue en dimension~$3$.} \medskip Cette proposition termine la description des situations possibles~; cependant nous ne connaissons pas \`a l'heure actuelle d'exemple explicite o\`u le point (ii) est r\'ealis\'e. \medskip \noindent {\bf ``D\'emonstration"} Tout d'abord, mentionnons que la contraction divisorielle que nous \'etudions est tr\`es particuli\`ere car nous savons {\em a priori} que le diviseur exceptionnel a une structure de fibration en espaces projectifs sur une base lisse. Pour le point (i), supposons par l'absurde que $f(E)$ est une surface. Dans ce cas, la fibre g\'en\'erale est un ${\Bbb P} ^1$ et M.\ Andreatta et J.A.\ Wi\'sniewski montrent qu'une \'eventuelle fibre particuli\`ere est soit ${\Bbb P} ^2$, soit la quadrique ${\cal Q}_2$, soit la quadrique singuli\`ere ${\cal Q}_2^0$. Dans notre situation, une \'eventuelle fibre particuli\`ere est donc ${\cal Q}_2$ et l'image $\pi ({\cal Q}_2)$ dans $Y$ est une courbe rationnelle $C$ d'auto-intersection $-1$. \noindent Montrons que ceci n'est pas possible, \`a nouveau par un argument de d\'eformation. En effet, $K_X$ est trivial sur $C$, donc $$ N_{C/X} = {\cal O}_{{\Bbb P} ^1}(-1) \oplus {\cal O}_{{\Bbb P} ^1}(a) \oplus {\cal O}_{{\Bbb P} ^1}(b)$$ o\`u $a$ et $b$ sont deux entiers satisfaisant la relation $a + b = -1$. De l\`a $$ \dim \operatorname{Hilb} _{[C]}(X) \geq \dim H^0(C,N_{C/X}) - \dim H^1(C,N_{C/X}) = a+b+2 = 1 > 0$$ d'o\`u l'on d\'eduit que $C$ se d\'eforme dans $X$ et ce hors de $Y$. Pour le point (ii), nous sommes dans la situation ``facile" du travail de M.\ Andreatta et J.A.\ Wi\'sniewski car les fibres de $f$ restreinte \`a $E$ sont \'equi-dimensionnelles. Dans notre situation, la fibre g\'en\'erale est une quadrique ${\cal Q}_2$ et il n'y a pas de fibres particuli\`eres~: la situtation est, transversalement \`a $f(E)$, la r\'esolution d'une singularit\'e nodale $3$-dimensionnelle.\hskip 3pt \vrule height6pt width6pt depth 0pt \newpage
1992-10-14T18:03:43
9209
alg-geom/9209001
en
https://arxiv.org/abs/alg-geom/9209001
[ "alg-geom", "math.AG" ]
alg-geom/9209001
Dr Roger Brussee
Rogier Brussee
On the $(-1)$-curve conjecture of Friedman and Morgan
13 pages, LaTeX 2.09
null
null
null
null
Main difference with previous version: we prove that every differentiably embedded sphere with self intersection $-1$ in a simply connected algebraic surface with $p_g >0$ is homologous to a $(-1)$-curve if $|K_{\min}|$ contains a smooth irreducible curve of genus at least 2 and $p_g$ is even or $K_{\min}^2 \not\equiv 7 \pmod8$ (here $K_{\min}$ is the canonical class of the minimal model).
[ { "version": "v1", "created": "Sat, 12 Sep 1992 12:01:29 GMT" }, { "version": "v2", "created": "Wed, 14 Oct 1992 16:46:38 GMT" } ]
2008-02-03T00:00:00
[ [ "Brussee", "Rogier", "" ] ]
alg-geom
\section{Introduction.} It is now well known that the deformation type of an algebraic surface is determined by its oriented diffeomorphism type up to a finite number of choices \cite{F&M:ellipticI}\,\cite[theorem S.2]{F&M}. It is therefore natural to ask if a deformation invariant is in fact an invariant of the underlying oriented differentiable manifold. For example, Van de Ven conjectured that this is true for the Kodaira dimension \cite{O&V:overview}\,\cite{F&M}\,\cite{Pidstrigach&Tyurin:specialinst}. In this paper we study whether the deformation invariant decomposition \[ {alggeodecomp} H_2(X) = H_2(X_{\min}) \oplus^\perp \mathop\oplus {\Bbb Z} E_i, \] in the homology of the minimal model and the span of the $(-1)$-curves is invariant under orientation preserving diffeomorphisms (cf. \cite[conj. 2]{F&M:BAMS}) A {\sl $(-1)$-curve} on a complex surface is a smooth holomorphically embedded 2-sphere with self-intersection $-1$. A $(-1)$-curve can be blown down to obtain a new smooth complex surface. Successively contracting all $(-1)$-curves gives the minimal model~$X_{\min}$ which is unique if $p_g >0$. More generally we will call the total transform of a $(-1)$-curve on some intermediate blow-down a $(-1)$-curve as well. It is in this sense that the decomposition~\ref{alggeodecomp} is a deformation invariant. A {\sl $(-1)$-sphere} on a $4$-manifold is a smooth differentiably embedded 2 sphere with self-intersection~$-1$. A classical $(-1)$-curve is obviously a $(-1)$-sphere, a reducible one can be deformed to a $(-1)$-sphere by smoothing out the double points. Moreover if two $(-1)$-curves are orthogonal, they can be deformed in disjoint $(-1)$-spheres. Friedman and Morgan conjectured that if a surface has a unique minimal model, then modulo homological equivalence the relation between its $(-1)$-spheres and its $(-1)$-curves is the strongest possible. \pr@claim{\bf}{ \thetheorem}{\sl} Conjecture (-1)conj. (Friedman and Morgan \cite[conj. 2,3 prop. 4]{F&M:BAMS}) Let $X$ be a simply connected algebraic surface with Kodaira dimension~$\kappa \ge 0$. Then every $(-1)$-sphere is homologous to a $(-1)$-curve up to orientation. In particular the decomposition~(\ref{alggeodecomp}) is invariant under orientation preserving diffeomorphisms. (I have slightly reformulated the conjecture, and added the simply connectedness hypothesis). Now, $X$ contains $n$ disjoint $(-1)$-spheres if and only if there is a differentiable connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \#n{\bar\P}$. The decomposition \ref{alggeodecomp} can be thought of as being induced by this special connected sum decomposition. Friedman and Morgan also made a conjecture about more general connected sum decompositions, which would imply conjecture~\ref{(-1)conj} above. \pr@claim{\bf}{ \thetheorem}{\sl} Conjecture consumconj. (Friedman and Morgan \cite[conj. 9]{F&M:BAMS}) Let $X$ be a simply connected algebraic surface with $\kappa \ge 0$. Suppose $X$ admits a connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$ for a negative definite manifold~$N$, then $H_2(N,{\Bbb Z})$ is generated by $(-1)$-curves. Note that by theorems of Donaldson (\cite[th. 1.3.1, 9.3.4, 10.1.1]{D&K}, $N$ has automatically a standard negative definite intersection form if $p_g(X)>0$. Conjecture \ref{consumconj} (and hence conjecture \ref{(-1)conj}) has been proved for blow-ups of simply connected surfaces with $p_g >0$ and big monodromy (like elliptic surfaces or complete intersections), and simply connected surfaces with $p_g >0$ whose minimal model admits a spin structure (i.e. $K_{\min} \equiv 0(2)$) \cite[cor. 4.5.4]{F&M}. Conjecture~\ref{(-1)conj} has been proved for the Dolgachev surfaces (i.e. $\kappa=p_g = 0$). For minimal surfaces, conjecture \ref{(-1)conj} would imply strong minimality. A $4$-manifold is called {\sl strongly minimal} if for every diffeomorphism $X \# N_1 \buildrel\scriptscriptstyle \rm diff\over\iso Y \# N_2$ with $N_i \buildrel\scriptscriptstyle \rm diff\over\iso n_i {\bar\P}^2$, we have $H_2(N_2) \subset H_2(N_1)$ (c.f. \cite[def. IV.4.6]{F&M}). Conjecture \ref{(-1)conj} would also imply that the canonical class of the minimal model $K_{\min}$ is invariant mod 2 under orientation preserving diffeomorphisms. Conjecture \ref{consumconj} would imply that a minimal surface is {\sl irreducible} i.e. for every decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$, say $N$ is {\relax homeomorphic} to $S^4$, thereby avoiding the Poincar\'e conjecture. In this paper we will show that under the stronger assumption $p_g >0$, $(-1)$-spheres must give rise, if not to $(-1)$-curves then in any case to special algebraic 1-cycles. Indeed, the main theorem~\ref{main} below leaves little room for $(-1)$-spheres not homologous to a $(-1)$-curve. Furthermore we will reduce a similar statement for general connected sum decompositions, to a technical problem in gauge theory. To state the theorem we need some notation. Let $N_1(X)_{\Bbb Z} \subset H_2(X,{\Bbb Z})$ be the preferred subgroup of algebraic classes i.e. the subgroup generated by algebraic curves. Its rank $\rho$ is the Picard number. The effective cone $\rmmath{NE}(X) \subset N_1(X)_{\Bbb Q}$ is the cone spanned by positive rational multiples of algebraic curves. The subcone $\rmmath{NE}(X_{\min}) = \rmmath{NE}(X) \cap H_2(X_{\min},{\Bbb Q})$ is the cone spanned by the pullbacks of rational curves on the minimal model, i.e. the effective cone in $H_2(X_{\min})$. Finally we note that since $N_1(X)_{\Bbb Q}$ is a finite dimensional vector space, the closure of the effective cone $\overline{\rmmath{NE}}(X)$ is well defined. \pr@claim{\bf}{ \thetheorem}{\sl} Theorem main. Let $X$ be a simply connected algebraic surface with $p_g >0$ and let~$K$ be its canonical divisor. Then for every $(-1)$-sphere in $X$, there is an orientation such that $e$ is either represented by a $(-1)$-curve or $e \in \overline{\rmmath{NE}}(X_{\min})$, depending on whether $K\cdot e$ is negative or positive respectively. Note that $K\cdot e \ne 0$ since $K\cdot e \equiv e^2 \!\!\pmod 2$. I have no examples where $K\cdot e$ is positive (i.e. a counter-example to conjecture~\ref{(-1)conj}) but without further assumptions neither can I exclude this case. \pr@claim{\bf}{ \thetheorem}{\sl} Corollary maincor. In addition to the assumptions of the theorem suppose that the minimal model $X_{\min}$ has Picard number $1$ or that the linear system $|K_{\min}|$ contains a smooth irreducible curve of genus at least two, and that $p_g$ is even or $K_{\min}^2 \not\equiv 7 \pmod 8$, then every $(-1)$-sphere is homologous to a $(-1)$-curve (i.e. conjecture~\ref{(-1)conj} is true for $X$). we will use this corollary to prove conjecture~\ref{(-1)conj} for blow-\-ups of Horikawa surfaces with $K_{\min}^2$ even and zero-sets of general sections in sufficiently ample $n-2$-bundles on $n$-folds with $\rho =1$ generalising Friedman and Morgan's result for complete intersections in ${\Bbb P}^n$. The proof of theorem~\ref{main} is based on two very general properties of the ${\rm SO}(3)$ Donaldson-Kotschick invariant~$\phi_k$. Kotschick observed that it follows from the invariance properties of the $\phi_k$~polynomial, that it is divisible by the Poincare dual of a $(-1)$-sphere. On the other hand, using Morgan's algebro geometric description of the Donaldson polynomials we show that $\phi_k$ has pure Hodge type for $k \gg 0$. The theorem and the corollary then follow by using Donaldson's and O'Grady's non-triviality results. \sloppy \pr@claim{\sl}{\relax}{\relax}Acknowledgement *. I have greatly benefitted from discussions with Stephan Bauer, Chris Peters, Victor Pidstrigach, Jeroen Spandaw, Kieran O'Grady, Gang Xiao and Ping Zhang, who I would all like to thank heartily. Special thanks for Simon Donaldson for helping me with gauge theory, and for inviting me to Oxford university. Its mathematics institute has proved to be a very friendly and stimulating environment. This paper grew out of work in my thesis \cite{RB:thesis}. It is a pleasure to thank my thesis advisors Martin L\"ubke and Van de Ven for their help and insight. (here they can not remove such words !) \fussy \section{The $\phi_k$~polynomials.} We will need ${\rm SO}(3)$ Donaldson polynomials $q_{L,k,\Omega}$ and in particular the $\phi_k$~invariant introduced by Kotschick \cite{Kotschick:SO(3)}. Let $X$ be a simply connected 4-manifold with odd $b_+ \ge 3$. The polynomial $q_{L,k,\Omega}$ on $H_2(X)$ corresponds to the moduli space $\M^{\rm asd}(L,k)$ of ASD ${\rm SO}(3)$-connections on the ${\rm SO}(3)$-bundle~$P_k$ with $w_2(P_k) \equiv L \pmod 2$, and $p_1(P_k) = -4k$, oriented by the choice of the lift~$L$ of $w_2(P_k)$, and an orientation $\Omega$ of a maximal positive subspace in $H^2(X,{\Bbb R})$ \cite[\S 9.2]{D&K}. We choose $\Omega$ once and for all (e.g. using a complex structure if present), and we will suppress it in the notation. $q_{L,k}$ has degree $ d = 4k - \numfrac32 (1 + b_+) $. Note that the ${\rm SO}(3)$ bundle~$P_k$ exists if and only if $p_1 \equiv K^2 \pmod 4$, and that $k \in {\numfrac14} {\Bbb Z}$. To define $\phi_k(X)$ we lift the second Stiefel Whitney class~$w_2(X)$ of the manifold to an integral class~$K$. For complex surfaces, the canonical divisor is such a lift. Now define $\phi_k = q_{K,k} $. $\phi_k(X)$ is invariant under orientation preserving diffeomorphisms up to sign. Now suppose that $X$ has a decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$ for a negative definite manifold $N$, necessarily with standard intersection form. Then we have a decomposition $H_2(X) = H_2(X') \mathop\oplus H_2(N)$. Choose generators $e_1,\ldots,e_n$ of $H_2(N)$, such that $K\cdot e_i \equiv -1 \pmod 4$. This fixes the generators up to permutation. By Poincar\'e duality we can consider the generators~$e_i$ as linear forms on $H_2(X)$. Any polynomial $Q$ on $H_2(X)$ can be uniquely written as a polynomial in the dual classes of $e_i$ with polynomials on $H_2(X')$ as coefficients ({\sl the $e$-expansion}). \pr@claim{\bf}{ \thetheorem}{\relax} Definition good. A $4$-manifold is said to have a {\sl good} connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$ if for every generator $e_i$ of $H_2(N)$, $e_i$ divides $q_{L,k}(X)$ for all $L$ with $L\cdot e_i$ odd and all $k \gg 0$. \pr@claim{\bf}{ \thetheorem}{\sl} Proposition {or.princip}. (Kotschick \cite[prop. 8.1]{Kotschick:SO(3)}) A connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# n{\bar\P}^2$ is a good connected sum decomposition. \proof\ (sketch). For notational simplicity we only prove divisibility for the $\phi_k$ invariant. The reflection~$R_e$ in the hyperplane defined by a generator~$e$ of $H_2(n{\bar\P}^2)$ can be represented by an orientation preserving diffeomorphism, for example ${\rm id} \# {\Bbb C}$-conjugation (c.f. \cite[prop 2.4]{F&M:ellipticI}). Then it follows from the general invariance properties of the ${\rm SO}(3)$-polynomials \cite[9.2.2]{D&K} that $R_e^*\phi_k(X) = - \phi_k(X)$. Hence $\phi_k(X)$ is an odd polynomial in the dual class of $e$, in particular $\phi_k$ is divisible by $e$. \ifcomment\bgroup\par\medskip\noindent\small $$ R_e^* \phi_k = R_e^*q_{\Omega,K,k} = q_{R_e^*\Omega,R_e^*K,k} = (-1)^{\(K-R_e^*K \over 2\)^2}q_{\Omega,K,k} = -\phi_k. $$ \par\medskip\noindent\egroup\fi \endproof One should expect that any connected sum decomposition is good. This is because the coefficients $q_{L,k,N,I}$ of the $e$-expansion have the same invariance properties as $q_{L,k}$ under orientation preserving diffeomorphisms of $X'$. Conjecturally, these invariants depend only on the homotopy type of $N$, (c.f. \cite[conjecture above lemma 4.5.6]{F&M}) and so the argument for $N = n{\bar\P}^2$ would give the divisibility by the generators in general. A naive gauge theoretic analysis seems to confirm this conjecture, but some technical difficulties remain to be overcome. In any case we will state and prove our results for surfaces admitting a good connected sum decomposition. \section{Pureness of the Donaldson polynomials.} Now we come to the algebraic geometric part of the proof. The Hodge structure on $H^2(X,{\Bbb Z})$ induces a natural Hodge structure on $S^d H^2(X)$. Let $$ S^d H^2(X) \lhook\nobreak\joinrel\nobreak\m@p--\rightarrow{j} H^{2d}(X\times\cdots\times X) $$ be the natural injection in the cohomology of the $d$ fold product of $X$. Then $j$ is a map of Hodge structures. Hence a polynomial $q\in S^d H^2(X)$ is of pure Hodge type~$(d,d)$ if and only if $j(q)$ is pure of Hodge type~$(d,d)$. Now clearly a sufficient condition for $j(q)$ to be of type~$(d,d)$ is that it is represented by an algebraic cycle. We will prove the rather natural statement that those Donaldson polynomials that can be computed completely by algebraic geometry give rise to algebraic cycles. However to make this statement precise requires serious work (as so often in mathematics). Fortunately almost all of the work has already been done by J. Morgan \cite{Morgan}. \pr@claim{\bf}{ \thetheorem}{\sl} Proposition (d,d). Let $X$ be a simply connected algebraic surface with $p_g >0$. Then if $L \in \rmmath{NS}(X)$, there is constant $k_0 >0$ such that all Donaldson polynomials $q_{L,k}$ with $k > k_0$ and the integer ${\numfrac12} (L^2 -L\cdot K) - {\numfrac14}(L^2 + 4k)$ odd are represented by algebraic cycles. In particular these polynomials are of Hodge type $(d,d)$, where $d = \deg(q_{L,k}) = 4k - 3(1 + p_g)$. \proof. First suppose that $L \equiv 0 \pmod 2$, then $q_{L,k}$ is up to sign just the ${\rm SU}(2)$ polynomial $q_k$. Now the lemma follows directly from recent results of Morgan \cite{Morgan}. He shows that for odd $k\gg 0$, $q_k$ can be computed as follows. Let ${\mkern4mu\overline{\mkern-4mu\M}}^G_k = {\mkern4mu\overline{\mkern-4mu\M}}^G_k(H,0,k)$ be the closure of the moduli space of $H$-slope stable bundles in the moduli space of Gieseker $H$-stable sheaves with $c_1 = 0$, $c_2 = k$. For odd~$c_2$, there exists a universal sheaf~$\xi$ on ${\mkern4mu\overline{\mkern-4mu\M}}^G_k$ (cf. \cite[Remark A7]{Mukai:K3I} and \cite[prop. 2.2]{OGrady}) which determines a correspondence $$ \eqalign{ \nu:H_2(X) &\to H^2({\mkern4mu\overline{\mkern-4mu\M}}^G_k) \\ \Sigma &\to c_2(\xi) \slant \Sigma. } $$ Then if $H$ is $k$-generic (in a sense to be made more precise below) and $k \gg 0$, we have \cite[theorem 1]{Morgan} \[ {nupol} q_k(\Sigma) = \<\nu(\Sigma)^d,[{\mkern4mu\overline{\mkern-4mu\M}}^G_k]>. \] Now since the universal sheaf~$\xi$ is algebraic, it actually determines Chow cohomology classes $c_2(\xi) \in A^2(X\times{\mkern4mu\overline{\mkern-4mu\M}}^G_k)$ (cf. \cite[Definition 17.3]{Fulton}). Consider the diagram $$ \cdalign{ X^d\times{\mkern4mu\overline{\mkern-4mu\M}}^G_k \\ \llap{$\scriptstyle \pi_{X^d}$} \swarrow \quad \searrow\rlap{$\scriptstyle\pi_i$} \\ X^d \hskip 5em (X\times{\mkern4mu\overline{\mkern-4mu\M}}^G_k)_i. \hskip -3em } $$ Then by equation~(\ref{nupol}), the algebraic cycle $$ j(q_k) = \int_{[{\mkern4mu\overline{\mkern-4mu\M}}^G_k]} \pi_1^*c_2(\xi)\cdots \pi_d^*c_2(\xi) \in A^d(X^d) \iso A_d(X^d), $$ represents the image of the Donaldson polynomial~$q_k$ on the level of Chow groups. (Integration over the fibre $\int_{[{\mkern4mu\overline{\mkern-4mu\M}}^G_k]}$ is defined formally as the composition $$ A^i (X^d \times {\mkern4mu\overline{\mkern-4mu\M}}^G_k) \m@p--\rightarrow{[\pi_{X^d}]} A^{i-d}(X^d \times {\mkern4mu\overline{\mkern-4mu\M}}^G_k \to X^d) \m@p--\rightarrow{\pi_*} A^{i-d}(X^d), $$ where $[\pi_{X^d}]$ is the orientation class of the flat map $\pi_{X^d}$ (cf. \cite[section 17.4]{Fulton}). This proves the lemma if $L \equiv 0 \pmod 2$. In case $L \not \equiv 0 \pmod 2$, the results of Morgan carry over virtually unchanged, in fact the corresponding results are rather easier. To be more precise, for a Hodge metric~$g_H$, the moduli space of irreducible ASD ${\rm SO}(3)$-connections with $w_2 \equiv L \pmod 2$ and $-p_1 = 4k$ can be identified with the moduli space ${\cal M}_k$ of $H$-slope stable bundles with $c_1 = L$ and $4c_2 - c_1^2 = 4k$. For $k \gg 0$, the closure of the moduli space of $H$-stable bundles in the moduli space of Gieseker stable sheaves ${\mkern4mu\overline{\mkern-4mu\M}}^G_k$ has the proper complex dimension $d=4k - 3(1+p_g)$ and is generically smooth. Moreover ${\mkern4mu\overline{\mkern-4mu\M}}_k^G$ carries a universal sheaf~$\xi$ \cite[prop. 2.2]{OGrady}, and the class $c_2(\xi) - {\numfrac14} c_1^2(\xi)$ defines a $\nu$ correspondence and Chow cohomology classes just as in the discussion above. Finally, we choose a polarisation~$H$ which is $k$-generic in the sense that $H\cdot(L-2N) \ne 0$ for all $N \in \rmmath{NS}(X)$ with $-4k \le (L -2N)^2 < 0$ (i.e. $H$ is not on a wall). Then since $L \ne 0 \pmod 2$ every Gieseker $H$-semistable sheaf is actually slope $H$-stable. Now all of the discussion in \cite{Morgan} to relate the Gieseker and the Uhlenbeck compactification, and the $\mu$ and $\nu$ correspondence as far as it is concerned with slope stable sheaves and bundles carries over. \ifcomment\bgroup\par\medskip\noindent\small Actually the discussion of relative classes in 6.4.2 and 6.4.3 does not carry over, but is only needed to deal with the trivial connection anyway. \par\medskip\noindent\egroup\fi \endproof \pr@claim{\sl}{ \thetheorem}{\relax}Remark (d,d)diff. The pureness of the polynomials $q_{L,k}$ is also suggested by a differential geometric argument, which seems to be the point of view taken by Tyurin \cite[\S 4.22]{Tyurin:algaspect}. The complex structure on $X$ induces the complex structure $$ T^{10}{\cal B}^*_X = \{a \in A^{10}(\mathop{{\cal E}\mkern-3mu{\it nd}}\nolimits_0(V)),\ d^* a =0\} $$ on the space of irreducible connections modulo gauge ${\cal B}^*_X$. The space of irreducible ASD connections with respect to a K\"ahler metric is then an analytic subspace. Now the explicit formulas in \cite[Proposition 5.2.18]{D&K} for the forms representing $\mu({\rm Pd}(\omega))$ for an harmonic form $\omega \in H^2(X,{\Bbb C})$, show that $\mu$ preserves the Hodge structure. If we write formally $$ \phi_k({\rm Pd}(\omega_1), \ldots, {\rm Pd}(\omega_d)) = \int_{\M^{\rm asd}_k}\mu({\rm Pd}(\omega_1))\cdots \mu({\rm Pd}(\omega_d)), $$ then it is clear that $\phi_k(X) \ne 0$ only if the total Hodge type of $\omega_1,\ldots,\omega_d$ is $(d,d)$. The (probably inessential) problem is that it is not {\relax a priori} obvious whether integrating the form representatives over the non compact manifold ${\cal M}_k^{\rm asd}$ gives a valid way of computing $\phi_k$. \section{Proof of theorem~\protect\ref{main}.} We can now give proofs of the results stated in the introduction. The main theorem \ref{main} is the special case of theorem \ref{main'} below for $N = n{\bar\P}^2$ (see definition~\ref{good} for the definition of good connected sum decomposition). \pr@claim{\bf}{ \thetheorem}{\sl} Theorem main'. Let $X$ be a simply connected algebraic surface with $p_g >0$. Let $X_{\min}$ be its minimal model and let $K$ be its canonical divisor. Suppose $X$ admits a good smooth connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$ for a negative definite manifold~$N$. Then $H_2(N,{\Bbb Z})$ is generated by classes $e_1,\ldots,e_n$, with $e_i^2 = -1$ such that either $e_i$ is represented by a $(-1)$-curve, or $e_i \in \overline{\rmmath{NE}}(X_{\min})$ depending on whether $K\cdot e_i$ is negative or positive respectively. Here $\overline{\rmmath{NE}}(X_{\min})$ is the closure of the cone spanned by positive rational multiples of algebraic curves on the minimal model \proof. Choose a generator~$e$ of~$H_2(N)$. We first prove that $e$ is homologous to an algebraic cycle. Since $e$ is certainly integral, it suffices by the Lefschetz~$(1,1)$ theorem \cite[p. 163]{G&H} to prove that its Poincar\'e dual is of pure type~$(1,1)$. Choose $k$ sufficiently large as in proposition~\ref{(d,d)} (with $L= K$), and the definition \ref{good} of good. Then $\phi_k(X)$ is non trivial and has pure Hodge type~$(d,d)$. On the other hand we have $\phi_k = e \psi$. Since the number of Hodge types of~$e\psi$ is at least the number of Hodge types of~$e$, $e$ has to be of pure type as well. Since $e$ is a real class, it is then of type~$(1,1)$. To show that for the proper orientation $e$ lies on the closure of the full effective cone $\overline{\rmmath{NE}}(X)$, it is enough to show that $e\cdot H \ne 0$ for all ample divisors $H$. In fact, since the closure of the effective cone and the nef cone are in duality \cite[proposition 2.3]{Wilson:birational}, $e \in \pm \overline{\rmmath{NE}}(X)$ if and only if $e$ defines a strictly positive or strictly negative form on the ample cone. But since the ample cone is connected, it suffices to show that the form~$e$ has no sign change i.e. does not vanish on the ample cone. Since $e$ is a rational class we need to check this only for integral ample classes. Now for a fixed ample divisor~$H$ there is a $k_0 = k_0(H)$ such that $\phi_k(X)(H) \ne 0$ for $k > k_0$ \cite[th. 10.1.1]{D&K}. Since $e$ divides $\phi_k(X)$, it follows that $e\cdot H\ne 0$. By the orthogonality result \cite[th. 4.5.3]{F&M}, for every $(-1)$-sphere~$S$ in $X$ we have either $e\cdot S = 0$ or $e = \pm [S]$. Hence for the ``effective orientation'' of $e$ found above, $e$ is either homologous to a $(-1)$-curve, or $e$ is orthogonal to all $(-1)$-curves, i.e. $e \in H_2(X_{\min})$. In the first case $e\cdot K = -1 < 0$, in the latter case we have $e\cdot K = e\cdot K_{\min}> 0$ for as $p_g$ is positive, $K_{\min}$ is nef, and $K\cdot e \equiv e^2 \equiv 1 \pmod 2$. Since a divisor on the minimal model is effective if and only if its pullback to $X$ is effective, we have $\overline{\rmmath{NE}}(X_{\min})= \overline{\rmmath{NE}}(X) \cap H_2(X_{\min})$, and the result follows \endproof Theorem \ref{main'} gives the following technical refinement of corollary~\ref{maincor}, proving conjecture~\ref{consumconj} for the pair $(X,N)$ under an additional hypothesis. \pr@claim{\bf}{ \thetheorem}{\sl} Corollary maincor'. In addition to the assumptions of theorem \ref{main'}, suppose $X$ has a deformation~$Y$ with a minimal model~$Y_{\min}$ such that there are no classes $C\in\overline{\rmmath{NE}}(Y_{\min})$ with $C^2 = -1$ dividing all Donaldson polynomials $q_{L,k}$ with $L\cdot C$ odd and $k \gg 0$. Then $H^2(N)$ is generated by $(-1)$-curves. In particular this is true if \itm{(a)} the linear system $|K_{Y_{\min}}|$ contains a smooth irreducible curve of genus at least 2, and either $p_g$ is even or $K_{\min}^2 \not\equiv 7 \pmod 8$, or \itm{(b)} the Picard number $\rho(Y_{\min}) = 1$. As mentioned in the introduction this corollary has already been proved without the goodness condition under the assumptions $X_{\min}$ is spin or $X_{\min}$ has big monodromy \cite[cor 5.4]{F&M}. \proof. Since the deformations of a surface are all oriented diffeomorphic, we conclude that if $X$ admits a good connected sum decomposition $X\buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$, so does its deformation~$Y$. Moreover, the subgroup generated by $(-1)$-curves is stable under deformation by \cite[IV.3.1]{BPV}. Hence if $H_2(N)\subset H_2(Y)$ is generated by $(-1)$-curves, so is $H_2(N) \subset H_2(X)$. Thus we can assume $X = Y$. By definition~\ref{good} of a good decomposition, it is clear that no generator of $H_2(N)$ can be in $\overline{\rmmath{NE}}(X_{\min})$. Hence by theorem~\ref{main'}, $H^2(N)$ is generated by $(-1)$-curves. It remains to see that the extra condition is satisfied in the given special cases. In case (b), $\rmmath{NS}(X_{\min})$ is positive definite. For case (a) we argue by contradiction. Suppose there is a class $C\in\overline{\rmmath{NE}}(X_{\min})$, $C^2 = -1$ and $C$ divides $q_{C,k}$ for all $k \gg 0$. Since $C$ is orthogonal to all $(-1)$-curves, the proof of theorem \cite[th. 4.8]{Donaldson:pol},\,\cite[th. 9.3.14]{D&K} gives that $$ q_{C,k}(X)|_{H_2(X_{\min})} = \pm q_{C,k}(X_{\min}). $$ Since $C \in \rmmath{NS}(X_{\min})$, Morgan's comparison formula~\ref{nupol}, and O'Grady's non triviality result \cite[cor. 2.4, th.2.4]{OGrady}, give that for every $\omega \in H^0(K_{\min})$, which vanishes on a smooth irreducible curve of genus $g \ge 2$ we have $$ q_{C,k}(X_{\min})({\rm Pd}(\omega + \bar \omega)) \ne 0 $$ if $4k - 3(1+p_g)$ is even and $k\gg 0$ with ${\numfrac12}(C^2-C\cdot K) - {\numfrac14}(C^2 + 4k)$ odd. (Strictly speaking O'Grady uses a slightly different polynomial defined on $C^\perp \subset H_2(X)$, but it is easy to see that on $C^\perp$, $q_{C,k}$ coincides with his polynomial). Since $4k \equiv -C^2 \pmod 4$, and $\<C,\omega> =0$ this contradicts the divisibility of $q_{C,k}$ by $C$ if $p_g$ is even. If $p_g$ is odd, the same argument gives a contradiction if there is a polynomial $q_{L,k}$ with $L\in \rmmath{NS}(X_{\min})$, $L \cdot C \equiv 1$, and $L^2\equiv K_{\min}\cdot L \equiv 0$. This an affine equation for $L \pmod 2$ in $\rmmath{NS}(X_{\min}) \tensor {\Bbb Z}/2{\Bbb Z}$, so it has a solution if $C \not\equiv K_{\min} \pmod 2$. But if $C \equiv K_{\min}$, then $C^2 = -1 \equiv K_{\min}^2 \pmod 8$ contrary to assumption. \endproof \pr@claim{\sl}{ \thetheorem}{\relax}Remark *. If $|K_{\min}|$ contains a smooth irreducible curve but $p_g$ is odd and $K_{\min}^2\equiv 7 \pmod 8$ the proof above shows that there is up to orientation at most one generator $e_0$ of $H^2(N)$ which is not homologous to a $(-1)$-curve. Hence all $(-2)$-spheres in $H_2(X_{\min})$ are orthogonal to $e_0$, because the reflections they generate are represented by diffeomorphisms. We also get that if $e_0$ exists, $w_2(X)$ is represented by the sum of the generators of $H_2(N)$, hence $X'$ is spin. \pr@claim{\sl}{ \thetheorem}{\relax}Remark *. It follows from results in \cite{RB:e-exp} that if $C$ is orthogonal to all $(-1)$-curves, then the divisibility of say $\phi_k(X)$ by $C$ implies the divisibility of $\phi_k(X')$ by $C$ for all blow-downs of $X'$ intermediate between $X$ and $X_{\min}$. Hence the extra condition in corollary \ref{maincor'} gets stronger as we blow-up more points. It is interesting to see how the minimality of a surface plays a role here. For a non minimal surface every curve in $|K|$ is reducible since it contains an exceptional divisor, unless $X$ is a K3-surface blown-up once. Indeed in the non minimal case $q_{E,k}$ is divisible by the exceptional divisor $E$, and so O'Grady's theorem could not possibly be true. It would be very interesting if O'Grady's results would be true assuming the existence of an irreducible curve in $|K|$, or stretching things even further, a smooth irreducible curve in a pluricanonical system $|nK|$. \ifcomment\bgroup\par\medskip\noindent\small Suppose that $X$ is of general type and minimal. Then say $|13K|$ contains a smooth curve of genus at least 192, and we would still have a $\theta$-characteristic that extends over the whole surface, which seems to be one of the main points of O'Grady's construction. Since for elliptic surfaces the $(-1)$-curve conjecture~\ref{(-1)conj} is already known, such a result would confirm the conjecture for all simply connected surfaces with $p_g >0$ except those of general type with $p_g$ odd and $K_{\min}^2\equiv 7 \pmod 8$. \par\medskip\noindent\egroup\fi \section{Examples.} We give two examples of the use of corollary~\ref{maincor'}. The first example follows basically by leafing through \cite{BPV}. For the other we use Noether-Lefschetz theory to reprove and generalise Friedman and Morgan's result that for complete intersections with $p_g >0$ conjecture~\ref{(-1)conj} is true. Since one approach to Noether-Lefschetz theory is through monodromy groups, it is not surprising that this approach gives results similar to those using big monodromy. However the proof may be interesting because the Noether Lefschetz theorems we will use are proved using the ``infinitesimal method'', based on Hodge and deformation theory. \pr@claim{\bf}{ \thetheorem}{\sl} Proposition examples. Suppose $X$ is a smooth simply connected surface admitting a good connected sum decomposition $X \buildrel\scriptscriptstyle \rm diff\over\iso X' \# N$. Then $H_2(N)$ is generated by $(-1)$-curves if \itm(a) $X$ is the blow-up of a Horikawa surface with $K^2$ even (see \cite[table~10]{BPV}), or \itm(b) $(Y,{\cal O}(1))$ is a simply connected projective local complete intersection of dimension $r+2$ with Picard number $\rho =1$, $V$ is an ample $r$-bundle, and $X$ is the blow up of the smooth zero locus $X_a$ of a section in $V(a)$ with $a \gg 0$. \proof. Case (a). By \cite[th. 10.1, remark VII.10.1]{BPV} and Bertini's theorem, a Horikawa surface with $K^2$ even is simply connected and the linear system $|K|$ contains a smooth curve. \smallskip\noindent Case (b). For $r=0$ the proposition is a special case of corollary~\ref{maincor'} so we assume $r >0$. Choose $a$ so large that $V(a)$ is globally generated. By an application of the vector bundle version of the Lefschetz hyperplane theorem \cite{Sommese&VdV}, \cite[cor. 22]{Okonek:Barth-Lefschetz}, $X_a$ is simply connected \cite[p.158]{G&H}. To prove that $p_g(X_a) >0$ consider the sequence $$ 0 \to I_X \det(V(a))\tensor {\cal O}_Y(K_Y) \to \det( V(a))\tensor {\cal O}_Y(K_Y) \to {\cal O}_X(K_X) \to 0. $$ Now choose $a$ so large that $\det V\tensor {\cal O}_Y(K_Y) \tensor {\cal O}_Y(ra)$ has a section non-vanishing on~$X$. Finally, if $a$ is sufficiently large then $\rho(X_a) = 1$ for the general section by Ein's generalization of the Noether-Lefschetz theorem in ${\Bbb P}^3$ to ample vector bundles on projective varieties \cite[th. 2.4]{Ein}. \endproof \pr@claim{\sl}{ \thetheorem}{\relax}Remark *. Suppose that in case (b), $Y$ is smooth, and $V = \mathop\oplus_{i=1}^r {\cal O}(d_i)$. We can then be more precise since there is no need to twist up. It suffices that $V$ and $\det(V)\tensor {\cal O}_Y(K_Y)$ are spanned by global sections, $H^{1,1}(\det(V))= H^{1,1}(V\tensor \det(V)) = 0$, and $$ H^0(V) \tensor H^0(\det(V)\tensor {\cal O}_Y(K_Y)) \to H^0(V\tensor \det(V)\tensor{\cal O}_Y(K_Y))) $$ is surjective (e.g. if the $d_i \gg 0$ or $Y ={\Bbb P}^n$, with the exception of $n=3$, $V = {\cal O}(2)$ or ${\cal O}(3)$, and $n=4$, $V ={\cal O}(2) \mathop\oplus {\cal O}(2)$). This follows from judicially checking the cohomological conditions \cite[lemma 3.2.1,3.2.2,3.2.3]{Spandaw:thesis} using Kodaira-Nakano vanishing. For ${\Bbb P}^n$ the statement follows from the classical Noether Lefschetz theorem. Also note that by choosing $a$ sufficiently large we can make $K_{X_a}$ very ample, and so we can find a smooth irreducible curve of genus at least 2 in its linear system. Hence case (b) follows directly from corollary~\ref{maincor'} if $p_g(X_a)$ or $K^2_{X_a}$ are even.
1995-03-23T06:20:39
9503
alg-geom/9503014
en
https://arxiv.org/abs/alg-geom/9503014
[ "alg-geom", "math.AG" ]
alg-geom/9503014
Christian Gantz
Christian Gantz and Brian Steer
Stable Parabolic Bundles over Elliptic Surfaces and over Orbifold Riemann Surfaces
12 pages, LaTeX
null
null
null
null
For an elliptic surface $q:Y \to \Sigma$, with prescribed singular fibres, Stefan Bauer proved directly via algebraic geometry that the stable bundles over $Y$, whose chern classes are pull backs from $\Sigma$, correspond to the stable (V-)bundles over $\Sigma$. We show, via a short proof in differential geometry, a generalisation to stable parabolic bundles. This uses extensions of Donaldson's deep result, giving the existence of Hermitian-Yang-Mills (or anti-self-dual) connections on stable parabolic bundles. In our cases these connections are flat and hence, correspond to representations of certain fundamental groups, which in turn are isomorphic, by Ue's work. To generalize Bauer's equivalence of the corresponding moduli spaces of stable bundles, we combine his arguments with Kronheimer & Mrowka's construction of the moduli spaces of stable parabolic bundles. Finally, we consider the pulling back of smooth parabolic bundles via $q$.
[ { "version": "v1", "created": "Wed, 22 Mar 1995 15:21:07 GMT" } ]
2008-02-03T00:00:00
[ [ "Gantz", "Christian", "" ], [ "Steer", "Brian", "" ] ]
alg-geom
\section{Introduction} If $q:Y \rightarrow \Sigma$ is an elliptic surface (to be made precise) then the induced map of fundamental groups is an isomorphism if we consider $\Sigma$ as an orbifold, \cite{ue}, \cite{dol}. Hence, we obtain a correspondence of flat bundles (by bundles we always mean complex vector bundles). Donaldson showed that each stable degree zero bundle over $Y$ or $\Sigma$ admits a unique Hermitian-Yang-Mills (or anti-self-dual) connection. In fact, the obvious conditions on a bundle $E' \rightarrow Y$ to come from $\Sigma$, namely $\mbox{$\cal C$}_{1}(E') \in q^{*}\mbox{$\Lambda$}^{2}(\Sigma)$ and $\mbox{$\cal C$}_{2}(E')=0$, imply that this H.Y.M. connection is flat. So, there is a correspondence of stable degree zero bundles over $Y$ and $\Sigma$. The generalisation to all degrees has been shown by Bauer, \cite{bau}, via a direct proof in algebraic geometry. Donaldson's result has been extended by several authors to parabolic bundles, theorem \ref{b}. This and the use of extension results for flat bundles, theorem \ref{y}, gives an identification of stable parabolic degree zero bundles over $(\Sigma,P)$ and $(Y,P')$, where $P$ is a finite collection of generic points in $\Sigma$ and $P'=q^{-1}(P)$, with representations of the fundamental groups of $\Sigma-P$ and $Y-P'$, respectively. These groups are again isomorphic and so, we have a correspondence of stable parabolic degree zero bundles over $(Y,P')$ and over $(\Sigma,P)$. This extends readily to all degrees by tensoring with a parabolic line bundle. Another proof for genuine bundles of any degree, not using parabolic bundles at all, relies on the correspondence between stable bundles over $\Sigma$ ($Y$) and representations of the fundamental group of circle bundles (i.e. Seifert fibred spaces) over $\Sigma$ ($Y$), \cite[p 63]{fas}, \cite{bao}. These groups are also isomorphic, see the proof of proposition \ref{v}. Combining Bauer's arguments with Kronheimer \& Mrowka's description of the moduli spaces of stable parabolic bundles, we show that these are complex manifolds if we fix determinants and if they are pull backs from $\Sigma$. Finally, we consider smooth parabolic bundles and produce some details about the correspondence of stable ones. After recalling elliptic surfaces and parabolic bundles in the following two sections we take one section to state the results. More details of our work can be found in \cite{gan}. \hfill \noindent {\bf Acknowledgements} \vspace{0.2cm} We are grateful to P. Kronheimer, S. Bauer, M. L\"ubke, T. Peternell, C. Okonek, S. Agnihotri and R. Plantiko for helpful remarks. \section{Elliptic surfaces} Throughout, let $q:Y \rightarrow \Sigma$ be an elliptic surface, i.e. $Y$ is a compact complex surface, $\Sigma$ a compact Riemann surface and $q^{-1}(\sigma)$ an elliptic curve for generic, i.e. all but finitely many, $\sigma \in \Sigma$, c.f. \cite{gri}, \cite{bpv}. We will assume that any non-generic fibre is either a rational curve of multiplicity one with one self-intersection (called singular fibre) or a multiple elliptic curve and furthermore, that there is at least one singular fibre. Moishezon shows that all elliptic surfaces are deformation equivalent to the ones we consider, \cite{moi}. \begin{theo}[Ue, Dolgachev] \showlabel{k} If $q:Y \rightarrow \Sigma$ has singular fibres and if $U_{0} \subseteq \Sigma$ is an open ball such that $\pi^{-1}(U_{0})$ contains all singular fibres but no multiple elliptic curves then $\pi^{-1}(U_{0})$ is simply connected. \end{theo} If $Y_{\sigma}:=q^{-1}(\sigma)$ has multiplicity $m>1$, let $\tilde{U}$ and $B$ be open discs in $\C$, $\phi:B \rightarrow U$ a chart with $U \subseteq \Sigma$ and $\phi(0)=\sigma$ and construct a uniformization of $U$ by \[ \threehorss{\tilde{U}}{B}{U}{z^{m}}{\phi} \] where $\langle \eta=e^{2 \pi i/m} \rangle = \Z_{m} \subseteq \C$ acts on $\tilde{U}$ in the standard way. After this construction is done for all multiple points $\sigma \in \Sigma$ we think of $\Sigma$ as an orbifold, cf. \cite{fas}. Bauer points out that for all orbifold Riemann surfaces there exists an elliptic surface over it. The natural gauge-theoretic objects on orbifolds are V-bundles: a (local, complex) rank $r$ V-bundle $E|_{U}$ is isomorphic to \[ (\tilde{U} \times \C^{r},\Z_{m}) \rightarrow (\tilde{U},\Z_{m}) \,\,\,\,\,\,\, \mbox{with}\] \[ \eta (\tilde{u},z_{1},...,z_{r})=( \eta \tilde{u}, \eta^{a_{1}} z_{1},..., \eta^{ a_{r}} z_{r}) \] for some isotropies $(a_{1},...,a_{r}) \in \{ 0,...,m-1 \}^{r}$. \begin{theo}[Furuta \& Steer, Seifert] \showlabel{i} Smooth V-bundles over $\Sigma$ are classified by rank, degree (which is rational) and isotropies. \end{theo} For any $y \in Y_{\sigma}$ we can choose coordinates $(z_{1},z_{2})$ on $U' \ni y$ such that $\phi^{-1} \circ q (z_{1},z_{2})=z^{m}_{2}$. Hence we can lift $\phi^{-1} \circ q$ locally to a regular map \[ \sqrt[m]{\phi^{-1} \circ q}=z_{2}:U' \rightarrow \tilde{U} \] uniquely up to the action of $\Z_{m}$. In particular, $q$ is a map of orbifolds. A divisor on $\Sigma$ can be represented by a finite sum $D=\sum_{i \in I} \sigma_{i}n_{i}/m_{i}$ where $n_{i} \in \Z$ and $m_{i}$ the multiplicity of $\sigma_{i} \in \Sigma$. The vertical divisors on $Y$ are precisely the pull backs of divisors on $\Sigma$. Hence, \cite{nas}, the line bundles $\mbox{$\cal O$} (D')$ over $Y$ with vertical divisor correspond to the line V-bundles over $\Sigma$. As $q$ is regular away from finitely many points theorem \ref{k} implies that $q_{*}:\pi_{1}(Y) \rightarrow \pi_{1}^{V}(\Sigma)$ is an isomorphism. Here, the orbifold fundamental group $\pi_{1}^{V}(\Sigma)$ is an extension of $\pi_{1}(\Sigma)$ by a torsion group with unipotent elements corresponding to the multiple points. The isomorphism implies that the first betti number of $Y$ is even and therefore $Y$ is Kahler, by \cite{miy}. The correspondence between representations of the fundamental group and flat bundles on manifolds extends to orbifolds via the construction with simply connected coverings. Hence, the flat bundles over $Y$ correspond to the flat V-bundles over $\Sigma$. \section{Parabolic bundles} Let $P:=\{ p_{1},...,p_{n} \}$ be a collection of generic points in $\Sigma$ and put $P':= \{ P_{j}':=q^{-1}(p_{j}) \}_{j=1...n} \subseteq Y$. A parabolic V-bundle $E$ over the pair $(\Sigma,P)$ is a V-bundle $E \rightarrow \Sigma$ with proper filtrations \[ E|_{P_{j}}=E_{j,1} \supset E_{j,2} \supset ... \supset E_{j,l_{j}} \neq 0 \] and weights \[ 0 \leq \alpha_{j,1} < \alpha_{j,2} < ... < \alpha_{j,l_{j}} <1 \] for all $j=1,...,n$. We call $\mu_{j,k}:=\mbox{rank}\, (E_{j,k}/E_{j,k+1})$ the multiplicity of $\alpha_{j,k}$. Let $\alpha_{j}$ be the diagonal matrix of rank equal to $\mbox{rank}\, E$ and with entries $\alpha_{j,k}$ ($k=1,...,l_{j}$) with multiplicities. Then put $\alpha(E):=\{ \alpha_{1},...,\alpha_{n} \}$. We will sometimes write $|E|$ for the underlying genuine V-bundle. Correspondingly, one defines parabolic bundles $E'\rightarrow (Y,P')$ where the $\{E'_{j,k} \}$ are bundles over $P_{j}'$ with weights $\{ \alpha'_{j,k}\}$ of multiplicities $\{ \mu'_{j,k} \}$, encoded in $\alpha'(E')$. If the bundle and the filtrations are holomorphic we speak of a holomorphic parabolic bundle. Thinking of vectors in $E|_{P}$ as having (the obvious) weights ($+\infty$ being the weight of zero vectors), a morphism of parabolic bundles is a bundle map not decreasing the weight of any vector. In a direct sum of parabolic bundles the weight of a vector is the minimum of the weights of it's projections. Let us fix a Kahler metric on $Y$ with (1,1)-form $\omega'$. When we use ordinary bundle invariants and operators (like $\mbox{$\cal C$}_{i}$, $\det$ or $\deg$) on parabolic bundles we mean to apply them to the underlying bundles. Then we have \[ \mbox{par}\, \mbox{$\cal C$}_{1} (E):= \mbox{$\cal C$}_{1}(E) + \sum_{j=1}^{n} \mbox{Tr\,} (\alpha_{j}) \mbox{PD}\, (p_{j} ) \in \mbox{$\Lambda$}^{2}(\Sigma,\R) \] and similarly, $\mbox{par}\, \mbox{$\cal C$}_{1}(E') \in \mbox{$\Lambda$}^{2}(Y,\R)$. Let $\mbox{par}\, \deg E:= \langle \mbox{par}\, \mbox{$\cal C$}_{1}(E), \Sigma \rangle \in \R$ and $\mbox{par}\, \deg E':=\langle \mbox{par}\, \mbox{$\cal C$}_{1}(E') \cup \omega',Y \rangle \in \R$. We also have (since $P \cdot P=0$) \[ \mbox{par}\, \mbox{$\cal C$}_{2}(E'):=\mbox{$\cal C$}_{2}(E') + 2 \sum_{j=1}^{n} \sum_{k=1}^{l_{j}} \alpha'_{j,k} \mbox{PD}\,(d_{j,k}') +\frac{1}{2} \mbox{par}\, \mbox{$\cal C$}_{1}^{2}(E') \in \mbox{$\Lambda$}^{4}(Y,\R) \] where $d'_{j,k}:= \deg (E'_{j,k}/E'_{j,k+1})$. \begin{defi} A holomorphic parabolic V-bundle $\mbox{$\cal E$} \rightarrow (\Sigma,P)$ is called {\bf stable} if for all non-zero parabolic maps $\mbox{$\cal F$} \rightarrow \mbox{$\cal E$}$, injective over some point in $\Sigma$ and with $\mbox{rank}\, \mbox{$\cal F$} < \mbox{rank}\, \mbox{$\cal E$}$ we have \[\mbox{deg}\, \mbox{$\cal F$} / \mbox{rank}\, \mbox{$\cal F$} < \mbox{deg}\, \mbox{$\cal E$} / \mbox{rank}\, \mbox{$\cal E$}. \] Similarly for $\mbox{$\cal E$}' \rightarrow (Y,P')$. \end{defi} \section{Statement of results} Being a map of orbifolds, $q:Y \rightarrow \Sigma$ induces pull backs of holomorphic parabolic V-bundles over $\Sigma$ to holomorphic parabolic bundles over $Y$. Our main result is: \begin{theo} \showlabel{z} Pulling back induces a correspondence between stable parabo\-lic V-bundles $\mbox{$\cal E$} \rightarrow (\Sigma,P)$ and those stable parabolic bundles $\mbox{$\cal E$}' \rightarrow (Y,P')$ with $\mbox{par}\, \mbox{$\cal C$}_{2}(\mbox{$\cal E$}')=0$ and $\mbox{par}\, \mbox{$\cal C$}_{1}(\mbox{$\cal E$}') \in q^{*} \mbox{$\Lambda$}^{2}(\Sigma,\R)$. \end{theo} In particular: \begin{theo} \showlabel{a} Pulling back induces a bijection between stable V-bundles $\mbox{$\cal E$}$ over $\Sigma$ and those stable bundles $\mbox{$\cal E$}'$ over $Y$ with $\mbox{$\cal C$}_{2}(\mbox{$\cal E$}')=0$ and $\det \mbox{$\cal E$}' = \mbox{$\cal O$} (D')$ for a vertical divisor $D'$. \end{theo} Theorem \ref{a} has been shown (under some assumptions on the Kahler metric of $Y$, \cite[p 511]{bau}) by Bauer first, using algebraic geometry. Our goal is a differential-geometric proof of theorem \ref{z} using the correspondence of stable parabolic bundles and parabolic H.Y.M. connections. Let $\mbox{$\cal E$} \rightarrow (\Sigma,P)$ be a stable parabolic bundle, $\mbox{$\cal E$}':=q^{*} \mbox{$\cal E$}$, $\mbox{$\cal T$}:=\,\, \mbox{Par End}_{0} \mbox{$\cal E$}$ and $\mbox{$\cal T$}':=q^{*} \mbox{$\cal T$} = \,\, \mbox{Par End}_{0} \mbox{$\cal E$}'$. The important thing to note is that $\mbox{$\cal T$}$ and $\mbox{$\cal T$}'$ are holomorphic (V-)bundles. We have the deformation complexes (of smooth sections) \begin{center} \setlength{\unitlength}{1.5mm} \begin{picture}(0,30)(0,-15) \putcc{-26}{10}{\Omega_{\Sigma}^{0} (\mbox{$\cal T$})} \putcc{0}{10}{\Omega_{\Sigma}^{0,1} (\mbox{$\cal T$})} \putcc{26}{10}{0} \putcc{-26}{-10}{\Omega_{Y}^{0} (\mbox{$\cal T$}')} \putcc{0}{-10}{\Omega_{Y}^{0,1} (\mbox{$\cal T$}')} \putcc{26}{-10}{\Omega_{Y}^{0,2} (\mbox{$\cal T$}')} \put(-26,6){\vector(0,-1){12}} \put(0,6){\vector(0,-1){12}} \put(26,6){\vector(0,-1){12}} \put(-18,10){\vector(1,0){9}} \putbc{-15}{12}{\mbox{$\bar{\partial}$}_{\cal{T}}} \put(-18,-10){\vector(1,0){8}} \putbc{-13}{-8}{\mbox{$\bar{\partial}$}_{\cal{T}'}} \put(9,10){\vector(1,0){10}} \putbc{13}{12}{\mbox{$\bar{\partial}$}_{\cal{T}}} \put(8,-10){\vector(1,0){9}} \putbc{13}{-8}{\mbox{$\bar{\partial}$}_{\cal{T}'}} \putcl{2}{0}{q^{*}} \end{picture} \end{center} Let $E:=_{C^{\infty}} \mbox{$\cal E$}$, $\mbox{$\cal O$}(D)=\det \mbox{$\cal E$}$ and $\mbox{$\cal M$}(E,D)$ be the space of stable structures on $E$ with determinant $\mbox{$\cal O$}(D)$. Similarly, define $\mbox{$\cal M$}(E', D')$ for $E':=q^{*}E$ and $D':=q^{*}D$. Using the extension of standard deformation theory, \cite{dak}, to parabolic bundles makes $\mbox{$\cal M$}(E',D')$ into a Hausdorff complex space, see \cite[p 83]{km2} and \cite{mun}, with a description near $\mbox{$\cal E$}'$ given by the zero set of a holomorphic map \[ \mbox{$\Lambda$}^{0,1}_{\bar{\partial}}(Y,\mbox{$\cal T$}') \rightarrow \mbox{$\Lambda$}^{0,2}_{\bar{\partial}}(Y,\mbox{$\cal T$}'). \] By the vanishing of the second cohomology over $\Sigma$, $\mbox{$\cal M$}(E,D)$ is even a complex manifold with a chart near $\mbox{$\cal E$}$ given by $\mbox{$\Lambda$}^{0,1}_{\bar{\partial}}(\Sigma,\mbox{$\cal T$})$. \begin{prop}[Bauer] \showlabel{t} Stability of $\mbox{$\cal E$}$ implies that \[ q^{*}:\mbox{$\Lambda$}^{0,1}_{\bar{\partial}}(\Sigma,\mbox{$\cal T$}) \rightarrow \mbox{$\Lambda$}^{0,1}_{\bar{\partial}}(Y,\mbox{$\cal T$}') \] is an isomorphism. \end{prop} This proposition and theorem \ref{z} imply: \begin{coro} If $\mbox{$\cal M$}_{\cal{E}}(E,D)$ is the connected component of $\mbox{$\cal E$}$, correspondingly for $\mbox{$\cal E$}'$, then \[ q^{*}:\mbox{$\cal M$}_{\cal{E}}(E,D) \rightarrow \mbox{$\cal M$}_{\cal{E}'}(E',D') \] is an isomorphism of complex manifolds. \end{coro} \begin{prop} \showlabel{u} If $E' \rightarrow (Y,P')$ is a rank $r$ parabolic bundle satisfying $\mbox{par}\, \mbox{$\cal C$}_{2} (E')=0$, $\mbox{par}\, \mbox{$\cal C$}_{1} (E') \in q^{*} \mbox{$\Lambda$}^{2}(\Sigma,\R)$ and if the space $\mbox{$\cal M$}(E')$ of stable structures on $E'$ is non-empty then \begin{description} \item[(i)] there exists a unique line V-bundle $L \rightarrow \Sigma$ with $q^{*} L=\det E'$. Also, $E'$ is uniquely determined by $L$ and it's weights $\alpha'$; \item[(ii)] we have \[ \mbox{$\cal M$}(E') = \bigsqcup_{a} \mbox{$\cal M$} (L,r,a,\alpha') \] where the union is over isotropies $a \in \Z_{m_{1}} \times ... \times \Z_{m_{k}}$ ($k=\sharp \{ $ marked points on $\Sigma \}$) compatible with $L$ and $\mbox{$\cal M$}(L,r,a,\alpha')$ is the space of stable structures on the unique rank $r$ parabolic V-bundle over $(\Sigma,P)$ with determinant $L$, isotropies $a$ and weights $\alpha'$. \end{description} \end{prop} \section{Stable parabolic bundles} \begin{defi} A {\bf (parabolic) hermitian metric} $h'$ on a parabolic bundle $E'\rightarrow (Y,P')$ (similarly for $E\rightarrow (\Sigma,P)$) is a hermitian metric on $E'|_{Y-P'}$ such that \[ \forall y \in P'_{j} \,\,\,\,\,\,\,\, E'_{j,k}(y)=\{ s(y) \,\, | \,\, s \in \Gamma_{loc}(E') \mbox{ s.t. }\, h'(s(-))=O(\mbox{d} (P'_{j},-)^{\alpha'_{j,k}}) \}. \] \end{defi} For a holomorphic parabolic bundle $\mbox{$\cal E$}'$, $h'$ induces a Chern connection on $\mbox{$\cal E$}'|_{Y=P'}$. So, one can talk of H.Y.M.-connections. Around $P'_{j}$, a parabolic connection has holonomy conjugated to $\alpha'_{j}$, \cite{kro1}, and can therefore not be extended over $P'_{j}$, in general. A parabolic connection is called reducible if the parabolic bundle decomposes together with the connection. The bridge between algebraic and differential geometry is \begin{theo}[\cite{mas}, \cite{saw}, \cite{mun}, \cite{b93}, \cite{nas}] \showlabel{b} Any degree zero holomorphic parabolic bundle over a complex Kahler surface or orbifold of dimension one is stable if and only if it admits an irreducible H.Y.M. metric, unique up to isomorphism. Furthermore, an H.Y.M. connection has finite action. \end{theo} The primary result, for genuine bundles, is due to Donaldson; Munari's proof relies on Simpson's work. A H.Y.M. connection over $\Sigma$ is obviously flat. \begin{lemm} \showlabel{c} If $\mbox{$\cal E$}' \rightarrow (Y,P')$ has $\mbox{par}\, \mbox{$\cal C$}_{2}(\mbox{$\cal E$}')=0$ and $\mbox{par}\, \mbox{$\cal C$}_{1}(\mbox{$\cal E$}') \in q^{*} \mbox{$\Lambda$}^{2}(\Sigma,\R)$ then a H.Y.M. connection is neccessarily flat. \end{lemm} \noindent{\em Proof:} By \cite[p 100]{mun} or \cite{gan} we obtain the second (in particular the existence of) and, from \cite{dak}, the third equality in \[0=8 \pi^{2} (\mbox{par}\, \mbox{$\cal C$}_{2}(\mbox{$\cal E$}') - \frac{1}{2} \mbox{par}\, \mbox{$\cal C$}_{1}^{2}(\mbox{$\cal E$}'))= \int_{Y-P'} \mbox{Tr\,} (F^{2}) = ||F^{-}||^{2} - ||F^{+}||^{2}. \] $\Box$ \hfill\vspace{0.5cm} \begin{theo}[Munari, Biquard, Simpson] \showlabel{y} A holomorphic bundle over $Y-P'$ ($\Sigma-P$) with a flat hermitian metric extends uniquely (up to isomorphism) to a holomorphic parabolic bundle over $(Y,P)$ ($(\Sigma,P)$) such that the hermitian metric becomes a parabolic metric. \end{theo} Now we prove theorem \ref{z}: \noindent{\em Proof:} At first, we treat the special case of degree zero bundles. If $\mbox{$\cal E$} \rightarrow (\Sigma,P)$ is stable and of degree zero, let $A$ be the unique irreducible flat parabolic connection on $\mbox{$\cal E$}$. By the uniqueness in theorem \ref{y}, $A|_{\Sigma-P}$ is still irreducible. The regularity of $q|_{Y-P'}$ away from finitely many points and theorem \ref{k} imply that $(q|_{Y-P'})_{*}:\pi_{1}(Y-P') \rightarrow \pi_{1}^{V}(\Sigma-P)$ is an isomorphism. Hence, if $\mbox{$\cal E$}':=q^{*} \mbox{$\cal E$}$ and $A':=q^{*}A$ then $A'$ is an irreducible flat parabolic connection and hence $\mbox{$\cal E$}'$ is stable. Conversely, if $\mbox{$\cal E$}' \rightarrow (Y,P')$ is stable and of degree zero, let $A'$ be the unique parabolic H.Y.M. connection, which is flat by lemma \ref{c} and $A'|_{Y-P'}$ is irreducible. We push forward $(\mbox{$\cal E$}',A')|_{Y-P'}$ to $\Sigma-P$. By theorem \ref{y}, this push forward extends uniquely to a holomorphic parabolic bundle with irreducible parabolic connection $(\mbox{$\cal E$},A) \rightarrow (\Sigma,P)$. As $q^{*}(\mbox{$\cal E$},A)$ and $(\mbox{$\cal E$}',A')$ are isomorphic over $Y-P'$ the uniqueness of theorem \ref{y} implies that they are isomorphic over $(Y,P')$. We are finished with the degree zero case. To show the general case, fix a generic point $p \in \Sigma-P$ and let $\Delta:=\int_{q^{-1}(p)} \omega'$. For each $d \in \R$ let $[d]$ be it's integer part, let $[p]$ be the holomorphic line bundle with divisor $p$ and define the holomorphic parabolic line bundle $\mbox{$\cal L$}_{d} \rightarrow (\Sigma,p)$ to be $\mbox{$\cal L$}_{d}:=[p]^{[d]}$ with weight $d-[d]$ over $p$. Let $\mbox{$\cal L$}'_{d}:=q^{*} \mbox{$\cal L$}_{d}$ be the parabolic pull back. We have $\mbox{par}\, \deg \mbox{$\cal L$}_{d}=d$ and $\mbox{par}\, \deg \mbox{$\cal L$}'_{d}=\Delta d$. Tensoring any rank $r$ stable parabolic bundle $\mbox{$\cal E$} \rightarrow (\Sigma,P)$ ($\mbox{$\cal E$}' \rightarrow (Y,P')$) of degree $-dr$ ($-\Delta d r$) with $\mbox{$\cal L$}_{d}$ ($\mbox{$\cal L$}'_{d}$) gives a stable parabolic bundle of degree zero over $(\Sigma,P \cup \{ p\} )$ ($(Y,P' \cup \{ q^{-1}(p) \} )$). Furthermore, we have $\mbox{par}\, \mbox{$\cal C$}_{2}(\mbox{$\cal E$}' \otimes \mbox{$\cal L$}'_{d})=0$. So we are reduced to the degree zero case. $\Box$ \hfill\vspace{0.5cm} We use Bauer's arguments, \cite[p 514]{bau}, to prove proposition \ref{t}: \noindent{\em Proof:} The sheaf $\mbox{$\cal T$}$ can be considered as a sheaf on $\Sigma$ or on $|\Sigma|$, the underlying Riemann surface of the orbifold $\Sigma$, see the correspondence between V-bundles and parabolic bundles in section 5 of \cite{fas}. As sheaves, $q_{*} \mbox{$\cal T$}' = \mbox{$\cal T$}$. The Leray spectral sequence induces an exact sequence, \cite[p 10]{bpv}, \[ 0 \rightarrow \threehorsb{\mbox{$\Lambda$}^{1}(\Sigma,\mbox{$\cal T$})}% {\mbox{$\Lambda$}^{1}(Y,\mbox{$\cal T$}')}{\mbox{$\Lambda$}^{0}(\Sigma,q_{*1}\mbox{$\cal T$}')}{q^{*}} \rightarrow ... \] where $q_{*1} \mbox{$\cal T$}'$ is the first direct image sheaf of $\mbox{$\cal T$}'$. It suffices to see that $\mbox{$\Lambda$}^{0}(\Sigma,q_{*1}\mbox{$\cal T$}')=0$. Relative duality, \cite[p 99]{bpv}, gives \[ q_{*1}\mbox{$\cal T$}' = q_{*}(\mbox{$\cal T$}'^{*} \otimes \mbox{$\cal K$}_{Y} \otimes q^{*}\mbox{$\cal K$}^{*}_{|\Sigma |})^{*}= \mbox{$\cal T$} \otimes q_{*}(\mbox{$\cal K$}_{Y} \otimes q^{*} \mbox{$\cal K$}^{*}_{|\Sigma |})^{*} \] where $\mbox{$\cal K$}_{|\Sigma |}$ is the canonical bundle. By \cite[p 98, 161-162]{bpv}, \[ \mbox{$\cal K$}_{Y} \otimes q^{*} \mbox{$\cal K$}^{*}_{|\Sigma |} = q^{*} q_{*1} \mbox{$\cal O$}_{Y}^{*} \otimes \mbox{$\cal O$}_{Y}(\sum (m_{i}-1)Y_{\sigma_{i}}) \] where the sum is over the singular points $\sigma_{i}$ of $\Sigma$, $q_{*1}\mbox{$\cal O$}_{Y}$ is locally free of rank one since all the other sheaves in this identity are, and \[ \deg (q_{*1} \mbox{$\cal O$}_{Y})=- \chi (\mbox{$\cal O$}_{Y}). \] In particular, $\mbox{$\cal K$}_{Y} \cdot \mbox{$\cal K$}_{Y}=0$. Hence, $\chi (Y)=12 \chi(\mbox{$\cal O$}_{Y})$, \cite[p 472]{gri}, which is equal to the positive number of singular fibres, cf. \cite{ue}. Now, $\mbox{$\cal O$}_{Y}((m_{i}-1)Y_{\sigma_{i}})= q^{*}\mbox{$\cal O$}_{\Sigma}(\frac{m_{i}-1}{m_{i}}\sigma_{i})$. (This is in fact a trivial sheaf over $\Sigma$.) We obtain \[ q_{*1}\mbox{$\cal T$}'=\mbox{$\cal T$} \otimes q_{*1} \mbox{$\cal O$}_{Y} \otimes \mbox{$\cal O$}_{\Sigma}(\sum \frac{1-m_{i}}{m_{i}} \sigma_{i})\] and any non zero section of this induces a non zero map $\mbox{$\cal E$} \rightarrow \mbox{$\cal E$} \otimes \mbox{$\cal L$}$ for a negative line V-bundle $\mbox{$\cal L$}$. This is ruled out by stability of $\mbox{$\cal E$}$. $\Box$ \hfill\vspace{0.5cm} \section{Smooth parabolic bundles} \begin{prop} \showlabel{v} Two smooth line V-bundles over $\Sigma$ are isomorphic if their pull backs to $Y$ are isomorphic. \end{prop} There is an equivalence between smooth line V-bundles over $\Sigma$ and $\mbox{$\Lambda$}^{2}_{V}(\Sigma,\Z)$, \cite{fas}. As we don't have a sufficient theory of V-cohomology however, our proof is not by showing injectivity of $\mbox{$\Lambda$}_{V}^{2}(\Sigma,\Z) \rightarrow \mbox{$\Lambda$}^{2}(Y,\Z)$. \noindent{\em Proof:} It suffices to show that a smooth line V-bundle $L \rightarrow \Sigma$ is trivial if $L':=q^{*}L \rightarrow Y$ is trivial. Let us write $q':SL' \rightarrow SL$ for the induced map of circle bundles. We use the isomorphism $q_{*}:\pi_{1}(Y) \rightarrow \pi_{1}^{V}(\Sigma)$ and the commutative diagram \setlength{\unitlength}{1.0mm} \begin{center} \begin{picture}(0,40)(0,-20) \putcc{-52}{13}{0} \putcc{-26}{13}{K'} \putcc{0}{13}{\pi_{1}(SL')} \putcc{26}{13}{\pi_{1}(Y)} \putcc{52}{13}{0} \putcc{-52}{-13}{0} \putcc{-26}{-13}{K} \putcc{0}{-13}{\pi_{1}^{V}(SL)} \putcc{26}{-13}{\pi_{1}^{V}(\Sigma)} \putcc{52}{-13}{0} \put(-45,13){\vector(1,0){12}} \put(-19,13){\vector(1,0){10}} \put(9,13){\vector(1,0){10}} \put(33,13){\vector(1,0){12}} \put(-45,-13){\vector(1,0){12}} \put(-19,-13){\vector(1,0){10}} \put(9,-13){\vector(1,0){10}} \put(33,-13){\vector(1,0){12}} \put(-52,8){\vector(0,-1){16}} \put(-26,8){\vector(0,-1){16}} \put(0,8){\vector(0,-1){16}} \put(52,8){\vector(0,-1){16}} \put(26,8){\vector(0,-1){16}} \putbc{13}{15}{\pi'_{*}} \putbc{13}{-11}{\pi_{*}} \putcl{2}{0}{q'_{*}} \putcl{28}{0}{q_{*}} \end{picture} \end{center} where $K=\langle k \rangle$, $K'=\langle k' \rangle$ for regular fibres $k$ of $SL$ and $k'$ of $SL'$. The rows are exact because bundles are always regular maps. By the five-lemma, $q'_{*}$ is an isomorphism if $(k' \mapsto k):K' \rightarrow K$ is an isomorphism. Certainly, it is surjective. Assume there is a V-homotopy $H: [0,1] \times [0,1] \rightarrow SL$ with boundary $k^{n}$. Since $q$ is regular away from finitely many points, there exists $H':[0,1] \times [0,1] \rightarrow Y$ lifting $\pi \circ H$. Hence there exists $\tilde{H}:[0,1] \times [0,1] \rightarrow SL'$ lifting $H$ and $H'$. If $k \in S(L_{x})$ then $\mbox{$\partial$} H' = \pi' \circ \mbox{$\partial$} \tilde{H} \subseteq Y_{x}$ and $\mbox{$\partial$} H =q' \circ \mbox{$\partial$} \tilde{H} =k^{n}$. W.l.o.g. $x \in U_{0}$, where $U_{0}$ is as in theorem \ref{k}, and we are working on $SL'|_{q^{-1}(U_{0})}=S^{1} \times q^{-1}(U_{0})$. So we can lift some homotopy (inside $q^{-1}(U_{0})$) with boundary $\mbox{$\partial$} H'$ to one relating $\mbox{$\partial$} \tilde{H}$ to $(k')^{n}$. Hence, $q'_{*}$ is an isomorphism. Seifert proved that $\pi_{1}^{V}(SL)=$ \[ \langle a_{j}, b_{j},g_{i},k \,\, : \,\, [a_{j},k]= [b_{j},k]=[g_{i},k]=1=% g_{i}^{m_{i}}k^{\beta_{i}}=k^{-b} \prod_{j=1}^{g} [a_{j},b_{j}] \prod_{i=1}^{n} g_{i} \rangle \] where $g$ is the genus of $\Sigma$ and $m_{i}$ is the multiplicity of $\sigma_{i}$. Furthermore, the isotropy $\beta_{i} \,\,\, \mbox{mod}\,\,\, m_{i}$ of $L$ at $\sigma_{i}$ and $\deg L=b+\sum_{1}^{n} \beta_{i}/m_{i}$ are independent of the choices of lifts $g_{i}, a_{j}$ and $b_{j}$ of the generators of $\pi_{1}^{V} (\Sigma)$. By the isomorphism of the above extensions and if $SL'$ is trivial, we can choose lifts such that all $\beta_{i}$'s are zero as well as $b$. By theorem \ref{i}, this implies that $L$ is trivial. $\Box$ \hfill\vspace{0.5cm} Now we prove proposition \ref{u}: \noindent{\em Proof:} Theorem \ref{z} implies the existence of some $L \rightarrow \Sigma$ with $\det E'=q^{*}L$ and that the parabolic filtration of $E'$ along $P'$ is by trivial bundles; in particluar $\mbox{$\cal C$}_{2} (E')=0$. Proposition \ref{v} gives the uniqueness of $L$. Now, $|E'|$ is uniquely determined by $\det E'$. Two different filtrations of $E'|_{P'_{j}}$ by trivial subbundles are related by a map $P_{j}' \rightarrow \mbox{Sl}(r,\C)$ which can be extended to an isomorphism of $E'$ being the identity outside a tubular neighbourhood of $P'_{j}$ since $\mbox{Sl}(r,\C)$ is simply connected. This shows (i). After the last argument and by theorem \ref{i}, part (ii) follows from proposition \ref{v} and theorem \ref{z}. $\Box$ \hfill\vspace{0.5cm}
1996-08-13T11:05:17
9503
alg-geom/9503023
en
https://arxiv.org/abs/alg-geom/9503023
[ "alg-geom", "math.AG" ]
alg-geom/9503023
Richard Earl
Richard Earl
The Mumford relations and the moduli of rank three stable bundles
35 Pages, no figures LaTeX v 2.09
null
null
null
null
We find a complete set of relations for the rational cohomology ring of the moduli space of rank three stable bundles over a Riemann surface of genus g and also show that the Pontryagin ring vanishes in degree 12g-8 and greater. The results are obtained by introducing some 'dual' Mumford relations and generalising Kirwan's proofs of the Mumford and Newstead conjectures in the rank two case. (In this revised version of the paper the vanishing degree of the Pontryagin ring of the moduli space has been improved from `in and above degree 12g-4' to `in and above degree 12g-8'. This degree is now known to be sharp.)
[ { "version": "v1", "created": "Wed, 29 Mar 1995 18:43:54 GMT" }, { "version": "v2", "created": "Thu, 14 Dec 1995 13:13:43 GMT" }, { "version": "v3", "created": "Tue, 13 Aug 1996 09:00:19 GMT" } ]
2008-02-03T00:00:00
[ [ "Earl", "Richard", "" ] ]
alg-geom
\section{Introduction.} Let ${\cal M}(n,d)$ denote the moduli space of semistable holomorphic vector bundles of coprime rank $n$ and degree $d$ over a Riemann surface $M$ of genus $g \geq 2.$ Throughout this article we will write \[ \bar{g} = g-1. \] Recall that a holomorphic vector bundle $E$ over $M$ is said to be semistable (resp. stable) if every proper subbundle $F$ of $E$ satisfies \[ \mu(F) \leq \mu(E) \indent (\mbox{resp. } \mu(F) < \mu(E)) \] where $\mu(F) = \mbox{degree}(F)/\mbox{rank}(F)$ is the slope of $F$. Non-semistable bundles are said to be unstable. When $n$ and $d$ are coprime the stable and semistable bundles coincide.\\ \indent Let ${\cal E}$ be a fixed $C^{\infty}$ complex vector bundle of rank $n$ and degree $d$ over $M$. Let ${\cal C}$ be the space of all holomorphic structures on ${\cal E}$ and let ${\cal G}_{c}$ denote the group of all $C^{\infty}$ complex automorphisms of ${\cal E}$. Atiyah and Bott \cite{AB} identify the moduli space ${\cal M}(n,d)$ with the quotient ${\cal C}^{ss}/{\cal G}_{c}$ where ${\cal C}^{ss}$ is the open subset of ${\cal C}$ consisting of all semistable holomorphic structures on ${\cal E}$. In this construction both ${\cal C}$ and ${\cal G}_{c}$ are infinite dimensional; there exist other constructions \cite{K3} of the moduli space ${\cal M}(n,d)$ as genuine geometric invariant theoretic quotients which are in a sense finite dimensional approximations of Atiyah and Bott's construction.\\ \indent There is a known set of generators \cite{N2,AB} for the rational cohomology ring of ${\cal M}(n,d)$ as follows. Let $V$ denote a universal bundle over ${\cal M}(n,d) \times M$. Atiyah and Bott then define elements \begin{equation} a_{r} \in H^{2r}({\cal M}(n,d);{\bf Q}), \quad b_{r}^{s} \in H^{2r-1}({\cal M}(n,d);{\bf Q}), \quad f_{r} \in H^{2r-2}({\cal M}(n,d);{\bf Q}) \label{000} \end{equation} where $1 \leq r \leq n,1 \leq s \leq 2g$ by writing \begin{equation} c_{r}(V) = a_{r} \otimes 1 + \sum_{s=1}^{2g} b_{r}^{s} \otimes \alpha_{s} + f_{r} \otimes \omega \indent 1 \leq r \leq n \label{0} \end{equation} where $\omega$ is the standard generator of $H^{2}(M;{\bf Q})$ and $\alpha_{1},...,\alpha_{2g}$ form a fixed canonical cohomology basis for $H^{1}(M;{\bf Q})$. The ring $H^{*}({\cal M}(n,d);{\bf Q})$ is freely generated as a graded algebra over ${\bf Q}$ by the elements (\ref{000}). Notice from the definition that $f_{1}=d$. We further introduce the notation \[ \xi_{i,j} = \sum_{s=1}^{g} b_{i}^{s} b_{j}^{s+g}. \] \indent The universal bundle $V$ is not unique, although its projective class is. We may tensor $V$ by the pullback to ${\cal M}(n,d) \times M$ of any holomorphic line bundle $K$ over ${\cal M}(n,d)$ to give another bundle with the same universal property. This process changes the generators of $H^{*}({\cal M}(n,d);{\bf Q})$. In particular it changes $a_{1}$ by $n c_{1}(K)$ and $c_{1}(\pi_{!}V)$ by $(d-n\bar{g})c_{1}(K)$ where $\pi:{\cal M}(n,d) \times M \rightarrow {\cal M}(n,d)$ is the first projection and $\pi_{!}$ is the direct image map from K-theory \cite[p.436]{H}. Since $n$ and $d$ are coprime there exist integers $u$ and $v$ such that \[ u n + v (d-n\bar{g}) =1. \] Thus if we take $K$ to be \[ \left. \mbox{det}(V \right|_{{\cal M}(n,d)}) ^{u} \otimes (\mbox{det} \pi_{!}V)^{v} \] then $V \otimes \pi^{*}(K^{-1})$ is a new universal bundle such that \begin{equation} u a_{1} + v c_{1}(\pi_{!}V) = 0. \label{NORM} \end{equation} Following Atiyah and Bott \cite[p.582]{AB} we replace $V$ by this normalised universal bundle.\\ \indent The normalised bundle $V$ is universal in the sense that its restriction to $\{[E]\} \times M$ is isomorphic to $E$ for each semistable holomorphic bundle $E$ over $M$ of rank $n$ and degree $d$ and where $[E]$ is the class of $E$ in ${\cal M}(n,d).$ Then the stalk of the $i$th higher direct image sheaf $R^{i}\pi_{*}V$ (see \cite[$\S 3.8$]{H}) at $[E]$ is \[ H^{i}(\pi^{-1}([E]),V_{|\pi^{-1}([E])}) = H^{i}(M,V_{|[E] \times M}) \cong H^{i}(M,E). \] \indent Tensoring $E$ with a holomorphic line bundle over $M$ of degree $D$ gives an isomorphism between ${\cal M}(n,d)$ and ${\cal M}(n,d+nD)$. Since $n$ and $d$ are coprime we may assume without any loss of generality that $2\bar{g} n<d<(2\bar{g}+1)n$ and so we will write \[ d=2n\bar{g}+\delta \indent (0 < \delta < n) \] from now on. From \cite[lemma 5.2]{N} we know that $H^{1}(M,E)=0$ for any semistable holomorphic bundle $E$ of slope greater than $2\bar{g}$. Thus $\pi_{!}V$ is in fact a vector bundle over ${\cal M}(n,d)$ with fibre $H^{0}(M,E)$ over $[E] \in {\cal M}(n,d)$ and, by the Riemann-Roch theorem, of rank $d-n\bar{g}=n\bar{g}+\delta$.\\ \indent In particular if we express the Chern classes $c_{r}(\pi_{!}V)$ in terms of the generators $a_{r},b_{r}^{s}$ and $f_{r}$ of $H^{*}({\cal M}(n,d);{\bf Q})$ then knowing the images of the $r$th Chern classes in $H^{*}({\cal M}(n,d);{\bf Q})$ vanish for $r>n\bar{g}+\delta$ gives us relations in terms of the images of the generators in $H^{*}({\cal M}(n,d);{\bf Q}).$ Now from \cite[prop. 9.7]{AB} we know that \begin{equation} H^{*}({\cal M}(n,d);{\bf Q}) \cong H^{*}({\cal M}_{0}(n,d);{\bf Q}) \otimes H^{*}(\mbox{Jac}(M);{\bf Q}) \label{21} \end{equation} where $\mbox{Jac}(M)$ is the Jacobian of the Riemann surface $M$ and ${\cal M}_{0}(n,d)$ is the moduli space of rank $n$ bundles with degree $d$ and fixed determinant line bundle. $H^{*}(\mbox{Jac}(M);{\bf Q})$ is an exterior algebra on $2g$ generators and we can choose the isomorphism (\ref{21}) so that these generators correspond to $b_{1}^{1},...,b_{1}^{2g}$ and the elements $a_{2},...,a_{n},b_{2}^{1},...,b_{n}^{2g},f_{2},...,f_{n}$ correspond to the generators of $H^{*}({\cal M}_{0}(n,d);{\bf Q})$. So we can find relations in terms of $a_{2},...,a_{n},b_{2}^{1},...,b_{n}^{2g},$ and $f_{2},...,f_{n}$ by equating to zero the coefficients of $\prod_{s \in S}b_{1}^{s}$ in the Chern classes $c_{r}(\pi_{!}V)$ for $r>n\bar{g}+\delta$ and for every subset $S \subseteq \{1,...,2g\}.$\\ \indent Mumford's conjecture, as proven by Kirwan \cite[$\S$2]{K2}, was that when the rank $n$ is two then these relations together with the relation (\ref{NORM}) from normalising the universal bundle $V$ provide a complete set of relations in $H^{*}({\cal M}_{0}(2,d);{\bf Q})$. Subsequently a stronger version of Mumford's conjecture has been proven \cite{E} showing the relations coming from the first vanishing Chern class $c_{2g}(\pi_{!}V)$ generate the relation ideal of $H^{*}({\cal M}_{0}(2,d))$ as a ${\bf Q}[a_{2},f_{2}]$-module.\\ \begin{rem} In the rank two case the Mumford relations above differ somewhat from the relations $\xi_{r}$ introduced by Zagier and studied in \cite{B,KN,ST,Z}. In the notation of \cite{Z} \[ \Psi_{\{1,...,2g\}} \left( \frac{-t-a_{1}}{2} \right) = \frac{(-1)^{g\bar{g}/2 + g}}{2^{2g-1}} t^{\bar{g}} F_{0}(t^{-1}) \] where $\Psi_{\{1,...2g\}}(x)$ denotes the coefficient of $\prod_{s=1}^{2g} b_{1}^{s}$ in $\Psi(x) =\sum_{r \geq 0} c_{r}(\pi_{!}V) x^{2g-1-r}$ and $F_{0}(t) = \sum_{r=0}^{\infty} \xi_{r} t^{r}$. In the notation of \cite{KN} $\xi_{r}$ appears as $\zeta_{r}/r!$ and in \cite{ST} as $\Phi^{(r)}/r!.$ \end{rem} \indent We will demonstrate later (remark \ref{inadequacy}) that the Mumford relations are not complete when the rank $n$ is greater than two. For now we introduce a new set of relations. Let $L$ be a fixed line bundle over $M$ of degree $4\bar{g}+1$ and let $\phi:{\cal M}(n,d) \times M \rightarrow M$ be the second projection. Then $\pi_{!}(V^{*} \otimes \phi^{*}L)$ is a vector bundle over ${\cal M}(n,d)$ of rank $(3\bar{g}+1)n-d =ng-\delta$ with fibre $H^{0}(M,E^{*} \otimes L)$ over $[E]$. By equating to zero the coefficients of $\prod_{s \in S}b_{1}^{s}$ in the Chern classes $c_{r}(\pi_{!}(V^{*} \otimes \phi^{*}L))$ for $r>ng- \delta$ and for every subset $S \subseteq \{1,...,2g\}$ we may find relations in terms of the generators $a_{2},...,a_{n},b_{2}^{1},...,b_{n}^{2g},$ and $f_{2},...,f_{n}$. We will refer to these new relations as the dual Mumford relations. \begin{rem} \label{dualise} The map $E \mapsto E^{*} \otimes L$ induces an automorphism of $H^{*}({\cal M}(2,d);{\bf Q})$ mapping the Mumford relations to the dual Mumford relations and vice versa. Hence we can deduce that the dual Mumford relations are complete when the rank is two from Kirwan's proof of Mumford's conjecture \cite[$\S$ 2]{K2}. \end{rem} \indent Our first result (to be proved in $\S 4$) now reads as:\\[\baselineskip] {\bf THEOREM 1.} {\em The Mumford and dual Mumford relations together with the relation (\ref{NORM}) due to the normalisation of the universal bundle $V$ form a complete set of relations for $H^{*}({\cal M}(3,d);{\bf Q}).$}\\[\baselineskip] \indent The Newstead-Ramanan conjecture states \cite[$\S$5a]{N2} that the Pontryagin ring of the tangent bundle to ${\cal M}(2,d)$ vanishes in degrees $4g$ and higher. The conjecture was proven independently by Thaddeus \cite{T} and Kirwan \cite[$\S$4]{K2}, and has been proven more recently by King and Newstead \cite{KN} and Weitsman \cite{W}. In $\S 5$ we will use a similar method to Kirwan's but now also involving the dual Mumford relations to prove:\\[\baselineskip] {\bf THEOREM 2.} {\em The Pontryagin ring of the moduli space ${\cal M}(3,d)$ vanishes in degrees $12g-8$ and above.} \section{Kirwan's Approach.} \indent The group ${\cal G}_{c}$ is the complexification of the gauge group ${\cal G}$ of all smooth automorphisms of ${\cal E}$ which are unitary with respect to a fixed Hermitian structure on ${\cal E}$ \cite[p.570]{AB}. We shall write $\overline{{\cal G}}$ for the quotient of ${\cal G}$ by its $U(1)$-centre and $\overline{{\cal G}}_{c}$ for the quotient of ${\cal G}_{c}$ by its ${\bf C}^{*}$-centre. \\ \indent There are natural isomorphisms \cite[9.1]{AB} \[ H^{*}({\cal C}^{ss}/{\cal G}_{c};{\bf Q}) = H^{*}({\cal C}^{ss}/\overline{{\cal G}}_{c};{\bf Q}) \cong H^{*}_{\overline{{\cal G}}_{c}}({\cal C}^{ss};{\bf Q}) \cong H^{*}_{\overline{{\cal G}}}({\cal C}^{ss};{\bf Q}) \] since the ${\bf C}^{*}$-centre of ${\cal G}_{c}$ acts trivially on ${\cal C}^{ss}$, $\overline{{\cal G}}_{c}$ acts freely on ${\cal C}^{ss}$ and $\overline{{\cal G}}_{c}$ is the complexification of $\overline{{\cal G}}$. Atiyah and Bott \cite[thm. 7.14]{AB} show that the restriction map $H^{*}_{\overline{{\cal G}}}({\cal C};{\bf Q}) \rightarrow H^{*}_{\overline{{\cal G}}}({\cal C}^{ss};{\bf Q})$ is surjective. Further $H^{*}_{\overline{{\cal G}}}({\cal C};{\bf Q}) \cong H^{*}(B\overline{{\cal G}};{\bf Q})$ since ${\cal C}$ is an affine space \cite[p.565]{AB}. So putting this all together we have \begin{equation} H^{*}(B\overline{{\cal G}};{\bf Q}) \cong H^{*}_{\overline{{\cal G}}}({\cal C};{\bf Q}) \rightarrow H^{*}_{\overline{{\cal G}}}({\cal C}^{ss};{\bf Q}) \cong H^{*}({\cal M}(n,d);{\bf Q}) \label{14} \end{equation} is a surjection.\\ \indent As shown in \cite[prop. 2.4]{AB} the classifying space $B{\cal G}$ can be identified with the space $\mbox{Map}_{d}(M,BU(n))$ of all smooth maps $f:M \rightarrow BU(n)$ such that the pullback to $M$ of the universal vector bundle over $BU(n)$ has degree $d$. If we pull back this universal bundle using the evaluation map \[ \mbox{Map}_{d}(M,BU(n)) \times M \rightarrow BU(n): (f,m) \mapsto f(m) \] then we obtain a rank $n$ vector bundle ${\cal V}$ over $B{\cal G} \times M$. If we restrict the pullback bundle induced by the maps \[ {\cal C}^{ss} \times E{\cal G} \times M \rightarrow {\cal C} \times E{\cal G} \times M \rightarrow {\cal C} \times_{{\cal G}} E{\cal G} \times M \stackrel{\simeq}{\rightarrow} B{\cal G} \times M \] to ${\cal C}^{ss} \times \{e\} \times M$ for some $e \in E{\cal G}$ then we obtain a ${\cal G}$-equivariant holomorphic bundle on ${\cal C}^{ss} \times M$. The $U(1)$-centre of ${\cal G}$ acts as scalar multiplication on the fibres, and the associated projective bundle descends to a holomorphic projective bundle over ${\cal M}(n,d) \times M$ which is in fact the projective bundle of $V$ \cite[pp.579-580]{AB}.\\ \indent By a slight abuse of notation we define elements $a_{r}, b_{r}^{s}, f_{r}$ in $H^{*}(B{\cal G};{\bf Q})$ by writing \[ c_{r}({\cal V}) = a_{r} \otimes 1 + \sum_{s=1}^{2g} b_{r}^{s} \otimes \alpha_{s} + f_{r} \otimes \omega \indent 1 \leq r \leq n. \] Atiyah and Bott show \cite[prop. 2.20]{AB} that the ring $H^{*}(B{\cal G};{\bf Q})$ is freely generated as a graded algebra over ${\bf Q}$ by the elements $a_{r}, b_{r}^{s}, f_{r}$. The only relations amongst these generators are that the $a_{r}$ and $f_{r}$ commute with everything else and that the $b_{r}^{s}$ anticommute with each other.\\ \indent The fibration $BU(1) \rightarrow B{\cal G} \rightarrow B\overline{\cal G}$ induces an isomorphism \cite[p.577]{AB} \[ H^{*}(B{\cal G};{\bf Q}) \cong H^{*}(B\overline{{\cal G}};{\bf Q}) \otimes H^{*}(BU(1);{\bf Q}). \] The generators $a_{r},b_{r}^{s}$ and $f_{r}$ of $H^{*}(B{\cal G};{\bf Q})$ can be pulled back via a section of this fibration to give rational generators of the cohomology ring of $B\overline{\cal G}$. We may if we wish omit $a_{1}$ since its image in $H^{*}(B\overline{{\cal G}};{\bf Q})$ can be expressed in terms of the other generators. The only other relations are again the commuting of the $a_{r}$ and $f_{r}$, and the anticommuting of the $b_{r}^{s}$. We may then normalise ${\cal V}$ suitably so that these generators for $H^{*}(B\overline{{\cal G}};{\bf Q})$ restrict to the generators $a_{r}, b_{r}^{s}, f_{r}$ for $H^{*}({\cal M}(n,d);{\bf Q})$ under the surjection (\ref{14}).\\ \indent The relations amongst these generators for $H^{*}({\cal M}(n,d);{\bf Q})$ are then given by the kernel of the restriction map (\ref{14}) which in turn is determined by the map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \cong H^{*}_{\overline{{\cal G}}}({\cal C};{\bf Q}) \otimes H^{*}(BU(1);{\bf Q}) \rightarrow \indent\indent\indent\indent\indent\indent \] \[ \indent\indent\indent\indent\indent H^{*}_{\overline{{\cal G}}}({\cal C}^{ss};{\bf Q}) \otimes H^{*}(BU(1);{\bf Q}) \cong H^{*}_{\cal G}({\cal C}^{ss};{\bf Q}). \] In order to describe this kernel we consider Shatz's stratification of ${\cal C}$, the space of holomorphic structures on ${\cal E}$ \cite{Sh}. The stratification $\{{\cal C}_{\mu} : \mu \in {\cal M} \}$ is indexed by the partially ordered set ${\cal M}$, consisting of all the types of holomorphic bundles of rank $n$ and degree $d$, as follows.\\ \indent Any holomorphic bundle $E$ over $M$ of rank $n$ and degree $d$ has a canonical filtration (or flag) \cite[p.221]{HN} \[ 0 = E_{0} \subset E_{1} \subset \cdot \cdot \cdot \subset E_{P} = E \] of sub-bundles such that the quotient bundles $Q_{p}= E_{p}/E_{p-1}$ are semi-stable and $\mu(Q_{p}) > \mu(Q_{p+1})$. We will write $d_{p}$ and $n_{p}$ respectively for the degree and rank of $Q_{p}$. Given such a filtration we define the type of $E$ to be \[ \mu = (\mu(Q_{1}),...,\mu(Q_{P})) \in {\bf Q}^{n} \] where the entry $\mu(Q_{p})$ is repeated $n_{p}$ times. When there is no chance of confusion we will also refer collectively to the strata of type $(n_{1},...,n_{s})$ and we will write $\Delta$ for the collection of strata with $n_{p}=1$ for each $p$. The semistable bundles have type $\mu_{0} = (d/n,...,d/n)$ and form the unique open stratum. The set ${\cal M}$ of all possible types of holomorphic vector bundles over $M$ will provide our indexing set. A partial order on ${\cal M}$ is defined as follows. Let $\sigma=(\sigma_{1},...,\sigma_{n})$ and $\tau=(\tau_{1},...,\tau_ {n})$ be two types; we say that $\sigma \geq \tau$ if and only if \[ \sum_{j \leq i} \sigma_{j} \geq \sum_{j \leq i} \tau_{j} \mbox{ for } 1 \leq i \leq n-1. \] The set ${\cal C}_{\mu} \subseteq {\cal C},$ $\mu \in {\cal M}$, is defined to be the set of all holomorphic vector bundles of type $\mu$.\\ \indent The stratification also has the following properties:-\\ \indent (i) The stratification is smooth. That is each stratum ${\cal C}_{\mu}$ is a locally closed ${\cal G}_{c}$-invariant submanifold. Further for any $\mu \in {\cal M}$ \cite[7.8]{AB}\\ \begin{equation} \overline{{\cal C}_{\mu}} \subseteq \bigcup_{\nu \geq \mu} {\cal C}_{\nu}. \label{11} \end{equation} \indent (ii) Each stratum ${\cal C}_{\mu}$ is connected and has finite (complex) codimension $d_{\mu}$ in ${\cal C}$. Moreover given any integer $N$ there are only finitely many $\mu \in {\cal M}$ such that $d_{\mu} \leq N$. Further $d_{\mu}$ is given by the formula \cite[7.16]{AB} \begin{equation} d_{\mu}= \sum_{i>j} (n_{i}d_{j}-n_{j}d_{i}+n_{i}n_{j}\bar{g}) \label{12} \end{equation} where $d_{k}$ and $n_{k}$ are the degree and rank, respectively, of $Q_{k}$.\\ \indent (iii) The gauge group ${\cal G}$ acts on ${\cal C}$ preserving the stratification which is equivariantly perfect with respect to this action \cite[thm. 7.14]{AB}. In particular there is an isomorphism of vector spaces \[ H^{k}_{{\cal G}}({\cal C};{\bf Q}) \cong \bigoplus_{\mu \in {\cal M}} H^{k-2d_{\mu}}_{{\cal G}}({\cal C}_{\mu};{\bf Q}) = H^{k}_{{\cal G}}({\cal C}^{ss};{\bf Q}) \oplus \bigoplus_{\mu \neq \mu_{0}} H^{k-2d_{\mu}}_{{\cal G}} ({\cal C}_{\mu};{\bf Q}). \] The restriction map $H^{*}_{{\cal G}}({\cal C};{\bf Q}) \rightarrow H^{*}_{{\cal G}}({\cal C}^{ss};{\bf Q})$ is the projection onto the summand $H^{*}_{{\cal G}}({\cal C}^{ss};{\bf Q})$ and so the kernel is isomorphic as a vector space to \begin{equation} \bigoplus_{k \geq 0} \bigoplus_{\mu \neq \mu_{0}} H^{k-2d_{\mu}}_{{\cal G}} ({\cal C}_{\mu};{\bf Q}). \label{13} \end{equation} \begin{rem} \label{inadequacy} We can at this point use a dimension argument to show that the Mumford relations are generally not complete when the rank $n$ is greater than two. From the isomorphism (\ref{13}) we can see that for the Mumford relations to be complete it is necessary that the least degree of a Mumford relation must be less than or equal to the smallest real codimension of an unstable stratum. The degree of $\sigma_{r,S}^{k}$ equals $2(n\bar{g}+\delta -nr -k) -|S|$ which is least when $r=-1,k=n-1,$ and $S=\{1,...,2g\}.$ So the smallest degree of a Mumford relation is $2(\delta + (n-1)\bar{g}).$ However a simple calculation minimising the codimension formula (\ref{12}) shows that the least real codimension of an unstable stratum is $2(\delta + (n-1)\bar{g})$ when $\delta < n/2$ and is $2(n-\delta +(n-1)\bar{g})$ when $\delta>n/2$. Hence the Mumford relations are not complete when $n \geq 3$ and $\delta>n/2.$ A similar argument shows that the dual Mumford relations are not complete when $\delta<n/2$ since the smallest degree of a dual Mumford relation is $2(n-\delta +(n-1)\bar{g}).$ Clearly however this simple argument does not tell us anything concerning the union of the Mumford and dual Mumford relations. \end{rem} \indent To conclude this section we will describe a set of criteria for the completeness of a set of relations in $H^{*}({\cal M}(n,d);{\bf Q})$ and reformulate the Mumford and dual Mumford relations in a way more suited to these criteria. Consider the formal power series \[ c(\pi_{!}{\cal V})(t) = \sum_{r \geq 0} c_{r}(\pi_{!}{\cal V}) \cdot t^{r} \in H^{*}_{\cal G}({\cal C};{\bf Q})[[t]]. \] The vanishing of the image of $c_{r}(\pi_{!}{\cal V})$ in $H^{*}({\cal M}(n,d);{\bf Q})$ for $r > n\bar{g}+\delta$ is equivalent to the image of $c(\pi_{!}{\cal V})(t)$ being a polynomial of degree at most $n\bar{g}+\delta$ or equally to the image of \[ \Psi(t) = t^{n\bar{g}+\delta}c(\pi_{!}{\cal V})(t^{-1}) \] being a polynomial of degree at most $n\bar{g}+\delta$ in $H^{*}({\cal M}(n,d);{\bf Q})[t].$ If we write $\Psi(t)$ as the series \[ \Psi(t)=\sum_{r=-\infty}^{\bar{g}} (\sigma_{r}^{0} + \sigma_{r}^{1}t + \cdot \cdot \cdot + \sigma_{r}^{n-1}t^{n-1} ) (\tilde{\Omega}(t))^{r} \] where $\tilde{\Omega}(t)= t^{n}+a_{1}t^{n-1}+ \cdot \cdot \cdot +a_{n}$ then the Mumford relations are equivalent to the vanishing of the images of $\sigma_{r,S}^{k} (r<0,0 \leq k \leq n-1,S \subseteq \{1,...,2g\})$ in $H^{*}({\cal M}_{0}(n,d);{\bf Q})$ when we write \begin{equation} \sigma_{r}^{k} = \sum_{S \subseteq \{1,...,2g\}} \sigma_{r,S}^{k} \prod_{s \in S} b_{1}^{s}. \label{998} \end{equation} We will refer to $\sigma_{r,S}^{k} (r<0,0 \leq k \leq n-1,S \subseteq \{1,...,2g\})$ as the Mumford relations.\\ \indent Similarly we know that the restriction of \[ \Psi^{*}(t) = t^{ng - \delta}c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t^{-1}) \] to $H^{*}({\cal M}(n,d);{\bf Q})$ is a polynomial. As before we may put $\Psi^{*}(t)$ in the form \[ \Psi^{*}(t) = \sum_{r=-\infty}^{\bar{g}}(\tau_{r}^{0}+\tau_{r}^{1}t+ \cdot \cdot \cdot + \tau_{r}^{n-1}t^{n-1})(\tilde{\Omega}(t))^{r} \] where $\tilde{\Omega}(t) = t^{n}+a_{1}t^{n-1}+ \cdot \cdot \cdot +a_{n}$ and similarly we write \begin{equation} \tau_{r}^{k}=\sum_{S \subseteq \{1,...,2g\}} \tau_{r,S}^{k} \prod_{s \in S} b_{1}^{s}. \label{999} \end{equation} We will refer to $\tau_{r,S}^{k} (r<0,0 \leq k \leq n-1,S \subseteq \{1,...,2g\})$ as the dual Mumford relations.\\ \indent The motivation for this is that the restrictions of $\sigma_{r,S}^{k}$ and $\tau_{r,S}^{k}$ to the strata ${\cal C}_{\mu}$ are easier to calculate in this form. This is a crucial step in applying the following completeness criteria.\\ \indent Given $\mu=(\mu_{1},...,\mu_{n}),\nu=(\nu_{1},...,\nu_{n}) \in {\cal M}$ then we write $\nu \prec \mu$ if there exists $T$, $1 \leq T \leq n$, such that \[ \nu_{i} = \mu_{i} \mbox{ for } T < i \leq n \mbox{ and } \nu_{T} > \mu_{T}. \] We write $\nu \preceq \mu$ if $\nu \prec \mu$ or $\nu = \mu.$ A few easy calculations verify that $\preceq$ is a total order on ${\cal M}$ with minimal element $\mu_{0}$, the semistable type. For an unstable type $\mu$ we will write $\mu-1$ for the type previous to $\mu$ with respect to $\preceq.$ \begin{prop} \label{KCC} (Completeness Criteria) Let ${\cal R}$ be a subset of the kernel of the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}^{ss};{\bf Q}). \] Suppose that for each unstable type $\mu$ there is a subset ${\cal R}_{\mu}$ of the ideal generated by ${\cal R}$ such that the image of ${\cal R}_{\mu}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \] is zero when $\nu \prec \mu$ and when $\nu = \mu$ contains the ideal of $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ generated by $e_{\mu}$, the equivariant Euler class of ${\cal N}_{\mu}$, the normal bundle to the stratum ${\cal C}_{\mu}$ in ${\cal C}.$ Then ${\cal R}$ generates the kernel of the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}^{ss};{\bf Q}) \] as an ideal of $H^{*}_{\cal G}({\cal C};{\bf Q}).$ \end{prop} \begin{rem} The proof of proposition \ref{KCC} below follows similar lines to the proof of \cite[prop.1]{K2}. However there are some differences -- the order $\preceq$ does not generally coincide with $\leq$ -- and further the proof of \cite[p.867]{K2} as given is true only for the rank two case. For these reasons we include a proof of proposition \ref{KCC} below although it clearly owes many of its origins to \cite{K2}. \end{rem} {\bf Proof} Let $\mu \in {\cal M}$ and define \[ V_{\mu} = \bigcup_{\nu \preceq \mu} {\cal C}_{\nu}. \] We will firstly show that $V_{\mu}$ is an open subset of ${\cal C}$ containing ${\cal C}_{\mu}$ as a closed submanifold. Note that if $\nu \leq \mu$ then $\nu \preceq \mu$ and thus by property (\ref{11}) if $\nu \succ \mu$ then $\overline{{\cal C}}_{\nu} \subseteq {\cal C} - V_{\mu}$. The stratification is locally finite and hence $V_{\mu}$ is open. Further note that the closure of ${\cal C}_{\mu}$ in $V_{\mu}$ equals \[ V_{\mu} \cap \bigcup_{\nu \geq \mu} {\cal C}_{\nu} = {\cal C}_{\mu} \] as required.\\ \indent Recall now that the composition of the Thom-Gysin map \[ H_{{\cal G}}^{*-2d_{\mu}}({\cal C}_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}(V_{\mu};{\bf Q}) \] with the restriction map \[ H^{*}_{\cal G}(V_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] is given by multiplication by the Euler class $e_{\mu}$ which is not a zero-divisor in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ \cite[p.569]{AB}. It follows from the exactness of the Thom-Gysin sequence \[ \cdot \cdot \cdot \rightarrow H_{{\cal G}}^{*-2d_{\mu}}({\cal C}_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}(V_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}(V_{\mu-1};{\bf Q}) \rightarrow \cdot \cdot \cdot \] that the direct sum of the restriction maps \[ H^{*}_{\cal G}(V_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \oplus H^{*}_{\cal G}(V_{\mu-1};{\bf Q}) \] is injective. Hence inductively the direct sum of restriction maps \[ H^{*}_{\cal G}(V_{\mu-1};{\bf Q}) \rightarrow \bigoplus_{\nu \prec \mu} H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \] is injective and in particular the image of any element of ${\cal R}_{\mu}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}(V_{\mu-1};{\bf Q}) \] is zero.\\ \indent For any given $i \geq 0$ there are only finitely many $\nu \in {\cal M}$ such that $2d_{\nu} \leq i$ and so for each $i \geq 0$ there exists some $\mu$ such that \[ H_{{\cal G}}^{i}({\cal C};{\bf Q}) = H_{{\cal G}}^{i}(V_{\mu};{\bf Q}). \] Hence it is enough to show that for each $\mu$ the image in $H^{*}_{\cal G}(V_{\mu};{\bf Q})$ of the ideal generated by ${\cal R}$ contains the image in $H^{*}_{\cal G}(V_{\mu};{\bf Q})$ of the kernel of the restriction map \begin{equation} H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}^{ss};{\bf Q}). \label{QQQ} \end{equation} Note that the above is clearly true for $\mu=\mu_{0}$ as $V_{\mu_{0}}={\cal C}^{ss}.$ We will proceed by induction with respect to $\preceq$.\\ \indent Assume now that $\mu \neq \mu_{0}$ and that $\alpha \in H^{*}_{\cal G}({\cal C};{\bf Q})$ lies in the kernel of (\ref{QQQ}). Suppose that the image of $\alpha$ in $H^{*}_{\cal G}(V_{\mu-1};{\bf Q})$ is in the image of the ideal generated by ${\cal R}.$ We may, without any loss of generality, assume that the image of $\alpha$ in $H^{*}_{\cal G}(V_{\mu-1};{\bf Q})$ is zero. Thus by the exactness of the Thom-Gysin sequence there exists an element $\beta \in H_{{\cal G}}^{*-2d_{\mu}}({\cal C}_{\mu};{\bf Q})$ which is mapped to the image of $\alpha$ in $H^{*}_{\cal G}(V_{\mu};{\bf Q})$ by the Thom-Gysin map \[ H_{{\cal G}}^{*-2d_{\mu}}({\cal C}_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}(V_{\mu};{\bf Q}). \] Hence the image of $\alpha$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] is $\beta e_{\mu}$, and by hypothesis there is an element $\gamma$ of ${\cal R}_{\mu}$ which maps under the restriction map \[ H^{*}_{\cal G}(V_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] to $\beta e_{\mu}.$ Now the images of $\gamma$ and $\alpha$ in $H^{*}_{\cal G}(V_{\mu-1};{\bf Q})$ are both zero and we also know the direct sum of the restriction maps \[ H^{*}_{\cal G}(V_{\mu};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \oplus H^{*}_{\cal G}(V_{\mu-1};{\bf Q}) \] to be injective. Thus the images of $\gamma$ and $\alpha$ in $H^{*}_{\cal G}(V_{\mu};{\bf Q})$ are the same, completing the proof. $\indent \Box$ \begin{rem} Kirwan's completeness criteria follow from the above criteria since for each $\mu$ \[ V_{\mu-1} \subseteq {\cal C} - \bigcup_{\nu \geq \mu} {\cal C}_{\nu}. \] So if the restriction of a relation to $H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q})$ vanishes for every $\nu \not \geq \mu$ then certainly the same relation restricts to zero in $H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q})$ for any $\nu \prec \mu.$ \end{rem} \begin{rem} Kirwan's proof of Mumford's conjecture \cite[$\S$ 2]{K2} amounts to showing that for each unstable type $\mu=(d_{1},d_{2})$ the set \[ {\cal R}_{\mu} = \bigcup \{ \sigma_{d_{2}-2g+1,S}^{0},\sigma_{d_{2}-2g+1,S}^{1} \}, \] where the union is taken over all subsets $S \subseteq \{1,...,2g\}$, satisfies the above criteria. In the rank two case the criteria of proposition \ref{KCC} are in fact equivalent to Kirwan's completeness criteria since $\preceq$ and $\leq$ coincide. \end{rem} \section{Chern Class Computations.} \indent We first describe the restriction maps $H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ and our preferred generators for $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$. Let $\mu=(d_{1}/n_{1},...,d_{P}/n_{P})$. Let ${\cal C}(n_{p},d_{p})^{ss}$ denote the space of all semistable holomorphic structures on a fixed Hermitian vector bundle of rank $n_{p}$ and degree $d_{p}$ and let ${\cal G}(n_{p},d_{p})$ be the gauge group of that bundle. Atiyah and Bott \cite[prop. 7.12]{AB} show that the map \[ \prod_{p=1}^{P} {\cal C}(n_{p},d_{p})^{ss} \rightarrow {\cal C}_{\mu}, \] which sends a sequence of semistable bundles $(F_{1},...,F_{P})$ to the direct sum $F_{1} \oplus \cdot \cdot \cdot \oplus F_{P}$, induces an isomorphism \[ H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \cong \bigotimes_{1 \leq p \leq P} H^{*}_{{\cal G}(n_{p},d_{p})}({\cal C}(n_{p},d_{p})^{ss};{\bf Q}). \] Thus we can find generators \begin{equation} \bigcup_{p=1}^{P} \left( \{a_{r}^{p} | 1 \leq r \leq n_{p} \} \cup \{ b_{r}^{p,s} | 1 \leq r \leq n_{p},1 \leq s \leq 2g \} \cup \{ f_{r}^{p} | 2 \leq r \leq n_{p} \} \right) \label{A} \end{equation} corresponding to the generators of $H^{*}_{\cal G}({\cal C}^{ss};{\bf Q})$ described earlier in (\ref{0}). As before we also define \[ \xi_{i,j}^{p,q} = \sum_{s=1}^{g} b_{i}^{p,s} b_{j}^{q, s+g}. \] To explicitly describe the restriction map note that $c_{r}({\cal V})$ restricts to $c_{r}(\bigoplus_{p=1}^{P} {\cal V}_{p})$ where ${\cal V}_{p}$ is the universal bundle on ${\cal C}(n_{p},d_{p})$. The restrictions of the generators of $H^{*}_{\cal G}({\cal C};{\bf Q})$ can be written in terms of the generators of $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ by taking the appropriate coefficients in the K\"{u}nneth decomposition.\\ \indent One problem that we will be faced with in due course is how to calculate the coefficients of $\prod_{s \in S} b_{1}^{s}$ once we have restricted to a stratum. Suppose first that the stratum concerned is of type $\mu=(d_{1},...,d_{n}) \in \Delta$ and take $\zeta \in H^{*}_{\cal G}({\cal C};{\bf Q}).$ We can express $\zeta$ in terms of the generators \[ \{a_{r} | 1 \leq r \leq n \} \cup \{ b_{r}^{s} | 1 \leq r \leq n,1 \leq s \leq 2g \} \cup \{ f_{r} | 2 \leq r \leq n \} \] but equally we could write $\zeta$ in terms of \[ \{a_{r} | 1 \leq r \leq n \} \cup \{ n b_{r}^{s} -(n-r+1)a_{r-1}b_{1}^{s} | 2 \leq r \leq n,1 \leq s \leq 2g \} \] \begin{equation} \cup \{ n^{2} f_{r} - n(n-r+1)(\xi_{r-1,1} + \xi_{1,r-1}) + (n-r+1)(n-r+2) a_{r-2} \xi_{1,1} | 2 \leq r \leq n \} \label{1000} \end{equation} and $\{b_{1}^{s} | 1 \leq s \leq 2g\}.$ We shall take the coefficients of $\prod_{s \in S}b_{1}^{s}$ when $\zeta$ is expressed in this latter form. The reason for this is that the restrictions of the elements (\ref{1000}) in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ can then be written in terms of \begin{equation} \{a_{1}^{r} | 1 \leq r \leq n \} \cup \{b_{1}^{p,s}-b_{1}^{n,s} | 1 \leq p \leq n-1, 1 \leq s \leq 2g \} \label{1001} \end{equation} (see remark \ref{tedious}.) We can uniquely write the restriction of $\zeta$ in terms of the elements (\ref{1001}) and the restrictions of $b_{1}^{s},(1 \leq s \leq 2g)$. Hence we may calculate the restrictions of the coefficients of $\prod_{s \in S} b_{1}^{s}$ in $\zeta$ by taking the coefficients of \[ \prod_{s \in S} (b_{1}^{1,s} + \cdots + b_{1}^{n,s}) \] in the restriction of $\zeta$.\\ \indent We deal with a general type stratum in a similar way. Let $\mu = (d_{1}/n_{1},...,d_{P}/n_{P})$. We define formal symbols $a^{p,k},b^{p,k,s}$ and $d^{p,k}$ such that the $r$th Chern class $c_{r}({\cal V}_{p})$ is given by the $r$th elementary symmetric polynomial in \begin{equation} a^{p,k} + \sum_{s=1}^{2g} b^{p,k,s} \otimes \alpha_{s} + d^{p,k} \otimes \omega \indent (1 \leq k \leq n_{p}) \label{QQ} \end{equation} when $1 \leq r \leq n_{p}$ and $1 \leq p \leq P$. In terms of $a^{p,k},b^{p,k,s}$ and $d^{p,k}$ the restriction map to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ is formally the same as the restriction map when $\mu \in \Delta$. Again we may uniquely write the restriction of $\zeta$ in terms of \begin{equation} \bigcup_{p=1}^{P} \bigcup_{k=1}^{n_{p}} \{a^{p,k}, d^{p,k} \} \cup \bigcup_{p=1}^{P-1} \bigcup_{k= 1}^{n_{p}} \bigcup_{s=1}^{2g} \{ b^{p,k,s} - b^{P,n_{P},s} \} \cup \bigcup_{k= 1}^{n_{P}-1} \bigcup_{s=1}^{2g} \{ b^{P,k,s} - b^{P,n_{P},s} \} \label{NEW1} \end{equation} and the restrictions of $b_{1}^{s}, (1 \leq s \leq 2g)$, and we take the coefficients of \[ \prod_{s \in S}(b_{1}^{1,s} + \cdots + b_{1}^{P,s}) \] as before.\\ \indent So in our definitions of the Mumford and dual Mumford relations, (\ref{998}) and (\ref{999}), we assume first that $\sigma_{r}^{k}$ and $\tau_{r}^{k}$ have first been written in terms of the elements (\ref{1000}) before taking the appropriate coefficient. \begin{rem} \label{tedious} It is a trivial but tedious calculation to show that the restrictions of the elements (\ref{1000}) in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ for $\mu \in \Delta$ can indeed be written in terms of the elements (\ref{1001}). Let $a_{r}^{\mu}$ denote the restriction of $a_{r}$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$; this equals the $r$th elementary symmetric product in $a_{1}^{1},...,a_{1}^{n}$. The restrictions of $b_{r}^{s}$ and $f_{r}$ in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equal \[ \sum_{i=1}^{n} b_{1}^{i,s} \frac{\partial a_{r}^{\mu}}{\partial a_{1}^{i}}, \indent \sum_{i=1}^{n} d_{i} \frac{\partial a_{r}^{\mu}}{\partial a_{1}^{i}} + \sum_{i=1}^{n} \sum_{j=1}^{n} \xi_{1,1}^{i,j} \frac{\partial^{2} a_{r}^{\mu}}{\partial a_{1}^{i} \partial a_{1}^{j}}. \] The restrictions of the elements (\ref{1000}) can then be seen to equal \[ a_{r}^{\mu}, \indent \sum_{i=1}^{n-1} \left( n \frac{\partial a_{r}^{\mu}}{\partial a_{1}^{i}} - (n-r+1) a_{r-1}^{\mu} \right) ( b_{1}^{i,s} - b_{1}^{n,s}), \] and \[ n^{2} \sum_{i=1}^{n} d_{i} \frac{\partial a_{r}^{\mu}}{\partial a_{1}^{i}} + \sum_{i=1}^{n-1} \sum_{j=1}^{n-1} \sum_{s=1}^{g} (b_{1}^{i,s} - b_{1}^{n,s})(b_{1}^{j,s+g} - b_{1}^{n,s+g}) \left( n^{2} \frac{\partial^{2}a_{r}^{\mu}}{\partial a_{1}^{i} \partial a_{1}^{j}} \right. \] \[ \left. - n(n-r+1) \left(\frac{\partial a_{r-1}^{\mu}}{\partial a_{1}^{i}} + \frac{\partial a_{r-1}^{\mu}}{\partial a_{1}^{j}} \right) + (n-r+1)(n-r+2)a_{r-2}^{\mu} \right). \] \end{rem} \indent The remains of this section are given over to calculating the Mumford and dual Mumford relations. Our first problem is to obtain their generating functions from their respective Chern characters which we can evaluate using the Grothendieck-Riemann-Roch theorem (GRR). \begin{lem} \label{Chernlemma} Suppose that \begin{equation} {\rm ch}(E) = \sum_{i=1}^{M} \alpha_{i} e^{\delta_{i}} + \sum_{i=1}^{N} \beta_{i} e^{\epsilon_{i}} \label{230} \end{equation} where the $\beta_{i},\delta_{i}$ and the $\epsilon_{i}$ are formal degree two classes and the $\alpha_{i}$ are formal degree zero classes. Then as a formal power series \begin{equation} c(E)(t) = \sum_{r=0}^{\infty} c_{r}(E) \cdot t^{r} = \prod_{i=1}^{M} (1+\delta_{i}t)^{\alpha_{i}} \prod_{i=1}^{N} \exp \left\{ \frac{\beta_{i}t}{1+\epsilon_{i}t} \right\}. \label{231} \end{equation} \end{lem} {\bf Proof} The relationship between the Chern character and Chern polynomial is as follows. If $\mbox{ch}(E)= \sum_{i=1}^{K} e^{\gamma_{i}}$ where $\gamma_{i}$ are formal degree two classes then \[ c(E)(t) = \prod_{i=1}^{K} (1+\gamma_{i}t). \] If $\mbox{ch}(E)$ is in the form of (\ref{230}) then by comparing degrees we find that \[ \sum_{i=1}^{M} \alpha_{i}(\delta_{i})^{n} + \sum_{i=1}^{N} n \beta_{i} (\epsilon_{i})^{n-1} = \sum_{i=1}^{K} (\gamma_{i})^{n} \] for each $n \geq 0$. Thus on the level of formal power series $\log c(E)(t)$ equals \[ \sum_{i=1}^{K} \sum_{r=1}^{\infty} (-1)^{r+1} \frac{(\gamma_{i}t)^{r}}{r} = \sum_{i=1}^{M} \alpha_{i} \log(1+\delta_{i}t) + \sum_{i=1}^{N} \frac{\beta_{i}t}{1+\epsilon_{i}t} \] and hence the result (\ref{231}). $\indent \Box$.\\[\baselineskip] \indent Armed with the above lemma we are now in a position to determine the Chern polynomials $c(\pi_{!}{\cal V})(t)$ and $c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t)$. We can, and will, calculate these Chern polynomials in terms of the generators $a_{r},b_{r}^{s}$ and $f_{r}$ of $H^{*}_{\cal G}({\cal C};{\bf Q})$ (see (\ref{26}) and (\ref{210})). However the expressions obtained are somewhat cumbersome and for ease of calculation we will find the formal expressions, (\ref{25}) and (\ref{29}), calculated directly from the above lemma of more use. \begin{prop} \label{Chernprop} The Chern polynomial $c(\pi_{!}{\cal V})(t)$ equals \begin{equation} \Omega(t)^{-\bar{g}} \prod_{k=1}^{n} (1+\delta_{k}t)^{W_{k}} \exp \left\{ \frac{X_{k}t}{1+\delta_{k}t} \right\} \label{25} \end{equation} and $c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t)$ equals \begin{equation} \Omega(t)^{3\bar{g}+1} \prod_{k=1}^{n} (1+\delta_{k}t)^{-W_{k}} \exp \left\{ \frac{-X_{k}t}{1+\delta_{k}t} \right\}, \label{29} \end{equation} where $\delta_{1},...,\delta_{n}$ are formal degree two classes such that their $r$th elementary symmetric polynomial equals $a_{r},$ and \[ \Omega(t)= \prod_{k=1}^{n}(1+\delta_{k}t) = 1 + a_{1}t + \cdot \cdot \cdot + a_{n}t^{n}, \indent \xi_{i,j}= \sum_{s=1}^{g} b_{i}^{s} b_{j}^{s+g}, \] \[ W_{k}= \sum_{i=1}^{n} f_{i} \frac{\partial \delta_{k}}{\partial a_{i}} + \sum_{i=1}^{n} \sum_{j=1}^{n} \xi_{i,j} \frac{\partial^{2} \delta_{k}}{\partial a_{i} \partial a_{j}}, \indent X_{k} = \sum_{i=1}^{n} \sum_{j=1}^{n} \xi_{i,j} \frac{\partial \delta_{ k}}{\partial a_{i}} \frac{\partial \delta_{k}}{\partial a_{j}}. \] In terms of the generators $a_{r},b_{r}^{s}$ and $f_{r}$ for $H^{*}_{\cal G}({\cal C};{\bf Q})$ then $c(\pi_{!}{\cal V})(t)$ equals \begin{equation} \Omega(t)^{-\bar{g}} \exp \left\{ \int_{0}^{t} \left( \frac{d}{u} - \sum_{i=1}^{n} \frac{f_{i} u^{i-2}}{\Omega(u)} + \sum_{i=1}^{n} \sum_{j=1}^{n} \frac{\xi_{i,j}u^{i+j-2}}{\Omega(u)^{2}} \right) {\rm d}u \right\} \label{26} \end{equation} and $c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t)$ equals \begin{equation} \Omega(t)^{3\bar{g}+1} \exp \left\{ \int_{0}^{t} \left( -\frac{d}{u} + \sum_{i=1}^{n} \frac{f_{i} u^{i-2}}{\Omega(u)} - \sum_{i=1}^{n} \sum_{j=1}^{n} \frac{\xi_{i,j}u^{i+j-2}}{\Omega(u)^{2}} \right) {\rm d}u \right\}. \label{210} \end{equation} \end{prop} {\bf Proof} Now $\mbox{ch}({\cal V}) = e^{\gamma_{1}} + \cdot \cdot \cdot + e^{\gamma_{n}}$ where $\gamma_{1},...,\gamma_{n}$ are formal degree two classes such that their $r$th elementary symmetric polynomial equals \[ c_{r}({\cal V}) = a_{r} \otimes 1 + \sum_{s=1}^{2g} b_{r}^{s} \otimes \alpha_{s} + f_{r} \otimes \omega \indent (1 \leq r \leq n) . \] For each $k \geq 0$ there exist coefficients $\rho_{r_{1},...,r_{n}}^{(k)}$ such that \[ (\gamma_{1})^{k} + \cdot \cdot \cdot + (\gamma_{n})^{k} = \sum \rho_{r_{1},...,r_{n}}^{(k)} (c_{1}({\cal V}))^{r_{1}} \cdot \cdot \cdot (c_{n}({\cal V}))^{r_{n}} \] where the sum is taken over all non-negative $r_{1},...,r_{n}$ such that $r_{1}+2r_{2}+ \cdots +nr_{n} = k$. Now \[ (a_{1} \otimes 1 + \sum_{s=1}^{2g} b_{1}^{s} \otimes \alpha_{s} + f_{1} \otimes \omega)^{r_{1}} \cdot \cdot \cdot (a_{n} \otimes 1 + \sum_{s=1}^{2g} b_{n}^{s} \otimes \alpha_{s} + f_{n} \otimes \omega)^{r_{n}} \] equals \[ (a_{1})^{r_{1}} \cdot \cdot \cdot (a_{n})^{r_{n}} \otimes 1 + \sum_{i=1}^{n} \sum_{s=1}^{2g} b_{i}^{s} \frac{\partial}{\partial a_{i}} (a_{1})^{r_{1}} \cdot \cdot \cdot (a_{n})^{r_{n}} \otimes \alpha_{s} \] \[ +\sum_{i=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} (a_{1})^{r_{1}} \cdot \cdot \cdot (a_{n})^{r_{n}} \otimes \omega + \sum_{i=1}^{n} \sum_{j=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} (a_{1})^{r_{1}} \cdot \cdot \cdot (a_{n})^{r_{n}} \otimes \omega. \] Since \[ \sum \rho_{r_{1},...,r_{n}}^{(k)} (a_{1})^{r_{1}} \cdot \cdot \cdot (a_{n})^{r_{n}} = (\delta_{1})^{k} + \cdot \cdot \cdot + (\delta_{n})^{k} \] we find that $\mbox{ch}({\cal V})$ equals \[ \sum_{k=1}^{n} e^{\delta_{k}} \otimes 1 + \sum_{i=1}^{n} \sum_{s=1}^{2g} \sum_{k=1}^{n} b_{i}^{s} \frac{\partial}{\partial a_{i}} e^{\delta_{k}} \otimes \alpha_{s} \] \begin{equation} +\sum_{i=1}^{n} \sum_{k=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} e^{\delta_{k}} \otimes \omega + \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} e^{\delta_{k}} \otimes \omega. \label{212} \end{equation} {}From GRR we have $\mbox{ch}(\pi_{!}{\cal V}) = \pi_{*} ( \mbox{ch}({\cal V}) \cdot 1 \otimes (1-\bar{g} \omega))$ and hence $\mbox{ch}(\pi_{!}{\cal V})$ equals \[ \sum_{i=1}^{n} \sum_{k=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} e^{\delta_{k}} + \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} e^{\delta_{k}} - \bar{g} \sum_{k=1}^{n} e^{\delta_{k}} = \sum_{k=1}^{n} (-\bar{g} + W_{k} + X_{k}) e^{\delta_{k}}. \] Note that $W_{k}$ has degree zero and $X_{k}$ has degree two. Hence by lemma \ref{Chernlemma} we see that $c(\pi_{!}{\cal V})(t)$ equals \[ (\Omega(t))^{-\bar{g}} \prod_{k=1}^{n} (1+\delta_{k}t)^{W_{k}} \exp \left\{ \frac{X_{k}t}{1+\delta_{k}t} \right\} \] to give equation (\ref{25}).\\ \indent Now $\frac{{\rm d}}{{\rm d}t} \log ( \Omega(t)^{\bar{g}} c(\pi_{!}{\cal V})(t))$ equals \[ \sum_{i=1}^{n} \sum_{k=1}^{n} f_{i} \frac{\partial \delta_{k}}{\partial a_{i}} \frac{\delta_{k}}{1+\delta_{k}t} + \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \xi_{i,j} \left( \frac{\partial^{2} \delta_{k}}{\partial a_{i} \partial a_{j}} \frac{\delta_{k}}{1+\delta_{k} t} + \frac{\partial \delta_{k}}{\partial a_{i}} \frac{\partial \delta_{k}}{\partial a_{j}} \frac{1}{(1+ \delta_{k} t)^{2}} \right) \] \[ = \sum_{i=1}^{n} \sum_{j=1}^{n} \frac{\xi_{i,j}}{t^{2}} \left( \sum_{k=1}^{n} t \frac{\partial^{2} \delta_{k}}{\partial a_{i} \partial a_{j}} - \sum_{k=1}^{n} \left( \frac{t}{1+ \delta_{k}t} \frac{\partial^{2} \delta_{k}}{\partial a_{i} \partial a_{j}} - \frac{t^{2}}{(1+\delta_{k} t)^{2}} \frac{\partial \delta_{k}}{\partial a_{i}} \frac{\partial \delta_{k}}{\partial a_{j}} \right) \right) \] \begin{equation} + \sum_{i=1}^{n} \frac{f_{i}}{t} \left( \sum_{k=1}^{n} \frac{\partial \delta_{k}}{\partial a_{i}} - \sum_{k=1}^{n} \frac{\partial \delta_{k}}{\partial a_{i}} \frac{1}{1+ \delta_{k} t} \right) \label{219} \end{equation} Since $\sum_{k=1}^{n} \frac{\partial \delta_{k}}{\partial a_{i}} = \frac{ \partial a_{1}}{\partial a_{i}}$, $f_{1}=d$, and $\sum_{k=1}^{n} \frac{\partial^{2} \delta_{k}}{\partial a_{i} \partial a_{j}} = \frac{\partial^{2} a_{1}}{\partial a_{i} \partial a_{j}} = 0$ then (\ref{219}) reduces to \[ \frac{d}{t} - \sum_{i=1}^{n} \sum_{k=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} \frac{\log (1+ \delta_{k} t)}{t^{2}} - \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} \frac{\log (1 + \delta_{k}t)}{t^{2}} \] \[ = \frac{d}{t} - \left( \sum_{i=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} + \sum_{i=1}^{n} \sum_{j=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} \right) \frac{\log \Omega(t)}{t^{2}} \] to give equality (\ref{26}).\\ \indent The calculations for the dual case follow in a similar fashion. We have that $\mbox{ch}({\cal V}^{*}) = e^{-\gamma_{1}} + \cdot \cdot \cdot + e^{-\gamma_{n}}$ with $\gamma_{1},...,\gamma_{n}$ as before and arguing as in the calculation of (\ref{212}) we determine that $\mbox{ch}({\cal V}^{*})$ equals \[ \sum_{k=1}^{n} e^{-\delta_{k}} \otimes 1 + \sum_{i=1}^{n} \sum_{s=1}^{2g} \sum_{k=1}^{n} b_{i}^{s} \frac{\partial}{\partial a_{i}} e^{-\delta_{k}} \otimes \alpha_{s} \] \begin{equation} +\sum_{i=1}^{n} \sum_{k=1}^{n} f_{i} \frac{\partial}{\partial a_{i}} e^{-\delta_{k}} \otimes \omega + \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \xi_{i,j} \frac{\partial^{2}}{\partial a_{i} \partial a_{j}} e^{-\delta_{k}} \otimes \omega. \label{213} \end{equation} We know that $\mbox{ch}(\phi^{*}L) = \phi^{*}(e^{(4\bar{g}+1)\omega}) = 1 \otimes (1+ (4\bar{g}+1)\omega)$ and GRR shows that $\mbox{ch}(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))$ equals \[ \pi_{*}(\mbox{ch}({\cal V}^{*}) \cdot \mbox{ch}(\phi^{*}L) \cdot 1 \otimes (1 - \bar{g} \omega)) = \pi_{*}(\mbox{ch}({\cal V}^{*}) \cdot 1 \otimes (1 + (3\bar{g}+1) \omega)) \] which gives \begin{equation} \mbox{ch}(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L)) = \sum_{k=1}^{n} ( (3\bar{g}+1) -W_{k} +X_{k} ) e^{-\delta_{k}}. \label{215} \end{equation} Applying lemma \ref{Chernlemma} to expression (\ref{215}) gives equation (\ref{29}). Expression (\ref{210}) is arrived at by calculating $\frac{{\rm d}}{{\rm d}t} \log ((\Omega(t))^{-3\bar{g}-1}c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(t))$ and grouping the terms in a similar manner to expression (\ref{219}). $\indent \Box$ \begin{rem} Note that $\delta_{k}, W_{k}$ and $X_{k}$ are not elements of $H^{*}_{\cal G}({\cal C};{\bf Q})$. However the direct sum of the restriction maps \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow \bigoplus_{\mu \in \Delta} H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] is injective and so we may consider $\delta_{k},W_{k}$ and $X_{k}$ as elements of $\bigoplus_{\mu \in \Delta} H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ corresponding respectively to $a_{1}^{k},d_{k}$ and $\xi_{1,1}^{k,k}$ in each summand $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}).$ \end{rem} \begin{rem} From (\ref{26}) we can find an expression for \[ \frac{\Psi'(t)}{\Psi(t)} = \frac{d-n\bar{g}}{t} - \frac{c(\pi_{!}{\cal V})'(t^{-1})}{t^{2} c(\pi_{!}{\cal V})(t^{-1})}. \] In fact we may write $\Psi'(t)/\Psi(t)$ as a rational function with denominator $(\tilde{\Omega}(t))^{2}$ and a numerator of degree at most $2n-1$. By multiplying by $\Psi(t)$ and comparing coefficients of $t^{k} (\tilde{\Omega}(t))^{r}, (r \leq \bar{g}, 0 \leq k <n)$ we may derive recurrence relations amongst the Mumford relations which determine $\{\sigma_{r}^{k} : 0 \leq k < n\}$ in terms of $\{\sigma_{r+1}^{k},\sigma_{r+2}^{k} : 0 \leq k < n \}$. Similar recurrence relations exist among the dual Mumford relations which determine $\{\tau_{r}^{k} : 0 \leq k < n\}$ in terms of $\{\tau_{r+1}^{k},\tau_{r+2}^{k} : 0 \leq k < n \}$. \end{rem} \indent The calculation of the restriction of $c(\pi_{!}{\cal V})(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})[[t]]$ follows easily from the previous proposition. As in \cite[prop. 2]{K2} this restriction can be expressed in terms of elementary functions of the generators of $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ when $\mu \in \Delta$. However for a general type $\mu$ this restriction cannot be expressed so easily and we will find formal expressions similar to (\ref{25}) of more use. \begin{cor} \label{Chernresn} Let $\mu= (d_{1}/n_{1},...,d_{P}/n_{P})$. The restriction to $H_{\cal G}^{*}({\cal C}_{\mu};{\bf Q})[[t]]$ of $c(\pi_{!}{\cal V})(t)$ equals the formal power series \begin{equation} \Omega_{\mu}(t)^{-\bar{g}} \prod_{p=1}^{P} \prod_{k=1}^{n_{p}} (1+\delta_{k}^{p}t)^{W_{k}^{p}} \exp \left\{ \frac{X_{k}^{p}t}{1+\delta_{k}^{p}t} \right\} \label{27} \end{equation} and similarly the restriction of $c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t)$ to $H_{\cal G}^{*}({\cal C}_{\mu};{\bf Q})[[t]]$ equals \begin{equation} \Omega_{\mu}(t)^{3\bar{g}+1} \prod_{p=1}^{P} \prod_{k=1}^{n_{p}} (1+\delta_{k}^{p}t)^{-W_{k}^{p}} \exp \left\{ \frac{-X_{k}^{p}t}{1+\delta_{k}^{p}t} \right\}, \label{211} \end{equation} where $\delta_{1}^{p},...,\delta_{n_{p}}^{p}$ are formal degree two classes such that their $r$th elementary symmetric polynomial equals $a_{r}^{p}$, where $\Omega_{\mu}(t)= \prod_{p=1}^{P} \prod_{k=1}^{n_{p}} (1+ \delta_{k}^{p}t)$ is the restriction of $\Omega(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})[t],$ and where $\xi_{i,j}^{p,p},W_{k}^{p}$ and $X_{k}^{p}$ correspond to the expressions defined in the statement of proposition \ref{Chernprop}. \end{cor} {\bf Proof} Expression (\ref{27}) is immediate from the previous proposition once we note that the restriction of $\mbox{ch}(\pi_{!}{\cal V})$ to $H^{*}_{\cal G} ( {\cal C}_{\mu};{\bf Q})$ equals \[ \sum_{p=1}^{P} \pi_{*}(\mbox{ch}({\cal V}_{p}) \cdot 1 \otimes (1-\bar{g} \omega)) \] and recall that the Chern polynomial is multiplicative. The dual expression (\ref{211}) follows in a similar fashion. $\Box$ \begin{cor} \label{Deltaresn} Let $\mu=(d_{1},...,d_{n}) \in \Delta$. Then the restriction of $c(\pi_{!}{\cal V})(t)$ to $H^{*}_{\cal G} ({\cal C}_{\mu};{\bf Q})[[t]]$ equals \[ \prod_{p=1}^{n} (1+ a_{1}^{p} t)^{d_{p} -\bar{g}} \exp \left\{ \frac{\xi_{1,1}^{p,p} t}{1 + a_{1}^{p} t} \right\}. \] Also the restriction of $c(\pi_{!}({\cal V}^{*} \otimes \phi^{*}L))(-t)$ to $H^{*}_{\cal G} ({\cal C}_{\mu};{\bf Q})[[t]]$ equals \[ \prod_{p=1}^{n} (1+ a_{1}^{p} t)^{3\bar{g}+1-d_{p}} \exp \left\{ \frac{-\xi_{1,1}^{p,p} t}{1 + a_{1}^{p} t} \right\}. \] \end{cor} {\bf Proof} Simply note that in this case $\delta_{1}^{p} = a_{1}^{p},W_{1}^{p} = d_{p}$ and $X_{1}^{p}=\xi_{1,1}^{p,p}. \indent \Box$ \begin{rem} Let $\mu=(d_{1}/n_{1},...,d_{P}/n_{P})$. From the calculation (\ref{212}) and since the Chern character is additive we know that the restriction of $\mbox{ch}({\cal V})$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ \sum_{p=1}^{P} \sum_{k=1}^{n_{p}} \exp \left\{ \delta_{k}^{p} + \sum_{s=1}^{2g} \left( \sum_{i=1}^{n_{p}} b_{i}^{p,s}\frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} \right) \otimes \alpha_{s} + W_{k}^{p} \otimes \omega \right\}. \] Thus in terms of our earlier notation (\ref{QQ}) we have \[ a^{p,k} = \delta_{k}^{p}, \indent b^{p,k,s} = \sum_{i=1}^{n_{p}} b_{i}^{p,s} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}}, \indent d^{p,k}=W_{k}^{p}. \] \end{rem} \indent We end this section with two further calculations, namely the Chern polynomials of the normal bundle ${\cal N}_{\mu}$ to the stratum ${\cal C}_{\mu}$ in ${\cal C}$ (necessary to the completeness criteria) and of the tangent bundle $T$ to the moduli space ${\cal M}(n,d)$ (needed for generalising the proof of the Newstead-Ramanan conjecture). \begin{lem} \label{normal} Let $\mu=(d_{1}/n_{1},...,d_{P}/n_{P})$. Then the Chern polynomial $c({\cal N}_{\mu})(t)$ of the normal bundle in ${\cal C}$ to the stratum ${\cal C}_{\mu}$ equals \begin{equation} {\cal P}_{\mu}(t)^{\bar{g}} \prod_{I<J} \prod_{k=1}^{n_{I}} \prod_{l=1}^{n_{J}} (1+(\delta_{l}^{J}-\delta_{k}^{I})t)^{W_{k}^{I}-W_{l}^{J}} \exp \left\{ \frac{-\Xi_{k,l}^{I,J}t}{1+(\delta_{l}^{J}-\delta_{k}^{I})t} \right\} \label{216} \end{equation} where \[ \Xi_{k,l}^{I,J}= \sum_{s=1}^{g} \left( \sum_{i=1}^{n_{I}} b_{i}^{I,s} \frac{\partial \delta_{k}^{I}}{\partial a_{i}^{I}} -\sum_{j=1}^{n_{J}} b_{j}^{J,s} \frac{\partial \delta_{l}^{J}}{\partial a_{j}^{J}} \right) \left( \sum_{i=1}^{n_{I}} b_{i}^{I,s+g} \frac{\partial \delta_{k}^{I}}{\partial a_{i}^{I}} -\sum_{j=1}^{n_{J}} b_{j}^{J,s+g} \frac{\partial \delta_{l}^{J}}{\partial a_{j}^{J}} \right) \] and \[ {\cal P}_{\mu}(t)= \prod_{I<J} \prod_{k=1}^{n_{I}} \prod_{l=1}^{n_{J}} (1+(\delta_{l}^{J} - \delta_{k}^{I})t). \] \end{lem} {\bf Proof} Kirwan \cite[lemma 2]{K2} showed that the normal bundle ${\cal N}_{\mu}$ to ${\cal C}_{\mu}$ in ${\cal C}$, equals \[ -\pi_{!} \left( \bigoplus_{I < J} {\cal V}^{*}_{I} \otimes {\cal V}_{J} \right). \] {}From the proof of the proposition \ref{Chernprop} we can find expressions for $\mbox{ch}({\cal V}_{J})$ and $\mbox{ch}({\cal V}_{I}^{*})$ corresponding to (\ref{212}) and (\ref{213}). The GRR implies that \[ \mbox{ch}({\cal N}_{\mu}) = \sum_{I<J} \pi_{*} ( \mbox{ch}({\cal V}_{I}^{*}) \cdot \mbox{ch}({\cal V}_{J}) \cdot 1 \otimes (\bar{g} \omega -1)). \] Substituting in these expressions for $\mbox{ch}({\cal V}_{J})$ and $\mbox{ch}({\cal V}_{I}^{*})$ we find that $\mbox{ch}({\cal N}_{\mu})$ equals \[ \sum_{I<J} \left\{ \sum_{k=1}^{n_{I}} \sum_{l=1}^{n_{J}} (\bar{g}+W_{k}^{I}-W_{l}^{J}-\Xi_{k,l}^{I,J})e^{\delta_{l}^{J}-\delta_{k}^{I}} \right\}. \] Applying lemma \ref{Chernlemma} produces the required result (\ref{216}). $\indent \Box$ \begin{lem} \label{Pont} The total Pontryagin class of ${\cal M}(n,d)$ equals \[ \prod_{1 \leq k < l \leq n} (1 + (\delta_{k} - \delta_{l})^{2})^{2\bar{g}}. \] In particular the Pontryagin ring of ${\cal M}(n,d)$ is generated by the elementary symmetric polynomials in \[ \{ ( \delta_{k} -\delta_{l}) ^{2} : 1 \leq k < l \leq n \}. \] \end{lem} {\bf Proof} Let $T$ denote the tangent bundle of ${\cal M}(n,d)$. From \cite[p.582]{AB} we know that \[ T + T^{*} -2 = \pi_{!}({\rm End} V \otimes (\Omega_{M}^{1}-1)). \] Applying GRR we find \[ \mbox{ch} T + \mbox{ch} T^{*} - 2 = 2 \bar{g} \mbox{ch} ({\rm End V}|{\cal M}(n,d)) \] which we know to equal \[ 2 \bar{g} \left(\sum_{k=1}^{n} e^{\delta_{k}} \right) \left( \sum_{l=1}^{n} e^{-\delta_{l}} \right) \] from expressions (\ref{212}) and (\ref{213}).\\ \indent Now let $p(T)(t) = \sum_{r \geq 0} p_{r}(T) t^{r}$ denote the Pontryagin polynomial. The relationship between the Pontryagin classes and the Chern classes is given by \[ p(T)(-1) = c(T)(1) \cdot c(T)(-1) \indent \cite[\mbox{Cor. } 15.5]{MS}. \] Hence $p(T)(-1)$ equals \[ \prod_{k \neq l} (1+ \delta_{k} -\delta_{l})^{2\bar{g}} = \prod_{k<l} (1 - (\delta_{k} - \delta_{l})^{2})^{2\bar{g}}. \] The total Pontryagin class of ${\cal M}(n,d)$ then equals $p(T)(1)$ and hence the result. $\Box$ \section{A Complete Set of Relations.} Whilst we observed in remark \ref{inadequacy} that neither the Mumford relations nor the dual Mumford relations are in themselves a complete set of relations when the rank is greater than two, it is still possible to put these relations into the context of the completeness criteria. In terms of these criteria we will show how the Mumford relations contain subsets corresponding to all strata of the form \[ \mu= (d_{1}/n_{1},...,d_{P}/n_{P}) \] where $n_{P}=1.$ Similarly the dual Mumford relations contain subsets corresponding to all those strata with $n_{1}=1.$ From this we shall deduce that in the rank three case the Mumford and dual Mumford relations form a complete set.\\ \indent Before we continue with the main proposition we need a lemma on the vanishing of the Mumford and dual Mumford relations on restriction to a stratum. \begin{lem} \label{vanishing} Let $\mu=(d_{1}/n_{1},...,d_{P}/n_{P})$. The image of the Mumford relation $\sigma_{r,S}^{k}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] vanishes when $r < d_{P}/n_{P} -2g +1$. The image of the dual Mumford relation $\tau_{r,S}^{k}$ under the restriction map vanishes when $r < 2\bar{g}-d_{1}/n_{1}.$ \end{lem} {\bf Proof} Recall that the Mumford relations are given by $\sigma_{r,S}^{k} (r<0,0 \leq k \leq n-1,S \subseteq \{1,...,2g\})$ when $\Psi(t)= t^{d-n\bar{g}}c(\pi_{!}{\cal V})(t^{-1})$ is written in the form \[ \sum_{r=-\infty}^{\bar{g}} (\sigma_{r}^{0} + \sigma_{r}^{1} t + \cdot \cdot \cdot + \sigma_{r}^{n-1} t^{n-1})(\tilde{\Omega}(t))^{r}, \indent \sigma_{r}^{k}=\sum_{S \subseteq \{1,...,2g\}} \sigma_{r,S}^{k} \prod_{s \in S} b_{1}^{s}. \] For $1 \leq k \leq n$ and any fixed integer $R$ the power $t^{-k}$ appears in \[ \sum_{r=-\infty}^{\bar{g}} (\sigma_{r}^{0} + \sigma_{r}^{1} t + \cdot \cdot \cdot + \sigma_{r}^{n-1} t^{n-1})(\tilde{\Omega}(t))^{r-R-1} \] only when $r=R.$ Let $C_{r}^{i}$ denote the coefficient of $t^{-i}$ in $\Psi(t)(\tilde{\Omega}(t))^{-r-1}$. Then \[ (\sigma_{r}^{0} + \sigma_{r}^{1} t + \cdot \cdot \cdot + \sigma_{r}^{n-1} t^{n-1}) = (t^{n} + a_{1} t^{n-1} + \cdot \cdot \cdot + a_{n}) \sum_{i=1}^{n} C_{r}^{i} t^{-i} \] modulo negative powers of $t$ and hence \begin{equation} \sigma_{r}^{n-k} = \sum_{i=1}^{k} a_{k-i} C_{r}^{i} \indent (r<0,1 \leq k \leq n) \label{31}. \end{equation} \indent Now let $K$ be a fixed line bundle over $M$ of degree $D$ where $D$ is the smallest integer such that \[ \mu(Q_{P} \otimes K) = \frac{d_{P}}{n_{P}} + D > 2\bar{g} \] where $Q_{P}=E_{P}/E_{P-1}.$ Since $\mu(Q_{p} \otimes K) \geq \mu(Q_{P} \otimes K) > 2\bar{g}$ then $\pi_{!}({\cal V}_{p} \otimes \phi^{*}K)$ is a bundle over ${\cal C}(n_{p},d_{p})^{ss}$ of rank $d_{p} +(D-\bar{g})n_{p}$ for each $1 \leq p \leq P.$ In particular \[ \Psi(\pi_{!}({\cal V}_{p} \otimes \phi^{*}K))(t) = t^{d_{p}+n_{p}(D-\bar{g})} c (\pi_{!}({\cal V}_{p} \otimes \phi^{*}K))(t^{-1}) \] is a polynomial modulo relations in $H^{*}_{{\cal G}(n_{p},d_{p})} ({\cal C}(n_{p},d_{p})^{ss};{\bf Q})$. From GRR we have that $\mbox{ch}(\pi_{!}({\cal V}_{p} \otimes \phi^{*}K))$ equals \begin{equation} \mbox{ch}(\pi_{!}{\cal V}_{p}) + \pi_{*}(\mbox{ch} {\cal V}_{p} \cdot 1 \otimes D\omega) = \mbox{ch}(\pi_{!}{\cal V}_{p}) +D \sum_{k=1}^{n_{p}} e^{\delta_{k}^{p}}. \label{32} \end{equation} In terms of Chern polynomials (\ref{32}) gives \[ c(\pi_{!}({\cal V}_{p} \otimes \phi^{*}K))(t) = (\Omega_{p}(t))^{D} c(\pi_{!}{\cal V}_{p})(t) \] where $\Omega_{p}(t) = \prod_{k=1}^{n_{p}} ( 1 + \delta_{k}^{p}t)$. Hence \begin{equation} \prod_{p=1}^{P} \Psi(\pi_{!}({\cal V}_{p} \otimes \phi^{*}K))(t) = (\tilde{\Omega}_{\mu}(t))^{D} \Psi_{\mu}(t) \label{51} \end{equation} is a polynomial modulo relations in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ where $\Psi_{\mu}(t),$ and $\tilde{\Omega}_{\mu}(t)$ are respectively the restrictions to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ of $\Psi(t)$ and $\tilde{\Omega}(t)$. Thus the coefficient of $ t^{-k}$ in $\Psi_{\mu}(t)\tilde{\Omega}_{\mu}(t)^{-r-1}$ is a relation when $r \leq -1 -D$. So by (\ref{31}) the restriction of $\sigma_{r}^{k}$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ vanishes when $r \leq d_{P}/n_{P} - 2g.$ The dual calculation follows by a similar argument. $\indent \Box$\\[\baselineskip] \indent Thus finally we come to \begin{prop} \label{biggy} Let $\mu = (d_{1}/n_{1},...,d_{P}/n_{P})$ with $n_{P}=1$. Then there is a subset ${\cal R}_{\mu}$ of the ideal generated by the Mumford relations such that the image of the ideal generated by ${\cal R}_{\mu}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \indent \nu=(\tilde{d}_{1}/\tilde{n}_{1},...,\tilde{d}_{T}/\tilde{n}_{T}) \] is zero when either \[ \mbox{(i) } \tilde{d}_{T}/\tilde{n}_{T}> d_{P} \indent \mbox{or} \indent \mbox{(ii) } \tilde{n}_{T}=1, \tilde{d}_{T} = d_{P}, \mbox{ and } \nu \not \geq \mu \] and contains the ideal of $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ generated by $e_{\mu}$ when $\nu=\mu.$\\ \indent Let $\mu = (d_{1}/n_{1},...,d_{P}/n_{P})$ with $n_{1}=1$. Then there is a subset ${\cal R}_{\mu}$ of the ideal generated by the dual Mumford relations such that the image of the ideal generated by ${\cal R}_{\mu}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \indent \nu=(\tilde{d}_{1}/\tilde{n}_{1},...,\tilde{d}_{T}/\tilde{n}_{T}) \] is zero when either \[ \mbox{(i) } \tilde{d}_{1}/\tilde{n}_{1}< d_{1}/n_{1} \indent \mbox{or} \indent \mbox{(ii) } \tilde{n}_{1}=1, \tilde{d}_{1} = d_{1} \mbox{ and } \nu \not \geq \mu \] and contains the ideal of $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ generated by $e_{\mu}$ when $\nu=\mu.$ \end{prop} {\bf Proof} Let $\Psi(t)= t^{d-n\bar{g}}c(\pi_{!}{\cal V})(t^{-1})$ and let $C^{R}_{K}, (R<0,1 \leq K \leq n)$ denote the coefficient of $t^{-K}$ in $\Psi(t)(\tilde{\Omega}(t))^{-R-1}$. Let \[ \mu=(d_{1}/n_{1},...,d_{P-1}/n_{P-1},d_{P}) \] so that $n_{P}=1.$\\ \indent Since the Chern polynomial is multiplicative the restriction in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ of $C_{R}^{K}$, which we will write as $C_{R}^{K,\mu}$, equals the coefficient of $t^{-1}$ in \begin{equation} t^{K-1} \prod_{p=1}^{P} \Psi_{p}(t)(\tilde{\Omega}_{p}(t))^{-R-1} \label{50} \end{equation} where \[ \Psi_{p}(t)= t^{d_{p}-n_{p}\bar{g}}c(\pi_{!}{\cal V}_{p})(t^{-1}), \indent \tilde{\Omega}_{p}(t)=t^{n_{p}}+a_{1}^{p} t^{n_{p}-1}+ \cdot \cdot \cdot + a_{n_{p}}^{p} \] for $1 \leq p \leq P$. Further from the previous lemma we know that $C_{R}^{K,\mu}$ vanishes when $R<-D=d_{P}-2g+1.$\\ \indent We facilitate the proof of proposition \ref{biggy} with the following lemma and corollaries \begin{lem} Let $\theta(t)$ equal \begin{equation} t^{d-nd_{P}+(n-1)\bar{g}} \prod_{p=1}^{P-1} \prod_{k=1}^{n_{p}} ( 1 +(\delta_{k}^{p}-a_{1}^{P})/t)^{W_{k}^{p}+\bar{g}-d_{P}} \exp \left\{ \frac{\Xi_{k,1}^{p,P}}{t+\delta_{k}^{p}-a_{1}^{P}} \right\}. \label{52} \end{equation} Then modulo relations in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$, \[ C_{-D}^{K,\mu}= (-a_{1}^{P})^{K-1} (\xi_{1,1}^{P,P})^{g} \Theta \] where $\Theta$ is the constant coefficient of $\theta(t).$ \end{lem} {\bf Proof} From corollary \ref{Deltaresn} we know that \[ \Psi_{P}(t) (\tilde{\Omega}_{P}(t))^{D-1} = (t+a_{1}^{P})^{\bar{g}} \exp \left\{ \frac{ \xi_{1,1}^{P,P}}{t+a_{1}^{P}} \right\} \] where $\xi_{1,1}^{P,P} = \sum_{s=1}^{g} b_{1}^{P,s} b_{1}^{P,s+g}.$ Also in a Laurent series the coefficient of $t^{-1}$ is invariant under transformations such as $t \mapsto t-a_{1}^{P}.$ So from (\ref{50}) $C_{-D}^{K,\mu}$ equals the coefficient of $t^{-1}$ in \begin{equation} (t-a_{1}^{P})^{K-1} t^{\bar{g}} \exp (\xi_{1,1}^{P,P}/t) \prod_{p=1}^{P-1} \Psi_{p}(t-a_{1}^{P})(\tilde{\Omega}_{p}(t-a_{1}^{P}))^{D-1}. \label{D} \end{equation} \indent From the proof of lemma \ref{vanishing} (\ref{51}) we know that \[ \Psi_{p}(t)(\tilde{\Omega}_{p}(t))^{D-1} = \Psi(\pi_{!}({\cal V}_{p} \otimes \phi^{*}{\cal L}))(t) \] where ${\cal L}$ is a fixed line bundle over $M$ of degree $D-1.$ For each $p \neq P$, $Q_{p} \otimes {\cal L}$ is a semistable bundle of slope \[ \frac{d_{p}}{n_{p}} - d_{P} + 2\bar{g} > 2\bar{g}. \] Hence $\pi_{!}({\cal V}_{p} \otimes \phi^{*}{\cal L})$ is a bundle over ${\cal C}(n_{p},d_{p})^{ss}$ and $\Psi_{p}(t)(\tilde{\Omega}_{p}(t))^{D-1}$ is a polynomial modulo relations in $H^{*}_{{\cal G} (n_{p},d_{p})}({\cal C}(n_{p},d_{p})^{ss};{\bf Q}).$ As $(\xi_{1,1}^{P.P})^{g+1} = 0$ it follows from (\ref{D}) that $C_{-D}^{K,\mu}$ equals the constant coefficient of \begin{equation} (\xi_{1,1}^{P,P})^{g} (t-a_{1}^{P})^{K-1} \prod_{p=1}^{P-1} \Psi_{p}(t-a_{1}^{P}) (\tilde{\Omega}_{p}(t-a_{1}^{P}))^{D-1} \label{E} \end{equation} modulo relations in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$.\\ \indent Since $\sum_{k=1}^{n_{p}} W_{k}^{p} =d_{p}$ then we know from corollary \ref{Chernresn} that $\Psi_{p}(t-a_{1}^{P})$ equals \[ (\tilde{\Omega}_{p}(t-a_{1}^{P}))^{-\bar{g}} t^{d_{p}} \prod_{k=1}^{n_{p}} ( 1 +( \delta_{k}^{p} -a_{1}^{P})/t)^{W_{k}^{p}} \exp \left\{ \frac{X_{k}^{p}}{t+\delta_{k}^{p} -a_{1}^{P}} \right\}. \] Recall from lemma \ref{normal} that \[ \Xi_{k,1}^{p,P} = \sum_{s=1}^{g} \left( \sum_{i=1}^{n_{p}} b_{i}^{p,s} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} - b_{1}^{P,s} \right) \left( \sum_{i=1}^{n_{p}} b_{i}^{p,s+g} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} - b_{1}^{P,s+g} \right) \] and we also have that \[ X_{k}^{p} = \sum_{s=1}^{g} \left( \sum_{i=1}^{n_{p}} b_{i}^{p,s} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} \right) \left( \sum_{i=1}^{n_{p}} b_{i}^{p,s+g} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} \right). \] Since \[ (\xi_{1,1}^{P,P})^{g} = (-1)^{g\bar{g}/2} g! \prod_{s=1}^{2g} b_{1}^{P,s} \] then \[ (\xi_{1,1}^{P,P})^{g}(\Xi_{k,1}^{p,P})^{q} = (\xi_{1,1}^{P,P})^{g}(X_{k}^{p})^{q} \indent (q \geq 0). \] Thus by (\ref{E}) and the identity $\tilde{\Omega}_{p}(t-a_{1}^{P}) = t^{n_{p}}\prod_{k=1}^{n_{p}}(1+(\delta_{k}^{p}-a_{1}^{P})/t),$ we have that $C_{-D}^{K,\mu}$ equals the constant coefficient of \[ (\xi_{1,1}^{P,P})^{g} (t-a_{1}^{P})^{K-1} \theta(t). \] Since $(\xi_{1,1}^{P,P})^{g}\theta(t)$ is a polynomial modulo relations in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ then the lemma follows. $\indent \Box.$ \begin{cor} Define $C_{R,S}^{K} (R<0,1 \leq K \leq n, S \subseteq \{1,...,2g\})$ by \[ C_{R}^{K} = \sum_{S \subseteq \{1,...,2g\}} C_{R,S}^{K} \prod_{s \in S} b_{1}^{s} \] writing $C_{R,S}^{K}$ in terms of the elements (\ref{1000}) and also define $\tilde{a}_{r},\tilde{b}_{r}^{s}$ and $\tilde{f}_{r}$ by \[ c_{r}(\bigoplus_{p=1}^{P-1} {\cal V}_{p}) = \tilde{a}_{r} \otimes 1 + \sum_{s=1}^{2g} \tilde{b}_{r}^{s} \otimes \alpha_{s} + \tilde{f}_{r} \otimes \omega. \] Then the restriction of $C_{-D,S}^{K}$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals a non-zero constant multiple of \begin{equation} (a_{1}^{P})^{K-1} \prod_{s \not \in S} (\tilde{b}_{1}^{s}-(n-1) b_{1}^{P,s}) \Theta \label{M} \end{equation} for any subset $S \subseteq \{1,...,2g\}.$ \end{cor} {\bf Proof} We know that $(\xi_{1,1}^{P,P})^{g}$ equals \[ (-1)^{g\bar{g}/2} g! \prod_{s=1}^{2g} b_{1}^{P,s} = (-1)^{g\bar{g}/2} n^{-2g} g! \prod_{s=1}^{2g} ((\tilde{b}_{1}^{s} + b_{1}^{P,s})-(\tilde{b}_{1}^{s}-(n-1) b_{1}^{P,s})) \] and also that the restriction of $b_{1}^{s}$ in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals $\tilde{b}_{1}^{s}+ b_{1}^{P,s}.$ Further \[ \tilde{b}_{1}^{s} - (n-1)b_{1}^{P,s} = \sum_{p=1}^{P-1} \sum_{k=1}^{n_{p}} \left( \sum_{i=1}^{n_{p}} b_{1}^{p,s} \frac{\partial \delta_{k}^{p}}{\partial a_{i}^{p}} - b_{1}^{P,s} \right). \] So the corollary follows once we note from (\ref{52}) that $\theta(t)$, and hence $\Theta$, can be written in terms of the elements (\ref{NEW1}). $\indent \Box$. \begin{cor} Let $\Lambda$ equal \begin{equation} \bigcup \{ \sigma_{-D,S}^{n-1},...,\sigma_{-D,S}^{0} \} \label{54} \end{equation} where the union varies over all subsets $S \subseteq \{1,...,2g\}$. Then all elements of the form \begin{equation} \prod_{k=2}^{n-1} (\tilde{f}_{k})^{m_{k}} \prod_{k=1}^{n-1} \prod_{s \in S_{k}} \tilde{b}_{k}^{s} \prod_{k=1}^{n-1} (\tilde{a}_{k})^{r_{k}} (a_{1}^{P})^{r} \prod_{s \in S} b_{1}^{P,s} \Theta \label{81} \end{equation} lie in the restriction of the ideal generated by $\Lambda$, where $r,r_{1},...,r_{n-1},m_{2},...,m_{n-1}$ are arbitrary non-negative integers and $S,S_{1},...,S_{n-1}$ are subsets of $\{1,...,2g\}$. \end{cor} {\bf Proof} Let $(\Lambda)$ denote the ideal of $H^{*}_{\cal G}({\cal C};{\bf Q})$ generated by $\Lambda$. Using induction on (\ref{31}) we know that the restriction of $C_{-D,S}^{K}$ lies in the image of $(\Lambda)$. From (\ref{M}) and since $b_{1}^{s}$ restricts to $\tilde{b}_{1}^{s} + b_{1}^{P,s}$ it follows that all elements of the form \[ (a_{1}^{P})^{K-1} \prod_{s \in S_{1}} \tilde{b}_{1}^{s} \prod_{s \in S_{2}} b_{1}^{P,s} \Theta \] for arbitrary $S_{1},S_{2} \subseteq \{1,...,2g\}$ and $1 \leq K \leq n$, lie in the restriction of $(\Lambda).$ The restriction of $a_{k}$ in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals $\tilde{a}_{k}+\tilde{a}_{k-1}a_{1}^{P}$. By noting that $(a_{1}^{P})^{r}$ equals \[ (\tilde{a}_{1}+a_{1}^{P})(a_{1}^{P})^{r-1} -(\tilde{a}_{2} + \tilde{a}_{1} a_{1}^{P})(a_{1}^{P})^{r-2} + \cdot \cdot \cdot + (-1)^{n-1}(\tilde{a}_{n-1}a_{1}^{P})(a_{1}^{P})^{r-n} \] for $r \geq n$ we see that all elements of the form \[ (a_{1}^{P})^{r} \prod_{s \in S_{1}} \tilde{b}_{1}^{s} \prod_{s \in S_{2}} b_{1}^{P,s} \cdot \Theta \indent (r \geq 0) \] lie in the restriction of $(\Lambda)$. Finally working inductively on the variables $r_{1},...,r_{n-1},$ $S_{2},S_{3},...,S_{n-1}$ and $m_{2},m_{3},...m_{n-1}$ in that order we find that all elements of the form (\ref{81}) lie in the image of $(\Lambda)$ since under the restriction map $H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ \[ a_{k} \mapsto \tilde{a}_{k}+\tilde{a}_{k-1}a_{1}^{P} \indent b_{k}^{s} \mapsto \tilde{b}_{k}^{s}+ a_{1}^{P} \tilde{b}_{k-1}^{s} + \tilde{a}_{k-1}b_{1}^{P,s} \] and \begin{equation} f_{k} \mapsto \tilde{f}_{k} + d_{P}\tilde{a}_{k-1} + a_{1}^{P} \tilde{f}_{k-1} + \sum_{s=1}^{g} ( \tilde{b}_{k-1}^{s} b_{1}^{P,s+g} + b_{1}^{P,s} \tilde{b}_{k-1}^{s+g}). \quad \Box \label{60} \end{equation} \indent We now continue with the proof of proposition \ref{biggy}. Let ${\cal C}'={\cal C}(n-1,d-d_{P})$ and let ${\cal G}'= {\cal G}(n-1,d-d_{P}).$ Let $\mu' = (d_{1}/n_{1},...,d_{P-1}/n_{P-1})$ and let $e_{\mu'}$ denote the equivariant Euler class of the normal bundle to ${\cal C}'_{\mu'}$ in ${\cal C}'.$ Let \[ U_{\mu'} = {\cal C}' - \bigcup_{\nu'>\mu'} {\cal C}'_{\nu'}. \] Then $U_{\mu'}$ is an open subset of ${\cal C}'$ which contains ${\cal C}'_{\mu'}$ as a closed submanifold. So we have the maps\\ \begin{picture}(400,120) \put(190,60){\makebox(0,0){$H^{*-2d_{\mu'}}_{{\cal G}'}({\cal C}'_{\mu'};{\bf Q}) \rightarrow H^{*}_{{\cal G}'}(U_{\mu'};{\bf Q}) \rightarrow H^{*}_{{\cal G}'}(U_{\mu'}-{\cal C}'_{\mu'};{\bf Q})$}} \put(190,110){\makebox(0,0){$H^{*}_{{\cal G}'}({\cal C}';{\bf Q})$}} \put(190,10){\makebox(0,0){$H^{*}_{{\cal G}'}({\cal C}'_{\mu'};{\bf Q})$}} \put(190,102){\vector(0,-1){34}} \put(190,50){\vector(0,-1){32}} \put(100,50){\vector(2,-1){65}} \put(150,30){\makebox(0,0)[tr]{multiplication by $e_{\mu'}$}} \end{picture} \\ Let $a'_{r},{b^{s}_{r}}'$ and $f'_{r}$ denote the generators of $H^{*}_{{\cal G}'}({\cal C}';{\bf Q})$. Also take $\nu' \not \geq \mu'$ and let $\hat{a}_{r},\hat{b}_{r}^{s},\hat{f}_{r}$ denote the restrictions of $a'_{r},{b^{s}_{r}}',f'_{r}$ in $H^{*}_{{\cal G}'}({\cal C}'_{\nu'};{\bf Q})$. Since the stratification is equivariantly perfect then the restriction map \[ H^{*}_{{\cal G}'}({\cal C}';{\bf Q}) \rightarrow H^{*}_{{\cal G}'}(U_{\mu'};{\bf Q}) \] is surjective \cite[p.859]{K2}. From the exactness of the Thom-Gysin sequence we have that for every element of the form $\alpha e_{\mu'}$ in $H^{*}_{{\cal G}'}({\cal C}'_{\mu'};{\bf Q})e_{\mu'}$ there is some $\beta (a'_{r},{b_{r}^{s}}' ,f'_{r})$ in $H^{*}_{{\cal G}'}({\cal C}';{\bf Q})$ such that \[ \beta(\tilde{a}_{r},\tilde{b}_{r}^{s},\tilde{f}_{r}) = \alpha e_{\mu'} \mbox{ and } \beta(\hat{a}_{r},\hat{b}_{r}^{s},\hat{f}_{r}) = 0. \] Since every element of the form (\ref{81}) lies in the restriction of $(\Lambda)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ then every element of the form \begin{equation} \beta(\tilde{a}_{r},\tilde{b}_{r}^{s},\tilde{f}_{r}) (a_{1}^{P})^{r} \prod_{s \in S} b_{1}^{P,s} \Theta \indent (r \geq 0,S \subseteq \{1,...,2g\}) \label{z1} \end{equation} similarly lies in the restriction of $(\Lambda)$. Now let $\nu=(\nu',d_{P})$ with $\nu' \not \geq \mu'$. Note that the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \] is formally the same as (\ref{60}) but with $\hat{a}_{r},\hat{b}_{r}^{s},\hat{f}_{r}$ replacing $\tilde{a}_{r},\tilde{b}_{r}^{s},\tilde{f}_{r}$. Thus there are elements of $(\Lambda)$ which restrict to (\ref{z1}) under (\ref{60}) and have restriction \[ \beta(\hat{a}_{r},\hat{b}_{r}^{s},\hat{f}_{r}) (a_{1}^{P})^{r} \prod_{s \in S} b_{1}^{P,s} \hat{\Theta} = 0 \] in $H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q})$.\\[\baselineskip] \indent Define ${\cal R}_{\mu}$ to be all those elements of $(\Lambda)$ which restrict to an element of the form \[ \alpha e_{\mu'} (a_{1}^{P})^{r} \prod_{s \in S} b_{1}^{P,s} \Theta \indent (r \geq 0, S \subseteq \{1,...,2g\}, \alpha \in H^{*}_{{\cal G}'}({\cal C}'_{\mu'};{\bf Q})) \] in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ and which restrict to zero in $H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q})$ for any $\nu = (\nu',d_{P})$ with $\nu' \not \geq \mu'$.\\[\baselineskip] \indent From the definition of $\Theta$ (\ref{52}) we know that $e_{\mu'} \Theta$ is the constant coefficient of \begin{equation} (-1)^{d_{\mu'}} t^{d_{\mu'}} c({\cal N}_{\mu'})(-t^{-1}) \theta(t) \label{53} \end{equation} where ${\cal N}_{\mu'}$ is the normal bundle to ${\cal C}'_{\mu'}$ in ${\cal C}'$ and $d_{\mu'}$ is the codimension of ${\cal C}'_{\mu'}$ in ${\cal C}'$. From lemma \ref{normal} and the fact that \[ d_{\mu'} + d - nd_{P} + (n-1)\bar{g} = d_{\mu} \] we know (\ref{53}) equals \[ (-1)^{d_{\mu'}} t^{d_{\mu}} c({\cal N}_{\mu})(-t^{-1}) \] which has constant coefficient $(-1)^{d_{\mu'}+d_{\mu}}e_{\mu}.$ Hence the ideal \[ H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})e_{\mu} \] lies in the restriction of ${\cal R}_{\mu}$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}).$\\ \indent Finally from lemma \ref{vanishing} and the definition of $\Lambda$ (\ref{54}) we know that the image of ${\cal R}_{\mu}$ under the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}_{\nu};{\bf Q}) \indent \nu = (\tilde{d}_{1}/\tilde{n}_{1},...,\tilde{d}_{T}/\tilde{n}_{T}) \] vanishes when $\tilde{d}_{T}/\tilde{n}_{T} > d_{P}/n_{P}$ proving the first half of proposition \ref{biggy}.\\ \indent The proof of the dual case follows in a similar fashion.$ \indent \Box$\\[\baselineskip] \indent In the general rank case there are strata of types not covered in the previous proposition. Moreover the strata on which the restrictions of the relations have been demonstrated to vanish do not generally coincide with the strata mentioned in the hypotheses of the completeness criteria. However in the rank two and rank three cases all unstable strata are covered by the above proposition. In the rank two case proposition \ref{biggy} shows that the Mumford relations and the dual Mumford relations both form complete sets, simply duplicating Kirwan's work \cite{K2} and remark \ref{dualise}. In the rank three case we have the following:\\[\baselineskip] {\bf THEOREM 1.} {\em The Mumford and dual Mumford relations together with the relation (\ref{NORM}) due to the normalisation of the universal bundle $V$ form a complete set of relations for $H^{*}({\cal M}(3,d);{\bf Q}).$} {\bf Proof} The unstable strata are now of types (2,1),(1,1,1) and (1,2). From the previous proposition we may meet the completeness criteria for the (2,1) and (1,1,1) strata using the Mumford relations. In these cases those strata where the restriction of ${\cal R}_{\mu}$ have been shown to vanish are those strata ${\cal C}_{\nu}$ such that $\nu \prec \mu$. The criteria for the (1,2) types may be met using the dual Mumford relations. In this case those strata where the restriction of ${\cal R}_{\mu}$ vanishes (according to proposition \ref{biggy}) are those strata ${\cal C}_{\nu}$ such that $\nu \not \geq \mu$ which certainly includes those strata such that $\nu \prec \mu. \indent \Box$ \begin{rem} As remarked earlier it was shown in \cite[thm.4]{E} that the Mumford relations $\sigma_{-1,S}^{1}$ for $S \subseteq \{1,...,2g\}$ generate the relation ideal of $H^{*}({\cal M}_{0}(2,1);{\bf Q})$ as a ${\bf Q}[a_{2},f_{2}]$-module. Evidence for this theorem appears in the Poincar\'{e} polynomial of the relation ideal which equals \cite[p.593]{AB} \[ \frac{t^{2g}(1+t)^{2g}}{(1-t^{2})(1-t^{4})}. \] \indent Similarly in the rank three case the Poincar\'{e} polynomial of the ideal of relations among our generators for $H^{*}({\cal M}_{0}(3,1);{\bf Q})$ equals \[ \frac{(1+t^{2})^{2} t^{4g-2}(1+t)^{2g}(1+t^{3})^{2g} - (1+t^{2}+t^{4}) t^{6g-2}(1+t)^{4g}}{(1-t^{2})(1-t^{4})^{2}(1-t^{6})}, \] The first Mumford relation $\sigma_{-1,\{1,...,2g\}}^{2}$ has degree $4g-2$ and the first dual Mumford relation $\tau_{-1,\{1,...,2g\}]}^{2}$ has degree $4g$. This strongly suggests that the relations \[ \{ \sigma_{-1,S}^{i}, \tau_{-1,S}^{i} : i=1,2, S \subseteq \{1,...,2g\} \} \] generate the relation ideal of $H^{*}({\cal N}(3,d);{\bf Q})$ as a \[ {\bf Q}[a_{2},a_{3},f_{2},f_{3}] \otimes \Lambda^{*}\{b_{2}^{1},...,b_{2}^{2g}\} \] module. \end{rem} \section{On the Vanishing of the Pontryagin Ring.} \indent We now move on to discuss the Pontryagin ring of the moduli space in the rank three case. For each $S \subseteq \{1,...,2g\}$ we define $\Psi_{S}(t)$ and $\Psi^{*}_{S}(t)$ by writing \[ \Psi(t) = \sum_{S \subseteq \{1,...,2g\}} \Psi_{S}(t) \prod_{s \in S} b_{1}^{s}, \indent \Psi^{*}(t) = \sum_{S \subseteq \{1,...,2g\}} \Psi^{*}_{S}(t) \prod_{s \in S} b_{1}^{s}. \] Kirwan proved the Newstead-Ramanan conjecture \cite[$\S$ 4]{K2} by considering relations derived from the expression \[ \Psi_{\{1,...,2g\}}(t)\Psi_{\{1,...,2g\}}(-t-a_{1}).\] Arguing along similar lines but now considering the expression \[ \Phi(t) = \Psi_{\{1,...,2g\}}(t)\Psi^{*}_{\{1,...,2g\}}(t) \] we will show that in the rank three case the Pontryagin ring vanishes in degree $12g-8$ and above -- theorem 2 below. \begin{lem} \label{Pontlemma} Let $\mu = (d_{1}, d_{2},...,d_{n}) \in \Delta$. The restriction of $\Phi(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ (-1)^{g} \frac{A(t)^{2g}}{n^{4g}\tilde{\Omega}_{\mu}(t)} \] where \[ \tilde{\Omega}_{\mu}(t) = \prod_{p=1}^{n} (t+a_{1}^{p}), \indent A(t) = \sum_{p=1}^{n} \prod_{q \neq p} (t+a_{1}^{q}). \] \end{lem} {\bf Proof} From corollary \ref{Deltaresn} we know that the restriction of $\Psi(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ \prod_{p=1}^{n} (t+a_{1}^{p})^{d_{p}-\bar{g}} \exp \left\{ \frac{\xi_{p}}{t+a_{1}^{p}} \right\} \] where $\xi_{p} = \xi_{1,1}^{p,p} = \sum_{s=1}^{g} b_{1}^{p,s} b_{1}^{p,s+g}.$ Let $v_{s} = b_{1}^{1,s} + \cdot \cdot \cdot + b_{1}^{n,s}$ denote the restriction of $b_{1}^{s}$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ and let $w_{i,j}^{s} = b_{1}^{i,s} - b_{1}^{j,s}$ (see (\ref{1001})). Then $nb_{1}^{i,s} = v_{s} + \sum_{j=1}^{n} w_{i,j}^{s}$ and hence \[ n^{2} \xi_{i} = \sum_{s=1}^{g} v_{s}v_{s+g} + \sum_{s=1}^{g} \left( v_{s} \sum_{j=1}^{n} w_{i,j}^{s+g} + \sum_{j=1}^{n} w_{i,j}^{s} v_{s+g} \right) + \sum_{s=1}^{g} \sum_{j=1}^{n} \sum_{k=1}^{n} w_{i,j}^{s} w_{i,k}^{s+g}. \] Note that \begin{equation} \sum_{p=1}^{n} \frac{\xi_{p}}{t+a_{1}^{p}} = \frac{1}{\tilde{\Omega}_{\mu}(t)} \sum_{i=1}^{n} \sum_{q \neq i} \xi_{i} (t+a_{1}^{q}) \label{N} \end{equation} Thus (\ref{N}) equals \[ \frac{1}{n^{2} \tilde{\Omega}_{\mu}(t)} \left\{ A(t) \sum_{s=1}^{g} v_{s}v_{s+g} + \sum_{s=1}^{g} (B_{s}(t) v_{s+g} + v_{s}B_{s+g}(t)) + \Gamma(t) \right\} \] where \[ A(t) = \sum_{i=1}^{n} \prod_{q \neq i} (t+a_{1}^{q}) , \indent B_{s}(t) = \sum_{i=1}^{n} \sum_{j=1}^{n} w_{i,j}^{s} \prod_{q \neq i} (t+a_{1}^{q}), \] \[ \Gamma(t) = \sum_{i=1}^{n} \sum_{j=1}^{n} \sum_{k=1}^{n} \sum_{s=1}^{g} w_{i,j}^{s} w_{i,k}^{s+g} \prod_{q \neq i}(t+a_{1}^{q}). \] The exponential of (\ref{N}) equals \[ \exp \left\{ \frac{\Gamma(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right\} \prod_{s=1}^{g} \left[ 1+ \frac{B_{s}(t)v_{s+g} + v_{s} B_{s+g}}{n^{2}\tilde{\Omega}_{\mu}(t)} + \left( \frac{A(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} - \frac{B_{s}B_{s+g}}{n^{4}\tilde{\Omega}_{\mu}(t)^{2}}\right) v_{s}v_{s+g} \right]. \] The coefficient of $\prod_{s=1}^{2g} v_{s}$ in the above then equals \[ (-1)^{g\bar{g}/2} \exp \left\{ \frac{\Gamma(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right\} \prod_{s=1}^{g} \left( \frac{A(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} - \frac{B_{s}B_{s+g}}{n^{4}\tilde{\Omega}_{\mu}(t)^{2}} \right) \] or equivalently \[ (-1)^{g\bar{g}/2} \exp \left\{ \frac{\Gamma(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right\} \left( \frac{A(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right)^{g} \exp\left\{\frac{-\xi(t)}{n^{2}A(t)\tilde{\Omega}_{\mu}(t)} \right\} \] where $\xi(t) = \sum_{s=1}^{g} B_{s}(t)B_{s+g}(t)$. Thus the restriction of $\Psi_{\{1,...,2g\}}(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ (-1)^{g\bar{g}/2}\left(\prod_{p=1}^{n} (t+a_{1}^{p})^{d_{p}-\bar{g}} \right) \exp \left\{ \frac{\Gamma(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right\} \left( \frac{A(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right)^{g} \exp\left\{\frac{-\xi(t)}{n^{2}A(t)\tilde{\Omega}_{\mu}(t)} \right\} \] and similarly the restriction of $\Psi^{*}_{\{1,...,2g\}}(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ (-1)^{g\bar{g}/2} \left(\prod_{p=1}^{n} (t+a_{1}^{p})^{3\bar{g}+1-d_{p}} \right) \exp \left\{ \frac{- \Gamma(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right\} \left(\frac{-A(t)}{n^{2}\tilde{\Omega}_{\mu}(t)} \right)^{g} \exp\left\{\frac{\xi(t)}{n^{2}A(t)\tilde{\Omega}_{\mu}(t)} \right\}. \] The result then follows. $\indent \Box$\\[\baselineskip] \indent Now if we write $\Phi(t)$ in the form \[ \sum_{r = -\infty}^{2g-1} ( \rho_{r}^{0} + \rho_{r}^{1} t + \cdot \cdot \cdot + \rho_{r}^{n-1} t^{n-1}) ( \tilde{\Omega}(t))^{r} \] where $\tilde{\Omega}(t) = t^{n} + a_{1}t^{n-1} + \cdot \cdot \cdot +a_{n}$ then we know that the elements $\rho_{r}^{k},(r < 0 ,0 \leq k \leq n-1)$ lie in the kernel of the restriction map \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow H^{*}_{\cal G}({\cal C}^{ss};{\bf Q}). \] \indent From lemma \ref{Pontlemma} we know that the restriction of $\Phi(t)$ to $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q})$ equals \[ (-1)^{g} \frac{A(t)^{2g}}{n^{4g}\tilde{\Omega}_{\mu}(t)} \] for any $\mu \in \Delta.$ Let $\rho_{r}^{k,\mu}$ denote the restriction of $\rho_{r}^{k}$ in $H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}).$ Thus we have that \[ \frac{(-1)^{g}}{n^{4g}} A(t)^{2g} = \sum_{k=0}^{n-1} \rho_{-1}^{k,\mu} t^{k} \indent \mbox{mod } \tilde{\Omega}_{\mu}(t). \] Hence by substituting $t = -a_{1}^{i}$ for each $i$ we obtain \[ \frac{(-1)^{g}}{n^{4g}} \left( \prod_{p=1,p \neq i}^{n} (a_{1}^{i} - a_{1}^{p}) \right)^{2g} = \sum_{k=0}^{n-1} \rho_{-1}^{k,\mu}(-a_{1}^{i})^{k}. \] Since the direct sum of restriction maps \[ H^{*}_{\cal G}({\cal C};{\bf Q}) \rightarrow \bigoplus_{\mu \in \Delta} H^{*}_{\cal G}({\cal C}_{\mu};{\bf Q}) \] is injective \cite[prop. 3]{K2} we have that \begin{equation} \frac{(-1)^{g}}{n^{4g}} \left( \prod_{p=1,p \neq i}^{n} (\delta_{i} - \delta_{p}) \right)^{2g} = \sum_{k=0}^{n-1} \rho_{-1}^{k}(-\delta_{i})^{k} \label{321}. \end{equation} Solving the equations (\ref{321}) we obtain \begin{equation} \rho_{-1}^{k} = \frac{(-1)^{g+n}}{n^{4g}} \sum_{i=1}^{n} S_{i}^{k} \left( \prod_{p=1,p \neq i}^{n} (\delta_{i} - \delta_{p}) \right)^{2g-1} \label{rev5} \end{equation} where $S_{i}^{k}$ equals the $k$th elementary symmetric polynomial in $\{\delta_{p}: p \neq i\}$.\\ \indent We will show later, in proposition \ref{Pontinad}, that the relations $\rho_{-1}^{k}$ above are insufficient to prove any vanishing of the Pontryagin ring in ranks greater than three. For now consider the rank three case. We write \[ \alpha = \delta_{1}-\delta_{2},\indent \beta = \delta_{2}-\delta_{3}, \indent \gamma = \delta_{3} - \delta_{1}. \] We know from lemma \ref{Pont} that the Pontryagin ring is generated by the elementary symmetric polynomials in $\alpha^{2},\beta^{2}$ and $\gamma^{2}.$ The relations $\rho_{-1}^{0}, \rho_{-1}^{1},\rho_{-1}^{2}$ read as \begin{eqnarray} (\alpha \beta)^{2g-1} + (\beta \gamma)^{2g-1} + (\gamma \alpha)^{2g-1} = 0, \label{reva}\\ (\delta_{1}+\delta_{3}) (\alpha \beta)^{2g-1} +(\delta_{2}+\delta_{1}) (\beta \gamma)^{2g-1} + (\delta_{3}+\delta_{2}) (\gamma \alpha)^{2g-1} = 0, \label{revb}\\ (\delta_{1} \delta_{3})(\alpha \beta)^{2g-1} +(\delta_{2} \delta_{1}) (\beta \gamma)^{2g-1} + (\delta_{3} \delta_{2}) (\gamma \alpha)^{2g-1} = 0. \label{revc} \end{eqnarray} The equations (\ref{reva}), $a_{1} \times$ (\ref{reva}) -- (\ref{revb}), and (\ref{revc}) $+ a_{1} \times$ (\ref{revb}) $-a_{2} \times$ (\ref{reva}) then show \begin{equation} (\delta_{2})^{k} (\alpha \beta)^{2g-1} + (\delta_{3})^{k} (\beta \gamma)^{2g-1} +(\delta_{1})^{k} (\gamma \alpha)^{2g-1} = 0, \label{rev2} \end{equation} for $k = 0,1,2$. Note that \[ (\delta_{i})^{r+3} = a_{1} (\delta_{i})^{r+2} - a_{2} (\delta_{i})^{r+1} + a_{3} (\delta_{i})^{r} \] and hence equation (\ref{rev2}) holds for all non-negative $k$. Further note that \begin{equation} \gamma^{2} = (a_{1})^{2} - 4 a_{2} + 2 a_{1} \delta_{2} - 3 (\delta_{2})^{2} \label{rev1} \end{equation} and so combining equation (\ref{rev2}) with equation (\ref{rev1}) and two similar equations for $\alpha^{2}$ and $\beta^{2}$ we see that \[ \gamma^{2l} (\delta_{2})^{k} (\alpha \beta)^{2g-1} + \alpha^{2l} (\delta_{3})^{k} (\beta \gamma)^{2g-1} + \beta^{2l} (\delta_{1})^{k} (\gamma \alpha)^{2g-1} = 0, \] for any non-negative $k,l$. Let $r,s,t$ be three non-negative integers with an even sum. Note \[ 2 \alpha = (a_{1} - 3 \delta_{2}) - \gamma, \indent 2 \beta = (3 \delta_{2} - a_{1}) - \gamma, \] and hence $(\alpha^{r} \beta^{s} + \alpha^{s} \beta^{r}) \gamma^{t},$ when written in terms of $a_{1}, \delta_{2}$ and $\gamma$ is an even function in $\gamma$.\\ \indent Now any element of the Pontryagin ring can be written as a sum of elements of the form \[ F(u,v,w) = \alpha^{u}\beta^{v}\gamma^{w} + \alpha^{v}\beta^{w}\gamma^{u} + \alpha^{w}\beta^{u}\gamma^{v} + \alpha^{u}\beta^{w}\gamma^{v} + \alpha^{v}\beta^{u}\gamma^{w} + \alpha^{w}\beta^{v}\gamma^{u}, \] where $u+v+w$ is even. From the argument above we know that \begin{equation} F(2g-1+r,2g-1+s,t) = 0 \label{rev4} \end{equation} for $r,s,t \geq 0$ and $r+s+t$ even. If $u \geq 1$ then we have \begin{equation} F(u,v,w) = -F(u-1,v,w+1) - F(u-1,v+1,w) \label{rev3} \end{equation} since $\alpha+\beta+\gamma=0$.\\ \indent Suppose now that $ u \geq v \geq w$. We claim $F(u,v,w)=0$ if $u+v+w \geq 6g-4.$ Note that \[ \mbox{max}\{u,v,w\} > \mbox{max}\{u-1,v+1,w+1\} \] unless $u-v$ equals zero or one. In either case we find that $u \geq v \geq 2g-1$ and hence $F(u,v,w) = 0$ by (\ref{rev4}). Hence by repeated applications of identity (\ref{rev3}) we see that $F(u,v,w) = 0$ when $u+v+w \geq 6g-4$ and so we have:\\[\baselineskip] {\bf THEOREM 2.} {\em The Pontryagin ring of the moduli space ${\cal M}(3,d)$ vanishes in degrees $12g-8$ and above.} \begin{rem} Theorem 2 falls short of Neeman's conjecture \cite{NE} which states that the Pontryagin ring of ${\cal M}(n,d)$ should vanish in degrees above $2gn^{2}-4g(n-1)+2$. When $n=3$ this gives $10g+2$. \end{rem} \begin{rem} In the rank two case the relations (\ref{rev5}) show that \[ ((a_{1})^{2} - 4a_{2})^{g}=0 \] and that the Pontryagin ring of ${\cal M}(2,d)$ vanishes in degrees greater than or equal to $4g$, duplicating Kirwan's proof of the Newstead-Ramanan conjecture. \end{rem} \indent To conclude we show now that the relations $\rho_{-1}^{k}$ are inadequate to show any vanishing of the Pontryagin ring when $n \geq 4$. From equation (\ref{rev5}) we see that the ideal of the Pontryagin ring is contained in the ideal generated by the formal expressions \begin{equation} \left( \prod_{p=1,p \neq i}^{n} (\delta_{i} - \delta_{p}) \right)^{2g-1} \label{rev6}. \end{equation} Let $I$ denote the ideal generated by the relations (\ref{rev6}) and consider this as an ideal of ${\bf C}[\delta_{1},...,\delta_{n}]$. By Hilbert's Nullstellensatz the radical $\sqrt{I}$ of $I$ consists of those elements of the Pontryagin ring which vanish on the intersection of the subspaces given by \begin{equation} \prod_{p \neq i} (\delta_{i} - \delta_{p}) = 0, \indent i=1,...,n. \label{Q16} \end{equation} We shall consider the even and odd cases for $n$ separately.\\ \indent (i) $n$ is even -- write $n=2m.$ The intersection of the subspaces (\ref{Q16}) consists of $(2m)!/(2^{m}m!)$ distinct $m$-dimensional subspaces of ${\bf C}^{n}.$ One of these subspaces is given by the equations \begin{equation} \delta_{2k-1} = \delta_{2k}, \indent k=1,...,m. \label{plane} \end{equation} We know from lemma \ref{Pont} that the total Pontryagin class $p(T)$ of ${\cal M}(n,d)$ equals \[ \prod_{1 \leq k < l \leq n} (1 + (\delta_{k}-\delta_{l})^{2})^{2 \bar{g}} \] and in the subspace (\ref{plane}) $p(T)$ then equals \[ \prod_{1 \leq k < l \leq m} (1 + (\delta_{2 k -1} - \delta_{2 l -1})^{2})^{8 \bar{g}}. \] In particular we see that none of the Pontryagin classes of ${\cal M}(n,d)$ vanish on the subspace (\ref{plane}).\\ \indent (ii) $n$ is odd -- write $n=2m+1$. The intersection of the subspaces (\ref{Q16}) consists of $(2k+1)!/(3 \cdot 2^{k}(k-1)!)$ distinct $k$-dimensional subspaces of ${\bf C}^{n}$. One of these subspaces is given by the equations \begin{equation} \delta_{1} = \delta_{2} = \delta_{3}, \quad \delta_{2k} = \delta_{2k+1}, \quad k=2,...,m. \label{Plane} \end{equation} In the subspace (\ref{Plane}) the total Pontryagin class of ${\cal M}(n,d)$ equals \[ \left( \prod_{2 \leq k \leq m} (1+(\delta_{1}-\delta_{2k})^{2})^{12\bar{g}} \right) \left( \prod_{2 \leq k < l \leq m} (1+(\delta_{2k}-\delta_{2l})^{2})^{8\bar{g}} \right). \] In particular we see that none of the Pontryagin classes of ${\cal M}(n,d)$ vanish on the subspace (\ref{Plane}).\\ \indent Thus we see that none of the Pontryagin classes $p_{r}(T)$ are nilpotent modulo the formal relations (\ref{rev6}). Hence: \begin{prop} \label{Pontinad} For $n \geq 4$ the Pontryagin classes $p_{r}(T) \in H^{4r}({\cal M}(n,d);{\bf Q})$ are not nilpotent modulo $\rho_{-1}^{k}$ for $0 \leq k \leq n-1$. In particular these relations are inadequate to prove any non-trivial vanishing of the Pontryagin ring. \end{prop}
1995-03-30T07:20:33
9503
alg-geom/9503022
en
https://arxiv.org/abs/alg-geom/9503022
[ "alg-geom", "math.AG" ]
alg-geom/9503022
Luca Barbieri-Viale
L. Barbieri-Viale, C. Pedrini and C. Weibel
Roitman's theorem for singular complex projective surfaces
36 pages, LaTeX
Duke Math. J. 84 (1996), 155-190
null
null
null
Let $X$ be a complex projective surface with arbitrary singularities. We construct a generalized Abel--Jacobi map $A_0(X)\to J^2(X)$ and show that it is an isomorphism on torsion subgroups. Here $A_0(X)$ is the appropriate Chow group of smooth 0-cycles of degree 0 on $X$, and $J^2(X)$ is the intermediate Jacobian associated with the mixed Hodge structure on $H^3(X)$. Our result generalizes a theorem of Roitman for smooth surfaces: if $X$ is smooth then the torsion in the usual Chow group $A_0(X)$ is isomorphic to the torsion in the usual Albanese variety $J^2(X)\cong Alb(X)$ by the classical Abel-Jacobi map.
[ { "version": "v1", "created": "Wed, 29 Mar 1995 10:13:35 GMT" } ]
2008-02-03T00:00:00
[ [ "Barbieri-Viale", "L.", "" ], [ "Pedrini", "C.", "" ], [ "Weibel", "C.", "" ] ]
alg-geom
\section*{Introduction} If $X$ is a smooth projective surface over the complex numbers $\C$, the classical Abel--Jacobi map goes from the Chow group $A_0(X)$ of cycles of degree 0 to the (group underlying the) Albanese Variety $Alb(X)$. Roitman's Theorem \cite{Roit} states that this map induces an isomorphism on torsion subgroups. (See \cite{CT} for a nice compendium). The goal of this paper is to remove the word ``smooth'' from Roitman's theorem. For this we shall modify the definition of $A_0(X)$, replace $Alb(X)$ with Griffiths' intermediate Jacobian $J^2(X)$, and construct a generalization of the Abel--Jacobi map. \medskip \noindent{\bf Main Theorem. }\ {\it Let $X$ be a reduced projective surface over $\C$. Then there is a natural map from $A_0(X)$ to $J^2(X)$ inducing an isomorphism on torsion: $$A_0(X)_{tors}\cong J^2(X)_{tors}.$$ In particular, the torsion subgroup is a finite direct sum of copies of $\Q/\Z$.} \smallskip If $X$ is a normal surface, this theorem is a reformulation of a theorem of Collino and Levine \cite{C2} \cite{L-Alb}, because (as we will show in Corollary~\ref{abnorm}), $J^2(X)$ is isomorphic to the Albanese of any desingularization of $X$. Gillet studied the Abel--Jacobi map in \cite{GDuke} when $X$ is a singular surface with ``ordinary multiple curves'' ({\it e.g.\/}\ a seminormal surface with smooth normalization $\tilde X$). He proved in \cite[Theorem B]{GDuke} that if $\tilde X$ satisfied some extra hypotheses ($p_g=0$, etc.) then the Abel--Jacobi map is surjective with finite kernel. Thus we deduce: \medskip \noindent{\bf Corollary. }\ {\it Let $X$ be a surface with ordinary multiple curves such that $H^2(X,{\cal O}_X)=0$. Assume that Bloch's conjecture holds for the normalization $\tilde X$ of $X$. Then the Abel--Jacobi map is an isomorphism. $$A_0(X) \cong J^2(X)$$} We now describe the ingredients in our main theorem. If $X$ is a proper surface over $\C$, the intermediate Jacobian $J^2(X)$ is defined to be $$J^2(X) \mbox{\,$\stackrel{\pp{\rm def}}{=}$}\,\frac{H^{3}(X,\C)}{F^2H^{3}+ H^{3}(X,\Z(2))}.$$ Here $F^2H^3$ refers to the Hodge filtration of \cite{D} and the coefficients $\Z(2)$ refer to the embedding of $\Z$ in $\C$ sending 1 to $(2\pi i)^2$. When $X$ is a smooth surface, it is well known that $J^2(X)$ is isomorphic to the Albanese Variety $Alb(X)$. Now suppose that $X$ is a singular surface. We will show in Corollary~\ref{1-motive} that $J^2(X)$ is a complex torus, and that if $X'$ is a resolution of singularities for $X$ then $J^2(X)$ is an extension of $Alb(X')$ by a torus. That is, the map $J^2(X)\to Alb(X')$ forms a 1--motive in the sense of Deligne \cite{D}; we call it the {\it Albanese 1--motive} of $X$. Given this, the torsion subgroup of $J^2(X)$ is a finite direct sum of copies of $\Q/\Z$. The modified version of $A_0(X)$ is defined as a subgroup of the Levine--Weibel Chow group $CH_0(X)$ of zero-cycles on $X$ \cite{LW}. By definition, $CH_0(X)$ is the abelian group generated by the smooth closed points on $X$, modulo the subgroup generated by all terms $D=\sum n_iP_i$ (with $P_i$ smooth on $X$) such that $D = (f)$ for some rational function $f$ on some curve $C$, the curve being locally defined by a single equation on the surface $X$. If $X$ is a surface with $c$ proper components, there is a natural surjection $CH_0(X)\to\Z^c$, called the degree map. By definition, $A_0(X)$ is the kernel of the degree map. In order to prove our Main Theorem, we need to reinterpret $CH_0(X)$ in terms of algebraic $K$-theory. Let $SK_0(X)$ denote the subgroup of $K_0(X)$ generated by the classes of smooth points on the surface $X$. Then $CH_0(X)$ is isomorphic to $SK_0(X)$, by the map sending a smooth point to its class in $K_0(X)$. This is the Riemann--Roch Theorem if $X$ is smooth. It was proven for affine surfaces in \cite[Theorem 2.3]{LW}. For arbitrary quasi-projective surfaces it is due to Levine \cite{L1}, who proved that both groups are isomorphic to $H^2(X,{\cal K}_2)$ ({\it cf.\/}\ \cite{PW1}, \cite{C1}). The isomorphism $$CH_0(X) \cong H^2(X,{\cal K}_2) \cong SK_0(X)$$ is often called ``Bloch's formula'' for surfaces. We have laid this paper out as follows. In \S1 we present some basic facts about Deligne cohomology of a proper but singular scheme. The corresponding Deligne Chern classes which will be used in later sections is introduced in \S2. In \S3 we construct and compare the Mayer-Vietoris sequences for $K$-theory and Deligne cohomology that we shall need. In \S4 we compute $J^2(X)$ for any proper surface $X$. Our computation shows that $J^2(X)$ is part of a 1-motive $Alb(X)$ which we call the {\it Albanese 1-motive of $X$}. In \S5 we describe the structure of $SK_1$ of any curve over any algebraically closed field. In \S6 and \S7 we establish some technical results about ${\cal K}_2$-cohomology, ending with the exact sequence of Theorem~\ref{NH3} for a normal surface $X$ over any field containing $\frac1n$. $$0\to H^1(X,{\cal K}_2)/n\to N\Het3(X)\to{}_nCH_0(X)\to 0.$$ Finally we prove the Main Theorem in \S8. \medskip \section*{Notation} All schemes we consider will be separated and of finite type over a field $k$. We call such a scheme a {\it curve} if it is 1-dimensional, and a {\it surface} if it is 2-dimensional. When $X$ is an algebraic scheme over $\C$, we will write $H^*(X,\Z)$ and $H^*(X,\C)$ for the singular cohomology of the associated analytic space $X_{an}$ as well as for the mixed Hodge structure on it, given by Deligne \cite{D}. The weight filtration on $H^*(X,\Z)$ will be written as $W_iH^*$, and the Hodge filtration on $H^*(X,\C)$ will be written as $F^iH^*$. The notation $\Z(r)$ denotes the subgroup $(2\pi i)^r\Z$ of $\C$. Unless we wish to call attention to the relation with $H^*(X,\C)$, we will write $H^*(X,\Z)$ instead of $H^*(X,\Z(r))$. The notation $\Z(r)_{\cal D}$ denotes the Deligne complex on a smooth scheme $X$ over $\C$ (see \S1). We will use the Deligne complex to define the Deligne cohomology of proper schemes; in the affine case the definition of Deligne-Beilinson cohomology is different (one needs to consider logarithmic poles), and we remain silent about this. Similarly, the Zariski sheaves ${\cal H}_{\cal D}^*(r)$ (defined as the higher direct images of $\Z(r)_{\cal D}$) are used only for proper schemes, as a technical device. (See \S1, (2.4), \ref{square} and \ref{crux}.) The Zariski sheaf ${\cal K}_q$ on $X$ is obtained by sheafifying the Quillen higher $K$-theory functor $U\mapsto K_q(U)$. The ${\cal K}$-cohomology groups $H^p(X,{\cal K}_q)$ are just the Zariski cohomology of these sheaves. As indicated in the introduction, when $X$ is a surface the most important ${\cal K}$-cohomology group is $H^2(X,{\cal K}_2)\cong CH_0(X)$. Similarly, we shall write ${\cal K}_q(\Z/n)$ and ${\cal H}^q(\mu_n^{\otimes i})$ for the Zariski sheaves associated to the presheaves sending $U$ to $K_q(U;\Z/n)$ and $\Het{q}(U,\mu_n^{\otimes i})$, respectively. In general, we will always use calligraphic letters for Zariski sheaves. Finally, we will use some standard notation. Let $H$ be an abelian group or sheaf of abelian groups. Then $H_{tors}$ will denote its torsion subgroup. For each integer $n$ we will write $H/n$ for $H/nH$, and ${}_nH$ for the subgroup $\{ x\in H\colon nx=0\}$ of $H$. \goodbreak \section{Deligne cohomology groups} For $X$ smooth (possibly affine) over $\C$ we let $\Z(r)_{{\cal D}}$ denote the ``Deligne complex'' $$0\to \Z(r)\to {\cal O}_{X_{an}}\by{d} \cdots \by{d} \Omega^{r-1}_{X_{an}} \to0$$ of sheaves on the complex analytic manifold $X_{an}$, where $\Z(r)\mbox{\,$\stackrel{\pp{\rm def}}{=}$}\, (2\pi i)^r\Z$ is in degree 0. The {\it analytic} Deligne cohomology groups of the smooth scheme $X$ are defined to be $$H_{\cal D}^q(X,\Z(r))=H^q(X,\Z(r)_{{\cal D}}) \mbox{\,$\stackrel{\pp{\rm def}}{=}$}\,\H_{an}^q(X,\Z(r)_{{\cal D}}).$$ We then have exact sequences of complexes of sheaves on $X_{an}$: \B{equation}\label{augment} 0\to \Omega^{<r}_{X_{an}}[-1]\to \Z(r)_{{\cal D}}\by{\varepsilon_X} \Z\to 0. \E{equation} We can also define the Deligne cohomology groups of a smooth simplicial scheme $X_{\bul}$ by considering $\Z(r)_{{\cal D}}$ as a complex of analytic sheaves on $X_{\bul}$. This yields an exact sequence of complexes parallel to (\ref{augment}) by \cite[5.1.9.(II)]{D}. Now let $X$ be a singular scheme. A {\it smooth proper hypercovering} $X_{\bul}\to X$ of $X$ ({\it cf.\/}\ \cite[6.2.5--6.2.8]{D}) is a simplicial scheme $X_{\bul}$ with smooth components $X_i$, each proper over $X$, together with a morphism to $X$ satisfying ``universal cohomological descent.'' We define the Deligne cohomology of $X$ to be: $$H_{{\cal D}}^q(X,\Z(r))\mbox{\,$\stackrel{\pp{\rm def}}{=}$}\,\H_{an}^q(X_{\bul},\Z(r)_{{\cal D}}).$$ This definition is independent of the choice of smooth proper hypercovering by \cite[Expos\'e V{\it bis}, 5.1.7 and 5.2.4]{SGA4}. There is a canonical descent isomorphism $H^*(X,\Z)\cong H^*(X_{\bul},\Z)$, so the map $\varepsilon$ in (\ref{augment}) induces a natural map $\varepsilon_X\colon H_{\cal D}^*(X,\Z(r))\to H_{an}^*(X,\Z)$. It is well-known (see \cite[1.6.4]{Bei}{}) that $\varepsilon_X$ preserves products. For $X$ proper with arbitrary singularities we have a standard long exact sequence \B{equation}\label{modf} \cdots\by{\varepsilon}\kern-2pt H^q(X,\Z)\to\kern-2pt H^q(X,\C)/F^r\kern-2pt\to\kern-2pt H_{{\cal D}}^{q+1}(X,\Z(r)) \by{\varepsilon}\kern-2pt H^{q+1}(X,\Z)\to\cdots \E{equation} induced by (\ref{augment}) and $\Z\cong\Z(r)\subset\C$, as well as \B{equation}\label{star} \cdots\to F^rH^q(X,\C)\to H^q(X,\C/\Z(r)) \to H_{{\cal D}}^{q+1}(X,\Z(r)) \to F^rH^{q+1}(X,\C)\to\cdots \E{equation} If $X$ is a proper surface then from (\ref{modf}) we have an exact sequence \B{equation}\label{extJZ} 0\to J^2(X) \to H_{{\cal D}}^{4}(X,\Z(2)) \by{\varepsilon} H^{4}(X,\Z(2))\to0. \E{equation} Any map $i\colon Y\to X$ lifts to a morphism $i\colon Y_{\mbox{\Large $\cdot $}}\to X_{\bul}$ between hypercoverings; see \cite[Expos\'e V{\it bis}, 5.1.7 and 5.2.4]{SGA4} or \cite[6.2.8]{D}. The {\it relative} Deligne cohomology of this map is defined in the notation of \cite[6.3.3]{D} to be: $$H_{{\cal D}}^q(X{\rm mod\,}Y,\Z(r))\mbox{\,$\stackrel{\pp{\rm def}}{=}$}\, \H_{an}^q(X_{\bul}{\rm\,mod\,}Y_{\mbox{\Large $\cdot $}},\ \Z(r)_{{\cal D}} {\rm\, mod\,}\Z(r)_{{\cal D}}) $$ By \cite[6.3.2.2]{D} we have a functorial long exact sequence \B{equation} \cdots\to H_{{\cal D}}^q(X{\rm\,mod\,}Y,\Z(r))\to H_{{\cal D}}^q(X,\Z(r))\to H_{{\cal D}}^q(Y,\Z(r)) \to\cdots \E{equation} and of course there are relative versions of (\ref{modf}) and (\ref{star}) which depend functorially on the pair $(X,Y)$, such as \B{equation}\label{modf-rel} \cdots\by\varepsilon\kern-3pt H^q(X{\rm mod\,}Y,\Z)\to\kern-3pt H^q(X{\rm mod\,}Y )/F^r\to\kern-3pt H_{{\cal D}}^{q+1}(X{\rm mod\,}Y,\Z(r))\by\varepsilon\cdots. \E{equation} \subsection*{Low degree Deligne cohomology} We will need the following calculation of $H_{\cal D}^q(X,\Z(2))$ for $q\le2$. Given a proper scheme $X$ over $\C$, we fix a smooth proper hypercovering $X_{\bul}\by\pi X$. By abuse of notation, we shall write ${\cal H}_{\cal D}^q(2)$ for the complexes $R^q\omega_*\Z(2)_{\cal D}$ of Zariski sheaves on either $X_{\bul}$ or $X$, $\omega$ denoting either $\omega_{\mbox{\Large $\cdot $}}\colon X_{\mbox{\Large $\cdot $} an}\to X_{\mbox{\Large $\cdot $} zar}$ or $\omega=\pi\omega_{\mbox{\Large $\cdot $}}\colon X_{\mbox{\Large $\cdot $} an}\to X_{zar}$. \B{prop}\label{sheafHD} For $X$ proper and connected over $\C$ we have: \kern-8pt \B{description} \item[{\it (i)}] $H_{\cal D}^0(X,\Z(2)) = 0$; \item[{\it (ii)}] $H_{\cal D}^1(X,\Z(2))\cong\C/\Z(2) \cong\C^*$; \item[{\it (iii)}] $H_{\cal D}^2(X,\Z(2))_{tors}\cong H^1(X,\Q/\Z)$. \E{description}\kern-8pt Moreover, if $X$ is irreducible then we have \B{description} \item[{\it (iv)}] $H_{\cal D}^2(X,\Z(2)) = H^0_{zar}(X,{\cal H}_{\cal D}^2(2)) =H^0_{zar}(X_{\bul},{\cal H}_{\cal D}^2(2))$ \E{description} and there are edge homomorphisms: \B{description} \item[{\it (v)}] $H^1_{zar}(X,{\cal H}_{\cal D}^2(2)) \hookrightarrow H^1_{Zar}(X_{\bul},{\cal H}_{\cal D}^2(2))\hookrightarrow H_{\cal D}^3(X,\Z(2))$ \quad (these are injections); \item[{\it (vi)}] $H^2_{zar}(X,{\cal H}_{\cal D}^2(2)) \longby{} H^2_{Zar}(X_{\bul},{\cal H}_{\cal D}^2(2))\to H_{\cal D}^4(X,\Z(2))$. \E{description} \E{prop} \B{proof}It is well-known that $H_{an}^1(X,\Z)$ is torsion-free. Hence $(i)$ and $(ii)$ follow immediately from (\ref{modf}). (Cf. the proof of Lemma~2.17 in \cite{GDuke}.) Part $(iii)$ follows from this and (\ref{star}), since $H^1(X,\C/\Z(2))_{tors}\cong H^1(X,\Q/\Z)$. In order to prove parts $(iv)$, $(v)$ and $(vi)$ we use the Leray spectral sequences for $\omega$ and $\omega_{\mbox{\Large $\cdot $}}$: \B{equation} \label{simps} \B{array}{ccccc} E^{p,q}_2 &=& H^p_{zar}(X,\,{\cal H}_{\cal D}^q(i)) &\mbox{$\Rightarrow$}& H_{\cal D}^{p+q}(X,\Z(i))_{\strut} \\ E^{p,q}_2 &=& H^p_{zar}(X_{\bul},{\cal H}_{\cal D}^q(i)) &\mbox{$\Rightarrow$}& H_{\cal D}^{p+q}(X,\Z(i)) \E{array}\E{equation} with $i=2$ ({\it cf.\/}\ \cite[(2.13)]{GDuke}). For this, we need to compute ${\cal H}_{\cal D}^0(2)$ and ${\cal H}_{\cal D}^1(2)$. When $U$ is smooth we may identify the analytic sheaf ${\cal O}_U/\Z(2)$ with ${\cal O}_U^*$ and obtain a quasi-isomorphism between $\Z(2)_{\cal D}$ and the complex $0\to{\cal O}_U^*\longby{dlog}\Omega_U^1$. It follows that there is a distinguished triangle of complexes of analytic sheaves on $X_{\bul}$ $$\C^*[-1] \to \Z(2)_{\cal D} \to \Omega_{X_{\bul}}^1/dlog({\cal O}_{X_{\bul}}^*)[-2] \to \C^*.$$ Applying $\omega_*$ and $R^1\omega_*$ immediately yields ${\cal H}_{\cal D}^0(2)=0$ and ${\cal H}_{\cal D}^1(2)=\omega_*\C^*=\C^*$ on both $X_{\mbox{\Large $\cdot $} zar}$ and $X_{zar}$. Therefore in either spectral sequence (\ref{simps}) the row $q=0$ vanishes and in row $q=1$ we have $H_{zar}^p(X,\C^*)=H_{zar}^p(X_{\bul},\C^*)$. The exact sequences of low degree terms in (\ref{simps}) become: $$\B{array}{c} 0\to H_{zar}^1(X,\C^*)\to H_{\cal D}^2(X,\Z(2)) \to H^0(X,{\cal H}_{\cal D}^2) \longby{d_2} H_{zar}^2(X,\C^*) \to H_{\cal D}^3(X,\Z(2))\\ 0\to H_{zar}^1(X,\C^*)\to H_{\cal D}^2(X,\Z(2)) \to H^0(X_{\bul},{\cal H}_{\cal D}^2) \longby{d_2} H_{zar}^2(X,\C^*) \to H_{\cal D}^3(X,\Z(2)). \E{array}$$ The map between these sequences identifies them, and $H^0(X,{\cal H}_{\cal D}^2)\cong H^0(X_{\bul},{\cal H}_{\cal D}^2)$ by the 5-lemma. If $X$ is irreducible then $H_{zar}^p(X,\C^*)=0$ for $p\ne0$. Hence $H_{\cal D}^2(X,\Z(2))$ is isomorphic to $H^0(X,{\cal H}_{\cal D}^2)$. Parts $(v)$ and $(vi)$ follows similarly. \E{proof} For each $n$ there is a distinguished triangle of complexes of analytic sheaves on $X_{\bul}$: \B{equation}\label{triangle} \Z/n[-1]\by\delta\Z(2)_{\cal D}\by{n}\Z(2)_{\cal D} \by{\bar\varepsilon}\Z/n \longby{\delta[1]}\Z(2)_{\cal D}[1]. \E{equation} The comparison theorem between the analytic and \'etale sites, together with universal cohomological descent, yields $H^q(X_{\bul},\Z/n)\cong H^q(X,\Z/n)\cong\Het{q}(X,\Z/n)$. Fixing an $n^{th}$ root of unity allows us to identify $\mu_n$, $\mu_n^{\otimes2}$ and $\Z/n$ on $\eX$. If $X$ is proper, the cohomology of the triangle (\ref{triangle}) yields ``Kummer sequences'' \B{equation}\label{Kum-etHD} 0\to H_{\cal D}^{q}(X,\Z(2))/n \longby{\bar\varepsilon} \Het{q}(X,\mu_n^{\otimes2})\longby\delta H_{\cal D}^{q+1}(X,\Z(2))_{n-tors} \to0. \E{equation} By \ref{sheafHD} this identifies $\mu_n\cong\Het0(X,\Z/n)$ with the $n$-torsion in $\C^*\cong H_{\cal D}^1(X,\Z(2))$, and identifies $\Het1(X,\mu_n^{\otimes2})\cong\Het1(X,\Z/n)$ with the $n$-torsion in $H_{\cal D}^2(X,\Z(2))$. \smallskip Now consider the morphism $\omega\colon X_{\mbox{\Large $\cdot $} an}\to X_{zar}$. Applying the higher direct image $R^q\omega_*$ to (\ref{triangle}) yields an exact sequence of Zariski sheaves: \B{equation}\label{Rq} \longby{\bar\varepsilon} {\cal H}^{q-1}(\mu_n^{\otimes2}) \longby\delta {\cal H}_{\cal D}^q(2) \longby{n} {\cal H}_{\cal D}^q(2) \longby{\bar\varepsilon} {\cal H}^q(\mu_n^{\otimes2})\longby\delta. \E{equation} In particular, $\delta$ identifies ${\cal H}^1(\mu_n^{\otimes2})$ with the $n$-torsion subsheaf of ${\cal H}_{\cal D}^2(2)$. The map $\Het{*-1}(X,\mu_n^{\otimes2})\by\delta H_{\cal D}^*(X,\Z(2))$ is also the abutment of a morphism of Leray spectral sequences. At the $E_2$-level it is $H_{zar}^p(X,{\cal H}^{q-1}(\mu_n^{\otimes2})) \by\delta H_{zar}^p(X,{\cal H}_{\cal D}^q(2))$. If $X$ is proper and irreducible then the bottom row of both spectral sequences degenerates ({\it e.g.\/}\ $H_{zar}^p(X,\Z/n)=0$ for $p\ne0$) and we obtain the following result. \B{cor}\label{morph}If $X$ is proper and irreducible, there is a commutative diagram whose rows are the exact sequences of low degree terms of Leray spectral sequences: $$\B{array}{ccccccccc} 0&\to& \kern-5pt H^1(X,{\cal H}^1(\mu_n^{\otimes2}))&\to&\kern-3pt\Het2(X,\mu_n^{\otimes2}) &\to& \kern-3pt H^0(X,{\cal H}^2(\mu_n^{\otimes2})) &\kern-4pt\longby{d_2}&\kern-8pt H^2(X,{\cal H}^1(\mu_n^{\otimes2}))\\ &&\delta\downarrow &&\delta\downarrow&&\delta\downarrow&& \delta\downarrow\\ 0&\to& \kern-5pt H^1(X,{\cal H}_{\cal D}^2(2))&\to&\kern-3pt H_{\cal D}^3(X,\Z(2)) &\to& \kern-3pt H^0(X,{\cal H}_{\cal D}^3(2)) &\kern-4pt\longby{d_2}& \kern-8pt H^2(X,{\cal H}_{\cal D}^2(2)). \E{array}$$ \E{cor} \section{Chern classes in Deligne cohomology} For each scheme $X$ of finite type over $\C$ the exponential map ${\cal O}_{X_{an}}\to{\cal O}_{X_{an}}^*$ induces a quasi-isomorphism between $\Z(1)_{{\cal D}}=(\Z\to{\cal O}_{X_{an}})$ and ${\cal O}_{X_{an}}^*[-1]$. This quasi-isomorphism also holds over a simplicial scheme $X_{\bul}$ by naturality, so $\H^q(X_{\bul},\Z(1)_{\cal D})\cong H^{q-1}(X_{\bul},{\cal O}^*_{X_{\bul}})$. This gives a natural map from $H^1_{an}(X,{\cal O}_X^*)$ to $\H^2(X_{\bul},\Z(1)_{\cal D})\cong H_{an}^1(X_{\bul},{\cal O}^*_{X_{\bul}})$ for every smooth proper hypercovering $X_{\bul}\to X$. Composing with the determinant map $K_0(X)\to \mbox{Pic}(X)$ and the natural map $\mbox{Pic}(X)\to H_{an}^1(X,{\cal O}_X^*)$ yields a map $c_1\colon K_0(X) \to \H^2(X_{\bul},\Z(1)_{\cal D})$. Now the splitting principle holds for Deligne cohomology by \cite[5.2]{GDuke}. (Warning: if $X$ is not proper this differs slightly from the splitting principle proven in \cite[1.7.2]{Bei}!) Thus the map $c_1$ extends to Chern classes $c_i\colon K_0(X) \to \H^{2i}(X_{\bul},\Z(i)_{\cal D})$ for vector bundles. When $X$ is proper, these are the Deligne-Beilinson Chern classes $$c_i\colon K_0(X) \to H_{\cal D}^{2i}(X,\Z(i)).$$ Recall from (\ref{augment}) that there is a map $\varepsilon_X\colon \H^{2i}(X_{\bul},\Z(i)_{\cal D})\to H_{an}^{2i}(X,\Z)$, and that it is product-preserving. \B{lemma}\label{c2-an} ({\it cf.\/}\ Beilinson \cite[1.7]{Bei}) The composition of $c_i$ with the map $\varepsilon_X$ is the classical Chern class of the associated topological vector bundle \cite{MCC}: $$c_i^{an}\colon K_0(X)\to H_{an}^{2i}(X,\Z)=H_{top}^{2i}(X,\Z).$$ \E{lemma} \B{proof} Since $\varepsilon_X$ preserves cup products, the splitting principle shows that it suffices to establish the result for $c_1$. If $X$ is smooth then $c_1$ is the analytic determinant map, and $\varepsilon_X$ is just the usual map $\partial_X\colon H_{an}^1(X,{\cal O}_X^*)\to H_{an}^2(X,\Z)$ used to define $c_1^{an}$ on analytic vector bundles, so it is clear that $c_1^{an}=\varepsilon_X\circ c_1$. To deduce the result for general $X$, choose a smooth proper hypercover $u\colon X_{\bul}\to X$. Composing $\partial_X$ (which is $c_1^{an}$) with the descent isomorphism $H_{an}^{2i}(X,\Z)\cong H_{an}^{2i}(X_{\bul},\Z)$ is the descent map $H^1(X,{\cal O}_X^*)\to H^1(X_{\bul},{\cal O}_{X_{\bul}}^*)$ (which is $c_1$) composed with $\partial_{X_{\bul}}$, {\it i.e.,\/}\ with $\varepsilon_{X_{\bul}}$. \E{proof} Reduction of $\varepsilon_X$ mod $n$ yields a map $\bar\varepsilon_X\colon H_{\cal D}^{2i}(X,\Z(i))\to H_{an}^{2i}(X,\Z/n)$. Since reduction mod $n$ is product-preserving and sends $c_1^{an}$ to the \'etale Chern class $c_1^{et}$, we deduce the \B{cor}\label{c-etale} The composition of $c_i$ with $\bar\varepsilon_X$ is the \'etale Chern class $$c_i^{et}\colon K_0(X) \to H_{an}^{2i}(X,\Z/n) \cong\Het{2i}(X,\mu_n^{\otimes i}).$$ \E{cor} In this paper we shall be mostly concerned with the class $c_2\colon K_0(X)\to H^4_{{\cal D}}(X,\Z(2))$ when $X$ is a projective surface. Recall from the introduction (or \cite{LW}) that the Chow group $CH_0(X)$ of zero-cycles on $X$ is isomorphic to the subgroup $SK_0(X)$ of $K_0(X)$. If $X$ has $c$ irreducible components then there is a natural degree map $CH_0(X)\to\Z^c$, and $A_0(X)$ is defined to be the kernel of this map. The following cohomological interpretation of the degree map will be useful. \B{lemma}\label{degree} (Beilinson \cite[1.9]{Bei}) If $X$ is a projective surface, the degree map is the same (up to sign) as the classical Chern class $$ CH_0(X) \hookrightarrow K_0(X) \by{c_2^{an}} H_{an}^4(X,\Z) \cong \Z^c.$$ By (\ref{extJZ}), the Deligne Chern class $c_2$ induces a natural map $\rho\colon A_0(X)\to J^2(X)$ fitting into the diagram $$ \begin{array}{ccccccc} 0 \to & A_0(X) &\to& CH_0(X) &\by{deg}& \Z^c &\to0 \\ &\rho\downarrow& &c_2\downarrow& & \mbox{\large $\parallel$} & \\ 0 \to & J^2(X) &\to& H_{\cal D}^4(X,\Z(2)) &\by{\varepsilon}& H^4_{an}(X,\Z) &\to0 \\ \end{array}$$ \E{lemma} \smallskip\noindent{\bf Definition: } We shall refer to the map $\rho$ as the {\it Abel--Jacobi map}, because if $X$ is a smooth surface then $J^2(X)$ is the usual Albanese variety and the map $\rho$ coincides with the classical Abel--Jacobi map by \cite[1.9.1]{Bei} or \cite[2.24]{GDuke}. \medskip \B{proof} Observe that if $X$ has $c$ proper irreducible components then $H^4(X,\Z)\cong\Z^c$, because the singular locus of $X$ has real analytic dimension $\le2$. Given Lemma~\ref{c2-an}, the second assertion follows from the first. If $X$ is a smooth projective surface the result is classical; one way to see it is to use the product formula for two divisors on $X$: $$c_2^{an}(D\otimes E) = -c_1^{an}(D)\cup c_1^{an}(E) = - (D\cdot E)[X].$$ In general, choose a resolution of singularities $X'\to X$. Since $X'$ has $c$ disjoint components, the degree map on $X$ factors through the degree map on $X'$ as $CH_0(X)\to CH_0(\tilde X)\to\Z^c$. By naturality of $c_2^{an}$, the isomorphism $H^4_{an}(X,\Z)\cong H^4_{an}(X',\Z)$ allows us to deduce the result for $X$ from the result for $X'$. \E{proof} \B{num}\label{simplicial} As observed by Beilinson \cite[2.3]{Bei} ({\it cf.\/}\ \cite[\S5]{GDuke}), the formalism of Deligne cohomology allows us to extend the Chern classes from $K_0(X)$ to higher $K$-theory as well. The higher Deligne Chern classes are homomorphisms $$c_i\colon K_q(X) \to H^{2i-q}_{\cal D}(X,\Z(i)).$$ Composition with $\varepsilon_X$ yields the higher analytic Chern classes $c^{an}_i$, and reduction mod $n$ yields the higher \'etale Chern classes $c^{et}_i$. Moreover, the following holds. \B{description} \item{(\thethm.1)} There is a connected simplicial presheaf $K\simeq\Omega_0BQP$ and a simplicial sheaf ${\cal D}$ on $X_{zar}$ such that $\pi_q K(U) = K_q(U)$ for $q\ge1$, and $\pi_q{\cal D}(U) = H^{2i-q}(U_{\mbox{\Large $\cdot $}},\Z(i)_{\cal D})$ for $q\ge0$. Moreover, there is a map of simplicial presheaves $C_i^{ss}\colon K \to {\cal D}$ such that $\pi_q C_i^{ss}(X)$ is the Deligne cohomology Chern class $c_i$ on $K_q(X)$. ({\it cf.\/}\ \cite[5.4]{GDuke}, which differs somewhat from \cite{Bei} and \cite{GBAMS}.) Indeed, ${\cal D}$ is the simplicial sheaf of abelian groups associated by the Dold-Kan theorem (\cite[8.4.1]{W-homo}) to the good truncation $\tau^{\le0}\R\omega_*\Z(i)_{\cal D}[2i]$ of the total derived direct image of $\Z(i)_{\cal D}[2i]$ under $\omega\colon {X_{\bul}}{}_{,an}\to X_{zar}$. \item{(\thethm.2)} Let ${\cal E}$ denote the simplicial sheaf associated by the Dold-Kan theorem to the good truncation $\tau^{\le0}\R\omega_*\Z/n[2i]$ of the total derived direct image of $\Z/n[2i]$. Then $\pi_q{\cal E}(U)=H_{an}^{2i-q}(U,\Z/n) \cong\Het{2i-q}(U,\mu_n^{\otimes i})$. If we define $L$ to be the homotopy fiber of $K\by{n}K$ then we have $\pi_q L(U) = K_{q+1}(U;\Z/n)$. This all gives a homotopy commutative diagram whose rows are homotopy fibration sequences \B{equation}\label{fib}\B{array}{ccccccccc} \Omega K &\to& L &\to& K &\by{n}& K && \\ \kern18pt\downarrow\Omega C^{ss}_2\kern-10pt && \kern13pt\downarrow C^{ss}_2 \kern-10pt&& \kern13pt\downarrow C^{ss}_2 \kern-10pt&& \kern13pt\downarrow C^{ss}_2 \kern-8pt&& \\ \Omega{\cal D} &\to& \Omega{\cal E} &\by\delta&{\cal D}&\by{n}&{\cal D}& \by{\bar\varepsilon}&{\cal E}.\\ \E{array}\E{equation} From Corollary~\ref{c-etale} and a standard argument with $\Het{*}(X,G,\mu_n^{\otimes i})$ it is easy to see that not only does $K\to{\cal E}$ induce the higher \'etale Chern class $c^{et}_i$ on $K_*(X)$ but the map $L\to\Omega{\cal E}$ induces the usual \'etale Chern classes on $K$-theory with coefficients mod $n$. $$c^{et}_i\colon K_q(X;\Z/n)\longby{} \Het{2i-q}(X,\mu_n^{\otimes i}) $$ Applying $\pi_2$ to (\ref{fib}) with $i=2$ and $U=X$ yields the commutative diagram \B{equation}\label{et-HD}\B{array}{ccccccc} K_3(X)&\to&K_3(X;\Z/n) &\longby{}& K_2(X) &\by{n}& K_2(X) \\ &&\downarrow c^{et}_2 && \downarrow c_2 && \downarrow c_2 \\ 0&\to& \Het1(X,\mu_n^{\otimes2}) &\longby{\delta}& H_{\cal D}^2(X,\Z(2)) &\by{n}& H_{\cal D}^2(X,\Z(2)). \E{array}\E{equation} By (\ref{Kum-etHD}) we see that $c^{et}_2$ vanishes on $K_3(X)$ and factors through ${}_nK_2(X)$. Applying $\pi_2$ to (\ref{fib}) with $i=2$ and sheafifying yields the commutative diagram of sheaves in which the bottom row is part of (\ref{Rq}): \B{equation}\label{et-HD-sheaf}\B{array}{ccccccc} {\cal K}_3&\to&{\cal K}_3(\Z/n) &\longby{}& \kern-5pt{\cal K}_2 &\by{n}& \kern-5pt{\cal K}_2 \\ &&\downarrow c^{et}_2 && \downarrow c_2 && \downarrow c_2 \\ 0&\to& {\cal H}^1(\mu_n^{\otimes2}) &\longby{\delta}& {\cal H}_{\cal D}^2(2) &\by{n}& {\cal H}_{\cal D}^2(2) \E{array}\E{equation} By (\ref{Rq}) we see that $c^{et}_2$ vanishes on ${\cal K}_3$ and factors through the sheaf ${}_n{\cal K}_2$. \item{(\thethm.3)} There is a morphism of spectral sequences between the Brown--Gersten spectral sequence for $K_{-*}(X)$ and the Leray spectral sequence in (\ref{simps}) converging to $H^{2i+*}_{\cal D}(X,\Z(i))$. At the $E_2^{pq}$-level the morphisms are the cohomology of $c_i$: $$H^p_{zar}(X,{\cal K}_{-q}) \by{c_i} H^p_{zar}(X,{\cal H}_{\cal D}^{2i+q}(i)) .$$ Here ${\cal K}_q$ is the sheaf on $X_{zar}$ associated to the presheaf $K_q$ and the sheaves ${\cal H}_{\cal D}^j(i)$ are $R^j\omega_*\Z(i)_{\cal D}$, as in the proof of Proposition~\ref{sheafHD}. By \cite{TT}, the first spectral sequence converges to $K_{-p-q}(X)$ whenever $X$ is quasi-projective. The second spectral sequence is an obvious reindexing of (\ref{simps}) and converges to $H_{\cal D}^{2i+p+q}(X,\Z(i))$. \E{description} \E{num} Here are three applications of the morphism of spectral sequences in (2.4.3). First, if $X$ is a projective surface we have a commutative diagram $$\begin{array}{ccccc} CH_0(X)&\cong\kern-8pt& H^2(X,{\cal K}_2) &\hookrightarrow& K_0(X)\\ \strut & & c_2\downarrow& & c_2\downarrow \\ & & H^2(X,{\cal H}_{\cal D}^2(2)) &\to & H_{\cal D}^4(X,\Z(2)).\\ \end{array}$$ where the bottom horizontal map is given by Proposition~\ref{sheafHD}$(vi)$. Second, suppose that $Y$ is 1-dimensional. \kern-1pt Then we may identify the group $H^1\kern-1pt(X\kern-1pt,{\cal K}_2\kern-1pt)$ with the subgroup $SK_1(X)$ of $K_1(X)$, and $c_2\colon SK_1(X)\to H_{\cal D}^3(X,\Z(2))$ is identified with the composite $H^1(X,{\cal K}_2)\by{c_2}H^1(X,{\cal H}_{\cal D}^2(2))\to H_{\cal D}^3(X,\Z(2))$. Third, suppose that $X$ is an irreducible projective surface. Then $c_2$ vanishes on the image of $H^2(X,{\cal K}_3)$ in $K_1(X)$ because it factors through $H^2(X,{\cal H}_{\cal D}^1(2))=H_{zar}^2(X,\C^*)$, which is zero because $X$ is irreducible, as we saw in the proof of Proposition~\ref{sheafHD}. Since $SK_1(X)$ is an extension of $H^1(X,{\cal K}_2)$ by this image, we may summarize this as follows. \B{lemma}\label{c2-SK1} Let $X$ be an irreducible projective surface over $\C$. Then the Chern class $c_2\colon SK_1(X)\to H_{\cal D}^3(X,\Z(2))$ factors as: $$SK_1(X)\vlongby{\mbox{\rm onto}}H^1(X,{\cal K}_2) \by{c_2}H^1(X,{\cal H}_{\cal D}^2(2))\hookrightarrow H_{\cal D}^3(X,\Z(2)). $$ \E{lemma} \goodbreak \section{Mayer--Vietoris sequences} Since we are going to deal with resolutions of singularities or normalizations we will need some Mayer--Vietoris sequences. In this section we do this for mixed Hodge structures, Deligne cohomology and $K$-theory. Associated to a proper birational morphism $f\colon X'\to X$ of $\C$-algebraic schemes, and every closed subscheme $i\colon Y\hookrightarrow X$ we have the commutative square \B{equation}\label{birsquare} \B{array}{rcl} Y'& {\stackrel{i'}{\hookrightarrow}}& X'\\ \p{f'}\downarrow & &\downarrow\p{f}\\ Y &{\stackrel{i}{\hookrightarrow}}& X \E{array} \E{equation} where $Y' = f^{-1}(Y)\ (\ =Y\times_X X')$. We shall always assume that $Y$ is chosen so that the restriction $f\colon X'-Y'\by{\simeq} X-Y$ is an isomorphism. \goodbreak \B{prop}\label{M-V-H} {\rm (Mayer--Vietoris for mixed Hodge structures)\,} Associated with any square (\ref{birsquare}) we have a long exact sequence of mixed Hodge structures $$\cdots\to H^n(X,\Z)\by{u}H^n(X',\Z)\oplus H^n(Y,\Z)\by{v} H^n(Y',\Z)\by{\partial} H^{n+1}(X,\Z)\to\cdots $$ in which $$ u=\left(\begin{array}{c}{f^*}\\{i^*}\end{array} \right) \qquad{\rm and}\qquad v= (i'^*,-f'^*)$$ \E{prop} \B{proof} We have a map of long exact sequences $$ \begin{array}{ccccccccc} \cdots & \to & H^n(X{\rm mod\,} Y,\Z) & \to & H^n(X,\Z) & \by{i^*} & H^{n}(Y,\Z) & \to & \cdots \\ & &\p{f^*}\downarrow\cong & &{\p{f^*}\downarrow} & & {\p{f'^*}\downarrow} & &\\ \cdots & \to & H^n(X'{\rm mod\,} Y',\Z) & \to & H^n(X',\Z) & \by{i'^*} & H^{n}(Y',\Z) & \to & \cdots \end{array} $$ where $H^*(-{\rm mod\,}\dag ,\Z)$ is the relative singular cohomology functor (defined in \cite[8.3.8]{D}). By excision $f^*\colon H^*(X{\rm \,mod\,}Y,\Z)\cong H^*(X'{\rm\,mod\,}Y',\Z)$ ({\it cf.\/}\ \cite[8.3.10]{D}). By \cite[8.3.9 and 8.2.2]{D} the diagram above is a diagram in the abelian category of mixed Hodge structures. The Mayer--Vietoris exact sequence now follows by a standard diagram chase. \E{proof} We then have, as well: \B{schol} \label{M-V-D} {\rm (Mayer--Vietoris for Deligne cohomology)\,} Associated with any square (\ref{birsquare}) we have a long exact sequence in Deligne cohomology $$\cdots\to H_{{\cal D}}^n(X',\Z(r))\oplus H_{{\cal D}}^n(Y,\Z(r)) \to H_{{\cal D}}^n(Y',\Z(r))\to H_{{\cal D}}^{n+1}(X,\Z(r))\to\cdots $$ \E{schol} \B{proof} The proof of Proposition~\ref{M-V-H} goes through, once we know that Deligne cohomology satisfies excision. But since we have excision for the mixed Hodge structure on relative singular cohomology, one can see it holds for Deligne cohomology by arguing with the relative cohomology sequence (\ref{modf-rel}). \E{proof} \B{thm}\label{M-V-K} {\rm (Mayer--Vietoris for $K$-theory)\,} Let $X$ be a reduced quasiprojective surface over a field with normalization $\tilde X$. Then there is a 1-dimensional subscheme $Y$ with $Y_{red}={\rm Sing}\, X$ such that the normalization square ({\it cf.\/}\ (\ref{birsquare})) $$\B{array}{rcl} \tilde Y& \hookrightarrow & \tilde X\\ \p{\tilde \pi}\downarrow & &\downarrow\p{\pi}\\ Y&\hookrightarrow & X \E{array}$$ induces exact sequences in $K$-theory: $ K_1(\tilde X)\oplus K_1(Y)\to K_1(\tilde Y)\by{\partial} K_0(X)\to K_0(\tilde X)\oplus K_0(Y) \to K_0(\tilde Y) $ $ SK_1(\tilde X)\oplus SK_1(Y)\to SK_1(\tilde Y)\by{\partial} SK_0(X)\to SK_0(\tilde X)\to 0 $ \E{thm} \B{proof} Let $K_*(X,\tilde X)$ and $K_*(Y,\tilde Y)$ be the relative groups fitting into the long exact sequences in the commutative diagram $$\begin{array}{ccccccccccc} K_1(X) &\to& K_1(\tilde X) &\to& K_0(X,\tilde X) &\to& K_0(X) &\to& K_0(\tilde X) &\to& K_{-1}(X,\tilde X)\\ \downarrow && \downarrow&&\downarrow &&\downarrow && \downarrow&&\mbox{\large $\parallel$}\\ K_1(Y) &\to& K_1(\tilde Y) &\to& K_0(Y,\tilde Y) &\to& K_0(Y) &\to& K_0(\tilde Y) &\to& K_{-1}(Y,\tilde Y). \\ \end{array}$$ (The far right terms are isomorphic by \cite[A.6]{PW3}.) To establish the existence and exactness of the $K_1-K_0$ sequence we must show that ``excision'' holds for $K_0$, {\it i.e.,\/}\ that $K_0(X,\tilde X)\cong K_0(Y,\tilde Y)$ for some $Y$ with $Y_{red}={\rm Sing}\, X$ (see \cite[5.1]{GW}). If $Y$ is a subscheme of $X$ defined by an ${\cal O}_{\tilde X}$-ideal ${\cal I}\subset{\cal O}_X$ then by \cite[A.6]{PW3} there is a natural exact sequence \B{equation}\label{K0rel} H^1(Y,{{\cal I}}/{{\cal I}}^2\otimes\Omega_{\tilde X/X}) \vlongby{\eta(Y)} K_0(X,\tilde X) \to K_0(Y,\tilde Y) \to0. \E{equation} We define $Y_1$ using the conductor ideal ${\cal J}$, and $Y$ using the ideal ${{\cal I}}={{\cal J}}^2$. Then $Y_{red}={\rm Sing}\, X$, and the map from ${\cal I}/{\cal I}^2$ to ${\cal J}/{\cal J}^2$ is zero. By naturality in $Y\to Y_1$, the map $\eta(Y)$ in (\ref{K0rel}) is the composite map $ H^1(Y,{\cal I}/{\cal I}^2\otimes\Omega_{\tilde X/X}) \by0 H^1(Y,{\cal J}/{\cal J}^2\otimes\Omega_{\tilde X/X}) \vlongby{\eta(Y_1)} K_0(X,\tilde X)$$ so $\eta(Y)=0$ in (\ref{K0rel}). Hence excision holds for $Y$, as claimed. There is a natural map from the $K_1-K_0$ sequence onto the ``Units-Pic'' sequence, and the kernel is the $SK_1-SK_0$ sequence. A standard diagram chase, described in \cite[8.6]{PW1}, shows that the latter sequence sequence is also exact. \E{proof} We remark that if $Y$ is reduced, or 0-dimensional, or even affine, then the obstruction $H^1(Y)$ in (\ref{K0rel}) automatically vanishes, and excision is immediate. Theorem \ref{M-V-K} was proven in these special cases in \cite[7.5]{PW1} and \cite[A.3]{PW3}. \B{cor}\label{KtoD} With the notation of \ref{M-V-K}, the following diagram commutes. $ \begin{array}{cccccccc} SK_1(\tilde X)\oplus SK_1(Y) &\kern-8pt\to\kern-9pt& SK_1(\tilde Y) &\kern-7pt\to\kern-8pt &SK_0(X) & \kern-7pt\to\kern-8pt & SK_0(\tilde X) &\kern-7pt\to0\\ \downarrow&&\downarrow&&\downarrow&&\downarrow&\\ K_1(\tilde X)\oplus K_1(Y) &\kern-8pt\to\kern-9pt& K_1(\tilde Y) &\kern-7pt\to\kern-8pt &K_0(X) & \kern-7pt\to\kern-8pt & K_0(\tilde X) &\kern-7pt\to0\\ c_2\downarrow\ &&c_2\downarrow\ &&c_2\downarrow\ &&c_2\downarrow\ &\\ H_{\cal D}^3(\tilde X,\Z(2))\oplus H_{\cal D}^3(Y,\Z(2)) & \kern-8pt\to\kern-9pt & H_{\cal D}^3(\tilde Y,\Z(2))& \kern-7pt\to\kern-8pt & H_{\cal D}^4(X,\Z(2)) &\kern-7pt\to\kern-8pt & H_{\cal D}^4(\tilde X,\Z(2)) &\kern-7pt\to0 \end{array} $ \E{cor} \B{proof} We use the notation of (2.4.1). For each open $U$ in $X$, let $F(U)$ denote the homotopy fiber of $K(U\times_X Y)\times K(U\times_X\tilde X) \to K(U\times_X\tilde Y)$. By Proposition \ref{M-V-D} the corresponding homotopy fiber for Deligne cohomology is ${\cal D}(U)$. In addition, there is a natural map from $K(U)$ to $F(U)$ which is an isomorphism on $\pi_0$ by Theorem \ref{M-V-K}. Therefore the natural map $C_2^{ss}$ of (2.4.1) induces a map $F(U)\to{\cal D}(U)$ on homotopy fibers, making the diagram $$\begin{array}{ccccc} K_1(\tilde Y) &\by{\partial}& \pi_0F(X) &{\stackrel{\simeq}{\leftarrow}}& K_0(X)\\ \strut C_2^{ss}\downarrow\ &&\ \downarrow C_2^{ss}&&\ \downarrow C_2^{ss}\\ H_{\cal D}^3(\tilde Y,\Z(2)))&\by{\partial}& \pi_0{\cal D}(X) &{\stackrel{\simeq}{\leftarrow}}& H_{\cal D}^4(X,\Z(2)) \end{array}$$ commute. But the top composite is the $K$-theory boundary map in Theorem \ref{M-V-K}. \E{proof} Using (2.4) and Lemma~\ref{c2-SK1}, we may refine Corollary~\ref{KtoD} as follows. \B{schol}\label{SKtoD} With the notation of \ref{M-V-K}, the following diagram commutes. $ \begin{array}{cccccccc} SK_1(\tilde X)\oplus SK_1(Y) &\kern-8pt\to\kern-9pt& SK_1(\tilde Y) &\kern-7pt\to\kern-8pt &SK_0(X) & \kern-7pt\to\kern-8pt & SK_0(\tilde X) &\kern-7pt\to0\\ \mbox{\rm onto}\downarrow\ \ &&\cong\downarrow&&\cong\downarrow&& \cong\downarrow&\\ H^1(\tilde X,{\cal K}_2)\oplus H^1(Y,{\cal K}_2) &\kern-8pt\to\kern-9pt& H^{1}(\tilde Y,{\cal K}_2) &\kern-7pt\to\kern-8pt & H^{2}(X,{\cal K}_2) & \kern-7pt\to\kern-8pt & H^{2}(\tilde X,{\cal K}_2)& \\ c_2\downarrow\ &&c_2\downarrow\ &&c_2\downarrow\ &&c_2\downarrow\ &\\ H_{\cal D}^3(\tilde X,\Z(2))\oplus H_{\cal D}^3(Y,\Z(2)) & \kern-8pt\to\kern-9pt & H_{\cal D}^3(\tilde Y,\Z(2))& \kern-7pt\to\kern-8pt & H_{\cal D}^4(X,\Z(2)) &\kern-7pt\to\kern-8pt & H_{\cal D}^4(\tilde X,\Z(2)) &\kern-7pt\to0 \end{array} $ \E{schol} \goodbreak \section{The Albanese 1-motive of a proper surface} In this section a {\it surface}\, will mean a \underline{proper} reduced $2$-dimensional scheme $X$ of finite type over the complex numbers $\C$. We will consider the intermediate jacobian $$J^2(X) \mbox{\,$\stackrel{\pp{\rm def}}{=}$}\, \frac{H^{3}(X,\C)}{F^2H^{3}+ H^{3}(X,\Z(2))}$$ This is the mixed Hodge theoretic generalization of the classical Albanese group variety of a smooth surface. We begin with an elementary result (cf. \cite[Remark 5.5]{GDuke}). \B{lemma}\label{filt} Suppose that $X$ is a proper surface. Then $ F^2H^i(X,\C)\cap H^i(X,\R)=0 \mbox{\ \ for \ } i=2,3. $ Hence in sequence (\ref{modf}) we have $$\B{array}{ccc} H^i(X,\Z)_{tors} \strut&=&\mbox{kernel of } H^i(X,\Z)\to H^i(X,\C)/F^2H^i\\ \strut&=&\mbox{image of } H_{\cal D}^i(X,\Z(2))\by{\varepsilon}H_{an}^i(X,\Z). \E{array}$$ \E{lemma} \kern-6pt \B{proof} We will show that $H^i(X,\R)$ injects into $H^i(X,\C)/F^2$. When $X$ is smooth, then $H^i(X)$ has pure weight $i$. In this case complex conjugation on $H^i(X,\C)$ fixes $H^i(X,\R)$ but the subspace $F^2H^i(X,\C)$ meets its conjugate in $0$. If $X$ is a singular surface, choose a resolution of singularities $X'\to X$. If $Y$ is a curve containing the singular locus of $X$, then we are in the situation of square (\ref{birsquare}). Since $F^2H^1=F^2H^2=0$ for the curves $Y$ and $Y'$, the Mayer--Vietoris sequence in Proposition~\ref{M-V-H} yields $F^2H^i(X,\C)=F^2H^i(X',\C)$ for $i=2,3$. Comparing the $\R$ and $\C$ structures in the Mayer--Vietoris long exact sequence of Proposition~\ref{M-V-H} yields the following diagram has exact rows: $$\begin{array}{ccccc} H^1(Y',\R)&\to&H^2(X,\R) & \to & H^2(X',\R)\\ \downarrow & &\downarrow & &\downarrow \\ H^1(Y',\C)&\to& \displaystyle\frac{H^2(X,\C)}{F^2H^2}& \to &\displaystyle\frac{H^2(X',\C)}{F^2H^2} \end{array}$$ The right-most vertical arrow in the diagram is injective because $X'$ is smooth. A diagram chase shows that the middle vertical arrow is injective, whence the lemma. \E{proof} \subsection*{Normal surfaces} Consider a surface $X$ with normal singularities; its singular locus $\Sigma$ is a finite set of closed points. Choose a desingularization $f\colon X' \to X$ and consider the exceptional divisor $E=f^{-1}(\Sigma)$; $E$ is the finite disjoint union of the inverse images $E_{\sigma}\mbox{\,$\stackrel{\pp{\rm def}}{=}$}\, f^{-1}(\sigma)$ of the $\sigma\in \Sigma$. Associated to $f$ is the square (\ref{birsquare}), with $Y=\Sigma$ and $Y'=E$. Because the fibers $E_\sigma$ of $f$ are connected (by Zariski's Main Theorem), we have $H^0(\Sigma,\Z)\cong H^0(E,\Z)$. From Proposition~\ref{M-V-H} we get a long exact sequence of mixed Hodge structures: \B{equation}\label{exnorm} \B{array}{c} 0\to H^1(X,\Z)\by{f^*} H^1(X',\Z)\to H^1(E,\Z)\to H^2(X,\Z)\by{f^*} \qquad\\ \qquad\qquad H^2(X',\Z)\to H^2(E,\Z)\to H^3(X,\Z)\by{f^*} H^3(X',\Z) \to 0 \E{array} \E{equation} \noindent Recall that if $X$ is proper then each $H^n(X')$ has pure weight $n$, and that $W_{n-1}H^n(X,\Q)$ is the kernel of $f^*\colon H^n(X,\Q)\to H^n(X',\Q)$ by \cite[8.2.5]{D}. That is, $H^n(X)$ has pure weight $n$ if and only if $H^n(X,\Q)$ injects into $H^n(X',\Q)$. There are examples of normal surfaces for which $H^2(X)$ does not have pure weight 2, i.e., with $W_1H^2\neq 0$ ({\it cf.\/}\ \cite{BVS1},\cite{BVS2}). The following result quantifies this impurity. \B{prop}\label{pure} Let $X$ be a proper normal surface. If $n\ne2$ then $H^n(X)$ has pure weight $n$. If $n=2$ and $E$ is the exceptional divisor in a desingularization $X'$, then $$W_{1}H^2(X,\Q)=\coker H^1(X',\Q)\to H^1(E,\Q).$$ \E{prop} \B{proof} If $n\ne3$ this follows from the sequence of mixed Hodge structures (\ref{exnorm}). For $n=3$ we must show that $H^3(X,\Q)$ embeds in $H^3(X',\Q)$. Nothing is lost if we replace $X'$ by a quadratic transformation, so we may assume that all the irreducible components $E_1,\dots,E_n$ of $E$ are non-singular and that if $i\neq j$ and $E_i\cap E_j \neq\emptyset$ then $E_i$ and $E_j$ intersect transversally in exactly one point not belonging to any other $E_k$. From (\ref{exnorm}) and the commutative square $$\B{array}{ccc} \Pic(X') &\by{\cap E} & \Pic(E) \\ \p{c_1}\downarrow& & \downarrow \p{c_1}\\ H^2(X',\Z) &\to & H^2(E,\Z) \E{array}$$ we see that it suffices to prove that $\Pic(X')\otimes\Q \to H^2(E,\Q)$ is a surjection. Now $H^2(E,\Q)=\oplus H^2(E_i,\Q) \cong \Q^n$, and $\Pic(E)\otimes\Q \by{c_1} H^2(E,\Q)$ is just the degree map. Moreover, the intersection pairing on the divisors on $X'$ satisfies $(D\mbox{\Large $\cdot $} E_i) ={\rm deg}(D\cap E_i)$. Thus if we represent an element of $\Pic(X')$ by a divisor $D$, its image in $H^2(E,\Q)\cong\Q^n$ is given by the intersection vector $(D\mbox{\Large $\cdot $} E_1,\dots,D\mbox{\Large $\cdot $} E_n)$. Now each $E_i$ represents an element of $\Pic(X')$, and their intersection vectors form a basis of $H^2(E,\Q)$ because the intersection matrix $(E_i\mbox{\Large $\cdot $} E_j)$ is negative definite (see \cite[\S 1]{M} or \cite[14.1]{Lip}). \E{proof} \B{cor}\label{abnorm} Let $f\colon X' \to X$ be a resolution of singularities of a proper normal surface. Then $J^2(X)$ is an abelian variety, because there is an isomorphism $$f^*\colon J^2(X)\by{\simeq} J^2(X')$$ \E{cor} \B{proof} By (\ref{exnorm}) and \ref{pure}, $f^*\colon H^3(X,\Z)\by{} H^3(X',\Z)$ is onto with torsion kernel. \E{proof} \subsection*{Normalization} Now let $X$ be a non-normal surface. The singular locus $\Sigma$ of $X$ is $1$-dimensional. Letting $\pi: \tilde X \to X$ denote its normalization, we have $\pi\colon\tilde X-\tilde\Sigma\by{\simeq}X-\Sigma$, where $\tilde\Sigma=\pi^{-1}\Sigma$. By Proposition~\ref{M-V-H}, $\pi$ induces a long exact sequence of mixed Hodge structures. \B{equation}\label{exnon-norm} \B{array}{c} H^1(X,\Z)\to H^1(\tilde X,\Z)\oplus H^1(\Sigma ,\Z)\to H^1(\tilde\Sigma ,\Z)\to H^2(X,\Z)\to\\ H^2(\tilde X,\Z)\oplus H^2(\Sigma,\Z)\to H^2(\tilde\Sigma ,\Z) \to H^3(X,\Z) \to H^3(\tilde X,\Z) \to 0 \E{array} \E{equation} Since the Hodge structure on $H^2$ of a curve is pure of type $(1,1)$, the abelian group $$M = \frac{\coker H^2(\Sigma,\Z)\by{\pi^*}H^2(\tilde\Sigma,\Z)} {\coker H^2(X,\Z)\by{\pi^*}H^2(\tilde X,\Z)}$$ has a mixed Hodge structure which is pure of type $(1,1)$, and there is an extension of mixed Hodge structures \B{equation}\label{exMHS} 0\to M \to H^3(X,\Z) \to H^3(\tilde X,\Z) \to0. \E{equation} \B{prop}\label{non-norm} Let $X$ be a proper surface, with normalization $\pi\colon\tilde X \to X$. Then we have an extension $$0\to (\C/\Z)^s \to J^2(X)\by{\pi^*} J^2(\tilde X)\to 0$$ where $s$ is the rank of the abelian group $M$. \E{prop} \B{proof} $J^2(X)$ is the cokernel of the natural map $H^3(X,\Z(2))\to H^3(X,\C)/F^2H^3$. Given this, the result is a formal consequence of (\ref{exMHS}) and the fact that $F^2M=0$, which implies that $F^2H^3(X,\C) \cong F^2H^3(\tilde X,\C)$. We remark that the complex torus $(\C/\Z)^s$ that arises in this extension is a quotient of the complex torus $(\C/\Z(2))^s = M\otimes(\C/\Z(2))$ by a finite group. \E{proof} \B{cor}\label{1-motive} Let $f\colon X'\to X$ be a desingularization of a proper surface $X$, obtained by resolving the singularities of its normalization $\tilde X$. Then there is an exact sequence $$0\to (\C/\Z)^s \to J^2(X)\by{f^*} J^2(X')\to 0$$ where $s$ is the rank of $M$, as in Proposition~\ref{non-norm}. In particular, if $X'$ has irregularity $q$ then the torsion subgroup of $J^2(X)$ is isomorphic to $(\Q/\Z)^{2q+s}$. \E{cor} Recall from \cite[10.1.2]{D} that a ``1-motive'' $M=(L,A,T,J,u)$ is defined to be an extension $J$ of an abelian variety $A$ by a complex torus $T$, a lattice $L$ and a homomorphism $L\by{u} J$. Since we may canonically identify the group of $\C$-points of the abelian variety $Alb(X')$ with $J^2(X')$, the conclusion of Corollary~\ref{1-motive} is just that $J^2(X)$ is part of a 1-motive $Alb(X)$ in which the lattice $L$ is zero. \B{defi} Let $X$ be a proper surface over $\C$. The {\it Albanese 1-motive} of $X$ is the 1-motive $Alb(X)$ given by $$(0,Alb(X'),(\C/\Z(2))^s,J^2(X),{\rm zero}).$$ As the construction in \ref{non-norm} shows, $Alb$ is a functor from proper surfaces to 1-motives. \E{defi} \goodbreak \subsection*{Torsion in $J^2(X)$} For simplicity, let us write $\Q/\Z$ for the torsion subgroup $\Q(2)/\Z(2)$ of $\C/\Z(2)$, so that $H^i(-,\Q/\Z)\cong H^i(-,\C/\Z(2))_{tors}$. The maps $H^i(-,\C/\Z(2))\to H_{\cal D}^{i+1}(-,\Z(2))$ of (\ref{star}) induce canonical maps $$H^i(-,\Q/\Z)\cong H^i(-,\C/\Z(2))_{tors}\to H_{\cal D}^{i+1}(-,\Z(2))_{tors}.$$ These are the maps in the following Proposition: \B{prop}\label{HD-tors} Let $Z$ be a proper scheme over $\C$. Then \B{description} \item[{\it i)}] $H^1(Z,\Q/\Z)\by{\cong} H_{\cal D}^2(Z,\Z(2))_{tors}$ \item[{\it ii)}] If $Z$ is either a curve or a surface then $$H^2(Z,\Q/\Z)\by{\cong} H_{\cal D}^3(Z,\Z(2))_{tors} $$ \item[{\it iii)}] If $Z$ is a surface then $$H^3(Z,\Q/\Z)\by{\cong}H_{\cal D}^4(Z,\Z(2))_{tors}\cong J^2(Z)_{tors}$$ \E{description} \E{prop} \B{proof} The first assertion was proven in \ref{sheafHD}. If $Z$ is a curve then $F^2H^2(Z,\C)=0$, so by (\ref{star}) we have $H_{\cal D}^3(Z,\Z(2))\cong H^2(Z,\C/\Z(2))$, and the result is immediate. We may therefore suppose that $Z$ is a surface, say with $c$ irreducible components, so that $H^4(Z,\Z)=\Z^c$. We deduce from (\ref{extJZ}) that $J^2(Z)_{tors}\cong H_{\cal D}^4(Z,\Z(2))_{tors}$. Moreover, since $F^2H^4(Z,\C)=H^4(Z,\C)=\C^c$ the sequence (\ref{star}) ends in \B{equation}\label{star3} H_{\cal D}^3(Z,\Z(2))\to F^2H^3(Z,\C)\to H^3(Z,\C/\Z(2))\to H_{\cal D}^4(Z,\Z(2)) \by{\varepsilon}\Z^c\to0. \E{equation} Lemma \ref{filt} states that for $i=2,3$ the image of $\varepsilon$ in the exact sequence $$H^{i-1}(Z,\C)/F^2H^{i-1}\to H_{\cal D}^i(Z,\Z(2))\by{\varepsilon} H^i(Z,\Z)\to H^i(Z,\C)/F^2H^i$$ of (\ref{modf}) is the torsion subgroup $H^i(Z,\Z)_{tors}$. Combining this with the universal coefficient theorem, we have a commutative diagram with exact rows: $$\begin{array}{ccccccc} 0\to\kern-7pt&H^{2}(Z,\C)/F^2H^{2}\oplus H^{2}(Z,\Z) &\to&H_{\cal D}^3(Z,\Z(2)) &\by{\varepsilon}&H^3(Z,\Z)_{tors}&\to0\\ &\uparrow & &\uparrow & &\mbox{\large $\parallel$} & \\ 0\to\kern-7pt&H^{2}(Z,\Z)\otimes\C/\Z(2) &\to & H^{2}(Z,\C/\Z(2))&\to & H^3(Z,\Z)_{tors}&\to 0. \end{array}$$ By the five-lemma, sequence (\ref{star}) and (\ref{star3}) we get the extensions \B{equation} 0\to F^2H^2(Z,\C)\to H^{2}(Z,\C/\Z(2))\to H_{\cal D}^3(Z,\Z(2))\to0, \E{equation} \B{equation} 0\to F^2H^3(Z,\C)\to H^3(Z,\C/\Z(2))\to J^2(Z)\to0. \E{equation} Since $F^2H^i(Z)$ is uniquely divisible, we may pass to torsion subgroups. This proves the remainder of the proposition. \E{proof} \medski \section{Curves} The singular locus of a reduced surface is usually an (unreduced) curve. For this reason, we need information about $K_1$ and $K_2$ of curves in order to study surfaces. This information is given by theorems \ref{smooth} and \ref{sing} below. Part {\it i)\,} of Theorem \ref{smooth} is of course well-known and almost classical; a reference is \cite[1.1]{R}. Since these results are of independent interest, we have expanded our exposition to include the case of characteristic $p$. By a `curve' over a field $k$ we mean a $1$-dimensional quasiprojective scheme $Y$ over $k$; a curve is not necessarily reduced. There is a natural map from $K_1(Y)$ to the group $H^0(Y,{\cal O}_Y)$ of global units of $Y$; the kernel of this map is usually written as $SK_1(Y)$. When $Y$ is a curve there is a natural isomorphism $SK_1(Y)\cong H^1(Y,{\cal K}_2)$, as well as a natural short exact sequence $$0\to H^1(Y,{\cal K}_3)\to K_2(Y)\to H^0(Y,{\cal K}_2)\to 0$$ given by the Brown-Gersten spectral sequence \cite{TT}. \B{thm}\label{smooth} Let $Y$ be a smooth curve over an algebraically closed field $k$. Let $r\geq 0$ denote the number of irreducible components of $Y$ which are proper. Then \B{description} \item[{\it i)}] $SK_1(Y)\cong(k^*)^r\oplus V_1$ where $V_1$ is a uniquely divisible group; \item[{\it ii)}] $K_2(Y)$ and $H^0(Y,{\cal K}_2)$ are both divisible abelian groups. \E{description} \E{thm} \noindent\B{proof If $Y$ is a smooth connected curve over an algebraically closed field $k$ then the localization sequence is \B{equation}\label{localization} \coprod_{y\in Y(k)}^{} K_2(k) \to K_2(Y) \to K_2(k(Y))\by{tame} \coprod_{y\in Y(k)}^{} k^* \to SK_1(Y)\to 0 \E{equation} and the image of $K_2(Y)$ in $K_2(k(Y))$ is $H^0(Y,{\cal K}_2)$. Since $\coprod K_2(k)$ is divisible \cite[1.3]{BT}, the divisibility of $K_2(Y)$ is equivalent to the divisibility of $H^0(Y,{\cal K}_2)$. If char$(k)=0$, the result now follows from Suslin's exact sequence \cite[4.4]{S2} for $n$ invertible in $k$: $$0\to H^0(Y,{\cal K}_2)/n\to \Het{2}(Y,\mu_n^{\otimes2})\to {}_nSK_1(Y) \to0$$ Indeed, if $Y$ is affine then $\Het{2}(Y)=0$, and if $Y$ is projective then the composite $\mu_n \cong \Het{2}(Y,\mu_n^{\otimes2})\to SK_1(Y)\to k^*$ is the standard inclusion. If char$(k)=p>0$, we need only a slight additional argument. Because $k(Y)$ is the function field of a curve, we know from \cite[p.391]{BT} that $K_2(k(Y))$ is $p$-divisible, and from \cite[1.10]{S2} (which is implicit in \cite[p.397]{BT}) that it has no $p$-torsion. Hence both $K_2(k(Y))$ and $\coprod k^*$ are uniquely $p$-divisible groups, {\it i.e.,\/}\ $\Z[\frac{1}{p}]$-modules. It follows that both the kernel $H^0(Y,{\cal K}_2)$ and cokernel $SK_1(Y)$ of the `tame symbol' map in (\ref{localization}) must be uniquely $p$-divisible. This proves Theorem~\ref{smooth} in characteristic $p$. \E{proof} \B{lemma} Let $Y$ be a smooth connected projective curve over $\C$. Then $$c_2\colon SK_1(Y) \to H_{\cal D}^3(Y,\Z(2))\cong \C^*$$ is a split surjection. In particular, it is an isomorphism on torsion subgroups. \E{lemma} \B{proof} This is implicit in p.219 of Gillet's paper \cite{GDuke}. The isomorphism $H_{\cal D}^3(Y,\Z(2))\cong \C/\Z(2)\cong\C^*$ follows from (\ref{star}) or \ref{HD-tors}. If $y\in Y(\C)$ is considered as an element of $Pic(Y)$ and $z\in\C^*$ then we can form $\{ y,z\}\in SK_1(Y)$ and the product formula yields $c_2(\{ y,z\})=-c_1(y)\cup c_1(z) = z^{-1}\in \C^*$. \E{proof} \B{thm}\label{sing} Let $Y$ be any curve over an algebraically closed field $k$, and let $r\geq 0$ denote the number of irreducible components of $Y$ which are proper. Then \B{description} \item[{\it i)}] If char$(k)=0$, or if $Y$ is reduced, then $$SK_1(Y)\cong (k^*)^r\oplus V_1,$$ where $V_1$ is a uniquely divisible abelian group; \item[{\it ii)}] If char$(k)=p>0$ then $$SK_1(Y)\cong (k^*)^r\oplus V_1\oplus P,$$ where $V_1$ is uniquely divisible and $P$ is a $p$-group of bounded exponent. \item[{\it iii)}] If $k=\C$ then the Chern class $$c_2\colon SK_1(Y) \to H_{\cal D}^3(Y,\Z(2))\cong (\C^*)^r$$ is a split surjection. In particular, it is an isomorphism on torsion subgroups. \E{description} \E{thm} \noindent\B{proof We proceed in three steps.\\ {\it Step 1.} Suppose that $Y$ is any reduced curve over $k$. If we pick $r$ smooth points $y_i$ on $Y$, one on each proper component of $Y$, then $Y_0=Y-\{y_1,\ldots ,y_r\}$ is affine. The localization sequence for $Y_0\subset Y$ is \B{equation}\label{notloc} \coprod_{i=1}^{r} K_2(k) \to K_2(Y) \to K_2(Y_0)\to \coprod_{i=1}^{r} k^* \to SK_1(Y)\to SK_1(Y_0)\to 0 \E{equation} If $\tilde Y$ is the normalization of $Y$, then we may indentify the $y_i$ with points on the smooth curve $\tilde Y$. By the smooth case \ref{smooth}, the composition of $$\coprod k^* \to SK_1(Y)\to SK_1(\tilde Y)$$ is an injection. Hence $SK_1(Y)$ is the direct sum of $\coprod k^*$ and $SK_1(Y_0)$ while $K_2(Y)$ is the direct sum of the image of the divisible group $\coprod K_2(k)$ and the group $K_2(Y_0)$. Part (iii) now follows from the above Lemma. This argument also shows that we may replace $Y$ by $Y_0$ in proving parts (i) and (ii) of Theorem~\ref{sing} for reduced curves.\\ {\it Step 2.} Now suppose that $Y= \mbox{Spec}(A)$ is any reduced affine curve over $k$. Let $B$ be the normalization of $A$, and $I$ the conductor ideal from $B$ to $A$. By \cite[3.1 and 4.2]{GR}, excision holds for $K_1$ and there is an exact sequence $$K_2(B)\oplus K_2(A/I)\to K_2(B/I)\to SK_1(A)\to SK_1(B)\to 0$$ Since $B$ is a finite product of Dedekind domains, $B/I$ is a finite principal ideal ring. By Corollary~\ref{k2div} below, $K_2(B/I)$ is uniquely divisible. By Theorem~\ref{smooth}, $SK_1(B)$ is uniquely divisible and $K_2(B)$ is divisible. Finally, since $A/I$ is finite dimensional, we know from Corollary~\ref{k2div} that $K_2(A/I)$ is uniquely divisible (modulo bounded $p$-torsion if char$(k)=p\neq 0$). A diagram chase shows that $SK_1(A)$ is uniquely divisible (modulo bounded $p$-torsion if $p\ne0$). This proves theorem \ref{sing} for reduced curves. \B{lemma}\label{nilpo} Let $I$ be a nilpotent ideal in an algebra $A$ over a field $k$. \B{description} \item[{\it (a)}] If char$(k)=0$, $K_n(A,I)$ is a uniquely divisible group for every $n$. \item[{\it (b)}] If char$(k)=p>0$, $K_2(A,I)$ is a $p$-group of bounded exponent. \E{description} \E{lemma} \B{proof} Part {\it (a)\,} is proven in \cite[1.4]{W966}. If char$(k)=p$, chose $m$ such that $I^{p^m}=0$; we will show that $p^mK_2(A,I)$. Indeed, $K_2(A,I)$ is generated by Steinberg symbols $\{a,1+x\}$ with $a\in A$ and $x\in I$, and $p^m\,\{a,1+x\}$ is $\{a,1+x^{p^m}\}=\{a,1\}=0$. \E{proof} \B{cor}\label{k2div} Let $A$ be a finite algebra over an algebraically closed field $k$. \B{description} \item[{\it (a)}] If char$(k)=0$ or if $A$ is a principal ideal ring, the group $K_2(A)$ is uniquely divisible. \item[{\it (b)}] If char$(k)=p$, $K_2(A)$ is the sum of the uniquely divisible group $K_2(A_{red})$ and a $p$-group of bounded exponent. \E{description} \E{cor} \B{proof} Let $I$ be the nilradical of $A$, so that $A_{red}=A/I$ is semisimple and hence $A\to A_{red}$ splits. Then $K_2(A)\cong K_2(A_{red})\oplus K_2(A,I)$, and $K_2(A_{red})$ is uniquely divisible by \cite[1.3]{BT}. Finally, if $A$ is a principal ideal ring then $A$ is a product of truncated polynomial rings $k[s]/(s^n)$ and a direct calculation (\cite[p.485]{Graham}) shows that $K_2(k[s]/(s^n))\cong K_2(k)$. \E{proof} \noindent{\it Step 3.} Finally, suppose that $Y$ is a curve which is not reduced. Let ${\cal I}$ denote the nilradical ideal sheaf of ${\cal O}_Y$, and write ${\cal K}_2{\cal I}$ for the sheafification of the presheaf $U\mapsto K_2({\cal O}_Y(U),{\cal I}(U))$. If ${\cal K}_{2,red}$ denotes the sheafification of $U\mapsto K_2(U_{red})$, there is an exact sequence of sheaves on $Y_{Zar}$ \B{equation}\label{Ired} {\cal K}_2{\cal I}\to {\cal K}_2\to {\cal K}_{2,red}\to 0 \E{equation} Let $U$ denote the smooth locus of $Y_{red}$. Since $U_{red}$ is smooth, the ring map ${\cal O}_U\to{\cal O}_{U_{red}}$ splits. Therefore ${\cal K}_2{\cal I}\mid_U$ injects into ${\cal K}_2\mid_U$, {\it i.e.,\/}\ the kernel of ${\cal K}_2{\cal I}\to{\cal K}_2$ is a skyscraper sheaf supported on $Y-U$. It follows that we have an exact sequence $$H^0(Y,{\cal K}_{2,red})\to H^1(Y,{\cal K}_2{\cal I})\to H^1(Y,{\cal K}_2)\to H^1(Y_{red},{\cal K}_2)\to 0$$ which we may rewrite as follows \B{equation}\label{reduced} K_2(Y_{red})\to H^1(Y,{\cal K}_2{\cal I})\to SK_1(Y)\to SK_1(Y_{red})\to 0 \E{equation} By Step 2, $SK_1(Y_{red})$ is uniquely divisible. If char$(k)=p$, we know by Lemma~\ref{nilpo}(b) that $H^1(Y,{\cal K}_2{\cal I})$ is a $p$-group of bounded exponent, and this part {\it (ii)\,} of Theorem~\ref{sing} because a uniquely divisible group has no nontrivial extensions by a $p$-group. Finally, suppose that char$(k)=0$. By Lemma~\ref{nilpo}(a), $H^1(Y,{\cal K}_2{\cal I})$ is uniquely divisible. By Proposition~\ref{n-divisible} below, $K_2(Y_{red})$ is divisible. In this case part {\it (i)\,} of Theorem~\ref{sing} follows from Step 2 and the exact sequence (\ref{reduced}). \E{proof} \B{prop}\label{n-divisible} If $\frac{1}{n}\in k$ and $Y$ is a curve then $K_2(Y)$ is $n$-divisible. \E{prop} \B{proof} We consider the $K$-theory of $Y$ with coefficients $\Z/n$, which is related to the usual Quillen $K$-theory of $Y$ by exact sequences such as $$0\to K_2(Y)\otimes\Z/n\to K_2(Y;\Z/n)\to K_1(Y)_{n-tors}\to 0$$ We know by \cite[1.4]{W966} that $K_2(Y;\Z/n)\cong K_2(Y_{red};\Z/n)$, and hence that $K_2(Y)\otimes\Z/n$ is a subgroup of $K_2(Y_{red})\otimes\Z/n$. Thus we may assume that $Y$ is reduced. Let $\tilde Y$ be the normalization of $Y$. The conductor ideal defines a zero-dimensional subscheme Spec($C$) of $Y$, and also its preimage Spec($D$) in $\tilde Y$. Because excision holds (see \cite[1.2]{W966}) we have an exact sequence $$K_3(D;\Z/n)\to K_2(Y;\Z/n)\to K_2(\tilde Y;\Z/n)\oplus K_2(C;\Z/n)$$ Now $K_3(D;\Z/n)\cong K_3(D_{red};\Z/n)$, again by \cite[1.4]{W966}. Since $D_{red}$ is a finite product of copies of $k$, and $K_3(k;\Z/n)=0$ by Suslin \cite{Suslin}, we have $K_3(D;\Z/n)=0$. Hence $K_2(Y;\Z/n)$ injects into $K_2(\tilde Y;\Z/n)\oplus K_2(C;\Z/n)$. By naturality, the subgroup $K_2(Y)\otimes\Z/n$ of $K_2(Y;\Z/n)$ injects into the corresponding subgroup $K_2(\tilde Y)\otimes\Z/n \oplus K_2(C)\otimes\Z/n$ of $K_2(\tilde Y;\Z/n)\oplus K_2(C;\Z/n)$, but: $K_2(\tilde Y)$ is divisible by Theorem~\ref{smooth} and $K_2(C)$ is divisible by Corollary~\ref{k2div}, so this latter subgroup is zero, hence $K_2(Y)\otimes\Z/n =0$ as claimed. \E{proof} \section{$K$-theory results} In this section we collect some results on the relation between the Zariski sheaves ${\cal K}_2$ and ${\cal H}^q(\mu_n^{\otimes2})$, namely \ref{HoobH2}, \ref{nK2} and \ref{square}, which will be used in the proof of the main theorem. In this section, our field $k$ will always contain $\frac1n$. The first result, which we cite without proof, concerns the sheafification of the \'etale Chern class $c_2^{et}\colon K_2(X)/n\to\Het2(X,\mu_n^{\otimes2})$. It is an extension by Hoobler of a well known result for smooth schemes. \B{prop}\label{HoobH2} (Hoobler \cite{Hoob}; {\it cf.\/}\ \cite[Thm. 0.2]{PW3}) Let $X$ be a scheme of finite type over a field containing $\frac1n$. Then the \'etale Chern class $c_2^{et}$ induces an isomorphism of Zariski sheaves on $X$: \quad ${\cal K}_2/n \by{\simeq} {\cal H}^2(\mu_n^{\otimes2})$. \E{prop} Our other results concern the n-torsion subsheaf ${}_n{\cal K}_2$ of ${\cal K}_2$. We begin with the local version. \B{lemma}\label{HoobH1} Let $A$ be a semilocal ring essentially of finite type over a field $k$. Assume $k$ contains a primitive $n^{th}$ root of unity $\zeta$. Define a map $$\varphi: \Het{1}(A,\mu_n^{\otimes2})\cong A^*/A^{*n}\to{}_nK_2(A)$$ by $\varphi(a)=\{a,\zeta\}$, $a\in A^*$. Then $\varphi$ is surjective. If $A$ is regular and $k$ contains an algebraically closed field then $\varphi$ is an isomorphism. \E{lemma} \B{proof} $\varphi$ is well defined because $\{ a^n,\zeta\}=\{ a,1\}=0$. Suppose first that $A$ is regular. Then $\varphi$ is onto by the Merkurev-Suslin Theorem. If in addition its field of fractions $F$ contains an algebraically closed field, then $\Het{1}(F,\mu_n^{\otimes2})\cong{}_nK_2(F)$ by \cite[3.7]{S2}. Comparing the Bloch-Ogus resolution of $\Het{1}(A,\mu_n^{\otimes2})$ to the Gersten--Quillen resolution of ${}_nK_2(A)$, one gets that $\Het{1}(A,\mu_n^{\otimes2})\cong {}_nK_2(A)$. The promotion to any semilocal ring $A$ follows from the same arguments used by R.Hoobler in \cite{Hoob}. Since $A$ is a localization of a finitely generated $k$-algebra, there exists localization $B$ of a polynomial ring over $k$ and an ideal $I$ in $B$ such that $A=B/I$. Let $(B^h,I^h)$ be the henselization of the pair $(B,I)$. As $B^h$ is a direct limit of semilocal regular rings finite over $B$, the map $\Het{1}(B^h,\mu_n^{\otimes2})\by{\varphi}{}_n{\cal K}_2(B^h)$ is an isomorphism. By a result of O. Gabber \cite[Th.1]{Gabber} we have $K_3(B^h;\Z/n)\cong K_3(B^h/I^h;\Z/n)\cong K_3(A;\Z/n)$. By proper base change $$\Het{1}(B^h,\mu_n^{\otimes2})\cong \Het{1}(B^h/I^h,\mu_n^{\otimes2})\cong \Het{1}(A,\mu_n^{\otimes2}).$$ The universal exact sequence for $K$-theory with coefficients yields a commutative diagram $$\begin{array}{ccccccc}0\to & K_3(B^h)\otimes\Z/n&\to &K_3(B^h;\Z/n)&\to &{}_nK_2(B^h)&\to 0\\ &\downarrow & &\mbox{\large $\parallel$} & &\downarrow & \\ 0\to&K_3(A)\otimes\Z/n &\to& K_3(A;\Z/n)&\by{}&{}_nK_2(A)&\to 0 \end{array}$$ Thus the right-most vertical arrow is surjective. We then conclude from commutativity of the diagram: $$\begin{array}{ccc} \Het{1}(B^h,\mu_n^{\otimes2}) &\by{\simeq}&{}_nK_2(B^h)\\ \mbox{\large $\parallel$} & &\quad\downarrow\mbox{\rm onto}\\ \Het{1}(A,\mu_n^{\otimes2}) &\by{\varphi}&{}_nK_2(A)\ \end{array}$$ \kern-14pt \E{proof} \B{schol}\label{phi-bar} Let $A$ and $k$ be as in Lemma~\ref{HoobH1}. If $n$ is even, assume that $k$ contains a square root of $-1$. If $\beta$ is a Bott element in $K_2(k;\Z/n)$ mapping to $\zeta\in K_1(k)=k^*$, then multiplication by $\beta$ lifts the map $\varphi$ to a map $$\bar\varphi\colon\Het{1}(A,\mu_n^{\otimes2})\to K_3(A;\Z/n).$$ This map is a split injection, because the \'etale Chern class satisfies $c^{et}_2\bar\varphi=-1$. \E{schol} \B{proof} The assumption on $k$ implies that $\beta$ exists and has order $n$, so $\bar\varphi(a)=\{ a,\beta\}$ is well-defined and lifts $\varphi$. The product formula (see \cite[Theorem 3.2(ii)]{W-chern}) states that $c_2(\{a,\beta\})=-[a]\otimes\zeta$ in $A^*/A^{*n}\otimes\mu_n(k) \cong \Het{1}({\rm Spec}\,(A),\mu_n^{\otimes2})$ for every $a\in A^*$. Thus up to sign $c_2$ is a left inverse of $\bar\varphi$. \E{proof} \B{thm}\label{nK2} Let $X$ be a scheme of finite type over $\C$. Then the \'etale Chern class defines an isomorphism of Zariski sheaves: $$c^{et}_2\colon\; {}_n{\cal K}_2 \by{\simeq} {\cal H}^1(\mu_n^{\otimes2}).$$ \E{thm} \B{proof} We saw in (2.4.2) that $c^{et}_2\colon {\cal K}_3(\Z/n) \to {\cal H}^1(\mu_n^{\otimes2})$ vanishes on ${\cal K}_3$. Hence $c^{et}_2$ is well-defined on ${}_n{\cal K}_2$. To verify that it is an isomorphism, we check the stalks at a point $x\in X$. If $A={\cal O}_{X,x}$ we see from \ref{phi-bar} that the surjection $\varphi\colon \Het1(A,\mu_n^{\otimes2})\to{}_nK_2(A)$ of Lemma~\ref{HoobH1} satisfies $c^{et}_2\ \varphi=-1$. Elementary algebra now implies that $c^{et}_2$ is an isomorphism on ${}_nK_2(A)$ and hence on ${}_n{\cal K}_2$. \E{proof} \B{cor}\label{square} By (\ref{et-HD-sheaf}), the following diagram commutes: $$ \B{array}{ccc} {}_n{\cal K}_2 &\longby\tau& \kern-4pt{\cal K}_2 \\ c^{et}_2\downarrow\cong & &\downarrow{c_2}\\ {\cal H}^1(\mu_n^{\otimes2}) &\longby{\delta}& {\cal H}_{\cal D}^2(2) \E{array}$$ where $\tau$ is the obvious inclusion and $\delta$ is defined in (\ref{triangle}) and (\ref{Rq}). \E{cor} \B{rmk} This gives the following explicit formula for $\delta$. Given a unit $a\in A^*$, where $U={\rm Spec}\,(A)$, write $[a]$ for the class of $c_1(a)$ in $H_{an}^1(U_{\mbox{\Large $\cdot $}},\Z(1)_{\cal D})\cong H_{an}^0(U_{\mbox{\Large $\cdot $}},{\cal O}_{U_{\mbox{\Large $\cdot $}}}^*)$. Then the product formula for $c_2$ shows that $\delta$ sends $a\otimes\zeta\in{\cal O}_X^*(U)\otimes\mu_n\cong{\cal H}^1(\mu_n^{\otimes2})(U)$ to $[\zeta]\cup[a]\in H_{an}^2(U_{\mbox{\Large $\cdot $}},\Z(2)_{\cal D})={\cal H}_{\cal D}^2(U)$. \E{rmk} \section{An exact sequence for ${\cal K}_2$-cohomology} We now give some exact sequences relating $H^1(X,{\cal K}_2)$ and $H^2(X,{\cal K}_2)$. The first is a reinterpretation of \cite[Theorem D]{PW3} in terms of hypercohomology. Let ${\cal K}_2^\bullet$ denote the complex ${\cal K}_2\by{n}{\cal K}_2$ concentrated in degrees 0 and 1. The short exact sequence $0\to{\cal K}_2[-1]\to{\cal K}_2^\bullet\to{\cal K}_2\to0$ gives rise to a long exact sequence, reminiscient of \cite[(4.4)]{S2}: \B{equation}\B{array}{c} 0\to H^0(X,{\cal K}_2)/n \by\iota \H^1(X,{\cal K}_2^\bullet) \by{} H^1(X,{\cal K}_2)\by{n}H^1(X,{\cal K}_2)_{\mathstrut} \by\iota \\ \qquad\H^2(X,{\cal K}_2^\bullet)\to H^2(X,{\cal K}_2) \by{n} H^2(X,{\cal K}_2) \to\cdots^{\mathstrut}\E{array} \E{equation} From this we extract short exact ``Kummer'' sequences, such as \B{equation}\label{Kummer} 0\to H^1(X,{\cal K}_2)/n \by\iota \H^2(X,{\cal K}_2^\bullet) \by{\pi} {}_nH^2(X,{\cal K}_2) \to0. \E{equation} We also have the exact sequence of low degree terms in the hypercohomology spectral sequence for ${\cal K}_2^\bullet$, the relevant part of which is: \B{equation}\label{lowdeg} H^0(X,{\cal K}_2/n)\vlongby{d_2} H^2(X,{}_n{\cal K}_2) \to \H^2(X,{\cal K}_2^\bullet)\by\eta H^1(X,{\cal K}_2/n) \to H^3(X,{}_n{\cal K}_2). \E{equation} \B{prop}\label{hyperK2}(\cite[Theorem D]{PW3}) Let $X$ be a quasi-projective over a field $k$ containing $\frac1n$. Then: \B{description} \item[{\it (a)}] The $d_2$-differential $H^0(X,{\cal K}_2/n) \longby{d_2} H^2(X,{}_n{\cal K}_2)$ in the hypercohomology spectral sequence (\ref{lowdeg}) is the composite $$H^0(X,{\cal K}_2/n) \by{\partial} H^1(X,n\cdot{\cal K}_2) \by{\partial} H^2(X,{}_n{\cal K}_2)$$ of the boundary maps in the usual interlocking sequences for ${\cal K}_2$. \item[{\it (b)}] If $X$ is a surface with isolated singularities, the map $\pi$ in the Kummer sequence (\ref{Kummer}) for $\H^2(X,{\cal K}_2^\bullet)$ factors through the surjection $\eta$ in the hypercohomology spectral sequence (\ref{lowdeg}). \E{description} \E{prop} \B{proof} Part (a) is a special case of a more general result which we have isolated in Lemma~\ref{hyper} below. For part (b), it suffices to show that the following diagram commutes. $$\begin{array}{ccccccccc} H^0(X,{\cal K}_2/n)\kern-5pt& \by{\gamma}\kern-5pt& H^2(X,{}_n{\cal K}_2)\kern-8pt &\by{\beta}\kern-7pt& H^1(X,{\cal K}_2)/n\kern-7pt&\to\kern-7pt&H^1(X,{\cal K}_2/n)\kern-5pt & \to\kern-5pt& {}_nH^2(X,{\cal K}_2) \kern-2pt\to0\\ \mbox{\large $\parallel$} & &\mbox{\large $\parallel$} &&\iota\downarrow &&\mbox{\large $\parallel$} &&\\ H^0(X,{\cal K}_2/n)\kern-5pt& \by{d_2}\kern-5pt& H^2(X,{}_n{\cal K}_2)\kern-8pt &\to\kern-7pt& \H^2(X,{\cal K}_2^\bullet \kern-7pt&\to\kern-7pt& H^1(X,{\cal K}_2/n) \kern-5pt&\to\kern-5pt&\ 0\hfill\\ \end{array} $$ The top row is the exact sequence of \cite[Theorem D]{PW3}, the bottom row is the exact sequence of low degree terms (\ref{lowdeg}) and the vertical arrow $\iota$ comes from the Kummer sequence (\ref{Kummer}). Since ${\cal K}_2\to{\cal K}_2/n$ factors through ${\cal K}^\bullet[1]$, the right square commutes. The left square commutes by part (a). The map $\beta$ is constructed as follows. Let ${\cal L}^\bullet$ denote the subcomplex ${\cal K}_2\to n\cdot{\cal K}_2$ of ${\cal K}_2^\bullet$; ${\cal L}^\bullet$ is quasi-isomorphic to ${}_n{\cal K}_2$. The inclusion of $n\cdot{\cal K}_2[-1]$ into ${\cal L}^\bullet$ induces a natural map $H^1(X,n\cdot{\cal K}_2)\to H^2(X,{}_n{\cal K}_2)$, and we know that this map is onto by \cite[4.8.1]{PW3}. We showed in \cite[Proposition 4.9]{PW3} that $H^1(X,n\cdot{\cal K}_2)\to H^1(X,{\cal K}_2)$ factors through this surjection, and the induced map is $\beta$. Thus $\iota\beta$ is induced from the composite map $n\cdot{\cal K}_2\to{\cal K}_2\to{\cal K}_2^\bullet[1]$ upon taking $H^1$. But this composite map factors through the subcomplex ${\cal L}^\bullet[1]$ of ${\cal K}^\bullet[1]$, so it follows that the left square commutes. \E{proof} Here is the general result about hypercohomology spectral sequences used to prove part (a) above. It works for any topos $X$. \B{lemma}\label{hyper} For any sheaf ${\cal F}$, let ${\cal C}^\bullet$ denote the complex ${\cal F} \by{n} {\cal F}$ concentrated in degrees 0 and 1. Then up to the sign $(-1)^{p}$, the $d_2$-differentials $$H^p(X,{\cal F}/n)=H^p(X,H^1{\cal C}) \longrightarrow H^{p+2}(X,H^0{\cal C})=H^{p+2}(X,{}_n{\cal F})$$ in the hypercohomology spectral sequence of ${\cal C}^\bullet$ are the composites $$H^p(X,{\cal F}/n)\by{\partial}H^{p+1}(X,n\cdot{\cal F}) \by{\partial}H^{p+2}(X,{}_n{\cal F})$$ of the boundary maps $\partial$ associated respectively to the exact sequences $$0\to n\cdot{\cal F}\to{\cal F}\to{\cal F}/n\to0,\qquad 0\to{}_n{\cal F}\to{\cal F}\to n\cdot{\cal F}\to0$$ \E{lemma} \B{proof} Given injective resolutions ${}_n{\cal F}\to{\cal I}^\bullet$, $n\cdot{\cal F}\to{\cal J}^\bullet$ and ${\cal F}/n{\cal F}\to{\cal K}^\bullet$ we can form injective resolutions ${\cal F}\to{\cal E}^{0\bullet}={\cal I}^\bullet\oplus{\cal J}^\bullet$ and ${\cal F}\to{\cal E}^{1\bullet}={\cal J}^\bullet\oplus{\cal K}^\bullet$ using the Horseshoe Lemma. These form the two columns of a Cartan-Eilenberg resolution ${\cal E}^{\bullet\bullet}$ of the complex ${\cal F}\by{n}{\cal F}$; by the sign trick, the single horizontal differential in this complex is $(-1)^{p}$ times the projection/inclusion $I^p\oplus{\cal J}^p\to {\cal J}^p\to {\cal J}^p\oplus{\cal K}^p$. Given a class $[s]\in H^p(X,{\cal F}/n)$, represent it by $s\in H^0(X,K^p)$ with $\partial s=0$ in $H^0(X,K^{p+1})$. Applying $\partial^v$ to $(0,s)\in H^0(X,J^p\oplus K^p)$ gives an element $(t,0)$ of $H^0(X,J^{p+1}\oplus K^{p+1})$. Thus $\partial\colon H^p(X,{\cal F}/n)\to H^{p+1}(X,n\cdot{\cal F})$ sends $[s]$ to $[t]$. Applying $\partial^v$ to $(0,t)\in H^0(X,I^p\oplus J^p)$ gives $(u,0)$ for some $u\in H^0(X,I^{p+1})$. By construction, $u$ is a cycle in $I^\bullet$ and $\partial\colon H^{p+1}(X,n\cdot{\cal F})\to H^{p+2}(X,{}_n{\cal F})$ sends $[t]$ to $[u]$. Now the hypercohomology spectral sequence arises from the row filtration on the Cartan-Eilenberg resolution ${\cal E}^{\bullet\bullet}$. Since the pair $((0,(-1)^{p}t),(0,s))\in\mbox{Tot}^p({\cal I})$ has $(((-1)^pu,0),(0,0))$ for its boundary, the $d_2$-differential in the spectral sequence takes $[s]$ to $(-1)^{p}[u]$. \E{proof} We are now going to connect Proposition~\ref{hyperK2} with \'etale cohomology using $c^{et}_2$. For this we need to resort to some standard topological constructions. Our main result will be Theorem~\ref{NH3} below. Recall from (2.4.1) that there is a simplicial presheaf $K$ on $X_{zar}$ such that $\pi_qK(U)=K_{q}(U)$. Let $\tilde K(U)$ be the universal covering space of the basepoint component of $K(U)$; $\tilde K$ is a simplicial presheaf by \cite[8.3 or 16.4]{May}. Let $\tilde K^{(2)}(U)$ denote the second layer of the Postnikov tower of $\tilde K(U)$, defined in \cite[8.1]{May}; it is an Eilenberg-MacLane complex of type $(K_2U,2)$ and $\tilde K^{(2)}$ is a simplicial presheaf. Moreover by \cite[8.2]{May} there are Kan fibrations $\tilde K^{(2)} \leftarrow \tilde K \to K$. Now let $\tilde L(U)$ denote the homotopy fiber of the map $\tilde K(U) \by{n} \tilde K(U)$, and let $M(U)$ denote the homotopy fiber of the map $\tilde K^{(2)}(U) \by{n} \tilde K^{(2)}(U)$. Each $\tilde L(U)$ is a connected space with $\pi_1\tilde L(U)=K_2(U)/n$ and $\pi_q\tilde L(U)=K_{q+1}(U;\Z/n)$ for $q\ge2$, while $M(U)$ has only two nontrivial homotopy groups: $\pi_1M(U)=K_2(U)/n$ and $\pi_2M(U)={}_nK_2(U)$. In fact, it is not hard to see that $M(U)$ is homotopy equivalent to the simplicial space obtained by applying the Dold-Kan theorem to the presheaf of chain complexes $K_2 \by{n} K_2$ concentrated in degrees 2 and 1. We can perform the above constructions so that there is a commutative diagram of simplicial presheaves (in which the diagram (\ref{fib}) forms the right side): \B{equation}\label{grid}\B{array}{ccccccc} M &\leftarrow& \tilde L &\to& L &\vlongby{C^{ss}_2}& \Omega{\cal E} \\ \downarrow && \downarrow && \downarrow &&\delta\downarrow\enskip\\ \tilde K^{(2)}&\leftarrow& \tilde K &\to& K &\longby{C_2^{ss}}&{\cal D}\\ n\downarrow\enskip&&n\downarrow\enskip&&n\downarrow\enskip&& n\downarrow\enskip\\ \tilde K^{(2)}&\leftarrow& \tilde K &\to& K &\vlongby{C_2^{ss}}&{\cal D} \E{array}\E{equation} \B{num} Given any simplicial presheaf $F$ on $X$, the {\it generalized sheaf cohomology groups} $\H^q(X,F)$ were defined for $q\le0$ by Brown and Gersten \cite[p.280]{BG}. (The homotopy categories of simplicial presheaves and simplicial sheaves are equivalent by \cite[2.8]{J}. In particular, if $\tilde F$ is the simplicial sheaf associated to $F$ then $\H^q(X,F)=\H^q(X,\tilde F)$.) If $F$ is the simplicial sheaf associated by the Dold-Kan theorem to a cochain complex ${\cal F}$ (concentrated in negative degrees), then $\H^q(X,F)\cong \H_{zar}^{q}(X,{\cal F})$ for $q\le0$ by \cite[p.281]{BG}. Since the simplicial sheaf associated to ${\cal K}_2^\bullet[2]$ is the sheafification of $M$ we have $\H^q(X,M)=\H_{zar}^{q+2}(X,{\cal K}_2^\bullet)$. Similarly, by (2.4.2) we have $$\H^q(X,\Omega{\cal E})=\H^{q-1}(X,{\cal E})\cong \Het{3-q}(X,\mu_n^{\otimes2}).$$ In particular, diagram (\ref{grid}) induces maps \B{equation}\label{branch} \H^2(X,{\cal K}_2^\bullet)=\H^0(X,M) \stackrel\lambda\leftarrow \H^0(X,\tilde L)\longby{c_2^{et}} \H^0(X,\Omega{\cal E}) = \Het3(X,\mu_n^{\otimes2}). \E{equation} If $F$ is a simplicial presheaf on $X$, we write $\tilde\pi_qF$ for the sheaf associated to the presheaf $U\mapsto \pi_qF(U)$. For example, we have $$\tilde\pi_qM = \cases{{\cal K}_2/n& if $q=1$; \cr {}_n{\cal K}_2& if $q=2$; \cr 0 & else.\cr} \qquad \tilde\pi_q\tilde L=\cases{{\cal K}_2/n& if $q=1$; \cr {\cal K}_{q+1}(\Z/n)& if $q\ge2$; \cr 0 & else.\cr} $$ Now recall that $X$ is quasi-projective over $\C$. By \cite[Theorem 3]{BG} there is a ``Brown-Gersten'' spectral sequence in the fourth quadrant: $$E_2^{pq}=H_{zar}^p(X,\tilde\pi_{-q}F) \Rightarrow\H^{p+q}(X,F).$$ In general, this spectral sequence is ``fringed'' \cite[p.285]{BG}, but since all the $F$ we consider here are infinite loop spaces this fringing does not affect $\H^q(X,F)$ for $q\le0$. \E{num} \B{example} Here is an example of the fringing phenomenon. If $F$ is associated to a cochain complex ${\cal F}$, with ${\cal F}^q=0$ for $q>0$, then it is well known that the Brown-Gersten spectral sequence for $F$ is the same as the hypercohomology spectral sequence for ${\cal F}$. For example, the simplicial sheaf ${\cal E}$ was defined in (2.4.2) as being associated to $\tau^{\le0}\R\omega_*\Z/n[2i]$. The hypercohomology spectral sequence of this complex coincides with the Leray spectral sequence for $\Het{2i+*}(X,\mu_n^{\otimes i})$ in the region $q\le0$. Thus it is a fringed spectral sequence converging in the region $p+q\le0$. The line $p+q=+1$ converges to the kernel of $\Het{2i+1}(X,\mu_n^{\otimes2})\to H^0(X,{\cal H}^{2i+1}(\mu_n^{\otimes i}))$. On the other hand, the sheafification $\tilde M$ of $M$ is associated to the complex of sheaves ${\cal K}_2^\bullet[2]$. Hence the Brown-Gersten spectral sequence for $M$ is the same as the hypercohomology spectral sequence for ${\cal K}_2^\bullet[2]$, and there is no fringing effect. \E{example} \medskip Any morphism $E\to F$ of simplicial presheaves induces a morphism of Brown-Gersten spectral sequences. Thus (\ref{branch}) gives us a commutative diagram: \def\rod{\stackrel{c^{et}_2\quad\cong}{\kern-20pt \hbox to115pt{\rightarrowfill}\kern-35pt}} \B{equation}\label{maze}\B{array}{ccccc} H^0(X,{\cal K}_2/n) && \rod && H^0(X,{\cal H}^2(\mu_n^{\otimes2})) \\ \mbox{\large $\parallel$} && && \mbox{\large $\parallel$} \\ H^0(X,\tilde\pi_1M) &\stackrel\cong\leftarrow& H^0(X,\tilde\pi_1\tilde L) &\vlongby{c^{et}_2\enspace\cong}& H^0(X,\tilde\pi_2{\cal E}) \\ d_2\downarrow\quad && d_2\downarrow\quad && d_2\downarrow\quad \\ H^2(X,\tilde\pi_2M) &\leftarrow& H^2(X,\tilde\pi_2\tilde L) &\vlongby{c^{et}_2}& H^2(X,\tilde\pi_3{\cal E}) \\ \mbox{\large $\parallel$} && && \mbox{\large $\parallel$} \\ H^2(X,{}_n{\cal K}_2)&& \rod && H^2(X,{\cal H}^1(\mu_n^{\otimes2})) \\ \E{array}\E{equation} (The bottom square of (\ref{maze}) commutes because, as noted in (2.4.2), the Chern class map $c_2^{et}\colon{\cal K}_3(\Z/n)\to{\cal H}^1(\mu_n^{\otimes2})$ factors through ${}_n{\cal K}_2$.) The following description of the differential in the Leray spectral sequence was suggested in \cite[(0.4)]{PW3}. \B{prop}\label{diff} If we identify ${\cal K}_2/n$ with ${\cal H}^2(\mu_n^{\otimes2})$ by \ref{HoobH2} and ${}_n{\cal K}_2$ with ${\cal H}^1(\mu_n^{\otimes2})$ by \ref{nK2}, then the differential $d_2\colon H^0(X,{\cal H}^2(\mu_n^{\otimes2})\to H^2(X,{\cal H}^1(\mu_n^{\otimes2})$ in the Leray spectral sequence for $\Het*(X,\mu_n^{\otimes2})$ becomes identified with the differential in \ref{hyperK2}(a), {\it i.e.,\/}\ $$H^0(X,{\cal H}^2(\mu_n^{\otimes2}))\cong H^0(X,{\cal K}_2/n) \by{\partial} H^1(X,n\cdot{\cal K}_2)\by{\partial} H^2(X,{}_n{\cal K}_2) \cong H^2(X,{\cal H}^1(\mu_n^{\otimes2})). $$ \E{prop} \B{proof} The left vertical map in (\ref{maze}) is the differential in the hypercohomology spectral sequence for $\H^{*}(X,{\cal K}_2^\bullet)$ by Example 7.4, and was described in Proposition~\ref{hyperK2}(a). Again by Example~7.4, the right vertical map in (\ref{maze}) is the corresponding differential in the Leray spectral sequence for $\Het{4+*}(X,\mu_n^{\otimes2})$. A diagram chase on (\ref{maze}), starting at $H^0(X,\tilde\pi_1\tilde L)$, yields the result. \E{proof} \medskip \B{defi} Following Suslin \cite{S2}, we define $N\Het3(X)$ to be the kernel of the natural map $\Het{3}(X,\mu_n^{\otimes 2})\to H^0(X,{\cal H}^3(\mu_n^{\otimes2}))$. Here $X$ can be any scheme in which $n$ is invertible. Of course, when $X$ is a surface over an algebraically closed field the sheaf ${\cal H}^3(\mu_n^{\otimes2})$ vanishes and we have $N\Het3(X)=\Het3(X,\mu_n^{\otimes2})$. \E{defi} \smallskip The following result was proven by Suslin \cite[p. 19]{S2} for smooth varieties. It is a partial answer to \cite[Question~2]{BV1} and was conjectured in \cite[(0.4)]{PW3}. \B{thm} \label{NH3} Let $X$ be a surface with isolated singularities over a field $k$ containing an algebraically closed field and $\frac1n$. Then $$N\Het3(X)\cong\H^2(X,{\cal K}_2\kern-3pt\by{n}\kern-3pt{\cal K}_2).$$ In particular, by (\ref{Kummer}) there is a functorial short exact sequence: $$0\to H^1(X,{\cal K}_2)/n\to N\Het3(X)\to{}_nCH_0(X)\to 0.$$ \E{thm} \B{proof} Since $X$ is a surface, the Brown-Gersten spectral sequences associated to the simplicial presheaves in (\ref{branch}) have only three nonzero columns. Using the computations given in (7.3) for $\tilde\pi_qM$ and $\tilde\pi_q\tilde L$, the resulting exact sequences form the rows of a commutative diagram. \B{equation}\label{M-L-et}\B{array}{cccccccc} \kern-8pt H^0(X,{\cal K}_2/n)&\kern-3pt\by{d_2}& \kern-3pt H^2(X,{}_n{\cal K}_2) \kern-6pt&\to& \kern-6pt \H^2(X,{\cal K}_2^\bullet) \kern-2pt&\to&\kern-6pt H^1(X,{\cal K}_2/n) \kern-6pt&\to0\\ \kern-8pt\uparrow\cong&&\uparrow\mbox{\rm onto}&& \uparrow\lambda&&\uparrow\cong&\\ \kern-8pt H^0(X,{\cal K}_2/n)&\kern-3pt\by{d_2}& \kern-3pt H^2(X,{\cal K}_3(\Z/n)) \kern-6pt&\to& \kern-6pt \H^0(X,\tilde L) \kern-2pt&\to&\kern-6pt H^1(X,{\cal K}_2/n) \kern-6pt&\to0\\ \kern-8pt\downarrow\cong&& c_2^{et}\downarrow\mbox{\rm onto}\kern1.5em &&\downarrow\kern1em&&\downarrow\cong&\\ \kern-8pt H^0(X,{\cal H}^2(\mu_n^{\otimes2}))&\kern-3pt\by{d_2}& \kern-3pt H^2(X,{\cal H}^1(\mu_n^{\otimes2})) \kern-6pt&\to&\kern-6pt N\Het3(X) \kern-2pt&\to&\kern-6pt H^1(X,{\cal H}^2(\mu_n^{\otimes2})) \kern-6pt&\to0 \E{array}\E{equation} The outside vertical maps are isomorphisms by \ref{HoobH2}. The two vertical maps marked `onto' in (\ref{M-L-et}) are actually split surjections with the same kernel, and are identified by Lemma~\ref{HoobH1} since $\varphi\colon{\cal H}^1(\mu_n^{\otimes2})\to{}_n{\cal K}_2$ yields an isomorphism on $H^2$. Indeed, by \ref{phi-bar} we know that the map $c_2^{et}\colon{\cal K}_3(\Z/n)\to{\cal H}^1(\mu_n^{\otimes2})$ is a surjection, split up to sign by $\bar\varphi\colon {\cal H}^1(\mu_n^{\otimes2})\to{\cal K}_3(\Z/n)$. A diagram chase on (\ref{M-L-et}) shows that the two maps $\H^0(X,\tilde L)\longby{\lambda}\H^2(X,{\cal K}_2^\bullet)$ and $\H^0(X,\tilde L)\longby{c_2^{et}} N\Het3(X)$ are both onto with the same kernel. Thus the quotients $\H^2(X,{\cal K}_2^\bullet)$ and $N\Het3(X)$ are isomorphic. \E{proof} \B{cor}\label{H3} If $k$ is algebraically closed then the short exact sequence is: $$0\to H^1(X,{\cal K}_2)/n\to \Het{3}(X,\mu_n^{\otimes 2}) \to{}_nCH_0(X)\to 0.$$ \E{cor} \B{thm} \label{indivisible} Let $X$ be a normal projective surface over an algebraically closed field $k$. Let $\ell$ a prime number, $\ell\neq$char$(k)$. Then $$H^1(X,{\cal K}_2)\otimes \Q_{\ell}/\Z_{\ell} =0\quad\mbox{and} \quad \Het{3}(X,\Q_\ell/\Z_\ell) \cong CH_0(X)_{\ell-tors}$$ \E{thm} \B{proof} Choose a resolution of singularities $\pi\colon X'\to X$. Passing to the direct limit as $\nu\to\infty$, with $n=\ell^\nu$, the short exact sequences of Corollary~\ref{H3} become the rows of the commutative diagram $$\begin{array}{ccccccc} 0\to & H^1(X,{\cal K}_2)\otimes\Q_\ell/\Z_\ell &\to&\Het{3}(X,\Q_\ell/\Z_\ell)&\to &CH_0(X)_{\ell-tors}&\to 0\\ &\downarrow & &\downarrow & &\downarrow\cong & \\ 0\to & H^1(X',{\cal K}_2)\otimes\Q_\ell/\Z_\ell &\to&\Het{3}(X',\Q_\ell/\Z_\ell)&\to &CH_0(X')_{\ell-tors}&\to 0. \end{array}$$ The right-hand vertical map is an isomorphism by the Collino-Levine Theorem \cite{C2} \cite{L-Alb}. By \cite{CTR}, we have $H^1(X',{\cal K}_2)\otimes\Q_{\ell}/\Z_{\ell} =0$. Therefore it suffices to show that $$\Het3(X,\Q_{\ell}/\Z_{\ell})\cong \Het3(X',\Q_{\ell}/\Z_{\ell}).$$ There is a Mayer--Vietoris sequence for $\ell$-adic cohomology similar to (\ref{exnorm}) for the square (\ref{birsquare}). This yields an exact sequence $$0\to T\to\Het{3}(X,\Z_{\ell})\to \Het{3}(X',\Z_{\ell})\to0$$ with $T = \Het{2}(E,\Z_{\ell})/\im(\Het{2}(X',\Z_{\ell}))$. The proof of Proposition~\ref{pure} goes through in the $\ell$-adic setting as well to show that $T$ is a torsion group ({\it cf.\/}\ \cite[2.1]{C2}). Since we also have $\Het{4}(X,\Z_\ell)\cong \Het{4}(X',\Z)\cong\Z_\ell^c$, the universal coefficient theorem yields the result: $$\Het{3}(X,\Q_{\ell}/\Z_{\ell})\cong \Het{3}(X,\Z_\ell)\otimes\Q_{\ell}/\Z_{\ell}\cong \Het{3}(X',\Z_\ell)\otimes\Q_{\ell}/\Z_{\ell}\cong \Het{3}(X',\Q_{\ell}/\Z_{\ell}).$$ \kern-24pt \E{proof} \section{Proof of the Main Theorem} Let $X$ be a complex projective surface. In Lemma~\ref{degree} we constructed the Abel--Jacobi map $\rho\colon A_0(X)\to J^2(X)$. Our Main Theorem, stated in the Introduction, states that $\rho$ induces an isomorphism $A_0(X)_{tors}\cong J^2(X)_{tors}$. We now proceed to prove the Main Theorem. If $X$ is a normal surface then the result $A_0(X)_{tors}\cong J^2(X)_{tors}$ is a paraphrase of the theorem of Levine and Collino (see \cite{C2}, \cite{L-Alb}) that $A_0(X)_{tors}\cong J^2(\tilde X)_{tors}$ for any resolution of singularities $\tilde X\to X$, because $J^2(X)\cong J^2(\tilde X)$ by Corollary~\ref{abnorm}. Granting the normal case, we shall establish the general case of our Main Theorem by comparing a singular surface $X$ with its normalization $\tilde X$. For this, we need the following crucial Lemma. Let ${\cal H}_{an}^2(\Z)$ denote the Zariski sheaf on $X$ associated to the presheaf $U\mapsto H_{an}^2(U,\Z)$ \B{lemma}\label{BVS} Let $X$ be an irreducible proper surface over $\C$. Then the following composition is zero. $$H_{{\cal D}}^2(X,\Z(2))\by\varepsilon H_{an}^2(X,\Z)\to H^0(X,{\cal H}_{an}^2(\Z))$$ \E{lemma} \B{proof} By Lemma~\ref{filt} the image of $\varepsilon$ is the torsion subgroup of $H_{an}^2(X,\Z)$. However, the sheaf ${\cal H}_{an}^2(\Z)$ and hence its global sections are torsion free by \cite[Cor. 3]{BVS1}. \E{proof} \B{prop}\label{crux} If $X$ is an irreducible proper surface over $\C$, the following natural map is zero. $$H^0(X,{\cal K}_2)\to H^0(X,{\cal K}_2/n)$$ \E{prop} \B{proof} By Proposition~\ref{sheafHD} the natural map $H_{\cal D}^2(X,\Z(2))\to H^0_{zar}(X,{\cal H}_{\cal D}^2(2))$ is an isomorphism. The Proposition follows from Lemma~\ref{BVS} and a chase on the following diagram, the left part of which commutes by (2.4.1). $$\B{array}{ccccccc} K_2(X) &\by{c_2}& H_{\cal D}^2(X,\Z(2))&\by\varepsilon& H_{an}^2(X,\Z) &\to& H_{an}^2(X,\Z/n) \\ \downarrow&&\downarrow\cong &&\downarrow&&\downarrow\\ H^0(X,{\cal K}_2) &\by{c_2}& H^0(X,{\cal H}_{\cal D}^2(2))&\to& H^0(X,{\cal H}_{an}^2(\Z)) &\to& H^0(X,{\cal H}^2(Z/n)). \E{array}$$ \vskip-17pt \E{proof} \B{rmk} When $X$ is a {\it smooth} proper variety over an algebraically closed field of characteristic zero, Proposition~\ref{crux} was proven by Colliot-Th\'el\`ene and Raskind \cite{CTR}, and also by H. Esnault \cite{Esn} over $\C$. \E{rmk} \B{prop}\label{SK1} Let $Z$ be a scheme which is proper over $\C$. If $Z$ is either a curve or a normal surface then \B{description} \item[{\it i)}] $c_2\colon H^1(Z,{\cal K}_2)_{tors}\cong H^2(Z,\Q/\Z)$ \item[{\it ii)}] $H^1(Z,{\cal K}_2)\otimes \Q/\Z = 0$ \E{description} \E{prop} \goodbreak \B{proof} The hypothesis on $Z$ allows us to use \ref{HD-tors} for the isomorphism $H^2(Z,\Q/\Z)\cong H_{\cal D}^3(X,\Z(2))_{tors}$. When $Z$ is a curve both assertions follow from Theorem~\ref{sing} and this remark. When $Z$ is a normal surface, part {\it ii)\,} was proven in Theorem~\ref{indivisible}. In order to prove part {\it i)\,} for a normal surface $Z$, we apply $H^1$ to Corollary~\ref{square} and combine with the diagram of Corollary~\ref{morph} to get a commutative diagram for each $n$: \B{equation}\label{square2}\B{array}{ccc} H^1(Z,{}_n{\cal K}_2)&\by{\strut\tau_n} &H^1(Z,{\cal K}_2)_{n-tors}\\ c_2^{et}\downarrow\cong & &\downarrow{c_2}\quad\\ H^1(Z,{\cal H}^1(\mu_n^{\otimes2}))&\by\delta & H^1(Z,{\cal H}_{\cal D}^2(2))_{n-tors}\\ \downarrow && \downarrow\qquad \\ \Het2(Z,\mu_n^{\otimes2}) &\by\delta & H_{\cal D}^3(Z,\Z(2))_{n-tors}. \E{array}\E{equation} Taking the direct limit as $n\to\infty$ turns $\mu_n^{\otimes2}$ into $\Q/\Z$. Since $H^2(Z,\Q/\Z)$ is the torsion subgroup of $H_{\cal D}^3(Z,\Z(2))$ by Proposition~\ref{HD-tors}, we have a commutative diagram $$\B{array}{cccccccc} &\kern-5pt H^1(Z,{\cal K}_{2,tors})\kern-5pt &\by{\tau}&\kern-3pt H^1(Z,{\cal K}_2)_{tors}& \kern-5pt\to & \kern-6pt\coker(\tau)&\to0&\\ &c_2^{et}\downarrow\;\cong&&\downarrow c_2&&\kern-5pt\downarrow&&\\ 0\to\kern-3pt&\kern-5pt H^1(Z,{\cal H}^1(\Q/\Z))\kern-5pt&\to&\kern-3pt \Het2(Z,\Q/\Z)&\kern-5pt\to& \kern-6pt H^0(Z,{\cal H}^2(\Q/\Z)) &\kern-7pt\longby{d_2}&\kern-8pt H^2(Z,{\cal H}^1(\Q/\Z))\\ \E{array}$$ in which the bottom row is exact by Corollary~\ref{morph}. Therefore in order to prove $(i)$ we are reduced to the claim that $$\coker\ta \cong \ker H^0(Z,{\cal H}^2(\Q/\Z))\by{d_2} H^2(Z,{\cal H}^1(\Q/\Z))$$ For each $n$, let $\gamma_n$ denote the composition $H^0(Z,{\cal K}_2/n)\kern-1.5pt\by\partial\kern-2pt H^1(Z,n\cdot{\cal K}_2) \kern-1.5pt\by\partial\kern-2pt H^2(Z,{}_n{\cal K}_2)$ in the usual interlocking long exact sequences \B{equation}\label{interlock}\B{array}{ccccccc} H^1(Z,{}_n{\cal K}_2)&\by{\mathstrut\tau_n}&H^1(Z,{\cal K}_2) &\to&H^1(Z,n\cdot{\cal K}_2)&\by\partial&H^2(Z,{}_n{\cal K}_2)\\ &&&&\mbox{\large $\parallel$}&&\\ H^0(Z,{\cal K}_2)&\by0& H^0(Z,{\cal K}_2/n) &\stackrel\partial\hookrightarrow& H^1(Z,n\cdot{\cal K}_2) &\to &H^1(Z,{\cal K}_2). \E{array}\E{equation} The arrow marked `0' in this diagram is the zero map by Proposition~\ref{crux}. The other zig-zag composition in (\ref{interlock}), from $H^1(Z,{\cal K}_2)$ to $H^1(Z,{\cal K}_2)$, is multiplication by $n$. It follows from (\ref{interlock}) that $$\ker(\gamma_n) \cong H^0(Z,{\cal K}_2/n)\cap \im(H^1(Z,{\cal K}_2)) \cong \frac{H^1(Z,{\cal K}_2)_{n-tors}}{H^1(Z,{}_n{\cal K}_2)} = \coker(\tau_n). $$ By Proposition \ref{hyperK2}(a), $\gamma_n$ is the differential $d_2$ in the hypercohomology spectral sequence for ${\cal K}_2\by{n}{\cal K}_2$. By Proposition~\ref{diff}, we may also identify $\gamma_n$ with the $d_2$-differential in the Leray spectral sequence for $\Het{*}(Z,\mu_n^{\otimes2})$. Passing to the direct limit we obtain the claimed formula: $\coker\tau = \lim\limits_{n\to\infty}\coker\tau_n\cong \lim\limits_{n\to\infty} \ker\gamma_n=\ker(d_2)$. \E{proof} We are now ready to prove our Main Theorem for an arbitrary projective surface $X$. Letting $\tilde X$ be its normalization and $Y$ a subscheme chosen as in Theorem \ref{M-V-K}, we have a Mayer-Vietoris Sequence in $K$-theory, and also for Deligne cohomology by \ref{M-V-D}. Taking the torsion subgroups of the diagram in Corollary~\ref{SKtoD} yields the following commutative diagram (in which we have abbreviated the left-hand terms for legibility). \B{equation}\label{diagram}\begin{array}{cccccccc} \kern-5pt\biggl\{{SK_1(\tilde X)\oplus\atop SK_1(Y)}\biggr\}_{tors} \kern-7pt&\to\kern-6pt & SK_1(\tilde Y)_{tors} \kern-5pt&\to\kern-5pt& SK_0(X)_{tors} \kern-5pt&\to\kern-7pt& SK_0(\tilde X)_{tors} & \kern-5pt\to0\\ \kern-5pt\downarrow\mbox{}\kern-5pt & &\downarrow\cong & &\downarrow\cong&&\downarrow\cong&\\ \kern-5pt\biggl\{{H^1(\tilde X)\oplus\atop H^1(Y,{\cal K}_2)\ } \biggr\}_{tors} \kern-7pt&\to\kern-6pt & H^{1}(\tilde Y\kern-2pt,{\cal K}_2)_{tors} \kern-5pt&\to\kern-5pt& H^{2}(X,{\cal K}_2)_{tors} \kern-5pt&\to\kern-7pt& H^{2}(\tilde X,{\cal K}_2)_{tors} & \kern-5pt\to0\\ \kern-5ptc_2\downarrow\cong\ &&c_2\downarrow\cong &&c_2\downarrow\ \ &&c_2\downarrow\cong \ &\\ \kern-5pt\biggl\{{H_{\cal D}^3(\tilde X)\atop H_{\cal D}^3(Y)}\biggr\}_{tors} \kern-7pt&\to\kern-6pt & H_{\cal D}^3(\tilde Y\kern-2pt,\Z(2))_{tors} \kern-5pt&\to\kern-5pt& H_{\cal D}^4(X,\Z(2))_{tors}\kern-5pt&\to\kern-7pt& H_{\cal D}^4(\tilde X,\Z(2))_{tors}& \kern-5pt\to0 \end{array} \E{equation} Some discussion of diagram (\ref{diagram}) is in order. The 3 isomorphisms between the terms in the top two rows come from \ref{SKtoD}. The 2 vertical maps in the lower left of (\ref{diagram}) are isomorphisms by Proposition \ref{SK1}. The lower right vertical map $H^{2}(\tilde X,{\cal K}_2)_{tors}\cong J^2(\tilde X)_{tors} \cong H_{\cal D}^4(\tilde X,\Z(2))_{tors}$ is an isomorphism because $\tilde X$ is normal. The bottom row of (\ref{diagram}) is exact, because by Proposition \ref{HD-tors} it is isomorphic to $$H^2(\tilde X,\Q/\Z)\oplus H^2(Y,\Q/\Z) \to H^2(\tilde Y,\Q/\Z) \to H^3(X,\Q/\Z) \to H^3(\tilde X,\Q/\Z) \to0. $$ The top two rows of (\ref{diagram}) are exact except at $SK_1(\tilde Y)_{tors}$ and $H^1(\tilde Y,{\cal K}_2)_{tors}$ by Proposition~\ref{SK1} and the elementary lemma~\ref{tors-exact} below, whose proof is left as an exercise. The 5-lemma implies that we have an isomorphism $$c_2\colon H^{2}(X,{\cal K}_2)_{tors}\cong H_{\cal D}^4(X,\Z(2))_{tors}$$ and this finishes the proof of our Main Theorem. \hfil$\bullet$ \B{rmk}In order for the diagram chase of (\ref{diagram}) to work, it suffices to know the crude surjectivity of the left vertical map as $n\to\infty$: $$H^1(\tilde X,{\cal K}_2)_{tors}\oplus H^{1}(Y,{\cal K}_2)_{tors} \longby{c_2} H^2(\tilde X,\Q/\Z)\oplus H^2(Y,\Q/\Z).$$ \E{rmk} \B{lemma}\label{tors-exact} Let $A\to B\to C\to D$ be an exact sequence of abelian groups. If $A\otimes\Q/\Z=0$ then the following sequence is exact. $$B_{tors} \to C_{tors} \to D_{tors}$$ \E{lemma} Here is a motivic version of our Main Theorem. For a 1-motive $M=(L,A,T,J,u)$ we let $M_{tors}$ denote the extension of torsion subgroups. $$0\to T_{tors}\to J_{tors}\to A_{tors}\to 0$$ Then our Main Theorem says that $Alb(X)_{tors}$ can be described via {\it algebraic} zero-cycles, {\it i.e.,\/}\ that $J^2(X)_{tors}$ is isomorphic to $A_0(X)_{tors}$ in a way compatible with normalization and desingularization. \B{schol} Let $Alb(X)$ the Albanese 1--motive of a projective surface. We then have the following identification of $Alb(X)_{tors}$: $$\begin{array}{ccccccc} 0\to&(\Q/\Z)^s &\to&A_0(X)_{tors} & \to & A_0(\tilde X)_{tors}&\to 0\\ &\mbox{\large $\parallel$} &&\downarrow\cong&&\downarrow\cong& \\ 0\to&(\Q/\Z)^s &\to & J^2(X)_{tors}& \to & J^2(\tilde X)_{tors} &\to 0. \end{array}$$ \E{schol} \B{rmk} If $X$ is an {\it affine} surface over $\C$ then $CH_0(X)=A_0(X)$ is uniquely divisible. Indeed, the fact that $A_0(X)_{tors} =0$ was proven in \cite[Theorem 2.6]{L2}. And divisibility of $CH_0(X)=SK_0(X)$ is classical, probably attributable to Murthy: Every smooth point $x$ on $X$ is in the image of a map $j\colon C\to X$ in which $C$ is a smooth affine curve. The group $\Pic(C)$ is divisible, and the class of $x$ is in the image of the map $j_*\colon \Pic(C)\to SK_0(X)$. Since $H^3(X,\C)=0$ as well, we also have $J^2(X)=0$. Thus Roitman's Theorem holds by default in the affine case. \E{rmk}
1995-03-22T06:20:14
9503
alg-geom/9503011
en
https://arxiv.org/abs/alg-geom/9503011
[ "alg-geom", "math.AG" ]
alg-geom/9503011
Christoph Lossen
Gert-Martin Greuel, Christoph Lossen
Equianalytic and equisingular families of curves on surfaces
LaTeX v 2.09
null
null
null
null
We consider flat families of reduced curves on a smooth surface S such that each member C has the same number of singularities of fixed singularity types and the corresponding (locally closed) subscheme H of the Hilbert scheme of S. We are mainly concerned with analytic resp. topological singularity types and give a sufficient condition for the smoothness of H (at C). Our results for S=P^2 seem to be quite sharp for families of cuves of small degree d.
[ { "version": "v1", "created": "Tue, 21 Mar 1995 14:42:27 GMT" } ]
2008-02-03T00:00:00
[ [ "Greuel", "Gert-Martin", "" ], [ "Lossen", "Christoph", "" ] ]
alg-geom
\section*{Introduction}\addcontentsline{toc}{section}{Introduction} We consider flat families of reduced curves on a smooth surface $S$ such that for each member $C$ of the family the number of singular points of $C$ and for each singular point $x \in C$ the ``singularity type'' of $(C,x)$ is fixed. Fixing these data imposes conditions on the space of all curves and we obtain in this way a locally closed subscheme of the Hilbert scheme $H_S$ of $S$. We are mainly concerned with the study of the equianalytic $(H^{ea}_S)$ respectively the equisingular Hilbert scheme $(H^{es}_S)$, which are defined by fixing the analytic respectively the (embedded) topological type of the singularities. We show that fixing the analytical (respectively topological) type of $(C,x)$ imposes, at most, $\tau (C,x) =$ Tjurina number of $(C,x)$ (respectively, at most, $ \mu (C,x)-$ mod$(C,x)$, where $\mu$ denote the Milnor number, mod the modality in the sense of \cite{AGV}) conditions with equality if $H^1(C,{\cal N}^{ea}_{C/S})$ (respectively $H^1(C,{\cal N}^{es}_{C/S})$) vanish. Here ${\cal N}^{ea}_{C/S}$ (respectively ${\cal N}^{es}_{C/S}$) denote the equianalytic (respectively equisingular) normal bundle. The vanishing of $H^1$ implies the independence of the imposed conditions and the smoothness of $H^{ea}_S$ (respectively $H^{es}_S$) at $C$ (cf.\ \S 3). In Theorem 3.7 we prove sufficient conditions for the vanishing of $H^1(C,{\cal N}^{ea}_{C/S})$ (respectively $H^1(C,{\cal N}^{es}_{C/S})$); for the special case $S = \P^2$ we obtain an additional criterion in Corollary 3.12. For the proof we use a vanishing theorem of \cite{GrK} which is an improvement upon the usual vanishing theorem for sheaves which are not locally free. The local isomorphism defect isod$_x ({\cal N}^{ea}_{C/S}, {\cal O}_C)$, which is introduced in 3.4, measures how much ${\cal N}^{ea}_{C/S}$ fails to be free at $x$, and similar for ${\cal N}^{es}_{C/S}$. In many cases of interest, in particular for ${\cal N}^{es}_{C/S}$ and related sheaves, the isomorphism defect is quite big and gives a considerable improvement of the desired vanishing results. Therefore, we make some effort to compute respectively estimate it for certain classes of singularities in \S 4. In \S 5 we give some explicit examples and applications. The present work is, in some sense, a continuation of some part of \cite{GrK}, where only equianalytic families were considered. Our results about the smoothness of $H^{es}_{\P^2}$, which are valid for arbitrary singularities, contain the previously known facts for curves with only ordinary multiple points (cf.\ \cite{Gia}) as a special case; concerning the smoothness of $H^{ea}_{\P^2}$ they are an improvement of \cite{Sh1}. For concrete applications of the theorems of this paper it is important to have good estimates for the isomorphism defects. Most of the formulas concerning these, together with further refinements and detailed proofs, appeared in \cite{Lo}. Our results for $\P^2$ seem to be quite sharp for small $d$ but are asymptotically weaker than those of Shustin \cite{Sh3} which are quadratic in $d$. However, the methods presented here work for arbitrary surfaces and may be combined with Shustin's to provide asymptotically optimal results for curves on (some classes of) rational surfaces. This will be the subject of a forthcoming joint paper. \newpage \section{Equisingular deformations of plane curve singularities} In this paragraph we recall some definitions and results due to J.\ Wahl in the framework of formal deformation theory (cf.\ \cite{Wa2}), transfer them to the complex analytic category and obtain some additional results which are used later. \begin{sub}\label{1.1}{\rm Let $(C,0) \subset ({\Bbb C}^2,0)$ be a reduced plane curve singularity and $f \in {\cal O}_{{\Bbb C}^2,0} = {\Bbb C}\{u,v\}$ a convergent power series defining the germ $(C,0)$. Furthermore, let $m = \mbox{ mult}_0(C)$ denote the multiplicity of $(C,0)$, that is $f \in \frak{m}^m_{{\Bbb C}^2,0}\backslash\frak{m}^{m+1}_{{\Bbb C}^2,0}$ where $\frak{m}_{X,x}$ denotes the maximal ideal of a germ $(X,x)$. Consider a deformation $\varphi : ({\cal C},0) \to (T,0)$ of $(C,0)$ over an arbitrary complex germ $(T,0)$ together with a section $\sigma : (T,0) \to ({\cal C},0)$. Without loss of generality we may assume $\varphi$ to be embedded, that is $\varphi$ is given by a commutative diagram \unitlength1cm \begin{picture}(5,2.5) \put(3.5,2){$(C,0)$} \put(4,1.75){\vector(0,-1){1}} \put(3.9,0.25){0} \put(4.75,2){$\hookrightarrow$} \put(4.75,0.25){$\in$} \put(5.5,2){$({\cal C},0)$} \put(5.75,0.75){\vector(0,1){1}} \put(6,1.75){\vector(0,-1){1}} \put(5.4,1.25){$\sigma$} \put(6.15,1.25){$\varphi$} \put(5.5,0.25){$(T,0)$} \put(6.75,2){$\hookrightarrow$} \put(7.5,2){$({\Bbb C}^2 \times T,0)$} \put(7.8,1.75){\vector(-3,-2){1.5}} \put(7,1){$pr$} \end{picture} where $pr$ is the (natural) projection, $\sigma$ maps to the trivial section and $({\cal C},0)$ is a hypersurface germ of $({\Bbb C}^2 \times T,0)$ defined by a power series $F \in {\cal O}_{{\Bbb C}^2 \times T,0}$. Let $I_{\sigma(T)}$ denote the ideal of $\sigma(T,0) \subset ({\Bbb C}^2 \times T,0)$, then we call the deformation with section $(\varphi,\sigma)$ {\sl equimultiple}, if $F \in I^m_{\sigma(T)}$ (which is, of course, independent of the choice of the embedding and the choice of $F$). } \end{sub} \begin{sub}\label{1.2}{\rm Before defining equisingular deformations, let us recall that the {\sl equisingularity type}\/ (or {\sl topological type}) of $(C,0)$ may be defined as follows: consider an embedded resolution of $(C,0) \subset ({\Bbb C}^2,0)$ given by a sequence of blowing up points $(N \ge 1)$: \begin{equation} M_N \buildrel\pi_N\over \to M_{N-1} \to \cdots \to M_1 \buildrel\pi_1\over\to M_0 = ({\Bbb C}^2,0). \end{equation} Let $C_i \subset M_i$ be the strict and $\hat{C}_i \subset M_i$ the reduced total transform of $C_0 := (C,0) \subset M_0$ under $\psi_i := \pi_1 \circ \cdots \circ \pi_i$. Assume that \begin{itemize} \item $\pi_1$ blows up $0$; \item for $i = 2, \ldots, N$, $\pi_i$ blows up singular points of $\hat{C}_i$; \item $\hat{C}_N$ has only singularities of type $A_1$. \end{itemize} Hence, $C_N$ is smooth and $\psi_N$ induces a resolution of $(C,0)$. If we choose the minimal resolution (that is blowing up only non--nodes of $\hat{C}_i$) we obtain a well--defined system of multiplicity sequences (cf.\ \cite{BrK}), which defines the equisingularity type of $(C,0)$. It is well--known (and was proved by Zariski in \cite{Zar}) that this system of multiplicity sequences determines the embedded topological type of $(C,0)$ and vice versa (cf.\ \cite{BrK}, 8.4 and \cite{Zar} for further characterizations). } \end{sub} \begin{sub}\label{1.3}{\rm Consider a deformation $\varphi : ({\cal C},0) \to (T,0)$ of $(C,0)$ (without section) and assume it to be embedded \unitlength1cm \begin{picture}(4,2.5) \put(4.5,2){$({\cal C},0)$} \put(5,1.75){\vector(0,-1){1}} \put(4.7,1.25){$\varphi$} \put(4.5,0.25){$(T,0)$} \put(5.75,2){$\hookrightarrow$} \put(6.5,2){$({\Bbb C}^2 \times T,0)$} \put(6.75,1.75){\vector(-3,-2){1.5}} \put(6.2,1.15){$pr$} \end{picture} $\varphi$ is called {\sl equisingular}\/ (cf.\ \cite{Wa2}, \S 3, \S 7) if there exists a sequence of blowing up subspaces \begin{equation} {\cal M}_N \buildrel\tilde{\pi}_N\over\to {\cal M}_{N-1} \to \cdots \to {\cal M}_1 \buildrel\tilde{\pi}_1\over \to {\cal M}_0 = ({\Bbb C}^2 \times T,0) \end{equation} such that if ${\cal C}_i \subset {\cal M}_i$ denotes the strict and $\hat{{\cal C}}_i \subset {\cal M}_i$ the reduced total transform of ${\cal C}_0 = ({\cal C},0)$ under $\tilde{\psi}_i := \tilde{\pi}_1 \circ \cdots \circ \tilde{\pi}_i$, the following holds: \begin{itemize} \item[(i)] sequence (1) is induced by (2) via the base change $0 \mapsto T$; \item[(ii)] there is a section $\sigma : (T,0) \to {\cal C}_0$ of $\varphi_0 = \varphi$ such that $\varphi$ is equimultiple along $\sigma$ and $\tilde{\pi}_1$ blows up $\sigma(T,0) \subset {\cal M}_0$; \item[(iii)] for $i = 1, \ldots, N$ there are sections $(T,0) \to {\cal C}_i$ of $\varphi_i = \varphi \circ \tilde{\psi}_i \mid_{{\cal C}_i} : {\cal C}_i \to (T,0)$ through all singular points of $\hat{C}_i$ (each of those sections being mapped via $\tilde{\pi}_i$ to such a section of $\varphi_{i-1}$) such that $\varphi_i$ is equimultiple along them. $\tilde{\pi}_{i+1}: {\cal M}_{i+1} \to {\cal M}_i$ blows up the sections going through those singular points of $\hat{C}_i$ which are blown up by $\pi_{i+1}$ ($i \le N-1$). \end{itemize} The sections of (ii), (iii) are called a (compatible) system of {\sl equimultiple sections} of $\varphi$ through all infinitely near points of $(C,0)$. This definition is obviously independent of the embedding of $\varphi$. Moreover, since an equimultiple deformation of an $A_1$--singularity is trivial, the definition is also independent of the embedded resolution (1). The section $\sigma$ of (ii) is called a {\sl singular section}\/ of $\varphi$. If $(T,0)$ is reduced, then $\varphi : ({\cal C},0) \to (T,0)$ is equisingular if and only if for a small good representative $\varphi : {\cal C} \to T$ and for all $t \in T$ there exists an $x \in \varphi^{-1}(t)$ such that the Milnor number $\mu(\varphi^{-1}(t),x)$ is equal to the Milnor number $\mu (C,0)$. The existence of a singular section was shown by B.\ Teissier (\cite{Te}, \S 5). } \end{sub} The following theorem is basically due to J.\ Wahl (\cite{Wa2}, Theorem 7.4): \begin{theorem} Let $\varphi : ({\cal C},0) \to (T,0)$ be any equisingular deformation of the reduced plane curve singularity $(C,0) \subset ({\Bbb C}^2,0)$. \begin{itemize} \item[(i)] The equimultiple sections through all infinitely near points of $(C,0)$ which are required to exist for $\varphi$ are uniquely determined. \item[(ii)] Let $\phi : {\cal C}_{(C,0)} \to S_{(C,0)}$ be the semiuniversal deformation of $(C,0)$. Then there exists a smooth subgerm $S^{es}_{(C,0)} \subset S_{(C,0)}$ such that if $\varphi$ is induced from $\phi$ via the base change $\psi : (T,0) \to S_{(C,0)}$, then $\psi$ factors through $S^{es}_{(C,0)}$. In particular, the restriction of $\phi$ to $S^{es}_{(C,0)}$ is a {\rm semiuniversal equisingular deformation} of $(C,0)$. \item[(iii)] Let $T_\varepsilon := \mbox{ Spec}({\Bbb C}[\varepsilon]/\varepsilon^2)$ be the base space of first order infinitesimal deformations. The set \begin{center} $I^{es}(C,0) := \{g \in {\Bbb C}\{u,v\}\; \Big|$ \raisebox{1.5ex}{$F = f + \varepsilon g$ defines an equisingular}\hspace{-5.25cm}\raisebox{-1.5ex}{deformation of $(C,0)$ over $T_\varepsilon$}\hspace{1cm}$\Big\}$ \end{center} is an ideal, the {\rm equisingularity ideal}\/ of $(C,0)$. Especially it contains the Jacobian ideal \[ j(C,0) = (f,\frac{\partial f}{\partial u},\; \frac{\partial f}{\partial v}) \cdot {\Bbb C} \{u,v\} \] and the vector space $I^{es}(C,0)/j(C,0)$ is isomorphic to the tangent space of $S^{es}_{(C,0)} \subset S_{(C,0)}$. \end{itemize} \end{theorem} {\bf Proof}: Wahl considers only deformations over Artinian spaces $(T,0)$ but the above facts follow easily from his results: \begin{itemize} \item[(i)] Since we require the existence of holomorphic sections over arbitrary complex germs $(T,0)$, by Wahl these are unique modulo arbitrary powers of the maximal ideal $\frak{m}_{T,0}$, hence unique. \item[(ii)] The existence of a smooth formal semiuniversal equisingular deformation of $(C,0)$ was proved by Wahl. The existence of a convergent representative can be deduced from his result by applying Artin's and Elkik's algebraization theorems. A simple direct proof, using the deformation of the parametrization, is given in \cite{Gr}. \item[(iii)] follows directly from (\cite{Wa2}, Proposition 6.1).\hfill $\Box$ \end{itemize} \begin{proposition}\label{1.5} Openness of versality holds for equisingular deformations, that is if $\varphi : ({\cal C},0) \to (T,0)$ is an equisingular deformation of $(C,0)$, then for any equisingular representative $\varphi : {\cal C} \to T$ together with the singular section $\sigma : T \to {\cal C}$ the set of points $t \in T$ such that $({\cal C}, \sigma(t)) \to (T,t)$ is a versal deformation of $(\varphi^{-1}(t), \sigma(t))$ is a Zariski--open subspace of $T$. \end{proposition} {\bf Proof}: This follows quite formally from a criterion for openness of versality due to Flenner (\cite{Fl}, Satz 4.3).\hfill $\Box$ \begin{sub}\label{1.6}{\rm Let $\mu(c,0) = \dim_{\Bbb C}({\Bbb C}\{u,v\}/(\frac{\partial f}{\partial u}, \frac{\partial f}{\partial v}))$ respectively $\tau(C,0) = \dim_{\Bbb C}({\Bbb C}\{u,v\}/(f, \frac{\partial f}{\partial u},\; \frac{\partial f}{\partial v}))$ denote the Milnor respectively Tjurina number of $(C,0)$. It is well--known that a deformation of $(C,0)$ over a reduced base $(T,0)$ is equisingular if and only if the Milnor number is constant along the (unique) singular section. Hence $S^{es}_{(C,0)}$, being smooth, coincides with the $\mu$--constant stratum of $S_{(C,0)}$. The codimension of $S^{es}_{(C,0)}$ in $S_{(C,0)}$ is (by Theorem 1.4 (ii) and (iii)) equal to \[ \tau^{es}(C,0) = \dim_{\Bbb C}({\Bbb C}\{u,v\}/I^{es}(C,0)). \] Together with a result of Gabrielov (\cite{Gab}), which states that the {\sl modality}\/ mod$(f)$ of the function $f$ with respect to right equivalence (cf.\ \cite{AGV}) is equal to the dimension of the $\mu$--constant stratum of $f$ in the ($\mu$--dimensional) semiuniversal unfolding of $f$, we obtain the following } \end{sub} \begin{lemma} For any reduced plane curve singularity $(C,0)$ defined by $f \in {\Bbb C}\{u,v\}$, we have \[ \tau^{es}(C,0) = \mu(C,0) - \mbox{ mod}(f). \] \end{lemma} \begin{sub}\label{1.9}{\rm If $\sim$ denotes any equivalence relation of plane curve singularities, a $\sim$--{\sl singularity type} is by definition a (not ordered) tuple ${\cal S} = ((C_1,x_1)/\sim, \ldots, (C_m,x_m)/\sim)$ of $\sim$-equivalence classes with $m$ a non--negative integer. In this paper we are mainly interested in the following two cases: \begin{itemize} \item $\sim =$ analytic equivalence (isomorphism of complex space germs), in this case we call the $\sim$--singularity type {\sl analytic type} and denote it by ${\cal A}$. \item $\sim =$ topological equivalence (embedded homeomorphism of complex space germs (cf.\ 1.3)), the corresponding singularity type is called {\sl equisingularity type} or {\sl topological type} and denoted by ${\cal T}$. \end{itemize} If $(C,x)$ is a reduced plane curve singularity, then ${\cal S}(C,x) = (C,x)/\sim$ denotes its singularity type and if $C$ is a reduced curve with finitely many singular points $x_1, \ldots, x_m$ which are all planar, then ${\cal S}(C) = ((C,x_1)/\sim, \ldots, (C,x_m)/\sim)$ is the $\sim$--singularity type of $C$. For ${\cal S} = {\cal A}$ we obtain ${\cal A}(C)$, the {\sl equianalytic type} of $C$, and for ${\cal S} = {\cal T}$ we obtain ${\cal T}(C)$, the {\sl equisingular type} of $C$. As equisingular deformations preserve the topological type, the equianalytic deformations preserve the analytic type of each fibre, where a deformation $\varphi : ({\cal C},0) \to (T,0)$ of $(C,0)$ is called {\sl equianalytic} if $({\cal C},0)$ is analytic isomorphic to $(C \times T,0)$ over $(T,0)$, that is, $\varphi$ is analytically trivial.} \end{sub} \newpage \section{The equianalytic and equisingular Hilbert scheme} \begin{sub}\label{2.1}{\rm Let $S$ be a smooth surface, $T$ a complex space, then by a {\sl family of embedded (reduced) curves over} $T$ we mean a commutative diagram \[ \begin{array}{lcl} {\cal C} & \buildrel j\over\hookrightarrow & \;\;S \times T\\ \varphi\searrow & & \swarrow pr\\ & T & \end{array} \] where $\varphi$ is a proper and flat morphism such that for all points $t \in T$ the fibre $\varphi^{-1}(t)$ is a {\sl curve}\/ (that is a reduced pure 1--dimensional complex space), moreover, $j : {\cal C} \hookrightarrow S \times T$ is a closed embedding and $pr$ denotes the natural projection. Such a family is called {\sl equianalytic} (respectively {\sl equisingular}) if for all $t \in T$ the induced (embedded) deformation of each singular point of $\varphi^{-1}(t)$ over ($T,t$) is equianalytic (respectively equisingular) --- along the unique singular section $\sigma$ (cf.\ \ref{1.3}). } \end{sub} \begin{sub}\label{2.2}{\rm The Hilbert functor ${\cal H} ilb_S$ on the category of complex spaces defined by \[ {\cal H} ilb_S(T) := \{\mbox{ subspaces of } S \times T, \mbox{ proper and flat over } T\} \] is well--known to be representable by a complex space $H_S$ (cf.\ \cite{Bin}). This means there is a universal family \[ \begin{array}{lcl} {\cal U} & \buildrel j\over\hookrightarrow & \;\; S \times H_S\\ \varphi\searrow & & \swarrow pr\\ & H_S & \end{array} \] such that each element of ${\cal H} ilb_S(T)$, $T$ a complex space, can be induced from $\varphi$ via base change by a {\sl unique} map $T \to H_S$. We define the {\sl equianalytic} (respectively {\sl equisingular}) {\sl Hilbert functor} ${\cal H} ilb^{ea}_S$ (respectively ${{\cal H}}ilb^{es}_S$) to be the subfunctor of ${\cal H} ilb_S$ with \[ \begin{array}{lcl} {\cal H} ilb^{ea}_S(T) & := & \{\mbox{equianalytic families of embedded curves over }T\}\\[0.5ex] {\cal H} ilb^{es}_S(T) & := & \{\mbox{equisingular families of embedded curves over } T\} \end{array} \] Moreover, fixing the analytic (respectively topological) singularity type (cf.\ 1.8), we define \[ \begin{array}{lcl} {\cal H} ilb^{\cal A}_S(T) & := & \{\mbox{families in } {\cal H} ilb^{ea}_S(T) \mbox{ whose fibres have (the fixed) analytic singularity type } {\cal A}\}\\[0.5ex] {\cal H} ilb^{\cal T}_S(T) & := & \{ \mbox{families in } {\cal H} ilb^{es}_S(T) \mbox{ whose fibres have (the fixed) topological singularity type } {\cal T}\} \end{array} \] } \end{sub} \begin{proposition} Let ${\cal A}$ be an analytic singularity type corresponding to the topological type ${\cal T}$, then the functors ${\cal H} ilb^{\cal A}_S$ and ${\cal H} ilb^{\cal T}_S$ are representable by locally closed subspaces $H^{\cal A}_S \subset H^{\cal T}_S \subset H_S$. \end{proposition} \begin{corollary} The functor ${\cal H} ilb^{ea}_S$ (respectively ${\cal H} ilb^{es}_S$) is representable by a complex space $H^{ea}_S$ (respectively $H^{es}_S$) which is given as the disjoint union of all $H^{{\cal A}}_S$ (respectively $H^{\cal T}_S$). \end{corollary} \begin{remark}{\rm If $S = \P^2$ and if we fix the degree of all fibres to be $d$, then $H^{es,d}_{\P^2} \subset \P^N$ (where $N = \frac{d^2 + 3d}{2}$) is given as a {\sl finite}\/ disjoint union of locally closed subspaces, while in general $H^{ea,d}_{\P^2} \subset \P^N$ is an {\sl infinite}\/ union. } \end{remark} \begin{sub}\label{2.3}{\rm For the proof of Proposition 2.3 we need the following Lemma, which, for the equianalytic case, is proven in (\cite{GrK}, Lemma 1.4). A proof for the equisingular case is given in \cite{Gr}. } \end{sub} \begin{lemma} Let $(C,x)$ be the germ of an isolated plane curve singularity and $\varphi :({\cal C},x) \to (B,b)$ a deformation of $(C,x)$, then there are unique closed subgerms $(B^{ea},b) \subset (B^{es},b) \subset (B,b)$ such that for any morphism $f : (T,t) \to (B,b)$: \[ \begin{array}{l} f^\ast\varphi \mbox{ is an equianalytic deformation if and only if } f(T,t) \subset (B^{ea},b)\\ f^\ast\varphi \mbox{ is an equisingular deformation if and only if } f(T,t) \subset (B^{es},b) \end{array} \] Moreover, if $\phi :{\cal C}_{(C,x)} \to S_{(C,x)}$ denotes the semiuniversal deformation of $(C,x)$ and if $\psi : (B,b) \to S_{(C,x)}$ is any morphism inducing $\varphi$ via pull--back, then $(B^{ea},b) = (\psi^{-1}(0),b)$ and $(B^{es},b) = (\psi^{-1}(S^{es}_{(C,x)}),b)$. \end{lemma} \begin{sub}\label{2.4}{\rm {\bf Proof of Proposition 2.3}: First we have to remark that the condition for all fibres to be reduced curves defines an open subspace $\widetilde{H}_S \subset H_S$. Now, let $b \in \widetilde{H}_S$ be such that the fibre $\varphi^{-1}(b)$ in the universal family has topological type ${\cal T}$, then by Lemma 2.7 for each $x \in \varphi^{-1}(b)$ there is a unique closed subspace $(H_x^{es},b) \subset (\widetilde{H}_S,b)$ such that a morphism $f : (T,t) \to (\widetilde{H}_S,b)$ factors through $(H^{es}_x,b)$ if and only if $f^\ast\varphi$ is an equisingular deformation. Let \[ (H^{es},b) := \bigcap\limits_{x \in \varphi^{-1}(b)} (H_x^{es},b) \subset (\widetilde{H}_S,b) \] and $H^{es}(b) \subset \widetilde{H}_S$ be a small (unique) representative, then $\cup H^{es}(b)$, where the union is taken over all $b$ whose fibre $\varphi^{-1}(b)$ has topological type ${\cal T}$, defines a locally closed subspace of $H_S$ which obviously represents ${\cal H} ilb^{\cal T}_S$. The statement for ${\cal H} ilb^{\cal A}_S$ follows in the same manner. \hfill$\Box$ } \end{sub} \newpage \section{Completeness of the equianalytic and equisingular characteristic linear series} \begin{sub}\label{3.1}{\rm Let $S$ be a smooth complex surface and $C \subset S$ a reduced compact curve. Then a {\sl deformation of} $C/S$ over the pointed complex space $T$, $0 \in T$, is a triple $({\cal C}, \tilde{i}, j)$ such that we obtain a Cartesian diagram \unitlength1cm \begin{center} \begin{picture}(7,2.5) \put(2.5,2){$C$} \put(3,2){$\hookrightarrow$} \put(3.15,2.2){$j$} \put(3.75,2){${\cal C}$} \put(4.3,2.1){\line(1,0){0.5}} \put(2.3,1.5){$i\;\cap$} \put(3.75,1.5){$\cap\;\tilde{i}$} \put(2.5,1){$S$} \put(3,1){$\hookrightarrow$} \put(3.75,1){$S \times T$} \put(5,1){$\Phi$ flat} \put(2.5,0.5){$\downarrow$} \put(3.75,0.5){$\downarrow\; \pi$} \put(2.5,0){$0$} \put(3.2,0){$\in$} \put(3.75,0){$T$} \put(4.8,0.1){\vector(-1,0){0.5}} \put(4.8,0.1){\line(0,1){2}} \end{picture} \end{center} where $j$ is a closed embedding and the composed morphism $\Phi := \pi \circ \tilde{i}$ is flat ($S \hookrightarrow S \times T$ denotes the canonical embedding with image $S \times \{0\}$ and $\pi$ is the projection). Two deformations $({\cal C}, \tilde{i}, j)$ and $({\cal C}', \tilde{i}', j')$ of $C/S$ over $T$ are {\sl isomorphic} if there exists an isomorphism ${\cal C} \simeq {\cal C}'$ such that the obvious diagram (with the identity on $S \times T$) commutes. ${\cal D} e\!f_{C/S}$ denotes the deformation functor from pointed complex spaces to sets defined by \[ {\cal D} e\!f_{C/S} (T) := \{\mbox{isomorphism classes of deformations of } C/S \mbox{ over } T\} \] and we have the natural forgetful morphism ${\cal D} e\!f_{C/S} \to {\cal D} e\!f_C$ given by $({\cal C}, \tilde{i}, j) \mapsto$ \hbox{$({\cal C}, \Phi = \pi \circ \tilde{i}, j)$}, where ${\cal D} e\!f_C$ denotes the functor of isomorphism classes of deformations of $C$ (forgetting the embedding). Furthermore, for each point $x \in C$, we consider the morphism ${\cal D} e\!f_C \to {\cal D} e\!f_{C,x}$ where ${\cal D} e\!f_{C,x}$ denotes the functor of isomorphism classes of deformations of the analytic germ $(C,x)$. Let $T_\varepsilon =$ Spec $({\Bbb C}[\varepsilon]/\varepsilon^2)$ be the base space of first order infinitesimal deformations. We turn our attention to a subfunctor ${\cal D} e\!f^\prime_{C,x} \subset {\cal D} e\!f_{C,x}$ such that \[ (T^1)^\prime := {\cal D} e\!f_{C,x}^{\prime}(T_\varepsilon) \] is an ideal in ${\cal D} e\!f_{C,x} (T_\varepsilon) \cong {\Bbb C}\{u,v\}/j(C,x)$ and the corresponding ``global'' subfunctor ${\cal D} e\!f^\prime_{C/S} \subset {\cal D} e\!f_{C/S}$ where ${\cal D} e\!f^\prime_{C/S}(T)$ consists exactly of all those elements of ${\cal D} e\!f_{C/S}(T)$ which are mapped to ${\cal D} e\!f^\prime_{C,x}(T)$ for all points $x \in C$. } \end{sub} {\bf Examples}: \begin{enumerate} \ite ${\cal D} e\!f^{ea}_{C/S}$ the subfunctor of ${\cal D} e\!f_{C/S}$ consisting of all isomorphism classes of equianalytic deformations of $C/S$, that is of those deformations whose induced deformations of the analytic germs $(C,x)$ happen to be equianalytic for all $x \in C$. Here ${\cal D} e\!f^{ea}_{C,x}(T_\varepsilon) = 0$ in ${\Bbb C}\{u,v\}/j(C,x)$. \ite ${\cal D} e\!f^{es}_{C/S}$ the subfunctor of ${\cal D} e\!f_{C/S}$ consisting of all isomorphism classes of equisingular deformations of $C/S$, that is of those deformations whose induced deformations of $(C,x)$ are equisingular for all $x \in C$. Here ${\cal D} e\!f^{es}_{C,x}(T_\varepsilon) = I^{es} (C,x)/j(C,x)$. \ite Further examples are the equimultiple, equigeneric and equiclassical deformation functors (cf.\ \cite{DH}). \end{enumerate} \begin{remark}{\rm ${\cal D} e\!f^{ea}_{C/S}$ coincides with ${\cal D} e\!f^{es}_{C/S}$ if (and only if) $C$ has only simple (ADE)--singularities. } \end{remark} Now, let $J_C$ be the ideal sheaf of $C$ in ${\cal O}_S$, then we have the natural exact sequence \[ 0 \to J_C/J^2_C \to \Omega^1_S \otimes_{{\cal O}_S} {\cal O}_C \to \Omega^1_C \to 0 \] respectively its dual \[ 0 \to \theta_C \to \theta_S \otimes_{{\cal O}_S} {\cal O}_C \buildrel\Psi\over\to {\cal N}_{C/S} \to {\cal T}^1_C \to 0. \] Here ${\cal N}_{C/S} = {\cal O}_S(C) \otimes_{{\cal O}_S} {\cal O}_C$ denotes the normal sheaf of $C$ in $S$ and ${\cal T}^1_C := \mbox{ Coker } (\Psi)$ is a skyscraper sheaf concentrated in the singular points of $C$ with $H^0(C,{\cal T}^1_C) \cong {\cal D} e\!f_C(T_\varepsilon)$ and with stalk in $x \in C$ equal to ${\cal T}^1_{C,x} \cong {\cal D} e\!f_{C,x}(T_\varepsilon) = T^1_{(C,x)}$ (for details cf.\ \cite{Art}). Furthermore, for each subfunctor ${\cal D} e\!f^\prime_{C/S}$ as above, let $({\cal T}^1_C)'$ denote the subsheaf of ${\cal T}^1_C$ with stalk in $x$ isomorphic to $(T^1)^\prime \subset T^1$ and \[ {\cal N}'_{C/S} := \mbox{ Ker}({\cal N}_{C/S} \to {\cal T}^1_C/({\cal T}^1_C)'). \] In particular, \vspace{-1cm} \begin{eqnarray*} {\cal N}^{ea}_{C/S} & = & \mbox{ Ker }({\cal N}_{C/S} \to {\cal T}^1_C),\\ {\cal N}^{es}_{C/S} & = & \mbox{ Ker } ({\cal N}_{C/S} \to {\cal T}^1_C/({\cal T}^1_C)^{es}) \end{eqnarray*} where ${\cal T}^1_{C,x} \cong {\Bbb C} \{u,v\}/j(C,x),\; ({\cal T}^1_C)^{es}_x \cong I^{es}(C,x)/j(C,x)$. \begin{lemma}\label{3.2} There is a canonical isomorphism \[ \Phi: \quad {\cal D} e\!f^\prime_{C/S}(T_\varepsilon) \buildrel\cong\over\longrightarrow H^0(C,{\cal N}'_{C/S}). \] \end{lemma} {\bf Proof}: Each representative of an element in ${\cal D} e\!f^\prime_{C/S}(T_\varepsilon)$ is given by local equations $(f_i + \varepsilon g_i = 0)_{i\in I}$ (where $f_i,g_i \in \Gamma(U_i,{\cal O}_S)$ for an open covering $(U_i) _{i\in I}$ of $S$), which satisfy \begin{itemize} \item $(f_i = 0)_{i\in I}$ are local equations for $C \subset S$ \item $f_i + \varepsilon g_i = (a_{ij} + \varepsilon b_{ij}) \cdot (f_j + \varepsilon g_j)$ on $U_i \cap U_j =: U_{ij}$ with $a_{ij}$ a unit in $\Gamma(U_{ij}, {\cal O}_S)$ and $b_{ij} \in \Gamma(U_{ij}, {\cal O}_S)$ \item the germ $g_{i,x}$ of $g_i$ projects to an element of ${\cal D} e\!f^\prime_{C,x}(T_\varepsilon) = (T^1)^\prime$ for all $x \in C \cap U_i$. \end{itemize} For the induced sections $\frac{g_i}{f_i} \in \Gamma(U_i, {\cal O}_S(C))$ it follows immediately \[ \frac{g_i}{f_i} - \frac{g_j}{f_j} = \frac{a_{ij}g_j + b_{ij}f_j}{a_{ij}f_j} - \frac{g_j}{f_j} = \frac{b_{ij}}{a_{ij}} \equiv 0 \in \Gamma(U_{ij}, {\cal N}_{C/S}) \] and $\frac{g_i}{f_i}$ maps to an element of $({\cal T}^1_C)' \subset {\cal T}^1_C$. Hence, $(\frac{g_i}{f_i})_{i \in I}$ defines a global section in ${\cal N}'_{C/S}$. It is easy to check that in this way we get the isomorphism we were looking for (cf.\ \cite{Mu}, \cite{Lo}).\hfill $\Box$ \begin{sub}\label{3.3} {\rm Let $C$ be a compact reduced curve, ${\cal F}$ and ${\cal G}$ coherent torsion--free sheaves on $C$, which have rank 1 on each irreducible component $C_i$ of $C$ and $x \in C$. Then we define the {\sl local isomorphism defect} of ${\cal F}$ in ${\cal G}$ in $x$ as \[ \mbox{isod}_x ({\cal F}, {\cal G}) := \min(\dim_{{\Bbb C}} \mbox{ Coker} (\varphi : {\cal F}_x \to {\cal G}_x)) \] where the minimum is taken over all (injective) local homomorphisms $\varphi : {\cal F}_x \to {\cal G}_x$. In particular, isod$_x({\cal F},{\cal G})$ is a non--negative integer and not zero only in finitely many points (in \cite{GrK} isod$_x({\cal F},{\cal G})$ was denoted by $-$ind$_x({\cal F},{\cal G})$). We call \[ \mbox{isod}({\cal F},{\cal G}) := \sum_{x \in C} \mbox{ isod}_x ({\cal F},{\cal G}) \] the {\sl total (local) isomorphism defect} of ${\cal F}$ in ${\cal G}$. For an irreducible component $C_i$ of $C$ and $x \in C_i$ we set \[ \begin{array}{l} {\cal F}_{C_i} := {\cal F} \otimes {\cal O}_{C_i}\mbox{ modulo torsion}\\ \mbox{isod}_{C_i,x}({\cal F},{\cal G}) := \min(\dim_{\Bbb C} \mbox{ Coker} (\varphi_{C_i} : {\cal F}_{C_i,x} \to {\cal G}_{C_i,x})) \end{array} \] where the minimum is taken over all $\varphi_{C_i}$, which are induced by local homomorphisms $\varphi : {\cal F}_x \to {\cal G}_x$, and \[ \mbox{isod}_{C_i}({\cal F}, {\cal G}) := \sum_{x\in C_i} \mbox{ isod}_{C_i,x} ({\cal F}, {\cal G}). \] Note that this is again a non--negative integer. In Chapter 4 we present some explicit calculations.} \end{sub} \begin{proposition}\label{3.4} (\cite{GrK}, Proposition 5.2) Let $S$ be a smooth surface, $C \subset S$ a compact reduced curve and ${\cal F}$ a torsion--free coherent ${\cal O}_C$--module which has rank 1 on each irreducible component $C_i$ of $C$ $(i = 1, \ldots, s)$. Then $H^1(C,{\cal F}) = 0$ if for $i = 1, \ldots, s$ \[ \chi({\cal F}_{C_i}) > \chi (w_{C,C_i}) - \mbox{ isod}_{C_i} ({\cal F},w_C). \] Here $\chi({\cal M}) = \dim H^0((C,{\cal M}) - \dim H^1(C,{\cal M})$ for a coherent sheaf ${\cal M}$ on $C$ and $w_C$ denotes the dualizing sheaf, $w_{C,C_i} := w_C \otimes {\cal O}_{C_i}$. \end{proposition} \begin{remark}{\rm Using Riemann--Roch and the adjunction formula, the condition above reads \[ \deg({\cal F}_{C_i}) > (K_S + C) \cdot C_i - \mbox{ isod}_{C_i}({\cal F},{\cal O}_C) \] where $K_S$ is the canonical divisor on $S$. Since isod is a local invariant and since $C$ has planar singularities, we can replace $w_C$ by ${\cal O}_C$. } \end{remark} \begin{theorem}\label{3.5} Let $S$ be a smooth complex surface and $C \subset S$ a reduced compact curve, $H^{ea}_S$ respectively $H^{es}_S$ be the representing spaces for the equianalytic respectively equisingular Hilbert functor, then \begin{itemize} \item[(i)] $\dim(H^{ea}_S, C) \ge C^2 + 1 - p_a(C) - \tau(C)$, with $\tau(C) = \sum_{x \in Sing(C)} \tau(C,x),$\\[1.0ex] $\dim(H^{es}_S, C) \ge C^2 + 1 - p_a(C) - \tau^{es}(C)$, with $\tau^{es}(C) = \sum_{x \in Sing(C)} \tau^{es}(C,x),$\\[1.0ex] where $\tau(C,x) = \dim_{\Bbb C}({\cal O}_{C,x}/j(C,x))$ and $\tau^{es}(C,x) = \dim_{\Bbb C}({\cal O}_{C,x}/I^{es}(C,x))$ \item[(ii)] If $H^1(C, {\cal N}^{ea}_{C/S}) = 0$ (respectively $H^1(C,{\cal N}^{es}_{C/S}) = 0)$ then $H^{ea}_S$ (respectively $H^{es}_S$) is smooth at $C$ of dimension \[ C^2 + 1 - p_a (C) - \tau(C) \quad (\mbox{respectively } C^2 + 1 - p_a(C) - \tau^{es}(C)). \] \item[(iii)] Let $C = C_1 \cup \ldots \cup C_s$ be the decomposition into irreducible components, then \begin{itemize} \item[$\bullet$] $H^1(C, {\cal N}^{ea}_{C/S}) = 0$ if for $i = 1, \ldots, s$ \[ - K_S \cdot C_i > D \cdot C_i + \tau(C_i) - \mbox{ isod}_{C_i}({\cal N}^{ea}_{C/S}, {\cal O}_C) \] \item[$\bullet$] $H^1(C,{\cal N}^{es}_{C/S}) = 0$ if for $i = 1, \ldots, s$ \[ - K_S \cdot C_i > \sum\limits_{x \in\; Sing (C)} \dim_{\Bbb C} (({\cal O}_{C,x}/I^{es}(C,x)) \otimes {\cal O}_{C_i,x}) - \mbox{ isod}_{C_i} ({\cal N}^{es}_{C/S}, {\cal O}_C) \] \end{itemize} where $D = \cup_{j\not= i} C_j$ and $K_S$ denotes the canonical divisor on $S$. Moreover, the isomorphism defects isod$_{C_i}({\cal N}^{ea}_{C/S}, {\cal O}_C)$ (respectively isod$_{C_i}({\cal N}^{es}_{C/S}, {\cal O}_C))$ have the lower bound $\#(C_i \cap \mbox{ Sing } C)$. \end{itemize} \end{theorem} \begin{remark}{\rm \begin{enumerate} \item If all singularities of $C$ are quasi--homogeneous or {\sl ordinary $k$--tuple points} (all branches are smooth with distinct tangents) then we obtain as an equivalent criterium for the vanishing of $H^1(C,{\cal N}^{es}_{C/S})$ \[ - - - K_S \cdot C_i > D \cdot C_i + \tau^{es}(C_i) - \mbox{ isod}_{C_i} ({\cal N}^{es}_{C/S},{\cal O}_C). \] \item {}From the adjunction formula we obtain \[ - - - K_S \cdot C_i = C^2_i - 2 p_a (C_i) + 2. \] \item If $C$ is irreducible, we have \begin{itemize} \item $H^1(C,{\cal N}^{ea}_{C/S}) = 0$ if $-K_SC > \tau(C) - \mbox{ isod}({\cal N}^{ea}_{C/S}, {\cal O}_C)$ \item $H^1(C,{\cal N}^{es}_{C/S}) = 0$ if $-K_SC > \tau^{es}(C) - \mbox{ isod}({\cal N}^{es}_{C/S}, {\cal O}_C)$. \end{itemize} \end{enumerate}} \end{remark} \begin{sub}\label{3.6} {\rm {\bf Proof}: Most parts of the proof are identical for the equianalytic $(ea)$ and the equisingular $(es)$ case, there we use again the notation $H'$ respectively ${\cal N}'_{C/S}$ as above: \vspace{-0.5cm} \begin{itemize} \item[(ii)] Let $H^1(C, {\cal N}'_{C/S}) = 0$ and $A \twoheadrightarrow A/(\eta) = \bar{A}$ be a small extension of Artinian ${\Bbb C}$--algebras. For the smoothness of $(H', C)$, we have to show that each equianalytic (respectively equisingular) family $\bar{{\cal C}}$ over $\bar{A}$ lifts to an equianalytic (respectively equisingular) family ${\cal C}$ over $A$. $\bar{{\cal C}}$ is given by local equations $\bar{F_i} \in \Gamma(U_i, {\cal O}_S \otimes \bar{A})$, where $(U_i)_{i \in I}$ is an open covering of $S$, such that $\bullet$ on $U_{ij} := U_i \cap U_j$, $\bar{F}_i = \bar{G}_{ij} \cdot \bar{F}_j$ with a unit $\bar{G}_{ij}$ $\bullet$ the image $F^{(0)}_i$ of $\bar{F}_i$ in $\Gamma(U_i, {\cal O}_S \otimes {\Bbb C})$ is a local equation for $C \subset S$ $\bullet$ the germs $\bar{F}_{i,x}$ describe an equianalytic (respectively equisingular) deformation of $(C,x)$. On the other hand, we know that the equianalytic (respectively equisingular) functor $E'$, which associates to each Artinian local ${\Bbb C}$--algebra the set of all equianalytic (respectively equisingular) deformations of $(C,x)$ over Spec $A$, is smooth and has a very good deformation theory (cf.\ \cite{Wa1}, for the equisingular case). Using the results of M.\ Schlessinger (\cite{Schl}, Remark 2.17), this guarantees in particular \begin{itemize} \item[$\bullet$] the existence of an equianalytic (respectively equisingular) lifting $F_i \in \Gamma(U_i, {\cal O}_S \otimes A)$ of $\bar{F}_i$ \item[$\bullet$] for any lifting $G_{ij} \in \Gamma(U_{ij}, {\cal O}_S \otimes A)$ of $\bar{G}_{ij}$ the existence of $h_{ij} \in \Gamma(U_{ij}, {\cal O}_S \otimes A)$ with $F_i = G_{ij} \cdot F_j + \eta \cdot h_{ij}$ and $(h_{ij})_x \in j(C,x)$ (respectively $(h_{ij})_x \in I^{es}(C,x)$). \end{itemize} To obtain the lifted family we are looking for, we have to modify the $F_i$ and $G_{ij}$ in a suitable way, such that the $h_{ij}$ become 0. We know \begin{eqnarray*} \eta \cdot h_{ij} + \eta \cdot G_{ij} \cdot h_{jk} & = & F_i - G_{ij} \cdot F_j + G_{ij} \cdot (F_j - G_{jk} \cdot F_k)\\ & = & \eta \cdot h_{ik} + (G_{ik} - G_{ij} \cdot G_{jk}) \cdot F_k \end{eqnarray*} where $(G_{ik} - G_{ij} \cdot G_{jk}) \in \Gamma(U_{ijk}, {\cal O}_S \otimes (\eta))$ and $(\eta) \cdot \frak m_A = 0$. As sections in ${\cal O}_S \otimes A/\frak m_A = {\cal O}_S \otimes {\Bbb C}$ we obtain \[ h_{ij} + G_{ij}^{(0)} \cdot h_{jk} = h_{ik} + \left[\frac{1-G_{ij} \cdot G_{jk} \cdot G_{ik}^{-1}}{\eta}\right] \cdot G^{(0)}_{ik} \cdot F_k^{(0)}. \] Furthermore, $F_i^{(0)} = G_{ij}^{(0)} \cdot F_j^{(0)}$, which implies in $\Gamma(U_{ijk}, {\cal N}_{C/S})$ the cocycle condition \[ \frac{h_{ij}}{F_i^{(0)}} + \frac{h_{jk}}{F_j^{(0)}} = \frac{h_{ik}}{F_i^{(0)}}. \] From the definition of the $h_{ij}$ it follows that $\left(\frac{h_{ij}}{F_i^{(0)}} \mid i, j \in I\right)$ represents an element in $H^1(C,{\cal N}'_{C/S}) = 0$. Hence, there exist $f_i \in \Gamma(U_i, {\cal O}_S \otimes {\Bbb C})$ such that \[ \frac{h_{ij}}{F_i^{(0)}} = \frac{f_j}{F_j^{(0)}} - \frac{f_i}{F_i^{(0)}} \] as sections in ${\cal N}'_{C/S}$, especially $h_{ij} + f_i - f_j \cdot G_{ij}^{(0)} \in \Gamma(U_{ij}, J_C)$ and all germs $(f_i)_x$ lie in the Jacobian (respectively equisingularity) ideal. Defining $g_{ij} := \frac{h_{ij} + f_i - f_j \cdot G_{ij}^{(0)}}{F_j^{(0)}},\; \tilde{F}_i := F_i + \eta \cdot f_i$ and $\widetilde{G}_{ij} := G_{ij} + \eta \cdot g_{ij}$ we obtain the lifted family we were looking for. \item[(i)] The germ $(H'_S, C)$ is the fibre over the origin of a (non--linear) obstruction map $H^0(C, {\cal N}'_{C/S}) \to H^1(C, {\cal N}'_{C/S})$ (cf.\ \cite{La}, Theorem 4.2.4). Hence \vspace{-0.5cm} \begin{eqnarray*} \dim\; H^0(C, {\cal N}'_{C/S}) & \ge & \dim(H'_S, C)\\ & \ge & \dim(H^0(C, {\cal N}'_{C/S})) - \dim(H^1(C, {\cal N}'_{C/S}))\\ & = & \chi({\cal N}_{C/S}) - \chi({\cal T}^1_C/({\cal T}^1_C)')\\ & = & \deg({\cal N}_{C/S}) + \chi({\cal O}_C) - \tau'(C)\\ & = & C^2 + 1 -p_a(C) - \tau'(C) \end{eqnarray*} where $\tau'(C)$ denotes the total Tjurina number of $C$ (respectively $\tau^{es}(C))$. Both inequalities become an equality if $H^1(C,{\cal N}'_{C/S}) = 0$. \item[(iii)] By Proposition \ref{3.4}, $H^1(C, {\cal N}'_{C/S}) = 0$, if for $i = 1, \ldots, s$ \[ \deg(\overline{{\cal N}'_{C/S} \otimes {\cal O}_{C_i}}) > (K_S + C) \cdot C_i - \mbox{ isod}_{C_i} ({\cal N}'_{C/S}, {\cal O}_C) \] where $\overline{\phantom{xxx}}$ denotes reduction modulo torsion. On the other hand, we have the exact sequence \[ 0 \to \overline{{\cal N}'_{C/S} \otimes {\cal O}_{C_i}} \to {\cal N}_{C/S} \otimes {\cal O}_{C_i} \to ({\cal T}^1_C/({\cal T}^1_C)^\prime), \otimes {\cal O}_{C_i} \to 0 \] which implies \[ \deg(\overline{{\cal N}'_{C/S} \otimes {\cal O}_{C_i}}) = C \cdot C_i - \dim_{\Bbb C} H^0(C, ({\cal T}^1_C/({\cal T}^1_C)^\prime) \otimes {\cal O}_{C_i}). \] Finally, we obtain the above criteria by \[ \dim_{\Bbb C} H^0(C, {\cal T}^1_C/({\cal T}^1_C)^{es} \otimes {\cal O}_{C_i}) = \sum_{x \in C} \dim_{\Bbb C}(({\cal O}_{C,x}/I^{es}(C,x)) \otimes {\cal O}_{C_i,x}) \] respectively using the Leibniz rule (with $g$ as the equation of $(D,x))$ by \vspace{-0.5cm} \begin{eqnarray*} \dim_{\Bbb C} H^0(C, {\cal T}^1_C \otimes {\cal O}_{C_i}) & = & \sum_{x \in C} \dim_{\Bbb C} (({\cal O}_{C,x}/j(C,x)) \otimes {\cal O}_{C_i,x})\\ & = & \sum_{x \in C} \dim_{\Bbb C}({\cal O}_{C_i,x}/g \cdot j(C_i,x))\\ & = & \sum_{x \in C} (\dim_{\Bbb C}({\cal O}_{C_i,x}/j(C_i,x)) + \dim_{\Bbb C}({\cal O}_{C_i,x}/g \cdot {\cal O}_{C_i,x}))\\ & = & \tau(C_i) + C_i \cdot D. \end{eqnarray*} \end{itemize} \begin{flushright} $\Box$\end{flushright} } \end{sub} \begin{sub}\label{3.7}{\rm {\bf Curves in} $\P^2({\Bbb C})$ Let $C \subset \P^2 := \P^2({\Bbb C})$ be a reduced curve of degree $d$ with (homogeneous) equation $F(X,Y,Z) = 0$, then we define the {\sl polar} of $C$ relative to the point ($\alpha : \beta : \gamma) \in \P^2$ to be the curve $C_{\alpha\beta\gamma}$ with equation $\alpha \cdot F_X(X,Y,Z) + \beta \cdot F_Y(X,Y,Z) + \gamma \cdot F_Z(X,Y,Z) = 0$. }\end{sub} \begin{lemma} The {\sl generic polar} $C_{\alpha\beta\gamma}$ (with $(\alpha:\beta:\gamma) \in \P^2$ a generic point) is an irreducible curve of degree $d-1$, if and only if $C$ is not the union of $d \ge 3$ lines through the same point. \end{lemma} {\bf Proof}: Applying Bertini's theorem, we have irreducibility if there is no algebraic relation between $F_X, F_Y$ and $F_Z$. Considering such a (homogeneous) relation of minimal degree \[ \sum_\alpha a_\alpha F^{\alpha_1}_X \cdot F^{\alpha_2}_Y \cdot F^{\alpha_3}_Z = 0 \] and differentiating, we obtain a system of equations \[ A \cdot \;\left(\begin{array}{c}F_{XX}\{\Bbb F}_{XY}\{\Bbb F}_{XZ}\end{array}\right) \; + B \cdot\;\left(\begin{array}{c}F_{YX}\{\Bbb F}_{YY}\{\Bbb F}_{YZ}\end{array}\right) \; + \Gamma \cdot\;\left(\begin{array}{c}F_{ZX}\{\Bbb F}_{ZY}\{\Bbb F}_{ZZ}\end{array}\right) \; = 0 \] where $A, B$ and $\Gamma$ generically do not vanish. Now the lemma follows from the fact that the Hessian covariant vanishes identically only if $C$ is the union of $d$ lines through one point (cf.\ \cite{He}, Lehrsatz 6). \hfill $\Box$\\ In the following we choose suitable coordinates such that Sing $C$ lies in the affine plane $Z \not= 0$ and a generic polar $C' \subset P^2$ relative to $(\alpha : \beta : 0)$ with equation $\alpha \cdot F_X + \beta \cdot F_Y = 0$ is irreducible. Then we have an obvious morphism ${\cal O}_{C'}(d) \to {\cal T}^1_C$ given by the natural projections \[ {\cal O}_{C',x} = {\cal O}_{\P^2,x}/(\alpha f_X + \beta f_Y) \to {\cal O}_{\P^2,x}/j(C,x) = {\cal T}^1_{C,x} \] where $f(X,Y) = F(X,Y, 1)$ is the affine equation of $C$. We define \vspace{-0.5cm} \begin{eqnarray*} \tilde{{\cal N}}^{ea}_{C'/\P^2} & := & Ker({\cal O}_{C'}(d) \to {\cal T}^1_C)\\ \tilde{{\cal N}}^{es}_{C'/\P^2} & := & Ker({\cal O}_{C'}(d) \to {\cal T}^1_C/({\cal T}_C^1)^{es}). \end{eqnarray*} \begin{corollary}\label{3.8} Let $C \subset \P^2$ be a reduced projective curve of degree $d$, $C_i(i = 1,\ldots, s)$ its irreducible components and $d_i$ the degree of $C_i$. \begin{enumerate} \item[(i)] $H^1(C, {\cal N}^{ea}_{C/\P^2}) = 0$ if and only if the forgetful morphism ${\cal D} e\!f_{C/\P^2} \to \prod_{x \in Sing\, C}$ ${\cal D} e\!f_{C,x}$ is surjective. \item[(ii)] If $H^1(C, {\cal N}^{ea}_{C/\P^2}) = 0$ (respectively $H^1(C, {\cal N}^{es}_{C/S}) = 0)$ then $(H^{ea}_{\P^2}, C)$ (respectively $(H^{es}_{\P^2},C)$) is smooth of dimension $\frac{1}{2} d(d+3) - \tau(C)$ (respectively $\frac{1}{2} d \cdot (d+3) - \tau^{es}(C))$. \item[(iii)] $H^1(C, {\cal N}^{ea}_{C/\P^2}) = 0$ if for $i = 1, \ldots, s$ \[ 3d_i > d_i \cdot (d - d_i) + \tau(C_i) - \mbox{ isod}_{C_i} ({\cal N}^{ea}_{C/\P^2}, {\cal O}_C), \] moreover, isod$_{C_i}({\cal N}^{ea}_{C/\P^2}, {\cal O}_C) \ge \#$ (Sing$(C) \cap C_i)$,\\ $H^1 (C,{\cal N}^{es}_{C/S}) = 0$ if for $i = 1, \ldots, s$ \[ 3 d_i > d_i \cdot (d-d_i) + \tau^{es}(C_i) - \mbox{ isod}_{C_i}({\cal N}^{es}_{C/\P^2}, {\cal O}_C). \] \item[(iv)] If $C$ is not the union of $d \ge 3$ lines through one point $H^1(C, {\cal N}^{ea}_{C/\P^2})$ (respectively $H^1(C, {\cal N}^{es}_{C/\P^2})$) vanishes, if \[ 4 d > 4 + \tau(C) - \mbox{ isod}(\tilde{{\cal N}}^{ea}_{C'/\P^2}, {\cal O}_{C'}) \] \[ (\mbox{respectively } 4 d > 4 + \tau^{es}(C) - \mbox{ isod} (\tilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'})) \] where $C'$ denotes the generic polar as defined above. \end{enumerate} \end{corollary} {\bf Proof}: (ii) and (iii) follow immediately from Theorem \ref{3.5}. To prove (iv), we consider the exact sequences \[ \begin{array}{ccccccccc} 0 & \to & {\cal N}'_{C/\P^2} & \to & {\cal N}_{C/\P^2} & \to & {\cal T}^1_C/({\cal T}^1_C)' & \to & 0\\ 0 & \to & \tilde{{\cal N}}'_{C'/\P^2} & \to & {\cal O}_{C'}(d) & \to & {\cal T}^1_C/({\cal T}^1_C)' & \to & 0 \end{array} \] and the corresponding long exact cohomology sequences where $'$ represents again both the equianalytic and the equisingular case. We know that ${\cal N}_{C/\P^2} \cong {\cal O}_C(d)$ and (by Proposition \ref{3.4}) $H^1(C,{\cal O}_C(d)) = 0$. Furthermore, if $C$ is not the union of $d \ge 3$ lines through one point, $C'$ is irreducible and \vspace{-0.5cm} \begin{eqnarray*} \deg(\tilde{{\cal N}}'_{C'/\P^2}) - (K_{\P^2} + C') \cdot C' & = & \deg({\cal O}_{C'}(d)) - \tau'(C) - (d-4) \cdot (d-1)\\ & = & 4 \cdot (d-1) - \tau'(C). \end{eqnarray*} Hence, applying Proposition \ref{3.4} the conditions in (iv) guarantee the vanishing of $H^1(C', \tilde{{\cal N}}'_{C'/\P^2})$. Additionally, the exact sequence \[ 0 \to {\cal O}_{\P_2} \to {\cal O}_{\P_2} (d) \to {\cal O}_C (d) \to 0 \] respectively an analogous sequence for $C'$ induce surjective mappings \[ \Phi : H^0(\P^2, {\cal O}_{\P^2}(d)) \twoheadrightarrow H^0(C, {\cal O}_C(d)) \mbox{ respectively } \Phi' : H^0(\P^2, {\cal O}_{\P^2}(d))\twoheadrightarrow H^0(C', {\cal O}_{C'}(d)) \] which lead to a commutative diagram with exact horizontal rows \unitlength1cm \begin{picture}(10,3) \put(0.5,1.5){$H^0(\P^2, {\cal O}_{\P^2}(d))$} \put(3.5,1.8){\vector(2,1){1.2}} \put(3.8,2.2){$\Phi$} \put(5.0,2.4){$H^0(C, {\cal O}_C(d)) \to H^0(C, {\cal T}^1_C/({\cal T}^1_C)') \to H^1(C, {\cal N}'_{C/\P^2}) \to 0$} \put(9.0,1.5){$\|$} \put(3.8,0.75){$\Phi'$} \put(3.5,1.4){\vector(2,-1){1.2}} \put(5.0,0.5){$H^0(C', {\cal O}_{C'}(d)) \to H^0(C, {\cal T}^1_C/({\cal T}^1_C)') \to 0.$} \end{picture} This shows that $H^0(C,{\cal O}_C(d)) \to H^0(C,{\cal T}^1_C/({\cal T}^1_C)')$ is surjective and, hence, $H^1(C, {\cal N}'_{C/\P^2}) = 0$. The same argument shows that $H^0(C, {\cal O}_{C}(d)) \to H^0(C,{\cal T}^1_C)$ is surjective if and only if $H^1(C, {\cal N}^{ea}_{C/\P^2}) = 0$. Since ${\cal D} e\!f_{C/\P^2}(T_\varepsilon) = H^0(C, {\cal O}_C(d))$ and $H^0(C, {\cal T}^1_C) \cong \prod_{x\in Sing\, C}$ ${\cal D} e\!f_{C,x}(T_\varepsilon)$ this is equivalent to ${\cal D} e\!f_{C/\P^2}(T_\varepsilon) \to \prod_{x \in Sing(C)}$ ${\cal D} e\!f_{C,x}(T_\varepsilon)$ being surjective (and ${\cal D} e\!f_{C/\P^2}$ being unobstructed). But ${\cal D} e\!f_{C/\P^2}$ and $\prod$ ${\cal D} e\!f_{C,x}$ being unobstructed, the surjectivity on the tangent level implies the surjectivity of the functors.\hfill $\Box$ \begin{remark}{\rm \begin{itemize} \item[(i)] We call the inequalities in \ref{3.8} (iii) respectively \ref{3.8} (iv) the $3d$-- respectively $4d$--criteria. \item[(ii)] The use of a generic polar is due to Shustin \cite{Sh1}, who obtained (with a different proof) the weaker inequality $4d > 4 + \mu(C)$ instead of \ref{3.8} (iv), where $\mu(C)$ is the total Milnor number of $C$. \end{itemize} } \end{remark} \begin{sub}\label{3.9}{\rm {\bf A generalization}: We are also interested in families of curves in $\P^2$ of degree $d$ where for some singularities the analytic type is fixed, for others only the topological type is fixed and for the remaining singularities any deformation is allowed. Let $C \subset \P^2$ be of degree $d$ and Sing$(C) = \{x_1, \ldots, x_k\} \cup \{y_1, \ldots, y_\ell\} \cup \{z_1, \ldots, z_m\}$. We define the subsheaf $({\cal T}^1_C)^\prime$ of ${\cal T}^1_C$ by \[ ({\cal T}^1_C)'_x = \left\{ \begin{array}{ll} 0 & \mbox{ if } x \in \{x_1, \ldots, x_k\}\\ I^{es}(C,x)/j(C,x) & \mbox{ if } x \in \{y_1, \ldots, y_\ell\}\\ T^1_{(C,x)} & \mbox{else} \end{array}\right. \] and put \[ \tau'(C) := \dim_{\Bbb C} H^0(C,{\cal T}^1_C/({\cal T}^1_C)^\prime) = \sum\limits^k_ {i=1} \tau(C,x_i) + \sum\limits^\ell_{j=1} \tau^{es}(C,y_j). \] Assume there exists a reduced curve $C' \subset \P^2$ of degree $d'$ with the following properties: \begin{itemize} \item[(a)] $C'$ is irreducible, \item[(b)] $\{x_1, \ldots, x_k\} \cup \{y_1, \ldots, y_\ell\} \subset C'$, \item[(c)] if $f_j$ is a local equation of $(C', x_j),\; j = 1, \ldots, k$, then $f_j \in j(C,x_j)$; if $f_j$ is a local equation of $(C',y_j),\; j = 1, \ldots, \ell$, then $f_j \in I^{es}(C,y_j)$. \end{itemize} Define \vspace{-0.5cm} \begin{eqnarray*} {\cal N}' & := & \mbox{Ker} ({\cal N}_{C/\P^2} = {\cal O}_C(d) \to {\cal T}^1_C/({\cal T}^1_C) ^\prime),\\ \widetilde{{\cal N}}' & := & \mbox{Ker} ({\cal O}_{C'}(d) \to {\cal T}^1_C/({\cal T}^1_C)^\prime). \end{eqnarray*} Let ${\cal A}$ be the analytic singularity type defined by $(C,x_1), \ldots, (C,x_k), {\cal T}$ the topological singularity type defined by $(C,y_1), \ldots, (C,y_\ell)$ and let ${\cal H} ilb^{{\cal A},{\cal T}}_{\P^2}$ denote the functor parametrising proper and flat families of reduced curves in $\P^2$ which have $k$ singular points of fixed analytic type ${\cal A}$ and $\ell$ singular points of fixed topological type ${\cal T}$ (see \ref{2.2} for a precise definition). This functor is represented by a locally closed subspace $H^{{\cal A},{\cal T}}_{\P^2} \subset H_{\P^2}$ (cf.\ Proposition 2.3). } \end{sub} \begin{proposition} Let $C \subset \P^2$ be a reduced projective curve of degree $d$, Sing$(C) = \{x_1, \ldots, x_k\} \cup \{y_1, \ldots, y_\ell\} \cup \{z_1, \ldots, z_m\}$ and assume that there exists a curve $C' \subset \P^2$ of degree $d'$ satisfying (a) -- (c) above. \vspace{-0.5cm} \begin{itemize} \item[(i)] If $\ell = 0$, then $H^1(C,{\cal N}') = 0$ if and only if ${\cal D} e\!f_{C/\P^2} \to \prod\limits^k_{i=1}{\cal D} e\!f_{(C,x_i)}$ is surjective. \item[(ii)] If $H^1(C,{\cal N}') = 0$ then $H^{{\cal A},{\cal T}}_{\P^2}$ is smooth at $C$ of dimension $\frac{1}{2} d (d+3) - \tau'(C)$. \item[(iii)] $H^1(C,{\cal N}') = 0$ if $d'(d-d'+3) > \tau'(C) -$ isod$(\widetilde{{\cal N}}',{\cal O}_{C'})$. \end{itemize} \end{proposition} {\bf Proof}: Use the same argumentation as for Corollary 3.12.\hfill $\Box$ \begin{remark}{\rm \begin{enumerate} \item If $\ell = m = 0$, (i) is the same as 3.12 (i). If $\ell = m = 0$ (respectively $k = m = 0)$ (ii) is the same as 3.12\ (ii). If $C$ is irreducible we may take $C' = C$ and then (iii) is equivalent to 3.12 (iii) for $\ell = m = 0$ respectively $k = m = 0$. If $C$ is not the union of $d$ lines through one point, we may take $C'$ to be a generic polar and then (iii) is the same as 3.12 (iv) for $\ell = m = 0$ (respectively $k = m = 0$). \item We obtain the best possible result for a curve $C'$ of degree $d' = \frac{d+3}{2}$ satisfying (a) -- (c). In \cite{Sh3}, Shustin has proven the existence of such an irreducible curve $C'$ of degree \[ d' \le (2\kappa^2 + \sqrt{\kappa}) \sqrt{\mu(C)} + (1 - \frac{1}{\kappa}) d, \] where $\mu(C)$ denotes the total Milnor number and \[ \begin{array}{ll} \kappa = &\; \max\;\{\mu(x_i, C) + \mbox{ mult}_{x_i}(C),\; \mu(y_j,C) + \mbox{ mult}_{y_j}(C)\} - 1\\[-0.5ex] & {1 \le i \le k\atop 1 \le j \le \ell} \end{array} \] \end{enumerate} } \end{remark} \newpage \section{Local isomorphism defects of plane curve singularities} Let $u, v$ be local coordinates of the smooth surface $S$ in a singular point $x$ of the reduced compact curve $C \subset S,\; C = C_1 \cup \ldots \cup C_s$ the decomposition into irreducible components and $f(u,v) = 0$ (respectively $f_i(u,v) = 0$) be local equations of $(C,x)$ (respectively ($C_i,x$)). In the following, we give estimations for the (local) isomorphism defects occurring in Chapter 3: \begin{lemma}\label{4.1} For a reduced plane curve singularity $(C,x) \subset (S,x)$ we have: \begin{itemize} \item [(i)] isod$_x ({\cal N}^{ea}_{C/S}, {\cal O}_C) = 1$ if $(C,x)$ is quasihomogeneous.\\ isod$_x ({\cal N}^{ea}_{C/S}, {\cal O}_C) > 1$ if $(C,x)$ is not quasihomogeneous. \item[(ii)] isod$_{C_i,x} ({\cal N}^{ea}_{C/S}, {\cal O}_C) \ge 1$ for $i = 1, \ldots, s$. \end{itemize} \end{lemma} {\bf Proof}: \vspace{-0.5cm} \begin{itemize} \item[(i)] By definition isod$_x ({\cal N}^{ea}_{C/S}, {\cal O}_C) = \min\; \dim_{\Bbb C}(\mbox{ coker }\, \varphi : j(C,x) \to {\cal O}_{C,x})$ where the minimum is taken over all ${\cal O}_{C,x}$--linear maps. Now the Jacobian ideal $j(C,x)$ of an isolated singularity cannot be generated by a single element and there is an isomorphism $\varphi : j(C,x) \buildrel \cong\over\to \frak{m}_{C,x}$ exactly if $(C,x)$ is quasihomogeneous. \item[(ii)] We have to look for an ${\cal O}_{C,x}$--linear map $\varphi : j(C,x) \to {\cal O}_{C,x}$ whose restriction to $(C_i,x)$ has minimal cokernel. The above statement follows immediately. \hfill $\Box$ \end{itemize} \begin{sub}\label{4.2}{\rm Let $(C,x)$ be quasihomogeneous with positive weight vector $w = (w_1, w_2)$ and (weighted) degree $d$, then (as an ${\cal O}_{C,x}$--ideal) $I^{es}(C,x)$ is generated by the Jacobian ideal $j(C,x)$ and all monomials $u^\alpha v^\beta$ with $w_1 \cdot \alpha + w_2 \cdot \beta \ge d$. Furthermore we have the normalization \[ n : {\cal O}_{C,x} \hookrightarrow \bar{{\cal O}} := \prod\limits^r_{i=1} {\Bbb C}\{t\} \] where $r$ denotes the number of local irreducible components $(C^{(i)},x)$ of $(C,x)$ (not to be confused with the global components $C_i$). In the following, we use the notations cond$({\cal O})$, cond$(j)$ respectively cond$(I^{es})$ for the {\sl conductor ideals} of ${\cal O}_{C,x}$, the Jacobian respectively the equisingularity ideal in ${\cal O}_{C,x}$, where, for an ${\cal O}_{C,x}$--ideal $I$, \[ \mbox{ cond} (I) := \{ g \in I \mid g \cdot \bar{{\cal O}} \subset I\}. \] Furthermore, for all these ${\cal O}_{C,x}$--ideals, we denote by $\Gamma(I) \subset {\Bbb N}^r$ the {\sl set of values} of $I$ and by $\underline{c} (I) \in {\Bbb N}^r$ the {\sl conductor} of $I$, that is $\Gamma(\mbox{cond}(I)) = \underline{c} (I) + {\Bbb N}^r$.} \end{sub} \begin{lemma} Let $(C,x)$ be quasihomogeneous of degree $d$, then \begin{itemize} \item[(i)] isod$_x ({\cal N}^{es}_{C/S}, {\cal O}_C) = \delta(C,x) - \dim_{\Bbb C}(I^{es}(C,x)/$cond$(I^{es}))$\\ {\it especially}: isod$_x({\cal N}^{es}_{C/S}, {\cal O}_C) \ge 1$ with equality if and only if $j(C,x) = I^{es}(C,x)$. \item[(ii)] isod$_{C_i,x} ({\cal N}^{es}_{C/S}, {\cal O}_C) = \dim_{\Bbb C}(({\cal O}_{C,x}/\mbox{cond}({\cal O})) \otimes {\cal O}_{C_i,x}) - \dim_{\Bbb C}((I^{es}(C,x)/\mbox{cond}(I^{es})) \otimes {\cal O}_{C_i,x})$. \end{itemize} \end{lemma} \newpage {\bf Proof}: \begin{itemize} \item[(i)] To calculate isod$_x({\cal N}^{es}_{C/S}, {\cal O}_C)$ we have to consider an ${\cal O}_{C,x}$--linear mapping \[ \Psi : I^{es}(C,x) \to {\cal O}_{C,x} \] with minimal cokernel. We know that such a $\Psi$ maps cond$(I^{es})$ to cond$({\cal O})$, hence we obtain the estimate \vspace{-0.5cm} \begin{eqnarray*} \mbox{isod}_x({\cal N}^{es}_{C/S}, {\cal O}_C) & \ge & \dim_{\Bbb C}({\cal O}_{C,x}/\mbox{cond}({\cal O})) - - - \dim_C(I^{es}(C,x)/\mbox{cond}(I^{es}))\\ & = & \delta(C,x) - \dim_{\Bbb C}(I^{es}(C,x)/\mbox{cond}(I^{es})) \end{eqnarray*} with equality if and only if there exists a $\Psi$ that maps cond$(I^{es})$ onto cond$({\cal O})$, or equivalently (using $\varphi : j(C,x) \buildrel \cong\over\to \frak m_{C,x})$ if and only if we can find an ${\cal O}_{C,x}$--linear mapping \[ \Phi : I^{es}(C,x) \to j(C,x) \] of weighted degree $\underline{c} (j) - \underline{c} (I^{es})$. Now $\underline{c}(I^{es})-\underline{1}$ is a maximal (in the sense of \cite{De}) in the semigroup $\Gamma(I^{es}) \supset \Gamma(j)$. Hence, $(\underline{c}({\cal O}) - \underline{c}(j)) + \underline{c} (I^{es}) - \underline{1}$ is a maximal in $\Gamma({\cal O})$ and using the symmetry of $\Gamma({\cal O})$ we see that \[ (\underline{c}({\cal O}) - \underline{1}) - (\underline{c}({\cal O}) - \underline{c}(j) + \underline{c} (I^{es}) - \underline{1}) = \underline{c}(j) - \underline{c}(I^{es}) \in \Gamma({\cal O}). \] The additional statement is an immediate consequence from the fact that all monomials of degree at least $d$ are contained in cond$(I^{es})$; thus, \vspace{-0.5cm} \begin{eqnarray*} \dim_{\Bbb C}(I^{es}(C,x)/\mbox{cond}(I^{es})) & = & \dim_{\Bbb C}(j(C,x)/\mbox{cond}(I^{es}) \cap j(C,x))\\ & \le & \dim_{\Bbb C}(j(C,x)/\mbox{cond}(j))\\ & = & \delta(C,x) - 1. \end{eqnarray*} \item[(ii)] Follows from the considerations above.\hfill $\Box$ \end{itemize} \begin{remark}{\rm If $(C,x)$ is quasihomogeneous, then it is easy to see that \vspace{-0.5cm} \begin{eqnarray*} \dim_{\Bbb C}(({\cal O}_{C,x}/\mbox{cond}({\cal O})) \otimes {\cal O}_{C_i,x}) & = & \dim_{\Bbb C}({\cal O}_{C_i,x}/\mbox{cond}({\cal O}_{C_i})) + (D \cdot C_i,x)\\ \dim_{\Bbb C}((I^{es}(C,x)/\mbox{cond}(I^{es})) \otimes {\cal O}_{C_i,x}) & \ge & \dim_{\Bbb C}(I^{es}(C_i,x)/\mbox{cond}(I^{es}(C_i,x))) \end{eqnarray*} (where $C_j$, $j = 1, \ldots, s)$ are the irreducible components of $C$ and ($D = \bigcup\limits_{j\not= i} C_j)$. Thus we obtain as an upper bound \[ \mbox{isod}_{C_i,x}({\cal N}^{es}_{C/S}, {\cal O}_C) \le \mbox{ isod}_x({\cal N}^{es}_{C_i/S}, {\cal O}_{C_i}) + (D \cdot C_i,x). \] Furthermore, for integer weights $w_1 \ge w_2$, gcd$(w_1, w_2) = 1$, the difference is bounded by $\frac{w_1 - 1}{w_2} + 2$. } \end{remark} \begin{sub}\label{4.3}{\rm {\bf Examples}: \begin{itemize} \item[(i)] If $(C,x)$ is an ADE--singularity, then isod$_x({\cal N}^{es}_{C/S}, {\cal O}_C) = 1$. \item[(ii)] If $(C,x)$ is homogeneous of degree $r \ge 3$, then \[ \mbox{isod}_x({\cal N}^{es}_{C/S}, {\cal O}_C) = \frac{r \cdot(r-1)}{2} -2 = \tau^{es}(C,x) - \mbox{ mult}_x(C) \] furthermore, let $C = C_i \cup D$ as above, then we obtain for $(C_i,x)$ a smooth branch \[ \mbox{isod}_{C_i,x} ({\cal N}^{es}_{C/S}, {\cal O}_C) = r-2, \] while in the singular case \[ \mbox{isod}_{C_i,x} ({\cal N}^{es}_{C/S}, {\cal O}_C) = (C_i \cdot D, x) + \mbox{ isod}_x ({\cal N}^{es}_{C_i/S}, {\cal O}_{C_i}). \] More generally, these statements are valid, if $(C,x)$ is an ordinary $r$--tuple point ($r$ smooth branches with different tangents, $r \ge 3$); in this case each equimultiple deformation is equisingular. \item[(iii)] If $(C,x)$ has the local equation $u^p - v^q = 0$ where $q \ge p \ge 3$ and $(C_i,x)$ consist of $b \le r = \mbox{ gcd } (p,q)$ irreducible branches, then \[ \mbox{isod}_{C_i,x} ({\cal N}^{es}_{C/S}, {\cal O}_C) = \frac{b}{2r} \cdot (pq(2-\frac{b}{r}) + (r-p-q)) - \Big[\frac{q-2}{p}\Big] - M, \] where $M = 2$ unless $b = 1$ and $q = k \cdot p\;\; (k \in {\Bbb N})$, then $M = 1$. \end{itemize} } \end{sub} \begin{sub}\label{4.4}{\rm Now let $S = \P^2$ and $C \subset \P^2$ be different from the union of $d \ge 3$ lines through one point and $C' \subset \P^2$ denote the (irreducible) generic polar of $C$ (cf.\ \ref{3.7}) with affine equation $\alpha \cdot f_X + \beta f_Y = 0$. } \end{sub} \begin{lemma} \begin{itemize} \item[(i)] isod$_x (\widetilde{{\cal N}}^{ea}_{C'/\P^2}, {\cal O}_{C'}) \ge 0$ with equality if and only if $(C,x)$ is quasihomogeneous. \item[(ii)] isod$_x (\widetilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'}) \ge \delta(C',x) - - - \dim_{\Bbb C}(I^{es}(C,x) \otimes {\cal O}_{C',x}/\mbox{cond}(I^{es}(C,x) \otimes {\cal O}_{C',x}))$. \end{itemize} \end{lemma} {\bf Proof}: By definition, equality in (i) is equivalent to the statement that the ${\cal O}_{C',x}$--ideal generated by $f, f_X$ and $f_Y$ is generated by one single element. Obviously, this holds exactly if $(C,x)$ is quasihomogeneous, (ii) follows from the considerations in the proof of Lemma 4.3.\hfill $\Box$ \begin{remark}{\rm The generic polar $C'$ depends on the whole curve $C$ and not only on the germ $(C,x)$, hence, in general it is {\sl not} enough to know the local equation of $(C,x)$ to be able to calculate isod$_x(\widetilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'})$. For example, if $(C,x)$ is homogeneous of degree $d \ge 6$, $(C',x)$ need not be quasihomogeneous. But, in special cases, we are able to give explicit formulas: } \end{remark} \begin{sub}\label{4.5}{\rm {\bf Examples}: \begin{itemize} \item[(i)] If $(C,x)$ is an ADE--singularity, then isod$_x(\widetilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'}) = 0$. \item[(ii)] If $(C,x)$ has the (homogeneous) local equation $(u^r - v^r = 0)\; (r \ge 3)$ then $(C',x)$ has an equation $(\tilde{u}^{r-1} - \tilde{v}^{r-1} = 0)$ and \[ \mbox{isod}_x (\widetilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'}) = \frac{r \cdot (r-3)}{2}. \] Moreover, in the case of a (not homogeneous) local equation $(u^p - v^q = 0)\; (q > p \ge 3)$, $(C',x)$ has an equation $(\tilde{u}^{p-1} - \tilde{v}^{q-1} = 0)$ and we obtain the estimate \[ \mbox{isod}_x(\widetilde{{\cal N}}^{es}_{C'/\P^2}, {\cal O}_{C'}) \ge \frac{(p-3)(q-1)}{2} + \frac{2gcd(p-1, q-1) - gcd(p,q) -1}{2}- \Big[\frac{q}{p}\Big] + \varepsilon \] where $\varepsilon = 0$ unless $p$ divides $q$, then $\varepsilon = 1$. \end{itemize} } \end{sub} \begin{sub}{\rm {\bf Problem}: Do the different isomorphism defects considered above behave (lower) semicontinuous under equianalytic respectively equisingular deformations of $C/S$? } \end{sub} \newpage \section{Applications and Examples} \begin{corollary}\label{5.1} \begin{itemize} \item[(i)] Let $C \subset \P^2$ be a curve of degree $d$ with {\sl ordinary} $( k_i${--}) {\sl multiple points} $(i = 1, \ldots, N)$ as the only singularities, $C = C_1 \cup \ldots \cup C_s$ its decomposition into irreducible components (deg $C_i = d_i)$ then \[ (H^{es}_{\P^2}, C) \mbox{ is smooth of dimension }\frac{d \cdot(d+3)}{2} - \sum\limits^N_{j=1} \bigg(\frac{k_j \cdot(k_j+1)}{2} - 2\bigg), \mbox{ if for } i = 1, \ldots, s \] \[ 3 \cdot d_i > \sum_{{x \in C_i \cap Sing\, C\atop mult_x(C)>2}} \mbox{mult}_x(C_i) \] where mult$_x(C)$ (respectively mult$_x (C_i)$) denote the multiplicity of $C$ (respectively $C_i$) at $x$. \item[(ii)] Let $C \subset \P^2$ be a curve of degree $d$ whose singularities are all of local equations $(u^{p_i} - v^{q_i} = 0)\; (q_i \ge p_i)$ or ADE--singularities, then \[ (H^{es}_{\P^2}, C)\mbox{ is smooth of dimension } \frac{d(d+3)}{2} - \sum\limits^N_{i=1} \bigg(\frac{(p_i + 1) \cdot (q_i + 1) - gcd(p_i, q_i) - 5}{2} - \Big[\frac{q_i}{p_i}\Big] + \varepsilon_i\bigg) \] where $\varepsilon_i = 0$ unless $p_i$ divides $q_i$, then $\varepsilon_i = 1$, if \[ 4d > 4 + \sum_{\{ADE\}} \mu(C,x) + \sum_{\{not\; ADE\}} (p_i + 2q_i -3 - gcd(p_i -1, q_i - 1)).\] \end{itemize} \end{corollary} {\bf Proof}: The statements follow immediately from Corollary 3.12 and Example 4.5 (ii), respectively Example 4.9 (ii). \hfill $\Box$ \begin{remark}\label{5.2}{\rm The result in (i) was already obtained by C.\ Giacinti-Diebolt (\cite{Gia}) using vanishing theorems on the normalization of $C$. It implies the ancient result of Severi \cite{Sev} that for a curve $C$ with no other singularities but ordinary double points $(H^{es}_{\P^2}, C)$ is smooth. Another consequence of the calculations in Chapter 4 is the following: the contribution of a quasihomogeneous singularity $(C,x)$ with local equation $(u^p - v^q + uv \cdot \tilde{f}(u,v) = 0)$, $q \ge p \ge 3$, to the right--hand side in the $3d$--criterion for an irreducible curve $C$ is \[ \tau^{es}(C,x) - \mbox{ isod}_x({\cal N}^{es}_{C/S}, {\cal O}_C) = p + q - gcd(p,q) - \varepsilon \] where $\varepsilon = 0$, unless $q \equiv 1 \mbox{ mod } p$, then $\varepsilon = 1$, while the contribution of an $A_k$--singularity is $k-1$. This corresponds to the result of E.\ Shustin in \cite{Sh2}. Nevertheless, in some cases the new $4d$--criterion gives more information: } \end{remark} \begin{sub}{\rm {\bf Example}: \begin{itemize} \item[(a)] $(x^4 - x^2 z^2 + y^2 z^2 + y^3 z) \cdot y \cdot (x + 2y + z) \cdot (x-2y-z) = 0$ defines a {\it reducible}\/ curve $C \subset \P^2$ having exactly 3 ordinary triple points lying on 1 line (hence, Corollary \ref{5.1} does {\it not}\/ apply) and 7 ordinary double points. But $4d = 28 > 23 = 4 + \tau(C)$; hence, $(H^{ea}_{\P^2}, C)$, respectively $(H^{es}_{\P^2}, C)$ are smooth of dimension 16. \end{itemize}} \end{sub} \begin{sub}{\rm In general, it is a difficult problem to determine for a given $d$ whether there exists a projective plane curve of degree $d$ having a fixed number of singularities of given {\sl analytic} type. On the other hand, the local deformations of a plane curve singularity are well understood. Hence, knowing about the existence of one low degree curve with ``big'' singularities our $4d$--criterion allows us to give positive answers to some of the above existing problems (answers which we did not obtain with the $3d$--criterion in \cite{GrK}).} \end{sub} \begin{remark}{\rm \begin{enumerate} \item The surjectivity statement in \ref{3.8} (i) implies: let $C \subset \P^2$ be of degree $d$ such that $H^1(C,{\cal N}^{ea}_{C/\P^2}) = 0$ and let $\{x_1, \ldots, x_n\}$ be any subset of Sing$(C)$. If, for $i = 1, \ldots, n$, the germ $(C,x_i)$ admits a deformation with nearby fibre having singularities $y^1_i, \ldots, y^{s_i}_i$, then $C$ admits an embedded deformation with nearby curve $C_t \subset \P^2$ having $y^1_1, \ldots, y^{s_1}_1, \ldots, y^1_n, \ldots, y^{s_n}_n$ as singularities. Hence, there exists a curve of degree $d$ with the $y$'s as singularities. \item The $4d$--criterion in \ref{3.8} has the advantage that $C$ need not be irreducible. On the other hand, in the $3d$--criterion in \ref{3.5} and \ref{3.8} we can completely forget about $A_1$--singularities on $C$. By Lemma 4.1, for a node we have $\tau(C_i,x) -$ isod$_{C_i,x}({\cal N}^{ea}_{C/\P^2}, {\cal O}_C) \le 0$, which can be neglected in the right--hand side of the $3d$--criterion (we actually obtain $-1$ if the node results from the intersection of two global components of $C$). \item Since, for a node ${\cal N}^{ea}_{C/S,x} = {\cal N}^{es}_{C/S,x}$, $\tau_{C,x} = \tau^{es}_{C,x}$ and since the isomorphism defect is a local invariant, we can neglect nodes also in the $3d$--formulas for $es$. For the $4d$--criterion, however, nodes have to be counted with 1 (Lemma 4.7). \end{enumerate}} \end{remark} \begin{sub}{\rm {\bf Examples}: \begin{itemize} \item[(b)] the irreducible curve $C \subset \P^2$ with affine equation \vspace{-0.5cm} \begin{eqnarray*} f(x,y)& = & y^2 - 2x^2y + c_1 xy^2 + c_2y^3+x^4-2c_1x^3y+c_3x^2y^2+c_4xy^3+c_5y^4+c_1x^5\\ & & \phantom{y^2} -(3c_2+2c_3)x^4y - - -(2c_4+2)x^3y^2-(2c_1+2c_5)x^2y^3+c_6xy^4-(c_4+2)y^5\\ & & \phantom{y^2}+(2c_2+c_3)x^6+(c_4+2)x^5y+ (2c_1+c_5) x^4y^2-c_6x^3y^3+(c_4+3)x^2y^4\\ & & \phantom{y^2}+ c_1xy^5-(3c_2+c_3+c_6)y^6 \end{eqnarray*} where $c_1 := 16\alpha\beta^2 - 66\beta^2,\; c_2 := 3 \alpha\beta - \frac{23}{2}\beta,\; c_3 := -(\beta + 7c_2),\; c_4 := -4\alpha + 13,\; c_5 := \beta^2-c_1,\; c_6 := 24\alpha\beta - 92\beta$ and $\alpha,\beta$ (complex) solutions of $4\alpha^2 - 30\alpha + 55 = 0$ and $\beta^3 = \alpha^2 - 7\alpha + 12$ of degree 6 has exactly one singularity, which is of type $A_{19}$. (The equation of this curve was found by H.\ Yoshihara (cf.\ \cite{Yos})). Now $4d = 24 > 23 = \tau(C) + 4$ and, hence, each combination of $A$--singularities given by an adjacent subdiagram of $A_{19}$ occurs on a curve of degree 6 (this is a very simple proof of a well--known result which was previously proved by using moduli theory of $K3$--surfaces). \item[(c)] The curve $C \subset \P^2$ with homogeneous equation $x^9 + zx^8 + z(xz^3+y^4)^2 = 0$ has exactly one singularity at $(0:0:1)$ which is of type $A_{31}$. Again we have $4d=36 > 35 = \tau(C) + 4$, hence $H^{ea}_{\P^2}$ is smooth at $C$. $C$ is obtained by a small deformation of Luengo's example of a degree 9 curve ($x^9 + z(xz^3+y^4)^2=0$, having an $A_{35}$--singularity) with non--smooth $(H^{ea}_{\P^2}, C)$. Of course, our criterion supports also this non--smoothness since $4d = 36 < 39 = \tau(C) + 4$. \item[(d)] $x^7 + y^7 + (x-y)^2 x^2y^2z = 0$ defines an irreducible curve $C \subset \P^2$ which has 3 transverse cusps (not quasihomogeneous!) at $(0:0:1)$ and no other singularities. Since $4d = 28 > 24 + 4 - 1 \ge \tau(C) + 4 -$ isod$_{(0:0:1)} (\widetilde{{\cal N}}^{ea}_{C'/\P^2}, w_{C'})$ we see that every local deformation of 3 transverse cusps can be realized by curves of degree 7. \end{itemize} } \end{sub} \newpage \addcontentsline{toc}{section}{References}
1997-08-22T23:14:19
9503
alg-geom/9503013
en
https://arxiv.org/abs/alg-geom/9503013
[ "alg-geom", "math.AG" ]
alg-geom/9503013
Claus Hertling
G.-M. Greuel, C. Hertling, and G. Pfister
Moduli spaces of semiquasihomogeneous singularities with fixed principal part
31 pages. AMSLaTeX
null
null
null
null
We construct coarse moduli spaces of semiquasihomogeneous hypersurface singularities with respect to right equivalence and contact equivalence. We have to fix the principal part of the semiquasihomogeneous singularities. For the moduli spaces with respect to contact equivalence we also fix the Hilbert function of the Tjurina algebra induced by the weights.
[ { "version": "v1", "created": "Wed, 22 Mar 1995 12:58:16 GMT" } ]
2008-02-03T00:00:00
[ [ "Greuel", "G. -M.", "" ], [ "Hertling", "C.", "" ], [ "Pfister", "G.", "" ] ]
alg-geom
\section*{Introduction}\addcontentsline{toc}{section}{Introduction} One of the important achievements of singularity theory is the explicit classification of certain ``generic'' classes of isolated hypersurface singularities via normal forms and the analysis of its properties (cf.\ \cite{AGV}). More complicated singularities deform into a collection of singularities from these classes and deformation theory is a powerful tool in studying specific singularities. For a further classification of more complicated classes of singularities the explicit determination of normal forms seems to be impossible and not appropriate. The aim of this article is to start towards a classification of isolated hypersurface singularities of any dimension via geometric methods, that is by explicitely constructing a (coarse) moduli space for such singularities with certain invariants being fixed. Our method starts from deformation theory and leads to the construction of geometric quotients of quasiaffine spaces by certain algebraic groups whose main part is unipotent. This last part is a major ingredient and uses the general results of \cite{GP 2}. In projective algebraic geometry, the theory of moduli spaces is highly developed but in singularity theory only a few attempts have been made so far, for example by Ebey, Zariski, Laudal, Pfister, Luengo, Greuel (cf.\ \cite{LP} for a systematic approach and \cite{GP 1} for a short survey). In this paper we consider only semiquasihomogeneous singularities given as a power series $f \in {\Bbb C} \{x_1, \ldots, x_n\}$ or as a complex space germ $(X,0) = (f^{-1}(0),0) \subset ({\Bbb C}^n,0)$, together with positive weights $w_1, \ldots, w_n$ of the variables such that the principal part $f_0$ of $f$ (terms of lowest degree) has an isolated singularity. For the classification we first fix the Milnor number, probably the most basic invariant of an isolated hypersurface singularity. Fixing the Milnor number is known (for $n \not= 3)$ to be equivalent to fixing, in a family, the embedded topological type of the singularity. If the Milnor number is fixed, the classification of semiquasihomogeneous singularities falls naturally into two parts. Firstly, the classification of the quasihomogeneous principal parts or, which amounts to the same, the classification of hypersurfaces in a weighted projective space. Secondly, the classification of semiquasihomogeneous hypersurface singularities with fixed principal part. These two parts differ substantially, since the group actions whose orbits describe isomorphism classes of singularities are of a completely different nature. This article is devoted to the second task. The most important equivalence relations for hypersurface singularities are right equivalence (change of coordinates in the source) and contact equivalence (change of coordinates and multiplication with a unit or, equivalently, preserving the isomorphism class of space germs). It turns out that right equivalence, which is really a classification of functions, is easier to handle. We prove the existence of a finite group $E_{f_0}$ acting on the affine space $T_-$, the base space of the semiuniversal $\mu$--constant deformation of $f_0$ of strictly negative weight, such that $T_-/E_{f_0}$ is the desired coarse moduli space. We also show that a fine moduli space almost never exists. See \S 1 for definitions and precise statements. Hence, $T_-/E_{f_0}$ classifies, up to right equivalence, semiquasihomogeneous power series with fixed principal part. An important step in the construction of moduli spaces with respect to right equivalence as well as with respect to contact equivalence is to prove that isomorphisms between two semiquasihomogeneous functions have necessarily non--negative degree. This is proved in \S 2 and uses the fact that the filtration on the Brieskorn lattice $H''_0(f)$ induced by the weights coincides with the $V$--filtration, which is independent of the coordinates. The proof relies on an analysis of this filtration given in \cite{He}. In order to obtain a moduli space with respect to contact equivalence we have to fix, in addition to the Milnor number, also the Tjurina number. This is clear because the dimensions of the orbits of the contact group acting on $T_-$ depend on the Tjurina number. But fixing the Tjurina number is not sufficient. The orbit space of the contact group for fixed Tjurina number is, as a topological space, in general not separated, hence, cannot carry the structure of a complex space. It turns out, however, that if we fix the whole Hilbert function of the Tjurina algebra induced by the weights, the orbit space is a complex space and a coarse moduli space which classifies, up to contact equivalence, semiquasihomogeneous hypersurface singularities with fixed principal part and fixed Hilbert function of the Tjurina algebra. For precise statements see \S 4. These moduli spaces are actually locally closed algebraic varieties in a weighted projective space. The orbits of the contact group acting on $T_-$ can also be described as orbits of an algebraic group $G = U \rtimes (E_{f_0} \cdot {\Bbb C}^\ast)$ where $E_{f_0}$ is the finite group mentioned above and $U$ is a unipotent algebraic group. The main ingredient for the proof in the case of contact equivalence is the theorem on the existence of geometric quotients for unipotent groups in \cite{GP 2}. But, in order to give the above simple description of the strata, we have to use, in a non--trivial way, also the symmetry of the Milnor algebra, a fact which was already noticed in \cite{LP}. The stratification with respect to the Hilbert function of the Tjurina algebra and the proof for the existence of a geometric quotient are constructive and allow the explicit determination of the moduli spaces and families of normal forms for specific examples. \newpage \section{Moduli spaces with respect to right equivalence} Let ${\Bbb C}\{x_1, \ldots, x_n\} = {\Bbb C} \{x\}$ be the convergent power series ring. Two power series $f, g \in {\Bbb C}\{x\}$ are called {\bf right equivalent} $(\buildrel r\over\sim)$ if there exists a $\psi \in$ Aut$({\Bbb C}\{x\})$ such that $f = \psi(g)$; $f$ and $g$ are called {\bf contact equivalent} $(\buildrel c\over\sim)$ if there exists a $\psi \in$ Aut$({\Bbb C}\{x\})$ and $u \in {\Bbb C}\{x\}^\ast$ such that $f = u \psi(g)$. (Equivalently, the local algebras ${\Bbb C}\{x\}/(f)$ and ${\Bbb C}\{x\}/(g)$ are isomorphic respectively the complex germs $(X,0) \subset ({\Bbb C}^n,0)$ and $(Y,0) \subset ({\Bbb C}^n,0)$ defined by $f$ and $g$ are isomorphic.) Let $d$ and $w_1, \ldots, w_n$ be any integers. A polynomial $f_0 \in {\Bbb C}[x_1, \ldots, x_n] = {\Bbb C}[x]$ is {\bf quasihomogeneous} of {\bf type} $(d; w_1, \ldots, w_n)$ if for any monomial $x^\alpha = x_1^{\alpha_1} \cdot \ldots \cdot x_n^{\alpha_n}$ occurring in $f_0$, \[ \deg\, x^\alpha := |\alpha| := w_1 \alpha_1 + \cdots + w_n\alpha_n \] is equal to $d$. $w_1, \ldots, w_n$ are called {\bf weights} and $\deg\, x^\alpha$ is called the (weighted) {\bf degree} of $x^\alpha$. For an arbitrary power series $f = \sum c_\alpha x^\alpha,\; f \not= 0$, we set \[ \deg\, f = \inf \{|\alpha|\, \mid\, c_\alpha \not= 0\}, \] and call it the degree of $f$. For a family of power series $F = \sum c_{\alpha, \beta} x^\alpha s^\beta \in {\Bbb C} \{x,s\}$, parametrized by ${\Bbb C}\{s\}$, we put $\deg_x F= \inf \{|\alpha| \mid \exists\, \beta$ such that $c_{\alpha,\beta} \not= 0\}$. $f$ is called quasihomogeneous if it is a quasihomogeneous polynomial (of some type). $f$ is called {\bf semiquasihomogeneous} of type $(d; w_1, \ldots, w_n)$, if \[ f = f_0 + f_1, \] where $f_0$ is a quasihomogeneous polynomial of type $(d; w_1, \ldots, w_n)$, $f_1$ is a power series such that $\deg\, f_1 > \deg\, f_0$ and, moreover, $f_0$ has an isolated singularity at the origin. $f_0$ is called the {\bf principal part} of $f$. Two right equivalent semiquasihomogeneous power series of the same type have right equivalent principal parts Recall (\cite{SaK 1}) that a power series $f$ with isolated singularity is right equivalent to a quasihomogeneous polynomial with respect to positive weights if and only if \[ f \in j(f) := (\partial f/\partial x_1, \ldots, \partial f / \partial x_n). \] Moreover, in this case the {\bf normalized weights} $\overline{w}_i = \frac{w_i}{d} \in {\Bbb Q} \cap (0,\frac{1}{2}]$ are uniquely determined. We may consider $f \in {\Bbb C}\{x\}, f(0) = 0$ as a map germ $f : ({\Bbb C}^n,0) \to ({\Bbb C},0)$. An {\bf unfolding} of $f$ over a complex germ or a pointed complex space $(S,0)$ is by definition a cartesian diagram \[ \begin{array}{ccc} ({\Bbb C}^n,0) & \hookrightarrow & ({\Bbb C}^n,0) \times (S,0)\\[1.0ex] f\; \downarrow & & \downarrow \; \phi\\[1.0ex] ({\Bbb C},0) & \hookrightarrow & ({\Bbb C},0) \times (S,0)\\[1.0ex] \downarrow & & \downarrow \\[1.0ex] 0& \hookrightarrow & (S,0). \end{array} \] Hence, $\phi(x,s) = (F(x,s),s)$ and the unfolding $\phi$ is determined by $F : ({\Bbb C}^n,0) \times (S,0) \to ({\Bbb C},0),\; F(x,s) = f(x) + g(x,s),\; g(x,0) = 0$, and we say that $F$ defines an unfolding of $f$. Two unfoldings $\phi$ and $\phi'$ defined by $F$ and $F'$ over $(S,0)$ are called right equivalent if there is an isomorphism $\Psi :({\Bbb C}^n,0) \times (S,0) \buildrel \cong\over\to ({\Bbb C}^n,0) \times (S,0),\; \Psi(x,s) = (\psi(x,s),s)$, such that $\phi \circ \Psi = \phi'$. For the construction of moduli spaces we have to consider, more generally, families of unfoldings over arbitrary complex base spaces. Let $S$ denote a {\bf category of base spaces}, for example the category of complex germs or of pointed complex spaces or of complex spaces. A {\bf family of unfoldings} over $S \in {\cal S}$ is a commutative diagram \[ \begin{array}{rcl} ({\Bbb C}^n,0) \times S & \buildrel\phi\over\longrightarrow & ({\Bbb C},0) \times S\\ \searrow & & \swarrow\\ & S &. \end{array} \] Hence, $\phi (x,s) = (G(x,s),s) = (G_s(x),s)$ and for each $s \in S$, the germ $\phi : ({\Bbb C}^n,0) \times (S,s) \to ({\Bbb C},0) \times (S,s)$ is an unfolding of $G_s : ({\Bbb C}^n,0) \to ({\Bbb C},0)$. A morphism of two families of unfoldings $\phi$ and $\phi' = (G', id_s)$ over $S$ is a morphism $\Psi : ({\Bbb C}^n,0) \times S \to ({\Bbb C}^n,0) \times S,\; \Psi (x,s) = (\psi(x,s),s) = (\psi_s(x), s)$ such that $\phi \circ \Psi = \phi'$ (equivalently : $G_s(\psi(x,s)) = G'_s(x)$). $\phi$ and $\phi'$ are called {\bf right equivalent families of unfoldings} if there is a morphism $\Psi$ of $\phi$ and $\phi'$ such that for each fixed $s \in S$, $\psi_s \in \mbox{ Aut}({\Bbb C}^n,0)$.\\ {}From now on let $f_0 \in {\Bbb C}[x_1, \ldots, x_n]$ denote a quasihomogeneous polynomial with isolated singularity of type $(d; w_1, \ldots, w_n)$ with $w_i > 0$ for $i = 1, \ldots, n$. Consider a power series $f$ which is right equivalent to a semiquasihomogeneous power series $f'$ of type $(d; w_1, \ldots, w_n)$. We say that an unfolding $F$ defines an {\bf unfolding of $\bf f$ of negative weight} over $(S,0)$ if $F$ is right equivalent to $f'(x) + g(x,s)$ with $g(x,0) = 0$ and $\deg_x g > d$. This holds, for instance, if there exists a ${\Bbb C}^\ast$--action with (strictly) negative weights on $(S,0)$ such that $\deg\, g = d$, with respect to the ${\Bbb C}^\ast$--actions on $({\Bbb C}^n,0)$ and on $(S,0)$. By Theorem 2.1 the definition is independent of the choice of $f'$. We shall now describe the semiuniversal unfolding of $f_0$ of negative weight. Let $x^\alpha$, $\alpha \in B \subset {\Bbb N}^n$, be a monomial basis of the Milnor algebra ${\Bbb C}\{x\}/(\partial f_0/\partial x_1, \ldots, \partial f_0/\partial x_n)$ which is of ${\Bbb C}$--dimension $\mu$ (the Milnor number of $f_0$), and let $\bar{F} (x,t) = f_0(x) + \sum_{\alpha \in B} x^\alpha s_\alpha,\; s = (s_\alpha)_{\alpha \in B} \in {\Bbb C}^\mu$ be the semiuniversal unfolding of $f_0$. We are mainly interested in the sub--unfolding over the affine pointed space $T_- = ({\Bbb C}^k,0)$, \[ F(x,t) = f_0(x) + \sum^k_{i=1} t_i m_i,\;\;\; t = (t_1, \ldots, t_k) \in T_-, \] where the $m_i$ are the ``upper'' monomials, that is \[ \{m_1, \ldots, m_k\} = \{ x^\alpha \,\mid\, \alpha\in B,\, |\alpha| > d\}. \] For fixed $t \in T_-, F_t(x) = F(x,t) \in {\Bbb C}[x]$ is a semiquasihomogeneous polynomial with principal part $f_0$. Let $A = {\Bbb C}[(s_\alpha)_{\alpha \in B}]$ and $A_- = {\Bbb C}[t_1, \ldots, t_k]$. If we give weights to $s_\alpha$ and $t_i$ by $w(s_\alpha) = d- |\alpha|$ and $w(t_i) = d - \deg(m_i)$, then $\bar{F}$ and $F$ are quasihomogeneous polynomials in ${\Bbb C}[x,s]$ respectively ${\Bbb C}[x,t]$ and $F$ is the restriction of $\bar{F}$ to $T_-$, the negative weight part of $T =$ Spec\,$A$, defined by $\{t_1, \ldots, t_k\} = \{s_\alpha\mid w(s_\alpha) < 0\}$. {\bf Example}: $f_0 = x^3 + y^3 + z^7$ is quasihomogeneous of type $(d; w_1, w_2, w_3) = (21; 7, 7, 3)$ with Milnor number $\mu = 24$. The upper monomials of a monomial basis of the Milnor algebra ${\Bbb C}\{x,y,z\}/(x^2, y^2, z^6)$ are $m_1 = x z^5,\; m_2 = yz^5,\; m_3 = xyz^3,\; m_4= xyz^4,\; m_5 = xyz^5$ and, hence, $A_- = {\Bbb C}[t_1, \ldots, t_5],\; T_- = {\Bbb C}^5$, \[ F(x,y,z,t) = f_0 + \sum^5_{i=1}t_i m_i = f_0 + t_1xz^5 + t_2yz^5 + t_3xyz^3 + t_4xyz^4 + t_5xyz^5, \] $w(t_1, \ldots, t_5) = (-1, -1, -2, -5, -8)$. \begin{remark}\label{1.1}{\rm Fix any $t \in T_-$. $F$ defines an unfolding of $F_t$ of negative weight over the pointed space $(T_-, t)$. If we restrict this unfolding to the germ $(T_-,t)$ this is actually a semiuniversal unfolding of $F_t$ of negative weight because of the following: The monomials $m_1, \ldots, m_k$ represent certainly a basis of ${\Bbb C} \{x\} /(\frac{\partial F_t}{\partial x_1}, \ldots, \frac{\partial F_t} {\partial x_n})$ for $t$ sufficiently close to $0$, since $\mu(F_t) = \mu(f_0)$. But, using the ${\Bbb C}^\ast$--actions on $T_-$ and on ${\Bbb C}^n$, we see that any $F_t$ is contact equivalent to some $F_{t'}$, $t'$ close to $0$. Hence, ${\cal O}_{{\Bbb C}^n \times T_-,0\times T_-}/(\frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n})$ is actually free over $T_-$ with basis $m_1, \ldots, m_k$ and the result follows. We call the affine family \[ F : {\Bbb C}^n \times T_- \to {\Bbb C}, \] $(x,t) \mapsto f_0(x) + \sum^k_{i=1} t_i m_i$ the {\bf semiuniversal family of unfoldings of negative weight of semiquasihomogeneous power series with fixed principal part} $\bf f_0$. } \end{remark} \begin{lemma}\label{1.2} The family of unfoldings $F$ has the following property. If $f$ is any semiquasihomogeneous power series with principal part $f_0$, then: \begin{enumerate} \item[(i)] $T_- = \{0\}$ if and only if $f_0$ is simple or simple elliptic. \item[(ii)] There exists a $t \in T_-$ such that $f \buildrel r\over\sim F_t$. \item[(iii)] Let $f \buildrel r\over\sim F_t$ and let $G(x,s) = f(x) + g(x,s)$ be any unfolding of $f$ of negative weight over the germ $(S,0)$. Then there exists a morphism, unique on the tangent level, of germs $\varphi : (S,0) \to (T_-, t)$ such that $\varphi^\ast F$ is right equivalent to $G$ (that is $T_-$ does not contain trivial subfamilies of unfoldings). \item[(iv)] Assume $f_0$ is neither simple nor simple elliptic. There exist $t, t' \in T_- , t\neq t'$, arbitrarily close to $0$, such that $F_t \buildrel r\over\sim F_t'$ (that is $F$ is not universal in any neighbourhood of $0 \in T_-$). \end{enumerate} \end{lemma} \newpage {\bf Proof}: \begin{enumerate} \item[(i)] is due to Saito \cite{SaK 2}. \item[(ii)] follows from \cite{AGV}, 12.6, Theorem (p.\ 209). \item[(iii)] If $T_-$ would contain trivial subfamilies of unfoldings there must be a $t \in T_-$ with $\mu(F_t) < \mu(f_0)$, which is not the case. \item[(iv)] The group $\mu_d$ of $d$--th roots of unit acts on $T_-$, has $0$ as fixed point and a non--trivial orbit for any $t \not= 0$. Since for $\xi \in \mu_d,\; F_{\xi\circ t}(\xi \circ x) = \xi^d F_t(x) = F_t(x)$, two different points of an orbit of $\mu_\alpha$ correspond to right equivalent functions, we obtain (iv). \end{enumerate} Let us introduce the notion of a fine and coarse moduli space for unfoldings of negative weight with principal part $f_0$ (the weights $w_1, \ldots, w_n$ and $f_0$ are given as above): let ${\cal S}$ be a category of base spaces. For $S \in {\cal S}$, a {\bf family of unfoldings of negative weight with principal part} $\bf f_0$ over $S$ is a family of unfoldings \[ \phi : ({\Bbb C}^n,0) \times S \to ({\Bbb C},0)\times S,\; (x,s) \mapsto (G(x,s),s) = (G_s(x),s) \] such that: for any $s \in S$, $G_s :({\Bbb C}^n,0) \to ({\Bbb C},0)$ is right equivalent to a semiquasihomogeneous power series with principal part $f_0$ and the germ of $G$ at $s$, $G :({\Bbb C}^n,0) \times (S,s) \to ({\Bbb C},0)$, is an unfolding of $G_s$ of negative weight. For any morphism of base spaces $\varphi : T \to S$, the induced map $\varphi^\ast \phi : ({\Bbb C}^n,0) \times T \to ({\Bbb C},0)\times T,\; (x,t) \mapsto (G(x,\varphi(t)),t)$, is an unfolding of negative weight with principal part $f_0$ over $T$. Hence, we obtain a functor \[ \mbox{Unf}^-_{f_0} : {\cal S} \to \mbox{ sets} \] which associates to $S \in {\cal S}$ the set of right equivalence classes of families of unfoldings of negative weight with principal part $f_0$ over $S$. If $pt \in {\cal S}$ denotes the base space consisting of one reduced point, then \begin{tabular}{lp{12cm}} Unf$^-_{f_0}(pt) =$ & $\{$ right equivalence classes of power series $f \in {\Bbb C}\{x_1, \ldots, x_n\}$ which are right equivalent to a semiquasihomogeneous power series with principal part $f_0\}$. \end{tabular} A {\bf fine moduli space} for the functor Unf$^-_{f_0}$ consists of a base space $T$ and a natural transformation of functors \[ \psi : \mbox{ Unf}^-_{f_0} \to \mbox{ Hom}(-,T) \] such that the pair $(T, \psi)$ represents the functor Unf$^-_{f_0}$. The pair $(T,\psi)$ is a {\bf coarse moduli space} for Unf$^-_{f_0}$ if \begin{enumerate} \item[(i)] if $\psi(pt)$ is bijective, and \item[(ii)] given the solid arrows (natural transformations) in the following diagram \vspace{-0.5cm} \unitlength1cm \begin{picture}(8,3) \put(4,0){Hom$(-,T)$} \put(6.5,0){\line(1,0){0.15}} \put(6.7,0){\line(1,0){0.15}} \put(6.9,0){\line(1,0){0.15}} \put(7.1,0){\line(1,0){0.15}} \put(7.3,0){\line(1,0){0.15}} \put(7.5,0){\line(1,0){0.15}} \put(7.7,-0.1){$>$} \put(8.5,0){Hom$(-,T')$,} \put(6.8,1.5){\vector(-3,-2){1.3}} \put(7.8,1.5){\vector(3,-2){1.3}} \put(6.8,2){Unf$^-_{f_0}$} \end{picture} there exists a unique dotted arrow (natural transformation) making the diagram commutative. \end{enumerate} A fine moduli space is, of course, coarse. The definitions of fine and coarse moduli spaces still depend on the category of base spaces ${\cal S}$. If ${\cal S}$ is the category of complex germs and if $(S,0) \in {\cal S}$, then Hom$((S,0), T)$ denotes the set of morphisms of germs $(S,0) \to (T,t)$ where $t$ may be any point of $T$. In this case, if $(T, \psi)$ is a fine moduli space, given any $t \in T$, there exists a unique (up to right equivalence) universal unfolding of negative weight with principal part $f_0$ over the germ $(T,t)$ which corresponds to id $\in$ Hom$((T,t), (T,t))$. But we may not have a universal family over all of $T$. If ${\cal S}$ is the category of all complex spaces, the existence of a fine moduli space implies the existence of a global universal family over $T$. But we shall see that even for complex germs as base spaces a fine moduli space may not exist. A coarse moduli space, however, does exist even if ${\cal S}$ is the category of all complex spaces. The reason is that for a coarse moduli space we do not require any kind of a universal family. \begin{theorem}\label{1.3} Let $E_{f_0}$ be the finite group defined in Definition \ref{2.5}, acting on $T_-$. The geometric quotient $T_-/E_{f_0}$ is a coarse moduli space for the functor Unf$^-_{f_0} : \mbox{ complex spaces } \to$ sets. \end{theorem} {\bf Proof}: Since $E_{f_0}$ is finite, and the action is holomorphic, the geometric quotient $T_-/E_{f_0}$ exists as a complex space. According to Theorem 2.1, Proposition \ref{2.3} and Corollary \ref{2.5}, for any semiquasihomogeneous power series $f$ with principal part $f_0$ there exists a unique point $\underline{t} \in T_-/E_{f_0}$ such that if $f_t \buildrel r\over\sim f$, $t \in T_-$ maps to $\underline{t}$. In this way we obtain a bijection $\psi(pt)$ from the set of right equivalence classes of semiquasihomogeneous power series with principal part $f_0$ to $T_-$. Now let $G : ({\Bbb C}^n,0) \times S \to ({\Bbb C},0)$ define an element of Unf$^-_{f_0}(S)$ for some complex space $S$. We may cover $S$ by open sets $U_i$ such that there exist morphisms $\varphi_i : U_i \to T_-$ with $\varphi^\ast F \buildrel r\over\sim G|_{U_i}$. Even if the $\varphi_i$ are not unique, by the properties of a quotient the compositions $U_i \buildrel \varphi_i\over \to T_- \to T_-/E_{f_0}$ glue together to give a morphism $S \to T_-/E_{f_0}$. This construction is functorial and provides the desired natural transformation Unf$^-_{f_0} \to \mbox{ Hom}(-,T_-/E_{f_0})$. This finishes the proof of Theorem \ref{1.3} (for further details for construction of moduli spaces via geometric quotients cf.\ \cite{Ne}). \begin{remark}{\rm (i)\quad If $f_0$ is simple or simple elliptic, then the coarse moduli space constructed above consists of one reduced point. Hence, it is even a fine moduli space. (ii)\quad If $f_0$ is neither simple nor simple elliptic, Unf$^-_{f_0}$ does not admit a fine moduli space, even not if we take complex germs as base spaces. This can be seen as follows: assume there exists such a fine moduli space then, since it is also coarse, it must be isomorphic to $T_-/E_{f_0}$. Moreover, there exists a universal unfolding over the germ $(T_-/E_{f_0}, 0)$ which can be induced from the semiuniversal unfolding $F$ over the germ $(T_-,0)$ and vice versa. Since $T_-$ does not contain trivial subfamilies, the semiuniversal family $F$ over $(T_-,0)$ would be universal, which contradicts Lemma \ref{1.2} (iv). } \end{remark} {\bf Example}: Let $f_0(x,y) = x^4 + y^5$. We obtain $T_- = {\Bbb C}$ and $F(x,y,t) = x^4 + y^5 + tx^2y^3,\; (d;w_1,w_2;w(t)) = (20;4,5;-2)$. In this case $E_{f_0} = \mu_d$ and the ring of invariant functions on $T_-$ is ${\Bbb C}[t^{10}]$, hence $T_-/E_{f_0} \cong {\Bbb C}$. We give a computational argument that a fine moduli space does not exist:\\ A local universal family over $(T_-/E_{f_0},0)$ would be given by $G :({\Bbb C}^n,0) \times (T_-/E_{f_0}) \to ({\Bbb C},0),\; (x,y,s) \mapsto G(x,y,s)$. The proof of Theorem \ref{1.3} shows that then $F$ would be induced from $G$ by the canonical map $T_- \to T_-/E_{f_0}$, which is not an isomorphism. Moreover, the fibre $F^{-1}(0)$ would be isomorphic to $G^{-1}(0)$ under the map $(x,y,t) \mapsto (x,y,s = t^{10})$. The image of this map can be computed by eliminating $t$ from $F(x,y,t) = 0,\; s - t^{10} = 0$. The result is the hypersurface defined by $G = (x^4 + y^5)^{10} - sx^{20}y^{30}$. The special fibre for $s = 0$ has a non--isolated singularity, hence is not isomorphic to $f_0 = 0$. \begin{remark}\label{1.5}{\rm Since the group $E_{f_0}$ acts even algebraically on $T_-$ by Proposition \ref{2.4}, $T_-/E_{f_0}$ is an algebraic variety. We may take the category of base spaces ${\cal S}$ to be the category of (separated) algebraic spaces and define (families of) unfoldings in the same manner as above, replacing the analytic local ring ${\Bbb C}\{x\}$ by the henselization of ${\Bbb C}[x]$. With the same proof as above we obtain that $T_-/E_{f_0}$ is a coarse moduli space for the functor \[ \mbox{Unf}^-_{f_0} : \mbox{ algebraic spaces } \to \mbox{ sets.} \] } \end{remark} \newpage \section{Isomorphism of semiquasihomogeneous singularities} We fix weights $w_1,...,w_n \in {\Bbb N}$ and a degree $d\in {\Bbb N}$ such that the normalized weights $\overline{w}_i = {w_i \over d}$ fulfill $0 < \overline{w}_i \leq {1\over 2}$. The weights induce a filtration on ${\Bbb C} \{x\}$. An automorphism $\varphi \neq id$ of ${\Bbb C} \{x\} $ has degree $m=\deg \varphi$ if $m$ is the maximal number such that \[ \deg (\varphi (x_i) - x_i) \geq w_i + m \ \ \forall \ i=1,...,n. \] The automorphisms of degree $\geq 0$ form the group $\mbox{Aut}_{\geq 0}({\Bbb C} \{x\} )$ of all automorphisms of ${\Bbb C} \{x\} $ which respect the filtration. The automorphisms of degree $>0$ form a normal subgroup $\mbox{Aut}_{>0}({\Bbb C} \{x\} )$ in Aut$_{\geq 0}({\Bbb C} \{x\} )$. Automorphisms will be called quasihomogeneous if they map each quasihomogeneous polynomial to a quasihomogeneous polynomial of the same degree. They form a group $G_w \subset \mbox{ Aut}_{\geq 0}({\Bbb C} \{x\} ),$ which is isomorphic to the quotient Aut$_{\geq 0}({\Bbb C} \{x\} )/\mbox{Aut}_{>0}({\Bbb C} \{x\} )$. The image $\varphi (f)$ of a semiquasihomogeneous power series $f$ of degree $d$ by an automorphism $\varphi$ of ${\Bbb C} \{x\} $ is semiquasihomogeneous of the same degree if $\deg \varphi \geq 0 $. The converse is true, too: \begin{theorem} \label{2.1} Let $f$ and $g$ be semiquasihomogeneous of degree $d$, and let $\varphi$ be an automorphism of ${\Bbb C} \{x\} $ such that $\varphi (f) = g$. Then $\deg \varphi \geq 0$. \end{theorem} {\bf Proof:} The proof uses some facts which come from the Gauss--Manin connection for isolated hypersurface singularities (\cite{SS}, \cite{SaM}, \cite{AGVII}, \cite{He}). The main idea is the following: in the case of a semiquasihomogeneous singularity the weights $\overline{w}_i$ induce a filtration on ${\Bbb C} \{x\}$ and a filtration on the Brieskorn lattice $H_0''(f)$. This last filtration coincides with the $V$--filtration and is independent of the coordinates. The Brieskorn lattice $H_0''(f)$ is \[ H_0'' = \Omega^n / df\land d\Omega^{n-1} . \] Here $\Omega^k = \Omega^k_{{\Bbb C}^n,0}$ denotes the space of germs of holomorphic $k$--forms. The class of $\omega \in \Omega^n$ in $H_0''(f)$ is denoted by $s[\omega ]_0 \in H_0''(f)$. The $V$--filtration on $H_0''(f)$ is determined by the orders $\alpha_f (\omega) = \alpha_f(s[\omega]_0)$ of $n$--forms $\omega \in \Omega^n$. The most explicit description of the order $\alpha_f (\omega)$ might be the following (\cite{AGVII}, \cite{He}): \newpage \begin{eqnarray*} \alpha_f(\omega) = \min\,\{ \alpha & | & \exists \mbox{ (manyvalued) continuous family of cycles}\\ &&\delta (t)\in H_{n-1}(X_t,{\Bbb Z}) \mbox{ on the Milnor fibers }X_t \\ &&\mbox{of the singularity }f:({\Bbb C}^n,0) \to ({\Bbb C},0),\\ &&\mbox{such that } a_{\alpha,k} \neq 0 \mbox{ in } \\ &&\int\limits_{\delta (t)} {\omega \over df} = \sum_{\beta,k} a_{\beta,k}\cdot t^\beta \cdot (\ln t)^k \\ &&\mbox{for a } k \mbox{ with } 0\leq k \leq n-1\ \}. \end{eqnarray*} The description shows that we have \[ \alpha_f (\omega) = \alpha_g (\varphi (\omega)) = \alpha_g (\varphi(h) d\varphi(x)) \] for $\omega = h(x)dx_1...dx_n = hdx \in \Omega^n$. Since $f$ is semiquasihomogeneous it is possible to give a simple algebraic description of the order $\alpha_f (\omega)$. Indeed, we define mappings \vspace{-0.5cm} \begin{eqnarray*} \nu_C & : & {\Bbb C}\{ x_1,...,x_n\} \to {\Bbb Q}_{\geq 0}\cup \{\infty\},\\ \nu_\Omega & : & \Omega^n \to {\Bbb Q}_{>-1}\cup \{\infty\},\\ \nu_f & : & H_0''(f) \to {\Bbb Q}_{>-1}\cup \{\infty\} \end{eqnarray*} by \[ \nu_C (x^\alpha) = \sum_{i=1}^n \overline{w}_i \alpha_i\ , \ \nu_C(0) = \infty, \ \nu_C (\sum b_\alpha x^\alpha) = \min \{\nu_C (x^\alpha)\ |\ b_\alpha\neq 0\} \] and \[ \nu_\Omega (hdx) = \nu_C (hx_1...x_n) -1 \] and \[ \nu_f (s[\omega]_0) = \nu_f(\omega) = \max \{\nu_\Omega (\eta)\ |\ s[\eta]_0 = s[\omega]_0\}. \] Then, from \cite{He}, Chapter 2.4, it follows that \[ \nu_f (\omega) = \alpha_f (\omega) = \alpha_g (\varphi(\omega)) = \nu_g (\varphi(\omega)). \] For all $\eta\in \Omega^{n-2}$ we have \[ \nu_f (df\land d\eta) \geq -1 + \sum_j \overline{w}_j + (1-\max( \overline{w}_i )) \geq \sum_j \overline{w}_j - \frac{1}{2}. \] For $\omega$ with \[ \min \{\nu_\Omega (\omega ), \nu_f (\omega ),\nu_g (\varphi (\omega )), \nu_\Omega (\varphi (\omega )) \} < \sum_j \overline{w}_j - \frac{1}{2} \] this implies \[ \nu_\Omega (\omega ) = \nu_f (\omega ) = \nu_g (\varphi (\omega)) = \nu_\Omega (\varphi (\omega )). \] We obtain \[ \sum_j \overline{w}_j -1 = \nu_\Omega (dx) = \nu_f (dx) = \nu_g (d\varphi(x)) = \nu_\Omega (d\varphi(x)). \] For $i$ with $\overline{w}_i < \frac{1}{2}$ we obtain \vspace{-0.5cm} \begin{eqnarray*} \overline{w}_i + \nu_\Omega (dx) & = & \nu_C(x_i) + \nu_\Omega (dx) = \nu_\Omega (x_idx) = \nu_f (x_idx) \\ & = & \nu_g (\varphi (x_i) d\varphi(x)) = \nu_\Omega (\varphi (x_i) d\varphi(x)) = \nu_C (\varphi (x_i)) + \nu_\Omega (d\varphi(x)) \\ & = & \nu_C (\varphi (x_i)) + \nu_\Omega (dx) \end{eqnarray*} and $ \nu_C (\varphi (x_i)) = \overline{w}_i $. For $i$ with $\overline{w}_i = \frac{1}{2}$ the equality $\nu_\Omega (x_idx) = \sum \overline{w}_i - \frac{1}{2} $ implies \[ \nu_\Omega (\varphi (x_idx)) \geq \sum \overline{w}_i - \frac{1}{2} \] and $ \nu_C (\varphi (x_i)) \geq \frac{1}{2} $. Therefore, $\nu_C (\varphi (x_i)) \geq \nu_C (x_i) = w_i \ \ \forall i=1,...,n,$ and thus $\deg \varphi \geq 0$. \hfill \begin{remark}\label{2.2} \rm In the following, Theorem \ref{2.1} will be used to describe a finite group $E_{f_0}\subset \mbox{ Aut}(T_{-})$ which operates transitively on each set of parameters in $T_{-}$ which belong to one right equivalence class. Theorem \ref{2.1} also shows that the Hilbert function of the Tjurina algebra (cf.\ Chapter 4) is an invariant of the contact equivalence class. \end{remark} Now let $f_0\in {\Bbb C}[x_1,...,x_n]$ be quasihomogeneous of degree $d$ with an isolated singularity in 0. Let $m_1,...,m_k$ denote the monomials of degree $>d$ in a monomial base of the Milnor algebra of $f_0$. Consider the semiuniversal unfolding of $f_0$ of negative weight, \[ F = f_0 + \sum_{i=1}^k m_it_i. \] For a fixed value of $t$ we write $F_t = f_0 + \sum m_it_i$. With $\deg t_i = w(t_i) = d - \deg m_i < 0$ we obtain a filtration on ${\Bbb C}[t_1,...,t_k] = A_{-}$ such that $F \in {\Bbb C}[x,t]$ is quasihomogeneous of degree $d$ in $x$ and $t$. We write $T_{-}=Spec\ A_{-}$ (cf.\ \S 1). \begin{proposition} \label{2.3} For any semiquasihomogeneous power series $f$ with principal part $f_0$ there exist an automorphism $\varphi \in Aut_{>0}({\Bbb C}\{x\})$ and a parameter $t\in T_{-}$ such that $\varphi (f) = F_t$. The $t\in T_{-}$ is uniquely determined. \end{proposition} {\bf Proof}: The existence of $\varphi$ and $t$ is proved in \cite{AGV}, 12.6, Theorem (p.\ 209). The following proves the uniqueness of $t$. Let $t$ and $t'\in T_{-}$ and $\psi \in \mbox{ Aut}_{>0}({\Bbb C}\{x\})$ be given such that $\psi (F_t) = F_{t'}$. With $\psi_s(x_i) = x_i + s(\psi(x_i)-x_i)$ we obtain a family $\psi_s$ of automorphisms in Aut$_{>0}({\Bbb C}\{x\})$. The family $\psi_s(F_t)$ of semiquasihomogeneous functions with principal part $f_0$ connects $\psi_0(F_t) = F_t$ and $\psi_1(F_t) = F_{t'}$. The family may not be contained in $T_{-},$ but can be induced from $T_{-}$ by a suitable base change: Following the proof of the theorem in [AGV], 12.6 (p. 209), we can find a family $\chi_s$ of automorphisms and a holomorphic map $\sigma: {\Bbb C} \to T_-$ such that $\chi_s \circ \psi_s (F_t) = F_{\sigma (s)}$ and $\chi_s\in Aut_{>0}({\Bbb C} \{x\})$ and even $\chi_0 = id = \chi_1,\ \sigma (0)=t,\, \sigma (1)=t'$. But since $T_{-}$ is part of the semiuniversal deformation, which is miniversal on the $\mu$--constant stratum, and since $T_{-}$ does not contain trivial subfamilies with respect to right equivalence, $t=t'$ as desired. \hfill \begin{proposition} \label{2.4} \begin{enumerate} \item For any $\varphi \in G_w^{f_0} = \{\psi\in G_w\ |\ \psi(f_0)=f_0\}$ and any $t\in T_{-}$ there exist $s = \theta(\varphi)(t) \in T_{-}$ and an automorphism $\psi\in Aut_{>0}({\Bbb C}\{x\})$ such that $\psi \circ \varphi (F_t) = F_s$. \item The function $\theta (\varphi) : T_{-} \to T_{-}$ is uniquely determined, bijective and fulfills $\theta(\varphi^{-1}) = \theta^{-1}(\varphi)$ and $\theta(\varphi)\circ \theta(\psi) =\theta(\varphi \circ \psi)$ for any $\psi \in G_w^{f_0}$. \item The components $\theta (\varphi)(t_i)$ are quasihomogeneous polynomials in $A_{-}$ of degree $\deg (t_i)$. \end{enumerate} \end{proposition} {\bf Proof}: The statements 1.\ and 2.\ follow from Proposition \ref{2.3} and from the fact that Aut$_{>0}({\Bbb C}\{x\})$ is a normal subgroup of Aut$_{\geq 0}({\Bbb C}\{x\})$. Statement 3.\ follows from the proof of the theorem in \cite{AGV}, 12.6 (p.\ 209). Along the lines of this proof one can construct power series $\psi_1,...,\psi_n \in {\Bbb C}\{x,t\}$ and a family of automorphisms $\psi (t)$ such that $\psi (t)(x_i) = \psi_i(t)$ with the following properties: \begin{quote} $\psi_i$ is quasihomogeneous in $x$ and $t$ of degree $w_i,$ \\ $\psi_i - x_i$ has degree $>w_i$ in $x,$ \\ for any fixed $t$ the automorphism $\psi (t)\in \mbox{ Aut}_{>0}({\Bbb C}\{x\})$ with $\psi (t)(x_i) = \psi_i(t)$ gives $\psi (t) \circ \varphi (F_t) = F_{\theta (\varphi)(t)}$. \end{quote} The power series $F=f_0+\sum m_it_i,$ and $\varphi (F)=f_0+\ldots$ and $\psi (t)\circ \varphi (F) = f_0 + \sum m_i\theta (\varphi)(t_i)$ are all quasihomogeneous of degree $d$ with respect to $x$ and $t$. This proves 3. \hfill \\ The functions $\theta (\varphi)$ are biholomorphic. \begin{definition} The image $\theta (G_w^{f_0})$ in Aut$(T_{-})$ will be denoted by $E_{f_0}$. \end{definition} \begin{corollary}\label{2.5} The map $\theta : G_w^{f_0} \to E_{f_0} \subset Aut (T_{-})$ is a group homomorphism. The automorphisms $\theta (\varphi)$ of $T_{-}$ commute with the ${\Bbb C}^{*}$-operation on $T_{-}$. Each orbit of $E_{f_0}$ consists of all parameters in $T_{-}$ which belong to one right equivalence class. \end{corollary} {\bf Proof}: The first two statements follow from Proposition \ref{2.4}, the third statement follows from Theorem \ref{2.1}. \hfill \begin{proposition}\label{2.6} \begin{enumerate} \item The group $G_w^{f_0}$ is finite if $\overline{w}_1,...,\overline{w}_{n-1} < \frac{1}{2}$ and $\overline{w}_n \leq \frac{1}{2}$. \item The group $E_{f_0}$ is finite. \end{enumerate} \end{proposition} {\bf Proof}:\\ {\bf 1.}\quad The dimension of the algebraic group $G_w$ is \[ \dim G_w = \sum_{i=1}^n \# (\mbox{ monomials } x^\alpha \mbox{ of degree }w_i \ ). \] The group $G_w$ operates on \[ V = \bigoplus_{\deg x^\alpha = d} {\Bbb C}\cdot x^\alpha. \] Let $j(f_0)$ denote the Jacobi ideal of $f_0$ and $j_i(f_0)$ the ideal \[ j_i(f_0) = (\frac{\partial f_0}{\partial x_{1}},\ldots, \frac{\partial f_0}{\partial x_{i-1}}, \frac{\partial f_0}{\partial x_{i+1}},\ldots, \frac{\partial f_0}{\partial x_{n}}). \] The tangent space $T_{f_0}G_w f_0 \subset T_{f_0}V$ of $G_w f_0$ in $f_0$ is \[ T_{f_0} G_w f_0 \cong j(f_0)\cap V. \] For any relation \[ 0 = \sum_{i=1}^n \sum_{\deg x^\alpha = w_i} a_{\alpha,i}\cdot x^\alpha \cdot \frac{\partial f_0}{\partial x_{i}} = \sum_{i=1}^n b_i \frac{\partial f_0}{\partial x_{i}} \] with $a_{\alpha,i}\in {\Bbb C}$ and $b_i = \sum_{\deg x^\alpha = w_i} a_{\alpha,i} \cdot x^\alpha $ we have $\deg b_i = w_i$ and $\deg \frac{\partial f_0}{\partial x_{j}} = d - w_j > w_i$ for $j\neq i$. Therefore, $b_i \not\in j_i(f_0)$ or $b_i = 0$. But since $f_0$ has an isolated singularity, the sequence $(\frac{\partial f_0}{\partial x_{1}},\ldots, \frac{\partial f_0}{\partial x_{n}}) $ is a regular sequence and $\frac{\partial f_0}{\partial x_{i}}$ is not a zero divisor in $j_i(f_0)$. This implies $b_i = 0 $ for any $i,$ and \[ j(f_0)\cap V = \bigoplus_{i=1}^n\; \bigoplus_{\deg x^\alpha = w_i} {\Bbb C} \cdot x^\alpha \cdot {\partial f_0 \over \partial x_{i}}, \] and \[ \dim G_w^{f_0} = \dim G_w - \dim j(f_0) \cap V = 0. \] {\bf 2.}\quad One can order the weights $w_i$ such that $\overline{w}_1,\ldots,\overline{w}_r < \frac{1}{2},$ $\overline{w}_{r+1},\ldots,\overline{w}_n = \frac{1}{2}$. The generalized Morse lemma and Theorem \ref{2.1} imply the existence of an automorphism $\varphi \in G_w$ and of a quasihomogeneous polynomial $g_0 \in {\Bbb C}[x_1,\ldots,x_r]$ of degree $d$ such that $\varphi (f_0) = g_0 + x^2_{r+1} + \ldots + x^2_n $. Now let $\widetilde{m}_1,\ldots,\widetilde{m}_k$ be the monomials of degree $> d$ in a monomial base of the Jacobi algebra of $g_0$. Analogously to $F$ we obtain families \vspace{-0.5cm} \begin{eqnarray*} \widetilde{G} & = & g_0 + \sum_{i=1}^k \widetilde{m}_i \widetilde{t}_i \\ \mbox{and } G & = & \widetilde{G} + x^2_{r+1} + \ldots + x^2_n . \end{eqnarray*} It is well known that $G_t$ and $G_{t'}$ are right equivalent if and only if $\widetilde{G}_t$ and $\widetilde{G}_{t'}$ are right equivalent. Let $\widetilde{w}$ be the tuple of weights $\widetilde{w} = (w_1,\ldots,w_r)$. The group $G_{\widetilde{w}}^{g_0}$ is finite by the first part of this proposition and induces a finite group $\widetilde{E}_{\widetilde{w}}$ of automorphisms of $\widetilde{T}_{-} = Spec\ {\Bbb C}[\tilde{t}]$. In fact this is the largest subgroup of Aut$(\widetilde{T}_{-})$ which respects the right equivalence classes. Similarly to Proposition \ref{2.4} one can prove that $\varphi$ induces a biholomorphic mapping from $T_{-}$ to $\widetilde{T}_{-}$ which respects the right equivalence classes. This gives an injective (in fact bijective) mapping from $E_{f_0}$ to $\widetilde{E}_{\widetilde{w}}$. Hence, $E_{f_0}$ is finite. \hfill \begin{example}\label{2.7} \rm $f_0 = x^3 + y^3 + z^7,\ (d;w_1,w_2,w_3) = (21;7,7,3),\ T_{-} = {\Bbb C}^5$, $F = f_0 + \sum_{i=1}^5 t_im_i = f_0 + t_1xz^5 + t_2yz^5 + + t_3xyz^3 + t_4xyz^4 + t_5xyz^5$, the weights of $(t_1,\ldots,t_5)$ are $(-1,-1,-2,-5,-8)$. $G_w^{f_0}$ contains $6\cdot 3 \cdot 7$ elements: obviously, $G_w^{f_0} \cong G_{(1,1)}^{g_0} \times {\Bbb Z}_7$ where $g_0 = x^3 + y^3$. The group $G_{(1,1)}^{g_0}$ is isomorphic to a subgroup of ${\bf Gl}(2,{\Bbb C})$. The image in ${\bf PGl}(2,{\Bbb C})$ permutes three points in ${\bf P^1C}$ and is isomorphic to $S_3,$ the kernel is isomorphic to $\{ id , \xi \cdot id , \xi^2 \cdot id\},$ where $\xi = e^{2\pi i/3}$. Therefore $G_{(1,1)}^{g_0}$ is \[ G_{(1,1)}^{g_0} = (\langle\alpha\rangle \ltimes \langle\beta\rangle) \times \langle\gamma\rangle \times \langle\delta\rangle \cong S_3 \times {\Bbb Z}_3 \times {\Bbb Z}_7 \] with \vspace{-0.5cm} \begin{eqnarray*} \alpha & : \ (x,y,z)\ \to & ( y, x,z), \\ \beta & : \ (x,y,z)\ \to & (\xi x,\xi^2 y,z), \\ \gamma & : \ (x,y,z)\ \to & (\xi x,\xi y,z), \\ \delta & : \ (x,y,z)\ \to & ( x, y,e^{2\pi i/7}z). \end{eqnarray*} The mapping $\theta: G_w^{f_0} \to E_{f_0}$ is an isomorphism with \vspace{-0.5cm} \begin{eqnarray*} \theta(\alpha) & :\ (t_1,t_2,t_3,t_4,t_5)\ \to & ( t_2, t_1, t_3, t_4 , t_5) ,\\ \theta(\beta ) & :\ (t_1,t_2,t_3,t_4,t_5)\ \to & (\xi t_1,\xi^2 t_2, t_3, t_4, t_5) ,\\ \theta(\gamma) & :\ (t_1,t_2,t_3,t_4,t_5)\ \to & (\xi t_1,\xi t_2,\xi^2 t_3,\xi^2 t_4 ,\xi^2 t_5) ,\\ \theta(\delta) & :\ (t_1,t_2,t_3,t_4,t_5)\ \to & (\zeta^5 t_1,\zeta^5 t_2,\zeta^3 t_3,\zeta^4 t_4 ,\zeta^5 t_5) \mbox{ with } \zeta = e^{2\pi i/7}. \end{eqnarray*} Let ${\Bbb C}^{\ast}$ denote the group of ${\Bbb C}^{\ast}$-operations on $T_{-}$. Then $E_{f_0} \cap {\Bbb C}^{\ast} = \langle \theta(\gamma),\theta(\delta)\rangle$ and $E_{f_0}\cdot {\Bbb C}^{\ast} \cong \langle\theta(\alpha),\theta(\beta)\rangle \times {\Bbb C}^{\ast} \cong S_3\times{\Bbb C}^{\ast}$. \end{example} \newpage \section{Kodaira--Spencer map and integral manifolds} Let $f_0$ be semiquasihomogeneous of type $(d; w_1, \ldots, w_n)$, $w_i > 0$, and $F : {\Bbb C}^n \times T_- \to {\Bbb C},\; (x,t) \mapsto f_0(x) + \sum\limits^k_{i=1} t_im_i$,\enspace the semiuniversal family of unfoldings of negative weight as in \S 1. In order to describe the orbits of the contact group acting on $T_-$ we study the Kodaira--Spencer map of the induced semiuniversal family of deformations (of space germs) defined as follows. Let \[ {\cal X} = \{(x,t) \in {\Bbb C}^n \times T_- \mid F(x,t) = 0\} \] and let $({\cal X}, 0 \times T_-)$ denote the germ of ${\cal X}$ along the trivial section $0 \times T_-$ which is a subgerm of $({\Bbb C}^n \times T_- , 0 \times T_-) = ({\Bbb C}^n,0) \times T_-$. The composition with the projection gives a morphism \[ \phi : ({\cal X}, 0 \times T_-) \hookrightarrow ({\Bbb C}^n,0) \times T_- \to T_- \] such that, for any $t \in T_-$, $(\phi^{-1}(t), (0,t)) \cong ({\cal X}_t,0)\subset ({\Bbb C}^n,0)$ is a semiquasihomogeneous hypersurface singularity with principal part equal to $({\cal X}_0,0) = (f_0^{-1} (0),0) =: (X_0,0)$. We call this family the {\bf semiuniversal family of deformations of negative weight of semiquasihomogeneous hypersurface singularities with fixed principal part} $(\bf X_0,0)$ (see also \S 4). For the study of the Kodaira--Spencer map of $({\cal X},0 \times T_-) \to T_-$ it is more convenient to work on the ring level $A_- \to A_-\{x\}/F$. The Kodaira--Spencer map (cf.\ \cite{LP}) of the family $A_- \to A_-\{x\}/F$, \[ \rho : \mbox{Der}_{\Bbb C} A_- \to (x)A_-\{x\}/\left(F + (x) (\frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n})\right), \] is defined by $\rho(\delta) = \mbox{ class}(\delta F) = \mbox{ class}(\sum\limits^k_{i=1} \delta(t_i)m_i)$. Let ${\cal L}$ be the kernel of $\rho$. ${\cal L}$ is a Lie--algebra and along the integral manifolds of ${\cal L}$ the family is analytically trivial (cf.\ \cite{LP}). In our situation it is possible to give generators of ${\cal L}$ as $A_-$--module: Let $I = A_-\{x\}/(\frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n})$, then $I$ is a free $A_-$--module and $\{m_i\}_{i=1, \ldots, k}$ can be extended to a free basis. Multiplication by $F$ defines an endomorphism of $I$ and $F I \subseteq \bigoplus\limits^k_{i=1} m_iA_-$. Define $h_{\alpha,j}$ by \[ x^\alpha F = \sum h_{\alpha,j} m_j \mbox{ in } I. \] Then $h_{\alpha j}$ is homogeneous of degree $|\alpha| + \deg(t_j) = |\alpha|+ d - \deg(m_j)$. This implies $h_{\alpha j} = 0$ if $|\alpha| + \deg(t_j) \ge 0$, in particular $h_{\alpha j} = 0$ if $|\alpha| \ge (n-1)d -2 \sum w_i$. For $\alpha$ and $|\alpha| < (n-1)d -2 \sum w_i$ let $\delta_\alpha := \sum h_{\alpha,j} \frac{\partial}{\partial t_j}$. \begin{proposition}\label{3.1} (cf.\ \cite{LP}, Proposition 4.5): \begin{enumerate} \item $\delta_{\alpha}$ is homogeneous of degree $|\alpha|$. \item ${\cal L} = \sum A_- \delta_\alpha$. \end{enumerate} \end{proposition} Now there is a non--degenerate pairing on $I$ (the residue pairing) which is defined by $\langle h, k\rangle = \mbox{ hess} (h \cdot k)$. Here $hess(h)$ is the evaluation of $h$ at the socle (the hessian of $f$). Using the pairing one can prove the following: \begin{proposition}\label{3.2} There are homogeneous elements $n_1, \ldots, n_k \in A_- \{x\}$ with the following properties: \begin{enumerate} \item If $n_i F = \sum^k_{j = 1} h_{ij} m_j$ in $I$ then $h_{ij} = h_{k-j+1, k-i+1}$. \item If $\delta_i := \sum^k_{j=1} h_{ij} \frac{\partial}{\partial t_j}$ then $\delta_i$ is homogeneous of degree $\deg(n_i)$ and ${\cal L} = \sum^k_{i = 1} A_-\delta_i$. \end{enumerate} \end{proposition} In \cite{LP} (Proposition 5.6) this proposition is proved for $n = 2$. The proof can easily be extended to arbitrary $n$. The important fact is the symmetry, expressed in 1.\\ Let $L_+$ be the Lie--algebra of all vector fields of ${\cal L}$ of degree $\ge w = \min\{w_i\}$. Then $L$ is finite dimensional and nilpotent. $\delta_2, \ldots, \delta_k \in L_+$ and $\delta_1 = \sum\limits^k_{i=1} \deg(t_i) t_i \frac{\partial}{\partial t_i}$ is the Euler vector field (cf. \cite{LP}). Let $L = L_+ \oplus {\Bbb C} \delta_1$ then $L$ is a finite dimensional and solvable Lie--algebra and ${\cal L} = \sum A_- L,\; L/L_+ \cong {\Bbb C}\delta_1$. \begin{corollary}\label{3.3} The integral manifolds of ${\cal L}$ coincide with the orbits of the algebraic group $exp(L)$. \end{corollary} Now consider the matrix $M(t) := (\delta_i(t_j))_{i, j = 1, \ldots, k} = (h_{ij})_{i,j=1, \ldots, k}$. Evaluating this matrix at $t \in T_-$ we have \begin{eqnarray*} \mbox{rank } M(t) & = & \mbox{dimension of a maximal integral manifold of } {\cal L}\\ & & \mbox{(resp.\ of the orbit of exp(L)) at}\; t\\ & = & \mu - \tau(t), \end{eqnarray*} where $\tau(t)$ denotes the Tjurina number of the singularity defined by $t$ \enspace i.e.\ of $F(x,t)$. \begin{example}\label{3.4}{\rm We continue with Example \ref{2.7}, $f_0 = x^3 + y^3 + z^7$. Let \vspace{-0.5cm} \begin{eqnarray*} n_1 & = & -21\\ n_2 & = & -21 z + \left(\frac{250}{49} t_1^3 t_2 + \frac{55}{7} t_1^2 t_3 - \frac{250}{49} t_2^4\right) y - \frac{55}{7} t_2^2 t_3 x\\ n_3 & = & -21 z^2 - 30 t_2 y\\ n_4 & = & -21x\\ n_5 & = & -21y \end{eqnarray*} then the matrix defined by Proposition \ref{3.2} is \[ (\delta_i(t_j)) = \left( \begin{array}{ccccc} t_1 & t_2 & 2t_3 & 5t_4 & 8t_5\\ 0 & 0 & 0 & 2t_3-\frac{10}{7}t_1t_2 & 5t_4\\ 0 & 0 & 0 & 0 & 2t_3\\ 0 & 0 & 0 & 0 & t_2\\ 0 & 0 & 0 & 0 & t_1 \end{array} \right). \] We have $\mu = 24$ and \begin{tabular}{lp{14cm}} $\tau = 21$ & if and only if $2t_3 - \frac{10}{7} t_1t_2 \not= 0$,\\ $\tau = 22$ & if and only if $2t_3 - \frac{10}{7} t_1t_2 = 0$ and $t_1 \not= 0$ or $t_2 \not= 0$ or $t_3 \not= 0$ or $t_4 \not= 0$,\\ $\tau = 23$ & if and only if $t_1 = t_2 = t_3 = t_4 = 0$ and $t_5 \not= 0$,\\ $\tau = 24$ & if and only if $t_1 = t_2 = t_3 = t_4 = t_5 = 0$. \end{tabular}} \end{example} \newpage \section{Moduli spaces with respect to contact equivalence} In this section we want to construct a coarse moduli space for semiquasihomogeneous hypersurface singularities with fixed principal part with respect to contact equivalence, that is isomorphism of space germs. Such a moduli space does only exist if we fix further numerical invariants. We shall use the Hilbert function of the Tjurina algebra induced by the given weights. Let us first define the functor for which we are going to construct the moduli space. A complex germ $(X,0) \subset ({\Bbb C}^n,0)$ is called a {\bf quasihomogeneous} (respectively {\bf semiquasihomogeneous}) {\bf hypersurface singularity} of type $(d; w_1, \ldots, w_n)$ if there exists a quasihomogeneous polynomial $f \in {\Bbb C}[x_1, \ldots, x_n]$ (respectively a semiquasihomogeneous power series $f \in {\Bbb C}\{x_1, \ldots, x_n\})$ of type $(d; w_1, \ldots, w_n)$ such that $(X,0) = (f^{-1} (0),0)$. If $f_0$ is the principal part of $f$ then $(X_0,0) = (f_0^{-1}(0),0)$ is called the {\bf principal part} of $(X,0)$. Multiplying $f$ with a unit changes $f_0$ by a constant, hence the principal part if well--defined. Two power series are contact equivalent if and only if the corresponding space germs are isomorphic. A {\bf deformation} ({\bf with section}) of $(X,0)$ over a complex germ or a pointed complex space $(S,0)$ is a cartesian diagram \[ \begin{array}{ccc} 0 & \hookrightarrow & (S,0)\\ \downarrow & & \downarrow\; \sigma\\ (X,0) & \hookrightarrow & ({\cal X},0)\\ \downarrow & & \downarrow\; \phi\\ 0 & \hookrightarrow & (S,0) \end{array} \] such that $\phi$ is flat and $\phi \circ \sigma =$ id. Two deformations $(\phi, \sigma)$ and $(\phi', \sigma')$ of $(X,0)$ over $(S,0)$ are isomorphic if there is an isomorphism $({\cal X},0) \buildrel\cong\over\to ({\cal X}',0)$ such that the obvious diagram commutes. We shall only consider deformations with section. If $(X,0) = (f^{-1}(0),0)$ and if $F : ({\Bbb C}^n,0) \times (S,0) \to ({\Bbb C},0)$ is an unfolding of $f$ then the projection $({\cal X},0) = (F^{-1}(0),0) \to (S,0)$ is a deformation of $(X,0) \hookrightarrow ({\cal X},0)$ with trivial section $\sigma(s) = (0,s)$. Conversely, any deformation of $(X,0)$ is isomorphic to a deformation induced by an unfolding in this way. A deformation ($\phi,\sigma$) of a hypersurface singularity $(X,0)$, which is isomorphic to a semiquasihomogeneous hypersurface singularity $(X',0) = (f^{-1}(0),0)$ of type $(d; w_1, \ldots, w_n)$ over $(S,0)$, is called {\bf deformation of negative weight} if it is isomorphic to a deformation induced by an unfolding of $f$ of negative weight. We have to show that the definition is independent of the chosen unfolding: two inducing unfoldings differ by a right equivalence and a multiplication with a unit. We have shown in \S 1 that the definition depends only on the right equivalence class. Hence, we have to show the following: if $f(x)$ is a semiquasihomogeneous power series, $f(x) + g(x,s),\; g(x,0) = 0,\; \deg_x g > d$, an unfolding of negative weight and $u(x,s) \in {\cal O}^\ast_{{\Bbb C}^n \times S,0}$ a unit, then $u(f+g) \buildrel r\over\sim f'(x) + g'(x,s)$ with $f^{-1}(0) = f'^{-1}(0),\; g'(x,0) = 0$ and $\deg_x g'>d$. Replacing $u(x,s)$ by $(u(x,0))^{-1} u(x,s)$ we may assume that $u(x,s) = u_0(s) + su_1(x,s),\; u_0(0) = 1,\; u_1(0,s) = 0$. If $\nu \in {\cal O}_{S,0}$ is a $d$--th root of $u_0$ and if $\psi$ denotes the automorphism of degree 0, $\psi (x,s) = (\nu(s)^{w_1}x_1, \ldots, \nu(s)^ {w_n}x_n)$, then $u_0(s) f(x) = f(\psi(x,s)) + s \tilde{f}(x,s),\; \deg_x \tilde{f} > d$. But this implies $u(f+g) \circ \psi^{-1} = f + g'$ with $g'(x,0) = 0$ and $\deg\, g'_x > d$ as desired. Again, we have to consider not only germs but also arbitrary complex spaces as base spaces. A {\bf family of deformations} of hypersurface singularities over a base space $S\in {\cal S}$ is a morphism $\phi : {\cal X} \to S$ of complex spaces together with a section $\sigma : S \to {\cal X}$ such that for each $s \in S$ the morphism of germs $\phi : ({\cal X}, \sigma(s)) \to (S,s)$ is flat and the fibre $({\cal X}_s, \sigma(s)) = (\phi^{-1} (s), \sigma(s))$ is a hypersurface singularity. This is, of course, only a condition on the germ $({\cal X}, \sigma(S))$ of ${\cal X}$ along $\sigma(S)$. A morphism of two families $(\phi,\sigma)$ and $(\phi',\sigma')$ over $S$ is a morphism $\psi : {\cal X} \to {\cal X}'$ such that $\phi = \phi' \circ \psi$ and $\sigma' = \psi \circ \sigma$. $(\phi,\sigma)$ and $(\phi',\sigma')$ are called {\bf contact equivalent} or {\bf isomorphic families of deformations} if there exists a morphism $\psi$ such that for any $s \in S$, $\psi$ induces an isomorphism of the germs of the fibres $({\cal X}_s,\sigma(s)) \cong ({\cal X}'_s, \sigma'(s))$. Let us fix a quasihomogeneous hypersurface singularity $(X_0,0) \subset ({\Bbb C}^n,0)$ of type $(d; w_1, \ldots, w_n)$. For $S \in {\cal S}$, a {\bf family of deformations of negative weight with principal part} $(\bf X_0, 0)$ over $S$ is a family of deformations \[ S \buildrel \sigma\over\to ({\cal X}, \sigma(S)) \buildrel\phi\over\to S \] with section such that: for any $s \in S$ the fibre $({\cal X}_s, \sigma(s))$ is isomorphic to a semiquasihomogeneous hypersurface singularity with principal part $(X_0,0)$ and the germ $(S,s) \buildrel\sigma\over\to ({\cal X},\sigma(s)) \buildrel \phi\over\to (S,s)$ is a deformation of $({\cal X}_s,\sigma(s))$ of negative weight. For any morphism of base spaces $\varphi : T \to S$, the induced deformation $T \to (\varphi^\ast {\cal X},\; \varphi^\ast \sigma(T)) \to T$ is a family of deformations with negative weight and principal part $(X_0,0)$. We obtain a functor \[ \mbox{Def}^-_{X_0} : {\cal S} \to \mbox{ sets} \] which associates to $S \in {\cal S}$ the set of isomorphism classes of families of deformations of negative weight with principal part $(X_0,0)$ over $S$. The notations of {\bf fine} and {\bf coarse moduli space} for the functor Def$^-_{X_0}$ are defined in the same manner as for the functor Unf$^-_{f_0}$ in \S 1. The objects we are going to classify are elements of \begin{tabular}{lp{10cm}} Def$^-_{X_0}(pt) =$ & $\{$ isomorphism classes of complex space germs $(X,0)$ which are isomorphic to a semiquasihomogeneous hypersurface singularity with principal part $X_0\}$. \end{tabular} Again, as for Unf$^-_{f_0}$, we cannot expect to obtain fine moduli spaces in general. In order to obtain a coarse moduli space, we have to stratify $T_-$ into $G$--invariant strata on which the geometric quotient with respect to $G$ exists, where $G = \exp\, L_+ \rtimes (E_{f_0} \cdot {\Bbb C}^\ast) \subset \mbox{ Aut}(T_-)$. Once we have this, the proof is the same as for Theorem \ref{1.3}. We want to apply Theorem 4.7 from \cite{GP 2} to the action of $L_+$ on $T_-$. \begin{theorem}\label{4.1} (\cite{GP 2})\quad Let $A$ be a noetherian ${\Bbb C}$--algebra and $L_+ \subseteq \mbox{ Der}_{\Bbb C}^{nil}A$ a finite dimensional nilpotent Lie algebra. Suppose $A$ has a filtration \[ F^\bullet :\; 0 = F^{-1}(A) \subset F^0 (A) \subset F^1 (A) \subset \ldots \] by subvector spaces $F^i(A)$ such that\\ $({\bf F})\qquad\qquad\qquad\qquad\qquad \delta F^i(A) \subseteq F^{i-1} (A)\, \mbox{ for all } i \in {\Bbb Z},\; \delta \in L_+$.\\ Suppose, moreover, $L_+$ has a filtration \[ Z_\bullet : L_+ = Z_1 (L_+) \supseteq Z_2 (L_+) \supseteq \ldots \supseteq Z_e(L_+) \supseteq Z_{e+1}(L_+) = 0 \] by sub Lie algebras $Z_j(L_+)$ such that\\ $({\bf Z})\qquad\qquad\qquad\qquad\qquad\qquad [L_+, Z_j(L_+)] \subseteq Z_{j+1}(L_+)\, \mbox{for all}\, j \in {\Bbb Z}$.\\ Let $d : A \to \mbox{ Hom}_{\Bbb C} (L_+, A)$ be the differential defined by $d(a) (\delta) = \delta (a)$ and let \hbox{Spec $A = \cup U_\alpha$} be the flattening stratification of the modules \[ \mbox{Hom}_{\Bbb C} (L_+, A) / A d (F^i(A))\quad i = 1, 2,\ldots \] and \[ \mbox{Hom}_{\Bbb C} (Z_j(L_+), A) / \pi_j(A(dA))\quad j = 1, \ldots, e, \] where $\pi_j$ denotes the projection Hom$_{\Bbb C}(L_+, A) \to$ Hom$_{\Bbb C}(Z_j(L_+), A)$.\\ Then $U_\alpha$ is invariant under the action of $L_+$ and $U_\alpha \to U_\alpha / L_+$ is a geometric quotient which is a principal fibre bundle with fibre $\exp(L_+)$. Furthermore, the closure $\bar{U}_\alpha$ of $U_\alpha$ is affine, $\bar{U}_\alpha = \mbox{ Spec } A_\alpha$, and the canonical map $U_\alpha/L_+ \to \mbox{ Spec } A_\alpha^{L_+}$ is an open embedding. \end{theorem} To apply the theorem we have to construct these filtrations and interpret the corresponding stratification in terms of the Hilbert function of the Tjurina algebra.\\ There are natural filtrations $H^\bullet ({\Bbb C}\{x\})$ respectively $F^\bullet (A_-)$ on ${\Bbb C}\{x\}$ respectively $A_-$ defined as follows:\\ Let $F^i(A_-) \subseteq A_-$ be the ${\Bbb C}$--vectorspace generated by all quasihomogeneous polynomials of degree $> - (i+1)w$ and $H^i({\Bbb C}\{x\})$ be the ideal generated by all quasihomogeneous polynomials of degree $\ge i w$, where \[w := \min\{w_1, \ldots, w_n\}.\] The filtration $F^\bullet(A_-)$ has the property (${\bf F}$) because every homogeneous vector field of $L_+$ is of degree $\ge w$. We also have $A_- dA_- = A_- dF^sA_-$ with $s = \left[\frac{(n-1)d-2\sum w_i}{w}\right]$, since $nd - 2 \sum w_i$ is the degree of the Hessian of $f$ and $t_k$ is the variable of smallest degree. To define $Z_\bullet$ let $Z_i(L_+) :=$ the Lie algebra generated by the vectorfields $\delta \in L_+,\; \delta$ homogeneous and $\deg(\delta) \ge r_i$, \[ r_i := \min\{\deg(\delta_j) \mid t_{k+1-j} \in F^{s-i}(A_-)\}. \] $Z_\bullet(L_+)$ has the property $({\bf Z})$ because $\deg([\delta,\delta']) \ge \deg(\delta) + \deg(\delta')$ for all $\delta, \delta' \in L_+$. \begin{example}\label{4.2} {\rm We continue with Example \ref{3.4}, $f_0 = x^3 + y^3 + z^7$. $w = 3$.\\ $F^\circ(A_-)$ is the ${\Bbb C}$--vector space generated by $t_1, t_2, t_3, t_1^2, t_1t_2, t_2^2$.\\ $F^1(A_-)$ is the ${\Bbb C}$--vector space generated by $t_4, \{t_1^\nu t_2^\mu t_3^\lambda\}_{\nu+\mu+2\lambda \le 5}$.\\ $F^2(A_-)$ is the ${\Bbb C}$--vector space generated by $t_5, \{t_1^\nu t_2^\mu t_3^\lambda t_4\}_{\nu+\mu+2\lambda \le 3}, \{t_1^\nu t_2^\mu t_3^\lambda\}_{\nu + \mu + 2\lambda \le 8}$.\\ We have $s = 2 = \left[\frac{2 \cdot 21 - 2 \cdot 17}{3}\right]$.\\ $A_-dF^\circ(A_-) = \bigoplus\limits^3_{i=1} A_-dt_i$.\\ $A_-dF^1 (A_-) = \bigoplus\limits^4_{i=1} A_-dt_i$.\\ $A_-dF^2 (A_-) = A_-dA_-$.\\ $r_1 = 3, r_2 = 6$.\\ $L_+ = Z_1(L_+)$.\\ $Z_2(L_+)$ generated by the homogeneous vector fields $\delta \in L_+$ with $\deg(\delta) \ge 6$.\\ Especially $A_-Z_2(L_+) = \sum\limits^5_{i=3} A_- \delta_i$.\\ $Z_3(L_+) = 0$. } \end{example} We can use Theorem \ref{4.1} to obtain a geometric quotient of the action of $L_+$ on the flattening stratification defined by the filtrations $F^\bullet$ and $Z_\bullet$. Before doing this we shall prove that this flattening stratification is also the flattening stratification of the modules defining the Hilbert function of the Tjurina algebra.\\ For $t \in T_-$ the {\bf Hilbert function of the Tjurina algebra} \[ {\Bbb C}\{x\}/\left(F(t),\frac{\partial F(t)}{\partial x_1}, \ldots, \frac{\partial F(t)}{\partial x_n}\right) \] corresponding to the singularity defined by $t$ with respect to $H^\bullet$ is by definition the function \[ m \mapsto \tau_m(t) := \dim_{\Bbb C} {\Bbb C}\{x\}/\left(F(t), \frac{\partial F(t)}{\partial x_1}, \ldots, \frac{\partial F(t)}{\partial x_n}, H^m\right). \] Notice that $\tau_m(t) = \tau(t)$ if $m$ is large and $\tau_m(t)$ does not depend on $t$ for small $m$. On the other hand, $\mu_m := \mu_m(t) := \dim_{\Bbb C} {\Bbb C}\{x\}/(\frac{\partial F(t)}{\partial x_1}, \ldots, \frac{\partial F(t)}{\partial x_n}, H^m)$ does not depend on $t \in T_-$ and \[ \mu_m - \tau_m(t) = \mbox{ rank } (\delta_i(t_j) (t))_{\deg(t_j) > d- mw}. \] This is an immediate consequence of the following fact:\\ Let \[ T^m := A_-\{x\}/\left(F, \frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n}, H^m\right), \] then the following sequence is exact and splits: let $\{X^\alpha\}_{\alpha \in B}$ be a monomial base of $A_-\{x\}/\left(\frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n}\right)$. \[ \begin{array}{rcrllllll} 0 & \to & \bigoplus\limits_{\buildrel |\alpha|\le d\over \alpha \in B} A_-x^\alpha & \to & T^{\frac{d}{w} + i} & \to & \mbox{Der}_{\Bbb C} A_-/\left({\cal L} + \sum_{\deg(t_j) \le - iw} A_- \frac{\partial}{\partial t_j}\right) & \to & 0\\ & & x^\alpha & \mapsto & \mbox{class}(x^\alpha) & & & & \\ & & & &\mbox{class}(m_j) & \mapsto & \mbox{class}(\frac{\partial}{\partial t_j}), & & \end{array} \] and with the identification $\sum\limits_{\deg(t_j) > -iw} A_- \frac{\partial}{\partial t_j} \simeq A_-^{N_i}$ we obtain\\ Der$_{\Bbb C} A_-/({\cal L} + \sum\limits_{\deg(t_j) \le -iw} A_- \frac{\partial}{\partial t_j}) \simeq A_-^{N_i}/M_i$, where $M_i$ is the $A_-$--submodule generated by the rows of the matrix $(\delta_i(t_j))_{\deg(t_j) > -iw}$. We have $F \in H^m$, hence $\mu_m = \tau_m$, if $m \le \frac{d}{w}$ and $H^m \subset (\frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n})$, hence $\mu_m - \tau_m(t)$ is independent of $m$ and equal to $\mu - \tau(t)$, if $m \ge \frac{d}{w} + s + 1$ . Therefore, we have $s + 1$ relevant values for $\tau_i$, and we denote \vspace{-0.5cm} \begin{eqnarray*} \underline{\tau}(t) & := & (\tau_{\frac{d}{w} + 1}(t), \ldots, \tau_{\frac{d}{w} + s + 1}(t)),\\ \underline{\mu} & := & (\mu_{\frac{d}{w}+1}, \ldots, \mu_{\frac{d}{w}+s+1}). \end{eqnarray*} Moreover, let $\Sigma = \{\underline{r} := (r_1, \ldots, r_{s+1}) \mid \exists\, t \in T_-$ so that $\underline{\mu} - \underline{\tau}(t) = \underline{r}\}$ and $T_- = \cup_{\underline{r} \in \Sigma}U_{\underline{r}}$ be the flattening stratification of the modules $T^{\frac{d}{w}+1}, \ldots, T^{\frac{d}{w}+s+1}$. That is, $\{U_{\underline{r}}\}$ is the stratification of $T_-$ defined by fixing the Hilbert function $\underline{\tau} = \underline{\mu} - \underline{r}$ with the scheme structure defined by the flattening property. Let us now consider an arbitrary deformation $\phi : ({\cal X},\{0\} \times S) \hookrightarrow ({\Bbb C}^n,0) \times S \to S$ of $(X,0) \subset ({\Bbb C}^n,0)$ of negative weight over a base space $S \in {\cal S}$ where, for each $s \in S$, the ideal of the germ $({\cal X},(0,s)) \subset ({\Bbb C}^n \times S, (0,s))$ is defined by $F(x,s) = f(x) + g(x,s),\; g(x,0) = 0$. Let us denote by ${\cal O}_S \{x\} = {\cal O}_{{\Bbb C}^n \times S, 0 \times S}$ the topological restriction of ${\cal O}_{{\Bbb C}^n \times S}$ to $0 \times S$, considered as a sheaf on $S$. Then $J(I_{{\cal X} , 0 \times S})$, the Jacobian ideal sheaf of $({\cal X}, \{0\} \times S) \subset ({\Bbb C}^n,0) \times S$, is locally defined by $(F, \frac{\partial F}{\partial x_1}, \ldots, \frac{\partial F}{\partial x_n}) \subset {\cal O}_S \{x\}$ and $H^m_S \subset {\cal O}_S\{x\}$ is the ideal sheaf generated by $g \in {\cal O}_S \{x\}$ such that $\deg_x g \ge mw,\; w = \min\{w_1, \ldots, w_n\}$ as above. We say that the {\bf family} $\phi$ is {\bf $\underline{\tau}$--constant} if the coherent ${\cal O}_S$--sheaves \[ T^m_S := {\cal O}_S \{x\}/J(I_{{\cal X},\{0\}\times S}) + H^m_S \] are flat for $\frac{d}{w} + 1\le m \le \frac{d}{w} + s + 1$ (equivalently, for all $m$). Of course, if $T^m_S$ is flat, then \[ \tau_m(s) := \dim_{\Bbb C} T^m_{S,s} \otimes {\cal O}_{S,s}/\frak m_{S,s} \] is independent of $s \in S$. The converse holds for reduced base spaces: \begin{lemma} If $S$ is reduced, then the sheaf $T^m_S$ is flat if and only if $\tau_m(s)$ is independent of $s \in S$. \end{lemma} The proof is standard (cf.\ \cite{GP 3}). Hence, over a reduced base space $S$, $\underline{\tau}$--constant means just that the Hilbert function $\underline{\tau}(s) = (\tau_{\frac{d}{w} + 1} (s), \ldots, \tau_{\frac{d}{w} + s + 1} (s))$ of the Tjurina algebra is constant. But for arbitrary base spaces we have to require flatness of the corresponding $T^m_S$. {\bf Example} ($f_0 = x^3 + y^3 + z^7$, continued) \begin{eqnarray*} \underline{\tau}(t) & = & (\tau_8(t), \tau_9(t), \tau_{10}(t))\\ \underline{\mu} & = & ( \mu_8, \mu_9, \mu_{10}) = (22, 23, 24)\\ \Sigma & = & \{(0, 0, 0), (0, 0,1), (0, 1, 2), (1, 1, 2), (1, 2, 3)\}\\ U_{(1,2,3)} & = & D(2t_3 - \frac{10}{7} t_1 t_2) \subseteq T_- = {\Bbb C}^5\\ U_{(1, 1, 2)} & = & V(2t_3 - \frac{10}{7}t_1t_2) \cap D(t_1, t_2) \subseteq T_-\\ U_{(0, 1, 2)} & = & V(t_1, t_2, t_3) \cap D(t_4) \subseteq T_-\\ U_{(0, 0, 1)} & = & V(t_1, t_2, t_3, t_4) \cap D(t_5) \subseteq T_-\\ U_{(0, 0, 0)} & = & \{(0, 0, 0, 0, 0)\}. \end{eqnarray*} \begin{lemma}\label{4.3} \begin{enumerate} \item $(0, \ldots, 0,1)$ and $(0, \ldots, 0) \in \Sigma$. $U_{(0, \ldots,0)} = \{0\}$ is a smooth point and $U_{(0, \ldots, 1)}$ is defined by $t_1 = \cdots = t_{k-1} = 0$ and $t_k \not= 0$. \item Let $\bar{\Sigma} = \Sigma\backslash\{(0, \ldots, 0)\}$ and for $\underline{r} \in \bar{\Sigma}$ put \[ \widetilde{U}_{\underline{r}} = \left\{ \begin{array}{ll} U_{\underline{r}} & \mbox{ if } \underline{r} \not= (0, \ldots, 0, 1)\\ U_{(0, \ldots, 0, 1)} \cup U_{(0, \ldots, 0)} & \mbox{ if } \underline{r} = (0, \ldots, 0, 1). \end{array}\right. \] \end{enumerate} Then $\{\widetilde{U}_{\underline{r}}\}_{\underline{r} \in \bar{\Sigma}}$ is the flattening stratification of the modules $\{\mbox{Hom}_{\Bbb C}(L_+, A_-)/A_- dF^i A_-\}$ and $\{\mbox{Hom}_{\Bbb C}(Z_i(L_+), A_-)/\pi_i(A_-dA_-)\}$. \end{lemma} {\bf Proof of Lemma \ref{4.3}}: Because of the exact sequence above the flattening stratification of the modules $\{T^{\frac{d}{w}+i}\}$ is also the flattening stratification of $\{\mbox{Der}_{\Bbb C} A_-/({\cal L} + \sum_{\deg(t_j) \le -iw} A_- \frac{\partial}{\partial t_j})\}$ respectively the flattening stratification of $\{A_-^{N_i}/M_i\},\; M_i$ the submodule generated by the rows of the matrix $(\delta_i(t_j))_{\deg(t_j) > -iw}$. Now we have $(\ast)$\hspace{4.5cm}$\delta_i(t_j) = \delta_{k-j+1}(t_{k-i+1})$. By definition of $Z_i(L_+)$ we have \[ A_-Z_i(L_+) = \sum_{t_{k+1-j} \in F^{s-i}} A_- \delta_j \] and with the identification \[ \sum\limits A_- \frac{\partial}{\partial t_j} = A_-^k, \] and $M^i$ the submodule generated by the rows of the matrix $(\delta_\ell(t_j))_{\ell \ge r_i}$ we obtain \[ \mbox{Der}_{\Bbb C} A_- / A_- Z_i(L_+) \cong A_-^k/M^i. \] (*) implies that the flattening stratification of the modules $\{T^{\frac{d}{w}+1}, \ldots, T^{\frac{d}{w}+s}\}$, which is $T_- = \cup_{\underline{r} \in \bar{\Sigma}} \widetilde{U}_{\underline{r}}$, is the flattening stratification of the modules $\{\mbox{Der}_{\Bbb C} A_-/A_- Z_i(L_+)\}_{i=1, \ldots, s}$. Furthermore the modules $\{\mbox{Hom}_{\Bbb C}(L_+, A_-)/A_-dF^iA_-\}$ and\\ $\{\mbox{Der}_{\Bbb C} A_-/A_- L_+ + \sum_{\deg(t_j)\le -iw} A_-\frac{\partial}{\partial t_j}\}$ have the same flattening stratification and they are flat on $U_{\underline{r}}$, because \[ 0 \to A_- \to \mbox{ Der}_{\Bbb C} A_-/A_-L_+ + \sum_{\deg(t_j)\le -iw} A_-\frac{\partial}{\partial t_j} \to \mbox{ Der}_{\Bbb C} A_- /{\cal L} + \sum_{\deg(t_j)\le -iw} A_-\frac{\partial}{\partial t_j} \to 0 \] is exact and splits on $T_-\backslash\{0\}$. This proves the lemma. \begin{remark}\label{4.4}{\rm The main point of the lemma is that the flattening stratification of the modules $\{\mbox{Hom}_{\Bbb C}(L_+, A_-)/A_- dF^iA_-\}$ is equal to the flattening stratification of the modules $\{\mbox{Hom}_{\Bbb C}(Z_i(L_+), A_-)/\pi_i(A_- d A_-)\}$, hence, is defined by the Hilbert function of the Tjurina algebra alone, without any reference to the action of $L$. This is a consequence of the symmetry expressed in Proposition \ref{3.2}.} \end{remark} As a corollary we obtain the following \begin{theorem}\label{4.5} For $\underline{r} \in \Sigma,\; \widetilde{U}_{\underline{r}}$ is invariant under the action of $L_+$. Let Spec $A_{\underline{r}}$ be the closure of $\widetilde{U}_{\underline{r}}$ then $\widetilde U_{\underline{r}} \to \widetilde{U}_{\underline{r}}/L_+$ is a geometric quotient contained in Spec $A^{L_+}_{\underline{r}}$ as an open subscheme of Spec $A^{L_+}_{\underline{r}}$. \end{theorem} {\bf Example} ($f_0 = x^3 + y^3 + z^7$, continued) \[ \begin{array}{lccc} 1) & \widetilde{U}_{(1, 2, 3)} = D(2t_3 - \frac{10}{7} t_1 t_2) & \longrightarrow & \widetilde{U}_{(1, 2, 3)}/L_+ = \mbox{ Spec}\, {\Bbb C}[t_1, t_2, t_3]_{2t_3-\frac{10}{7} t_1t_2}\\[1.0ex] & \bigcap\mid & & \bigcap\mid\\[1.0ex] & \mbox{Spec }{\Bbb C}[t_1, \ldots, t_5] & \longrightarrow & \mbox{Spec }{\Bbb C}[t_1, t_2, t_3]\\[2.0ex] 2) & \widetilde{U}_{(1, 1, 2)} & \longrightarrow & \widetilde{U}_{(1, 1, 2)}/L_+ = D(t_1, t_2)\\[1.0ex] & \bigcap\mid & & \bigcap\mid\\[1.0ex] & \mbox{Spec }{\Bbb C}[t_1, t_2, t_4, t_5] & \longrightarrow & \mbox{Spec }{\Bbb C}[t_1, t_2, t_4]\\[1.0ex] \multicolumn{4}{l}{(\mbox{identifiying } {\Bbb C}[t_1, \ldots, t_5]/2t_3 - \frac{10}{7} t_1t_2 = {\Bbb C}[t_1, t_2, t_4, t_5].)}\\[2.0ex] 3) & \widetilde{U}_{(0,1,2)} & \longrightarrow & \widetilde{U}_{(0,1,2)}/L_+ = D(t_4)\\[1.0ex] & \bigcap\mid & & \bigcap\mid\\[1.0ex] & \mbox{Spec }{\Bbb C}[t_4, t_5] & \longrightarrow & \mbox{Spec }{\Bbb C}[t_4]\\[2.0ex] 4) & \widetilde{U}_{(0,0,1)} & = & \widetilde{U}_{(0,0,1)}/L_+\\[1.0ex] & \| & & \|\\[1.0ex] & \mbox{Spec }{\Bbb C}[t_5] & = & \mbox{Spec }{\Bbb C}[t_5] \end{array} \] \vspace{1cm} Now $L/L_+ \simeq {\Bbb C} \delta_1$ acts on the geometric quotients $\widetilde{U}_{\underline{r}}/L_+$ (the ${\Bbb C}^\ast$--action defined by the Euler vector field $\delta_1$). Also the group $E_{f_0}$ acts and this action commutes with the ${\Bbb C}^\ast$--action (cf.\ \ref{2.5}). If we combine this fact with Theorem 4.6 we obtain the main theorem of this article. In order to formulate it properly let us denote by \[ \mbox{Def}^-_{X_0, \underline{\tau}} : {\cal S} \to \mbox{ sets} \] the subfunctor of Def$^-_{X_0}$ which associates to a base space $S \in {\cal S}$ the set of isomorphism classes of $\underline{\tau}$--constant families of deformations of negative weight with principal part $(X_0,0)$ over $S$. For such a family $\underline{\tau}(s)$ is constant and equal to some tuple $\underline{\mu} - \underline{r} \in {\Bbb N}^{s+1}$. \begin{theorem}\label{4.6} Let $G = \exp L_+ \rtimes (E_{f_0} \cdot {\Bbb C}^\ast) \subseteq \mbox{ Aut }(T_-)$. \begin{enumerate} \item The orbits of $G$ are unions of finitely many integral manifolds of ${\cal L}$. \item Let $T_- = \cup_{\underline{r}\in \Sigma} U_{\underline{r}}$ be the stratification fixing the Hilbert function $\underline{\tau}$ of the Tjurina algebra described above. $U_{\underline{r}}$ is invariant under the action of $G$ and the geometric quotient $U_{\underline{r}} \to U_{\underline{r}}/G$ exists and is locally closed in a weighted projective space. \item $U_{\underline{r}}/G$ is the coarse moduli space for the functor Def$^-_{X_0, \underline{\tau}} : $ complex spaces $\to$ sets with $\underline{\tau} = \underline{\mu} - \underline{r}$. \end{enumerate} \end{theorem} \begin{remark}{\rm As in the case of right equivalence (see Remark 1.5) we may take (separated) algebraic spaces as category of base spaces. That is, $U_{\underline{r}}/G$ is a coarse moduli space for the functor \[ \mbox{Def}^-_{X_0, \underline{\tau}} : \mbox{ algebraic spaces }\to \mbox{ sets}. \] } \end{remark} {\bf Proof} (of Theorem 4.7): We first prove that $U_{\underline{r}}$ is invariant under the action of $G$ and that $U_{\underline{r}} \to U_{\underline{r}}/G$ is a geometric quotient. To prove that $U_{\underline{r}}$ is invariant under the action of $G$ it is enough by definition of $U_{\underline{r}}$ that it is invariant under the action of $E_{f_0}$. The Hilbert function $\underline{\tau}$ of the Tjurina algebra is invariant under contact equivalence. This is a consequence of Theorem \ref{2.1} because an automorphism $\varphi$ of ${\Bbb C}\{x\}$ inducing the isomorphy of two semiquasihomogeneous singularities with principal part $f_0$ has degree $\ge 0$. More precisely, let $f,g$ be semiquasihomogeneous with principal part $f_0$ and $uf = \varphi(g)$ for a unit $u$ then $\deg(\varphi) \ge 0$ and consequently $(f, \frac{\partial f}{\partial x_1}, \ldots, \frac{\partial f}{\partial x_n}, H^m)$ is mapped isomorphically to $(g, \frac{\partial g} {\partial x_1}, \ldots, \frac{\partial g}{\partial x_n}, H^m)$ for all $m$, in particular $\underline{\tau}(f) = \underline{\tau}(g)$. Moreover, let $\sigma \in E_{f_0}$, then there is a $\varphi : A_-\{x\} \to A_-\{x\},\; \deg_x(\varphi) \ge 0$ and $\varphi|_{A_-} = id_{A_-}$ such that \[\varphi(F(x,t)) \equiv F(x,\sigma (t))\hbox{ mod } A_- H^N \hbox{ \ for sufficiently large }N\] (cf. proof of Proposition 2.4). This implies $\sigma(T^m) = T^m$ for all $m$ and proves that $E_{f_0}$ and, therefore, $G$ acts on the strata $U_{\underline{r}}$ of the flattening stratification of the modules $\{T^m\}$. Now we prove that $U_{\underline{r}} \to U_{\underline{r}}/G$ is a geometric quotient. First of all it is obvious that the geometric quotients \[ U_{(0, \ldots, 0,1)} \to U_{(0, \ldots, 0,1)}/G = \{\ast\} \] and \[ U_{(0, \ldots, 0)} = \{\ast\} = U_{(0, \ldots, 0)}/G = \{\ast\} \] exist. Let $\underline{r} \not= (0, \ldots, 0,1),\; (0,\ldots, 0)$ then $\widetilde{U}_{\underline{r}} = U_{\underline{r}}$. Let $U_{\le\underline{r}} = \mbox{ Spec}A_{\underline{r}}$ be the closure of $U_{\underline{r}}$ then we obtain \[ \begin{array}{ccc} \mbox{Spec}A_{\underline{r}} & \buildrel \pi\over\longrightarrow & \mbox{Spec}A_{\underline{r}}^{L_+}\\[1.0ex] \cup|\; i & & \cup|\; j\\[1.0ex] U_{\underline{r}} & \buildrel\pi|_{U_{\underline{r}}}\over\longrightarrow & U_{\underline{r}}/L_+. \end{array} \] $\pi|_{U_{\underline{r}}}$ defines a geometric quotient and $i,j$ are open embeddings (Theorem \ref{4.5}). Notice that $\pi$ itself is not necessarily a geometric quotient. Now Spec$A_{\underline{r}}^{L_+}$ is affine and $E_{f_0}$ acts on Spec$A_{\underline{r}}^{L_+}$ and also on $U_{\underline{r}}/L_+$. This implies (cf.\ \cite{MF}) that \[ \mbox{Spec}A_{\underline{r}}^{L_+} \buildrel\lambda\over\to \mbox{ Spec}(A_{\underline{r}}^{L_+})^{E_{f_0}} \] is a geometric quotient (not necessarily as algebraic schemes since $A^{L_+}_{\underline{r}}$ need not be of finite type over ${\Bbb C}$) and consequently \[ \lambda|_{U_{\underline{r}}/L_+} : U_{\underline{r}}/L_+ \to (U_{\underline{r}}/L_+)/E_{f_0} \] is a geometric quotient which is an algebraic scheme. Especially $(U_{\underline{r}}/L_+)/E_{f_0} \subseteq \mbox{ Spec}(A_{\underline{r}}^{L_+})^{E_{f_0}}$ is an open subset. Finally, ${\Bbb C}^\ast$ acts on Spec$(A_{\underline{r}}^{L_+})^{E_{f_0}}$. It has one fixed point $\{\ast\}$ corresponding to $U_{(0, \ldots, 0)} \subseteq \bar{U}_r =$ Spec$A_{\underline{r}}$. Outside this fixed point the ${\Bbb C}^\ast$--action leads to a geometric quotient: \[ \begin{array}{ccc} \mbox{Spec}(A_{\underline{r}}^{L_+})^{E_{f_0}}\backslash\{\ast\} & \longrightarrow & \mbox{Proj}(A_{\underline{r}}^{L_+})^{E_{f_0}}\\[1.0ex] \cup & & \cup\\[1.0ex] (U_{\underline{r}}/L_+)/E_{f_0} & \longrightarrow & ((U_{\underline{r}}/L_+)/E_{f_0})/{\Bbb C}^\ast\\[1.0ex] & & \|\\[1.0ex] & & U_{\underline{r}}/G. \end{array} \] This proves part (1) and (2) of the theorem. It remains to prove that if $t,t' \in T_-$ define isomorphic singularities then $t$ and $t'$ are in the same orbit of $G$. Let $F_t = u\varphi(F_{t'})$ for $t, t' \in T_-, u \in {\Bbb C}\{x\}^\ast$ a unit and $\varphi$ an automorphism of ${\Bbb C}\{x\}$. Using the ${\Bbb C}^\ast$--action we find $t''\in T_-,\ u_1 =\frac{u}{u(0)}\in {\Bbb C}\{x\}^\ast$ and an automorphism $\varphi_1$ of ${\Bbb C}\{x\}$ such that $F_t = u_1\varphi_1 (F_{t''}),\ u_1(0)=1$ and $t'$ and $t''$ are in one ${\Bbb C}^\ast$--orbit. Then \[ G(z) := (1+z(u_1-1))\varphi_1(F_{t''}) \] is an unfolding of $G(0) = F_t$ of negative weight. This unfolding can be induced by the semiuniversal unfolding, that is there exists a family of coordinate transformations $\underline{\psi}(z, -)$ and a path $v$ in $T_-$ such that \[ G(z) = F(\psi_1(z,x), \ldots, \psi_n(z,x), v(z)) \] and $v(0) = t$ and $F_{t''} \buildrel r\over\sim F(\psi(1,x), v(1))$. Now $t=v(0)$ and $v(1)$ are in one orbit of $\exp L$, and $v(1)$ and $t''$ are in one orbit of $E_{f_0}$. Hence the result. Now (3) follows in the same manner as the proof of Theorem 1.3. {\bf Example} ($f_0 = x^3 + y^3 + z^7$, continued) \begin{enumerate} \item $U_{(1,2,3)} \longrightarrow U_{(1,2,3)}/G \simeq {\Bbb C}^2,\; \underline{\tau} = (21, 21, 21),\; \tau = 21$ \\ normal form: $f_0 + t_1 xz^5 + t_2 yz^5 + t_3xyz^3,\\ (t_1:t_2:t_3) \in D_+ (2t_3 - \frac{10}{7} t_1t_2)/S_3 \subset \P^2_{(1:1:2)}/S_3$\\ $(D_+(2t_3 - \frac{10}{7} t_1t_2)/S_3 \simeq {\Bbb C}^2$, the $S_3$--action being explained in Example 2.8). \item $U_{(1,1,2)} \longrightarrow U_{(1,1,2)}/G \simeq \P^2_{(2,3,5)}\backslash(0:0:1),\; \underline{\tau} = (21, 22, 22),\; \tau = 22$\\ normal form: $f_0 + t_1xz^5+t_2yz^5+\frac{10}{7} t_1t_2xyz^3 +t_4 xyz^4$,\\ $(t_1:t_2:t_4) \in \P^2_{(1:1:5)}/S_3 \;\; (\simeq \P^2_{(2,3,5)})$ \item $U_{(0,1,2)} \longrightarrow U_{(0,1,2,)}/G = \{\ast\},\; \underline{\tau} = (22, 22, 22),\; \tau = 22$\\ normal form: $f_0 + xyz^4$ \item $U_{(0,0,1)} \longrightarrow U_{(0,0,1,)}/G = \{\ast\},\; \underline{\tau} = (22, 23, 23),\, \tau = 23$\\ normal form: $f_0 + xyz^5$ \item $U_{(0,0,0)} \longrightarrow U_{(0,0,0,)}/G = \{\ast\},\; \underline{\tau} = (22, 23, 24),\; \tau = 24$\\ normal form: $f_0$ \end{enumerate} Hence the moduli space of semiquasihomogeneous hypersurface singularities $X = \{(x,y,z) \mid f (x,y,z) = 0\}$ with principal part $X_0 = \{(x,y,z) \mid x^3 + y^3 + z^7 = 0\}$ consists of 5 strata (${\Bbb C}^2,\; \P^2_{(2,3,5)} \backslash (0 : 0 : 1)$, and 3 isolated points) corresponding to 5 possible Hilbert functions $\underline{\tau}$ of the Tjurina algebra ${\Bbb C}\{x,y,z\}/(f, \frac{\partial f}{\partial x}, \frac{\partial f}{\partial y}, \frac{\partial f}{\partial z})$. The generic stratum $U_{(1,2,3)}$ (minimal $\underline{\tau}$) is an open subset in ${\Bbb C}^5$, the quotient being 2--dimensional, as well as the quotient of the 4--dimensional ``subgeneric'' stratum $U_{(1,1,2)}$. Note that the families of normal forms are not universal. It just means that each semiquasihomogeneous singularity with principal part $f_0$ occurs and that different parameters do not give contact equivalent singularities, except modulo the ${\Bbb C}^\ast$-- and $S_3$--action. We see that $U_{(1,1,2)}/G$ can be compactified by $U_{(0,1,2)}/G$, that is \[ U_{(1,1,2)} \cup U_{(0,1,2)} \to (U_{(1,1,2)} \cup U_{(0,1,2)})/G = \P^2_{(2,3,5)} \] is a geometric quotient. So in this example there exist geometric quotients of the strata with constant Tjurina number and, hence, a coarse moduli space for fixed principal part and fixed Tjurina number. In general this is false (cf.\ \cite{LP}, \S 7). \begin{remark}{\rm 1.\quad The generic stratum $U_{\underline{\tau}\min}$ corresponding to minimal Hilbert function $\underline{\tau}$ (with respect to lexicographical ordering) is an open, quasiaffine subset of $T_-$ and, hence, $U_{\underline{\tau}\min}/L_+$ is smooth by Theorem 4.1. In particular, the generic moduli space $U_{\underline{\tau}\min}/G$ has, at most, quotient singularities (coming from the ${\Bbb C}^\ast$--action and the finite group $E_{f_0}$). It is not known whether the bigger stratum $U_{\tau\min}$ corresponding to minimal Tjurina number $\tau$ admits a geometric quotient, except for $n = 2$ (cf.\ \cite{LP}). 2.\quad We always have two special strata, the most special $U_{(0, \ldots,0)} = \{\ast\}$ (corresponding to $f_0$) and the ``subspecial'' $U_{(0, \ldots, 1)} \cong {\Bbb C}\backslash \{\ast\}$ (corresponding to the singularity $f_0 + m_k,\;\; m_k$ generating the socle of ${\Bbb C}\{x\}/j(f_0)$, that is the monomial of maximal degree). The $G$--quotients of these strata give two reduced, isolated points. 3.\quad As we have seen for $x^3 + y^3 + z^7$, the finite group $E_{f_0}$ need not be abelian. If $f_0 = x_1^{a_1} + \cdots + x_n^{a_n}$ is of Brieskorn--Pham type and gcd$(a_i, a_j) = 1$ for $i \not= j$, then $E_{f_0} \cong \mu_d$, the group of $d$'th roots of unity, $d = \deg\, f_0$. 4.\quad Note that a coarse moduli space is more than just a bijection between its points and the corresponding set of isomorphism classes. For instance, let $U_{\underline{r}}/G$ be affine and let $S \buildrel\sigma\over\to ({\cal X}, \sigma(S)) \buildrel\phi\over\to S$ be a family of deformations from Def$^-_{X_0} (S)$ with $\underline{\tau}({\cal X}_s, \sigma(s)) = \underline{\mu} - \underline{r}$. If $S$ is compact then $\phi$ must be locally trivial since any morphism from $S$ to $U_{\underline{r}}/G$ maps $S$ onto finitely many points.} \end{remark} \newpage \addcontentsline{toc}{section}{References}
1996-02-14T06:20:08
9503
alg-geom/9503010
en
https://arxiv.org/abs/alg-geom/9503010
[ "alg-geom", "hep-th", "math.AG", "nlin.SI", "solv-int" ]
alg-geom/9503010
Benjamin Enriquez
B. Enriquez and V. Rubtsov
Hitchin systems, higher Gaudin operators and $r$-matrices
null
null
null
null
null
We adapt Hitchin's integrable systems to the case of a punctured curve. In the case of $\CC P^{1}$ and $SL_{n}$-bundles, they are equivalent to systems studied by Garnier. The corresponding quantum systems were identified by B. Feigin, E. Frenkel and N. Reshetikhin with Gaudin systems. We give a formula for the higher Gaudin operators, using results of R. Goodman and N. Wallach on the center of the enveloping algebras of affine algebras at the critical level. Finally we construct a dynamical $r$-matrix for Hitchin systems for a punctured elliptic curve, and $GL_{n}$-bundles, and (for $n=2$) the corresponding quantum system.
[ { "version": "v1", "created": "Mon, 20 Mar 1995 20:46:36 GMT" }, { "version": "v2", "created": "Tue, 13 Feb 1996 11:35:19 GMT" } ]
2015-06-30T00:00:00
[ [ "Enriquez", "B.", "" ], [ "Rubtsov", "V.", "" ] ]
alg-geom
\section{Introduction.}{} In [13], N. Hitchin introduced a class of integrable systems, attached to a complex curve $X$ and a semisimple Lie group $G$. The problem of quantization of these systems was then connected by A. Beilinson and V. Drinfel'd to the Langlands program. This quantization makes use of differential operators on the moduli space of $G$-bundles on $X$, obtained from the action of the center of the local completion of the enveloping algebra of a Kac-Moody algebra, at the critical level. This program can also be carried out in the case of curves with marked points. In the particular case of the punctured ${\bf C} P^{1}$, the action of the center of the enveloping algebra was studied by B. Feigin, E. Frenkel and N. Reshetikhin in [6]; they obtained an integrable system whose first operators are identical to Gaudin's operators ([9]). In this paper, we consider the question of expressing the action of higher central elements. We first note, that the Adler-Kostant-Symes (AKS) scheme, which allows to write families of commuting operators ([2], [14], [21]), can be applied in the present situation, and then show that the higher Hamiltonians obtained in [6], coincide with those. So our problem turns out to be equivalent to expressing higher central elements in the enveloping algebras at critical level, a problem which was solved by several authors ([10], [12]). Here we show how the results of [10] can be used to derive a simple expression of higher Gaudin Hamiltonians. We then turn to the case of punctured elliptic curves. We show that the integrability of Hitchin's system can be deduced from an $r$-matrix argument. Here $r$-matrix relations contain additional terms, due to an invariance under the Cartan algebra action. The $r$-matrix depends on the moduli parameters, so it reminds dynamical $r$-matrices. In the case of one puncture, our $L$-operator and $r$-matrix seem connected with those considered respectively by I. Krichever and A. Gorsky and N. Nekrasov in [15] and [11], and H. Braden, T. Suzuki and E. Sklyanin [5], [19]. It is also analogous to the $r$-matrix appearing in the work of G. Felder and C. Wieczerkowski on the Knizhnik-Zamolodchikov-Bernard equation on elliptic curves ([7]). We give the form of the first Hamiltonians in this case; one of them contains a Weierstass potential, and so is analogous to the Calogero-Moser Hamiltonian. We compute the corresponding quantum Hamiltonians, in the case $G=GL_{2}$. We would like to thank V. Drinfel'd, B. Feigin, G. Felder, E. Frenkel, A. Gorsky, N. Nekrasov, A. Reyman, and A. Stoyanovsky for discussions connected with the subject of this work. We are thankful to A. Beilinson and V. Drinfel'd for sending us their paper [4]. The work of V.R. was supported by the CNRS, and partially by grant RFFI 95-01-01101; he is grateful to the Centre de Math\'ematiques de l'\'Ecole Polytechnique, where this work was done, for hospitality. \section{1.}{Hitchin and Beilinson-Drinfeld systems in the case of a general punctured curve.} \medskip\noindent \it 1.1. Hitchin systems. \rm Let $\overline X$ be a smooth compact complex curve, $x_{i}$ ($i=1,\cdots,N$) be distinct points on $\overline X$. Set $X=\overline X-\{x_{i}\}$. Let $G$ be a reductive complex Lie group, $B\subset G$ and $T\subset B$ Borel and Cartan subgroups of $G$; let ${\bf g}$, ${\bf b}$ and ${\bf t}$ be their Lie algebras. Let ${\cal M}_{G}(X)$ be the moduli space of principal $G$-bundles on $\overline X$ with choices of a $B$-orbit in each fibre over $x_{i}$. Let us identify ${\bf g}$ with its dual, using a non-degenerate invariant form $\langle, \rangle_{{\bf g}}$. Let $P \in{\cal M}_{G}(X)$, then $T^{*}_{P}{\cal M}_{G}(X)$ is formed of the $\xi\in H^{0}(\overline X,\Omega_{\overline X}(\sum_{i=1}^{N}(x_{i}))\otimes {\bf g}_{P})$, such that $\xi$ has the expansion near $x_{i}$, $\xi= {1\over{u_{i}}}\xi_{i}+ {\rm regular}$, and $\xi_{i}\in ({\bf b}_{x_{i}})^{\perp}$; ${\bf b}_{x_{i}}$ is the subspace of the fibre of ${\bf g}_{P}$ at $x_{i}$, corresponding to the $B$-orbit at $P_{x_{i}}$, $u_{i}$ is a local coordinate at $x_{i}$. The Hitchin map $$ H : T^{*}{\cal M}_{G}(X)\to {\cal H}_{X}= \bigoplus_{i=1}^{r}H^{0}\big(\overline X,\Omega_{\overline X}^{d_{i}} ((d_{i}-1)\sum_{l=1}^{N}(x_{l}))\big), $$ is then defined by $(H(P,\xi))_{l}= P_{d_{l}}(\xi)$; $r$ is the rank of $G$, $d_{l}$ ($1\le l\le r$) are the characteristic degrees of ${\bf g}$ and $P_{d_{l}}$ corresponding invariant polynomials. The generic fiber of the natural projection ${\cal M}_{G}(X)\to{\cal M}_{G}(\overline X)$ is $(G/B)^{N}$ if genus$(\overline X)>1$, the generic bundle having no automorphisms; on the other hand, we have for genus$(\overline X)>0$, $\dimm {\cal H}_{X} =\dimm{\cal H}_{\overline X}+\sum_{l=1}^{r}(d_{l}-1)N =\dimm{\cal H}_{X}+N(\dimm B-r)$. If genus$(\overline X)=1$, an open subset of ${\cal M}_{G}(X)$ is identified with $T/W$ ($W$ is the Weyl group of $G$) if $N=0$, and with $T\rtimes W\setminus [T\times(G/B)^{N}]$ for $N\ge 1$ (only $W$ acts in the first factor, and $T\rtimes W$ acts diagonally on $(G/B)^{N}$); on the other hand, $\dimm{\cal H}_{X}=\sum_{i=1}^{r}(d_{i}-1)N$ if $N\ge 1$, and $r$ if $N=0$. If genus$(\overline X)=0$ and $N\ge 3$, an open subset of ${\cal M}_{G}(X)$ is identified with $G\setminus(G/B)^{N}$, whereas $\dimm {\cal H}_{X}=\sum_{l=1}^{r} [(d_{l}-1)(N-2)-1]$. The cases $N\le 2$ give trivial moduli spaces and ${\cal H}_{X}$. So in all cases $$ \dimm{\cal M}_{G}(X)= \dimm{\cal H}_{X}. $$ We can see as in [13] that the functions on $T^{*}{\cal M}_{G}(X)$, defined by the coordinates of $H$, Poisson commute. Moreover, for $G=GL_{n}({\bf C})$ we can consider the spectral cover of $\overline X$, defined as $\{(x,\lambda)| \lambda^{n}+ \sum_{l\ge 1}H_{i}\lambda^{n-l} = 0\}$, for $(H_{l})\in {\cal H}$ fixed, in the total space of $\Omega_{\overline X}(\sum_{i=1}^{N}(x_{i}))$; it has ramification of order $n$ at the points $x_{i}$, in the generic situation. It is possible to build a line bundle over the spectral cover, and to study the integrability of the system as in [13]. \medskip\noindent \it 1.2. Beilinson-Drinfeld systems. \rm To quantize the Hitchin systems, Beilinson and Drinfeld ([4]) define $\dimm{\cal M}_{G}(X)$ commuting differential operators on ${\cal M}_{G}(X)$, with symbols the coordinates of the map $H$ (here we assume no marked points). They are constructed as follows: a base point $x$ on $X$ being fixed, ${\cal M}_{G}(X)$ is identified with $G({\cal O}_{x})\setminus G(k_{x})/G(A)$ (Siegel-Weil); ${\cal O}_{x}$ and $k_{x}$ are respectively the local ring and field at $x$, and $A=H^{0}(X-\{x\}, {\cal O}_{X})$. Then the center $Z(U_{-h^{\vee}}({\bf g}(k_{x}))_{loc})$ of $U_{-h^{\vee}}({\bf g}(k_{x}))_{loc}$ (local completion of the enveloping algebra of the critical level extension of ${\bf g}(k_{x})$) acts by differential operators on the line bundle $(\det)^{-h^{\vee}}$ over ${\cal M}_{G}(X)$. This procedure can easily be extended to the punctured case: remark that ${\cal M}_{G}(X)= G({\cal O}_{x})\setminus G(k_{x})/\Gamma$, where $\Gamma\subset G(A)$ is formed of the regular maps from $\overline X-\{x\}$ to $G$, taking values in $B$ at points $x_{i}$. Let $(\lambda_{i})_{1\le i\le N}$ be a system of dominant weights of $G$. We define a line bundle ${\cal L}_{(\lambda_{i})}$ on ${\cal M}_{G}(X)$ as follows: $(\lambda_{i})$ defines a character of $\Gamma$ (by the maps $\Gamma\to B^{N}\to T^{N}$) and so a line bundle ${\cal L}'_{(\lambda_{i})}$ on $G(k_{x})/\Gamma$, then ${\cal L}'_{(\lambda_{i})}\otimes (\det)^{-h^{\vee}}$ has a natural action of $G({\cal O}_{x})$; ${\cal L}_{(\lambda_{i})}$ is then the quotient bundle. The center of $U_{-h^{\vee}}({\bf g}(k_{x}))_{loc}$ then acts on this bundle by differential operators as before. \section{2.}{Hitchin systems in the rational case.} In this section and the following, we set $\overline X={\bf C} P^{1}$, and denote by $z_{i}$ the coordinate of the marked point $x_{i}$ ($i=1,...,N$); we assume that no $x_{i}$ coincides with $\infty$. We will express the corresponding Hitchin systems, and recall an $r$-matrix result of Semenov about them. An open subset of ${\cal M}_{G}(X)$ is formed by parabolic structures on the trivial bundle; this subset, that we call ${\cal M}_{G}^{(0)}(X) $ is isomorphic to $G\setminus (G/B)^{N}$ [the left action of $G$ is diagonal]. Recall the Springer resolution $T^{*}(G/B)\to{\cal N}$, ${\cal N}$ the nilpotent cone of ${\bf g}$ ([20]). Then we construct, by reduction, the resolution $$ T^{*}[G\setminus (G/B)^{N}]\to\{(\eta^{(i)})\in{\cal N}^{N}|\sum_{i=1}^{N}\eta^{(i)}=0\}/G $$ (the action of $G$ on the last term is by conjugation). When the $\eta^{(i)}$ are regular, the parabolic structure corresponding to $(\eta^{(i)})_{i=1,\cdots,N}$ is $(g_{i}B)_{i=1,\cdots,N}$, where $g_{i}$ are elements of $G$ conjugating $\eta^{(i)}$ to elements of ${\bf b}\subset {\bf g}$. The $1$-form $\xi$ is then $$ \xi=\sum_{i=1}^{N}{\eta^{(i)}\over{z-z_{i}}}dz.\leqno(1) $$ The Poisson structure on $T^{*}{\cal M}_{G}^{(0)}(X)$ corresponds, in terms of the $(\eta^{(i)})$, to the product of Kostant-Kirillov structures on each ${\cal N}$. In tensor notation: $\{ \eta^{(i)}\otimes_{,}\eta^{(j)}\}=\delta_{ij}[P,1\otimes\eta^{(j)}] =-\delta_{ij}[P,\eta^{(i)}\otimes 1]$, $P$ the permutation operator. We deduce from this: $$ \{\eta(z)\otimes_{,}\eta(w)\}=[{P\over{z-w}},\eta(z)\otimes 1+1\otimes \eta(w)],\leqno(2) $$ where $\eta(z)=\sum_{i=1}^{N}{\eta^{(i)}\over{z-z_{i}}}$. So we have: \proclaim{Proposition 2.1} ([18]) Let us endow ${\cal N}^{N}$ with the product of Kostant-Kirillov structures on each factor. Then the mapping ${\cal N}^{N}\to {\bf g}[[z^{-1}]]$, $(\eta^{(i)})_{1\le i\le N} \mapsto \eta(z)=\sum_{i=1}^{N}{\eta^{(i)}\over{z-z_{i}}}$, is Poisson. \endgroup\par\medbreak We deduce from this that the coefficients of the forms $P_{d_{i}}(\eta(z))$ are in involution, because all the central functions on ${\bf g}[[z^{-1}]]$ are in involution. (This gives another proof of involutivity of Hitchin's Hamiltonians.) Let us give now the expression of the corresponding flows: \proclaim{Proposition 2.2} Decompose $P_{d_{l}}(\eta(z))$ as $\sum_{a_{1}+\cdots+a_{N}=d_{l}-1} {H_{d_{l},(a_{i})}\over{ \prod_{i=1}^{N}(z-z_{i})^{a_{i}} }}(dz)^{d_{l}}$, and denote by $\pr_{d_{l},(a_{i})}$ the flow generated by $H_{d_{l},(a_{i})}$. Then we have the identity of rational functions in $z$ $$ \sum_{a_{1}+\cdots+a_{N}=d_{l}-1} {\pr_{d_{l},(a_{j})}(\eta^{(i)}) \over{\prod_{j=1}^{N}(z-z_{j})^{a_{j}}}} = [P'_{d_{l}}(\eta(z)),{\eta^{(i)}\over{z-z_{i}}}]. \leqno(3) $$ For ${\bf g}=sl_{n}({\bf C})$, the r.h.s. is $[d_{l}(\sum_{j=1}^{N}{\eta^{(j)}\over{z-z_{j}}})^{d_{l}-1},\eta^{(i)}]$. For ${\bf g}$ arbitrary, the flows corresponding to $d_{1}=2$ are $$ \pr_{i}\eta^{(j)}=-{[\eta^{(i)},\eta^{(j)}]\over{z_{i}-z_{j}}} {\rm \ \ for \ \ }j\ne i, {\rm \ \ and \ \ } \pr_{i}\eta^{(i)}=\sum_{j\ne i}{[\eta^{(i)},\eta^{(j)}] \over{z_{i}-z_{j}}}. \leqno(4) $$ \endgroup\par\medbreak We note that in the case $g=sl_{n}({\bf C})$, the flows $\pr_{i}$ already appeared in [8] (we thank J. Harnad for pointing out this reference to us). Their integration was studied by many authors (cf. e.g. [1], [3]). \section{3.}{Gaudin systems.} \medskip\noindent \it 1. The moduli stack in the rational case. \rm Let again ${\bf g}$ be an arbitrary reductive complex Lie algebra, $G$ be the adjoint group. Let $\Delta$ be the set of the roots of ${\bf g}$ w.r.t. ${\bf t}$, ${\bf g}={\bf t}\oplus \bigoplus_{\alpha\in\Delta}{\bf g}_{\alpha}$ the associated decomposition of ${\bf g}$, $R$ be the root lattice of ${\bf g}$, $W$ the Weyl group of ${\bf g}$. Classes of principal $G$-bundles on $\overline X={\bf C} P^{1}$ are parametrized by $\mathop{\rm Hom}\limits(R,{\bf Z})/W$; to $\varpi\in\mathop{\rm Hom}\limits(R,{\bf Z})$ we associate the Lie algebra bundle on ${\bf C} P^{1}$ $$ {\bf g}(\varpi)={\bf t} \oplus\bigoplus_{\alpha\in\Delta}{\bf g}_{\alpha}(\varpi(\alpha)\infty),\leqno(5) $$ and the associated $G$-bundle $G(\varpi)={\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec} {\bf g}(\varpi)$. Its automorphism group is a subgroup of $G({\bf C}[z])$, $$ P_{\varpi}=L_{\varpi}U_{\varpi}, \quad U_{\varpi}=\prod_{\alpha\in\Delta}N_{\alpha}(H^{0}(\varpi(\alpha)\infty)), \leqno(6) $$ $L_{\varpi}$ the subgroup of ${\bf g}$ with Lie algebra ${\bf l}_{\varpi}={\bf t}\oplus\bigoplus_{\varpi(\alpha)=0}{\bf g}_{\alpha}$. The moduli space of $G$-bundles on ${\bf C} P^{1}-\{z_{i}\}$ is $$ {\cal M}_{G}(X)=\prod_{[\varpi]\in\mathop{\rm Hom}\limits(R,{\bf Z})/W} {\cal M}_{G}^{\varpi}(X), \leqno(7) $$ where ${\cal M}_{G}^{\varpi}(X)$ is identified with $P_{\varpi}\setminus (G/B)^{N}$, where the action of $G({\bf C}[z])$ is the composition of the morphism $G({\bf C}[z])\to G^{N}$, $g(z)\mapsto (g(z_{i}))_{i}$, and the left translation. Let $(\lambda_{i})_{i}$ be integral dominant weights of $G$, ${\cal L}_{\lambda_{i}}$ be the associated line bundles on $G/B$; $\boxtimes_{i=1}^{N}{\cal L}_{\lambda_{i}}$ is a $G^{N}$-equivariant bundle on $(G/B)^{N}$, so it is $P_{\varpi}$-equivariant; let ${\cal L}_{(\lambda_{i})}$ be the quotient bundle on ${\cal M}_{G}^{\varpi}(X)$. \medskip\noindent \it 2. The FFR scheme. \rm The procedure of sect. 1.2 was applied in [6] to the case of the punctured ${\bf C} P^{1}$. Let us set some notations. Let for each $i$, $k_{i}$ and ${\cal O}_{i}$ be the local field and ring at $z_{i}$; let $\tilde {\bf g}$ the central extension of $\oplus_{i}{\bf g}(k_{i})$ by the cocycle $c((a_{i}),(b_{i}))=\sum_{i=1}^{N}\res_{x_{i}} \langle a_{i}, db_{i}\rangle_{{\bf g}} K$, with values in the abelian algebra ${\bf C} K$. Let $\tilde {\bf g}_{+}$ be the preimage of ${\bf g}(\oplus_{i}{\cal O}_{i})$ in $\tilde {\bf g}$; $\tilde {\bf g}_{+}$ is then isomorphic to ${\bf g}(\oplus_{i}{\cal O}_{i})\oplus {\bf C} K$. Let for $\lambda$ integral dominant weight of ${\bf g}$, $V_{\lambda}$ be the corresponding irreducible representation of ${\bf g}$; and let for $k\in {\bf C}$, and $\lambda_{1},...,\lambda_{N}$ integral dominant weights of ${\bf g}$, $V_{(\lambda_{i})}^{k}$ be the representation of $\tilde {\bf g}_{+}$ in $V_{\lambda_{1}}\otimes...\otimes V_{\lambda_{N}}$, where elements of ${\bf g}(\oplus_{i}{\cal O}_{i})$ act as their images in ${\bf g}^{\oplus N}$, and $K$ by $k$. Let $\bar {\bf g}_{(z_{i})}$ be the Lie algebra of regular maps from $X$ to ${\bf g}$; choose and denote the same way a lifting of this algebra to $\tilde {\bf g}$. Let $\hat {\bf g}$ be the universal central extension of ${\bf g}((u))$, and ${\bf V}_{0}^{-h^{\vee}}$ be the critical level vacuum module over it. Central fields $T(\zeta)\in Z(U_{-h^{\vee}}(\hat{\bf g})_{loc})[[\zeta^{\pm1}]]$ are in correspondance with imaginary weight singular vectors $\sum I(-l)J(-k)...v_{0}\in{\bf V}_{0}^{-h^{\vee}}$, $I,J,...\in {\bf g}$. Following [6], the action of $T(\zeta)$ on $H^{0}({\cal M}_{G}(X), {\cal L}_{(\lambda_{i})})$ can be described as follows. We have an identification $$ H^{0}({\cal M}_{G}(X), {\cal L}_{(\lambda_{i})})=\bar H_{(\lambda_{i})}^{-h^{\vee}}=\{\mu\in ({\bf V}_{(\lambda_{i})}^{-h^{\vee}})^{*}|\mu {\rm \ is \ }\bar {\bf g}_{(z_{i})}{\rm -invariant}\},\leqno(8) $$ where for any $k$, ${\bf V}^{k}_{(\lambda_{i})}$ is the induced representation $\mathop{\rm Ind}_{\tilde {\bf g}_{+}}^{\tilde {\bf g}} V^{k}_{(\lambda_{i})}$. According to [6], 3, lemma 1, $$ \bar H_{(\lambda_{i})}^{-h^{\vee}}\simeq (V_{\lambda_{1}}\otimes\cdots\otimes V_{\lambda_{N}})^{*}. \leqno(9) $$ The field $T$ corresponds to an imaginary weight singular vector $\sum I(-l)J(-k)...v_{0}\in{\bf V}_{0}^{-h^{\vee}}$, $I,J,...\in {\bf g}$. Due to the ``swapping'' identity (3.1) of [6], the action of this singular vector on $(V_{\lambda_{1}}\otimes\cdots\otimes V_{\lambda_{N}})^{*}$ is $$ \sum{1\over{(l-1)!}}\pr^{l-1}I(u){1\over{(k-1)!}}\pr^{l-1}J(u)..., \leqno(10) $$ where $I(u)=\sum_{i=1}^{N}{{I^{(i)}}\over{u-z_{i}}}$. For example, the operators corresponding to the degree two Casimir element are the Gaudin Hamiltonians $H_{2,i}$, such that the combination $H_{2}(\zeta)=\sum_{i=1}^{N}{H_{2,i}\over{\zeta-z_{i}}}$ satisfies $$ H_{2}(\zeta)=\sum_{i} e_{i}(z)e_{i}(z), \leqno(11) $$ with $(e_{i})$ an orthomormal basis of ${\bf g}$. \medskip\noindent \it 3. The AKS scheme. \rm On the other hand, the expression (1) gives a realization of the Lie algebra $u^{-1}{\bf g}[[u^{-1}]]$. More precisely, we have a Lie algebra morphism $\pi:u^{-1}{\bf g}[[u^{-1}]]\to {\bf g}^{\oplus N}$, defined by $\pi(Iu^{-k})=\sum_{i=1}^{N}I^{(i)}z_{i}^{k-1}$. Let us show how the AKS scheme enables us to construct a commuting family in $U(u^{-1}{\bf g}[[u^{-1}]])$. Let us decompose the central extension ${\bf C} K\to \hat {\bf g}\to {\bf g}((u))$ as $\hat {\bf g}={\bf a}\oplus {\bf b}$, ${\bf a}=\sigma(u^{-1}{\bf g}[[u^{-1}]])$ and ${\bf b}=\alpha^{-1}({\bf g}[u])$, $\alpha$ being the projection and $\sigma$ being a section of $u^{-1}{\bf g}[[u^{-1}]]$ to $\hat {\bf g}$. Then, $U\hat {\bf g}=U{\bf a}\oplus (U\hat {\bf g}){\bf b}$. We have then an algebra morphism $Z(U\hat {\bf g})\to U{\bf a}$, given by the projection to the first factor, whose image is a commuting family in $U{\bf a}$. Let us specialize this construction to the critial level. We have then a sequence of morphisms $$ Z(U_{-h^{\vee}}\hat {\bf g})\to U(u^{-1}{\bf g}[[u^{-1}]])\to (U{\bf g})^{\otimes N}, $$ the last one being given by $\pi$. This gives a family of commuting differential operators on $(G/B)^{N}$. \noindent \it Remark. \rm According to [17], Gaudin systems can be obtained from quantum tops systems by a reduction procedure, which explains that the AKS scheme can be applied to them. \medskip\noindent \it 4. Coincidence of the AKS and FFR systems. \rm To see that these operators are the same as those obtained by the previous construction, let us work out the AKS scheme more explicitly. The central field $T(u)$ associated to $\sum I(-l)J(-k)...$, is the normally ordered product $$ \sum{1\over{(l-1)!}}{1\over{(k-1)!}}...(\partial^{l-1}\bar I(u)(\partial^{k-1}\bar J(u)...()))\leqno(12) $$ (where the parenthesis stand for the normal ordering operation); here $$ \bar I(u)=\sum_{n\in{\bf Z}}I(-n-1)u^{n}=I_{+}(u)+I_{-}(u),\leqno(13) $$ $I_{+}(u)=\sum_{n\ge 0}I(-n-1)u^{n}$. The transform of this expression by the AKS procedure will be, due to the conventions $(AB)(u)=(A_{+}B+BA_{-})(u)$, $$ \sum{1\over{(l-1)!}}{1\over{(k-1)!}}...(\partial^{l-1} I_{+}(u)(\partial^{k-1}J_{+}(u)...())). \leqno(14) $$ But the image by $\pi$ of $I_{+}(u)$ is $I(u)=\sum_{i=1}^{N}{{I^{(i)}}\over{u-z_{i}}}$; this shows \proclaim{Proposition 3.1} The expressions (10) and (14) for AKS and FFR Hamiltonians, coincide. \endgroup\par\medbreak \medskip\noindent \it 5. Application: expression of the higher Gaudin operators in the $sl_{n}$ case. \rm The following can then be deduced from [10], using the Newton identities. \proclaim{Proposition 3.2} Let $(s_{p})_{p\ge 0}$ be the sequence of polynomials in $n$, defined by $s_{1}=0$, $s_{2}=n/2$, $s_{3}=-2n/3$, and for $p\ge 2$, $$ (n-p)s_{p}-2(p+1)s_{p+1}-(p+2)s_{p+2}=0;\leqno(15) $$ let $\lambda_{1},\cdots,\lambda_{n}$ be the solutions of the equation $$ \lambda_{n}-s_{1}(n)\lambda^{n-1}+s_{2}(n)\lambda^{n-2}-s_{3}(n)\lambda^{n-3}\cdots=0, \leqno(16) $$ and let $H=\diag(\lambda_{1},\cdots,\lambda_{n})$. Let us set, ${}^{k}H={\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec}(k)H$, for $k\in K=SU(n,{\bf C})$; and let $dk$ be a Haar measure on $K$. Then the higher Gaudin Hamiltonians are the operators $H_{l,a_{i}}$, ($\sum_{i=1}^{N}a_{i}=l-1$), defined for each $l=2,...,N$ by $$ \sum_{a_{1}+...+a_{N}=l-1} {H_{l,a_{i}}\over{\prod_{i=1}^{N}(\zeta-z_{i})^{a_{i}}}}=\int_{K}\bigg( {{({}^{k}H)^{(i)}}\over{\zeta-z_{i}}}\bigg)^{l}dk.\leqno(17) $$ \endgroup\par\medbreak \section{4.}{An $r$-matrix for the case of punctured elliptic curves.} Let us turn now to the case where $\overline X$ is an elliptic curve ${\bf C}^{\times}/q^{{\bf Z}}$ ($q<1$). We denote by $z_{i}$ ($i=1,\cdots,N$) the coordinates of the marked points. We fix from now on, $G=GL_{n}({\bf C})$. Consider the open subset ${\cal M}^{(0)}_{G}(\overline X)$ of ${\cal M}_{G}(\overline X)$, formed of the space of bundles on $\overline X$, direct sums of line bundles of degree $0$. These bundles are parametrized by the symmetric product $\overline X^{(n)}$; to $(t_{1},\cdots,t_{n})\in ({\bf C}^{\times})^{n}$, we associate the bundle ${\cal E}_{(t_{\alpha})}={\bf C}^{\times}\times{\bf C}^{n}/\{(z,\xi)\sim (qz,\diag(t_{\alpha})\xi)\}$ over $X$; changing $(t_{\alpha})$ into $(q^{a_{\alpha}}t_{\alpha})$ (with the $a_{\alpha}$ integers) gives an isomorphic bundle, the isomorphism being $(z,\xi)\mapsto (z,\diag(z^{a_{\alpha}})\xi)$. Now, the preimage in ${\cal M}_{G}(X)$ of this open subset can be identified with $$ T\rtimes S_{n}\setminus ({\bf C}^{\times})^{n}\times (G/B)^{N}/[(t_{\alpha},g_{i}B) \sim (q^{a_{\alpha}}t_{\alpha},\diag(z_{i}^{a_{\alpha}})g_{i}B)], $$ $T\rtimes S_{n}$ acting diagonally on $(G/B)^{N}$, and by permutations on $({\bf C}^{\times})^{n}$. We denote it by ${\cal M}^{(0)}_{G}(X)$. The cotangent to ${\cal M}^{(0)}_{G}(X)$ is now the quotient by $S_{n}$ of the reduction by $T$ of $T^{*}(({\bf C}^{\times})^{n}\times (G/B)^{N})$. The Springer resolution gives now a mapping from $T^{*}{\cal M}^{(0)}_{G}(X)$ to $$ \eqalign{ T\rtimes S_{n}\setminus\{(p_{\alpha},t_{\alpha},\eta_{i}) \in{\bf C}^{n}\times({\bf C}^{\times})^{n} & \times{\cal N}^{N} | (\sum_{i=1}^{N}\eta_{i})_{t}=0\}/ \parskip0pt\par\noindent}\noindent#1}}} & / \{(p_{\alpha},t_{\alpha},\eta_{i})\sim(p_{\alpha}, q^{a_{\alpha}}t_{\alpha}, {\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec}(\diag(z_{i}^{a_{\alpha}}))\eta_{i})\}, \parskip0pt\par\noindent}\noindent#1}}}}$$ bijective over the open subset, defined by the condition that each $\eta_{i}$ be regular. This map is compatible with the Poisson bracket given by the product of $\{p_{\alpha},t_{\beta}\}=\delta_{\alpha\beta}t_{\beta}$, and Kostant-Kirillov on each copy of ${\cal N}$. The corresponding $1$-form $\xi\in H^{0}(X,gl({\cal E}_{(t_{\alpha})})(-\sum_{i=1}^{N}(z_{i})))$ can be seen as a $1$-form $\tilde{\xi}$ on ${\bf C}^{\times}$ with values in $gl_{n}({\bf C})$, with simple poles at $z_{i}q^{{\bf Z}}$, and such that $\tilde{\xi}(qz)={\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec}(t_{\alpha})\tilde{\xi}(z)$; it is given by $\tilde{\xi}(z)=\bar{\xi}(z){{dz}\over z}$, with $$ \bar{\xi}(z)_{\alpha\beta}=\sum_{i=1}^{N}\eta_{\alpha\beta}^{(i)} {{\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta}z z_{i}^{-1})}\over {\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta})\theta}\def\vare{\varepsilon(z z_{i}^{-1})}} {\rm \ if } \alpha\ne \beta, \bar{\xi}(z)_{i}={1\over\theta}\def\vare{\varepsilon'(1)}p_{\alpha}+\sum_{i=1}^{N}{1\over\theta}\def\vare{\varepsilon'(1)} {\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1})\eta_{i}^{i}. \leqno(18) $$ Here $\theta}\def\vare{\varepsilon(z)=\prod_{i\ge 0}(1-q^{i}z)\prod_{i\ge 1}(1-q^{i}z^{-1})$; $\theta}\def\vare{\varepsilon$ has the properties $\theta}\def\vare{\varepsilon(qz)=-z^{-1}\theta}\def\vare{\varepsilon(z)$, $\theta}\def\vare{\varepsilon(z^{-1})=-z^{-1}\theta}\def\vare{\varepsilon(qz)$; we denote $\dot\theta}\def\vare{\varepsilon(z)=z{d\theta}\def\vare{\varepsilon\over dz}(z)$. We will show that the commutativity of the coordinates of the $\mathop{\rm tr}\limits\tilde{\xi}(z)^{k}$ in a basis of the space of $k$-forms on $X-\{z_{i}\}$ can be deduced, as in prop. 2.1, from an $r$-matrix argument: \proclaim{Proposition 4.1} Let $r(z,w,t_{\alpha})$ and $\rho(z,w,t_{\alpha})$ be the matrices acting on ${\bf C}^{n}\otimes{\bf C}^{n}$, with elements $$ r(z,w,t_{\alpha})_{\alpha\beta}^{\gamma\delta}=\bigg( -{\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta}zw^{-1})\over {\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta})\theta}\def\vare{\varepsilon(z w^{-1})}}\delta_{\alpha\beta}^{\delta\gamma} +{1\over\theta}\def\vare{\varepsilon'(1)}{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zw^{-1})\delta_{\alpha\beta}^{\gamma\delta} \bigg)(1-\delta_{\alpha\beta})\leqno(19) $$ and $$ \rho(z,w,t_{\alpha})_{\alpha\beta}^{\gamma\delta}={1\over\theta}\def\vare{\varepsilon'(1)} {\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta}zw^{-1})\over {\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta})\theta}\def\vare{\varepsilon(z w^{-1})}} \bigg[ {\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t_{\alpha}t_{\beta}^{-1})+{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t_{\alpha}^{-1}t_{\beta} zw^{-1}) \bigg] \delta_{\alpha\beta}^{\delta\gamma}(1-\delta_{\alpha\beta});\leqno(20) $$ let $\bar{\xi}(z)$ be given by formula (18); let us endow the system of variables $(p_{\alpha},t_{\alpha},\eta_{i})$ with the Poisson brackets, product of $\{p_{\alpha},t_{\beta}\}=\delta_{\alpha\beta}t_{\beta}$ and Kostant-Kirillov on each copy of ${\cal N}$; then we have $$ \eqalign{ \{\bar\xi(z,t_{\alpha})\otimes_{,}\bar\xi(w,t_{\alpha})\}= [r(z,w,t_{\alpha}), & \bar\xi(z,t_{\alpha})\otimes 1+1\otimes \bar\xi (w,t_{\alpha})] \parskip0pt\par\noindent}\noindent#1}}} & +\rho(z,w,t_{\alpha})[(\sum_{i=1}^{N}\eta_{i})_{t}\otimes 1 -1\otimes(\sum_{i=1}^{N}\eta_{i})_{t}].}\leqno(21) $$ \endgroup\par\medbreak \noindent {\bf Proof.} In the case of the brackets $\{\bar\xi_{\alpha\beta}, \bar\xi_{\beta\gamma}\}$, $\alpha,\beta,\gamma$ all different, it is a consequence of the formula $$ {\theta}\def\vare{\varepsilon(tzw^{-1})\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(zw^{-1})}} {\theta}\def\vare{\varepsilon(tt'w)\over{\theta}\def\vare{\varepsilon(tt')\theta}\def\vare{\varepsilon(w)}} -{\theta}\def\vare{\varepsilon(t^{\prime -1}zw^{-1})\over{\theta}\def\vare{\varepsilon(t^{\prime -1})\theta}\def\vare{\varepsilon(zw^{-1})}} {\theta}\def\vare{\varepsilon(tt'z)\over{\theta}\def\vare{\varepsilon(tt')\theta}\def\vare{\varepsilon(z)}} = {\theta}\def\vare{\varepsilon(tz)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(z)}} {\theta}\def\vare{\varepsilon(t'w)\over{\theta}\def\vare{\varepsilon(t')\theta}\def\vare{\varepsilon(w)}}; $$ to show it, let $F(z,w,t,t')$ be the difference of both sides. We have $F(qz,w,t,t')=t^{-1}F(z,w,t,t')$; moreover $F$ has no poles for $z\to 0$ or $z\to w$; since $t\notin q^{{\bf Z}}$, this shows $F=0$. In the case of the brackets $\{\bar\xi_{\alpha\a},\bar\xi_{\alpha\beta}\}$, $\alpha\ne \beta$, it follows from $$ \eqalign{ {1\over\theta}\def\vare{\varepsilon'(1)}t{d\over dt}[{\theta}\def\vare{\varepsilon(tw)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(w)}}] +{1\over\theta}\def\vare{\varepsilon'(1)}{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z){\theta}\def\vare{\varepsilon(tw)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(w)}} = & -{\theta}\def\vare{\varepsilon(tzw^{-1})\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(zw^{-1})}} {\theta}\def\vare{\varepsilon(tz)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(z)}} \parskip0pt\par\noindent}\noindent#1}}} & +{1\over\theta}\def\vare{\varepsilon'(1)}{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zw^{-1}){\theta}\def\vare{\varepsilon(tw)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(w)}}; }$$ this equality is proven as follows; let $F(z,w)$ be the difference of the two sides, then $F(qz,w)=F(z,w)$ and $F(z,w)$ has no poles for $z\to w $ or $z\to 0$, which shows that $F(z,w)$ does not depend on $z$; pose $F(z,w)=\varphi(w)$, then $\varphi(qw)=t\varphi(w)$ (this follows from ${\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(qz)=-1+{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z)$, obtained by derivation of the functional equation in $\theta}\def\vare{\varepsilon$); and $\varphi(z)$ has no poles either, so $F(z,w)=0$. In the case of the brackets $\{\bar\xi_{\alpha\beta},\bar\xi_{\beta\alpha}\}$, $\alpha\ne \beta$, it follows from the fact, that if we pose $$ F(z,w)= {\theta}\def\vare{\varepsilon(t^{-1}z)\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(z)}}{\theta}\def\vare{\varepsilon(tw)\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(w)}} +{\theta}\def\vare{\varepsilon(t^{-1}zw^{-1})\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(zw^{-1})}}{1\over\theta}\def\vare{\varepsilon'(1)} [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z)-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(w)], $$ we have $F(z,w)=F(z\zeta,w\zeta)$ for any $\zeta\in{\bf C}^{\times}$. Indeed, $F(qz,w)=tF(z,w)-{1\over\theta}\def\vare{\varepsilon'(1)}t {\theta}\def\vare{\varepsilon(t^{-1}zw^{-1})\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(zw^{-1})}}$, so with $\varphi(z,w)=F(z,w)-F(z\zeta,w\zeta)$, we have $\varphi(qz,w)=t\varphi(z,w)$; as $F(z,w)$ has no poles in $z$ (or in $w$), $\varphi$ has no poles either, and so it vanishes. So $F$ is only a function of $zw^{-1}$, that we can evaluate when $w\to 1$; this evaluation gives the matrix elements of $\rho$. The brackets $\{\bar\xi_{\alpha\a},\bar\xi_{\beta\b}\}$ are all zero, and the $[r,\bar\xi\otimes 1 + 1\otimes\bar\xi]_{\alpha\a}^{\beta\b}$ also; finally, the brackets $\{\bar\xi_{\alpha\beta},\bar\xi_{\gamma\delta}\}$ ($\alpha,\beta,\gamma,\delta$ all different) are all zero, as well as the matrix elements $[r,\bar\xi\otimes 1 + 1\otimes\bar\xi]_{\alpha\gamma}^{\beta\delta}$. \hfill $~\vrule height .9ex width .8ex depth -.1ex$ Now, after the reduction by $T$, the $\mathop{\rm tr}\limits\bar\xi(z)^{s}$ will be in involution. Let us give now the explicit form of the Hamiltonians generated by $\mathop{\rm tr}\limits\bar\xi(z)^{2}$. We have $$\eqalign{ \mathop{\rm tr}\limits\bar\xi(z)^{2} & =\sum_{\alpha=1}^{n}(p_{\alpha}+\sum_{i=1}^{N}\eta_{\alpha\a}^{(i)} {\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1}))^{2} +2\sum_{1\le \alpha<\beta\le n} \bigg( \sum_{i=1}^{N}\eta_{\alpha\beta}^{(i)}{\theta}\def\vare{\varepsilon(t_{\alpha}t_{\beta}^{-1}zz_{i}^{-1})\over \theta}\def\vare{\varepsilon(t_{\alpha}t_{\beta}^{-1})\theta}\def\vare{\varepsilon(zz_{i}^{-1})} \bigg) \cdot \parskip0pt\par\noindent}\noindent#1}}} &\cdot \bigg( \sum_{i=1}^{N}\eta_{\beta\alpha}^{(i)}{\theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1}zz_{i}^{-1})\over \theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1})\theta}\def\vare{\varepsilon(zz_{i}^{-1})} \bigg);\parskip0pt\par\noindent}\noindent#1}}}} $$ since $$ \bar\xi(qz)={\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec}(t_{\alpha})\bar\xi(z)-(\sum_{i=1}^{N}\eta^{(i)})_{t}, $$ we have $$ (\mathop{\rm tr}\limits\bar\xi^{2})(qz)=(\mathop{\rm tr}\limits\bar\xi^{2})(z)+\mathop{\rm tr}\limits(\sum_{i=1}^{N} \eta^{(i)})_{t}^{2}-2\sum_{i=1}^{N}{{\dot\theta}\over{\theta}}(zz_{i}^{-1}) \{ \sum_{\alpha=1}^{n}\eta_{\alpha\a}(\sum_{i}\eta_{\alpha\a}^{(i)})\}, $$ so that $$ \mathop{\rm tr}\limits\bar\xi(z)^{2}=H_{0}+\sum_{i=1}^{N}H_{i} {{\dot\theta}\over{\theta}}(zz_{i}^{-1})+\sum_{i=1}^{N}\bigg( {{\dot\theta}\over{\theta}}(zz_{i}^{-1})\bigg)^{2} \{\sum_{\alpha=1}^{n}\eta_{\alpha\a}(\sum_{i}\eta_{\alpha\a}^{(i)})\}; $$ using $$ \eqalign{ ({\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{j}^{-1}))^{2} & =\wp(\ln (zz_{i}^{-1})) +\wp(\ln (zz_{j}^{-1})) \parskip0pt\par\noindent}\noindent#1}}} & -2{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1}) [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{j}^{-1})] +[{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1})]^{2}, } $$ $$\eqalign{ {\theta}\def\vare{\varepsilon(tzz_{i}^{-1})\over{\theta}\def\vare{\varepsilon(t)\theta}\def\vare{\varepsilon(zz_{i}^{-1})}} {\theta}\def\vare{\varepsilon(t^{-1}zz_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(zz_{j}^{-1})}} & =[{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{j}^{-1})] {\theta}\def\vare{\varepsilon(t^{-1}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(z_{i}z_{j}^{-1})}} \parskip0pt\par\noindent}\noindent#1}}} & -[{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t^{-1}z_{i}z_{j}^{-1})] {\theta}\def\vare{\varepsilon(t^{-1}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{-1})\theta}\def\vare{\varepsilon(z_{i}z_{j}^{-1})}} {\rm\ if\ } j\ne i, \parskip0pt\par\noindent}\noindent#1}}} & =\wp(\ln (zz_{i}^{-1}))-\wp(\ln (tz_{i})) {\rm\ \ else } \parskip0pt\par\noindent}\noindent#1}}}}$$ [we set $z=e^{\tau}$, so near $\tau=0$, ${\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z)\sim{1\over\tau}$, $\wp(\tau)\sim{1\over{\tau^{2}}}$], we find $$\eqalign{ H_{i} & =2\sum_{\alpha=1}^{n}p_{\alpha}\eta_{\alpha\a}^{(i)}+2\sum_{\alpha=1}^{n} \sum_{j\ne i}\eta_{\alpha\a}^{(i)} \eta_{\alpha\a}^{(j)} [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{j}z_{i}^{-1})] \parskip0pt\par\noindent}\noindent#1}}} & +2\sum_{\alpha\ne \beta}\sum_{j\ne i}\eta_{\alpha\beta}^{(i)}\eta_{\beta\alpha}^{(j)} {\theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1})}\theta}\def\vare{\varepsilon(z_{i} z_{j}^{-1})} \parskip0pt\par\noindent}\noindent#1}}} }\leqno(22) $$ (a less symmetric form could be obtained using ${\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z)+{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z^{-1})=1$, and the irrelevance of combinations of the $\sum_{i=1}^{N}\eta_{\alpha\a}^{(i)}$), and $$ \eqalign{ H_{0} & =\sum_{\alpha=1}^{n}p_{\alpha}^{2}+\sum_{j<i} \eta_{\alpha\a}^{(i)} \eta_{\alpha\a}^{(j)}[{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1})]^{2} -2\sum_{\alpha<\beta}\sum_{i=1}^{N} \eta_{\alpha\beta}^{(i)} \eta_{\beta\alpha}^{(i)}\wp(\ln(t_{\alpha}t_{\beta}^{-1})) \parskip0pt\par\noindent}\noindent#1}}} & -2\sum_{\alpha<\beta}\sum_{i\ne j}\eta_{\alpha\beta}^{(i)}\eta_{\beta\alpha}^{(j)} [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t_{\beta}t_{\alpha}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon} (t_{\beta}t_{\alpha}^{-1}z_{i}z_{j}^{-1})] {\theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t_{\beta}t_{\alpha}^{-1}) \theta}\def\vare{\varepsilon(z_{i}z_{j}^{-1})}}. \parskip0pt\par\noindent}\noindent#1}}}}\leqno(23) $$ \section{Remark.}{} It is interesting to compare these results with those of [15], [5], [19], [11]. The system considered in these papers is connected with the case $N=1$. Also there should be some connection between the $r$-matrix (19) and the ones from [5] and [19]. \section{5.}{Gaudin-Calogero system in the $sl_{2}$ case.} Let us see now how to construct a quantization of the system of the last section when $G=GL_{2}$. We will construct differential operators on $$ {\cal M}^{(0)}_{G}(X)= T\rtimes S_{n}\setminus ({\bf C}^{\times})^{n}\times (G/B)^{N}/[(t_{\alpha},g_{i}B) \sim (q^{a_{\alpha}}t_{\alpha},\diag(z_{i}^{a_{\alpha}})g_{i}B)], $$ whose symbols will be the Hitchin's Hamiltonians, $\mathop{\rm tr}\limits\bar\xi(z)^{s}$, for $n=2$. For this, we consider an integer $k$ and a system of dominant weights $(\lambda_{i})_{i=1,\cdots,N}$, and the algebra ${\cal A}=\Diff(({\bf C}^{\times})^{n},{\cal L}_{k}^{\boxtimes n})\otimes \otimes_{i=1}^{N}\Diff(G/B,{\cal L}_{\lambda_{i}})$ [here ${\cal L}_{k}=\pi^{*}{\cal O}(k(1))$, $\pi$ the projection ${\bf C}^{\times}\to X$, $\Diff(X,{\cal L})=H^{0}(X,{\cal L}\otimes{\cal D}_{X}\otimes{\cal L}^{-1})$, for $X$ an analytic variety and ${\cal L}$ a line bundle on $X$]. Let $(t_{\alpha})_{1\le \alpha\le n}$ be the coordinates on $({\bf C}^{\times})^{n}$, and $\hat{p}_{\alpha}=t_{\alpha}{\pr\over{\pr t_{\alpha}}}+k{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t_{\alpha})$; let again, $e_{\alpha\beta}^{(i)}$ denote the action of $e_{\alpha\beta}\in gl_{n}({\bf C})$ on the $i$-th factor of the second part of ${\cal A}$. Consider now the matrix $L(z)\in gl_{n}({\bf C})\otimes {\cal A}$, defined by $$ \eqalign{ L(z)_{\alpha\beta}=\sum_{i=1}^{N} e_{\alpha\beta}^{(i)} {{\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta}z z_{i}^{-1})}\over {\theta}\def\vare{\varepsilon(t_{\alpha}^{-1}t_{\beta})\theta}\def\vare{\varepsilon(z z_{i}^{-1})}} & {\rm \ if\ } \alpha\ne \beta, \parskip0pt\par\noindent}\noindent#1}}} & L(z)_{\alpha\a}={1\over\theta}\def\vare{\varepsilon'(1)}\hat{p}_{\alpha} +\sum_{i=1}^{N}{1\over\theta}\def\vare{\varepsilon'(1)}{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(zz_{i}^{-1})e_{\alpha\a}^{i}. }\leqno(24) $$ Let us perform now the reduction of ${\cal A}$ w.r.t. $T$. It can be done as follows: let ${\cal A}[0]$ be the weight zero subalgebra of ${\cal A}$, ${\cal A}[0]=\{x\in {\cal A}| [h_{\alpha\a},x]=0, 1\le \alpha \le n\}$ and ${\cal A}^{red}={\cal A}[0]/(h_{\alpha\a})_{1\le \alpha \le n}$ (where $(h_{\alpha\a})_{1\le \alpha\le n}$ is the left, or right ideal generated by the $h_{\alpha\a}$ in ${\cal A}[0]$). Then ${\cal A}^{red}$ is the algebra of globally defined differential operators on $({\bf C}^{\times})^{n}\times [T\setminus(G/B)^{N}]$, twisted by the quotient of ${\cal L}_{k}^{\boxtimes n}\boxtimes\boxtimes_{i=1}^{N} {\cal L}_{\lambda_{i}}$. {}From $\mathop{\rm tr}\limits L(qz)^{2}=\mathop{\rm tr}\limits L(z)^{2}+\mathop{\rm tr}\limits(L(z)h+hL(z))+\mathop{\rm tr}\limits(h^{2})$, we see that $\mathop{\rm tr}\limits L(z)^{s}$, $s=1,2$ define elements of $[{\cal A}/\sum_{\alpha=1}^{2} {\cal A} h_{\alpha\a}]\otimes H^{0}(X,{\cal O}(s\sum_{i=1}^{N}(z_{i})))$, which also belong to ${\cal A}^{red}\otimes H^{0}(X,{\cal O}(s\sum_{i=1}^{N}(z_{i})))$. Then \proclaim{Proposition 5.1} The expansions of $\mathop{\rm tr}\limits L(z)^{s}$, $s=1,2$, along bases of $ H^{0}(X,{\cal O}(s\break\sum_{i=1}^{N}(z_{i})))$, form a commutative family in ${\cal A}^{red}$. These operators are $S_{2}$-invariant and invariant under the action of ${\bf Z}^{2}$ defined by $(a_{\alpha})\cdot (t_{\alpha},g_{i}B)=(q^{a_{\alpha}}t_{\alpha},\diag(z_{i}^{a_{\alpha}})g_{i}B)$, and hence define operators on ${\cal M}^{(0)}_{GL_{2}}(X)$, twisted by the line bundle associated with $(k,\lambda_{i})$. Their symbols coincide with Hitchin's Hamiltonians. \endgroup\par\medbreak \noindent {\bf Proof.} If $w\in S_{2}$, then $w^{*}L(z)=L(z)$; let $\vare_{\alpha}$ be the $\alpha$-th basis vector of ${\bf Z}^{2}$, then $\vare_{\alpha}^{*}L(z)={\rm Ad}}\def\Diff{{\rm Diff}}\def\Spec{{\rm Spec}(\diag(1,\cdots,z,\cdots,1))L(z)$ ($z$ in $\alpha$-th position). The last statement follows from the fact that the symbol of $\hat{p}_{\alpha}$ is $p_{\alpha}$, and the symbol of $e_{\alpha\beta}^{(i)}$ is $\eta_{\alpha\beta}^{(i)}$. The first statement follows from a direct computation, using the explicit form of the Hamiltonians: $$ \eqalign{ \hat{H}_{i} & = \hat{p}h^{(i)} +2\sum_{j\ne i}h^{(i)}h^{(j)} [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{j}z_{i}^{-1})] \parskip0pt\par\noindent}\noindent#1}}} & +2\sum_{j\ne i} \big(e^{(i)}f^{(j)} {\theta}\def\vare{\varepsilon(t^{2}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{2})}\theta}\def\vare{\varepsilon(z_{i} z_{j}^{-1})} +e^{(j)}f^{(i)} {\theta}\def\vare{\varepsilon(t^{-2}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{-2})}\theta}\def\vare{\varepsilon(z_{i} z_{j}^{-1})}\big) }\leqno(25) $$ and $$ \eqalign{ \hat{H}_{0} & =\hat{p}^{2}+\sum_{j<i} h^{(i)} h^{(j)}[{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(z_{i}z_{j}^{-1})]^{2} -2\sum_{i=1}^{N}e^{(i)}f^{(i)}\wp(\ln(t^{2})) \parskip0pt\par\noindent}\noindent#1}}} & -2\sum_{i\ne j}e^{(i)}f^{(j)} [{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t^{2})-{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon} (t^{2}z_{i}z_{j}^{-1})] {\theta}\def\vare{\varepsilon(t^{2}z_{i}z_{j}^{-1})\over{\theta}\def\vare{\varepsilon(t^{2}) \theta}\def\vare{\varepsilon(z_{i}z_{j}^{-1})}}. \parskip0pt\par\noindent}\noindent#1}}}}\leqno(26) $$ \hfill $~\vrule height .9ex width .8ex depth -.1ex$ Equations (25) and (26) define differential operators acting on ${\bf C}^{\times}\times [T\setminus ({\bf C} P^{1})^{N}]$; $(t,t_{i})$ being the product coordinates on ${\bf C}^{\times}\times ({\bf C} P^{1})^{N}$, we have $\hat{p}=2t{\pr\over{\pr t}}+2k{\dot\theta}\def\vare{\varepsilon\over\theta}\def\vare{\varepsilon}(t^{2})$, $h^{(i)}=2(t_{i}{\pr\over{\pr t_{i}}}+\lambda_{i})$, $e^{(i)}=t_{i}^{2}{\pr\over{\pr t_{i}}}+2\lambda_{i}t_{i}$, $f^{(i)}=-{\pr\over{\pr t_{i}}}$. For $N=1$, this system is reduced to $$ \hat{H}_{0}=\hat{p}^{2}-2e^{(1)}f^{(1)}\wp(\ln t^{2}) , \quad\hat{H}_{1}=e^{(1)}f^{(1)}. $$ \section{Remarks.}{} 1. A natural module for ${\cal A}^{red}$ is ${\rm Fun}({\bf C}^{\times})\otimes (V_{\lambda_{1}}\otimes\cdots\otimes V_{\lambda_{N}})[0]$. More precisely, we can pose the eigenvalue problem $\hat{H}_{i}\psi=\mu_{i}\psi$, $\hat{H}_{0}\psi=\mu_{0}\psi$, $\psi$ a function of ${\bf C}^{\times}$, with values in $\otimes_{i=1}^{N}V_{\lambda_{i}}$, whose component in $(\otimes_{i=1}^{N}V_{\lambda_{i}})[\bar\lambda_{i}]$, $\psi_{\bar\lambda_{i}}(t)$, satisfies $\psi_{\bar\lambda_{i}}(qt)=z_{1}^{\bar\lambda_{1}}\cdots z_{N}^{\bar\lambda_{N}}z^{\ell} \psi_{\bar\lambda_{i}}(t)$, for each system of weights $(\bar\lambda_{i})$, $\ell$ being a fixed integer. The space of such functions, with only poles at $q^{{\bf Z}}$, is stable under the actions of $\hat{H}_{0}$ and the $\hat{H}_{i}$. 2. Prop. 5.1 suggests that the operators constructed here coincide with the result of the action of the center of the enveloping algebra at the critical level, when $k=2$. Indeed in this case, after [22], the quotient of ${\cal L}_{k}^{\boxtimes 2}$ by $S_{2}$ coincides with $(\det_{|{\cal M}^{(0)}_{GL_{2}}(X)})^{-2}$, on which this center should act. After obtaining the main results of this paper, we learnt about the paper of N. Nekrasov [16], where Hitchin systems for degenerate curves are described as many-body problems. \vskip 1truecm \noindent {\bf References} \bigskip \item{[1]} M.R. Adams, J. Harnad, E. Previato, {\sl Isospectral Hamiltonian flows in finite and infinite dimensions II. Integration of flows,} Commun. Math. Phys. 134 (1990), 555-85. \medskip \item{[2]} M. Adler, {\sl On a trace functional for formal pseudo-differential operators and the symplectic structure for the KdV type equations,} Invent. Math. 50 (1979), 219-48. \medskip \item{[3]} A. Beauville, {\sl Jacobiennes des courbes spectrales et syst\`emes compl\`etement int\'e-grables,} Acta Math., 169 (1990), 211-35. \medskip \item{[4]} A.A. Beilinson, V.G. Drinfeld, {\sl Quantization of Hitchin's fibration and Langlands program,} preprint. \medskip \item{[5]} H.W. Braden, T. Suzuki, {\sl $R$-matrices for Elliptic Calogero-Moser Models}, Lett. Math. Phys. 30, 147-59 (1994). \medskip \item{[6]} B.L. Feigin, E.V. Frenkel, N. Reshetikhin, {\sl Gaudin model, Bethe ansatz and correlation functions at the critical level,} Commun. Math. Phys. 166 (1), 27-62 (1995). \medskip \item{[7]} G. Felder, C. Wieczerkowski, {\sl Conformal field theory on elliptic curves and Knizhnik-Zamolodchikov-Bernard equations,} hep-th/9411004. \medskip \item{[8]} R. Garnier, Rend. Circ. Mat. Palermo 43, 155-91 (1919). \medskip \item{[9]} M. Gaudin, Jour. Physique 37 (1976), 1087-1098. \medskip \item{[10]} R. Goodman, N.R. Wallach, {\sl Higher-order Sugawara operators for affine Lie algebras,} Trans. AMS, 315:1 (1989), 1-55. \medskip \item{[11]} A.S. Gorsky, N.A. Nekrasov,{\sl Elliptic Calogero-Moser system from two-dimensio-nal current algebra,} hep-th/9401021. \medskip \item{[12]} N. Hayashi, {\sl Sugawara operators and Kac-Kazhdan conjecture,} Invent. Math. 54 (1988), 13-52. \medskip \item{[13]} N. Hitchin, {\sl Stable bundles and integrable systems,} Duke Math. Jour., 54 (1), 91-114 (1987). \medskip \item{[14]} B. Kostant, {\sl The solution to a generalized Toda lattice and representation theory,} Adv. Math. 34 (1980), 13-53. \medskip \item{[15]} I.M. Krichever, {\sl Elliptic solutions of the Kadomtsev-Petviashvili equation and integrable systems of particles,} Funct. An. Appl., 14 (1), 282-90 (1990). \medskip \item{[16]} N. Nekrasov, {\sl Holomorphic bundles and many-body systems,} PUPT-1534, ITEP-N95/1, hep-th/9503157. \medskip \item{[17]} A.G. Reyman, {\sl Quantum tops,} Int. J. Mod. Phys. B, 7:20-21 (1993), 3707-13. \medskip \item{[18]} M.A. Semenov-Tian-Shansky, {\sl D. Sci. thesis,} LOMI, Leningrad (1985). \medskip \item{[19]} E.K. Sklyanin, {\sl Dynamical $r$-matrices for the elliptic Calogero-Moser system,} LPTHE 93-42, hep-th/9308060. \medskip \item{[20]} T.A. Springer, {\sl Trigonometric sums, Green functions of finite groups and representations of Weyl groups,} Inv. Math. 36, 173-207 (1976). \medskip \item{[21]} W. Symes, {\sl Systems of Toda type, inverse spectral problems and representation theory,} Invent. Math. 59 (1990), 195-338. \medskip \item{[22]} L.W. Tu, {\sl Semistable bundles over an Elliptic Curve,} Adv. Math. 98, 1-26 (1993). \medskip \medskip\medskip \section{}{} B.E., V.R.: Centre de Math\'{e}matiques, URA 169 du CNRS, Ecole Polytechnique, 91128 Palaiseau, France V.R.: ITEP, Bol. Cheremushkinskaya, 25, 117259, Moscow, Russia. \bye
1995-08-08T07:51:28
9503
alg-geom/9503021
en
https://arxiv.org/abs/alg-geom/9503021
[ "alg-geom", "math.AG" ]
alg-geom/9503021
Nitin Nitsure
Nitin Nitsure and Claude Sabbah
Moduli of pre-$\cal D$-modules, perverse sheaves and the Riemann-Hilbert morphism -I
LaTeX, 28 pages
null
null
null
null
We construct a moduli scheme for semistable pre-$\D$-modules with prescribed singularities and numerical data on a smooth projective variety. These pre-$\D$-modules are to be viewed as regular holonomic $\D$-modules with `level structure'. We also construct a moduli scheme for perverse sheaves on the variety with prescribed singularities and other numerical data, and represent the de Rham functor (which gives the Riemann-Hilbert correspondence) by an analytic morphism between the two moduli schemes.
[ { "version": "v1", "created": "Tue, 28 Mar 1995 11:35:00 GMT" } ]
2008-02-03T00:00:00
[ [ "Nitsure", "Nitin", "" ], [ "Sabbah", "Claude", "" ] ]
alg-geom
\subsection*{\hbox{}\hfill{\normalsize\sl #1}\hfill\hbox{}}} \textheight 23truecm \textwidth 15truecm \addtolength{\oddsidemargin}{-1.05truecm} \addtolength{\topmargin}{-1.5truecm} \makeatletter \def\l@section{\@dottedtocline{1}{0em}{1.2em}} \makeatother \begin{document} \title{Moduli of pre-${\cal D}$-modules, perverse sheaves\\ and the Riemann-Hilbert morphism -I} \author{Nitin Nitsure\thanks{Tata Institute of Fundamental Research, Bombay} \and Claude Sabbah\thanks{CNRS, URA D0169, Ecole Polytechnique, Palaiseau}} \date{March 28, 1995} \maketitle \begin{abstract} We construct a moduli scheme for semistable pre-${\cal D}$-modules with prescribed singularities and numerical data on a smooth projective variety. These pre-${\cal D}$-modules are to be viewed as regular holonomic ${\cal D}$-modules with `level structure'. We also construct a moduli scheme for perverse sheaves on the variety with prescribed singularities and other numerical data, and represent the de Rham functor (which gives the Riemann-Hilbert correspondence) by an analytic morphism between the two moduli schemes. \end{abstract} \vfill \tableofcontents \vfill \newpage \section{Introduction} This paper is devoted to the moduli problem for regular holonomic ${\cal D}$-modules and perverse sheaves on a complex projective variety $X$. It treats the case where the singular locus of the ${\cal D}$-module is a smooth divisor $S$ and the characteristic variety is contained in the union of the zero section $T^*_XX$ of the cotangent bundle of $X$ and the conormal bundle $N^*_{S,X}$ of $S$ in $X$ (also denoted $T_S^*X$). The sequel (part II) will treat the general case of arbitrary singularities. A moduli space for ${\cal O}$-coherent ${\cal D}$-modules on a smooth projective variety was constructed by Simpson [S]. These are vector bundles with integrable connections, and they are the simplest case of ${\cal D}$-modules. In this moduli construction, the requirement of semistability is automatically fulfilled by all the objects. Next in order of complexity are the so called `regular meromorphic connections'. These ${\cal D}$-modules can be generated by vector bundles with connections which have logarithmic singularities on divisors with normal crossing. These ${\cal D}$-modules are not ${\cal O}$-coherent, but are torsion free as ${\cal O}$-modules. A moduli scheme does not exist for these ${\cal D}$-modules themselves (see section 1 of [N]), but it is possible to define a notion of stability and construct a moduli for vector bundles with logarithmic connections. This was done in [N]. Though many logarithmic connections give rise to the same meromorphic connection, the choice of a logarithmic connection is infinitesimally rigid if its residual eigenvalues do not differ by nonzero integers (see section 5 of [N]). In the present paper and its sequel, we deal with general regular holonomic ${\cal D}$-modules. Such modules are in general neither ${\cal O}$-coherent, nor ${\cal O}$-torsion free or pure dimensional. We define objects called pre-${\cal D}$-modules, which play the same role for regular holonomic ${\cal D}$-modules that logarithmic connections played for regular meromorphic connections. We define a notion of (semi-)stability, and construct a moduli scheme for (semi-) stable pre-${\cal D}$-modules with prescribed singularity stratification and other numerical data. We also construct a moduli scheme for perverse sheaves with prescribed singularity stratification and other numerical data on a nonsingular variety, and show that the Riemann-Hilbert correspondence defines an analytic morphism between (an open set of) the moduli of pre-${\cal D}$-modules and the moduli of perverse sheaves. The contents of this paper are as follows. Let $X$ be a smooth projective variety, and let $S$ be a smooth hypersurface on $X$. In section 2, we define pre-${\cal D}$-modules on $(X,S)$ which may be regarded as ${\cal O}_X$-coherent descriptions of those regular holonomic ${\cal D}_X$-modules whose characteristic variety is contained in $T^*_XX\cup T^*_SX$. The pre-${\cal D}$-modules form an algebraic stack in the sense of Artin, which is a property that does not hold for the corresponding ${\cal D}$-modules. In section 3, we define a functor from the pre-${\cal D}$-modules to ${\cal D}$-modules (in fact we mainly use the presentation of holonomic ${\cal D}$-modules given by Malgrange [Mal], that we call Malgrange objects). This is a surjective functor, and though not injective, it has an infinitesimal rigidity property (see proposition \ref{prop4}) which generalizes the corresponding result for meromorphic connections. In section 4, we introduce a notion of (semi-)stability for pre-${\cal D}$-modules, and show that semistable pre-${\cal D}$-modules with fixed numerical data form a moduli scheme. Next, we consider perverse sheaves on $X$ which are constructible with respect to the stratification $(X-S)\cup S$. These have finite descriptions through the work Verdier, recalled in section 5. We observe that these finite descriptions are objects which naturally form an Artin algebraic stack. Moreover, we show in section 6 that S-equivalence classes (Jordan-H\"older classes) of finite descriptions with given numerical data form a coarse moduli space which is an affine scheme. Here, no hypothesis about stability is necessary. In section 7, we consider the Riemann-Hilbert correspondence. When a pre-${\cal D}$-module has an underlying logarithmic connection for which eigenvalues do not differ by nonzero integers, we functorially associate to it a finite description, which is the finite description of the perverse sheaf associated to the corresponding ${\cal D}$-module by the Riemann-Hilbert correspondence from regular holonomic ${\cal D}$-modules to perverse sheaves. We show that this gives an analytic morphism of stacks from the analytic open subset of the stack (or moduli) of pre-${\cal D}$-modules on $(X,S)$ where the `residual eigenvalues are good', to the stack (or moduli) of finite descriptions on $(X,S)$. In section 8, we show that the above morphism of analytic stacks is in fact a spread (surjective local isomorphism) in the analytic category. We also show that it has removable singularities in codimension 1, that is, is can be defined outside codimension two on any parameter space which is smooth in codimension 1. \paragraph{Acknowledgement} The authors thank the exchange programme in mathematics of the Indo-French Center for the Promotion of Advanced Research, New Delhi, the Ecole Polytechnique, Paris, and the Tata Institute of Fundamental Research, Bombay, for supporting their collaboration. The first author also thanks ICTP Trieste and the University of Kaiserslautern for their hospitality while part of this work was done. \section{Pre-${\cal D}$-modules} Let $X$ be a nonsingular variety and let $S\subset X$ be a smooth divisor (reduced). Let ${\cal I}_S\subset {\cal O}_X$ be the ideal sheaf of $S$, and let $T_X[\log S]\subset T_X$ be the sheaf of all tangent vector fields on $X$ which preserve ${\cal I}_S$. Let ${\cal D}_X[\log S]\subset {\cal D}_X$ be the algebra of all partial differential operators which preserve $I_S$; it is generated as an ${\cal O}_X$ algebra by $T_X[\log S]$. The ${\cal I}_S$-adic filtration on ${\cal O}_X$ gives rise to a (decreasing) filtration of ${\cal D}_X$ as follows: for $k\inZ\!\!\!Z$ define $V^k{\cal D}_X$ as the subsheaf of ${\cal D}_X$ whose local sections consist of operators $P$ which satisfy $P\cdot {\cal I}_S^j\subset {\cal I}_S^{k+j}$ for all $j$. By construction, one has ${\cal D}_X[\log S]=V^0{\cal D}_X$ and every $V^k({\cal D}_X)$ is a coherent ${\cal D}_X[\log S]$-module. Let $p:N_{S,X}\to S$ denote the normal bundle of $S$ in $X$. The graded ring $\mathop{\rm gr}\nolimits_V{\cal D}_X$ is naturally identified with $p_*{\cal D}_{N_{S,X}}$. Its $V$-filtration (corresponding to the inclusion of $S$ in $N_{S,X}$ as the zero section) is then split. There exists a canonical section $\theta$ of the quotient ring ${\cal D}_X[\log S]/{\cal I}_S{\cal D}_X[\log S]=\mathop{\rm gr}\nolimits^0_V{\cal D}_X$, which is locally induced by $x\partial_x$, where $x$ is a local equation for $S$. It is a central element in $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$. This ring contains ${\cal O}_S$ as a subring and ${\cal D}_S$ as a quotient (one has ${\cal D}_S=\mathop{\rm gr}\nolimits^0_V{\cal D}_X/\theta\mathop{\rm gr}\nolimits^0_V{\cal D}_X$). One can identify locally on $S$ the ring $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$ with ${\cal D}_S[\theta ]$. A coherent $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$-module on which $\theta$ acts by $0$ is a coherent ${\cal D}_S$-module. The locally free rank one ${\cal O}_S$-module ${\cal N}_{S,X}={\cal O}_X(S)/{\cal O}_X$ is a $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$-module on which $\theta$ acts by $-1$. \begin{definition}\rm A {\sl logarithmic module} on $(X,S)$ will mean a sheaf of ${\cal D}_X[\log S]$-modules, which is coherent as an ${\cal O}_X$-module. A {\sl logarithmic connection} on $(X,S)$ will mean a logarithmic module which is coherent and torsion-free as an ${\cal O}_X$-module. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem} It is known that when $S$ is nonsingular, any logarithmic connection on $(X,S)$ is locally free as an ${\cal O}_X$-module. \begin{definition}[Family of logarithmic modules]\rm Let $f:Z\to T$ be a smooth morphism of schemes. Let $Y\subset Z$ be a divisor such that $Y\to T$ is smooth. Let $T_{Z/T}[\log Y]\subset T_{X/Y}$ be the sheaf of germs of vertical vector fields which preserve the ideal sheaf of $Y$ in ${\cal O}_Z$. This generates the algebra ${\cal D}_{Z/T}[\log Y]$. A family of logarithmic modules on $Z/T$ is a ${\cal D}_{Z/T}[\log Y]$-module which is coherent as an ${\cal O}_Z$-module, and is flat over ${\cal O}_T$. When $f:Z\to T$ is the projection $X\times T\to T$, and $Y=S\times T$, we get a {\sl family of logarithmic modules on $(X,S)$ parametrized by $T$}. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem} The restriction to $S$ of a logarithmic module is acted on by $\theta$: for a logarithmic connection, this is the action of the residue of the connection, which is an ${\cal O}_S$-linear morphism. \refstepcounter{theorem}\paragraph{Remark \thetheorem} There is an equivalence (restriction to $S$) between logarithmic modules supported on the reduced scheme $S$ and $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$-modules which are ${\cal O}_S$-coherent, (hence locally free ${\cal O}_S$-modules, since they are locally ${\cal D}_S$-modules). In the following, we shall not make any difference between the corresponding objects. \bigskip We give two definitions of pre-${\cal D}$-modules. The two definitions are `equivalent' in the sense that they give not only equivalent objects, but also equivalent families, or more precisely, the two definitions give rise to isomorphic algebraic stacks. To give a familier example of such an equivalence, this is the way how vector bundles and locally free sheaves are `equivalent'. Note also that mere equivalence of objects is not enough to give equivalence of families --- for example, the category of flat vector bundles is equivalent to the category of $\pi_1$ representations, but an algebraic family of flat bundles gives in general only a holomorphic (not algebraic) family of $\pi_1$ representations. In their first version, pre-${\cal D}$-modules are objects that live on $X$, and the functor from pre-${\cal D}$-modules to ${\cal D}$-modules has a direct description in their terms. The second version of pre-${\cal D}$-modules is more closely related to the Malgrange description of ${\cal D}$-modules and the Verdier description of perverse sheaves, and the Riemann-Hilbert morphism to the stack of perverse sheaves has direct description in its terms. \begin{definition}[Pre-${\cal D}$-module of first kind on $(X,S)$]\rm Let $X$ be a nonsingular variety, and $S\subset X$ a smooth divisor. A pre-${\cal D}$-module ${\bf E} = (E,F,t,s)$ on $(X,S)$ consists of the following data (1) $E$ is a logarithmic connection on $(X,S)$. (2) $F$ is a logarithmic module on $(X,S)$ supported on the reduced scheme $S$ (hence a flat connection on $S$). (3) $t:(E\vert S) \to F$ and $s:F \to (E\vert S)$ are ${\cal D}_X[\log S]$ linear maps, which satisfies the following conditions: (4) On $E\vert S$, we have $st = R$ where $R\in End(E\vert S)$ is the residue of $E$. (5) On $F$, we have $ts = \theta_F$ where $\theta_F:F\to F$ is the ${\cal D}_X[\log S]$ linear endomorphism induced by any Eulerian vector field $x\partial /\partial x$. \end{definition} If $(E,F,t,s)$ and $(E',F',t',s')$ are two pre-${\cal D}$-modules, a morphism between them consists of ${\cal D}_X[\log S]$ linear morphisms $f_0:E\to E'$ and $f_1:F\to F'$ which commute with $t,t'$ and with $s,s'$. \refstepcounter{theorem}\paragraph{Remark \thetheorem} It follows from the definition of a pre-${\cal D}$-module $(E,F,t,s)$ that $E$ and $F$ are locally free on $X$ and $S$ respectively, and the vector bundle morphisms $R$, $s$ and $t$ all have constant ranks on irreducible components of $S$. \paragraph{Example} Let $E$ be a logarithmic connection on $(X,S)$. We can associate functorially to $E$ the following three pre-${\cal D}$-modules. Take $F_1$ to be the restriction of $E$ to $S$ as an ${\cal O}$-module. Let $t_1 = R$ (the residue) and $s_1 = 1_F$. Then ${\bf E}_1=(E,F_1,t_1,s_1)$ is a pre-${\cal D}$-module, which under the functor from pre-${\cal D}$-modules to ${\cal D}$-modules defined later will give rise to the meromorphic connection corresponding to $E$. For another choice, take $F_2 = E\vert S$, $t_2=1_F$ and $s_2=R$. This gives a pre-${\cal D}$-module ${\bf E}_2 = (E,F_2,t_2,s_2)$ which will give rise to a ${\cal D}$-module which has nonzero torsion part when $R$ is not invertible. For the third choice (which is in some precise sense the minimal choice), take $F_3$ to be the image vector bundle of $R$. Let $t_3 =R:(E\vert S)\to F_3$, and let $s_3:F_3\hookrightarrow (E\vert S)$. This gives a pre-${\cal D}$-module ${\bf E}_3 = (E,F_3,t_3,s_3)$. We have functorial morphisms ${\bf E}_3\to {\bf E}_2 \to {\bf E}_1$ of pre-${\cal D}$-modules. \begin{definition}[Families of pre-${\cal D}$-modules]\rm Let $T$ be a complex scheme. A family ${\bf E}_T$ of pre-${\cal D}$-modules on $(X,S)$ parametrized by the scheme $T$, a morphism between two such families, and pullback of a family under a base change $T'\to T$ have obvious definitions (starting from definition of families of ${\cal D}_X[\log S]$-modules), which we leave to the reader. This gives us a fibered category $PD$ of pre-${\cal D}$-modules over the base category of $C\!\!\!\!I$ schemes. Let $\cal PD$ be the largest (nonfull) subcategory of $PD$ in which all morphisms are isomorphisms. This is a groupoid over $C\!\!\!\!I$ schemes. \end{definition} \begin{proposition} The groupoid $\cal PD$ is an algebraic stack in the sense of Artin. \end{proposition} \paragraph{Proof} It can be directly checked that $\cal PD$ is a sheaf, that is, descent and effective descent are valid for faithfully flat morphisms of parameter schemes of families of pre-${\cal D}$-modules. Let $Bun_X$ be the stack of vector bundles on $X$, and let $Bun_S$ be the stack of vector bundles on $S$. Then $\cal PD$ has a forgetful morphism to the product stack $Bun_X\times_{C\!\!\!\!I} Bun_S$. The later stack is algebraic and the forgetful morphism is representable, hence the desired conclusion follows. \bigskip Before giving the definition of pre-${\cal D}$-modules of the second kind, we observe the following. \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem1} Let $N$ be any line bundle on a smooth variety $S$, and let $\ov{N} = P(N\oplus {\cal O}_S)$ be its projective completion, with projection $\pi : \ov{N} \to S$. Let $S^{\infty} = P(N)$ be the divisor at infinity. For any logarithmic connection $E$ on $(\ov{N} ,S\cup S^{\infty})$, the restriction $E\vert S$ is of course a ${\cal D}_{\ov{N}}[\log S\cup S^{\infty}]$-module. But conversely, for any ${\cal O} $-coherent ${\cal D}_{\ov{N}}[\log S\cup S^{\infty}]$-module $F$ scheme theoretically supported on $S$, there is a natural structure of a logarithmic connection on $(\ov{N} ,S\cup S^{\infty})$ on its pullup $\pi ^*(F)$ to $\ov{N}$. The above correspondence is well behaved in families, giving an isomorphism between the algebraic stack of ${\cal D}_{\ov{N}}[\log S\cup S^{\infty}]$-modules $F$ supported on $S$ and the algebraic stack of logarithmic connections $E$ on $(\ov{N} ,S\cup S^{\infty})$ such that the vector bundle $E$ is trivial on the fibers of $\pi :\ov{N} \to S$. The functors $\pi ^*(-)$ and $(-)\vert S$ are fully faithful. \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem2} Let $S\subset X$ be a smooth divisor, and let $N=N_{S,X}$ be its normal bundle. Then the following are equivalent in the sense that we have fully faithful functors between the corresponding categories, which give naturally isomorphic stacks. (1) ${\cal D}_X[\log S]$-modules which are scheme theoretically supported on $S$. (2) ${\cal D}_N[\log S]$-modules which are scheme theoretically supported on $S$. (3) ${\cal D}_{\ov{N}}[\log S \cup S^{\infty}]$-modules which are scheme theoretically supported on $S$. The equivalence between (2) and (3) is obvious, while the equivalence between (1) and (2) is obtained as follows. The Poincar\'e residue map $res:\Omega ^1_X[\log S] \to {\cal O}_S$ gives the following short exact sequence of ${\cal O}_S$-modules. $$0\to \Omega ^1_S \to \Omega ^1_X[\log S]|S \to {\cal O}_S\to 0$$ By taking duals, this gives $$0 \to {\cal O}_S \to T_X[\log S]|S \to T_S\to 0.$$ It can be shown that there exists a unique isomorphism $T_X[\log S]\vert S \to T_N[\log S]\vert S$ which makes the following diagram commute, where the rows are exact. $$\matrix{ 0 & \to & {\cal O}_S & \to & T_N[\log S]|S & \to & T_S & \to & 0 \cr & & \Vert & & \downarrow & & \Vert & & \cr 0 & \to & {\cal O}_S & \to & T_X[\log S]|S & \to & T_S & \to & 0 \cr }$$ \refstepcounter{theorem}\paragraph{Remarks \thetheorem} (1) Observe that the element $\theta$ is just the image of $1$ under the map ${\cal O}_S \to T_X[\log S]\vert S$. (2) Using the notations of the beginning of this section, one can identify the ring $\pi_*{\cal D}_{\ov{N}}[\log S \cup S^{\infty}]$ with $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$. Hence $\theta$ is a global section of ${\cal D}_{\ov{N}}[\log S \cup S^{\infty}]$. \bigskip We now make the following important definition. \begin{definition}[Specialization of a logarithmic module]\rm Let $E$ be a logarithmic module on $(X,S)$. Then the specialization $\mathop{\rm sp}\nolimits_SE$ will mean the logarithmic connection $\pi ^*(E\vert S)$ on $(\ov{N_{S,X}} , S\cup S^{\infty})$. \end{definition} Now we are ready to define the second version of pre-${\cal D}$-modules. \begin{definition}[Pre-${\cal D}$-modules of the second kind on $(X,S)$]\label{def1}\rm A pre-${\cal D}$-mo\-dule (of the second kind) ${\bf E} = (E_0,E_1,c,v)$ on $(X,S)$ consists of the following data (1) $E_0$ is a logarithmic connection on $(X,S)$, (2) $E_1$ is a logarithmic connection on $(\ov{N_{S,X}},S\cup S^\infty)$, (3) $c:\mathop{\rm sp}\nolimits_SE_0 \to E_1$ and $v:E_1 \to \mathop{\rm sp}\nolimits_SE_0$ are ${\cal D}_{\ov{N_{S,X}}}[\log S\cup S^\infty]$-linear maps, which satisfies the following conditions: (4) on $\mathop{\rm sp}\nolimits_SE_0$, we have $cv = \theta_{\mathop{\rm sp}\nolimits_SE_0}$, (5) on $E_1$, we have $vc = \theta_{E_1}$, (6) the vector bundle underlying $E_1$ is {\sl trivial} in the fibers of $\pi:\ov{N_{S,X}}\to S$ (that is, $E_1$ is locally over $S$ isomorphic to $\pi^*(E_1|S)$). \end{definition} If $(E_0,E_1,c,v)$ and $(E'_0,E'_1,c',v')$ are two pre-${\cal D}$-modules, a morphism between them consists of ${\cal D}_X[\log S]$ linear morphisms $f_0:E_0\to E'_0$ and $f_1:E_1\to E'_1$ such that $\mathop{\rm sp}\nolimits_Sf_0$ and $f_1$ commute with $v,v'$ and with $c,c'$. \begin{definition}[Families of pre-${\cal D}$-modules of the second kind]\rm Let $T$ be a complex scheme. A family ${\bf E}_T$ of pre-${\cal D}$-modules on $(X,S)$ parametrized by the scheme $T$, a morphism between two such families, and pullback of a family under a base change $T'\to T$ have obvious definitions which we leave to the reader. This gives us a fibered category $PM$ of pre-${\cal D}$-modules of second kind over the base category of $C\!\!\!\!I$ schemes. \end{definition} \begin{proposition} The functor which associates to each family of pre-${\cal D}$-module $(E_0,E_1,c,v)$ of the second kind parametrized by $T$ the family of pre-${\cal D}$-module of the first kind $(E_0,E_1|S, c|S, v|S)$ is an equivalence of fibered categories. \end{proposition} \paragraph{Proof} This follows from remarks \ref{rem1} and \ref{rem2} above. \section{From pre-${\cal D}$-modules to ${\cal D}$-modules} In this section we first recall the description of regular holonomic ${\cal D}$-modules due to Malgrange [Mal] and we associate a `Malgrange object' to a pre-${\cal D}$-module of the second kind (Proposition \ref{prop2}), which has good residual eigenvalues (definition \ref{goodres}), each component of $S$ do not differ by positive integers. Having such a direct description of the Malgrange object enables us to prove that every regular holonomic ${\cal D}$-module with characteristic variety contained in $T^*_XX\cup T^*_SX$ arises from a pre-${\cal D}$-module (Corollary \ref{cor3}), and also helps us to prove an infinitesimal rigidity property for the pre-${\cal D}$-modules over a given ${\cal D}$-module (Proposition \ref{prop4}). \inter{Malgrange objects} Regular holonomic ${\cal D}$-modules on $X$ whose characteristic variety is contained in $T^*_XX\cup T^*_SX$ have an equivalent presentation due to Malgrange and Verdier, which we now describe. Let us recall the definition of the {\sl specialization} $\mathop{\rm sp}\nolimits_S(M)$ of a regular holonomic ${\cal D}_X$-module $M$: fix a section $\sigma$ of the projection $C\!\!\!\!I\toC\!\!\!\!I/Z\!\!\!Z$ and denote $A$ its image; every such module admits a unique (decreasing) filtration $V^kM$ ($k\inZ\!\!\!Z$) by ${\cal D}_X[\log S]$-submodules which is good with respect to $V{\cal D}_X$ and satisfies the following property: on $\mathop{\rm gr}\nolimits^k_VM$, the action of $\theta$ admits a minimal polynomial all of whose roots are in $A+k$. Then by definition one puts $\mathop{\rm sp}\nolimits_SM=\oplus_{k\inZ\!\!\!Z}\mathop{\rm gr}\nolimits^k_VM$. One has $(\mathop{\rm sp}\nolimits_SM)[*S]=\mathop{\rm sp}\nolimits_S(M[*S])=(\mathop{\rm gr}\nolimits_{V}^{\geq k}M)[*S]$ for all $k\geq 1$, if we put $\mathop{\rm gr}\nolimits_{V}^{\geq k}M=\oplus_{\ell\geq k}\mathop{\rm gr}\nolimits^\ell_VM$. The $p_*{\cal D}_{N_SX}$-module $\mathop{\rm sp}\nolimits_SM$ does not depend on the choice of $\sigma$ (if one forgets its gradation). If $\theta$ acts in a locally finite way on a $\mathop{\rm gr}\nolimits^0_V{\cal D}_X$ or a $p_*{\cal D}_{N_{S,X}}$-module, we denote $\Theta$ the action of $\exp(-2i\pi\theta)$. Given a regular holonomic ${\cal D}_X$-module, we can functorially associate to it the following modules: (1) $M[*S]={\cal O}_X[*S]\otimes_{{\cal O}_X}M$ is the $S$-localized ${\cal D}_X$-module; it is also regular holonomic; (2) $\mathop{\rm sp}\nolimits_S M$ is the specialized module; this is a regular holonomic $p_*{\cal D}_{N_SX}$-module, which is also {\sl monodromic}, i.e. the action of $\theta$ on each local section is locally (on S) finite. The particular case that we shall use of the result proved in [Mal] is then the following: \begin{theorem} There is an equivalence between the category of regular holonomic ${\cal D}_X$-modules and the category which objects are triples $({\cal M},\overline M,\alpha)$, where ${\cal M}$ is a $S$-localized regular holonomic ${\cal D}_X$-module, $\overline M$ is a monodromic regular holonomic $p_*{\cal D}_{N_SX}$-module and $\alpha$ is an isomorphism (of localized $p_*{\cal D}_{N_SX}$-modules) between $\mathop{\rm sp}\nolimits_S{\cal M}[*S]$ and $\overline M[*S]$. \end{theorem} In fact, the result of [Mal] does mention neither holonomicity nor regularity. Nevertheless, using standard facts of the theory, one obtains the previous proposition. Regularity includes here regularity at infinity, i.e. along $S^\infty$. This statement can be simplified when restricted to the category of regular holonomic ${\cal D}$-modules which characteristic variety is contained in the union $T^*_XX\cup T^*_SX$. \begin{definition}\rm A {\sl Malgrange object} on $(X,S)$ is a tuple $(M_0,M_1,C,V)$ where (1) $M_0$ is an $S$-localized regular holonomic ${\cal D}_X$-module which is a regular meromorphic connection on $X$ with poles on $S$, (2) $M_1$ is a $S$-localized monodromic regular holonomic $p_*{\cal D}_{N_SX}$-module which is a regular meromorphic connection on $N_{S,X}$ (or $\ov{N_{S,X}}$) with poles on $S$ (or on $S\cup S^\infty$), (3) $C,V$ are morphisms (of $p_*{\cal D}_{N_{S,X}}$-modules) between $\mathop{\rm sp}\nolimits_SM_0$ and $M_1$ satisfying $VC=\Theta-\mathop{\rm id}\nolimits$ on $\mathop{\rm sp}\nolimits_SM_0$ and $CV=\Theta-\mathop{\rm id}\nolimits$ on $M_1$. \end{definition} The morphisms between two Malgrange objects are defined in an obvious way, making them an abelian category. The previous result can be translated in the following way, using [Ve]: \begin{corollary} There is an equivalence between the category of regular holonomic ${\cal D}$-modules which characteristic variety is contained in $T^*_XX\cup T^*_SX$ and the category of Malgrange objects on $(X,S)$. \end{corollary} \inter{From pre-${\cal D}$-modules to Malgrange objects} \begin{definition}\label{goodres}\rm (1) We say that a logarithmic connection $F$ on $(X,S)$ has {\sl good residual eigenvalues} if for each connected component $S_a$ of the divisor $S$, the residual eigenvalues $(\lambda _{a,k})$ of $F$ along $S_a$ do not include a pair $\lambda _{a,i},\lambda _{a,j}$ such that $\lambda _{a,i}-\lambda _{a,j}$ is a nonzero integer. (2) We say that a pre-${\cal D}$-module ${\bf E} =(E_0,E_1,s,t)$ has {\sl good residual eigenvalues} if the logarithmic connection $E_0$ has good residual eigenvalues as defined above. \end{definition} We now functorially associate a Malgrange object ${\bf M}=\eta ({\bf E})=(M_0,M_1,C,V)$ to each pre-${\cal D}$-module ${\bf E} = (E_0,E_1,c,v)$ on $(X,S)$ with $E_0$ having good residual eigenvalues. \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem3} By definition of a pre-${\cal D}$-module it follows that the nonzero eigenvalues of $\theta_a$ on $E_0|S_a$ (the residue along $S_a$) are the same as the nonzero eigenvalues of $\theta_a$ on $E_{1,a}$. \begin{proposition}{\bf(The Malgrange object associated to a pre-${\cal D}$-module with good residual eigenvalues)}\quad\label{prop2} Let ${\bf E}=(E_0,E_1,c,v)$ be a pre-${\cal D}$-module on $(X,S)$ of the second kind (definition \ref{def1}), such that $E_0$ has good residual eigenvalues. Let $\eta ({\bf E} ) = (M_0,M_1,C,V)$ where (1) $M_0=E_0[*S]$, (2) $M_1=E_1[*S]$, (3) $C = c\circ \displaystyle{e_{}^{-2i\pi\theta_{E_0}}- 1\over\theta_{E_0}}$. (4) $V=v$ Then $\eta ({\bf E})$ is a Malgrange object, and $\eta$ is functorial in an obvious way. \end{proposition} \paragraph{Proof} Because $E_0$ has good residual eigenvalues, one can use the filtration $V^kE_0[*S]$ $=I_{S}^{k}E_0\subset E_0[*S]$ in order to compute $\mathop{\rm sp}\nolimits_SE_0[*S]$. It follows that the specialization of $E_0[*S]$ when restricted to $N_{S,X}-S$ is canonically isomorphic to the restriction of $\mathop{\rm sp}\nolimits_SE_0=\pi ^*(E_0\vert S)$ to $N_{X,S}-S$. \inter{Essential surjectivity} \begin{proposition}\label{prop3} Every Malgrange object $(M_0,M_1,C,V)$ on $(X,S)$ can be obtained in this way from a pre-${\cal D}$-module. \end{proposition} \paragraph{Proof} This follows from [Ve]: one chooses Deligne lattices in $M_0$ and $M_1$. One uses the fact that every ${\cal D}$-linear map between holonomic ${\cal D}$-modules is compatible with the $V$-filtration, so sends the specialized Deligne lattice of $M_0$ to the one of $M_1$. Moreover, the map $v$ can be obtained from $V$ because the only integral eigenvalue of $\theta$ on the Deligne lattice is $0$, so $\displaystyle{e_{}^{-2i\pi\theta}- 1\over\theta}$ is invertible on it. The previous two propositions give the following. \begin{corollary}\label{cor3} The functor from pre-${\cal D}$-modules on $(X,S)$ to regular holonomic ${\cal D}$-modules on $X$ with characteristic variety contained in $T^*_XX\cup T^*_SX$ is essentially surjective. \end{corollary} \inter{Infinitesimal rigidity} For a regular holonomic ${\cal D}$-module ${\bf M}$ with characteristic variety $T^*_XX\cup T^*_SX$, there exist several nonisomorphic pre-${\cal D}$-modules ${\bf E}$ which give rise to the Malgrange object associated to ${\bf M}$. However, we have the following infinitesimal rigidity result, which generalizes the corresponding results in [N]. \begin{proposition}[Infinitesimal rigidity]\label{prop4} Let $T=\mathop{\rm Spec}\nolimits\displaystyle{C\!\!\!\!I [\epsilon ]\over (\epsilon ^2)}$. Let ${\bf E}_T$ be a family of pre-${\cal D}$-modules on $(X,S)$ parametrized by $T$. Let the associated family ${\bf M}_T$ of ${\cal D}$-modules on $X$ be constant (pulled back from $X$). Let ${\bf E}$ (which is the specialization at $\epsilon =0$) be of the form ${\bf E} = (E,F,s,t)$ where along any component of $S$, no two distinct eigenvalues of the residue of the logarithmic connection $E$ differ by an integer. Then the family ${\bf E}_T$ is also constant. \end{proposition} \paragraph{Proof} By [N], the family $E_{T}$ is constant, as well as the specialization $\mathop{\rm sp}\nolimits_SE_{T}$. As a consequence, the residue $\theta_{E_T}$ is constant. Let us now prove that the family $F_T$ is constant. Let $S_a$ be a component of $S$ along which the only possible integral eigenvalue of $\theta_E$ is $0$. Then it is also the only possible integral eigenvalue of $\theta_F$ along $S_a$ because the generalized eigenspaces of $\theta_E$ and $\theta_F$ corresponding to a nonzero eigenvalue are isomorphic by $s$ and $t$ (see remark \ref{rem3}). We also deduce from [N] that $F_T$ is constant as a logarithmic module along this component. Assume now that $0$ is not an eigenvalue of $\theta_E$ along $S_a$ but is an eigenvalue of $\theta_F$ along this component. Then $\theta_F$ may have two distinct integral eigenvalues, one of which is $0$. Note that, in this case, $\theta_E$ is an isomorphism (along $S_a$), as well as $\theta_{E_T}$ which is obtained by pullback from $\theta_E$. It follows that on $S_a$ we have an isomorphism $F_T\simeq E_T|S_a\oplus \mathop{\rm Ker}\nolimits\theta_{F_T}$. Consequently $\mathop{\rm Ker}\nolimits\theta_{F_T}$ is itself a family. It is enough to show that this family is constant. But the corresponding meromorphic connection on $N_{S,X}^{}-S$ is constant, being the cokernel of the constant map $C_T:M_{0T}\to M_{1T}$. We can then apply the result of [N] because the only eigenvalue on $\mathop{\rm Ker}\nolimits\theta_F$ is $0$. The maps $s_T$ and $t_T$ are constant if and only if for each component $S_a$ of $S$ and for some point $x_a\in S_a$ their restriction to $F_T|{x_a}\times T$ and $E_T|{x_a}\times T$ are constant. This fact is a consequence of the following lemma. \begin{lemma} Let $E$ and $F$ be finite dimensional complex vector spaces, and let $\theta_E\in End (E)$ and $\theta_F\in End (F)$ be given. Let $V\subset W=Hom(F,E) \times Hom(E,F)$ be the closed subscheme consisting of $(s,t)$ with $st=\theta_E$ and $ts=\theta_F$. Let $\phi :W\to W$ be the holomorphic map defined by $$\phi (s,t) = (s, t {e^{st} -1 \over st}).$$ Then the differential $d\phi$ is injective on the Zariski tangent space to $V$ at any closed point $(s,t)$. \end{lemma} \paragraph{Proof} Let $(a, b)$ be a tangent vector to $V$ at $(s,t)$. Then by definition of $V$, we must have $at+sb=0$ and $ta+bs=0$. Using $at+sb=0$, we can see that $d\phi (a, b) = (a, bf(st))$ where $f$ is the entire function on $End(E_0)$ defined by the power series $(e^x-1)/x$. Suppose $(a,bf(st))=0$. Then $a=0$ and so the condition $ta+bs=0$ implies $bs=0$. As the constant term of the power series $f$ is $1$ and as $bs=0$, we have $bf(st)=b$. Hence $b=0$, and so $d\phi$ is injective. \section{Semistability and moduli for pre-${\cal D}$-modules.} In order to construct a moduli scheme for pre-${\cal D}$-modules, one needs a notion of semistability. This can be defined in more than one way. What we have chosen below is a particularly simple and canonical definition of semistability. (In an earlier version of this paper, we had employed a definition of semistability in terms of parabolic structures, in which we had to fix the ranks of $s:E_1\to E_0|S$ and $t:E_0|S \to E_1$ and a set of parabolic weights.) Let $S_a$ be the irreducible components of the smooth divisor $S\subset X$. For a pre-${\cal D}$-module ${\bf E} =(E_0,E_1,s,t)$, we denote by $E_a$ the restriction of $E_1$ to $S_a$, and we denote by $s_a$ and $t_a$ the restrictions of $s$ and $t$. \inter{Definition of semistability} We fix an ample line bundle on $X$, and denote the resulting Hilbert polynomial of a coherent sheaf $F$ by $p(F,n)$. For constructing a moduli, we fix the Hilbert polynomials of $E_0$ and $E_a$, which we denote by $p_0(n)$ and $p_a(n)$. Recall (see [S]) that an ${\cal O} _X$-coherent ${\cal D} _X[\log S]$-module $F$ is by definition {\sl semistable} if it is pure dimensional, and for each ${\cal O} _X$ coherent ${\cal D} _X[\log S]$ submodule $F'$, we have the inequality $p(F',n)/rank (F') \le p(F,n)/rank (F) $ for large enough $n$. We call $p(F,n)/rank (F)$ the {\sl normalized Hilbert polynomial} of $F$. \begin{definition}\rm We say that the pre-${\cal D}$-module ${\bf E}$ is {\sl semistable} if the ${\cal D}_X[\log S]$-modules $E_0$ and $E_a$ are semistable. \end{definition} \refstepcounter{theorem}\paragraph{Remarks \thetheorem} (1) It is easy to prove that the semistability of the ${\cal D} _X[\log S]$-module $E_a$ is equivalent to the semistability of the logarithmic connection $\pi ^*_a(E_a)$ on $P(N_{S_a,X}\oplus 1)$ with respect to a natural choice of polarization. (2) When $X$ is a curve, a pre-${\cal D}$-module ${\bf E}$ is semistable if and only if the logarithmic connection $E_0$ on $(X,S)$ is semistable, for then $E_1$ is always semistable. (3) Let the dimension of $X$ be more than one. Then even when a pre-${\cal D}$-module ${\bf E}$ is a pre meromorphic connection (equivalently, when $s:E_1 \to E_0\vert S$ is an isomorphism), the definition of semistability of ${\bf E}$ does not reduce to the semistability of the underlying logarithmic connection $E_0$ on $(X,S)$. This is to be expected because we do not fix the rank of $s$ (or $t$) when we consider families of pre-${\cal D}$-modules. Also note that even in dimension one, meromorphic connections are not a good subcategory of the abelian category of all regular holonomic ${\cal D}$-modules with characteristic variety contained in $T^*_XX\cup T^*_SX$, in the sense that a submodule or a quotient module of a meromorphic connection is not necessarily a meromorphic connection. \inter{Boundedness and local universal family} We let the index $i$ vary over $0$ and over the $a$. \begin{proposition}[Boundedness] Semistable pre-${\cal D}$-modules with given Hilbert po\-lynomials $p_i$ form a bounded set, that is, there exists a family of pre-${\cal D}$-modules parametrized by a noetherian scheme of finite type over $C\!\!\!\!I$ in which each semistable pre-${\cal D}$-module with given Hilbert polynomials occurs. \end{proposition} \paragraph{Proof} This is obvious as each $E_i$ (where $i=0,a$) being semistable with fixed Hilbert polynomial, is bounded. Next, we construct a local universal family. By boundedness, there exists a positive integer $N$ such that for all $n\ge N$, the sheaves $E_0(N)$ and $E_1(N)$ are generated by global sections and have vanishing higher cohomology. Let $\Lambda = D_X[\log S]$. Let ${\cal O} _X =\Lambda_0 \subset \Lambda_1 \subset \cdots \subset \Lambda$ be the increasing filtration of $\Lambda$ by the order of the differential operators. Note that each $\Lambda_k$ is an ${\cal O}_X$ bimodule, coherent on each side. Let $r$ be a positive integer larger than the ranks of the $E_i$. Let $Q_i$ be the quot scheme of quotients $q_i:\Lambda_r\otimes {\cal O}_X (-N)^{p_i(N)}\to\!\!\!\!\to E_i$ where the right ${\cal O}_X$-module structure on $\Lambda_r$ is used for making the tensor product. Note that $G_i=PGL(p_i(N))$ has a natural action on $Q_i$. Simpson defines a locally closed subscheme $C_i\subset Q_i$ which is invariant under $G_i$, and a local universal family $E$ of $\Lambda$-modules parametrized by $C_i$ with the property that for two morphisms $T\to C_i$, the pull back families are isomorphic over an open cover $T'\to T$ if and only if the two morphisms define $T'$ valued points of $C_i$ which are in a common orbit of $G_i(T')$. On the product $C_0\times C_a$, consider the linear schemes $A_a$ and $B_a$ which respectively correspond to $Hom_{\Lambda}(E_1,E_0)$ and $Hom_{\Lambda}(E_0,E_1)$ (see Lemma 2.7 in [N] for the existence and universal property of such linear schemes). Let $F_a$ be the fibered product of $A_a$ and $B_a$ over $C_0\times C_a$. Let $H_a$ be the closed subscheme of $F_a$ where the tuples $(q_0,q_1,t,s)$ satisfy $st=\theta$ and $ts=\theta$. Finally let $H$ be the fibered product of the pullbacks of the $H_a$ to $C= C_0 \times \prod_a C_a$. Note that $H$ parametrizes a tautological family of pre-${\cal D}$-modules on $(X,S)$ in which every semistable pre-${\cal D}$-module with given Hilbert polynomials occurs. The group $${\cal G} = G_0 \times \prod_a (G_a \times GL(1))$$ has a natural action on $H$, with $$(q_0,q_a,t_a,s_a)\cdot (g_0,g_a,\lambda_a) = (q_0g_0,q_ag_a,(1/\lambda_a)t_a,\lambda_a s_a)$$ It is clear from the definitions of $H$ and this action that two points of $H$ parametrise isomorphic pre-${\cal D}$-modules if and only if they lie in the same $G$ orbit. The morphism $H\to C\times \prod _aC_a$ is an affine morphism which is ${\cal G}$-equivariant, and by Simpson's construction of moduli for $\Lambda$-modules, the action of ${\cal G}$ on $C\times \prod _aC_a$ admits a good quotient in the sense of geometric invariant theory. Hence a good quotient $H//{\cal G}$ exists by Ramanathan's lemma (see Proposition 3.12 in [Ne]), which by construction and universal properties of good quotients is the coarse moduli scheme of semistable pre-${\cal D}$-modules with given Hilbert polynomials. Note that under a good quotient in the sense of geometric invariant theory, two different orbits can in some cases get mapped to the same point (get identified in the quotient). In the rest of this section, we determine what are the closed points of the quotient $H//{\cal G}$. \refstepcounter{theorem}\paragraph{Remark \thetheorem} For simplicity of notation, we assume in the rest of this section that $S$ has only one connected component. It will be clear to the reader how to generalize the discussion when $S$ has more components. \inter{Reduced modules} Assuming for simplicity that $S$ has only one connected component, so that ${\cal G} = {\cal H} \times GL(1)$ where $H=G_0 \times G_1$, we can construct the quotient $H//{\cal G}$ in two steps: first we go modulo the factor $GL(1)$, and then take the quotient of $R=H//GL(1)$ by the remaining factor ${\cal H}$. The following lemma is obvious. \begin{lemma}\label{lem4.5} Let $T$ be a scheme of finite type over $k$, and let $V\to T$ and $W\to T$ be linear schemes over $T$. Let $V\times W$ be their fibered product (direct sum) over $T$, and let $V\otimes W$ be their tensor product. Let $\phi :V\times W\to V\otimes W$ be the tensor product morphism. Then its schematic image $D\subset V\otimes W$ is a closed subscheme which (i) parametrizes all decomposable tensors, and (ii) base changes correctly. Let $GL(1,k)$ act on $V\times W$ by the formula $\lambda \cdot (v,w) = (\lambda v, (1/\lambda )w)$. Then $\phi :V\times W\to D$ is a good quotient for this action. \end{lemma} \paragraph{Proof} The statement is local on the base, so we can assume that (i) the base $T$ is an affine scheme, and (ii) both the linear schemes are closed linear subschemes of trivial vector bundles on the base, that is, $V\subset A^m_T$ and $W\subset A^n_T$ are subschemes defined respectively by homogeneous linear equations $f_p(x_i)=0$ and $g_q(y_j)=0$ in the coordinates $x_i$ on $A^m_T$ and $y_j$ on $A^n_T$. Let $z_{i,j}$ be the coordinates on $A^{mn}_T$, so that the map $\otimes :A^m_T\times _T A^n_T \to A^{mn}$ sends $(x_i,y_j) \mapsto (z_{i,j})$ where $z_{i,j}=x_iy_j$. Its schematic image is the subscheme $C$ of $A^{mn}_T$ defined by the equations $z_{a,b}z_{c,d} - z_{a,d}z_{b,c} = 0$, that is, the matrix $(z_{i,j})$ should have rank $ < 2$. Take $D$ to be the subscheme of $C$ defined by the equations $f_p(z_{1,j},\ldots ,z_{m,j}) = 0$ and $g_q(z_{i,1},\ldots ,z_{i,n}) = 0$. Now the lemma \ref{lem4.5} follows trivially from this local coordinate description. \paragraph{} The above lemma implies the following. To get the quotient $H//GL(1)$, we just have to replace the fibered product $A\times B$ over $C_0\times C_1$ by the closed subscheme $Z\subset D\subset A\otimes B$, where $D$ is the closed subscheme consisting of decomposable tensors $u$, and $Z$ is the closed subscheme of $D$ defined as follows. Let $\mu _0$ and $\mu _1$ be the canonical morphisms from $A\otimes B$ to the linear schemes representing $End_{\Lambda} (E_0|S)$ and $End_{\Lambda} (E_1)$ respectively. Then $Z$ is defined to consist of all $u$ such that $\mu _0(u)=\theta \in End _{\Lambda}(E_0|S) $ and $\mu _1(u) = \theta \in End_{\Lambda}(E_1)$. There is a canonical $GL(1)$ quotient morphism $A\times B \to D$ over $C_0\times C_1$, which sends $(s,t)\mapsto u=s\otimes t$. These give the $GL(1)$ quotient map $H\to Z$. Note that the map $H\to C_0\times C_1$ is ${\cal G}$ equivariant, and the action of $GL(1)$ on $C_0\times C_1$ is trivial, so we get a ${\cal H}$-equivariant map $Z\to C_0\times C_1$. In order to describe the identifications brought about by the above quotient, we make the following definition. \begin{definition}\rm A {\sl reduced module} is a tuple $(E_0,E_1,u)$ where $E_0$ and $E_1$ are as in a pre-${\cal D}$-module, and $u\in Hom_{\Lambda}(E_1,E_0|S)\otimes Hom_{\Lambda}(E_0,E_1)$ is a decomposable tensor, such that the canonical maps $\mu _0:Hom_{\Lambda}(E_1,E_0|S)\otimes Hom_{\Lambda}(E_0,E_1) \to End_{\Lambda}(E_0|S)$ and $\mu _1: Hom_{\Lambda}(E_1,E_0|S)\otimes Hom_{\Lambda}(E_0,E_1) \to End_{\Lambda}(E_1)$, both map $u$ to the endomorphism $\theta$ of the appropriate module. In other words, there exist $s$ and $t$ such that $(E_0,E_1,s,t)$ is a pre-${\cal D}$-module, and $u=s\otimes t$. We say that the reduced module $(E_0,E_1,s\otimes t)$ is the associated reduced module of the pre-${\cal D}$-module $(E_0,E_1,s,t)$. Moreover, we say that a reduced module is semistable if it is associated to a semistable pre-${\cal D}$-module. \end{definition} \begin{lemma} Let $V$ and $W$ are two vector spaces, $v,v'\in V$ and $w,w'\in W$, then (1) If $v\otimes w=0$ then $v=0$ or $w=0$. (2) If $v\otimes w=v'\otimes w'\ne 0$, then there exists a scalar $\lambda \ne 0$ such that $v=\lambda v'$ and $w = (1/\lambda ) w'$. \end{lemma} \refstepcounter{theorem}\paragraph{Remark \thetheorem} The above lemma shows that if ${\bf E}$ and ${\bf E} '$ are two non-isomorphic pre-${\cal D}$-modules whose associated reduced modules are isomorphic, then we must have $s\otimes t =s'\otimes t'=0$. In particular, $\theta$ will act by zero on $E_0|S$ and also on $E_1$ for such pre-${\cal D}$-modules as $st=0$ and $ts=0$. \inter{S-equivalence and stability} \begin{definition}\rm By a {\sl filtration} of a logarithmic connection $E$ we shall mean an increasing filtration $E_p$ indexed by $Z\!\!\!Z$ by subvector bundles which are logarithmic connections. Similarly, a filtration on a ${\cal D} _X[\log S]$-module $F$ supported on $S$ will mean a filtration of the vector bundle $F\vert S$ by subbundles $F_p$ which are ${\cal D} _X[\log S]$-submodules. A filtration of a pre-${\cal D}$-module $(E_0,E_1,s,t)$ is an increasing filtration $(E_i)_p$ of the logarithmic connection $E_i$ ($i=0,1$) such that $s$ and $t$ are filtered morphisms with respect to the specialized filtration of $E_0$ and the filtration of $E_1$. A filtration of a reduced module $(E_0,E_1,u)$, with $u=s\otimes t$ where we take $s=0$ and $t=0$ if $u=0$, is a filtration of the pre-${\cal D}$-module $(E_0,E_1,s,t)$. We shall always assume that this filtration is exhaustive, that is, $(E_i)_p=0$ for $p\ll0$ and $(E_i)_p=E_i$ for $p\gg0$. A filtration is {\sl nontrivial} if some $(E_i)_p$ is a proper subbundle of $E_i$ for $i=0$ or $1$. \end{definition} For a filtered pre-${\cal D}$-module (or reduced module), each step of the filtration as well as the graded object are pre-${\cal D}$-modules (or reduced modules). \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{deform} There is a natural family $({\bf E}_\tau)_{\tau\in A^1}^{}$ of pre-${\cal D}$-modules or reduced modules parametrized by the affine line $A^1=\specC\!\!\!\!I[\tau]$, which fibre at $\tau=0$ is the graded object ${\bf E}'$ and the fibre at $\tau_0\neq0$ is isomorphic to the original pre-${\cal D}$-module or reduced module ${\bf E}$: put (for $i=0,1$) ${\cal E}_i=\oplus_{p\inZ\!\!\!Z}^{}(E_i)_p\tau^p\subset E_i\otimes C\!\!\!\!I[\tau,\tau_{}^{-1}]$ and the relative ${\cal D}\log$-structure is the natural one. \begin{definition}\rm A {\sl special filtration of a coherent ${\cal O} _X$-module} $E$ is a filtration for which each $E_p$ has the same normalized Hilbert polynomial as $E$. A {\sl special filtration of a reduced module} $(E_0,E_1,u)$ is a filtration of this reduced module which is special on $E_0$ and on $E_1$. \end{definition} The graded reduced module ${\bf E}'$ associated with a special filtration of a semistable reduced module ${\bf E}$ is again semistable. \begin{definition}\rm The equivalence relation on the set of isomorphism classes of all semistable reduced modules generated by this relation (by which ${\bf E} '$ is related to ${\bf E}$) will be called S-equivalence. \end{definition} \begin{definition}\label{defstable}\rm We say that a semistable reduced module is {\sl stable} if it does not admit any nontrivial special filtration. \end{definition} \refstepcounter{theorem}\paragraph{Remarks \thetheorem} (1) Note in particular that if each $E_0$, $E_a$ is stable as a $\Lambda$-module, then the reduced module ${\bf E} $ is stable. Consequently we have the following. Though the definition of stability depends on the ample line bundle $L$ on $X$, irrespective of the choice of the ample bundle, for any pre-${\cal D}$-module such that the monodromy representation of $E_0\vert (X-S)$ is irreducible, and the monodromy representation of $\pi _a ^*E_a \vert (N_{S_a,X}-S_a)$ is irreducible for each component $S_a$, the corresponding reduced module is stable. The converse is not true -- a pre-${\cal D}$-module, whose reduced module is stable, need not have irreducible monodromies. The example 2.4.1 in [N] gives a logarithmic connection, whose associated pre-${\cal D}$-module in which $s$ is identity and $t$ is the residue, gives a stable reduced module, but the monodromies are not irreducible. (2) If $u=0$, the reduced module is stable if and only $E_0$ and each $E_a$ is stable. (3) When $X$ is a curve, a reduced module with $u\ne 0$ is stable if and only if the logarithmic connection $E_0$ is stable. If $u=0$, each $E_a$ must moreover have length at most one as an ${\cal O}_X$-module. Hence over curves, there is a plentiful supply of stable reduced modules. \begin{lemma}\label{uisflat} Let $(E_0,E_1,u)$ be a reduced module and let $(E_i)_p$ be filtrations of $E_i$ ($i=0,1$). Then $s$ and $t$ are filtered morphisms with respect to the specialized filtrations if and only if there exists some point $P\in S$ such that the restrictions of $s$ and $t$ to the fibre $E_{i,P}$ at $P$ are filtered with respect to the restricted filtrations. \end{lemma} \paragraph{Proof} This follows from the fact that if a section $\sigma$ of a vector bundle with a flat connection has a value $\sigma(P)$ in the fibre at $P$ of a sub flat connection, then it is a section of this subbundle: we apply this to $s$ (resp. $t$) as a section of $Hom((E_0)_{p|S},(E_1)_{|S})$ (resp. $Hom((E_1)_{p|S},(E_0)_{|S})$). \inter{A criterion for stability} Let ${\bf E}=(E_0,E_1,u=s\otimes t)$ be a reduced module. Assume that we are given filtrations $0=F_0(E_i)\subset F_1(E_i)\subset\cdots\subset F_{\ell_i}(E_i)=E_i$ of $E_i$ ($i=0,1$) by vector subbundles which are ${\cal D}_X[\log S]$-submodules. For $j=0,\ldots,\ell_i$ let $k(j)$ be the smallest $k$ such that $s(\mathop{\rm sp}\nolimits_SF_j(E_0))\subset F_k(E_1)$ and let $J(s)$ be the graph of the map $j\to k(j)$. A {\sl jump point} is a point $(j,k(j))$ on this graph such that $k(j-1)<k(j)$. Consider also the set $G_s$ made by points under the graph: $G_s=\{ (j,k)\mid k\leq k(j)\}$. For $t$ there is an equivalent construction: we have a map $k\to j(k)$ and a set $G_t$ on the left of the graph $I(t)$: $G_t=\{ (j,k)\mid j\leq j(k)\}$. \begin{definition}\rm $u=s\otimes t$ is {\sl compatible} with the filtrations if the two sets $G_s$ and $G_t$ intersect at most at (common) jump points (where if $u=0$, take $s=0$ and $t=0$). \end{definition} \begin{proposition}\label{nonstable} Let ${\bf E}=(E_0,E_1,u)$ be a semistable reduced module. The following conditions are equivalent: (1) ${\bf E}$ is not stable, (2) there exists a nontrivial special filtration $F_j(E_i)$ ($j=0,\ldots\ell_i$) of each $E_i$ where all inclusions are proper and $u$ is compatible with these filtrations. \end{proposition} \paragraph{Proof} $(1)\Rightarrow(2)$: If ${\bf E}$ is not stable, we can find two nontrivial special filtrations $(E_0)_p$ and $(E_1)_q$ such that $s$ and $t$ are filtered morphisms. Let $p_j$ ($j=1,\ldots ,\ell_0$) be the set of jumping indices for $(E_0)_p$ and $q_k$ ($k=1,\ldots ,\ell_1$) for $(E_1)_q$. For each $j_0$ and $k_0$ we have $j(k(j_0))\leq j_0$ and $k(j(k_0))\leq k_0$. We define $F_j(E_0)=(E_0)_{p_j}$ and $F_k(E_1)=(E_1)_{q_k}$. We get nontrivial filtrations of $E_0$ and $E_1$ where all inclusions are proper. Moreover there cannot exist two distinct points $(j_0,k(j_0))$ and $(j(k_0),k_0)$ with $j_0\leq j(k_0)$ and $k_0\leq k(j_0)$ otherwise we would have $j_0\leq j(k_0)\leq j(k(j_0))\leq j_0$ and the same for $k_0$ so the two points would be the equal. Consequently $u$ is compatible with these filtrations. $(2)\Rightarrow(1)$: We shall construct a special filtration $((E_0)_p,(E_1)_q)$ of the reduced module from the filtrations $F_j(E_i)$ of each $E_i$. Choose a polygonal line with only positive slopes, going through each jump point of $G_s$ and for which each jump point of $G_t$ is on or above this line (see figure \ref{fig1}). \setlength{\unitlength}{.5truecm} \begin{figure}[htb] \begin{center} \begin{picture}(10,8)(0,0) \put(0,0){\line(1,0){10}} \put(0,0){\line(0,1){8}} \put(9.5,-.7){$j$} \put(-.5,7.5){$k$} \put(3,4){\circle*{.2}} \put(6,6){\circle*{.2}} \put(8,8){\circle*{.2}} \put(3,4){\line(1,0){3}} \put(6,6){\line(1,0){2}} \put(8,8){\line(1,0){2}} \put(3,0){\line(0,1){4}} \put(6,4){\line(0,1){2}} \put(8,6){\line(0,1){2}} \put(1,1){\circle{.2}} \put(2,3){\circle{.2}} \put(3,4){\circle{.2}} \put(5,5){\circle{.2}} \put(7,7){\circle{.2}} \put(0,1){\line(1,0){1}} \put(1,3){\line(1,0){1}} \put(2,4){\line(1,0){1}} \put(3,5){\line(1,0){2}} \put(5,7){\line(1,0){2}} \put(1,1){\line(0,1){2}} \put(2,3){\line(0,1){1}} \put(3,4){\line(0,1){1}} \put(5,5){\line(0,1){2}} \put(7,7){\line(0,1){1}} \put(6,3){$G_s$} \put(2,6){$G_t$} \multiput(.2,.2)(.2,.2){4}{\circle*{.1}} \multiput(1,1)(.2,.3){10}{\circle*{.1}} \multiput(3,4)(.4,.2){5}{\circle*{.1}} \multiput(5,5)(.2,.2){15}{\circle*{.1}} \end{picture} \caption{\label{fig1}$\bullet=$ jump points of $s$, $\circ=$ jump points of $t$} \end{center} \end{figure} Choose increasing functions $p(j)$ and $q(k)$ such that $p(j)-q(k)$ is identically $0$ on this polygonal line, is $<0$ above it and $>0$ below it (for instance, on each segment $[(j_0,k_0),(j_1,k_1)]$ of this polygonal line, parametrised by $j=j_0+m\varepsilon_1$, $k=k_0+m\varepsilon_2$, put $p(j)=p(j_0)+\varepsilon_2(j-j_0)$ and $q(k)=q(k_0)+\varepsilon_1(k-k_0)$, and $p(0)=q(0)=0$). For $p(j)\leq p<p(j+1)$ put $(E_0)_p=F_j(E_0)$ and for $q(k)\leq q<q(k+1)$ put $(E_1)_q=F_k(E_1)$. The filtration $((E_0)_p,(E_1)_q,u)$ is then a nontrivial special filtration of the reduced module ${\bf E}$. \begin{proposition}\label{open} Semistability and stability are Zariski open conditions on the parameter scheme of any family of reduced modules. \end{proposition} \paragraph{Proof} As semistability is an open condition on ${\cal D} _X[\log S]$-modules, it follows it is an open condition on reduced modules. Now, for any family of semistable reduced modules parametrised by a scheme $T$, all possible special filtrations of the form given by \ref{nonstable} on the specializations of the family are parametrised by a scheme $U$ which is projective over $T$. The image of $U$ in $T$ is the set of non stable points in $T$, hence its complement is open. \inter{Points of the moduli} We are now ready to prove the following theorem. \begin{theorem} Let $X$ be a projective variety together with an ample line bundle, and let $S\subset X$ be a smooth divisor. (1) There exists a coarse moduli scheme ${\cal P}$ for semistable pre-${\cal D}$-modules on $(X,S)$ with given Hilbert polynomials $p_i$. The scheme ${\cal P}$ is quasiprojective, in particular, separated and of finite type over $C\!\!\!\!I$. (2) The points of ${\cal P}$ correspond to S-equivalence classes of semistable pre-${\cal D}$-modules. (3) The S-equivalence class of a semistable reduced module ${\bf E} $ equals its isomorphism class if and only if ${\bf E} $ is stable. (4) ${\cal P}$ has an open subscheme ${\cal P} ^s$ whose points are the isomorphism classes of all stable reduced modules. This is a coarse moduli for (isomorphism classes of) stable reduced modules. \end{theorem} \paragraph{Proof} Let ${\cal P} = H//{\cal G}$. Then (1) follows by the construction of ${\cal P}$. To prove (2), first note that by the existence of the deformation ${\bf E} _t$ (see \ref{deform}) of any reduced module ${\bf E}$ corresponding to a weighted special filtration, and by the separatedness of ${\cal P}$, the reduced module ${\bf E} $ and its limit ${\bf E} '$ go to the same point of ${\cal P}$. Hence an S-equivalence class goes to a common point of ${\cal P}$. For the converse, first recall that ${\cal G} = {\cal H} \times GL(1)$, and the quotient ${\cal P}$ can be constructed in two steps: ${\cal P} = R//{\cal H}$ where $R=H/{\cal G}$. The scheme $R$ parametrizes a canonical family of reduced modules. Let the ${\cal H}$ orbit of a point $x$ of $R$ corresponding the reduced module ${\bf E} $ not be closed in $R$. Let $x_0$ be any of its limit points. Then there exists a 1-parameter subgroup $\lambda$ of ${\cal H}$ such that $x_0 = \lim _{t\to 0} \lambda (t) x$. This defines a map from the affine line $A^1$ to $R$, which sends $t\mapsto \lambda (t)x$. Let ${\bf E} _t$ be the pull back of the tautological family of reduced modules parametrized by $R$. Then from the description of the limits of the actions of 1-parameter subgroups on a quot scheme given in section 1 of Simpson [S], it follows that ${\bf E}$ has a special filtration such that the family ${\bf E} _t$ is isomorphic to a deformation of the type constructed in \ref{deform} above. Hence the reduced modules parametrized by $x$ and $x_0$ are S-equivalent. This proves (2). If the orbit of $x$ is not closed, then it has a limit $x_0$ outside it under a 1-parameter subgroup, which by above represents a reduced module ${\bf E} '$ which is the limit of ${\bf E} $ under a special filtration. As by assumption ${\bf E} '$ is not isomorphic to ${\bf E} $, the special filtration must be nontrivial. Hence ${\bf E} $ is not stable. Hence stable points have closed orbits in $R$. If $x$ represents a stable reduced module, then $x$ cannot be the limit point of any other orbit. For, if $x$ is a limit point of the orbit of $y$, then by openness of stability (see \ref{open}), $y$ should again represent a stable reduced module. But then by above, the orbit of $y$ is closed. This proves (3). Let $H^s\subset H$ be the open subscheme where the corresponding pre-${\cal D}$-module is stable. By (2) and (3) above, $H^s$ is saturated under the quotient map $H \to {\cal P}$, hence by properties of a good quotient, its image ${\cal P} ^s$ is open in ${\cal P}$. Moreover by (2) and (3) above, $H^s$ is the inverse image of ${\cal P}^ s$. Hence $H^s \to {\cal P} ^s$ is a good quotient, which again by (2) and (3) is an orbit space. Hence points of ${\cal P} ^s$ are exactly the isomorphism classes of stable reduced modules, which proves (4). \section{Perverse sheaves, Verdier objects and finite descriptions} Let $X$ be a nonsingular projective variety and let $S$ be a smooth divisor. The abelian category of perverse sheaves constructible with respect to the stratification $(X-S,S)$ of $X$ is equivalent to the category of `Verdier objects' on $(X,S)$. Before defining this category, let us recall the notion of specialization along $S$. Let ${\cal E}$ be a local system (of finite dimensional vector spaces) on $X-S$. The {\sl specialization} $\mathop{\rm sp}\nolimits_S{\cal E}$ is a local system (of the same rank) on $N_{S,X}^{}-S$ equipped with an endomorphism $\tau_{\cal E}$. It is constructed using the nearby cycle functor $\psi$ defined by Deligne applied to the morphism which describes the canonical deformation from $X$ to the normal bundle $N_{S,X}^{}$. A local system ${\cal F}$ on $N_{S,X}^{}-S$ equipped with an endomorphism $\tau_{\cal F}$ is said to be {\sl monodromic} if $\tau_{\cal F}$ is equal to the monodromy of ${\cal F}$ around $S$. Then $\mathop{\rm sp}\nolimits_S{\cal E}$ is monodromic. \begin{definition}\rm A {\sl Verdier object} on $(X,S)$ is a tuple ${\bf V}=({\cal E},{\cal F},C,V)$ where (1) ${\cal E}$ is a local system on $X-S$, (2) ${\cal F}$ is a monodromic local system on $N_{S,X}^{}-S$, (3) $C:\mathop{\rm sp}\nolimits_S{\cal E}\to{\cal F}$ and $V:{\cal F}\to\mathop{\rm sp}\nolimits_S{\cal E}$ are morphisms of (monodromic) local systems on $N_{S,X}^{}-S$ satisfying (4) $CV=\tau_{\cal F}-\mathop{\rm id}\nolimits$ and $VC=\tau_{\cal E}-\mathop{\rm id}\nolimits$. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem} The morphisms between Verdier objects on $(X,S)$ are defined in an obvious way, and the category of Verdier objects is an abelian category in which each object has finite length. Hence the following definition makes sense. \begin{definition}\rm We say that two Verdier objects are {\sl S-equivalent} if they admit Jordan-H\"older filtrations such that the corresponding graded objects are isomorphic. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem} Let $B$ be a tubular neighbourhood of $S$ in $X$, diffeomorphic to a tubular neighbourhood of $S$ in $N_{S,X}^{}$. Put $B^*=B-S$. The specialized local system $\mathop{\rm sp}\nolimits_S{\cal E}$ can be realized as the restriction of ${\cal E}$ to $B^*$, its monodromy $\tau_{\cal E}$ at some point $x\in B^*$ being the monodromy along the circle normal to $S$ going through $x$. Hence a Verdier object can also be described as a tuple ${\bf V}$ where ${\cal F}$ is a local system on $B^*$ and $C$, $V$ are morphisms between ${\cal E}|B^*$ and ${\cal F}$ subject to the same condition (4). \bigskip The notion of a family of perverse sheaves is not straightforward. We can however define the notion of a family of Verdier objects. Let us define first a family of local systems on $X-S$ (or on $N_{S,X}^{}-S$) parametrized by a scheme $T$. This is a locally free $p^{-1}{\cal O}_T$-module of finite rank, where $p$ denotes the projection $X-S\times T\to T$. Morphisms between such objects are $p^{-1}{\cal O}_T$-linear. The notion of a family of Verdier objects is then straightforward. In order make a moduli space for Verdier objects, we shall introduce the category of `finite descriptions' on $(X,S)$. Let us fix the following data (D): (D1) finitely generated groups $G$ and $G_a$ for each component $S_a$ of $S$, (D2) for each $a$ an element $\tau_a$ which lies in the center of $G_a$ and a group homomorphism $\phi_a:G_a\to G$. \begin{definition}\label{def2}\rm A finite description ${\bf D}$ (with respect to the data (D)) is a tuple $(E,\rho,F_a,\rho_a,C_a,V_a)$ where (1) $\rho:G\to GL(E)$ is a finite dimensional complex representation of the group $G$; for each $a$ we will regard $E$ as a representation of $G_a$ via the homomorphism $\phi_a:G_a\to G$; (2) for each $a$, $\rho_a:G_a\to GL(F_a)$ is a finite dimensional complex representation of the group $G$; (3) for each $a$, $C_a:E\to F_a$ and $V_a: F_a\to E$ are $G_a$-equivariant linear maps such that $V_aC_a=\rho(\tau_a)-\mathop{\rm id}\nolimits$ in $GL(E)$ and $C_aV_a=\rho_a(\tau_a)-\mathop{\rm id}\nolimits$ in $GL(F_a)$. \end{definition} A morphism between two finite descriptions has an obvious definition. \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem4} Let $P_0\in X-S$ and let $P_a$ be a point in the component $B*_a$ of $B^*$. Choose paths $\sigma_a:[0,1]\to X-S$ with $\sigma_a(0)=P_0$ and $\sigma_a(1)=P_a$. Let $G$ be the fundamental group $\pi_1(X-S,P_0)$, and let $G_a = \pi_1 (B^*_a,P_a )$. Let $\tau_a \in G_a $ be the positive loop based at $P_a $ in the fiber of $B^*_a\to S_a $. Finally, let $\phi_a:G_a\to G$ be induced by the inclusion $B^*_a\hookrightarrow X-S$ by using the path $\sigma_a$ to change base points. Then, under the equivalence between representations of fundamental group and local system, the category of finite description with respect to the previous data is equivalent to the category of Verdier objects on $(X,S)$. \refstepcounter{theorem}\paragraph{Remark \thetheorem} The category of finite descriptions is an abelian category in which each object has finite length. Therefore the notion of S-equivalence as in definition 5.3 above makes sense for finite descriptions. \begin{definition}\rm A family of finite descriptions parametrized by a scheme $T$ is a tuple $(E_T, \rho_T,F_{T,a}, \rho_{T,a}, C_{T,a}, V_{T,a})$ where $E_T$ and the $F_{T,a}$ are locally free sheaves on $T$, $\rho$ and $\rho_{T,a}$ are families of representations into these, and the $C_{T,a}$ and $V_{T,a}$ are ${\cal O}_T$-homomorphisms of sheaves satisfying the analogues of condition \ref{def2}.3 over $T$. The pullback of a family under a morphism $T'\to T$ is defined in an obvious way, giving a fibered category. Let $PS$ denote the corresponding groupoid. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem} It can be checked (we omit the details) that the groupoid $PS$ is an Artin algebraic stack. \section{Moduli for perverse sheaves} Let us fix data (D) as above. \begin{theorem} There exists an affine scheme of finite type over $C\!\!\!\!I$, which is a coarse moduli scheme for finite descriptions ${\bf D}=(E,\rho,F_a,\rho_a,C_a,V_a)$ relative to {\rm (D)} with fixed numerical data $n=\dim E$ and $n_a=\dim F_a$. The closed points of this moduli scheme are the S-equivalence classes of finite descriptions with given numerical data $(n,n_a)$. \end{theorem} Using remark \ref{rem4} we get \begin{corollary} There exists an affine scheme of finite type over $C\!\!\!\!I$, which is a coarse moduli scheme for Verdier objects ${\bf V}=({\cal E},{\cal F},C,V)$ (or perverse sheaves on $(X,S)$) with fixed numerical data $n={\rm rank} {\cal E}$ and $n_a={\rm rank}{\cal F}|B^*_a$. The closed points of this moduli scheme are the S-equivalence classes of Verdier objects with given numerical data $(n,n_a)$. \end{corollary} The above corollary and its proof does not need $X$ to be a complex projective variety, and the algebraic structure of $X$ does not matter. All that is needed is that the fundamental group of $X-S$ and that of each $S_a$ is finitely generated. The rest of this section contains the proof of the above theorem. \begin{proposition}\label{prop5} (1) Let ${\bf D}$ be a finite description, and let $\mathop{\rm gr}\nolimits({\bf D})$ be its semisimplification. Then there exists a family ${\bf D}_T$ of finite descriptions parametrized by the affine line $T=A^1$ such that the specialization ${\bf D}_0$ at the origin $0\in T$ is isomorphic to $\mathop{\rm gr}\nolimits({\bf D})$, while ${\bf D}_t$ is isomorphic to ${\bf D}$ at any $t\ne 0$. (2) In any family of finite descriptions parametrized by a scheme $T$, each S-equiva\-len\-ce class (Jordan-H\"older class) is Zariski closed in $T$. \end{proposition} \paragraph{Proof} The statement (1) has a proof by standard arguments which we omit. To prove (2), first note that if ${\bf D}_T$ is any family and ${\bf D}'$ a simple finite description, then the condition that ${\bf D}' \times \{ t \}$ is a quotient of ${\bf D}_t$ defines a closed subscheme of $T$. From this, (2) follows easily. \paragraph{Construction of Moduli} Let $E$ and $F_a$ be vector spaces with $\dim(E)=n$ and $\dim(F_a) =n_a$. Let $\cal R$ be the affine scheme of all representations $\rho$ of $G$ in $E$, made as follows. Let $h_1,\ldots , h_r$ be generators of $G$. Then $\cal R$ is the closed subscheme of the product $GL(E)^r$ defined by the relations between the generators. Similarly, choose generators for each $G_a$, and let ${\cal R}_i$ be the corresponding affine scheme of all representations $\rho_a$ of $G_a$ in $F_a$. Let $$A \subset {\cal R} \times \prod_a ({\cal R}_a \times Hom(E,F_a) \times Hom(F_a,E))$$ be the closed subscheme defined by condition \ref{def2}.3 above. Its closed points are tuples $(\rho,\rho_a, C_a, V_a)$ where the linear maps $C_a:E\to F_a$ and $V_a:F_a\to E$ are $G_a$-equivariant under the representations $\rho \phi_a: G_a\to GL(E)$ and $\rho_a: G_a\to GL(F_a)$, and satisfy $V_aC_a = \rho (\tau_a )-1$ in $GL(E)$, and $C_aV_a = \rho_a(\tau_a) -1$ in $GL(F_a)$ for each $a$. The product group ${\cal G} =GL(E) \times (\prod_a GL(F_a))$ acts on the affine scheme $A$ by the formula $$(\rho ,\rho_a, C_a,V_a)\cdot (g,g_a) = (g^{-1}\rho g, g_a^{-1}\rho_a g_a,g_a^{-1}C_ag, g^{-1}V_ag_a).$$ The orbits under this action are exactly the isomorphism classes of finite descriptions. The moduli of finite descriptions is the good quotient ${\cal F} =A//{\cal G}$, which exists as $A$ is affine and ${\cal G}$ is reductive. It is an affine scheme of finite type over $C\!\!\!\!I$. It follows from \ref{prop5}.1 and \ref{prop5}.2 and properties of a good quotient that the Zariski closures of two orbits intersect if and only if the two finite descriptions are S-equivalent. Hence closed points of ${\cal F}$ are S-equivalence classes (Jordan-H\"older classes) of finite descriptions. \section{Riemann-Hilbert morphism} To any Malgrange object ${\bf M}$, there is an obvious associated Verdier object ${\bf V}({\bf M})$ obtained by applying the de~Rham functor to each component of ${\bf M}$. This defines a functor, which is in fact an equivalence of categories from Malgrange objects to Verdier objects. We have already defined a functor $\eta$ from pre-${\cal D}$-modules with good residual eigenvalues to Malgrange objects. Composing, we get an exact functor from pre-${\cal D}$-modules with good residual eigenvalues to Verdier objects. Choosing base points in $X$ and paths as in remark \ref{rem4} we get an exact functor ${\cal R\!\!H}$ from pre-${\cal D}$-modules to finite descriptions. This construction works equally well for families of pre-${\cal D}$-modules, giving us a holomorphic family ${\cal R\!\!H} ({\bf E}_T)$ of Verdier objects (or finite descriptions) starting from a holomorphic family ${\bf E}_T$ of pre-${\cal D}$-modules with good residual eigenvalues. \refstepcounter{theorem}\paragraph{Remark \thetheorem} Even if ${\bf E}_T$ is an algebraic family of pre-${\cal D}$-modules with good residual eigenvalues, the associated family ${\cal R\!\!H} ({\bf E}_T)$ of Verdier objects may not be algebraic. \refstepcounter{theorem}\paragraph{Remark \thetheorem} If a semistable pre-${\cal D}$-module has good residual eigenvalues, then any other semistable pre-${\cal D}$-module in its S-equivalence class has (the same) good residual eigenvalues. Hence the analytic open subset $T_g$ of the parameter space $T$ of any analytic family of semistable pre-${\cal D}$-modules defined by the condition that residual eigenvalues are good is saturated under S-equivalence. \begin{lemma} If two semistable pre-${\cal D}$-modules with good residual eigenvalues are S-equivalent (in the sense of definition \ref{defstable} above), then the associated finite descriptions are S-equivalent (that is, Jordan-H\"older equivalent). \end{lemma} \paragraph{Proof} Let ${\bf E} =(E_0,E_1,s,t)$ be a pre-${\cal D}$-module with good residual eigenvalues (that is, the logarithmic connection $E_0$ has good residual eigenvalues on each component of $S$) such that $s\otimes t=0$. Then one can easily construct a family of pre-${\cal D}$-modules parametrized by the affine line $A^1$ which is the constant family ${\bf E} $ outside some point $P\in A^1$, and specializes at $P$ to ${\bf E} '=(E_0,E_1,0,0)$. Let $\phi :A^1 \to F$ be the resulting morphism to the moduli ${\cal F}$ of finite descriptions. By construction, $\phi$ is constant on $A^1 -P$, and so as ${\cal F}$ is separated, $\phi$ is constant. As the points of ${\cal F}$ are the S-equivalence classes of finite descriptions, it follows that the finite descriptions corresponding to ${\bf E} $ and ${\bf E} '$ are S-equivalent. Hence the S-equivalence class of the finite description associated to a pre-${\cal D}$-module depends only on the reduced module made from the pre-${\cal D}$-module. Now we must show that any two S-equivalent (in the sense of \ref{defstable}) reduced semistable modules have associated finite descriptions which are again S-equivalent (Jordan-H\"older equivalent). This follows from the deformation given in \ref{deform} by using the separatedness of ${\cal F}$ as above. \bigskip Now consider the moduli ${\cal P} = H//{\cal G}$ of semistable pre-${\cal D}$-modules. Let $H_g$ be the analytic open subspace of $H$ where the family parametrized by $H$ has good residual eigenvalues. By the above remark, $H_g$ is saturated under $H\to {\cal P}$. Hence its image ${\cal P} _g\subset {\cal P}$ is analytic open. Let $\phi :H_g \to {\cal F}$ be the classifying map to the moduli ${\cal F}$ of finite descriptions for the tautological family of pre-${\cal D}$-modules parametrized by $H$, which is defined because of the the above lemma. By the analytic universal property of GIT quotients (see Proposition 5.5 of Simpson [S] and the remark below), $\phi$ factors through an analytic map ${\cal R\!\!H} :P_g \to {\cal F}$, which we call as the {\sl Riemann-Hilbert morphism}. \refstepcounter{theorem}\paragraph{Remark \thetheorem} In order to apply Proposition 5.5 of [S], note that a ${\cal G}$-linear ample line bundle can be given on $H$ such that all points of $H$ are semistable. Moreover, though the proposition 5.5 in [S] is stated for semisimple groups, its proof works for reductive groups. \refstepcounter{theorem}\paragraph{Remark \thetheorem} The Riemann-Hilbert morphism can also be thought of as a morphism from the analytic stack of pre-${\cal D}$-modules with good residual eigenvalues to the analytic stack of perverse sheaves. \section{Some properties of the Riemann-Hilbert morphism} In this section we prove some basic properties of the morphism ${\cal R\!\!H}$, which can be interpreted either at stack or at moduli level. \begin{lemma}[Relative Deligne construction]\label{lemdel} (1) Let $T$ be the spectrum of an Artin local algebra of finite type over $C\!\!\!\!I$, and let $\rho_T$ be a family of representations of $G$ (the fundamental group of $X-S$ at base point $P_0$) parametrized by $T$. Let $E$ be a logarithmic connection with eigenvalue not differing by nonzero integers, such that the monodromy of $E$ equals $\rho$, the specialization of $\rho_T$. Then there exists a family $E_T$ of logarithmic connections parametrized by $T$ such that $E_0=E$ and $E_T$ has monodromy $\rho_T$. (2) A similar statement is true for analytic germs of $G$-representations. \end{lemma} \paragraph{Proof} For each $a$, choose a fundamental domain $\Omega_a$ for the exponential map ($z\mapsto \exp (2\pi \sqrt{-1} z)$) such that the eigenvalues of the residue $R_a(E)$ of $E$ along $S_a$ are in the interior of the set $\Omega_a$. As the differential of the exponential map $M(n,C\!\!\!\!I )\to GL(n,C\!\!\!\!I)$ is an isomorphism at all those points of $M(n,C\!\!\!\!I )$ where the eigenvalues do not differ by nonzero integers, using the fundamental domains $\Omega_a$ we can carry out the Deligne construction locally to define a family $E_T$ of logarithmic connections on $(X,S)$ with $E_0=E$, which has the given family of monodromies. Note that for the above to work, we needed the inverse function theorem, which is valid for Artin local algebras. \refstepcounter{theorem}\paragraph{Remark \thetheorem} If in the above, the family $\rho_T$ of monodromies is a constant family (that is, pulled back under $T\to \mathop{\rm Spec}\nolimits (C\!\!\!\!I )$), then $E_T$ is also a constant family as follows from Proposition 5.3 of [N]. \begin{proposition}[`Injectivity' of ${\cal R\!\!H} $]\label{propinj} Let ${\bf E}=(E,F,t,s )$ and ${\bf E}'=(E',F',t',s')$ be pre-${\cal D}$-modules having good residual eigenvalues, such that for each $a$, the eigenvalues of the residues of $E$ and $E'$ over $S_a$ belong a common fundamental domain $\Omega_a$ for the exponential map $exp :C\!\!\!\!I \to C\!\!\!\!I ^*:z\mapsto \exp (2\pi \sqrt{-1}z)$. Then ${\bf E}$ and ${\bf E}'$ are isomorphic if and only if the finite descriptions ${\cal R\!\!H}({\bf E})$ and ${\cal R\!\!H}({\bf E}')$ are isomorphic. \end{proposition} \paragraph{Proof} It is enough to prove that if the Malgrange objects ${\bf M}$ and ${\bf M}'$ are isomorphic, then so are the pre-${\cal D}$-modules ${\bf E}$ and ${\bf E}'$. First use the fact that, in a given meromorphic connection $M$ on $X-S$ (or on $N_{S,X}^{}-S$), there exists one and only one logarithmic connection having its residue along $S_a$ in $\Omega_a$ for each $a$, to conclude that $E$ and $E'$ (resp. $F$ and $F'$) are isomorphic logarithmic modules. To obtain the identification between $s$ and $s'$ (resp. $t$ and $t'$), use the fact that these maps are determined by their value at a point in each connected component $N_{S_a,X}^{}-S_a$ of $N_{S,X}^{}-S$ and this value is determined by the corresponding $C_a$ or $C'_a$ (resp. $V_a$ or $V'_a$). \begin{proposition}[Surjectivity of ${\cal R\!\!H}$]\label{propsurj} Let ${\bf D}$ be a finite description, and let $\sigma_a:C\!\!\!\!I ^*\to C\!\!\!\!I $ be set theoretic sections of $z\mapsto \exp (2\pi \sqrt{-1}z)$. Then there exists a pre-${\cal D}$-module ${\bf E}$ whose eigenvalues of residue over $S_a$ are in image$(\sigma_a)$, for which ${\cal R\!\!H}({\bf E})$ is isomorphic to ${\bf D}$. \end{proposition} \paragraph{Proof} This follows from proposition \ref{prop3}. \refstepcounter{theorem}\paragraph{Remark \thetheorem} The propositions \ref{propinj} and \ref{propsurj} together say that the set theoretic fiber of ${\cal R\!\!H}$ over a given finite description is in bijection with the choices of `good' logarithms for the local monodromies of the finite description (here `good' means eigenvalues do not differ by nonzero integers). \begin{proposition}[Tangent level injectivity for ${\cal R\!\!H}$]\label{propinfinj} Let $(E,F,t,s)_T$ be a family of pre-${\cal D}$-modules having good residual eigenvalues parametrized by the spectrum $T$ of an Artinian local algebra. Let the family ${\cal R\!\!H}(E,F,t,s)_T$ of finite descriptions parametrized by $T$ be constant (pulled back under $T\to \specC\!\!\!\!I$). Then the family $(E,F,t,s)_T$ is also constant. \end{proposition} \paragraph{Proof} This is just the rigidity result of proposition \ref{prop4}. \begin{proposition}[Infinitesimal surjectivity for ${\cal R\!\!H}$]\label{propinfsurj} Let $T$ be the spectrum of an Artin local algebra of finite type over $C\!\!\!\!I$, and let ${\bf D}$ be a family of finite descriptions parametrized by $T$. Let ${\bf E}$ be a pre-${\cal D}$-module having good residual eigenvalues such that ${\cal R\!\!H}({\bf E})={\bf D}_{\xi}$, the restriction of ${\bf D}$ over the closed point $\xi$ of $T$. Then there exists a family ${\bf E}'_T$ of pre-${\cal D}$-modules having good residual eigenvalues with ${\bf E}'_{\xi}={\bf E}$ and ${\cal R\!\!H}({\bf E}_T)={\bf D}_T$. \end{proposition} \paragraph{Proof} This follows from lemma \ref{lemdel} and the proof of proposition \ref{prop3} which works for families over Artin local algebras. \begin{theorem} The analytic open substack of the stack (or analytic open subset of the moduli) of pre-${\cal D}$-modules on $(X,S)$, where ${\bf E} $ has good residual eigenvalues, is an analytic spread over the stack (or moduli) of perverse sheaves on $(X,S)$ under the Riemann-Hilbert morphism. \end{theorem} \paragraph{Proof} This follows from propositions \ref{propsurj}, \ref{propinfinj} and \ref{propinfsurj} above. Note that we have not defined ${\cal R\!\!H}$ on the closed analytic subset $T_o$ of the parameter space of a family where ${\bf E} $ does not have good residual eigenvalues. Note that $T_o$ is defined by a `codimension one' analytic condition, that is, if $T$ is nonsingular, and if $T_o$ is a nonempty and proper subset of $T$, then $T_o$ has codimension 1 in $T$. However, it follows from Proposition \ref{propremov} below that the morphism ${\cal R\!\!H}$ on $T-T_o$ can be extended to an open subset of $T$ of complementary codimension at least two. However, on the extra points to which it gets extended, it may not represent the de Rham functor. \begin{proposition}[Removable singularities for ${\cal R\!\!H}$]\label{propremov} Let $T$ be an open disk in $C\!\!\!\!I$ centered at $0$. Let ${\bf E}_T=(E,F,t,s)_T$ be a holomorphic family of pre-${\cal D}$-modules parametrized by $T$. Let the restriction $E_z$ have good residual eigenvalues for all $z\in T-\{0\}$. Then there exists a holomorphic family ${\bf D}_U$ of finite descriptions parametrized by a neighbourhood $U$ of $0\in T$ such that on $U-\{0\}$, the families ${\cal R\!\!H}({\bf E}_U\vert U-\{0\}) $ and ${\bf D}_{U- \{0\}}$ are isomorphic. \end{proposition} If at $z=0$ the logarihmic connection $E$ does not have good residual eigenvalues, it is possible to change it to obtain a new logarithmic connection having good residual eigenvalues. This is done by the classical `shearing transformation' that we adapt below ({\sl inferior and superior modifications} for pre-${\cal D}$-modules). This can be done in family and has no effect on the Malgrange object at least locally. \begin{definition}\rm If $E$ is a vector bundle on $X$, and $V$ a subbundle of the restriction $E\vert S$, then the inferior modification ${_VE}$ is the sheaf of all sections of $E$ which lie in $V$ at points of $S$. This is a locally free subsheaf of $E$ (but not generally a subbundle). The superior modification $^VE$ is the vector bundle ${\cal O}_X(S)\otimes {_VE}$. \end{definition} \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem6} If $E\vert S =V\oplus V'$, then we have a canonical isomorphism $${_VE}\vert S \to V \oplus ({\cal N}^*_{S,X}\otimes V')$$ and hence also a canonical isomorphism $$^VE|S\to ({\cal N}_{S,X}\otimes V)\oplus V'$$ \refstepcounter{theorem}\paragraph{Remark \thetheorem}\label{rem7} If $(E,\nabla)$ is a logarithmic connection on $(X,S)$ and $V$ is invariant under the residue, then it can be seen that $_VE$ is invariant under $\nabla$, so is again a logarithmic connection. We call it the inferior modification of the logarithmic connection $E$ along the residue invariant subbundle $V\subset E\vert S$. It has the effect that the residual eigenvalues along $V$ get increased by $1$ when going from $E$ to $_VE$. As ${\cal O}_X(S)$ is canonically a logarithmic connection, the superior modification $^VE$ is also a logarithmic connection, with the residual eigenvalues along $V$ getting decreased by $1$. \bigskip Let $(E,F,t,s)$ be pre-${\cal D}$-module on $(X,S)$ such that $E$ has good residual eigenvalues. Let us for simplicity of writing assume that $S$ is connected. Let $E|S = \oplus_\alpha E^\alpha$ and $F=\oplus_\alpha F^\alpha$ be the respective direct sum decompositions into generalized eigen subbundles for the action of $\theta$. Then (see also remark \ref{rem3}) as $\theta$ commutes with $s$ and $t$, it follows that $t(E^\alpha) \subset F^\alpha$ and $s(F_\alpha)\subset E^\alpha$. Moreover, when $\alpha\ne 0$, the maps $s$ and $t$ are isomorphisms between $E^\alpha$ and $F^\alpha$. Now let $\alpha\ne 0$. Let $V=E^\alpha$ and $V'=\oplus_{\beta\ne \alpha}E^\beta$. Let $F'' = \oplus_{\beta\ne \alpha}F^\beta$. Let $F' = F^\alpha \oplus {\cal N}^*_{S,X}\otimes F''$. Let $E'={_VE}$. Then using \ref{rem6} and the above, we get maps $t':E'|S \to F'$ and $s':F'\to E'|S$ such that $(E',F',s',t')$ is a pre-${\cal D}$-module. \begin{definition}\label{definfmod}\rm We call the pre-${\cal D}$-module $(E',F',s',t')$ constructed above as the inferior modification of $(E,F,s,t)$ along the generalized eigenvalue $\alpha\ne 0$. \end{definition} Similarly, we can define the superior modification along a generalized eigenvalue $\alpha\ne 0$ by tensoring with ${\cal O}_X(S)$. \refstepcounter{theorem}\paragraph{Remark \thetheorem} The construction of inferior or superior modification of pre-${\cal D}$-modules can be carried out over a parameter space $T$ (that is, for families) provided the subbundles $V$ and $V'$ form vector subbundles over the parameter space $T$ (their ranks are constant). \paragraph{Proof of \protect\ref{propremov}} If the restriction $E= E_{T\vert z=0}$ has good residual eigenvalues, then ${\cal R\!\!H} {\bf E}_T$ has the desired property. So suppose $E$ does not have good residual eigenvalues. We first assume for simplicity of writing that $E$ fails to have good residual eigenvalues because its residue $R_a$ on $S_a$ has exactly one pair $(\alpha,\alpha-1)$ of distinct eigenvalues which differ by a positive integer, with $\alpha-1\ne 0$. Let $f_T$ be the characteristic polynomial of $R_{a,T}$. Then $f_0$ has a factorization $f_0 = gh$ such that the polynomials $g$ and $h$ are coprime, $g(\alpha)=0$ and $h(\alpha-1)=0$. On a neighbourhood $U$ of $0$ in $T$ we get a unique factorization $f_T\vert U = g_Uh_U$ where $g_U$ specializes to $g$ and $h_U$ specializes to $h$ at $0$. By taking $U$ small enough, we may assume that $g_U$ and $h_U$ have coprime specializations at all points of $U$. Let $V_U$ be the kernel of the endomorphism $g_U(R_{a,U})$ of the bundle $E_{a,U}$. If $U$ is small enough then $F_U$ is a subbundle. Now take the inferior modification ${\bf E}'= ({_V}E_U, F'_U,t'_U,s'_U)$ of the family $(E,F,t,s)_U$ as given by construction \ref{definfmod}. Then ${_VE}_U$ is a family of logarithmic connections having good residual eigenvalues, so by definition ${\bf E}'$ has good residual eigenvalues. If $(0,1)$ are the eigenvalues, then use superior modification along the eigenvalue $1$. If $R_a$ has eigenvalues $(\alpha,\alpha-k)$ for some integer $k\ge 1$, then repeat the above inferior (or superior) modification $k$ times (whether to choose an inferior or superior modification is governed by the following restriction : the multiplicity of the generalized eigenvalue $0$ should not decrease at any step). By construction, we arrive at the desired family $(E',F',s',t')$. \section*{References} \addcontentsline{toc}{section}{References} [L] Laumon, G. : Champs alg\'ebriques. Preprint no. 88-33, Universit\'e Paris Sud, 1988. [Mal] Malgrange, B. : Extension of holonomic ${\cal D}$-modules, in Algebraic Analysis (dedicated to M. Sato), M. Kashiwara and T. Kawai eds., Academic Press, 1988. [Ne] Newstead, P.E. : {\sl Introduction to moduli problems and orbit spaces}, TIFR lecture notes, Bombay (1978). [N] Nitsure, N. : Moduli of semistable logarithmic connections. J. Amer. Math. Soc. 6 (1993) 597-609. [S] Simpson, C. : Moduli of representations of the fundamental group of a smooth projective variety - I, Publ. Math. I.H.E.S. 79 (1994) 47-129. [Ve] Verdier, J.-L. : Prolongements de faisceaux pervers monodromiques, Ast\'erisque 130 (1985) 218-236. \bigskip Addresses: School of Mathematics, Tata Institute of Fundamental Research, Homi Bhabha Road, Bombay 400 005, India. e-mail: [email protected] Centre de Mathematiques, CNRS ura169, Ecole Polytechnique, Palaiseau cedex, France. e-mail: [email protected] \end{document} 
1995-08-28T06:20:09
9503
alg-geom/9503004
en
https://arxiv.org/abs/alg-geom/9503004
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9503004
Rogier Brussee
Rogier Brussee
The canonical class and the $C^\infty$ properties of K\"ahler surfaces
38 pages. Hardcopy available upon request. Latex 2e with amsart v. 1.2 or AMSlaTeX version 1.1. reason for resubmission: Changed title, corrected serious error in the argument for $p_g = 0$, major technical improvements in the handling of the localised Euler class of infinite dimensional bundles, referred some analysis to the literature and made some general improvements in the exposition
null
null
Bielefeld Preprint 95-037
null
We give a self contained proof using Seiberg Witten invariants that for K\"ahler surfaces with non negative Kodaira dimension (including those with $p_g = 0$) the canonical class of the minimal model and the $(-1)$-curves, are oriented diffeomorphism invariants up to sign. This implies that the Kodaira dimension is determined by the underlying differentiable manifold (Van de Ven Conjecture). We use a set up that replaces generic metrics by the construction of a localised Euler class of an infinite dimensional bundle with a Fredholm section. This allows us to compute the Seiberg Witten invariants of all elliptic surfaces with excess intersection theory. We then reprove that the multiplicities of the elliptic fibration are determined by the underlying oriented manifold, and that the plurigenera of a surface are oriented diffeomorphism invariants.
[ { "version": "v1", "created": "Fri, 10 Mar 1995 21:48:55 GMT" }, { "version": "v2", "created": "Fri, 25 Aug 1995 16:23:58 GMT" } ]
2008-02-03T00:00:00
[ [ "Brussee", "Rogier", "" ] ]
alg-geom
\section{Preparation} We first prove the corollaries from the main theorems~\ref{main} and~\ref{ellmult}. \begin{pf} Corollary \ref{spheres}. Let $S$ be a positive sphere on a surface with $\kappa \ge 0$. Blow up $n =S^2 + 1$ times. Now $e = S + E_1 + \cdots +E_n$ is a $(-1)$-sphere. Hence there is a $(-1)$-curve $E_0$ such that $e = \pm E_0 \in H_2(X,{\Bbb Z})$. Then $S = \pm E_0$ , or $e = E_0 = E_1$ say. The first possibility leads to the contradiction $E_0^2 \ge 0$, the second to $S = 0 \in H_2(X,{\Bbb Z})$. (Reducing non negative spheres to $(-1)$-spheres is a well known trick, but I forgot where I read it precisely.) Corollary~\ref{-infty} follows from corollary~\ref{spheres}. Corollary~\ref{Kodaira}.By the above, a K\"ahler surface is of Kodaira dimension $-\infty$ if it contains a non trivial $(0)$-sphere. Clearly all ruled surfaces contain one. To deal with ${\Bbb P}^2$, note that there is no surface with $b_+ =b_1=0$. Thus diffeomorphisms between surfaces with $b_2=1$, $b_1=0$ are automatically orientation preserving. Then a surface diffeomorphic to ${\Bbb P}^2$ must contain a $(+1)$-sphere, and is therefore of Kodaira dimension $-\infty$. Since $b_2 = 1$ it must in fact be equal to ${\Bbb P}^2$ (alternatively use Yau's result that ${\Bbb P}^2$ is the only surface with the homotopy type of ${\Bbb P}^2$ \cite[Theorem 1.1]{BPV}, but this is a deep theorem). We conclude that Kodaira dimension $-\infty$ can be characterised by just diffeomorphism type. Without loss of generality we can therefore assume that $\kappa \ge 0$. If $\Kmin^2 > 0$, then $X$ is of general type. If $\Kmin^2 = 0 $ and $\Kmin$ is not torsion, then $\kappa(X) = 1$, finally if $\Kmin$ is torsion, $\kappa(X) = 0$. This proves that Kodaira dimension is determined by the oriented diffeomorphism type. If $X$ and $Y$ are orientation reversing diffeomorphic, both are minimal, otherwise one of them would contain a positive sphere. Then necessarily either $K_X^2 = K_Y^2 = 0$, or both have $K_X^2 ,K_Y^2 >0$, i.e. $X$ and $Y$ are of general type. Now copy the argument of \cite[lemma S.4]{FM}: for minimal surfaces with $\kappa = 0,1$, the signature $\sign = \numfrac13(K^2 - 2e)\le 0$. Thus $\sigma(X) = -\sigma(Y) = 0$, and $e(X) = e(Y) = 0$. In Kodaira dimension $0$, this leaves only tori and hyperelliptic surfaces, which can fortunately be recognised by homotopy type \cite[lemma 2.7]{FM}. Corollary~\ref{plurigenera}. Since $P_1=p_g$ is an oriented topological invariant we will whence assume that $n\ge 2$. We have to distinguish between the different Kodaira dimensions. For surfaces of general type (i.e $\kappa = 2$) we argue as follows. The plurigenera $P_n$ and $\chi(O_X)$ are birational invariants. Then by Ramanujan vanishing and Riemann Roch (cf. \cite[corollary VII.5.6]{BPV}) we have \begin{equation}\label{gtpluri} P_n(X) = P_n(\Xmin) = {\numfrac12} n(n-1) \Kmin^2 + \chi({\mathcal O}_X) \end{equation} Since $\chi({\mathcal O}_X)$ is an oriented topological invariant the $P_n$ are oriented diffeomorphism invariants in this case. For surfaces with Kodaira dimension $0$ or $1$ with a fundamental group that is not finite cyclic, we simply quote \cite[S.7]{FM}. For surfaces with finite cyclic fundamental group, it follows from the invariance of the multiplicities and the canonical bundle formula which gives an explicit formula for $P_n(X)$ in terms of the multiplicities and $\chi(O_X)$. (see \cite[lemma I.3.18, prop. I.3.22]{FM}). \end{pf} Here is an other easy corollary \begin{Corollary} Every $(-2)$-sphere $\tau$ is orthogonal to $\Kmin$ . If there is a $(-1)$-curve $E_1$ such that $\tau\cdot E_1 \ne 0$, then there is a $(-1)$-curve $E_2$ such that $\tau = \pm E_1 \pm E_2 \in H_2(X,{\Bbb Z})$. \end{Corollary} \begin{pf} Let $R_\tau $ be the reflection in $\tau$. It is represented by a diffeomorphism with support in a neighborhood of $\tau$. By the invariance of $\Kmin$ up to sign, $R_\tau \Kmin = \Kmin + (\tau \cdot \Kmin) \tau = \pm \Kmin$. But if $\Kmin \ne 0 \in H^2(X,{\Bbb Q})$, then $\tau$ and $\Kmin$ are indepent, since $\tau^2 = -2 $ and $\Kmin^2 \ge 0$. Thus in either case $(\tau, \Kmin) = 0$. Moreover if $E_1$ is a $(-1)$-curve then either $R_\tau E_1 = E_1$, $R_\tau E_1 = -E_1$, or there is a different $(-1)$-curve $E_2$ such that $R_\tau(E_1) = \pm E_2$. The first possibility gives $\tau\cdot E_1 = 0$, the second $(\tau \cdot E_1)^2 = 2$ i.e. is impossible, and the third $(\tau\cdot E_1) = \pm 1$. The statement follows. \end{pf} It will be convenient to first prove the main theorem~\ref{main} with (co)homology groups with ${\Bbb Q}$ coefficients, and later mop up to prove the theorem over ${\Bbb Z}$. Theorem~\ref{main} mod torsion is a formal consequence of the existence of a set of basic classes $$ \K(X) = \{K_1, K_2 \ldots \} \subset H^2(X,{\Bbb Z}) $$ functorial under oriented diffeomorphism between $4$-manifolds with $b_+ \ge 1$, and having the following properties: \begin{Properties} \label{*} For every K\"ahler surface $X$ of non negative Kodaira dimension \begin{enumerate} \item \label{i} the $K_i$ are of type $(1,1)$ i.e. represented by divisors, \item \label{ii} if $X$ is minimal, then for every K\"ahler form $\Phi$, $\deg_\Phi(K_X )\ge |\deg_\Phi(K_i)|$, \item \label{iii} if $\~X \m@p--\rightarrow{\sigma} X$ is the blow-up of a point $x \in X$, then $\sigma_*\K(\~X) = \K(X)$. \item \label{iv} every $K_i$ is characteristic i.e. $K_i \equiv w_2(X) \pmod 2$, \item \label{v} $K_X \in \K$. \end{enumerate} \end{Properties} In the case that $X$ is an algebraic surface we could replace item~\ref{ii} by weaker and more geometric requirement that $2g(H) - 2 \ge H^2 + |K_i\cdot H|$ for every very ample divisor $H$ without changing the results. We will see later that Seiberg Witten theory will give such an inequality for all surfaces minimal or not. This should not be confused with a Thom conjecture type of statement, since our methods do not give information about the minimal genus for arbitrary smooth real surfaces in a homology class. It is also clearly impossible to have a degree inequality like property \ref{ii} for all K\"ahler forms if $X$ is rational or ruled. Recall that for algebraic surfaces, the Mori cone $\NEbar(X) \subset H_2(X,{\Bbb R})$ is the closure of the cone generated by effective curves. It is dual to the nef (or K\"ahler) cone. In other words, the numerical equivalence class of a curve $D$ lies in $\NEbar(X)$ if and only if $H\cdot D \ge 0$ for all $H$ ample. For a K\"ahler surface $(X,\Phi)$, it will be convenient to define the nef cone as closure of the positive cone in $H^{1,1}(X) \subset H^2(X,{\Bbb R})$ spanned by all K\"ahler forms, and containing $\Phi$. The Mori cone $\NEbar$ is then just the dual cone in $H_2(X,{\Bbb R})\cap {H^{1,1}}^{\scriptscriptstyle\vee}$ i.e. $$ \NEbar= \{C \in {H^{1,1}}^{\scriptscriptstyle\vee} \subset H_2(X,{\Bbb R}) \mid \int_C \omega \ge 0, \txt{for all K\"ahler forms $\omega$}\}. $$ (With this definition, a line bundle is nef iff for all $\epsilon > 0$, it admits a metric such that the curvature form $F$ has $\numfrac{\sqrt{-1}}{2\pi} F \ge -\epsilon \Phi$. A class $\omega \in \NEbar$ if there exists a sequence of closed positive currents of type $(1,1)$ converging to the dual of $\omega$, i.e is $\NEbar$ dual to $N_{\text{psef}}$ in \cite[proposition 6.6]{Demailly}. I am grateful to Demailly for explaining this to me). We will freely identify homology and cohomology by Poincar\'e duality. \begin{Lemma} \label{decomp} If a class $L \in H^{1,1}(X)$ satisfies $\deg_\Phi(K_X)\ge |\deg_\Phi(L)|$ for all K\"ahler forms $\Phi$, then there is a unique decomposition of the canonical divisor $K_X = D_+ + D_- $ with $D_+$, $D_- \in \NEbar(X)$ such that $L= D_+ - D_-$. \end{Lemma} \begin{pf} Define $D_\pm = {\numfrac12}(K_X \pm L)$. Then $K_X = D_+ + D_-$, $L= D_+ - D_- $, and $D_\pm \in \NEbar$. \end{pf} The following simple lemma is a minor generalisation of the fact that the canonical divisor of a surface of general type is numerically connected \cite[VII.6.1]{BPV}. \begin{Lemma} \label{connectedness} Let $X$ be a minimal K\"ahler surface of non negative Kodaira dimension. Suppose there is a decomposition $ K_X = D_+ + D_-$ with $D_+$, $D_- \in \NEbar(X)\subset H^{11}(X)$. Then $D_+ \cdot D_- \ge 0 $, with equality if and only if say $K_X \cdot D_+ = D_+^2 = 0$. Thus if $X$ is of general type then $D_+ = 0$, if $\kappa(X) = 1$, then $D_+ = \lambda K_X$ with $0 \le \lambda \le 1$, and finally $D_+ = D_- = 0$ if $\kappa(X) = 0$. \end{Lemma} \begin{pf} First assume that $D_+^2 \le 0$. Since $K_X$ is nef, $D_+ \cdot D_- = (K_X-D_+)\cdot D_+ \ge -D_+^2 \ge 0$, with equality iff $K_X \cdot D_+ = D_+^2 = 0$. If $D_+^2 >0$ and $D_-^2 >0 $, then using the K\"ahler form $\Phi$, we can write $D_+ = \alpha \Phi + C_+$ and $D_- = \beta \Phi + C_-$ with $\alpha$, $\beta >0$ and $C_\pm \in \Phi^\perp$. By the Hodge index theorem, $$ D_+ \cdot D_- = \alpha \beta \Phi^2 + C_+\cdot C_- \ge \alpha\beta \Phi^2 - \sqrt{-C_+^2}\sqrt{-C_-^2} >0. $$ The statement for surfaces of general type follows directly from Hodge index and the fact that $K_X^2 > 0$. If $\kappa(X) = 1$, then $K_X$ is a generator of the unique isotropic subspace of $K_X^\perp$, so $D_+ = \lambda K_X$, and $D_- = (1-\lambda) K_X$. Since $K_X$, $D_+$ and $D_- \in \NEbar(X)$, $\lambda$ is bounded by $0 \le \lambda \le 1$. Finally if $\kappa(X) = 0$, $K_X$ is numerically trivial and, $D_+$ and $D_-$ must be zero as well. \end{pf} \begin{Lemma} \label{inequality} Let $X$ be a surface of non negative Kodaira dimension with $(-1)$-curves $E_1, \ldots E_m$. Assume that $\K$ has properties~\ref{*}. Then $K_i^2 \le K_X^2$ for all $K_i \in \K$, with equality if and only if $$ K_i = \lambda \Kmin + \sum \pm E_i \in H^2(X,{\Bbb Q}) $$ where $\lambda = \pm 1$ if $X$ is of general type, $\lambda$ is a rational number with $|\lambda | \le 1 $ if $\kappa(X) = 1$, and where $\lambda = 0$ if $\kappa(X) = 0$. \end{Lemma} \begin{pf} By property \eqref{iii}, and \eqref{iv}, $K_i = K_{i,\min} + \sum_j (2a_{ij} + 1)E_j$. Thus $$ K_i^2 \le K_{i,\min}^2 - \#(-1)\hbox{-curves}, $$ with equality if and only if $a_{ij} =0$, or $-1$ for all $i,j$. Since $K_X^2 = \Kmin^2 - \#(-1)$-curves, we can assume that $X$ is minimal. Using property~\eqref{i} and \eqref{ii} and lemma~\ref{decomp}, write $K_X = D_+ + D_-$ and $K_i = D_+ - D_-$, with $D_\pm \in \NEbar(X)$. Then by lemma~\ref{connectedness} $K_i^2 = K_X^2 - 4 D_+\cdot D_- \le K_X^2 $ with equality under the stated condition. Note that this lemma does not use diffeomorphism invariance, nor that $K_X \in \K$. \end{pf} We are now in a position to formulate and prove half of the main theorem \begin{Proposition} \label{Kcharacterisation} Assume that for all 4-manifolds $X$ with $b_+ \ge 1$ there is a set of basic classes $\K(X) = \{K_1, K_2, \ldots\} \subset H^2(X,{\Bbb Z})$ functorial under oriented diffeomorphism having properties~\ref{*}. Then $\Kmin$ is an oriented $C^\infty$ invariant up to sign and torsion, and every $(-1)$-sphere is represented by a $(-1)$-curve up to sign and torsion. \end{Proposition} \begin{pf} Using lemma~\ref{inequality} we can easily reduce the invariance of $\Kmin$ up to sign and torsion to showing that $(-1)$-spheres are represented by $(-1)$-curves up to sign and torsion. Since $K_X \in \K$, there is a nonempty subset $\K_0 = \{ K_j\} \subset \K$ with $K_j^2 = K_X^2 = 2e(X) + 3\sign(X)$. Consider the projection $K_{j,\min}$ of $K_j$ to the minimal model i.e. the projection to the orthogonal complement of the $(-1)$-spheres. If $K_{j,\min}^2 >0$, then by lemma~\ref{inequality}, $X$ is of general type, and $K_{j,\min} = \pm \Kmin$ up to torsion. If $K_{j,\min}^2 = 0$, there are two possibilities. If $K_{j,\min}$ is torsion for all $j$, then again by lemma~\ref{inequality}, $X$ is of Kodaira dimension $0$ i.e. $\Kmin$ is also torsion. Otherwise we choose $j$ such that $K_{j,\min} \ne 0$ has maximal divisibility. Since $K_X \in \K_0$ our little lemma shows that, the Kodaira dimension is $1$ and $K_{j,\min} = \pm \Kmin$. Now let $e$ be the class of a $(-1)$-sphere in $H^2(X,{\Bbb Q})$. Without loss of generality, we can assume that $K_X\cdot e < 0$. Consider $R_e$ the reflection generated by a $(-1)$-sphere $e$. It is represented by an orientation preserving diffeomorphism. Since $\K$ is invariant under oriented diffeomorphisms, the characterisation of basic classes with square $K_X^2$ tells us that \begin{align} R_e K_X &= \Kmin + \sum E_i + 2 (K_X\cdot e)e \label{line1} \\ &= \lambda \Kmin + \sum \pm E_i \label{line2} \end{align} with $|\lambda| \le 1$. Taking intersection with $E_i$ we find that $(E_i\cdot e)( e\cdot K_X) =0$~or~$1$. Since $ K_X\cdot e \equiv e^2$ is odd, $e$ is either orthogonal to all $(-1)$ curves (i.e. $ e\in H^2(\Xmin,{\Bbb Q})$) or there is a $(-1)$-curve, say $E_1$, such that $K_X\cdot e =E_1 \cdot e= -1$. However, $e \in H^2(\Xmin)$ implies that $e = {\lambda -1 \over 2K_X \cdot e}\Kmin$, which is impossible because $\Kmin^2 \ge 0$. Thus, after renumbering the $(-1)$-curves, \eqref{line1} and \eqref{line2} can be rewritten to \begin{equation}\label{(-1)-sphere} e = {\numfrac12}(1-\lambda) \Kmin + \sum _{i=1}^N E_i \end{equation} with $N = {\numfrac14}(1-\lambda)^2 \Kmin^2 + 1$. Now reflect $e$ in $E_1^\perp$. $R_{E_1} e$ is yet another $(-1)$-sphere, so it has a representation as in equation~\eqref{(-1)-sphere}, except possibly for an overall sign \begin{align*} R_{E_1}e &= {\numfrac12}(1-\lambda)\Kmin - E_1 + \sum _{i=2}^N E_i \\ &= \pm \big({\numfrac12}(1-\mu) \Kmin + \sum_{j=1}^M E_{i_j}\big). \end{align*} Upon comparison, we see that the sign is minus, that $N=M =1$, and that $0 \le 1-\lambda = \mu - 1 \le 0$ unless $\Kmin = 0$. In other words $ e = E_1 \in H^2(X,{\Bbb Q})$. \end{pf} \section{The localised Euler class of a Banach bundle.}\label{top} We will use a construction pioneered by Pidstrigatch and Pidstrigatch Tjurin \cite{Pidstrigatch:instanton}, \cite[\S 2]{PT}, which is a convenient and general way to define fundamental cycles for moduli spaces arising from elliptic equations. Unfortunately their construction is not quite in the generality we will need it, and we will therefore set it up in fairly large generality here. The cycle is the localised homological Euler class of an infinite dimensional bundle. It can be used to give definitions that avoid transversality arguments needing small deformations, generic metrics etcetera, although transversality will be extremely useful for computations and proofs. The construction is modeled on Fultons intersection theory and in the complex case it makes the whole machinery of excess intersection theory available. However, although the construction is very simple in principle, the whole thing has turned a bit technical. On first reading it is best to ignore the difference between \v Cech and singular homology, and continue to proposition \ref{locEuler}, the construction of the Euler class in the proof of this proposition and corollary~\ref{locChern}. Some readers might even want to continue to the next section, since we will use rather little of the general machinery for the proofs of the theorems and corollaries in the introduction. \smallskip\noindent We first make some algebraic topological preparations. For any pair of topological spaces $A \subset X$, homology with closed support and with local coefficients $\locsys$ is defined as $$ H_i^{\cl}(X,A;\locsys) = {\displaystyle \lim_{\leftarrow}}_K H_i(X,A \cup (X-K);\locsys) $$ where we take the limit over all compacta $K \subset X - \o A$. $H^{\cl}_*$ is functorial under proper maps. Unfortunately this ``homology theory'' suffers the same tautness problems that singular homology has. To be able to work with well behaved cap products we will have to complete it. The following works well enough for our purposes but is a bit clumsy. Suppose that $X$ is {\sl locally modelable} i.e. is locally compact Hausdorff and has local models which are each subsets of some ${\Bbb R}^n$. Obviously locally compact subsets of locally modelable spaces are locally modelable, in particular a closed subset of a local modelable space is locally modelable. Then for every compact subset $K \subset X - \o A$ there is a neighborhood $U_K \supset K$ in $X$ which embeds in ${\Bbb R}^N$. We now define $$ \cH_i^{\cl}(X,A,\locsys) = {\displaystyle \lim_{\leftarrow}}_K \cH_i(U_K,A\cap U_K \cup (U_K -K); \locsys) $$ where for every pair $(Y,B)$ in a manifold $M$, \v Cech homology is defined as $$ \cH_i(Y,B) = {\displaystyle \lim_{\leftarrow}} \{H_i(V,W),\ (V,W) \txt{neighborhoods of} (Y,B) \txt{in} M\} $$ This definition depends neither on the choice of $U_K$, nor on the embedding $U_K \lhook\nobreak\joinrel\nobreak\to {\Bbb R}^N$, since two embeddings are dominated by the diagonal embedding, and $\cH_*(Y,B)$ does not depend on $M$ but only on $(Y,B)$ (c.f \cite[VIII.13.16]{Dold}). Fortunately we do not usually have to bother with \v Cech homology. Suppose in addition that $X$ is locally contractible e.g. locally a sub analytic set (c.f. \cite[\S I.1.7]{GoreskyMcPherson}, and the fact that Whitney stratified spaces admit a triangulation). Then $X$ is locally an Euclidean neighborhood retract (ENR) by \cite[IV 8.12]{Dold} and since in a Hausdorff space a finite union of ENR's is an ENR by \cite[IV 8.10]{Dold} we can assume that $U_K$ is an ENR. Now assume that $A$ is open. Then by \cite[prop. VIII 13.17]{Dold} $$ \cH_* (U_K,U_K\cap A \cup (U_K-K)) \iso H_*(U_K,U_K \cap A \cup (U_K - K)) \iso H_i(X,A \cup X-K). $$ Thus in this case $\cH_*^{\cl}(X,A) = H_*^{\cl}(X,A)$. If $A$ is closed and locally contractible then one should be able to organise things such that $U_K \cap A$ is an ENR and the same conclusion would hold. \begin{Lemma}\label{capproduct} Let $X$ be a locally modelable space, and $Z$ a locally compact (e.g. closed) subspace, then there are cap products $$ \cH^i(X,X - Z,\locsys) \tensor \cH^{\cl}_j(X,\locsys') \m@p--\rightarrow{\cap} \cH^{\cl}_{j-i}(Z, \locsys\tensor\locsys') $$ with the following properties. \begin{enumerate} \item If $Y$ is locally embeddable, $f: Y\to X$ is proper, and $\sigma' \in \cH^{\cl}_j(Y, Y-f^{-1}(Z))$, then the push-pull formula holds: $$ f_*(f^* c \cap \sigma') = c \cap f_* \sigma'. $$ \item If $Z \lhook\nobreak\joinrel\nobreak\m@p--\rightarrow{i} W$ is proper and $W$ is locally compact, we can increase supports i.e. $$ c|_{(X,X-W)} \cap \sigma = i_*(c \cap \sigma). $$ \end{enumerate} \end{Lemma} \begin{pf} For every $c \in \cH^i(X,X-Z)$ and $\sigma \in \cH_j^{\cl}(X)$, we have to construct a class $c \cap \sigma \in \cH_{i-j}(Z, Z-K)$ for a cofinal family of compacta $\{K\}$. Since $Z$ is locally compact, every compactum $K$ is contained in a compactum $L \subset Z$ with $L \supset \o L \supset K$. Likewise there exists a compactum $L' \supset\!\supset L$. By excision it suffices to construct a class in $\cH_{i-j}(L,L-K)$. Let $U_{L'}$ be a neighborhood of $L'$ in $X$ which embeds in ${\Bbb R}^N$. Let $V_L$, $W_{L-K} \subset V_L$, and $V_K \subset V_K$ be neighborhoods of respectively $L$, $L_K$ and $K$ in ${\Bbb R}^N$. Define $U_L = V_L \cap U_{L'}$. We can assume that $U_L \cap Z = U_L \cap L'$, $V_K \cap Z = V_K \cap L$, and after replacing $V_{L-K}$ by $(V_{L-K} - (L' \cap W_{L-K}^c) \cup V_K$, that $V_L \cap (L'-K) = W_{L-K} \cap (L' - K)$. Then our task is to construct a class $c_L\cap \sigma_L \in H_{i-j}(V_L,W_{L-K})$ possibly after shrinking $V_L$ and $W_{L-K}$. We have a restriction map $\cH^i(X,X-Z) \to \cH^i(U_L,U_L-L')$. After shrinking $V_L$ if necessary, $c|_{(U_L,U_L-L')}$ comes from a class $c_L \in H^i(V_L,V_L -L')$. By definition there is map $$ \cH^{\cl}_j(X) \to \cH_j(U_L,U_L-K) \to H_j(V_L,V_L-K). $$ Let $\sigma_L \in H_j(V_L,V_L-K)$ be the image of $\sigma$. Now write $V_L- K = (V_L - L') \cup (W_{L-K} - K)$. Then the standard cap product \cite[VII Def. 12.1]{Dold} gives a map $$ H^i(V_L,V_L-L') \tensor H_j(V_L,V_L-K) \m@p--\rightarrow{\cap} H_{j-i}(V_L,W_{L-K}-K) $$ so we get a class $c_L \cap \sigma_L \in H_{j-i}(V_L, W_{L-K})$ as required. Since if $K' \supset K$, choices for $K'$ will work a fortiori for $K$, we can pass to the limit. To prove the first property, note that since $f$ is proper, $f^{-1} Z$ is locally compact. Choose compacta $K \subset\!\subset L \subset\!\subset L' \subset Z$ giving compacta $f^{-1}K \subset\!\subset f^{-1}L \subset\!\subset f^{-1}L'$. Note that compacta of the form $f^{-1}K$ are a cofinal family of compacta in $f^{-1}(Z)$. Embed neighborhoods $U_{L'} \subset V_{L'} \subset {\Bbb R}^N$ and $U_{f^{-1}L'} \subset {\Bbb R}^M$. Now we carry out the construction above with the diagonal embedding of $U_{f^{-1}L'}$ in $ {\Bbb R}^{N + M}$. Let $V_{f^{-1}L'}$ be a neighborhood of $U_{f^{-1}L'} \in {\Bbb R}^{N+M}$. We can assume that $V_{f^{-1}L'} \to V_{L'}$ under the projection $\pi$ to ${\Bbb R}^N$. We can also assume that $c|_{(U_L,U_L-L')}$ comes from a class $c_L \in H^i(V_L,V_L - L')$. Finally let $\sigma_{f^{-1}L'}$ be an image of $\sigma$ in $H_j(V_{f^{-1}L'}, \pi^{-1}W_{K-L})$. Then the first property follows from the identity $$ \pi_*(\pi^* c_L \cap \sigma'_{f^{-1}L'}) = c_L \cap \pi_* \sigma'_{f^{-1}L'} $$ in $H_j(V_l, W_{K-L})$. The second property is left to reader. \end{pf} A smooth manifold $X$ of dimension $n$, has an orientation system $\orr(X)$, the sheafification of the presheaf $U \to H^n(X,X-U)$. Equivalently, we can define $\orr(X)$ as the sheaf $R^d\pi_*(X\times X, X\times X -\Delta, {\Bbb Z})$, where $\Delta$ is the diagonal of $X\times X$, $\pi$ the projection on the first coordinate, and $R^d\pi_*$ the parametrised version of the $d$ th cohomology. Likewise for a real vector bundle $E$ of rank $r$ there is an orientation system $\orr(E)$, the sheafification of $H_q(E|_U,E|_U - U)$. We have $\orr(X) = \orr (TX)^{\scriptscriptstyle\vee}$, as can be seen immediately from the alternative description of $\orr(X)$ and excision. A manifold $X$ has a unique fundamental class $[X] \in H^{cl}_n(X,\orr(X))$ in singular or \v Cech homology such that for small $U$ $$ [X]|_{X-U}\in H_d(X,X-U,H^d(X,X-U))= \rmmath{Hom}(H^d(X,X-U),H^d(X,X-U)) $$ is identified with the identity (cf \cite[p. 357]{Spanier}). Similarly, a bundle $E$ has a Thom class $\Phi_E \in \cH^r(E,E-X, \orr(E))$ \cite[p. 283]{Spanier}. In turn for every section $s$ in $E$ with zero set $Z(s)$, the Thom class defines a localised cohomological Euler class $e(E,s) = s^*\Phi_E \in \cH^r(X, X - Z(s), \orr(E))$. \smallskip\noindent Let $M$ be a Banach manifold, $E$ a real Banach vector bundle on $M$ and $s$ a section of $E$ with zero set $Z(s)$. The section induces an exact sequence \begin{equation}\label{tosplit} 0 \m@p--\rightarrow{} E \m@p--\rightarrow{} s^* TE \m@p--\rightarrow{\pi} TM \to 0, \end{equation} which expresses that the vertical tangent bundle of the total space of $E$ is canonically isomorphic to the bundle $E$. On $Z(s)$ we have a canonical splitting of this sequence, given by the sequence $$ 0 \m@p--\rightarrow{} TM \m@p--\rightarrow{Ts_0} s_0^*TE \m@p--\rightarrow{} E \m@p--\rightarrow{} 0 $$ defined by the zero section $s_0$, and the identification $s^*TE|_{Z(s)} = s_0^*TE|_{Z(s)}$ over $Z(s)$. This gives a canonical map $$ Ds: TM|_{Z(s)} \m@p--\rightarrow{Ts} s^*TE = s_0^*TE \m@p--\rightarrow{} E|_{Z(s)}. $$ If $D$ is a connection on $E$ then $D(s)$ is a splitting that extends the canonical splitting over $Z(S)$ (hence the notation) but in general connections need not exist on Banach manifolds. We will avoid choosing non canonical splittings. \begin{Proposition}\label{locEuler} Let $M$ be a smooth Banach manifold, $E$ a banach bundle over $M$ and $s$ a section in $E$. Assume that \begin{enumerate} \item {\sloppy The map $Ds$ is a section in the bundle $\Fred^d(TM|_{Z(s)},E|_{Z(s)})$ of Fredholm maps of index $d$. We say that $Z(s)$ has virtual dimension $d$, and that $Ds$ is Fredholm of index $d$. } \item The real line bundle $\det(\Ind(Ds))$ is trivialised over $Z(s)$. \end{enumerate} Then these data define a \v Cech homology class with closed support $$ {\Bbb Z}(E,s)= {\Bbb Z}(s) \in \cH^{\cl}_d(Z(s),{\Bbb Z}) $$ with the following properties. \begin{enumerate} \item\label{smoothcase} The class ${\Bbb Z}(s) = [Z(s)]$ if $Z(s)$ is smooth of dimension $d$ and carries the natural orientation defined by the trivialisation of $\det(\Ind Ds)$, \item\label{homotopy} if $\{C\}$ is a family of closed subsets of $M$ such that $C \cap Z(s)$ is compact for all $C$, then there is a natural map $\cH_j(Z(s)) \to {\displaystyle \lim_{\leftarrow}}_C H_j(M,M-C,{\Bbb Z})$, and if $s_t$ is a one parameter family of sections with this property then ${\Bbb Z}(s_0) = {\Bbb Z}(s_1)) \in {\displaystyle \lim_{\leftarrow}}_C H_d(M,M-C,{\Bbb Z})$. \end{enumerate} For every exact sequence $$ 0 \to E' \to E \to E'' \to 0, $$ defined over a neighborhood of $Z(s)$, let $s''$ be the induced section in $E''$, and $s'$ the induced section of $E'|_{Z(s'')}$ with zero set $Z(s)$. Then \begin{enumerate} \setcounter{enumi}{2} \item \label{fdeuler} if $E'$ has finite rank $$ {\Bbb Z}(s) = e(E'|_{Z(s'')},s')\cap {\Bbb Z}(s''), $$ \item \label{stability} if $Ds''|_{Z(s)}$ is surjective, then $Z(s'')$ is smooth in a neighborhood of $Z(s)$, $Ds':TZ(s'')|_{Z(s)} \to E'|_{Z(s)}$ is Fredholm, with $\Ind Ds' \iso \Ind Ds$ and $$ {\Bbb Z}(E,s) = {\Bbb Z}(E'|_{Z(s'')},s'). $$ \end{enumerate} \end{Proposition} For property \ref{homotopy} there are two typical situations we have in mind. One is that we have a natural connected family of sections $s_t$ such that $Z(s_t)$ is compact for all $t\in T$. In this situation we get a homology class ${\Bbb Z}(s_{t_0}) \in H_d(M)$ independent of the choice of $t_0$ (take $\{C\} = \{M\}$). Such will be the case in Seiberg Witten theory. In the other case we again have a family of sections $s_t$ but there is ``bubbling'' which invariably means we lack some a priori estimate. For example in Donaldson theory, the moduli space of ASD connections with curvature bounded in the $L^4$ norm is compact. Therefore it is natural to define a family of subsets $\{\B^{\le C}\}_{C \in {\Bbb R}^+}$ in the space $\B^*$ of all irreducible $L^2_2$ connections mod gauge, where $\B^{\le C}$ the subset of connections with $L^4$ norm of the curvature bounded by $C$. \begin{pf} If $M$ (hence $E$) is a finite dimensional manifold of dimension $N+d$ then $E$ is a real vector bundle of rank $N$ with an isomorphism $\det(E) = \det(TM)$ over $Z(s)$. Let $[M] \in H^{\cl}_N(M,\orr(M))$ be the fundamental class, and $\Phi_E$ the twisted Thom class of $E$ in $H^{N-d}(E,E-M,\orr(E))$. Define $$ {\Bbb Z}(s) = e(E,s) \cap [M] \in \cH^{\cl}_d(Z, \orr(E)\tensor \orr(M)) = \cH^{\cl}_d(Z(s),{\Bbb Z}) $$ i.e. ${\Bbb Z}(s)$ is the Poincar\'e dual of the localised cohomological Euler class. In the last step we used the chosen trivialisation of $\orr(E) \tensor \orr(M) = \orr(\det TM^{\scriptscriptstyle\vee}\tensor\det E) =\orr (\det(\Ind(Ds)))$ given by the trivialisation of the index. In the infinite dimensional case we proceed similarly but we have to go through a limiting process and use that we know what to do when the section is regular. For each compactum $K \subset Z$ we have to construct a class ${\Bbb Z}_K \in \cH_d(Z,Z-K)$ such that for $K' \supset K$ the class ${\Bbb Z}_{K'}|_{Z-K} = {\Bbb Z}_K$ under the restriction map $H_d(Z,Z-K') \to H_d(Z,Z-K)$. Over a neighborhood $U$ of $K$ in $M$ we can find a finite rank $N$ subbundle $F$ of $E$ such that $\rmmath{Im}(Ds)|_K + F|_K = E|_K$. Such a bundle certainly exists: we can choose a finite number of sections $s_1, \ldots s_N$ such that the $s_i$ span $\rmmath{Coker}(Ds_x)$ for every $x \in K$, and possibly after perturbing we can assume that the $s_i$ are linearly independent in a neighborhood. Let $\~E$ be the quotient bundle $E/F$ defined over $U$, and $\~s$ the induced section with zero set $M_f = Z(\~s)$ ($f$ is for finite, $M$ is for, well, manifold). Clearly the map $TM|_{Z(s)} \m@p--\rightarrow{Ds} E|_{Z(s)} \m@p--\rightarrow{} \~E $ is surjective. Since the canonical map $D\~s$ on $M_f$ restricts to this composition on $Z(s)$, $D\~s$ is surjective on $M_f$ possibly after shrinking $U$. Hence $M_f$ is a smooth manifold. Let $T = \ker(TM|_{M_f} \to \~E)$. There is a canonical identification $T \iso TM_f$. Now $T$ is a bundle of rank $N+d$ since \begin{equation}\label{theindex} \Ind(Ds)|_K = T - F. \end{equation} Thus $M_f$ has dimension $N+d$. On $M_f$, the section $s$ in $E$ lifts to a section $s_f$ of the subbundle $F$. Clearly $Z(s_f) = Z(s) \cap U$. Define $$ {\Bbb Z}_K = e(F|_{M_f},s_f) \cap [M_f] \in \cH_d(Z(s), Z(s) - K;{\Bbb Z})). $$ Here we have used the restriction map $$ \cH^{\cl}_d(Z(s)\cap U; \orr(F)\tensor \orr(M_f)) \to\cH_d(Z(s),Z(s)-K; \orr(F) \tensor \orr(M_f)), $$ the identification $\orr(\det(\Ind(Ds))) = \orr(F) \tensor \orr(M_f)$ and the chosen trivialisation of $\det(\Ind(Ds))$ as in the finite dimensional case. This construction does not depend on the choices. If $F$ and $F'$ are two choices of subbundles of $E$ then there is third bundle $F''$ containing $F + F'$. We can therefore assume that $F$ is a subbundle of $F'$. Then using primes to denote objects we get out of the construction above using $F'$ instead of $F$, we have a section $s'_f$ in $F'$, a section $s''_f$ in $F'/F$ cutting out $M_f$ in $M'_f$ and the identity \begin{align*} {\Bbb Z}_K' &= e(F'|_{M'_f} s'_f)\cap [M'_f] \\ &= e(F|_{M_f},s_f)\cap e(F'/F|_{M'_f}, s''_f) \cap [M'_f] \\ &= e(F|_{M_f},s_f) \cap [M_f] = {\Bbb Z}_K. \end{align*} Note that in the third step we have used the identification $\orr(M_f) = \orr(M'_f) \tensor \orr(F'/F))|_{M_f}$. In particular, if $K' \supset K$ all choices on $K'$ work a fortiori for $K$, so we can pass to the limit. The relation ${\Bbb Z}(s) = [Z(s)]$ for regular sections (property~\ref{smoothcase}), and the compatibility with Euler classes of finite rank bundles (property~\ref{fdeuler}) are now clear from the construction. The stability property~\ref{stability} also follows from the construction. For every compactum $K$, we can choose the finite rank subbundle $F$ as a subbundle of $E'$. Then $\~E \surj\to E''$. Now one checks that by a diagram chase that $$ Z(\~E,\~s) = Z(E'/F|_{Z(E'',s'')}, s'\bmod F) $$ and that { \let\to=\rightarrow \begin{align*} TZ(\~E,\~s) &= \rmmath{Ker}(TM \to \~E) \\ &= \rmmath{Ker}(\rmmath{Ker}(TM \to E'') \to E'/F) \\ &= \rmmath{Ker}(TZ(s'') \to E'/F) = TZ(E'/F). \end{align*} } In particular, the orientations agree. Thus we see that \begin{align*} {\Bbb Z}_K(E,s) &= e(F,s_f)\cap [Z(\~E,\~s)] \\ &= e(F,s_f)\cap [Z(E'/F|_{Z(E'',s'')}, s'\bmod F)] = {\Bbb Z}_K(E'|_{Z(s'')},s'). \end{align*} It only remains to pass to the limit over $K$. To see that $\cH^{\cl}_j(Z(s))$ maps to ${\displaystyle \lim_{\leftarrow}}_C H_j(M,M-C)$ note that for every compact subset $K = C \cap Z(s)$, we constructed a finite dimensional manifold $M_f \supset K$. Then we have maps \begin{align*} \cH_j^{\cl}(Z(s)) \to &\cH_j(Z(s),Z(s)-K) = \cH_j(Z(s) \cap M_f, Z(s) \cap M_f -K) \\ &\to H_j(M_f, M_f - C) \to H_j(M,M-C). \end{align*} Again this map is independent of choices, and we can pass to the limit. The homotopy property of ${\Bbb Z}$ is a formal consequence of the compatibility with finite dimensional Euler classes. Consider the trivial bundle ${\Bbb R}$ over the interval $[-1,2]$ with the one parameter family of sections $\theta - \tau$ where $\theta :[-1,2] \to {\Bbb R}$ is the inclusion and $0 \le \tau \le 1$. Then clearly $e({\Bbb R},\theta) = e({\Bbb R},\theta-1) \in H^1([-1,2],\{-1,2\})$ is the canonical generator. Consider $M\times [-1,2]$. Let $\pi:M \times [-1,2] \to M$ be the projection and $S: M\times [-1,2] \to \pi^*E$ an extension of our one parameter family of sections e.g. $S_t = s_0$ for $t \le 0$ and $S_t = s_1$ for $ t\ge 1$. The bundle $\pi^*E \mathop\oplus {\Bbb R}$ has a one parameter family of sections $(S,\theta- \tau)$. Now \begin{align*} {\Bbb Z}(s_0) &\buildrel \ref{stability}\over= \pi_* {\Bbb Z}(\pi^*E \mathop\oplus {\Bbb R}; (S,\theta)) \\ &\buildrel \ref{fdeuler}\over= \pi_* e({\Bbb R},\theta) \cap {\Bbb Z}(\pi^*E;S) \\ &= \pi_* e({\Bbb R},\theta - 1)\cap {\Bbb Z}(\pi^*E;S) \\ &= \pi_* {\Bbb Z}(\pi^*E \mathop\oplus {\Bbb R};(S,\theta-1)) = {\Bbb Z}(s_1) \end{align*} \end{pf} \begin{Corollary}\label{locChern} (compare \cite[prop. III.2.4]{PT}) Let $M$ be a complex Banach manifold, $E$ a holomorphic vector bundle and $s$ a holomorphic section with zero set $Z(s)$ Assume that $Ds$ is a section of $\Fred_{\Bbb C}^d(TM|_{Z(s)},E|_{Z(s)})$. We say that $Z(s)$ has complex virtual dimension $d$, and that $Ds$ is Fredholm of complex index $d$. Then the localised Euler class ${\Bbb Z}(s)= [Z(s)] \in H^{\cl}_{2d}(Z(s),{\Bbb Z})$, if $Z(s)$ is a local complete intersection of dimension $d$, and more generally \begin{equation}\label{magic} {\Bbb Z}(s) = [c(\Ind(Ds))^{-1} c_*(Z(s))]_{2d} \end{equation} where $c_*(Z(s))$ is the total homological chern class of $Z(s)$ defined analogous to \cite[example 4.2.6]{Fulton} by equation \eqref{homchernclass} and coincides with the Poincar\'e dual of the cohomological chern classes of the tangent bundle if $Z(s)$ is smooth. \end{Corollary} \begin{Remark} If $Z(s)$ is smooth we can even get away with an almost complex manifold $M$ and the assumption that $Ds$ is complex linear. \end{Remark} \begin{Remark} I have tacitly removed $M$ and $E$ from the notation of the homological Chern class $c_*(Z(s))$. I strongly believe that $c_*(Z(s))$ is independent of the embedding but I did not prove this. There is one case where independence of $c_*(Z(s))$ on the embedding can be proved completely analogous to \cite[Example 4.2.6]{Fulton} by simply replacing algebraic arguments by complex analytic ones: if for every $K \subset Z(s)$ compact, there exists a {\sl holomorphic} finite rank sub bundle $F \lhook\nobreak\joinrel\nobreak\to E$ defined over a neighborhood of $K$ such that $F|_K + \rmmath{Im}(Ds)|_K = E|_K$. Then a neighborhood $U_K$ of $K$ in $Z(s)$ sits in a complex rather then almost complex finite dimensional manifold $M_f$. Such a bundle should typically exist if $Z(s)$ has the structure of a quasi projective variety, and $\rmmath{Coker} DS$ has the interpretation of a coherent sheaf as in \cite[\S 5, \S6]{Pidstrigatch:instanton}. \end{Remark} \begin{pf} We will use Mac Phersons graph construction, that is we consider the limit $\lambda \to \infty$ of the map $(\lambda s: 1)$ in ${\Bbb P}(E \oplus {\mathcal O})$ or finite dimensional approximations thereof. We use the notations of the proof of proposition~\ref{locEuler}. For a compactum $K \subset Z(s)$ we choose the finite rank bundle $F$ as follows. It is a complex bundle, and in every point of $Z(s)$ there are sections of $F$ which restricted to a neighborhood are holomorphic sections of $E$ and which span locally a subbundle $F^{\hol} \lhook\nobreak\joinrel\nobreak\to F$, such that $Ds: TE|_{Z(s)} \surj\to E/F^{\hol}|_{Z(s)}$ is a surjection. We do not assume that $F$ is a holomorphic subbundle, because I do not see a reason why such a bundle should exist. However since $F$ is a complex bundle, the quotient $\~E = E/F$ and $$ TM_f|_{Z(s)} = T|_{Z(s)} = \rmmath{Ker} (TM|_{Z(s)} \m@p--\rightarrow{Ds} E|_{Z(s)} \to \~E|_{Z(s)}) $$ are complex bundles. We extend this complex structure on $TM_f$ over all of $M_f$, possibly after shrinking $M_f$, making it into an almost complex manifold of complex dimension $d+ N$. \def\overline{(0,1)}{\overline{(0,1)}} \def\overline{(\lambda s_f,1)}{\overline{(\lambda s_f,1)}} \def\overline{(s_f,0)}{\overline{(s_f,0)}} \def\overline{(s_f,1/\lambda)}{\overline{(s_f,1/\lambda)}} Consider the space ${\Bbb P}(F \oplus {\mathcal O}) \m@p--\rightarrow{\pi} M_f$. Then the total space of $F$ embeds in ${\Bbb P}(F \oplus {\mathcal O})$. The image of the zero section will also be called the zero section, and the complement of $F$ the divisor at infinity. Let $Q$ be the universal quotient bundle. The bundle $Q$ has sections $\overline{(0,1)}$ , and $\overline{(\lambda s_f,1)}$ cutting out the zero section and the graph of $\lambda s_f$ respectively. Equivalently we can cut out the graph of $\lambda s_f$ by $\overline{(s_f,1/\lambda)}$. Then clearly as $\lambda \to \infty$ the graph degenerates to a set contained in the zero set of $\overline{(s_f,0)}$. Now $Z(\overline{(s_f,0)})$ has two ``irreducible components''. One component $\~M_f \lhook\nobreak\joinrel\nobreak\to {\Bbb P} F$ is the closure of the image of $(s_f:0): M_f -Z(s) \to {\Bbb P} F|_{M_f - Z(s)} \subset {\Bbb P}(F\oplus {\mathcal O})$ It will be called the strict transform. The other component is just ${\Bbb P}(F \oplus {\mathcal O})|_{Z(s)}$. Let ${\mathcal E}_f = \~M_f \cap {\Bbb P}(F \oplus {\mathcal O})|_{Z(s)}$. It will be called the exceptional divisor. I claim that \begin{equation}\label{dimclaim} \cH^{\cl}_{2d+2N-1 + i}({\mathcal E}_f) =0 \txt{for} i \ge 0. \end{equation} Accepting this claim we see that $\~M_f$ carries a unique fundamental class $[\~M_f]$ restricting to $[\~M_f - {\mathcal E}_f]$ by the exact sequence $$ \cH_{2d + 2N}^{\cl}({\mathcal E}_f) \to \cH_{2d+2N}^{\cl}(\~M_f) \to H^{\cl}_{2d+2N}(\~M_f - {\mathcal E}_f) \to \cH^{\cl}_{2d+2N - 1}({\mathcal E}_f) $$ Consider $ C' = {\Bbb Z}(\overline{(s_f,0)}) - [\~M_f] \in \cH^{\cl}_{2d + 2N}(Z(\overline{(s_f,0)}))$. Then $C'$ comes from a unique class $C \in \cH^{\cl}_{2d+2N}({\Bbb P}(F \oplus {\mathcal O})|_{Z(s)})$ because of the sequence. $$ 0 \to \cH^{\cl}_{2d+2N}({\Bbb P}(F \oplus {\mathcal O})|_{Z(s)}) \to \cH^{\cl}_{2d+2N}(Z(\overline{(s_f,0)})) \to H^\cl_{2d+2N}(\~M_f - {\mathcal E}_f). $$ Now note that $Q$ restricted to the zero section is canonically isomorphic to $F$. We therefore have the following chain of equivalences \begin{align*} {\Bbb Z}(s)_K &= e(F,s_f) \cap [M_f] \\ &=\pi_* e(\pi^*F,\lambda s_f) \cap e(Q,\overline{(0,1)}) \cap [{\Bbb P}(F\oplus{\mathcal O})] \\ &=\pi_* \(e(Q,\overline{(\lambda s_f,1)})\cup e(Q,\overline{(0,1)})\) \cap [{\Bbb P}(F \oplus {\mathcal O})] \\ &=\pi_* e(Q,\overline{(0,1)}) \cap e(Q,\overline{(s_f,1/\lambda)}) \cap [{\Bbb P}(F \oplus {\mathcal O})] \\ &=\pi_* e(Q,\overline{(0,1)}) \cap \(e(Q,\overline{(s_f,0)})\cap [{\Bbb P}(F \oplus{\mathcal O})]\). \\ &=\pi_* e(Q,\overline{(0,1)}) \cap {\Bbb Z}(\overline{(s_f,0)}) \end{align*} If we accept the claim~\eqref{dimclaim} for a moment and we note that the support of $\~M_f$ and $e(Q,\overline{(0,1)})$ are disjoint we see further that $$ {\Bbb Z}(s)_K = \pi_*e(Q,\overline{(0,1)}) \cap C' = \pi_* e(Q) \cap C $$ If we use that $e(Q)= c_\Top(Q)$ this can be rewritten further to \begin{align*} {\Bbb Z}(s)_K &= [\pi_* c(Q)\cap C)]_{2d} \\ &= [c(F) \pi_*\((1-h)^{-1}\cap C \)]_{2d} \\ &= [c(F-T) \(c(T) s_*(Z(s),M_f)\)]_{2d} \end{align*} where we used the notation $h = c_1({\mathcal O}_{{\Bbb P}(F \oplus {\mathcal O})}(-1))$ and \begin{equation}\label{Segreclass} s_*(Z(s),M_f) \buildrel\mathrm{def}\over= \pi_* (1-h)^{-1} C \end{equation} for the total homological Segre class of the normal cone (this terminology will be justified in a minute). But $c(F-T)= c(\Ind Ds)^{-1}$ and since $T = TM_f$, \begin{equation}\label{homchernclass} c_*(Z(s))\buildrel\mathrm{def}\over= c(T) s_*(Z(s),M_f) \end{equation} is exactly the analogue of the homological chern classes of \cite[example 4.2.6]{Fulton}. We show that $c_*(Z(s))$ does not depend on the choice of $F$. Again it suffices to treat the case that $F'\subset F$. We use primes whenever an object is associated to $F'$. The independence follows directly from a formula for the Segre classes which expresses how it behaves under the extension $M'_f \subset M_f$ in terms of the normal bundle $F/F'$ of $M'_f \subset M_f$. \begin{equation}\label{coneext} s_*(Z(s),M_f) = c(F/F')^{-1} s_*(Z(s),M'_f). \end{equation} Assuming \eqref{coneext}, we see that \begin{align*} c_*(Z(s)) &= c(T)s_*(Z(s),M_f) \\ &= c(T) c(F/F')^{-1}s_*(Z(s),M'_f) = c(T')s_*(Z(s),M'_f). \end{align*} In particular we can take the limit over $K$. Formula \eqref{coneext} is well known for integrable complex manifolds \cite[example 4.1.5]{Fulton}, and we will follow the proof closely. There are two terms in the class $C$ occurring in the definition \eqref{Segreclass} of the Segre class, which we treat separately. \nc\sfprimenul{\overline{(s'_f,0)}} Note that there is a regular section $\sigma$ of $F/F'(1)$ on ${\Bbb P}(F\oplus{\mathcal O})|_{M_f}$ cutting out ${\Bbb P}(F' \oplus {\mathcal O})|_{M_f}$. Therefore \begin{align*} [{\Bbb P}(F'\oplus {\mathcal O})|_{M'_f}] &= e(F/F', s_f \bmod F') \cap [{\Bbb P}(F'\oplus {\mathcal O})|_{M_f}] \\ &= e(F/F',s_f \bmod F') \cap e(F/F'(1),\sigma) \cap [{\Bbb P}(F\oplus {\mathcal O})|_{M_f}]. \end{align*} Since on ${\Bbb P}(F'\oplus {\mathcal O})|_{M_f}$ there is an exact sequence $$ 0 \to Q'\to Q \to F/ F'\to 0, $$ we have $e(Q',\sfprimenul)\cup e(F/F', s_f \bmod F') = e(Q,\overline{(s_f,0)})$. Then the above implies that \begin{align*} {\Bbb Z}(Q',\sfprimenul) &= e(Q',\sfprimenul) \cap [{\Bbb P}(F'\oplus {\mathcal O})|_{M'_f}] \\ &=e(Q,\overline{(s_f,0)} )\cap e(F/F'(1),\sigma) \cap [{\Bbb P}(F\oplus {\mathcal O})|_{M_f}] \\ &= e(F/F'(1),\sigma)\cap {\Bbb Z}(Q,\overline{(s_f,0)}) \end{align*} As for the other term, on $\~M_f$ there is a smooth section in ${\mathcal O}(-1)$ given by $(s_f,0)$ which is an isomorphism ${\mathcal O} \iso {\mathcal O}(-1)$ on $\~M_f -{\mathcal E}$. It follows that $$ [\~M'_f - {\mathcal E}] = e(F/F',s_f \bmod F') \cap [\~M_f - {\mathcal E}] = e(F/F'(1),\sigma) \cap [\~M_f - {\mathcal E}]. $$ Then we have the equality $$ [\~M'_f] = e(F/F'(1),\sigma) \cap [\~M_f]. $$ because both left and right hand side are cycles supported on $\~M'_f -{\mathcal E} \cup {\Bbb P}(F'\oplus {\mathcal O})|_{Z(s)}$ restricting to $[{\mathcal M}'_f -{\mathcal E}]$. For the computation of the Segre class we can forget about the support given by $\sigma$ and use $$ e(F/F'(1)) = c_\Top(F/F'(1)) = \sum c_{\Top-j}(F/F')h^j. $$ Thus we finally get the expression \begin{align*} s_*(Z(s), M'_f) &= \pi_*\(\sum h^{i+j} c_{\Top-j}(F'/F) \cap ({\Bbb Z}(Q,\overline{(s_f,0)}) - [\~M_f])\) \\ &= c(F'/F) s_*(Z(s),M_f) \end{align*} which we set out to prove. \bgroup \def\eN{\eN} It remains to prove the claim~\eqref{dimclaim}. We first turn to the case that $Z(s)$ is smooth but possibly of the wrong dimension. This condition implies that $\rmmath{Im} Ds|_T \subset F$ has constant rank over $Z(s)$ because $\ker Ds|_{T}= \ker Ds = TZ(s)$. Then $\rmmath{Im} Ds|_T$ is just the normal bundle $\eN$ of $Z(s)$ in $M_f$. Now let us identify the limit set $(s_f: 1/\lambda)(M_f)$ when $\lambda \to \infty$. If we have a smooth path $\gamma$ with $\gamma(0) = x_0 \in Z(s)$, then we see that $ \lim_{t\to 0} (s_f:0)(\gamma(t)) = (Ds_f(\ddt|_0\gamma):0). $ Therefore $\~M_f$ is just the blowup $\^M_f$ of $Z(s)$ in $M_f$. This makes sense even though $M_f$ is only an almost complex manifold since the normal bundle $\eN$ has a complex structure. The blow up is obtained abstractly by identifying a tubular neighborhood $N_\epsilon$ of $Z(s)$ with the normal bundle, and replacing $N_\epsilon$ with $I = \{(l, x) \in {\Bbb P}\eN \times N_\epsilon \mid l \ni x\}$. It is an almost complex manifold, so certainly carries a fundamental class $[\~M_f]$. It is also clear that ${\mathcal E}_f = {\Bbb P}\eN$ is a submanifold of real codimension $2$, and certainly satisfies the claim~\eqref{dimclaim}. Let ${\mathcal O}({\mathcal E}_f)$ be the smooth complex line bundle on the blow-up $\^M_f$ defined by the exceptional divisor ${\mathcal E}_f$, and let $z \in A^0({\mathcal O}(E))$ be a section cutting out ${\mathcal E}_f= {\Bbb P}\eN$ with the proper orientation i.e. ${\Bbb Z}({\mathcal O}({\mathcal E}_f),z) = [{\mathcal E}_f]$. On $\^M_f$ the pulled back section is of the form $s_f = z \^s_f$ with $\^s$ nowhere vanishing. Therefore the limit set of $(s_f:1/\lambda)(\^M_f)$ in ${\Bbb P}(F \oplus {\mathcal O})|_{\^M_f}$ as $\lambda \to \infty$ is just $(\^s:0)(\^M_f) \cup D$ where $D \subset {\Bbb P}(F \oplus{\mathcal O})|_{{\mathcal E}_f}$ is the ${\Bbb P}^1$ bundle joining the zero section $(0:1)|_{{\mathcal E}_f}$ and the section $(\^s_f:0)$. Then down on $M_f$ the limit set of $(s_f: 1/\lambda)(M_f)$ is just $\~M_f \cup C{\mathcal E}_f$, where $C{\mathcal E}_f$ is cone bundle over $Z(s)$ joining ${\mathcal E}_f \subset \~M_f$ and the zero section. Now $C{\mathcal E}_f$ represents the homology class $C$. Thus $$ s_*(Z(s),M_f) = \pi_* (1-h)^{-1}C{\mathcal E}_f = \pi_* (1-h)^{-1} {\mathcal E}_f = \pi_* (1-h)^{-1} {\Bbb P}\eN = s(\eN) \cap [Z(s)] $$ Therefore if $Z(s)$ is smooth we find the expected formula $$ c_*(Z(s)) = c(TM_f)s(\eN) \cap [Z(s)] = c(TZ(s)) \cap [Z(s)]. $$ Note that in deriving this formula we have not really used the holomorphicity of $s$. It was sufficient that $M$ has an almost complex structure and that $Ds$ is complex linear. Replacing manifolds by stratified spaces the proof carries over essentially verbatim if $Z(s)$ is a local complete intersection since this condition implies that $Ds|_{T}$ has constant rank, and that we have a well defined normal bundle. \egroup In proving the claim \eqref{dimclaim} in the general case we use holomorphicity more strongly. We first blow up $Z(s)^{\mathrm{red}}$ in $M$ to get a new infinite dimensional analytic space $\^M$. That this is possible follows from the local analysis of the normal cone in~\cite[\S III.1]{PT}. Locally on $M$, the exceptional divisor ${\mathcal E} \subset \^M$ can be described as follows. Locally on $M$ we have an exact sequence of holomorphic bundles $$ 0 \to F^{\hol} \to E \to \~E^{\hol} \to 0, $$ such that $TM|_{Z(s)} \surj\to \~E|_{Z(s)}$ is surjective, i.e. locally $F^{\hol}$ can take the role of $F$. Further, locally we can split the sequence since $F^{\hol}$ has finite rank. Let the holomorphic subbundle $\~{\~E} \subset E$ be a lift of $\~E^{\hol}$. We write $s = s_f^{\hol} \oplus \~{\~s}$ corresponding to the decomposition $E = F^{\hol} \oplus \~{\~ E}$. Then locally ${\mathcal E} \iso {\mathcal E}^{\hol}_f \times_{Z(s)} {\Bbb P}\~{\~E}$, where ${\mathcal E}^{\hol}$ is the exceptional divisor of the blow up of $Z(s)$ in $M_f^{\hol}$, and where $M_f^{\hol}$ is the integrable finite dimensional complex manifold $Z(\~s^{\hol})$. Moreover ${\mathcal E}_f^{\hol}$ is naturally embedded in ${\Bbb P}(F^{\hol} \oplus {\mathcal O})|_{Z(s)} \subset {\Bbb P}(E \oplus {\mathcal O})|_{Z(s)}$. If we are a little more careful and choose $\~{\~E}$ such that ${\Bbb P}\~{\~E}|_{Z(s)} \subset {\mathcal E}$ then ${\mathcal E} = {\mathrm{Join}}({\mathcal E}_f^{\hol}, {\Bbb P}\~{\~E}|_{Z(s)}) \subset {\Bbb P} E|_{Z(s)}$. Let $z \in H^0({\mathcal O}({\mathcal E}))$ be a section vanishing exactly along ${\mathcal E}$. On $\^M$ we can decompose the section as $s = z^n \^s$. Therefore, just as in the previous finite dimensional case, $(s: 1/\lambda)(\^M) \to {\Bbb P}(E \oplus {\mathcal O})$ degenerates to $(\^s: 0)(\^M)\cup n D$ where $D$ is the ${\Bbb P}^1$ bundle over ${\mathcal E}$ joining the zero section $(0:1)|_{{\mathcal E}}$ and $(\^s_f:0)|_{{\mathcal E}}$. Down on $M$, this means that $(s: 1/\lambda)(M) \subset {\Bbb P}(E \oplus {\mathcal O})$ degenerates to $\~M \cup C{\mathcal E}$ where $\~M \subset {\Bbb P} E$ is isomorphic to $\^M$ with $\~M \cap {\Bbb P}(E \oplus {\mathcal O})|_{Z(s)} \iso {\mathcal E}$, and $C{\mathcal E}$ is the cone bundle over $Z(s)$ joining the zero section and ${\mathcal E}$. Now we finally come to our claim~\eqref{dimclaim}. The set ${\mathcal E}_f = {\Bbb P}(F \oplus {\mathcal O}) \cap {\mathcal E}$. At the very beginning we chose $F$ such that $F \supset F^{\hol}$. Locally we define $\~{\~F} = F \cap \~{\~E}$, then locally $F = F^{\hol} \oplus \~{\~F}$ and locally ${\mathcal E}_f = {\mathrm {Join}}({\mathcal E}^{\hol}_f , {\Bbb P}\~{\~F}|_{Z(s)})$. Thus ${\mathcal E}_f$ is a stratified space of real dimension $2d +2N-2$, and we are done. \end{pf} \begin{Remark}\label{Zhat} In the complex case we have obviously defined a class containing more information about the section. Let $$ \widehat{\Bbb Z}(s) = c(\Ind(Ds))^{-1}c_*(Z(s)). $$ \end{Remark} \section{Seiberg Witten classes} We will collect a few facts about Seiberg Witten basic classes in a formulation suitable for arbitrary K\"ahler surfaces. In the usual formulation, these classes are the support of a certain function on the set of $\Spin^c$-structures. However in the presence of 2-torsion, $\Spin^c$-structures cause endless confusion which is why I have chosen to base my exposition on SC-structures \cite{Karrer}. This notion catches the essence of $\Spin^c$-structures, the existence of spinors. It is well suited to the K\"ahler case and is equivalent to that of a $\Spin^c$-structure in dimension 4. For more details see \cite{Karrer}. Let $X$ be a closed oriented manifold of dimension $2n$. Choose a Riemannian metric $g$ with Levi-Civita connection $\nabla^g$, and Clifford algebra bundle $C(X,g) = C(T^{\scriptscriptstyle\vee} X,g)$. There is a natural isomorphism of bundles $c:\wedge ^* T^{\scriptscriptstyle\vee} X \to C(X,g)$ given by anti-symmetrisation. It induces a connection and metric on $C(X,g)$ also denoted $\nabla^g$ and $g$. An {\sl SC-structure} is a smooth complex vector bundle $W$ of rank $2^n$ together with an algebra bundle isomorphism $\rho: C(X,g) \to \mathop{{\mathcal E}\mkern-3mu{\mathit nd}}\nolimits(W)$. In other words an SC structure is a bundle with the irreducible Clifford algebra representation $\Delta$ in every fibre. A section $\phi \in A^0(W)$ is called a (smooth) spinor. An SC-structure exists if and only if $w_2(X)$ can be lifted to the integers \cite[\S 3.4]{Karrer}. Existence will be clear in the case of K\"ahler surfaces. SC-structures admit an invariant hermitian metric i.e. one such that Clifford multiplication by 1-forms is skew hermitian (sh). The chirality operator $\Gamma = (\sqrt{-1})^n c(\Vol_g)$ has square $1$, and is hermitian. Thus $\Gamma$ has an orthogonal eigenbundle decomposition $W = W^+ \oplus W^-$ with eigenvalue $\pm 1$, the positive and negative spinors of the SC-structure. A one form $\omega \in A^1(X)$ defines an skew hermitian map $c(\omega): W^\pm \to W^\mp$ which is an isomorphism away from the zero set of $\omega$. In this paragraph we assume $\dim(X)= 4$. Then $T^{\scriptscriptstyle\vee}_X \iso \mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits(W^+,W^-)^{\mathrm{sh}}$. Let $L_W = \det W^+$. Then $L_W \iso \det W^-$, by the isomorphism induced from Clifford multiplication by a generic $1$-form, which is an isomorphism outside codimension 4. Thus $W$ is a $\Spin^c(4)$-bundle if we identify $$ \Spin^c(4) = \{(U_1,U_2) \in U(2) \times U(2) \mid \det(U_1) = \det(U_2)\}. $$ We recover the usual definition $\Spin^c(4) = \Spin(4)\times_{{\Bbb Z}/2/Z} U(1)$ from the isomorphism $\Spin(4) = SU(2) \times SU(2)$. In any case by chasing around the cohomology sequences of the diagram $$ \arrowlen3em \begin{matrix} 0 &\m@p--\rightarrow{}&{\Bbb Z}/2{\Bbb Z}&\m@p--\rightarrow{}&\Spin^c(4)&\m@p--\rightarrow{}&\SO(4)\times U(1)&\m@p--\rightarrow{}1 \\ & &\vm@p\Vert{}& &\vm@p\uparrow{} & &\vm@p\uparrow{} \\ 0 &\m@p--\rightarrow{}&{\Bbb Z}/2{\Bbb Z}&\m@p--\rightarrow{}&\Spin(4) &\m@p--\rightarrow{}&\SO(4) &\m@p--\rightarrow{}1 \end{matrix} $$ we see that $L_W + w_2(X) \equiv 0 \pmod 2$, and that this is the only obstruction to lifting the $SO(4)\times U(1)$ bundle to $\Spin^c(4)$. If $H^2(X,{\Bbb Z})$ has no 2-torsion, the line bundle $L\equiv w_2(X)$ determines such a lift completely, and it is common to speak of the $\Spin^c$-structure $L$. An {\sl SC-Clifford module} $(S,\<,>,\nabla)$, is an SC-structure with a non-degenerate invariant hermitian metric $\<,>$ and a unitary Clifford connection $\nabla$ i.e. a unitary connection such that for all vector fields $X$, spinors $\phi \in A^0(S)$, and $1$-forms $\omega$ we have $$ \nabla_X (\omega \cdot \phi) = (\nabla^g_X \omega)\cdot \phi + \omega \cdot \nabla_X \phi. $$ The {\sl Dirac operator} $\delbar$ of a Clifford module is the composition $$ A^0(W) \m@p--\rightarrow{\nabla} A^1(W) \m@p--\rightarrow{\cdot} A^0(W). $$ It is an elliptic self adjoint first order differential operator, and it maps positive spinors to negative ones and vice versa (i.e. $\delbar:A^0(W^\pm) \to A^0(W^\mp)$). Since $\rho$ is parallel, $\nabla$ respects the decomposition $W = W^+ \oplus W^-$. Thus $\nabla$ induces a connection on $L_W$ with curvature $F$. Much of the usefulness of SC-structures is a consequence of the following easy lemma. \begin{Lemma} The set of isomorphism classes $\SC$ of SC-structures is an $H^2(X,{\Bbb Z})$ torsor i.e. if $\SC \ne \emptyset$ and we fix an SC-structure $W_0$, then for every SC-structures $W_1$, there exits a unique line bundle ${\mathcal L}$ such that $W_1 = W_0\tensor{\mathcal L}$. Every SC-structure $S$ admits a Clifford module structure $(W, \<,>, \nabla)$. If we fix one SC-Clifford module $(W_0, \<,>_0, \nabla_0)$, there is a unique triple $({\mathcal L},h,d)$ of a smooth line bundle ${\mathcal L}$, with hermitian metric $h$ and unitary connection $d$, such that \begin{equation}\label{SCrepr} (W,\<,>,\nabla) \iso (W_0, \<,>_0, \nabla_0) \tensor ({\mathcal L}, h,d). \end{equation} \end{Lemma} \begin{pf} Clearly if $W_0$ is an SC structure, so is $W_0\tensor {\mathcal L}$ for every line bundle ${\mathcal L}$. Conversely, the bundle of Clifford linear homomorphisms ${\mathcal L}(W_0,W) = \mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits_C(W_0,W)$ has rank~1, and the natural map $W_0 \tensor {\mathcal L}(W_0, W) \to W$ is an isomorphism. For existence of a Clifford module structure see \cite[prop. 4.2.1, 4.5.1]{Karrer}. It will be clear for K\"ahler surfaces. It follows directly from the definition of a Clifford module that the natural connection and metric on $\mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits(W_0,W)$ leaves ${\mathcal L}(W_0,W)$ invariant. Hence there is an induced metric and connection $(h,d)$ on ${\mathcal L}(W_0,W)$, which has property \eqref{SCrepr}. Conversely if $(W,\<,>,\nabla)$ is defined by equation \eqref{SCrepr}, then $$ ({\mathcal L},h,d) = \mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits_C\((W_0,\<,>_0,\nabla_0)\,,\, (W_0,\<,>_0,\nabla_0)\tensor({\mathcal L},h,d)\) $$ which proves uniqueness. \end{pf} If a base SC-structure is chosen, the line bundle ${\mathcal L}$ will be called the twisting line bundle. There is a natural gauge group $\G^{\Bbb C}$ acting on a Clifford module, the group of all smooth invertible Clifford linear endomorphisms. $\G^{\Bbb C}$ can be canonically identified with $C^\infty(X, {\Bbb C}^*)$. In the representation~\eqref{SCrepr}, $\G^{\Bbb C} = C^\infty(X, {\Bbb C}^*)$ acts in the usual way on the set of metrics and unitary connections on the twisting line bundle ${\mathcal L}$. Since every hermitian metric on a line bundle is gauge equivalent, so is every Clifford invariant metric on a Clifford module. Thus, up to gauge we can fix an invariant metric and we are left with a residual gauge group $\G = C^\infty(X, \mathrm{U}(1))$. The set of Clifford connections $\A$ on a fixed hermitian SC structure $(W,\<,>)$ (i.e. Clifford module structures) is an affine space modeled on $\sqrt{-1} A^1_{\Bbb R}(X)$. Using the representation~\eqref{SCrepr} and harmonic representatives, one shows that the set of connections mod gauge is $$ \B = \A / \G \iso \sqrt{-1} A^1_{\Bbb R}(X) / d\log C^\infty(X,\mathrm{U}(1)) \iso H^1_{DR}(X)/H^1(X,{\Bbb Z}) \oplus \ker d^* $$ We set $\Pee^* = \A \times A^0(W^+)^* / \G$. It is a ${\Bbb C}{\Bbb P}^\infty \times {\Bbb R}^+$ bundle over $\B$. Thus $\Pee^*$ has the homotopy type of $(S^1)^{b_1(X)} \times {\Bbb C}{\Bbb P}^\infty$. There is an alternative description of $\B$ and $\Pee^*$ that will be useful. Let $\A^{\Bbb C}$ be the set of all Clifford connections, and $\Herm$ the set of all hermitian metrics on ${\mathcal L}$. Let $$ \Amod = \{(\nabla,<,>),\ \nabla \txt{is} <,>\hbox{-unitary}\} \subset \A\times\Herm $$ be the set of Clifford module structures. Fix a metric $<,>_0$ and a $<,>_0$-unitary connection $\nabla_0$. The representation $\nabla = \nabla_0 + a$, models $\A^{\Bbb C}$ on $A^1_{\Bbb C}(X)$, and the representation $<,> = e^f < , >_0$ models $\Herm$ on $A^0_{\Bbb R}(X)$. A pair $(\nabla,<,> ) \in \Amod$ if and only if $a+ \bar a = df$. In particular $a$ is determined by $f$ and its imaginary part, so $\Amod$ is modeled on $A^0_{\Bbb R}(X)\times A^1_{\Bbb R}(X)$. Now the diagonal action of $\G^{\Bbb C}$ on $\A^{\Bbb C} \times \Herm$ leaves $\Amod$ invariant. Our alternative description of $\B$ and $\Pee^*$ is \begin{equation}\label{alternative} \Pee^* = \Amod \times A^0(W^+)^*/ \G^{\Bbb C} \to \B = \Amod/ \G^{\Bbb C} \end{equation} Finally, to do decent gauge theory we have to complete to Banach spaces and -manifolds. Seiberg Witten theory works fine with an $L^p_1$ completion of $\A$, $\A^{\Bbb C}$, and $A^0(W^+)$ and an $L^p_2$ completion of $\G$, $\G^{\Bbb C}$ and $\Herm$ if $p > \dim X$. In this range $L^p_1 \lhook\nobreak\joinrel\nobreak\to C^0$, and therefore the two possible $L_p$ descriptions of $\Pee^*$ and $\B$ coincide. On the other hand, the Sobolev range does not seem optimal: with more care and work one can probably use all $p$-completions with $2 - \dim(X) / p > 0$. We will suppress completions from the notation, explicitly mentioning completions if necessary. {}From now on we assume $\dim X = 4$. Fix an SC structure $W$ and choose an invariant hermitian metric $\<,>$. Choose a Riemannian metric $g$ and a real 2 form $\epsilon$, which are {\sl admissible} in the following sense: $L_W$ admits no connection with $F^+ = -2\pi \sqrt{-1}\epsilon^+$, where as usual $+$ means taking the self dual part. Admissible metrics and forms exist if $b_+ \ge 1$, since the condition is certainly satisfied if $c_1(L_W) \not\in \epsilon^{\mathrm{harm}} + H^-_g$ where $H^-_g$ is the space of $g$-anti-self-dual closed forms, and ``harm'' means projection to the harmonic part. Note that no use of Sard-Smale is made to define admissibility. Actually for most of our purposes it would be enough to let $\epsilon$ be a closed (hence harmonic) self-dual form. By a transversality argument \cite{Donaldson:intersectionform}, the admissible (metrics,forms) form a connected set if $b_+ \ge 2$. We say that a metric $g$ is admissible if $(g,0)$ is. Even if $b_+ = 1$, all metrics are admissible when $L_W^2 \ge 0$, and $L_W$ is not torsion. In dimension $4$, the anti-symmetrisation map gives an isomorphism $c:\Lambda ^+ \iso \rmmath{End}_0^{sh}(W^+)$ between the real self-dual forms and the traceless skew hermitian endomorphisms of $W^+$. This special phenomenon allows us (or rather Seiberg and Witten) to write down the monopole equations \cite{Witten} \begin{align} \label{SW1} \delbar \phi &= 0 \qquad \phi \in A^0(W^+) \\ \label{SW2} c(F^+) &= 2\pi \phi\<\phi,-> - \pi |\phi|^2 -2\pi\sqrt{-1} c(\eps^+). \end{align} Let ${\mathcal M} = {\mathcal M}(W,g,\epsilon) \subset \Pee^*$ be the space of solutions modulo gauge. As a technical remark, note that we use the conventions of \cite{BGV}, and that in their conventions the Weitzenb\"ock (Lichnerowitz) formula restricted to $W^+$ reads $$ \delbar^2 = \nabla^* \nabla + r /4 + c(F^+/2) $$ (\cite[th. 3.52]{BGV} and the observation that the twisting curvature of an SC structure is $1/\rmmath{rank}(W^+)$ times the curvature on $\det(W^+)$.) The sign difference in the $c(F^+)$ term in \cite[lemma 2]{KM:Thom} explains the relative change of sign with respect to \cite[formula $(*)$]{KM:Thom} in the Seiberg Witten equations. It is chosen in such a way that the Weitzenb\"ock formula gives $C^0$ control on the harmonic positive spinor $\phi$. A basic property of the monopole equation noted by Witten, which follows from the Weitzenb\"ock formula \cite[lemma 2]{KM:Thom} or a variational description \cite[Section 3]{Witten}, is the following \begin{Proposition}\label{Bochner} The monopole equations have no solution with $\phi \ne 0$ if the metric has positive scalar curvature. \end{Proposition} Alternatively we can define ${\mathcal M}$ as the zero of a Fredholm section in an infinite dimensional vector bundle. Let $\W^\pm = (\A\times A^0(W^+)^* \times_\G A^0(W^\pm) \to \Pee^*$. Then ${\mathcal M}$ is the zero of the section in $\W^- \oplus A^+(X)$ given by the monopole equations~\eqref{SW1}, and \eqref{SW2}. To see that it is actually a Fredholm section we linearise the equations, assuming that $(\nabla,\phi)$ is a solution, and $(\nabla + \eps a, \phi + \eps \psi)$ with $a \in \sqrt{-1}A_{\Bbb R}^1(X)$ and $\psi \in A^0(W^+)$ is a solution up to order 1 in $\eps$. We get (c.f \cite[eq.2.4]{Witten}) \begin{gather*} \delbar \psi + a\cdot \phi = 0 \\ c^{-1}(2\pi(\phi\<\psi,-> + \psi\<\phi,-> - \rmmath{Re}\<\phi,\psi>) -d^+a = 0. \end{gather*} The tangent space of the $\G$-orbit of $(\nabla,\phi)$ is $\{(a,\psi) = (- d u, u \phi), \ u \in \sqrt{-1}A^0_{\Bbb R}(X)\}$. Thus the Zariski tangent space of ${\mathcal M}$ in $(\nabla, \phi)$ is the first cohomology of the Fredholm complex $$ \sqrt{-1} A_{\Bbb R}^0(X) \to \sqrt{-1}A_{\Bbb R}^1(X) \oplus A^0(W^+) \to \sqrt{-1} A^+_{\Bbb R}(X) \oplus A^0(W^-), $$ where the maps are given by the left hand side of the linearised equations. The virtual dimension is given by Atiyah Singer index formula and is \begin{equation}\label{vdim} d(W) = \vdim_{\Bbb R}({\mathcal M}) = {\numfrac14}(L_W^2 - (2 e(X) + 3 \sign(X))), \end{equation} where $e(X)$ is the topological Euler characteristic, and $\sign(X)$ the signature \cite[eq. 2.5]{Witten}. The crucial property that makes Seiberg Witten theory so much easier than Donaldson theory is \begin{Proposition}\cite[Corollary 3]{KM:Thom},\cite[\S 3]{Witten} The moduli space ${\mathcal M}$ is compact. For fixed $c >0$ there are only finitely many SC-structure $W$ with $d({\mathcal M}(W)) \ge -c$ and ${\mathcal M}(W,g,\epsilon) \ne \emptyset$. \end{Proposition} Note that for generic pairs $(g,\epsilon)$, moduli spaces of negative virtual dimension are empty, but I do not see an a priori reason why moduli spaces of arbitrary negative virtual dimension should not exist for special pairs. Likewise for generic pairs the moduli space is smooth of dimension $d(W)$ \cite{KM:Thom}. However we have no need for this fact. The index bundle $\Ind(Ds)$ of the deformation complex can be deformed by compact operators (over a compact space !) into the sum of the index of the signature complex and the index of the complex dirac operator. Thus the determinant line bundle $\det(\Ind(Ds)$ of the index is naturally oriented by choosing an orientation for $\det H^1(X,{\Bbb R})^{\scriptscriptstyle\vee} \tensor H^+(X,{\Bbb R})$. We will in fact assume that an orientation for both $H^+$ and $H^1$ is chosen. Suppose further that the pair $(g,\epsilon)$ is admissable (i.e. ${\mathcal M}((W,g,\epsilon) \subset \Pee^*$), then proposition \ref{locEuler} in the previous section gives us a homology class $\MM \in H_{d(W)}(\Pee^*)$, i.e. a homology class of the proper virtual dimension even if ${\mathcal M}$ is not smooth, not reduced and not of the proper dimension (note that in our case the moduli space ${\mathcal M} = Z(s)$ is compact, and homology with closed support is just ordinary homology). In case ${\mathcal M}$ is smooth and has the proper dimension it is just the fundamental class. The class $\MM$ depends only on the connected component of $(g,\epsilon)$ in the space of admissable pairs, by the homotopy property of the localised Euler class proposition \ref{locEuler}.\ref{homotopy}. In particular ${\mathcal M}$ is independent of the admissable pair if $b_+ \ge 2$. If $b_+ = 1$ the choice of an orientation of $H^+$ is the choice of a connected component in $\{\omega^2 >0\} \subset H^2(X,{\Bbb R})$. It will be called the forward timelike cone. For every metric $g$ let $\omega_g$ be the unique self dual form in the forward timelike cone with $\int \omega^2 = 1$. For a pair $(g,\epsilon)$ and an SC-structure $W$ define the {\sl discriminant} \begin{equation}\label{discriminant} \Delta_W(g,\epsilon) = \int (c_1(L_W) - \epsilon) \omega_g \end{equation} A pair $(g,\epsilon)$ is admissable if the discriminant $\Delta_W(g,\epsilon) \ne 0$, because it means precisely that $c_1(L_W) \notin \epsilon^{\harm} + H^-$. Clearly the discriminant depends only on the period $(\omega_g, {\epsilon^+}^{\mathrm harm})$. \begin{Lemma} If $b_+ = 1$ a pair $(g,\epsilon)$ is admissable if and only if the discriminant $\Delta_W(g,\epsilon) \ne 0$. There are exactly two connected components of admissable pairs labeled by the sign of the discriminant. \end{Lemma} \begin{pf} Suppose two pairs $(g_i,\epsilon_i)$, $i=0,1$, have discriminants $\Delta_i$ of equal sign. Connect them by a path $(g_t, \epsilon_t)$ in the space of all pairs. Let $(\omega_t, \epsilon^{+,\harm}_t)$ be the corresponding path of periods. Then the discriminant $$ \Delta_t = \int (c_1(L_W) - \epsilon^{+,\harm}_t)\omega_t $$ is continuous in $t$ but may change sign. However if we modify the path by setting $$ \epsilon'_t = \epsilon_t + (\Delta_t - (1-t)\Delta_0- t\Delta_1)\omega_t $$ then using $\Delta_W(g,\epsilon + \delta) = \Delta_W(g,\epsilon) - \int \delta\wedge\omega_g$ and $\int \omega^2 = 1$ we see that $$ \Delta'_t= \Delta_W(g_t,\epsilon'_t) = (1-t)\Delta_0 + t\Delta_1. $$ In particular $\Delta'_t$ does not change sign, so that $(g_t,\epsilon'_t)$ is a path of admissable pairs. Conversely if $c_1(L_W) \in \epsilon^\harm + H^-$, then any connection $\nabla$ with induced Chern form $\epsilon^\harm$ determines a ``reducible'' solution $(\nabla,0) \in \Pee- \Pee^*$ of the monopole equations. \end{pf} \begin{Definition} If $b_+ \ge 2$, the {\sl SW-multiplicity} is the map \begin{align*} n:\SC &\to \Lambda^*H^1(X,{\Bbb Z})[t] \iso H_*(\Pee^*,{\Bbb Z}) \\ W &\mapstochar\nobreak\to \MM(W,g,\epsilon) \end{align*} where $(g,\epsilon)$ is any $W$-admissable pair. If $b_+ = 1$ the {\sl SW-multiplicities} $n_+$ and $n_-$ are defined similarly but with pairs $(g_\pm,\epsilon_\pm)$ having positive respectively negative discriminant. \end{Definition} It should be remarked that the SW-multiplicity (ies) depend(s) implicitly on the orientation of $H^+$ and $H^1$. For $b_+ > 1$ this is only a matter of sign, but for $b_+ = 1$ the orientation of $H^+$ determines in addition which invariant is $n_+$ and which is $n_-$. All known examples with $b_+ \ge 2$ have non trivial multiplicities only when the virtual dimension $d(W) = 0$. However for surfaces with $p_g =0$ it is easy to give examples with one of $n_\pm$ is non trivial for $d(W) >0$ we will in fact use such an invariant. If $b_1 \ne 0$, the $H^1$ part of the multiplicity becomes essential. \begin{Remark} Since $H_i(\Pee^*) = 0$ for $i <0$, a moduli space of negative virtual dimension never defines a nontrivial class. Thus if for a class $L \in H^2(X,{\Bbb Z})$ there exists an SC-structure $W$ with $L = c_1(L_W)$ and the multiplicity $n(W) \ne 0$ (respectively one of $n_{\pm}(W) \ne 0$ then $L^2 \ge 3 e(X) + 2\sign(X)$ (c.f. equation~\eqref{vdim}). \end{Remark} \begin{Remark}\label{specialchamber} In the case $b_+ = 1$ we can alternatively consider the multiplicity as depending in addition on a chamber structure in $$ \Gamma = \{(\omega,\epsilon) \in H^2(X,{\Bbb R})^2 \mid \omega^2 = 1,\ \omega_0 > 0\} $$ where a chamber is defined by walls which are in turn defined by all classes $L \equiv w_2(X)$ through equation~\eqref{discriminant}. This is particularly useful when we consider structures with $L_W^2 \ge 0$, $L_W$ is not torsion. Then all pairs $(g,0)$ are admissable and have discriminant of equal sign, because the forward timelike cone is strictly on one side of the hyperplane $L_W^\perp \subset H^2(X,{\Bbb R})$. Thus for this subset we have a preferred chamber. \end{Remark} We will say that $L\in H^2(X,{\Bbb Z})$ with $L \equiv w_2(X)$ has non trivial multiplicity if there is an SC-structure $W$ such that $L = c_1(L_W)$ and $W$ has non trivial multiplicity. If $b_+ =1$ we will further qualify which multiplicity is non trivial (i.e. $n_+$ or $n_-$) or which chamber is chosen. We will simply write $n(L) \ne 0$ or $n_+(L) \ne 0$ etc. A final and important piece of general theory is the following blow-up formula \cite{Stern:talk},\cite[\S 8]{FS:rational}. We will give a proof valid for K\"ahler surfaces in section~\ref{computations}. \begin{Theorem}\label{blowup4} Let $X$ be a closed oriented 4-manifold with $b_+ \ge 1$. An SC-structure $\~W$ on $X\#\Pbar^2$ can be decomposed as $\~W = W \# W_k^{\Pbar^2}$, with determinant lines $L_{\~W} = L_W + (2k+1)E$. If the multiplicity $n_{(\pm)}(\~W) \ne 0$ then $d(\~W) = d(W) - k(k+1) \ge 0$, and the multiplicity $n_{(\pm)}(W) \ne 0$. Moreover if $L_{W_{\Pbar^2}} = \pm E$ (i.e. $E \cdot L_{\~W} = \pm 1$) then $n_{(\pm)} (\~W) = n_{(\pm)}(W)$ under the identification $H^1(X,{\Bbb Z}) \iso H^1(\~X,{\Bbb Z})$. \end{Theorem} Here, $ n_{(\pm)} = n$ if $b_+ > 1$, and if $b_+ = 1$, it is understood that we compare say $n_+(W \# W_k^{\Pbar^2})$ with $n_+(W)$. \section{Seiberg Witten classes of K\"ahler surfaces} {}From now on, $(X, \Phi)$ denotes a K\"ahler surface. Then $X$ has a natural base SC-structure $$ W_0 = \Lambda^{0,*} X $$ with Clifford multiplication given by $$ c(\omega^{10} + \omega^{01}) = \sqrt2\(- i(\omega^{10}) +\eps(\omega^{01}) \), $$ where $i$ is contraction and $\eps$ is exterior multiplication. The metric and connection induced by the K\"ahler structure on $\Lambda^{0*} X$ define a Clifford module structure on $W_0$. For an arbitrary SC structure $W$ = $W({\mathcal L})$ the spinor bundles are of form $$ W^+ = (\Lambda^{00} \oplus \Lambda^{02})\tensor {\mathcal L}, \qquad W^- = \Lambda^{01}({\mathcal L}). $$ and $L_W = \det(W^+) = -K \tensor {\mathcal L}^2$ (c.f. lemma~\ref{SCrepr}). We call ${\mathcal L}$ the twisting line bundle. We now turn to the monopole equations (see also \cite[Section 4]{Witten}). In the decomposition of $W^+$, a positive spinor will be written $\phi = (\alpha,\beta)$. The Dirac equation is then \cite[Propos. 3.67]{BGV}. $$ \delbar \phi = \sqrt2 (\dbar\alpha + \dbar^* \beta)= 0. $$ \nc\dzdz{dz_1\wedge dz_2} \nc\dzbardzbar{d\bar z_1 \wedge d \bar z_2} \def\dzdzbar#1{dz_#1\wedge d\bar z_#1} Since $X$ is K\"ahler, we can locally choose holomorphic geodesic coordinates $(z_1,z_2)$. A basis of the self dual forms is then the K\"ahler form $\Phi = \numfrac {\sqrt{-1}}2(\dzdzbar1 + \dzdzbar2)$, $\dzdz$ and $\dzbardzbar$. Let $h$ be an hermitian metric on ${\mathcal L}$. Choose a unit generator $e$ for ${\mathcal L}$, then an orthonormal basis for $W^+ $ is $e$ and ${\numfrac12} e\dzbardzbar$. Using the definition of Clifford Multiplication we compute: \begin{align*} c(\Phi)e &= \numfrac{\sqrt{-1}}2 (-i(dz_1) \eps(d \bar z_1) + \eps(d \bar z_1)i(d z_1) -i(dz_2) \eps(d \bar z_2) + \eps(d \bar z_2)i(d z_2))e \\ &= - 2\sqrt{-1} e. \end{align*} In exactly the same way we compute $c(\Phi)$, ${\numfrac12} e \dzbardzbar$, and the action of $c(\dzdz)$ and $c(\dzbardzbar)$ on $e$ and ${\numfrac12} e \dzbardzbar$. The result in matrix form is given by $$ c(\Phi) = \begin{pmatrix} - 2\sqrt{-1} & 0 \\ 0 & 2\sqrt{-1} \end{pmatrix} \quad c(\dzdz) = \begin{pmatrix} 0 & -4 \\ 0 & 0 \end{pmatrix} \quad c(\dzbardzbar) = \begin{pmatrix} 0 & 0 \\ 4 & 0 \end{pmatrix}. \hskip 0pt minus 1 fil $$ \def\bet#1#2{\beta_{\dot #1\dot#2}} On the other hand, writing $\alpha = \alpha_e e$, and $\beta = {\numfrac12} \bet12 e \dzbardzbar$, $$ (\alpha + \beta) \< \alpha +\beta,-> = \begin{pmatrix} |\alpha_e|^2 & \alpha_e\bar\bet12 \\ \bar\alpha_e\bet12 & |\bet12|^2 \end{pmatrix}. $$ Thus if we define $\alpha^* = h(\alpha,-)$, $\beta^*= h(\beta,-)$ and take the trace free part, we get the healthy global expression $$ (2\pi(\alpha + \beta)\< \alpha + \beta, ->)_0= -2\pi\sqrt{-1} c\({\numfrac12} (|\beta|_h^2 - |\alpha|_h^2) \Phi + \sqrt{-1}(- \alpha \beta^* + \beta \alpha^*)) \) $$ Plug all this in the monopole equations~\eqref{SW1},\eqref{SW2}. Writing $c_1(F) =\numfrac{-1}{2 \pi i} F$, and using that $\Lambda \Phi =2$ the monopole equation for a K\"ahler metric and perturbation $\epsilon = \lambda \Phi$ can be rewritten to \begin{align} &\dbar \alpha + \dbar^* \beta = 0 \label{cSW1} \\ & F^{02} = 2\pi \beta \alpha^* \label{cSW2} \\ &F^{20} = -2\pi \alpha \beta^* \label{cSW3} \\ &\Lambda c_1(F)^{11} = (|\beta|^2 - |\alpha|^2) + 2\lambda. \label{cSW4} \end{align} Note that $F$ is the curvature on $L_W$, but that these are equations for a unitary connection $d =\dee + \dbar$ on ${\mathcal L}$ and sections $\alpha \in A^{00}({\mathcal L})$, and $\beta \in A^{02}({\mathcal L})$ through the identity $F = -F(K) + 2 F({\mathcal L},d)$. Here $F(K)$ is the curvature of the canonical line bundle i.e. minus the Ricci form. In terms of the twisting bundle the virtual (real) dimension of the moduli space reads \begin{equation}\label{cvdim} d({\mathcal L}) = d(\Lambda^{0*}({\mathcal L})) = {\numfrac14}(L^2 - K^2) = {\mathcal L}\cdot ({\mathcal L}- K). \end{equation} A more precise description is given by \begin{Proposition} \label{Kahlermonopoles} {\sloppy A necessary condition for the existence of solutions to the mono\-pole equations~\eqref{cSW1} to \eqref{cSW4}, is that $({\mathcal L},\dbar)$ is a holomorphic line bundle, and that } \begin{align} -\deg_\Phi(K) &\le \deg_\Phi(L) < \int(\lambda \Phi^2), \txt{or} \label{case0} \\ \int \lambda \Phi^2 &< \deg_\Phi(L) \le \deg_\Phi(K), \txt{or} \label{case2} \\ \int \lambda \Phi^2 &= \deg_\Phi(L) \label{singcase} \end{align} In particular $L_W = -K \tensor {\mathcal L}^2$ has a natural holomorphic structure. In case \eqref{case0} the moduli space ${\mathcal M} = {\mathcal M}({\mathcal L}, \Phi, \lambda)$ of solutions can be identified as a real analytic space with the moduli space of pairs of a holomorphic structures $\dbar$ on ${\mathcal L}$, and a divisor $\alpha \in |({\mathcal L},\dbar)|$, in particular the Zariski tangent space in $(\dbar,\alpha)$ is canonically identified with $H^0({\mathcal L}|_{Z(\alpha)})$. In case \eqref{case2} the moduli space ${\mathcal M}$ of solutions can be identified with the moduli space of pairs of a holomorphic structure $\dbar$ on ${\mathcal L}$, and an element $\beta \in {\Bbb P} H^2({\mathcal L}) = |K\tensor {\mathcal L}^{\scriptscriptstyle\vee}|^{\scriptscriptstyle\vee}$, in particular the Zariski tangent space at $(\dbar,\beta)$ is isomorphic to $\overline{H^0(K\tensor {\mathcal L}|_{Z(\bar\beta)})}$. In case \eqref{singcase} the ``moduli space'' ${\mathcal M} \subset \Pee - \Pee^*$ (i.e. $\alpha = \beta = 0$) can be identified with the space of holomorphic structures $\dbar$ on ${\mathcal L}$. \end{Proposition} \begin{pf} Combining~\eqref{cSW1} and ~\eqref{cSW2} yields \begin{equation}\label{positivity} \dbar\dbar^* \beta = -\dbar^2 \alpha = - F^{02}\alpha = -2\pi |\alpha|^2 \beta. \end{equation} Integrating both sides against $\<\beta,->$, immediately gives that $\alpha\beta= 0$ and $\dbar\beta = \dbar\alpha = 0$. Thus $F^{02} = F^{20} =0$, Since $F^{02} = 2 F^{02}({\mathcal L},d)$, $\dbar$ is a holomorphic structure on ${\mathcal L}$, and either $0 \ne \alpha \in H^0({\mathcal L})$ and $\beta= 0$ or $0 \ne \beta \in H^2({\mathcal L})$ and $\alpha = 0$, or $\alpha = \beta = 0$. Note that if for example $\alpha \ne 0$, then $\beta = 0$ is cut out transversely by equation~\eqref{positivity}. The last monopole equation~\eqref{cSW4} gives the condition $$ \deg(L) = -\deg(K) + 2\deg({\mathcal L}) = {\numfrac12}\int\Lambda c_1(F)\Phi^2 = {\numfrac12}\int ( |\beta|^2 - |\alpha|^2 + 2\lambda) \Phi^2 $$ which fixes the global $L_2$ norm of $\alpha$ and $\beta$, and determines whether $\alpha \ne 0$ or $\beta \ne 0$ or $\alpha = \beta = 0$. Finally we deal with equation~\eqref{cSW4}. If $\alpha \ne 0$ and $\beta =0$ then we are dealing essentially with the abelian vortex equation studied by Steve Bradlow \cite[\S 4]{Bradlow} Oscar Garcia-Prada and earlier in a different guise by Kazdan Warner \cite{KazdanWarner}. See also \cite{Bradlow:nonabelian} and \cite{OkonekTeleman:coupledSW}. I thank Steve Bradlow for pointing out that almost all of the work had already been done by him and Oscar Garcia-Prada. To identify the moduli space as a real analytic space we just jazz up Bradlow's results a bit. This is necessary because we have to to understand how the moduli space is cut out in order to apply the localised Euler class machinery in the next section. It is slightly more convenient to use our alternative description~\eqref{alternative} of $\Pee^*$, and solve for a pair $(d_L, h)$ where $h = e^f h_0$ is a hermitian metric on $L$ and $d_L= \dee + \dbar = d_0 + a$ is $h$ unitary, and mod out the full gauge group $\G^{\Bbb C}$ of all complex nowhere vanishing functions. To be precise we take $d_L$ in $L^p_1$, and $\G^{\Bbb C}$ and $f$ in $L^p_2$ with $p > 4$. The sections $\alpha$ and $\beta$, being disguised spinors, are as before in $L^p_1$. For an $h$-unitary connection we have , $\dee h(s,t) = h(\dee s, t) + h(s, \dbar t)$ for all sections $s,t \in A^0({\mathcal L})$. Thus $d_L$ is determined by $\dbar$ and $h$, or equivalently, $a^{01}$ and $f$. Expressed in $a^{01}$ and $f$, equation~\eqref{cSW4} becomes \begin{equation}\label{masterf} \laplace f = 2\pi (|\beta|^2_{h_0} - |\alpha|_{h_0}^2)e ^f - 2\sqrt{-1}\Lambda(\dee_0 a^{01} - \dbar_0\bar{a^{01}}) + \mu \end{equation} where $\mu = 2\pi(2\lambda + (\Lambda c_1(F(K)) - 2 \Lambda c_1({\mathcal L},\nabla_0)$ (compare \cite[lemma 4.1]{Bradlow}). If $\beta$ is small in $L^p_1$ hence in $C^0$, we can solve for $f$ in equation~\eqref{masterf} with the solution depending real analytically on $(a^{01},\alpha)$ by the analytical lemma~\ref{fsoln}. Moreover, variation of ~\eqref{masterf} with respect to $f$ when $\beta = 0$ gives \begin{equation}\label{deltamasterf} \delta\hbox{``eqn \eqref{masterf}''} = (\laplace + 2\pi|\alpha|^2e^f)\delta f. \end{equation} Thus, equation~\eqref{masterf} cuts out this solution transversely. More invariantly, if $\beta$ is small, there is a unique metric $h(\dbar,\alpha,\beta) = h_0e^{f(\dbar-\dbar_0,\alpha,\beta)}$ solving the last monopole equation~\eqref{cSW4}. \begin{Lemma} \label{fsoln} Let $X$ be a compact Riemannian manifold, and $\dim(X) < p < \infty$ a Sobolev weight. Then for every real non negative function $0 \le w_0 \in L^p$, with $\int w_0 > 0$ and real function $\mu_0 \in L^p$, with $\int \mu_0 > 0$, there exists a neighborhood $U_{(w_0,\mu_0)} \subset L^p \times L^p$ such that for all $(w,\mu)\in U_{(w_0,\mu_0)}$ the equation \begin{equation}\label{rawfeq} \laplace f = - w e^f + \mu \end{equation} has a unique $L^p_2$ solution depending analytically on $w$ and $\mu$. The solution is smooth if $w$ and $\mu$ are smooth. \end{Lemma} \begin{pf} As in \cite[lemma 4]{Bradlow} make the substitution $f = \~f - g$ where $g$ is the unique solution of $\laplace g = \int \mu - \mu $ to reduce to the case where $\mu$ is constant. Then apply \cite[theorem 10.5(a)]{KazdanWarner} to solve the equation for $w_0,\mu_0$ (note that Kazdan Warners Laplacian is negative definite and that the proof works fine with $w \in L^p$ instead of $C^\infty$). Since at a solution $f_0$ for $(w_0,\mu_0)$ we have $$ \delta \hbox{``eqn \eqref{rawfeq}''} = (\laplace + w_0 e^{f_0})\delta f $$ and $(\laplace + w_0 e^{f_0})$ is invertible, we conclude with the implicit function theorem that there continues to exist a solution for $(w,\mu)$ in a small neighborhood of $(w_0,\mu_0)$, and that this solution depends real analytically on $(w,\mu)$. Regularity follows from standard bootstrapping techniques. Uniqueness follows from the weak maximum principle (\cite[theorem 8.1]{GilbargTrudinger}, c.f. \cite[remark 10.12]{KazdanWarner}). \end{pf} In geometric terms, this has the following consequence. let $\A^{01}$ be (the $L^p_1$-completion) of the space of $\dbar$-operators on ${\mathcal L}$ modeled on $A^{01}(X)$ through $\dbar = \dbar_0 + a^{01}$. The complex gauge group $\G^{\Bbb C}$ acts naturally by conjugation. Let $$ \Pee^{01*} = \A^{01} \times (A^{00}({\mathcal L})\oplus A^{02}({\mathcal L}))^* /\G^{\Bbb C} $$ Clearly there is a projection $\Pee^* \to \Pee^{01*}$ forgetting $h$. What we have done is showing that there is section \begin{align*} \Pee^{01*} & \to \Pee^* \\ (\dbar,\alpha,\beta) &\to (\dbar,\alpha,\beta,h(\dbar,\alpha,\beta)) \end{align*} in a neighborhood of $\beta = 0$, whose image is cut out as a real analytic space by the last monopole equation~\eqref{cSW4}. So far we have not used the other equations. Suppose we are in case~\eqref{case0}, i.e. where a solution corresponds to sections. Then ${\mathcal M}$ is cut out by $\dbar^2 =0$, $\dbar \alpha = 0$, $\beta = 0$ and, by the preceding argument, $h = h(\dbar,\alpha,\beta)$. Thus projection identifies ${\mathcal M}$ with $$ \MBN = \{ (\dbar,\alpha,\beta) \in \Pee^{01}, \ \dbar^2 = 0,\ \dbar\alpha =0, \beta=0\} $$ For the Zariski tangent space it gives \begin{align*} T_{(\nabla,\alpha,0,h)}{\mathcal M} &= T_{(\dbar,\alpha,0)}\MBN \\ &= \rmmath{Ker}\left. \begin{pmatrix} \dbar & \alpha \\ & \dbar \end{pmatrix} \right/ \rmmath{Im} \begin{pmatrix} \alpha \\ -\dbar \end{pmatrix} \\ &= {\Bbb H}^1( 0 \m@p--\rightarrow{} {\mathcal O} \m@p--\rightarrow{\alpha} {\mathcal L} \m@p--\rightarrow{} 0) \\ &= H^0({\mathcal L}|_{Z(\alpha)}). \end{align*} It is easy to check that the linearised versions of equations~\eqref{cSW1}, \eqref{cSW2}, \eqref{cSW3}, and~\eqref{masterf} give the same result (as it should). \ifcomment\bgroup\par\medskip\noindent\small{ Substituting $(\nabla + \eps a,\alpha + \eps \xi,\eps \eta, h e^{\eps f})$ in \eqref{complexSW}, when $(\nabla,\alpha,0,h) \in {\mathcal M}$ yields \begin{align*} &\dbar^*\eta + \dbar \xi + a^{01} \alpha = 0 \\ &\dbar a^{01} = \bar\alpha \eta \\ &\dee a^{10} = - \alpha \bar \eta \\ &\numfrac{\sqrt{-1}}{2\pi}\Lambda (\dbar a^{10} + \dee a^{01}) + \numfrac1{4\pi} \laplace f = -|\alpha|^2_h f - 2 \rmmath{Re} h(\alpha,\xi). \end{align*} As above, the equations give $\eta = 0$ and we can just solve for $f$. We mod out the infinitesimal gauge transformations sending $u \in A^0_{\Bbb C}(X)$ to $(a,\xi,\eta,f) = (-du,u\alpha,0,2\rmmath{Re} u)$ The conclusion is the same. } Case \eqref{case2} is reduced to the previous case by Serre duality. In case \eqref{singcase} the metric $h$ we look for is an (almost) Hermite-Einstein metric. \end{pf} \begin{Corollary} \label{genus} Let $X$ be K\"ahler surface. and $L \equiv w_2(X)$ be a class in $H^2(X,{\Bbb Z})$ with $n(L) \ne 0$. Then $L$ is of type $(1,1)$. Moreover if $p_g >0$, then for all K\"ahler forms $\Phi$ on $X$, the class $L$ satisfies $$ \deg_\Phi(K_X) \ge \deg_\Phi(L) \ge -\deg_\Phi K_X $$ If $p_g = 0$, and $n_-(L) \ne 0$ (resp. $n_+(L) \ne 0$), then $$ \deg_\Phi(L) \ge -\deg_\Phi(K_X)\ \txt{(resp.} \deg_\Phi(L)\le \deg_\Phi(K_X)) $$ \end{Corollary} \begin{pf} First we consider the case $p_g >0$. Under the conditions of the corollary, there is an SC-structure $W$ with $L_W = L$ which admits at least one solution to the monopole equation for {\sl every} admissable pair $(g,\epsilon)$. In particular $W$ admits a solution for every K\"ahler metric and $\epsilon = \lambda \Phi$. Thus $L = L_W$ is of type $(1,1)$. Moreover the necessary condition for the existence of a solution of section {\sl or} cosection type (i.e. equation \ref{case0} {\sl or} \ref{case2} in proposition~\ref{Kahlermonopoles}) gives precisely the required inequality in the limit $\lambda \to 0$. If $p_g =0$, then $L$ is automatically of type $(1,1)$ and say the condition $n_-(L) \ne 0$ means that there is an SC structure $W$ with $L_W = L$ such that for any K\"ahler metric, $W$ admits solutions of section type (i.e. equation \ref{case0}) if $\lambda$ is sufficiently large. This gives a lower bound but no upper bound on $\deg_\Phi(L)$. \end{pf} \begin{Remark}\label{easyp_g=0} If $p_g =0$ and we restrict to perturbation $\epsilon =0$ (or small), then the same argument as in the $p_g >0$ case gives the stronger degree inequality if $L^2 \ge 0$, $L$ is not torsion, since in this case all metrics are admissable and have discriminant of equal sign. In particular on a Del Pezzo surface such classes do not exist. \end{Remark} \begin{Corollary} \label{Kisthere} Let $X$ be a K\"ahler surface with base SC structure $W_0 = \Lambda^{0*} X$. Then $n(W_0) = 1$ if $p_g >0 $ and $n_-(W_0) = 1$ if $p_g =0$, in particular $n(-K_X) \ne 0$ resp. $n_-(-K_X) \ne 0$. Likewise, $n(W_0(K_X) = \pm 1$ if $p_g>0$ and $n_+(W_0(K_X) = \pm 1$, in particular $n(K_X)\ne 0$ resp. $n_+(K_X) \ne 0$. Moreover $W_0$ is the only SC-structure $W$ with $L_W = -K_X$ mod torsion and non trivial multiplicity $n$ respectively $n_-$. In particular if $L \in H^2(X,{\Bbb Z})$, such that $L = -K \in H^2(X,{\Bbb Q})$ and $n(L) \ne 0$ resp. $n_-(L) \ne 0$ then $L = -K \in H^2(X,{\Bbb Z})$. \end{Corollary} \begin{pf} We will prove the statement for $-K_X$. Then we have to consider SC-structures $W = \Lambda^{0*}({\mathcal L})$ with $c_1({\mathcal L})$ torsion. Choose a K\"ahler metric and $\lambda \gg 0$. Then ${\mathcal M}(W) \iso \MBN({\mathcal L})$ the moduli space of line bundles with a section. But ${\mathcal M}^{BN}({\mathcal L})$ is just a reduced point if ${\mathcal L}$ is trivial, and empty if $c_1({\mathcal L})$ is non trivial torsion. Thus $W_0 = \Lambda^{0*} X$ is unique among the SC-structures $W$ with $L_W = -K_X$ mod torsion with $n(W) \ne 0$ (resp. $n_-(W) \ne 0$). In fact its multiplicity is $1$. The case $+K_X$ can be dealt similarly with Serre duality. Its multiplicity is $\pm 1$ because of the unpleasant orientation switches. \end{pf} \begin{Corollary}\label{slickdivisor} Let $D$ be an effective divisor with $D\cdot(D-K) = 0$, $h^0({\mathcal O}(D)) =1$, $h^0({\mathcal O}_D(D)) =0$, and $h^0({\mathcal L}(D)) = 0$ for every line bundle ${\mathcal L} \in \rmmath{Pic}^0(X)$. Then $n(-K_X + 2D) \ne 0$ if $p_g >0$ and $n_-(-K_X + 2D) \ne 0$ if $p_g =0$. Likewise, $n(K_X -2D) \ne 0$ if $p_g >0$ and $n_+(K_X - 2D) \ne 0$ if $p_g = 0$. \end{Corollary} \begin{pf} This corollary is proved just as the previous one, and reduces to it if $D =0$. The conditions of the corollary ensure precisely that $\MBN({\mathcal O}(D))$ consists of one smooth point and that $\vdim(\Lambda^{0*}(D)) = 0$. \end{pf} We are finally in the position to prove the main theorem~\ref{main} and corollary~\ref{poscurv}. Our first task is to define a set $\K$ of basic classes. \begin{Definition} If $b_+ \ge 2$ then the basic classes are defined by $$ \K = \{ K \in H^2(X,{\Bbb Z}) \mid n(K) \ne 0\} $$ If $b_+ = 1$ then $\K = \K_- \cup \K_+$ where \begin{align*} \K_- =\{ K &\in H^2(X,{\Bbb Z}) \mid n_-(K) \ne 0, \txt{and} \exists L \txt{with} n_-(L) \ne 0 \\ &\txt{such that} n_-(L-m(K+L))\ne 0 \txt{for some $m \ge 1$} \}. \end{align*} The set $\K_+$ is defined similarly in terms of $n_+$. Here we are allowed to take $m\ge 1$ rational as long as $m(K+L)$ is two divisible. \end{Definition} These basic classes are rightfully {\sl the} Seiberg-Witten basic classes when $b_+ \ge 2$, but for $b_+ =1$ the definition is geared towards the specific application we have in mind. We will show that $\K$ has all properties~\ref{*}. It is clear that $\K$ is an oriented diffeomorphism invariant, and that the basic classes are characteristic. The pushforward property \ref{*}.\ref{iii} follows immediately from the blow up formula theorem~\ref{blowup4} or~\ref{blowup}. For K\"ahler surfaces the classes are of type $(1,1)$ by corollary~\ref{genus}. The degree property~(\ref{*}.\ref{ii}) (for all surfaces minimal or not) follows also from corollary~\ref{genus}. This is immediate for $p_g >0$. If $p_g = 0$ assume that $K \in \K_+$ say, the case $K \in \K_-$ being essentially the same. Then the corollary gives the three inequalities \begin{align} \deg K &\le \deg K_X, \\ \deg L &\le \deg K_X, \\ -m\deg K &\le \deg K_X + (m-1) \deg L \le m\deg K_X. \end{align} If $p_g >0$ then $K_X \in \K$ by corollary~\ref{Kisthere}. Thus it remains to check that $K_X \in \K$ if $p_g =0$. In fact we will check that $-K_X \in \K$. We have already seen in corollary~\ref{Kisthere} that $n_-(-K_X) \ne 0$. Either directly from corollary~\ref{slickdivisor}, or using the invariance under the reflection in the exceptional curves $E_1, \ldots, E_n$ we see that $n_-(-K_X + 2\sum E_i) \ne 0$. Then denoting $$ {\mathcal L}_m = m\Kmin + \sum E_i, $$ we have to check that $n_-(-K_X + 2{\mathcal L}_m) \ne 0$. We will distinguish four cases. If $\kappa(X) = 0$, then $\Kmin$ is torsion and we can take $m=\rmmath{ord}(\Kmin)$, since $n_-(-K_X + 2\sum E_i) \ne 0$. If $\kappa(X) = 1$, then $\Xmin$ has a unique elliptic fibration $\Xmin \m@p--\rightarrow{\pi} C$. By the canonical bundle formula, $\Kmin = \pi^*{\mathcal L}_C(\pi^*K_C + \sum (p_i - 1)F_i)$, where ${\mathcal L}_C$ is a line bundle on $C$ of degree $\chi$. Since $p_g = 0$ and $\chi \ge 0$, we have $0 \le g\le q\le 1$, and we distinguish further between $g=0$ and $g =1$. If $g= 0$, then $c_1(\pi^*{\mathcal L}_C(K_C)) =(\chi-2) F$, where $F$ is a general fibre, and there are at least $3-\chi$ multiple fibers because $\Kmin >0$. Now the class $\Kmin + \sum_{i=1}^{2-\chi} F_i = \sum_{j=3-\chi}^n (p_j -1) F_j$ is of the form $m\Kmin$ with rational $m >1$. Again by corollary~\ref{slickdivisor}, we have $$ n_-(-K_X + 2{\mathcal L}_m) = n_-(-K_X + 2(\sum_{j=3-\chi}^n (p_j-1)F_j + \sum E_i)) \ne 0 $$ If $g=1$, then $\chi = 0$, and $K_C =0$. In this case we can take $m = 1$ since $c_1({\mathcal L}_C) =0\in H^2(X,{\Bbb Z})$ and by corollary~\ref{slickdivisor} $$ n_-(-K_X +2{\mathcal L}_1) = n_-(-K_X + 2(\sum (p_i-1)F_i + \sum E_i)) \ne 0. $$ The most instructive case is when $X$ is of general type. Then the irregularity $q=0$ since $p_g = 0$ and $\chi({\mathcal O}_X) >0$. Take $m=2$, then $\MBN({\mathcal L}_2) = |2\Kmin + \sum E_i|$. By formula~\eqref{gtpluri} (or directly by Ramanujan vanishing) $$ \dim_{\Bbb C}\MBN({\mathcal L}_2) = P_2 - 1 = \Kmin^2 = {\numfrac12} \vdim_{\Bbb R}(W_2). $$ Thus the moduli space is again smooth of the proper dimension and we conclude that $n_-(-K_X + 2{\mathcal L}_2) \ne 0$. In fact $n_-(\Lambda^{0*}({\mathcal L}_2)) = t^{\Kmin^2}$ since the ${\mathcal O}(1)$ on $\Pee^*$ corresponds to the ${\mathcal O}(1)$ on $\MBN$. This is because both measure the weight of the action of the constant gauge transformations on the spinors respectively sections. It now follows from lemma~\ref{inequality} that if $\kappa(X) \ge 0$, all SW-structure have a moduli space of virtual dimension $d=0$, and up to torsion, the basic classes are of type \begin{equation} K =\lambda \Kmin + \sum \pm E_i \bmod \hbox{Torsion}, \qquad |\lambda| \le 1. \end{equation} Moreover by proposition~\ref{Kcharacterisation}, $\Kmin$ is invariant up to sign and torsion and every $(-1)$-sphere is represented by a $(-1)$-curve up to sign and torsion. We first get rid of torsion in the $(-1)$-curve conjecture i.e. theorem {}~\ref{main} part~\ref{mainb}. Let $e$ be a $(-1)$-sphere, giving a connected sum decomposition $X = X'\# \Pbar^2$. As we have used before, there is a diffeomorphism $R_e =\rmmath{id} \# {\Bbb C}$-conjugation representing the reflection in $e$. I claim that for any SC-structure $W$ on a 4-manifold $$ R_e^*(W) = W \tensor{\mathcal O}((c_1(L_W),e)e), $$ where ${\mathcal O}(e)$ is the line bundle corresponding to the Poincar\'e dual of $e$. In fact if we write $R_e^* W = W \tensor {\mathcal L}$, then ${\mathcal L} = \mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits_C(W, R_e^*W)$ (c.f. the proof of \ref{SCrepr}). Now we can just identify $W$ and $R_e^* W$ on $X'$, i.e. ${\mathcal L}$ is trivialised on $X'$. Thus $$ c_1({\mathcal L}) \in \rmmath{Im} H^2(X, X- X', {\Bbb Z}) \iso H^2(\Pbar^2) \subset H^2(X,{\Bbb Z}). $$ Write ${\mathcal L} = {\mathcal O}(a e)$ for some integer $a$. Since $$ L_W + 2 a e = L_{R_e^* W} = R_e^*L_W = L + 2(e,L_W) e $$ the claim is proved. Going back to the K\"ahler case, we can assume that $e$ is homologous to a $(-1)$-curve $E$ up to torsion. Consider $W = R_e^*R_E^*(\Lambda^{0*} X) = \Lambda^{0*}(E-e)$. By oriented diffeomorphism invariance $n_{(-)}(W) \ne 0$ (in case $p_g =0$ we have tacitly used the fact that $R_e^*R_E^*$ induces the identity on rational cohomology so in particular does not change the orientation of $H^+$). Moreover $c_1(L_W) = -K_X$ up to torsion. By corollary~\ref{Kisthere}, we conclude that $W = \Lambda^{0*} X$, so $e = E \in H^2(X,{\Bbb Z})$. Finally for the invariance $\pm \Kmin$, consider any basic class of the form $L = \pm \Kmin + \sum \pm E_i$ up to torsion. After reflections in the $(-1)$-curves, we get a class $L'$ equal to $\pm K_X$ up to torsion. By corollary~\ref{Kisthere} $L' = \pm K_X \in H^2(X,{\Bbb Z})$. Now for any basis $E'_1, \ldots E'_n$ of the lattice in $H^2(X,{\Bbb Z})$ spanned by the $(-1)$-spheres (e.g. the $(-1)$-curves) we have the identity $$ \pm \Kmin = L' + \sum (E'_i,L') E'_i = L + \sum ( E'_i, L) E'_i \in H^2(X,{\Bbb Z}). $$ This finally proves theorem~\ref{main}. \begin{Remark} It is easy to give a definition of basic classes for $b_+ = 1$ that satisfies all properties ~\ref{*} except the invariance under blow down (i.e. property~\ref{*}.\ref{iii}). A class $K$ is then basic if there exists a metric $g$ such for all $\delta >0$ there exists an admissable pair $(g,\epsilon)$ with $\|\epsilon^{+,\harm}\| < \delta$ such that $n(g,\epsilon,K) \ne 0$. The degree inequality for minimal surfaces then follows from remark \ref{specialchamber}. But alas, if $K^2 <0$ one can not avoid the possibility that a chamber on the blow up realisable with small $\epsilon$ can only be realised for large $\epsilon$ on the blow down. In my original treatment I used this definition. I am grateful to Robert Friedman whose insistent questions about my definitions made me realise this mistake. \end{Remark} \begin{Remark}\label{whycastelnuovo} An easy application of the techniques of the next section gives the following. If ${\mathcal L}$ is a holomorphic line bundle on a surface with $p_g = q = 0$ with $h^0({\mathcal L}) \ge \chi({\mathcal L})\ge 1$, then $n_-(\Lambda^{0*}({\mathcal L})) = t^{{\mathcal L}({\mathcal L}-K_X) \over 2}$. If $p_g= q = 0$ and $\kappa(X) \ge 0$ we can apply this to ${\mathcal L}_2= 2\Kmin + \sum E_i$. Then by the Castelnuovo criterion and the above we conclude $n_-(-K_X + 2L_2) \ne 0$. This gives an alternative way to prove that $-K_X \in \K$ in this case. Conversely the degree inequality~\ref{*}.\ref{ii} cannot hold true for rational and ruled surfaces for K\"ahler forms $\Phi$ such that $\deg_\Phi(K_X) < 0$. Since in deriving the degree inequality we did not use that $\kappa(X) \ge 0$, we conclude that for $\kappa(X) = -\infty$ the set of the above defined basic classes $\K= \emptyset$. In particular we see that the following proposition is a rather direct analog of to the classical Castelnuovo criterion. \end{Remark} \begin{Proposition}\label{Castelnuovo} A K\"ahler surface is rational if and only if $b_1= 0$, and $\K = \emptyset$. \end{Proposition} \begin{Remark} After reading \cite{FM:SW} I realised the following. The blow up formula~\ref{blowup4} can be generalised to connected sum decompositions $X = X'\# N$ with $N$ negative definite and $H_1(N,{\Bbb Z}) = 0$. The latter condition is automatic for K\"ahler surfaces of non negative Kodaira dimension by a beautiful observation of Kotschick (an unramified covering $\~N \to N$ of degree $d$ gives an unramified covering $\~X = d X' \# \~N \to X'\# N$ which is an algebraic surface of non negative Kodaira dimension with a connected sum decomposition with a factor with $b_+ >0$). Such smooth negative definite manifolds $N$ have $H_2(N) = \mathop\oplus_{i=1}^n {\Bbb Z} n_i$. SC structures $W_N$ on $N$ are determined by $L_N = \sum (2a_i + 1) n_i$. Thus the reflections $R_{n_i}$ in $n_i^{\perp}$, act on the SC structures on $N$. SC -structures on $X'\# N$ are of the form $W= W_{X'} \# W_N$. Now the blow up formula is as if $N = n \Pbar^2$: $W = W_{X'} \# W_N$ is an SW-structure on $X'\# N$ if and only if $W_{X'}$ is a SW-structure on $X'$ and $d(W) \ge 0$. In particular the Seiberg Witten structures are invariant under the operation $R_{n_i}: W_{X'}\# W_N \to W_{X'} \# R_{n_i} W_N$, and $\mathop{{\mathcal H}\mkern-3mu{\mathit om}}\nolimits_C(W, R_{n_i}W)$ has a trivialisation over $X'$. With these remarks the arguments for $(-1)$-spheres carry over directly to prove that for K\"ahler surfaces $X$ with $\kappa(X)\ge 0$, with a connected sum decomposition $X = X'\# N$, $H_2(N) \subset H_2(X)$ is spanned by $(-1)$-curves. \end{Remark} Stefan Bauer showed me how to use the Seiberg Witten multiplicities and the basic classes to determine the multiplicities of the elliptic surface. If the surface does not have finite cyclic fundamental group, the multiplicities can be read off from the topology. Thus we consider a minimal elliptic surface $X_{pq}$ fibred over ${\Bbb P}^1$ with 2 multiple fibers of multiplicity $p$ and $q$ We will assume that $p\le q$. \begin{Corollary} (Bauer) The multiplicities $p$ and $q$ are determined by the underlying oriented differentiable manifold, unless $p_g = 0$, $p=1$ and $q$ arbitrary. The surfaces $X_{1q}$ are all rational and diffeomorphic. \end{Corollary} \begin{pf} If the canonical class $K_X$ is not torsion , we can write $K_X$ in terms of the primitive vector $\kappa$ in the ray spanned by $K_X$, normalised so that $\kappa \Phi >0$ $$ K_X = (p_g - 1)F + (p-1)F_p + (q-1) F_q = {(p_g + 1)pq - p -q \over \gcd(p,q)} \kappa \in H^2(X,{\Bbb Z})/\mathord{\hbox{Torsion}}. $$ Let $d(p,q) = \((p_g+1)pq-p-q\)/\gcd(p,q)$ be the oriented divisibility of $K_X$. If $K_X$ is torsion we simply set $d(p,q) = 0$. The divisibility $d(p,q)<0$ if and only if $p_g=0$, $p=1$ and $q$ is arbitrary. But this implies that $K_X$ is rational. We have already seen that we can recognise rationality as Kodaira dimension $-\infty$ and $b_1 =0$ (corollary~\ref{Kodaira} or proposition~\ref{Castelnuovo}). Thus we can assume that $X_{pq}$ has non negative Kodaira dimension. Then $\pm K_X \in \K$ are the basic classes with the highest divisibility (or torsion) and the oriented divisibility $d(p,q) \ge 0$ is just the unoriented divisibility of $\pm K_X$. The number $\gcd(p,q)$ is also determined by the oriented manifold, being the order of the fundamental group. Choose one of these classes, say $-K_X$. First consider the case $p_g >0$. Suppose that $K= -K_X + 2 F_q \le 0$, (i.e. on the same side of $0$ as $-K_X$), then it is the basic class with second largest divisibility since $F_q$ is the smallest effective vertical divisor, and $n(-K_X + 2F_q)) \ne 0$ by lemma~\ref{slickdivisor} above. Thus if there exist basic classes other then $\pm K_X$, we can reconstruct $p$ from $(2p/ \gcd(p,q))\kappa = K - (-K_X)$. Since $d(p,q)$, $p_g$ and $\gcd(p,q)$ are known, this determines $q$ as well. Obviously if we have chosen $+K_X$ the same arguments works with $K = K_X - 2F_q$, there is nothing that prefers $K_X$ over $-K_X$. In the case $p_g =0$ we make a small modification. We choose an orientation of $H^+$, which for a moment we assume is the standard one. Consider the classes $K \in H^2(X,{\Bbb Z})$ mod torsion in the half ray spanned by $0$ and $ -K_X$ with unoriented divisibility at most $d(p,q)$ (i.e. in between $0$ and $-K_X$) such that $n_-(K) \ne 0$. Note that $-K_X$ is just the basic class with largest divisibility in $\K_-$. Then if $K = -K_X + 2 F_q \le 0$ we can use exactly the same argument as in the case $p_g >0$. If we choose a different orientation of $H^+$, we replace $-K_X$ by $+K_X$ but just as above the conclusion is the same. If $\K = \pm K_X$ or for $p_g =0$ if $\{K \in [-K_X, 0] \mid n_-(K) \ne 0\} = -K_X $ then $d(p,q)\gcd(p,q)< 2p$. The few possibilities are listed in the following table $$ \begin{array}{|l|c|c|c|l|} \hline \strut&(p,q) &\gcd(p,q) &d(p,q) &\text{Type} \\ \hline p_g=0 &(2,2) &2 &0 &\text{Enriques} \\ &(2,3) &1 &1 & \\ &(2,4) &2 &1 & \\ &(2,5) &1 &3 & \\ &(3,3) &3 &1 & \\ &(3,4) &1 &5 & \\ \hline p_g=1 &(1,1) &1 &0 &\text{K3} \\ &(1,2) &1 &1 & \\ \hline \end{array} $$ Clearly, in this case the pair $(p,q)$ is determined by the oriented differentiable manifold as well. \end{pf} To prove that no surface with $\kappa \ge 0$ admits a metric with positive scalar curvature (corollary~\ref{poscurv}), first consider the case $p_g >0$. Then the statement is clear, and one of Witten's basic observations. By proposition~\ref{Bochner}, for 4-manifolds with positive scalar curvature $n(K) = 0$ for all $K \in H^2(X,{\Bbb Z})$, since for our metric with positive scalar curvature $g$ and small perturbations $\epsilon$, we have ${\mathcal M}(W,g,\epsilon) = \emptyset$ for all SC-structures $W$. On the other hand we just showed that $n(-K_X) \ne 0$ using a K\"ahler metric. The same argument works if $p_g = 0$ and $K_X^2 \ge 0$: $n(-K_X,g,\epsilon)$ is independent of the metric $g$ and of $\epsilon$ as long as $\epsilon$ is small, with the exception of the case $-K_X$ torsion in which case we have to choose $\epsilon$ in the forward light cone. But we can do better. For the general case $p_g =0$, we choose a perturbation $\epsilon = \lambda \Phi$ with $0<\lambda \ll 1$ say. Now suppose that the metric with positive scalar curvature $g$ has period $\omega_g = \omega_{\min} + \sum \eta_i E_i$ where $\omega_{\min} $ is the projection to the cohomology of minimal model. Then since $\omega_g$ is in the interior of the forward light cone, and $\Kmin$ is in the closure of the forward light cone, $\omega\cdot \Kmin = \omega_{\min}\cdot \Kmin\ge 0$ with equality iff $\Kmin$ is torsion. Then for {\sl some} choice of signs in $-\Kmin - \sum \pm E_i$ we have $$ \omega_g\cdot (-\Kmin - \sum \pm E_i)\le 0 < \lambda \int\omega_g \Phi $$ Thus for {\sl some} choice of signs we compute $n_-$ (rather than $n_+$) with our metric of positive scalar curvature and small perturbation. Hence $n_-(-\Kmin-\sum \pm E_i) = 0$. On the other hand $n_-(-\Kmin -\sum \pm E_i) = n_-(-K_X) \ne 0$, a contradiction just like before. \section{Some computations of Seiberg-Witten multiplicities}% \label{computations} In this section we will go beyond determining potential basic classes and compute the Seiberg Witten multiplicity of elliptic surfaces. We also prove an algebraic version of the blow up formula. It is here that our excess intersection formulas pay off. We first show how to go over to a fully complex point of view. Then we use the special geometry of elliptic surfaces to compute the multiplicities and finally we prove a blow up formula. {}From now on we identify an SC-structure with the corresponding twisting line bundle ${\mathcal L}$. We will consider the solutions of the monopole equations of section type, i.e. corresponding to equation~\eqref{case0}. We have already seen that the variation of the last monopole equation~\eqref{cSW4} with respect to the hermitian metric is $h$ is given by $(\laplace + |\alpha|_h^2)h^{-1}\delta h$ (c.f. equation~\ref{masterf}). Therefore the solutions to the fourth monopole equation ~\eqref{cSW4} is a smooth submanifold of $\Pee^*$ in a neighborhood of the moduli space ${\mathcal M}({\mathcal L})$. In the proof of proposition~\ref{Kahlermonopoles} we have seen that we can identify this submanifold with the ``vortex locus'' $\{h = h(\dbar,\alpha,\beta)\}$ i.e. the image of the section $\Pee^{01*} \to \Pee^*$. The vortex locus is well defined in a neighborhood of the moduli space ${\mathcal M}({\mathcal L})$ only, but this will not affect our arguments, as the construction of the localised Euler class in section \ref{top} depends only on a neighborhood of ${\mathcal M}({\mathcal L})$. Since the vortex locus is given by a function, we can identify it with its domain $\Pee^{01*}$ which carries a natural complex structure. By property \ref{stability} of proposition~\ref{locEuler} we are allowed to compute the localised Euler class $\MM({\mathcal L})$ of the moduli space by considering ${\mathcal M}({\mathcal L})$ as a zero set of a section $S$ over the vortex locus cut out by the remaining equations, which define the same ideal as $$ \dbar^2 = 0, \qquad \dbar \alpha =0, \quad \beta = 0 $$ i.e. complex equations ! Moreover the deformation complex of these equations on $\Pee^{01*}$ in a point $(\dbar,\alpha,0)$ is $$ A^{00}(X) \m@p--\rightarrow{} A^{01}(X) \oplus A^{00}({\mathcal L}) \oplus A^{02}({\mathcal L}) \m@p--\rightarrow{} A^{02}(X) \oplus A^{01}({\mathcal L}) $$ where the map is complex linear. We trivialise the determinant of the index using the complex structure. This has brought us safely in complex waters, and allows us to use proposition ~\ref{locChern} and in particular formula~\ref{magic}. {}From now on we identify ${\mathcal M}({\mathcal L})$ with $\MBN({\mathcal L})$. Define the vector bundles $$ \sfA^{pq}({\mathcal L}) =\(\A^{01} \times (A^{00}({\mathcal L} \oplus A^{02}({\mathcal L}))^*\)\times_{\G^{\Bbb C}} A^{pq}({\mathcal L}) $$ over $\Pee^{01*}$. Then $\MBN$ is given by a section $s$ in $E = A^{02}(X) \oplus \sfA^{01}({\mathcal L})$, and the tangent space is given by $$ T\Pee^{01*} \iso \(A^{01}(X) \oplus \sfA^{00}({\mathcal L}) \oplus \sfA^{02}({\mathcal L})\)/ A^{00}(X). $$ The deformation complex can be considered as a map $T\Pee^* \to E$ and is exactly what we called $Ds$ in section~\ref{top}. To identify the index $\Ind Ds$ we first make a compact perturbation, keeping only the differential operator part of the deformation complex. Then it splits naturally in the $\dbar$ complex on $X$ with index ${\Bbb C}^{\chi({\mathcal O}_X)}$ and the index of complex $$ 0 \m@p--\rightarrow{} \sfA^{00}({\mathcal L}) \m@p--\rightarrow{\dbar} \sfA^{01}({\mathcal L}) \m@p--\rightarrow{\dbar} \sfA^{02}({\mathcal L}) \to 0 $$ where $\dbar$ is the universal $\dbar$ operator descended to $\Pee^{01*}$. To rewrite this index in holomorphic terms, consider the universal divisor $$ \Delta = \{ (\dbar,\alpha,x) \mid \alpha(x) = 0\} $$ on the pull back of $X \times \MBN$. Now if $\Omega^{pq}$ is the sheaf of $C^{\infty}$ $(p,q)$-forms on $X$ considered as an ${\mathcal O}(X)$ module, then I claim that on the pull back of $X \times \MBN$ to $\A^{01} \times (A^{00}({\mathcal L} \oplus A^{02}({\mathcal L}))^*$, there is a $\G^{\Bbb C}$ equivariant exact sequence $$ 0 \m@p--\rightarrow{} {\mathcal O}(\Delta) \m@p--\rightarrow{\dbar} p_1^*\Omega^{00}({\mathcal L}) \m@p--\rightarrow{\dbar} p_1^*\Omega^{01}({\mathcal L}) \m@p--\rightarrow{\dbar} p_1^*\Omega^{02}({\mathcal L}) \to 0. $$ In fact this only says that $(\dbar,\alpha, x) \to \alpha(x)$ is a $\G^{\Bbb C}$ invariant section vanishing along $\Delta$ with multiplicity $1$ lying in the kernel of $\dbar$, which is obvious. Now descend this whole complex to $X \times \MBN$ and take push forward to $\MBN$. Then we get an exact sequence of complexes $$ 0 \m@p--\rightarrow{} Rp_*{\mathcal O}(\Delta) \m@p--\rightarrow{\dbar} \sfA^{00}({\mathcal L}) \m@p--\rightarrow{\dbar} A^{01}({\mathcal L}) \m@p--\rightarrow{\dbar} \sfA^{02}({\mathcal L}) \m@p--\rightarrow{} 0 $$ where we are considering $Rp_*{\mathcal O}(\Delta)$) as a complex with zero boundary operator and $\sfA^{pq}({\mathcal L})$ as a complex concentrated in degree $0$. Thus for the index we find \begin{equation}\label{index} \Ind(Ds) = \Ind\(Rp_* {\mathcal O}(\Delta)\) + {\Bbb C}^{\chi} \end{equation} A more precise description of $\MBN$ depends on the surface. Here we will do the case of elliptic surfaces. The author has succeeded in treating ruled surfaces in a similar way. \begin{Proposition}\label{multiplic} Let $X \m@p--\rightarrow{\pi} C$ be a K\"ahlerian elliptic surface over a curve $C$ of genus $g$, with multiple fibers $F_1, \ldots F_r$ of multiplicity $p_1, \ldots p_r$ of holomorphic Euler characteristic $\chi$. Consider the line bundle ${\mathcal L} = {\mathcal O}(\pi^*D + \sum a_i F_i)$ where $D$ is a divisor on $C$ of degree $d$, and $0 \le a_i < p_i$. Then the Seiberg Witten multiplicity is zero if $d < 0$, and if $d\ge 0$ it is given by $$ n_{(-)}(\Lambda^{0*}({\mathcal L})) = \begin{cases} (-1)^d{\chi + 2g-2\choose d} & \txt{if} \chi + g -2 \ge 0 \\ \sum (-1)^j { 1-g-\chi + d-j \choose d-j}{g \choose j} & \txt{if} \chi + g - 2 < 0 \end{cases} $$ \end{Proposition} Note that if the topological Euler characteristic $e >0$ (or equivalently $\chi >0$) then $g =q = {\numfrac12} b_1(X)$ \cite[corollary II.2.4]{FM}, so in this case $\chi + g-2 = p_g - 1$. Note that the second formula is just 1 if $p_g =q =0$ (i.e. $e>0$). This illustrates remark~\ref{whycastelnuovo}. If $p_g >0$ and $q = g = 0$, so in particular $e = 12 \chi > 0$, Witten proves this formula by choosing a general $\omega \in H^0(K_X)$ and using the perturbation $\epsilon = \omega + \bar\omega$. He then argues that the multiplicity $n({\mathcal L})$ is the number of ways we can decompose a fixed canonical divisor $K_0$ as $K_0 = D_+ + D_-$ with $D_+\in |({\mathcal L},\dbar_0)|$, and $D_- \in |K\tensor({\mathcal L},\dbar_0)^{\scriptscriptstyle\vee}|$, where $\dbar_0$ is the unique holomorphic structure that ${\mathcal L}$ admits \cite[eq. (4.23) e.v.]{Witten}. To be honest, this is what I read out of it. Note for example that his sheaf $R$ is just ${\mathcal L}|_Z(\alpha)$, and that $$ h^0(R) = h^0({\mathcal L}|_{Z(\alpha))} = \dim T_{(\dbar,\alpha)} = d $$ (the last equality we will see in a minute). Actually I think that the computations below are the mathematical version of (I paraphrase) ``integrating over the bosonic and fermionic collective coordinates in the path integral'' and ``computing the Euler class of the bundle of the cokernel of the operator describing the linearised monopole equations over the moduli space (the bundle of antighost zero modes)'' \cite[above (4.11)]{Witten}. In fact with hindsight, the latter seems a dual description of the localised Euler class in the case that the cokernel has constant rank. \begin{pf} We choose a K\"ahler metric and $\lambda$ such that $\deg_\Phi({\mathcal L}^{\tensor 2}(-K)) < \lambda \Vol(X)$. This means that if ${\mathcal L}$ has non zero multiplicity, it must carry a holomorphic structure with a section. In case $p_g =0$ it also means we are looking at $n_-$. But $({\mathcal L},\dbar)$ has a section if and only if $D$ is an effective divisor on $C$. In fact a family of vertical line bundles with a section gives a family of effective divisors on $C$ by pushforward of the line bundle, and conversely a family of effective divisors on $C$ gives a family of vertical line bundles with a section by pull back and multiplication with a fixed section in ${\mathcal O}(B) = {\mathcal O}(\sum a_i F_i)$ ($B$ for base locus). Thus there is a natural isomorphism $$ \MBN \iso \MBN_C = C^d $$ where $C^d$ is the $d^{\text{th}}$ symmetric power of $C$. The functorial isomorphism comes with an isomorphism ${\mathcal O}(\Delta_X) = {\mathcal O}(\pi^*\Delta_C + B)$. Next we use Grothendieck Riemann Roch (an alias of the family index theorem). Let $q: C \times C^d \to C^d$ be the projection map. Then the projection $p: X \times \MBN \to \MBN$ can be factored as $p = q\circ \pi\times \rmmath{id}$. Thus writing $\pi\times \rmmath{id}$ as $\pi$, \begin{align*} \rmmath{ch}(Rp_*{\mathcal O}(\Delta)) &= \rmmath{ch}\(Rq_* \({\mathcal O}(\Delta_C) \tensor R\pi_*{\mathcal O}(B)\)\) \\ &= q_*\(ch({\mathcal O}(\Delta_C) \rmmath{ch} R\pi_*{\mathcal O}(B) \rmmath{td}(C)\) \\ &= q_*\(ch({\mathcal O}(\Delta_C)) \pi_*\(e^B(1 - K/2 + \chi({\mathcal O}_X)(pt\times C^d)\)\) \\ &= \chi({\mathcal O}_X) q_*\(ch({\mathcal O}(\Delta_C))(pt\times C^d)\) \\ &= \rmmath{ch}({\mathcal O}(1)^\chi), \end{align*} where we have abbreviated the holomorphic Euler characteristic by $\chi$. If we denote by $x$ the chern class of ${\mathcal O}(1)$, then our computation shows that $$ c_t(\Ind Ds) = (1 + tx)^\chi, $$ at least over the rationals. The chern classes of the tangent bundle of $C^d$ are computed in \cite[eq. VII.5.4]{ACGH}. Denoting the pullback of the $\theta$ divisor on $\rmmath{Pic}^d$ to $C^d$ by $\theta$ the result is $$ c_t(T_{C^d}) = (1+ tx)^{d+1-g} e^{-t\theta / (1+ tx)} $$ Combining these two expressions, our multiplicity drops out \begin{align*} n(\Lambda^{0*}({\mathcal L})) &= c(\Ind Ds)^{-1} c(T_C^d) \cap [C^d] \\ &= [(1 + tx) ^{d +1-g - \chi} e^{-t\theta /1+tx}]_{t^d} \end{align*} With the following identity of formal power series \cite[eq. VIII.3.1]{ACGH} $$ [(1 + xt)^a f( - t/(1+ xt))]_{t^b} = [ (1 - xt)^{b - a - 1} f(-t)]_{t^b}. $$ the expression becomes $$ n(\Lambda^{0*}({\mathcal L})) = [(1-tx)^{\chi + g - 2} e^{-t\theta}]_{t^d} = \begin{cases} (-1)^d\sum_{j=0}^d {\chi + g - 2 \choose d-j} {\theta^j x^{d-j} \over j!} & \txt{if} \chi + g- 2 \ge 0 \\ \sum_{j=0}^d (-1)^j {1-g-\chi + d-j \over d-j}{\theta^j x^{d-j} \over j!} & \txt{if} \chi + g -2 < 0 \end{cases} \hskip 0pt minus 1fil $$ Now $\theta^jx^{d-j}\cap [C^d] = j! {g \choose j}$ \cite[below eq. VIII.3.3]{ACGH}. The elementary identity $\sum_j {a \choose j}{b \choose c-j} = {a + b \choose c}$ then gives the answer as stated. \end{pf} As a second application of the methods developed we give a complex analytic version of the blow up formula. \begin{Proposition}\label{blowup} Let $(X,\Phi)$ be a K\"ahler surface, and ${\mathcal L}$ a line bundle on $X$. Suppose that $\deg_\Phi({\mathcal L}^{\tensor 2}(-K) < \lambda \Vol(X)$. Let $\sigma:\~X \to X$ be the blow up of $X$ in a point, with K\"ahler form $\~\Phi$, and let $\~{\mathcal L} = {\mathcal L}(a E)$ be a line bundle on $\~X$ with $a\ge 0$. Suppose that the cohomology class of $\~\Phi$ is close to $\Phi$. Then there is a natural identification ${\mathcal M}(\~{\mathcal L}) = {\mathcal M}({\mathcal L})$, and $$ \MM(\~{\mathcal L}) = [(1+x)^{a(a-1)/2}\widehat \MM({\mathcal L})]_{\dim_{\Bbb R}= {\mathcal L}\cdot({\mathcal L}-K)- a(a-1)}. $$ Here $\widehat\MM$ is the class defined in remark~\ref{Zhat}, and $x$ the class of the natural bundle ${\mathcal O}(1)$ over ${\mathcal M}$. In particular if $a=0,1$ then $n(\~{\mathcal L}) = n({\mathcal L})$. \end{Proposition} Of course this proposition determines the multiplicity $$ n_{(-)}(\Lambda^{0*}({\mathcal L}(aE))) = n_{(-)}(\Lambda^{0*}({\mathcal L}(-aE)))). $$ Since quite in general $n_+(\Lambda^{0*}({\mathcal L})) = \pm n_-(\Lambda^{0*}(K \tensor {\mathcal L}^{{\scriptscriptstyle\vee}})$ it determines the corresponding relation for $n_+$ up to sign which is really all we need here. \begin{pf} The conditions on the degree imply that a solution of the monopole equations correspond to a holomorphic structure on ${\mathcal L}$ with a section. Since $\~\Phi$ is close to $\Phi$ we have (by definition of close) $\deg_{\~\Phi}(\~L) < \lambda\Vol(\~X)$, hence solutions on the blowup also correspond to holomorphic structures on $\~{\mathcal L}$ with a section. Now $aE$ is contained in the base locus of the sections. Thus similar to what we did for elliptic surfaces, we get an identification of ${\mathcal M}({\mathcal L})$ with ${\mathcal M}(\~{\mathcal L})$ by multiplication of the section with a section in ${\mathcal O}(aE)$, and the universal divisor on $\~X\times {\mathcal M}(\~{\mathcal L})$ is $\~\Delta = \Delta + aE$. Again, identify the chern class of the index of the deformation complex with formula~\eqref{index}. Let $\~p$ be the projection $\~X \times {\mathcal M}({\mathcal L}) \to {\mathcal M}({\mathcal L})$, and $p$ the projection $X\times {\mathcal M}({\mathcal L}) \to {\mathcal M}({\mathcal L})$. Then the total chern class of the index is $$ c(R\~p_*(\~\Delta)) = c\(Rp_*\({\mathcal O}(\Delta)\tensor R\sigma_*{\mathcal O}(aE)\) \). $$ By induction on $a$, one shows that $$ R\sigma_* {\mathcal O}(aE) = {\mathcal O} -{\mathcal O}_{pt}^{a(a-1)/2}. $$ Since ${\mathcal O}(\Delta|_{pt\times {\mathcal M}({\mathcal L})}) = {\mathcal O}(1)$ it gives $$ c(R\~p_*(\~\Delta)) = c(Rp_*{\mathcal O}(\Delta))/c({\mathcal O}(1))^{a(a-1)/2}. $$ Formula~\eqref{magic} gives us $$ \MM(\~L) = [(1+x)^{a(a-1)/2} \(c(Rp_*(\Delta))^{-1}c_*({\mathcal M}({\mathcal L})\)]_{d(\~L)} $$ Since the real virtual dimension of ${\mathcal M}(\~{\mathcal L})$ is $d(\~{\mathcal L}) = {\mathcal L}\cdot ({\mathcal L}-K) -a(a-1)$ and the term in brackets is exactly $\widehat\MM({\mathcal L})$, we have proved the formula. \end{pf}
1995-03-28T07:20:32
9503
alg-geom/9503018
en
https://arxiv.org/abs/alg-geom/9503018
[ "alg-geom", "math.AG" ]
alg-geom/9503018
Bruce Hunt
Bruce Hunt
A gem of the modular universe
LaTeX 2.09
null
null
null
null
We introduce one of the most beautiful algebraic varieties known, a quintic hypersurface in projective five-space, which is invariant under the action of the Weyl group of $E_6$. This variety is intricately related with many other moduli problems, some of which are: marked hyperelliptic curves of genus two, Picard curves of genus four with a $\sqrt{-3}$-level structure, six points on the projective line, abelian surfaces with (1,3) polarisations, quartic surfaces invariant under the action of the Heisenberg group in projective three-space, K3-surfaces which are double covers of the projective plane branched along six lines, and last but not least, cubic surfaces in projective three-space. These relationships are developed in some detail, with particular care on the birational aspects.
[ { "version": "v1", "created": "Mon, 27 Mar 1995 06:45:27 GMT" } ]
2008-02-03T00:00:00
[ [ "Hunt", "Bruce", "" ] ]
alg-geom
\part{Projective embeddings of modular varieties}\label{chapter10} \section{The tetrahedron in ${\Bbb P}^3$} \subsection{Arrangements defined by Weyl groups} Let $\Phi(G,T)\subset} \def\nni{\supset} \def\und{\underline \tt^*$ be a root system of a simple group $G$ (over $\komp$). Using notations as in Bourbaki we have the roots (for those systems which will be of interest to us in the sequel) \begin{equation}\label{e108.1}\begin{minipage}{14cm}\begin{itemize} \item[$\bf A_n$] $\pm (\ge_i-\ge_j),\ 1\leq i< j \leq n+1$; \item[$\bf B_n$] $\pm (\ge_i\pm \ge_j),\ \pm\ge_i, 1\leq i < j \leq n$; \item[$\bf C_n$] $\pm (\ge_i\pm \ge_j),\ \pm2\ge_i, 1\leq i < j \leq n$; \item[$\bf D_n$] $\pm (\ge_i\pm\ge_j),\ 1\leq i < j \leq n$; \item[$\bf F_4$] $\pm (\ge_i\pm \ge_j),\ \pm\ge_k,\ \pm{1\over 2}(\ge_1\pm\ge_2\pm\ge_3\pm\ge_4), 1\leq i < j \leq 4,\ k=1,\ldots,4$; \item[$\bf E_6$] $\pm (\ge_i \pm \ge_j), 1\leq i<j \leq 5,\ \pm{1\over 2}(\ge_1\pm\ge_2\pm\ge_3\pm\ge_4\pm\ge_5-\ge_6-\ge_7+\ge_8)$, with an even number of ``$-$'' signs in the parenthesis; \end{itemize} \end{minipage} \end{equation} Each root $\ga$ determines an orthogonal plane $\ga^{\perp}$, and for any arrangement $\bf X_n$, \begin{equation}\label{e108.10} \ifmmode {\cal A} \else$\cA$\fi({\bf X_n}):=\{\ga^{\perp} \Big| \ga \hbox{ a root}\} \end{equation} is a central arrangement in $\komp^n$, i.e., each of the planes passes through the origin. This induces a projective arrangement in ${\Bbb P}^{n-1}$, as follows. Blow up the origin of $\komp^n$; the exceptional divisor is a ${\Bbb P}^{n-1}$. The {\em projective arrangement} is the union of the intersections $[H]\cap {\Bbb P}^{n-1}$ in the exceptional divisor, where $[H]$ is the proper transform of the hyperplane $H=\ga^{\perp}$ under the blow up. The projective arrangements for $\bf B_n$ and $\bf C_n$ coincide, and these arrangements are given in ${\Bbb P}^{n-1}$ with projective coordinates $(x_1:\ldots :x_n)$ as follows: \begin{equation}\label{e108.2} \begin{minipage}{14cm}\begin{tabbing} $\ifmmode {\cal A} \else$\cA$\fi({\bf A_n})$: \quad \= $\{ x_i=0,\ i=1,\ldots, n;\ x_i=x_j,\ 1\leq i<j\leq n\}$;\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf B_n})$: \> $\{ x_i=0,\ i=1,\ldots, n;\ x_i=\pm x_j,\ 1\leq i<j\leq n\}$;\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf D_n})$: \> $\{ x_i=\pm x_j,\ 1\leq i<j\leq n\}$;\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf F_4})$: \> $\{ x_i=0,\ i=1,\ldots, n;\ x_i=\pm x_j,\ 1\leq i<j\leq 4,\ {1\over 2}(x_1\pm x_2\pm x_3\pm x_4)\}$;\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf E_6})$: \> $\{ x_i=\pm x_j,\ 1\leq i<j \leq 5,\ {1\over 2}(x_1\pm x_2 \pm x_3\pm x_4 \pm x_5 + x_6)\}$. \end{tabbing} \end{minipage} \end{equation} For the arrangement of type $\bf A_n$ we have made the coordinate transformation $x_1=\ge_1-\ge_{n+1}, \ldots, x_n=\ge_n-\ge_{n+1}$, so $x_i-x_j=\ge_i-\ge_j$ for $1\leq i< j\leq n$, and for $\bf E_6$ we have taken $x_6$ to replace $x_8-x_7-x_6$. The arrangements above are the arrangements defined by the projective reflection groups $PW({\bf X_n})$. Each hyperplane is the reflection plane for the reflection on the corresponding root. From this point of view these arrangements are studied in \cite{OS2}. \subsection{Rank 4 arrangements} As described above, the groups $W({\bf A_4})$, $W({\bf B_4})$, $W({\bf D_4})$ and $W({\bf F_4})$ give rise to projective arrangements in ${\Bbb P}^3$. They consists of ten, 16, 12 and 24 planes, respectively. They may also be described as follows (see \cite{GS})): \begin{tabbing} $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$: \quad \= four faces of a tetrahedron plus the six symmetry planes; \\ $\ifmmode {\cal A} \else$\cA$\fi({\bf B_4})$: \> \parbox{12cm}{six faces of a cube plus the nine symmetry planes plus the plane at infinity;}\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$: \> \parbox[t]{12cm}{six faces of a cube plus the six symmetry planes through two edges each, OR: eight faces of an octahedron plus three symmetry planes containing four vertices each plus the plane at infinity;}\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf F_4})$: \> \parbox[t]{12cm}{ ``desmic figure'': six faces of the cube, eight faces of an inscribed octahedron, nine symmetry planes and the plane at infinity; this is also determined by the regular 24-cell;} \end{tabbing} The combinatorial description of these arrangements can be encoded in numbers: \begin{equation}\label{e109.1} t_q(j):=\#\{{\Bbb P}^j \hbox{'s of the arrangement through which $q$ of the reflection planes pass}\}. \end{equation} In the case of the above arrangements we have the following data ($t_q:=t_q(0)$, the number of points): \begin{equation}\label{e109.2}\begin{minipage}{14cm} \begin{tabbing} $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$: \quad \= $t_6=5,\ t_4=10;\ t_3(1)=10,\ t_2(1)=15.$\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf B_4})$: \> $t_9=4, t_6=8,\ t_5=12,\ t_4=16;\ t_4(1)=6,\ t_3(1)=16,\ t_2(1)=36$. \\ $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$: \> $t_6=12,\ t_3=12;\ t_3(1)=16,\ t_2(1)=18.$ \\ $\ifmmode {\cal A} \else$\cA$\fi({\bf F_4})$: \> $t_9=24,\ t_4=96;\ t_4(1)=18,\ t_3(1)=32,\ t_2(1)=72.$ \end{tabbing} \end{minipage} \end{equation} \begin{definition}\label{d109.1} An arrangement $\ifmmode {\cal A} \else$\cA$\fi\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^n$ is said to be in (combinatorial) {\em general position}, if $t_q(j)=0$ for all $q>n-j$. All ${\Bbb P}^j$'s $\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal A} \else$\cA$\fi$ for which $j>n-q$ holds are the {\em singularities} of the arrangement. The singularities are {\it genuine} if they are not the intersection of higher-dimensional singular loci with one of the planes of the arrangement. The union of all genuine singularities is the {\em singular locus}. \end{definition} In the above arrangements we have the following singular loci: \begin{equation}\label{e109.3}\begin{minipage}{14cm} \begin{tabbing} $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$: \quad \= five singular points, ten singular lines;\\ $\ifmmode {\cal A} \else$\cA$\fi({\bf B_4})$: \> 12=4+8 (genuine) singular points, 22=6+16 singular lines; \\ $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$: \> 12 singular points, 16 singular lines; \\ $\ifmmode {\cal A} \else$\cA$\fi({\bf F_4})$: \> 24 (genuine) singular points, 50=18+32 singular lines. \end{tabbing} \end{minipage} \end{equation} \subsection{The tetrahedron} Consider now the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$ in (\ref{e109.2}). By (\ref{e109.3}) the singular locus consists of five points and ten lines. We introduce the following notation: $P_1=(1,0,0,0),\ P_2=(0,1,0,0),\ P_3=(0,0,1,0),\ P_4=(0,0,0,1),\ P_5=(1,1,1,1)$, and $l_{ij}$ will denote the line joining $P_i$ and $P_j$. Each line contains two of the five points, and at each of the points four of the ten lines meet. The arrangement is {\em resolved} by performing the following birational modification of ${\Bbb P}^3$: \begin{equation}\label{e109.4} \begin{minipage}{14cm}\begin{itemize}\item[a)] Blow up the five points, $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$; \item[b)] Blow up the proper transforms of the ten lines, $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2:\tilde{{\Bbb P}}^3\longrightarrow} \def\sura{\twoheadrightarrow \widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3,\ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta:\tilde{{\Bbb P}}^3\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$. \end{itemize} \end{minipage} \end{equation} In the resolution 15 exceptional divisors $E_1,\ldots,E_5$ and $L_{12},\ldots, L_{45}$ are introduced. The $E_i$ are the proper transforms of the exceptional divisors introduced under $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$, and are isomorphic to ${\Bbb P}^2$ blown up in the four points $(1:0:0),\ (0:1:0),\ (0:0:1),\ (1:1:1)$, as are the proper transforms $H_i$ of the ten planes of the arrangement. The ten exceptional divisors $L_{ij}$ are each isomorphic to ${\Bbb P}^1\times {\Bbb P}^1$. The symmetry group of $\tilde{{\Bbb P}}^3$ consists of projective linear transformations of ${\Bbb P}^3$ which preserve the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$, together with certain {\em birational} transformations of ${\Bbb P}^3$ which are {\em regular} on $\tilde{{\Bbb P}}^3$, i.e., which contain the singular locus (\ref{e109.3}) with simple multiplicity in their ramification locus. Hence the Weyl group itself, $W({\bf A_4})=\gS_5$ (symmetric group on five letters) is contained in the symmetry group. But in fact, $\gS_6$ is the symmetry group, and the extra generator is a permutation of one of the $E_i$ and $H_j$, which clearly can be done {\em on } $\tilde{{\Bbb P}}^3$. \subsection{A birational transformation} Note that since each of the ten lines in (\ref{e109.3}) contains two of the five points which are blown up under $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$, the normal bundle of the proper transform of each line on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3$ is $\ifmmode {\cal O} \else$\cO$\fi(1-2)\oplus\ifmmode {\cal O} \else$\cO$\fi(1-2)= \ifmmode {\cal O} \else$\cO$\fi(-1)\oplus \ifmmode {\cal O} \else$\cO$\fi(-1)$. By general results of threefold birational geometry, it follows that \begin{equation}\label{e110.1} \begin{minipage}{14cm}\begin{itemize}\item[a)] The divisors $L_{ij}$ on $\tilde{{\Bbb P}}^3$ may be blown down to an ordinary threefold rational point (node), i.e., a singularity given by the equation $x^2+y^2+z^2+t^2=0$, OR: \item[b)] The ten lines on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3$ may be blown down to the nodes mentioned in a). \end{itemize}\end{minipage}\end{equation} In other words, there is a threefold which we denote by $T$, which contains ten threefold nodes, with a birational triangle: \begin{equation}\label{e110.2} \unitlength1cm \begin{picture}(3,2)(0,-.5) \put(-.5,1){$\tilde{{\Bbb P}}^3$}\put(0,1.1){\vector(1,0){1.5}} \put(.7,1.3){$ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$} \put(1.8,1){$\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3 \stackrel{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1}{\longrightarrow} \def\sura{\twoheadrightarrow} {\Bbb P}^3$} \put(-.4,0){$\Pi_2$} \put(-.3,.9){\vector(1,-1){1}} \put(1.8,0){$\Pi_1$} \put(1.9,.9){\vector(-1,-1){1}} \put(.7,-.5){$T$} \end{picture} \end{equation} The map $\Pi_2$ blows down the union of ten disjoint ``quadric surfaces'' (i.e., divisors isomorphic to ${\Bbb P}^1\times {\Bbb P}^1$) to ordinary nodes, while $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ blows these quadric surfaces down to ten disjoint lines, which $\Pi_1$ then blows down to the same ten isolated nodes. The 5+10 divisors $E_i$ and $H_j$ on $\tilde{{\Bbb P}}^3$ have the following properties: \begin{equation}\label{e111.1}\begin{minipage}{14cm} \begin{itemize}\item[a)] Each is isomorphic to ${\Bbb P}^2$ blown up in four points; \item[b)] Each contains ten lines of intersection with the other 15, forming an arrangement in the blown up ${\Bbb P}^2$ of ten lines meeting in 15 points. \item[c)] Under the birational map $\Pi_2$ each of the divisors $E_i$ and $H_j$ are blown down to a ${\Bbb P}^2$; the image of the ten lines of b) lie four at a time in each of these ${\Bbb P}^2$'s, as the four $t_3$-points of the following arrangement, which is the union of the intersections of the given ${\Bbb P}^2$ with the others: $$\unitlength1cm \begin{picture}(8,3.5)(1,0.8) \thinlines \put(2,1.5){\line(1,0){6.5}} \put(2,1.5){\line(4,1){5}} \put(2,1.5){\line(6,5){3.5}} \put(2,1.5){\line(-1,0){0.5}} \put(2,1.5){\line(-4,-1){0.5}} \put(2,1.5){\line(-6,-5){0.5}} \put(5,.75){\line(0,1){4}} \put(5,4){\line(6,-5){3.5}} \put(5,4){\line(-6,5){0.5}} \put(8,1.5){\line(-4,1){5}} \put(8,1.5){\line(4,-1){0.5}} \put(2,1.5){\circle*{.2}} \put(5,2.25){\circle*{.2}} \put(5,4){\circle*{.2}} \put(8,1.5){\circle*{.2}} \end{picture}$$ \item[d)] The composition $\Pi_2\circ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^{-1}$ restricted to each of the planes $H_j$ is a usual Cremona transformation, blowing up three non-colinear points and blowing down the three joining lines. {\it Proof:} Take a face $H_j$ of the tetrahedron; $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$ blows up the three vertices it contains, so $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}(H_j)$ (the proper transform of $H_j$) is ${\Bbb P}^2$ blown up in three points. Under $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$, a fourth point of $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}(H_j)$ is blown up, but it is blown down again under $\Pi_2$, as are the proper transforms of the three lines (in the plane $H_j$) joining the three vertices. By symmetry the same holds for all the $H_j$. \end{itemize} \end{minipage} \end{equation} It follows that on $T$, the images $\tilde{H}_j=\Pi_2(H_j)$ and $\tilde{E}_i=\Pi_2(E_i)$ are copies of ${\Bbb P}^2$, each containing four of the ten nodes of $T$. Furthermore, in each of $\tilde{H}_j$ and $\tilde{E}_j$ we have the four $t_3$-points of the arrangement (\ref{e111.1}), which are these four nodes of $T$. Finally, since there are 15 ${\Bbb P}^2$'s, ten nodes and four of them in each of the 15 ${\Bbb P}^2$'s, there are five of these divisors passing through a given node. Explicitly, take the node $n_{ij}$ corresponding to the line $l_{ij}$ in (\ref{e109.3}). Then it meets the exceptional divisors $\~E_i,\ \~E_j$, as well as the three of the $\~H_{\nu}$ for which $H_{\nu}$ contains the line $l_{ij}$. \subsection{Fermat covers associated with arrangements} Let $\ifmmode {\cal A} \else$\cA$\fi\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^n$ be an arrangement of hyperplanes, i.e., a union $\ifmmode {\cal A} \else$\cA$\fi=\cup _{i=1}^k H_i$ of $k$ hyperplanes, and let $d\geq2 $ be an integer. To the pair $(\ifmmode {\cal A} \else$\cA$\fi,d)$ there is an associated function field $\ifmmode {\cal L} \else$\cL$\fi(\ifmmode {\cal A} \else$\cA$\fi,d)$, an algebraic extension of the rational function field $\ifmmode {\cal M} \else$\cM$\fi({\Bbb P}^n)$. It defines, in a unique way, a branched cover $Y(\ifmmode {\cal A} \else$\cA$\fi,d)\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^n$, and a unique desingularisation $\~Y(\ifmmode {\cal A} \else$\cA$\fi,d)$. The function field is defined by: \begin{equation}\label{e111a.1} \ifmmode {\cal L} \else$\cL$\fi(\ifmmode {\cal A} \else$\cA$\fi,d)=\komp\left({x_1 \over x_0},\ldots ,{x_n\over x_0}\right)\left[ \sqrt{H_2/H_1}\hspace{-1.5cm}\raisebox{.2cm}{$\scriptstyle d$} \hspace{1.5cm},\ldots, \sqrt{H_k/H_1}\hspace{-1.5cm}\raisebox{.2cm}{$\scriptstyle d$} \hspace{1.5cm}\right], \end{equation} and the cover $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ is the so-called Fox closure of the \'etale cover over ${\Bbb P}^n-\ifmmode {\cal A} \else$\cA$\fi$ which is defined by (\ref{e111a.1}). $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ is smooth outside of the {\em singular locus} of $\ifmmode {\cal A} \else$\cA$\fi$ (Definition \ref{d109.1}), and the singularities of $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ are resolved by resolving the singularities of $\ifmmode {\cal A} \else$\cA$\fi$. This is done by first blowing up all (genuine, i.e., not near pencil) singular points, then all singular lines, and so forth. The resolution (\ref{e109.4}) is a typical example. This is described in more detail in the author's thesis; the desingularisation $\~Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ is the fibre product in the following diagram: \begin{equation}\label{e111a.2} \begin{array}{ccc}\~Y(\ifmmode {\cal A} \else$\cA$\fi,d) & \longrightarrow} \def\sura{\twoheadrightarrow & Y(\ifmmode {\cal A} \else$\cA$\fi,d) \\ \~{\pi}\downarrow & & \downarrow\pi \\ \~{{\Bbb P}}^n & \stackrel{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}{\longrightarrow} \def\sura{\twoheadrightarrow} & {\Bbb P}^n \end{array} \end{equation} where the horizontal arrows are modifications and the vertical arrows are Galois covers with Galois group $(\integer/d\integer)^{k-1}$, which is the Galois group of the field extension of (\ref{e111a.1}). $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$ is the modification of ${\Bbb P}^n$ which resolves the singularities of $\ifmmode {\cal A} \else$\cA$\fi$. For example, each singular point $P$ on $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ is resolved by an {\em irreducible} divisor $D_P$, which itself is a Fermat cover $Y(\ifmmode {\cal A} \else$\cA$\fi',d)$, where $\ifmmode {\cal A} \else$\cA$\fi'\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^{n-1}$ is the arrangement induced in the exceptional ${\Bbb P}^{n-1}$ which resolves the point $P'=\pi(P)$. It consists of $k'$ planes, where $k'$ is the number of the $k$ hyperplanes which meet at the point $P'$, and in the process of resolving $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$, the cover $Y(\ifmmode {\cal A} \else$\cA$\fi',d)$ is resolved also. Hence on $\~Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ there is a {\em smooth} divisor $D_P$ which resolves the singular point $P$ of $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$. The singular covers $Y(\ifmmode {\cal A} \else$\cA$\fi,d)$ can also be realised as complete intersections, namely the intersections of $N=k-n$ Fermat hypersurfaces in ${\Bbb P}^{k-1}$: \begin{eqnarray}\label{e111a.3} F_1 & = & a_{11}x_1^d+\cdots + a_{1k}x_k^d \nonumber\\ \vdots & & \vdots \\ F_N & = & a_{N1}x_1^d+\cdots +a_{Nk}x_k^d \nonumber \end{eqnarray} where $a_{11}H_1+\cdots +a_{1k}H_k, \ldots , a_{N1}H_1+\cdots + a_{Nk}H_k$ are the $(k-n)$ linear relations among the $k$ hyperplanes $H_i$. The map $Y(\ifmmode {\cal A} \else$\cA$\fi,d)\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^n$ is realised explicitly by the map $(x_1,\ldots,x_k)\mapsto (x_1^d,\ldots,x_k^d)$. \subsection{The hypergeometric differential equation} The Fermat covers for the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$ are closely related to solutions of the hypergeometric differential equation on ${\Bbb P}^3$, which is an algebraic differential equation with regular singular points, whose singular locus {\em coincides} with the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$, meaning that solutions are locally branched along the planes of the arrangement. First we introduce a new notation for the 15 surfaces $E_i,\ H_j$. These can be numbered by pairs $(i,j),\ i<j\in \{0,\ldots,5\}$, with $E_i=H_{0i}$ and \begin{equation}\label{e111b.1} H_{ij}\cap H_{kl}\neq \emptyset \iff i\neq j\neq k\neq l. \end{equation} We denote by $0i$ the point $P_i$ in ${\Bbb P}^3$, and by $0ij$ the singular line joining $0i$ and $0j$ in ${\Bbb P}^3$. We then let $L_{0ij}$ denote the exceptional divisor on $\~{{\Bbb P}}^3$. We have (in ${\Bbb P}^3$) \begin{equation}\label{e111b.2} H_{ij}\cap H_{kl}=0mn \iff \{i,j,k,l\}\cap \{0,m,n\}=\emptyset. \end{equation} We want to consider branched covers $Y\longrightarrow} \def\sura{\twoheadrightarrow \~{{\Bbb P}}^3$ (with $\~{{\Bbb P}}^3$ as in (\ref{e109.4})), which are branched along the $H_{ij}$ and the $L_{0ij}$. Hence we let \begin{equation}\label{e111b.3} n_{ij}:=\hbox{ branching degree along $H_{ij}$};\quad n_{0ij}:=\hbox{ branching degree along $L_{0ij}$}, \end{equation} and of course $n_{ij},\ n_{0ij}\in \integer\cup \infty$. (It makes sense to allow negative branching degrees, as we will see below.) To define the hypergeometric differential equation we may just as well work on ${\Bbb P}^n$ with homogenous coordinates $(x_0:\ldots:x_n)$, and consider the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf A_n})$ of (\ref{e108.2}). Let $\gl_i\in \rat,\ \ i=0,\ldots,n+1, \infty,\ \sum_i\gl_i=n+1$. The hypergeometric differential equation is: \begin{equation}\label{e111b.4} \left\{ \begin{minipage}{14cm}$\ds(x_i-x_j)\del_i\del_jF + (\gl_i-1)(\del_iF-\del_jF) = 0,\ 1\leq i<j \leq n$ \\ $x_i(x_i-1)\del_i^2F+P_i(x,\gl)\del_iF +(\gl_i-1)\sum{x_{\ga}(x_{\ga}-1) \over (x_i-x_{\ga})}\del_{\ga}F + \gl_{\infty}(1-\gl_i)F=0,\ 1\leq i \leq n$. \end{minipage} \right. \end{equation} where $$P_i(x,\gl)=x_i(x_i-1)\sum{1-\gl_{\ga} \over x_i-x_{\ga}} + \gl_0+\gl_i -3-(2\gl_i+\gl_0+\gl_{n+1})x_i.$$ A solution of (\ref{e111b.4}) turns out to be a period of an algebraic curve (the periods are many valued, as are the solutions of (\ref{e111b.4})). The curve is \begin{equation}\label{e111b.5} y^{\nu}=x^{\mu_0}(x-1)^{\mu_{n+1}}(x-t_1)^{\mu_1}\cdots (x-t_n)^{\mu_n}, \end{equation} where the $\mu_i,\ \nu$ are related to the $\gl_i$ by the relation \begin{equation}\label{e111b.6} {\mu_i \over \nu} = 1-\gl_i. \end{equation} The equation (\ref{e111b.4}) has an $(n+1)$-dimensional solution space, spanned by $(n+1)$ periods of differentials of the curve (\ref{e111b.5}): \begin{equation}\label{e111c.1} \go_i=\int_{\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta_i}{dx \over y},\quad <\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta_0,\ldots,\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta_n>=H^1(C,\integer).\end{equation} Taking these gives a homogenous many valued map \begin{equation}\label{e111c.2} (\go_0,\ldots, \go_n):D\subset} \def\nni{\supset} \def\und{\underline \~{{\Bbb P}}^n \stackrel{\phi}{\longrightarrow} \def\sura{\twoheadrightarrow} {\Bbb P}^n, \end{equation} where $D$ is some Zariski open set (see (\ref{e111d.2}) below). The map is well-defined, since not all $\go_i$ vanish simultaneously. For very special values of the parameters $\gl_i$, the image of $\phi$ is the complex ball $\ball_n\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^n$ (this is just the Borel embedding of $\ball_n$ in its compact dual). In fact, one has the following theorem: \begin{theorem}[\cite{DM}, \cite{T}]\label{t111c.1} If the following conditions are satisfied, then $\phi(D)=\ball_n$: $$\sum\mu_i=2,\quad \forall_{i,j}:\ (1-\mu_i-\mu_j)^{-1}\in \integer\cup \infty.$$ In this case there exists a finite cover $$Y\longrightarrow} \def\sura{\twoheadrightarrow D$$ branched along the total transform of $\ifmmode {\cal A} \else$\cA$\fi({\bf A_n})$, which is a quotient $\gG\backslash \ball_n$ with $\gG$ torsion free. \end{theorem} The integers $n_{ij}:=(1-\mu_i-\mu_j)^{-1}$ are then just the branching degrees of $Y\longrightarrow} \def\sura{\twoheadrightarrow D$ along the divisor $H_{ij}$. In fact the numbering introduced in (\ref{e111b.1}) can be done analogously for any $n$. In the special case of $\ifmmode {\cal A} \else$\cA$\fi({\bf A_4})$ on ${\Bbb P}^3$, the integers $n_{0ij}$ of (\ref{e111b.3}) are determined by the relation $$n_{0ij}=2\left({1\over n_{kl}}+{1\over n_{lm}}+{1\over n_{km}}\right)^{-1}, $$ where the line $0ij$ is the intersection of $H_{kl},\ H_{lm},\ H_{km}$, and these together with the $n_{ij}$ describe the branching degrees along the entire branch locus. The solutions of the equations in Theorem \ref{t111c.1} are as follows: \renewcommand{\arraystretch}{1.5} \begin{equation}\label{e111d.1}\begin{array}{cl} 1) & {1\over 3}, {1\over 3}, {1\over 3}, {1\over 3}, {1\over 3}, {1\over 3}, \\ 2) & {1 \over 2}, {1\over 2}, {1\over 4}, {1\over 4}, {1\over 4}, {1\over 4} \\ 3) & {3 \over 4}, {1\over 4}, {1\over 4}, {1\over 4}, {1\over 4}, {1\over 4} \\ 4) & {1\over 2}, {1\over 3}, {1\over 3}, {1\over 3}, {1\over 3}, {1\over 6}\\ 5) & {3\over 8}, {3\over 8}, {3\over 8}, {3\over 8}, {3\over 8}, {1\over 8} \\ 6) & {5\over 12}, {5\over 12}, {5\over 12}, {1\over 4}, {1\over 4}, {1\over 4} \\ 7) & {7 \over 12}, {5 \over 12}, {1\over 4}, {1\over 4}, {1\over 4}, {1\over 4} \end{array} \end{equation} \renewcommand{\arraystretch}{1.2} The set $D$ of (\ref{e111c.2}) is determined as the complement of \begin{equation}\label{e111d.2} H_{\infty}=\{H_{ij}\Big| n_{ij}=\infty; \ L_{0ij}\Big| n_{0ij}=\infty\}\subset} \def\nni{\supset} \def\und{\underline \~{{\Bbb P}}^3. \end{equation} This is the locus which the uniformizing map (\ref{e111c.2}) maps onto the {\em boundary} of $\ball_3\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^3$, i.e., $\phi(D)=\ball_n,\ \phi(H_{\infty})\subset} \def\nni{\supset} \def\und{\underline \del\ball_n$. This requires of course that the corresponding covers of the divisors on the cover $\~Y$ be abelian varieties (as these are the compactification divisors on ball quotients). This can happen as follows \begin{equation}\label{e111d.3}\begin{minipage}{12cm} \begin{itemize}\item[(i)] On one of the $H_{ij}$, this can occur if the branching degrees are: 2 for the six lines of (\ref{e111.1}), and $-4$ for the four exceptional curves. \item[(ii)] On $L_{0ij}$, this can happen if $\mu_k+\mu_l+\mu_m=1,\ \mu_0+\mu_i+\mu_j=1$. \end{itemize} \end{minipage} \end{equation} In the second case, the surface $S_{0ij}$ covering $L_{0ij}$ is of the form $C_1\times C_2$, where $C_1\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$ (respectively $C_2\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$) is a cover, with branching determined by $(\mu_k,\mu_l,\mu_m)$ (respectively determined by $(\mu_0,\mu_i,\mu_j)$). It is an abelian variety $\iff$ both curves $C_i$ are elliptic. Note that $Y\longrightarrow} \def\sura{\twoheadrightarrow \~{{\Bbb P}}^3$ will be a Fermat cover $\iff$ all $n_{ij}$ coincide $\iff$ all $\mu_i$ conicide. In particular, \begin{proposition}\label{p111d.1} The only ball quotient in the list (\ref{e111d.1}) which is a Fermat cover which is a ball quotient is the solution 1), namely $Y(\ifmmode {\cal A} \else$\cA$\fi({\bf A_4}),3)$ is a ball quotient. \end{proposition} \begin{remark} We will see later (see I3 following Lemma \ref{lq4.1} below) that the solution 4) gives rise also to a Fermat cover which is a ball quotient, namely $Y(\ifmmode {\cal A} \else$\cA$\fi({\bf D_4}),3)$. \end{remark} \section{The Segre cubic ${\cal S}_3$} In this section we will show that the variety $T$ of (\ref{e110.2}) has a projective embedding as a cubic hypersurface known as the Segre cubic, which we denote by $\ifmmode {\cal S} \else$\cS$\fi_3$. \subsection{Segre's cubic primal} In ${\Bbb P}^5$ with homogenous coordinates $(x_0:\ldots:x_5)$ consider the locus \begin{equation}\label{e111.3} \ifmmode {\cal S} \else$\cS$\fi_3:=\{\sum_{i=0}^5x_i=0;\quad \sum_{i=0}^5x_i^3=0\}. \end{equation} As the first equation is linear, this shows that $\ifmmode {\cal S} \else$\cS$\fi_3$ is a hypersurface, i.e., $\ifmmode {\cal S} \else$\cS$\fi_3\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^4=\{x\in {\Bbb P}^5\big|\sum x_i=0\}$. Using $(x_0:\ldots:x_5)$ as projective coordinates, the relation $x_5=-x_0-\cdots -x_4$ gives the equation of $\ifmmode {\cal S} \else$\cS$\fi_3$ as a hypersurface; however, the equation in ${\Bbb P}^5$ shows that $\ifmmode {\cal S} \else$\cS$\fi_3$ is invariant under the symmetry group $\gS_6$, acting on ${\Bbb P}^5$ by permuting coordinates, which is not so immediate from the hypersurface equation. It is known that for any degree $d$ there is an upper bound on the number of ordinary double points which a hypersurface of degree $d$ can have, the so-called Varchenko bound. For cubic threefolds this number is ten, and it has been known since the last century that $\ifmmode {\cal S} \else$\cS$\fi_3$ is the {\em unique} (up to isomorphism) cubic with ten nodes. The nodes on $\ifmmode {\cal S} \else$\cS$\fi_3$ are given by the points of ${\Bbb P}^5$ for which three of the coordinates are 1 and the other three are $-1$. This is just the $\gS_6$-orbit of \begin{equation}\label{e112.0} (1,1,1,-1,-1,-1). \end{equation} There is another interesting locus on $\ifmmode {\cal S} \else$\cS$\fi_3$. Consider, in ${\Bbb P}^5$, the planes $P_{\gs}$ given by \begin{equation}\label{e112.1} P_{\gs}=\{x_{\gs(0)}+x_{\gs(3)}=x_{\gs(1)} + x_{\gs(4)}=x_{\gs(2)}+x_{\gs(5)}=0\}, \end{equation} where $\gs\in \gS_6$. There are 15 such $P_{\gs}$'s, the $\gS_6$-orbit of \begin{equation} P_{id}=\{x_0+x_3=x_1+x_4=x_3+x_5=0\}. \end{equation} One checks easily that each $P_{\gs}$ contains four of the double points; for example $P_{id}$ contains the following: $$(1,1,-1,1,-1,-1),\ (1,-1,1,1,-1,-1),\ (1,-1,-1,1,1,-1),\ (1,-1,-1,1,-1,1).$$ Furthermore, the intersection of $P_{id}$ with the other $P_{\gs}$ is the line arrangement (\ref{e111.1}). It is easily checked that each $P_{\gs}$ is contained entirely in $\ifmmode {\cal S} \else$\cS$\fi_3$. One can also argue as follows. Any line in ${\Bbb P}^5$ which contains two of the nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$ meets $\ifmmode {\cal S} \else$\cS$\fi_3$ with multiplicity 4, hence is contained in $\ifmmode {\cal S} \else$\cS$\fi_3$. Similarly, each $P_{\gs}$ meets $\ifmmode {\cal S} \else$\cS$\fi_3$ in the six lines of the arrangement (\ref{e111.1}), hence is contained in $\ifmmode {\cal S} \else$\cS$\fi_3$. We just remark that the hyperplane sections $\{x_i=0\}$ of $\ifmmode {\cal S} \else$\cS$\fi_3$ are cubic surfaces with equation \begin{equation}\label{e112.3} S_3 = \{ \sum_{i=0}^4x_i=\sum_{i=0}^4x_i^3=0\}. \end{equation} This cubic surface is known as the Clebsch diagonal surface and is a remarkably beautiful object. It is the unique cubic surface having $\gS_5$ as symmetry group. The relation between $S_3$ and the icosahedral group was studied by Hirzebruch. It turns out that $S_3$ is $A_5$-equivariantly birational to the Hilbert modular surface for $\ifmmode {\cal O} \else$\cO$\fi_{\rat(\sqrt{5})}$, of level $\sqrt{5}$. Other interesting hyperplane sections are given by the hyperplanes $\ifmmode {\cal T} \else$\cT$\fi_{ij}=\{x_i-x_j=0\}$; indeed, $\ifmmode {\cal T} \else$\cT$\fi_{ij}$ also contains four of the ten nodes, hence $\ifmmode {\cal T} \else$\cT$\fi_{ij}\cap \ifmmode {\cal S} \else$\cS$\fi_3$ is a four-nodal cubic surface. This four-nodal cubic surface is projectively unique, and is called the Cayley cubic. \subsection{A birational transformation} \begin{theorem}\label{t113.1} The variety $T$ of equation (\ref{e110.2}) is biregular to $\ifmmode {\cal S} \else$\cS$\fi_3$; the isomorphism $\psi:T\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal S} \else$\cS$\fi_3$ defined below is $\gS_6$-equivariant. \end{theorem} {\bf Proof:} Following Baker \cite{Baker}, IV, p.~152, we define a birational map $$\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda:{\Bbb P}^3- - \ra \ifmmode {\cal S} \else$\cS$\fi_3.$$ Consider all quadric surfaces in ${\Bbb P}^3$ passing through the points $P_i$ of (\ref{e109.3}). A base of this linear system is given by the following degenerate quadrics. Let $(z_0:\ldots:z_3)$ be homogenous coordinates on ${\Bbb P}^3$, and set \begin{equation}\label{e113.1} \begin{array}{lll} \xi=z_0(z_3-z_1), & \eta=z_1(z_3-z_2), & \gz=z_2(z_3-z_0); \\ \xi'=z_1(z_3-z_0), & \eta'=z_2(z_3-z_1), & \gz'=z_0(z_3-z_2). \end{array} \end{equation} These quadrics satisfy the relations $\xi+\eta+\gz=\xi'+\eta'+\gz'$ and $\xi\eta\gz=\xi'\eta'\gz'$. Now change coordinates by setting \begin{equation}\label{e113.a} \begin{array}{lll} \xi=X+Y, & \eta=Y+Z, & \gz=X+Z; \\ \xi'=-(X'+Y'), & \eta'=-(Y'+Z'), & \gz'=-(X'+Z'). \end{array}\end{equation} Then the relations $\xi+\eta+\gz=\xi'+\eta'+\gz'$ and $\xi\eta\gz=\xi'\eta'\gz'$ become \begin{eqnarray}\label{e113.2} X+Y+Z+X'+Y'+Z' & = & 0 \\ X^3+Y^3+Z^3+(X')^3+(Y')^3+(Z')^3 & = & 0. \nonumber \end{eqnarray} One sees this is just equation (\ref{e111.3}) of the Segre cubic. This yields a rational map $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda:{\Bbb P}^3- - \ra \ifmmode {\cal S} \else$\cS$\fi_3$, $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda(z_0:z_1:z_2:z_3)= (X,Y,Z,X',Y',Z')$. The base locus of the linear system of quadrics defining $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda$ (\ref{e113.1}) is the five points of (\ref{e109.3}), as the quadrics all contain these points. It follows that $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda$ blows up all five points, the exceptional divisors $E_1,\ldots ,E_5$ being projective planes. Now consider one of the ten lines of (\ref{e109.3}); for example, the one given by $z_2=z_3=0$. Then $\eta=\gz=\eta'=\gz'=0$ and $\xi=\xi'=-z_0z_1$. In other words, $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda$ maps that line to the point $(1,0,0,1,0,0)$ in the $(\xi,\eta,\gz,\xi',\eta',\gz')$ space, which is the point $(1,1,-1,-1,-1,1)$ in the $(X,Y,Z,X',Y',Z')$ space. But that is just one of the ten nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$. From $\gS_6$-symmetry we conclude that $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda:{\Bbb P}^3- - \ra \ifmmode {\cal S} \else$\cS$\fi_3$ coincides with the map $\Pi=\Pi_1\circ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}$, with $\Pi_1$ as in (\ref{e110.2}) and $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$ as in (\ref{e109.4}). In other words, $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda=\Pi$ is the composition of morphisms \begin{equation}\label{e113.3} \unitlength1cm \begin{picture}(3,2) \put(1.5,1.66){$\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^3$} \put(1.5,1.5){\vector(-1,-1){.9}} \put(1.7,1.5){\vector(1,-1){.9}} \put(.2,.33){${\Bbb P}^3$} \put(2.76,.33){$T,$} \put(.76,.3){$- - - - \ra$} \put(1.5,.66){$\Pi$} \put(.66,1.33){$\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$} \put(2.33,1.33){$\Pi_2$} \end{picture} \end{equation} and since (\ref{e113.2}) states that $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda=\Pi$ maps onto $\ifmmode {\cal S} \else$\cS$\fi_3$, this gives an isomorphism $T\ifmmode\ \cong\ \else$\isom$\fi \ifmmode {\cal S} \else$\cS$\fi_3$. Explicitly, $t\in T,\ t\mapsto (\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1\circ\Pi^{-1}_2)(t)\mapsto \beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda((\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1\circ\Pi^{-1}_2)(t))=\psi(t)\in \ifmmode {\cal S} \else$\cS$\fi_3$ is the desired map. The $\gS_6$-equivariance follows from the fact that the whole diagram (\ref{e113.3}) is $\gS_6$-equivariant. \hfill $\Box$ \vskip0.25cm Now consider the Picard group of $\ifmmode {\cal S} \else$\cS$\fi_3$. From the explicit form of birational map as given by Theorem \ref{t113.1} and (\ref{e110.2}), we see that $\hbox{Pic}} \def\Jac{\hbox{Jac}(\ifmmode {\cal S} \else$\cS$\fi_3)$ is generated by the image of the hyperplane class, call it $H$, and the five exceptional classes $E_i$. It follows that $\hbox{Pic}} \def\Jac{\hbox{Jac}(\ifmmode {\cal S} \else$\cS$\fi_3)$ has rank 6, and the primitive part $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal S} \else$\cS$\fi_3)$, i.e., the complement of the hyperplane class, has rank 5. The 15 classes $H_{ij}$ introduced in (\ref{e111b.1}) (these are the 15 linear ${\Bbb P}^2$'s on $\ifmmode {\cal S} \else$\cS$\fi_3$ noted in (\ref{e112.1})) give classes in $\hbox{Pic}} \def\Jac{\hbox{Jac}(\ifmmode {\cal S} \else$\cS$\fi_3)$ and in $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal S} \else$\cS$\fi_3)$. The 15 hyperplanes \begin{equation} \label{e112b.3} \ifmmode {\cal H} \else$\cH$\fi_{ij}=\{x_i+x_j=0\}, \end{equation} each of which meets $\ifmmode {\cal S} \else$\cS$\fi_3$ in three of the 15 ${\Bbb P}^2$'s, give 15 {\em relations} in $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal S} \else$\cS$\fi_3)$: since $\ifmmode {\cal H} \else$\cH$\fi_{ij}\cap \ifmmode {\cal S} \else$\cS$\fi_3$ is a hyperplane section, the sum of the three ${\Bbb P}^2$'s cut out by $\ifmmode {\cal H} \else$\cH$\fi_{ij}$, i.e., $H_{i_1,j_1}+H_{i_2,j_2}+H_{i_3.j_3}=\ifmmode {\cal H} \else$\cH$\fi_{ij}\cap \ifmmode {\cal S} \else$\cS$\fi_3$, is linearly equivalent to the hyperplane class. This yields the following exact sequence of $\integer$-modules: \begin{equation}\label{e112b.1} \begin{array}{ccccccccccc} 1 & \longrightarrow} \def\sura{\twoheadrightarrow & K & \longrightarrow} \def\sura{\twoheadrightarrow & \integer\{\ifmmode {\cal H} \else$\cH$\fi_{ij}\} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer\{H_{ij}\} & \longrightarrow} \def\sura{\twoheadrightarrow & \hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal S} \else$\cS$\fi_3) & \longrightarrow} \def\sura{\twoheadrightarrow & 1 \\ & & \Big\|\wr & & \Big\|\wr & & \Big\|\wr & & \Big\|\wr & & \\ 1 & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^5 & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{15} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{15} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^5 & \longrightarrow} \def\sura{\twoheadrightarrow & 1. \end{array}\end{equation} \begin{lemma}\label{l112b.1} In the sequence (\ref{e112b.1}), all $\integer$-modules are $\gS_6$-modules, i.e., the exact sequence is one of $\gS_6$-modules. \end{lemma} {\bf Proof:} This is visible for the right three entries of the first sequence in (\ref{e112b.1}), and it then follows for $K$. \hfill $\Box$ \vskip0.25cm Now consider a generic hyperplane section of $\ifmmode {\cal S} \else$\cS$\fi_3$; this is a smooth cubic surface. Let $\nu:S=\ifmmode {\cal S} \else$\cS$\fi_3\cap H\hookrightarrow} \def\hla{\hookleftarrow \ifmmode {\cal S} \else$\cS$\fi_3$ denote the inclusion of the section, and let $\nu*:H^2(\ifmmode {\cal S} \else$\cS$\fi_3,\integer)\longrightarrow} \def\sura{\twoheadrightarrow H^2(S,\integer)$ be the induced map on cohomology. Then by the Lefschetz hyperplane theorem, this map is {\em injective}, and since both $S$ and $\ifmmode {\cal S} \else$\cS$\fi_3$ are regular (i.e., not irregular, that is, have no holomorphic one forms), we may view this as an injective map of the Picard groups: $\hbox{Pic}} \def\Jac{\hbox{Jac}(\ifmmode {\cal S} \else$\cS$\fi_3)\hookrightarrow} \def\hla{\hookleftarrow \hbox{Pic}} \def\Jac{\hbox{Jac}(S)$, and a corresponding inclusion on the primitive part. Recall also that we have on the cubic surface 27 generators (the 27 lines), 45 relations among these (the 45 tritangents), and an exact sequence on $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(S)$ as in (\ref{eB3.2}). All in all we get the following map of sequences as in (\ref{e112b.1}): \begin{equation}\label{e112b.2}\begin{array}{ccccccccccc} 1 & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^5 & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{15} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{15} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^5 & \longrightarrow} \def\sura{\twoheadrightarrow & 1 \\ & & \downarrow & & \downarrow & & \downarrow & & \downarrow & & \\ 1 & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{24} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{45} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^{27} & \longrightarrow} \def\sura{\twoheadrightarrow & \integer^6 & \longrightarrow} \def\sura{\twoheadrightarrow & 1. \end{array}\end{equation} where the right hand groups are $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal S} \else$\cS$\fi_3)$ and $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(S)$, respectively, and the down arrows are inclusions (by Lefschetz). Note that this corresponds to a symmetry breaking. Indeed, on the first sequence there is a symmetry group $\gS_6$ acting, as already noted, while on the group $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(S)$, in fact on the whole second sequence, the group $W(E_6)$ acts naturally, as is well-known. \begin{proposition}\label{p112b.1} The ideal $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$ of the ten nodes is generated by the five quadrics $\ifmmode {\cal R} \else$\cR$\fi_{\gl}$ of the Jacobian ideal of $\ifmmode {\cal S} \else$\cS$\fi_3$. \end{proposition} {\bf Proof:} The inclusion $Jac(\ifmmode {\cal S} \else$\cS$\fi_3)\subset} \def\nni{\supset} \def\und{\underline \ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$ is obvious, and the five elements of $Jac(\ifmmode {\cal S} \else$\cS$\fi_3)$ are clearly independent. The fact that $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$ has rank 5 has been verified by standard basis computations (with the algebra program Macaulay). \hfill $\Box$ \vskip0.25cm \begin{corollary}\label{c112b.1} The ideal of the ten nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$, $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$, is the Jacobian ideal of $\ifmmode {\cal S} \else$\cS$\fi_3$. \hfill $\Box$ \vskip0.25cm \end{corollary} \subsection{Uniformisation} In this section we will show that the Segre cubic $\ifmmode {\cal S} \else$\cS$\fi_3$ is actually the Satake compactification of a Picard modular variety. Let $K=\rat(\sqrt{-3})$ be the field of Eisenstein numbers, and consider the $\rat$-group $G=U(3,1;K)$, the unitary group of a hermitian form on a four-dimensional $K$-vector space with signature (3,1). Consider the arithmetic group $\gG:=G_{\integer}=U(3,1;\ifmmode {\cal O} \else$\cO$\fi_K)\subset} \def\nni{\supset} \def\und{\underline G(K)$, where $\ifmmode {\cal O} \else$\cO$\fi_K$ denotes the ring of integers in $K$. It acts on the three-ball with non-compact quotient $\ifmmode {X_{\gG}} \else$\xg$\fi$. Consider the principal congruence subgroups $\gG(\sqrt{-3})$ and $\gG(3)$, as defined in \cite{J}. These determine a corresponding level structure in the sense of Definition 2.5 of \cite{J}. Now note the following well-known isomorphisms: \begin{equation}\label{e114.1} \gG/\gG(\sqrt{-3})=\gS_6, \quad \gG(3)/\gG(\sqrt{-3})=(\integer/3\integer)^9. \end{equation} It follows from this that the corresponding quotients $X(a):=X_{\gG(a)}$, $a=1,\sqrt{-3},3$, yield Galois covers \begin{equation}\label{e114.2} X(3)\stackrel{(\integer/3\integer)^9}{\longrightarrow} \def\sura{\twoheadrightarrow} X(\sqrt{-3}) \stackrel{\gS_6}{\longrightarrow} \def\sura{\twoheadrightarrow} X(1), \end{equation} which explicitly describe the level structures involved. As usual let $X(a)^*$ denote the Satake compactification. \begin{theorem}\label{t114.1} There is a commutative diagram \begin{equation}\label{e115.0} \unitlength.4cm \begin{picture}(18,18) \put(8,1){$\ifmmode {\cal S} \else$\cS$\fi_3$} \put(3,5.7){$\tilde{{\Bbb P}}^3$} \put(14,2.5){$X(\sqrt{-3})^*$} \put(8.8,7.7){$\-X(\sqrt{-3})$} \put(7.6,12){$Y^{\wedge}$} \put(3,17){$\tilde{Y}$} \put(14.5,14){$X(3)^*$} \put(9,19){$\-X(3)$} \put(3.5,16.5){\vector(1,-1){4}} \put(8.5,18.5){\vector(-3,-1){4.3}} \put(10,18.5){\vector(1,-1){4}} \put(14,14){\vector(-3,-1){4.3}} \put(9,18.5){\vector(0,-1){9.5}} \put(3,16.5){\vector(0,-1){9.5}} \put(8,11.5){\vector(0,-1){9.5}} \put(15,13.5){\vector(0,-1){9.5}} \put(8.5,7.5){\vector(-3,-1){4}} \put(9.5,7.5){\vector(1,-1){4}} \put(13.5, 2.5){\vector(-3,-1){4}} \put(3.5,5.5){\vector(1,-1){4}} \end{picture} \end{equation} where the horizontal maps from right to left are isomorphisms, those from left to right are birational, and the vertical maps are $(\integer/3\integer)^9$ covers. \end{theorem} {\bf Proof:} First we have, over ${\Bbb P}^3$, a singular cover $T_{DM}$ defined by the solution 1) of (\ref{e111d.1}). This is desingularised by blowing up the ${\Bbb P}^3$ along the singular locus of the arrangement, $\tilde{{\Bbb P}}^3 \longleftarrow} \def\rar{\rightarrow \tilde{T}_{DM}$. From the fact that all $\mu_i=1/3$, we see that all $n_{ij}$ and $n_{0ij}$ are equal to three, that is $T_{DM}$ is the Fermat cover $Y(\ifmmode {\cal A} \else$\cA$\fi({\bf A_4}),3)$, and $\tilde{T}_{DM}$ is its desingularisation $\tilde{Y}:=\tilde{Y}(\ifmmode {\cal A} \else$\cA$\fi({\bf A_4}),3)$ as in (\ref{e111a.2}); see also Proposition \ref{p111d.1}. By Theorems \ref{t111c.1} and \ref{t113.1}, $\tilde{Y}$ is the desingularisation of the ball quotient $\gG'\backslash \ball_3$, for some torsion free group $\gG'$. Blowing $\tilde{Y}$ down from $\tilde{{\Bbb P}}^3$ to $\ifmmode {\cal S} \else$\cS$\fi_3$ gives the singular variety $Y^{\wedge}$, which we will see in a minite is the Satake compactification of the ball quotient. Hence we only need to identify the groups and check the compactifications coincide. As to the first, we start with \begin{Lemma}\label{l115.1} Let $\GQ$ be an isotropic $\rat$-form of $U(3,1)$, $G_{\rat}\sim U(3,1;L)$, $L$ imaginary quadratic over $\rat$, and let $\gG\subset} \def\nni{\supset} \def\und{\underline G_{\rat}$ a torsion free arithmetic subgroup with arithmetic quotient $\ifmmode {X_{\gG}} \else$\xg$\fi$, Baily-Borel compactification $\ifmmode {X_{\gG}^*} \else$\xgs$\fi$ and toroidal compactification $\ifmmode {\overline{X}_{\gG}} \else$\xgc$\fi$. Then the isomorphism class of a single compactification divisor determines the field $L$, and hence $G_{\rat}$ up to isogeny. \end{Lemma} {\bf Proof:} First note that for $U(3,1)$ the parabolic (there is only one conjugacy class of parabolics, as the $\fR$-rank is one) takes on the particularly simple form \begin{equation}\label{e115.1} \begin{minipage}{14cm} \hspace*{\fill} $ P \ifmmode\ \cong\ \else$\isom$\fi (\ifmmode {\cal R} \else$\cR$\fi\ifmmode {\cal K} \else$\cK$\fi)\sdprod \ifmmode {\cal Z} \else$\cZ$\fi V, \quad \ifmmode {\cal R} \else$\cR$\fi\ifmmode\ \cong\ \else$\isom$\fi \fR^{\times}, \quad \ifmmode {\cal K} \else$\cK$\fi=SU(2)\times U(1)$ \hspace*{\fill} \hspace*{\fill} $\ifmmode {\cal Z} \else$\cZ$\fi= \fR, \quad V=\komp^2 $\hspace*{\fill} \end{minipage} \end{equation} For the $\rat$-form of $P$, it follows that $V_{\rat}\ifmmode\ \cong\ \else$\isom$\fi L^2$ for some imaginary quadratic field $L$, and for the arithmetic parabolic $\gG_P\subset} \def\nni{\supset} \def\und{\underline P,\ \gG_P\cap V_{\rat}\subset} \def\nni{\supset} \def\und{\underline (\ifmmode {\cal O} \else$\cO$\fi_L)^2$ is some lattice. Furthermore, the theory of toroidal embeddings shows that a compactification divisor of $\ifmmode {\overline{X}_{\gG}} \else$\xgc$\fi$ is of the form $\komp^2/(\gG_P\cap V_{\rat})$, which has complex multiplication by $L$, so its isomorphism class determines $L$, which was to be shown. \hfill $\Box$ \vskip0.25cm Now an easy calculation shows what the compactification divisors on $\tilde{Y}$ are. Namely, these are the irreducible components of the inverse image in $\tilde{Y}$ of the exceptional divisors $L_{0ij}\subset} \def\nni{\supset} \def\und{\underline \tilde{{\Bbb P}}^3,\ L_{0ij}\ifmmode\ \cong\ \else$\isom$\fi {\Bbb P}^1\times {\Bbb P}^1$. The local geometry of the arrangement shows the branch locus in $L_{0ij}$ is of the form $p_1^*(\ifmmode {\cal O} \else$\cO$\fi(3)) \otimes p_2^*(\ifmmode {\cal O} \else$\cO$\fi(3))$, i.e., of the form $\{0\}\times {\Bbb P}^1,\ \{1\}\times {\Bbb P}^1,\ \{\infty\}\times {\Bbb P}^1$ and ${\Bbb P}^1\times \{0\},\ {\Bbb P}^1\times \{1\},\ {\Bbb P}^1\times \{\infty\}$. It is well-known that the elliptic curve $E\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$, branched at $(0,1,\infty)$ to degree 3, with Galois group $\integer/3\integer$, is the elliptic curve $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}=\komp/\integer\oplus \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta\integer,\ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=e^{2\pi i / 3}$. From this it follows \begin{Lemma}\label{l115.2} The compactification divisors $\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi_i$ of $\tilde{Y}$ are products $$ \Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi_i\ifmmode\ \cong\ \else$\isom$\fi E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta},$$ where $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}$ is the unique elliptic curve with $\integer/6\integer$ as automorphism group, i.e., $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}=\komp/\integer\oplus \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta\integer=\{x^3+y^3+z^3=0\}$ and $\hbox{Aut}} \def\Im{\hbox{Im}(E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta})=<\pm 1,\pm\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta, \pm\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^2>$.\hfill $\Box$ \vskip0.25cm \end{Lemma} Note that the morphism $\tilde{Y}\longrightarrow} \def\sura{\twoheadrightarrow Y^{\wedge}$ blows down the $\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi_i$ to singular points (just as $\tilde{{\Bbb P}}^3\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal S} \else$\cS$\fi_3$ blows down the $L_{0ij}$ to the nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$) which lie over the nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$. From this and the well-known fact that the Satake compactification of a ball quotient has only isolated, zero-dimensional singularities, which are resolved in a torus embedding by means of complex tori or quotients thereof, we get the following \begin{Corollary}\label{c115.1} The variety $Y^{\wedge}$ is the Satake compactification of the quotient $\gG'\backslash \ball_3$, with $\gG'\subset} \def\nni{\supset} \def\und{\underline G_{\rat}$ and $G_{\rat}$ isogenous to $U(3,1;K)$, $K$ the field of Eisenstein numbers as above. \hfill $\Box$ \vskip0.25cm \end{Corollary} Now that it is established that $\gG'$ is (isogenous to) an arithmetic subgroup of $U(3,1;K)$, group-theoretic methods can be applied to determine the arithmetic subgroup. This is done in detail in \cite{J}, Lemma 2.9 and Theorem 2.11. The result is: $\gG'=P\gG(3)$, and the group $\gG_{\ifmmode {\cal S} \else$\cS$\fi_3}$ giving rise to the Segre cubic is $\gG_{\ifmmode {\cal S} \else$\cS$\fi_3}=P\gG(\sqrt{-3})$. This yields the statements of the theorem on the arithmetic groups. The compactification divisors of $\tilde{Y}$ coincide by Lemma \ref{l115.2} with those of $\-X_{\gG'}$, and these are blown down under $\tilde{Y}\longrightarrow} \def\sura{\twoheadrightarrow Y^{\wedge}$ to the singularities on the Satake compactification which is $Y^{\wedge}$, $Y^{\wedge}\ifmmode\ \cong\ \else$\isom$\fi X_{\gG'}^*$. The cover $\tilde{Y}\longrightarrow} \def\sura{\twoheadrightarrow \tilde{{\Bbb P}}^3$ (respectively $Y^{\wedge}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal S} \else$\cS$\fi_3$) is now readily identified with $\-X(3)\longrightarrow} \def\sura{\twoheadrightarrow \-X(\sqrt{-3})$ (respectively with $X(3)^*\longrightarrow} \def\sura{\twoheadrightarrow X(\sqrt{-3})^*$) of (\ref{e114.2}), from the fact that the branching loci, degrees and group actions coincide. Details can be found in \cite{J}. \hfill $\Box$ \vskip0.25cm \subsection{Moduli interpretation} Now applying Shimura's theory we get the following moduli description of $\ifmmode {\cal S} \else$\cS$\fi_3$ (see \cite{J}, \S2 for details). \begin{theorem}\label{t116.1} Any point $x\in \ifmmode {\cal S} \else$\cS$\fi_3-\{\hbox{ten nodes}\}$ determines a unique isomorphism class of principally polarised abelian fourfolds with complex multiplication by $K=\rat(\sqrt{-3})$ and a level $\sqrt{-3}$ structure. The signature of the complex multiplication is (3,1). Any point $x\in Y^{\wedge}-\{\hbox{inverse image under $\phi$ of (\ref{e115.0}) of the ten nodes}\}$ determines a unique isomorphy class of abelian fourfolds as above with a level 3 structure. \end{theorem} Moreover, the moduli interpretation of the 15 ${\Bbb P}^2$'s on $\ifmmode {\cal S} \else$\cS$\fi_3$ is given in \cite{J}. \begin{proposition}\label{p116.1} The 15 ${\Bbb P}^2$'s on $\ifmmode {\cal S} \else$\cS$\fi_3$ are compactifications of two-dimensional ball quotients which are moduli spaces of those abelian fourfolds $A_x^4$ as above which split: $$A_x^4\ifmmode\ \cong\ \else$\isom$\fi A_x^3\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}.$$ The intersections of the 15 planes determine moduli points of $A_x^4$ which further decompose, i.e., $A_x^3$ splits. \end{proposition} \begin{remark} It is natural to ask whether, given a point $x\in \ifmmode {\cal S} \else$\cS$\fi_3$, one can give the equations defining the abelian variety $A_x$ occuring in Theorem \ref{t116.1}. In some sense one can. First it turns out the $A_x$ is the Jacobian of an algebraic curve, as described by the hypergeometric equation as in equation (\ref{e111b.5}). Since the parameters are by Proposition \ref{p111d.1} the set 1) in (\ref{e111d.1}), these curves have the form: \begin{equation}\label{e116.1} C_{\tau}=\{y^3=\prod_{i=1}^6(x-t_i(\tau))\}; \end{equation} $C_{\tau}$ obviously has an automorphism of order 3, given by $y\mapsto \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta y$ with the third root of unity $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$. This yields an automorphism of the Jacobian of $C_{\tau}$. Without much difficulty one finds \begin{itemize}\item[(i)] $\Jac(C_{\tau})=A_{\tau}$ has complex multiplication by $\ifmmode {\cal O} \else$\cO$\fi_K$, the signature is (3,1). \item[(ii)] The automorphism group is $\ifmmode {\cal O} \else$\cO$\fi_K^*$, and is given by multiplication by $\pm\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$ in $\ifmmode {\cal O} \else$\cO$\fi_K$. \end{itemize} The most direct way to see this is to write down the Jacobian of the curve (\ref{e116.1}) and show that its periods have the complex multiplication. A basis of the (1,0) differentials on $C_{\tau}$ written in the normal form \begin{equation}\label{e116a.1} y^3=x(x-1)(x-t_1)(x-t_2)(x-t_3) \end{equation} is given by \begin{equation}\label{e116a.3} \int{dx \over \sqrt[3]{x(x-1)(x-t_1)(x-t_2)(x-t_3)}}; \end{equation} choosing a base of $H_1(C_{\tau},\integer)$ and taking the integrals over the elements of that base gives the Jacobian; the multiplication by $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$ is then evident. Hence one may invoke Shimura's theory to conclude: \begin{lemma}\label{l116a.1} The isomorphism classes of the Jacobians of the curves (\ref{e116a.1}) are given as the points of the arithmetic quotient $PU(3,1;\ifmmode {\cal O} \else$\cO$\fi_K)\backslash \ball_3$. Putting a $\sqrt{-3}$ level structure on the Jacobians yields the moduli space $\gG(\sqrt{-3})\backslash \ball_3$. \end{lemma} The latter space has already been identified with the open subset of smooth points on $\ifmmode {\cal S} \else$\cS$\fi_3$. The precise relation between the moduli {\em point} $\tau\in \ball_3$ and the {\em values} of the $t_i$ has been derived for surfaces, i.e., for $\tau$ in one of the subballs covering one of the 15 ${\Bbb P}^2$'s on $\ifmmode {\cal S} \else$\cS$\fi_3$, by Holzapfel. The result is: there are automorphic forms $G_2,G_3$ and $G_4$ of indicated weights on $\ball_2$ such that \begin{equation}\label{e116a.2} C_{\tau}=\{y^3=x^4-G_2(\tau)x^2-G_3(\tau)x-G_4(\tau)\}, \end{equation} much akin to the Weierstra\ss\ equation for an elliptic curve. (The variable $x$ in (\ref{e116a.2}) is of course different than that in (\ref{e116a.1})). There is no doubt a similar expression for $\tau\in \ball_3$. \end{remark} \section{The Igusa quartic ${\cal I}_4$} This variety has been known since the last century, and it is related to the configuration in ${\Bbb P}^4$ which is dual to the 15 hyperplanes of (\ref{e112b.3}) and the 15 planes of (\ref{e112.1}) which they cut out on $\ifmmode {\cal S} \else$\cS$\fi_3$, and in fact $\ifmmode {\cal I} \else$\cI$\fi_4$ is just the dual variety of $\ifmmode {\cal S} \else$\cS$\fi_3$. It was also known in the last century that the tangent hyperplane sections of $\ifmmode {\cal I} \else$\cI$\fi_4$ are Kummer surfaces, giving $\ifmmode {\cal I} \else$\cI$\fi_4$ a moduli interpretation. Igusa, in the 1960's, made this rigorous and showed that $\ifmmode {\cal I} \else$\cI$\fi_4$ is the Satake compactification of $\gG(2)\backslash \sieg_2$, the Siegel modular threefold of level 2. We begin by discussing the projective variety, then turn to Igusa's results. \subsection{The quartic locus associated to a configuration of 15 lines} Let $l_{\gs}$ be the line dual in ${\Bbb P}^4$ to the ${\Bbb P}^2$ of (\ref{e112.1}), and let $h_{ij}$ denote the point dual to $\ifmmode {\cal H} \else$\cH$\fi_{ij}$ of (\ref{e112b.3}). Then these 15 lines meet at the 15 points $h_{ij}$, and three of the 15 lines meet at each, corresponding to the three ${\Bbb P}^2$'s which are contained in each $\ifmmode {\cal H} \else$\cH$\fi_{ij}$. Furthermore, each of the 15 lines contains three of the 15 points, as each ${\Bbb P}^2$ is contained in three of the $\ifmmode {\cal H} \else$\cH$\fi_{ij}$. It is useful to introduce the following notation: each line is given a notation $(ij)$, and two such lines $(ij),\ (kl)$ meet if and only if the sets $(ij),\ (kl)$ are disjoint. Hence the 15 points are numbered by {\em synthemes} $(ij,kl,mn)$ and the three lines meeting each point are the indicated {\it duads} (pairs) $(ij),\ (kl),\ (mn)$. Then there are ten sets such as 23, 31, 12 and 56, 64, 45 with the property that the first and last three do not meet, but each of the first meets each of the last. Therefore the six lines are generators of a quadric surface \begin{equation}\label{e117a.1} Q_{ijk}=\parbox{12cm}{quadric with $(ij), (jk), (ik)$ in one ruling and $(lm), (mn), (ln)$ in the other ruling} \end{equation} Then $Q_{ijk}$ lies in a ${\Bbb P}^3$, and there are ten such, corresponding to the ways of dividing the six numbers into two {\it triads} (triples). Let us denote the corresponding ${\Bbb P}^3$ by $K_{ijk}$, so \begin{equation}\label{e117a.1a} Q_{ijk}\subset} \def\nni{\supset} \def\und{\underline K_{ijk}. \end{equation} Then each of the 15 lines is contained in four of the $K_{ijk}$, and six of the $K_{ijk}$ meet at each of the 15 points. Consider now a set of four mutually skew of the 15 lines, for example 12, 23, 24, 25. Then there will be a two-dimensional space of ${\Bbb P}^2$'s which meet all four lines (as we are in ${\Bbb P}^4$, generically a plane and a line will not intersect). Of all of these planes, there are exactly two passing through a given point of space $x\in {\Bbb P}^5$. The locus we are interested in is: \begin{equation}\label{e117a.2} \ifmmode {\cal Q} \else$\cQ$\fi:=\left\{x\in {\Bbb P}^5 \left| \parbox{6cm}{ the two planes meeting four skew lines of the 15 $(ij)$ and passing through $x$ {\em coincide}}\right.\right\}. \end{equation} If, as in (\ref{e113.1}), we take coordinates $\xi,\eta,\gz,\xi',\eta',\gz'$ satisfying $\xi+\eta+\gz=\xi'+\eta'+\gz'$ as coordinates on ${\Bbb P}^4$, then the condition (\ref{e117a.2}) yields a locus with equation (\cite{Baker}, p.~125): \begin{equation}\label{e117a.3} \sqrt{(\eta-\gz')(\eta'-\gz)} + \sqrt{(\gz-\xi')(\gz'-\xi)} + \sqrt{(\xi-\eta')(\xi'-\eta)}=0. \end{equation} To find the dual variety of the locus $\ifmmode {\cal Q} \else$\cQ$\fi$, Baker does the following. Letting $a,b,c$ be variables, $a'=(1-a),\ b'=(1-b),\ c'=(1-c)$, consider the six points which are the vertices of a coordinate simplex in ${\Bbb P}^5$, and call them $A, B, C, A', B', C'$. Then any point of our ${\Bbb P}^4$ can be written as $x=A/bc'+B/ca'+ C/ab'+ A'/b'c+ B'/c'a+ C'/a'b$. Calculating the tangent plane of $\ifmmode {\cal Q} \else$\cQ$\fi$ at a point $x\in \ifmmode {\cal Q} \else$\cQ$\fi$ which satisfies (\ref{e117a.3}), in terms of the coordinates used in (\ref{e117a.3}), one gets: \begin{equation}\label{e117a.4} bc'\xi+ca'\eta+ab'\gz-b'c\xi'-c'a\eta'-a'b\gz'=0. \end{equation} Now putting $u=bc', v=ca', w=ab', u'=-b'c, v'=-c'a, w'=-a'b$, the equation becomes \begin{equation}\label{e117a.5} u\xi+v\eta+w\gz+u'\xi'+v'\eta'+w'\gz'=0, \end{equation} with the two identities \begin{equation}\label{e117a.6} u+v+w+u'+v'+w'=0,\quad uvw+u'v'w'=0. \end{equation} Since the identities (\ref{e117a.6}) do not depend on the point, it follows that these equations define the dual variety. Now comparing with (\ref{e113.1}), we have \begin{proposition}\label{p117a.1} The dual variety of the quartic locus $\ifmmode {\cal Q} \else$\cQ$\fi$ is the Segre cubic $\ifmmode {\cal S} \else$\cS$\fi_3$. \end{proposition} It is easy to see that $\ifmmode {\cal Q} \else$\cQ$\fi$ is singular along the 15 lines. It was also noted classically that a tangent hyperplane section of $\ifmmode {\cal Q} \else$\cQ$\fi$ is a Kummer quartic surface, with 16 nodes, 15 from the intersections with the 15 singular lines, and one from the point of tangency. \subsection{Igusa's results} The relation to the Kummer quartic surfaces is correctly understood by studying theta constants for the theta functions with 1/2-characteristics. This was done by Igusa in \cite{igusa}, and we now recall some of his results. \subsubsection{Theta functions} Let $\tau\in \sieg_g=\{M\in M_g(\komp)\big| \tau={^t(\tau)},\, \Im(\tau) \hbox{ positive definite}\}$, $z\in \komp^g$, and $m=(m',m'')\in \rat^{2g}$. Note that $\sieg_g$ is a hermitian symmetric space of type $\bf III_g$. \begin{definition}\label{d117.1} The {\em theta function of degree $g$ and characteristic $m$} is defined by the power series $$\gt_m(\tau,z)=\sum_{n\in \integer^g}\exp\left({1\over 2}{^t(n+m')}\tau(n+m') + {^t(n+m')}(z+m'')\right).$$ \end{definition} As a function of $\tau$ the series $\gt_m$ converges precisely for $\tau\in \sieg_g$, while as functions of $z$ by fixed $\tau$ these are theta functions on $A_{\tau}=\komp^g/(\integer^g+\tau\integer^g)$. As such the zeros on $A_{\tau}$ are determined by the characteristic $m$. The corresponding {\em theta constant} is \begin{equation}\label{e117.2} \gt_m(\tau):=\gt_m(\tau,0). \end{equation} Igusa has studied in \cite{igusa} these theta constants, in particular the theta functions with characteristics $m\in {1\over 2^n}\integer$. Some of his results are the following. \begin{lemma}\label{l3.1.2} $\gt_m(\tau)\equiv 0 \iff m\hbox{\em mod}(1)$ satisfies $\exp(4\pi i{(^tm')}m'')=-1.$ \end{lemma} The Siegel modular group $\gG_g(1)=Sp(2g,\integer)$ acts on the arguments $(\tau, z)$ as follows: \begin{equation}\label{e117.3} M=\left(\begin{array}{cc}A & B \\ C & D \end{array}\right),\ M(\tau,z)=\left((A\tau+B)(C\tau+D)^{-1}, (C\tau + D)^{-1}z\right), \end{equation} and on the characteristic itself by \begin{equation}\label{e117.4} M(m)=\left(\begin{array}{cc}D & -C \\ -B & A \end{array}\right)\cdot m +{1 \over 2} \left(\begin{array}{c} \diag(C^tD) \\ \diag(A^tB)\end{array} \right). \end{equation} The behavior of the theta functions under $M$ is given by \begin{lemma}\label{l117.1} Let $M\in \gG_g(1)$ act on $(\tau,z)$ as in (\ref{e117.3}) and on the characteristic $m$ as in (\ref{e117.4}). Then the theta functions transform according to the rule: \begin{equation} \gt_{M(m)}(M(\tau,z)) = \gk(M)\exp(2\pi i \phi_m(M))\hbox{det}} \def\Ker{\hbox{Ker}(C\tau+D)^{1/2} \times \exp(\pi i {^tz}(C\tau +D)^{-1}Cz)\gt_m(\tau,z), \end{equation} where $\gk(M)$ is some eighth root of unity and $\phi_m(M)$ is defined by the formula $$\phi_m(M)=-{1\over2}{^tm'}BDm' + {^tm''}{^tA}Cm'' - 2{^tm'}{^tB}Cm'' -{^t\diag}(A^tB)(Dm'-Cm'').$$ \end{lemma} In particular for the theta constants the formula becomes \begin{equation}\label{e117.5} \gt_{M(m)}(M\tau)=\gk(M)\exp(2\pi i \phi_m(M))\hbox{det}} \def\Ker{\hbox{Ker}(C\tau+D)^{1/2}\gt_m(\tau). \end{equation} What the equation (\ref{e117.5}) says for $g=2$ is that up to an eighth root of unity, non-vanishing theta constants with 1/2-characteristics are automorphic forms of weight 1/2 for the main congruence subgroup of level 2 in $Sp(4,\integer)$. Indeed, for $M\in \gG(2)$, it holds that $e^{2\pi i\phi_m(M)}=1$, as Igusa shows. There are 16 characterstics $m$; six are {\em odd} (i.e., $\gt_m(\tau,z)=-\gt_m(\tau,-z)$) so give rise to vanishing theta constants, while ten are even. The fourth powers $\gt_m^4$ are genuine automorphic forms for $\gG(2)$, and determine a morphism \begin{equation}\label{e118.2} f:\gG(2)\backslash \sieg_2 \longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^9=(\gt_{m_1}^4: \cdots :\gt_{m_{10}}^4), \end{equation} where $m_1,\ldots,m_{10}$ are the ten even characteristics. \subsubsection{The ring of automorphic forms} Among the ten coordinate theta functions there are five linear relations, the Riemann relations. This implies that the map $f$ in (\ref{e118.2}) maps into a ${\Bbb P}^4$, displaying the quotient $X_{\gG(2)}$ as a hypersurface. In fact, since this is an embedding by means of automorphic functions whose closure $X_{\gG(2)}^*\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^4$ is normal (see below), it follows that $f$ gives a Baily-Borel embedding of the arithmetic quotient. The proof that $f$ is an embedding given by Igusa is quite deep, involving showing that the ring of modular forms of $\gG(2)$ is the integral closure of the ring generated by the said theta functions. More precisely, his result is \begin{theorem}[\cite{igusa},p.~397]\label{t118.1} Take as coordinates in ${\Bbb P}^4$ the following theta constants: $$y_0=\gt^4_{(0110)}(\tau),\ y_1=\gt^4_{(0100)}(\tau),\ y_2=\gt^4_{(0000)}(\tau),$$ $$y_3=\gt^4_{(1000)}(\tau)-\gt^4_{(0000)}(\tau),\ y_4=-\gt^4_{(1100)}(\tau)-\gt^4_{(0000)}(\tau),$$ where we let $(ijkl)$ denote the characteristic $({i\over 2}{j\over 2}{k\over 2}{l\over 2})$. Set also $$\chi_{10}=\prod_{\hbox{even $m$}}\gt_m^2.$$ Then the ring of modular forms of $\gG(2)$ is given by: $$R(\gG(2))=\komp[y_0,\ldots,y_4,\chi_{10}]/\ifmmode {\cal E} \else$\cE$\fi,$$ where $\ifmmode {\cal E} \else$\cE$\fi$ is the ideal generated by the following two relations: \begin{minipage}{14cm}$\ifmmode {\cal E} \else$\cE$\fi=\left\{\begin{array}{rcl} R_1 & = & (y_0y_1+y_0y_2+y_1y_2-y_3y_4)^2-4y_0y_1y_2(\sum y_i) \\ R_2 & = & \chi_{10}^2-{1\over 4}s(y_0,\ldots,y_4),\ s \hbox{ homogenous of degree 5} \end{array}\right.$ \end{minipage} \end{theorem} However, the formula $R_1$ relating the theta functions was known long before Igusa. Since the five linear relations determining the image ${\Bbb P}^4$ of $f$ are known, it is sufficient to give a single relation of minimal degree among the $\gt^4_m$ to determine the image. This relation can be found as early as in the 1887 paper of Maschke \cite{maschke}, p.~505\footnote{the equation is somewhat hidden: ``...da\ss\ dagegen die symmetrische Function vierter Dimension sich bis auf einen Zahlenfactor als das Quadrat der zweiten Dimension erweist.''}. In terms of the theta constants above, this equation is \begin{equation}\label{e118.1} \left(\sum\gt^8_m\right)^2-4\left(\sum \gt_m^{16}\right)=0, \end{equation} which, as can be checked, is the same quartic as that given by $R_1$ in \ref{t118.1}, as well as that given by (\ref{e117a.3}). \begin{definition}\label{d118.1} The {\em Igusa quartic} $\ifmmode {\cal I} \else$\cI$\fi_4$ is the quartic threefold defined in ${\Bbb P}^4$ by the relation $R_1$ of Theorem \ref{t118.1} or the equation (\ref{e118.1}). \end{definition} As a corollary we have \begin{corollary}\label{c118.1} The Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$ and the quartic locus $\ifmmode {\cal Q} \else$\cQ$\fi$ of (\ref{e117a.3}) coincide, and this quartic is the Satake compactification of $X_{\gG(2)}$. \end{corollary} Hence we have described $X_{\gG(2)}^*$ as a singular quartic hypersurface in ${\Bbb P}^4$. There are the two interesting loci: \begin{itemize}\item[(i)] the singular locus, which is the boundary of the Baily-Borel embedding of $X_{\gG(2)}$; \item[(ii)] the intersection of $\ifmmode {\cal I} \else$\cI$\fi_4$ with the coordinate hyperplanes in ${\Bbb P}^9$, which are the modular subvarieties $\-Y_m(2)$ of \cite{J}, Thm. 3.19; these are quotients of symmetric subdomains isomorphic to a product of discs. \end{itemize} As already mentioned, the singular locus of $\ifmmode {\cal I} \else$\cI$\fi_4$ consists of 15 lines; this can be directly calculated from the equation. Alternatively, applying general formula for the number of cusps (see for example \cite{Yam}) we see that $X_{\gG(2)}$ has 15 one-dimensional boundary components and 15 zero-dimensional boundary components; by \ref{c118.1} this is then the singular locus of $\ifmmode {\cal I} \else$\cI$\fi_4$. (That these boundary components are rational curves is obvious ($\gG(2)\backslash \sieg_1$ is rational); that they are actually {\em lines} is not so obvious, but an easy calculation). This line of reasoning also requires the result, also due to Igusa, that, although $\gG(2)$ is not torsion-free, there are nonetheless no singularities on $X_{\gG(2)}$. \subsection{Moduli interpretation} The embedding (\ref{e118.2}) of $X_{\gG(2)}^*$ as the quartic $\ifmmode {\cal I} \else$\cI$\fi_4$ shows that $\ifmmode {\cal I} \else$\cI$\fi_4$ has a moduli interpretation. In fact, $X_{\gG(2)}$ is a rough moduli space of principally polarised abelian surfaces with a level 2 structure. However, $\gG(2)$ contains torsion, namely the element $-1$, so $X_{\gG(2)}$ is {\em not} a fine moduli variety. This corresponds to the fact that the automorphism $z\mapsto -z$ of $A_{\tau}$ {\em preserves} the level 2 structure, hence the actual {\em object} which is parameterised by $X_{\gG(2)}$ is the {\em quotient} $A_{\tau}/(z\mapsto -z)$. This is just the Kummer quartic surface which already occured above. The precise relation is given by \begin{theorem}\label{t119.1} For a point $x\in \ifmmode {\cal I} \else$\cI$\fi_4-\{\hbox{intersections of $\ifmmode {\cal I} \else$\cI$\fi_4$ with the ten coordinate planes in (\ref{e118.2})}\}$, the corresponding Kummer quartic surface $K_x=A_{\tau}/\{\pm1\}$, where $x=p(\tau)$ for the natural projection $p:\sieg_2\longrightarrow} \def\sura{\twoheadrightarrow X_{\gG(2)}$, is the intersection of $\ifmmode {\cal I} \else$\cI$\fi_4$ with the tangent hyperplane at $x$, $T_x\ifmmode {\cal I} \else$\cI$\fi_4$: $$K_x=\ifmmode {\cal I} \else$\cI$\fi_4\cap T_x\ifmmode {\cal I} \else$\cI$\fi_4.$$ \end{theorem} This statement can be found for example in \cite{Baker}. It amounts to the fact, true in any dimension, that for $n\geq3$ the theta functions with characteristics $\in \integer/n\integer$ on a fixed $A_{\tau}$ give an embedding of $A_{\tau}$, while for $n=2$ they map onto the Kummer variety. The reason one must exclude the ten hyperplane sections in Theorem \ref{t119.1} is the following result. \begin{proposition}\label{p120.1} The ten hyperplane sections $\{\gt_m^4=0\} \cap \ifmmode {\cal I} \else$\cI$\fi_4$ are tangent hyperplane sections, i.e., the intersection is of degree 2 and multiplicity 2. \end{proposition} A proof, based only on the equation of $\ifmmode {\cal I} \else$\cI$\fi_4$, can be found in \cite{Baker}. To understand the meaning of this, note that a general hyperplane section meets $\ifmmode {\cal I} \else$\cI$\fi_4$ in a quartic surface, while the intersections here are quadric surfaces, hence to preserve degree must be counted twice (i.e., multiplicity 2). Consider the symmetric subdomain $\sieg_1\times \sieg_1\subset} \def\nni{\supset} \def\und{\underline \sieg_2$, which in this case is the set of reducible matrices: $$\sieg_1\times \sieg_1 = \left\{\left(\begin{array}{cc} \tau_1 & 0 \\ 0 & \tau_2\end{array} \right)\right\}\subset} \def\nni{\supset} \def\und{\underline \left\{\left(\begin{array}{cc} \tau_1 & \tau_{12} \\ \tau_{12} & \tau_2 \end{array}\right)\right\} =\sieg_2.$$ Then an easy calculation shows that the theta function of Definition \ref{d117.1} is a {\em product} of two theta functions of a single variable (i.e., $z\in \komp$). This is equivalent to the fact that for reducible $\tau \in \sieg_2$, the abelian surface $A_{\tau}$ is a product of two elliptic curves, $A_{\tau}=E_1\times E_2$. In this case, the map given onto the ``product Kummer'' variety is a map $s:E_1\times E_2\longrightarrow} \def\sura{\twoheadrightarrow E_1/\{\pm1\}\times E_2/\{\pm1\}={\Bbb P}^1\times {\Bbb P}^1$, and this ${\Bbb P}^1\times {\Bbb P}^1$ is the quadric surface occuring in \ref{p120.1}. Since ${\Bbb P}^1\times {\Bbb P}^1$ has no moduli, we see that {\em formally} the statement of Theorem \ref{t119.1} remains true for all $x\in \ifmmode {\cal I} \else$\cI$\fi_4-\{\hbox{15 singular lines}\}$, if we consider product Kummer varieties instead of the usual ones, and the hyperplane section is the quadric surface of Proposition \ref{p120.1}. Note however, that this quadric surface, being a modular subvariety, can also be described as: \begin{equation}\label{e120.1} E_1/\{\pm1\}\times E_2/\{\pm1\}\ifmmode\ \cong\ \else$\isom$\fi (\gG_1(2)\backslash \sieg_1)^*\times (\gG_1(2)\backslash \sieg_1)^*, \end{equation} describing the product Kummer surface of a reducible abelian surface as a compactification of an arithmetic quotient, that is, as a Janus-like variety. We then get the following moduli interpretation of the quadric surfaces. \begin{proposition}\label{p120.2} The ten quadric surfaces of Proposition \ref{p120.1} are modular subvarieties which correspond to abelian surfaces which split. More precisely, for any $x$ on one of the quadric surfaces, but not on any of the singular lines (there are six such singular lines on each quadric surface, see (\ref{e117a.1})), determines a smooth abelian surface which splits, with a level 2 structure. \end{proposition} Finally we note that this geometry can be described, as discussed already in \cite{J} and many other places, in terms of the finite geometry of \begin{equation}\label{e121.1} V=(\integer/2\integer)^4. \end{equation} Let $<\ ,\ >$ denote the induced symplectic form on $V$; every vector $v\in V$ is isotropic with respect to $<\ ,\ >$. Since there are 15 non-zero vectors, there are 15 one-dimensional boundary components. Similarly, there are 15 isotropic planes in $V$, giving 15 zero-dimensional boundary components. The modular subvarieties of Proposition \ref{p120.2} correspond in this setting to non-singular pairs $\{\gd,\gd^{\perp}\}$, where $\gd$ is a two-dimensional subspace of $V$ on which $<\ ,\ >$ is non degenerate, and $\gd^{\perp}$ denotes the orthocomplement with respect to $<\ ,\ >$. Of these there are exactly ten, as is easily checked. We leave further details to the reader. \subsection{Birational transformations}\label{section3.4} We have seen above in Proposition \ref{p117a.1} that $\ifmmode {\cal S} \else$\cS$\fi_3$ and $\ifmmode {\cal I} \else$\cI$\fi_4$ are dual varieties. It follows from general theory that they are then in fact {\em birational}. In this section we describe the ensuing birational map explicitly. We consider the following modifications of ${\Bbb P}^4$. \begin{itemize}\item[a)] Blow up the ten nodes (\ref{e112.0}) of $\ifmmode {\cal S} \else$\cS$\fi_3$; denote this by $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^4\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$. There are ten exceptional ${\Bbb P}^3$'s, each with normal bundle $\ifmmode {\cal O} \else$\cO$\fi_{{\Bbb P}^3}(-1)$. Consider one of the 15 hyperplanes $\ifmmode {\cal H} \else$\cH$\fi_{ij}$ of (\ref{e112b.3}). Since each hyperplane contains $4+2+1=7$ nodes, its proper transform on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^4$ is a ${\Bbb P}^3$ blown up in those seven points; each of the 15 ${\Bbb P}^2$'s of (\ref{e112.1}) lying on $\ifmmode {\cal S} \else$\cS$\fi_3$ contains four of the nodes, so their proper transforms are copies of ${\Bbb P}^2$ blown up in four points. Finally, let $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3\subset} \def\nni{\supset} \def\und{\underline \widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^4$ denote the proper transform of $\ifmmode {\cal S} \else$\cS$\fi_3$ in $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^4$; $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3$ is smooth, and ${\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1}_{|\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3}:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal S} \else$\cS$\fi_3$ is a desingularisation of $\ifmmode {\cal S} \else$\cS$\fi_3$, replacing each node with a quadric surface $\ifmmode\ \cong\ \else$\isom$\fi {\Bbb P}^1\times {\Bbb P}^1$. \item[b)] Let $\ifmmode \hbox{{\script I}} \else$\scI$\fi(15)$ denote the ideal of the 15 singular lines of $\ifmmode {\cal I} \else$\cI$\fi_4$; blow up $\ifmmode \hbox{{\script I}} \else$\scI$\fi(15)$, and let $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}}^4\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ denote this modification. Under $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$, each of the lines is replaced by a ${\Bbb P}^2$-bundle over that line, and each point is replaced by a union of ${\Bbb P}^1$'s, one each for each {\em pair} $(l_1,l_2)$ of {\em lines} meeting at the point; this mentioned ${\Bbb P}^1$ is then the intersection of the fibre ${\Bbb P}^2$ of $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ at that point with the (two) exceptional ${\Bbb P}^2$-bundles over the lines $l_1$ and $l_2$. Note that the proper transforms of the ten quadrics of Proposition \ref{p120.1} on $\ifmmode {\cal I} \else$\cI$\fi_4$ are still biregular to ${\Bbb P}^1\times {\Bbb P}^1$, while the proper transforms of each of the lines turns out to be a {\em Kummer modular surface}, that is, ${\Bbb P}^2$ blown up in four points. Let $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4$ be the proper transform of $\ifmmode {\cal I} \else$\cI$\fi_4$ in $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}}^4$; then ${\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2}_{|\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4}:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_4$ is a desingularisation of $\ifmmode {\cal I} \else$\cI$\fi_4$. \end{itemize} \begin{theorem}\label{t123.1} The varieties $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3$ and $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4$ are biregular, and the explicit birational map $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha:\ifmmode {\cal S} \else$\cS$\fi_3- - \ra \ifmmode {\cal I} \else$\cI$\fi_4$ is the birational morphism completing the following diagram: $$\begin{array}{ccccc} & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3 & \stackrel{\tilde{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha}}{\longrightarrow} \def\sura{\twoheadrightarrow} & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4 & \\ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1 & \downarrow & & \downarrow & \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2 \\ & \ifmmode {\cal S} \else$\cS$\fi_3 & \stackrel{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha}{- - \ra} & \ifmmode {\cal I} \else$\cI$\fi_4. & \end{array}$$ Moreover, $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$ is $\gS_6$-equivariant. \end{theorem} {\bf Proof:} As $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$ and $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ are $\gS_6$-equivariant, the second statement follows from the first. Let $D\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal S} \else$\cS$\fi_3$ be the open set: \begin{equation}\label{e123.1} D=\ifmmode {\cal S} \else$\cS$\fi_3-\{\hbox{15 hyperplanes $P_{\gs}$ of (\ref{e112.1})}\}; \end{equation} here we may take the regular map of $D$ onto the set of tangent hyperplanes (now viewing $\ifmmode {\cal I} \else$\cI$\fi_4$ as the projective dual of $\ifmmode {\cal S} \else$\cS$\fi_3$), and set \begin{eqnarray}\label{e123.2} \varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha_{|D}:D & \longrightarrow} \def\sura{\twoheadrightarrow & D'\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal I} \else$\cI$\fi_4 \\ x & \mapsto & ({\Bbb P}^3)_x=\hbox{tangent hyperplane to $\ifmmode {\cal S} \else$\cS$\fi_3$ at $x$} \nonumber \end{eqnarray} \begin{Lemma} The subset $D'\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal I} \else$\cI$\fi_4$ is: $D'=\ifmmode {\cal I} \else$\cI$\fi_4-\{\hbox{10 quadric surfaces of Proposition \ref{p120.2}}\}$. \end{Lemma} {\bf Proof:} Suppose $x\in D$; then $({\Bbb P}^3)_x$ meets $D$ in an irreducible cubic (the union of the $P_{\gs}$ are {\em all} the linear subspaces contained in $\ifmmode {\cal S} \else$\cS$\fi_3$, so outside of this locus $({\Bbb P}^3)_x\cap \ifmmode {\cal S} \else$\cS$\fi_3$ cannot have a linear factor, so, being cubic, must be irreducible), while the ten quadric surfaces are the locus of the tangent hyperplanes meeting $\ifmmode {\cal S} \else$\cS$\fi_3$ in one of the nodes, all of which are excluded in $D$. \hfill $\Box$ \vskip0.25cm Now we glue $D$ onto the rest of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3$, and $D'$ onto the rest of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4$. The locus $\gL_1=\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3-D$ coincides with $\gL_2=\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4-D'$, as follows from the descriptions of the rational maps $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$ and $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ above. Both the $\gL_i$ consist of ten ${\Bbb P}^1\times {\Bbb P}^1$'s and 15 rational surfaces, each isomorphic to ${\Bbb P}^2$ blown up in four points. Hence we can complete $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha_{|D}$ to a biregular isomorphism $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi}_3\longrightarrow} \def\sura{\twoheadrightarrow \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4$, by fixing an isomorphism $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha_{\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi}:\gL_1\longrightarrow} \def\sura{\twoheadrightarrow \gL_2$, and setting $$\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha(x)=\left\{\parbox{6cm}{$\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha_{|D}(x),$ if $x\in D$ \\ $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha_{\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi}(x)$, if $x\in \gL_1$}\right.$$ completing the proof of Theorem \ref{t123.1}. \hfill $\Box$ \vskip0.25cm The following description is more concrete. If $x$ is one of the nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$, there is a quadric cone of tangent (to $\ifmmode {\cal S} \else$\cS$\fi_3$) hyperplanes at $x$; so closing up $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$ maps $x$ to the quadric surface over which the above is a cone, i.e., $x$ is blown up. If $x$ is {\em not} a node, then there is a unique tangent hyperplane $T_x\ifmmode {\cal S} \else$\cS$\fi_3$, determining a point of $\ifmmode {\cal I} \else$\cI$\fi_4$. Furthermore, $T_x\ifmmode {\cal S} \else$\cS$\fi_3$ and $T_y\ifmmode {\cal S} \else$\cS$\fi_3$ {\em coincide} for $x\neq y$, if and only if $x$ and $y$ are contained in a common Segre plane (\ref{e112.1}), and the line joining $x$ and $y$ in that Segre plane passes through one of the four nodes, say $N$, in that Segre plane. This is because $T_x\ifmmode {\cal S} \else$\cS$\fi_3\cap \ifmmode {\cal S} \else$\cS$\fi_3=\ifmmode {\cal H} \else$\cH$\fi\cup Q_x$, where $Q_x$ is a residual quadric cone, and the quadric cone is the intersection of $T_x\ifmmode {\cal S} \else$\cS$\fi_3$ with the cone $C_N$ which is the tangent cone of the node $N$ in the Segre plane. So if $x$ and $y$ lie on a line through $N$, $Q_x$ and $Q_y$ coincide, so $T_x\ifmmode {\cal S} \else$\cS$\fi_3$ and $T_y\ifmmode {\cal S} \else$\cS$\fi_3$ coincide also. \begin{theorem}\label{t122a.1} The duality map $d:\ifmmode {\cal S} \else$\cS$\fi_3- - \ra \ifmmode {\cal I} \else$\cI$\fi_4$ is given by the linear system of quadrics \ref{p112b.1}, i.e., by the elements of the ideal $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$ of the ten nodes: $d=\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$. \end{theorem} {\bf Proof:} It suffices to check that $d$, viewed as a modification of $\ifmmode {\cal S} \else$\cS$\fi_3$, coincides with the birational map $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$ of Theorem \ref{t123.1}. But this is easy. As the base locus is the set of nodes, these are blown up. As just explained, $x$ and $y$ in one of the Segre planes map to the same point on the image line precisely when the line joining them passes through one of the nodes in the Segre plane. As these lines are precisely what the map $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$ blows down, $d$ certainly coincides with $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$. \hfill $\Box$ \vskip0.25cm We also have the following analogue of Corollary \ref{c112b.1}. \begin{lemma}\label{l122a.1} The ideal $\ifmmode \hbox{{\script I}} \else$\scI$\fi(15)$ of the 15 singular lines of the Igusa quartic coincides with the Jacobian ideal of $\ifmmode {\cal I} \else$\cI$\fi_4$. \end{lemma} {\bf Proof:} Once again the inclusion $\ifmmode \hbox{{\script J}} \else$\scJ$\fi ac(\ifmmode {\cal I} \else$\cI$\fi_4)\subset} \def\nni{\supset} \def\und{\underline \ifmmode \hbox{{\script I}} \else$\scI$\fi(15)$ is obvious, and the inverse inclusion can be verified by means of standard basis computations, namely that $\ifmmode \hbox{{\script I}} \else$\scI$\fi(15)$ is generated by five cubics. \hfill $\Box$ \vskip0.25cm Along the same lines as Theorem \ref{t122a.1} we then get \begin{theorem} \label{t122b.1} The duality map $d:\ifmmode {\cal I} \else$\cI$\fi_4- - \ra\ifmmode {\cal S} \else$\cS$\fi_3$ is given by the system of cubics containing the 15 lines, i.e., by the Jacobian ideal of $\ifmmode {\cal I} \else$\cI$\fi_4$: $d=\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha^{-1}$. \end{theorem} {\bf Proof:} As above, it suffices to show that $d$, viewed as a modification of $\ifmmode {\cal I} \else$\cI$\fi_4$, coincides with the map $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha^{-1}$ of Theorem \ref{t123.1}. This is readily verified, as the base locus, the 15 lines, are blown up, while the tangent planes for any two points $x$ and $y$ in a common quadric of $\ifmmode {\cal I} \else$\cI$\fi_4$ (of Proposition \ref{p120.1}) coincide, blowing down the quadric surface to a node. \hfill $\Box$ \vskip0.25cm \subsection{The Siegel modular threefold of level 4} {}From the general theory of congruence subgroups, $X_{\gG(4)}\longrightarrow} \def\sura{\twoheadrightarrow X_{\gG(2)}$ is a Galois cover, with Galois group $\gG(2)/\gG(4)\ifmmode\ \cong\ \else$\isom$\fi (\integer/2\integer)^9$. Indentifying $X_{\gG(2)}^*$ with $\ifmmode {\cal I} \else$\cI$\fi_4$ and identifying $\ifmmode {\cal I} \else$\cI$\fi_4$ birationally with $\ifmmode {\cal S} \else$\cS$\fi_3$, we can consider Fermat covers over $X_{\gG(2)}^*$, i.e., given by a diagram \begin{equation}\label{e122b.1} \begin{array}{cccccc} Z({\bf A_4},n) & \stackrel{\-{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha}^{-1}}{- - \ra} & Y^{\wedge}({\bf A_4},n) & \longleftarrow} \def\rar{\rightarrow & \tilde{Y}({\bf A_4},n) & \\ \downarrow & & \downarrow & & \downarrow & (\integer/2\integer)^9 \\ \ifmmode {\cal I} \else$\cI$\fi_4 & \stackrel{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha^{-1}}{- - \ra} & \ifmmode {\cal S} \else$\cS$\fi_3 & \longleftarrow} \def\rar{\rightarrow & \tilde{\ifmmode {\cal S} \else$\cS$\fi}_3 & \end{array} \end{equation} where $\-{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha}^{-1}$ is {\em induced} by $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha^{-1}$, that is, (\ref{e122b.1}) is a fibre square (cf. (\ref{e115.0}), where $\tilde{\ifmmode {\cal S} \else$\cS$\fi}_3$ is denoted $\tilde{{\Bbb P}}^3$). \begin{theorem}\label{t122b.1} The Fermat cover $Z({\bf A_4},2)$ is the Satake compactification of the Siegel modular threefold of level 4. \end{theorem} {\bf Proof:} It suffices to show that $\~Y({\bf A_4},2)$ is the induced cover over $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4$, where $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_4\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_4$ is the desingularisation of $\ifmmode {\cal I} \else$\cI$\fi_4$ of Theorem \ref{t123.1}. Now the identification can be reduced to identifying what is in the branch locus of $\~Y({\bf A_4},2)\longrightarrow} \def\sura{\twoheadrightarrow \~{\ifmmode {\cal S} \else$\cS$\fi}_3$. There are two kinds of components: \begin{itemize}\item[a)] covers $\~Y({\bf A_3},2)$ of blown up ${\Bbb P}^2$'s, the $H_{ij}$ of (\ref{e111b.1}); \item[b)] covers $\~Y({\bf A_2},2)\times \~Y({\bf A_2},2)$ of ${\Bbb P}^1\times {\Bbb P}^1$'s, the $L_{0ij}$. \end{itemize} \begin{Lemma}\label{l122b.1} $\~Y({\bf A_3},2)\ifmmode\ \cong\ \else$\isom$\fi S(4)$, Shioda's elliptic modular surface of level 4. \end{Lemma} {\bf Proof:} This is well-known. $\~Y({\bf A_3},2)$ is K3 since it is a Fermat cover branched over six lines. One constructs structures of fibre space $\~Y({\bf A_3},2)\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$ with elliptic curves as fibres by taking the cover of the pencil of lines through a node (each such line meets four of the six lines outside the node, so the cover is branched at four points, i.e., is elliptic). The six fibres of type $I_4$ are readily identified, as are the 16 sections. \hfill $\Box$ \vskip0.25cm \begin{Lemma}\label{l122c.1} The cover $\~Y({\bf A_2},2)\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$ coincides with the cover $(\gG_1(4)\backslash \sieg_1)^*\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1$, by which we mean the Galois actions coincide. \end{Lemma} {\bf Proof:} This is even more well-known. \hfill $\Box$ \vskip0.25cm The theorem now follows, provided we accept that $\~Y({\bf A_4},2)$ is a quotient of $\sieg_2$ at all, i.e., that the cover $\sieg_2\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_4^0$ factorises (here $\ifmmode {\cal I} \else$\cI$\fi_4^0=\ifmmode {\cal I} \else$\cI$\fi_4-\{\hbox{15 lines}\}$), $\tilde{Y}({\bf A_4},2)^0:= \~Y({\bf A_4},2)-q^{-1}(\hbox{15 lines})$: \begin{equation}\label{e122c.1}\unitlength1.6cm \begin{picture}(1.6,1.4) \put(.2,1.33){$\sieg_2$} \put(.6,1.4){\vector(1,0){1}} \put(1.8,1.33){$\ifmmode {\cal I} \else$\cI$\fi_4^0$} \put(.4,1.25){\vector(1,-1){.6}} \put(.7,.45){$\~Y({\bf A_4},2)^0.$} \put(1.2,.66){\vector(1,1){.6}} \put(1.7,.8){$q$} \end{picture} \end{equation} But there is an easy way to see that this is the case: we can, for any given $x\in \~Y({\bf A_4},2)-\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi$ and $y=q(x)$, put a level 4 structure on $A_y$, such that the Galois group just permutes the level 4 over level 2 structures, that is, we make the identification $\gG(2)\backslash\gG(4)\ifmmode\ \cong\ \else$\isom$\fi (\integer/2\integer)^9\ifmmode\ \cong\ \else$\isom$\fi$ the Galois group of the cover. So $\~Y({\bf A_4},2)$, being a moduli space as in Shimura's theory, is a quotient of $\sieg_2$. \hfill $\Box$ \vskip0.25cm One could also imagine arguing with uniqueness of Galois covers, since we know the branch locus, branch degrees and Galois group. However there is in general no such uniqueness of covers, so we have to be careful. In our situation, there are two possible approaches to show uniqueness: \begin{itemize}\item[1)] Since the modular subvarieties determine, on the group-theoretic side, generators of the corresponding arithmetic group, we could conclude, from the isomorphisms \ref{l122b.1} and \ref{l122c.1}, the desired result. \item[2)] Since the branch divisors are totally geodesic with respect to the Bergmann metric, on the cover the metric retains its symmetry property. \end{itemize} Method 1) has been applied in \cite{J}, and 2) can be carried out for ball quotients. \section{The Hessian varieties of ${\cal S}_3$ and ${\cal I}_4$} \subsection{The Nieto quintic}\label{section4.1} Let $(x_0:\ldots:x_5)$ be the projective coordinates on ${\Bbb P}^5$ used to define $\ifmmode {\cal S} \else$\cS$\fi_3$ in (\ref{e111.3}), and let $\gs_i=\gs_i(x_0,\ldots,x_5)$ be the $i$-th elementary symmetric function $\gs_{\gl}=\sum_{i_1<\ldots < i_{\gl}}x_{i_1}\cdots x_{i_{\gl}}$ in $(x_0:\ldots:x_5)$. Define the {\em Nieto quintic} $\ifmmode {\cal N} \else$\cN$\fi_5$ by the equations\begin{equation}\label{e124.1} \ifmmode {\cal N} \else$\cN$\fi_5=\left\{\begin{array}{l}\gs_1 = 0 \\ \gs_5=0\end{array} \right. \subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^4=\{\gs_1=0\}\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^5. \end{equation} The symmetry of $\ifmmode {\cal N} \else$\cN$\fi_5$ under the symmetric group $\gS_6$ is evident from the equation. This quintic was discovered in the thesis \cite{Nie} and further studied in \cite{BN}, which will be our general reference for this section. We just briefly describe the geometry of $\ifmmode {\cal N} \else$\cN$\fi_5$ without discussing details. The singular locus is relatively easy to determine, just by calculating the Jacobian of (\ref{e124.1}). The result is \begin{proposition}[\cite{BN}, 3.1]\label{p124.1} $\ifmmode {\cal N} \else$\cN$\fi_5$ has the following singular locus: \begin{itemize}\item[(i)] 20 lines $L_{ijk}=\{x_i=x_j=x_k=0=\sum x_i\}$; \item[(ii)] ten isolated points, the $\gS_6$-orbit of $(1,1,1,-1,-1,-1)$, which are the points $P_{ij}=(1,\pm1,\ldots \pm 1)$, with $+1$ in the $i$-th and $j$-th positions. \end{itemize} \end{proposition} We will give a different proof of this below, see the discussion following Proposition \ref{piq8.1}. Note that the ten points occuring in (ii) are just the ten nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$ (see (\ref{e112.0}), cf.~also Remark \ref{r126.2} below). Furthermore, a local calculation shows that the singularities of $\ifmmode {\cal N} \else$\cN$\fi_5$ along the lines of (i) are of the type $\{\hbox{disc}\}\times A_1$, and at the points of (ii) are ordinary double points. Hence the former are resolved by a ${\Bbb P}^1$-bundle over the line $L_{ijk}$, while the points are resolved, as with the case of $\ifmmode {\cal S} \else$\cS$\fi_3$, by quadric surfaces. The 20 lines $L_{ijk}$ of \ref{p124.1} meet at the following 15 points: \begin{equation}\label{e124.2} Q_{ij}=(0,\ldots,1,\ldots,-1,\ldots) =\{\gS_6-\hbox{orbit of } Q_{56}=(0:0:0:0:1:-1)\}. \end{equation} \begin{lemma}\label{l124.1} The 20 lines $L_{ijk}$ of Proposition \ref{p124.1} meet four at a time at the 15 points $Q_{ij}$; each line $L_{ijk}$ contains three of the points, namely we have $Q_{ij}\in L_{klm}\iff \{i,j\}\cap \{k,l,m\}=\emptyset$. \end{lemma} {\bf Proof:} The line $L_{123}$ contains the three points $Q_{46},\ Q_{45}$ and $Q_{56}$, so by $\gS_6$-invariance each line contains three of the $Q_{ij}$. The point $Q_{56}$ is contained in the four lines $L_{123},\ L_{124},\ L_{134}$ and $L_{234}$, so by $\gS_6$-invariance, each point is contained in four lines. \hfill $\Box$ \vskip0.25cm Also, $\ifmmode {\cal N} \else$\cN$\fi_5$ contains a finite number of linear planes. \begin{lemma}\label{l125.1} $\ifmmode {\cal N} \else$\cN$\fi_5$ contains the following 30 ${\Bbb P}^2$'s: \begin{itemize}\item[(i)] 15 planes $N_{ijkl}=\{x_i+x_j=x_k+x_l=x_m+x_n=0\}$; \item[(ii)] 15 planes $N_{ij}=\{x_i=x_j=0=\sum_{k\neq i,j}x_k\}$; \end{itemize} \end{lemma} {\bf Proof:} It is immediately verified that these planes satisfy the equation (\ref{e124.1}).\hfill $\Box$ \vskip0.25cm Presumably these are in fact {\em all} the linear planes contained in $\ifmmode {\cal N} \else$\cN$\fi_5$. Note that the $N_{ijkl}$ are just the 15 planes (\ref{e112.1}) lying on the Segre cubic $\ifmmode {\cal S} \else$\cS$\fi_3$. Among the 15 planes $N_{ijkl}$ the common intersections were described in the discussion of the planes $P_{\gs}$ on the Segre cubic (see (\ref{e111.1})). \begin{lemma}\label{l125.2} Each plane $N_{ijkl}$ contains the following four of the ten points of \ref{p124.1}, (ii): $$P_{km},\ P_{kn},\ P_{lm} \hbox{ and } P_{ln};$$ it also contains the following three of the 15 points $Q_{ij}$ of (\ref{e124.2}): $Q_{ij},\ Q_{kl}$ and $Q_{mn}$. \end{lemma} {\bf Proof:} Consider $N_{0123}$; it contains the four nodes $(1:-1:1:-1:1:-1),\ (1:-1:-1:1:1:-1),\ (1:-1:1:-1:-1:1)$ and $(1:-1:-1:1:-1:1)$ which are the points $P_{24},\ P_{25},\ P_{34}$ and $P_{35}$, which gives the first statement by $\gS_6$-symmetry (there is an asymmetry in the notation, since we may take $i<j,k<l$ in the notation for $N_{ijkl}$, and since the first coordinate of $P_{ij}$ may be assumed to be $+1$). Similarly, $N_{0123}$ contains the three points $Q_{01},\ Q_{23}$ and $Q_{45}$, giving the second statement by $\gS_6$-symmetry.\hfill $\Box$ \vskip0.25cm We now note that these seven points lie in the plane $N_{0123}$ as in Figure \ref{Figure1}. \begin{figure}[thb] $$\fbox{\unitlength1.5cm \begin{picture}(6,4) \put(.5,.5){\line(1,1){3}} \put(1,1){\line(3,1){4.5}}\put(1,1){\line(-3,-1){.5}} \put(1,1){\circle*{.2}}\put(1,.7){$P_{35}$} \put(.5,2.5){\line(3,-1){5}} \put(.7,3.1){$N_{01}$} \put(2,2){\circle{.2}}\put(1.9,2.2){$Q_{23}$} \put(3,1){\circle{.2}}\put(2.4,.9){$Q_{45}$} \put(2.5,3.5){\line(1,-1){3}}\put(3,3.5){\line(0,-1){3}} \put(3,1.75){\circle*{.2}} \put(2.8,1,4){$P_{24}$} \put(3,3){\circle*{.2}}\put(3.2,3){$P_{25}$} \put(4,2){\circle{.2}}\put(3.8,2.2){$Q_{01}$} \put(5,1){\circle*{.2}}\put(4.6,.7){$P_{34}$} \put(5.2,3.1){$N_{23}$}\put(5.6,1.8){$N_{45}$} \put(5,3.5){in $N_{0123}$} \put(.5,1){\line(1,0){5}} \thicklines \put(1,3){\line(1,-1){2.5}} \put(2.5,.5){\line(1,1){2.5}} \put(.5,2){\line(1,0){5}} \end{picture}}$$ \caption[15 planes of type 1 on ${\cal N}_5$]{\label{Figure1}\small The plane $N_{0123}$} \end{figure} This is in fact easily checked. Note that the lines in $N_{0123}$, i.e., the intersections with the other $N_{ijkl}$ are {\em not} the lines of Proposition \ref{p124.1}; those lines have equations such as $x_0=x_1+x_2=x_1+x_3=x_4+x_5$. However, in the 15 planes $N_{ij}$ of \ref{l125.1}, several of the 20 singular lines $L_{ijk}$ {\em do} lie. In fact, we have \begin{lemma}\label{l125.3} Each $N_{ij}$ contains the four lines $L_{ijk}.\ L_{ijl},\ L_{ijm}$ and $L_{ijn}$. There are three planes passing through $L_{ijk}$, namely $N_{ij},\ N_{ik}$ and $N_{jk}$. $N_{ij}$ contains none of the nodes of \ref{p124.1}, (i), but contains six of the points $Q_{ij}$ of (\ref{e124.2}), namely $Q_{kl},\ Q_{km},\ Q_{kn},\ Q_{lm},\ Q_{ln}$ and $Q_{mn}$. These six points lie three at a time on the $L_{ijk}$ and form in each $N_{ij}$ a configuration as shown in Figure \ref{Figure2}. \begin{figure}[tbh] $$\fbox{\unitlength1.5cm \begin{picture}(6,4)\put(.5,1){\line(1,0){5}} \put(.5,2){\line(1,0){5}} \put(3,3.5){\line(0,-1){3}} \put(3,3){\circle{.2}}\put(1,1){\circle{.2}}\put(2,2){\circle{.2}} \put(3,1.66){\circle{.2}}\put(4,2){\circle{.2}}\put(5,1){\circle{.2}} \thicklines \put(.5,.5){\line(1,1){3}}\put(5.5,2.5){\line(-3,-1){5}} \put(.5,2.5){\line(3,-1){5}}\put(2.5,3.5){\line(1,-1){3}} \put(.5,2.7){$L_{015}$}\put(1.6,3.6){$L_{012}$}\put(3.6,3.6){$L_{013}$} \put(5,2.7){$L_{014}$}\put(5.5,1.2){$N_{0125}$} \put(5.5,1.8){$N_{0124}$}\put(2.6,.2){$N_{0123}$} \put(1.1,.7){$Q_{25}$}\put(4.6,.7){$Q_{34}$}\put(2.6,1.3){$Q_{23}$} \put(1.8,2.2){$Q_{24}$}\put(3.8,2.2){$Q_{35}$}\put(3.2,2.8){$Q_{45}$} \put(5,3.5){in $N_{01}$} \end{picture}}$$ \caption[15 planes of type 2 on ${\cal N}_5$]{\label{Figure2}\small The plane $N_{01}$} \end{figure} The three light lines are intersection of $N_{01}$ with $N_{ijkl}$ as indicated. \end{lemma} {\bf Proof:} This is once again easily verified. \hfill $\Box$ \vskip0.25cm Finally we note that there are hyperplanes in ${\Bbb P}^4$ cutting out these ${\Bbb P}^2$'s on $\ifmmode {\cal N} \else$\cN$\fi_5$. \begin{lemma}\label{l126.1} The six hyperplanes $\ifmmode {\cal H} \else$\cH$\fi_{ij}=\{x_i+x_j=0\}$ meet $\ifmmode {\cal N} \else$\cN$\fi_5$ each in the union of the three planes $N_{ijkl},\ N_{ijkm}$ and $N_{ijkn}$, and a residual quadric; the six hyperplanes $x_i=0$ meet $\ifmmode {\cal N} \else$\cN$\fi_5$ each in the union of five planes $N_{ij},\ N_{ik},\ N_{il},\ N_{im}$ and $N_{in}$. \end{lemma} {\bf Proof:} This is once again just a computation. \hfill $\Box$ \vskip0.25cm Now let us consider the intersection of $\ifmmode {\cal S} \else$\cS$\fi_3$ and $\ifmmode {\cal N} \else$\cN$\fi_5$. As is obvious from the above description, they both contain the 15 planes $N_{ijkl}$, and, the intersection being of degree 15, this is the entire intersection. From general arguments on projective varieties, from the fact that the dual of $\ifmmode {\cal S} \else$\cS$\fi_3$, namely the Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$, is {\em normal}, it follows that the parabolic divisor on $\ifmmode {\cal S} \else$\cS$\fi_3$, which is the intersection of $\ifmmode {\cal S} \else$\cS$\fi_3$ with the {\em Hessian variety}, must get blown down under the duality map, i.e., the intersection $Hess(\ifmmode {\cal S} \else$\cS$\fi_3)\cap \ifmmode {\cal S} \else$\cS$\fi_3$ consists of the 15 planes on $\ifmmode {\cal S} \else$\cS$\fi_3$! Since the Hessian has degree 5, this is the entire intersection, and it is natural to ask whether $\ifmmode {\cal N} \else$\cN$\fi_5$ and $Hess(\ifmmode {\cal S} \else$\cS$\fi_3)$ are related. In fact, we have \begin{lemma}\label{l126.2} The Nieto quintic is the Hessian of $\ifmmode {\cal S} \else$\cS$\fi_3$, i.e., $\ifmmode {\cal N} \else$\cN$\fi_5= Hess(\ifmmode {\cal S} \else$\cS$\fi_3)$, with equality, not just isomorphism. \end{lemma} {\bf Proof:} This is an easy computation (at least for a computer).\hfill $\Box$ \vskip0.25cm \begin{remark}\label{r126.2} Since the Hessian variety $\hbox{Hess}} \def\rank{\hbox{rank}(V)$ aquires nodes where $V$ has nodes, this ``explains'' the ten isolated singularities on $\ifmmode {\cal N} \else$\cN$\fi_5$. \end{remark} \subsection{Two birational transformations} We consider in this section two particularly interesting birational maps from $\ifmmode {\cal N} \else$\cN$\fi_5$. \subsubsection{ \ } The first arises through the duality map. Consider the birational map of ${\Bbb P}^4$ given in the following diagram:\begin{equation}\label{e126.1} \begin{array}{ccccccc} {\Bbb P}^4 & \stackrel{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1}{\longleftarrow} \def\rar{\rightarrow} & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}^4 & \stackrel{\~{\ga}}{- - - \ra} & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\widehat} \def\tilde{\widetilde} \def\nin{\not\in{{\Bbb P}}}^4 & \stackrel{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2}{\longrightarrow} \def\sura{\twoheadrightarrow} & {\Bbb P}^4 \\ \cup & & \cup & & \cup & & \cup \\ {\cal S}_3 & \longleftarrow} \def\rar{\rightarrow & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\cal S}_3 & \stackrel{\sim}{\longrightarrow} \def\sura{\twoheadrightarrow} & \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\cal I}_4 & \longrightarrow} \def\sura{\twoheadrightarrow & {\cal I}_4 \end{array}\quad; \end{equation} $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1$ and $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ were described in section \ref{section3.4}, and this diagram extends the one of Theorem \ref{t123.1} to the ambient rational fourfolds. It is easily seen that the ensuing rational map of ${\Bbb P}^4$, $\ga:=\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2\circ \~{\ga}\circ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}:{\Bbb P}^4\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$, is the map given by the Jacobian ideal of $\ifmmode {\cal S} \else$\cS$\fi_3$, that is, by the linear system of quadrics on the ten nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$ (see Corollary \ref{c112b.1}). This ``defines'' the map $\tilde{\ga}$; although we could in principle take any extension of $\tilde{\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha}:\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal S} \else$\cS$\fi_3}\longrightarrow} \def\sura{\twoheadrightarrow \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi_4}$, for our purposes it is convenient to use $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$. Then set \begin{equation}\label{e127.1} \ifmmode {\cal W} \else$\cW$\fi_{10}:=\ga(\ifmmode {\cal N} \else$\cN$\fi_5) \subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^4. \end{equation} We now describe $\ifmmode {\cal W} \else$\cW$\fi_{10}$ and show it is a hypersurface of degree 10, explaining the notation. First we have the \begin{lemma}\label{l25aux} The map $\ga:{\Bbb P}^4\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ blows up the ten nodes $P_{ij}$ of $\ifmmode {\cal S} \else$\cS$\fi_3$, with exceptional divisors $\ifmmode {\cal E} \else$\cE$\fi_{ij}$. Let $\ifmmode {\cal C} \else$\cC$\fi_{ij}$ denote the tangent cone at the point $P_{ij}$, a quadric cone fibred in lines passing through $P_{ij}$. Then each line of the cone gets blown down to the corresponding point in $\ifmmode {\cal E} \else$\cE$\fi_{ij}$. \end{lemma} {\bf Proof:} Since all the quadrics of $\ifmmode \hbox{{\script I}} \else$\scI$\fi(10)$ vanish at the $P_{ij}$, these points are blown up. To say the lines of $\ifmmode {\cal C} \else$\cC$\fi_{ij}$ get blown down is to say the ratios of the quadrics are constant along the line. This follows from the fact that the quadrics are the partial derivatives of $f$ (the defining polynomial of $\ifmmode {\cal S} \else$\cS$\fi_3$), and the line is tangent to the zero locus of $f$. \hfill $\Box$ \vskip0.25cm From this we get \begin{lemma} $\ga$, restricted to $\ifmmode {\cal N} \else$\cN$\fi_5$, is an isomorphism on the complement in $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}(\ifmmode {\cal N} \else$\cN$\fi_5)$ of the intersection of $\ifmmode {\cal N} \else$\cN$\fi_5$ with the tangent cones at the ten isolated singularities $P_{ij}$. \end{lemma} {\bf Proof:} This is clear from construction, taking into account the following fact, proved in \cite{BN}: the intersection of $\ifmmode {\cal N} \else$\cN$\fi_5$ with the tangent cone of one of the nodes $P_{ij}$ consists of the six Segre planes through the node, and an irreducible quartic {\em ruled} surface. It is then clear that these ruled surfaces get blown down, and that outside the Segre cubic and the ruled quartics, the birational map is a morphism. \hfill $\Box$ \vskip0.25cm {}From this it follows: \begin{lemma}\label{l127.1} The birational map $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2\circ \~{\ga} \circ \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_1^{-1}:\ifmmode {\cal N} \else$\cN$\fi_5- - \ra \ifmmode {\cal W} \else$\cW$\fi_{10}$ has image $\ifmmode {\cal W} \else$\cW$\fi_{10}$, whose singular locus contains the following. \begin{itemize}\item[(i)] ten singular quadric surfaces (the tangent hyperplane intersections of $\ifmmode {\cal I} \else$\cI$\fi_4$); \item[(ii)] 20 singular lines, coming from the singular locus of $\ifmmode {\cal N} \else$\cN$\fi_5$. \end{itemize} \end{lemma} {\bf Proof:} This follows from the description above. \hfill $\Box$ \vskip0.25cm Now note that both $Hess(\ifmmode {\cal I} \else$\cI$\fi_4)$ and $\ifmmode {\cal W} \else$\cW$\fi_{10}$ are symmetric under $\gS_6$, and both are singular along the ten quadrics of the Igusa $\ifmmode {\cal I} \else$\cI$\fi_4$. It may very well be that the two coincide, but we have not checked this. At any rate, $\ifmmode {\cal W} \else$\cW$\fi_{10}$ meets $\ifmmode {\cal I} \else$\cI$\fi_4$ in the union of quadric surfaces, each with multiplicity 2. Hence \begin{theorem}\label{t127.1} $\ifmmode {\cal I} \else$\cI$\fi_4\cap\ifmmode {\cal W} \else$\cW$\fi_{10}$ consists of the ten quadric surfaces (\ref{e117a.1}), each with multiplicity 2. Consequently, the degree of $\ifmmode {\cal W} \else$\cW$\fi_{10}$ is 10, justifying the notation. \end{theorem} {\bf Proof:} The intersection has reduced degree 20, and each surface component is counted twice, hence the degree of the intersection is 40, so the degree of $\ifmmode {\cal W} \else$\cW$\fi_{10}$ is 10. \hfill $\Box$ \vskip0.25cm \begin{problem}\label{p131.1} Is $\ifmmode {\cal W} \else$\cW$\fi_{10}$ also a compactification of an arithmetic quotient? \end{problem} \subsubsection{ \ } The other birational transformation is the following. \begin{equation}\label{e127.2}\begin{minipage}{14cm}\begin{itemize} \item[a)] Blow up the 15 points $Q_{ij}$ of (\ref{e124.2}); let $p_1:\~{\ifmmode {\cal N} \else$\cN$\fi}_5\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal N} \else$\cN$\fi_5$ denote this blow up. \item[b)] As each of the lines $L_{ijk}$ contains three points (see Lemma \ref{l124.1}), each $L_{ijk}$ can be blown down to an isolated singular point (the normal bundle is $\ifmmode {\cal O} \else$\cO$\fi(-2)\oplus\ifmmode {\cal O} \else$\cO$\fi(-2)$, cf.~(\ref{e110.1})). Let $p_2:\~{\ifmmode {\cal N} \else$\cN$\fi}_5\longrightarrow} \def\sura{\twoheadrightarrow \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$ denote this blow down. \end{itemize} \end{minipage} \end{equation} The following is easy to see (see Figures \ref{Figure1} and \ref{Figure2}). \begin{lemma}\label{l127.3} The singular locus of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$ consists of the 20 isolated cusps from (\ref{e127.2}) b), and the ten cusps, the images of the singular points $P_{ij}$ of Proposition \ref{p124.1}, (ii). The proper transforms of the $N_{ijkl}$ of Lemma \ref{l125.1}, (i) on $\ifmmode {\cal N} \else$\cN$\fi_5$ are ${\Bbb P}^2$'s blown up in three points, a del Pezzo surface; the proper transforms of the $N_{ij}$ of Lemma \ref{l125.1}, (ii) are ${\Bbb P}^2$'s blown up in six points, then the $L_{ijk}$ are blown down to four nodes, so this is the singular cubic surface with four nodes, the Cayley cubic. \end{lemma} \subsection{Moduli interpretation} The Nieto quintic was discovered as the solution of a certain moduli problem, and we briefly state the results of \cite{BN} describing this. The point of departure is the action of the Heisenberg group $H_{2,2}$ on ${\Bbb P}^3$, and the study of quartics which are invariant under the action. $H_{2,2}$ is a group of order 32 generated by the following linear transformations of ${\Bbb P}^3$ with coordinates $(z_0:z_1:z_2:z_3)$: \begin{equation}\label{e128.1}\begin{array}{rclcl} \gs_1 & : & (z_0:z_1:z_2:z_3) & \mapsto & (z_2:z_3:z_0:z_1) \\ \gs_2 & : & (z_0:z_1:z_2:z_3) & \mapsto & (z_1:z_0:z_3:z_2) \\ \tau_1 & : & (z_0:z_1:z_2:z_3) & \mapsto & (z_0:z_1:-z_2:-z_3) \\ \tau_2 & : & (z_0:z_1:z_2:z_3) & \mapsto & (z_0:-z_1:z_2:-z_3) \end{array} \end{equation} The center of the group is $\pm1$ and $PH_{2,2}=H_{2,2}/\pm1$ has a nice interpretation: \begin{equation}\label{e128.2} PH_{2,2}\ifmmode\ \cong\ \else$\isom$\fi (\integer/2\integer)^4, \end{equation} which carries, as in (\ref{e121.1}), an induced symplectic form. This means that one can speak of isotropic elements of the {\em group} $PH_{2,2}$. The normaliser of $H_{2,2}$ in $SL(4,\komp)$ maps surjectively to $\gS_6\ifmmode\ \cong\ \else$\isom$\fi Sp(4,\integer/2\integer)$, which acts transitively on diverse geometric loci of the sympectic form {\em inside} the group $PH_{2,2}$. These loci are: \begin{equation}\label{e128.3}\begin{minipage}{14cm}\begin{itemize} \item[a)] 15 pairs of skew lines \item[b)] 15 invariant tetrahedra \item[c)] ten fundamental quadrics. \end{itemize}\end{minipage}\end{equation} The moduli problem considered is a special set of quartics which are invariant under (\ref{e128.2}). The set of {\em all} invariant quartics is just a ${\Bbb P}^4$, spanned for example by the five quartics: $$\begin{array}{ccc} & g_0:=z_0^4+z_1^4+z_2^4+z_3^4 & \\ g_1:=2(z_0^2z_1^2+z_2^2z_3^2) & g_2=2(z_0^2z_2^2+z_1^2z_3^2) & g_3:=2(z_0^2z_3^2+z_1^2z_2^2) \\ & g_4:=4z_0z_1z_2z_3. & \end{array}$$ Let $(A,B,C,D,E)$ denote the coordinates of a particular quartic $Q_{(A,B,C,D,E)}=\{Ag_0+Bg_1+Cg_2+Dg_3+Eg_4=0\}$. The generic quartic $Q_{(A,B,C,D,E)}$ is smooth, and the locus of singular quartics can be determined as an equation in $(A,\ldots,E)$. Note that the $(A,\ldots,E)$ are functions of $(z_0:\cdots:z_3)$, so the answer as to whether $Q_{(A,B,C,D,E)}$ is singular depends on the point $z\in {\Bbb P}^3$. This is discussed in detail in \cite{BN}. The result is given in Table \ref{table19}. \begin{table} \begin{center} \caption{\label{table19}Singular Heisenberg invariant quartics } \begin{tabular}{|l|l|l|l|}\hline $z\in {\Bbb P}^3$ & dim$Q^{sing}$ & $Q_{(A,B,C,D,E)}$ & $S_{(A,B,C,D,E)}$ \\ \hline\hline $\notin$ fix line & 0 & Kummer surface & Segre cubic \\ \hline $\in$ one fix line & 2 & singular in four coordinate vertices & $A=0$ \\ \hline $\in$ the intersection of two fixed lines & 3 & singular along two fixed lines & $A=B=0$ \\ \hline \end{tabular} \end{center} {\small Notations: $Q^{sing}$ denotes the space of quartics singular at $z$, $S_{(A,B,C,D,E)}$ denotes the equation of the locus $Q^{sing}$ in the coordinates $(A,B,C,D,E)$.} \end{table} As one sees, the first row of the table is equivalent to Theorem \ref{t119.1} above! The special class of quartics to be considered here is, however, a quite different set, consisting of generically smooth quartics. This is the set of Kummer surfaces of (1,3)-polarised abelian surfaces, which, as it turns out, can be smoothly embedded in ${\Bbb P}^3$. This was discovered independently by Naruki and Nieto (see \cite{na} and \cite{Nie}). The 16 exceptional ${\Bbb P}^1$'s resolving the 16 double points of the Kummer surface are 16 disjoint {\em lines} on the quartics. Also, by a result of Nikulin \cite{Ni}, the converse is true, i.e., any quartic containing 16 lines is a Kummer surface. Furthermore, the quartic being invariant under $PH_{2,2}$, if it contains one line, it contains all 16 transforms, so the moduli involved is the condition: \begin{equation}\label{e129.1} L\ \parbox[t]{12cm}{ is a line in ${\Bbb P}^3$ lying on a smooth Heisenberg invariant quartic surface} \end{equation} The equation describing this in the Grassmannian $\fG(2,2)=\{x_0^2+\cdots + x_5^2=0\}$ is calculated in \cite{Nie}. It is \begin{equation}\label{e130.1} \ifmmode {\cal M} \else$\cM$\fi_{20}=\{\gs_5(x_0^2,\ldots,x_5^2)=0=\gs_1(x_0^2,\ldots,x_5^2)\}. \end{equation} Now one considers the natural 2-power map \begin{eqnarray}\label{e130.2} m_2:{\Bbb P}^5 & \longrightarrow} \def\sura{\twoheadrightarrow & {\Bbb P}^5 \\ (x_0,\ldots,x_5) & \mapsto & (x_0^2,\ldots,x_5^2)=(u_0,\ldots,u_5) \nonumber \end{eqnarray} and the image of $\ifmmode {\cal M} \else$\cM$\fi_{20}$ in ${\Bbb P}^5$. Comparing the equations (\ref{e124.1}) and (\ref{e130.1}) we have \begin{lemma}\label{l130.1} $m_2(\ifmmode {\cal M} \else$\cM$\fi_{20})=\ifmmode {\cal N} \else$\cN$\fi_5$. \end{lemma} The main results of \cite{BN} can be described as follows. First we define a Zariski open subset $M^s\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal M} \else$\cM$\fi_{20}$. The following 15 quadric surfaces $q_{ij}\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^5$ actually lie on $\ifmmode {\cal M} \else$\cM$\fi_{20}$, as is easily verified: \begin{equation}\label{e130a.1} q_{ij}=\{x_i=x_j=0=\sum_{m\neq i,j}x_m^2\}. \end{equation} Under the squaring map $m_2$ (\ref{e130.2}) the quadric $q_{ij}$ maps to the plane \begin{equation} N_{ij}=\{u_i=u_j=0=\sum_{m\neq i,j}u_m\}; \end{equation} so the image of ${\bf Q}:=\cup_{i,j}q_{ij}$ is ${\bf N}:=\cup_{i,j}N_{ij}$, and the planes $N_{ij}$ are the 15 planes of Lemma \ref{l125.1} (ii). Furthermore the $N_{ij}$ are contained in the branch locus of ${m_2}_{|\ifmmode {\cal M} \else$\cM$\fi_{20}}:\ifmmode {\cal M} \else$\cM$\fi_{20}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal N} \else$\cN$\fi_5$; this locus is {\em singular} on $\ifmmode {\cal M} \else$\cM$\fi_{20}$ because $\ifmmode {\cal N} \else$\cN$\fi_5$ is tangent to $u_i=0$ and $u_j=0$ in all of $N_{ij}$. Next consider the inverse image under $m_2$ of the ten nodes; since the nodes lie on {\em none} of the branch planes $u_i=0$, each node has $\deg(m_2)=32$ inverse images, so $\ifmmode {\cal M} \else$\cM$\fi_{20}$ has 320 singular points (clearly also nodes), which are the $\gS_6$-orbit, call it ${\bf P}$, of the points \begin{equation}\label{e130a.2} (\pm1:\pm1:\pm1:\pm i:\pm i: \pm i). \end{equation} Finally consider the inverse images of the 15 Segre planes of Lemma \ref{l125.1} (i). This locus is given by the 15 equations which are the $\gS_6$-orbit of \begin{equation}\label{e130a.3} x_0^2+x_1^2=x_2^2+x_3^2=x_4^2+x_5^2=0. \end{equation} Inspection shows that this degree 8 surface on $\ifmmode {\cal M} \else$\cM$\fi_{20}$ splits into eight planes, giving altogether 120=15.8 planes on $\ifmmode {\cal M} \else$\cM$\fi_{20}$; let ${\bf R}$ denote their union. Now define: \begin{equation}\label{e130a.5} M^s:=\ifmmode {\cal M} \else$\cM$\fi_{20}-{\bf Q}-{\bf P}-{\bf N},\ \ifmmode {\cal N} \else$\cN$\fi_5^s:=m_2(M^s). \end{equation} Then the statement proved in \cite{BN} is \begin{theorem}\label{t130.1} \begin{itemize}\item[a)] $M^s$ is isomorphic to a Zariski open subset of the moduli space $\ifmmode {\cal A} \else$\cA$\fi_{(1,3)}(2)$ of abelian surfaces with a $(1,3)$ polarisation and a level $2$ structure; \item[b)] There is a double cover $p:\~{\ifmmode {\cal N} \else$\cN$\fi}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal N} \else$\cN$\fi_5$ for which $p^{-1}(\ifmmode {\cal N} \else$\cN$\fi_5^s)$ is isomorphic to a Zariski open set of the moduli space $\ifmmode {\cal A} \else$\cA$\fi_{(2,6)}(2)$; \item[c)] $\ifmmode {\cal N} \else$\cN$\fi_5^s$ is the moduli space of $PH_{2,2}$-invariant smooth quartic surfaces containing $16$ skew lines. \end{itemize} \end{theorem} Since the varieties $\ifmmode {\cal M} \else$\cM$\fi_{20}$, $\~{\ifmmode {\cal N} \else$\cN$\fi}$ and $\ifmmode {\cal N} \else$\cN$\fi_5$ are compactifications of the Zariski open sets of (\ref{e130a.5}), we have the following: \begin{corollary} \label{c130.1} There are birational equivalences: $$\ifmmode {\cal M} \else$\cM$\fi_{20}- - \ra (\gG_{(1,3)}(2)\backslash \sieg_2)^*,\quad \~{\ifmmode {\cal N} \else$\cN$\fi}- - \ra (\gG_{(2,6)}(2)\backslash \sieg_2)^*,\quad \ifmmode {\cal N} \else$\cN$\fi_5- - \ra (\gG\backslash \sieg_2)^*,$$ where $ \gG_{(1,3)} \subset} \def\nni{\supset} \def\und{\underline \gG_{(2,6)} \subset} \def\nni{\supset} \def\und{\underline \gG,\ [\gG:\gG_{(2,6)}(2)]=2.$ \end{corollary} As is shown in \cite{BN}, the map $\~{\ifmmode {\cal N} \else$\cN$\fi}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal N} \else$\cN$\fi_5$ is given in the following way. It just happens to turn out the {\em any} of the $PH_{2,2}$-invariant quartics with 16 skew lines actually contains 32 lines, the first skew set of 16 and a second set of 16 skew lines. The second set of sixteen is found as the image of the first set under the involution \begin{equation}\label{e130.3} (x_0:\ldots:x_5)\mapsto \left({-1 \over x_0}: {1\over x_1}: \cdots :{1\over x_5}\right), \end{equation} which can be adjoined to the group $PH_{2,2}$ to form a group of order 32. Altogether the 32 lines have the following properties. \begin{equation}\label{e130.4} \begin{minipage}{14cm}\begin{itemize} \item[a)] The 32 lines intersect in 32 points; \item[b)] Each line contains ten of the 32 intersection points; \item[c)] Each intersection point is contained in ten of the 32 lines. \end{itemize}\end{minipage}\end{equation} A configuration with the properties (\ref{e130.4}) is called a $(32_{10})$-configuration. {}From Nikulin's results just mentioned, it follows that the second set of 16 lines are also the images of blown-up torsion points on another abelian surface, so there are {\em two} abelian surfaces with $(2,6)$ polarisation and level $2$ structure giving rise to the {\em same} resolved Kummer surface, i.e., the map is given by \begin{eqnarray}\label{e130.5} \~{\ifmmode {\cal N} \else$\cN$\fi} & \longrightarrow} \def\sura{\twoheadrightarrow & {\cal N}_5 \\ (A_{\tau_1},A_{\tau_2}) & \mapsto & \-{(A_{\tau_1}/\{\pm 1\})}\ifmmode\ \cong\ \else$\isom$\fi \-{(A_{\tau_2}/\{\pm 1\})},\nonumber \end{eqnarray} where the isomorphism permutes the two sets of 16 skew lines. The next step is to identify the modular subvarieties on the arithmetic quotients of Corollary \ref{c130.1}. From the structure of the periods we know that in terms of abelian surfaces, these modular subvarieties parameterise the abelian surfaces which split. These loci are described to some extent in \cite{BN}. \begin{theorem}\label{t131.1}\begin{itemize}\item[a)] Points on $\ifmmode {\cal N} \else$\cN$\fi_5$ parameterise smooth quartic surfaces unless they lie on one of the 30 planes of Lemma \ref{l125.1}; \item[b)] points on $\ifmmode {\cal N} \else$\cN$\fi_5$ parameterise quartic surfaces containing more that 32 lines if an only if the corresponding abelian surfaces are products. Furthermore, a line on a surface of this set of quartic surfaces has coordinates in ${\Bbb P}^5$ which is in the $\gS_6$-orbit of $$x_0^4(x_1^2+x_2^2)+x_1^4(x_2^2+x_0^2)+x_2^4(x_0^2+x_1^2)-6x_0^2x_1^2x_2^2 =0.$$ \end{itemize} \end{theorem} Unfortunately, these result do not allow us to explicitly describe the relation between the compactification $\ifmmode {\cal M} \else$\cM$\fi_{20}$ and compactifications of $\ifmmode {\cal A} \else$\cA$\fi_{(1,3)}(2)$, in particular the Baily-Borel embedding. This must be considered an interesting open problem. \subsection{A conjecture} To end this section we make a conjecture on one of the birational models of the variety $\ifmmode {\cal N} \else$\cN$\fi_5$. Consider the birational map $\ifmmode {\cal N} \else$\cN$\fi_5- - \ra \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$ of (\ref{e127.2}). Recalling now the Janus-like isomorphism between the Picard modular variety $\-{X}_{\gG_{\sqrt{-3}}(\sqrt{-3})}$ and the Siegel modular variety $\-{X}_{\gG(2)}$ (see \cite{J}), it is natural to ask about an analogue here, since the involved Siegel modular varieties of Corollary \ref{c130.1} all are related to level 2, albeit with different polarisations. So consider abelian fourfolds with complex multiplication by $\rat(\sqrt{-3})$ of signature (3,1), with a level $\sqrt{-3}$ structure, but with (1,1,1,3) polarisations. \begin{problem}\label{p132.1} Is $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$ the Satake compactification of $X_{(1,1,1,3)}(\sqrt{-3}):=\gG_{(1,1,1,3)}(\sqrt{-3})\backslash \ball_3$, where $\gG_{(1,1,1,3)}(\sqrt{-3})$ denotes the arithmetic group giving equivalence of complex multiplication by $\rat(\sqrt{-3})$, signature $(3,1)$, with a level $\sqrt{-3}$ structure and a $(1,1,1,3)$-polarisation? \end{problem} I conjecture that for {\em some} subgroup of $\gG_{(1,1,1,3)}(\sqrt{-3})$, this does in fact hold. Evidence for the conjecture: \begin{itemize}\item[i)] The proper transforms of the 15 Segre planes are by Proposition \ref{p116.1} the moduli space of principally polarised abelian threefolds with complex multiplication by $\rat(\sqrt{-3})$, signature (2,1), with a level $\sqrt{-3}$ structure, (although these moduli spaces are blown up in three points on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$). These could parameterize abelian fourfolds with said CM, signature (3,1) with a level $\sqrt{-3}$ structure and polarisation $(1,1,1,3)$ which split: $$A_4\ifmmode\ \cong\ \else$\isom$\fi A_3\times A_1,$$ where $A_3$ has CM, signature (2,1), polarisation (1,1,1), and $A_1$ has CM, but a polarisation 3. \item[ii)] The proper transforms of the 15 planes $N_{ij}$ of Lemma \ref{l125.1}, (ii), are four nodal cubic surfaces (Lemma \ref{l127.3}). These surfaces occur also on the ball quotient $\ifmmode {\cal S} \else$\cS$\fi_3$ above: pick any four of the nodes which are not coplanar; they determine a unique ${\Bbb P}^3$ in ${\Bbb P}^4$, and its intersection with $\ifmmode {\cal S} \else$\cS$\fi_3$, a cubic surface, has four nodes in the four nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$ in that ${\Bbb P}^3$. (Note that there is a unique four-nodal cubic surface, as it is ${\Bbb P}^2$ blown up in the six intersection points of four (general) lines, a complete quadrilateral in ${\Bbb P}^2$, and any two such quadrilaterals are projectively equivalent. This cubic surface is usually called the Cayley cubic, mentioned above.) \item[iii)] The singular locus consists of isolated singular points, resolved by quadric surfaces, so these singularities are rational. Recall that at each $P_{ij}$, six of the 15 Segre planes meet. At each $Q_{ij}$ (the 15 points (\ref{e124.2})), three of the Segre planes and six of the $N_{ij}$ of Lemma \ref{l125.1}, (ii) meet. In both cases, the quadric ${\Bbb P}^1\times {\Bbb P}^1$ can be covered equivariantly by a product $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}$ of the elliptic curve $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}$ with branching only at the intersection with the proper transforms of the 30 planes above, as follows: \begin{itemize}\item $P_{ij}$: $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1\times {\Bbb P}^1$ a Galois $\integer/3\integer$-quotient; \item $Q_{ij}$: $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^1\times{\Bbb P}^1$ is the product of two double covers branched at $0,1,\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta, \varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^2$. \end{itemize} This supports by Lemma \ref{l115.1} the idea that this could be the compactification locus of $X_{(1,1,1,3)}(\sqrt{-3})$. \end{itemize} \section{The Coble variety $\cal Y$} \subsection{Arithmetic quotients of domains of type $\bf IV_n$} Let $V$ be a $k$ vector space of dimension $n$, $k$ a totally real field, and $b$ a bilinear symmetric form on $V$. Let $G(V,b)$ be the symmetry group, and $G_{\rat}=Res_{k|\rat}G(V,b)$ the $\rat$-group it defines. We assume that $G_{\rat}$ is of hermitian type, so that for every infinite prime $\nu$ of $k$ the signature of $b_{\nu}$ is $(n-2,2)$ or $b_{\nu}$ is definite. $G_{\rat}$ is (absolutely) simple (defines an irreducible domain) only if $k=\rat$ (or if $b_{\nu}$ is definite for all but a single $\nu$, in which case the $\rat$-group is anisotropic, but we will not consider this situation), and the corresponding real group gives rise to a bounded symmetric domain only if $b$ has Witt index 2. This is the case we consider here. The classification of such forms is well-known; since we require the Witt index to be 2, two such forms $b$ and $b'$ are equivalent over $\rat$ $\iff$ $det(b)=det(b')$, where $det(b)$ is to be viewed as an element of $\rat^{\times}/(\rat^{\times})^2$. Now let $\ifmmode {\cal L} \else$\cL$\fi\subset} \def\nni{\supset} \def\und{\underline V$ be a (maximal) lattice, and let $G_{\ifmmode {\cal L} \else$\cL$\fi}$ be the arithmetic group it defines, $\gG\subset} \def\nni{\supset} \def\und{\underline G_{\ifmmode {\cal L} \else$\cL$\fi}$ a subgroup of finite index. We first remark on the moduli interpretation of the arithmetic quotient $\ifmmode {X_{\gG}} \else$\xg$\fi$. \begin{proposition}\label{p106.1} $\ifmmode {X_{\gG}} \else$\xg$\fi$ is a moduli space of (pure) Hodge structures of weight 2 on $V$ with $h^{2,0}=1$ (and $h^{1,1}=dim(\ifmmode {X_{\gG}} \else$\xg$\fi)$) with respect to the lattice $\ifmmode {\cal L} \else$\cL$\fi\subset} \def\nni{\supset} \def\und{\underline V$. \end{proposition} {\bf Proof:} The symmetry group of such a Hodge structure is of real type $SO(n-2,2)$, and $G(V,b)$ is a $\rat$-form in which $G_{\ifmmode {\cal L} \else$\cL$\fi}$ is an arithmetic subgroup. Since the corresponding ``period'' (i.e., position of the varying complex subspace $H^{1,1}$ in $H^2_{\komp}$) is clearly the same exactly when the two periods differ by an element of $G_{\ifmmode {\cal L} \else$\cL$\fi}$, while $\gG$ defines a level structure of some kind, the result follows. \hfill $\Box$ \vskip0.25cm This proposition is often used in the study of polarised K3-surfaces, which have a pure Hodge structure of type (1,19,1). In fact, for each polarisation degree (i.e., the number $C^2$ for the ample divisor $C$ on the K3-surface which gives the projective embedding) $2e,\ (e\geq1)$ one has an arithmetic group $\gG_e$ such that the arithmetic quotient $X_{\gG_e}$ is the moduli space of K3 surfaces with the given polarisation. Recall the {\em Picard number} $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$ is the rank of the group of algebraic cycles, i.e., $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=rk_{\integer}H^2(S,\integer)\cap H^{1,1}$. Then one has the following. \begin{proposition}\label{p106.2} Let $S$ be a K3 surface with $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$= the Picard number of $S$. Then the dimension of the moduli space of K3's which are in the family preserving the lattice of algebraic cycles $H^2(S,\integer)\cap H^{1,1}$ is 20-$\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$. \end{proposition} {\bf Proof:} Recall that for a K3 surface $H^1(S,\gT)\ifmmode\ \cong\ \else$\isom$\fi H^1(S,\Omega} \def\go{\omega} \def\gm{\mu} \def\gn{\nu} \def\gr{\rho^1)$, so $H^{1,1}(S)$ may be viewed as the tangent space of the local deformation space, which should be thought of as a varying complex subspace of $H^2(S,\komp)_{prim}$, while $H^2(S,\integer)$ is fixed. Let $\ifmmode \hbox{{\script A}} \else$\scA$\fi=H^2(S,\integer)\cap H^{1,1}$ be the lattice of algebraic cycles, $\ifmmode \hbox{{\script T}} \else$\scT$\fi=H^2(S,\integer)\cap (H^{2,0}(S)\oplus H^{0,2}(S))$ the lattice of transcendental cycles. We have $rk_{\integer}\ifmmode \hbox{{\script A}} \else$\scA$\fi=rk_{\integer}(H^2(S,\integer))-rk_{\integer}\ifmmode \hbox{{\script T}} \else$\scT$\fi$ in general and $rk_{\integer}\ifmmode \hbox{{\script A}} \else$\scA$\fi=\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta$ by assumption, so $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=22-rk_{\integer}\ifmmode \hbox{{\script T}} \else$\scT$\fi$, while the moduli space is defined by the group $G(V',b')$, where $V'=\ifmmode \hbox{{\script A}} \else$\scA$\fi^{\perp}\otimes \rat$, since we are requiring $\ifmmode \hbox{{\script A}} \else$\scA$\fi$ to be preserved. (Recall that for an algebraic cycle $\ifmmode {\cal C} \else$\cC$\fi$ the integral $\int_{\ifmmode {\cal C} \else$\cC$\fi}\go$ over the holomorphic two-form $\go$ vanishes, hence the algebraic cycles contribute nothing to the periods). Thus $G(\fR)=SO(20-\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta,2)$, giving rise to a domain of type $\bf IV_{20-\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}$. \hfill $\Box$ \vskip0.25cm Of course in this particular case, the lattice $\ifmmode {\cal L} \else$\cL$\fi\subset} \def\nni{\supset} \def\und{\underline V$ is very special; the ``intersection form'' $b$ restricted to $\ifmmode {\cal L} \else$\cL$\fi$ is even and unimodular, and as is well-known, decomposes as \begin{equation}\label{e106.1} \ifmmode {\cal L} \else$\cL$\fi\ifmmode\ \cong\ \else$\isom$\fi <-2e>\oplus {\bf H}^2 \oplus {\bf E_8}^2, \end{equation} where ${\bf H}$ is the two-dimensional hyperbolic lattice, and ${\bf E_8}$ is the root lattice of type ${\bf E_8}$. Let us remark that the compactification of these arithmetic quotients has been carried out in the thesis \cite{scat}, but we will not need this. We will very quickly describe a particulary interesting family of K3 surfaces which has been thoroughly studied by Yoshida and his collaborators, see \cite{MSY} for details on all matters here. \subsection{A four-dimensional family of K3's}\label{s106a.1} The family of K3 surfaces to be described here is the set of surfaces which are double covers of ${\Bbb P}^2$ branched along the union of six disjoint lines. Recall that there is a 19-dimensional family of K3 surfaces which are double covers of the plane branched along a sextic curve; they are smooth as long as the sextic is smooth, and generically have Picard number $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=1$. An arrangement of six lines in ${\Bbb P}^2$ is a maximally singular sextic; there are 15 intersection points of the six lines (if they are in general position), and each such gives rise to an $A_1$-singularity on the double cover. Resolving the 15 double points introduces 15 exceptional curves with self intersection number $-2$, so together with the pullback of the generic line, this gives 16 independent cycles on the surface: $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=16$. Hence the transcendental lattice $\ifmmode \hbox{{\script T}} \else$\scT$\fi$ has rank 4, so by Proposition \ref{p106.2}, the moduli space is four-dimensional. Let \begin{equation}\label{e106a.1} \gG=\{g\in G(\ifmmode \hbox{{\script T}} \else$\scT$\fi_{\fR},Q)(\ifmmode\ \cong\ \else$\isom$\fi SO(4,2))\Big| g(\ifmmode \hbox{{\script T}} \else$\scT$\fi)\subset} \def\nni{\supset} \def\und{\underline \ifmmode \hbox{{\script T}} \else$\scT$\fi\}, \end{equation} where $Q$ is the intersection form on $H^2(S,\integer)$, extended to $\fR$, then restricted to $\ifmmode \hbox{{\script T}} \else$\scT$\fi_{\fR}$. This is clearly an arithmetic subgroup, and by Proposition \ref{p106.1}, the arithmetic quotient $\ifmmode {\xg = \gG\bs\cD} \else$\xgeq$\fi$ is the four-dimensional moduli space. We list some of the interesting loci for this family. Let $L$ be the given arrangement, $L=l_1\cup \ldots \cup l_6$, and let $t_p:=$ the number of $p$-fold points of the arrangement, i.e., the number of points at which $p$ of the lines meet (see (\ref{e109.1})), and let $\pi:S\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^2$ denote the (singular) double cover. \subsubsection{Three-dimensional loci} \begin{itemize} \item[1)] Suppose there is a conic which is {\em tangent} to all six lines. Then the inverse image of this quadric is a ${\Bbb P}^1$, which, as is easily checked, has self-intersection number $4-6=-2$, so the double cover has 16 exceptional cycles, hence $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=17$. It is in fact easy to see that the surface $S$ is in this case a classical Kummer surface, i.e., a quartic surface in ${\Bbb P}^3$ with 16 nodes which is the Kummer variety of a principally polarised abelian surface $A_S$. The projection from a node gives the double cover $\pi:S\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^2$, and the tangent conic is the image of the (blown up) node used to project. The abelian surface is the Jacobian of a genus 2 curve, and this curve is the double cover of the conic, {\em branched at the six points of tangency}. This is well-known. \item[2)] If $t_3=1,\ t_2=12$, then the threefold point induces an $A_2$-singularity on the double cover which is resolved by two ${\Bbb P}^1$'s, so there are now 2+12 exceptional ${\Bbb P}^1$'s and the hyperplane section. We have the following picture. \\ \fbox{ \begin{minipage}{5cm} \unitlength1cm \begin{picture}(5,4)(0.5,0.3) \put(1,3){\line(2,-1){4}} \put(2,2.5){\circle*{.2}} \put(1,2.5){\line(1,0){4}} \put(1,2){\line(2,1){4}} \put(1,4){\line(1,-1){3.5}} \put(2.5,4){\line(1,-3){1.2}} \put(1.7,.5){\line(1,2){1.75}} \end{picture} \end{minipage}} \begin{minipage}{10cm} There are in fact three more exceptional ${\Bbb P}^1$'s, which are the inverse image on the double cover of the three lines which pass through the triple point and one of the three double points not lying on a line through the triple point. It is easy to see that these three double points are independent parameters of such arrangements, so this defines a three-dimensional family, so by Proposition \ref{p106.2}, we have $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=17$ for the generic member of this family. \end{minipage} \end{itemize} \subsubsection{Two-dimensional loci} \begin{itemize} \item[3)] \fbox{ \begin{minipage}{5cm} \unitlength1cm \begin{picture}(5,4)(0.5,0.3) \put(1,2.5){\line(2,-1){4}} \put(2,2){\circle*{.2}} \put(1,2){\line(1,0){4}} \put(1,1.5){\line(2,1){4}} \put(2.5,.5){\line(1,2){1.75}} \put(3.35,2.3){\circle*{.2}} \put(1,3.5){\line(2,-1){4}} \put(2.55,4){\line(1,-2){1.75}} \end{picture} \end{minipage}} \begin{minipage}{10cm}If $t_3=2, t_2=9$, there are two possibilities. Suppose first that the two threefold points do {\em not} lie on one of the six lines. Then we have the picture to the left. This gives rise to two isolated $A_2-$singularities. The inverse image of the line joining the two threefold points is also an exceptional ${\Bbb P}^1$. In this case the generic double cover has $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=18$, and as parameters one can take two double ratios: consider two of the lines $l_1, l_2$, both passing through one of the threefold points $p$; the three intersection points with the other lines, together with $p$, give four points on each line -- hence two double ratios. \end{minipage} \item[4)] It may also occur that both threefold points lie on a line, but in this case we also have $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=18$, i.e., a two-dimensional family. \end{itemize} \subsubsection{One-dimensional loci} \begin{itemize} \item[5)] If $t_4=1$, then the double cover has an {\em elliptic} singularity over the point, so is not K3. Hence this is a genuine {\em degeneration} of the K3, i.e., belongs to the boundary of the compactification. It turns out that then a line must also be double, so that the double cover has two components. \item[6)] As a further specialisation of 4) it may happen that there are three triple points. Since four of the lines may be choosen fixed (for example $x_0=x_1=x_2=0,\ x_0-x_1=0$), there is only one modulus, given for example by the intersection point of the two variable lines. Here we have $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=19$. \end{itemize} \subsubsection{Zero-dimensional loci} \begin{itemize} \item[7)] If three lines are taken, each {\em double}, then the double cover splits into two copies of ${\Bbb P}^2$. This is in the closure of the set of degenerations of type 5). \item[8)] The arrangement is the complete quadrilateral. The picture is: \\ \fbox{ \begin{minipage}{6.5cm} \unitlength.8cm \begin{picture}(8,5)(1.5,0) \thinlines \put(2,1.5){\line(1,0){6.5}} \put(2,1.5){\line(4,1){5}} \put(2,1.5){\line(6,5){3.5}} \put(2,1.5){\line(-1,0){0.5}} \put(2,1.5){\line(-4,-1){0.5}} \put(2,1.5){\line(-6,-5){0.5}} \put(5,.75){\line(0,1){4}} \put(5,4){\line(6,-5){3.5}} \put(5,4){\line(-6,5){0.5}} \put(8,1.5){\line(-4,1){5}} \put(8,1.5){\line(4,-1){0.5}} \end{picture} \end{minipage}} \begin{minipage}{9cm}It is known that the {\em Fermat} cover (not the double cover) of this arrangement is Shioda's elliptic modular surface of level 4, $S(4)$, so it follows that the double cover is isogenous to $S(4)$, i.e., a quotient of $S(4)$ by a group isomorphic to $(\integer/2\integer)^4$. This is the most special K3 surface in the family and has $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta=20$. \end{minipage} \end{itemize} \subsubsection{Level 2 structure} Now consider, in addition to the above data, a level 2 structure. Geometrically this amounts to fixing an {\em order} of the six lines. In terms of the lattice $\ifmmode \hbox{{\script T}} \else$\scT$\fi$ it is not so easy to see what it means. In \cite{MSY} it is shown by explicit computation that the subgroup $\gG(2)$ is the group generated by reflections on the ``roots'' of $\ifmmode \hbox{{\script T}} \else$\scT$\fi$, that is the integral elements of norm $-2$. Furthermore it is shown there that $\gG$ is generated by the reflections on the elements of norm $-2$ or $-4$, and that $\gG/\gG(2)\ifmmode\ \cong\ \else$\isom$\fi \gS_6\times \integer/(2)$. Hence by the results 2.7.1, 2.7.7, 2.8.2 of \cite{MSY} we have \begin{proposition}\label{p106b.1} The arithmetic quotient $\gG(2)\backslash \ifmmode {\cal D} \else$\cD$\fi$ is the moduli space for K3 surfaces which are double covers of ${\Bbb P}^2$, branched over an {\em ordered} set of six lines. \end{proposition} \begin{table} \caption{\label{table18} Loci of a four-dimensional family of K3 surfaces } \vspace*{.5cm} \begin{minipage}{16.5cm} \hspace*{2.5cm}\fbox{\begin{minipage}{5cm} Locus 1) Igusa quartic \unitlength1cm \begin{picture}(2,2)(0,.2) \put(0,1){\circle{.2}} \put(1,0.5){\circle{.2}} \put(2,1){\circle{.2}} \put(0,2){\circle{.2}} \put(1,2.5){\circle{.2}} \put(2,2){\circle{.2}} \bezier{100}(0,1)(-.2,1.5)(0,2) \bezier{100}(0,2)(0.5,2.5)(1,2.5) \bezier{100}(1,2.5)(1.5,2.5)(2,2) \bezier{100}(2,2)(2.3,1.5)(2,1) \bezier{100}(2,1)(1.5,0.45)(1,.5) \bezier{100}(1,.5)(0.5,0.45)(0,1) \end{picture} {\hspace*{\fill} \fbox{1} \hfill} \end{minipage}} \hspace*{\fill} \fbox{\begin{minipage}{5cm} Locus 2) $X^{\{ijk\}}$ \unitlength1cm \begin{picture}(2,2)(0,.2) \put(1,1){\circle{.2}} \put(2,1.5){\circle{.2}} \put(1,2){\circle{.2}} \put(3,1.8){\circle{.2}} \put(1,1){\circle{.2}} \put(1.1,1){\line(1,0){.9}} \put(2,1){\circle{.2}} \put(2.1,1){\line(1,0){.9}} \put(3,1){\circle{.2}} \end{picture} {\hspace*{\fill} \fbox{20} \hfill} \end{minipage}} \hspace{\fill} \begin{minipage}{15cm}\unitlength1cm \begin{picture}(15,2) \put(5,.7){\vector(-1,1){.8}} \put(8,.7){\vector(1,1){.8}} \put(13,.7){\vector(-1,1){.8}} \end{picture} \end{minipage} \hspace*{3cm} \fbox{\begin{minipage}{5cm} Locus 3) $X^{\{ijk;lmn\}}$ \unitlength1cm \begin{picture}(2,2)(.2,.3) \put(0,1){\put(1,1){\circle{.2}} \put(1.1,1){\line(1,0){.9}} \put(2,1){\circle{.2}} \put(2.1,1){\line(1,0){.9}} \put(3,1){\circle{.2}} } \put(1,1){\circle{.2}} \put(1.1,1){\line(1,0){.9}} \put(2,1){\circle{.2}} \put(2.1,1){\line(1,0){.9}} \put(3,1){\circle{.2}} \end{picture} {\hspace*{\fill} \fbox{10} \hfill} \end{minipage}} \hspace{\fill} \fbox{\begin{minipage}{5cm} Locus 4) $X^{\{ijk;imn\}}$ \unitlength1cm \begin{picture}(2,2) \put(.5,1){\circle{.2}} \put(.6,1.1){\line(2,1){.9}} \put(.6,1.1){\line(2,-1){.9}} \put(1.5,.5){\circle{.2}} \put(1.5,1.5){\circle{.2}} \put(1.6,.6){\line(2,-1){.9}} \put(1.6,1.6){\line(2,1){.9}} \put(2.5,1.4){\circle{.2}} \put(2.5,0){\circle{.2}} \put(2.5,2){\circle{.2}} \end{picture} {\hspace*{\fill} \fbox{90} \hfill} \end{minipage}} \begin{minipage}{15cm} \unitlength1cm \begin{picture}(15,2) \put(5,.7){\vector(0,1){.8}} \put(9,.7){\vector(1,1){.8}} \put(13,.7){\vector(0,1){.8}} \end{picture} \end{minipage} \hspace*{3cm} \fbox{\begin{minipage}{5cm} Locus 5) $X^{\{ij\}}$ \unitlength1cm \begin{picture}(2,2) \put(.5,1){\circle{.2}}\put(.6,1){\line(1,0){.9}} \put(1.5,1){\circle{.2}}\put(1.6,1){\line(1,0){.9}} \put(2.5,1){\circle{.2}}\put(2.6,1){\line(1,0){.9}} \put(2,2){\circle{.3}}\put(2,2){\circle*{.2}} \end{picture} {\hspace*{\fill} \fbox{15} \hfill} \end{minipage}} \hspace{\fill} \fbox{\begin{minipage}{5cm} Locus 6) $X^{\{ijk;klm;mni\}}$ \unitlength1cm \begin{picture}(2,2)(0,.2) \put(.5,.5){\circle{.2}}\put(.6,.5){\line(1,0){.9}} \put(.5,.5){\line(1,2){.4}} \put(1.5,.5){\circle{.2}}\put(1.6,.5){\line(1,0){.9}} \put(1,1.5){\circle{.2}}\put(1,1.5){\line(1,2){.4}} \put(1.5,2.5){\circle{.2}}\put(1.5,2.5){\line(1,-2){.4}} \put(2,1.5){\circle{.2}} \put(2,1.5){\line(1,-2){.4}} \put(2.5,.5){\circle{.2}} \end{picture} {\hspace*{\fill} \fbox{120} \hfill} \end{minipage}} \begin{minipage}{15cm} \unitlength1cm \begin{picture}(15,2) \put(5,.7){\vector(0,1){.8}} \put(9,.7){\vector(1,1){.8}} \put(13,.7){\vector(0,1){.8}} \end{picture} \end{minipage} \hspace*{3cm} \fbox{\begin{minipage}{5cm} Locus 7) $X^{\{ijk;kl;mn\}}$ \unitlength1cm \begin{picture}(2,2) \put(1,1){\circle{.3}}\put(1,1){\circle*{.2}} \put(2,1){\circle{.3}}\put(2,1){\circle*{.2}} \put(1.5,2){\circle{.3}}\put(1.5,2){\circle*{.2}} \end{picture} {\hspace*{\fill} \fbox{15} \hfill} \end{minipage}} \hspace{\fill} \fbox{\begin{minipage}{5cm} Locus 8) $X^{\{ijk;klm;mni;jln\}}$ \unitlength1cm \begin{picture}(2,2)(.9,.3) \put(.5,1.5){\circle{.2}}\put(1.6,.9){\circle{.2}}\put(1.6,2.1){\circle{.2}} \put(2,1.5){\circle{.2}}\put(2.7,.3){\circle{.2}}\put(2.7,2.7){\circle{.2}} \put(0,1.25){\line(2,1){3}}\put(0,1.75){\line(2,-1){3}} \put(1.33,.5){\line(2,3){1.5}}\put(1.33,2.5){\line(2,-3){1.5}} \end{picture} {\hspace*{\fill} \fbox{30} \hfill} \end{minipage}} \end{minipage} \vspace*{.5cm} {\small The notations $X^{\{ijk;klm;mni;jln\}}$, etc, are taken from \cite{MSY}; the arrows indicate inclusions among the various loci. \hfill\newline \vspace*{-1.2cm}The symbol \unitlength1cm \begin{picture}(.15,.1)\put(0.2,0.1){\circle{.3}}\put(0.2,0.1){\circle*{.2}} \end{picture} \quad means a double line. The number of each kind of loci is indicated by \fbox{x}; the dimensions are three in the top row down to zero in the last row. Locus 1) is where the six points lie on a conic, while the 20 $X^{\{ijk\}}$ are the loci where there are three of the six points on a line. The 15 $X^{\{ijk;kl;mn\}}$ lie on the boundary of the moduli space, while the 30 $X^{\{ijk;klm;mni;jln\}}$ lie in ``the farthest interior'' of the domain. A more complete description is given in Corollary \ref{c133.1}.} \end{table} We refer the reader to \cite{MSY} for a detailed description of the loci described above, of the periods and of the corresponding Picard-Fuchs equations (and much more). We give in Table \ref{table18} a description of the loci, giving the dual graph of the six lines (i.e., a vertex for each line, two vertices lying on a line $\iff$ the corresponding lines meet), as well as the number of loci, and the names given to them in \cite{MSY}. We now give an explicit projective description of the Baily-Borel compactification of the arithemetic quotient $\gG(2)\backslash \ifmmode {\cal D} \else$\cD$\fi$ of Proposition \ref{p106b.1}. All the facts presented here were proved originally by Coble \cite{C} or by Yoshida and his collaborators in \cite{MSY}. We have the four-dimensional family of K3 surfaces just discussed, defined in terms of a set of (ordered) six lines in the plane. Dual to the six lines are six points, and so the relation with the moduli space of cubic surfaces is evident. Let two ordered sets of six lines, $(l_1,\ldots,l_6),\ (l_1',\ldots,l_6')$ be given. \begin{definition}\label{d133.0} The two sets of lines $(l_1,\ldots,l_6),\ (l_1',\ldots,l_6')$ are said to be {\em associated}, if the following relation holds. Since the set $(l_1,\ldots,l_6)$ is ordered, we can form two triangles, $$\Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi(l_1,l_2,l_3),\quad \Delta} \def\gd{\delta} \def\gG{\ifmmode {\Gamma} \else$\gG$\fi(l_4,l_5,l_6);$$ these two triangles have together six vertices, which come equipped with a numbering, say $(p_1,\ldots,p_6)$, and these correspond dually to another ordered set of six lines, $(l_{p_1},\ldots,l_{p_6})$. Then $(l_1,\ldots,l_6)$ and $(l_1',\ldots,l_6')$ are associated, if: $(l_{p_1},\ldots,l_{p_6})=(l_1',\ldots,l_6')$, as a set of six ordered lines. \end{definition} Of course, starting with two sets of ordered six points, one can define in the same way the notion of association. Since, as abstract moduli spaces, the space of ordered sets of six lines is the ``same'' (by duality) as the set of ordered sets of six points, we see that we are dealing here with the space of sets of six ordered points in ${\Bbb P}^2$. This problem was dealt with in the papers of Coble \cite{C}, and has been given a modern treatment in \cite{DO}. It can be described as follows. The relevant moduli space is easy to describe: let $(p_1,...,p_n)$ be a set of $n$ points in ${\Bbb P}^k$; this is represented by $M$, the $n\times (k+1)$ matrix whose $i^{th}$ column gives the coordinates of the point $p_i$. The moduli space is then the GIT quotient \begin{equation}\label{e133.2} \hbox{{\helv P}$^k_n$\ } \def\helva{\hbox{\helv A}} = GL(k+1) \backslash M(n,k+1)/(\komp^*)^n. \end{equation} By taking the set of semistable points in $M(n,k+1)$ the above quotient is compact, although singular. It is classical that \ifmmode \hbox{\helv P}^1_6 \else $\pos$\fi\ is a threefold whose compactification can be identified with a cubic threefold in ${\Bbb P}^4$ with ten ordinary double points, which is just the Segre cubic $\ifmmode {\cal S} \else$\cS$\fi_3$. Note that the similar moduli problem, namely six points on a conic in ${\Bbb P}^2$, is realised by the Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$, so these are very closely related, but not identical moduli problems. Our interest here is in \pts, a fourfold. In this case we may represent elements by matrices \begin{equation}\label{e133.3} \hbox{{\helv P}$^k_n$\ } \def\helva{\hbox{\helv A}} \ni M = \left[ \begin{array}{c c c c c c} 1 & 0 & 0 & 1 & x & w \\ 0 & 1 & 0 & 1 & y & z \\ 0 & 0 & 1 & 1 & u & u \\ \end{array} \right] , \end{equation} and as Coble shows, the map \pts $\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$, $M \mapsto [x:y:w:z:u]$ is a birational map (it is clear that $\pts$ is rational, this map simply gives an explicit birationalisation). The GIT theory here consists of finding $G$-invariant functions on \pts, and these turn out to be generated by $3\times 3$ minors of $M$. In terms of the matrix $M$ the process of association can be described as follows. Each such matrix $M$ determines a second one: since the six points are ordered, one can define six lines by $l_{12}=\overline{p_1p_2}$, $l_{13}=\overline{p_1p_3}$, $l_{23}=\overline{p_2p_3}$, $l_{45}=\overline{p_4p_5}$, $l_{46}=\overline{p_4p_6}$, $l_{56}=\overline{p_5p_6}$; these six lines determine dually six points, whose coordinates are then brought into the normal form given above. It turns out that the entries of the second matrix are determined by the fact that the maximal minors are proportional to the maximal minors of the first. More precisely, if we let $(ijk)$ denote the $3\times 3$ minor of $M$ which is given by the columns $i,j,k$, and if we let $M'$ be the associated matrix, $(ijk)'$ the corresponding minor, then the minors of $M$ and $M'$ are related by: \begin{equation}\label{e133.1} (123)(145)(246)(356)=(124)'(135)'(236)'(456)'. \end{equation} Now association is an involution on $\pts$, and one can take the {\em quotient} by this involution. \begin{definition}\label{d133.1} Let $\ifmmode {\cal Y} \else$\cY$\fi$ be the double cover of ${\Bbb P}^4$ branched along the Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$, $\pi:\ifmmode {\cal Y} \else$\cY$\fi\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$. \end{definition} Clearly $\ifmmode {\cal Y} \else$\cY$\fi$ will be singular precisely along the singular locus of $\ifmmode {\cal I} \else$\cI$\fi_4$, i.e., \begin{lemma}\label{l133.1} The singular locus of $\ifmmode {\cal Y} \else$\cY$\fi$ consists of 15 lines, the inverse images of the 15 singular lines of $\ifmmode {\cal I} \else$\cI$\fi_4$. \end{lemma} \begin{theorem}[\cite{DO},Example 4, p.~37]\label{t133.0} The moduli space of six ordered points in ${\Bbb P}^2$ is equal to the double cover $\ifmmode {\cal Y} \else$\cY$\fi$, and the double cover involution on $\ifmmode {\cal Y} \else$\cY$\fi$ coincides with the association involution on $\pts$. \end{theorem} In other words, a set $(p_1,\ldots,p_6)$ is {\em associated to itself}, if and only if the six points lie on a conic in ${\Bbb P}^2$. Consider one of the hyperplanes $H$ in ${\Bbb P}^4$, $H=\{\gt_m^4=0\}$ of Proposition \ref{p120.1}. Since $H$ is {\em tangent} to $\ifmmode {\cal I} \else$\cI$\fi_4$, the inverse image $\pi^{-1}(H)$ in $\ifmmode {\cal Y} \else$\cY$\fi$ will {\em split into two copies of ${\Bbb P}^3$}. In this way, we get a union of 20 ${\Bbb P}^3$'s on $\ifmmode {\cal Y} \else$\cY$\fi$, \begin{lemma}\label{l133.2} The inverse images $\pi^{-1}(H)$ of the tangent hyperplanes $H=\{\gt_m^4=0\}$ consist of two copies each of ${\Bbb P}^3$, and these two ${\Bbb P}^3$'s on $\ifmmode {\cal Y} \else$\cY$\fi$ meet in the quadric surface which is the inverse image under $\pi$ of the quadric on $\ifmmode {\cal I} \else$\cI$\fi_4$ to which $H$ is tangent. This gives a total of 20 such ${\Bbb P}^3$'s on $\ifmmode {\cal Y} \else$\cY$\fi$. \end{lemma} A resolution of singularities of $\pts$ is affected by resolving the Igusa quartic by blowing up the ideal of the 15 lines; this is the map $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_2$ of Theorem \ref{t123.1}. Let $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}}$ denote this desingularisation $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}} \longrightarrow} \def\sura{\twoheadrightarrow \hbox{\pts}$. On $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}}$ we have a set of 36 divisors, the {\em discriminant locus}, the proper transforms of the Igusa quartic, the 20 ${\Bbb P}^3$'s and the 15 exceptional divisors of the blow up. It is clear how this variety is the moduli space of cubic surfaces: blow up ${\Bbb P}^2$ in the six points, and embed by the linear system of cubic curves through the six points. The ordering of the six points of course determines a marking of the 27 lines in the well-known manner. The symmetry group of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}}$ is $\gS_6\times \integer/2\integer$; although the Weyl group $W(E_6)$ acts birationally on it, the action is not regular. For that it is neccessary to modify $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}}$ even more. Dolgachev mentions in \cite{DO} that he suspects it is sufficient to blow up $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\hbox{\pts}}$ in the intersection of the 36 divisors. One of the many things proved in \cite{MSY} is the following. \begin{theorem}\label{t133.1} The variety $\ifmmode {\cal Y} \else$\cY$\fi$ is the Baily-Borel compactification of the arithmetic quotient $\gG(2)\backslash \ifmmode {\cal D} \else$\cD$\fi$ of Proposition \ref{p106.1}. \end{theorem} The proof given in \cite{MSY} of this fact simply (!) calculates the image of the period map, and in determining when the periods lie on the boundary of the period domain $\ifmmode {\cal D} \else$\cD$\fi$, the authors find that this locus coincides with the set of K3 surfaces whose set of six lines correspond to those singularities of Lemma \ref{l133.1} of $\ifmmode {\cal Y} \else$\cY$\fi$. \begin{corollary}\label{c133.1} The Loci 5) and 7) of Table \ref{table18} are the inverse images on $\ifmmode {\cal Y} \else$\cY$\fi$ of the 15 singular lines and 15 singular points, respectively, of the branch locus $\ifmmode {\cal I} \else$\cI$\fi_4$. The Loci 3) of Table 18 are the inverse images of the ten special hyperplane sections of Lemma \ref{l133.2}, i.e., the quadrics. The loci 2) of Table 18 are the 20 ${\Bbb P}^3$'s of Lemma \ref{l133.2}, and Locus 1) is just the branch locus of the double cover. \end{corollary} \part{A Gem of the modular universe}\label{chapter13} \renewcommand{\arraystretch}{1} \section{The Weyl group $W(E_6)$} \subsection{Notations} We use the same notation as above for the 27 lines on a cubic surface in ${\Bbb P}^3$: $a_1,...a_6,\ b_1,...,b_6,\ c_{12},...,c_{56}$. The 36 double sixes are: $$N= \left[ \matrix{ a_1 & a_2 & a_3 & a_4 & a_5 & a_6 \cr b_1 & b_2 & b_3 & b_4 & b_5 & b_6 \cr } \right], \hspace{1cm} (1) $$ $$N_{ij}=\left[ \matrix{ a_i & b_i & c_{jk} & c_{jl} & c_{jm} & c_{jn} \cr a_j & b_j & c_{ik} & c_{il} & c_{im} & c_{in} \cr } \right], \hspace{1cm} (15)$$ $$N_{lmn}\footnote{here we switch notations from $N_{ijk}$ in equation (\ref{eB2.2}) to $N_{lmn}$ for convenience} =\left[ \matrix{ a_i & a_j & a_k & c_{mn} & c_{ln} & c_{lm} \cr c_{jk} & c_{ik} & c_{ij} & b_l & b_m & b_n \cr} \right] \hspace{1cm} (20). $$ The 45 tritangents are:\begin{equation}\label{eQ1.1}\begin{minipage}{6cm} \begin{center} $(ij)=<a_i\ b_j\ c_{ij}>,\ \ i\neq j \ \ (30)$ $(ij.kl.mn)=<c_{ij}\ c_{kl}\ c_{mn}>\ \ (15).$ \end{center} \end{minipage} \end{equation} Two double sixes are {\em syzygetic} it they contain four lines in common, for example: $$N\ \ \hbox{and} \ \ N_{12}\ \ \hbox{have } a_1,\ a_2,\ b_1,\ b_2\ \hbox{in common},$$ and {\em azygetic} if they have six lines in common, for example: $$N\ \ \hbox{and} \ \ N_{456}\ \ \hbox{have } a_1,\ a_2,\ a_3,\ \ b_4,\ b_5,\ b_6\ \hbox{in common}.$$ Two azygetic double sixes have six lines in common and contain 12 other lines; these 12 lines form another double six, azygetic with respect to both, for example $N,\ N_{123},\ N_{456}$. Such triples are refered to as triples of azygetic double sixes or, because of the interpretation in terms of tritangents, a trihedral pair. Each double six is syzygetic to 15 others, forming 270 such pairs, and azygetic to 20 others, forming 120 triples. Our notation for the 120 triples are: \begin{equation}\label{eQ1.2} \begin{array}{rclr} \{ijk\} & = & <N,N_{ijk},N_{lmn}>, & (10) \\ \{ij.jk\} & = & <N_{ij},N_{ik},N_{jk}>, & (20) \\ \{ij.kl\} & = & <N_{ij},N_{ikl},N_{jkl}> & (90).\\ \end{array} \end{equation} We recognize these as the trihedral pairs of (\ref{eB2.1}) under the correspondence $$\left[\begin{array}{ccc} a_i & b_j & c_{ij} \\ b_k & c_{jk} & a_j \\ c_{ik} & a_k & b_i \end{array}\right] \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow <N_{ij},N_{ik},N_{jk}>,\quad \left[\begin{array}{ccc} c_{il} & c_{jm} & c_{kn} \\ c_{mn} & c_{ik} & c_{jl} \\ c_{jk} & c_{ln} & c_{im} \end{array}\right] \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow <N,N_{ijk},N_{lmn}>, $$ $$\left[\begin{array}{ccc} a_i & b_j & c_{ij} \\ b_l & a_k & c_{kl} \\ c_{il} & c_{jk} & c_{mn} \end{array}\right] \longleftrightarrow} \def\ra{\rightarrow} \def\Ra{\Rightarrow <N_{ij},N_{ikl},N_{jkl}>.$$ Hence the triads of trihedral pairs discussed there are expressed in condensed form as follows: \begin{equation}\label{eQ1.3}\begin{minipage}{14cm} $$ [ijk.lmn]=\left[ \matrix{ N_{ij} & N_{jk} & N_{ik} \cr N_{lm} & N_{mn} & N_{ln} \cr N & N_{ijk} & N_{lmn} \cr } \right],\hspace{1cm} (10)$$ $$ [ij.kl.mn]=\left[ \matrix{ N_{ij} & N_{ikl} & N_{jkl} \cr N_{kl} & N_{kmn} & N_{lmn} \cr N_{mn} & N_{nij} & N_{mij} \cr } \right], \hspace{1cm} (30).$$ \end{minipage} \end{equation} The group of incidence preserving permutations of the 27 lines, a group of order 51840, can be generated by the following six operations: $$(i,i+1),\ i=1,...,5:\hbox{transposition of the indices},$$ and $$ (123) :\hbox{map } N \mapsto N_{123},$$ and the graph of this presentation is shown in Figure \ref{Figure4}. \begin{figure}[hbt] $$ \unitlength0.8cm \begin{picture}(10.5,3.5) \put(0.25,3.0){\circle*{0.5}} \put(2.75,3.0){\circle*{0.5}} \put(5.25,3.0){\circle*{0.5}} \put(7.75,3.0){\circle*{0.5}} \put(10.25,3.0){\circle*{0.5}} \put(5.25,0.5){\circle*{0.5}} \linethickness{0.4mm} \put(0.25,3.0){\line(1,0){2.5}} \put(2.75,3.0){\line(1,0){2.5}} \put(5.25,3.0){\line(1,0){2.5}} \put(7.75,3.0){\line(1,0){2.5}} \put(5.25,0.5){\line(0,1){2.5}} \put(0.25,3.6){\makebox(0,0){(12)}} \put(2.75,3.6){\makebox(0,0){(23)}} \put(5.25,3.6){\makebox(0,0){(34)}} \put(7.75,3.6){\makebox(0,0){(45)}} \put(10.25,3.6){\makebox(0,0){(56)}} \put(5.25,-0.1){\makebox(0,0){(123)}} \end{picture} $$ \caption[The graph of the group of the 27 lines]{\label{Figure4}\small The graph of the group of the permutations of the 27 lines} \end{figure} This is the graph whose vertices correspond to generators, two vertices A, B being connected if ABA=BAB and not connected if AB=BA. \subsection{Roots} Let $\tt$ be a maximal abelian subalgebra of the compact Lie algebra $\ee_{6,u}$ over $\fR$, i.e., $\tt \ifmmode\ \cong\ \else$\isom$\fi \fR^6$. Let $x_1,...,x_6$ be coordinates such that the root forms of $E_6$ are: \begin{eqnarray*} (40) & & \pm(x_i\pm x_j), \hspace{1cm} 1\leq i <j\leq 5 \\ (32) & & \pm{1 \over 2}(\pm x_1\pm x_2\pm x_3\pm x_4 \pm x_5 + x_6), \ \ \hbox{even number of ``$-$'' signs inside the parenthesis.} \end{eqnarray*} (Note that in Bourbaki notation, our variables $x_i=\ge_i,\ i=1,..,5$, while our coordinate $x_6$ is denoted $\ge_8-\ge_7-\ge_6$ there). The 36 positive root forms are given by $\pm x_i+x_j$ and ${1\over 2}(\pm x_1\pm x_2\pm x_3\pm x_4 \pm x_5 + x_6)$, and they correspond to the 36 double sixes of the 27 lines on a cubic surface. We use the following notations for these forms \begin{equation}\label{eQ2.1} \begin{array}{lcl} h & = & {1 \over 2}(x_1+...+x_6), \\ h_{1j} & = & x_{j-1}-{1\over 2}(x_1+...+x_5-x_6),\ \ j=2,...,6 \\ h_{jk} & = & -x_{j-1}+x_{k-1},\ \ 1\neq j<k \\ h_{1jk} & = & x_{j-1}+x_{k-1}, \ \ j,k=2,...,6 \\ h_{jkl} & = &+x_{j-1}+x_{k-1}+x_{l-1}-{1 \over 2}(x_1+...+x_5-x_6),\ \ j,k,l\neq 1. \\ \end{array} \end{equation} The Weyl group of $E_6$ is generated by the reflections on these 36 hyperplanes; we denote these reflections by $s,\ s_{ij},$ and $s_{ijk}$. As a system of simple roots we take : \begin{equation}\label{eQ2.q} \begin{array}{cclcl} \ga_1 & = & -{1 \over 2}(-x_1+...+x_5-x_6) & = & h_{12} \\ \ga_2 & = & x_1+x_2 & = & h_{123} \\ \ga_3 & = & -x_1+x_2 & = & h_{23} \\ \ga_4 & = & -x_2+x_3 & = & h_{34} \\ \ga_5 & = & -x_3+x_4 & = & h_{45} \\ \ga_6 & = & -x_4+x_5 & = & h_{56}. \\ \end{array} \end{equation} Then the Dynkin diagram is as shown in Figure \ref{Figure5}; we recover Figure \ref{Figure4} by replacing $\ga_i$ by the corresponding {\it reflection} $s, s_{ij}, s_{ijk}$ on the hyperplanes where $h, h_{ij}, h_{ijk}$, respectively, vanish. This shows clearly the isomorphism of $W(E_6)$ and the group of the permutations of the 27 lines, $$Aut(\ifmmode {\cal L} \else$\cL$\fi)\ifmmode\ \cong\ \else$\isom$\fi W(E_6).$$ \begin{figure}[hbt] $$ \unitlength0.8cm \begin{picture}(10.5,3.5) \put(0.25,3.0){\circle*{0.5}} \put(2.75,3.0){\circle*{0.5}} \put(5.25,3.0){\circle*{0.5}} \put(7.75,3.0){\circle*{0.5}} \put(10.25,3.0){\circle*{0.5}} \put(5.25,0.5){\circle*{0.5}} \linethickness{0.4mm} \put(0.25,3.0){\line(1,0){2.5}} \put(2.75,3.0){\line(1,0){2.5}} \put(5.25,3.0){\line(1,0){2.5}} \put(7.75,3.0){\line(1,0){2.5}} \put(5.25,0.5){\line(0,1){2.5}} \put(0.25,3.6){\makebox(0,0){$\ga_1$}} \put(2.75,3.6){\makebox(0,0){$\ga_3$}} \put(5.25,3.6){\makebox(0,0){$\ga_4$}} \put(7.75,3.6){\makebox(0,0){$\ga_5$}} \put(10.25,3.6){\makebox(0,0){$\ga_6$}} \put(5.25,-0.1){\makebox(0,0){$\ga_2$}} \end{picture} $$ \caption[The Dynkin diagram of $E_6$]{\label{Figure5}\small The Dynkin diagram of the Weyl group of $E_6$} \end{figure} The action of the reflections on the root forms can be described as follows:\par \vspace{.5cm} \renewcommand{\arraystretch}{1.4} $\begin{array}{ccc|ccc|ccc|ccc} s(h_{ij}) & = & h_{ij} & s(h_{ijk}) & = & h_{lmn} & s_{ijk}(h) & = & h_{lmn} & s_{ij}(h) & = & h \\ \hline s_{ijk}(h_{lmn}) & = & h & s_{ijk}(h_{kmn}) & = & h_{kmn} & s_{ijk}(h_{jkn}) & = & h_{in} & s_{ijk}(h_{ln}) & = & h_{ln} \\ \hline s_{ijk}(h_{kn}) & = & h_{ijn} & s_{ijk}(h_{jk}) & = & h_{jk} & s_{ij}(h_{klm}) & = & h_{klm} & s_{ij}(h_{jlm}) & = & h_{ilm} \\ \hline s_{ij}(h_{ijm}) & = & h_{ijm} & s_{ij}(h_{ij}) & = & h_{ij} & s_{ij}(h_{jk}) & = & h_{ik} & s_{ij}(h_{lk}) & = & h_{lk} \\ \hline \end{array}$\renewcommand{\arraystretch}{1} \subsection{Vectors} The Killing form of $E_6$, a quadratic invariant, can be calculated as the sum of the squares of all roots, and evaluates to (a constant times): $$I_2=x_1^2+x_2^2+x_3^2+x_4^2+x_5^2+{1 \over 3}x_6^2.$$ With respect to the Killing form we have the vectors dual to the root forms: \begin{equation}\label{eQ3.1} \begin{array}{lcll} H & = & {1 \over 2}(1,1,1,1,1,3); & \\ H_{1j} & = & -{1 \over 2}(1,...,-1,...,-3), & 1 \hbox{ in the $(j-1)^{st}$ spot},\ j=2,\ldots,6; \\ H_{jk} & = & -{1 \over 2}(0,..1,..,-1,..,0), & \pm 1 \hbox{ in the $(j-1)^{st}, (k-1)^{st}$ spot}, 1<j<k\leq 6; \\ H_{1jk} & = & {1 \over 2}(0,..1,..,1,..,0), & 1 \hbox{ in the $(j-1)^{st}, (k-1)^{st}$ spot},\ 1<j<k\leq 6; \\ H_{jkl} & = & -{1 \over 2}(1,-1..,-1,..,-1,..,-3), & 1's \hbox{ in the $(j-1),\ (k-1),\ (l-1)$ spots}, \\ & & & \ 1<j<k<l\leq 6; \\ \end{array} \end{equation} which may be thought of as the root vectors (of the positive roots; the negative roots have a ``$-$'' sign in front). \begin{table} \caption{\label{table23} The arrangement in ${\Bbb P}^5$} \renewcommand{\tabcolsep}{4pt} \begin{tabular}{|l|r|rr|rrr|rrrrr|rrrr|} \hline $N(\ifmmode {\cal O} \else$\cO$\fi)$ & $A_1$ & $A_1^2$ & $A_2$ & $A_1^3$ &$A_{1,2}$ & $A_3$ & $A_{1^2,2}$ & $A_2^2$ & $A_{1,3}$ & $A_4$ & $D_4$ & $A_{1,2^2}$ & $A_{1,4}$ & $A_5$ & $D_5$ \\ \hline $\ifmmode {\cal O} \else$\cO$\fi$ & k & $t_2(3)$ & $t_3(3)$ & $t_3(2)$ &$t_4(2)$ & $t_6(2)$ & $t_5(1)$ & $t_6(1)$ & $t_7(1)$ & $t_{10}(1)$ & $t_{12}(1)$ & $t_7$ & $t_{11}$ & $t_{15}$ & $t_{20}$ \\ \hline $\#$ & 36 & 270 & 120 & 540 & 720 & 270 & 1080 & 120 & 540 & 216 & 45 & 360 & 216 & 36 & 27 \\ \hline $t(4)$ & 1 & 15 & 10 & 45 & 80 & 45 & 150 & 20 & 105 & 60 & 15 & 70 & 66 & 15 &15 \\ \hline $t_2(3)$ & & 1 & 0 & 6 & 8 & 3 & 28 & 4 & 18 & 12 & 3 & 20 & 20 & 6 & 7 \\ $t_3(3)$ & & & 1 & 0 & 6 & 9 & 9 & 2 & 18 & 18 & 6 & 6 & 18 & 6 & 9 \\ \hline $t_3(2)$ & & & & 1 & 0 & 0 & 6 & 0 & 3 & 0 & 1 & 6 & 6 & 1 & 3 \\ $t_4(2)$ & & & & & 1 & 0 & 3 & 1 & 3 & 3 & 0 & 4 & 6 & 3 & 3 \\ $t_6(2)$ & & & & & & 1 & 0 & 0 & 2 & 4 & 2 & 0 & 4 & 2 & 5 \\ \hline $t_5(1)$ & & & & & & & 1 & 0 & 0 & 0 & 0 & 2 & 2 & 0 & 1 \\ $t_6(1)$ & & & & & & & & 1 & 0 & 0 & 0 & 3 & 0 & 3 & 0 \\ $t_7(1)$ & & & & & & & & & 1 & 0 & 0 & 0 & 2 & 1 & 1 \\ $t_{10}(1)$ & & & & & & & & & & 1 & 0 & 0 & 1 & 1 & 2 \\ $t_{12}(1)$ & & & & & & & & & & & 1 & 0 & 0 & 0 & 3 \\ \hline \end{tabular} \end{table} As is well-known, there is also a set of 27 fundamental weights which form an orbit of $W(E_6)$, namely: \begin{equation}\label{eQ3.3} \begin{array}{cclccl} a_1 & = & -{2 \over 3}x_6; & a_j & = & x_{j-1}-{1 \over 2}(x_1+...+x_5+{1 \over 3}x_6); \\ b_1 & = & {1 \over 2}(x_1+...+x_5-{1 \over 3}x_6); & b_j & = & x_{j-1}+{1 \over 3}x_6; \\ c_{1j} & = & -x_{j-1}+{1 \over 3}x_6; & c_{ij} & = & -x_{j-1}-x_{i-1}+{1 \over 2}(x_1+...+x_5-{1 \over 3}x_6) .\\ \end{array} \end{equation} These form the $W(E_6)$ orbit of the fundamental weights denoted $\overline{\go}_1$ and $\overline{\go}_6$ in Bourbaki, which are just our $-a_1$ and $b_6$, respectively. Note that the following relation holds: \begin{equation}\label{eQ3.4} \sum_{i=1}^6 a_i = -3h =-3(\sum_{i=1}^6 x_i) =-\sum_{i=1}^6 b_i. \end{equation} Also note that the $a_i$ and $b_i$ are related by \begin{equation}\label{eQ3.5} b_i=a_i-{1\over 3}(a_1+\cdots +a_6). \end{equation} The corresponding vectors which are dual with respect to the Killing form are: \begin{equation}\label{eQ3.2} \begin{array}{cclcclr} A_1 & = & (0,...,0,-2); & A_j & = & {1 \over 2}(-1,...,+1,..,-1) & +1\ \hbox{in the $j-1$ spot;} \\ B_1 & = & {1 \over 2}(1,...,1,-1); & B_j & = & (0,...,1,..,1) & +1\ \hbox{in the $j-1$ spot;} \\ C_{1j} & = & (0,..,-1,...,1); & C_{ij} & = & {1 \over 2}(1,..,-1,..,-1,..,-1) & -1\ \hbox{in the $j-1$, $i-1$ spots.} \\ \end{array} \end{equation} \subsection{The arrangement defined by $W(E_6)$} The 36 hyperplanes in ${\Bbb P}^5$ defined by the vanishing of the 36 root forms form the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf E_6})$ of (\ref{e108.2}). For later reference we give the combinatorial data of the arrangement here. We denote as in (\ref{e109.1}) a ${\Bbb P}^m$ through which $k$ of the hyperplanes pass by $t_k(m)$. For the normalisers we use the notation $A_{i^k,j^l}$ for $A_i^k\times A_j^l$. The data of the arrangement is given in Table \ref{table23}. \subsection{Special Loci} In Table \ref{table24} we give a list of special loci which will be particularly important in what follows, so we give a brief description of each. \begin{table}[htb] \caption{\label{table24} Special loci in ${\Bbb P}^5$ \hfill} $$\begin{array}{|r|l| l| c| l|}\hline \# & \hbox{space} & \hbox{Symmetry} & N(\ifmmode {\cal O} \else$\cO$\fi) & \hbox{notation in Table \ref{table23}} \\ \hline \hline 36 & {\Bbb P}^4 & A_5 & A_1 & - \\ \hline 120 & {\Bbb P}^3 & D_4 & A_2 & t_3(3) \\ \hline 120 & {\Bbb P}^1 & A_2 & A_2\times A_2 & t_6(1) \\ \hline 216 & {\Bbb P}^1 & A_2 & A_4 & t_{10}(1) \\ \hline 45 & {\Bbb P}^1 & A_1 & D_4 & t_{12}(1) \\ \hline 36 & \hbox{point} & - & A_5 & t_{15} \\ \hline 27 & \hbox{point} & - & D_5 & t_{20} \\ \hline \end{array}$$ \end{table} \subsubsection{36 ${\Bbb P}^4$'s}\label{i1} In each of the 36 hyperplanes given by the vanishing of one of the 36 forms (\ref{eQ2.1}), $h$ say, the induced group is $\gS_6$, and as a reflection group on ${\Bbb P}^4$ it defines a projective arrangement of 15 planes; since each double six is syzygetic to 15 and azygetic to 20 others, there are 15 hyperplanes through which one of the other 35 intersect $h$, and ten planes through which two others of the 35 meet $h$. We immediately recognize this geometry as that in ${\Bbb P}^4$ discussed in the first part of the paper. The 15 hyperplanes are the 15 $\ifmmode {\cal H} \else$\cH$\fi_{ij}$ of (\ref{e112b.3}), each of which cuts out three planes on $\ifmmode {\cal S} \else$\cS$\fi_3$, and the ten are the hyperplanes mentioned in (\ref{e117a.1a}) and Proposition \ref{p120.1}. These in turn are the dual hyperplanes to the ten nodes on $\ifmmode {\cal S} \else$\cS$\fi_3$. \subsubsection{120 ${\Bbb P}^3$'s}\label{i2} These ${\Bbb P}^3$'s correspond to the 120 triples of azygetic double sixes, i.e., each is cut out by three of the 36 hyperplanes of \ref{i1}. In each such hyperplane, these ${\Bbb P}^3$'s correspond to the ten hyperplanes in $h$ just mentioned, given by the $K_{ijk}$ of (\ref{e117a.1a}). Each of these contains 15 planes, and one can check that these are just the faces and symmetry planes of a cube. The six lines in $K_{ijk}$ which are the singular locus $\ifmmode {\cal I} \else$\cI$\fi_4\cap K_{ijk}$, are easily identified with the six 12-fold lines $t_{12}(1)$ which are contained in $K_{ijk}$\footnote{The arrangement is $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$, minus the plane at infinity. Of the 16 $t_3(1)$ of $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$ (see (\ref{e109.2})), four lie in the plane at infinity.}, and the nine points $t_{20}$ contained in $K_{ijk}$ are the intersection points of those six lines\footnote{Likewise, nine of the 12 $t_6$ of $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$ lie in the plane at infinity}. Equations of the 120 ${\Bbb P}^3$'s are given by a triple of azygetic double sixes, e.g., by $<h,h_{ijk},h_{lmn}>$. \subsubsection{120 ${\Bbb P}^1$'s} \label{i3} The 120 lines correspond exactly to $A_2$ subroot systems, each containing three (positive) roots, so that each line contains three of the 36 points. The 120 lines are determined as follows. Consider a triad of triples of azygetic double sixes and the corresponding matrix of linear forms (see (\ref{eQ1.3})), say $$[ijk.lmn]=\left[ \matrix{ h_{ij} & h_{jk} & h_{ik} \cr h_{lm} & h_{mn} & h_{ln} \cr h & h_{ijk} & h_{lmn} \cr } \right].$$ Taking the ideal defined by the vanishing of two rows defines the corresponding line, i.e., \begin{equation}\label{eQ5.1} \begin{array}{lcl} L_{\{ij.jk\}} & = & <h_{lm}, h_{mn}, h_{ln}, h, h_{ijk}, h_{lmn}> \\ L_{\{lm.mn\}} & = & <h_{ij}, h_{jk}, h_{ik}, h, h_{ijk}, h_{lmn}> \\ L_{\{ijk\}} & = & <h_{ij}, h_{jk}, h_{ik},h_{lm}, h_{mn}, h_{ln}>.\\ \end{array} \end{equation} Each of the 120 lines contains three of the nodes, so for example, \begin{equation}\label{eQ5.2} H_{ij},\ H_{jk},\ H_{ik} \in L_{\{ij.jk\}}. \end{equation} There are 40 such triples of the 120 lines, which have the characterising property that they span ${\Bbb P}^5$. These correspond to subroot systems of the type $A_2 \times A_2 \times A_2$, where all three copies are orthogonal to one another. Note that given an $A_2$ subroot system, there is a unique $A_2 \times A_2$ subroot system orthogonal to it. Thus the $A_2$ subroot system is defined by the vanishing of the six root forms of the complementary $A_2 \times A_2$. There are 120 of each of both types of subroot systems. Summing up, there are six of the 36 hyperplanes passing through each of these 120 lines while each such line contains three of the 36 nodes. The {\it induced arrangement} is as follows. Blowing up along the line introduces an exceptional ${\Bbb P}^3$ over each point of the line; the intersection of it with the proper transforms of the six planes passing through it is the induced arrangement. It is of type $A_2\times A_2$, i.e., is given by two skew lines in ${\Bbb P}^3$ and two sets of three hyperplanes through each line. \subsubsection{216 ${\Bbb P}^1$'s}\label{i4} Consider a pair of {\em skew} lines, say $a_1, a_2.$ There is a unique double six containing the given pair as a column, e.g., $$N_{12}=\left[ \matrix{ a_1 & b_1 & c_{23} & c_{24} & c_{25} & c_{26} \cr a_2 & b_2 & c_{13} & c_{14} & c_{15} & c_{16} \cr } \right]. $$ There are 216 lines in ${\Bbb P}^5$ which join points such as $A_1, A_2, H_{12}$ (see (\ref{eQ3.1}) and (\ref{eQ3.2})). The ideal of these 216 lines is generated by 24 sextics, forming the irreducible representation denoted 24$_p$ in \cite{BL}. We can exhibit these sextics explicitly, as follows. The 216 lines are given by the equations: \begin{equation}\label{eQ7.1} \begin{array}{lclc} <A_i,A_j,H_{ij}> & = & <h_{kl}|k,l\neq i,j; h_{ijk}> & (15) \\ <B_i,B_j,H_{ij}> & = & <h_{kl}|k,l\neq i,j; h_{klm}|k,l,m\neq i,j> & (15) \\ <A_i,B_i,H> & = & <h_{kl}|k,l\neq i> & (6) \\ <A_i,C_{jk},H_{lmn}> & = & <h_{jk}, h_{\gl \gm}|\gl,\gm\neq i,j,k; h_{ij\gl}, h_{ik\gl}|\gl\neq i,j,k> & (60) \\ <B_i,C_{jk},H_{ijk}> & = & <h_{jk}, h_{\gl \gm}|\gl,\gm\neq i,j,k; h_{\gl\gm\gn}|\gl\neq i,j,k,\gm\neq i,k, \gn\neq i> & (60) \\ <C_{ik},C_{jk},H_{ij}> & = & <h_{ijm},h_{mn},h,h_{kmn}|m,n\neq i,j,k> & (60), \\ \end{array} \end{equation} i.e., each is defined by the vanishing of ten of the $h$'s; these lines are the $t_{10}(1)$ listed in the table of the arrangement. We claim the sextics are the products of the six root forms of an $A_2\times A_2$ subroot system. To see this, pick one, say $\Phi=h_{12}\cdot h_{13}\cdot h_{23}\cdot h_{45}\cdot h_{46}\cdot h_{56}$. It will suffice to check that for any of the 216 lines listed in (\ref{eQ7.1}), at least one of the hyperplanes on the right hand side is among the set $h_{12}, h_{13}, h_{23},h_{45}, h_{46}, h_{56}$. This is at most a tedious, but straightforward task. The dual ${\Bbb P}^3$'s, which are defined by the vanishing of the forms which are dual to the points of the left-hand sides, each {\em contain} the ten points which are dual to the forms on the right, for example $$P_{<A_i,A_j>}=\{a_i=a_j=h_{ij}=0\}\ni (H_{kl},H_{ijk}).$$ The induced arrangement over each line is the arrangement $\ifmmode {\cal A} \else$\cA$\fi(W({\bf A_4}))$ of (\ref{e109.2}). Among the ten hyperplanes defining the line, say $<A_1,B_1,H>$, there are ten triples of azygetic double sixes, $\{ij.jk\}$ in (\ref{eQ1.2}), with $i\neq 1$: $\{23.34\},\ \{23.35\},\ \{23.36\},\ \{24.45\},\ \{24.46\},\ \{25.56\},\ \{34.45\},\ \{34.46\},\ \{35.56\}$, and $\{45.56\}$, and these determine the ten $t_3(1)$ of (\ref{e109.2}). \subsubsection{45 ${\Bbb P}^1$'s}\label{i5} The 45 lines are the lines joining the 27 points of (\ref{eQ3.2}) in threes, for example, $$L_{(ij)}=<A_i,B_j,C_{ij}>.$$ These lines are defined by the vanishing of 12 of the $h$'s, so for example \begin{equation}\label{eQ9.1} L_{(12)}=<h_{34},h_{35},h_{36},h_{45},h_{46},h_{56},h_{234},h_{235},h_{236}, h_{245},h_{246},h_{256}>; \end{equation} these are the hyperplanes corresponding to the 12 double sixes {\em not} containing any of $a_1, b_2, c_{12}$. The induced arrangement is $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_4}))$, with 12 planes corresponding to the 12 hyperplanes through the line. Again there will be $t_2(1)$'s and $t_3(1)$'s, corresponding to triples of azygetic double sixes (respectively to syzygetic double sixes). The ideal of the 45 lines is generated by 15 quartics which form the irreducible representation denoted 15$_q$ in \cite{BL}. It is easy to see that this space of quartics is generated by a product of four pairwise azygetic $h$'s, for example by $h_{24}\cdot h_{124}\cdot h_{35}\cdot h_{135}$. In fact, each hyperplane of type $h_{ij}$ contains the 15 lines numbered (like the tritangents) by: \par $$\begin{array}{lccl} (kl) & \hbox{for} & k,l\neq i,j & \hbox{(12 of these)}\\ (ij.kl.mn) & \hbox{for} & k,l,m,n \neq i,j\ & \hbox{(3 of these)} \end{array}$$ \\ while the hyperplanes of type $h_{ijk}$ contain the 15 lines numbered by: $$\begin{array}{lccl} (mn) & \hbox{for} & n=i\ or \ j, m\neq i,j & \hbox{(9 of these)}\\ (il.jm.kn) & & & \hbox{(6 of these).} \end{array}$$\\ It is now easy to check that every line is contained in at least one of the four hyperplanes. Alternatively we can argue as follows: each $h$ contains 15 of the lines; there are six ${\Bbb P}^3$'s which are the intersection of two of the four, three of which are contained in each $h$. These three meet in a common line in the $h$, so the number of lines contained in the union is: $4\cdot(15-7)+2\cdot 6+1=45$, where the 7=number of lines in the union of the three ${\Bbb P}^2$'s in each $h$, 2=3-1 is the number of lines in each such ${\Bbb P}^2$, not in the others, and one is the common line. Note that this is the Macdonald representation corresponding to the four roots of an $A_1\times A_1\times A_1\times A_1\subset} \def\nni{\supset} \def\und{\underline D_4$ subroot system. Five of these lines meet at each of the 27 points, corresponding to the five tritangents through each of the 27 lines. \subsubsection{36 points}\label{i6} These are the 36 points (\ref{eQ3.1}) dual to the 36 hyperplanes of \ref{i1}. The induced arrangement is of course just the arrangement in ${\Bbb P}^4$ above. There are 15 hyperplanes passing through each of the 36 points, and these are just the hyperplanes which are coded by the double sixes which are syzygetic to the one with the notation of the point as in (\ref{eQ3.1}). So, for example, the 15 ${\Bbb P}^4$'s through the point $H$ are the 15 $h_{ij}$. These points correspond to ($\pm$) the roots of $E_6$. The orthogonal complement in $\fR^6$ of the root $\ga$ is projectively equivalent to the {\em dual} hyperplane to the point. For example, $H$ is dual to $h$, and one of the hyperplanes $P$ will contain $H$ $\iff$ the dual point $p$ is contained in $h$. The ideal of the 36 points is generated by 20 cubics, forming the irreducible representation of $W(E_6)$ denoted 20$_p$ in \cite{BL}. We can find these cubics explicitly as follows. Consider the hyperplanes $a_1, b_2, c_{12}$ corresponding to a tritangent. From Table \ref{table23} above we see that each of these hyperplanes contains 20 of the 36 points (actually, the table contains the dual information: there are 20 of the $h_{ij}$, etc, passing through each of the 27 points), and the ${\Bbb P}^3$ which is the common intersection of these three contains 12 of the 36 (the dual information is contained in the table: the 45 lines are 12-fold lines). Hence the product $a_1\cdot b_2\cdot c_{12}$ contains 3.(20-12)+12 = 36 of the 36 points. Through each of the 36 points, also 15 of the 27 hyperplanes of (\ref{eQ3.3}) pass, corresponding to the 15 lines {\em not} contained in the double six whose notation the point has. For example, the point $H={1\over 2}(1,\ldots,1,3)$ is contained in all the $c_{ij}$. In the exceptional ${\Bbb P}^4$ at the point, both sets of 15 hyperplanes (coming from the 36, respectively 27 hyperplanes) {\it coincide}. \subsubsection{27 points}\label{i7} These are the points $A_i,B_i,C_{ij}$ of (\ref{eQ3.2}). There are 20 of the 36 hyperplanes meeting at each, so the induced arrangement is one of 20 ${\Bbb P}^3$'s in ${\Bbb P}^4$, and one sees easily that it is $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_5}))$. This arrangement is also induced in any of the hyperplanes $a_i, b_i,c_{ij}$ of (\ref{eQ3.3}); we note that there are two kinds of ${\Bbb P}^2$, namely $t_2(2)$'s, corresponding to pairs of skew lines, and $t_3(2)$'s, corresponding to tritangents. Since each line is contained in five tritangents, there are five of the latter and 15 of the former (in each $a_i$, etc.). These 15 form an arrangement of type $\ifmmode {\cal A} \else$\cA$\fi(W({\bf A_5}))$ as discussed above. The ideal of these 27 points is generated by 30 cubics, forming the irreducible representation denoted 30$_p$ in \cite{BL}. It is easy to see that this space of cubics is generated by a product of three members of a syzgetic triple as $h_{12}\cdot h_{13}\cdot h_{23}$, for example: Each of the hyperplanes contains 15 of the 27, the ${\Bbb P}^3$ which is thier common intersection contains nine, so the union contains 3.(15-9)+9=27, or all of the points. Note that this is just the Macdonald representation corresponding to the (3) roots of an $A_2$ subroot system. We need, in addition to the above, certain information on the dual spaces. \subsubsection{45 ${\Bbb P}^3$'s} \label{i8} Consider one of the 45 ${\Bbb P}^3$'s which is dual to one of the 45 lines of \ref{i5}; it is cut out by three of the forms (\ref{eQ3.3}), and can be denoted as one of the 45 tritangents, for example, if $L_{(ij)}$ denotes the line $<A_i,B_j,C_{ij}>$ as in (\ref{eQ9.1}), the dual ${\Bbb P}^3$ may be denoted by $l_{(ij)}$, and \begin{eqnarray}\label{eq3.1} l_{(ij)} & = & a_i\cap b_j \cap c_{ij} \\ l_{(ij:kl:mn)} & = & c_{ij}\cap c_{kl}\cap c_{mn}.\nonumber \end{eqnarray} Consider the ${\Bbb P}^3$ $l_{(12)}$, given by $a_1=b_2=c_{12}=0$, or $x_1=x_6=0$. Then one checks easily that the hyperplanes (\ref{eQ2.1}) reduce in $l_{(12)}$ to the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf F_4})$ of (\ref{e108.2}). Considering how the 27 hyperplanes (\ref{eQ3.3}) intersect $l_{(12)}$, we find that these restrict to the set of short roots, that is, give a subarrangement of type $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$. See also Proposition \ref{pQ7.1} below. \subsection{Invariants} Since the 27 forms (\ref{eQ3.3}) are (as a set) invariant under the Weyl group the expression \begin{equation}\label{invariantsQ3} I_k:=\sum_{i,j} \left\{ a_i^k + b_i^k + c_{ij}^k \right\}, \end{equation} if non-vanishing, is an invariant of degree $k$. The ring of invariants of $W(E_6)$ is generated by elements in degrees 2, 5, 6, 8, 9 and 12, which can be taken to be $I_2,\ldots, I_{12}$. We note that while $I_2$ and $I_5$ are {\em unique}, the other invariants are only defined up to addition of terms coming from lower degrees. \section{The invariant quintic} \subsection{Equation} There is a unique (up to scalars) $W(E_6)$-invariant polynomial of degree 5. Written with integer coefficients in the variables $x_i$ it is \begin{equation}\label{eiq1.1} f(x_1,\ldots,x_6)=x_6^5-6x_6^3\gs_1(x) -27x_6\left(\gs_1^2(x)-4\gs_2(x)\right) -648\sqrt{\gs_5(x)}, \end{equation} where $\gs_i(x)$ is the $i$th elementary symmetric polynomial of the $x_1^2,\ldots,x_5^2$, so in particular $\sqrt{\gs_5(x)}=x_1x_2x_3x_4x_5$. The polynomial $f(x)$ displays manifestly the $W(D_5)$-invariance of the quintic. Under the change of variables from the $x_i$ to the $a_i$ of (\ref{eQ3.3}), the equation $g(a)$ can be derived as follows. By (\ref{eQ3.5}), we have $b_i=a_i-{1\over3}(a_1+\cdots+a_6)$, which by equation (\ref{eQ3.4}) can be written $b_i=a_i-h$. The following trick was shown to me by I. Naruki. Consider $\prod_{i=1}^6 a_i -\prod_{i=1}^6 b_i$. This sextic divides the root $h$, and the quotient is $W(E_6)$-invariant. To see this, calculate \begin{eqnarray} a_1\cdot \cdots \cdot a_6 -(b_1\cdot \cdots \cdot b_6) & = & \prod a_i -\prod (a_i-h) \label{eiq1.2} \\ & = & \gs_6(a)-\left[ \gs_6(a)-h\gs_5(a) +h^2\gs_4(a) -h^3 \gs_3(a) +h^4\gs_2(a) -h^5\gs_1(a)\right], \nonumber \end{eqnarray} where here $\gs_i(a)$ are the elementary symmetric functions of the $a_i$. Consequently, \begin{eqnarray*} a_1\cdot \cdots \cdot a_6 -(b_1\cdot \cdots \cdot b_6) & = & h\left(\gs_5(a) -h\gs_4(a) +h^2\gs_3(a)-h^3\gs_2(a) +h^4\gs_1(a)\right), \end{eqnarray*} and since by (\ref{eQ3.4}) $h=-{1\over 3}\gs_1(a)$, this yields \begin{eqnarray} g(a) & = & 81\gs_5(a)+27\gs_4(a)\gs_1(a) +9\gs_3(a)\gs_1^2(a) +3\gs_2(a)\gs_1^3(a) +\gs_1^5(a),\label{eiq1.3} \end{eqnarray} giving the expression of the invariant quintic expressing manifestly the $W(A_5)=\gS_6$-invariance. \begin{definition} \label{diq1.1} The {\em invariant quintic} $\ifmmode {\cal I} \else$\cI$\fi_5$ is the hypersurface of degree 5 $$\ifmmode {\cal I} \else$\cI$\fi_5:=\{x\in {\Bbb P}^5\Big| f(x)=0\}\ifmmode\ \cong\ \else$\isom$\fi \{a\in {\Bbb P}^5\Big| g(a)=0\},$$ where the isomorphism is given by the change of coordinates from the $x_i$ to the $a_i$. \end{definition} \subsection{Singular locus} Because of the equivalence of the coordinates $x_i,\ i=1,\ldots,5$, there are essentially two different partial derivatives of $f$, namely \begin{eqnarray}\label{eiq2.1} j_1: & = & {\del f \over \del x_1}\ifmmode\ \cong\ \else$\isom$\fi \cdots \ifmmode\ \cong\ \else$\isom$\fi {\del f \over \del x_5} \\ j_2: & = & {\del f \over \del x_6}. \nonumber \end{eqnarray} Calculating these forms gives \begin{eqnarray}\label{eiq2.2} -{\del f \over \del x_i} & = & 12x_6^3x_i +54x_6x_i(x_j^2+x_k^2+x_l^2+x_m^2)+648 x_jx_kx_lx_m, \\ {\del f\over \del x_6} & = & 5x_6^4-18x_6^2\gs_1(x) +27(\gs_1^2(x) -4\gs_4(x)). \nonumber \end{eqnarray} These are quartics with manifest $W(D_4)$ and $W(D_5)$ symmetry, respectively. \begin{theorem}\label{tiq4.1} The singular locus of $\ifmmode {\cal I} \else$\cI$\fi_5$ consists of the 120 lines of \ref{i3}, which meet ten at a time in the 36 points of \ref{i6}. \end{theorem} {\bf Proof:} Consider first the hyperplane section $x_6=0$. Then the equations to be solved are \begin{eqnarray}\label{eiq4.1} x_ix_jx_kx_l & = & 0,\quad (i,j,k,l<6); \\ \label{eiq4.2} \gs_1^2(x)-4\gs_2(x) & = & 0. \end{eqnarray} {}From (\ref{eiq4.1}) we get: two of the $x_i$ must vanish, say $x_4,\ x_5$, and then (\ref{eiq4.2}) takes the form \begin{eqnarray} \left(x_1^2+x_2^2+x_3^2\right)^2 -4\left(x_1^2x_2^2+x_1^2x_3^2+x_2^2x_3^2\right) & = & 0, \nonumber \\ (x_1+x_2+x_3)(x_1-x_2-x_3)(x_1+x_2-x_3)(x_1-x_2+x_3) & = & 0 \label{eiq4.3} \end{eqnarray} which splits into a product of four lines. Since the product $x_1\cdot \cdots \cdot x_5=0$ is a coordinate simplex in ${\Bbb P}^4=\{x_6=0\}$, it follows that the 2-simplices of this simplex correspond to planes where two of the coordinates vanish, hence there are ${ 5 \choose 2}=10 $ such 2-simplices; in each we have the four lines given by (\ref{eiq4.3}). This gives the 40 of the 120 lines contained in $x_6=0$. This implies that, in the union of the 27 hyperplanes (\ref{eQ3.3}), the singular locus of $\ifmmode {\cal I} \else$\cI$\fi_5$ consists of 120 lines. Suppose that $x_6\neq 0$. Then the simultaneous vanishing of the partials ${\del f \over \del x_i},\ i=1,\ldots,5$ imply that four of the $x_i$ must vanish, say $x_2=x_3=x_4=x_5=0$. But the intersection of $\ifmmode {\cal I} \else$\cI$\fi_5$ with the line $\{x_2=x_3=x_4=x_5=0\}$ is given by \begin{equation}\label{eiq4.4} x_6^5-6x_6^3x_1^2-27x_6x_1^4=x_6(x_6+i\sqrt{3}x_1)(x_6-i\sqrt{3}x_1) (x_6+3x_1) (x_6-3x_1), \end{equation} and the last two terms are the equations of $b_2$ and $c_{12}$, two other of the 27 hyperplanes of (\ref{eQ3.3}). From this we conclude that any singular point is contained in one of the 27 hyperplanes, and by the above, that the singular locus of $\ifmmode {\cal I} \else$\cI$\fi_5$ consists of the 120 lines, as stated. \hfill $\Box$ \vskip0.25cm The types of singularities are given by the following. \begin{proposition}\label{piq5.1} The singularities of $\ifmmode {\cal I} \else$\cI$\fi_5$ can be described as follows. \begin{itemize}\item[i)] At a generic point $x\in$ one of the 120 lines, a transversal hyperplane section has an ordinary $A_1$-singularity, so the singularity is of type disc$\times A_1$. \item[ii)] At one of the 36 intersection points $p$, the singularity has multiplicity 3, and the tangent cone is of the form $$s_5+s_4t+s_3t^2,$$ where $s_3$ is the cone over the Segre cubic, $s_4=s_3\cdot h_p$, where $h_p$ is the hyperplane of \ref{i1} dual to $p$, and $s_5$ is the cone over the intersection $\ifmmode {\cal I} \else$\cI$\fi_5\cap h_p$. \end{itemize} \end{proposition} {\bf Proof:} i) follows from consideration of generic hyperplane sections of $\ifmmode {\cal I} \else$\cI$\fi_5$, which are quintic threefolds with 120 isolated singularities, so singularities worse than $disc\times A_1$ are impossible. ii) is just a computation, done as follows. Suppose the point is $p=H_{23}=(1,-1,0,0,0,0)$. Then inhomogenizing by setting $t_i=x_i/x_1-tp_i$ (where $p_i$ denotes the $i$th coordinate of $p$), inserting into the equation of $\ifmmode {\cal I} \else$\cI$\fi_5$ gives the stated result. The fact that $s_3$ is the cone over $\ifmmode {\cal S} \else$\cS$\fi_3$ can be seen as follows. We can write \begin{equation}\label{eiq6.1} f=s_5+s_3(th_p +t^2) \end{equation} and it follows that blowing up $\ifmmode {\cal I} \else$\cI$\fi_5$ at $p$ is given by setting $t=\infty$ and that the proper transform of $\ifmmode {\cal I} \else$\cI$\fi_5$ in the exceptional ${\Bbb P}^4$ (of the blow up of ${\Bbb P}^5$ at the point $p$) is a cubic $S_3=0$, where $s_3=0$ is the cone over $S_3=0$. Since there are ten of the 120 lines meeting at $p$, the resolving divisor of the blow up, which is a cubic threefold, will have ten isolated singularities. As mentioned already above, this implies the cubic threefold is isomorphic to $\ifmmode {\cal S} \else$\cS$\fi_3$. One can also see the 15 special hyperplane sections of $\ifmmode {\cal S} \else$\cS$\fi_3$: these are the proper transforms, under the blowing up of $p$, of the 15 of the 36 hyperplanes \ref{i1} passing through the point. The rest is calculation. \hfill $\Box$ \vskip0.25cm We have (using Macaulay) calculated the ideal $\ifmmode \hbox{{\script I}} \else$\scI$\fi(120)$ of the 120 lines, and it turns out to be just the Jacobian ideal of $\ifmmode {\cal I} \else$\cI$\fi_5$. I know of no simple proof of this fact. \subsection{Resolution of singularities} It turns out to be very easy to desingularize $\ifmmode {\cal I} \else$\cI$\fi_5$. By the proof of Proposition \ref{piq5.1}, we know the 36 triple points can be resolved by blowing up each such point $p$. Let $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^{(1)}:\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_5$ denote this blow up of $\ifmmode {\cal I} \else$\cI$\fi_5$. This has the effect of seperating all 120 lines of $\ifmmode {\cal I} \else$\cI$\fi_5$, and the singularities along the lines are just $A_1$, by Proposition \ref{piq5.1}, i). Hence a desingularisation is achieved by resolving each of the 120 lines. There are two possible ways to do this. First, one can blow up the lines in ${{\Bbb P}^5}^{(1)}$=${\Bbb P}^5$ blown up in the 36 points, and take the proper transform of $\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}$; this has the effect of replacing each singular line by a quadric surface bundle, a ${\Bbb P}^1\times {\Bbb P}^1$-bundle, over the line. Hence there are 120 exceptional divisors, each isomorphic to ${\Bbb P}^1\times {\Bbb P}^1\times {\Bbb P}^1$. We call this resolution of $\ifmmode {\cal I} \else$\cI$\fi_5$ the {\em big resolution} and denote it by $\~{\ifmmode {\cal I} \else$\cI$\fi}_5$. Secondly, we can take a small resolution by blowing down one of the fiberings in the exceptional ${\Bbb P}^1\times {\Bbb P}^1$ over a point of the line. In this way, each of the singular lines is replaced by a ${\Bbb P}^1$-bundle over the line, in other words by a ${\Bbb P}^1\times {\Bbb P}^1$. We call this the {\em small resolution} and denote it by $\ifmmode {\cal I} \else$\cI$\fi_5^{(s)}$. Here no further (beyond the 36 on $\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}$) exceptional divisors are introduced. \begin{lemma}\label{liq7.1} The quintic $\ifmmode {\cal I} \else$\cI$\fi_5$ has two resolutions, which we denote by $\~{\ifmmode {\cal I} \else$\cI$\fi}_5$ and $\ifmmode {\cal I} \else$\cI$\fi_5^{(s)}$. On $\~{\ifmmode {\cal I} \else$\cI$\fi}_5$ there are 36+120 exceptional divisors, 36 copies of the resolution of the Segre cubic, and 120 copies of ${\Bbb P}^1\times {\Bbb P}^1\times {\Bbb P}^1$. On $\ifmmode {\cal I} \else$\cI$\fi_5^{(s)}$ there are only 36 exceptional divisors, each a small resolution of the Segre cubic. \end{lemma} \subsection{${\cal I}_5$ is rational} Quite generally, in ${\Bbb P}^5$, taking four ${\Bbb P}^3$'s which meet only in lines, there is a unique line which meets each of them and passes through a given point $P\in {\Bbb P}^5$, namely the line $<\ga,P>\cap<\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda,P>\cap <\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta,P>\cap <\gd,P>$, if $\ga,\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda,\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta,\gd$ denote the ${\Bbb P}^3$'s and $<\ga,P>$ is the hyperplane spanned by $\ga$ and $P$. Now let $P\in \ifmmode {\cal I} \else$\cI$\fi_5$, and choose four of the 15 of the 45 ${\Bbb P}^3$'s through one of the triple points $p$, such that the four ${\Bbb P}^2$'s on $(\ifmmode {\cal S} \else$\cS$\fi_3)_p$ meet each other only in points; then the four ${\Bbb P}^3$'s meet only in lines, and we may apply this reasoning to conclude: \begin{center} \parbox{14cm}{for each $P\in \ifmmode {\cal I} \else$\cI$\fi_5-\{4\ {\Bbb P}^3$'s \}, there is a unique line $L_p$, which joins $P$ and $\ga,\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda,\gamma} \def\gs{\sigma} \def\gS{\Sigma} \def\gz{\zeta,\gd$.} \end{center} Then, fixing a generic hyperplane $F\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^5$. the line $L_p$ intersects $F$ in a single point; we get a rational map: $$\ifmmode {\cal I} \else$\cI$\fi_5 -\ -\ -\ \rightarrow F$$ $$P\mapsto L_p\cap F.$$ We now carry out this argument to derive an explicit rationalisation. I am indebted to B. v. Geemen for help in performing this. We {\em choose} four convienient ${\Bbb P}^3$'s which only meet in lines (although these do not all pass through a point). The four ${\Bbb P}^3$'s will be defined as follows: \begin{equation}\label{eQR.1} \begin{array}{clccc} P_1 & = & \{a_1=b_4=0\} & = & \{l_1=m_1=0\} \\ P_2 & = & \{a_4=b_5=0\} & = & \{l_2=m_2=0\} \\ P_3 & = & \{a_5=b_6=0\} & = & \{l_3=m_3=0\} \\ P_4 & = & \{c_{35}=c_{12}=0\} & = & \{l_4=m_4=0\} \end{array} \end{equation} Letting $F$ be an auxilliary ${\Bbb P}^4$ with homogenous coordinates $(y_0:\ldots:y_4)$, the intersection of the line $<P_1,\ga>\cap \cdots \cap <P_4,\ga>$ with $F$ is given by $$y_0l_i-y_im_i=0,$$ which leads to \begin{equation}\label{eQR.2} \begin{array}{ccl} y_0 & = & m_1\cdots m_4 \\ y_i & = & l_i\cdot m_1\cdots \widehat} \def\tilde{\widetilde} \def\nin{\not\in{m}_i \cdots m_4, \end{array}\end{equation} a system of quartics in ${\Bbb P}^5$, which, when restricted to $\ifmmode {\cal I} \else$\cI$\fi_5$, give the rational map $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha:\ifmmode {\cal I} \else$\cI$\fi_5 - - \ra {\Bbb P}^4(=F)$. Inverting the equations for $x_1,\ldots,x_6$ we get \begin{equation}\label{eQR.3} \begin{array}{ccl} x_1 & = & y_0^6y_1y_3+y_0^5y_1^2y_3-2y_0^6y_2y_3-y_0^5y_1y_2y_3 +y_0^4y_1^2y_2y_3+2y_0^4y_1y_2y_3^2+2y_0^3y_1^2y_2y_3^2 \\ & & \ -y_0^6y_1y_4-y_0^5y_1^2y_4-y_0^5y_1y_2y_4-2y_0^4y_1^2y_2y_4 -y_0^3y_1^2y_2^2y_4-y_0^5y_1y_3y_4-y_0^4y_1^2y_3y_4 \\ & & \ \ -2y_0^4y_1y_2y_3y_4-4y_0^3y_1^2y_2y_3y_4-2y_0^3y_1y_2^2y_3y_4 -3y_0^2y_1^2y_2^2y_3y_4-2y_0^3y_1y_2y_3^2y_4 \\ & & \ \ \ -3y_0^2y_1^2y_2y_3^2y_4-2y_0^2y_1y_2^2y_3^2y_4 -2y_0y_1^2y_2^2y_3^2y_4+y_0^2y_1^2y_2y_3y_4^2 \\ & & \ \ \ \ +y_0y_1^2y_2^2y_3y_4^2+y_0y_1^2y_2y_3^2y_4^2+y_1^2y_2^2y_3^2y_4^2 \\ x_2 & = & y_0^6y_1y_3+y_0^5y_1^2y_3+y_0^5y_1y_2y_3+y_0^4y_1^2y_2y_3 +2y_0^5y_1y_3^2+2y_0^4y_1^2y_3^2+2y_0^4y_1y_2y_3^2 \\ & & \ +2y_0^3y_1^2y_2y_3^2-y_0^6y_1y_4-y_0^5y_1^2y_4-y_0^5y_1y_2y_4 -2y_0^4y_1^2y_2y_4-y_0^3y_1^2y_2^2y_4-5y_0^5y_1y_3y_4 \\ & & \ \ -5y_0^4y_1^2y_3y_4-6y_0^4y_1y_2y_3y_4-8y_0^3y_1^2y_2y_3y_4 -2y_0^3y_1y_2^2y_3y_4-3y_0^2y_1^2y_2^2y_3y_4-4y_0^4y_1y_3^2y_4 \\ & & \ \ \ -4y_0^3y_1^2y_3^2y_4-6y_0^3y_1y_2y_3^2y_4 -7y_0^2y_1^2y_2y_3^2y_4-2y_0^2y_1y_2^2y_3^2y_4 -2y_0y_1^2y_2^2y_3^2y_4+2y_0^5y_1y_4^2 \\ & & \ \ \ \ +2y_0^4y_1^2y_4^2+2y_0^4y_1y_2y_4^2+4y_0^3y_1^2y_2y_4^2+2y_0^2y_1^2y_2^2y_4^2 +4y_0^4y_1y_3y_4^2+4y_0^3y_1^2y_3y_4^2 \\ & & \ \ \ \ \ +6y_0^3y_1y_2y_3y_4^2+9y_0^2y_1^2y_2y_3y_4^2+2y_0^2y_1y_2^2y_3y_4^2 +5y_0y_1^2y_2^2y_3y_4^2+2y_0^3y_1y_3^2y_4^2 +2y_0^2y_1^2y_3^2y_4^2 \\ & & \ \ \ \ \ \ +4y_0^2y_1y_2y_3^2y_4^2+5y_0y_1^2y_2y_3^2y_4^2+2y_0y_1y_2^2y_3^2y_4^2 +3y_1^2y_2^2y_3^2y_4^2 \\ x_3 & = & 2y_0^7y_3+3y_0^6y_1y_3+y_0^5y_1^2y_3+2y_0^6y_2y_3 +3y_0^5y_1y_2y_3+y_0^4y_1^2y_2y_3-2y_0^7y_4-3y_0^6y_1y_4 \\ & & \ -y_0^5y_1^2y_4-2y_0^6y_2y_4-5y_0^5y_1y_2y_4-2y_0^4y_1^2y_2y_4 -2y_0^4y_1y_2^2y_4-y_0^3y_1^2y_2^2y_4-2y_0^6y_3y_4 \\ & & \ \ -3y_0^5y_1y_3y_4-y_0^4y_1^2y_3y_4-4y_0^5y_2y_3y_4 -6y_0^4y_1y_2y_3y_4-2y_0^3y_1^2y_2y_3y_4-2y_0^3y_1y_2^2y_3y_4 \\ & & \ \ \ -y_0^2y_1^2y_2^2y_3y_4+2y_0^3y_1y_2y_3^2y_4 +y_0^2y_1^2y_2y_3^2y_4-2y_0^3y_1y_2y_3y_4^2 \\ & & \ \ \ \ -y_0^2y_1^2y_2y_3y_4^2-2y_0^2y_1y_2^2y_3y_4^2 -y_0y_1^2y_2^2y_3y_4^2-2y_0^2y_1y_2y_3^2y_4^2-y_0y_1^2y_2y_3^2y_4^2 \\ & & \ \ \ \ \ -2y_0y_1y_2^2y_3^2y_4^2-y_1^2y_2^2y_3^2y_4^2 \\ x_4 & = & 2y_0^7y_3+3y_0^6y_1y_3+y_0^5y_1^2y_3+y_0^5y_1y_2y_3 +y_0^4y_1^2y_2y_3-2y_0^5y_1y_3^2-2y_0^4y_1^2y_3^2-2y_0^7y_4 \\ & & \ -3y_0^6y_1y_4-y_0^5y_1^2y_4-3y_0^5y_1y_2y_4-2y_0^4y_1^2y_2y_4 -y_0^3y_1^2y_2^2y_4-2y_0^6y_3y_4-y_0^5y_1y_3y_4 \\ & & \ \ +y_0^4y_1^2y_3y_4-2y_0^4y_1y_2y_3y_4-y_0^2y_1^2y_2^2y_3y_4+4y_0^4y_1y_3^2y_4 +4y_0^3y_1^2y_3^2y_4+2y_0^3y_1y_2y_3^2y_4 \\ & & \ \ \ +3y_0^2y_1^2y_2y_3^2y_4-2y_0^4y_1y_3y_4^2-2y_0^3y_1^2y_3y_4^2 -2y_0^3y_1y_2y_3y_4^2-3y_0^2y_1^2y_2y_3y_4^2 \\ & & \ \ \ \ -y_0y_1^2y_2^2y_3y_4^2-2y_0^3y_1y_3^2y_4^2-2y_0^2y_1^2y_3^2y_4^2 -2y_0^2y_1y_2y_3^2y_4^2-3y_0y_1^2y_2y_3^2y_4^2-y_1^2y_2^2y_3^2y_4^2 \\ x_5 & = & -y_0^6y_1y_3-y_0^5y_1^2y_3-y_0^5y_1y_2y_3-y_0^4y_1^2y_2y_3 +y_0^6y_1y_4+y_0^5y_1^2y_4+2y_0^6y_2y_4+3y_0^5y_1y_2y_4 \\ & & \ +2y_0^4y_1^2y_2y_4+2y_0^4y_1y_2^2y_4+y_0^3y_1^2y_2^2y_4 +3y_0^5y_1y_3y_4+3y_0^4y_1^2y_3y_4+2y_0^4y_1y_2y_3y_4 \\ & & \ \ +4y_0^3y_1^2y_2y_3y_4+2y_0^3y_1y_2^2y_3y_4+y_0^2y_1^2y_2^2y_3y_4 +y_0^2y_1^2y_2y_3^2y_4-2y_0^5y_1y_4^2-2y_0^4y_1^2y_4^2 \\ & & \ \ \ -2y_0^4y_1y_2y_4^2-4y_0^3y_1^2y_2y_4^2-2y_0^2y_1^2y_2^2y_4^2 -2y_0^4y_1y_3y_4^2-2y_0^3y_1^2y_3y_4^2-2y_0^3y_1y_2y_3y_4^2 \\ & & \ \ \ \ -5y_0^2y_1^2y_2y_3y_4^2-3y_0y_1^2y_2^2y_3y_4^2-y_0y_1^2y_2y_3^2y_4^2 -y_1^2y_2^2y_3^2y_4^2 \\ x_6 & = & -3y_0^6y_1y_3-3y_0^5y_1^2y_3-3y_0^5y_1y_2y_3-3y_0^4y_1^2y_2y_3 +3y_0^6y_1y_4+3y_0^5y_1^2y_4+3y_0^5y_1y_2y_4 \\ & & \ +6y_0^4y_1^2y_2y_4+3y_0^3y_1^2y_2^2y_4+3y_0^5y_1y_3y_4+3y_0^4y_1^2y_3y_4 +6y_0^4y_1y_2y_3y_4+6y_0^3y_1^2y_2y_3y_4 \\ & & \ \ +3y_0^2y_1^2y_2^2y_3y_4-3y_0^2y_1^2y_2y_3^2y_4+3y_0^2y_1^2y_2y_3y_4^2 +3y_0y_1^2y_2^2y_3y_4^2+3y_0y_1^2y_2y_3^2y_4^2 \\ & & \ \ \ +3y_1^2y_2^2y_3^2y_4^2 \end{array} \end{equation} a system of octics in ${\Bbb P}^4$, yielding the rational map $$ \psi:{\Bbb P}^4\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_5.$$ These rational maps are morphisms outside of the base locus. \begin{lemma}\label{lQR.1} The base locus of $\varphi} \def\gt{\theta} \def\gT{\Theta} \def\ga{\alpha$ consists of the four ${\Bbb P}^3$'s $P_1,\ P_2,\ P_3,\ P_4$. The base locus of $\psi$ is a surface of degree 32.\hfill $\Box$ \vskip0.25cm \end{lemma} The first statement is clear from construction, while the second is a computation. We performed this with the help of Macaulay to calculate a standard basis of the ideal; the base locus is the intersection of the six octics. \section{Hyperplane sections} \subsection{Reducible hyperplane sections} Consider the hyperplane section $H_5:=\ifmmode {\cal I} \else$\cI$\fi_5\cap \{x_6=0\}$; it is the union of five ${\Bbb P}^3$'s which form a coordinate simplex $x_1\cdot x_2\cdot x_3\cdot x_4\cdot x_5$ in the ${\Bbb P}^4$ given by $\{x_6=0\}$. Now $a_1=-2/3x_6$ and invariance implies that the 27 hyperplane sections $a_i=0,\ b_i=0,\ c_{ij}=0$ all have the same property. Each such hyperplane contains 40 of the 120 ${\Bbb P}^1$'s, which meet six at a time in 20 of the 36 points. Consider three lines in a tritangent, say $(a_1,\ b_2,\ c_{12})$. These three hyperplanes pass through a common ${\Bbb P}^3$, namely $$l_{(12)}:=\{x_6=0,\ x_1=0\}.$$ Such ${\Bbb P}^3$'s therefore correspond to the tritangents and there are 45 such on $\ifmmode {\cal I} \else$\cI$\fi_5$; these are just the 45 ${\Bbb P}^3$'s of \ref{i8}. Hence we have \begin{proposition}\label{pQ7.1} The quintic $\ifmmode {\cal I} \else$\cI$\fi_5$ contains 45 ${\Bbb P}^3$'s, which are cut out by the 27 hyperplane sections (\ref{eQ3.3}), and each such hyperplane section meets $\ifmmode {\cal I} \else$\cI$\fi_5$ in the union of five of the 45 ${\Bbb P}^3$'s. These can be numbered in terms of the tritangents of a cubic surface, i.e., for any 3 lines in a tritagent plane of a cubic surface, the corresponding hyperplanes of (\ref{eQ3.3}) intersect in a common ${\Bbb P}^3$, and this ${\Bbb P}^3$ lies on $\ifmmode {\cal I} \else$\cI$\fi_5$. \end{proposition} Also the intersections of the 45 ${\Bbb P}^3$'s can be described. Each such ${\Bbb P}^3$ contains 16 of the 120 lines which meet in 12 of the 36 points; these 12 points are the vertices of a triad of desmic tetrahedra. Consider the ${\Bbb P}^3$ $l_{(12)}$; the corresponding tritangent meets 12 others, namely (13), (14), (15), (16), (32), (42), (52), (62), (12.34.56), (12.35.46), (12.36.45) and (21), and the 12 ${\Bbb P}^3$'s corresponding to them meet $l_{(12)}$ in a ${\Bbb P}^2$ (the generic intersection has dimension 1). These 12 planes in $l_{(12)}$ form the arrangement $\ifmmode {\cal A} \else$\cA$\fi({\bf D_4})$ in ${\Bbb P}^3$. \subsection{Special hyperplane sections} We now consider the intersections of $\ifmmode {\cal I} \else$\cI$\fi_5$ with the 36 reflection hyperplanes \ref{i1}. Take for example the reflection hyperplane $\{h=0\}$; since $h$ is just a multiple of $\gs_1(a)$ (see (\ref{eQ3.4})), it follows from the equation (\ref{eiq1.3}) that the intersection $\{h=0\}\cap \ifmmode {\cal I} \else$\cI$\fi_5$ is a quintic hypersurface in ${\Bbb P}^4$ with the equation: \begin{equation}\label{eiq8.1} Q_1:=\{h=0\}\cap \ifmmode {\cal I} \else$\cI$\fi_5 = \left\{ \begin{array}{c} \gs_1(a) = 0 \\ \gs_5(a) = 0 \end{array} \right. \end{equation} Comparing with the equation (\ref{e124.1}), we see that this is a copy of the Nieto quintic! By symmetry, each of the 36 hyperplane sections is isomorphic to this one, and we denote them by \begin{equation}\label{eiq8.2} T=\{h=0\}\cap \ifmmode {\cal I} \else$\cI$\fi_5,\ T_{ij}=\{h_{ij}=0\}\cap \ifmmode {\cal I} \else$\cI$\fi_5,\ T_{ijk}=\{ h_{ijk}=0\}\cap \ifmmode {\cal I} \else$\cI$\fi_5. \end{equation} So we have: \begin{proposition}\label{piq8.1} There are 36 copies of the Nieto quintic (\ref{e124.1}) on $\ifmmode {\cal I} \else$\cI$\fi_5$. \end{proposition} We can determine the singular locus of these hyperplane sections, independently of the discussion given in section \ref{section4.1}. The reflection hyperplane contains 20 of the 120 lines, which meet in 15 of the 36 points (corresponding to the 15 roots of an ${\bf A_5}$ subsystem), so the quintic threefold has 20 singular lines, with 15 singular points of multiplicity 3. In fact, the resolving divisor of each of these 15 points is a four-nodal cubic surface, which is a hyperplane section of the Segre cubic $\ifmmode {\cal S} \else$\cS$\fi_3$ (see the discussion following Problem \ref{p132.1}, ii)). Furthermore, recalling that there are ten of the 120 lines which pass through the triple point which is {\em dual} to the given reflection hyperplane, each such intersects the reflection hyperplane transversally, giving the ten isolated ordinary double points on that quintic (see Proposition \ref{p124.1}), and in some sense ``explains'' these isolated singularities. \subsection{Generic hyperplane sections} A generic hyperplane section is a quintic threefold in ${\Bbb P}^4$ with 120 nodes. This is a fascinating family of Calabi-Yau threefolds, which has a beautiful geometric configuration associated with it, in some sense ``dual'' to the configuration of the 27 lines on a cubic surface. \begin{proposition} Let $H\in {\Bbb P}^5$ be a generic hyperplane and let $Q_H=\ifmmode {\cal I} \else$\cI$\fi_5\cap H$ be the hyperplane section. Then we have \begin{itemize}\item[1)] There are 45 ${\Bbb P}^2$'s on $Q_H$, which are cut out by 27 hyperplanes; these could appropriately be called {\em quintangent planes}. \item[2)] The group of incidence preserving permutations of the 45 ${\Bbb P}^2$'s is $W(E_6)$; this is also the group of incidence preserving permutations of the 27 hyperplanes. \item[3)] There are 36 hyperplane sections of $Q_H$, each of which is a 20-nodal quintic surface. \item[4)] The 120 nodes of $Q_H$ form an orbit under $W(E_6)$. \end{itemize}\end{proposition} {\bf Proof:} For any of the 45 ${\Bbb P}^3$'s in $\ifmmode {\cal I} \else$\cI$\fi_5$ and hyperplane section $H$, it holds that ${\Bbb P}^3\cap H={\Bbb P}^2\subset} \def\nni{\supset} \def\und{\underline H\cap \ifmmode {\cal I} \else$\cI$\fi_5=Q_H$, showing 1). The second point is evident, and in a sense ``dual'' to the situation with cubic surfaces. We have seen that a special hyperplane section as in (\ref{eiq8.2}) is isomorphic to the Nieto quintic and has 20 singular lines in its singular locus; therefore any generic hyperplane section has exactly 20 nodes. 4) follows since the $W(E_6)$ orbit consisting of the 120 lines, restricted to the hyperplane section is still an orbit. \hfill $\Box$ \vskip0.25cm We now consider some of the invariants of the nodal quintic threefolds. Let $V$ denote a nodal quintic, $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{V}\longrightarrow} \def\sura{\twoheadrightarrow V$ a small resolution and $\tilde{V}\longrightarrow} \def\sura{\twoheadrightarrow V$ a big resolution. Letting $s$ denote the number of nodes, the betti numbers are \begin{equation}\label{e8.3.1}\begin{minipage}{15cm}$$\begin{array}{ll} b_1(V)=1=b_1(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{V}), & b_2(\tilde{V})=1+d+s; \\ b_2(V)=1+d=b_2(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{V}), & b_4(\tilde{V})=1+d+s; \\ b_3(V)=b_3(V_t)-s+d, & b_3(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{V})=b_3(V)-s+d=b_3(\tilde{V}), \end{array}$$ \end{minipage}\end{equation} where $V_t$ is a smooth hypersurface of same degree as $V$ and $d$ is the {\it defect}. The defect may be calculated by the following result. \begin{theorem}[\cite{W}, p.~27]\label{twerner} Let $V\in {\Bbb P}^4$ be a nodal hypersurface of degree $n\geq 3$. Then $$\hbox{dim}} \def\deg{\hbox{deg}(\ifmmode \hbox{{\script P}} \else$\scP$\fi_{2n-5}(V))=\hbox{dim}} \def\deg{\hbox{deg}\left\{ \parbox{6cm}{homogenous polynomials of degree $2n-5$ in ${\Bbb P}^4$, containing all nodes of $V$}\right\} = {2n-1 \choose 4} -s +d.$$ \end{theorem} Applied to the case at hand, we need the dimension of the space of {\it quintics} vanishing at all the nodes. Clearly this is the degree five component in the ideal of the 120 points. As we mentioned above, we {\it know} the ideal of the 120 lines (it is the Jacobian ideal of $\ifmmode {\cal I} \else$\cI$\fi_5$ $\ifmmode \hbox{{\script J}} \else$\scJ$\fi ac(\ifmmode {\cal I} \else$\cI$\fi_5)$), so we know also the ideal of the 120 points; it is the restriction of $\ifmmode \hbox{{\script J}} \else$\scJ$\fi ac(\ifmmode {\cal I} \else$\cI$\fi_5)$ to the hyperplane, generated by six quartics. \begin{proposition} The dimension of the space $\ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_H)$ is 30. \end{proposition} {\bf Proof:} Each of the quartics (which are clearly independent for a generic hyperplane $H$) of $\ifmmode \hbox{{\script J}} \else$\scJ$\fi ac(\ifmmode {\cal I} \else$\cI$\fi_5)$ can be multiplied by any hyperplane, giving a quintic which contains the 120 nodes. The set of hyperplanes is $({\Bbb P}^5)^{\vee}$, so the dimension of $\ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_H)$ is $6\cdot 5=30$. \hfill $\Box$ \vskip0.25cm We can now apply Theorem \ref{twerner} to calculate the defect $d$ for $Q_H$. The formula is $126-120+d=30$, from which is follows that $d=24$. \begin{corollary} The small resolutions $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H$ of the quintic threefolds $Q_H$ have the following betti and Hodge numbers: $$\begin{array}{lll} b_2(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H) = 25, & h^{1,1}=25, & \\ b_3(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H)=12=2+2h^{2,1}, & h^{2,1} =5, & e=2h^{1,1}-2h^{2,1}=40. \end{array}$$ \end{corollary} {\bf Proof:} Insertion of $d=24$ in (\ref{e8.3.1}).\hfill $\Box$ \vskip0.25cm In the well known manner for Calabi-Yau threefolds the isomorphism $H^2(V,\Omega} \def\go{\omega} \def\gm{\mu} \def\gn{\nu} \def\gr{\rho^1)\ifmmode\ \cong\ \else$\isom$\fi H^1(V,\Theta)$ identifies the Hodge space $H^{2,1}$ with the space of infinitesimal deformations of $V$, $H^1(V,\Theta)$. This is by the above five-dimensional, hence the moduli space of these 120 nodal quintics (a Zariski open subset of $({\Bbb P}^5)^{\vee}$) is also a global space of complex deformations of the small resolution. We can describe the space $H^{2,1}$ more concretely as follows. Consider the space $\ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_H)$; let $\ifmmode \hbox{{\script J}} \else$\scJ$\fi\subset} \def\nni{\supset} \def\und{\underline \ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_H)$ be the subspace generated by the Jacobi ideal of $Q_H$; since $Q_H$ has five partial derivatives, $\ifmmode \hbox{{\script J}} \else$\scJ$\fi$ is $5\cdot 5=25$ dimensional, and $\ifmmode \hbox{{\script J}} \else$\scJ$\fi$ cannot contribute to infinitesimal deformations, so we have $$H^{2,1}(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H)\ifmmode\ \cong\ \else$\isom$\fi \ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_H)/\ifmmode \hbox{{\script J}} \else$\scJ$\fi.$$ As a final remark consider the Picard group $\hbox{Pic}} \def\Jac{\hbox{Jac}(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H)$ and the orthocomplement of the hyperplane section $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H)$. Then $rk_{\integer}\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H)=24$, and the 45 ${\Bbb P}^2$'s give us privledged representatives in $\hbox{Pic}} \def\Jac{\hbox{Jac}^0$; the 27 hyperplanes represent relations, so we have an exact sequence $$\integer^{27}\longrightarrow} \def\sura{\twoheadrightarrow \integer^{45}\longrightarrow} \def\sura{\twoheadrightarrow \hbox{Pic}} \def\Jac{\hbox{Jac}^0(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_H) \longrightarrow} \def\sura{\twoheadrightarrow 1,$$ and the kernel is six-dimensional. The sum sequence is then \begin{equation}\label{e48a.1} 1 \longrightarrow} \def\sura{\twoheadrightarrow \integer^6 \longrightarrow} \def\sura{\twoheadrightarrow \integer^{27}\longrightarrow} \def\sura{\twoheadrightarrow \integer^{45} \longrightarrow} \def\sura{\twoheadrightarrow \integer^{24}\longrightarrow} \def\sura{\twoheadrightarrow 1, \end{equation} and this is really dual to the sequence (\ref{eB3.2}) for cubic surfaces. \begin{remark} The period map for this five-dimensional family of Calabi-Yau threefolds maps to the domain $\ifmmode {\cal D} \else$\cD$\fi = Sp(6,\fR)/U(1)\times U(5)$. Note that any hyperplane passing through one of the 45\ ${\Bbb P}^3$'s will intersect $\ifmmode {\cal I} \else$\cI$\fi_5$ in the union of that ${\Bbb P}^3$ and a residual quartic; clearly these constitute the set of cusps for the period map, i.e., on the 45 lines in $({\Bbb P}^5)^{\vee}$ (the dual ${\Bbb P}^5$) which parameterise the set of hyperplanes passing through one of the 45 ${\Bbb P}^3$'s, the period map maps to the boundary of the domain $\ifmmode {\cal D} \else$\cD$\fi$ above. These 45 one-dimensional cusps meet in 27 points, i.e., zero-dimensional cusps, which correspond to the 27 hyperplane sections which split into the union of five ${\Bbb P}^3$'s. But we can say more. Noting that, excepting the hyperplanes above, all hyperplane sections of $\ifmmode {\cal I} \else$\cI$\fi_5$ are irreducible quintics, the worst that can happen is that the hyperplane passes through one of the 36 triple points of $\ifmmode {\cal I} \else$\cI$\fi_5$. We will see below that these are still Calabi-Yau (Proposition \ref{p158.1}), hence {\it not} contained in the boundary. \end{remark} \subsection{Tangent hyperplane sections} We now consider the case of a hyperplane tangent to $\ifmmode {\cal I} \else$\cI$\fi_5$ at a point $p\in \ifmmode {\cal I} \else$\cI$\fi_5$. In this case the section $Q_p$ aquires an additional node. Note that the 121 nodes fall into two ``orbits'', one set of 120 on which $W(E_6)$ acts as a permutation group, and the additional point $p$. For a 121-nodal quintic the same calculation as above gives $e(Q_p)=42$, $h^{2,1}=4,\ h^{1,1}=25$. It follows that the $H_4(Q_p,\rat)$ is the same as for $Q_x,\ x\in {\Bbb P}^5$ generic. The difference to the generic case is in $H_3$, more precisely in $H^{2,1}$. Indeed, we now require $\ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_p)$, that is, quintics through all 121 nodes, so as opposed to the general case, we now only have, for each of the five quartics in the Jacobi ideal of $Q_p$, since each contains $p$, a five-dimensional family of quintics, as above. But for the quartics through the 120 nodes which are {\it not} in the Jacobi ideal, we must take a hyperplane {\it through the point} $p$, so \begin{proposition} For a 121-nodal quintic $Q_p$, $p\in \ifmmode {\cal I} \else$\cI$\fi_5$, we have $\hbox{dim}} \def\deg{\hbox{deg}\ifmmode \hbox{{\script P}} \else$\scP$\fi_5(Q_p)=5\cdot 5 + 1\cdot 4 =29$. \hfill $\Box$ \vskip0.25cm \end{proposition} We can now apply Theorem \ref{twerner} to calculate the defect: $$d=29-126+121 = 24.$$ \begin{corollary} The betti numbers for the small resolution $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_p$ are $$\begin{array}{ll}b_2(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_p) =25, & h^{1,1} = 25, \\ b_3(\widehat} \def\tilde{\widetilde} \def\nin{\not\in{Q}_p) = 10, & h^{2,1}=4,\ \ \ \ e=42. \\ \end{array}$$ \end{corollary} We remark that since $h^{1,1}$ is still 25, the sequence (\ref{e48a.1}) still holds for $Q_p$. \section{Birational maps and the projection from a triple point} \subsection{The cuspidal model} First we recall the notations $\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}$ for the blow up of $\ifmmode {\cal I} \else$\cI$\fi_5$ at the 36 triple points, $\~{\ifmmode {\cal I} \else$\cI$\fi}_5$ for the big resolution of $\ifmmode {\cal I} \else$\cI$\fi_5$, and $\ifmmode {\cal I} \else$\cI$\fi_5^{(s)}$ for the small resolution. Note that on $\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}$, each of the 120 lines has normal bundle $\ifmmode {\cal O} \else$\cO$\fi(-2)^{\oplus 3}$, hence each line can be blown down to an isolated singular point. \begin{definition}\label{dq4.1} Consider the following birational transformation of $\ifmmode {\cal I} \else$\cI$\fi_5$: \begin{itemize}\item[i)] Blow up the 36 triple points, $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^{(1)}:\ifmmode {\cal I} \else$\cI$\fi_5^{(1)} \longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_5$; \item[ii)] Blow down the proper transforms of the 120 lines to 120 isolated singularities, $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta^{(2)}:\ifmmode {\cal I} \else$\cI$\fi_5^{(1)}\longrightarrow} \def\sura{\twoheadrightarrow \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$. \end{itemize} Step ii) defines the {\em cuspidal model} $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$. \end{definition} This is a four-dimensional analogue of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$ of (\ref{e127.2}). Indeed, for each of the 36 hyperplane sections of Proposition \ref{piq8.1}, the proper transform on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$ is isomorphic to $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$: \begin{lemma}\label{lq4.1} Let $T\ifmmode\ \cong\ \else$\isom$\fi \ifmmode {\cal N} \else$\cN$\fi_5$ be one of the 36 special hyperplane sections of (\ref{eiq8.2}), and let $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{T}$ denote its proper transform on $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$. Then $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{T}\ifmmode\ \cong\ \else$\isom$\fi \widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$. \end{lemma} {\bf Proof:} Just check that the steps i) and ii) of Definition \ref{dq4.1}, when restricted to $T$, coincide with those of (\ref{e127.2}).\hfill $\Box$ \vskip0.25cm Let us mention that $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$ ``looks like'' a ball quotient too, at least assuming a positive answer to Problem \ref{p132.1}. We explain what ``looks like'' means in the following items. \begin{itemize} \item[I1] Each isolated singularity is resolved by a ${\Bbb P}^1\times {\Bbb P}^1\times {\Bbb P}^1$; the arrangement induced in each by the proper transforms of the 36 hyperplanes and 36 exceptional divisors is a {\em product}, consisting of three fibres in each fibering (i.e., $\{\hbox{3 points}\}\times {\Bbb P}^1\times {\Bbb P}^1,\ {\Bbb P}^1\times \{\hbox{3 points}\}\times {\Bbb P}^1,\ {\Bbb P}^1\times {\Bbb P}^1\times \{\hbox{3 points}\}$, see \ref{i3}). Hence this can be covered in an equivariant way by $E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}\times E_{\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta}$ (see Lemma \ref{l115.2}). \item[I2] The proper transforms of the 36 hyperplane sections of Proposition \ref{piq8.1} are by Lemma \ref{lq4.1} isomorphic to $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal N} \else$\cN$\fi}_5$, so, if the Problem \ref{p132.1} has an affirmative solution, these are ball quotients, with cusps being those isolated singularities of $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{\ifmmode {\cal I} \else$\cI$\fi}_5$ which are contained in the given $\widehat} \def\tilde{\widetilde} \def\nin{\not\in{T}$. \item[I3] Consider the 45 ${\Bbb P}^3$'s of Proposition \ref{pQ7.1}. These are (the proper transforms of) the 45 ${\Bbb P}^3$'s of \ref{i8}. These ${\Bbb P}^3$'s are also ball quotients, in fact in two different ways. \begin{itemize}\item[1)] There is a cover $Y\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$, branched over the arrangement $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_4}))$ in ${\Bbb P}^3$, which is a ball quotient. This example can be derived from the solution 4) of (\ref{e111d.1}) by means of the natural squaring map $m_2:{\Bbb P}^3\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3,\ (x_0:\ldots :x_3)\mapsto (x_0^2:\ldots : x_3^2)$. Then the arrangement $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_4}))$ is the pullback under $m_2$ of the six symmetry planes of the tetrahedron in the arrangement $\ifmmode {\cal A} \else$\cA$\fi(W({\bf A_4}))$, and pulling back the solution 4), we get the cover $Y\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$, branched along $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_4}))$ (with branching degree 3 at each hyperplane), which is a ball quotient by a fix point free group. \item[2)] There is a cover $Z\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$, branched along the arrangement $\ifmmode {\cal A} \else$\cA$\fi(W({\bf F_4}))$ in ${\Bbb P}^3$ (but not a Fermat cover), which is a ball quotient; this example is explained in \cite{hunt}, Thm.~7.6.5, and is the {\it only} known ball quotient related to a plane arrangement in ${\Bbb P}^4$ which does {\it not} derive from those given by solutions of the hypergeometric differential equation. \end{itemize} Both of the arrangements mentioned, $\ifmmode {\cal A} \else$\cA$\fi(W({\bf D_4}))$ and $\ifmmode {\cal A} \else$\cA$\fi(W({\bf F_4}))$, arise naturally on the 45 ${\Bbb P}^3$'s: the first is the intersection with the 27 hyperplanes, the second is the intersection with the 36 hyperplanes. \end{itemize} \subsection{Projection from a triple point} Let $p\in \ifmmode {\cal I} \else$\cI$\fi_5$ be one of the 36 triple points, and let $h_p$ be the dual hyperplane (one of the 36 ${\Bbb P}^4$'s of \ref{i1}). The projection of ${\Bbb P}^5$ from $p$ is defined as follows. Consider the ${\Bbb P}^4$ of all lines through $p$; this is just the dual $h_p$, and each line $l_p$ through $p$ corresponds to a unique point of $h_p$ (its intersection with $h_p$). Since any point $x$ of ${\Bbb P}^5$ is on a unique line $(l_x)_p$ through $p$, the map \begin{eqnarray}\label{e154.1} \pi_p:{\Bbb P}^5 & \longrightarrow} \def\sura{\twoheadrightarrow & h_p \\ x & \mapsto & (l_x)_p\cap h_p \nonumber \end{eqnarray} gives the {\em projection from $p$}. Restricting to $\ifmmode {\cal I} \else$\cI$\fi_5$ this gives a generically finite (rational) map, which we also denote by $\pi_p$, $\pi_p:\ifmmode {\cal I} \else$\cI$\fi_5- - \ra h_p$. \begin{lemma}\label{l154.1} $\pi_p:\ifmmode {\cal I} \else$\cI$\fi_5- - \ra h_p$ is generically a double cover. \end{lemma} {\bf Proof:} Since the triple point has multiplicity 3, a generic line will meet $\ifmmode {\cal I} \else$\cI$\fi_5$ in $(5-3)=2$ further points. \hfill $\Box$ \vskip0.25cm \begin{lemma}\label{l154.2} $\pi_p:\ifmmode {\cal I} \else$\cI$\fi_5- - \ra h_p$ is a quotient map by the group $G_p\ifmmode\ \cong\ \else$\isom$\fi \integer/2\integer$ generated by the reflection $\gs_p$ on the root $p$. \end{lemma} {\bf Proof:} The reflection $\gs_p$ fixes $h_p$; it is the inversion ($(z_0:z_1) \mapsto (z_1:z_0)$) on any line $l_p$ through $p$, where the homogenous coordinates are choosen such that $l_p\cap h_p=(1:1)$. Since $\ifmmode {\cal I} \else$\cI$\fi_5$ is mapped by $\gs_p$ onto itself, it follows that two points of $\ifmmode {\cal I} \else$\cI$\fi_5\cap l_p$ are related by inversion on $l_p$. So the group action is manifest. \hfill $\Box$ \vskip0.25cm We now describe how to make $\pi_p$ into a {\em morphism}. First of all, one must blow up $p$; let $\varrho} \def\ge{\varepsilon} \def\gx{\xi} \def\gy{\eta_p:\ifmmode {\cal I} \else$\cI$\fi_{5,p}\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_5$ denote this blow up. Let $(\ifmmode {\cal S} \else$\cS$\fi_3)_p$ be the copy of $\ifmmode {\cal S} \else$\cS$\fi_3$ which is the exceptional divisor at $p$. For any $x\in (\ifmmode {\cal S} \else$\cS$\fi_3)_p$, the line $(l_x)_p$ through $p$ and intersecting $h_p$ in the Segre cubic there, is tangent to $\ifmmode {\cal I} \else$\cI$\fi_5$ {\em at the triple point} $p$. Secondly, certain subvarieties get {\em blown down}. Indeed, suppose $(l_x)_p$ is {\em contained in} $\ifmmode {\cal I} \else$\cI$\fi_5$ for some $x\in \ifmmode {\cal I} \else$\cI$\fi_5$. Then, clearly, $(l_x)_p\mapsto (l_x)_p\cap h_p$, the whole line maps to a point, or in other words, gets blown down. \begin{lemma}\label{l155.1} The projection $\pi_p:\ifmmode {\cal I} \else$\cI$\fi_5- - \ra h_p$, which is well-defined on $\ifmmode {\cal I} \else$\cI$\fi_{5,p}$, blows down all linear subspaces on $\ifmmode {\cal I} \else$\cI$\fi_5$ which pass through $p$, and is a double cover outside the union $\ifmmode \hbox{{\script L}} \else$\scL$\fi_p$ of all such linear subspaces on $\ifmmode {\cal I} \else$\cI$\fi_5$ passing through $p$.\hfill $\Box$ \vskip0.25cm \end{lemma} We now describe $\ifmmode \hbox{{\script L}} \else$\scL$\fi_p$. Recall that the linear subspaces on $\ifmmode {\cal I} \else$\cI$\fi_5$ are the 45 ${\Bbb P}^3$'s and their intersections. Hence $\ifmmode \hbox{{\script L}} \else$\scL$\fi_p$ consists of all the ${\Bbb P}^3$'s and their intersections, which pass through $p$. Recall from \ref{i6} that this is the set of 15 of the 45 ${\Bbb P}^3$'s of Proposition \ref{pQ7.1}. Therefore, we get \begin{lemma}\label{l155.2} The projection $\pi_p,p:\ifmmode {\cal I} \else$\cI$\fi_{5,p}\longrightarrow} \def\sura{\twoheadrightarrow h_p$ blows down the union of 15 ${\Bbb P}^3$'s to the 15 planes in $h_p$ which are the intersection of $\ifmmode {\cal S} \else$\cS$\fi_3$ and $\ifmmode {\cal N} \else$\cN$\fi_5$. \end{lemma} Now let $X=\ifmmode {\cal I} \else$\cI$\fi_{5,p}^{\%}$, the double cover of $h_p$ branched along the union of $\ifmmode {\cal S} \else$\cS$\fi_3$ and $\ifmmode {\cal N} \else$\cN$\fi_5$ (which is of degree 8, so a double cover exists). $X$ is clearly {\em singular along} the 15 planes. Indeed: \begin{lemma}\label{l155.3} $\pi_p,p:\ifmmode {\cal I} \else$\cI$\fi_{5,p}\longrightarrow} \def\sura{\twoheadrightarrow h_p$ factors over $\ifmmode {\cal I} \else$\cI$\fi_{5,p}^{\%}$, and $\Pi:X\longrightarrow} \def\sura{\twoheadrightarrow h_p$ is the double cover of ${\Bbb P}^4$ branched along the union $\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5$. \end{lemma} {\bf Proof:} This follows from the discussion above; the branch locus $\ifmmode {\cal R} \else$\cR$\fi$ is the set: $$\ifmmode {\cal R} \else$\cR$\fi=\{x\in \ifmmode {\cal I} \else$\cI$\fi_{5,p}\Big| (l_x)_p \hbox{ is tangent to $\ifmmode {\cal I} \else$\cI$\fi_5$ at $x$}\}.$$ This happens if either \begin{itemize}\item[i)] $x\in h_p$, since then $x$ is fixed by $\gs_p$; \item[ii)] $x\in (\ifmmode {\cal S} \else$\cS$\fi_3)_p$, the exceptional divisor over $p$. \end{itemize} Therefore $\ifmmode {\cal R} \else$\cR$\fi=\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5$. By Lemmas \ref{l155.1} and \ref{l155.2}, 15 ${\Bbb P}^3$'s are blown down to ${\Bbb P}^2$'s, and outside of this locus, $\Pi$ is 2:1. \hfill $\Box$ \vskip0.25cm \subsection{Double octics and quintic hypersurfaces} With the result of Lemma \ref{l155.3} at hand, we can get a new slant on the quintic threefolds which are hyperplane sections of $\ifmmode {\cal I} \else$\cI$\fi_5$. For this, consider a hyperplane section of the cover $\Pi:X\longrightarrow} \def\sura{\twoheadrightarrow h_p$, that is, let $H\subset} \def\nni{\supset} \def\und{\underline h_p$ be a hyperplane, and let $X_H$ be its inverse image in $X$: $$\Pi_H:X_H\longrightarrow} \def\sura{\twoheadrightarrow H,$$ a double cover of ${\Bbb P}^3$. The branch locus is $H\cap (\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5)$, which is the union of a cubic and a quintic surface in ${\Bbb P}^3$. Note the $H\cap \{\hbox{ one of the 15 ${\Bbb P}^2$'s $\subset} \def\nni{\supset} \def\und{\underline \ifmmode {\cal S} \else$\cS$\fi_3\cap \ifmmode {\cal N} \else$\cN$\fi_5$}\}$ is a {\em line}, contained in both $H\cap \ifmmode {\cal S} \else$\cS$\fi_3$ and in $H\cap \ifmmode {\cal N} \else$\cN$\fi_5$. In other words, $H\cap (\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5)=S_H\cup Q_H$, where $S_H$ is the cubic surface, $Q_H$ is the quintic surface, and $$S_H\cap Q_H=\{\hbox{15 lines}\}.$$ \begin{proposition}\label{p158.1} Let $X_H=\Pi^{-1}(H)$, the double cover of ${\Bbb P}^3$ branched along $S_H\cup Q_H$. Then there is a canonical model $\-X_H$ of $X_H$ which is Calabi-Yau. \end{proposition} {\bf Proof:} We know the resolution of $X$; it is given by ``inverting'' the projection from the node, by blowing up along the 15 planes $\ifmmode {\cal S} \else$\cS$\fi_3\cap \ifmmode {\cal N} \else$\cN$\fi_5$, yielding $\ifmmode {\cal I} \else$\cI$\fi_{5,p}$. Let $\-X_H$ be the proper transform of $X_H$ in $\ifmmode {\cal I} \else$\cI$\fi_{5,p}$. Assuming $H$ to be sufficiently general, $\-X_H$ clearly has canonical singularities (as $\ifmmode {\cal I} \else$\cI$\fi_{5,p}$ does), so we must only show that it is Calabi-Yau. We note, however, that $\-X_H$ is (the proper transform on $\ifmmode {\cal I} \else$\cI$\fi_{5,p}$ of) a hyperplane section of $\ifmmode {\cal I} \else$\cI$\fi_5$! This is because the degree is invariant under projection, hence under $\Pi$. But this is a hyperplane section of $\ifmmode {\cal I} \else$\cI$\fi_5$ through the triple point $p$. Hence the proper transform on $\-X_H$ of the exceptional divisor $(\ifmmode {\cal S} \else$\cS$\fi_3)_p$ is a hyperplane section of $\ifmmode {\cal S} \else$\cS$\fi_3$, i.e., a (generically smooth) cubic surface. This singularity is known to be canonical, and $\-X_H$ is, just as a nodal quintic, canonically Calabi-Yau. \hfill $\Box$ \vskip0.25cm \begin{table} \caption{\label{table25} Degenerations of double octics and quintic hypersurfaces} \begin{minipage}{16.5cm} \begin{center} \fbox{\begin{minipage}{12cm}\begin{center} Space of all quintic hypersurfaces 101-dimensional \end{center} \end{minipage}} $$\cup$$ \fbox{\begin{minipage}{6cm}\begin{center} 120-nodal quintics 5-dimensional \end{center} \end{minipage}} $$\cup$$ \fbox{\begin{minipage}{5cm}\begin{center} quintic hypersurfaces with 111 nodes and one multiplicity 3 singular point 4-dimensional \end{center}\end{minipage}} $$\|$$ \fbox{\begin{minipage}{5cm}\begin{center} double cover $Y\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^3$, branched over $S\cup Q$ $S\cap Q=\{\hbox{ 15 lines}\}$ \end{center} \end{minipage}} $$\cap$$ \fbox{\parbox{6cm}{double cover branched over cubic and quintic, such that $S\cup Q$ is stable}} $$\cap$$ \fbox{\begin{minipage}{12cm}\begin{center} Space of all double octics 149-dimensional \end{center}\end{minipage}} \end{center} \end{minipage} \end{table} \begin{corollary}\label{c158.1} The family of hyperplane sections of $\ifmmode {\cal I} \else$\cI$\fi_5$ passing through one of the 36 triple points $p$ is, via projection, a family of Calabi-Yau threefolds which are degenerations of double octics. \end{corollary} It is natural to ask the meaning of this in terms of variations of Hodge structures. Recalling that a Type III degeneration of a K3 surface, corresponding to a zero-dimensional boundary component of the period domain, is one like a quartic degenerating into four planes, it is natural to ask \begin{question} Is a double cover of ${\Bbb P}^3$ branched over the union of a cubic and a quintic a semistable degeneration of a double octic? \end{question} \begin{remark} There is a notion of ``connecting'' moduli spaces of CY threefolds by degenerations, and the Corollary \ref{c158.1} shows that the moduli space of quintic hypersurfaces in ${\Bbb P}^4$ and the moduli space of double octics are connected; the birational transformations which are required for such ``connections'' are given here by projection in projective space, very geometric. \end{remark} In Table \ref{table25} we give a rough description of these relations. \subsection{The dual picture} Now consider $\ga(\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5)$, with $\ga$ the map (\ref{e126.1}) given by the quadrics on the ten nodes of $\ifmmode {\cal S} \else$\cS$\fi_3$. By Theorem \ref{t122a.1} and by definition of $\ifmmode {\cal W} \else$\cW$\fi_{10}$ (\ref{e127.1}), we have \begin{equation}\label{e156.1} \ga(\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5) = \ifmmode {\cal I} \else$\cI$\fi_4\cup \ifmmode {\cal W} \else$\cW$\fi_{10}, \end{equation} and by Theorem \ref{t127.1}, the intersection $\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10}$ consist of 10 quadric surfaces. Define $\ifmmode {\cal W} \else$\cW$\fi$ to be the double cover of ${\Bbb P}^4$ branched along $\ifmmode {\cal W} \else$\cW$\fi_{10}$: \begin{equation}\label{e156.2} \tau:\ifmmode {\cal W} \else$\cW$\fi\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4=(h_p)^{\vee}. \end{equation} We may consider the fibre square: \begin{equation}\label{e156.3} \begin{array}{rcl} {\cal Z} & \longrightarrow} \def\sura{\twoheadrightarrow & {\cal Y} \\ \downarrow & & \downarrow \pi \\ \tau:{\cal W} & \longrightarrow} \def\sura{\twoheadrightarrow & {\Bbb P}^4 \end{array} \end{equation} where $\pi:\ifmmode {\cal Y} \else$\cY$\fi\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ is defined in Definition \ref{d133.1}. Then $\pi_{\ifmmode {\cal Z} \else$\cZ$\fi}:\ifmmode {\cal Z} \else$\cZ$\fi\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ is a Galois cover with Galois group $G_{\ifmmode {\cal Z} \else$\cZ$\fi}\ifmmode\ \cong\ \else$\isom$\fi \integer/2\integer\times \integer/2\integer$. Let $H\ifmmode\ \cong\ \else$\isom$\fi \integer/2\integer\subset} \def\nni{\supset} \def\und{\underline G_{\ifmmode {\cal Z} \else$\cZ$\fi}$ be the diagonal subgroup; it is a normal subgroup, and we may form the quotient $$\eta:\ifmmode {\cal Z} \else$\cZ$\fi\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal Z} \else$\cZ$\fi',\quad \ifmmode {\cal Z} \else$\cZ$\fi'=\ifmmode {\cal Z} \else$\cZ$\fi/H.$$ \begin{lemma}\label{l156.1} $\pi_{\ifmmode {\cal Z} \else$\cZ$\fi}:\ifmmode {\cal Z} \else$\cZ$\fi\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ factors over $\eta$, and $\eta':\ifmmode {\cal Z} \else$\cZ$\fi'\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ is a double cover, hence Galois. \end{lemma} {\bf Proof:} This is a general fact about fibre squares of double covers like (\ref{e156.3}). \hfill $\Box$ \vskip0.25cm \begin{theorem}\label{t156.1} The rational map $\ga$ induces a rational map of the double covers $\Pi:X\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ of Lemma \ref{l155.3} and $\eta':\ifmmode {\cal Z} \else$\cZ$\fi'\longrightarrow} \def\sura{\twoheadrightarrow {\Bbb P}^4$ of Lemma \ref{l156.1}. Furthermore, the rational map $$\Xi:X- - \ra \ifmmode {\cal Z} \else$\cZ$\fi'$$ is $\gS_6$-equivariant. \end{theorem} {\bf Proof:} Recall from Lemma \ref{l25aux} that $\ga$ blows up the ten nodes and blows down the tangent cones of the nodes to the quadric surfaces (on $\ifmmode {\cal I} \else$\cI$\fi_4$). $\Pi:X\longrightarrow} \def\sura{\twoheadrightarrow h_p$ is a double cover branched along $\ifmmode {\cal S} \else$\cS$\fi_3\cup \ifmmode {\cal N} \else$\cN$\fi_5$, and we can calculate the image of the branch locus under $\ga$. The ten nodes get blown up, the ten quadric cones (in ${\Bbb P}^4$) get blown down to quadric surfaces (in the exceptional ${\Bbb P}^3$'s). Let $\~C\subset} \def\nni{\supset} \def\und{\underline X$ be the inverse image in $X$ of the union of the ten quadric cones; then on $X\backslash \~C$, $\ga$ is {\it biregular}. On the other hand, $\ga(\~C)$ is just the union of the ten quadric surfaces of the intersection $\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10}$. Consequently $$\ga_{|X\backslash \~C}:X\backslash\~C \longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal Z} \else$\cZ$\fi'\backslash(\eta')^{-1}(\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10})$$ is a reguar morphism of double covers, and letting $C\subset} \def\nni{\supset} \def\und{\underline h_p$ denote the ten quadric cones, $X\backslash \~C\longrightarrow} \def\sura{\twoheadrightarrow h_p\backslash C$ is a double cover, as is also $$\ifmmode {\cal Z} \else$\cZ$\fi'\backslash(\eta')^{-1}(\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10}) \longrightarrow} \def\sura{\twoheadrightarrow (h_p)^{\vee}\backslash \ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10}, $$ while $\ga(C)=\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10}$. Hence in the diagram $$\begin{array}{rcl} X & \stackrel{\Xi}{\longrightarrow} \def\sura{\twoheadrightarrow} & \ifmmode {\cal Z} \else$\cZ$\fi' \\ \downarrow & & \downarrow \\ h_p & \stackrel{\ga}{\longrightarrow} \def\sura{\twoheadrightarrow} & (h_p)^{\vee} \end{array}$$ $\Xi$ is regular outside of $\~C$ and maps $\~C$ to $(\eta')^{-1}(\ifmmode {\cal I} \else$\cI$\fi_4\cap \ifmmode {\cal W} \else$\cW$\fi_{10})$. Furthermore, everything is defined $\gS_6$-equivariantly. This proves the Theorem. \hfill $\Box$ \vskip0.25cm \begin{corollary}\label{c157.1} $\ifmmode {\cal I} \else$\cI$\fi_5$ sits $\gS_6$-equivariantly birationally in the center of the diagram \unitlength1cm $$\begin{picture}(2,2)\put(0,0){${\cal W}$} \put(.5,0.15){\vector(1,0){1.3}} \put(1.9,0){${\Bbb P}^4$.} \put(0,1.6){${\cal Z}$} \put(.5,1.75){\vector(1,0){1.3}} \put(1.9,1.6){${\cal Y}$} \put(.25,1.5){\vector(0,-1){1}}\put(1.95,1.5){\vector(0,-1){1}} \put(.4,1.6){\vector(1,-1){.5}} \put(.9,.9){${\cal Z}'$} \put(1.2,.8){\vector(1,-1){.5}} \end{picture}$$ This shows the relation between the quintic $\ifmmode {\cal I} \else$\cI$\fi_5$ and the Coble variety $\ifmmode {\cal Y} \else$\cY$\fi$. \end{corollary} {\bf Proof:} We have the series of modifications $$\begin{array}{ccccccc} \ifmmode {\cal I} \else$\cI$\fi_5 & \longrightarrow} \def\sura{\twoheadrightarrow & \ifmmode {\cal I} \else$\cI$\fi_{5,p} & \stackrel{\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda}{\longrightarrow} \def\sura{\twoheadrightarrow} & X & \stackrel{\Xi}{\longrightarrow} \def\sura{\twoheadrightarrow} & \ifmmode {\cal Z} \else$\cZ$\fi' \\ & & & & \downarrow & & \downarrow \\ & & & & {\Bbb P}^4 & \stackrel{\ga}{\longrightarrow} \def\sura{\twoheadrightarrow} & {\Bbb P}^4, \end{array}$$ where $\ifmmode {\cal I} \else$\cI$\fi_5\longrightarrow} \def\sura{\twoheadrightarrow \ifmmode {\cal I} \else$\cI$\fi_{5,p}$ blows up the node $p$, $\beta} \def\gk{\kappa} \def\gl{\lambda} \def\gL{\Lambda$ blows down the 15 ${\Bbb P}^3$'s through the node to the 15 ${\Bbb P}^2$'s of the intersection $\ifmmode {\cal S} \else$\cS$\fi_3\cap \ifmmode {\cal N} \else$\cN$\fi_5$, $\ga$ and $\Xi$ are as described above. Since $\ifmmode {\cal Z} \else$\cZ$\fi'$ clearly sits in the center of the diagram and all modifications are $\gS_6$-equivariant, the Corollary follows. \hfill $\Box$ \vskip0.25cm \section{${\cal I}_5$ and cubic surfaces} \subsection{The Picard group} Let $A_1(\ifmmode {\cal I} \else$\cI$\fi_5)$ be the Chow group of Weil divisors modulo algebraic equivalence. Clearly a generic hyperplane section yields an element in $A_1(\ifmmode {\cal I} \else$\cI$\fi_5)$, which we denote by $n$. Recall the reducible hyperplane sections of Proposition \ref{pQ7.1} which split each into the union of five copies of ${\Bbb P}^3$. These subvarieties are divisors on $\ifmmode {\cal I} \else$\cI$\fi_5$, hence also yield classes in the Chow group. These 45 divisors are related by 27 relations, the sum of the five classes in the Chow group being equivalent to $n$. Since $\ifmmode {\cal I} \else$\cI$\fi_5$ is normal, we have an injection $\hbox{Pic}} \def\Jac{\hbox{Jac}(\ifmmode {\cal I} \else$\cI$\fi_5)\hookrightarrow} \def\hla{\hookleftarrow A_1(\ifmmode {\cal I} \else$\cI$\fi_5)$. Let $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal I} \else$\cI$\fi_5)$ denote the orthogonal complement of the class $n$ in $A_1(\ifmmode {\cal I} \else$\cI$\fi_5)$ with respect to this injection. Then we have \begin{lemma}\label{liq6.1} We have an exact sequence of $\integer$-modules, $$0\longrightarrow} \def\sura{\twoheadrightarrow \integer^6\longrightarrow} \def\sura{\twoheadrightarrow \integer^{27}\longrightarrow} \def\sura{\twoheadrightarrow \integer^{45} \longrightarrow} \def\sura{\twoheadrightarrow {\em\hbox{Pic}} \def\Jac{\hbox{Jac}}^0(\ifmmode {\cal I} \else$\cI$\fi_5) \longrightarrow} \def\sura{\twoheadrightarrow 0.$$ \end{lemma} {\bf Proof:} The 45 ${\Bbb P}^3$'s are classes in $A_1(\ifmmode {\cal I} \else$\cI$\fi_5)$ which generate $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal I} \else$\cI$\fi_5)$ (as they contain all singularities), and the 27 relations are those just mentioned, given by the 27 hyperplane sections. So the sequence is clear as soon as we have shown that the rank of $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal I} \else$\cI$\fi_5)$ is 24 (see the sequence (\ref{e48a.1})). This now follows from the Lefschetz hyperplane theorem, as the dimension of $\ifmmode {\cal I} \else$\cI$\fi_5$ is four, so there is an isomorphism between the $H^2$'s of $\ifmmode {\cal I} \else$\cI$\fi_5$ and a hyperplane section. We may apply the Lefschetz theorem because the singularities of $\ifmmode {\cal I} \else$\cI$\fi_5$ and of a hyperplane section are local complete intersections (see the book by Goresky \& MacPherson for details). \hfill $\Box$ \vskip0.25cm Note that this sequence displays $\hbox{Pic}} \def\Jac{\hbox{Jac}^0(\ifmmode {\cal I} \else$\cI$\fi_5)$ as an {\em irreducible} $W(E_6)$-module. Furthermore, we see that just as in (\ref{e48a.1}), this sequence is dual to the corresponding sequence for cubic surfaces. \subsection{${\cal I}_5$ and cubic surfaces: combinatorics} We collect the facts relating the combinatorics of the 27 lines with those of $\ifmmode {\cal I} \else$\cI$\fi_5$ in Table \ref{table26}. \begin{table}[htb] \caption{\label{table26} Combinatorics of ${\cal I}_5$ and the 27 lines} \vspace*{.5cm} \renewcommand{\arraystretch}{1.5} \begin{tabular}{|l|l|}\hline Locus on a cubic surface (see Table \ref{table20}) & Locus on ${\cal I}_5$ \\ \hline \hline 27 lines $a_i, b_i, c_{ij}$ & 27 hyperplane sections $\{a_i=0\}\cap {\cal I}_5$, etc. \\ \hline 2 lines are skew & \parbox{7cm}{the hyperplanes intersect in one of 216 ${\Bbb P}^3$'s dual to the lines of \ref{i4}; this ${\Bbb P}^3$ intersects ${\cal I}_5$ in the union of three planes and a quadric (see Lemma \ref{l126.1})} \\ \hline \parbox{7cm}{two lines are in a tritangent} & \parbox{7cm}{the hyperplanes intersect in one of the 45 ${\Bbb P}^3$'s of \ref{i8} } \\ \hline 45 tritangents & the 45 ${\Bbb P}^3$'s of \ref{i8} \\ \hline \parbox{7cm}{Two tritangents meet in a line of the cubic surface} & \parbox{7cm}{two of the 45 ${\Bbb P}^3$'s meet in a ${\Bbb P}^2$; this is one of the planes in the ${\Bbb P}^3$ defining the arrangement ${\cal A}(W({\bf D_4}))$ as discussed in \ref{i8}} \\ \hline \parbox{7cm}{Two tritangents meet in a line outside of the cubic surface} & \parbox{7cm}{two of the 45 ${\Bbb P}^3$'s are {\em skew}, i.e., meet only in a line; this line is part of the singular locus of the arrangement ${\cal A}(W({\bf D_4}))$ just mentioned} \\ \hline 36 double sixes & \parbox{7cm}{36 triple points of ${\cal I}_5$ AND 36 copies of the Nieto quintic ${\cal N}_5$} \\ \hline \parbox{7cm}{Two double sixes are azygetic} & \parbox{7cm}{two of the triple points lie on one of the 120 lines of the singular locus of ${\cal I}_5$} \\ \hline \parbox{7cm}{Two double sixes are syzygetic} & \parbox{7cm}{two of the triple points do not lie on one of the 120 lines} \\ \hline \parbox{7cm}{A line is {\em not} contained in a double six} & \parbox{7cm}{the hyperplane dual to the line contains the triple point which corresponds to the double six} \\ \hline \end{tabular} \end{table} \section{The dual variety} Let $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ be the projective dual variety to $\ifmmode {\cal I} \else$\cI$\fi_5$; since $\ifmmode {\cal I} \else$\cI$\fi_5$ is invariant under $W(E_6)$, so is $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$. Although we do not have explicit equations for $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$, we can say quite a bit about its geometry, just from the fact that it is dual to $\ifmmode {\cal I} \else$\cI$\fi_5$. \subsection{Degree} First we show that {\em degree of\ $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$=10m+4k}. Quite generally, one can say the following. Suppose we are given a variety $X\subset} \def\nni{\supset} \def\und{\underline {\Bbb P}^n$ which has singular locus consisting of a set of {\em lines}, meeting each other in a set of {\em points}, and let us further assume that the situation is symmetric, i.e., each line contains the same number of points, each point being hit by the same number of lines; let us denote these numbers by $N=\#$ lines, $M=\#$ points, $\gn=\#$ points in each line and $\gm=\#$ lines meeting at each point. Consider the dual variety $X^{\vee}$. We claim: \begin{itemize} \item[-] There are $N$ ${\Bbb P}^{n-2}$'s $\subset} \def\nni{\supset} \def\und{\underline X^{\vee}$. \item[-] Each ${\Bbb P}^{n-2}$ is cut out by $\gn$ hyperplanes. \item[-] There are $M$ such special hyperplane sections of $X^{\vee}$. \item[-] Each of the $M$ hyperplanes meets $X^{\vee}$ in $\gm$ of the $N\ {\Bbb P}^{n-2}$'s. \item[-] Hence, deg($X^{\vee}$)=$m\gm+rest$, \end{itemize} where the $rest$ is given in terms of the local geometry around the given point. The proofs of these are immediate: each of the points corresponds to a hyperplane (=set of all hyperplanes through the point), each line defines dually a ${\Bbb P}^{n-2}$, and since $X$ is singular along the line, each hyperplane through the line is {\em tangent} to $X$ there $\Rightarrow$ the dual ${\Bbb P}^{n-2}\subset} \def\nni{\supset} \def\und{\underline X^{\vee}$. The other statements are then clear. To determine $rest$, consider the following. The set theoretic image of the given point in the dual variety is the {\em total} transform ({\em not} the proper transform) of the given point. This is set theoretically easy to compute, but there may be a multiplicity coming in. We apply these considerations to $\ifmmode {\cal I} \else$\cI$\fi_5$ and $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$: on $\ifmmode {\cal I} \else$\cI$\fi_5$ we have singular lines, $N$=120, $M$=36, $\gn$=3, $\gm$=10, and hence $deg(\ifmmode {\cal I} \else$\cI$\fi_5^{\vee})=m10+rest.$ In our case $rest$ is easy to figure out: recall that we resolved the singularities of \ifmmode {\cal I} \else$\cI$\fi$_5$ by blowing up the 36 points, then the 120 lines; the resolving divisors over the points were copies of the Segre cubic. The variety dual to the Segre cubic is the Igusa quartic, and the image of the ten nodes on the Segre cubic are ten quadric surfaces (\ref{e117a.1}) which are {\em tangent hyperplane sections}, i.e., the hyperplanes which meet the Igusa quartic in one such quadric and are tangent to it there. These ten hyperplanes are of course just the 10 ${\Bbb P}^3$'s on the dual variety being cut out by the chosen hyperplane section (see Proposition \ref{p160.1} below). This hyperplane section of $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ may be {\em tangent} to $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ along the Igusa quartic, hence $$deg(\ifmmode {\cal I} \else$\cI$\fi_5^{\vee})=10m+k\cdot4.$$ \subsection{Singular locus} Consider the 45 ${\Bbb P}^3$'s on \cIf; since there is a pencil of hyperplanes through each, the dual variety $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$\ will have 45 singular lines, which meet in 27 points (which are dual to the 27 hyperplanes cutting out the 45 ${\Bbb P}^3$'s). These 27 points are of course $A_i, B_i, C_{ij}$. Applying our reasoning from above to this we see that deg(${\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}}^{\vee}$)=5+ rest. We conclude rest=0, or in other words, {\em a resolution of singularities of ${\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}}$ is affected by blowing up the 45 lines simultaneously; there is no exceptional divisor over the 27 points.} However, since we are dealing with fourfolds, $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ could even be normal and still have a singular locus of dimension two. For example, it is reasonable to believe that the ten quadrics on each copy of the Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$ on the reducible hyperplane sections discussed below might be {\em singular} on $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$, but that is of course just a guess. Furthermore, there is no reason whatsoever why the dual variety should be normal. In fact, it is a case of great exception when the dual variety is normal, the general case being that there is a singular parabolic divisor (coming from the intersection $\hbox{Hess}} \def\rank{\hbox{rank}(X)\cap X$), as well as a double point locus, also (in general) a divisor, coming from the set of bitangents. In our case, however, since $\hbox{Hess}} \def\rank{\hbox{rank}(\ifmmode {\cal I} \else$\cI$\fi_5)\cap \ifmmode {\cal I} \else$\cI$\fi_5$ consists of the union of the 45 ${\Bbb P}^3$'s, all of which get blown down, there is no parabolic {\em divisor}. But there is no easy way to exclude a double point divisor. \subsection{Reducible hyperplane sections} As already mentioned, $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$\ contains 120 ${\Bbb P}^3$'s, each being cut out by three of the $h$'s, (in fact by a triple of azygetic double sixes), and each such intersection $h\cap \ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ consists of ten such ${\Bbb P}^3$'s, plus a copy of the Igusa quartic. There are 36 such hyperplane sections which decompose into ten ${\Bbb P}^3$'s and a copy of the Igusa quartic: \begin{proposition}\label{p160.1} The 36 hyperplane sections ${\bf h}\cap \ifmmode {\cal I} \else$\cI$\fi_5$, ${\bf h}=h,\ h_{ij},\ h_{ijk}$, are reducible, consisting of ten ${\Bbb P}^3$'s and a copy of the Igusa quartic $\ifmmode {\cal I} \else$\cI$\fi_4$. The ten ${\Bbb P}^3$'s are just the $K_{ijk}$ of (\ref{e117a.1a}), each a bitangent plane to $\ifmmode {\cal I} \else$\cI$\fi_4$. \end{proposition} {\bf Proof:} These are the 36 hyperplanes dual to the 36 triple points of $\ifmmode {\cal I} \else$\cI$\fi_5$; at each such $p$ ten of the 120 lines meet, and the triple point itself yields the copy of $\ifmmode {\cal I} \else$\cI$\fi_4$ (it is blown up with exceptional divisor $(\ifmmode {\cal S} \else$\cS$\fi_3)_p$, which is dual to $(\ifmmode {\cal I} \else$\cI$\fi_4)_p$, a copy of $\ifmmode {\cal I} \else$\cI$\fi_4$). \hfill $\Box$ \vskip0.25cm So restricted to the triple point, the duality $\ifmmode {\cal I} \else$\cI$\fi_5- - \ra \ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$ yields precisely the dual map $\ga$ of (\ref{e126.1})! The 120 ${\Bbb P}^3$'s meet two at a time in 270 ${\Bbb P}^2$'s, each of which is cut out by six of the $h$'s (2 triples of azygetic double sixes, two rows in a triple). Note that these 270 ${\Bbb P}^2$'s are the $t_6(2)$ of Table \ref{table23}. Each ${\Bbb P}^2$ contains two nodes and five of the 27 points, as well as two of the 45 lines. Through each such line two of these ${\Bbb P}^2$'s pass (as each line is cut out by 12 of the $h$'s). Therefore in each $h$ we have ten ${\Bbb P}^3$'s meeting in ${10 \choose 2}=45\ {\Bbb P}^2$'s which meet in 15 of the 45 ${\Bbb P}^1$'s, and contain 15 of the 27 points. The 15 lines and 15 points are just the singular locus of the Igusa quartic, and the ten ${\Bbb P}^3$'s are tangent to the Igusa quartic along quadrics, as mentioned earlier. \subsection{Special hyperplane sections} Inspection of the 27 forms and 27 points in ${\Bbb P}^5$ shows that each of the 27 hyperplanes contains {\em none} of the 27 points and {\em none} of the 45 lines; it follows that hyperplane sections such as $\ifmmode {\cal K} \else$\cK$\fi:=\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}\cap \{a_1=0\}$ are irreducible hypersurfaces in ${\Bbb P}^4$ with 45 isolated singularities, coming from the intersections with the singular lines of $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$. As mentioned above, there may also be a singular locus coming from other singularities on $\ifmmode {\cal I} \else$\cI$\fi_5^{\vee}$. Furthermore, there are 40 ${\Bbb P}^2$'s lying on this threefold, and 16 hyperplanes in $a_i$ which cut out ten of these on $\ifmmode {\cal K} \else$\cK$\fi$. The 16 hyperplanes are those 16 of the 216 ${\Bbb P}^3$'s which lie in $a_i$, corresponding to the 16 lines which $a_i$ is skew to. The symmetry group of this threefold is $W(D_5)$. This is a {\em degeneration} of a {\em generic} hyperplane section, which will contain 120 ${\Bbb P}^2$'s. \bigskip
1995-03-30T07:20:48
9503
alg-geom/9503024
en
https://arxiv.org/abs/alg-geom/9503024
[ "alg-geom", "math.AG" ]
alg-geom/9503024
Heath Martin
Heath M. Martin and Juan C. Migliore
Degrees of generators of ideals defining subschemes of projective space
27 pages, LaTeX v 2.09
null
null
null
null
For an arithmetically Cohen--Macaulay subscheme of projective space, there is a well-known bound for the highest degree of a minimal generator for the defining ideal of the subscheme, in terms of the Hilbert function. We prove a natural extension of this bound for arbitrary locally Cohen--Macaulay subschemes. We then specialize to curves in $\pthree$, and show that the curves whose defining ideals have generators of maximal degree satisfy an interesting cohomological property. The even liaison classes possessing such curves are characterized, and we show that within an even liaison class, all curves with the property satisfy a strong Lazarsfeld--Rao structure theorem. This allows us to give relatively complete conditions for when a liaison class contains curves whose ideals have maximal degree generators, and where within the liaison class they occur.
[ { "version": "v1", "created": "Wed, 29 Mar 1995 18:57:19 GMT" } ]
2008-02-03T00:00:00
[ [ "Martin", "Heath M.", "" ], [ "Migliore", "Juan C.", "" ] ]
alg-geom
\section{Degrees of Generators} Throughout this paper, we will work over an algebraically closed field $k$, of arbitrary characteristic. Let $S = k[x_0, \dots, x_n]$ be the polynomial ring over $k$, and ${\Bbb P}^n$ the $n$-dimensional projective space over $k$. We will furthermore consider only subschemes of ${\Bbb P}^n$ which are locally Cohen--Macaulay and equidimensional. It is well-known that this is equivalent to requiring that all the intermediate cohomology modules $H^i_*({\Bbb P}^n, V)$, $1 \leq i \leq \dim V$, have finite length. Given a subscheme $V$ of ${\Bbb P}^n$, with $\dim V = d$, let $I_V$ denote its homogeneous, saturated defining ideal in $S$. Thus $S/I_V$ is a standard graded $k$-algebra, and so we can define the Hilbert function of $S/I_V$ by $$ H(S/I_V, t) = \dim_k [S/I_V]_t. $$ Alternatively, we sometimes write $H(V, t)$ for $H(S/I_V, t)$. It is a standard fact that there is a polynomial $P(S/I_V, t)$, having degree $d$, such that $H(S/I_V, t) = P(S/I_V, t)$ for all $t \gg 0$. We furthermore define the $n$th difference of $H(S/I_V, t)$ inductively as follows: \begin{eqnarray*} \Delta^1 H(S/I_V,t) &=& H(S/I_V, t) - H(S/I_V, t-1) \\ \Delta^n H(S/I_V,t) &=& \Delta^1 (\Delta^{n-1} H(S/I_V, t)). \end{eqnarray*} Now, since $H(S/I_V, t)$ is eventually a polynomial of degree $d$, the function $\Delta^{d+1} H(S/I_V, t)$ is eventually zero, and we define $$ \sigma(S/I_V) = \min \{\, k : \Delta^{d+1} H(S/I_V, t) = 0 \mbox{ for all $t \geq k$} \, \}. $$ Again, we will sometimes write $\sigma(V)$ for $\sigma(S/I_V)$. It is worth noting that if $\dim V = d$, then the Hilbert function of $V$ and the Hilbert polynomial of $V$ are equal in all degrees $\geq t + d + 1$ if and only if $\sigma(V) = t$. The degree at which the Hilbert function and the Hilbert polynomial agree from then on is sometimes called, at least in the context of local algebra, the postulation number, and has played an important role in questions about Cohen--Macaulayness and related invariants in local rings. Given an ideal $I$, we will write $\alpha(I)$ for the minimal degree of a minimal generator, and $\omega(I)$ for the maximal degree of a minimal generator of $I$. Next, let $H^i_*({\Bbb P}^n, {\cal I}_V) = \oplus_{s\in {\Bbb Z}} H^i({\Bbb P}^n, {\cal I}_V(s))$ be the cohomology modules of $V$. We will put $h^i({\Bbb P}^n, {\cal I}_V(t)) = \dim_k H^i({\Bbb P}^n, {\cal I}_V(t))$. We let $e(V) = \max\{\, t : h^{d+1}({\Bbb P}^n, {\cal I}_V(t)) \not=0\,\}$ denote the index of speciality of $V$. Also, note that by our assumption that subschemes be locally Cohen--Macaulay and equidimensional, $h^i({\Bbb P}^n, {\cal I}_V(t))$ is non-zero for only finitely many $t$, when $1 \leq i \leq d$. Hence for a non-arithmetically Cohen--Macaulay curve $C$ in ${\Bbb P}^n$, with notation following Martin-Deschamps and Perrin, we can define $$ r_a(C) = \min \{\, n \in {\Bbb Z} : h^1({\Bbb P}^n, {\cal I}_C(n)) \not = 0 \,\} \quad\quad r_o(C) = \max \{\, n \in {\Bbb Z} : h^1({\Bbb P}^n, {\cal I}_C(n)) \not = 0 \,\}, $$ and $\mathop{\rm diam\,} H^1_*({\Bbb P}^n, C) = r_o(C) - r_a(C) + 1$, the number of components between the first and the last non-zero components, inclusive. Note that some of the intermediate components may have dimension zero, but we also count them. If $C$ is arithmetically Cohen--Macaulay, then $H^1_*({\Bbb P}^3, C) = 0$, and we will put $\mathop{\rm diam\,} H^1_*({\Bbb P}^3, C) = 0$. In this section, we prove a statement about the maximal degree of a generator for the defining ideal of a curve, in terms of the Hilbert function and the cohomology of the curve. The relationship between these two objects is well-known, and we spell it out explicitly in the first lemma. \begin{lemma} \label{hilbfcn_eqn} Let $C$ be a curve in ${\Bbb P}^n$. Then \begin{eqnarray*} \Delta^2 H(C,t) &=& h^2({\Bbb P}^n, {\cal I}_C(t)) - 2h^2({\Bbb P}^n, {\cal I}_C(t-1)) + h^2({\Bbb P}^n, {\cal I}_C(t-2)) \\ && \quad\quad \mbox{} - h^1({\Bbb P}^n, {\cal I}_C(t)) + 2h^1({\Bbb P}^n, {\cal I}_C(t-1)) - h^1({\Bbb P}^n, {\cal I}_C(t-2)) \end{eqnarray*} \end{lemma} \begin{proof} First, recall that the Hilbert polynomial is given by $P(C,t) = h^0({\Bbb P}^n, {\cal O}_C(t)) - h^1({\Bbb P}^n, {\cal O}_C(t))$ (see \cite[Exercise III.5.2]{hart}). Thus, from the short exact sequence $$ 0 \rightarrow I_t \rightarrow S_t \rightarrow H^0({\Bbb P}^n, {\cal O}_C(t)) \rightarrow H^1({\Bbb P}^n, {\cal I}_C(t)) \rightarrow 0, $$ we obtain \begin{eqnarray*} H(C,t) &=& h^0({\Bbb P}^n, {\cal O}_C(t)) - h^1({\Bbb P}^n, {\cal I}_C(t)) \\ &=& P(C,t) + h^2({\Bbb P}^n, {\cal I}_C(t)) - h^1({\Bbb P}^n, {\cal I}_C(t)). \end{eqnarray*} Now, since $P(C, t)$ is a polynomial of degree $1$, on taking second differences we obtain \begin{eqnarray*} \Delta^2 H(C,t) &=& h^2({\Bbb P}^n, {\cal I}_C(t)) - 2h^2({\Bbb P}^n, {\cal I}_C(t-1)) + h^2({\Bbb P}^n, {\cal I}_C(t-2)) \\ && \quad\quad \mbox{} - h^1({\Bbb P}^n, {\cal I}_C(t)) + 2h^1({\Bbb P}^n, {\cal I}_C(t-1)) - h^1({\Bbb P}^n, {\cal I}_C(t-2)) \end{eqnarray*} which is what we wanted to show. \end{proof} The following corollary is an immediate consequence. \begin{cor} \label{Delta-coh} $\Delta^2 H(C, t) = 0$ for all $t \geq k+2$ if and only if $h^2({\Bbb P}^n, {\cal I}_C(t)) = h^1({\Bbb P}^n, {\cal I}_C(t))$ for all $t \geq k$. \mbox{\hskip 1cm $\rlap{$\sqcap$}\sqcup$} \end{cor} \begin{prop} \label{equal-coh} Suppose $C$ is a curve in ${\Bbb P}^n$ defined by an ideal $I = I_C$. If $\omega(I) = \sigma(C) + k$ for some $k \geq 1$, then $e(C) = r_o(C)$ and $h^2({\Bbb P}^n, {\cal I}_C(t)) = h^1({\Bbb P}^n, {\cal I}_C(t))$ for $t \geq e(C) -k+1$. \end{prop} \begin{proof} We first show that $e(C) = r_o(C)$. If not, let $m = \max \{\, e(C), r_o(C) \,\}$. Then by Castelnuovo--Mumford regularity \cite{mumford}, we have $$ \sigma(C) < \sigma(C) + k = \omega(I) \leq \mathop{\rm reg\,}(I) \leq m+3. $$ But $\Delta^2 H(C, m+2) \not=0$, because of Lemma~\ref{hilbfcn_eqn} and so $\sigma(C) = m+3$, which is a contradiction. Thus, we must have $e(C) = r_o(C) = m$. In particular, $\mathop{\rm reg\,}(I) = m+3$, and so again by Castelnuovo--Mumford regularity, we have $\sigma(C) = \omega(I) - k \leq \mathop{\rm reg\,}(I) - k = m + 3 - k$. Thus, we have $\Delta^2 H(C, t) = 0$ for $t \geq m+3-k$, and so by Lemma~\ref{hilbfcn_eqn}, $h^1({\Bbb P}^n, {\cal I}_C(t)) = h^2({\Bbb P}^n, {\cal I}_C(t))$ for $t \geq m-k+1$. \end{proof} \begin{prop} \label{maxdegree} Suppose $I = I_C$ defines a curve $C$ in ${\Bbb P}^n$. Then $\omega(I) \leq \sigma(S/I) + \mathop{\rm diam\,} H^1_*({\Bbb P}^n, {\cal I}_C)$. \end{prop} \begin{proof} This follows immediately from the previous proposition, since $h^1({\Bbb P}^n, {\cal I}_C(t))$ and $h^2({\Bbb P}^n, {\cal I}_C(t))$ can only possibly be non-zero and equal for $k=\mathop{\rm diam\,} H^1_*({\Bbb P}^n, {\cal I}_C)$ degrees. \end{proof} \begin{remark} Also note that Proposition~\ref{maxdegree} includes the case that $C$ is arithmetically Cohen--Macaulay, and says that $\omega(I_C) \leq \sigma(S/I_C)$. This is a result in \cite{DGM}; see also \cite[Proposition 1.2]{CGO}. \end{remark} \begin{remark} Some comments about the proof of this result are in order. First of all, if either $e(C) \not= r_o(C)$ or if $e(C) = r_o(C)$ and $h^1({\Bbb P}^n, {\cal I}_C(r_o)) \not= h^2({\Bbb P}^n, {\cal I}_C(r_o))$, we get $\omega(I) \leq \sigma(S/I)$. This is the same bound as when $C$ is assumed to be arithmetically Cohen--Macaulay. Thus, the cases of most interest occur when $H^1_*({\Bbb P}^n, {\cal I}_C)$ and $H^2_*({\Bbb P}^n, {\cal I}_C)$ both become zero at the same degree, and moreover have equal dimensions for some number of preceding degrees. Essentially, Proposition~\ref{equal-coh} says that having a generator of high degree forces $h^1({\Bbb P}^n, {\cal I}_C(t))$ and $h^2({\Bbb P}^n, {\cal I}_C(t))$ to be equal in a large number of degrees. \end{remark} Curves which are not arithmetically Cohen--Macaulay and have a generator of maximum degree in the sense of the above proposition, must be ``almost Buchsbaum.'' This means that a general linear form $L$ induces a multiplication on $H^1_*({\Bbb P}^n, {\cal I}_C)$ which has non-trivial kernel in each degree. As notation, if $L$ is a linear form, let $K_L$ be the kernel of the multiplication on $H^1_*({\Bbb P}^n, {\cal I}_C)$ induced by $L$. \begin{prop} \label{almostBuchsbaum} Let $C$ be a non-arithmetically Cohen--Macaulay curve in ${\Bbb P}^3$, and suppose that $h^1({\Bbb P}^n, {\cal I}_C(t)) = h^2({\Bbb P}^n, {\cal I}_C(t))$ in the last $r$ degrees. Let $L$ be a general linear form and let $K = K_L$. Then $\dim_k K_{t} > \dim_k K_{t+1} > 0$ for all $t = r_o(C) - r + 1, \dots, r_o(C) - 1$. \end{prop} \begin{proof} Let $L$ be a general general linear form defining a hyperplane $H$, and let $Z = C \cap H$ be the hyperplane section of $C$ considered as a subscheme of ${\Bbb P}^{n-1}$. Then it is easy to see that $$ \Delta^2 H(C, t) = \Delta^1 H(Z, t) + \Delta^1 \dim_k K_{t-1}. $$ Let $p = r_o(C)$, and note that by the condition on cohomology and by Lemma~\ref{hilbfcn_eqn}, we have $$ 0 = \Delta^2 H(C, p+2) = \Delta^1 H(Z, p+2) + \Delta^1 K_{p+1}. $$ But $\dim_k K_{p+1} = 0$ and $\dim_k K_p > 0$, so $\Delta^1 K_{p+1} < 0$. This implies that $\Delta^1 H(Z, p+2) > 0$, and since $\Delta^1 H(Z, t)$ is non-increasing in the range in which we are interested (see \cite{DGM}), then $\Delta^1 H(Z,t) > 0$ for all $t = r_o(C) - r + 3, \dots, r_o(C) + 2$. But the assumptions on the cohomology of $C$ then imply $$ 0 = \Delta^2 H(C, t) = \Delta^1 H(Z,t) + \Delta^1 \dim _k K_{t-1} $$ for $t = r_o(C) - r + 3, \dots, r_o(C) + 2$, and so $\Delta^1 \dim_k K_{t-1} < 0$. That is, $\dim_k K_{t} > \dim_k K_{t+1} > 0$ for $t = r_o(C) - r+1, \dots, r_o(C) - 1$. \end{proof} As an immediate corollary, we obtain the following statement: \begin{cor} \label{almostBuchs-deg} Suppose $C$ is a curve in ${\Bbb P}^n$ having a generator of degree $\sigma(C) + \mathop{\rm diam\,} H^1_*({\Bbb P}^n, {\cal I}_C)$. Then for each general linear form $L$, $\dim K_t > \dim K_{t+1} > 0$, for $t = r_a(C), \dots, r_o(C) - 1$. \mbox{\hskip 1cm $\rlap{$\sqcap$}\sqcup$} \end{cor} We can also use Proposition~\ref{almostBuchsbaum} to refine the bound on the maximum degree of a generator. \begin{cor} If $C$ is a curve in ${\Bbb P}^n$ defined by an ideal $I = I_C$, then $\omega(I) \leq \sigma(C) + \mathop{\rm diam\,} K$. \end{cor} \begin{proof} Again, this follows from Proposition~\ref{equal-coh} and Proposition~\ref{almostBuchsbaum}. \end{proof} More generally, we have the following result for subschemes of dimension $d$ in ${\Bbb P}^n$: \begin{prop}\label{general_maxdegree} Let $V$ be a subscheme of ${\Bbb P}^n$ having dimension $d$, defined by an ideal $I = I_V$. Then $\omega(I) \leq \sigma(V) + \max \{\, \mathop{\rm diam\,} H^i_*({\Bbb P}^n, {\cal I}_V) : i = 1, \dots, d\,\}$. \end{prop} \begin{proof} Since the proof of this result is quite similar to the case of curves, we will only give an outline. The Hilbert polynomial of $V$ is given by $$ P(V, t) = \sum_{i=0}^d (-1)^i h^i({\Bbb P}^n, {\cal O}_V(t)), $$ and so from the exact sequence $$ 0 \rightarrow I_t \rightarrow S_t \rightarrow H^0({\Bbb P}^n, {\cal O}_V(t)) \rightarrow H^1({\Bbb P}^n, {\cal I}_V(t)) \rightarrow 0, $$ we see that the Hilbert function of $V$ is $$ H(V, t) = P(V, t) + \sum_{i=1}^d (-1)^i h^i({\Bbb P}^n, {\cal I}_V(t)). $$ Since $P(V,t)$ is a polynomial of degree $d$, when we take $(d+1)$th differences, we get $$ \Delta^{d+1} H(V, t) = \sum_{i=1}^d (-1)^i \sum_{j=0}^{d+1} (-1)^j{{d+1} \choose j} h^i({\Bbb P}^n, {\cal I}_V(t-j)). $$ Now we argue by cases. If none of the cohomology modules end in the same place, it is easy to see by Castelnuovo--Mumford regularity that $\omega(I) \leq \sigma(V)$. Suppose, on the other hand, that some of the cohomology modules end in the same degree $t$, say, and the others end in degrees $< t$, and let $m$ be the maximum of the diameters of the intermediate cohomologies. Then $\mathop{\rm reg\,}(V) = t + d + 1$, and by the formula above, $\sigma(V) \geq t - m +d + 1$, and since $\omega(I) \leq \mathop{\rm reg\,}(I)$, the required inequality follows. \end{proof} We can be a bit more precise in a few cases. For instance, suppose $V$ is a surface in ${\Bbb P}^n$. Then there are only three non-zero cohomology modules, and as in the proof above the Hilbert function of $V$ is given by $$ H(V, t) = P(V, t) - h^3({\Bbb P}^n, {\cal I}_V(t)) + h^2({\Bbb P}^n, {\cal I}_V(t)) - h^1({\Bbb P}^n, {\cal I}_V(t)). $$ Because the $h^3$ term and the $h^1$ term have the same sign, the only cancellation that can occur comes from the $h^2$ term, and using the same argument as above, we get the inequality $\omega(I) \leq \sigma(I) + \mathop{\rm diam\,} H^2_*({\Bbb P}^n, {\cal I}_V)$. In particular, if $V$ were a non-arithmetically Cohen--Macaulay surface, with $H^2_*({\Bbb P}^n, {\cal I}_V) = 0$, then $\omega(I) \leq \sigma(V)$, which is the same bound as in the arithmetically Cohen--Macaulay case. \section{Curves with equal cohomology} The previous section showed that the property of having a generator of high degree is very closely related to having equal (non-zero) cohomology dimensionally in a large number of degrees. This section is devoted to characterizing when an even liaison class of curves in ${\Bbb P}^3$ has this property. By a slight abuse of terminology, we make the following definition: \begin{defn} A curve $C$ in ${\Bbb P}^3$ is said to have {\em equal cohomology} if $e(C) = r_o(C)$ and $h^1({\Bbb P}^3, {\cal I}_C(t)) = h^2({\Bbb P}^3, {\cal I}_C(t))$ for $t = r_a(C), \dots, r_o(C)$. \end{defn} We first recall the structure theory for curves in ${\Bbb P}^3$ (which holds more generally for codimension $2$ subschemes of ${\Bbb P}^n$) initiated by Lazarsfeld and Rao, and developed in a series of papers, of which \cite{BBM} contains the most general statement and proof. See also the book \cite{juan-book} for a comprehensive overview of liaison theory and the Lazarsfeld--Rao structure theory. First, let $C$ be a curve, and choose a form $F \in I_C$ of degree $f$ and a form $G \in S$ of degree $g$ so that $F$ and $G$ have no common components. Then $I_Z = G \cdot I_C + (F)$ defines a curve $Z$ in ${\Bbb P}^3$, called a basic double link of $C$, and denoted $$ C:(g,f) \rightarrow Z. $$ It is easy to see that $Z$ is evenly linked to $C$, and that there is a short exact sequence $$ 0 \rightarrow S(-g-f) \stackrel{[F\;G]}{\longrightarrow} I_C(-g) \oplus S(-f) \stackrel{\phi}{\longrightarrow} I_Z \rightarrow 0, $$ where $\phi(r,s) = rG+sF$. The Lazarsfeld--Rao property says essentially that even liaison classes of curves are built up by this process of basic double linkage. More precisely, let ${\cal L}$ be an even liaison class of curves in ${\Bbb P}^3$. Then the cohomology module $M = H^1_*(C)$ is invariant up to shifts in grading as $C$ varies in ${\cal L}$, and there is a leftmost shift of $M$ which actually occurs as the deficiency module of a curve in ${\cal L}$, and every rightward shift is realized. Thus ${\cal L}$ is parameterized by shifts of $M$, and a curve $C_0$ which has $H^1_*(C_0)$ in the leftmost shift is called a {\em minimal curve}. Every other curve $C$ in ${\cal L}$ is obtained from $C_0$ as follows: there is a curve $C_m$ which is a deformation of $C$ through curves having constant cohomology, and a series of basic double links \begin{equation}\label{bdl-chain} C_0 : (1, d_0) \rightarrow C_1 : (1, d_1) \rightarrow \cdots \rightarrow C_{m-1} : (1, d_{m-1}) \rightarrow C_m. \end{equation} We can moreover choose the degrees to satisfy $d_0 = \cdots = d_s < d_{s+1} < \cdots < d_{m-1}$. Note furthermore that $C_i$ is in the $i$th shift of ${\cal L}$; that is, $H^1_*(C_i)$ is a rightward shift by $i$ degrees of $H^1_*(C_0)$. Also, for curves in ${\Bbb P}^3$, the book \cite{mdp1} gives much more information about the behavior of invariants along liaison classes, and moreover gives an algorithm for computing the minimal curve in the liaison class from the deficiency module associated to the class. We will need to have some information about how Hilbert functions change as we move along an even liaison class by basic double linkage. The following result is quite elementary. \begin{lemma} \label{hilbfcn-bdl} Suppose $I = I_C$ defines a curve $C$ in ${\Bbb P}^3$. Let $L$ be a general hyperplane and let $F$ define a surface of degree $d$ containing $C$. Form the basic double link $Z$ of $C$ by $L$ and $F$. Then $$ \Delta^2 H(Z, t) = \left\{ \begin{array}{ll} \Delta^2 H(C, t-1) + 1 & \mbox{\rm if $1 \leq t \leq d-1$} \\ \Delta^2 H(C, t-1) & \mbox{\rm if $t \geq d$.} \end{array}\right. $$ \end{lemma} \begin{proof} Since $Z$ is a basic double link of $C$, we have a short exact sequence $$ 0 \rightarrow S(-d-1) \rightarrow I_C(-1) \oplus S(-d) \rightarrow I_Z \rightarrow 0. $$ Using the additive properties of Hilbert functions, it is easy to see that $\Delta^2 H(Z, t) = \Delta^2 H(C, t-1) + \Delta^3 H(F, t)$. Then the statement follows, since $\Delta^3 H(F, t) = 1$ for $0 \leq t \leq d-1$, and is zero otherwise. Note that this is also in \cite[Corollary 2.3.5]{nollet}, in terms of postulation characters. \end{proof} \begin{definition} \label{deltadef} Suppose there is a chain $$ {\cal C} : C_0 : (1, d_0) \rightarrow C_1 : (1, d_1) \rightarrow \cdots \rightarrow C_{m-1} : (1, d_{m-1}) \rightarrow C_m $$ of basic double links by surfaces $F_i$ having degrees $d_i$. Then define $\delta({\cal C}, t, s)$ to be the number of $F_i$ such that $d_i \geq t-s+i+1$, for $i = 1, \dots, s-1$. \end{definition} It is easy to see that $\delta({\cal C}, t, s) \leq \delta({\cal C}, t-1, s)$. The following result is a straightforward calculation using Lemma~\ref{hilbfcn-bdl}. \begin{cor} \label{hilbfcn-chain} Let ${\cal C}$ as above be a chain of basic double links. Then for each $s = 0, \dots, m$, $$ \Delta^2 H(C_s, t) = \Delta^2 H(C_0, t-s) + \delta({\cal C},t,s).\mbox{\hskip 1cm $\rlap{$\sqcap$}\sqcup$} $$ \end{cor} The main result of this section concerns when an even liaison class has a curve with $h^1$ and $h^2$ equal for some number of places, and gives a characterization in terms of the Hilbert function of the minimal curve in the liaison class. We first state a more general version, from which we can trivially make a statement about minimal curves. \begin{thm} \label{coh-prop} Let ${\cal L}$ be an even liaison and let $C_0$ be a curve in ${\cal L}$. Then a sequence $$ C_0 \rightarrow \dots \rightarrow C_m = C $$ of basic double links can be constructed with $C$ having $h^1({\Bbb P}^3, {\cal I}_C(t)) = h^2({\Bbb P}^3, {\cal I}_C(t))$ in the last $r$ places if and only if $C_0$ has $e(C_0) \leq r_o(C_0)$ and has Hilbert function satisfying $$ \Delta^2 H(C_0, r_o(C_0) - r + 3) \leq \Delta^2 H(C_0, r_o(C_0) - r + 4) \leq \cdots \leq \Delta^2 H(C_0, r_o(C_0) + 2) \leq 0. $$ \end{thm} \begin{proof} Suppose that $C_0$ has $e(C_0) \leq r_o(C_0)$ and Hilbert function $$ \Delta^2 H(C_0, t) = \quad \cdots \quad t_1 \quad t_2 \quad \cdots \quad t_r \quad 0 \quad \cdots $$ where $t_1 \leq t_2 \leq \cdots \leq t_r \leq 0$, and the term $t_r$ occurs in degree $r_o(C_0) + 2$. If not all the $t_i$ are equal, let $s$, $1 \leq s \leq r$, be the first integer for which $t_{s-1} < t_s$. Otherwise, let $s=r+1$. Note that $I_{C_0}$ must have elements in degree $\leq r_o(C_0)-r+3$, so we can form the basic double link $C_0:(1, r_o(C_0)-r+s+1) \rightarrow C_1$. Then by Lemma~\ref{hilbfcn-bdl}, $C_1$ has Hilbert function $$ \Delta^2 H(C_1, t) = \quad \cdots \quad t_1 + 1 \quad \cdots \quad t_{s-1}+1 \quad t_s \quad \cdots \quad t_r \quad 0 \quad \cdots, $$ where now the term $t_r$ occurs in degree $r_o(C_0)+3$. Continuing by induction, we construct a sequence of basic double links to a curve $C_m$ having Hilbert function $$ \Delta^2 H(C_m, t) = \quad \cdots \quad u_1 \quad \cdots \quad u_r \quad 0 \quad \cdots, $$ where $u_1 = \cdots = u_r = 0$, and where the term $u_r$ occurs in degree $r_o(C_0) + 2 +m$. We claim that $C_m$ has the cohomology property. To see this, note that $r_o(C_m) = r_o(C_0) + m$, and that $\sigma(C_m) \leq r_o(C_0)+m-r+3$, so that $\Delta^2 H(C_m, t) = 0$ for all $t \geq r_o(C_0)+m-r+3$. But by Corollary~\ref{Delta-coh}, we see that $h^2({\Bbb P}^3, {\cal I}_{C_m}(t)) = h^1({\Bbb P}^3, {\cal I}_{C_m}(t))$ for all $t \geq r_0(C_0)+m-r-1 = r_o(C_m) - r + 1$. That is, $h^2({\Bbb P}^3, {\cal I}_{C_m}(t)) = h^1({\Bbb P}^3, {\cal I}_{C_m}(t))$ in the last $r$ places. For the converse, suppose $C \in {\cal L}$ has the cohomology property and that there is a sequence of basic double links $$ {\cal C} : C_0:(1, d_0) \rightarrow C_1:(1, d_1) \rightarrow \dots \rightarrow C_m = C. $$ First, note that by \cite[Lemma 1.14]{BM1}, $e(Z)$ increases by at least one each time we move up in the liaison class. But $r_o(Z)$ increases by exactly one each time. Since the cohomology property for $C_m$ in particular means that $e(C_m) = r_o(C_m)$, then we must have $e(C_0) \leq r_o(C_0)$. Next we show that the Hilbert function of $C_0$ has the given form. First, $$ \Delta^2 H(C_0, r_o(C_0)+2)= \Delta^2 H(C_m, r_o(C_0) + 2 +m) - \delta({\cal C}, r_o(C_0)+m, m), $$ and since $r_o(C_0) + 2 + m = r_o(C_m)+2$, we have $\Delta^2 H(C_m, r_o(C_0)+2+m) = 0$, because of Corollary~\ref{Delta-coh} and the fact that $h^1({\Bbb P}^3, {\cal I}_C(t)) = h^2({\Bbb P}^3, {\cal I}_C(t))$ for $t\geq r_0(C_m)$. Thus, $\Delta^2 H(C_0, r_o(C_0)+2) \leq 0$. Moreover, for each $i = 1, \dots, r-1$, we have \begin{eqnarray*} \Delta^2 H(C_0, r_o(C_0)+2-i) &=& \Delta^2 H(C_m, r_o(C_0)+2-i+m) - \delta({\cal C}, r_o(C_0)+2-i+m, m) \\ &=& -\delta({\cal C}, r_o(C_0)+2-i+m, m). \end{eqnarray*} As above, the second equality follows from the fact that $r_o(C_0) + 2 -i + m = r_o(C_m) + 2 - i$ and because the cohomology property for $C_m$ implies $\Delta^2 H(C_m, r_o(C_m) + 2 - i) = 0$ via Corollary~\ref{Delta-coh}. But by our observation following Definition~\ref{deltadef}, we have \begin{eqnarray*} \Delta^2 H(C_0, r_o(C_0)+2-i) &=& - \delta({\cal C}, r_o(C_0)+2-i+m, m) \\ &\leq& -\delta({\cal C}, r_o(C_0)+2-(i-1)+m, m) \\ &=& \Delta^2 H(C_0, r_o(C_0)+2-(i-1)), \end{eqnarray*} and so the proof is finished. \end{proof} Since for any curve $C$ in an even liaison class ${\cal L}$ there is a sequence of basic double links from a minimal curve in ${\cal L}$ to a curve $C_m$, followed by a deformation with constant cohomology to $C$, the Theorem has the immediate consequence: \begin{cor} \label{mincoh-prop} An even liaison class of curves in ${\Bbb P}^3$ contains a curve $C$ having $h^1({\Bbb P}^3, {\cal I}_C(t)) = h^2({\Bbb P}^3, {\cal I}_C(t))$ in the last $r$ places if and only if the minimal curve $C_0$ in ${\cal L}$ has $e(C_0) \leq r_o(C_0)$ and has Hilbert function satisfying $$ \Delta^2 H(C_0, r_o(C_0) - r + 3) \leq \Delta^2 H(C_0, r_o(C_0) - r + 4) \leq \dots \leq \Delta^2 H(C_0, r_o(C_0) + 2) \leq 0. $$ \end{cor} \section{Liaison Classes of Curves with Equal Cohomology} As we saw in the previous section, the presence or non-presence of a curve in an even liaison class ${\cal L}$ having equal cohomology can be detected by looking at the Hilbert function of the minimal curve in ${\cal L}$. This raises some immediate questions: if an even liaison class ${\cal L}$ contains curves with equal cohomology, in what shifts of ${\cal L}$ do they occur, and ``how many'' such curves are there? To answer these questions, we first make some remarks concerning how the equal cohomology property behaves with respect to basic double linkage. \begin{remark}\label{eqcoh&bdls} \begin{enumerate} \item[(a.)] \label{bdl-rem1} Let ${\cal L}$ be an even liaison class with associated deficiency module of diameter $r$. Suppose ${\cal L}$ contains curves with equal cohomology. Then by the previous corollary, the minimal curve in the liaison class has Hilbert function ending in a non-decreasing sequence of $r$ non-positive terms, beginning in degree $r_a(C_0) + 2$. Say $\Delta^2H(C_0, t) = \cdots\quad t_1 \quad \cdots\quad t_r$ is this sequence. Let $$ C_0:(1,b_0) \rightarrow C_1:(1, b_1) \rightarrow C_m:(1,b_{m-1}) \rightarrow C_m $$ be a sequence of basic double links. By Lemma~\ref{hilbfcn-bdl}, if $\Delta^2 H(C_i, t)$ ends in negative terms, and if the basic double linkage $C_i:(1,b_i) \rightarrow C_{i+1}$ changes one of these negatives, then it also changes every term preceding. In particular, it must change the left-most negative term. Moreover, each basic double link which changes negatives increases the left-most negative term by exactly one. Note that a link $C_i:(1, b_i) \rightarrow C_{i+1}$ changes negative terms if and only if $b_i \geq r_a(C_i) + 2 = r_a(C_0) + i + 2$. \item[(b.)] \label{bdl-rem2} Related to this is the observation that if there are more than $-t_1$ basic double links which change negative terms, then $C_m$ cannot have equal cohomology. More precisely, if we have $b_i \geq r_a(C_0) + i + 2$ for more than $-t_1$ indices $i$, then the $t_1$ term eventually becomes positive, and this forces $h^1({\Bbb P}^3, {\cal I}_C(t))$ and $h^2({\Bbb P}^3, {\cal I}_C(t))$ to be non-equal in at least the leftmost degree. \item[(c.)] \label{bdl-rem3} Continuing with that theme, it follows rather trivially that if we have more than $-t_1$ links which change negative terms, then no further basic double link can possibly produce a curve with equal cohomology. \item[(d.)] \label{bdl-rem4} As a final remark, note that if $C :(1, d) \rightarrow D$ is a basic double link, and if $C$ has equal cohomology, then $D$ has equal cohomology if and only if $d \leq r_a(C) + 3$. This follows directly from Corollary~\ref{Delta-coh} and Lemma~\ref{hilbfcn-bdl}. \end{enumerate} \end{remark} Now, we begin our description of which curves in the liaison class have equal cohomology by showing that there is a unique minimal such curve. \begin{prop} Suppose ${\cal L}$ is an even liaison class which contains curves having equal cohomology. Then up to deformation through curves with constant cohomology, there is a unique minimal curve with equal cohomology. \end{prop} \begin{proof} This is quite easy, given our remarks above. If $\Delta^2 H(C_0, t) = \cdots\quad t_1\quad\cdots\quad t_r$ are the final non-decreasing non-positive terms in the Hilbert function of $C_0$, as guaranteed by Corollary~\ref{mincoh-prop}, then it requires at least $-t_1$ basic double links to reach a curve $D$ for which $\Delta^2 H(D, t) = 0$ for $t \geq r_a(D) + 2$, and by Corollary~\ref{Delta-coh}, $D$ then has equal cohomology. On the other hand, the construction in Theorem~\ref{coh-prop} produces a curve with equal cohomology in exactly $-t_1$ steps. Also, if we reach a curve in equal cohomology in $-t_1$ steps, then every basic double link of degree $b_i$ satisfies $b_i \geq r_a(C_0) + i + 2$. Thus, by Lemma~\ref{hilbfcn-bdl}, no matter what sequence of basic double links we take, the Hilbert function of the resulting curve is invariant. Thus any two curves in this shift with equal cohomology are deformations through curves with constant cohomology; see \cite[Proposition 3.1]{BM2}. \end{proof} Now, we need to recall a result from \cite{BM2} on ``equivalence'' of basic double links. In the proof of \cite[Lemma 5.2]{BM2}, they show the following: suppose $$ C_1:(1, b_1) \rightarrow C_2:(1, b_2) \rightarrow C_3 $$ are basic double links with $b_1 < b_2$. Then the sequence $b_1, b_2$ is equivalent to the sequence $b_2 - 1, b_1 + 1$ in the sense that if we make the basic double links $$ C_1:(1, b_2 - 1) \rightarrow C_2':(1, b_1+1) \rightarrow C_3', $$ then $C_3$ and $C_3'$ are deformations of each other, through curves with constant cohomology. Note that implicit in this is the fact that the basic double linkage of degree $b_2 - 1$ can actually be made; this is also noted in their proof. We will use this idea of ``flipping'' adjacent degrees in the next result. \begin{prop} Suppose ${\cal L}$ is an even liaison class containing curves with equal cohomology, and that $C$ is a curve in ${\cal L}$ having equal cohomology. Then $C$ is obtained from the minimal curve having equal cohomology by a sequence of basic double linkages followed by a deformation through curves with constant cohomology, if necessary. Each curve in the sequence also has equal cohomology. \end{prop} \begin{proof} Assume that the deficiency module associated to ${\cal L}$ has diameter $r$. First, since $C$ is in ${\cal L}$, then by the Lazarsfeld--Rao property, after deforming $C$ through curves with constant cohomology if necessary, we may assume that there is a sequence of basic double links \begin{equation}\label{bdl1} C_0:(1,b_0) \rightarrow C_1:(1,b_1) \rightarrow\cdots \rightarrow C_m = C \end{equation} where $C_0$ is the (absolute) minimal curve in ${\cal L}$, and we can assume that $b_0 \leq \dots \leq b_{m-1}$. Since ${\cal L}$ possesses curves with equal cohomology, $C_0$ has Hilbert function given by $$ \Delta^2 H(C_0, t) = \quad\cdots\quad t_1 \quad\cdots\quad t_r $$ where $t_1\leq\dots\leq t_r \leq 0$, and $t_1$ is in degree $r_a(C_0) + 2$. Now, since $C$ has equal cohomology, then exactly $-t_1$ of the basic double links in the sequence~(\ref{bdl1}) change negatives. That is, exactly $-t_1$ of the $b_i$ satisfy $b_i \geq r_a(C_0) + i + 2$. This follows from Remark~\ref{bdl-rem2}(a). Choose the first index $s \geq 0$ for which $b_i \geq r_a(C_0) + i + 2$, and using the equivalence outlined above, flip this degree down to the first position. Note that this is possible since our original sequence of $b_i$'s is non-decreasing and $b_{s-j} < r_a(C_0) + s + 2 - j \leq b_s - j$ for $0 \leq j \leq s$. This creates an equivalent sequence of basic double links of degrees $b_0', \dots, b_{m-1}'$, where $b_0' = b_s - s$, $b_i' = b_{i-1} + 1$ for $0 < i \leq s$, and $b_i' = b_i$ for $i \geq s+1$. In particular, exactly $-t_1$ of the $b_i'$ satisfy $b_i' \geq r_a(C_0) + i + 2$, and moreover $b_0' \geq r_a(C_0) + 2$. Continue in the same manner: find the second time that a $b_i'$ changes negatives, and flip it down to the second position, and so forth. Then we end up with a sequence $c_0, \dots, c_{m-1}$ of basic double links which is equivalent to the one we started with, and which moreover has $c_i \geq r_a(C_0) + i + 2$ for $i = 0, \dots, -t_1-1$. Hence, by Remark~\ref{eqcoh&bdls}, since we change exactly $-t_1$ negative terms, all in the first $-t_1$ links, and since we eventually end up with a curve having equal cohomology, then the curve $C_{-t_1}$ and each curve from $C_{-t_1}$ on, must also have equal cohomology. This follows from Remark~\ref{bdl-rem1}. \end{proof} We recapitulate what we have proven in the next statement: \begin{thm} \label{LR-prop} Suppose ${\cal L}$ is an even liaison class containing curves with equal cohomology. Then there is a minimal shift ${\cal L}^t$ which contains a curve with equal cohomology; the curves with equal cohomology in the minimal shift are unique up to deformation through curves with constant cohomology; every curve in ${\cal L}$ with equal cohomology is obtained from the minimal one by basic double linkage and deformation through curves with constant cohomology; and finally every rightward shift of ${\cal L}^t$ contains, up to deformation, a finite, non-zero number of curves with equal cohomology. \end{thm} \begin{proof} The only part which remains to be proven is the final statement. If $C$ is a curve with equal cohomology in some shift ${\cal L}^s$ of the liaison class, then $\Delta^2 H(C, t) = 0$ for all $t \geq r_a(C) + 2$. This implies in particular that $I_C$ contains non-zero elements of degree $\geq r_a(C) + 2$. Thus, we can make a basic double link $C:(1, r_a(C) + 2) \rightarrow D$, and $D \in {\cal L}^{s+1}$ also has equal cohomology. On the other hand, if we make a basic double link of degree $> r_a(C) + 3$, then the resulting curve does not have equal cohomology, so there is only a finite number of allowable degrees. \end{proof} Theorem~\ref{LR-prop} shows that the curves with equal cohomology have a strong Lazarsfeld-Rao property, in the sense that there are unique minimal curves, every other curve is obtained by basic double linkage, and in each allowable shift, there are only a finite number of curves, up to deformation. In the case of Buchsbaum liaison classes, we can actually count the number of curves in each shift which have equal cohomology. \begin{prop} \label{number-Buchsbaum} Suppose ${\cal L}$ is a Buchsbaum even liaison class having curves with equal cohomology, and let ${\cal L}^s$ be the minimal shift in which such a curve occurs. Then for each $t \geq s$, there are exactly $2^{t-s}$ curves, up to flat deformations, having equal cohomology. \end{prop} \begin{proof} This follows from the fact that if $D \in {\cal L}^h$, then $\alpha(I_D) = 2N + h$, where $N = \sum \dim H^1_*({\Bbb P}^3, {\cal I}_D(i))$, and the description of the Hilbert function of minimal Buchsbaum curves in \cite[Proposition 2.1]{BM1}. In particular, the minimal curve $C$ having equal cohomology has Hilbert function $$ \Delta^2 H(C, t) = 1 \quad 2 \quad \cdots \quad 2\alpha + s. $$ In order to move from ${\cal L}^s$ to ${\cal L}^{s+1}$ and preserve the cohomology property, we can only make basic double links of degree $2\alpha+s$ or $2\alpha+s+1$. Similarly, we can only move from ${\cal L}^{s+1}$ to ${\cal L}^{s+2}$ by basic double links of degree $2\alpha+s+1$ or $2\alpha+s+2$. Continuing inductively, the statement is proven. \end{proof} We can also give a proof more along the lines of the original proof that the liaison classes of curves in ${\Bbb P}^3$ have the Lazarsfeld--Rao property. It is much less constructive in nature, but, in some sense, points out the naturality of our cohomological criterion of equal cohomology. Since it is so different in spirit from our previous argument, we felt it necessary to include it here. \begin{lemma} Let $\cal F$ be a rank $(r+1)$ vector bundle on ${\Bbb P}^3$ with $H^2_* ({\Bbb P}^3, {\cal F}) = 0$. Let \[ \phi_1 : \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-a_i ) \rightarrow {\cal F}, \hskip .5in a_1 \leq \dots \leq a_r \] \[ \phi_1 : \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-b_i ) \rightarrow {\cal F}, \hskip .5in b_1 \leq \dots \leq b_r \] be morphisms whose degeneracy loci are curves $C_1$ and $C_2$ with equal cohomology. Then there exists a morphism \[ \phi : \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-c_i ) \rightarrow {\cal F}, \hskip .5in c_i = \min \{ a_i ,b_i \} \] whose degeneracy locus is also a curve with equal cohomology. \end{lemma} \begin{proof} Notice that $C_1$ and $C_2$ are evenly linked, in the even liaison class determined by the stable equivalence class of $\cal F$, according to Rao's classification \cite{rao}. By \cite[Lemma 2.1]{BBM}, there exists such a $\phi$ whose degeneracy locus is a curve $C$. We just have to prove that $C$ has equal cohomology. Twisting and relabeling if necessary, we may assume that we have locally free resolutions \[ 0 \rightarrow \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-a_i ) \rightarrow {\cal F} \rightarrow {\cal I}_{C_1} \rightarrow 0 \] \[ 0 \rightarrow \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-b_i ) \rightarrow {\cal F} \rightarrow {\cal I}_{C_2} (h) \rightarrow 0. \] Notice that the deficiency module of $C_2$ is shifted $h$ places to the right of that of $C_1$. We then get \[ 0 \rightarrow H^2 ({\Bbb P}^3, {\cal I}_{C_1} (t)) \rightarrow H^3 ({\Bbb P}^3, \bigoplus {\cal O}_{{\Bbb P}^3} (t-a_i )) \rightarrow H^3 ({\Bbb P}^3, {\cal F}(t)) \rightarrow 0 \] \[ 0 \rightarrow H^2 ({\Bbb P}^3, {\cal I}_{C_2} (t+h)) \rightarrow H^3 ({\Bbb P}^3, \bigoplus {\cal O}_{{\Bbb P}^3} (t-b_i )) \rightarrow H^3 ({\Bbb P}^3, {\cal F}(t)) \rightarrow 0. \] (The first 0 comes from the assumption on the vanishing of the cohomology of $\cal F$ and the second from the fact that $h \geq 0$.) By the assumption of equal cohomology, for $t \geq r_a(C_1)$ the first term in the first sequence has the same dimension as the first term in the second sequence. Hence for $t \geq r_a(C_1)$, also the second terms are equal. Therefore \[ \left \{ a_i \ | \ r_a(C_1) - a_i \leq -4 \right \} = \left \{ b_i \ | \ r_a(C_1) - b_i \leq -4 \right \} \] (since these are the terms which contribute to the middle cohomology space in the degrees $t \geq r_a$). That is, \[ \left \{ a_i \ | \ a_i \geq r_a + 4 \right \} = \left \{ b_i \ | \ b_i \geq r_a + 4 \right \} \] Call this set $A$. Now, for any curve $Y$ with locally free resolution \[ 0 \rightarrow \bigoplus_{i=1}^{r+1} {\cal O}_{{\Bbb P}^3} (-d_i ) \rightarrow {\cal F} \rightarrow {\cal I}_{Y} (\delta ) \rightarrow 0, \] $Y$ has equal cohomology if and only if $\{ d_i \ | \ d_i \geq r_a(Y) + 4 \} = A$ (since this set determines $h^3({\Bbb P}^3, \bigoplus {\cal O}(-d_i + t))$ and hence $h^2 ({\cal I}_Y (t+\delta )$ in the desired range). The proof of the lemma follows immediately from this fact. \end{proof} \begin{cor} \label{LRprop} The set of curves in a given even liaison class which have equal cohomology satisfy the Lazarsfeld--Rao property. \end{cor} \begin{proof} The proof is identical to that in \cite{BBM}. The lemma above replaces \cite[Lemma 2.1]{BBM}. Then \cite[Proposition 2.2]{BBM} goes through to prove the uniqueness of the minimal element. Similarly, \cite[Proposition 2.3]{BBM} goes through to show the relation between the minimal element and any other curve in the even liaison class with equal cohomology. Finally, \cite[Theorem 2.4]{BBM} still works to show how to produce a curve with equal cohomology as a sequence of basic double links followed by a deformation, starting with a minimal curve with equal cohomology. The proof in \cite{BBM} shows that such a sequence exists. The fact that we start with equal cohomology and end with equal cohomology shows that every step in between has equal cohomology too. \end{proof} \begin{remark}\label{mincurves} We do not yet know of any examples of even liaison classes for which the absolute minimal curve $C_0$ is also the minimal curve with equal cohomology. \end{remark} \section{Integral Curves with Equal Cohomology} \label{integral} There has been much recent progress on further clarifying the structure of even liaison classes by giving conditions for the presence within the liaison classes of nice curves. In particular, there is information on where in a given class one can find integral curves \cite{nollet}, or smooth and connected curves in Buchsbaum classes \cite{MDP2}, and on how these curves are related to each other and to the minimal curve in the class. The paper \cite{PR} shows that at least in Buchsbaum classes, smooth and connected curves share the same Lazarsfeld--Rao properties as irreducible curves, and imply that one can obtain the integral curves within a given shift of a liaison class by deforming irreducible curves. Their calculations are based also on the work of Nollet, as well as on \cite{MDP2}. In this section, we are interested in using some of the results of \cite{nollet} to obtain some information on when an even liaison class contains curves which are integral and have equal cohomology. We first recall the relevant definitions and some results from \cite{nollet}. Let $C$ be a curve in ${\Bbb P}^3$, defined by an ideal $I=I_C$. The postulation character of $C$ is given by $\gamma_C(n) = -\Delta^3 H(C, n)$. There are three natural invariants to attach to $C$: \begin{eqnarray*} s(C) &=& \min\{\,n : \gamma_C(n) \geq 0 \,\} \\ t(C) &=& \min\{\,n : \gamma_C(n) > 0\,\} \\ t_1(C) &=& \mbox{smallest degree of a surface containing $C$ which meets}\\ &&\quad\quad\mbox{a surface of degree $s(C)$ containing $C$ properly}. \end{eqnarray*} We note for clarity that $s(C) = \alpha(C)$, the minimal degree of a generator of $I_C$, and $t_1(C) = \beta(C)$, the minimal degree for which $I_{\leq t} = \oplus_{i \le t} [I]_i$ generates an ideal of codimension $2$. Next, say that $C$ dominates a curve $D$ at height $h$ if $C$ can be obtained from $D$ by a sequence of $h$ basic double links, followed by a deformation. The central definition for this section is the following: suppose $C$ dominates the minimal curve $C_0$ in ${\cal L}$ at height $h$. Then $$ \theta_C(n) = \left\{ \begin{array}{ll} \gamma_C(n), & \mbox{if $s(C) \le n < s(C_0) + h$} \\ \gamma_C(n) - \gamma_{C_0}(n-h), & \mbox{if $n \geq s(C_0) + h$} \\ 0, & \mbox{otherwise.} \end{array} \right. $$ (This definition appears in \cite{PR} and is clearly equivalent to the one in \cite{nollet}.) We say $\theta_C$ is connected in degrees $\ge a$ if $\theta_C(b) > 0$ for $b \geq a$ implies $\theta_C(n) > 0$ for all $a \leq n \leq b$, and similarly $\theta_C$ is connected in degrees $\le b$ if $\theta_C(a) > 0 $ for some $a \leq b$ implies $\theta_C(n) > 0$ for all $a \leq n \leq b$. Finally, $\theta_C$ is connected about an interval $[a,b]$ if it is connected in degrees $\geq a$ and in degrees $\leq b$, and if $\theta_C(n) > 0$ for all $a \leq n \leq b$. Now, Nollet proves the following theorem in \cite{nollet}: \begin{thm} Let ${\cal L}$ be an even liaison class of curves in ${\Bbb P}^3$ with minimal curve $C_0$. \begin{enumerate} \item[{\rm (}a.{\rm )}] {\rm (\cite[Theorem 5.2.1]{nollet})} If $C \in {\cal L}$ is an integral curve of height $h$, then $\theta_C$ is connected about $[t(C_0)+h, t_1(C_0) + h - 1]$. \item[{\rm (}b.{\rm )}] {\rm (\cite[Theorem 5.2.5]{nollet})} Conversely, suppose $C$ dominates at height $h$ an integral curve $D$ in ${\cal L}$, which is generically Cartier on a surface of minimal degree and has either $\theta_D \not=0$ or $t(D) \leq e(D) + 4$. If $\theta_C$ is connected about $[t(C_0) + h, t_1(C_0) + h - 1]$, then $C$ can be deformed to an integral curve. \end{enumerate} \end{thm} Thus, having $\theta_C$ connected about the interval $[t(C_0) + h, t_1(C_0) + h - 1]$ is very close to having $C$ integral. As it turns out, this condition is relatively easy to check for curves with equal cohomology. We begin with some elementary calculations. Throughout, let ${\cal L}$ be an even liaison class of curves, which contains curves having equal cohomology, and let $C_0$ be the (absolute) minimal curve in ${\cal L}$ and $C$ the minimal curve in ${\cal L}$ with equal cohomology. \begin{lemma}\label{s_t_and_theta} Suppose the minimal curve $C$ with equal cohomology has height $h$ over $C_0$. Then: \begin{eqnarray*} s(C) &=& s(C_0) + h \\ t(C) &=& t(C_0) + h \\ \theta_C(n) &=& \left\{ \begin{array}{ll} -\gamma_{C_0}(n-h) & \mbox{\rm if $r_a(C_0) + h + 2 < n \leq \sigma(C_0) + h$}\\ 0 & \mbox{\rm otherwise }.\end{array}\right. \end{eqnarray*} \end{lemma} \begin{proof} Note that in order to move up the liaison class from the minimal curve $C_0$ to the curve $C$, in order to get $C$ with equal cohomology, we must take basic double links $C_i \rightarrow C_{i+1}$ of degree large enough to change the negative signs in $\Delta^2 H(C_i, t)$. Clearly, this degree $d$, say, is strictly larger than $t(C_i)$. By \cite[Corollary 2.3.5]{nollet}, then, $\gamma_{C_{i+1}}(n) = \gamma_{C_i}(n-1)$ for $n \leq d$. In particular, $s(C_{i+1}) = s(C_i) + 1$ and $t(C_{i+1}) = t(C_i) + 1$, and so the first two statements are done by induction. The assertion about $\theta_C$ then follows from the definition of $\theta_C$ and the fact that $\gamma_C(n) = 0$ for $n \geq r_a(C_0) + h + 2$, since $C$ has equal cohomology, and using that $s(C_0) \leq r_a(C_0) + 2$ and $s(C) = s(C_0) + h$. \end{proof} Now we are ready to determine which curves have both equal cohomology and connected $\theta$. Our first result takes care of a rather trivial case. \begin{prop} Suppose the minimal curve $C$ with equal cohomology has height $h$ over $C_0$ and has $\theta_C$ connected about the interval $[t(C_0) + h, t_1(C_0) + h - 1]$. Then $C=C_0$, up to deformation, and $t(C_0) = t_1(C_0)$. Conversely, if $C_0$ has equal cohomology and $t(C_0) = t_1(C_0)$, then $\theta_{C_0}$ is connected about the interval $[t(C_0), t_1(C_0) - 1]$. \end{prop} \begin{proof} First note that since $C$ has equal cohomology, we must have $t(C) \leq r_a(C) + 2$, and this then implies by Lemma~\ref{s_t_and_theta} that $t(C_0) \leq r_a(C_0) + 2$. But again by Lemma~\ref{s_t_and_theta}, this means that $\theta_C(t(C_0)+h) = 0$, so $\theta_C > 0$ on the interval $[t(C_0)+h, t_1(C_0) + h -1]$ if and only if this interval is empty. Clearly, this is equivalent to $t(C_0) = t_1(C_0)$. Now, $\theta_C$ is connected in degrees $\geq t(C_0) + h$ if and only $\theta_C = 0$ in degrees $\geq t(C_0) + h$, and again by Lemma~\ref{s_t_and_theta}, this occurs if and only if $\sigma(C_0) = r_a(C_0) + 2$. By Corollary~\ref{Delta-coh}, this implies that $C_0$ has equal cohomology, and so $C = C_0$, up to deformation. The other direction is quite trivial, since $\theta_{C_0} = 0$ and $[t(C_0), t_1(C_0) - 1]$ is empty. \end{proof} Our next proposition is the main result of this section, and tells us when an even liaison class contains curves with equal cohomology and connected $\theta$. Note that it is identical in spirit to Theorem~\ref{mincoh-prop}. \begin{prop} \label{theta-coh} Suppose ${\cal L}$ is an even liaison class of curves with minimal curves satisfying $t(C_0) < t_1(C_0)$. Then ${\cal L}$ contains a curve $D$ of height $\overline{h}$, say, having equal cohomology and $\theta_D$ connected about $[t(C_0) + \overline{h}, t_1(C_0) + \overline{h} - 1]$ if and only if $t_1(C_0) \leq \sigma(C_0) + 1$ and the Hilbert function of $C_0$ satisfies $$ \Delta^2H(C_0, r_a(C_0) + 2) < \Delta^2H(C_0, r_a(C_0) + 3) < \cdots <\Delta^2H(C_0, \sigma(C_0) - 1) < 0. $$ \end{prop} \begin{proof} First, suppose $C_0$ has $t_1(C_0) \leq \sigma(C_0) + 1$ and satisfies the condition on the Hilbert function. Note that the condition on the Hilbert function implies that $t(C_0) \leq r_a(C_0) + 2$. Then by Lemma~\ref{s_t_and_theta}, the minimal curve $C$ with equal cohomology has $\theta_C(t) = 0$ for $t(C_0) + h \leq t \leq r_a(C_0) + h + 2$ and $\theta_C(t) > 0$ for $r_a(C_0)+h+2 < t \leq \sigma(C_0) + h$, where $C$ has height $h$ over $C_0$. Now perform a sequence of $r_a(C_0) + 2 - t(C_0)$ basic double links all of degree $r_a(C_0) + h + 3$ to reach a curve $D$ of height $\overline{h} = h + r_a(C_0) + 2 - t(C_0)$. Then by a repeated application of \cite[Corollary 2.3.5]{nollet}, it is easy to check that $$ \theta_D(t) = \left\{ \begin{array}{ll} 1 & \mbox{ for $t(C_0) + \overline{h} \leq t \leq r_a(C_0) + \overline{h} + 2$} \\ \theta_C(t+h-\overline{h}) & \mbox{ for $r_a(C_0) + \overline{h} + 2 < t \leq \sigma(C_0) + \overline{h}$} \\ 0 & \mbox{ otherwise. } \end{array} \right. $$ In particular, $\theta_D$ is connected about $[t(C_0) + \overline{h}, t_1(C_0) + \overline{h} - 1]$. Also, since $D$ was obtained from $C$ by basic double links of low degree, $D$ still has equal cohomology; see Remark~\ref{bdl-rem4}(d.). Conversely, suppose $D$ is a height $\overline{h}$ curve with equal cohomology, and with $\theta_D$ connected about the interval $[t(C_0) + \overline{h}, t_1(C_0) + \overline{h} - 1]$. Then there is a sequence of basic double links $$ C:(1,b_0) \rightarrow C_1:(1, b_1) \rightarrow \cdots \rightarrow D $$ where $C$ is the minimal curve in ${\cal L}$ with equal cohomology (of height $h$, say) and where each $b_i$ satisfies $b_i \le r_a(C_0) + h + i + 3$ (see Remark~\ref{eqcoh&bdls}(d.) or the proof of Corollary~\ref{LRprop}). If $C=C_0$, then $r_a(C_0) + 2 = \sigma(C_0)$, so there is nothing to show. Hence we may assume that $C_0$ does not have equal cohomology, and this means that $\Delta^2 H(C_0, r_a(C_0) + 2) < 0$, which in turn implies $\gamma_{C_0}(r_a(C_0) + 3) < 0$ and $t(C_0) \leq r_a(C_0) + 2$. By a repeated use of \cite[Corollary 2.3.5]{nollet}, $\gamma_D(n) = \gamma_C(n-\overline{h}+h)$ for $r_a(C_0) + \overline{h}+3 \leq n$. But by Lemma~\ref{s_t_and_theta}, $\gamma_C(n-\overline{h}+h) = -\gamma_{C_0}(n-\overline{h})$ for $r_a(C_0) + \overline{h} + 3 \leq n \leq \sigma(C_0) + \overline{h}$. Now, $\theta_D$ is connected in degree $\geq t(C_0) + \overline{h}$, and our assumption that $t(C_0) < t_1(C_0)$ implies in particular that $\theta_D(t(C_0) + \overline{h}) > 0$. Also, $\theta_D(\sigma(C_0) + \overline{h}) = -\gamma_{C_0}(\sigma(C_0)) > 0$, so $\theta_D > 0$ on the interval $[t(C_0) + \overline{h}, \sigma(C_0) + \overline{h}]$. Next, note that $\theta_D$ is positive on the interval $[r_a(C_0) + \overline{h} + 3, \sigma(C_0) + \overline{h}]$, since this interval is contained in the interval $[t(C_0) + \overline{h}, \sigma(C_0) + \overline{h}]$. Thus, we have $0 < \theta_D(t) = -\gamma_{C_0}(t-\overline{h})$ for $r_a(C_0) + \overline{h} + 3 \leq t \leq \sigma(C_0) + \overline{h}$. This clearly implies that $\gamma_{C_0}(t) < 0$ for $r_a(C_0) + 3 \leq t\leq \sigma(C_0)$, and this means that $\Delta^2 H(C_0, t)$ is strictly increasing in the given range. Finally, to see that $t_1(C_0) \leq \sigma(C_0) + 1$, note that the connectedness property of $\theta_D$ implies that $\theta_D > 0$ on $[r_a(C_0) + \overline{h} + 3, t_1(C_0) + \overline{h}]$, but clearly $\theta_D(t) = 0$ for $t >\sigma(C_0) + \overline{h}$, by the argument above. \end{proof} \section{Degrees of Generators and Liaison Classes of Curves} In this section, we go back to studying degrees of generators of the ideals defining space curves, by using the results on cohomology given in the previous sections. We are able to give some nice conditions on the degrees of the components of the deficiency module associated to a liaison class in order for the class to contain a curve whose ideal has a generator of maximal degree. Since knowledge about curves with equal cohomology in a given liaison class depends so crucially on knowing the Hilbert function of the minimal curve in the liaison class, our characterizations for when a liaison class contains curves with generators of high degree work best when we already know the Hilbert function of the minimal curve. To do this, we have to make some extra assumptions on the liaison class. We have concentrated on cohomological criteria, and two very clean statements are given in Proposition~\ref{degree-maxcorank} and in Proposition~\ref{degree-Buchsbaum}. Similarly, using our results on integral curves, we give some results on existence within a liaison class of integral curves with generators of maximal degree. As we showed in the previous sections, if $C$ is defined by an ideal $I = I_C$, then the property of $I_C$ having a generator of high degree is very closely related to having $h^1({\Bbb P}^3, {\cal I}_C(t)) = h^2({\Bbb P}^3, {\cal I}_C(t))$ in a large number of places, and furthermore, curves with this cohomology property are easily constructed by basic double linkage, as long as we know the minimal curve in a liaison class. However, we should remark that in order to construct a curve whose ideal has a high degree generator, we need to choose the degrees of the basic double links carefully, since it is possible to make $h^1 = h^2$ in the maximum number of places, without introducing a high degree generator. The following example should clarify this somewhat. \begin{example} Start with the Buchsbaum liaison class ${\cal L} = {\cal L}_{4,1}$, whose deficiency module has two consecutive components of dimensions $4$ and $1$, respectively. Then the minimal curve $C_0$ in ${\cal L}$ has Hilbert function (see \cite[Corollary 2.18]{BM1}) $$ \Delta^2 H(C_0, t) = 1 \quad 2 \quad \cdots \quad 10 \quad -2 \quad -1. $$ Taking the two basic double links $$ C_0:(1, 13) \rightarrow C_1:(1, 13) \rightarrow C_2 $$ produces the curve $C_2$, for which $h^1 = h^2$ in the last two places, and which has $\sigma(C_2) = 12$, but $\omega(I_{C_2}) = 13$. Thus the bound in Proposition~\ref{maxdegree} is not obtained, even though $C_2$ does have the cohomology property. On the other hand, note that if we take the basic double links $$ C_0:(1, 12) \rightarrow C'_1:(1,14) \rightarrow C'_2, $$ then $C'_2$ has $h^1 = h^2$ in the last two places, and also has $\sigma(C'_2) = 12 = \omega(I_{C'_2}) - 2$. So the bound is achieved in this case. \end{example} Generally speaking, this second sequence of basic double links produces curves whose defining ideals have a high degree generator. Note that it is essentially the procedure given by Theorem~\ref{coh-prop}. However, we need to require that there are no ``trailing zeroes'' on the end of the second difference of the Hilbert function. We formalize this in the next statement. \begin{prop}\label{degree} Suppose ${\cal L}$ is an even liaison class of curves in ${\Bbb P}^3$, with minimal curve $C_0$, and let $r = \mathop{\rm diam\,} H^1_*({\Bbb P}^3, {\cal I}_C)$. If the Hilbert function of $C_0$ satisfies $$ \Delta^2 H(C_0, r_a(C_0)+2) \leq \dots \leq \Delta^2 H(C_0, r_o(C_0)+2) < 0, $$ then there exists a curve $C$ in ${\cal L}$ whose defining ideal $I_C$ satisfies $\omega(I_C) = \sigma(S/I_C) + r$. \end{prop} \begin{proof} We follow the construction of basic double links in the first part of Theorem~\ref{coh-prop}. Perform the given sequence of basic double links, up to the next to the last stage. Since $\Delta^2 H(C_0, r_o(C_0) + 2) < 0$, then at this stage we see that $$ \Delta^2 H(C_{m-1}, t) = \cdots \quad -1 \quad -1 \quad \cdots \quad -1 \quad 0 \quad \cdots, $$ where there are $r$ terms of $-1$, and the rightmost term occurs in degree $r_o(C_0)+2+m$. Thus our final basic double link $C_{m-1}:(1, r_o(C_0)+m+4) \rightarrow C_m$ produces the curve $C_m$, whose ideal $I_{C_m}$ has a generator of degree $r_o(C_0) + m + 4$, and such that $\sigma(C_m) = r_o(C_0) + m - r + 4$. Hence $\omega(I_{C_m}) = \sigma(C_m) + r$, which is what we wanted to show. \end{proof} In fact, using the structure theory for curves with equal cohomology developed in the previous section, we can say more about the curves in an even liaison class having maximal degree generators. Namely, the minimal such curve occurs in the shift ${\cal L}^s$ of ${\cal L}$, where $s=-\Delta^2 H(C_0, r_a+2)$, and every other such curve is obtained by a sequence of basic double links from this minimal one, followed if necessary by a deformation. Moreover, up to a flat deformation, there are only a finite number of curves with a maximal degree generator in each allowable shift. \begin{remark} It is interesting to note that curves in a non-arithmetically Cohen--Macaulay liaison class which have a generator of maximal degree in fact have all of their generators of relatively high degree. Indeed, since the minimal such curve $C$ lies in the shift ${\cal L}^s$ as above, $\alpha(I_C) = \alpha(I_{C_0}) + s$. So, at least if the minimal curve does not already have a maximal degree generator (see Remark~\ref{mincurves}), we are forced to increase all the degrees of the generators. This is in some contrast with the arithmetically Cohen--Macaulay case, where there are curves $D$ with quadric generators and for which $\omega(D) = \sigma(D)$, the maximum possible. For example, let $D$ be the union of a plane curve of degree $m$ and a line, which meet at one point. Then it is easy to see that $I_D$ has quadric generators and $\omega(D) = \sigma(D) = m$. On the other hand, even within a non-arithmetically Cohen--Macaulay liaison class, we can make the difference $\omega(I_C) -\alpha(I_C)$ arbitrarily large when $\omega(I_C)$ is maximal. For this, simply take the minimal curve $C$ with a maximal degree generator, and form $m$ basic double links all of degree $\alpha(I_C)$. Then the resulting curve $C_m$ still has $\omega(I_{C_m})$ maximal (see Remark~\ref{eqcoh&bdls}(d.)), and has $\omega(I_{C_m}) - \alpha(I_{C_m}) = \omega(I_C) - \alpha(I_{C}) + m$. \end{remark} Next, we want to interpret the conditions on Hilbert functions only in terms of cohomology. As we noted above, to do this we need to make some extra assumptions to allow us to calculate the Hilbert function of the minimal curve. Our first result is for maximal corank curves, and the second for Buchsbaum curves. \begin{cor} \label{degree-maxcorank} Suppose ${\cal L}$ is an even liaison class of curves in ${\Bbb P}^3$, with minimal curve $C_0$. Assume that $e(C_0) < r_a(C_0)$ {\rm (}i.e., $C_0$ has maximal corank{\rm )}. Then ${\cal L}$ contains a curve $C$ such that $\omega(I_C) = \sigma(C) + \mathop{\rm diam\,} H^1_*({\Bbb P}^3, {\cal I}_C)$ if and only if $$ h^1({\Bbb P}^3, {\cal I}_{C_0}(t)) \geq 3 h^1({\Bbb P}^3, {\cal I}_{C_0}(t+1)) - 3 h^1({\Bbb P}^3, {\cal I}_{C_0}(t+2)) + h^1({\Bbb P}^3, {\cal I}_{C_0}(t+3)), $$ for $t=r_a(C_0), \dots, r_o(C_0)$. \end{cor} \begin{proof} The conditions on the cohomology modules guarantee via Lemma~\ref{hilbfcn_eqn} that $$ \Delta^2 H(C_0, r_o - r + 2) \leq \dots \leq \Delta^2(r_o + 2, C_0) < 0, $$ and so sufficiency follows from Proposition~\ref{degree}. To see necessity, note that if ${\cal L}$ contains a curve with a generator of maximal degree, then by Proposition~\ref{equal-coh}, that curve has equal cohomology, and so by Corollary~\ref{mincoh-prop}, the minimal curve in ${\cal L}$ has Hilbert function whose second difference ends in a sequence of non-decreasing negative terms, and then Lemma~\ref{hilbfcn_eqn} translates this back to the required statement about cohomology. \end{proof} Recall that a Buchsbaum liaison class is completely determined by the dimensions of the graded components of the associated deficiency module. We will write ${\cal L}_{n_1\dots n_r}$ for the Buchsbaum class associated to the graded module $M = \oplus [M]_i$, where $\dim_k [M]_i = n_i$ and is zero otherwise, and where we assume that $n_1, n_r > 0$ and $n_i \ge 0$ for $1 < i < r$. \begin{prop} \label{degree-Buchsbaum} Suppose ${\cal L} = {\cal L}_{n_1 \dots n_r}$ is a Buchsbaum even liaison class. Then ${\cal L}$ contains a curve $C$ such that $\omega(I_C) = \sigma(I_C) + r$ if and only if $n_i \geq 3 n_{i+1}$ for $i = 1, \dots r-1$. \end{prop} \begin{proof} It follows from \cite[Corollary 2.18]{BM1} that the conditions on the deficiency module guarantee that the Hilbert function of $C_0$ satisfies the conditions given in Proposition~\ref{degree}, and we can therefore use that result to prove sufficiency. On the other hand, if $C$ is a curve in ${\cal L}$ whose ideal has a generator of maximal degree, then by Proposition~\ref{equal-coh}, $C$ has equal cohomology. Thus by Corollary~\ref{mincoh-prop}, the minimal curve $C_0 \in {\cal L}$ has Hilbert function satisfying $$ \Delta^2 H(C_0, r_o(C_0) - r + 2) \leq \dots \leq \Delta^2 H(C_0, r_o(C_0) + 2) \leq 0. $$ But now it follows from the description of the Hilbert function for minimal Buchsbaum curves given in \cite[Corollary 2.18]{BM1} that $n_i \geq 3 n_{i+1}$, for each $i = 1, \dots, r-1$. \end{proof} \begin{remark} Proposition \ref{degree} is, in a sense, a complete answer to the problem of determining which even liaison classes $\cal L$ contain a curve $C$ whose defining ideal $I_C$ satisfies $\omega (I_C ) = \sigma (S/I_C ) + \hbox{diam } H^1_* ({\Bbb P}^3, {\cal I}_C )$. However, it is generally not easy to tell, for a given even liaison class, what the Hilbert function of the corresponding minimal curve is. So it is worth noting that there are also times when one can tell directly from the associated deficiency module $M$ (defined up to shift) that such a curve does not exist. If $M$ is annihilated by the maximal ideal (i.e. if $C$ is Buchsbaum), then the necessary and sufficient condition for the existence of such a curve is given in terms of the dimensions of the components of $M$ (Proposition~\ref{degree-Buchsbaum}). Similarly, if $\cal L$ contains any curve with maximal corank then the minimal curve has maximal corank as well, since any basic double link increases $r_a $ by exactly 1 and $e$ by at least 1 (\cite[Lemma~1.14]{BM1}). Hence again it reduces to a question of the dimensions of the module components. It is also true that $\cal L$ contains curves of maximal rank (i.e.\ $r_0 < \alpha$) if and only if the minimal curve in $\cal L$ has maximal rank, since a basic double link increases $r_o$ by exactly 1 and increases $\alpha$ by at most 1. (This was first observed in \cite[Theorem~2.1]{BM3}.) Then it follows from Theorem~\ref{coh-prop} that if the deficiency module $M$ associated to $\cal L$ has diameter 3 or more, and if $\cal L$ contains any curve of maximal rank, then $\cal L$ does not contain any curve achieving our bound on $\omega$. Finally, we observe that it follows from Proposition~\ref{equal-coh} and Corollary~\ref{almostBuchs-deg} that if $\mathop{\rm diam\,} K < \mathop{\rm diam\,} M$ then $\cal L$ contains no curve achieving our bound. This immediately rules out a huge number of even liaison classes, since ``most'' classes will have a module containing at least one pair of consecutive components, for which multiplication by a general linear form is injective. \end{remark} By using the results in Section~\ref{integral}, we can prove similar results for the existence of integral curves with generators of maximal degree. However, because the theorems in that section only go one direction, and because deformations do not in general preserve integrality or degrees of generators, we can only get necessity. \begin{thm} Suppose ${\cal L}$ is a liaison class whose minimal curve $C_0$ satisfies $t(C_0) < t_1(C_0)$. If ${\cal L}$ contains an integral curve with a generator of maximal degree, then the minimal curve $C_0$ has Hilbert function which satisfies $$ \Delta^2 H(C_0, r_a(C_0) + 2) < \cdots < \Delta^2 H(C_0, \sigma(C_0) - 1) < 0. $$ \end{thm} \begin{proof} This follows immediately from Proposition~\ref{theta-coh}. \end{proof} As before, with extra assumptions on the liaison class, we can give a cohomological criterion. \begin{prop} Suppose ${\cal L}$ is a liaison class whose minimal curve $C_0$ satisfies $t(C_0) < t_1(C_0)$. \begin{enumerate} \item[{\rm (}a.{\rm )}] If $C_0$ has maximal corank, and if ${\cal L}$ contains integral curves with generators of maximal degree, then $$ h^1({\Bbb P}^3, {\cal I}_{C_0}(t)) > 3 h^1({\Bbb P}^3, {\cal I}_{C_0}(t+1)) - 3 h^1({\Bbb P}^3, {\cal I}_{C_0}(t+2)) + h^1({\Bbb P}^3, {\cal I}_{C_0}(t+3)), $$ for $t = r_a(C_0), \dots, r_o(C_0)$. \item[{\rm (}b.{\rm )}] If ${\cal L}_{n_1\dots n_r}$ is a Buchsbaum liaison class containing integral curves with generators of maximal degree, then $n_i > 3n_{i+1}$ for $i = 1, \dots, n_r$. \end{enumerate} \end{prop} \begin{proof} This follows exactly as before, using Proposition~\ref{theta-coh}. \end{proof}
1995-03-28T07:20:38
9503
alg-geom/9503019
en
https://arxiv.org/abs/alg-geom/9503019
[ "alg-geom", "math.AG" ]
alg-geom/9503019
Martin Pikaart
Martin Pikaart
An orbifold partition of ${\overline{M}_g^n}$
16 pages, Latex Version 2.09, will appear in The Moduli space of Curves (eds. Dijkgraaf, Faber, van der Geer), Progress in Math., Birkh"auser
null
null
Utrecht preprint Nov. 1994, nr. 882
null
We define a partition of ${\overline{M}_g^n}$ and show that the cohomology of ${\overline{M}_g^n}$ in a given degree admits a filtration whose respective quotients are isomorphic to the shifted cohomology groups of the parts if $g$ is sufficiently large. This implies that the map $H^k({\overline{M}_g^n}) \ra H^k(M_g^n)$ is onto and that the Hodge structure of $H^k(M_g^n)$ is pure of weight $k$ if $g \geq 2k+1$. Our main ingredient is the stability theorem of Harer and Ivanov.
[ { "version": "v1", "created": "Mon, 27 Mar 1995 08:47:43 GMT" } ]
2008-02-03T00:00:00
[ [ "Pikaart", "Martin", "" ] ]
alg-geom
\section{Introduction} The moduli space of smooth $n$-pointed complex curves of genus $g$ is a quasi-projective orbifold $M_g^n$ of dimension $3g-3+n$ (where as usual, we assume that $2g-2+n>0$). It is compactified by the moduli space of stable pointed curves, $\Mgnbar$, which is a projective orbifold. We will write $M_g$ and $\Mgbar$ if $n=0$. We need some terminology in order to state our main result. Let $X$ be an irreducible orbifold. An {\it orbifold partition} of $X$ is a finite filtration by closed subvarieties $$X_\bullet:= (X =X_0 \supset X_1 \supset \cdots \supset X_{m+1} =\emptyset),$$ such that $ Y_\alpha:=X_\alpha \setminus X_{\alpha+1} $ is an orbifold. If $\dim Y_\alpha >\dim Y_{\alpha +1}$ this is called a {\it stratification}. A connected component of a $Y_\alpha $ is called a {\it part} (respectively a {\it stratum}). For example, a stratification of $\Mgnbar$ is defined by the subvarieties $X_i:=\{ \mbox{curves with at least $i$ singularities}\}$ for $i=0,1,\dots$; the strata are just the loci that parametrize stable pointed curves of a fixed topological type. We will therefore refer to this as the {\it stratification of $\Mgnbar$ by topological type}. A simple example of an orbifold partition (of $ {\bf C}^2$) that is not a stratification is given by ${\bf C}^2 \supset L_1 \cup L_2 \supset L_1$, where $L_i$ is the ith coordinate axis; the parts are ${\bf C}^2 \setminus (L_1\cup L_2)$, $L_2 \setminus \{ 0 \}$ and $L_1$. Let us say that an orbifold partition $X_\bullet $ of $X$ {\it filters cohomology up to degree $k$} if the Gysin map $$H^{j-2 {\it codim} Y_\alpha}(Y_\alpha)(-{\it codim} Y_\alpha) \rightarrow H^j(X \setminus X_{\alpha+1})$$ is injective for all $j \leq k$ and all $\alpha$. (Here and throughout this paper cohomology is taken with rational coefficients.) For $2{\it codim} Y_\alpha > k$ this condition is empty. Thus, if one is only interested in a partition filtering cohomology up to degree $k$, it suffices to consider $X_0 \supset \cdots \supset X_{\alpha+1}$ such that $X_{\alpha+1}$ has codimension at least $\frac{1}{2}k$. It is easily seen that if a partition filters cohomology up to degree $k$ there exists a filtration on $H^j(X)$ for $j \leq k$, such that there is a canonical morphism of mixed Hodge structures which maps the respective quotients isomorphically onto the groups $ H^{j -2 {\it codim} Y_\alpha}(Y_\alpha)(-{\it codim} Y_\alpha) \mbox{ for } j<k \mbox{ and all } \alpha$. We can now formulate our main result. \medbreak\noindent{\bf Corollary \ref{coho Mgnbar}} {\it Given $k \geq 0$, then for $g$ large enough, $\Mgnbar$ admits an orbifold partition which has $M_g^n$ as its open part, is coarser than the stratification by topological type and filters cohomology up to degree $k$.} \medbreak We shall see that the stratification by topological type does not have this property. We deduce from the main result the following corollary. \medbreak\noindent{\bf Corollary \ref{pure Hodge}} {\it If $g \geq 2k+1$, then the restriction map $H^k(\Mgnbar) \ra H^k(M_g^n)$ is surjective; consequently the mixed Hodge structure on $H^k(M_g)$ is pure of weight $k$.} \medbreak Corollary \ref{coho Mgnbar} enables us to define the ``stable cohomology'' of $\Mgnbar$ as $g$ goes to infinity and $n$ is fixed. This stable cohomology is not finitely generated. For example, if $n=0$, the generators in degree 2 are the tautological class and the boundary classes, naturally indexed by $0,1,2, \dots$. Using additional properties of the partition that we construct we prove that for any $n \geq 0$ the cohomology of $\Mgnbar$ is not of Tate type if $g$ is sufficiently large. Let us sketch the proof of corollary \ref{coho Mgnbar} and give the motivation for the partition we define. (See Sections \ref{graphs and partition} and \ref{results} for details.) We take $n=0$ for simplicity. Let $D_0$ be the locus in $\Mgbar$ parametrizing irreducible curves with one node and $D_1$ the locus in $\Mgbar$ parametrizing curves with one smooth component of genus $g-1$ and one of genus 1, joined in one point. Consider the maps $f_i:H^{l-2}(D_i) \ra H^l(D_i)$ for $i=0$ or 1, given by taking the cup product with the first Chern class of the normal bundle of $D_i$. The normal direction of $D_i$ corresponds to smoothening the unique singular point. Notice that $f_i$ can be obtained as the composite of the Gysin map $H^{l-2}(D_i)(-1) \ra H^l(M_g \cup D_i)$ and the obvious restriction map. The unique singular point lies in both cases on at least one local component of high genus. As we shall see in Section \ref{results}, this will imply that the $f_i$ are injective in low degree and a fortiori the Gysin maps are injective. However, this does not carry over to codimension 2. Let $D_{01}$ be the locus in $\Mgbar$ parametrizing curves with two nodes and two components, one smooth component of genus $g-1$ and one singular component of geometric genus 0. Clearly one of the nodes lies on two local components of genus 0. As we shall see, this implies that $H^{l-4}(D_{01})(-2) \ra H^l(D_{01})$ is not injective (in fact zero). A way to restore injectivity is by coarsening the partition. If we let $D$ be the union of $ D_1$ and $D_{01}$, then the normal direction of $D$ corresponds to smoothening the singular point on the component of genus $g-1$ and we get an injective map as we wanted. Smoothening the singularity on the genus 0 component has become a direction along the part. It will turn out that we can define the partition completely in terms of graphs and certain subgraphs. Our parts will be (irreducible) unions of strata of the stratification by topological type; locally they are like the parts in the given example of a non-stratification of ${\bf C}^2$. The unique open dense stratum of the stratification by topological type in a given part will be refered to as its {\it generic} stratum. For every part, its generic stratum will have a graph with the property that every edge corresponds to a singularity lying on at least one local component of high genus. Unfortunately, I was unable to define such a subgraph canonically; Definition \ref{specifiek} involves the choice of the number $\alpha$. The point is that smoothening many singularities lying on low genus components may yield a component of high genus. See Remark \ref{infinite genus} for further comments. In Section \ref{graphs and partition} we prove the existence of an orbifold partition using only some formal combinatorial properties of graphs. In Section \ref{stable and weak} we define weak subgraphs and prove the combinatorial properties needed in the previous section. Section \ref{results} contains the results. \smallskip I thank Eduard Looijenga for suggesting the problem to me and for many helpful conversations. I am grateful to the referee for useful comments. \section{Stable graphs and the orbifold partition}\label{graphs and partition} All graphs considered in this paper are weighted, that is, each vertex has a non-negative integer associated to it. For a stable $n$-pointed curve $C$, we define its {\it stable graph} as follows. Each irreducible component of $C$ defines a vertex, the weight of the vertex being the genus of the normalisaton of the component. Omitting from the normalisation of such an irreducible component the inverse images of the singular and marked points on $C$ yields a smooth projective curve minus a finite number of points. For each missing point we draw a half edge emanating from the vertex defined by that connected component and for every singular point on $C$ we join the two half edges associated to it to obtain a whole edge. We omit the word whole if it is clear that we consider a whole edge. The remaining half edges, which correspond to the $n$ marked points on the curve, will be called {\it loose half edges}. Thus, unless we specify, a half edge can be either half of a whole edge or a loose half edge. We define a {\it stable graph} ($n$-pointed of genus $g$) as the stable graph of some stable $n$-pointed curve of genus $g$. An automorphism of a stable graph is an automorphism of the underlying unweighted graph preserving the weights. Given a stable graph one can recover a topological model for its stable $n$-pointed curve by taking one curve for every vertex with genus equal to the weight of the vertex, omitting small open discs for every half edge emanating from that vertex, glueing the appropriate boundaries and contracting them. Notice that if we had incorporated the ordering of the $n$ points in the definition of a stable graph, then our graphs would correspond bijectively to strata of the stratification by topological type. A {\it subgraph} is a subset of a stable graph with the property that if it contains a half edge it contains its unique vertex. By a {\it full subgraph} we mean a subgraph which contains every half edge emanating from one of its vertices. A connected component of a full subgraph is a stable graph. In the same way as above, one can associate to a connected component of a full subgraph a topological model of a curve. We define the {\it genus} of a connected component of a full subgraph as the arithmetic genus of the curve associated to that connected component. The {\it genus} of a full subgraph is defined as the sum of the genera of its connected components. In other words, the genus of a full subgraph equals the first Betti number of that full subgrabh plus the sum of the weights taken over the vertices in the full subgraph. We denote the greatest integer less than or equal to a real number $a$ by $[a]$. Fix $k$ and let ${\cal G}_{k}(g,n)$ be the set of (isomorphism classes of) stable graphs, $n$-pointed of genus $g$ and at most $[k/2+1]$ whole edges. If there is no chance of confusion, we write ${\cal G}$ for ${\cal G}_{k}(g,n)$. (For the reason we only consider graphs with this many edges, see the remark made in the introduction following the definition of a orbifold partition which filters cohomology.) The maximal number of vertices for a graph in ${\cal G}$ is $[k/2+1]+1$. From now on we will only consider graphs in ${\cal G}$. We may regard a stable graph as a one dimensional topological space. If $e$ is a whole edge, we denote the contraction of $e$ by $\pi_e: \Gamma \rightarrow \Gamma /e$ and the image of its end vertices by $\pi_e(e)$. We make $\Gamma/e$ into a weighted graph as follows. If $p$ is a vertex of $\Gamma$ not incident with $e$, then the weight of $\pi_e(p)$ is that of $p$. The weight of $\pi_e(e)$ is defined as the sum of the weights of the two end vertices if $e$ is not a loop and the weight of its unique end vertex plus one otherwise. Clearly this makes $\Gamma/e$ into an element of ${\cal G}_k(g,n)$, corresponding to the stable curve obtained by smoothening the singularity corresponding to the edge $e$. If $\Delta$ is a full subgraph of $\Gamma$, we denote the quotient of $\Gamma$ obtained by contracting all whole edges of $\Delta$ by $\Gamma/ \Delta$. The image of a full subgraph under $\pi_e$ is full, provided that the contracted edge has either both or neither of its end vertices in that full subgraph. The permutation group on $n$ elements, $S_n$, acts on $\Mgnbar$ fixing the locus of points parametrizing curves with at least a fixed number of singularities. Such loci define the stratification by topological type, see introduction. Thus $S_n$ acts on the stratification by topological type. The orbits of the $S_n$-action on the stratification by topological type of $\Mgnbar $ minus loci of codimension at least $[k/2+1]+1$ are in one to one correspondence with the elements of ${\cal G}_k(g,n)$. Giving a partition of the first set is therefore equivalent to giving a partition of ${\cal G}_k(g,n)$. This, in turn, corresponds to a function $$\phi : {\cal G}_k(g,n) \ra {\cal G}_k(g,n),$$ as follows. Recall that we want the parts we seek to be irreducible unions of strata of the stratification by topological type. A graph will be mapped to the graph of the generic stratum of the part it will belong to. (Compare the example in the introduction: the graph of $\Delta_{01}$ is mapped to the graph of $\Delta_1$.) This implies that $\phi$ contracts a certain subgraph. So we have to assign to every graph in ${\cal G}_k(g,n)$ a subgraph in such a way that the function $\phi$, defined as contraction of that subgraph, corresponds to an orbifold partition. One important ingredient we need to obtain an orbifold partition is a partial order on the image of $\phi$, such that the union of all parts greater than or equal to a given part will be a Zariski open neighbourhood of that part. Now we will state the formal properties of the graphs needed to obtain an orbifold partition- which will turn out to filter cohomology. \begin{definition} Suppose we are given, for every $\Gamma$ in ${\cal G}_k(g,n)$, a subgraph $\Gamma_W$. Define $\phi:{\cal G}_k(g,n) \ra {\cal G}_k(g,n)$ by $\phi(\Gamma) =\Gamma/\Gamma_W$. Denote the image of $\phi$ by $I$. For $ \Gamma$ in $I$ define $S_\Gamma^0$ to be the locus in $\Mgnbar$ whose points correspond to stable $n$-pointed curves with graph $\Gamma$. Define $S_\Gamma$ (respectively, $S$) to be the locus in $\Mgnbar$ whose points correspond to stable $n$-pointed curves with graph in $\phi^{-1}(\Gamma)$ (respectively, ${\cal G}$). \end{definition} If $C$ is a curve with graph $\Gamma$, then we will call the irreducible components of $C$ corresponding to vertices in $\Gamma_W$ {\it weak components}. Analogously we define {\it strong components}. \begin{proposition} \label{formal} Notations as in the previous definition. Suppose a partial order is given on $I$, such that the following properties hold: \begin{enumerate} \item[a.] $\Gamma_W$ is a proper full subgraph of $\Gamma$ which is invariant under the automorphism group of the stable graph $\Gamma$; \item[b.] if $e$ is a whole edge of $\Gamma$ not contained in $\Gamma_W$, then $\Gamma/e > \Gamma$; \item[c.] if $e$ is a whole edge contained in $\Gamma_W$ , then the image of $\Gamma_W$ under contraction of $e$ is $(\Gamma/e)_W$; \item[d.] a whole edge $e$ is contained in $\Gamma_W$ if and only if $\pi_e(e)$ is contained in $(\Gamma/e)_W$; \item[e.] if $k=1$ (respectively $k \geq 2$) and $p$ a vertex of $\Gamma$ such that the weight of $p$ is 0 (respectively the weight of $p$ is at most $k+2$), then $\Gamma_W$ contains the vertex $p$. \end{enumerate} Then the following hold : \begin{enumerate} \item $S_{{\it max }(I)} = M_g^n$; \item for all $\Gamma \in I$: (the connected components of) $S_\Gamma$ and $S_\Gamma^0$ are orbifolds; the generic curve $C_\Gamma$ has graph $\Gamma$; \item for all $\Gamma \in I$: $\cup_{\Delta \geq \Gamma} S_\Delta$ is a Zariski open neighbourhood of $S_\Gamma$; \item the (connected components of the) $S_\Gamma$ are parts of an orbifold partition; \item for all $\Gamma \in I$: the strong (respectively, weak) components of $C_\Gamma$ do not (respectively, do) degenerate in $S_\Gamma$; \item for all $\Gamma \in I$: every singular point of $C_\Gamma$ lies on at least one strong component of $C_\Gamma$; \item for all $\Gamma \in I$: if $C'$ is a strong component of $C_\Gamma$ and $k=1$ (respectively $k \geq 2$) then the genus of $C'$ is at least 1 (respectively at least $k+3$). \end{enumerate} \end{proposition} \begin{proof} {\it 1}. Notice that by property {\it a}, $\Gamma_W$ is a proper full subgraph, which means that $\Gamma/\Gamma_W$ has no whole edges if and only if $\Gamma$ has no whole edges. By property {\it b}, the stable graph of a smooth $n$-pointed curve is larger than every other graph; consequently, $S_{max(I)}=M_g^n$. {\it 2}. We already remarked that (a connected component of) $S_\Gamma^0$ is an orbifold, namely it is a stratum of the stratification by topological type. By the definition of $\phi$ we have that $S_\Gamma$ is contained in the closure of $S_\Gamma^0$. So it suffices to prove that it is locally closed and without self intersection. It follows from property {\it c} that $S_\Gamma$ is locally closed. Suppose $S_\Gamma$ has self intersection; this is necessarily a normal crossing. Let $\Delta$ be the graph corresponding to the generic locus of self intersection of $S_\Gamma$ and suppose for simplicity that there are two local branches of $S_\Gamma$. Thus, locally in a neighbourhood of $S_\Delta$, there is an involution permuting the two branches of $S_\Gamma$. This involution induces an involution $i$ on $\Delta$ which keeps $\Delta_W$ fixed but not pointwise fixed. If $e$ is an edge of $\Delta_W$ such that $\Delta/e$ is the graph of one of the local branches, then $\Delta/i(e)$ is the (isomorphic) graph of the other local branch. But by property {\it a}, $\Delta_W$ contains both $e$ and $i(e)$, so $S_\Gamma$ contains the plane spanned by the two local branches, contradiction. {\it 3}. This follows directly from statement {\it 2} and property {\it b}. {\it 4}. Choose an order reversing injective map $m: I \ra {\bf N}$ and define $X_i:= \cup_{m(\Gamma) \leq i} S_\Gamma$. The $X_i$ form a filtration which defines an orbifold partition by statements {\it 2} and {\it 3} whose parts are the $S_\Gamma$. {\it 5}. Follows from property {\it d}. {\it 6}. By property {\it c}, we have that the image of $\phi$ equals the fix point set of $\phi$, which means that for $\Gamma$ in the image of $\phi$, every whole edge of $\Gamma$ has at least one end vertex not in $\Gamma_W$. {\it 7}. This is clear from property {\it e}. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} Now we have reduced the problem of finding a partition to a purely combinatorial one, which will be dealt with in the next section. \section{Stable graphs, weak subgraphs}\label{stable and weak} We will now start out to define a full subgraph $\Gamma_W$ for every $\Gamma$ in ${\cal G}_k(g,n)$ and prove the properties a,b,c,d and e of Proposition \ref{formal}. For any full subgraph $\Delta$ of a stable graph $\Gamma$, we denote its number of vertices respectively its genus by $v(\Delta)$ resp. $g(\Delta)$. We denote the weight of a vertex $P$ by $w(P)$. Let $k \in {\bf Z}_{\geq 0}$ be given. Let $\Phi_k :{\bf Z}_{\geq 0} \ra {\bf R}_{\geq 0} $ be a function satisfying $$2 \Phi_k(n) +[\frac{1}{2}k+1] \leq \Phi_k(n-1), \mbox{ if } n > 0.$$ In particular, it is a decreasing function. \begin{proposition} \label{union} Let $k \geq 0$ be given. Let $\Delta_1$ and $\Delta_2$ be full subgraphs of a stable graph $\Gamma$. Let $\Delta$ be the full subgraph on the vertices of $\Delta_1$ and $\Delta_2$. If $g(\Delta_i) \leq \Phi_k(v(\Gamma)-v(\Delta_i)) $ for $i=1$ and $i=2$, then also $g(\Delta) \leq \Phi_k(v(\Gamma)-v(\Delta)) $. \end{proposition} \begin{proof} We may suppose $ v(\Delta_1) \geq 1$ and $ v(\Delta_1) \geq v(\Delta_2)$. The statement is trivial if $\Delta_2$ is contained in $\Delta_1$, so we may assume that this is not the case. Thus $v(\Delta) > v(\Delta_1)$. We have: $$ g(\Delta) \leq g(\Delta_1) + g(\Delta_2) +[\frac{1}{2}k+1] \leq 2 \Phi_k(v(\Gamma)-v(\Delta_1)) +[\frac{1}{2}k+1] $$ $$ \leq \Phi_k(v(\Gamma)-v(\Delta)). \eqno{\Box}$$ \end{proof} Here and in the next proposition the term $[\frac{1}{2}k+1]$ comes from the fact that we take a {\it full} subgraph; compare the formula for the genus of a full subgraph given in Section \ref{graphs and partition}. \begin{definition} \label{weak} A full subgraph $\Delta$ of a stable graph $\Gamma$ is called $\Phi_k$-weak if $g(\Delta) \leq \Phi_k(v(\Gamma)-v(\Delta))$. \end{definition} Notice that this definition only depends on $k$ and not on $g$. By the proposition above the maximal $\Phi_k$-weak subgraph is well defined. \begin{definition} \label{maximal weak} Denote the maximal $\Phi_k$-weak subgraph of $\Gamma$ by $\Gamma_W$. A vertex is called strong if it is contained in the complement of $\Gamma_W$. Denote by $\Gamma_S$ maximal full subgraph on the strong vertices. \end{definition} Not every full subgraph of the maximal $\Phi_k$-weak subgraph is $\Phi_k$-weak, see the example following Proposition \ref{abcde hold}. Therefore we don't use the term $\Phi_k$-weak vertex. \begin{proposition}\label{full proper invariant} If $g > \Phi_k(0)$, then for every $\Gamma$ in ${\cal G}_k(g,n)$ we have that $\Gamma_W$ is a proper full subgraph which is invariant under the automorphisms of the stable graph $\Gamma$. \end{proposition} \begin{proof} By definition, $\Gamma_W$ is full. It is a proper subgraph because $g(\Gamma)>\Phi_k(v(\Gamma)-v(\Gamma))$ and thus $\Gamma$ is not $\Phi_k$-weak. Invariance under automorphisms follows because both the genus and the number of vertices are kept fixed by automorphisms of the stable graph. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} Recall that if $e$ is a whole edge of $\Gamma$, we write $\pi_e$ for the contraction of $e$ and $\pi_e(e)$ for the image of its end vertices. The inverse image of a full subgraph under $\pi_e$ is full. For $e$ a whole edge of $\Gamma$, we put $\delta(e):=0$ if $e$ is a loop, $\delta(e):=1$ otherwise. \begin{proposition} \label{sub} If $e$ is a whole edge contained in $\Gamma_W$ or in its complement, then we have $\pi_e(\Gamma_W) \subset (\Gamma /e)_W$. \end{proposition} \begin{proof} We have: $$g(\pi_e(\Gamma_W))=g (\Gamma_W) \leq \Phi_k(v(\Gamma)-v(\Gamma_W)) =$$ \[ \left\{ \begin{array}{ll} \Phi_k((v(\Gamma/e) +\delta(e)-v(\pi_e(\Gamma_W))) & \mbox{ $e$ outside $\Gamma_W$}\\ \Phi_k(v(\Gamma/e) +\delta(e)-[v(\pi_e(\Gamma_W))+\delta(e)]) & \mbox{ $e$ contained in $\Gamma_W$} \end{array} \right. \] $$ \leq \Phi_k(v(\Gamma/e)-v(\pi_e(\Gamma_W))). \eqno{\Box}$$ \end{proof} \begin{proposition} \label{gammaw maps to gamma/ew} If $e$ is a whole edge contained in $\Gamma_W$, then the image of $\Gamma_W$ under contraction of $e$ is $ (\Gamma /e)_W$. \end{proposition} \begin{proof} Proposition \ref{sub} yields us one inclusion. Put $\Delta :=\pi_e^{-1}((\Gamma/e)_W)$, then we have: $$g(\Delta) =g ((\Gamma/e)_W) \leq \Phi_k(v(\Gamma/e)-v((\Gamma/e)_W)) $$ $$=\Phi_k(v(\Gamma)-\delta(e)-[v(\Delta)-\delta(e)]) =\Phi_k(v(\Gamma)-v(\Delta)). \eqno{\Box}$$\end{proof} \begin{proposition}\label{not weak kept fixed under degeneration} A whole edge $e$ is contained in $\Gamma_W$ if and only if $\pi_e(e)$ is contained in $(\Gamma/e)_W$. \end{proposition} \begin{proof} The implication "$\Rightarrow$" follows from Proposition \ref{gammaw maps to gamma/ew}. For the other implication, suppose $(\Gamma/e)_W$ contains $\pi_e(e)$. Put $\Delta :=\pi_e^{-1}((\Gamma/e)_W)$, then the same computation as in the proof of Proposition \ref{gammaw maps to gamma/ew} implies that $\Delta$ is $\Phi_k$-weak and therefore contained in $\Gamma_W$. Consequently, $\Gamma_W$ contains $e$. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{definition} \label{phi} Define a function $\phi_k(g,n): {\cal G}_k(g,n) \rightarrow {\cal G}_k(g,n)$ by $\phi_k(\Gamma) =\Gamma /\Gamma_W$. When there is no chance of confusion, we will write $\phi$ instead of $\phi_k$. \end{definition} We call $\phi$ contraction of the maximal $\Phi_k$-weak subgraph. It is clear that we have: $\Gamma$ is contained in $Im(\phi)$ if and only if $ \Gamma_W$ does not contain any whole edges. It follows that the image of $\phi$ equals the fix point set of $\phi$. Put $I_k(g,n):=Im(\phi_k(g,n))$; we will write $I$ instead of $I_k(g,n)$ if there is no chance of confusion. The set $I$ will be the index set for our orbifold partition. Before we can define a partial order on $I$, we need two more propositions. \begin{proposition} \label{low inside} If $P$ is a vertex of $\Gamma$ and $w(P) \leq {\mbox max} \{ w(Q) \}$, where $Q$ runs over the vertices of $\Gamma_W$, then $\Gamma_W$ contains $P$. \end{proposition} \begin{proof} Suppose not, let $\Delta$ be the full subgraph on $P$ and $\Gamma_W$. We have : $$ g(\Delta) \leq g(\Gamma_W)+g(P) +[\frac{1}{2}k+1] \leq 2g(\Gamma_W) + [\frac{1}{2}k+1] $$ $$ \leq \Phi_k(v(\Gamma)-(v(\Gamma_W) +1)) =\Phi_k(v(\Gamma)-v(\Delta)). \eqno{\Box}$$ \end{proof} \begin{proposition}\label{joins} Let $e$ be an edge which joins $\Gamma_S$ and $\Gamma_W$. Let $\Lambda $ be the full subgraph of $\Gamma/e$ on the vertices of $\pi_e(\Gamma_W) -\pi_e(e)$. Then $\Lambda$ is $\Phi_k$-weak. \end{proposition} \begin{proof} We have: $$ g(\Lambda) \leq g(\Gamma_W) \leq \Phi_k(v(\Gamma)-v(\Gamma_W)) =\Phi_k(v(\Gamma)-[v(\Lambda)+1])$$ $$=\Phi_k(v(\Gamma/e)-v(\Lambda)). \eqno{\Box}$$ \end{proof} \begin{definition} Let $\Gamma$ be in ${\cal G}$. Let $s(\Gamma)$ be the number of strong vertices. Define $\{ w_i(\Gamma) \}_{i=1}^{s(\Gamma)}$ to be the set of weights of strong vertices, ordered such that $w_i(\Gamma) \geq w_{i+1}(\Gamma)$. For $i \in \{ s(\Gamma)+1,...,[k/2+1]+1 \}$, define $w_i(\Gamma):=g$. Let $n(\Gamma)$ be the number of half edges parting from strong vertices. Finally, define {\rm index}$(\Gamma)$ of a graph $\Gamma$ to be the vector $(w_1(\Gamma),....,w_{[k/2+1]+1}(\Gamma),n(\Gamma)).$ \end{definition} Notation: ${\rm index}(\Gamma)> {\rm index}(\Delta)$ refers to lexicographical ordering. \begin{proposition} \label{indeks} a) If $e$ is a whole edge of $\Gamma$ not contained in $\Gamma_W$, then ${\rm index}(\Gamma/e)>{\rm index}(\Gamma)$. \hfill \break b) If $e$ is a whole edge contained in $\Gamma_W$, then ${\rm index}(\Gamma/e) ={\rm index}(\Gamma)$. \end{proposition} \begin{proof} a) Suppose first that neither of the end vertices of $e$ is in $\Gamma_W$ and suppose that ${\rm index}(\Gamma) $ $= (\dots,w_i , \dots , w_j,\dots)$. If $e$ is a loop and its end vertex has weight $w_i$, then by Propositions \ref{low inside} and \ref{not weak kept fixed under degeneration} we have ${\rm index}(\Gamma/e)=(\dots,w_i+1,\dots) > {\rm index}(\Gamma)$. If $e$ is not a loop and its end vertices have weights $w_i$ and $w_j$, then ${\rm index}(\Gamma/e)=(\dots,w_i+w_j,\dots) > {\rm index}(\Gamma)$. Secondly suppose $e$ has end vertices $P \in \Gamma_S$ and $Q \in \Gamma_W$. By Proposition \ref{joins} we have $s(\Gamma) \geq s(\Gamma/e) $. If $s(\Gamma) > s(\Gamma/e)$, then we are done. If $s(\Gamma) = s(\Gamma/e)$ and $w(Q) >0$ the argument above applies. If $s(\Gamma) = s(\Gamma/e)$ and $w(Q)=0$, then the first $[k/2+1]+1$ coefficients of both indices are equal. We define the following numbers: \[ \begin{array}{l} a:=\# \{ \mbox{edges joining $P$ and $Q$} \},\\ b:=\# \{ \mbox{loops at $Q$} \},\\ c:=\# \{ \mbox{edges joining $Q$ and another vertex in $\Gamma_W$} \},\\ d:=\# \{ \mbox{edges joining $Q$ and a vertex in $\Gamma_S$ not equal to $P$}\}, \\ f:=\# \{ \mbox{loose half edges at $Q$}\}. \end{array} \] One sees easily that $n(\Gamma/e)=n(\Gamma) +a+2b+c+d+f-2$. Stability of the vertex $Q$ implies $a+2b+c+d+f\geq 3$, so we have ${\rm index}(\Gamma/e)> {\rm index}(\Gamma)$. \hfill \break b) Follows immediately from Proposition \ref{gammaw maps to gamma/ew}. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{corollary} The index is preserved under contraction of the maximal $\Phi_k$-weak subgraph. \end{corollary} \begin{definition}\label{partial order} We define a partial order on $I$ as follows: $\Gamma \geq \Delta$ if and only if ${\rm index}(\Gamma)$ $ > {\rm index}(\Delta)$ or $\Gamma =\Delta$. \end{definition} \begin{corollary} \label{order OK} If $e$ is a whole edge of $\Gamma$ not contained in $\Gamma_W$, then $\Gamma/e > \Gamma$. \end{corollary} Now we are almost ready to prove the remaining property e of Proposition \ref{formal}. We do this by making a suitable choice for $\Phi_k$. For a given $k$, define $ \alpha := 1/ [k/2+1]+2$ and $\beta:=[k/2+1]^2 \alpha^{[k/2+1]+1}$. We fix $\alpha$ and $\beta$ for the rest of this paper. \begin{definition} \label{specifiek} Define $\Phi_k :{\bf Z}_{\geq 0} \ra {\bf R}_{\geq 0} $ by $\Phi_k(n):= \alpha^{-n}\beta$. \end{definition} An easy calculation yiels that $$2 \Phi_k(n) +[\frac{1}{2}k+1] \leq \Phi_k(n-1), \mbox{ if } n > 0.$$ \begin{proposition}\label{gammaW bevat laag geslacht} If $k=1$ (respectively $k \geq 2$) and $P$ a vertex of $\Gamma$ such that the weight of $P$ is 0 (respectively the weight of $P$ is at most $k+2$), then $\Gamma_W$ contains the vertex $P$. \end{proposition} \begin{proof} Let $l$ be the number of loops at the vertex $P$. One has to check that $0+l \leq \Phi_k([k/2+1]-l) $ (respectively $k+2+l \leq \phi_k([k/2+1]-l)$), which is an easy computation. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{proposition} \label{abcde hold} Notations as above. Let $k$ be given, and let $g$ be at least $\beta$. Define $\Phi_k$ as in Definition \ref{specifiek}. Define $\phi$ and $I$ as in Definition \ref{phi}. Let a partial order on $I$ be defined by Definition \ref{partial order}. Then the properties a, b, c, d and e of Proposition \ref{formal} hold. \end{proposition} \begin{proof} The properties a, b, c, d and e are precisely the Propositions \ref{full proper invariant}, \ref{order OK}, \ref{gammaw maps to gamma/ew}, \ref{not weak kept fixed under degeneration} and \ref{gammaW bevat laag geslacht}. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} Not every full subgraph of the maximal $\Phi_k$-weak subgraph is $\Phi_k$-weak. Consider the following example: $k=9$, so $[k/2+1]=5$ and $\alpha=11/5$. We have $\beta=11^6/5^4 \approx 2834$ and thus we have to take $g > 2834$. Let $\Gamma$ be the graph with vertices $p_1, \dots , p_6$ and and edges $e_1, \dots , e_5$ such that $e_i$ joins $p_i$ and $p_{i+1}$. Suppose $w(p_1)=w(p_2)=11,~w(p_3)=22,~w(p_4)=44,~w(p_5)=88$ and $w(p_6)=g-176$. One can check that the maximal $\Phi_k$-weak subgraph is the full subgraph on the vertices $p_1, \dots,p_5$, but the full subgraph on the vertex $p_5$ is not $\Phi_k$-weak. \begin{remark} \label{infinite genus} {\rm If one is willing to consider graphs of infinite genus, i.e.\ several vertices can have infinite genus, then a canonical definition of weak subgraph is readily available: just take the full subgraph on the vertices of finite genus. (The definition of index needs to be adapted.) Properties analogous to those of the previous propositions, and in some cases even stronger results, can be proved.} \end{remark} \section{The results}\label{results} We keep the notations and assumptions of Proposition \ref{abcde hold}, unless the contrary is explicitely stated. As we have seen in Section \ref{graphs and partition}, a connected component of the $S_\Gamma^0$ is a stratum of the stratification by topological type; it is an orbifold of codimension equal to the number of singular points of its topological model which we will denote by $C_\Gamma$. In fact, a stratum is isomorphic to a product of lower dimensional moduli spaces $M_h^m$ modulo a finite group. Recall that the symmetric group on $n$ elements acts transitively on the connected components of the $S_\Gamma^0$. Furthermore, the orbifold normal bundle of $S_\Gamma^0$ has fibre over $(C,x_1, \dots, x_n)$ isomorphic to $\oplus(T_xC' \otimes T_xC'')$, where the sums runs over the singular points of $C$ and $C',~C''$ are the local components of $C$ in a suitable neighbourhood of $x$. This isomorphism globalizes to an isomorphism of bundles on $S_\Gamma^0$. It follows from Proposition \ref{formal} that the orbifold normal bundle of $S_\Gamma$ has fibre over $(C,x_1, \dots, x_n)$ isomorphic to $\oplus(T_xC' \otimes T_xC'')$. Here the sums runs over the singular points of $C$ which are specializations of singular points on the topological model of $S_\Gamma^0$ and $C',~C''$ are the local components of $C$ in a suitable neighbourhood of $x$. Thus, the sums runs over the edges of $\Gamma$. Notice that $S=\cup_{\Gamma \in I} S_\Gamma$ and that the real codimension of the complement of $S$ in $\Mgnbar$ is larger than $k$. Before going on we explain the Stability Theorem of Harer and Ivanov in an algebro-geometric way. These theorems are essential in the proof of Theorem \ref{filters coarse}. Let $S$ be the locus in $\overline{M_{g+1}}$ of irreducible stable curves of genus $g$ with one singularity; $S$ is a stratum of the stratification by topological type, it is isomorphic to $M_g^2$ modulo the involution permuting the two points. There is a map $p:S \rightarrow M_g $ which forgets the two points. Let $i:U_S \hookrightarrow \overline{M_{g+1}}$ be the inclusion of a suitable $C^\infty$ tubular neighbourhood of $S$ and let $\pi :U_S \rightarrow S$ be the natural retraction. Consider the diagram $$M_{g+1} \stackrel{i}{\longleftarrow} U_S \setminus S \stackrel{p \circ \pi}{\longrightarrow} M_g.$$ The Stability Theorem of Harer and Ivanov now says that if $g \geq 2k+1$, then both maps induce isomorphisms on cohomology in degree up to $k$ (see \cite{Harer1},\cite{Ivanov}). Furthermore $H^0(M_g) \cong {\bf Q}$ for $g \geq 0$ and $H^1(M_g) \cong 0$ for $g \geq 1$, see \cite[Ch. 7]{Harer2}. These facts account for the conditions of property e of Proposition \ref{formal}. Hence we can define the kth stable cohomology group $H^k(M_\infty)$ of the moduli space by $H^k(M_\infty):=H^k(M_g)$ when $g \geq 2k+1$. If $g \geq 2k+1$ (respectively $g \geq 0$ if $k=0$, respectively $g>0$ if $k=1$), we say $g$ is in the stable range with respect to $k$. Moreover $i^*$ and $(p \circ \pi)^*$ are morphisms of mixed Hodge structures, so $H^k(M_\infty)$ carries a well-defined mixed Hodge structure. \begin{remark} {\rm In \cite{Mumford} classes $\kappa_i \in H^{i,i}(M_\infty)$ are constructed and in \cite{Miller} it is proved that the symmetric algebra on these classes injects into $H^\bullet(M_\infty)$. Mumford conjectures that this is actually an isomorphism in low degree, see \cite[Introduction]{Mumford}. The conjecture would imply our corollary \ref{pure Hodge}. } \end{remark} We need a corollary of Harer's results. Assigning to a pointed curve the tangent space at its ith point defines a line bundle ${\cal L}_i$ on $\Mgnbar$. Consider the natural forgetful map $M_g^n \rightarrow M_g$. Define $H^\bullet (M_g)[u_1, \dots, u_n] \rightarrow H^\bullet(M_g^n)$, where the $u_i$ have degree 2, by sending $u_i$ to the first Chern class $c_1({\cal L}_i)$. Then we have (see \cite{Looijenga}): this is an isomorphism up to degree $k$ if $g \geq 2k+1$. \begin{theorem} \label{filters coarse} Let $k$ be given. If $g > \beta ,$ then the partition $S=\cup_{\Gamma \in I} S_\Gamma$ is coarser than the stratification by topological type, has $S_{{\it max }(I)} =M_g^n$ and filters cohomology up to degree $k$. \end{theorem} \begin{proof} We have already seen in Proposition \ref{formal}, properties 1 and 4, that the $S_\Gamma$ form a partition which has $M_g^n$ as open part and is coarser than the stratification by topological type. So it remains to show that for all $\Gamma \in I$ and all $l<k$, the Gysin maps on cohomology $H^{l-2codim S_\Gamma}(S_\Gamma)(-codim S_\Gamma) \ra H^l(\cup_{\Delta \geq \Gamma} S_\Delta)$ induced by the inclusions $S_\Gamma \rightarrow \cup_{\Delta \geq \Gamma} S_\Gamma$ are injections. When we write property x we mean property x of Proposition \ref{formal}. The orbifold $S_\Gamma^0$ is the quotient of $$ \prod M_{g_s}^{n_s} \times \prod M_{g_w}^{n_w}$$ by a finite group. Here the first product runs over the strong vertices of $\Gamma$ and the second over those which are not strong. Because of properties 2 and 5 we have that $S_\Gamma$ is contained in the quotient of $$\prod M_{g_s}^{n_s} \times \prod \overline{M_{g_w}^{n_w}}$$ by a finite group, where again the products run over the vertices which are strong respectively not strong. Property 7 implies that the $g_a$ are in the stable range w.r.t. $[k/2+1]$. By the result of Looijenga mentioned above we get : $$H^i(M_{g_s}^{n_s}) \cong \mbox{degree $i$ part of }H^i(M_\infty)[u_1, \dots, u_{n_s}], \eqno{(1)}$$ where the $u_i$ have degree 2. We have seen that the normal bundle of $S_\Gamma$ splits as a direct sum of line bundles, and thus its top Chern class becomes the product of the first Chern classes of these line bundles. We claim that these first Chern classes are all of the form $u_i+u_j$ or $u_i+a$, where the $u_i$ are as in $(1)$ and $a$ is an element of $H^*(\overline{M_{g_w}^{n_w}})$. We postpone the proof of the claim for a moment. By properties 2 and 3, $S_\Gamma$ is a closed suborbifold of $\cup_{\Delta \geq \Gamma}S_\Delta$. Consider the Gysin sequence for the inclusion $S_\Gamma$ in $\cup_{\Delta \geq \Gamma}S_\Delta$: $$ \dots \ra H^{l-2 {\it codim} S_\Gamma}(S_\Gamma)(-codim S_\Gamma) \ra H^l(\cup_{\Delta \geq \Gamma} S_\Delta) \ra H^l(\cup_{\Delta > \Gamma} S_\Gamma) \ra \dots $$ Composing $H^{l-2codimS_\Gamma}(S_\Gamma)(-codim S_\Gamma) \ra H^l(\cup_{\Delta \geq \Gamma} S_\Delta)$ with the restriction morphism to $ H^l(S_\Gamma)$ we get a morphism $H^{l-2codimS_\Gamma}(S_\Gamma)(-codim S_\Gamma) \ra H^l(S_\Gamma)$ which is given by taking the cup product with the top Chern class of the normal bundle of $S_\Gamma$. From what we have said above it follows that cupping with this Chern class is injective up to degree $k$. A fortiori the Gysin maps are injective. It remains to prove the claim. As explained, the line bundles of which we are taking the first Chern classes correspond to whole edges of the graph $\Gamma$. Let $f$ be a whole edge and let $P$ and $Q$ be its end vertices (which possibly coincide). The line bundle under consideration is the tensor product of the two line bundles corresponding to $P$ and half of $f$ respectively to $Q$ and the other half of $f$. By property 6 we have that either $P$ and $Q$ are both strong vertices or precisely one of them is a strong vertex. In the first case the first Chern classes of both line bundles are of the form $u_i$ and so we get $u_i+u_j$ as first Chern class of the tensor product. In the second case we get $u_i+a$ for some element $a$ of $H^*(\overline{M_{g_w}^{n_w}})$. This proves the claim. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{remark} {\rm The parts $S_\Gamma$ depend upon the definition of $\Phi_k$-weak subgraphs, which depends upon $\alpha$, which in turn depends upon $k$. This implies that the filtration in Theorem \ref{filters coarse} depends upon $k$. In this remark we will write $\alpha_k$ and $\beta_k$ to stress dependence. There is a natural inclusion $i: {\cal G}_{k}(g,n) \hookrightarrow {\cal G}_{k+1}(g,n)$ which is a bijection if $k$ is even. We claim that $i(\Gamma_W) \subset (i(\Gamma))_W$. This is clear if $k$ is even because then $\alpha_k^{v(\Delta)-v(\Gamma)}\beta_k =\alpha_{k+1}^{v(i(\Delta))-v(i(\Gamma))}\beta_{k+1}$ and we even have equality. If $k$ is odd one has to check that $\alpha_k^{v(\Delta)-v(\Gamma)}\beta_k \leq \alpha_{k+1}^{v(i(\Delta))-v(i(\Gamma))}\beta_{k+1}.$ If we put $l:=[k/2+1]$ and $n=v(\Delta)-v(\Gamma)$, then this amounts to checking the inequality $(\frac{2l+3}{l+1})^{l+1-n}(l+1)^2 \geq (\frac{2l+1}{l})^{l-n}l^2$, which is tedious but elementary. It follows that the parts $S'_\Delta$ for $k+1$ contain unions of parts $S_\Gamma$ for $k$: $S'_\Delta \supset \cup_{\Gamma \in J} S_\Gamma$. We can now apply Theorem \ref{filters coarse} to this union to get a filtration on $H^l(S_\Delta')$ some of whose subquotients are isomorphic to $H^{l+2 codimS_\Delta'-2 codimS_\Gamma}(S_\Gamma)(-codim S_\Delta' + codim S_\Gamma)$ for $l<k$. We conclude that enlarging $k$ amounts to taking the union of some parts. } \end{remark} \begin{remark} {\rm Theorem \ref{filters coarse} does not hold if we replace the partition by the stratification by topological type. To see this, let $k$ be given and take $g>5\alpha^{[k/2+1]+1}$. Take $n=0$ for simplicity. Choose natural numbers $d>c>b>a>\alpha^{[k/2+1]+1}$ such that $a+b+c+d=g$. Consider the graph $\Gamma$ which has five vertices, of weights $0,~a~,b,~c,~d$ and four edges, joining the weight $0$ component to the other four. The automorphism group of the stable graph $\Gamma$ clearly is trivial. One checks easily that $\Gamma_W$ is the full subgraph on the weight $0$ vertex, which implies that $\Gamma$ defines the open stratum $S_\Gamma^0$ of a part $S_\Gamma$. The part $S_\Gamma$ is obtained by letting the genus $0$ curve degenerate in all possible ways. There are three possible degenerations, corresponding to the three possibilities of partioning the four other vertices into two sets of two. So we have: $S_\Gamma^0 \cong N \times M_0^4$ and $S_\Gamma \cong N \times \overline{M_0^4}$, where $N:=M_a^1 \times M_b^1 \times M_c^1 \times M_d^1$. $M_0^4$ is a ${\bf P}^1$ minus 3 points and $\overline{M_0^4}$ is ${\bf P}^1$. Thus $S_\Gamma=S_\Gamma^0 \cup \cup_1^3 N \times {\it point}$. We have $H^2(N \times \overline{M_0^4})=H^2(N) \otimes H^0(\overline{M_0^4}) \oplus H^0(N) \otimes H^2(\overline{M_0^4})=H^2(N) \oplus {\bf Q}$. Suppose Theorem \ref{filters coarse} would hold with the partition replaced by the stratification by topological type. Then we would have, using the above: $H^2(N \times \overline{M_0^4})= H^2(N \times {M_0^4}) \oplus \oplus_{i=1}^3 H^0(N \times {\it point})= H^2(N) \oplus {\bf Q}^3$, which leads to a contradiction. } \end{remark} \begin{corollary} \label{coho Mgnbar} Given $k \geq 0$, then for $g>\beta$, $\Mgnbar$ admits an orbifold partition which has $M_g^n$ as its open part, is coarser than the stratification by topological type and filters cohomology up to degree $k$. \end{corollary} \begin{proof} Using Theorem \ref{filters coarse} we get the statement with $\Mgnbar$ replaced by $S$. The real codimension of the complement of $S$ in $\Mgnbar$ is $2[k/2+1]+2 >k$, thus $H^l(\Mgnbar) \cong H^l(S)$, for all $l<k$. (Compare the remark following the definition of a partition which filters cohomology in the introduction.) Combining these gives the desired result. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{corollary} \label{pure Hodge} If $g \geq 2k+1$, then $H^k(\Mgnbar) \ra H^k(M_g^n)$ is onto; consequently, the mixed Hodge structure on $H^k(M_g^n)$ is pure of weight $k$. \end{corollary} \begin{proof} By corollary \ref{coho Mgnbar}, we have that $H^k(M_g^n)$ is a quotient of $H^k(\Mgnbar)$, so the last statement holds for $g > \beta$. Now use that the image of $H^k(\Mgnbar)$ in $H^k(M_g^n)$ is $W^kH^k(M_g^n)$, which equals $H^k(M_g^n)$ if $g \geq 2k+1$. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{corollary}\label{not Tate} If $g >>0$, then the cohomology of $\Mgnbar$ is not of Tate type. \end{corollary} \begin{proof} We use the fact that the modular form $\Delta$ can be seen as a non vanishing holomorphic 11-form on $\overline{M_1^{11}}$ (see \cite{Deligne}). Thus, $H^{11,0}(\overline{M_1^{11}})$ is not zero. Take $g$ large enough and $n$ arbitrary. One of the strata in corollary \ref{coho Mgnbar} is the following: the generic graph consists of two vertices, one of weight $g-11$ and one of weight 1. There are 11 edges joining them and $n$ loose half edges at the vertex of weight $g-11$. Clearly, the maximal $\Phi_k$-weak subgraph of this graph consists of the vertex of weight 1 with its 11 half edges. This graph is therefore in the image of the map $\phi$ and defines a generic point of a part of codimension 11. The open dense topological stratum in it is $M_{g-11}^{n+11} \times {M_1^{11}}$. The corresponding part is obtained by letting the genus 1 curve degenerate, thus it is $M_{g-11}^{n+11} \times \overline{M_1^{11}}$. Consider the cohomology group $H^{22,11}(\Mgnbar) \subset H^{33}(\Mgnbar)$. By corollary \ref{coho Mgnbar} it has a subquotient $H^{11,0}(M_{g-11}^{n+11} \times \overline{M_1^{11}})$, which has as direct summand $H^0(M_{g-11}^{n+11}) \bigotimes H^{11,0}(\overline{M_1^{11}}) \neq 0 $. By considering different strata we find that other cohomology groups don't vanish either. For example: suppose $n \geq 10$ and take a graph consisting of two vertices, one of weight 1 and one of weight $g-1$, joined by one edge. Furthermore the weight one vertex has 10 loose half edges and the other vertex $n-10$. This defines the generic graph of a part of codimension 1. By the same argument as above, we see that $H^{12,1}(\Mgnbar)$ is not zero. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} \begin{proposition}\label{stabiele cohomologie voor Moneindignstreep} There exists a ``stable cohomology" of $\Mgnbar$ for $n$ fixed and $g \ra \infty$. \end{proposition} \begin{proof} Notation as before. Let $k$ be given and suppose $g > \beta$. Let $\Gamma$ be in the image of $\phi_k(g,n)$, that is, $\Gamma$ defines a generic stratum of a part $S$. Choose a strong vertex $P$ of $\Gamma$, let $h$ be the weight of $P$. Consider the graph we get by changing $h$ to $h+1$. Call this graph $\Delta$. Because the definition of $\Phi_k$-weak subgraph does not depend on $g$, we see that $\Gamma_W$ and $\Delta_W$ can be identified. This means that $\Delta$ is in the image of $\phi_k(g+1,n)$ and therefore defines a part $T$ in $\overline{M_{g+1}^n}$. Furthermore we get that $H^l(S) \cong H^l(T)$ for all $l <k$, because $h$ is in the stable range with respect to $k$. For every $x \in {\bf N}$ we define $\psi_x: {\cal G}_k(g+x,n) \ra {\cal G}_k(g+x+([k/2+1]+1)!,n)$ as follows: for any vertex $P$ in $\Gamma_S$ add $([k/2+1]+1)!/ \# \Gamma_S$ to its weight. $\psi$ clearly respects the parts and thus defines a map $\psi_x :I_k(g+x,n) \ra I_k(g+x+([k/2+1]+1)!,n)$. Because the maximal $\Phi_k$-weak subgraphs don't change, this map is an injection. Furthermore it induces isomorphisms on the cohomology of the parts, as explained above. So we get an injection induced by $ \psi_x$: $$H^l(M_{g+x}^n) \hookrightarrow H^l(M_{g+x+([\frac{1}{2}k+1]+1)!}^n),$$ for all $l<k$, which maps the subquotients isomorphically onto the corresponding subquotients. We claim that the inductive limit over these maps is independent of $x$. We will compare the inductive limits for 0 and $x>0$. For any $\Gamma$ in $Im(\phi_k(g,n))$ we choose one strong vertex $P_\Gamma$. Define $\chi_x(\Gamma) \in {\cal G}_k(g+x,n)$ by adding $x$ to the weight of $P$. $\chi_{x}(\Gamma)$ is in the image of $\phi_k(g+x,n)$. Define $\overline{\chi_x}: Im(\phi_k(g+([k/2+1]+1)!,n)) \ra Im(\phi_k(g+x+([k/2+1]+1)!,n))$ by adding $x$ to the weight of the strong vertex $\psi_x (P_\Gamma)$. We clearly have $\overline{\chi_x} \circ \psi_0=\psi_x \circ \chi_x$. Now we replace 0 by $x$ and $x$ by $([k/2+1]+1)!$. Playing the same trick and using that the inductive limits are canonically isomorphic if the indices differ by a multiple of $([k/2+1]+1)!$, we prove our claim. We define $H^l(\overline{M^n_\infty})$ as this inductive limit. \unskip\nolinebreak\hfill\hbox{\quad $\Box$} \end{proof} One may regard this as the ``stable cohomology" of $\Mgnbar$ for $g \ra \infty$. We note however that this stable cohomology is not finitely generated. For example, if $n=0$, the stable generators in degree $2$ are the tautological class, and the boundary classes naturally indexed by $0,1,2,\dots$. We note also that the part defined in Corollary \ref{not Tate} defines a part for every $g$ sufficiently large; consequently the proof of Theorem \ref{stabiele cohomologie voor Moneindignstreep} shows that even the stable cohomology of $\Mgnbar$ for $g \ra \infty$ is not of Tate type.
1996-12-06T08:03:17
9503
alg-geom/9503001
en
https://arxiv.org/abs/alg-geom/9503001
[ "alg-geom", "math.AG" ]
alg-geom/9503001
Nitin Nitsure
Nitin Nitsure
Quasi-parabolic Siegel Formula
LaTeX, 6 pages. Reason for re-submission : A factor that was missing in the first version is now included in the formula
null
null
null
null
The result of Siegel that the Tamagawa number of $SL_r$ over a function field is 1 has an expression purely in terms of vector bundles on a curve, which is known as the Siegel formula. We prove an analogous formula for vector bundles with quasi-parabolic structures. This formula can be used to calculate the Betti numbers of the moduli of parabolic vector bundles using the Weil conjucture.
[ { "version": "v1", "created": "Thu, 2 Mar 1995 06:49:00 GMT" }, { "version": "v2", "created": "Fri, 6 Dec 1996 07:04:00 GMT" } ]
2008-02-03T00:00:00
[ [ "Nitsure", "Nitin", "" ] ]
alg-geom
\subsection*{\hbox{}\hfill{\normalsize\sl #1}\hfill\hbox{}}} \textheight 23truecm \textwidth 15truecm \addtolength{\oddsidemargin}{-1.05truecm} \addtolength{\topmargin}{-1.5truecm} \makeatletter \def\l@section{\@dottedtocline{1}{0em}{1.2em}} \makeatother \title{Quasi-parabolic Siegel Formula} \author{Nitin Nitsure} \begin{document} \date{Corrected version, 6 December 1996} \maketitle School of Mathematics, Tata Institute of Fundamental Research, Homi Bhabha Road, Mumbai 400 005, India. e-mail: [email protected] \begin{abstract} The result of Siegel that the Tamagawa number of $SL_r$ over a function field is $1$ has an expression purely in terms of vector bundles on a curve, which is known as the Siegel formula. We prove an analogous formula for vector bundles with quasi-parabolic structures. This formula can be used to calculate the betti numbers of the moduli of parabolic vector bundles using the Weil conjucture. \end{abstract} \section{Introduction} The Betti numbers of the moduli of stable vector bundles on a complex curve, in all the cases where the rank and degree are coprime, were first determined by Harder and Narasimhan [H-N] as an application of the Weil conjuctures. For this, they made use of the result of Siegel that the Tamagawa number of the special linear group over a function field is 1. In their refinement of the same Betti number calculation in [D-R], Desale and Ramanan expressed the result of Siegel in purely vector bundle terms. This result about the Tamagawa number, called the Siegel formula, was later given a simple proof in the language of vector bundles by Ghione and Letizia [G-L], by introducing a notion of effective divisors of higher rank on a curve, and counting the number of effective divisors which correspond to a given vector bundle. This purpose of this note is to introduce the notion of a quasi-parabolic divisor of higher rank on a curve (Definition 3.1 below), and to prove a quasi-parabolic analogue (Theorem 3.4 below) of the Siegel formula, which is done here by suitable generalizing the method of [G-L]. In a note to follow, this formula is used to calculate the Zeta function and thereby the Betti numbers of the moduli of parabolic bundles in the case `stable = semistable' (these Betti numbers have already been calculated by a guage theoretic method for genus $\ge 2$ in [N] and for genus $0$ and $1$ by Furuta and Steer in [F-S]). {\bf Acknowledgement} I thank M. S. Narasimhan for suggesting the problem of extending [H-N] to parabolic bundles. \section{Divisors supported on $X-S$} Let $X$ be an absolutely irreducible, smooth projective curve over the finite field $k={\bf F}_q$, and let $S$ be any closed subset of $X$ whose points are $k$-rational. Let $K$ denote the function field of $X$, and let $K_X$ denote the constant sheaf $K$ on $X$. Let $g$ denote the genus of $X$. Let $r$ be a positive integer. Recall that (see [G-L]) a coherent subsheaf $D\subset K _X^r$ of generic rank $r$ is called an $r$-divisor, and the $r$-divisor is called effective (or positive) if ${\cal O} _X^r\subset D$. The support of the divisor is by definition the support of the quotient $D/{\cal O} _X^r$, which is a torsion sheaf. The lenght $n$ of $D/{\cal O} _X^r$ is called the degree of the divisor. Note that $D$ is a locally free sheaf of rank $r$ and degree $n$. \refstepcounter{theorem}\paragraph{Remark \thetheorem} Let $Z_X(t)$ be the zeta function of $X$. Then as $S$ consists of $k$-rational points, it can be seen that the zeta function $Z_{X-S}$ of $X-S$ is given by the formula $$Z_{X-S}(t) = (1-t)^sZ_X(t)\eqno(1)$$ where $s$ is the cardinality of $S$. Note that an effective $r$-divisor on $X-S$ is the same as an effective $r$-divisor on $X$ whose support is disjoint from $S$. The part (1) of the proposition 1 of [G-L] gives the following, with $X-S$ in place of $X$. \begin{proposition} Let $b_n^{(r)}$ be the number of effective $r$-divisors of degree $n$ on $X$ whose support is disjoint from $S$. Let $Z_{X-S}^{(r)}(t) = \sum _{n\ge 0} b_n^{(r)}t^n$. Then we have $$Z_{X-S}^{(r)}(t) = \prod _{1\le j\le r} Z_{X-S}(q^{j-1}t)\eqno(2)$$ \end{proposition} In order to have the analogue of the part (2) of the proposition 1 of [G-L], we need the following lemmas. \begin{lemma} Let $V$ be a finite dimensional vector space over $k={\bf F}_q$, and $s$ a positive integer. For any $1\le i\le s$, let $\pi _i:k^s\to k$ be the linear projection. For any surjective linear map $\phi :V\to k^s$, let $V_i$ be the kernel of $\pi _i\phi :V\to k$, which is a hyperplane in $V$ as $\phi$ is surjective. Let $P=P(V)$, and $P_i=P(V_i)$ denote the corresponding projective spaces. Let $N(\phi )$ denote the number of $k$-rational points of $P - \cup _{1\le i\le s} P_i$. Then for any other surjective $\psi : V\to k^s$, we have $N(\phi )=N(\psi )$. In other words, given $s$, this number depends only on $dim(V)$. \end{lemma} \paragraph{Proof} Given any two surjective maps $\phi ,\psi :V\to k^s$, there exists an $\eta \in GL(V)$ such that $\phi \eta = \psi$. From this, the result follows. \begin{lemma} Let $n$ be a positive integer, such that $n>2g-2+s$ where $g$ is the genus of $X$ and $s$ is the cardinality of $s$. Let $b_n$ is the total number of effective $1$-divisors of degree $n$ supported on $X-S$. Then for any line bundle $L$ on $X$ of degree $n$, the number of effective $1$-divisors supported on $X-S$ which define $L$ is $b_n/P_X(1)$, where $P_X(1)$ is the number of isomorphism classes of line bundles of any fixed degree on $X$. \end{lemma} (Here, $P_X(t)$ is the polynomial $(1-t)(1-qt)Z_X(t)$.) \paragraph{Proof} Let $L$ be any line bundle on $X$ of degree $n$, where $n>2g-2+s$. Then $H^1(X,L(-S))=0$, so the natural map $\phi :H^0(X,L)\to H^0(X,L|S)$ is surjective. Let $V=H^0(X,L)$. Then $dim(V)=n+1-g$. Choose a basis for each fiber $L_P$ for $P\in S$. This gives an identification of $H^0(X,L|S)$ with $k^s$. Now it follows that the number $N(\phi )$ defined in the preceeding lemma depends only on $n$, and is independent of the choice of $L$ as long as it has degree $n$. But $N(\phi )$ is precisely the number of effective $1$-divisors supported on $X-S$, which define the line bundle $L$ on $X$. Using the above lemma, the following proposition follows, by an argument similar to the proof of part (2) of proposition 1 in [G-L]. The proof in [G-L] expresses the number of $r$-divisors in terms of the number of $1$-divisors, and the above lemma tells us the number of $1$-divisors with support in $X-S$ corresponding to a given line bundle on $X$. \begin{proposition} For $L$ a line bundle of degree $n$, let $b_n^{(r,L)}$ be the number of effective $r$-divisors on $X$ supported on $X-S$, having determinant isomorphic to $L$. Then provided that $n>2g-2+s$, we have $$b_n^{(r,L)} = b_n^{(r)}/P_X(1)\eqno(3)$$ \end{proposition} \begin{proposition} $$\lim _{n\to\infty} {b_n^{(r)}\over{q^{rn}}} = P_X(1){(q-1)^{s-1}\over{q^{g-1+s}}} Z_{X-S}(q^{-2})\cdots Z_{X-S}(q^{-r})\eqno(4)$$ \end{proposition} \paragraph{Proof} The above statement is the analogue of proposition 2 of [G-L], with the following changes. Instead of all $r$-divisors on $X$ in [G-L], we consider only those which are supported over $X-S$, and instead of $Z_X(t)$, we use $Z_{X-S}(t)$. As $Z_{X-S}(t) = (1-t)^sZ_X(t)$, the property of $Z_X(t)$ that it has a simple pole at $t=q^{-1}$ and is regular at $1/q^j$ for $j\ge 2$ is shared by $Z_{X-S}(t)$. Hence the proof in [G-L] works also in our case, proving the proposition. \refstepcounter{theorem}\paragraph{Remark \thetheorem} There is a minor misprint in the equation labeled (1) in [G-L] (page 149); the factor $q^{g-1}$ should be read as $q^{1-g}$. Let $L$ be any given line bundle on $X$. Choose any closed point $P\in X-S$, and let $l$ denote its degree. For any ${\cal O} _X$ module $E$, set $E(m)=E\otimes {\cal O}_X(mP)$. If a vector bundle $E$ of rank $r$ degree $n$ has determinant $L$, then $E(m)$ has determinant $L(rm)$, degree $n+rml$ and Euler characteristic $\chi(m)=n+rml+r(1-g)$. The equations (3) and (4) above imply the following. $$\lim _{m\to\infty} {b_{n+rml}^{(r,L(rm))}\over{q^{r\chi(m)}}} = (q-1)^{s-1} q^{(r^2-1)(g-1)-s} Z_{X-S}(q^{-2})\cdots Z_{X-S}(q^{-r})\eqno(5)$$ \section{Quasi-parabolic divisors} For basic facts about parabolic bundles, see [S] and [M-S]. We now introduce the notion of a quasi-parabolic effective divisor of rank $r$. Let $S\subset X$ be a finite subset consisting of $k$-rational points. For each $P_i\in S$, let there be given positive integers $p_i$ and $r_{i,1},\ldots ,r_{i,p_i}$ with $r_{i,1}+\ldots +r_{i,p_i} =r$. This will be called, as usual, the quasi-parabolic data. Recall that a quasi-parabolic structure on a vector bundle $E$ of rank $r$ on $X$ by definition consists of flags $E_{P_i}=F_{i,1}\supset\ldots\supset F_{i,p_i}\supset F_{i,p_i+1}=0$ of vector subspaces in the fibers over the points of $S$ such that $dim(F_{i,j}/F_{i,j+1})=r_{i,j}$ for each $j$ from $1$ to $p_i$. \begin{definition}\rm Let $X$, $S$, and the numerical data $(r_{i,j})$ be as above. A positive quasi-parabolic divisor $(F,D)$ on $X$ consists of (i) a quasi-parabolic structure $F$ on the trivial bundle ${\cal O} _X^r$, consisting of flags $F_i$ in $k^r$ at points $P_i\in S$ of the given numerical type $(r_{i,j})$, together with (ii) an effective $r$-divisor $D$ on $X$, supported on $X-S$. \end{definition} Note that if $(F,D)$ is a quasi-parabolic $r$-divisor, then the rank $r$ vector bundle $D$ has a parabolic structure given by $F$. We denote by $P^{(r)}_E$ the set of all effective parabolic $r$-divisors whose associated parabolic bundle is isomorphic to a given parabolic bundle $E$. For any vector bundle $E$ of rank $r$, let $Hom^S_{inj}({\cal O} _X^r,E)$ denote the set of all injective sheaf homomorphisms ${\cal O} _X^r \to E$ which are injective when restricted to $S$. For any quasi-parabolic bundle $E$, the group of all quasi-parabolic automorphisms of $E$ will be denoted by $ParAut(E)$. Then $ParAut(E)$ acts on $Hom_{inj}^S({\cal O} _X^r,E)$ by composition. This action is free, and $P^{(r)}_E$ has a canonical bijection with the quotient set $Hom_{inj}^S({\cal O} _X^r,E)/ParAut(E)$. Hence the cardinality of $P^{(r)}_E$ is given by $$|P^{(r)}_E|= {{|Hom_{inj}^S({\cal O} _X^r,E)|}\over |{ParAut(E)|}}\eqno(6)$$ For $1\le i\le s$, let ${\rm Flag}_i$ be the variety of flags in $k^r$ of the numerical type $(r_{i,1},\ldots ,r_{i,p_i})$. Let ${\rm Flag}_S = \prod _{1\le i\le s}{\rm Flag}_i$. Let $f(q,r_{i,j})$ denote the number of $k$-rational points of ${\rm Flag}_S$. If $a_n^{(r,L)}$ denotes the number of quasi-parabolic divisors of flag data $(r_{i,j})$ with degree $n$, rank $r$ and determinant $L$, then we have $$a_n^{(r,L)} = f(q,r_{i,j})b_n^{(r,L)}\eqno(7)$$ Now let $J(r,L)$ denote the set of all isomorphism classes of quasi-parabolic vector bundles of rank $r$, degree $n$, determinant $L$ having the given quasi-parabolic data $(r_{i,j})$ over $S$. Hence the equation (6) above implies the following. $$a^{(r,L)}_n = \sum _{E\in J(r,L)} {{|Hom_{inj}^S({\cal O} _X^r,E)|}\over |{ParAut(E)|}}\eqno(8)$$ For any integer $m$, the map from $J(r,L) \to J(r, L(rm)$ which sends $E$ to $E(m) =E\otimes O_X(mP)$ is a bijection which preserves $|ParAut|$. Hence for each $m$, we have $$a^{(r,L(rm))}_{n+rml} = \sum _{E\in J(r,L)} {{|Hom_{inj}^S({\cal O} _X^r,E(m))|}\over |{ParAut(E)|}}\eqno(9) $$ \begin{lemma} With the above notations, $$\lim_{m\to \infty}{{|Hom_{inj}^S({\cal O} _X^r,E(m))|}\over {q^{r\chi (E(m))}}} = {(q^r-1)^s(q^r-q)^s\cdots(q^r-q^{r-1})^s\over q^{r^2s}}\eqno(10)$$ If $S$ is non-empty, the limit is already attained for all large enough $m$ (where `large enough' depends on $E$). \end{lemma} \paragraph{Proof} If $S$ is empty, the above lemma reduces to lemma 3 in [G-L]. If $S$ is nonempty, then any morphism of locally free sheaves on $X$ which is injective when restricted to $S$ is injective. Let $m$ be large enough, so that $E(m)$ is generated by global sections, $H^1(X,E(m))=0$, and $h^0(X,E(m))=\chi(E(m)) \ge rs$. Then $H^0(X,E(m))$ has a basis consisting of sections $\sigma_{i,P_j}$, $\tau_{\ell}$ for $i=1,\ldots, r$, $j= 1,\ldots,s$, and $\ell=1,\ldots, \chi(E(m))-rs$, such that (1) the sections $\tau_{\ell}$ are zero on $S$, (2) the sections $\sigma_{i,P_j}$ are zero at all other points of $S$ except $P_j$ (and hence $\sigma_{i,P_j}$ restrict at $P_j$ to a basis of the fiber of $E(m)$ at $P_j$. Any element of $Hom_{{\cal O}_X}({\cal O}_X^r,E(m)) = Hom_{{\bf F}_q}({\bf F_q}^r, H^0(X,E(m)))$ is given in terms of this basis by a $r\times q^{\chi(E(m))}$ matrix $A$. The condition that this lies in $$Hom_{inj}^S({\cal O} _X^r,E(m)) \subset Hom({\cal O} _X^r, E(m))$$ is the condition that each of the $s$ disjoint $r\times r$-minors, corresponding to the part $\sigma_{1,P_j},\ldots, \sigma_{r,P_j}$ of the basis, has nonzero determinant. This contributes the factor $${|GL_r({\bf F}_q)|\over |M_r({\bf F}_q)|}= {(q^r-1)(q^r-q)\cdots(q^r-q^{r-1})\over q^{r^2}} $$ for each $P_j$, which proves the lemma. \begin{lemma} The following sum and limit can be interchanged to give $$\sum _{E\in J(r,L)} \lim _{m\to\infty} {{|Hom_{inj}^S({\cal O} _X^r,E(m))|}\over {q^{r\chi (E(m))}|ParAut(E)|}} = \lim _{m\to\infty} \sum _{E\in J(r,L)} {{|Hom_{inj}^S({\cal O} _X^r,E(m))|}\over {q^{r\chi (E(m))} |ParAut(E)|}} $$ \end{lemma} This lemma has a proof entirely analogous to the corresponding statement in [G-L], so we omit the details. By equation (10), the left hand side in the above lemma equals $${(q^r-1)^s(q^r-q)^s\cdots(q^r-q^{r-1})^s\over q^{r^2s}} \sum _{E\in J(r,L)} {1\over{|ParAut(E)|}} $$ On the other hand, by (9), the right hand side is $\lim _{m\to\infty} a^{(r,L(rm))}_{n+rml}/q^{r\chi (m)}$. By equations (5) and (7), this limit has the following value. $$f(q,r_{i,j})(q-1)^{s-1} q^{(r^2-1)(g-1)-s} Z_{X-S}(q^{-2})\cdots Z_{X-S}(q^{-r})$$ By putting $Z_{X-S}(t) = (1-t)^sZ_X(t)$ in the above, and cancelling common factors from both sides, we get the following. \begin{theorem {\rm (Quasi-parabolic Siegel formula)} $$\sum _{E\in J(r,L)} {1\over{|ParAut(E)|}}= f(q,r_{i,j}) {q^{(r^2-1)(g-1)}\over q-1 } Z_X(q^{-2})\cdots Z_X(q^{-r}) $$ \end{theorem} \refstepcounter{theorem}\paragraph{Remark \thetheorem} If $S$ is empty or more generally if the quasi-parpbolic structure at each point of $S$ is trivial (that is, each flag consists only of the zero subspace and the whole space), then on one hand $ParAut(E)=Aut(E)$, and on the other hand each flag variety is a point, and so $f(q,r_{i,j})=1$. Hence in this situation the above formula reduces to the original Siegel formula $$\sum _{E\in J(r,L)} {1\over{|Aut(E)|}}= {q^{(r^2-1)(g-1)}\over q-1 } Z_X(q^{-2})\cdots Z_X(q^{-r}) $$ \section*{References} [D-R] Desale, U. V. and Ramanan, S. : Poincar\'e Polynomials of the Variety of Stable Bundles, {\sl Math. Annln.} {\bf 216} (1975), 233-244. [F-S] Furuta, M. and Steer, B. : Siefert-fibered homology 3-spheres and Yang-Mills equations on Riemann surfaces with marked points, {\sl Adv. Math.} {\bf 96} (1992) 38-102. [G-L] Ghione, F. and Letizia, M. : Effective divisors of higher rank on a curve and the Siegel formula, {\sl Composito Math.} {\bf 83} (1992), 147-159. [H-N] Harder, G. and Narasimhan, M. S. : On the Cohomology Groups of Moduli Spaces of Vector Bundles over Curves, {\sl Math. Annln.} {\bf 212} (1975), 215-248. [M-S] Mehta, V. B. and Seshadri, C. S. : Moduli of vector bundles on curves with parabolic structures, {\sl Math. Annln.} {\bf 248} (1980) 205-239. [N] Nitsure, N. : Cohomology of the moduli of parabolic vector bundles, {\sl Proc. Indian Acad. Sci. (Math. Sci.)} {\bf 95} (1986) 61-77. [S] Seshadri, C. S. : Fibres vectoriels sur les courbes algebriques, {\sl Asterisque} {\bf 96} (1982). \end{document}
1993-09-30T22:08:20
9309
alg-geom/9309007
en
https://arxiv.org/abs/alg-geom/9309007
[ "alg-geom", "hep-th", "math.AG" ]
alg-geom/9309007
David R. Morrison
Paul S. Aspinwall, Brian R. Greene and David R. Morrison
The Monomial-Divisor Mirror Map
22 pages, LaTeX
Internat. Math. Res. Notices (1993), 319-337
null
IASSNS-HEP-93/43, CLNS 93/1237
null
For each family of Calabi-Yau hypersurfaces in toric varieties, Batyrev has proposed a possible mirror partner (which is also a family of Calabi-Yau hypersurfaces). We explain a natural construction of the isomorphism between certain Hodge groups of these hypersurfaces, as predicted by mirror symmetry, which we call the monomial-divisor mirror map. We indicate how this map can be interpreted as the differential of the expected mirror isomorphism between the moduli spaces of the two Calabi-Yau manifolds. We formulate a very precise conjecture about the form of that mirror isomorphism, which when combined with some earlier conjectures of the third author would completely specify it. We then conclude that the moduli spaces of the nonlinear sigma models whose targets are the different birational models of a Calabi-Yau space should be connected by analytic continuation, and that further analytic continuation should lead to moduli spaces of other kinds of conformal field theories. (This last conclusion was first drawn by Witten.)
[ { "version": "v1", "created": "Thu, 30 Sep 1993 21:08:10 GMT" } ]
2008-02-03T00:00:00
[ [ "Aspinwall", "Paul S.", "" ], [ "Greene", "Brian R.", "" ], [ "Morrison", "David R.", "" ] ]
alg-geom
\section*{References\@mkboth {REFERENCES}{REFERENCES}}\small\list {\arabic{enumi}.}{\settowidth\labelwidth{[#1]}\leftmargin\labelwidth \advance\leftmargin\labelsep \usecounter{enumi}} \def\hskip .11em plus .33em minus .07em{\hskip .11em plus .33em minus .07em} \sloppy\clubpenalty4000\widowpenalty4000 \sfcode`\.=1000\relax} \let\endthebibliography=\endlist \makeatother \setlength{\textwidth}{425pt} \setlength{\textheight}{550pt} \setlength{\topmargin}{0pt} \setlength{\oddsidemargin}{35pt} \setlength{\evensidemargin}{0pt} \newif\iffn\fnfalse \makeatletter \@ifundefined{reset@font}{\let\reset@font\empty}{} \long\def\@footnotetext#1{\insert\footins{\reset@font\footnotesize \interlinepenalty\interfootnotelinepenalty \splittopskip\footnotesep \splitmaxdepth \dp\strutbox \floatingpenalty \@MM \hsize\columnwidth \@parboxrestore \edef\@currentlabel{\csname p@footnote\endcsname\@thefnmark}\@makefntext {\rule{\z@}{\footnotesep}\ignorespaces \fntrue#1\fnfalse\strut}}} \makeatother \newfont{\bbbfont}{msbm10 scaled\magstep1} \newfont{\smallbbbfont}{msbm8} \newfont{\tinybbbfont}{msbm6} \newfont{\footbbbfont}{msbm10} \newfont{\smallfootbbbfont}{msbm7} \newfont{\tinyfootbbbfont}{msbm5} \newcommand{\Bbb}[1]{\iffn \mathchoice{\mbox{\footbbbfont #1}}{\mbox{\footbbbfont #1}} {\mbox{\smallfootbbbfont #1}}{\mbox{\tinyfootbbbfont #1}}\else \mathchoice{\mbox{\bbbfont #1}}{\mbox{\bbbfont #1}} {\mbox{\smallbbbfont #1}}{\mbox{\tinybbbfont #1}}\fi} \newcommand{\mathbin{\mbox{\bbbfont\char"6E}}}{\mathbin{\mbox{\bbbfont\char"6E}}} \newcommand{\mathbin{\mbox{\bbbfont\char"6F}}}{\mathbin{\mbox{\bbbfont\char"6F}}} \newcommand{\rtimes}{\mathbin{\mbox{\bbbfont\char"6F}}} \newfont{\symfont}{msam10 scaled\magstep1} \newcommand{\mathrel{\dabar@\dabar@\mathchar"0\msafam@4B}}{\mathrel{\mbox% {\symfont\char"39\char"39\char"4B}}} \newcommand{\text}[1]{\mathchoice{\mbox{\rm #1}}{\mbox{\rm #1}} {\mbox{\scriptsize\rm #1}}{\mbox{\tiny\rm #1}}} \newcommand{\operatorname}[1]{\mathop{\rm #1}\nolimits} \newcommand{\stackrel}{\stackrel} \newcommand{\eqref}[1]{{\rm (\ref{#1})}} \newcommand{\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}}{\mathchoice {\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}} {\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}} {\mathrel{\mskip-3mu/\mskip-4.5mu/\mskip-3mu}} {\mathrel{\mskip-3mu/\mskip-4.5mu/\mskip-3mu}} } \newcommand{{\Upsilon}}{{\Upsilon}} \newcommand{\Bbb{C}}{\Bbb{C}} \newcommand{{\cal D}}{{\cal D}} \newcommand{{\cal F}}{{\cal F}} \newcommand{{\cal K}}{{\cal K}} \newcommand{{\cal M}}{{\cal M}} \renewcommand{\O}{{\cal O}} \renewcommand{\P}{\Bbb P} \newcommand{\Bbb Q}{\Bbb Q} \newcommand{\Bbb R}{\Bbb R} \newcommand{\Bbb Z}{\Bbb Z} \newcommand{\operatorname{Aut}}{\operatorname{Aut}} \newcommand{\mathop{\widetilde{\rm Aut}}\nolimits}{\mathop{\widetilde{\rm Aut}}\nolimits} \newcommand{\operatorname{Hom}}{\operatorname{Hom}} \newcommand{\operatorname{Ker}}{\operatorname{Ker}} \newcommand{\operatorname{Coker}}{\operatorname{Coker}} \newcommand{\ |\ }{\ |\ } \newcommand{\operatorname{Spec}}{\operatorname{Spec}} \newcommand{\operatorname{gen}}{\operatorname{gen}} \renewcommand{\div}{\operatorname{div}} \newcommand{\operatorname{Div}}{\operatorname{Div}} \newcommand{\operatorname{WDiv}}{\operatorname{WDiv}} \newcommand{\operatorname{ad}}{\operatorname{ad}} \newcommand{\operatorname{CPL}}{\operatorname{CPL}} \newcommand{\operatorname{cpl}}{\operatorname{cpl}} \newcommand{{\operatorname{DR}}}{{\operatorname{DR}}} \newcommand{\operatorname{Pic}}{\operatorname{Pic}} \newcommand{P}{P} \newcommand{\polyhedron^\circ}{P^\circ} \newcommand{\Delta}{\Delta} \newcommand{\fan^\circ}{\Delta^\circ} \newcommand{{\cal N}}{{\cal N}} \newcommand{{\Xi}}{{\Xi}} \newcommand{_{\text{toric}}}{_{\text{toric}}} \newcommand{_{\text{poly}}}{_{\text{poly}}} \newcommand{\widehat{V}^\circ}{\widehat{V}^\circ} \newcommand{GIT}{GIT} \newcommand{PL}{PL} \newtheorem{theorem}{Theorem} \renewcommand{\thetheorem}{} \newtheorem{lemma}{Lemma} \renewcommand{\thelemma}{} \newtheorem{definition}{Definition} \renewcommand{\thedefinition}{} \newtheorem{conjecture}{Conjecture} \renewcommand{\theconjecture}{} \begin{document} \@dblarg\@xtitle{The Monomial-Divisor Mirror Map} \@dblarg\@xauthor{Paul S. Aspinwall, Brian R. Greene and David R. Morrison} \address{Aspinwall:\ School of Natural Sciences, Institute for Advanced Study, Princeton, NJ 08540\\ Greene:\ F.R. Newman Laboratory of Nuclear Studies, Cornell University, Ithaca, NY 14853\\ Morrison:\ School of Mathematics, Institute for Advanced Study, Princeton, NJ 08540} \date{} \thanks{Research partially supported by DOE grant DE-FG02-90ER40542, a National Young Investigator award, NSF grant DMS-9103827, and an American Mathematical Society Centennial Fellowship. } \renewcommand{\LARGE}{\Large\bf} \maketitle \renewcommand{\Large}{\large} \begin{abstract} For each family of Calabi-Yau hypersurfaces in toric varieties, Batyrev has proposed a possible mirror partner (which is also a family of Calabi-Yau hypersurfaces). We explain a natural construction of the isomorphism between certain Hodge groups of these hypersurfaces, as predicted by mirror symmetry, which we call the {\em monomial-divisor mirror map}. We indicate how this map can be interpreted as the differential of the expected mirror isomorphism between the moduli spaces of the two Calabi-Yau manifolds. We formulate a very precise conjecture about the form of that mirror isomorphism, which when combined with some earlier conjectures of the third author would completely specify it. We then conclude that the moduli spaces of the nonlinear sigma models whose targets are the different birational models of a Calabi-Yau space should be connected by analytic continuation, and that further analytic continuation should lead to moduli spaces of other kinds of conformal field theories. (This last conclusion was first drawn by Witten.) \end{abstract} \section{Reflexive polyhedra} Mirror symmetry, which proposes that Calabi-Yau manifolds should come in pairs with certain remarkable properties, is a phenomenon that was first observed in the physics literature \cite{dixon,LVW,CLS,GP}.% \footnote{For general mathematical background on mirror symmetry and mirror pairs, we refer the reader to \cite{mirrorbook} and \cite{guide}.} The most concrete realization of this phenomenon---actually the only one in which there is a physical argument linking the conformal field theories associated to the pair of Calabi-Yau manifolds---is given by the Greene-Plesser orbifolding construction \cite{GP} for Fermat hypersurfaces in weighted projective spaces and certain quotients of them by finite groups. Roan \cite{roan-mirror,roan-topological} has given a natural description of this construction in terms of toric geometry, and he showed that the mirror phenomenon in that case can be interpreted as a kind of duality between toric hypersurfaces. This enabled him to give rigorous mathematical proofs of certain formulas discovered by physicists. Batyrev \cite{batyrev1} has recently found an elegant characterization of Calabi-Yau hypersurfaces which are ample Cartier divisors in (mildly singular) toric varieties. The characterization is stated in terms of the {\em Newton polyhedron} of the hypersurface, which is the convex hull of the monomials appearing in its equation. This is always an integral polyhedron, that is, a compact convex polyhedron $P$ whose vertices are elements of a lattice $M$ in a real affine space $M_{\Bbb R}:=M\otimes\Bbb R$. Batyrev's characterization states that the general hypersurface with Newton polyhedron $P$ is Calabi-Yau (that is, has trivial canonical bundle and at worst Gorenstein canonical singularities), provided that $0$ is in the interior of $P$, and that each affine hyperplane $H\subset M_{\Bbb R}$ which meets $P$ in a face of codimension one has the form \[H:=\{y\in M_{\Bbb R}\ |\ \langle \ell,y\rangle=-1\}\] for some $\ell=\ell(H)$ in the dual lattice $N:=\operatorname{Hom}(M,\Bbb Z)$. An integral polyhedron with this property is called {\em reflexive}. The normals $\ell(H)$ of supporting hyperplanes $H$ for codimension-one faces of a reflexive polyhedron $P$ have as their convex hull the {\em polar polyhedron} $\polyhedron^\circ$, which is defined to be \[\polyhedron^\circ:=\{x\in N_{\Bbb R}\ |\ \langle x,y\rangle\ge-1 \text{ for all } y\inP\}.\] Batyrev showed that the polar polyhedron $\polyhedron^\circ$ of a reflexive integral polyhedron $P$ is itself a reflexive integral polyhedron (with respect to the dual lattice $N$). This led him to propose that hypersurfaces $X$ and $Y$ with Newton polyhedra $P$ and $\polyhedron^\circ$, respectively, should form a mirror pair. The evidence for Batyrev's proposal is of several kinds. First and foremost is the fact that this polar polyhedron construction specializes to Roan's interpretation of the Greene-Plesser orbifolding construction in the case of quotients of Fermat hypersurfaces in weighted projective spaces. This is encouraging, since as noted above the Greene-Plesser construction provides the {\em only} complete example of mirror symmetry for hypersurfaces---the only example for which there is a physical argument for the existence of a mirror isomorphism of the corresponding conformal field theories. A second piece of evidence (which we discuss more fully below) is an isomorphism between certain Hodge groups associated to $X$ and $Y$ (extending the work of Roan), as would be predicted by mirror symmetry. And finally, Batyrev shows that his construction is compatible with the existence of certain ``quantum symmetries'' as expected based on physical reasoning. This quantum symmetry behavior looks somewhat unnatural mathematically, so verifying it is an important check. This evidence falls short of fully establishing a mirror symmetry relationship between $X$ and $Y$, since it does not link the corresponding conformal field theories. However, it does provide strong grounds for suspecting the existence of a mirror isomorphism. And the naturality of Batyrev's polar polyhedron construction is extremely compelling (at least to mathematicians). If mirror symmetry does hold between $X$ and $Y$, there will be an isomorphism between Hodge groups $H^{1,1}(\widehat{X})$ and $H^{d-1,1}(\widehat{Y})$, where $\widehat{X}\to X$ and $\widehat{Y}\to Y$ are appropriate (partial) resolutions of singularities, and $d$ is the common dimension of $X$ and $Y$. The existence of such an isomorphism had been shown quite explicitly by Roan \cite{roan-mirror,roan-topological} in the weighted Fermat hypersurface case---% the general case is addressed by Batyrev. In the earlier preprint versions of \cite{batyrev1}, Batyrev found an equality between the dimensions of certain subspaces $H^{1,1}_{\text{toric}}(\widehat{X})$ and $H^{d-1,1}_{\text{poly}}(\widehat{Y})$ of the Hodge groups, mistakenly believed to have been the entire spaces. In the final version of \cite{batyrev1}, he shows that the full Hodge groups are isomorphic, following suggestions made by the present authors. The error in the earlier version of the paper was fortuitous, however, as it revealed that the mirror isomorphism might be expected to preserve those subspaces. In this note, we explain a very natural construction of the isomorphism between $H^{1,1}_{\text{toric}}(\widehat{X})$ and $H^{d-1,1}_{\text{poly}}(\widehat{Y})$ , and indicate how it can be interpreted as the differential of the expected mirror map between the moduli spaces (when restricted to appropriate subspaces of those moduli spaces). The space $H^{d-1,1}_{\text{poly}}(\widehat{Y})$ is isomorphic to the space of first-order polynomial deformations of $Y$, and can be generated by {\em monomials}; the space $H^{1,1}_{\text{toric}}(\widehat{X})$ consists of that part of the second cohomology of $\widehat{X}$ coming from the ambient toric variety, and can be generated by toric {\em divisors}. Our map comes from a natural one-to-one correspondence between monomials and toric divisors whose definition is inspired by the constructions of Roan and Batyrev; we have named it the {\em monomial-divisor mirror map}. \section{Divisors} \label{divisorsection} Our first task is to describe the partial resolutions of singularities we will use, and the divisors on them. Let $\Delta$ be a fan determining a toric variety (see \cite{Oda} or \cite{Fulton} for the definitions, and for proofs of the facts we review below). The support $|\Delta|$ of $\Delta$ is a subset of a real vector space $N_{\Bbb R}$, and the convex cones $\sigma$ in the fan $\Delta$ are rational polyhedral cones with respect to a lattice $N$ in $N_{\Bbb R}$; the algebraic torus which acts on the toric variety is $T:=N\otimes\Bbb{C}^*$. We let $\Delta(1)$ denote the set of one-dimensional cones in $\Delta$. There is a natural {\em generator} map $\operatorname{gen}:\Delta(1)\to N$ which assigns to each one-dimensional cone $\rho$ the unique generator $\operatorname{gen}(\rho)$ of the semigroup $\rho\cap N$. Each such $\rho$ also has an associated $T$-invariant Weil divisor $D_\rho$ in the toric variety, which is the closure of the $T$-orbit corresponding to the cone $\rho$. One can always describe a {\em projective} toric variety by beginning with a compact convex polyhedron $P$ in a real affine space $M_{\Bbb R}$, integral with respect to a lattice $M$ in $M_{\Bbb R}$. The projective toric variety is then determined by the {\em normal fan} of the polyhedron $P$; this is the fan ${\cal N}(P)$ consisting of all cones ${\cal N}(P,p)$ to $P$ at $p\inP$, where \[{\cal N}(P,p):=\{x\in N_{\Bbb R}\ |\ \langle x,p\rangle\le\langle x,y\rangle \text{ for all } y\inP\},\] and where $N:=\operatorname{Hom}(M,\Bbb Z)$ is the dual lattice, and $N_{\Bbb R}:=N\otimes\Bbb R$. Each proper face of the polar polyhedron $\polyhedron^\circ$ of $P$ is contained in a unique cone in ${\cal N}(P)$, which is the cone over that face. The toric variety $V$ determined by the fan ${\cal N}(P)$ is the natural one in which the hypersurfaces $X$ with Newton polyhedron $P$ are ample divisors. In the case of a reflexive integral polyhedron $P$, the general such $X$ is an anti-canonical divisor in $V$ and will be a Calabi-Yau variety with canonical singularities, as proved by Batyrev \cite{batyrev1}. We need to partially resolve the singularities of $V$ while retaining the triviality of the canonical bundle of the hypersurface, getting as close as possible to a complete resolution. To do this, construct a blowup $\widehat{V}\to V$, determined by a fan $\Delta$ which is a subdivision of the fan ${\cal N}(P)$. There will be an induced blowup $\widehat{X}\to X$ of hypersurfaces, where $\widehat{X}$ is the proper transform of $X$ on $\widehat{V}$. In order to maintain the triviality of the canonical bundle of $\widehat{X}$, restrict the set $\Delta(1)$ of one-dimensional cones as follows: the image ${\Xi}:=\operatorname{gen}(\Delta(1))$ of $\Delta(1)$ in $N$ should lie in the set $\polyhedron^\circ\cap N$, where $\polyhedron^\circ$ is the polar polyhedron of $P$. (We will sometimes restrict ${\Xi}$ to lie in the subset $(\polyhedron^\circ\cap N)_0\subset{\polyhedron^\circ\cap N}$ consisting of those lattice points in $\polyhedron^\circ$ which do not lie in the interior of a codimension-one face of $\polyhedron^\circ$.) Oda and Park \cite{OP} (cf.\ also \cite{Stanley}) have shown the existence of a simplicial subdivision $\Delta$ of the fan ${\cal N}(P)$ such that $\operatorname{gen}(\Delta(1))=(\polyhedron^\circ\cap N){-}\{0\}$ (or any subset thereof), and such that the corresponding $\widehat{V}$ is projective. In general, there will be many such fans $\Delta$. Since the fan $\Delta$ is simplicial, the toric variety $\widehat{V}$ has the structure of an {\em orbifold} (formerly called {\em $V$-manifold} \cite{satake3}): it can be covered by open sets of the form $U/G_U$ where $G_U$ is a finite group acting on a manifold $U$ such that the fixed locus of any $1\ne g\in G_U$ has real codimension at least $2$. The open sets $U/G_U$ are used to define the notion of {\em orbifold-smooth differential forms}, pieced together from $G_U$-invariant smooth forms on the open sets $U$. Many of the theorems about the differential geometry of smooth algebraic varieties have natural orbifold versions. In particular, there are orbifold de~Rham cohomology groups $H^k_{\operatorname{DR}}(\widehat{V},\Bbb R)$ isomorphic to the real \v Cech cohomology \cite{satake3}, and orbifold Hodge groups $H^{p,q}(\widehat{V})$ which satisfy a version of the Dolbeault theorem \cite{baily1}. The general hypersurface $\widehat{X}\subset\widehat{V}$ is also an orbifold, and has orbifold de Rham and Hodge groups of its own. We can describe the group $\operatorname{WDiv}_T(\widehat{V})$ of toric Weil divisors on $\widehat{V}$ and their images in the Chow group $A_{n-1}(\widehat{V})$ (where $n=d+1$ is the dimension of $\widehat{V}$), as follows (cf.\ Cox \cite{cox}). There is a natural isomorphism\footnote{We use the notation $\Bbb Z\langle S\rangle$ for the free abelian group on the set $S$, and $\Bbb Z^S$ for the $\Bbb Z$-module of maps from $S$ to $\Bbb Z$, which is naturally isomorphic to the dual lattice $\operatorname{Hom}(\Bbb Z\langle S\rangle,\Bbb Z)$ of $\Bbb Z\langle S\rangle$. The map determined by $\varphi\in\Bbb Z^S$ is denoted by $s\mapsto\varphi_s$. } $\alpha:\Bbb Z^{\Xi}\to\operatorname{WDiv}_T(\widehat{V})$ which sends the function $\varphi\in\Bbb Z^{\Xi}$ to the divisor \[\sum\varphi_aD_{\operatorname{gen}^{-1}(a)}.\] Under this isomorphism, if we define an embedding $\operatorname{ad}_{\Xi}:M\to\Bbb Z^{\Xi}$ by sending $m\in M$ to the function $\operatorname{ad}_{\Xi}(m)$ defined by $\operatorname{ad}_{\Xi}(m):a\mapsto\langle a,m\rangle$, then \[\div(\chi^m)=-\alpha(\operatorname{ad}_{\Xi}(m)),\] where $\chi^m:T\to\Bbb{C}^*$ is the character of $T$ associated to $m$, regarded as a meromorphic function on $\widehat{V}$. Thus, $M$ gives rise to linear equivalences among toric divisors. In fact, there is an exact sequence \begin{equation} \label{exactseq} 0 \longrightarrow M \stackrel{\operatorname{ad}_{\Xi}}{\longrightarrow} \Bbb Z^{{\Xi}} \stackrel{\bar{\alpha}}{\longrightarrow} A_{n-1}(\widehat{V}) \longrightarrow 0 , \end{equation} where $\bar{\alpha}$ denotes the composite of $\alpha$ with the projection to the Chow group. This is nothing other than the usual description of $A_{n-1}$ as ``divisors modulo linear equivalence'', since $\Bbb Z^{\Xi}$ represents toric divisors and $M$ represents the linear equivalences among them. \medskip To compute the group of toric divisors on the hypersurface $\widehat{X}$, we use the natural restriction maps from divisors on $\widehat{V}$ (which exists since each toric divisor on $\widehat{V}$ meets $\widehat{X}$ in a subvariety of codimension $1$): \[\begin{array}{ccccccccc} 0&\longrightarrow&M&\longrightarrow&\operatorname{WDiv}_T(\widehat{V})&\longrightarrow& A_{n-1}(\widehat{V})&\longrightarrow&0\\ &&{\scriptstyle||}&&\downarrow&&\downarrow&&\\ 0&\longrightarrow&M&\longrightarrow&\operatorname{WDiv}_T(\widehat{X})&\longrightarrow& A_{d-1}(\widehat{X})&& \end{array}.\] This time, the toric divisors need not generate the entire Chow group; we denote the image of $\operatorname{WDiv}_T(\widehat{X})$ in $A_{d-1}(\widehat{X})$ by $A_{d-1}(\widehat{X})_{\text{toric}}$. Its complexification we call the {\em toric part of $H^{1,1}$}, denoted by $H^{1,1}_{\text{toric}}(\widehat{X}):=A_{d-1}(\widehat{X})_{\text{toric}}\otimes\Bbb{C}$. The kernel of the restriction map $\operatorname{WDiv}_T(\widehat{V})\to\operatorname{WDiv}_T(\widehat{X})$ is easy to describe. A divisor with trivial restriction must be supported on divisors which are disjoint from the general hypersurface $\widehat{X}\subset\widehat{V}$. Since the line bundle $\O_V(X)$ is generated by its global sections, the general hypersurface $X\subset V$ will not meet the zero-dimensional strata of $V$ (in the stratification by $T$-orbits). So any divisor on $\widehat{V}$ which maps to such a stratum will be disjoint from $\widehat{X}$, the proper transform of $X$. Such divisors are characterized by the property that the corresponding point in ${\Xi}$ lies in the interior of some codimension-one face of $\polyhedron^\circ$. Other toric divisors on $\widehat{V}$ cannot be disjoint from $\widehat{X}$, since they map to larger strata of $V$ which are not disjoint from $X$. Thus, if we let ${\Xi}_0={\Xi}\cap(\polyhedron^\circ\cap N)_0$ be the subset of ${\Xi}$ consisting of all points which do {\em not} lie in interiors of codimension-one faces of $\polyhedron^\circ$, we find that $\operatorname{WDiv}_T(\widehat{X})\cong\Bbb Z^{{\Xi}_0}$ and that \begin{equation} \label{divisorsA} A_{d-1}(\widehat{X})_{\text{toric}}\cong\operatorname{Coker}(\operatorname{ad}_{{\Xi}_0})\cong \Bbb Z^{{\Xi}_0}/M. \end{equation} In particular, if ${\Xi}\supset(\polyhedron^\circ\cap N)_0-\{0\}$ then $A_{d-1}(\widehat{X})_{\text{toric}}\cong\Bbb Z^{(\polyhedron^\circ\cap N)_0-\{0\}}/M$, and hence \begin{equation} \label{divisors} H^{1,1}_{\text{toric}}(\widehat{X})\cong(\Bbb Z^{(\polyhedron^\circ\cap N)_0-\{0\}}/M)\otimes\Bbb{C}. \end{equation} \section{Monomials} Our task in this section is to describe moduli spaces for hypersurfaces in $\widehat{V}$. We retain the notation of the previous section: $\widehat{V}$ is the toric variety associated to a subdivision $\Delta$ of the normal fan ${\cal N}(P)$ of a reflexive polyhedron $P$. We assume that $\Delta$ is simplicial, so that $\widehat{V}$ is $\Bbb Q$-factorial; we also assume that $\widehat{V}$ is projective. Given a hypersurface $\widehat{X}\subset\widehat{V}$, the space of first order deformations of complex structure of $\widehat{X}$ is isomorphic to $H^1(\Theta_{\widehat{X}})$. The simplest way to deform the complex structure on $\widehat{X}$ is to perturb the equation of the hypersurface; this leads to a subspace $H^1(\Theta_{\widehat{X}})_{\text{poly}}\subset H^1(\Theta_{\widehat{X}})$ of {\em polynomial} first-order deformations. (It is quite possible for this to be a proper subspace \cite{pdm}.) In the case that $\widehat{X}\subset\widehat{V}$ is a Calabi-Yau hypersurface, we can use the isomorphism $H^1(\Theta_{\widehat{X}}) \cong H^{d-1,1}(\widehat{X})$ to also specify a ``polynomial'' subspace $H^{d-1,1}_{\text{poly}}(\widehat{X})\subset H^{d-1,1}(\widehat{X})$ of the corresponding Hodge group. In principle, the moduli spaces of the hypersurfaces $\{\widehat{X}\}$ should be fairly easy to describe. Global sections of $\O_{\widehat{V}}(\widehat{X})$ provide equations for the hypersurfaces, and the entire family can be described as $\P H^0(\O_{\widehat{V}}(\widehat{X}))$. But we need to mod out by automorphisms of $\widehat{V}$, and this is where technical complications arise. Let $D$ be a toric divisor on $\widehat{V}$, and write $D=\sum_{a\in{\Xi}} d_a D_{\operatorname{gen}^{-1}(a)}$. There is a natural isomorphism between $H^0(\O(D))$ and the space $\Bbb{C}^{P_D\cap M}$, where $P_D$ is the polytope \[P_D:=\{y\in M_{\Bbb R} \ |\ \langle a,y\rangle\ge-d_a \text{ for all } a\in{\Xi}\} \] (cf.~\cite{Fulton}). In fact, if we identify $H^0(\O(D))$ with the space of meromorphic functions on $\widehat{V}$ which have (at worst) poles along $D$, then to each $m\in P_D\cap M$ we can associate the meromorphic function $\chi^m$: it has at worst poles along $D$ thanks to the definition of $P_D$. In the special case $D=\sum_{\rho\in\Delta(1)}D_\rho\in|-K_{\widehat{V}}|$, the polytope $P_{\Sigma\, D_\rho}$ coincides with the original polyhedron $P\subset M_{\Bbb R}$ used to describe $V$. The automorphism group $\operatorname{Aut}(\widehat{V})$ of a $\Bbb Q$-factorial toric variety $\widehat{V}$ has been described recently by Cox \cite{cox}, generalizing some results from the smooth case due to Demazure \cite{demazure}. Cox's description is in terms of a central extension of $\operatorname{Aut}(\widehat{V})$ by a torus $G$, which fits in an exact sequence \begin{equation}\label{auttilde} 1\longrightarrow G\longrightarrow \mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}) \longrightarrow \operatorname{Aut}(\widehat{V})\longrightarrow1, \end{equation} where $G:=\operatorname{Hom}(A_{n-1}(\widehat{V}),\Bbb{C}^*)$. The advantage of working with $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ is that it acts naturally on all cohomology groups $H^0(\O(D))$ at once. The clearest way to see these actions is to follow Cox again and introduce the {\em homogeneous coordinate ring}\/ $S:=\Bbb{C}[x_a]_{(a\in{\Xi})}$ of $\widehat{V}$. This ring can be (multi) graded by defining the {\em degree}\/ of the monomial $\prod x_a^{\varphi_a}$ to be the divisor class $[\sum \varphi_a D_{gen^{-1}(a)}]$ in $A_{n-1}(\widehat{V})$. For a fixed divisor $D$, the set of elements of degree $[D]$ in the homogeneous coordinate ring can be identified with $H^0(\O(D))$: the meromorphic function $\chi^m$ with $m\in P_D\cap M$ corresponds to the homogeneous monomial $x^{\div(\chi^m)+D}:=\prod x_a^{\langle a,m\rangle + d_a}$. The torus $T:=\operatorname{Hom}(M,\Bbb{C}^*)$ which acts on $\widehat{V}$ is naturally a subgroup of $\operatorname{Aut}(\widehat{V})$; the induced extension $\widetilde{T}$ of $T$ by $G$ has the form $\widetilde{T}:=\operatorname{Hom}(\Bbb Z^{\Xi},\Bbb{C}^*)$. In fact, if we apply the functor $\operatorname{Hom}(\mbox{---},\Bbb{C}^*)$ to the natural exact sequence \eqref{exactseq}, we get a sequence for $\widetilde{T}$ which fits as the first row in the commutative diagram \[\begin{array}{ccccccccc} 1&\longrightarrow&G&\longrightarrow&\widetilde{T}&\longrightarrow& T&\longrightarrow&1\\ &&{\scriptstyle||}&&\cap&&\cap&&\\ 1&\longrightarrow&G&\longrightarrow&\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})&\longrightarrow& \operatorname{Aut}(\widehat{V})&\longrightarrow&1 \end{array}.\] The grading of the homogeneous coordinate ring $S$ can also be described in terms of the action of $G$ on $S$. The torus $\widetilde{T}$ acts on $S$ in a transparent way: each monomial in $S$ can be written in the form $x^\varphi=\prod x_a^{\varphi_a}$ for some $\varphi\in\Bbb Z^{{\Xi}}$, and the action of $t\in\operatorname{Hom}(\Bbb Z^{{\Xi}},\Bbb{C}^*)$ sends $x^\varphi$ to $t(\varphi)\cdot x^\varphi$. When we restrict this action to the subgroup $G=\operatorname{Hom}(A_{n-1}(\widehat{V},\Bbb{C}^*))$, then for each divisor class $[D]$, the subspace of $S$ on which $G$ acts via the character $\gamma\mapsto\gamma([D])$ is precisely $H^0(\O(D))\cong\bigoplus\Bbb{C}\cdot x^{\div(\chi^m)+D}$. The induced action of $t\in\widetilde{T}$ on $H^0(\O(D))$ then sends $\chi^m$ to $t(\alpha^{-1}(\div(\chi^m)+D))\cdot\chi^m$, for every $m\in P_D\cap M$. This action can be described in terms of the map $\Bbb Z\langle P_D\cap M\rangle\to\Bbb Z^{\Xi}$ defined by \begin{equation}\label{action} m\mapsto(a\mapsto\langle a,m\rangle+d_a), \end{equation} which induces a homomorphism of tori $\widetilde{T}\to(\Bbb{C}^*)^{P_D\cap M}$ that determines the action of $\widetilde{T}$ on $\Bbb{C}^{P_D \cap M}$. Notice that the map \eqref{action} factors as a composite of two maps \begin{equation} \label{twomaps} \Bbb Z\langle P_D\cap M\rangle\to M\oplus\Bbb Z\to\Bbb Z^{\Xi} \end{equation} with the first map given by $m\mapsto(m,1)$ and the second map given by $(m,k)\mapsto \operatorname{ad}_{\Xi}(m)+k\cdot\alpha^{-1}(D)$. The corresponding homomorphism of tori factors as \begin{equation} \label{torusfactor} \widetilde{T}\to T\times\Bbb{C}^*\to(\Bbb{C}^*)^{P_D\cap M}. \end{equation} In the special case $D=\sum D_\rho$, the induced map $\Bbb{C}^*\to(\Bbb{C}^*)^{P\cap M}$ is simply the diagonal embedding. The groups $\operatorname{Aut}(\widehat{V})$ and $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ are not in general reductive. Thus, to construct moduli spaces for hypersurfaces\footnote{Batyrev \cite{batyrev2} has constructed moduli spaces for {\em affine}\/ hypersurfaces, obtaining a somewhat different space than ours if $(P\cap M)_0\ne(P\cap M)$ (it even has a different dimension). Batyrev and Cox \cite{BC} have recently considered a construction similar to the one described here.} on $\widehat{V}$, we should use Fauntleroy's extension \cite{fauntleroy1} of Mumford's Geometric Invariant Theory (GIT) \cite{GIT}, and attempt to construct a quotient for the action of $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ on $H^0(\O(D))$. It would be interesting to know if this construction of moduli spaces can be carried out in general---Fauntleroy has carried it out in some special cases \cite{fauntleroy2}. We can at least obtain a birational model of the desired moduli space by using a fairly standard result (cf.\ \cite{Rosenlicht,CDT}) which guarantees the existence of an $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$-stable Zariski-open set $U\subset H^0(\O(D))$ such that the geometric quotient $U/\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ exists. We indicate the birational class of such quotients with the notation $H^0(\O(D))\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu} \mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$. In the case of interest in this paper, $D=-K_{\widehat{V}}$. To study this particular moduli space, we take a simpler course of action, and restrict our attention to a subspace of $H^0(\O(-K_{\widehat{V}}))$ on which $\widetilde{T}$ acts in such a way that the quotient exists and has the ``expected'' dimension for the entire moduli space. In a wide class of examples, $\widetilde{T}$ is in fact the connected component of the identity in $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$, and the only differences between our moduli space and the ``true'' moduli space for hypersurfaces are a remaining quotient by a finite group, and a possible ambiguity in the choice of Zariski-open set used in constructing the quotient. In particular, the map from our space to the true moduli space is a dominant map between spaces of the same dimension. We hope that this latter property is true in general, but postpone that question to a future investigation. Our construction of a simplified model for the hypersurface moduli space relies on another result of Demazure and Cox about $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$. They show that \[\dim\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})=\dim\widetilde{T}+\#(R(N,\Delta)),\] where \begin{eqnarray*} R(N,\Delta):=\{m\in M&\ |\ &\langle \operatorname{gen}(\rho),m\rangle\le1 \text{ for all }\rho\in\Delta(1),\\ && \text{with equality for a unique } \rho=\rho_m\in\Delta(1)\} \end{eqnarray*} is the set of {\em roots} of the toric variety $\widehat{V}$ associated to the fan $\Delta$. Note that for each root $m\in R(N,\Delta)$, we have $-m\inP\cap M$. In fact, the set $-R(N,\Delta)$ can be characterized as the subset of $P\cap M$ consisting of lattice points which lie in the interiors of codimension-one faces of $P$. We can thus decompose \[P\cap M=-R(N,\Delta)\cup(P\cap M)_0,\] and write \[\Bbb{C}^{P\cap M}=\Bbb{C}^{-R(N,\Delta)}\oplus\Bbb{C}^{(P\cap M)_0}.\] The subgroup $\widetilde{T}\subset\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ preserves this direct sum decomposition, so we can let $\widetilde{T}$ act on the second factor $\Bbb{C}^{(P\cap M)_0}$ alone. Our ``simplified hypersurface moduli space'' will be the GIT\ quotient $\Bbb{C}^{(P\cap M)_0}_{\text{ss}}/\widetilde{T}$. (We regard the action of $\widetilde{T}$ on $\Bbb{C}^{(P\cap M)_0}$ as specifying a linearization of the action on $\P(\Bbb{C}^{(P\cap M)_0})$, so there is no ambiguity in the choice of GIT\ quotient.) There is then a natural rational map \begin{equation}\label{rationalmap} \Bbb{C}^{(P\cap M)_0}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\widetilde{T}\mathrel{\dabar@\dabar@\mathchar"0\msafam@4B} \Bbb{C}^{P\cap M}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}) \end{equation} which could be refined to a regular map \begin{equation}\label{map} \Bbb{C}^{(P\cap M)_0}_{\text{ss}}/\widetilde{T}\to U/\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}) \end{equation} if an appropriate set of ``semistable'' points $U\subset\Bbb{C}^{P\cap M}$ were available from (generalized) GIT. Note that $\operatorname{Ker}(\xi_{[-K]})$, which is a subgroup of both $\widetilde{T}$ and $\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$, acts trivially on both spaces. By equation \eqref{torusfactor}, $\widetilde{T}/\operatorname{Ker}(\xi_{[-K]})\cong T\times\Bbb{C}^*$. Note also that the two quotient spaces can be expected to have the same dimension. \begin{definition} We say that the family $\{\widehat{X}\}$ has the {\em dominance property} if the natural rational map $\Bbb{C}^{(P\cap M)_0}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\widetilde{T}\mathrel{\dabar@\dabar@\mathchar"0\msafam@4B} \Bbb{C}^{P\cap M}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ is a dominant map between two varieties of the same dimension. (Note that this property is independent of the choice of quotients.) \end{definition} This dominance property clearly holds if $R(N,\Delta)=\emptyset$; we expect that it should hold in general, but have not checked this. The ``simplified hypersurface moduli space'' parameterizes hypersurfaces with equations of the form \[\sum_{m\in(P\cap M)_0}c_m\chi^m=0\] modulo the equivalences given by the action of $\widetilde{T}/\operatorname{Ker}(\xi_{[-K]})\cong T\times\Bbb{C}^*$. The $\Bbb{C}^*$ factor is diagonally embedded in $(\Bbb{C}^*)^{(P\cap M)_0}$, and so gives an overall scaling of the equation. We can describe a Zariski-open subset\footnote{The apparent lack of naturality in this step of our construction---why restrict to a subset?---will be redressed later in the paper.} of our moduli space by restricting to equations with $c_0\ne0$, and using the overall scaling of the equation to set that coefficient $c_0$ equal to $1$. Thus, the open subset can be described as a quotient $\Bbb{C}^{(P\cap M)_0{-}\{0\}}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu} T$ with a point $c\in\Bbb{C}^{(P\cap M)_0{-}\{0\}}$ corresponding to the hypersurface with equation \[\chi^0+\sum_{m\in(P\cap M)_0{-}\{0\}}c_m\chi^m=0.\] Let ${\Upsilon}_0=(P\cap M)_0-\{0\}$ to simplify notation. Here is the crucial observation for the construction of the monomial-divisor mirror map: the action of $T=N\otimes\Bbb{C}^*$ on $\Bbb{C}^{{\Upsilon}_0}$ is induced by tensoring the homomorphism $\operatorname{ad}_{{\Upsilon}_0}:N\to \Bbb Z^{{\Upsilon}_0}$ with $\Bbb{C}^*$. (The explicit identification of $\operatorname{ad}_{{\Upsilon}_0}$ as the homomorphism needed to specify the $T$-action follows immediately from the definition of the maps in equation \eqref{twomaps}, since $\operatorname{ad}_{{\Upsilon}_0}$ is dual to the natural map $\Bbb Z\langle{{\Upsilon}_0}\rangle\to M$ induced by the inclusion ${{\Upsilon}_0}\subset M$.) In particular, the tangent space to the simplified moduli space $\Bbb{C}^{(P\cap M)_0{-}\{0\}}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu} T$ has the form $(\Bbb Z^{(P\cap M)_0{-}\{0\}}/N)\otimes\Bbb{C}$. When the family $\{\widehat{X}\}$ has the dominance property (e.g., when $R(N,\Delta)=\emptyset$), the induced rational map $\Bbb{C}^{(P\cap M)_0{-}\{0\}}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu} T\mathrel{\dabar@\dabar@\mathchar"0\msafam@4B} \Bbb{C}^{P\cap M}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$ is also dominant. In this case, we can describe the tangent space to the space of polynomial deformations of a general Calabi-Yau hypersurface $\widehat{X}\subset\widehat{V}$ as \begin{equation} \label{monomials} T_{[\widehat{X}],\,\,\Bbb{C}^{P\cap M}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})}= H^{d-1,1}_{\text{poly}}(\widehat{X})\cong (\Bbb Z^{(P\cap M)_0{-}\{0\}}/N)\otimes\Bbb{C}, \end{equation} where $[\widehat{X}]$ represents the class of $\widehat{X}$ in $\Bbb{C}^{P\cap M}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V})$. We can apply these same considerations to the family of hypersurfaces determined by the polar polyhedron $\polyhedron^\circ$, which Batyrev has proposed as a mirror partner for the family $\{\widehat{X}\}$. To do this, we need to choose a simplicial subdivision $\fan^\circ$ of ${\cal N}(\polyhedron^\circ)$ which determines a projective toric variety $\widehat{V}^\circ$ and a family of hypersurfaces $\widehat{Y}\subset\widehat{V}^\circ$. Replacing $\widehat{X}$, $P$, $M$, $N$ by $\widehat{Y}$, $\polyhedron^\circ$, $N$, $M$, respectively, in equation \eqref{monomials}, we find that \begin{equation} \label{Ymonomials} T_{[\widehat{Y}],\,\,\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)}= H^{d-1,1}_{\text{poly}}(\widehat{Y})\cong (\Bbb Z^{(\polyhedron^\circ\cap N)_0{-}\{0\}}/M)\otimes\Bbb{C}, \end{equation} whenever $\{\widehat{Y}\}$ has the dominance property. The monomial-divisor mirror map is now evident, when one compares equations \eqref{divisors} and \eqref{Ymonomials}. (Note that the same map $\operatorname{ad}_{(\polyhedron^\circ\cap N)_0-\{0\}}$ is used to define the embedding $M\to\Bbb Z^{(\polyhedron^\circ\cap N)_0-\{0\}}$ in both cases.) \begin{theorem} Let $P$ be a reflexive polyhedron, integral with respect to the lattice $M$, and let $\polyhedron^\circ$ be its polar polyhedron, which is integral with respect to the dual lattice $N$. Let $\Delta$ and $\fan^\circ$ be simplicial subdivisions of ${\cal N}(P)$ and ${\cal N}(\polyhedron^\circ)$, respectively, and let $\{\widehat{X}\}$ and $\{\widehat{Y}\}$ be the corresponding families of Calabi-Yau hypersurfaces. Assume that $(\polyhedron^\circ\cap N)_0{-}\{0\}\subset \operatorname{gen}(\Delta(1)) \subset (\polyhedron^\circ\cap N){-}\{0\}$. Assume also that $R(M,\fan^\circ)=\emptyset$, or more generally, simply assume that $\{\widehat{Y}\}$ has the dominance property. Then there is a natural isomorphism \begin{equation}\label{mdmm} H^{d-1,1}_{\text{poly}}(\widehat{Y})\stackrel{\cong}{\to} H^{1,1}_{\text{toric}}(\widehat{X}) \end{equation} induced by equations \eqref{divisors} and \eqref{Ymonomials}, since both spaces are naturally isomorphic to $\operatorname{Coker}(\operatorname{ad}_{(\polyhedron^\circ\cap N)_0-\{0\}})\otimes\Bbb{C}$. We call the isomorphism \eqref{mdmm} the {\em monomial-divisor mirror map}. \end{theorem} \section{K\"ahler cones} Let $\widehat{V}$ be a $\Bbb Q$-factorial toric variety, determined by a simplicial fan $\Delta$, and let ${\Xi}=\operatorname{gen}(\Delta(1))$. We can describe the group $\operatorname{Div}_T(\widehat{V})$ of toric Cartier divisors on $\widehat{V}$ as follows. In order for $D=\sum d_aD_{\operatorname{gen}^{-1}(a)}$ to be Cartier, there must be a continuous piecewise linear (PL) function $\psi_D:|\Delta|\to\Bbb R$ (called the {\em support function determined by $D$}) which is linear on each cone $\sigma\in\Delta$, which takes integer values on $|\Delta|\cap N$, and which satisfies \begin{equation}\label{PL} \psi_D(a)=- d_a\qquad\text{ for all }a\in{\Xi}. \end{equation} Since the fan $\Delta$ is simplicial, the PL\ function $\psi_D$ is completely determined by the values specified in equation \eqref{PL} and the fan $\Delta$: one just extends by linearity to each (simplicial) cone in $\Delta$. The integrality condition $\psi_D(|\Delta|\cap N)\subset\Bbb Z$ remains nontrivial, however. (If we require instead that $\psi_D(|\Delta|\cap N)\subset\Bbb Q$, we get the group of $\Bbb Q$-Cartier divisors.) Since $\Delta$ is simplicial, every Weil divisor is $\Bbb Q$-Cartier, that is, $\widehat{V}$ is $\Bbb Q$-factorial. Put another way, there is a natural isomorphism between $\operatorname{Pic}(\widehat{V})\otimes\Bbb Q$ and $A_{n-1}(\widehat{V})\otimes\Bbb Q$. The ample Cartier divisors (or ample $\Bbb Q$-Cartier divisors) are characterized by {\em strict convexity} of $-\psi_D$, where $\psi_D$ is the support function determined by $D$ (cf.\ \cite{Fulton}).\footnote{Our signs are chosen to conform to the literature as closely as possible, while giving the term ``convexity'' its conventional meaning.} This means the following: given a PL\ function $\eta$ which is linear on each cone $\sigma\in\Delta$, let $u_\sigma\in M=\operatorname{Hom}(N,\Bbb Z)$ be the linear function such that \[\langle x,u_\sigma\rangle =\eta(x)\quad\text{ for all } x\in\sigma.\] The convexity condition is: \[\eta(x)\ge\langle x,u_\sigma\rangle\quad\text{ for all } x\in|\Delta|;\] the convexity if {\em strict} if the inequality is strict for all $x\not\in\sigma$. (In fact, it suffices to check this at the points of ${\Xi}$: for convexity, we must have \[\eta(a)\ge\langle a,u_\sigma\rangle\quad\text{ for all } a\in{\Xi}\] with equality whenever $a\in\sigma$; strict convexity requires a strict inequality whenever $a\not\in\sigma$.) The cone of real convex PL\ functions is denoted by $\operatorname{CPL}(\Delta)$, following the notation of Oda and Park \cite{OP}. If there exists a strictly convex function in $\operatorname{CPL}(\Delta)$, the fan $\Delta$ is called {\em regular}. There is a related cone (introduced by Gel'fand, Zelevinski\v\i, and Kapranov \cite{GKZ}): \[\operatorname{CPL}^\sim(\Delta)=\{\varphi\in\Bbb R^{\Xi}\ |\ \exists\ \eta\in\operatorname{CPL}(\Delta) \text{ with } \varphi_a=\eta(a) \text{ for all } a\in{\Xi}\}.\] The definitions are constructed so that if $D$ is an ample $\Bbb Q$-Cartier divisor, then $\alpha^{-1}(D)\in\operatorname{CPL}^\sim(\Delta)$. (One can choose $\eta=-\psi_D$ in the definition.) Note that $M_{\Bbb R}$ is contained in $\operatorname{CPL}^\sim(\Delta)$, and the corresponding $\eta$'s are precisely the {\em smooth} PL\ functions. The image of $\operatorname{CPL}^\sim(\Delta)$ in $\Bbb R^{\Xi}/M_{\Bbb R}$, which we may think of as the set of convex PL\ functions modulo smooth PL\ functions, is denoted by $\operatorname{cpl}(\Delta)$. Under the isomorphism $\Bbb R^{\Xi}/M_{\Bbb R}\cong A_{n-1}(\widehat{V})\otimes\Bbb R (\cong\operatorname{Pic}(\widehat{V})\otimes\Bbb R)$, this cone $\operatorname{cpl}(\Delta)$ maps to the closed real cone generated by the ample divisor classes on $\widehat{V}$. An effective method of calculating all possible cones $\operatorname{cpl}(\Delta)$ in terms of the ``linear Gale transform'' is described in \cite{OP}. The exponential sheaf sequence gives rise to an isomorphism $\operatorname{Pic}(\widehat{V})\cong H^2(\widehat{V},\Bbb R)\cong H^{1,1}(\widehat{V},\Bbb R)$, since $h^i(\O_{\widehat{V}})=0$ for $i>0$. Now there is a natural notion of {\em positivity} for orbifold-smooth $(1,1)$-forms: one requires that the $G_U$-invariant $(1,1)$-forms on the local uniformizing sets $U$ be positive. The K\"ahler form of every orbifold-K\"ahler metric is easily seen to be a positive, orbifold-smooth $(1,1)$-form. The set ${\cal K}(\widehat{V})\subset H^{1,1}(\widehat{V},\Bbb R)$ consisting of orbifold de~Rham classes which have such a positive representative is called the {\em K\"ahler cone} of $\widehat{V}$. We could not find the following lemma in the literature (although it should be known). \begin{lemma} Under the natural map $\operatorname{Pic}(\widehat{V})\to H^2_{\operatorname{DR}}(\widehat{V},\Bbb R)$ which assigns to a line bundle the corresponding orbifold de Rham class, the ample line bundles map to positive de Rham classes. \end{lemma} Note that the lemma is not as obvious as is the analogous lemma in the smooth case, since even if we are given a projective embedding $\widehat{V}\to\P^N$, the pullback of the Fubini-Study form (which establishes the positivity of the de Rham class of $\O_{\P^N}(1)$) is {\em not} necessarily positive as an orbifold $2$-form. \medskip \begin{pf*}{Sketch of proof} We modify an argument of Guillemin and Sternberg \cite{GS1}. By a result of Delzant \cite{delzant} and Audin \cite{audin}, the toric variety $\widehat{V}$ can be described as a symplectic reduction of the action of $G=\operatorname{Hom}(A_{n-1}(\widehat{V}),\Bbb{C}^*)$ on $\Bbb{C}^{{\Xi}}$; the specific symplectic reduction which produces $\widehat{V}$ is $\Phi^{-1}(\alpha)/G_{\Bbb R}$, where $\Phi$ is the moment map for the action, $G_{\Bbb R}$ is the maximal compact subgroup of $G$, and $\alpha\in\operatorname{cpl}(\Delta)$. An ample line bundle $L$ on $\widehat{V}$ corresponds to a character $\chi^L$ of $G$, since $A_{n-1}(\widehat{V})$ is the character lattice of $G$; moreover, the corresponding point in $A_{n-1}(\widehat{V})\otimes\Bbb R$ lies in $\operatorname{cpl}(\Delta)$. If we apply the constructions described on pp.~520--521 of \cite{GS1}, we produce a line bundle on $\widehat{V}$ with a specific (orbifold-)connection, whose curvature is the symplectic form obtained by symplectic reduction from the standard form on $\Bbb{C}^{{\Xi}}$. That curvature form provides the desired positive orbifold $2$-form. \end{pf*} The converse statement---that if the class of a line bundle is represented by a positive, orbifold-smooth $2$-form then the line bundle is ample---% is a theorem of Baily \cite{baily2}. Putting the two together, we conclude that the image of the cone $\operatorname{cpl}(\Delta)$ in $H^2_{\operatorname{DR}}(\widehat{V},\Bbb R)$ is precisely the closure of the K\"ahler cone $\overline{{\cal K}(\widehat{V})}$. \medskip As remarked earlier, the hypersurface $\widehat{X}$ is itself an orbifold, so it has an orbifold K\"ahler cone ${\cal K}(\widehat{X})\subset H^{1,1}(\widehat{X},\Bbb R)$. The positive orbifold-smooth K\"ahler forms on $\widehat{V}$ corresponding to classes in the interior of $\operatorname{cpl}(\Delta)$ will restrict to positive orbifold-smooth forms on $\widehat{X}$, since $\widehat{X}$ meets all singular strata of $\widehat{V}$ transversally. We call the resulting cone ${\cal K}_{\text{toric}}(\widehat{X}) \subset{\cal K}(\widehat{X})\cap H^{1,1}_{\text{toric}}(\widehat{X})$ the {\em cone of toric K\"ahler classes} on $\widehat{X}$. \section{The K\"ahler moduli space} Mirror symmetry predicts a close relationship between the moduli space of complex structures on one Calabi-Yau manifold $\widehat{Y}$ and the so-called ``K\"ahler moduli space'' of its mirror partner $\widehat{X}$. This K\"ahler moduli space, which arises in the study of nonlinear sigma models with target $\widehat{X}$,\footnote{The physics of these models is believed to be as well-behaved on orbifolds as on manifolds \cite{DHVW,DFMS}.} is an open subset of ${\cal D} /\Gamma$, where \[{\cal D} := \{B+i\,J\in H^2(\widehat{X},\Bbb{C})\ |\ J\in{\cal K}(\widehat{X})\},\] and where \[\Gamma:= H^2(\widehat{X},\Bbb Z)\rtimes \operatorname{Aut}(\widehat{X})\] (cf.~\cite{compact}.) The precise open subset of ${\cal D} /\Gamma$ which constitutes the moduli space is difficult to determine, since general convergence criteria for the sigma model are unknown at present.\footnote{But as we will observe below, the region of convergence in this particular case can be inferred from mirror symmetry.} However, the open set is expected to include points sufficiently far out along any path in ${\cal D} $ whose imaginary part is moving towards infinity while staying away from the boundary of the K\"ahler cone. Such paths should approach a common point, called the {\em large radius limit}, in an appropriate partial compactification of ${\cal D} /\Gamma$. A general discussion of conditions under which such a limit point exists can be found in \cite{compact}. We will be interested in a ``toric subspace'' of the K\"ahler moduli space, defined by intersecting the moduli space with ${\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}$, where \[{\cal D}_{\text{toric}}(\widehat{X}):= \{B+i\,J\in H^{1,1}_{\text{toric}} (\widehat{X})\ |\ J\in{\cal K}_{\text{toric}}(\widehat{X})\},\] and $\Gamma_{\text{toric}}:=A_{d-1}(\widehat{X})_{\text{toric}} \rtimes \operatorname{Aut}(\widehat{X})_{\text{toric}}$. (The toric automorphisms $\operatorname{Aut}(\widehat{X})_{\text{toric}}$ are those automorphisms of $\widehat{X}$ induced by an automorphism of the ambient toric variety $\widehat{V}$.) If the GIT\ of the family of hypersurfaces is well-behaved (which will be the case for the families of primary interest to us), then for the general hypersurface $\widehat{X}$ the group $\operatorname{Aut}(\widehat{X})_{\text{toric}}$ will be finite. Let $L=A_{d-1}(\widehat{X})_{\text{toric}}$ and consider the torus $L\otimes\Bbb{C}^*$ which contains ${\cal D}/L$ as an open subset. The cone ${\cal K}_{\text{toric}}\subset L\otimes\Bbb R$, which is a rational polyhedral cone, determines an affine torus embedding ${\cal M}$ with a unique $0$-dimensional orbit $p\in{\cal M}$ under the action of the torus $L\otimes\Bbb{C}^*$. We let $\overline{{\cal D}/L}$ be the closure of ${\cal D}/L$ in ${\cal M}$, and let $({\cal D}/L)^-$ be the interior of $\overline{{\cal D}/L}$. (This contains the point $p$.) If $\operatorname{Aut}(\widehat{X})_{\text{toric}}$ is finite, then since everything in sight is $\operatorname{Aut}(\widehat{X})_{\text{toric}}$-equivariant, we may take the quotient and get a partial compactification $({\cal D}/\Gamma)^-$ of ${\cal D}/\Gamma$ with a distinguished boundary point (again denoted by $p$), the {\em large radius limit} point.\footnote{This construction is identical to the one described in \cite{compact}, since Looijenga's semi-toric compactification \cite{Looijenga} coincides with the ``toroidal embeddings'' of \cite{AMRT} when ${\cal K}$ is rational polyhedral and $\Gamma/L$ is finite.} This point is the common limit of the paths described earlier. The torus $L\otimes\Bbb{C}^*=A_{d-1}(\widehat{X})_{\text{toric}}\otimes\Bbb{C}^*$ which is being compactified can be described in the form \begin{eqnarray*} A_{d-1}(\widehat{X})_{\text{toric}}\otimes\Bbb{C}^*&=& (\Bbb Z^{{\Xi}_0}/M)\otimes\Bbb{C}^*\\&=& (\Bbb{C}^*)^{{\Xi}_0}/(M\otimes\Bbb{C}^*) \end{eqnarray*} using equation \eqref{divisorsA}. Now the orbits of $M\otimes\Bbb{C}^*$ on $(\Bbb{C}^*)^{{\Xi}_0}$ are all good orbits of the same dimension. The action of $M\otimes\Bbb{C}^*$ on the larger space $\Bbb{C}^{{\Xi}_0}$ may be less well-behaved, but in any case we can regard $(\Bbb{C}^*)^{{\Xi}_0}/(M\otimes\Bbb{C}^*)$ as a representative of the birational class of quotients $\Bbb{C}^{{\Xi}_0}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}(M\otimes\Bbb{C}^*)$. Suppose that the family $\{\widehat{Y}\subset\widehat{V}^\circ\}$ associated to the polar polyhedron of $\{\widehat{X}\}$ has the dominance property (introduced in section 3), and that $\Delta$ is chosen so that $(\polyhedron^\circ\cap N)_0{-}\{0\}\subset \operatorname{gen}(\Delta(1)) \subset (\polyhedron^\circ\cap N){-}\{0\}$. Then we deduce from the monomial-divisor mirror map a diagram in which the vertical maps are dominant: \[\begin{array}{ccccc} {\cal D}_{\text{toric}}(\widehat{X})/A_{d-1}(\widehat{X})_{\text{toric}}&\subset& (\Bbb{C}^*)^{(\polyhedron^\circ\cap N)_0{-}\{0\}}/(M\otimes\Bbb{C}^*)&=& \Bbb{C}^{(\polyhedron^\circ\cap N)_0{-}\{0\}}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}(M\otimes\Bbb{C}^*)\\ \raise6pt\hbox{$\Big\downarrow$}&& \raise6pt\hbox{$\Big\downarrow$}&& \stackrel{\stackrel{\scriptscriptstyle|}{\scriptscriptstyle|}} {\scriptscriptstyle\downarrow}\\ {\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}&\subset& (A_{d-1}(\widehat{X})\otimes\Bbb{C}^*)/\operatorname{Aut}(\widehat{X})_{\text{toric}}&& \Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ). \end{array}\] The left and center vertical maps are simply the quotient maps by $\operatorname{Aut}(\widehat{X})_{\text{toric}}$, and the right map is the dominant rational map from the simplified moduli space to the actual moduli space of the family $\{\widehat{Y}\}$. We can now formulate a ``mirror symmetry'' conjecture for these families, which generalizes some (less precise) earlier conjectures of Aspinwall and L\"utken \cite{AL} and Batyrev \cite{batyrev2}. \begin{conjecture} Suppose that $\operatorname{Aut}(\widehat{X})_{\text{toric}}$ is finite, and that the dominance property holds for $\{\widehat{Y}\}$. Then \begin{enumerate} \item there is an open set $U\subset({\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}})^-$ containing the large radius limit point $p$ such that $U\cap({\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}})$ is the toric part of the K\"ahler moduli space, \item there is an appropriate quotient $\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)$ and a {\em ``mirror map''}\footnote{We denote this map by $\mu^{-1}$ in order to match the conventions established in \cite{compact}.} \[\mu^{-1}:U\to \Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ),\] which is an isomorphism onto its image, and which, when restricted to $U\cap({\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}})$, serves to identify points whose conformal field theories are mirror-isomorphic, such that \item the differential of the inverse map $\mu$ at the ``large complex structure limit point'' $\mu^{-1}(p)$ \[d\mu: T_{\mu^{-1}(p),\,\,\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)} \to T_{p,\,\,U} \] coincides with the monomial-divisor mirror map \[ H^{d-1,1}_{\text{poly}}(\widehat{Y}) \to H^{1,1}_{\text{toric}}(\widehat{X}) \] up to signs, once we have made the canonical identifications \begin{eqnarray*} T_{\mu^{-1}(p),\,\,\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)}&=& H^{d-1,1}_{\text{poly}}(\widehat{Y})\\ T_{p,\,\,U}&=&H^{1,1}_{\text{toric}}(\widehat{X}). \end{eqnarray*} That is, there is some element \[ \theta_{\Delta} \in A_{d-1}(\widehat{X})_{\text{toric}} \otimes \Bbb{C}^* \subset \operatorname{Aut} ( H^{1,1}_{\text{toric}} (\widehat{X}) ) \] of order $2$, which when composed with the monomial-divisor mirror map yields $d\mu$. (When $d \ge 3$, the automorphism $\theta_{\Delta}$ which specifies the signs is unique.) \end{enumerate} In particular, the location of the large complex structure limit point $\mu^{-1}(p)$ can be calculated using the monomial-divisor mirror map and the knowledge of the cone ${\cal K}_{\text{toric}}(\widehat{X})$. \end{conjecture} In \cite{compact}, a very general conjecture was formulated which specifies the ``canonical coordinates'' to be used in the mirror map, up to some constants of integration. Those constants can be determined if one knows the differential of the mirror map at the large radius limit point---% for these toric hypersurfaces, that differential is supplied by the conjecture above. So the two conjectures together completely specify the canonical coordinates. A similar conjecture about canonical coordinates for toric hypersurfaces has been independently made by Batyrev and van Straten \cite{BvS}. If the conjecture stated above is true, then among other things the so-called ``$3$-point functions'' (part of the conformal field theories) must coincide under this mapping. The $3$-point function on the moduli space $\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)$ can be calculated in terms of the variation of Hodge structure of the family $\{\widehat{Y}\}$ (cf.~\cite{guide}), and this gives a method to identify precisely which subset of ${\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}$ constitutes the toric part of the K\"ahler moduli space. For that subset can be characterized as the domain of convergence of a power series expansion of the $3$-point functions, when calculated in the canonical coordinates on ${\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}$. Using the mirror map, this power series calculation can actually be made in $\Bbb{C}^{\polyhedron^\circ\cap N}\mathrel{\mskip-4.5mu/\!/\mskip-4.5mu}\mathop{\widetilde{\rm Aut}}\nolimits(\widehat{V}^\circ)$, by calculating the variation of Hodge structure of the family $\{\widehat{Y}\}$. The leading term in the power series expansion of the $3$-point function on ${\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}$ is the cubic form \[H^{1,1}_{\text{toric}}(\widehat{X})\times H^{1,1}_{\text{toric}}(\widehat{X})\times H^{1,1}_{\text{toric}}(\widehat{X})\to\Bbb{C}\] given by the cup product; higher terms are given by ``quantum corrections'' that depend on the point in ${\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}}$, and that all vanish at the large radius limit point $p\in({\cal D}_{\text{toric}}(\widehat{X})/\Gamma_{\text{toric}})^-$. One consequence of our conjecture would therefore be an agreement between the leading term in the variation of Hodge structure calculation for the family $\{\widehat{Y}\}$ near the large complex structure limit point, and the cup product cubic form on $\widehat{X}$. This consequence was first checked in an example by Aspinwall, L\"utken, and Ross \cite{ALR} several years ago.\footnote{It is possible to verify that the large complex structure limit point used in \cite{ALR} agrees with the one predicted by the monomial-divisor mirror map.} More recently, Batyrev \cite{batyrev2} checked his version of this statement in the case that $X$ itself is smooth, and the authors \cite{AGM} checked it in an example in which there are five different birational choices for $\widehat{X}$ (with the same $X\subset V$). After learning of the results of \cite{AGM} and of the present paper, Batyrev \cite{batyrev3} checked this consequence in general. The $3$-point function coming from variation of Hodge structure can be used to determine the choice of signs $\theta_{\Delta}$: the poles in that function should occur at {\em positive}\/ real values of the canonical coordinates (cf.~\cite{AGMiii}). The automorphism $\theta_{\Delta}$ with this property can be calculated explicitly using methods of Gel'fand, Zelevinski\v\i, and Kapranov \cite{GKZ}; we will discuss this in detail elsewhere. Another consequence of our conjecture is that the K\"ahler moduli spaces for different birational models $\widehat{X}$ of the function field of $X$ can naturally be regarded as analytic continuations of one another. (For after applying mirror symmetry, they are seen to occupy different regions in the same moduli space.) This was the principal conclusion of our earlier paper \cite{AGM}; a similar idea is due independently to Manin \cite{manin}. \section{Phases and the secondary fan} In the course of defining the monomial-divisor mirror map, we made a somewhat unnatural restriction to a Zariski-open subset of the simplified hypersurface moduli space. We now return to the study of the full moduli space. The ``simplified hypersurface moduli space'' will be birational to the quotient $(\Bbb{C}^*){}^{(P\cap M)_0}/\widetilde{T}$. In fact, the moduli space of primary interest is the space parameterizing those hypersurfaces whose singularities are no worse than generic. This is the complement of the ``principal discriminant'' of Gel'fand, Zelevinski\v\i, and Kapranov \cite{GKZ}. One natural compactification of the moduli space would be the one in which this ``principal discriminant'' is an ample divisor. Whatever compactification we use, the compactified moduli space is itself a toric variety (since it contains the torus $(\Bbb{C}^*){}^{(P\cap M)_0}/\widetilde{T}$ as a dense open subset). If we compactify so that the principal discriminant is ample, then the toric variety is determined by the Newton polyhedron for the principal discriminant which, as Gel'fand, Zelevinski\v\i, and Kapranov show, has a convenient combinatorial description as a so-called ``secondary polytope''. To explain the combinatorics, we note that the action of $\widetilde{T}$ on $(\Bbb{C}^*){}^{(P\cap M)_0}$ is induced by a homomorphism $\operatorname{ad}^+_{(P\cap M)_0}:N^+\to\Bbb Z^{(P\cap M)_0}$, dual to the map $\Bbb Z\langle{(P\cap M)_0}\rangle\to M^+$ given by equation \eqref{twomaps}, where $M^+=M\oplus\Bbb Z$. We should imagine embedding the set ${(P\cap M)_0}$ into $M^+$ via the map $b\mapsto (b,1)$; the image is a finite set of points in the affine hyperplane $\{(m,1)\}\subset M^+$. The convex cone spanned by these points we denote by $P^+$; it is the cone over the image of the original polyhedron $P$ (generated by the points ${(P\cap M)_0}$) in the affine hyperplane $\{(m,1)\}\subset M^+$. The dual cone to $P^+$ has the form $(\polyhedron^\circ)^+$, where $\polyhedron^\circ$ is the polar polyhedron of $P$. The {\em secondary fan} (which is the normal fan of the secondary polytope) is the fan consisting of all cones $\operatorname{cpl}(\Sigma)$, where $\Sigma$ is a regular refinement of the fan ${\cal N}(\,(\polyhedron^\circ)^+)$ (cf.~\cite{BFS}). Among the possible regular refinements $\Sigma$ we find fans of the form \[(\fan^\circ)^+:= \{\text{cone over }(\sigma\capP)\ |\ \sigma\in\fan^\circ\},\] for regular fans $\fan^\circ$ which refine ${\cal N}(\polyhedron^\circ)$. (But there are others, which do not have this form.) For such fans, it is easy to see that $\operatorname{cpl}(\,(\fan^\circ)^+)=\operatorname{cpl}(\fan^\circ)$, regarding both as cones in the same space $\Bbb Z^{{(P\cap M)_0}}/N^+\cong\Bbb Z^{{(P\cap M)_0}-\{0\}}/N$. So our chosen compactification of the moduli space is the toric variety which is specified by the secondary fan. It has the pleasant property that it includes all of the partial compactifications that were needed to describe the ``large complex structure limits'' coming from mirror symmetry of sigma models (since those were given by the cones $\operatorname{cpl}(\fan^\circ)$). It has another nice property as well, first observed by Kapranov et al.\ \cite{KSZ} and Batyrev \cite{batyrev2}: the compactification constructed in this way dominates all possible GIT\ compactifications, coming from different choices of linearization. What does this structure correspond to under mirror symmetry? The embedding $(\polyhedron^\circ\cap N)_0\subset N$, together with a (regular) refinement $\Delta$ of the fan ${\cal N}(P)$, was used to determine the projective toric variety $\widehat{V}$. The new extended embedding $(\polyhedron^\circ\cap N)_0\subset N^+$ (whose image lies in the affine hyperplane $\{(n,1)\}\subset N^+$), together with a regular refinement $\Sigma$ of the fan ${\cal N}(P^+)$, can also be used to determine a toric variety, of dimension one larger than the previous variety. Among these toric varieties we find the total spaces of canonical bundles over the various choices of $\widehat{V}$ (when we take $\Sigma$ of the form $\Delta^+$). This is precisely the structure that Witten has found to be relevant in his study of Landau-Ginzburg theories and their deformations \cite{phases}. Each choice of fan $\Sigma$ determines a different ``phase'' of the physical theory. When the fan $\Sigma$ is of the form $\Delta^+$, the physical theory is related to the nonlinear sigma model with target $\widehat{X}$ (where $\widehat{X}\subset\widehat{V}$ is generic). But for other fans $\Sigma$, the physical theory is quite different. (We refer the reader to \cite{phases} for more details.) Thus, not only can the different sigma models with birationally equivalent targets be viewed as analytic continuations of one another, there are further analytic continuations (to regions in the moduli space corresponding to $\operatorname{cpl}(\Sigma)$ for $\Sigma\ne\Delta^+$) to other kinds of physical theory. \ifx\undefined\leavevmode\hbox to3em{\hrulefill}\, \newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,} \fi
1993-09-02T13:47:17
9309
alg-geom/9309002
en
https://arxiv.org/abs/alg-geom/9309002
[ "alg-geom", "math.AG" ]
alg-geom/9309002
Kirti
Kirti Joshi (School of Mathematics, Tata Institute of Fundamental Research, Bombay)
A family of \'etale coverings of the affine line
The file is in standard LateX, eight pages
null
null
null
null
In the note we construct a family of \'etale coverings of the affine line. More specifically, let $F$ be a finite field of characteristic $p$ and suppose that the cardinality of $F$ is at least 4. Let $A = F[T]$ be the polynomial ring in one variable $T$, $K=F(T)$. Let $K_\infty$ be the completion of $K$ along the valuation given by $1/T$, and let $C$ be the completion of the algebraic closure of $K_\infty$. We prove in this note that there is a continous surjection $$\pi_1^{alg}(\A^1_C) \to \lim_{\leftarrow \atop I} SL_2( A/I )/{\pm 1},$$ where $\pi_1^{alg}(\A^1_C)$ is the algebraic fundamental group of the affine line over $C$, and the inverse limit on the right (above) is taken over all nonzero proper ideals in $A=F[T]$. We use the theory of Drinfel'd modular curves to obtain these coverings.
[ { "version": "v1", "created": "Thu, 2 Sep 1993 15:17:00 GMT" } ]
2015-06-30T00:00:00
[ [ "Joshi", "Kirti", "", "School of Mathematics, Tata Institute of Fundamental\n Research, Bombay" ] ]
alg-geom
\section{Introduction} This note was inspired by a colloquium talk given by S.~S.~Abhyankar at the Tata Institute\footnote{During his visit in the month of December, 1992}, on the work of Abhyankar, Popp and Seiler (see \cite{Popp}). It was pointed out in this talk that classical modular curves can be used to construct (by specialization) coverings of the affine line in positive characteristic. In this ``modular'' optic it seemed natural to consider Drinfel'd modular curves for constructing coverings of the affine line. This note is a direct outgrowth of this idea. While it is trivial to see that affine line in characteristic zero has no non-trivial \'etale coverings, in \cite{Abhyankar} it was shown that the situation in positive characteristic is radically different and far more interesting. Let us, for the sake of definiteness, work over a field $k$ of characteristic $p>0$. In \cite{Abhyankar}, Abhyankar conjectured that any finite group whose order is divisble by $p$ and which is generated by its $p$-Sylow subgroups (such a finite group is sometimes called a ``quasi-$p$-group''), occurs as quotient of the algebraic fundamental group of the affine line. It is customary to write $\pi_1^{\rm alg}({\bf A}^1_k)$ to denote the algebraic fundamental group of the affine line. While Abhyankar's conjecture indicates that the algebraic fundamental group of the affine line is quite complicated, our result perhaps illustrates its cyclopean proportions. The result we prove (see Theorem~\ref{main theorem} below) is the analogue of the following well-known classical result, which falls out of the theory of elliptic modular curves over the field of complex numbers: there is a continous quotient $\pi_1^{\rm alg}({\bf P}^1_{{\bf C}}-\{0,1728,\infty\}) \surjects \prod_{p} SL_2({\bf Z}_p)/\{\pm 1\}$, where ${\bf Z}_p$ denotes the ring of $p$-adic integers. We would like to thank N. Mohan Kumar for numerous suggestions and conversations; he also explained to us a variant of Abhyankar's Lemma, which is crucial to our argument. We would also like to thank E.-U.~Gekeler, Dipendra Prasad for electronic correspondence while this note was being written -- their remarks and suggestions have been extremely useful; in particular we would like to point out that the conjecture stated at the end of this note was formulated with the help of Dipendra Prasad. We would also like to thank M.~V.~Nori for useful comments and Dinesh Thakur for encouragement. \section{Resum\'e of Drinfel'd modules and their moduli} In this section we recall a few of the standard facts about Drinfel'd modules. Since the basic theory of Drinfel'd modules and their moduli is well documented we will be brief; all the facts which we will need can be found in the following standard references: \cite{Drinfel'd}, \cite{Deligne-Husemoller}, \cite{Gekeler2}. Since we do not need the full strength of Drinfel'd's work, we will work with a very special situation which is required for our purpose. In this section, we will outline this special situation. Let us fix some notations. Let ${\bf F}_q$ denote a finite field with $q=p^m$ elements and of characteristic $p$. We will write $A = {\bf F}_q[ T ]$, $K = {\bf F}_q(T)$. Further denote by $K_\infty$, the completion of $K$ along the valuation corresponding to $1/T$. Denote by $C$ the completion of the algebraic closure of $K_\infty$. The field $C$ is thus a ``universal domain'' of characteristic $p$. There is a natural inclusion $K \into C$. Let ${\cal G}_a$ denote the additive group scheme. Then it is easy to see that the ring of endomorphisms of ${\cal G}_a$ defined over $C$, denoted by $\mathop{\rm End}\nolimits_C({\cal G}_a)$, is a noncommutative ring generated by the Frobenius endomorphism. More precisely, for an indeterminate $\tau$, consider the noncommutative ring $C\{\tau_p\}$ of all polynomials in $\tau$ with coefficients in $C$, and where the multiplication rule is given by $\tau_p a = a^p \tau_p$ for all $a\in C$. Then one checks that $\mathop{\rm End}\nolimits_C({\cal G}_a) \isom C\{\tau_p\}$. Observe that this isomorphism gives rise to a ring homomorphism $\partial:\mathop{\rm End}\nolimits_C({\cal G}_a) \to C$ which sends an endomorphism of ${\cal G}_a$ to the constant term of the polynomial in $\tau$ associated to it. A {\em Drinfel'd module} $\phi$ over $C$ is a ring homomorphism $\phi:A \to \mathop{\rm End}\nolimits_C({\cal G}_a)$ such that composite map $\partial \phi $ is the natural inclusion of $A\to C$. It is easy to check that any such map factors through the subring $C\{\tau_p^m\}\subset C\{\tau_p\}$. For simplicity of notation, we will write $\tau=\tau_p^m$. Thus for any $a\in A$ we have an endomorphism $\phi_a$ of ${\cal G}_a$ defined over $C$. Moreover, note that, any such $\phi$ is ${\bf F}_q$ linear. So as $A$ is generated as an ${\bf F}_q$-algebra by $T$, giving a $\phi$ thus amounts to specifying a single endomorphism of ${\cal G}_a$ corresponding to $T\in A$, $\phi_T = \sum_{i=0}^r a_i \tau^i$, with $a_0 =T$. Then this can be extended to all of $A$. The $\tau$-degree of $\phi_T$ is called the {\em rank} of the Drinfel'd module $\phi$. If $\phi,\phi'$ are two Drinfel'd modules then a {\em morphism of Drinfel'd modules} is an endomorphism $u\in\mathop{\rm End}\nolimits_C({\cal G}_a)$ such that for all $a\in A$ we have $\phi_a u = u \phi_a' $. It is easy to see that any such $u$ is in fact contained in $C\{\tau\}$. For any $a\in A$ a Drinfeld module $\phi$ specifies a closed subgroup-scheme of ${\cal G}_a$: the kernel of the endomorphism $\phi_a$, $\ker(\phi_a)$. For example, if $a=T$, then the kernel of $\phi_T$ is simply the roots of the polynomial $T+\sum_{i=1}^r a_i X^{q^i-1} =0$ together with $0$. Moreover, for any ideal $I\subset A$, we can define a subgroup scheme of ${\cal G}_a$ using the ring structure. One checks that $\ker(\phi_I)(C)$ is a free $(A/I)$-module of rank $r$, where $r$ is the rank of $\phi$. This lets us define a notion of an $I$-level structure on $\phi$. Drinfel'd has shown (see \cite{Drinfel'd}) that there is a moduli of Drinfel'd modules with level structure (in general we have only a ``coarse moduli scheme''). Our main interest is the case of rank two Drinfel'd modules. And henceforth, we shall work with Drinfel'd modules of rank two. The theory of Drinfel'd modules of rank two behaves like the theory of elliptic curves. The fact that the moduli of Drinfel'd modules of rank two over $C$, is a smooth affine curve over $C$ is a very special case of a fundamental result of Drinfel'd (see \cite{Drinfel'd}). We need several facts about these Drinfel'd modular curves. The first fact we need is the following: \begin{thm} The ``coarse moduli'' of rank two Drinfel'd modules over $C$ is the affine line over $C$. \end{thm} \begin{proof} For a proof see \cite{Goss} paragraph 1.32, or \cite{Gekeler1}. \end{proof} Thus the above situation is analogous to the classical situation for elliptic curves. In fact the identification with the line is given by a ``j-invariant''. If $\phi_T = T+ a\tau + b \tau^2$, is a rank 2 module then $b\neq 0$ and then its $j$-invariant is $a^{q+1}/b$. Before we need some notations. For any ${\bf F}_q$ algebra $R$, let $$G_1(R) = \left\{g\in GL_2(R)\big| \det(g)\in {\bf F}_q^* \right\},$$ $$ G(R) = G_1(R)/Z({\bf F}_q),$$ where $Z({\bf F}_q)$ is the group of ${\bf F}_q$-valued scalar matrices with nonzero determinant. The following result is really the heart of our construction. \begin{thm}\label{Drinfel'd-Gekeler} For every non-zero ideal $I\subset A$, there is a ``coarse moduli'' of Drinfel'd modules of rank two with full $I$-level structure exists and is an affine curve over $C$. There is a ``forget the level structure'' morphism to ${\bf A}^1_C$. This map is branched over $0\in {\bf A}^1_C$. The covering is Galois and the Galois group is $G(A/I)$. The ramification index of any point lying over $0$ is $q+1$ and is independent of $I$. In particular, the ramification is tame. \end{thm} \begin{proof} As mentioned earlier, the existence of the moduli is due to Drinfel'd (see \cite{Drinfel'd}). These curves have been studied in great detail by Goss and Gekeler. In the our case the Galois group can easily be calculated, a convenient reference for it is \cite{Gekeler2}, for instance see Lemma 1.4 on page 79, also see the first section of \cite{Gekeler1}. The ramification information is computed in \cite{Goss}, Lemma 4.2. One also finds it computed in \cite{Gekeler2}, on page 87. As in the classical situation, the ramification takes place over the Drinfel'd module with extra automorphisms. From the definitions, it is clear that an automorphism of a Drinfel'd module of rank two over $C$ is firstly an automorphism $u$ of ${\cal G}_a$, defined over $C$. Clearly any such automorphism must be an invertible element of $C$. Then a simple calculation shows that if a nonzero element of $C$ is an automorphism of a Drinfel'd module then it must be a root of unity. One checks that with the exception of the module with $j$-invariant equal to zero, the automorphism group of the Drinfel'd module is ${\bf F}_q^*$. The module with $j$-invariant equal to zero has automorphism group $F_{q^2}^*$. Thus in particular, the ramification is tame. These facts are easily proved by explicit calculations. \end{proof} \section{The main theorem} We are now ready to state and prove our main theorem. One should note that most of the work has already gone in the construction and analysis the of the moduli of Drinfel'd modules. \begin{thm}\label{main theorem} If $p=2,3$ then assume that $q=p^m, m\geq 2$. There is a family of \'etale coverings, $Y_I$ of ${\bf A}^1_C$, indexed by the nonzero proper ideals $I\subset A$. The curves $Y_I$ are affine curves over $C$ and the covering $Y_I \to {\bf A}^1_C$ is Galois with Galois group $SL_2(A/I)/\{\pm1\}$. Moreover, these coverings form an inverse system indexed by $I$. Thus in particular we have a continous quotient $$\pi_1^{\rm alg}( {\bf A}^1_C) \to \lim_{\longleftarrow \atop I} \left( SL_2( A/I )/\{\pm 1\} \right).$$ \end{thm} \begin{proof} By Theorem~\ref{Drinfel'd-Gekeler}, we have a tamely ramified covering of the affine line which is branched over one point. Also note that the ramification index of any point over $0\in {\bf A}^1_C$ is $q+1$, independent of the ideal $I\subset A$. Now we apply a suitable variant of Abhyankar's Lemma to remove the tame ramification. The crucial thing is to ensure, if possible, that the ``pull back'' coverings remain irreducible. The following variant of Abhyankar's Lemma (see Lemma~\ref{Mohan's lemma}) which was pointed out to me by N. Mohan Kumar, gives an explicit criterion to check irreducibility, then we apply this criterion to the case at hand. This is an easy exercise in elementary group theory. We have stated all the necessary results as a sequence lemmas, and since the proofs of all the individual statements are easy, we leave the details to the reader. \end{proof} \begin{lemma}\label{Mohan's lemma} Let $k$ be an algebraically closed field of characteristic $p$. Let $X \to {\bf A}^1_k$ be a finite Galois cover defined over $k$, with Galois group $G$. Further assume that the cover is branched over $0\in {\bf A}^1_k$, and any point lying over it is tamely ramified with ramification index $n$. Let ${\bf A}^1_k \to {\bf A}^1_k$ be a ${\bf Z}/n$ covering ramified completely at $0$, and unramified elsewhere. Let $X'$ be the normalization of the fibre product $X \times_{{\bf A}^1_k} {\bf A}^1_k$. Suppose that there are no nontrivial homomorphisms $G \to {\bf Z}/n$. Then $X'$ is irreducible, and the Galois group of the covering $X' \to {\bf A}^1$ is $G$. \end{lemma} \begin{proof} Clearly one is reduced to proving the following field theory statement: Let $L/k(t)$ be a finite Galois extension with Galois group $G$. Let $E=k(t^{1/n})$. Then $L/k(t)$ and $E/k(t)$ are linearly disjoint over $k(t)$ if and only if there are no nontrivial homomorphisms $G \to {\bf Z}/n$. And the above statement is immediate from the fact that $E/k(t)$ is a Kummer extension. This finishes the proof. \end{proof} We now need some elementary group theoretic lemmas to apply the above criterion. Since the proofs are easy we will state the lemmas without proofs. \begin{lemma}\label{lemma1} Let $\wp \neq 0$ is a prime ideal of $A$. Then for all $k \geq 1$ and for all $n\geq 2$, the natural morphism $$GL_k(A/\wp^n) \to GL_k( A/\wp^{n-1})$$ is surjective. \end{lemma} Recall that for any ${\bf F}_q$ algebra $R$, we had defined the groups $$G_1(R) = \left\{g\in GL_2(R)\big| \det(g)\in {\bf F}_q^* \right\},$$ $$ G(R) = G_1(R)/Z({\bf F}_q),$$ where $Z({\bf F}_q)$ is the group of ${\bf F}_q$-valued scalar matrices with nonzero determinant. Further, for any prime ideal $\wp\neq 0$ of $A$, write $G_1^n = G_1(A/\wp^n)$, and $G^n = G(A/\wp^n)$, for all $n\geq 1$. The results which follow are valid for any non-zero prime ideal $\wp$, so we have supressed $\wp$ in our notations. \begin{lemma}\label{lemma2} The natural map $G_1^n \to G_1^{n-1}$ is surjective for all $n\geq 2$. \end{lemma} \begin{lemma}\label{lemma3} The natural map $G^n \to G^{n-1}$ is surjective for all $n\geq 2$. \end{lemma} Denote by $G^{(n,n-1)}$ the kernel of the map $G^n \to G^{n-1}$, similarly write $G_1^{(n,n-1)}$, for the kernel of the corresponding map for $G_1$. \begin{lemma}\label{lemma4} Let $q=2^m, m\geq 2$. Let $F/{\bf F}_q$ be any finite extension. Then there are no nontrivial morphisms $G(F) \to {\bf Z}/(q+1)$. \end{lemma} \begin{lemma} \label{lemma5} Let $q=2^m,m\geq 2$. Then for any $n\geq 1$ there are no nontrivial maps $G^n \to {\bf Z}/(q+1)$. \end{lemma} \begin{proof} The proof is by induction on $n$. For $n=1$, we are done by the previous lemma. Now show that the kernel of any map $G^n \to {\bf Z}/(q+1)$ contains $G^{(n,n-1)}$. Then we are done by induction. \end{proof} \begin{lemma}\label{lemma6} Let $q = p^m, p\neq 2$. If $p=3$ then $m\geq 2$. Then there is a canonical morphism $G^n \to {\bf F}_q^*/{\bf F}_q^{*2}$. In particular we have a natural map $G^n \to {\bf Z}/2$, obtained by identifying ${\bf F}_q^*/{\bf F}_q^{*2}$ with ${\bf Z}/2$. \end{lemma} \begin{lemma}\label{lemma7} Let $q=p^m, p\neq 2$. If $p=3$ then $m\geq2$. Let $F/{\bf F}_q$ be any finite extension. Then any morphism $G(F)\to {\bf Z}/(q+1)$ factors through the canonical morphism given by the above lemma, followed by the inclusion of ${\bf Z}/2\to {\bf Z}/(q+1)$. \end{lemma} \begin{lemma}\label{lemma8} Let $q = p^m, p\neq 2$. If $p=3, m\geq 2$. Then any nontrivial morphism $G^n \to {\bf Z}/(q+1)$ factors through the canonical morphism. \end{lemma} \begin{proof} This is again proved by induction on $n$. As before, one checks that any such map is trivial on $G^{(n,n-1)}$. \end{proof} Now we can identify our Galois groups. \begin{lemma}\label{lemma9} Let $q=2^m, m \geq 2$. For any non zero prime ideal $\wp\subset A$, we have: $$G^n = SL_2(A/\wp^n).$$ \end{lemma} \begin{lemma}\label{lemma10} Let $q=p^m, p\neq 2$, if $p=3$ then $m\geq 2$. let $\wp$ be any non zero prime ideal in $A$. For any $n\geq 1$, let $$\tilde{G^n} = ker(G^n \to {\bf Z}/2).$$ Then $\tilde{G^n} = SL_2(A/\wp^n)/\{\pm 1\}$. \end{lemma} Thus we can now prove our main theorem. If $p=2$, then the pull back coverings remain irreducible and the Galois group is $G^n$. If $p$ is odd, then there is a quadratic subfield in common. And the Galois group is $\tilde{G^n}$. Then we are done by the above lemmas. After the result of Madhav Nori (see \cite{Madhav}) and the one proved above, we would like to advance the following conjecture: \begin{conj} Let $G/K$ be any isotropic, semisimple, simply-connected algebraic group over $K={\bf F}_q(T)$, with center $Z$. Let ${\bf A}^{{\rm fin}}_K$ denote the finite adeles of $K$. Then any maximal compact subgroup of $G({\bf A}^{{\rm fin}}_K)/Z({\bf F}_q)$ occurs as a continous quotient of the fundamental group of the affine line over $C$. \end{conj}
1997-07-01T06:56:43
9707
alg-geom/9707002
en
https://arxiv.org/abs/alg-geom/9707002
[ "alg-geom", "math.AG" ]
alg-geom/9707002
Aaron Bertram
Aaron Bertram (University of Utah)
Stable pairs and log flips
17 pages, LaTeX2e. Will appear in the Proceedings of the AMS Santa Cruz conference (1995)
null
null
null
null
This paper has two parts. In the first part, we review stable pairs and triples on curves, leading up to Thaddeus' diagram of flips and contractions starting from the blow-up of projective space along a curve embedded by a complete linear series of the form K + ample. In the second part, we identify log canonical divisors which exhibit Thaddeus' flips and contractions as "log" flips and contractions in the sense of the log-minimal-model program.
[ { "version": "v1", "created": "Tue, 1 Jul 1997 04:56:34 GMT" } ]
2008-02-03T00:00:00
[ [ "Bertram", "Aaron", "", "University of Utah" ] ]
alg-geom
\section{Stable Bradlow Pairs.} Let ${\mathcal A}$ be a category with a zero object in which kernels and cokernels exist, as well as direct sums. Let $S \subset Ob({\mathcal A})$ be a subset which is closed under direct sums. \begin{definition} A function $\mu: S \rightarrow {{\bf R}}$ is called a {\it slope function} if for all short exact sequences $0 \rightarrow B \rightarrow A \rightarrow C \rightarrow 0$ of elements of $S$, $$\mu(B) < \mu(A) \Leftrightarrow \mu(A) < \mu(C)\ \mbox{and}\ \mu(B) = \mu(A) \Leftrightarrow \mu(A) = \mu(C)$$ \end{definition} Given a slope function $\mu:S \rightarrow {{\bf R}}$: \begin{definition} An object $A\in S$ is called {\it stable} if $\mu(B) < \mu(A)$ whenever $B\in S$ and there exists an injection $B \hookrightarrow A$ other than the identity. $A$ is called semistable if $\mu(B) \le \mu(A)$ above, strictly semistable if it is semistable but not stable, and unstable if it is not semistable. \end{definition} \begin{definition} (a) If $A\in S$ is unstable, then a filtration: $$0 \hookrightarrow A_1 \hookrightarrow A_2 \hookrightarrow ... \hookrightarrow A_n = A$$ by elements of $S$ is called a Harder-Narasimhan filtration if the $A_i/A_{i-1}$ are all in $S$ and semistable and $\mu(A_1) > \mu(A_2/A_1) > ... > \mu(A_n/A_{n-1})$. \medskip (b) If $A\in S$ is semistable, then a filtration $$0 \hookrightarrow A_1 \hookrightarrow A_2 \hookrightarrow ... \hookrightarrow A_n = A$$ is called a Jordan-H\"older filtration if the $A_i/A_{i-1}$ are all in $S$ and stable. Given a Jordan-H\"older filtration of $A$, the object $gr(A) := \oplus_{i=1}^nA_i/A_{i-1} \in S$ is called the associated graded of the filtration. Two Jordan-H\"older filtrations are called {\it s-equivalent} if their associated graded objects are isomorphic. \end{definition} \begin{example}{(Vector Bundles on $C$):} $\bullet$ ${\mathcal A}$ is the category of isomorphism classes of vector bundles on $C$. $\bullet$ $S = Ob({\mathcal A}) - \{0\}$. $\bullet$ $\mu:S \rightarrow {{\bf Q}}$ is the usual slope function $\mu(E) = \ \mbox{deg}(E)/\mbox{rk}(E)$. Then: (a) Harder-Narasimhan filtrations always exist (and up to isomorphism only depend upon the isomorphism class of $E$). (b) Jordan-H\"older filtrations always exist, producing associated gradeds which only depend upon the isomorphism class of $E$. In particular, $s$-equivalence becomes an equivalence relation on isomorphism classes of semistable bundles. (c) For fixed invariants $r$ (the rank) and either $d$ (the degree) or ${\mathcal O}_C(D)$ (the isomorphism class of the determinant), there are projective coarse moduli spaces $M_C(r,d)$ (respectively, $M_C(r,D)$) for the functors ``families of semistable vector bundles modulo s-equivalence with the given invariants''. (See \cite{S} for details.) \end{example} The next example is due to Bradlow and Garcia-Prada (\cite{BG}). \begin{example}{(Triples on $C$):} $\bullet$ ${\mathcal A}$ is the category of isomorphism classes of triples $(E,f,F)$, where $E,F$ are vector bundles on $C$ and $f:E \rightarrow F$ is a homomorphism. A triple is called {\it nontrivial} if $F\ne 0$ and {\it nondegenerate} if $f$ has maximal rank at some point. A morphism in this category is a pair $(\alpha,\beta)$ consisting of morphisms $\alpha:E\rightarrow E'$ and $\beta:F \rightarrow F'$ in the category of vector bundles, such that the following diagram commutes: $$\begin{array}{ccc}E & \stackrel\alpha\rightarrow & E' \\ f\downarrow & & f'\downarrow \\ F & \stackrel \beta\rightarrow & F' \end{array}$$ Both $\alpha$ and $\beta$ need to be injective to make $(\alpha,\beta): (E,f,F) \rightarrow (E',f',F')$ injective as a morphism of triples. Direct sums obviously exist as the ``free'' sum of triples: $(E,f,F) \oplus (E',f',F') = (E\oplus E',f\oplus f',F\oplus F')$. $\bullet$ $S$ is the set of nontrivial triples. $\bullet$ For each $\sigma \in {{\bf R}}$ and $(E,f,F) \in S$, let: $$\mu_\sigma(E,f,F) = \frac{\mbox{deg}(E) + \ \mbox{deg}(F) + \sigma (\mbox{rk}(E) + \ \mbox{rk}(F))}{\mbox{rk}(F)}.$$ Note the asymmetry in the slope function! One says $(E,f,F)$ is $\sigma$-stable if it is stable with respect to the slope function $\mu_\sigma$. \end{example} \begin{theorem}{(\cite{BG} Theorem 6.1)} For fixed invariants: $$r_1 = \ \mbox{rk}(E), r_2 = \ \mbox{rk}(F), d_1 = \ \mbox{deg}(E), d_2 = \ \mbox{deg}(F) \ \mbox{and}\ \sigma \in {{\bf R}}$$ a coarse moduli space exists for the functor ``families of non-degenerate $\sigma$-stable triples with the given invariants'', which is moreover projective if $r_1 + r_2$ is relatively prime to $d_1 + d_2$ and $\sigma$ is ``generic'' (see \cite{BG}). \end{theorem} \noindent {\bf Explanation of the Parameter:} The idea is to relate stable triples $(E,f,F)$ on $C$ to (equivariantly) stable equivariant bundles $G$ on $C\times {\bf P}^1$. (The action is the automorphism group of ${\bf P}^1$ acting on the second factor.) This is a consequence of K\"unneth, which gives an isomorphism $$\mbox{Hom}_{{\mathcal O}_C}(E,F) \cong \ \mbox{Ext}^1_{{\mathcal O}_{C\times {\bf P}^1}} (p^*E\otimes q^*{\mathcal O}_{{\bf P}^1}(2),p^*F),$$ telling us to look for $G$ in the corresponding extension: $$0 \rightarrow p^*F \rightarrow G \rightarrow p^*E \otimes q^*{\mathcal O}_{{\bf P}^1}(2) \rightarrow 0.$$ (this technique is often called ``dimensional reduction''.) The main point is now that ample line bundles on $C\times {\bf P}^1$ are of the form $p^*L\otimes q^*M$, and (an equivariant version of) Gieseker stability for bundles on $C\times {\bf P}^1$ depends upon a parameter, namely the {\bf ratio} deg$(L)$/deg$(M)$. Stability of $G$ with respect to a given ratio translates into $\sigma$-stability for triples for a fixed value of $\sigma$. \bigskip We are most concerned with the following, first considered by Bradlow in \cite{Br}. \begin{example}{(Pairs on $C$):} $\bullet$ Restrict the category of triples to the objects: $({\mathcal O}_C,f,E)$ and $(0,0,E)$. These objects we will call pairs, following the literature. This full subcategory is closed under kernels and cokernels. It is not closed under arbitrary direct sums, but if a set of pairs is given, with the property that at most one of them is of the form $({\mathcal O}_C,f,E)$, then their direct sum does lie in the subcategory. (This will be enough to construct associated gradeds for Jordan-H\"older filtrations!) $\bullet$ $S$ is the set of nontrivial pairs (i.e. $E \ne 0$) and a pair of the form $({\mathcal O}_C,f,E)$ is nondegenerate if and only if $f \ne 0$. $\bullet$ The slope functions $\mu_\sigma$ are the same as for triples. \end{example} With respect to this slope function, observe that a pair $({\mathcal O}_C,f,E)$ is $\sigma$-stable if and only if: \medskip (i)\ $\mu(F) < \mu(E) - \sigma(\frac 1{\mbox{\small rk}(F)} - \frac 1{\mbox{\small rk}(E)})$ for each $({\mathcal O}_C,g,F) \hookrightarrow ({\mathcal O}_C,f,E)$, and \medskip (ii)\ $\mu(F) < \mu(E) + \sigma(\frac 1{\mbox{\small rk}(E)})$ for each $(0,0,F) \hookrightarrow ({\mathcal O}_C,f,E)$. \begin{theorem}{(\cite{T1} (1.1)-(1.19))} Fix invariants: $$\mbox{rk}(E) = 2\ \mbox{and}\ \mbox{det}(E) \cong {\mathcal O}_C(D),\ \mbox{with} \ \mbox{deg}(D) = d.\ \mbox{Then:}$$ (a) Harder-Narasimhan and Jordan-H\"older filtrations exist if $\sigma > 0$ and yield a well-defined $s$-equivalence for $\sigma$-semi-stable pairs. (b) For each $\sigma > 0$, a projective variety $M_C(2,D,\sigma)$ (abbreviated $M_\sigma$) coarsely represents the functor: ``families of nondegenerate $\sigma$-semistable pairs $({\mathcal O}_C,f,E)$ modulo $s$-equivalence''. There is a universal family over the open locus parametrizing stable pairs, which is smooth and irreducible. \end{theorem} Full proofs of the properties listed below can be found in \cite{T1}. (Please note that our $\sigma$ differs from the $\sigma$ in \cite{T1} by a factor of $2$.) \bigskip \noindent {\bf Properties of Stable Pairs:} {\it Fix rk$(E) = 2$ and det$(E) \cong {\mathcal O}_C(D)$. Also assume that $g \ge 2$ (but see the note at the end of this section). Then: \medskip (a) There are no $\sigma$-semi-stable pairs if $\sigma < 0$ or $\sigma > d$. \medskip (b) There are always $0$-semi-stable pairs, though no $0$-stable pairs. \medskip (c) $M_{d}$ is a point.} \bigskip {\bf Proof:} If $\sigma < 0$, then by (ii) above, $(0,0,E) \hookrightarrow ({\mathcal O}_C,f,E)$ destabilizes any pair. If $\sigma > d$ and $({\mathcal O}_C,f,E)$ is given, let $L$ be the line-bundle image of ${\mathcal O}_C$ in $E$ with induced map $s:{\mathcal O}_C \rightarrow L$. Then using (i) above, the pair is destabilized by the natural inclusion $({\mathcal O}_C,s,L) \hookrightarrow ({\mathcal O}_C,f,E)$. \medskip If $\sigma = 0$, then conditions (i) and (ii) coincide, telling us that $({\mathcal O}_C,f,E)$ is semistable if and only if $E$ is semistable, and that no pair is $0$-stable. Note that there is no Jordan-H\"older filtration. \medskip If $\sigma = d$, then by the analysis in the proof of (a), $({\mathcal O}_C,f,E)$ is $\sigma$-semi-stable if and only if $f:{\mathcal O}_C \rightarrow E$ has no zeroes, and all such $({\mathcal O}_C,f,E)$ are $s$-equivalent, with associated graded $({\mathcal O}_C,\mbox{id},{\mathcal O}_C) \oplus (0,0,{\mathcal O}_C(D))$. So there are no $\sigma$-stable pairs, and the moduli space is a point. \medskip In contrast to the boundary cases presented here, the stable locus in $M_\sigma$ will be nonempty if $0 < \sigma < d$. \bigskip \noindent {\bf Critical Points and Local Triviality:} \begin{center} {\it Let $\Gamma = \{0 < c < d \ | \ c \equiv d \ (\mbox{mod}\ 2)\}$.} \end{center} {\it \noindent For each $\sigma$, let $Z_\sigma \subset M_\sigma$ be the locus of $\sigma$-strictly-semistable pairs. \medskip (d) If $\sigma \not \in \Gamma$, then $Z_\sigma = \emptyset$ (i.e. $\sigma$-semistable $\Rightarrow$ $\sigma$-stable). \medskip (e) If $c = d - 2n \in \Gamma$, then $Z_c \cong C_n$, the $n$-th symmetric product of $C$. \medskip (f) If $ I \subset (0,d) - \Gamma$ is an interval and $\sigma,\sigma' \in I$, then $M_\sigma \cong M_{\sigma'}$. \medskip (g) Suppose that $c\in \Gamma$ and $c^- < c < c^+$ are real numbers in the neighboring intervals of $(0,d) - \Gamma$. Then there are surjective morphisms: $$\begin{array}{ccccc}M_{c^-} &&&& M_{c^+} \\ & \stackrel{f^-}\searrow & & \stackrel{f^+}\swarrow \\ && M_c \end{array}$$ \noindent {\bf Key Point:} $f^-$ and $f^+$ are isomorphisms away from $Z_c \subset M_c$ and projective bundles over $Z_c$. (The projective bundles are identified in the proof).} \bigskip {\bf Proof:} If $\sigma > 0$ and if $F$ is the bundle in a destabilizing subpair of $({\mathcal O}_C,f,E)$, then it is easy to see that $F$ is a line bundle. But if $\sigma \not \in \Gamma$, then the right side of (i) and (ii) are not integers, whereas $\mu(F)$ is an integer. So we cannot have equality. This proves (d). \medskip Suppose $c = d - 2n \in \Gamma$. Then a $c$-strictly semistable pair $({\mathcal O}_C,f,E)$ has a subpair which is either isomorphic to $({\mathcal O}_C,s,{\mathcal O}_C(A))$ (where $s$ is the tautological section, deg$(A) = n$, and ${\mathcal O}_C(A)$ is the image of ${\mathcal O}_C$ in $E$) or else it is isomorphic to $(0,0,L)$, where deg$(L) = d - n$. But {\bf either} possibility forces the associated graded for the pair $({\mathcal O}_C,f,E)$ to be of the form $({\mathcal O}_C,s,{\mathcal O}_C(A)) \oplus (0,0,{\mathcal O}_C(D-A))$, and these are parametrized by $C_n$. \medskip The stability conditions do not change when $\sigma$ moves within an interval $I \subset (0,d) - \Gamma$ (again because $\mu(F) \in {\bf Z}$) so the moduli spaces are isomorphic by the universal property of a coarse moduli space. \medskip If $c\in \Gamma$, then apart from $Z_c$, the stability conditions do not change when $c$ is replaced by $c^-$ or $c^+$, so the first part of the key point follows as in the previous paragraph. \medskip Let $c = d - 2n$, and consider $({\mathcal O}_C,s,{\mathcal O}_C(A)) \oplus (0,0,{\mathcal O}_C(D-A)) \in Z_c$. Then it follows that among all pairs with this associated graded, exactly those pairs of the form: $$\begin{array}{ccccccccc}&&{\mathcal O}_C \\ && \downarrow & \searrow\\ 0 & \rightarrow & {\mathcal O}_C(A) & \rightarrow & E & \rightarrow & {\mathcal O}_C(D-A) \rightarrow 0\end{array}$$ are $c^-$-stable, and these are parametrized by $|K_C + D - 2A|$, which has dimension $d-2n+g-2$ (independent of $A$) since $d-2n > 0$. \medskip On the other hand, among all pairs with this associated graded, exactly those pairs of the form: $$\begin{array}{ccccccccc}&&&&&&{\mathcal O}_C \\ &&&&&\swarrow & \downarrow \\ 0 & \rightarrow & {\mathcal O}_C(D-A) & \rightarrow & E & \rightarrow & {\mathcal O}_C(A) &\rightarrow & 0\end{array}$$ are $c^+$-stable. Such pairs are parametrized by ${\bf P}(V)$, where $V$ sits in the long exact sequence: $$H^0(C,{\mathcal O}_C(D-A)) \rightarrow V \rightarrow H^1(C,O_C(D-2A)) \rightarrow H^1(C,{\mathcal O}_C(D-A)).$$ (in fact, ${\bf P}(V)$ is naturally isomorphic to ${\bf P}(H^0(C,{\mathcal O}_C(D-A) \otimes {\mathcal O}_A)^*))$. \medskip Thus the dimension of ${\bf P}(V)$ is $n-1$, independent of $A$. \bigskip So there are a {\bf finite} number of moduli spaces $M_\sigma$, linked by morphisms as in the following diagram: $$\begin{array}{cccccccccccc} &X_2&&&& X_1 & & & & X_0\\ \cdots & & \searrow & & \swarrow & & \searrow & & \swarrow & & \searrow\\ & & & M_{d-4} & & & & M_{d-2} & & & & M_d \\ \end{array}$$ where each $X_n \cong M_{(d-2n)^-} \cong M_{(d-2n-2)^+}$. \medskip Theorems ~\ref{old1} and ~\ref{old2} are embedded in this diagram because of: \medskip \noindent {\bf Large Values of $\sigma$:} \medskip {\it (h) If $d > 0$, then $X_0 \cong |K_C+D|^*$. \medskip \hskip .22in If $d > 2$, then $M_{d-2} \cong |K_C+D|^*$ and $X_1 \cong X = \ \mbox{bl}(|K_C+D|^*,C)$. \medskip \hskip .22in Moreover, the morphism $f^-: X_1 \rightarrow M_{d-2}$ is the blow-down. \medskip \hskip .22in If $d > 4$, then $f^+:X_1 \rightarrow M_{d-4}$ is the contraction $\gamma:X \rightarrow Y$.} \medskip {\bf Proof:} A special case of the proof of (g) shows that $X_0 \cong |K_C+D|^*$. \medskip Another special case of the proof of (g) shows that $f^+:X_0 \rightarrow M_{d-2}$ is an isomorphism, because the ``exceptional'' part of the map is a ${\bf P}^0$-bundle(!) \medskip When (g) is applied to the map $f^-:X_1 \rightarrow M_{d-2} = |K_C + D|^*$, one discovers that the exceptional locus is a divisor, which is a projective bundle over $C$, hence $f^-$ is the blow-down. \medskip Finally, when (g) is applied to the map $f^+:X_1 \rightarrow M_{d-4}$, the exceptional set consists of lines spanned by two points of $C$ (i.e. the secant lines) which are contracted to points. This means that the linear series which realizes $f^+$ must be a multiple of $|2H - E|$, so the fact that $f^+$ has connected fibers implies $f^+$ is equal to $\gamma$. \bigskip So we've got Theorem ~\ref{old1} and (a very precise) Theorem ~\ref{old2} when $d > 4$. To see what happens for $d = 4$ from this point of view (for example, in the case $g = 2$ and $d = 4$ considered in the introduction), we need to analyze: \medskip \noindent {\bf All Values of $\sigma$:} \medskip {\it (i) If $n > 1$, then $X_n$ (if defined) is isomorphic to $X_1$ off codimension $2$. \medskip \hskip .2in The maps $f^+:X_n \rightarrow M_{d-2n-2}$ are multiples of $|(n+1)H - nE|$. \medskip \hskip .2in The maps $f^-:X_n \rightarrow M_{d-2n}$ are multiples of $|nH - (n-1)E|$. \medskip (j) If $\sigma$ is in the first interval of $(0,d) - \Gamma$, then there is a morphism: $$f:M_\sigma = X_{[\frac {d-1}2]} \rightarrow M_C(2,D)$$ which is the contraction determined by high multiples of $|dH - (d-2)E|$.} \medskip {\bf Proof:} The first part of (i) is a dimension count using (g), which allows us to transfer linear series from $X_1 = X$ over to each $X_n$. The reader is referred to \cite{T1}, where the ample cone is constructed for each $X_n$, the boundary of which gives properties (i) and (j). Notice in particular, that each $f^+$ and $f^-$ is a very simple contraction by property (g), but that the final map $f$ in (j) can have rather more complicated behavior, as in the example of the introduction. \bigskip Thus the mirror image of the diagram following property (g) gives us: $$\begin{array}{cccccccccccc}X = X_1&&&&&&X_2 && & & X_{[\frac {d-1}2]}\\ \downarrow &&\searrow &&& \swarrow && \searrow &\cdots & \swarrow & \downarrow \\ |K_C+D|^*&&&&Y = M_{d-4}&&&&& &M_C(2,D) \end{array}$$ which is the advertised generalization of Theorems ~\ref{old1} and ~\ref{old2}. \bigskip \noindent Note: Most of this analysis also applies to curves of genus $0$ and $1$. \medskip {\bf genus 1:} All properties (a)-(j) apply. The only difference between this and the general case occurs when $d$ is even, in which case $M_C(2,D)$ is isomorphic to ${\bf P}^1$, rather than a point, as one would expect by a dimension count. For example, if $d = 4$, then $\gamma: X = \mbox{bl}({\bf P}^3,C) \rightarrow Y = {\bf P}^1$ is the contraction determined by the pencil of quadrics vanishing along $C$. \medskip {\bf genus 0:} Properties (a)-(j) apply if $d$ is even and $M_C(2,D)$ is a point, corresponding to the vector bundle ${\mathcal O}_{{\bf P}^1}(\frac d2) \oplus {\mathcal O}_{{\bf P}^1}(\frac d2)$. On the other hand, if $d = 2n + 1$, then $M_\sigma = \emptyset$ if $\sigma < 1$, because all bundles are unstable. Other than this, which forces obvious changes to properties (b),(c),(g) and (j), everything is as in the general case. Notice that in this case, $M_1$ is isomorphic to ${\bf P}^n = ({\bf P}^1)_n$, by property (e). For example, when $d = 5$, then $\gamma: X = \mbox{bl}({\bf P}^3,C) \rightarrow Y = M_1 = {\bf P}^2$ is the contraction determined by the web of quadrics vanishing along the twisted cubic. This contraction is a ${\bf P}^1$ bundle, a special case of the key point of property (g). \section{Log Flips.} The goal of this section is to interpret the birational maps: $$X = X_1 --\!\!\!> X_2 --\!\!\!> ... --\!\!\!> X_{[\frac {d-1}2]}$$ as flips, in the sense of the minimal model program. In fact, they are not flips, but rather their {\bf inverses} are flips (at least initially), in the traditional sense. While this is an interesting observation, it is not the one I want to pursue, because the inverses point in the wrong direction, from the point of view of Theorems ~\ref{old1} and ~\ref{old2}. For example, with this interpretation, the first contraction $\gamma:X\rightarrow Y$ is not a flipping contraction, but rather contracts curves whose intersection with $K_X$ are positive. Such contractions are hard to understand, in general. Fortunately, the theory of log minimal models provides a means for turning the flips around, provided we can find suitable divisors on the $X_k$. We begin with a quick tour of the parts of the log minimal model program relevant to our discussion. \bigskip Let $X$ be a smooth projective variety. \begin{definition} A ${{\bf Q}}$-divisor on $X$ is a finite sum of distinct prime divisors with rational coefficients. It is effective if all the coefficients are non-negative. Intersections with curves, self intersections and numerical equivalence are all defined as with ordinary divisors. Let $F$ be an effective ${{\bf Q}}$-divisor on $X$. If $F = \sum \alpha_iF_i$, then the support of $F$, denoted Supp$(F)$, is the union of the prime divisors $F_i$ which appear in $F$ with positive coefficients. \end{definition} \begin{definition} If $F$ is an effective divisor on $X$, then a {\bf log resolution} of $(X,F)$ is a morphism $f:\widetilde X \rightarrow X$ with the property that $\widetilde X$ is smooth, and $\sum f\mbox{-exceptional divisors}\ + f_*^{-1}(\mbox{Supp}(F))$ is a normal crossings divisor. \end{definition} \noindent Note: If $f:Y \rightarrow X$ is any birational morphism of smooth varieties, and if $D$ is a ${{\bf Q}}$-divisor on $X$, then $E_f$ will denote the sum of the $f$-exceptional divisors, and $f^*(D)$ and $f_*^{-1}(D)$ will denote the total transform and the strict transform of $D$ on $Y$, respectively. (They are well-defined by linearity.) \medskip Let $F = \sum \alpha_i F_i$ be an effective ${{\bf Q}}$-divisor. \medskip \begin{definition} $F$ is {\bf log canonical} if each coefficient $\alpha_i \le 1$, and there is a log resolution of $(X,F)$ with the property that all coefficients of the components of $E_f$ are at least $-1$ in the ${{\bf Q}}$-divisor: $$(K_{\widetilde X} -f^*K_X) + (f_*^{-1}(F) - f^*(F))$$ \end{definition} \noindent Note: This property is independent of the log resolution. \begin{definition} Suppose that $B \subset X$ is a curve (which we also identify with its image in $\mbox{H}_2(X,{\bf R})$). Then $B$ spans an {\bf extremal ray} of the cone of effective curves on $X$ if there is an element $\lambda \in \ \mbox{H}^2(X,{\bf R})$ such that: (i) $\lambda(B) = 0$ and (ii) if $\beta \in \ \mbox{H}_2(X,{\bf R})$ is a limit of sums $\sum c_iB_i$ of curves with positive (real) coefficients, then $\lambda(\beta) \ge 0$ with equality if and only if $\beta$ is a multiple of $B$. \end{definition} \begin{definition} Suppose that $B$ spans an extremal ray and $f:X \rightarrow Y$ is a morphism satisfying $f_*({\mathcal O}_X) = {\mathcal O}_Y$. If $B$ is contained in a fiber of $f$, and if moreover every curve contained in every fiber of $f$ is homologous to a (rational) multiple of $B$, then $f$ (which is uniquely determined if it exists) is called the {\bf extremal contraction} associated to $B$. \end{definition} Suppose that $F$ is a log-canonical divisor on $X$. A basic result of the log minimal model program is the following (see \cite{CKM} and \cite{Ketal}): \bigskip \noindent {\bf Contraction Theorem:} {\it If $B\subset X$ spans an extremal ray and $B.(K_X + F) < 0$, then there is an extremal contraction $\gamma:X \rightarrow Y$ associated to $B$.} \bigskip A central question of the minimal model program is: \bigskip \noindent {\bf Do Log Flips Exist?:} Suppose the contraction $f: X \rightarrow Y$ of the theorem is an isomorphism off codimension $2$ in $X$. Then does there exist a morphism $f^+:X^+ \rightarrow Y$ with the following properties: \medskip (a) $f^+$ is an isomorphism off codimension 2 in $X^+$. Let $(K_X + F)^+$ be the strict transform of $K_X+F$ in $X^+$. \medskip (b) If $B^+ \subset X^+$ is a curve lying in a fiber of $f^+$, then $B^+.(K_X+F)^+ > 0$. \medskip (c) The singularities of $X^+$ (or rather, of $(X^+,F^+)$) are not too bad (for example, so that we can even define the intersections $B^+.(K_X+F)^+$ in (b)). \bigskip When $F = \emptyset$ and the dimension of $X$ is $3$, then the affirmative answer to this question is a deep theorem of Mori (together with a definition of ``not too bad'', of course). The answer is also known to be yes for arbitrary $F$ and dimension $3$. The interested reader is urged to consult \cite{CKM} and \cite{Ketal}, as well as Koll\'ar's notes in this Proceedings for an introduction to the minimal and log minimal model programs and other applications. \bigskip Next, we construct a morphism which will eventually be a log resolution. \bigskip Let $M$ be a line bundle on $C$, let $C_k$ be the $k$-th symmetric product of $C$, and let $V = H^0(C,M)$. If $M$ has the following property: $$(*)_k: \ \mbox{For all} \ D \in C_k, \ \mbox{dim}(H^0(C,M(-D)) = \mbox{dim}(V) - k$$ then each such divisor $D$ determines a ${\bf P}^{k-1} \subset {\bf P}(V)$, which is called the span of $D$. Given that property $(*)_k$ holds, the $k$-th secant variety is: $$\Sigma_k(C) = \bigcup _{D\in C_k} \ \mbox{span}(D) \subset {\bf P}(V).$$ If $M$ satisfies $(*)_2$ (i.e. $M$ is very ample), let $X = \ \mbox{bl}({\bf P}(V),C)$ (as in \S 1). \bigskip \noindent Observation: ${\mathcal O}_C(K_C + D)$ satisfies $(*)_{d-1}$. (Riemann-Roch!) \medskip The following construction blows up the secant varieties of $C$. \begin{theorem}\label{logres} (\cite{B1} Theorem 1) (a) Suppose $n \ge 1$ and $M$ is a line bundle with property $(*)_{2n}$. Then there is a birational morphism $f:\widetilde X \rightarrow X$ which is a composition of the following blow-ups: \medskip $f^{(2)}: X^{(2)} \rightarrow X^{(1)} = X \ \mbox{blows up the strict transform of $\Sigma_2(C)$}$, \medskip $f^{(3)}:X^{(3)} \rightarrow X^{(2)} \ \mbox{blows up the strict transform of $\Sigma_3(C)$}$ \medskip \hskip .5in $\vdots$ $f^{(n)}:\widetilde X = X^{(n)} \rightarrow X^{(n-1)}\ \mbox{blows up the strict transform of $\Sigma_n(C)$}$ \medskip Moreover, the strict transform of each $\Sigma_k(C)$ in $X^{(k-1)}$ is smooth and irreducible of dimension $2k-1$, transverse to all exceptional divisors, so in particular $\widetilde X$ is smooth. \medskip For consistency, let $f^{(1)}: X \rightarrow {\bf P}(V)$ be the blow-down. Let $E^{(k)}$ be the strict transform in $\widetilde X$ of each $f^{(k)}$-exceptional divisor. Then $E^{(1)} + ... + E^{(n)}$ is a normal crossings divisor on $\widetilde X$ with $n$ smooth components. \medskip If $M$ is a line bundle that does not satisfy $(*)_2$, let $\widetilde X = {\bf P}(V)$. Then: \medskip (b) (Terracini recursiveness) Suppose $k \le n$ and $x\in \Sigma_k(C) - \Sigma_{k-1}(C)$. Then the fiber $$(f^{(k)})^{-1}(x) \subset X^{(k)}$$ is naturally isomorphic to ${\bf P}(H^0(C,M(-2A)))$, where $A$ is the unique divisor of degree $k$ whose span contains $x$. Moreover, the fiber $$f^{-1}(x) \subset E^{(k)} \subset \widetilde X$$ is isomorphic to $\widetilde X_A$, the variety obtained by applying (a) of the Theorem to the line bundle $M(-2A)$. \medskip (c) If $g \ge 2$ and if $M ={\mathcal O}_C(K_C+D)$, then there is a {\bf morphism} $$\widetilde \psi_{|K_C+D|}: \widetilde X \rightarrow M_C(2,D)$$ which extends $\psi_{|K_C+D|}$. When restricted to a fiber $f^{-1}(x)$ of part (b), $\widetilde \psi_{|K_C+D|}$ agrees with $\widetilde \psi_{|K_C+D-2A|}$ (and this property determines $\widetilde \psi_{|K_C+D|}$ uniquely!) \end{theorem} For the proof, see \cite{B1}. Notice that parts (a) and (b) make no reference to moduli, hence generalize to, for example, canonical embeddings, where condition $(*)_d$ is equivalent to the nonexistence of $g^1_d$'s. As for part (c), the idea is to construct a vector bundle on $C \times \widetilde X$ by a sequence of elementary modifications of the bundle (constructed from the universal extension) on $C \times |K_C+D|^*$ along the exceptional divisors for each $f^{(k)}$, and to use this bundle to get the map to moduli. \medskip In fact, though, the proof really constructs families of nondegenerate pairs $({\mathcal O}_C,f,E)$ parametrized by the $\widetilde X^{(k)}$ (in all genera) with the following property. For every $y\in \widetilde X$ and every $\sigma \in [0,d)$ (or $[1,d)$ if $g = 0$ and $d$ is odd) there is an $X^{(k)}$ such that the image of $y$ in $X^{(k)}$ parametrizes a $\sigma$-semistable pair. Thus, for each $\sigma$, there is a natural morphism: $$\psi_\sigma:\widetilde X \rightarrow M_\sigma.$$ \bigskip Now we construct log-canonical divisors on $X = \ \mbox{bl}(|K_C+D|^*,C)$. \medskip \noindent {\bf Linear Algebra Construction:} Given any vector bundle $F$ on $C$, the cup product gives rise to a linear map: $$c: \ \mbox{Ext}^1(F(D),F) \rightarrow \ \mbox{Hom}(H^0(C,F(D)),H^1(C,F))$$ Also, the summand ${\mathcal O}_C \hookrightarrow F \otimes F^*$ produces an inclusion of vector spaces: $$\iota: \ \mbox{Ext}^1({\mathcal O}_C(D),{\mathcal O}_C) \hookrightarrow \ \mbox{Ext}^1(F(D),F)$$ One can think of the composition $c\circ \iota$ pointwise as follows. Given $$\epsilon:\ 0 \rightarrow {\mathcal O}_C \rightarrow E \rightarrow {\mathcal O}_C(D) \rightarrow 0$$ one tensors each term by $F$, and $c(\iota(\epsilon))$ is the connecting homomorphism: $$c(\iota(\epsilon)) = \delta: H^0(C,F(D)) \rightarrow H^1(C,F).$$ When we lift $c\circ \iota$ to a map of trivial bundles on $|K_C + D|^*$, it determines a matrix $M(F)$ of linear forms on $|K_C+D|^*$ via: $${\mathcal O}_{|K_C+D|^*}(-1) \rightarrow \ \mbox{Hom}(H^0(C,F(D)),H^1(C,F)) \otimes {\mathcal O}_{|K_C+D|^*}.$$ \begin{proposition}\label{linalg} (a) For each $0 < k \le \frac d2$, there is a nonempty open subset $U \subset \ \mbox{Pic}^{k-(g-1)}(C)$ such that $$(*)\ \ L \in U \Rightarrow \ h^0(C,L^{-1}(D)) = d-k \ \mbox{and}\ h^1(C,L^{-1}) = k.$$ If $L \in U$, choose a basis for $H^0(C,L^{-1}(D))$, and let $I = (i_1,...,i_k)$ be a mutiindex with $1 \le i_1 < ... < i_k \le d-k$. Then $I$ determines a $k\times k$ minor $M_I(L^{-1})$ (choosing columns $i_1,...,i_k$ from the matrix $M(L^{-1})$) yielding a divisor on $\widetilde X$: $$D_{L,I} \in |kH - (k-1)E^{(1)} - (k-2)E^{(2)} - ... - E^{(k-1)}|.$$ (i.e. the generic multiplicity of $\mbox{det}(M_I(L^{-1}))$ along $\Sigma_i(C)$ is at least $k-i$). \medskip Finally, if we let $V_k$ be the sub-linear-series spanned by the $D_{L,I}$, then $V_k$ is base-point-free (and independent of choices of basis). \medskip (b) Suppose $g > 0$. Then for each $0 < l \le d$, there is a nonempty open subset $U \subset M_C(2,l-(2g-2))$ such that $$(*)\ \ F \in U \Rightarrow \ h^0(C,F^{-1}(D)) = 2d - l\ \mbox{and}\ h^1(C,F^{-1}) = l.$$ If $F \in U$, choose $J = (j_1,...,j_l)$ such that $1 \le j_1 < ... < j_l \le d-l$ and the minor $M_J(F^{-1})$ as in (a). Then det$(M_J(F^{-1}))$ determines a divisor: $$D_{F,J} \in |lH - (l-2)E^{(1)} - (l-4)E^{(2)} - ... |,$$ and the sub-linear-series $W_l$ spanned by the $D_{F,J}$ is base-point-free. \end{proposition} {\bf Proof:} The values for $h^0$ and $h^1$ in $(*)$ are generic in Pic and $M_C(2,*)$ respectively. $U$ is an intersection of two nonempty open subsets. For the next part, the following observation is crucial. Given an effective divisor $A$ on $C$, an extension $\epsilon \in \ \mbox{Ext}^1({\mathcal O}_C(D),{\mathcal O}_C)$ determines a point $\overline \epsilon \in \ \mbox{span}(A) \subset |K_C+D|^*$ if and only if the extension splits when pushed forward: $$\begin{array}{ccccccccc}0 & \rightarrow & {\mathcal O}_C & \rightarrow & E & \rightarrow & {\mathcal O}_C(D) & \rightarrow & 0 \\ &&\downarrow & \swarrow \\ &&{\mathcal O}_C(A) \end{array}$$ Now suppose $\overline \epsilon \in \Sigma_i(C)$, so is in the span of some divisor $A$ of degree $i < k$. It follows (tensoring the inclusion ${\mathcal O}_C(D-A) \hookrightarrow E$ by $L^{-1}$) that $H^0(C,L^{-1}(D-A)) \subset \ \mbox{ker}(c(\iota(\epsilon)))$, which by Riemann-Roch has dimension at least $(d-k)-i$. Thus the rank of each $M_I(L^{-1})$ is at most $i$, and therefore its determinant has multiplicity at least $k-i$ at $\overline \epsilon$, from which the linear series computation in (a) follows. The linear series in (b) is computed similarly. \medskip We prove base-point-freeness first when $g=0$ and $d = 2n+1$. Given $k$, if $y \in \widetilde X - (E^{(1)} \cup ...\cup E^{(k-1)})$, then the bundle $E$ associated to $\overline \epsilon = y$ is isomorphic to ${\mathcal O}_{{\bf P}^1}(m) \oplus {\mathcal O}_{{\bf P}^1}(d-m)$ where $d-m > m \ge k$. This is because of the crucial observation. It follows that $h^0(C,E(-k-1)) = d-2k$, so some $M_I({\mathcal O}_{{\bf P}^1}(-k-1))$ has full rank at $\overline \epsilon$, and thus $y$ is not a base point. If $y\in E^{(i)}$ for some (minimal) $i < k$, then Theorem ~\ref{logres} (b) allows us to place $y$ in a fiber over $\Sigma_i(C)$ isomorphic to $\widetilde X_A$ for some divisor $A$ satisfying deg$(A) = i$. The restriction to this fiber of $V_k$ is identified with the linear series $V_{k-i}$ under the isomorphism with $\widetilde X_A$, and so we can conclude base-point-freeness by induction. \medskip The proof of base-point-freeness is similar in general. Suppose $d = 2n+1$ or $2n+2$, so ${\mathcal O}_C(K_C+D)$ satisfies $(*)_{2n}$. Given $k$ (or $l = 2k$ or $2k-1$), first consider the points $y \in \widetilde X - (E^{(1)} \cup ... \cup E^{(k-1)})$. If $y \in E^{(k)} \cup ... \cup E^{(n)}$, then we can find an $L \in U$ (or $F\in U$) such that $h^0(C,E\otimes L^{-1}) = d - 2k$ (or $h^0(C,E\otimes F^{-1}) = 2d - 2l$) because the bundle $E$ associated to $\overline \epsilon = f(y)$ fits in an exact sequence: $0 \rightarrow {\mathcal O}_C(D-A) \rightarrow E \rightarrow {\mathcal O}_C(A) \rightarrow 0$ ($k \le \ \mbox{deg}(A) \le n$). (This is a consequence of the crucial observation.) If $y$ does not lie in an exceptional divisor, then the bundle $E$ associated to the extension $\overline \epsilon = y$ is semistable (Theorem ~\ref{logres} (c) or the crucial observation) in which case the same fact about $h^0(C,E\otimes L^{-1})$ (or $h^0(C,E\otimes F^{-1})$) is a standard result, for example, see \cite {B3}, Lemma 3.6. Once this is achieved, one has base-point-freeness off the exceptional divisors $E^{(1)} \cup ... \cup E^{(k-1)}$ and the points of these exceptional divisors are treated by induction using Theorem ~\ref{logres} (b) and the same identification of $V_k$ with $V_{k-i}$ (or $W_l$ with $W_{l-2i}$) as above. \bigskip There are two exceptional cases (given in detail below) where the linear series ($V_k$ in case (a) and $W_l$ in case (b)) are trivial, which occur when $\Sigma_n(C)$ is a divisor and $k$ (or $l$) is maximal. In all other cases, we can use Bertini to find smooth members of the linear series which meet all the exceptional divisors $E^{(1)},...,E^{(n)}$ transversally. \bigskip \noindent {\bf Exceptional Cases:} (a) Suppose $g = 0$ and $d = 2n+2 = 2k$. Then there is only one $k\times k$ matrix $M_I({\mathcal O}_{{\bf P}^1}(-k-1))$, and $V_k$ has only one element. (So since it is base-point-free, it has to be trivial!) With a suitable choice of basis, $M_I({\mathcal O}_{{\bf P}^1}(-k-1))$) is the standard square matrix: \medskip $$\left(\begin{array}{ccccc}z_1 & z_2 & z_3 & \cdots & z_k \\ z_2 & z_3 & z_4 & \cdots & z_{k+1} \\ \vdots & \vdots & \vdots & & \vdots \\ z_k & z_{k+1} & z_{k+2} & \cdots & z_{d-1}\end{array}\right)$$ \medskip \noindent whose $2\times 2$ minors cut out the rational normal curve $C \subset {\bf P}^{d-2} = |K_C+D|^*$ (see Proposition 9.7 in \cite{H}). Its determinant cuts out $\Sigma_n(C) \subset {\bf P}^{d-2}$. The linear series $V_k$ is trivial because here $E^{(n)} \equiv nH - (n-1)E^{(1)} - ... - 2E^{(n-1)}$ on $\widetilde X$. Notice that although we cannot use Bertini to find smooth divisors in $V_k$, we can use Theorem ~\ref{logres} (a) to conclude that $E^{(n)}$ itself is smooth and meets the other exceptional divisors transversally. \medskip (b) Suppose $g = 1$, and $d = l = 2n+1$. Then there is one $l\times l$ matrix $M_J(F^{-1})$ for each stable bundle $F$ of rank $2$ and degree $d = d-(2g-2)$. There is no reason a priori why this determines a trivial linear series, however, as in Exception (a), one computes that $\Sigma_n(C) \subset |K_C+D|^*$ is a divisor, of degree $d$, which must therefore be the zero locus of each determinant, and $E^{(n)} \equiv dH - (d-2)E^{(1)} - ... - 3E^{(n-1)}$. (The degree can be computed using Lemma 2.5 (Chapter VIII) from \cite{ACGH}). Again, we will use the fact that $E^{(n)}$ is smooth, intersecting the other exceptional divisors transversally. \bigskip \noindent {\bf Remarks:} If $l$ is even, the conditions $h^0(C,F^{-1}(D)) = 2d-l$ and $h^0(C,F^{-1}) = l$ may not be independent of the choice of a representative $F$ for a semistable point in $M_C(2,l-(2g-2))$. However, if these properties are true for the associated graded, then they hold for all representatives, as is easily checked. Moreover, if we let $l = 2k$, then the split bundles determine an inclusion of linear series: $V_k \cdot V_k \subseteq W_l$. (In genus $0$ and $1$, this is an equality!) \bigskip Next, we use the linear series to find: \bigskip \noindent {\bf Some Log Canonical Divisors on $X$:} Let $F = f_*A$, where: \medskip \noindent {\bf Genus 0:} (a) If $d = 2n+1$, then $A \in W_{2n}$ is a general member. \medskip (b) If $d = 2n+2$, then $A = E^{(n)} + A'$, where $A' \in V_{n}$ is a general member. \medskip \noindent {\bf Genus 1:} (a) If $d = 2n+1$, then $A = E^{(n)}$. \medskip (b) If $d = 2n+2$, then $A \in W_d$ is a general member. \medskip \noindent {\bf Genus $\ge$ 2:} $A \in W_d$ is general. \bigskip Alternatively, one can think of $F$ in each case as the strict transform in $X$ of a hypersurface in $|K_C+D|^*$ (highly singular along the secant varieties). However, when we think of $F$ as the push-forward of a divisor $A$ on $\widetilde X$, then the following becomes almost immediate. \bigskip \noindent {\bf Claim:} In all the cases above, $f:\widetilde X \rightarrow X$ is a log resolution of $(X,F)$ and $F$ is a log canonical divisor. \medskip {\bf Proof:} By Theorem ~\ref{logres} (a), all the $f$-exceptional divisors are smooth with normal crossings. In each case, the strict transform of the support of $F$ is the support of $A$, which is a sum of smooth divisors which intersect all others with normal crossings, either by Bertini or Theorem ~\ref{logres}(a) again. So $f$ is a log resolution of $(X,F)$. \medskip Since each blow-up $f^{(k)}$ was along a smooth center trasnverse to all exceptional divisors, it follows that the coefficient of $E^{(k)}$ in $K_{\widetilde X} - f^*K_X$ is the codimension of $\Sigma_k(C)$ in $|K_C+D|^*$ minus $1$, a consequence of Riemann-Hurwitz. It also follows that since we constructed $F$ as $f_*A$, the coefficient of $E^{(k)}$ in $f_*^{-1}F - f^*F = A-f^*F$ is the (negative of the) generic multiplicity of $F$ along the strict transform of $\Sigma_k(C)$ in $X$, which is computed directly from the linear series in which $A$ lies. This is the information we need to check that $F$ is log canonical. The computations in genus $0$ are left to the reader. Here is the data for genus $\ge 1$: \medskip $\bullet$ \ codimension of $\Sigma_k(C)$ in $|K_C+D|^*$:\ \ $d + g - 2k -1$. \medskip $\bullet$ \ multiplicity of $F$ along $E^{(k)}$:\ \ $d - 2k$. \medskip Since it follows that the coefficient of each $E^{(k)}$ in $(K_{\widetilde X} - f^*K_X) + (A - f^*F)$ is $g-2$, we see that $F$ is log canonical. \medskip We can (and need to!) do a little better when $g \ge 2$ if we use ${{\bf Q}}$-divisors. If $p,q$ are positive integers, let $(W_d)^{p}$ be the linear series spanned by products of $p$ elements of $W_d$, and given a smooth element $G \in (W_d)^p$ (this linear series is base-point free), consider $F' = \frac 1qf_*(G)$. This is not only numerically equivalent to $\frac pq F$ (as is easy to see), but all coefficients of the $E^{(k)}$ in the expression $f_*^{-1} F' - f^*F'$ are $\frac pq$ times the corresponding coefficients for $F$. We will abuse notation and say that this divisor is a member of $\frac pq F$, keeping in mind that if $p > q$, then the literal ${{\bf Q}}$-divisor $\frac pq F$ cannot be log canonical, by definition, while a member constructed in this way might be log canonical. In fact, if $d > 4$, then $$\left(\frac{d+g-5}{d-4}\right)F \ \mbox{has a log canonical member}$$ by the data above (keep in mind that $E^{(1)}$ is not $f$-exceptional). \bigskip Now we will relate these log canonical divisors to the diagram at the end of \S 2 constructed by stable pairs. Namely, recall that whenever $d > 2k$, there was a diagram: $$\begin{array}{ccccc} X_{k-1} & & & & X_k \\ &\stackrel{f^+}\searrow \ & & \stackrel{f^-}\swarrow \\ & & M_{d-2k}\ \ \ \ \end{array}$$ Moreover, $f^-$ and $f^+$ are obviously extremal ray contractions since each contracts a projective bundle over $C_k$ and dim$(H_2(X,{\bf R})) = 2$. (Take any curve in a projective-space fiber to span the extremal ray.) Finally, each contraction is an isomorphism off of codimension $2$. \bigskip \begin{proposition}\label{logflip} If $k = 2$ or $d > 2g -2$ and $k$ is arbitrary, then the diagram above is a log flip for (the strict transform on $X_{k-1}$ of) $K_X + \left(\frac {d+g-5}{d-4}\right)F$. \end{proposition} {\bf Proof:} We need to show: (a) the member of $\left(\frac {d+g-5}{d-4}\right)F$ constructed as above is log canonical on each $X_{k}$ (this will certainly suffice for the condition ``not too bad'' in the ``definition'' of log flips), and (b) If $B \subset X_{k-1}$ and $B^+ \subset X_k$ are curves spanning extremal rays, then $B.(K_X+\left(\frac {d+g-5}{d-4}\right)F) < 0$ and $B^+.(K_X+\left(\frac {d+g-5}{d-4}\right)F) > 0$. \bigskip We prove (b) first. Recall (property (i) from \S 2) that the map $f^+$ is a multiple of $|kH - (k-1)E|$. Thus, if $B \subset X_{k-1}$ is an extremal ray, then $B.(H - \frac{k-1}kE) = 0$. Moreover, $|(k-1)H - (k-2)E|$ is nef on $X_{k-1}$, and $B$ is {\bf not} contracted in this linear series, so it follows that $B.(H-\frac{k-2}{k-1}E) > 0$, and $B.E > 0$. \medskip \noindent From the data: \medskip $K_X \equiv -(d+g-1)H + (d+g-4)E$, and \medskip $F \equiv dH - (d-2)E$, \medskip \noindent we get $K_X+\left(\frac{d+g-5}{d-4}\right)F \equiv \frac {4g-4}{d-4}(H - \frac{d+2g-6}{4g-4}E)$, from which it follows that its intersection with $B$ is negative when $k = 2$ or $d > 2g-2$. Moreover, the mirror image of this argument shows that if $B^+ \subset X_k$ is an extremal ray for $f^-$, then its intersection with $K_X+\left(\frac{d+g-5}{d-4}\right)F$ is positive in the same cases. \medskip So the only thing left to see is the fact that the member of $\left(\frac {d+g-5}{d-4}\right)F$ we constructed is log canonical on all $X_k$, not just $X = X_1$, as was shown earlier. In fact, I claim a stronger result, which will explain all the maps as log flips: \begin{lemma} If $d > 2k$, then a general member of $\left(\frac{d+g-2k-1}{d-2k}\right)F$ is log canonical on $X_{k-1}$. \end{lemma} {\bf Proof:} The construction is as before, pushing down a general element of $(W_d)^{d+g-2k-1}$ and dividing by $d-2k$. After $k-1$ elementary modifications, the proof of Theorem ~\ref{logres}(c) (see \cite{B1}) produces a family of $\sigma = n-2k+1$-stable pairs on $C$ parametrized by $X^{(k-1)}$, hence a morphism $\psi^{(k-1)}:X^{(k-1)} \rightarrow X_{k-1}$ since $X_{k-1} = M_\sigma$. Moreover, the morphism $\psi_\sigma : \widetilde X \rightarrow X_{k-1}$ factors through $\psi^{(k-1)}$ via the composition of blow-downs $f_{k-1}:\widetilde X \rightarrow X^{(k-1)}$ from Theorem ~\ref{logres}(a). \medskip One checks that if $G \in (W_d)^p$ (for any $p > 0$), then $(f_{k-1})_*G$ descends to a divisor on $X_{k-1}$. Thus when we log resolve $(X_{k-1},\frac pq F)$ by the map $\psi_\sigma$, then only the exceptional divisors $E^{(k)}$ and above appear with a nonzero coefficient in $\psi_\sigma^*\frac pq F - (\psi_\sigma)_*^{-1}(\frac pq F)$, and those, it is easy to see, appear with the same coefficients as in the earlier computation. The lemma immediately follows. \medskip \begin{corollary}\label{logcan} Each rational map $X_{k-1} --\!\!\!> X_{k}$ is a log flip. \end{corollary} {\bf Proof:} Using the log canonical divisor on $X_{k-1}$ from the Lemma: $$K_X + \left(\frac{d+g-2k-1}{d-2k}\right)F \equiv \frac {k(2g-2)}{d-2k}\left(H - \frac{(2k-1)(2g-4) + d-2}{k(2g-2)}E\right)$$ has negative intersection with $B$ and positive intersection with $B^+$ (as in the proof of Proposition ~\ref{logflip}, keeping in mind the fact that $d > 2k$). \bigskip \noindent {\bf Final Remark:} I have split off Proposition ~\ref{logflip} from Corollary ~\ref{logcan} (which is in a sense more powerful!) to point out a curious fact. Namely, if $d > 2g-2$, which is precisely when $\psi_{|K_C+D|}:|K_C+D|^* --\!\!\!> M_C(2,D)$ is dominant, then we can construct a single ${{\bf Q}}$-divisor on $X$ for which all the maps $X_{k-1} --\!\!\!> X_k$ simultaneously become log flips. When $d \le 2g-2$, however, one needs to tailor the divisor to the variety $X_{k-1}$ and in fact it seems that no single ${{\bf Q}}$-divisor on $X$ will be log canonical and have the desired intersection properties with all the extremal rays. (At least the linear algebra construction does not produce such a divisor.) \medskip \noindent {\bf Acknowledgements:} I would like to thank Michael Thaddeus and the referee for their careful reading and useful comments on an earlier version of this paper.
1998-03-25T20:05:02
9707
alg-geom/9707010
en
https://arxiv.org/abs/alg-geom/9707010
[ "alg-geom", "math.AG" ]
alg-geom/9707010
Michael Finkelberg
Michael Finkelberg, Ivan Mirkovi\'c (Independent University of Moscow and University of Massachusetts at Amherst)
Semiinfinite Flags. I. Case of global curve $P^1$
References updated
null
null
null
null
The Semiinfinite Flag Space appeared in the works of B.Feigin and E.Frenkel, and under different disguises was found by V.Drinfeld and G.Lusztig in the early 80-s. Another recent discovery (Beilinson-Drinfeld Grassmannian) turned out to conceal a new incarnation of Semiinfinite Flags. We write down these and other results scattered in folklore. We define the local semiinfinite flag space attached to a semisimple group $G$ as the quotient $G((z))/HN((z))$ (an ind-scheme), where $H$ and $N$ are a Cartan subgroup and the unipotent radical of a Borel subgroup of $G$. The global semiinfinite flag space attached to a smooth complete curve $C$ is a union of Quasimaps from $C$ to the flag variety of $G$. In the present work we use $C=P^1$ to construct the category $PS$ of certain collections of perverse sheaves on Quasimaps spaces, with factorization isomorphisms. We construct an exact convolution functor from the category of perverse sheaves on affine Grassmannian, constant along Iwahori orbits, to the category $PS$. Conjecturally, this functor should correspond to the restriction functor from modules over quantum group with divided powers to modules over the small quantum group.
[ { "version": "v1", "created": "Wed, 9 Jul 1997 17:31:57 GMT" }, { "version": "v2", "created": "Wed, 25 Mar 1998 19:05:02 GMT" } ]
2008-02-03T00:00:00
[ [ "Finkelberg", "Michael", "", "Independent University of Moscow\n and University of Massachusetts at Amherst" ], [ "Mirković", "Ivan", "", "Independent University of Moscow\n and University of Massachusetts at Amherst" ] ]
alg-geom
\section{Introduction} \subsection{} We learnt of the {\em Semiinfinite Flag Space} from B.Feigin and E.Frenkel in the late 80-s. Since then we tried to understand this remarkable object. It appears that it was essentially constructed, but under different disguises, by V.Drinfeld and G.Lusztig in the early 80-s. Another recent discovery ({\em Beilinson-Drinfeld Grassmannian}) turned out to conceal a new incarnation of Semiinfinite Flags. We write down these and other results scattered in the folklore. \subsection{} Let $\bG$ be an almost simple simply-connected group with a Cartan datum $(I,\cdot)$ and a simply-connected simple root datum $(Y,X,\ldots)$ of finite type as in ~\cite{l}, ~2.2. We fix a Borel subgroup $\bB\subset\bG$, with a Cartan subgroup $\bH\subset\bB$, and the unipotent radical $\bN$. B.Feigin and E.Frenkel define the Semiinfinite Flag Space $\CZ$ as the quotient of $\bG((z))$ modulo the connected component of $\bB((z))$ (see ~\cite{ff}). Then they study the category $\PS$ of perverse sheaves on $\CZ$ equivariant with respect to the Iwahori subgroup $\bI\subset \bG[[z]]$. In the first two chapters we are trying to make sense of this definition. We encounter a number of versions of this space. In order to give it a structure of an ind-scheme, we define the (local) semiinfinite flag space as $ \widetilde{\bf Q}= \bG((z))/\bH\bN((z))$ (see section 4). The (global) semiinfinite space attached to a smooth complete curve $C$ is the system of varieties $\CQ^\al$ of ``quasimaps'' from $C$ to the flag variety of $\bG$ --- the Drinfeld compactifications of the degree $\al$ maps. In the present work we restrict ourselves to the case $C=\BP^1$. The main incarnation of the semiinfinite flag space in this paper is a collection $\CZ$ (for {\em zastava}) of (affine irreducible finite dimensional) algebraic varieties $\CZ^\alpha_\chi\sub \CQ^\al$, together with certain closed embeddings and {\em factorizations}. Our definition of $\CZ$ follows the scheme suggested by G.Lusztig in ~\cite{l2}, \S11: we approximate the ``closures'' of Iwahori orbits by their intersections with the transversal orbits of the opposite Iwahori subgroup. However, since the set-theoretic intersections of the above ``closures'' with the opposite Iwahori orbits can not be equipped with the structure of algebraic varieties, we postulate $\CZ^\alpha_\chi$ for the ``correct'' substitutes of such intersections. Having got the collection of $\CZ^\alpha_\chi$ with factorizations, we imitate the construction of ~\cite{fs} to define the category $\PS$ (for {\em polubeskrajni snopovi}) of certain collections of perverse sheaves with $\BC$-coefficients on $\CZ^\alpha_\chi$ with {\em factorization isomorphisms}. It is defined in chapter 2; this category is the main character of the present work. \subsection{} \label{quantum} If $\bG$ is of type $A,D,E$ we set $d=1$; if $\bG$ is of type $B,C,F$ we set $d=2$; if $\bG$ is of type $G_2$ we set $d=3$. Let $q$ be a root of unity of sufficiently large degree $\ell$ divisible by $2d$. Let $\fu$ be the small (finite-dimensional) quantum group associated to $q$ and the root datum $(Y,X,\ldots)$ as in ~\cite{l}. Let $\CC$ be the category of $X$-graded $\fu$-modules as defined in ~\cite{ajs}. Let $\CC^0$ be the block of $\CC$ containing the trivial $\fu$-module. B.Feigin and G.Lusztig conjectured (independently) that the category $\CC^0$ is equivalent to $\PS$. Let $\fU\supset\fu$ be the quantum group with divided powers associted to $q$ and the root datum $(Y,X,\ldots)$ as in ~\cite{l}. Let $\fC$ be the category of $X$-graded finite dimensional $\fU$-modules, and let $\fC^0$ be the block of $\fC$ containing the trivial $\fU$-module. The works ~\cite{kl}, ~\cite{l4} and ~\cite{kt} establish an equivalence of $\fC^0$ and the category $\CP(\CG,\bI)$. Here $\CG$ denotes the affine Grassmannian $\bG((z))/\bG[[z]]$, and $\CP(\CG,\bI)$ stands for the category of perverse sheaves on $\CG$ with finite-dimensional support constant along the orbits of $\bI$. \subsection{} \label{quantum res} The chapter 3 is devoted to the construction of the {\em convolution} functor $\bc_\CZ:\ \CP(\CG,\bI)\lra\PS$ which is the geometric counterpart of the restriction functor from $\fC^0$ to $\CC^0$, as suggested by V.Ginzburg (cf. ~\cite{gk} ~\S4). One of the main results of this chapter is the Theorem ~\ref{Satake} which is the sheaf-theoretic version of the classical Satake isomorphism. Recall that one has a {\em Frobenius homomorphism} $\fU\lra U(\fg^L)$ (see ~\cite{l}) where $U(\fg^L)$ stands for the universal enveloping algebra of the Langlands dual Lie algebra $\fg^L$. Thus the category of finite dimensional $\bG^L$-modules is naturally embedded into $\fC$ (and in fact, into $\fC^0$). On the geometric level this corresponds to the embedding $\CP(\CG,\bG[[z]]) \subset\CP(\CG,\bI)$. The Theorem ~\ref{Satake} gives a natural interpretation (suggested by V.Ginzburg) of the weight spaces of $\bG^L$-modules in terms of the composition $$\bG^L-mod\simeq\CP(\CG,\bG[[z]])\subset\CP(\CG,\bI)\stackrel{\bc_\CZ}{\lra} \PS.$$ \subsection{} Let us also mention here the following conjecture which might be known to specialists (characteristic $p$ analogue of conjecture in ~\ref{quantum}). Let $\bG^L$ stand for the Langlands dual Lie group. Let $p$ be a prime number bigger than the Coxeter number of $\fg^L$, and let $\overline\BF_p$ be the algebraic closure of finite field $\BF_p$. Let $\fC_p$ be the category of algebraic $\bG^L(\overline\BF_p)$-modules, and let $\fC^0_p$ be the block of $\fC_p$ containing the trivial module. Let $\CC_p$ be the category of graded modules over the Frobenius kernel of $\bG^L(\overline\BF_p)$, and let $\CC^0_p$ be the block of $\CC_p$ containing the trivial module (see ~\cite{ajs}). Finally, let $\PS_p$ be the category of snops {\em with coefficients in} $\overline\BF_p$, and let $\CP(\CG,\bI)_p$ be the category of perverse sheaves on $\CG$ constant along $\bI$-orbits {\em with coefficients in} $\overline\BF_p$. Then the categories $\CC^0_p$ and $\PS_p$ are equivalent, the categories $\fC^0_p$ and $\CP(\CG,\bI)_p$ are equivalent, and under these equivalences the restriction functor $\fC^0_p\lra\CC^0_p$ corresponds to the convolution functor $\CP(\CG,\bI)_p\lra\PS_p$ (cf. ~\ref{quantum res}). The equivalence $\CP(\CG,\bI)_p\iso\fC^0_p$ should be an extension of the equivalence between $\CP(\CG,\bG[[z]])_p\subset\CP(\CG,\bI)_p$ and the subcategory of $\fC^0_p$ formed by the $\bG^L(\overline\BF_p)$-modules which factor through the Frobenius homomorphism $Fr:\ \bG^L(\overline\BF_p)\lra\bG^L(\overline\BF_p)$. The latter equivalence is the subject of forthcoming paper of K.Vilonen and the second author. \subsection{} The Zastava space $\CZ$ organizing all the ``transversal slices" $\CZ^\alpha_\chi$ may seem cumbersome. At any rate the existence of various models of the slices $\CZ^\alpha_\chi$ (chapter 1), is undoubtedly beautiful by itself. Some of the wonderful properties of $\CQ^\al$ and $\CZ^\alpha_\chi$ are demonstrated in ~\cite{ku}, ~\cite{fk}, ~\cite{fkm} in the case $\bG=SL_n$. We expect all these properties to hold for the general $\bG$. \subsection{} To guide the patient reader through the notation, let us list the key points of this paper. The Theorem ~\ref{Z} identifies the different models of $\CZ^\alpha_\chi$ (all essentially due to V.Drinfeld) and states the factorization property. The exactness of the convolution functor $\bc_\CZ:\ \CP(\CG,\bI)\lra\PS$ is proved in the Theorem ~\ref{tough} and Corollary ~\ref{bunk}. The Theorem ~\ref{Satake} computes the value of the convolution functor on $\bG[[z]]$-equivariant sheaves modulo the parity vanishing conjecture ~\ref{parity}. \subsection{} In the next parts we plan to study $D$-modules on the local variety $\widetilde{\bf Q}$ (local construction of the category $\PS$, global sections as modules over affine Lie algebra $\hat\fg$, action of the affine Weyl group by Fourier transforms), the relation of the local and global varieties (local and global Whittaker sheaves, a version of the convolution functor twisted by a character of $N((z))$), and the sheaves on Drinfeld compactifications of maps into partial flag varieties. \subsection{} The present work owes its very existence to V.Drinfeld. It could not have appeared without the generous help of many people who shared their ideas with the authors. Thus, the idea of {\em factorization} (section 9) is due to V.Schechtman. A.Beilinson and V.Drinfeld taught us the {\em Pl\"ucker} picture of the (Beilinson-Drinfeld) affine Grassmannian (sections 6 and 10). G.Lusztig has computed the local singularities of the Schubert strata closures in the spaces $\CZ^\alpha_\chi$ (unpublished, cf ~\cite{l1}). B.Feigin and V.Ginzburg taught us their understanding of the Semiinfinite Flags for many years (in fact, we learnt of Drinfeld's Quasimaps' spaces from V.Ginzburg in the Summer 1995). A.Kuznetsov was always ready to help us whenever we were stuck in the geometric problems (in fact, for historical reasons, the section 3 has a lot in common with ~\cite{ku} \S1). We have also benefited from the discussions with R.Bezrukavnikov and M.Kapranov. Parts of this work were done while the authors were enjoying the hospitality and support of the University of Massachusetts at Amherst, the Independent Moscow University and the Sveu\v{c}ili\v{s}te u Zagrebu. It is a great pleasure to thank these institutions. \section{Notations} \subsection{} \label{group}{\bf Group $\bG$ and its Weyl group $\CW_f$.} We fix a Cartan datum $(I,\cdot)$ and a simply-connected simple root datum $(Y,X,\ldots)$ of finite type as in ~\cite{l}, ~2.2. Let $\bG$ be the corresponding simply-connected almost simple Lie group with the Cartan subgroup $\bH$ and the Borel subgroup $\bB\supset \bH$ corresponding to the set of simple roots $I\subset X$. We will denote by $\CR^+\subset X$ the set of positive roots. We will denote by $2\rho\in X$ the sum of all positive roots. Let $\bB_+=\bB$ and let $\bB_-\supset \bH$ be the opposite Borel subgroup. Let $\bN$ (resp. $\bN_-$) be the radical of $\bB$ (resp. $\bB_-$). Let $\bbH=\bB/\bN=\bB_-/\bN_-$ be the abstract Cartan group. The corresponding Lie algebras are denoted, respectively, by $\fb,\fb_-,\fn,\fn_-,\fh$. Let $\bX$ be the flag manifold $\bG/\bB$, and let $\bA=\bG/\bN$ be the principal affine space. We have canonically $H_2(\bX,\BZ)=Y;\ H^2(\bX,\BZ)=X$. For $\nu\in X$ let $\bL_\nu$ denote the corresponding $\bG$-equivariant line bundle on $\bX$. Let $\CW_f$ be the Weyl group of $\bG$. We have a canonical bijection $\bX^\bH=\CW_f$ such that the neutral element $e\in \CW_f=\bX^\bH\subset\bX$ forms a single $\bB$-orbit. We have a Schubert stratification of $\bX$ by $\bN$- (resp. $\bN_-$-)orbits: $\bX=\sqcup_{w\in \CW_f}\bX_w$ (resp. $\bX=\sqcup_{w\in \CW_f}\bX^w)$ such that for $w\in \CW_f=\bX^\bH\subset\bX$ we have $\bX^w\cap\bX_w=\{w\}$. We denote by $\ol\bX_w$ (resp. $\ol\bX^w$) the Schubert variety --- the closure of $\bX_w$ (resp. $\bX^w$). Note that $\ol\bX_w=\sqcup_{y\leq w} \bX_y$ while $\ol\bX^w=\sqcup_{z\geq w}\bX^z$ where $\leq$ denotes the standard Bruhat order on $\CW_f$. Let $e\in \CW_f$ be the shortest element (neutral element), let $w_0\in \CW_f$ be the longest element, and let $s_i,\ i\in I$, be the simple reflections in $\CW_f$. \subsection{} \label{reps} {\bf Irreducible representations of $\bG$.} We denote by $X^+$ the cone of positive weights (highest weights of finite dimensional $\bG$-modules). The fundamental weights $\omega_i:\ \langle i,\omega_j\rangle=\delta_{ij}$ form the basis of $X^+$. For $\lambda\in X^+$ we denote by $V_\lambda$ the finite dimensional irreducible representation of $\bG$ with highest weight $\lambda$. We denote by $V_\lambda^\vee$ the representation dual to $V_\lambda$; the pairing: $V_\lambda^\vee\times V_\lambda\lra\BC$ is denoted by $\langle,\rangle$. For each $\lambda\in X^+$ we choose a nonzero vector $y_\lambda\in V_\lambda^{\bN_-}$. We also choose a nonzero vector $x_\lambda\in (V_\lambda^\vee)^\bN$ such that $\langle x_\lambda,y_\lambda\rangle=1$. \subsection{} \label{config} {\bf Configurations of $I$-colored divisors.} Let us fix $\alpha\in\BN[I]\subset Y,\ \alpha=\sum_{i\in I}a_ii$. Given a curve $C$ we consider the configuration space $C^\alpha \df\prod_{i\in I} C^{(a_i)}$ of colored effective divisors of multidegree $\alpha$ (the set of colors is $I$). The dimension of $C^\alpha$ is equal to the length $|\alpha|=\sum_{i\in I}a_i$. Multisubsets of a set $S$ are defined as elements of some symmetric power $S^{(k)}$ and we denote the image of $(s_1,...,s_k)\in S^k$ in $S^{(k)}$ by $\{\{s_1,...,s_k\}\}$. We denote by $\fP(\alpha)$ the set of all partitions of $\alpha$, i.e multisubsets $\Ga= \{\{\ga_1,...,\ga_k\}\}$ of $\BN[I]$ with $\gamma_r\not=0$ and $\sum_{r=1}^k \ga_i=\al$. For $\Gamma\in\fP(\alpha)$ the corresponding stratum $C^\alpha_\Gamma$ is defined as follows. It is formed by configurations which can be subdivided into $m$ groups of points, the $r$-th group containing $\gamma_r$ points; all the points in one group equal to each other, the different groups being disjoint. For example, the main diagonal in $C^\alpha$ is the closed stratum given by partition $\alpha=\alpha$, while the complement to all diagonals in $C^\alpha$ is the open stratum given by partition $$ \alpha=\sum_{i\in I}(\underbrace{i_k+i_k+\ldots+i_k}_{a_k\operatorname{ times}}) $$ Evidently, $C^\alpha=\bigsqcup\limits_{\Gamma\in\fP(\alpha)}C^\alpha_\Gamma$. \bigskip \centerline{\bf CHAPTER 1. The spaces $Q$ and $Z$} \section{Quasimaps from a curve to a flag manifold} \subsection{} We fix a smooth projective curve $C$ and $\alpha\in\BN[I]$. \subsubsection{Definition} An algebraic map $f:\ C\lra\bX$ has degree $\alpha$ if the following equivalent conditions hold: a) For the fundamental class $[C]\in H_2(C,\BZ)$ we have $f_*[C]=\alpha\in Y=H_2(\bX,\BZ)$; b) For any $\nu\in X$ the line bundle $f^*\bL_\nu$ on $C$ has degree $\langle\alpha,\nu\rangle$. \subsection{} \label{maps} The Pl\"ucker embedding of the flag manifold $\bX$ gives rise to the following interpretation of algebraic maps of degree $\al$. For any irreducible $V_\lambda$ we consider the trivial vector bundle $\CV_\lambda=V_\lambda\otimes\CO$ over $C$. For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ we denote by the same letter the induced morphism $\phi:\ \CV_\lambda\otimes \CV_\mu \lra \CV_\nu$. Then a map of degree $\al$ is a collection of {\em line subbundles} $\fL_\lambda\subset\CV_\lambda,\ \lambda\in X^+$ such that: a) $\deg\fL_\lambda=-\langle\alpha,\lambda\rangle$; b) For any surjective $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$ we have $\phi(\fL_\lambda\otimes\fL_\mu)= \fL_\nu$; c) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(\fL_\lambda\otimes\fL_\mu)=0$. Since the surjections $V_\la\ten V_\mu\ra V_{\la+\mu}$ form one $\cs$-orbit, systems $\LL_\la$ satisfying (b) are determined by a choice of $\fL_{\om_i}$ for the fundamental weights $\om_i,\ i\in I$. If we replace the curve $C$ by a point, we get the Pl\"ucker description of the flag variety $\bX$ as the set of collections of lines $L_\la\sub V_\la$ satisfying conditions of type (b) and (c). Here, a Borel subgroup $B$ in $\bX$ corresponds to a system of lines $(L_\la,\ \la\in X^+)$ if the lines are the fixed points of the unipotent radical $N$ of $B$, $L_\la=(V_\la)^N$, or equivalently, if $N$ is the common stabilizer for all lines $N=\bb{\la\in X^+}\cap G_{L_\la}$. The space of degree $\alpha$ quasimaps from $C$ to $\bX$ will be denoted by $\qc^\alpha$. \subsection{Definition} \label{quasimaps} (V.Drinfeld) The space $\CQ^\al=\CQ^\al_C$ of {\em quasimaps} of degree $\alpha$ from $C$ to $\bX$ is the space of collections of {\em invertible subsheaves} $\fL_\lambda\subset\CV_\lambda,\ \lambda\in X^+$ such that: a) $\deg\fL_\lambda=-\langle\alpha,\lambda\rangle$; b) For any surjective $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$ we have $\phi(\fL_\lambda\otimes\fL_\mu)= \fL_\nu$; c) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(\fL_\lambda\otimes\fL_\mu)=0$. \subsubsection{Lemma} a) The evident inclusion $\qc^\alpha\subset\CQ^\alpha$ is an open embedding; b) $\CQ^\alpha$ is a projective variety. {\em Proof.} Obvious. $\Box$ \subsubsection{} \label{praf} Here is another version of the Definition, also due to V.Drinfeld. The principal affine space $\bA=\bG/\bN$ is an $\bH_a$-torsor over $\bX$. We consider its affine closure $\bbA$, that is, the spectrum of the ring of functions on $\bA$. Recall that $\bbA$ is the space of collections of vectors $v_\lambda\in V_\lambda,\ \lambda\in X^+$, satisfying the following Pl\"ucker relations: a) For any surjective $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$, and $\phi(y_\lambda\otimes y_\mu)=y_\nu$, we have $\phi(v_\lambda\otimes v_\mu)=v_\nu$; b) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(v_\lambda\otimes v_\mu)=0$. The action of $\bH_a$ extends to $\bbA$ but it is not free anymore. Consider the quotient stack $\hat\bX=\bbA/\bH_a$. The flag variety $\bX$ is an open substack in $\hat\bX$. A map $\hat{\phi}:\ C\to\hat\bX$ is nothing else than an $\bH_a$-torsor $\Phi$ over $C$ along with an $\bH_a$-equivariant morphism $f:\ \Phi\to\bbA$. The degree of this map is defined as follows. Let $\lambda:\ \bH_a\to\BC^*$ be the character of $\bH_a$ corresponding to a weight $\lambda\in X$. Let $\bH_\lambda\subset \bH_a$ be the kernel of the morphism $\lambda$. Consider the induced $\BC^*$-torsor $\Phi_\lambda=\Phi/\bH_\lambda$ over $C$. The map $\hat\phi$ has degree $\alpha\in\BN[I]$ if $$ \text{for any }\lambda\in X\quad\text{we have}\quad \deg(\Phi_\lambda)=\langle\lambda,\alpha\rangle. $$ {\bf Definition.} The space $\CQ^\alpha$ is the space of maps $\hat{\phi}:\ C\to\hat\bX$ of degree $\alpha$ such that the generic point of $C$ maps into $\bX\subset\hat\bX$. The equivalence of the two versions of Definition follows by comparing their Pl\"ucker descriptions. \subsection{} In this subsection we describe a stratification of $\CQ^\alpha$ according to the singularities of quasimaps. \subsubsection{} \label{sigma} Given $\beta,\gamma\in\BN[I]$ such that $\beta+\gamma=\alpha$, we define the proper map $ \sigma_{\beta,\gamma}:\ \CQ^\beta \times C^\gamma \lra \CQ^\alpha$. Namely, let $f=(\fL_\lambda)_{\lambda\in X^+}\in\ \CQ^\beta$ be a quasimap of degree $\beta$; and let $ D=\sum_{i\in I}D_i \cdd i $ be an effective colored divisor of multidegree $\gamma=\sum_{i\in I}d_ii$, that is, $\deg(D_i)=d_i$. We define $\sigma_{\beta,\gamma}(f,D)\df f(-D) \df ( \fL_\lambda(-\langle D,\lambda\rangle ) )_{\lambda\in X^+} \in\CQ^\alpha ,$ where we use the pairing $Div^I(C)\bb{\Z}\ten X\ra Div(C)$ given by $\langle D,\lambda \rangle = \sum_{i\in I}\langle i,\lambda\rangle \cdd D_i$. \subsubsection{} \label{strat M} {\bf Theorem.} ${\displaystyle \ \CQ^\alpha=\bigsqcup_{0\leq\beta\leq\alpha} \sigma_{\beta,\alpha-\beta}(\qc^\beta\times C^{\alpha-\beta})}$ {\em Proof.} Any invertible subsheaf $\fL_\la\sub \VV_\la$ lies in a unique line subbundle $\ti\fL_\la\sub \VV_\la$ called the {\em normalization} of $\fL$. So any quasimap $\fL$ defines a map $\ti\fL$ (called the {\em normalization} of $\fL$) of degree $\be\le\al$ and an $I$-colored effective divisor $D$ (called the {\em defect} of $\fL$) corresponding to the torsion sheaf $\ti\fL/\fL$, such that $\fL=\ti\fL(-D)$. $\Box$ \subsubsection{Definition} \label{domain} Given a quasimap $f=(\fL_\lambda)_{\lambda\in X^+} \in\ \CQ^\alpha$, its {\em domain of definition} $U(f)$ is the maximal Zariski open $U(f)\subset C$ such that for any $\lambda$ the invertible subsheaf $\fL_\lambda\subset\CV_\lambda$ restricted to $U(f)$ is actually a line subbundle. \subsubsection{Corollary} \label{big domain} For a quasimap $f=(\fL_\lambda)_{\lambda\in X^+}\in\ \CQ^\alpha$ of degree $\alpha$ the complement $C-U(f)$ of its domain of definition consists of at most $|\alpha|$ points. $\Box$ \subsection{} \label{C=line} From now on, unless explicitly stated otherwise, $C=\BP^1$. {\bf Proposition.} (V.Drinfeld) $\qc^\alpha$ is a smooth manifold of dimension $2|\alpha|+\dim(\bX)$. {\em Proof.} We have to check that at a map $f\in\qc^\alpha$ the first cohomology $H^1(\BP^1,f^*\CT\bX)$ vanishes (where $\CT\bX$ stands for the tangent bundle of $\bX$), and then the tangent space $\Theta_f\qc^\alpha$ equals $H^0(\BP^1,f^*\CT\bX)$. As $\CT\bX$ is generated by the global sections, $f^*\CT\bX$ is generated by global sections as well, hence $H^1(\BP^1,f^*\CT\bX)=0$. To compute the dimension of $\Theta_f\qc^\alpha= H^0(\BP^1,f^*\CT\bX)$ it remains to compute the Euler characteristic $\chi(\BP^1,f^*\CT\bX)$. To this end we may replace $\CT\bX$ with its associated graded bundle $\oplus_{\theta\in\CR^+}\bL_\theta$. Then $$ \chi(\BP^1,f^*(\bigoplus_{\theta\in\CR^+}\bL_\theta))=\sum_{\theta\in\CR^+} (\langle\alpha,\theta\rangle+1)=\langle\alpha,2\rho\rangle+\sharp\CR^+= 2|\alpha|+\dim\bX$$ $\Box$ \subsection{} \label{CZ} Now we are able to introduce our main character. First we consider the open subspace $U^\alpha\subset\ \CQ^\alpha$ formed by the quasimaps containing $\infty\in\BP^1$ in their domain of definition (see ~\ref{domain}). Next we define the closed subspace $\CZ^\alpha\subset U^\alpha$ formed by quasimaps with value at $\infty$ equal to $\bB_-\in\bX$: $$\CZ^\alpha\df\ \{f\in U^\alpha | f(\infty)=\bB_-\}$$ We will see below that $\CZ^\alpha$ is an affine algebraic variety. \subsubsection{} \label{dimension} It follows from Proposition ~\ref{C=line} that $\dim\CZ^\alpha=2|\alpha|$. \section{Local Flag space} In this section we define a version of $\CQ^\alpha$ where one replaces the global curve $C$ by the formal neighbourhood of a point. \subsection{} \label{SS} We set $\CO=\BC[[z]]\stackrel{p_n}{\lra}\CO_n=\BC[[z]]/z^n,\CK=\BC((z))$. We define the scheme $\bbA(\CO)$ (of infinite type): its points are the collections of vectors $v_\lambda\in V_\lambda\otimes\CO,\ \lambda\in X^+$, satisfying the Pl\"ucker equations like in ~\ref{praf}. It is a closed subscheme of $\prod_{i\in I}V_{\omega_i}\otimes\CO$. We define the open subscheme $\bbA(\CO)_n\subset\bbA(\CO)$: it is formed by the collections $(v_\lambda)_{\lambda\in X^+}$ such that $p_n(v_{\omega_i}) \not=0$ for all $i\in I$. Evidently, for $0\leq n\leq m$, one has $\bbA(\CO)_n\subset\bbA(\CO)_m$. We define the open subscheme $\CS\subset\bbA(\CO)$ as the union $\bigcup_{n\geq0}\bbA(\CO)_n$. One has $\CS=\bA(\CO)$. The scheme $\CS$ is equipped with the free action of $\bH_a:\ h(v_\lambda)_{\lambda\in X^+}=(\lambda(h)v_\lambda)_{\lambda\in X^+}$. The quotient scheme $\bQ=\CS/\bH_a$ is a closed subscheme in $\prod_{i\in I}\BP(V_{\omega_i}\otimes\CO)$. It is formed by the collections of lines satisfying the Pl\"ucker equations. We denote the natural projection $\CS\lra\bQ$ by $pr$. \subsection{} \label{SSeta} For $\eta\in\BN[I]$ we define the closed subscheme $\CS^{-\eta}\subset\CS$ formed by the collections $(v_\lambda)_{\lambda\in X^+}$ such that $v_\lambda=0\ \modul\ z^{\langle\eta,\lambda\rangle}$. We have the natural isomorphism $\CS\iso\CS^{-\eta},\ (v_\lambda)_{\lambda\in X^+}\mapsto (z^{\langle\eta,\lambda\rangle}v_\lambda)_{\lambda\in X^+}$. Now we can extend the definition of $\CS^\chi$ to arbitrary $\chi\in Y$. Namely, we define $\CS^\chi$ to be formed by the collections $(v_\lambda\in V_\lambda\otimes\CK)_{\lambda\in X^+}$ such that $(z^{\langle\chi,\lambda\rangle}v_\lambda)_{\lambda\in X^+}\in\CS$. Evidently, $\CS^\chi\subset\CS^\eta$ iff $\chi\leq\eta$, and then the inclusion is the closed embedding. The open subscheme $\CS^\eta- \bigcup_{\chi<\eta}\CS^\chi\subset\CS^\eta$ will be denoted by $\dCS^\eta\subset\CS^\eta$. The ind-scheme $\bigcup_{\eta\in Y}\CS^\eta$ will be denoted by $\tCS$. The ind-scheme $\tCS$ is equipped with the natural action of the proalgebraic group $\bG(\CO)$ (coming from the action on $\prod_{i\in I}V_{\omega_i}\otimes \CK$), and the orbits are exactly $\dCS^\eta,\ \eta\in Y$. \subsection{} \label{QQ} All the above (ind-)schemes are equipped with the free action of $\bH_a$, and taking quotients we obtain the schemes $\bQ^\eta=\CS^\eta/\bH_a,\ \eta\in Y$. They are all closed subschemes of the ind-scheme $\prod_{i\in I}\BP(V_{\omega_i}\otimes\CK)$. We have $\bQ^\chi\subset\bQ^\eta$ iff $\chi\leq\eta$, and then the inclusion is the closed embedding. The ind-scheme $\tbQ=\tCS/\bH_a$ is the union $\tbQ=\bigcup_{\eta\in Y}\bQ^\eta$. The ind-scheme $\tbQ$ is equipped with the natural action of the proalgebraic group $\bG(\CO)$ (coming from the action on $\prod_{i\in I}\BP(V_{\omega_i}\otimes\CK)$), and the orbits are exactly $\dbQ^\eta=\dCS^\eta/\bH_a,\ \eta\in Y$. \subsection{} We consider $C=\BP^1$ with two marked points $0,\infty\in C$. We choose a coordinate $z$ on $C$ such that $z(0)=0,z(\infty)=\infty$. \subsubsection{} For $\alpha\in\BN[I]$ we define the space $\hCQ^\alpha\stackrel{pr}{\lra} \CQ^\alpha$ formed by the collections $(v_\lambda\in\fL_\lambda\subset \CV_\lambda)_{\lambda\in X^+}$ such that a) $(\fL_\lambda\subset\CV_\lambda)_{\lambda\in X^+}\in\CQ^\alpha$; b) $v_\lambda$ is a regular nonvanishing section of $\fL_\lambda$ on $\BA^1=\BP^1-\infty$; c) $(v_\lambda)_{\lambda\in X^+}$ satisfy the Pl\"ucker equations like in ~\ref{praf}. It is easy to see that $\hCQ^\alpha\stackrel{pr}{\lra}\CQ^\alpha$ is a $\bH_a$-torsor: $h(v_\lambda,\fL_\lambda)=(\lambda(h)v_\lambda,\fL_\lambda)$. \subsubsection{} \label{m} Taking a formal expansion at $0\in C$ we obtain the closed embedding $\fs_\alpha:\ \hCQ^\alpha\hookrightarrow\CS$. Evidently, $\fs_\alpha$ is compatible with the $\bH_a$-action, so it descends to the same named closed embedding $\fs_\alpha:\ \CQ^\alpha\hookrightarrow\bQ$. \subsubsection{Lemma} \label{codime} Let $\beta\in\BN[I]$. Then $\codim_\bQ\bQ^{-\beta}\geq2|\beta|$. {\em Proof.} Choose $\alpha\geq\beta$, and consider the closed embedding $\fs_\alpha:\ \CQ^\alpha\hookrightarrow\bQ$. Then $\fs_\alpha^{-1}(\bQ^{-\beta})=\CQ^{\alpha-\beta}$ embedded into $\CQ^\alpha$ as follows: $(\fL_\lambda\subset\CV_\lambda)_{\lambda\in X^+} \mapsto(\fL_\lambda(-\langle\beta,\lambda\rangle0) \subset\CV_\lambda)_{\lambda\in X^+}$. Now $\codim_\bQ\bQ^{-\beta}\geq\codim_{\CQ^\alpha}\CQ^{\alpha-\beta}=2|\beta|$. $\Box$ \section{Pl\"ucker sections} In this section we describe another model of the space $\CZ^\alpha$ introduced in ~\ref{CZ}. \subsection{} \label{polynom} We fix a coordinate $z$ on the affine line $\BA^1=\BP^1-\infty$. We will also view the configuration space $\BA^\alpha\df(\BA^1)^\alpha$ (see ~\ref{config}) as the space of collections of unitary polynomials $(Q_\lambda)_{\lambda\in X^+}$ in $z$, such that (a) $\deg(Q_\lambda)=\langle\alpha,\lambda\rangle$, and (b) $Q_{\lambda+\mu}=Q_\lambda Q_\mu$. \subsection{} \label{fZ} Recall the notations of ~\ref{reps}. For each $\lambda\in X^+$ we will use the decomposition $V_\lambda=\BC y_\la\oplus(\Ker x_\la)= (V_\lambda)^\bN\oplus\fn_- V_\lambda$, compatible with the action of $\fh=\fb_-\cap\fb$, i.e., with the weight decomposition. For a section $v_\lambda\in\Ga(\BA^1,\CV_\lambda)= V_\la\ten\BC[z]\df V_\la[z]$, we will use a polynomial $Q_\la\df\ \langle x_\lambda,v_\lambda\rangle\in\BC[z]$, to write down the decomposition $v_\la=Q_\la\cdd y_\la\pl {v''}_\la\in \BC[z]\cdd y_\la\pl(\Ker x_\la)[z] = V_\la[z]$. {\bf Definition.} (V.Drinfeld) The space $\sZ^\alpha$ of {\em Pl\"ucker sections} of degree $\alpha$ is the space of collections of sections $v_\lambda\in\Ga(\BA^1,\CV_\lambda)= V_\la\ten\BC[z]\df V_\la[z] ,\ \lambda\in X^+$; such that for $v_\la=Q_\la\cdd y_\la\pl {v''}_\la\in \BC[z]\cdd y_\la\pl(\Ker x_\la)[z]$, one has a) Polynomial $Q_\lambda$ is unitary of degree $\langle\alpha,\lambda\rangle$; b) Component ${v''}_\la$ of $v_\la$ in $(\Ker x_\la)[z]$ has degree strictly less than $\langle\alpha,\lambda\rangle$; c) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$ and $\phi^\vee(x_\nu)=x_\lambda\otimes x_\mu$ we have $\phi(v_\lambda\otimes v_\mu)=v_\nu$; d) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(v_\lambda\otimes v_\mu)=0$. \subsubsection{} \label{affine} Collections $(v_\lambda)_{\lambda\in X^+}$ that satisfy (c), are determined by a choice of $v_{\omega_i},\ i\in I$. Hence $\sZ^\alpha$ is an affine algebraic variety. \subsubsection{} \label{pi} Due to the properties a),c) above, the collection of polynomials $Q_\lambda$ defined in a) satisfies the conditions of ~\ref{polynom}. Hence we have the map $$\pi_\alpha:\ \sZ^\alpha\lra\BA^\alpha$$ \section{Beilinson-Drinfeld Grassmannian} In this section we describe yet another model of the space $\CZ^\alpha$ introduced in ~\ref{CZ}. \subsection{} Let $C$ be an arbitrary smooth projective curve; let $\CT$ be a left $\bG$-torsor over $C$, and let $\tau$ be a section of $\CT$ defined over a Zariski open subset $U\subset C$, i.e., a trivialization of $\CT$ over $U$. We will define a $\bB$- (resp. $\bB_-$-) type $d(\tau)$ (resp. $d_-(\tau)$): a measure of singularity of $\tau$ at $C-U$. \subsubsection{} \label{type} Section $\tau$ defines a $\bB$-subtorsor $\bB\cdd\tau\sub\CT$. This reduction of $\CT$ to $\bB$ over $U$ is the same as a section of $\bB\bss\CT$ over $U$. Since $\bG/\bB$ is proper, this reduction (i.e. section), extends uniquely to the whole $C$. Thus we obtain a $\bB$-subtorsor $\barr{\bB\cdd\tau}\sub\CT$ (the closure of $\bB\cdd\tau\sub\CT|U$ in $\CT$), equipped with a section $\tau$ defined over $U$. Using the projection $\bB\lra\bbH$ we can induce $\barr{\bB\cdd\tau}$ to a torsor over $C$ for the abstract Cartan group $\bbH\cong \bB/\bN$ of $\bG$; namely, $\CT_{\tau,\bB}\df \bN\bss \barr{\bB\cdd\tau}$, equipped with a section $\tau_\bB$ defined over $U$. The choice of simple coroots (cocharacters of $\bbH$) $I\subset Y$ identifies $\bbH$ with $(\BC^*)^I$. Thus the section $\tau_\bB$ of $\CT_{\tau,\bB}$ produces an $I$-colored divisor $d(\tau)$ supported at $C-U$. We will call $d(\tau)$ the $\bB$-$type$ of $\tau$. Replacing $\bB$ by $\bB_-$ in the above construction we define the $\bB_-$-$type$ $d_-(\tau)$. \subsection{} \label{bZ} Recall that A.Beilinson and V.Drinfeld have introduced the {\em relative Grassmannian} $\CG_C^{(n)}$ over $C^n$ for any $n\in\BN$ (see ~\cite{todisappear}): its fiber $p_n^{-1}(x_1,\ldots,x_n)$ over an $n$-tuple $(x_1,\ldots,x_n)\in C^n$ is the space of isomorphism classes of $\bG$-torsors $\CT$ equipped with a section $\tau$ defined over $C-\{x_1,\ldots,x_n\}$. We will consider a certain finite-dimensional subspace of a partialy symmetrized version of the relative Grassmannian. {\bf Definition.} (A.Beilinson and V.Drinfeld) $\bZ^\alpha$ is the space of isomorphism classes of the following data: a) an $I$-colored effective divisor $D\in\BA^\alpha$; b) $\bG$-torsor $\CT$ over $\BP^1$ equipped with a section $\tau$ defined over $\BP^1-supp(D)$ such that: i) $\bB$-type $d(\tau)=0$; ii) $\bB_-$-type $d_-(\tau)$ is a negative divisor (opposite to effective) such that $d_-(\tau)+D$ is effective. \subsubsection{} \label{bZU} By the definition, the space $\bZ^\alpha$ is equipped with a projection $p_\alpha$ to $\BA^\alpha:\ (D,\CT,\tau)\mapsto D$. For a subset $U\subset\BA^1$ we will denote by $\bZ_U^\alpha$ the preimage $p_\alpha^{-1}(U)$. \subsubsection{} The reader may find another realization of $\bZ^\alpha$ in ~\ref{PBD} below. \subsection{} In this subsection we will formulate the crucial {\em factorization} property of $\bZ^\alpha$. \subsubsection{} Recall the following property of the Beilinson-Drinfeld relative Grassmannian $\CG_C^{(n)}\overset{p_n}{\lra}C^n$ (see ~\cite{todisappear}). Suppose an $n$-tuple $(x_1,\ldots,x_n)\in C^n$ is represented as a union of an $m$-tuple $(y_1,\ldots,y_m)\in C^m$ and a $k$-tuple $(z_1,\ldots,z_k)\in C^k,\ k+m=n$, such that all the points of the $m$-tuple are disjoint from all the points of the $k$-tuple. Then $p_n^{-1}(x_1,\ldots,x_n)$ is canonically isomorphic to the product $p_m^{-1}(y_1,\ldots,y_m)\times p_k^{-1}(z_1,\ldots,z_k)$ \subsubsection{} \label{factorization} Suppose we are given a decomposition $\alpha=\beta+\gamma,\ \beta,\gamma \in\BN[I]$ and two disjoint subsets $U,\Upsilon\subset\BA^1$. Then $U^\beta\times\Upsilon^\gamma$ lies in $\BA^\alpha$, and we will denote the preimage $p_\alpha^{-1}(U^\beta\times\Upsilon^\gamma)$ in $\bZ^\alpha$ by $\bZ^{\beta,\gamma}_{U,\Upsilon}= \bZ^\al|_{(U^\beta\times\Upsilon^\gamma)}$ (cf. ~\ref{bZU}). The above property of relative Grassmannian immediately implies the following {\bf Factorization property.} There is a canonical factorization isomorphism $\bZ^{\beta,\gamma}_{U,\Upsilon}\cong\bZ^\beta_U \times\bZ^\gamma_\Upsilon$, i.e., $$ \bZ^\al |_{(U^\beta\times\Upsilon^\gamma)}\cong \bZ^\be|_{U^\beta} \times \bZ^\ga|_{\Upsilon^\gamma} .$$ \subsection{Remark} Let us describe the fibers of $p_\alpha$ in terms of the normal slices to the semiinfinite Schubert cells in the loop Grassmannian. \subsubsection{} \label{Iwasawa} Let $\CG$ be the usual affine Grassmannian $\bG((z))/\bG[[z]]$. It is naturally identified with the fiber of $\CG^{(1)}_{\BP^1}$ over the point $0\in\BP^1$. Due to the Iwasawa decomposition in p-adic groups, there is a natural bijection between $Y$ and the set of orbits of the group $\bN((z))$ (resp. $\bN_-((z))$) in $\CG$; for $\gamma\in Y$ we will denote the corresponding orbit by $S_\gamma$ (resp. $T_\gamma$). We will denote by $\ol{T}_\gamma$ the ``closure'' of $T_\gamma$, that is, the union $\cup_{\beta\geq\gamma}T_\gamma$. It is proved in ~\cite{mv} that the intersection $\ol{T}_\gamma\cap S_\beta$ is not empty iff $\gamma\leq\beta$. Then it is an affine algebraic variety, a kind of a normal slice to $T_\be$ in $\barr T_\ga$. Let us call it $TS_{\gamma,\beta} \df \ol{T}_\gamma\cap S_\beta$ for short. If rank$(\bG)>1$ then $TS_{\gamma,\beta}= \ol{T}_\gamma\cap S_\beta$ is not necessarily irreducible. But it is always equidimensional of dimension $|\beta-\gamma|$. There is a natural bijection between the set of irreducible components of $TS_{\gamma,\beta}=\ol{T}_\gamma\cap S_\beta$ and the canonical basis of $U^+_{\beta-\gamma}$ (the weight $\beta-\gamma$ component of the quantum universal enveloping algebra of $\fn$) (see ~\cite{l} for the definition of canonical basis of $U^+$). \subsubsection{} Recall the diagonal stratification of $\BA^\alpha$ defined in ~\ref{config} and the map $p_\al:\bZ^\al\ra\BA^\alpha$. We consider a partition $\Gamma:\ \alpha=\sum_{k=1}^m\gamma_k$ and a divisor $D$ in the stratum $\BA_\Gamma^\alpha$. The interested reader will check readily the following {\em Claim.} $p_\alpha^{-1}(D)$ is isomorphic to the product $\prod_{k=1}^m TS_{-\gamma_k,0} =\prod_{k=1}^m \ol{T}_{-\ga_k}\cap S_0\cong \prod_{k=1}^m \ol{T}_0\cap S_{\ga_k}$. In particular, the fiber over a point in the closed stratum is isomorphic to $TS_{-\alpha,0}= \ol{T}_{-\al}\cap S_0 \cong \ol{T}_0\cap S_\al $, while the fiber over a generic point is isomorphic to the product of affine lines $TS_{-i,0}\cong \ol{T}_0\cap S_{-i}\cong\BA^1$, that is, the affine space $\BA^{|\alpha|}$. \subsubsection{Corollary} \label{irred} $\bZ^\alpha$ is irreducible. \section{Equivalence of the three constructions} \subsection{} In this subsection we construct an isomorphism $\varpi:\ \CZ^\alpha\iso\sZ^\alpha$, i.e., from the subsheaves $\fL_\la\sub\CV_\la$ we construct the sections $v_\la\in\Ga(\BA^1,\CV_\la)$. \subsubsection{} \label{tuda} Let $f\in\CZ^\alpha$ be a quasimap given by a collection $(\fL_\lambda \subset\CV_\lambda=V_\lambda\otimes\CO_{\BP^1})_{\lambda\in X^+}$. Since $\fL_\lambda|_{\BA^1}$ is trivial, it has a unique up to proportionality section $v_\lambda$ generating it over $\BA^1$. We claim that the pairing $\langle x_\lambda,v_\lambda\rangle$ does not vanish identically. In effect, since $\deg(f)=\alpha$, the meromorphic section $\frac{v_\lambda}{z^{\langle\alpha,\lambda\rangle}}$ of $\CV_\lambda$ is regular nonvanishing at $\infty\in\BP^1$. Moreover, since $f(\infty)=\bB_-$, we have $\frac{v_\lambda}{z^{\langle\alpha,\lambda\rangle}}(\infty)\in V_\lambda^{\bN_-}$. Thus, $\langle x_\lambda,\frac{v_\lambda}{z^{\langle\alpha,\lambda\rangle}} \rangle(\infty)\not=0$. Now we can normalize $v_\lambda$ (so far defined up to a multiplication by a constant) by the condition that $\langle x_\lambda,v_\lambda\rangle$ is a unitary polynomial. Let us denote this polynomial by $Q_\lambda$. It has degree $d_\lambda\leq\langle\alpha,\lambda\rangle$ since $\deg(f)=\alpha$. Since $\frac{v_\lambda}{z^{\langle\alpha,\lambda\rangle}}(\infty)\in V_\lambda^{\bN_-}$, we see that $\deg\langle e,v_\lambda\rangle<d_\lambda$ for any $e\perp y_\la$. Moreover, since $\deg(f)=\alpha$ we must then have $d_\lambda=\langle\alpha,\lambda\rangle$. Thus we have checked that the collection $(v_\lambda)_{\lambda\in X^+}$ satisfies the conditions a),b) of the Definition ~\ref{fZ}. The conditions c),d) of {\em loc. cit.} follow from the conditions b),c) of the Definition ~\ref{quasimaps}. In other words, we have defined the Pl\"ucker section $$\varpi(f)\df\ (v_\lambda)_{\lambda\in X^+}\in\sZ^\alpha$$ \subsubsection{} \label{obratno} Here is the inverse construction. Given a Pl\"ucker section $(v_\lambda)_{\lambda\in X^+}\in\sZ^\alpha$ we define the corresponding quasimap $f=(\fL_\lambda)_{\lambda\in X^+}\in\CZ^\alpha$ as follows. We can view $v_\lambda$ as a regular section of $\CV_\lambda(\langle\alpha,\lambda\rangle\infty)$ over the whole $\BP^1$. It generates an invertible subsheaf $\fL_\lambda'\subset \CV_\lambda(\langle\alpha,\lambda\rangle\infty)$. We define $$\fL_\lambda\df\ \fL_\lambda'(-\langle\alpha,\lambda\rangle\infty) \subset\CV_\lambda$$ \subsubsection{} It is immediate to see that the above constructions are inverse to each other, so that $\varpi:\ \CZ^\alpha\lra\sZ^\alpha$ is an isomorphism. \subsubsection{Remark} Note that the definition of the space $\CZ^\alpha$ depends only on the choice of Borel subgroup $\bB_-\subset \bG$, while the definition of $\sZ^\alpha$ depends also on the choice of the opposite Borel subgroup $\bB\subset \bG$ or, equivalently, on the choice of the Cartan subgroup $\bH\subset \bB_-$. We want to stress that the projection $\pi_\alpha:\ \CZ^\alpha=\sZ^\alpha\lra \BA^\alpha$ {\em does depend} on the choice of $\bB$. Let us describe $\pi_\alpha\varpi(f)$ for a genuine map (as opposed to quasimap) $f\in\CZ^\alpha$. To this end recall (see ~\ref{group}) that the $\bB$-invariant Schubert varieties $\ol\bX_{s_iw_0},\ i\in I$, are divisors in $\bX$. Their formal sum may be viewed as an $I$-colored divisor $\fD$ in $\bX$. Then $f^*\fD$ is a well defined $I$-colored divisor on $\BP^1$ since $f(\BP^1)\not\subset\fD$ since $f(\infty)=\bB_-\in\bX_{w_0}$. For the same reason the point $\infty$ does not lie in $f^*\fD$, so $f^*\fD$ is really a divisor in $\BA^1$. It is easy to see that $f^*\fD\in\BA^\alpha$ and $f^*\fD=\pi_\alpha\varpi(f)$. \subsection{} In this subsection we construct an isomorphism $\xi:\ \sZ^\alpha\iso\bZ^\alpha$, so from a system of sections $v_\la$ we construct a $\bG$-torsor $\CT$ with a section $\tau$ and an $I$-colored divisor $D$. \subsubsection{Lemma}({\em The Pl\"ucker picture of $\bG$.}) \label{drinf} The map $\psi:\ g\mapsto (gx_\lambda,gy_\lambda)_{\lambda\in X^+}$ is a bijection between $\bG$ and the space of collections $\{(u_\lambda\in V_\lambda^\vee, \upsilon_\lambda\in V_\lambda)_{\lambda\in X^+})\}$ satisfying the following conditions: a) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$ and $\phi^\vee(x_\nu)=x_\lambda\otimes x_\mu$ we have $\phi(\upsilon_\lambda\otimes \upsilon_\mu)=\upsilon_\nu$; b) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(\upsilon_\lambda\otimes\upsilon_\mu)=0$; c) For any $\bG$-morphism $\varphi:\ V_\lambda^\vee\otimes V_\mu^\vee\lra V_\nu^\vee$ such that $\nu=\lambda+\mu$ and $\varphi(x_\lambda\otimes x_\mu)=x_\nu$ we have $\varphi(u_\lambda\otimes u_\mu)=u_\nu$; d) For any $\bG$-morphism $\varphi:\ V_\lambda^\vee\otimes V_\mu^\vee\lra V_\nu^\vee$ such that $\nu<\lambda+\mu$ we have $\varphi(u_\lambda\otimes u_\mu)=0$; e) $\langle u_\lambda,\upsilon_\lambda\rangle=1$. {\em Proof.} We are considering the systems $(v,u)=\left(v_\la\in V_\lambda,\ u_\la\in V_\lambda^\vee, \ {\la\in X^+}\right)$ such that both $v$ and $u$ are Pl\"ucker sections and $\langle v,u\rangle =1$, i.e., $\langle v_\la,u_\la\rangle =1$ for each $\la$. These form a $\bG$-torsor and we have fixed its element $(y,x)$, which we will use to think of this torsor as a Pl\"ucker picture of $\bG$. The stabilizers $\bG_v$ and $\bG_u$ are the unipotent radicals of the opposite Borel subgroups, for instance $\bG_x=\bN$ and $\bG_y=\bN_-$. So this torsor canonicaly maps into the open $\bG$-orbit in $\bX\tim\bX$ and the fiber at $(\bB',{\bB}'')$ is a torsor for a Cartan subgroup $\bB'\cap\bB''$. $\Box$ \subsubsection{} \label{suda} Given a Pl\"ucker section $(v_\lambda)_{\lambda\in X^+}$, the collection of meromorphic sections $(x_\lambda\in\CV_\lambda^\vee,\frac{v_\lambda}{Q_\lambda}\in\CV_\lambda)$ evidently satisfies the conditions a)--e) of the above Lemma, and hence defines a meromorphic function $g:\ \BA^1\lra \bG$. Let us list the properties of this function. a) By the definition ~\ref{drinf} of the isomorphism $\psi$, since $g$ fixes the Pl\"ucker section $x$ the function $g$ actually takes values in $\bN\subset \bG$; b) The argument similar to that used in ~\ref{tuda} shows that $g$ can be extended to $\BP^1$, is regular at $\infty$, and $g(\infty)=1\in \bN$ (since $\frac{v_\lambda}{Q_\lambda}(\infty)=y_\la$, $g(\infty)$ stabilizes $x_\la$ and $y_\la$ so it lies in $\bN\cap\bN_-$); c) Let $D=\pi_\alpha(v_\lambda)$ (see ~\ref{pi}) be the $I$-colored divisor supported at the roots of $Q_\lambda$. Then $g$ is regular on $\BP^1-D$. \subsubsection{} \label{sjuda} We define $\xi(v_\lambda)=(D,\CT,\tau)\in\bZ^\alpha$ as follows: $D=\pi_\alpha(v_\lambda);\ \CT$ is the trivial $\bG$-torsor; the section $\tau$ is given by the function $g$. Let us describe the corresponding $\bbH$-torsor $\CT_{\tau,\bB_-}$ with meromorphic section $\tau_{\bB_-}$. To describe an $\bbH$-torsor $\fL$ with a section $s$ it suffices to describe the induced $\BC^*$-torsors $\fL_\lambda$ with sections $s_\lambda$ for all characters $\lambda:\ \bbH\lra\BC^*$. In fact, it suffices to consider only $\lambda\in X^+$. Then $\fL_\lambda$ is given by the construction of ~\ref{obratno}, and $s_\lambda=\frac{v_\lambda}{Q_\lambda}$. Thus, the conditions i),ii) of the Definition ~\ref{bZ} are evidently satisfied. \subsubsection{} \label{trivial} To proceed with the inverse construction, we will need the following {\em Lemma.} Suppose $(D,\CT,\tau)\in\bZ^\alpha$. Then $\CT$ is trivial and has a canonical section $\varsigma$. {\em Proof.} By the construction ~\ref{type}, $\CT$ is induced from the $\bB$-torsor $\barr{\bB\cdd \tau}$. By the Definition ~\ref{bZ}, the induced $\bbH$-torsor $\CT_{\tau,\bB}$ is trivial, that is, $\barr{\bB\cdd \tau}$ can be further reduced to an $\bN$-torsor. But any $\bN$-torsor over $\BP^1$ is trivial since $H^1(\BP^1,\bV)=0$ for any unipotent group $\bV$ (induction in the lower central series). $\Box$ \subsubsection{} According to the above Lemma, we can find a unique section $\varsigma$ of $\CT$ defined over the whole $\BP^1$ and such that $\varsigma(\infty)= \tau(\infty)$. Hence a triple $(D,\CT,\tau)\in\bZ^\alpha$ canonically defines a meromorphic function $$ g\df\ \tau\varsigma^{-1}:\ \BP^1\lra \bG ,$$ i.e., $g(x)\cdd\varsigma(x)=\tau(x),\ x\in\BP^1$. One sees immediately that $g$ enjoys the properties ~\ref{suda}a)--c). Now we can apply the Lemma ~\ref{drinf} in the opposite direction and obtain from $g$ a collection $\psi^{-1}(g)=(x_\lambda,\ti\upsilon_\lambda)_{\lambda\in X^+}$ with $\ti\upsilon_\la$ a certain meromorphic sections of $\CV_\la$. According to ~\ref{polynom}, the divisor $D$ defines a collection of unitary polynomials $(Q_\lambda)_{\lambda\in X^+}$, and we can define $v_\lambda\df\ Q_\lambda\ti\upsilon_\lambda$. One checks easily that $(v_\lambda)\in\sZ^\alpha$, and $(D,\CT,\tau)=\xi(v_\lambda)$. In particular, $\xi:\ \sZ^\alpha\lra\bZ^\alpha$ is an isomorphism. \subsection{} \label{Z} We conclude that $\CZ^\alpha,\sZ^\alpha,\bZ^\alpha$ are all the same and all maps to $\BA^\al$ coincide. We preserve the notation $\CZ^\alpha$ for this {\em Zastava} space, and $\pi_\alpha$ for its projection onto $\BA^\alpha$. We combine the properties ~\ref{dimension}, ~\ref{affine},~\ref{factorization},~\ref{irred} into the following {\bf Theorem.} a) $\CZ^\alpha$ is an irreducible affine algebraic variety of dimension $2|\alpha|$; b) For any decomposition $\alpha=\beta+\gamma,\ \beta,\gamma\in\BN[I]$, and a pair of disjoint subsets $U,\Upsilon\subset\BA^1$, we have the {\em factorization property} (notations of ~\ref{bZU} and ~\ref{factorization}): $$\CZ^{\beta,\gamma}_{U,\Upsilon}=\CZ^\beta_U\times\CZ^\gamma_\Upsilon$$ \bigskip \centerline{\bf CHAPTER 2. The category $\PS$} \section{Schubert stratification} \subsection{} \label{sig} We will stratify $\ZZ^\al$ in stages. We denote by $\CQ^\al\suppp\qp^\al\suppp \qc^\al $, respectively the variety of all quasimaps of degree $\al$ and the subvarieties of the quasimaps defined at $0$ and of genuine maps. In the same way we denote the varieties of based quasimaps $\ZZ^\al\suppp\zp^\al\df\ZZ^\al\cap\qp^\al \suppp \zc^\al\df\ZZ^\al\cap \qc^\al=$ based maps of degree $\al$. Recall (see ~\ref{sigma}) the map $\sigma_{\beta,\ga}: \CQ^\be\tim C\gaa\ra \CQ^{\be+\ga}, \ \si_{\be,\ga}(f,D)=f(-D)$. For $\be\le \al(=\be+\ga)$, it restricts to the embedding $ \CQ^\be\inj \CQ^{\al}, \ f\mm f(-(\al-\be)\cdd 0)=f(\ (\be-\al)\cdd 0)$, and in particular $\CZ^\beta\hra\CZ^\alpha$. \subsection{} \label{coarse} In the first step we stratify $\ZZ^\al$ according to the singularity at $0$. It follows immediately from the Theorem ~\ref{strat M} that $$ \CZ^\alpha\cong \bigsqcup_{0\leq\beta\leq\alpha}\dZ^\beta .$$ The closed embedding of a stratum $\dZ^\beta$ into $\CZ^\alpha$ will be denoted by $\sigma_{\beta,\alpha-\beta}$. \subsection{} Next, we stratify the quasimaps $\dZ^\al$ defined at $0$, according to the singularity on $\ccs$. Again, it follows immediately from the Theorem ~\ref{strat M} that $$ \zp^\alpha\cong \bigsqcup_{0\leq\beta\leq\alpha}\zc^\beta\tim (\ccs)^{\al-\be} .$$ \subsection{} \label{Schubert} One more refinement comes from the decomposition of the flag variety $\bX$ into the $\bB$-invariant Schubert cells. Given an element $w$ in the Weyl group $\CW_f$, we define the locally closed subvarieties ({\em Schubert strata}) $\dZ^{\alpha}_w\subset\dZ^\alpha$ and $\zc^{\alpha}_w\subset\zc^\alpha$, as the sets of quasimaps $f$ such that $f(0)\in\bX_w$. The closure of $\dZ^{\alpha}_w$ in $\CZ^\alpha$ will be denoted by $\oCZ^{\alpha}_w$. Evidently, $$\dZ^\alpha=\bigsqcup_{w\in \CW_f}\dZ^{\alpha}_w \aand \zc^\alpha=\bigsqcup_{w\in \CW_f}\zc^{\alpha}_w.$$ Beware that $\dZ^{\alpha}_w$ may happen to be empty: e.g. for $\alpha=0, w\not=w_0$. \subsubsection{} \label{fineS Q} Finally, the last refinement comes from the diagonal stratification of the configuration space $(\ccs)^\delta= \bigsqcup_{ \Ga\in\fP(\delta)}(\ccs)_\Ga^\delta$. Altogether, we obtain the following stratifications of $\ZZ^\al$: $$ \CZ^\alpha\cong \bigsqcup_{\alpha\ge \beta}\dZ^\beta \ \ (\text{{\em coarse stratification}}) $$ $$ \cong \bigsqcup^{\alpha\ge \beta\ge\ga}_{\Gamma\in\fP(\beta-\gamma)} \zc^\ga\tim (\ccs)^{\be-\ga}_\Gamma \ \ (\text{{\em fine stratification}}) $$ $$ \cong\bigsqcup^{\alpha\ge \beta\ge\ga}_{ w\in \CW_f,\ \Ga\in\fP(\be-\ga)} \zc^\ga_w \tim (\ccs)_\Ga^{\beta-\gamma} \ \ (\text{{\em fine Schubert stratification}}) .$$ Similarly, we have the {\em fine stratification} (resp. {\em fine Schubert stratification}) of $\CQ^\alpha$: $$\CQ^\alpha= \bigsqcup^{\alpha\ge \beta\ge\ga}_{\Gamma\in\fP(\beta-\gamma)} \qc^\ga\tim (\BP^1-0)^{\be-\ga}_\Gamma= \bigsqcup^{\alpha\ge \beta\ge\ga}_{ w\in \CW_f,\ \Ga\in\fP(\be-\ga)} \qc^\ga_w \tim (\BP^1-0)_\Ga^{\beta-\gamma}$$ Here $\qc^\gamma_w\subset\qc^\gamma$ denotes the locally closed subspace of maps $\BP^1\to\bX$ taking value in $\bX_w\subset\bX$ at $0\in\BP^1$. The strata $\qc^\gamma_w\times(\BP^1-0)_\Gamma^{\beta-\gamma}$ are evidently smooth. Note that the strata $\zp^\al_w$ are not necessarily smooth in general, e.g. for $\bG=SL_3,\ \alpha$ the sum of simple coroots, $w=w_0$. To understand the ``fine Schubert strata'' $\zc^\ga_w \tim (\ccs)_\Ga^\beta$ we need to understand the varieties $\zc^\ga_w$. \subsection{Conjecture} \label{conj} For $\gamma\in\BN[I],w\in\CW_f$ the variety $\zc^\gamma_w$ is smooth. Hence the ``fine Schubert stratification'' is really a stratification. \subsubsection{Lemma} \label{cheap} For $\gamma$ sufficiently dominant (i.e. $\langle\gamma,i'\rangle>10$) and arbitrary $w\in\CW_f$ the variety $\zc^\gamma_w$ is smooth. {\em Proof.} Let us consider the map $\varrho_\gamma:\ \qc^\gamma\lra\bX\times\bX,\ f\mapsto(f(0),f(\infty))$. We have $\oZ^\gamma=\varrho_\gamma^{-1}(\bX_w,\bB_-)$. It suffices to prove that $\varrho_\gamma$ is smooth and surjective. Recall that the tangent space $\Theta_f$ to $\qc^\gamma$ at $f\in\qc^\gamma$ is canonically isomorphic to $H^0(\BP^1,f^*\CT\bX)$. Let us interpret $\bX$ as a variety of Borel subalgebras of $\fg$. We denote $f(0)$ by $\fb_0$, and $f(\infty)$ by $\fb_\infty$. So we have to prove that the canonical map $H^0(\BP^1,f^*\CT\bX)\lra\CT_{\fb_0}\bX\oplus\CT_{\fb_\infty}\bX$ is surjective. To this end it is enough to have $H^1(\BP^1,f^*\CT\bX(-0-\infty))=0$. This in turn holds whenever $\gamma$ is sufficiently dominant. $\Box$ \subsubsection{Lemma} \label{dim Schubert} For $\gamma$ sufficiently dominant we have $\dim\oZ^\gamma_w=2|\gamma|-\dim\bX+\dim\bX_w$. {\em Proof.} The same as the proof of ~\ref{cheap}. $\Box$ \subsubsection{Remark} Unfortunately, one cannot prove the conjecture ~\ref{conj} for arbitrary $\gamma$ the same way as the Lemma ~\ref{cheap}: for arbitrary $\gamma$ the map $\varrho_\gamma$ is not smooth. The simplest example occurs for $\bG=SL_4$ when $\gamma$ is twice the sum of simple coroots. This example was found by A.Kuznetsov. \section{Factorization} This section follows closely \S4 of ~\cite{fs}. \subsection{} \label{fake} Now we replace the maps into the flag variety $\bX$ with the maps from $\BP^1$ to the product $\bX\tim Y=\sqcup_{\chi\in Y}\ \bX_\chi$. So for arbitrary $\chi\in Y$ and $\alpha\in\BN[I]$ we obtain the spaces $\CZ^\alpha_\chi$ of based maps into $\bX_\chi$ and it makes sense now to add the subscript $\chi$ to all the strata (coarse, Schubert, fine) defined in the previous section. We will consider a system $\ZZ$ of varieties $\CZ^\alpha_\chi,\ \al,\ga\in Y$, together with two kinds of maps defined for any $\beta,\ga\in\BN[I]$: a) closed embeddings, $$\sigma^{\beta,\gamma}_\chi:\ \CZ^\beta_\chi\hra\CZ^{\beta+\gamma}_{\chi+\gamma} ,$$ b) factorization identifications $$\CZ^{\beta,\gamma}_{\chi,\Ue,\Upe}= \CZ^\beta_{\chi,\Ue}\times\CZ^\gamma_{\chi-\beta,\Upe}$$ defined for $\varepsilon>0$ and $U_\varepsilon\df\ \{z\in\BC,\ |z|<\varepsilon\}$, and $\Upsilon_\varepsilon\df\ \{z\in\BC,\ |z|>\varepsilon\}$. Of course, without the subscript these are the factorizations from ~\ref{factorization} and the embeddings from ~\ref{sig}. \subsection{} \label{snop} We will denote by $\IC^\alpha_\chi$ the perverse $IC$-extension of the constant sheaf at the generic point of $\CZ^\alpha_\chi$. The following definition makes sense only modulo the validity of conjecture ~\ref{conj}. {\bf Definition.} A {\em snop} $\CK$ is the following collection of data: a) $\chi=\chi(\CK)\in Y$, called the {\em support estimate} of $\CK$; b) For any $\alpha\in\BN[I]$, a perverse sheaf $\CK^\alpha_\chi$ on $\CZ^\alpha_\chi$ smooth along the fine Schubert stratification; c) For any $\beta,\gamma\in\BN[I],\ \varepsilon>0$, a {\em factorization isomorphism} $$ \CK^{\beta+\gamma}_\chi|_{\CZ^{\beta,\gamma}_{\chi,\Ue,\Upe}} \iso \CK^\beta_\chi|_{\CZ^\beta_{\chi,\Ue}} \boxtimes \IC^\gamma_{\chi-\beta}|_{\CZ^\gamma_{\chi-\beta,\Upe}}$$ satisfying the {\em associativity constraints} as in ~\cite{fs}, \S\S 3,4. We spare the reader the explicit formulation of these constraints. \subsection{} \label{awful} Since at the moment the conjecture ~\ref{conj} is unavailable we will provide an ugly provisional substitute of the Definition ~\ref{snop}. Namely, recall that $\CZ^\alpha= \sqcup_{\alpha\geq\beta\geq\gamma}\oZ^\gamma\times(\BC^*)^{\beta-\gamma}$. We introduce an open subvariety $$\ddZ^\alpha=\bigsqcup_{\alpha\geq\beta\geq\gamma\gg0}\oZ^\gamma\times (\BC^*)^{\beta-\gamma}$$ The union is taken over sufficiently dominant $\gamma$, i.e. such that $\langle\gamma,i'\rangle>10$ for any $i\in I$. Certainly, if $\alpha$ itself is not sufficiently dominant, $\ddZ^\alpha$ may happen to be empty. We have the fine Schubert stratification $$\ddZ^\alpha=\bigsqcup^{\alpha\geq\beta\geq \gamma\gg0}_{w\in\CW_f,\ \Gamma\in\fP(\beta-\gamma)}\oZ^\gamma_w\times (\BC^*)^{\beta-\gamma}_\Gamma$$ with smooth strata (see the Lemma ~\ref{cheap}). Now we can repeat the Definition ~\ref{snop} replacing $\CZ^\alpha_\chi$ by $\ddZ^\alpha_\chi$. Thus in ~\ref{snop} b) we have to restrict ourselves to sufficiently dominant $\alpha$, and in ~\ref{snop} c) $\beta$ has to be sufficiently dominant as well. \subsubsection{} In what follows we use the Definition ~\ref{snop}. The reader unwilling to believe in the Conjecture ~\ref{conj} will readily substitute the conjectural Definition ~\ref{snop} with the provisional working Definition ~\ref{awful}. \subsection{Examples} We define the {\em irreducible} and {\em standard} snops. \subsubsection{} \label{CL} Let us describe a snop $\CL(w,\chi)$ for $\chi\in Y,\ w\in \CW_f$. a) The support of $\CL(w,\chi)$ is $\chi$. b) $\CL(w,\chi)^\alpha_\chi$ is the irreducible $IC$-extension $\IC(\oCZ^{\alpha}_{w,\chi})=j_{!*}\IC(\dZ^{\alpha}_{w,\chi})$ of the perverse $IC$-sheaf on the Schubert stratum $\dZ^{\alpha}_{w,\chi}\subset \dZ^\alpha_\chi$. Here $j$ stands for the affine open embedding $\dZ^{\alpha}_{w,\chi}\hra\oCZ^{\alpha}_w$. In particular, $\IC(\oCZ^\alpha_{w_0,\chi})=\IC^\alpha_\chi$. c) Evidently, $\oCZ^{\beta}_{w,\chi,\Ue}$ (resp. $\oCZ^{\beta,\gamma}_{w,\chi,\Ue,\Upe}$) is open in $\oCZ^{\beta}_{w,\chi}$ (resp. $\oCZ^{\alpha}_{w,\chi}$) for any $\beta+\gamma=\alpha$. Moreover, $\oCZ^{\beta,\gamma}_{w,\chi,\Ue,\Upe} =\oCZ^{\beta}_{w,\chi,\Ue}\times\CZ^\gamma_{\chi-\beta,\Upe}$. This induces the desired factorization isomorphism. \subsubsection{} \label{CM} If we replace in ~\ref{CL}b) above $j_{!*}\IC(\dZ^{\alpha}_{w,\chi})$ by $j_!\IC(\dZ^{\alpha}_{w,\chi})=:\CM(w,\chi)^\alpha_\chi$ (resp. $j_*\IC(\dZ^{\alpha}_{w,\chi})=:\CalD\CM(w,\chi)^\alpha_\chi$) we obtain the snop $\CM(w,\chi)$ (resp. $\CalD\CM(w,\chi)$). \subsection{} Given a snop $\CK$ with support $\chi$, and $\eta\geq\chi,\ \alpha\in\BN[I]$, we define a sheaf $'\CK^\alpha_\eta$ on $\CZ^\alpha_\chi$ as follows. We set $\gamma\df\ \eta-\chi$. If $\alpha\geq\gamma$ we set $$'\CK^\alpha_\eta\df\ (\sigma^{\alpha-\gamma,\gamma}_\chi)_* \CK^{\alpha-\gamma}_\chi$$ (for the definition of $\sigma$ see ~\ref{fake}). Otherwise we set $'\CK^\alpha_\eta\df\ 0$. It is easy to see that the factorization isomorphisms for $\CK$ induce similar isomorphisms for $'\CK$, and thus we obtain a snop $'\CK$ with support $\eta\geq\chi$. \subsection{} We define the category $\widetilde\PS$ of snops. \subsubsection{} \label{morphism} Given two snops $\CF,\CK$ we will define the morphisms $\Hom(\CF,\CK)$ as follows. Let $\eta\in Y$ be such that $\eta\geq\chi(\CF),\chi(\CK)$. For $\alpha=\beta+\gamma\in\BN[I]$ we consider the following composition: $$\vartheta^{\beta,\gamma}_\eta:\ \Hom_{\CZ^\alpha_\eta}('\CF^\alpha_\eta, '\CK^\alpha_\eta)\lra\Hom_{\CZ^{\beta,\gamma}_{\Ue,\Upe}} ('\CF^\alpha_\eta|_{\CZ^{\beta,\gamma}_{\Ue,\Upe}}, '\CK^\alpha_\eta|_{\CZ^{\beta,\gamma}_{\Ue,\Upe}})\iso$$ $$\Hom_{\CZ^\beta_{\eta,\Ue}\times\CZ^\gamma_{\eta-\beta,\Upe}} ('\CF^\beta_\eta|_{\CZ^\beta_{\eta,\Ue}}\boxtimes\IC^\gamma_{\eta-\beta} |_{\CZ^\gamma_{\eta-\beta,\Upe}}, '\CK^\beta_\eta|_{\CZ^\beta_{\eta,\Ue}}\boxtimes\IC^\gamma_{\eta-\beta} |_{\CZ^\gamma_{\eta-\beta,\Upe}})= \Hom_{\CZ^\beta_{\eta,\Ue}}('\CF^\beta_\eta|_{\CZ^\beta_{\eta,\Ue}}, '\CK^\beta_\eta|_{\CZ^\beta_{\eta,\Ue}})$$ (the second isomorphism is induced by the factorization isomorphisms for $'\CF$ and $'\CK$, and the third equality is just K\"unneth formula). Now we define $$\Hom(\CF,\CK)\df\ \dirlim_\eta\invlim_\alpha \Hom_{\CZ^\alpha_\eta}('\CF^\alpha_\eta,'\CK^\alpha_\eta)$$ Here the inverse limit is taken over $\alpha\in\BN[I]$, the transition maps being $\vartheta^{\beta,\alpha-\beta}_\eta$, and the direct limit is taken over $\eta\in Y$ such that $\eta\geq\chi(\CF),\chi(\CK)$, the transition maps being induced by the obvious isomorphisms $\Hom_{\CZ^\alpha_\eta}('\CF^\alpha_\eta,'\CK^\alpha_\eta)= \Hom_{\CZ^{\alpha+\gamma}_{\eta+\gamma}}('\CF^{\alpha+\gamma}_{\eta+\gamma}, '\CK^{\alpha+\gamma}_{\eta+\gamma})$. \subsubsection{} With the above definition of morphisms and obvious composition, the snops form a category which we will denote by $\widetilde\PS$. \subsection{} \label{PS} Evidently, the snops $\CL(w,\chi)$ are irreducible objects of $\widetilde\PS$. It is easy to see that any irreducible object of $\widetilde\PS$ is isomorphic to some $\CL(w,\chi)$. We define the category $\PS$ of {\em finite snops} as the full subcategory of $\widetilde\PS$ formed by the snops of finite length. It is an abelian category. We will see later that $\CM(w,\chi)$ and $\CalD\CM(w,\chi)$ (see ~\ref{CM}) lie in $\PS$ for any $w,\chi$. One can prove the following very useful technical lemma exactly as in ~\cite{fs}, 4.7. \subsubsection{Lemma} \label{stabilize} Let $\CF,\CK$ be two finite snops. Let $\eta\geq\chi(\CF),\chi(\CK)$. There exists $\beta\in\BN[I]$ such that for any $\alpha\geq\beta$ the canonical maps $\Hom(\CF,\CK)\lra \Hom_{\CZ^\alpha_\eta}('\CF^\alpha_\eta,'\CK^\alpha_\eta)$ are all isomorphisms. $\Box$ \bigskip \centerline{\bf CHAPTER 3. Convolution with affine Grassmannian} \section{Pl\"ucker model of affine Grassmannian} \subsection{} Let $\CG$ be the usual affine Grassmannian $\bG((z))/\bG[[z]]$. It is the ind-scheme representing the functor of isomorphism classes of pairs $(\CT,\tau)$ where $\CT$ is a $\bG$-torsor on $\BP^1$, and $\tau$ is its section (trivialization) defined off $0$ (see e.g. ~\cite{mv}). It is equipped with a natural action of proalgebraic group $\bG[[z]]$, and we are going to describe the orbits of this action. It is known (see e.g. {\em loc. cit.}) that these orbits are numbered by dominant cocharacters $\eta\in Y^+\subset Y$. Here $Y^+\subset Y$ stands for the set of cocharacters $\eta$ such that $\langle\eta,i'\rangle\geq0$ for any $i\in I$. For $\eta\in Y^+$ we denote the corresponding $\bG[[z]]$-orbit in $\CG$ by $\CG_\eta$, and we denote its closure by $\oCG_\eta$. Recall that for a dominant character $\lambda\in X^+$ we denote by $V_\lambda$ the corresponding irreducible $\bG$-module, and we denote by $\CV_\lambda$ the trivial vector bundle $V_\lambda\otimes\CO_{\BP^1}$ on $\BP^1$. \subsection{Proposition} \label{closure} The orbit closure $\oCG_\eta\subset\CG$ is the space of collections $(\CU_\lambda)_{\lambda\in X^+}$ of vector bundles on $\BP^1$ such that a) $\CV_\lambda(-\langle\eta,\lambda\rangle0)\subset\CU_\lambda\subset \CV_\lambda(\langle\eta,\lambda\rangle0)$, or equivalently, $\CU_\lambda(-\langle\eta,\lambda\rangle0)\subset\CV_\lambda\subset \CU_\lambda(\langle\eta,\lambda\rangle0)$; b) $\deg\CU_\lambda=\deg\CV_\lambda=0$, or in other words, $\dim\CV_\lambda(\langle\eta,\lambda\rangle0)/\CU_\lambda= \langle\eta,\lambda\rangle\dim V_\lambda$; c) For any surjective $G$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ and the corresponding morphism $\phi:\ \CV_\lambda\otimes\CV_\mu\lra\CV_\nu$ (hence $\phi:\ \CV_\lambda(\langle\eta,\lambda\rangle0)\otimes \CV_\mu(\langle\eta,\mu\rangle0)\lra \CV_\nu(\langle\eta,\lambda+\mu\rangle0)$) we have $\phi(\CU_\lambda\otimes\CU_\mu)=\CU_\nu$. {\em Proof.} $\bG$-torsor on a curve $C$ is the same as a tensor functor from the category of $\bG$-modules to the category of vector bundles on $C$. $\Box$ \subsection{} \label{loc} Let us give a local version of the above Proposition. Recall that $\CO=\BC[[z]]\subset\CK=\BC((z))$. For a finite-dimensional vector space $V$, a {\em lattice} $\fV$ in $V\otimes\CK$ is an $\CO$-submodule of $V\otimes\CK$ {\em commeasurable} with $V\otimes\CO$, that is, such that $(V\otimes\CO)\cap \fV$ is of finite codimension in both $V\otimes\CO$ and $\fV$. {\bf Proposition.} The orbit closure $\oCG_\eta\subset\CG$ is the space of collections $(\fV_\lambda)_{\lambda\in X^+}$ of lattices in $V_\lambda\otimes\CK$ such that a) $z^{\langle\eta,\lambda\rangle}(V_\lambda\otimes\CO)\subset \fV_\lambda\subset z^{-\langle\eta,\lambda\rangle}(V_\lambda\otimes\CO)$, or equivalently, $z^{\langle\eta,\lambda\rangle}\fV_\lambda\subset V_\lambda\otimes\CO\subset z^{-\langle\eta,\lambda\rangle}\fV_\lambda$; b) $\dim(z^{-\langle\eta,\lambda\rangle}(V_\lambda\otimes\CO)/\fV_\lambda)= \langle\eta,\lambda\rangle\dim V_\lambda$; c) For any surjective $G$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ and the corresponding morphism $\phi:\ (V_\lambda\otimes\CO)\otimes(V_\mu\otimes\CO)\lra(V_\nu\otimes\CO)$ (hence $\phi:\ z^{-\langle\eta,\lambda\rangle}(V_\lambda\otimes\CO)\otimes z^{-\langle\eta,\mu\rangle}(V_\mu\otimes\CO)\lra z^{-\langle\eta,\lambda+\mu\rangle}(V_\nu\otimes\CO)$), we have $\phi(\fV_\lambda\otimes \fV_\mu)=\fV_\nu$. $\Box$ \subsection{} \label{Iwahori} Let $\bI\subset\bG[[z]]$ be the Iwahori subgroup; it is formed by all $g(z)\in\bG[[z]]$ such that $g(0)\in\bB\subset\bG$. We will denote by $\CP(\CG,\bI)$ the category of perverse sheaves on $\CG$ with finite-dimensional support, constant along $\bI$-orbits. The stratification of $\CG$ by $\bI$-orbits is a certain refinement of the stratification $\CG=\sqcup_{\eta\in Y^+}\CG_\eta$. Namely, each $\CG_\eta$ decomposes into $\bI$-orbits numbered by $\CW_f/\CW_\eta$ where $\CW_\eta$ stands for the stabilizer of $\eta$ in $\CW_f$. For $w\in\CW_f/\CW_\eta$ we will denote the corresponding $\bI$-orbit by $\CG_{w,\eta}$. Let us introduce a Pl\"ucker model of $\CG_{w,\eta}$. \subsubsection{} For $\eta\in Y^+$ let $I_\eta\subset I$ be the set of all $i$ such that $\langle\eta,i'\rangle=0$ (thus for $i\not\in I_\eta$ we have $\langle\eta,i'\rangle>0$). Then $\CW_\eta$ is generated by the simple reflections $\{s_i, i\in I_\eta\}$. Let $\bP(I_\eta)$ be the corresponding parabolic subgroup (e.g. for $I_\eta=\emptyset$ we have $\bP(I_\eta)=\bB$, while for $I_\eta=I$ we have $\bP(I_\eta)=\bG$). Let $\bX(I_\eta)= \bG/\bP(I_\eta)$ be the corresponding partial flag variety. The $\bB$-orbits on $\bX(I_\eta)$ are naturally numbered by $\CW_f/\CW_\eta:\ \bX(I_\eta)= \sqcup_{w\in\CW_f/\CW_\eta}\bX(I_\eta)_w$. The Pl\"ucker embedding realizes $\bX$ as a closed subvariety in $\prod_{i\in I}\BP(V_{\omega_i})$. Its image under the projection $\prod_{i\in I}\BP(V_{\omega_i})\lra\prod_{i\not\in I_\eta} \BP(V_{\omega_i})$ exactly coincides with $\bX(I_\eta)$. \subsubsection{Lemma-Definition} a) For $\eta=\sum_{i\in I}n_ii$, and $(\CU_\lambda)_{\lambda\in X^+}\in\CG_\eta$ we have\\ $\dim(\CU_{\omega_i}+ \CV_{\omega_i}((n_i-1)\cdot0)/\CV_{\omega_i}((n_i-1)\cdot0))= \dim V_{\omega_i}^{{\bf U}(I_\eta)}$ where ${\bf U}(I_\eta)$ is the unipotent radical of $\bP(I_\eta)$. b) Thus $\CU_{\omega_i}, i\in I,$ defines a subspace $K_i$ in $\CV_{\omega_i}(n_i\cdot0)/\CV_{\omega_i}((n_i-1)\cdot0)= V_{\omega_i}$. This collection of subspaces $(K_i)_{i\in I}\in\prod_{i\in I}\operatorname{Gr}(V_{\omega_i})$ satisfies the Pl\"ucker relations and thus gives a point in $\bX(I_\eta)$; c) We will denote by $\br$ the map $\CG_\eta\lra\bX(I_\eta)$ defined in b); d) For $w\in\CW_f/\CW_\eta$ we have $\CG_{w,\eta}=\br^{-1}(\bX(I_\eta)_w)$. $\Box$ We are obliged to D.Gaitsgory who pointed out a mistake in the earlier version of the above Lemma. \subsubsection{} For $\theta\in Y$ we consider the corresponding homomorphism $\theta:\ \BC^*\lra\bH\subset\bG$ as a formal loop $\theta(z)\in\bG((z))$. It projects to the same named point $\theta(z)\in\bG((z))/\bG[[z]]=\CG$. There is a natural bijection between the set of $\theta(z),\ \theta\in Y$, and the set of Iwahori orbits: each Iwahori orbit $\CG_{w,\eta}$ contains exactly one of the above points, namely, the point $w\eta(z)$. \subsection{} \label{PBD} Recall the Beilinson-Drinfeld avatar $\bZ^\alpha$ of the Zastava space $\CZ^\alpha$ (see ~\ref{bZ}). In this subsection we will give a Pl\"ucker model of $\bZ^\alpha$. {\bf Proposition.} $\bZ^\alpha$ is the the space of pairs $(D,(\fU_\lambda)_{\lambda\in X^+})$ where $D\in\BA^\alpha$ is an $I$-colored effective divisor, and $(\fU_\lambda)_{\lambda\in X^+}$ is a collection of vector bundles on $\BP^1$ such that a) $\CV_\lambda(-\infty D)\subset\fU_\lambda\subset\CV_\lambda(+\infty D)$; b) $\CV_\lambda^\bN\subset\fU_\lambda\supset \CV_\lambda^{\bN_-}(-\langle D,\lambda\rangle)$ (notations of ~\ref{sigma}), the first inclusion being a {\em line subbundle} (and the second an invertible subsheaf); c) $\deg\fU_\lambda=0$; d) For any surjective $G$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ and the corresponding morphism $\phi:\ \CV_\lambda\otimes\CV_\mu\lra\CV_\nu$ (hence $\phi:\ \CV_\lambda(+\infty D)\otimes \CV_\mu(+\infty D)\lra \CV_\nu(+\infty D)$) we have $\phi(\fU_\lambda\otimes\fU_\mu)=\fU_\nu$. {\em Proof.} Obvious. $\Box$ \subsubsection{Remark} \label{?} Recall the isomorphism $\varpi^{-1}\xi:\ \bZ^\alpha\iso\CZ^\alpha$ constructed in section 7. Let us describe it in terms of ~\ref{PBD}. The Lemma ~\ref{trivial} says that there is a unique system of isomorphisms $\iota_\lambda:\ \fU_\lambda\iso\CV_\lambda,\ \lambda\in X^+$, identical at $\infty\in\BP^1$ and compatible with tensor multiplication. Then $\varpi^{-1}\xi(D,(\fU_\lambda)_{\lambda\in X^+})=(\fL_\lambda\subset \CV_\lambda)_{\lambda\in X^+}$ where $\fL_\lambda= \iota_\lambda(\CV_\lambda^{\bN_-}(-\langle D,\lambda\rangle))$. \subsection{} \label{stack} Let $\fM$ be the scheme representing the functor of isomorphism classes of $\bG$-torsors on $\BP^1$ equipped with trivialization in the formal neighbourhood of $\infty\in\BP^1$ (see ~\cite{ka} and ~\cite{kt1}). \subsubsection{} \label{strat stack} The scheme $\fM$ is stratified by the locally closed subschemes $\fM_\eta:\ \fM=\sqcup_{\eta\in Y^+}\fM_\eta$ according to the isomorphism types of $\bG$-torsors. Namely, due to Riemann's classification, for a $\bG$-torsor $\CT$ and any $\lambda\in X^+$ the associated vector bundle $\CV_\lambda^\CT$ decomposes as a direct sum of line bundles $\CO(r_k^\lambda)$ of well-defined degrees $r_1^\lambda\geq\ldots\geq r^\lambda_{\dim V_\lambda}$. Then $\CT$ lies in the stratum $\fM_\eta$ iff $r_1^\lambda=\langle\eta,\lambda\rangle$. For any $\eta\in Y^+$ the union of strata $\fM^\eta:= \sqcup_{Y^+\ni\chi\leq\eta}\fM_\chi$ forms an open subscheme of $\fM$. This subscheme is a projective limit of schemes of finite type, all the maps in projective system being fibrations with affine fibers. Moreover, $\fM^\eta$ is equipped with a free action of a prounipotent group $\bG^\eta$ (a congruence subgroup in $\bG[[z^{-1}]]$) such that the quotient $\ufM^\eta$ is a smooth scheme of finite type. The theory of perverse sheaves on $\fM$ smooth along the stratification by $\fM_\eta$ is developed in ~\cite{kt1}. We will refer the reader to this work, and will freely use such perverse sheaves, e.g. $\IC(\fM_\eta)$. \subsubsection{} \label{thin} Restricting a trivialization of a $\bG$-torsor from $\BP^1-0$ to the formal neighbourhood of $\infty\in\BP^1$ we obtain the closed embedding $\bi:\ \CG\hra\fM$. The intersection of $\fM_\eta$ and $\CG_\chi$ is nonempty iff $\eta\leq\chi$, and then it is transversal. Thus, $\oCG_\eta \subset\fM^\eta$. According to ~\cite{kt}, the composition $\oCG_\eta \hookrightarrow\fM^\eta\lra\ufM^\eta$ is a closed embedding. \subsubsection{} \label{mapstack} For a $\bG$-torsor $\CT$ and an irreducible $\bG$-module $V_\lambda$ we denote by $\CV_\lambda^\CT$ the associated vector bundle. Following ~\ref{maps} and ~\ref{quasimaps} we define for {\em arbitrary} $\alpha\in Y$ the scheme $\ofQ^\alpha$ (resp. $\fQ^\alpha$) representing the functor of isomorphism classes of pairs $(\CT,(\fL_\lambda)_{\lambda\in X^+})$ where $\CT$ is a $\bG$-torsor trivialized in the formal neighbourhood of $\infty\in\BP^1$, and $\fL_\lambda\subset\CV_\lambda^\CT,\ \lambda\in X^+$, is a collection of line subbundles (resp. invertible subsheaves) of degree $\langle-\alpha,\lambda\rangle$ satisfying the Pl\"ucker conditions (cf. {\em loc. cit.}). The evident projection $\ofQ^\alpha\lra\fM$ (resp. $\fQ^\alpha\lra\fM$) will be denoted by $\obp$ (resp. $\bp$). The open embedding $\ofQ^\alpha\hra\fQ^\alpha$ will be denoted by $\bj$. Clearly, $\bp$ is projective, and $\obp=\bp\circ\bj$. The free action of prounipotent group $\bG^\eta$ on $\fM^\eta$ lifts to the free action of $\bG^\eta$ on the open subscheme $\bp^{-1}(\fM^\eta)\subset\fQ^\alpha$. The quotient is a scheme of finite type $\ufQ^{\alpha,\eta}$ equipped with the projective morphism $\bp$ to $\ufM^\eta$. There exists a $\bG[[z^{-1}]]$-invariant stratification $\fS$ of $\fQ^\alpha$ such that $\bp$ is stratified with respect to $\fS$ and the stratification $\fM=\sqcup_{\eta\in Y^+}\fM_\eta$. One can define perverse sheaves on $\fQ^\alpha$ smooth along $\fS$ following the lines of ~\cite{kt1}. In particular, we have the irreducible Goresky-Macpherson sheaf $\IC(\fQ^\alpha)$. Following ~\ref{strat M} we introduce a decomposition of $\fQ^\alpha$ into a disjoint union of locally closed subschemes according to the isomorphism types of $\bG$-torsors and defects of invertible subsheaves: $$\fQ^\alpha=\bigsqcup^{\eta\in Y^+}_{\beta\leq\alpha}\ofQ^\beta_\eta \times C^{\alpha-\beta}$$ where $C=\BP^1$ and $\ofQ^\beta_\eta= \obp^{-1}(\fM_\eta)\subset\ofQ^\beta$. \subsection{IC sheaves} \subsubsection{} \label{GMZ} The Goresky-MacPherson sheaf $\IC^\alpha$ on $\CZ^\alpha$ is smooth along stratification $$\CZ^\alpha= \bigsqcup^{\alpha\geq\beta\geq\gamma\geq0}_{\Gamma\in\fP(\beta-\gamma)} \zc^\gamma\times(\BC^*)^{\beta-\gamma}_\Gamma$$ (cf. ~\ref{Schubert}). It is evidently constant along strata, so its stalk at a point in $\zc^\gamma\times(\BC^*)^{\beta-\gamma}_\Gamma$ depends on the stratum only. Moreover, due to factorization property, it depends not on $\alpha\geq\beta$ but only on their difference $\alpha-\beta\in \BN[I]$. We will denote it by $\IC^{\alpha-\beta}_\Gamma$. In case $\bG=SL_n$ these stalks were computed in ~\cite{ku}. \subsubsection{} \label{GMQ} Recall (see ~\ref{strat M}) that $\CQ^\beta,\ \beta\in\BN[I]$, is stratified by the type of defect: $$\CQ^\beta=\bigsqcup^{\beta\geq\gamma\geq0}_{\Gamma\in\fP(\beta-\gamma)} \oQ^\gamma\times C^{\beta-\gamma}_\Gamma$$ The Goresky-Macpherson sheaf $\IC(\CQ^\beta)$ on $\CQ^\beta$ is constant along the strata. It is immediate to see that its stalk at any point in the stratum $\oQ^\gamma\times C^{\beta-\gamma}_\Gamma$ is isomorphic, up to a shift, to $\IC^0_\Gamma$. In particular, it depends on the defect only. \subsubsection{} \label{GM} {\bf Proposition.} a) The Goresky-Macpherson sheaf $\IC(\fQ^\beta)$ on $\fQ^\beta,\ \beta\in Y$, is constant along the locally closed subschemes $$\fQ^\beta=\bigsqcup^{\beta\geq\gamma}_{\Gamma\in\fP(\beta-\gamma)} \ofQ^\gamma\times C^{\beta-\gamma}_\Gamma$$ b) The stalk of $\IC(\fQ^\beta)$ at any point in the $\ofQ^\gamma\times C^{\beta-\gamma}_\Gamma$ is isomorphic, up to a shift, to $\IC^0_\Gamma$. {\em Proof.} Will be given in ~\ref{later}. $\Box$ \subsubsection{} \label{parity} Let $\phi\in\fQ^\beta$. The stalk $\IC(\fQ^\beta)_\phi$ is a graded vector space. {\bf Conjecture.} {\em (Parity vanishing)} Nonzero graded parts of $\IC(\fQ^\beta)_\phi$ appear in cohomological degrees of the same parity. \subsubsection{Remark} In case $\bG=SL_n$ the conjecture follows from the Proposition ~\ref{GM} and ~\cite{ku} 2.5.2. In general case the conjecture follows from the unpublished results of G.Lusztig. We plan to prove it in the next part. \section{Convolution diagram} \subsection{Definition} For $\alpha\in Y,\eta\in Y^+$ we define the {\em convolution diagram} $\CG\CQ_\eta^\alpha$ as the space of collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ of vector bundles with invertible subsheaves such that a) $(\CU_\lambda)_{\lambda\in X^+}\in\oCG_\eta$, or in other words, $(\CU_\lambda)_{\lambda\in X^+}$ satisfies the conditions ~\ref{closure} a)-c); b) $\fL_\lambda\subset\CU_\lambda$ has degree $-\langle\alpha,\lambda\rangle$; c) For any surjective $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu=\lambda+\mu$ we have (cf. ~\ref{closure} c) $\phi(\CL_\lambda\otimes\CL_\mu)=\CL_\nu$; d) For any $\bG$-morphism $\phi:\ V_\lambda\otimes V_\mu\lra V_\nu$ such that $\nu<\lambda+\mu$ we have $\phi(\CL_\lambda\otimes\CL_\mu)=0$. \subsubsection{} Let us denote by $\oGQ_\eta^\alpha$ the open subvariety in $\CG\CQ_\eta^\alpha$ formed by all the collections $(\CU_\lambda,\fL_\lambda)$ such that $\fL_\lambda$ is a {\em line subbundle} in $\CU_\lambda$ for any $\lambda\in X^+$. The open embedding $\oGQ_\eta^\alpha\hra\CG\CQ_\eta^\alpha$ will be denoted by ~$\bj$. \subsection{Definition} a) We define the projection $\bp:\ \CG\CQ_\eta^\alpha \lra\oCG_\eta$ as $\bp(\CU_\lambda,\fL_\lambda)=(\CU_\lambda)$; b) We define the map $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\alpha+\eta}$ as follows: $$\bq(\CU_\lambda,\fL_\lambda)=(\fL_\lambda(-\langle\eta,\lambda\rangle0) \subset\CU_\lambda(-\langle\eta,\lambda\rangle0)\subset\CV_\lambda)$$ (cf. ~\ref{closure} a) and the Definition ~\ref{quasimaps}). \subsubsection{} \label{some} We will denote by $\obp$ the restriction of $\bp$ to the open subvariety $\oGQ_\eta^\alpha\stackrel{\bj}{\hookrightarrow}\CG\CQ_\eta^\alpha$. \subsubsection{Remark} Note that $\CG\CQ_0^\alpha=\CQ^\alpha,\ \oGQ_0^\alpha=\qc^\alpha,\ \bq=\id$, and $\bp$ is the projection to the point $\oCG_0$. Note also that while in the Definition ~\ref{quasimaps} we imposed the positivity condition $\alpha\in\BN[I]$ (otherwise $\CQ^\alpha$ would be empty) here we allow arbitrary $\alpha\in Y$. It is easy to see that $\CG\CQ_\eta^\alpha$ (as well as $\oGQ_\eta^\alpha$) is nonempty iff $\eta+\alpha\in\BN[I]$. \subsection{} Let us give a local version of the convolution diagram. Recall the notations of ~\ref{loc} and ~\ref{SSeta}. {\bf Definition.} For $\eta\in Y^+$ we define the {\em extended local convolution diagram} $\CG\CS_\eta$ as the ind-scheme formed by the collections $(\fV_\lambda,v_\lambda)_{\lambda\in X^+}$ of lattices in $V_\lambda\otimes\CK$ and vectors in $V_\lambda\otimes\CK$ such that a) $(\fV_\lambda)_{\lambda\in X^+}\in\oCG_\eta$, or in other words, $(\fV_\lambda)_{\lambda\in X^+}$ satisfies the condition ~\ref{loc}a)-c); b) $v_\lambda\in\fV_\lambda$; c) $(v_\lambda)_{\lambda\in X^+}\in\tCS$ (see ~\ref{SSeta}). \subsection{Definition} a) We define the projection $\bp:\ \CG\CS_\eta\lra\oCG_\eta$ as $\bp(\fV_\lambda,v_\lambda)=(\fV_\lambda)$; b) We define the map $\bq:\ \CG\CS_\eta\lra\CS$ as follows (cf. ~\ref{loc}a): $$\bq(\fV_\lambda,v_\lambda)_{\lambda\in X^+}= (z^{\langle\eta,\lambda\rangle}v_\lambda\in z^{\langle\eta,\lambda\rangle}\fV_\lambda \subset V_\lambda\otimes\CO)_{\lambda\in X^+}$$ \subsection{} \label{lcd} The torus $\bH_a$ acts in a natural way on $\CG\CS_\eta:\ h(\fV_\lambda,v_\lambda)=(\fV_\lambda,\lambda(h)v_\lambda)$. The action is evidently free, and we denote the quotient by $\CG\bQ_\eta$, the {\em local convolution diagram}. The map $\bp:\ \CG\CS_\eta\lra\oCG_\eta$ commutes with the action of $\bH_a$ (trivial on $\oCG_\eta$), so it descends to the same named map $\bp:\ \CG\bQ_\eta\lra\oCG_\eta$. The map $\bq:\ \CG\CS_\eta\lra\CS$ commutes with the action of $\bH_a$ (for the action on $\CS$ see ~\ref{SS}), so it descends to the same named map $\bq:\ \CG\bQ_\eta\lra\bQ$. The proalgebraic group $\bG(\CO)$ acts on $\CG\bQ_\eta$ and on $\bQ$, and the map $\bq$ is equivariant with respect to this action. \subsection{} \label{cartes} Let us compare the local convolution diagram with the global one. For $\alpha\in Y$, taking formal expansion at $0\in C$ as in ~\ref{m}, we obtain the closed embedding $\fs:\ \CG\CQ^\alpha_\eta\hookrightarrow \CG\bQ_\eta$. It is easy to see that the following diagram is cartesian: $$ \begin{CD} \CG\CQ^\alpha_\eta @>\fs>> \CG\bQ_\eta \\ @V{\bq}VV @V{\bq}VV \\ \CQ^{\eta+\alpha} @>{\fs}>> \bQ \end{CD} $$ \subsection{} Recall the locally closed embedding $\CZ^\alpha\subset\CQ^\alpha$. {\bf Definition.} We define the {\em restricted convolution diagram} $\CG\CZ_\eta^\alpha\subset\CG\CQ_\eta^\alpha$ as the preimage $\bq^{-1}(\CZ^{\eta+\alpha})$. The open subvariety $\CG\CZ_\eta^\alpha\cap \oGQ_\eta^\alpha$ will be denoted by $\oGZ_\eta^\alpha$. We will preserve the notations $\bp,\bq$ (resp. $\obp,\bj$) for the restrictions of these morphisms to $\CG\CZ_\eta^\alpha$ (resp. $\oGZ_\eta^\alpha$). \subsection{} \label{idiot} We construct the {\em Beilinson-Drinfeld} avatar $\CG\bZ^\alpha_\eta$ of the restricted convolution diagram $\CG\CZ_\eta^\alpha$. {\bf Definition.} $\CG\bZ^\alpha_\eta$ is the space of triples $(D,(\CU_\lambda)_{\lambda\in X^+},(\fU_\lambda)_{\lambda\in X^+})$ where $D\in\BA^{\eta+\alpha}$ is an effective $I$-colored divisor, and $(\CU_\lambda)_{\lambda\in X^+},(\fU_\lambda)_{\lambda\in X^+}$ are the collections of vector bundles on $\BP^1$ such that a) $(\CU_\lambda)_{\lambda\in X^+}\in\oCG_\eta$, or in other words, $(\CU_\lambda)_{\lambda\in X^+}$ satisfies the conditions ~\ref{closure} a)-c); b) $(D,(\fU_\lambda)_{\lambda\in X^+})\in\bZ^{\eta+\alpha}$, or in other words, $(D,(\fU_\lambda)_{\lambda\in X^+})$ satisfies the conditions ~\ref{PBD} a)-d); c) $\iota_\lambda(\CV_\lambda^{\bN_-}(-\langle D,\lambda\rangle))\subset \CU_\lambda(-\langle\eta,\lambda\rangle0)$ (notations of ~\ref{?}). \subsubsection{} The identification $\CG\CZ_\eta^\alpha=\CG\bZ_\eta^\alpha$ easily follows from ~\ref{PBD}. Under this identification, for $(D,(\CU_\lambda)_{\lambda\in X^+},(\fU_\lambda)_{\lambda\in X^+}) \in\CG\bZ_\eta^\alpha$, we have $$\bp(D,(\CU_\lambda)_{\lambda\in X^+},(\fU_\lambda)_{\lambda\in X^+})= (\CU_\lambda)_{\lambda\in X^+}$$ and $$\bq(D,(\CU_\lambda)_{\lambda\in X^+},(\fU_\lambda)_{\lambda\in X^+})= (D,(\fU_\lambda)_{\lambda\in X^+})$$ \subsection{} We will introduce the {\em fine} stratifications of $\CG\CQ_\eta^\alpha$ and $\CG\CZ_\eta^\alpha$ following the section 7. \subsubsection{Fine stratification of $\CG\CQ_\eta^\alpha$} \label{fine GQ} We have $$\CG\CQ_\eta^\alpha= \bigsqcup\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma$$ Here the union is taken over dominant $\chi\leq\eta$ in $Y^+$, arbitrary $\gamma\leq\beta\leq\alpha\in Y$, and partitions $\Gamma\in\fP(\beta-\gamma)$. Furthermore, $\ooGQ_\chi^\gamma\subset\oGQ_\chi^\gamma$ is an open subvariety formed by all the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $(\CU_\lambda)\in\CG_\chi\subset\oCG_\chi$, and $\fL_\lambda$ is a {\em line subbundle} in $\CU_\lambda$. The stratum $\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma$ is formed by the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $\bp(\CU_\lambda,\fL_\lambda)\in\CG_\chi\subset\oCG_\eta$; the normalization (see ~\ref{strat M}) of $\fL$ in $\CU$ has degree $\gamma$; and the defect $D$ (see {\em loc. cit.}) of $\fL$ in $\CU$ equals $(\alpha-\beta)0+D'$ where $D'\in(\BP^1-0)^{\beta-\gamma}_\Gamma$. {\bf Proposition.} $\ooGQ^\alpha_\eta$ is smooth for arbitrary $\alpha,\eta$, i.e. the fine strata are smooth. {\em Proof} will be given in ~\ref{sometimes}. \subsubsection{Fine Schubert stratification of $\CG\CQ_\eta^\alpha$} \label{fineS GQ} We have $$\CG\CQ_\eta^\alpha= \bigsqcup\ooGQ_{w,\chi}^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma$$ Here the union is taken over dominant $\chi\leq\eta$ in $Y^+$, representatives $w\in\CW_f/\CW_\chi$, arbitrary $\gamma\leq\beta\leq\alpha\in Y$, and partitions $\Gamma\in\fP(\beta-\gamma)$. Furthermore, $\ooGQ_{w,\chi}^\gamma\subset\oGQ_\chi^\gamma$ is an open subvariety formed by all the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $(\CU_\lambda)\in\CG_{w,\chi}\subset\oCG_\chi$, and $\fL_\lambda$ is a {\em line subbundle} in $\CU_\lambda$. The stratum $\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma$ is formed by the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $\bp(\CU_\lambda,\fL_\lambda)\in\CG_{w,\chi}\subset\oCG_\eta$; the normalization of $\fL$ in $\CU$ has degree $\gamma$; and the defect $D$ (see {\em loc. cit.}) of $\fL$ in $\CU$ equals $(\alpha-\beta)0+D'$ where $D'\in(\BP^1-0)^{\beta-\gamma}_\Gamma$. {\bf Proposition.} $\ooGQ^\alpha_{w,\eta}$ is smooth for arbitrary $\alpha,\eta,w$, i.e. the fine Schubert strata are smooth. {\em Proof} will be given in ~\ref{sometimes}. \subsubsection{Fine Schubert stratification of $\CG\CZ_\eta^\alpha$} \label{fine GZ} Similarly, we have $$\CG\CZ_\eta^\alpha= \bigsqcup\oGZ_{w,\chi}^\gamma\times(\BC^*)^{\beta-\gamma}_\Gamma$$ Here the union is taken over dominant $\chi\leq\eta$ in $Y^+$, representatives $w\in\CW_f/\CW_\chi$, arbitrary $\gamma\leq\beta\leq\alpha\in Y$, and partitions $\Gamma\in\fP(\beta-\gamma)$. Furthermore, $\oGZ_{w,\chi}^\gamma\subset\oGZ_\chi^\gamma$ is an open subvariety formed by all the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $(\CU_\lambda)\in\CG_{w,\chi}\subset\oCG_\chi$, and $\fL_\lambda$ is a {\em line subbundle} in $\CU_\lambda$. The stratum $\oGZ_{w,\chi}^\gamma\times(\BC^*)^{\beta-\gamma}_\Gamma$ is formed by the collections $(\CU_\lambda,\fL_\lambda)_{\lambda\in X^+}$ such that $\bp(\CU_\lambda,\fL_\lambda)\in\CG_{w,\chi}\subset\oCG_\eta$; the normalization of $\fL$ in $\CU$ has degree $\gamma$; and the defect $D$ of $\fL$ in $\CU$ equals $(\alpha-\beta)0+D'$ where $D'\in(\BC^*)^{\beta-\gamma}_\Gamma$. \subsubsection{} \label{unwilling} The reader unwilling to believe that $\oGZ^\gamma_{w,\chi}$ is smooth for arbitrary $\gamma\in Y$ may repeat the trick of ~\ref{awful}. Namely, one can replace $\CG\CZ^\alpha_\eta$ with an open subvariety ${\ddot{\CG\CZ}}^\alpha_\eta$ formed by the union of the above strata for sufficiently dominant $\gamma$; they are easily seen to be smooth. Moreover, $\bq({\ddot{\CG\CZ}}^\alpha_\eta)\supset\ddZ^{\eta+\alpha}$, and ${\ddot{\CG\CZ}}^\alpha_\eta\supset\bq^{-1}(\ddZ^{\eta+\alpha})$. \subsection{} \label{fine local} Let us introduce the {\em fine Schubert stratification} of the local convolution diagram (see ~\ref{lcd}). Iwahori subgroup $\bI\subset\bG(\CO)$ acts on $\tbQ$, and defines the {\em fine Schubert stratification} of $\tbQ$ by Iwahori orbits $\dbQ_w^\alpha\subset\dbQ^\alpha,\ \alpha\in Y,w\in\CW_f$. Furthermore, we have $$\CG\bQ_\eta=\bigsqcup\CG\dbQ_{w,\chi}^{-\alpha}$$ Here the union is taken over dominant $\chi\leq\eta$ in $Y^+$, representatives $w\in\CW_f/\CW_\chi$, and $\alpha\in\BN[I]$. The stratum $\CG\dbQ_{w,\chi}^{-\alpha}$ consists of collections $(\fV_\lambda,v_\lambda)_{\lambda\in X^+}$ (vectors $v_\lambda\in\fV_\lambda$ are defined up to multiplication by $\BC^*$) such that a) $(\fV_\lambda)_{\lambda\in X^+}\in\CG_{w,\chi}\subset\oCG_\eta$; b) $z^{-\langle\alpha,\lambda\rangle}v_\lambda\in\fV_\lambda$ for all $\lambda$, but $z^{-\langle\alpha,\lambda\rangle-1}v_\lambda\not\in\fV_\lambda$ for some $\lambda$. \section{Convolution} \subsection{} \label{hell} Let $\fA,\fB$ be smooth varieties, and let $\fp:\ \fA\lra\fB$ be a map. Suppose $\fA$ (resp. $\fB$) is equipped with a stratification $\fS$ (resp. $\fT$), and $\fp$ is stratified with respect to the stratifications. Let $\fR$ be another stratification of $\fB$, transversal to $\fT$. Let $\CB$ (resp. $\CA$) be a perverse sheaf on $\fB$ (resp. $\fA$) smooth along $\fR$ (resp. $\fS$). Let $b=\dim\fB$. {\em Lemma.} a) $\CA\otimes\fp^*\CB[-b]$ is perverse; b) $\CA\otimes\fp^*\CB[-b]=\CA\stackrel{!}{\otimes}\fp^!\CB[b]$; c) Let $\ol{R}$ (resp. $\ol{S}$) be the closure of a stratum $R$ in $\fR$ (resp. $S$ in $\fS$). Then $\IC(\ol{S})\otimes\fp^*\IC(\ol{R})[-b]= \IC(\ol{S}\cap\fp^{-1}\ol{R})$; d) Let $R$ be a stratum of stratification $\fR$. Then $\fp^{-1}R$ is smooth. {\em Proof.} a,b) Let $g:\ \fG\lra\fA\times\fB$ denote the closed embedding of the graph of $\fp$. The perverse sheaf $\CA\boxtimes\CB$ on $\fA\times\fB$ is smooth along the product stratification $\fS\times\fR$. The transversality of $\fR$ and $\fT$ implies that the embedding $g$ is noncharacteristic with respect to $\CA\boxtimes\CB$ (see ~\cite{ks}, Definition 5.4.12). Furthermore, by definition, $\CA\otimes\fp^*\CB=g^*(\CA\boxtimes\CB)$, and $\CA\stackrel{!}{\otimes}\fp^!\CB=g^!(\CA\boxtimes\CB)$. Now a) is nothing else than ~\cite{ks}, Corollary 10.3.16(iii), while b) is nothing else than ~\cite{ks}, Proposition 5.4.13(ii). $\Box$ c) We consider $\ol{S}\times\ol{R}$ as a subvariety of $\fA\times\fB$. Since $g:\ \fG\lra\fA\times\fB$ is noncharacteristic with respect to $\IC(\ol{S})\boxtimes\IC(\ol{R})=\IC(\ol{S}\times\ol{R})$, we conclude that $g^*\IC(\ol{S}\times\ol{R})[-b]=\IC(\fG\cap(\ol{S}\times\ol{R}))$. It remains to note that $\fG\cap(\ol{S}\times\ol{R})=\ol{S}\cap\fp^{-1}\ol{R}$, and $g^*\IC(\ol{S}\times\ol{R})=\IC(\ol{S})\otimes\fp^*\IC(\ol{R})$. $\Box$ d) We will view $\fp^{-1}R$ as a subscheme of $\fA$ (scheme-theoretic fiber over $R$). Let $a\in\fp^{-1}R\subset\fA$. The Zariski tangent space $T_a(\fp^{-1}R)$ equals $d\fp_a^{-1}(T_bR)$ where $b=\fp(a)$, and $d\fp_a:\ T_a\fA\lra T_b\fB$ stands for the differential of $\fp$ at $a$. Let $\CT$ be a stratum of $\fT$ containing $b=\fp(a)$. Then $T_b\CT$ is contained in $d\fp_a(T_a\fA)$ since $\fp$ is stratified with respect to $\fT$. Furthermore, $T_bR+T_b\CT=T_b\fB$ due to the transversality of $\CT$ and $R$. Hence $T_bR+d\fp_a(T_a\fA)=T_b\fB$. Hence $\dim(d\fp_a^{-1}(T_bR))= \dim\fA-\dim\fB+\dim R$. We conclude that the dimension of the Zariski tangent space $T_a(\fp^{-1}R)$ is independent of $a\in\fp^{-1}R$, and thus $\fp^{-1}R$ is smooth. $\Box$ \subsection{} \label{j} Consider the following cartesian diagram: $$ \begin{CD} \CG\CQ^\alpha_\eta @>\bi>> \fQ^\alpha \\ @V{\bp}VV @V{\bp}VV \\ \oCG_\eta @>{\bi}>> \fM \end{CD} $$ {\bf Proposition.} a) For a $\bG[[z]]$-equivariant perverse sheaf $\CF$ on $\oCG_\eta$ the sheaf $\IC(\fQ^\alpha)\otimes\bp^*\bi_*\CF[-\dim\ufM^\eta]$ is supported on $\CG\CQ^\alpha_\eta$ and is perverse; b) $\IC(\CG\CQ_\eta^\alpha)=\IC(\fQ^\alpha)\otimes\bp^*\bi_*\IC(\oCG_\eta) [-\dim\ufM^\eta]$. {\em Proof.} a) Let us restrict the right column of the above diagram to the open subscheme $\fM^\eta\subset\fM$, and take its quotient by $\bG^\eta$. We obtain the cartesian diagram $$ \begin{CD} \CG\CQ^\alpha_\eta @>\bi>> \ufQ^{\alpha,\eta} \\ @V{\bp}VV @V{\bp}VV \\ \oCG_\eta @>{\bi}>> \ufM^\eta \end{CD} $$ of schemes of finite type. Here the rows are closed embeddings, and $\ufM^\eta$ is smooth. The stratification $\fS$ of $\fQ^\alpha$ is invariant under the action of $\bG^\eta$, and descents to the same named quotient stratification of $\ufQ^{\alpha,\eta}$. Similarly, the stratification of $\fM$ by the isomorphism type of $\bG$-torsors descents to the stratification $\fT$ of $\ufM^\eta$. The sheaf $\bi_*\CF$ on $\ufM^\eta$ is smooth along the stratification $\fR$ transversal to $\fT$. We have to prove that $\IC(\ufQ^{\alpha,\eta})\otimes\bp^*\bi_*\CF[-\dim\ufM^\eta]$ is perverse. In order to apply the Lemma ~\ref{hell}a) we only have to find an embedding $\ufQ^{\alpha,\eta}\stackrel{u}{\hookrightarrow}\fA$ into a smooth scheme such that the map $\bp$ and stratification $\fS$ extend to $\fA$. Then we apply the Lemma ~\ref{hell}a) to the sheaf $\CA=u_*\IC(\ufQ^{\alpha,\eta})$ on $\fA$. We will construct $\fA$ as a projective bundle over $\ufM^\eta$. The points of $\ufM^\eta$ are the $\bG$-torsors $\CT$ over $C$ trivialized in some infinitesimal neighbourhood of $\infty\in C$. The points of $\ufQ^{\alpha,\eta}$ are the $\bG$-torsors $\CT$ over $C$ trivialized in some infinitesimal neighbourhood of $\infty\in C$ along with collections of invertible subsheaves $\fL_\lambda\subset\CV^\CT_\lambda, \lambda\in X^+$, satisfying Pl\"ucker relations. Now, if $\alpha=\sum_{i\in I}a_ii$ is dominant enough, the fiber of $\fA$ over $\CT\in\ufM^\eta$ is $\prod_{i\in I}\BP(\Gamma(C,\CV^\CT_{\omega_i}(a_i)))$. The map $u$ sends $(\CT,(\fL_\lambda\subset\CV^\CT_\lambda)_{\lambda\in X^+})$ to $(\fL_{\omega_i}(a_i))\in \prod_{i\in I}\BP(\Gamma(C,\CV^\CT_{\omega_i}(a_i)))$. If $\alpha$ is not dominant enough, we first embed $\ufQ^{\alpha,\eta}$ into $\ufQ^{\beta+\alpha,\eta}$ for dominant enough $\beta$ as follows: $(\CT,(\fL_\lambda\subset\CV^\CT_\lambda)_{\lambda\in X^+})\mapsto (\CT,(\fL_\lambda(-\langle\beta,\lambda\rangle0)\subset \CV^\CT_\lambda)_{\lambda\in X^+})$. Then we compose with the above projective embedding of $\ufQ^{\beta+\alpha,\eta}$. This completes the proof of a). $\Box$ b) We apply the Lemma ~\ref{hell}c) to $\ol{R}=\oCG_\eta,\ \ol{S}=\ufQ^{\alpha,\eta}$. $\Box$ \subsection{Proposition} \label{+} a) For a perverse sheaf $\CF$ in $\CP(\oCG_\eta,\bI)$ the sheaf $\IC(\fQ^\alpha)\otimes\bp^*\bi_*\CF[-\dim\ufM^\eta]$ is supported on $\CG\CQ^\alpha_\eta$ and is perverse; b) $\IC(\CG\CQ_{w,\eta}^\alpha)=\IC(\fQ^\alpha)\otimes\bp^*\bi_* \IC(\oCG_{w,\eta})[-\dim\ufM^\eta]$. {\em Proof.} The same as the proof of the Proposition ~\ref{j}; we need only to find a refinement $\fW$ of the stratification $\fM=\bigsqcup_{\eta\in Y^+}\fM_\eta$ which would be transversal to the Iwahori orbits $\CG_{w,\chi}$ in $\CG$. Now $\fM=\bigsqcup_{\eta\in Y^+}\fM_\eta$ is the stratification by the orbits of proalgebraic group $\bG[[z^{-1}]]$ acting naturally on $\fM$. The desired refinement is the stratification by the orbits of subgroup $\bI_-\subset\bG[[z^{-1}]]$ formed by the formal loops $g(z)\in\bG[[z^{-1}]]$ such that $g(\infty)\in\bB_-$. $\Box$ \subsubsection{} Recall the notations of ~\ref{some}. {\bf Conjecture.} a) The map $\obp:\ \oGQ^\alpha_\eta\lra\oCG_\eta$ is smooth onto its image; b) Up to a shift, $\IC(\fQ^\alpha)\otimes\bp^*\bi_*\CF[-\dim\ufM^\eta]=\bj_{!*}\obp^*\CF$ for any perverse sheaf $\CF$ in $\CP(\oCG_\eta,\bI)$. \subsection{Proof of the Propositions ~\ref{fine GQ} and ~\ref{fineS GQ}} \label{sometimes} We apply the Lemma ~\ref{hell}d) to the following situation: $\fA=\overset{\circ}\ufQ{}^{\alpha,\eta}\subset\ufQ^{\alpha,\eta},\ \fB=\ufM^\eta,\ R=\CG_{w,\eta},\ \fp=\obp$. The stratification $\fT$ is defined as follows. Recall the stratification $\fW$ of $\fM$ by $\bI_-$-orbits introduced in the proof of ~\ref{+}. It is invariant under the action of $\bG^\eta$ and descends to the desired stratification $\fT$ of $\ufM^\eta$ transversal to $R$. It remains to note that $\fA=\overset{\circ}\ufQ{}^{\alpha,\eta}=\obp^{-1}(\fM^\eta)/\bG^\eta$ is smooth being a quotient by the free group action of an open subscheme $\obp^{-1}(\fM^\eta)$ of the smooth scheme $\ofQ^\alpha$. Thus the assumptions of ~\ref{hell}d) are in force, and we conclude that $\ooGQ_{w,\eta}^\alpha=\obp^{-1}(\CG_{w,\eta})$ is smooth. The proof of smoothness of $\ooGQ_\eta^\alpha$ is absolutely similar. $\Box$ \subsection{} \label{z} We define the following locally closed subscheme $\fZ^\alpha\subset\fQ^\alpha$. Its points are the $\bG$-torsors $\CT$ over $C$ trivialized in the formal neighbourhood of $\infty\in C$ along with collections of invertible subsheaves $\fL_\lambda\subset\CV^\CT_\lambda, \lambda\in X^+$, satisfying Pl\"ucker relations plus two more conditions: a) in some neighbourhood of $\infty\in C$ the invertible subsheaves $\fL_\lambda\subset\CV^\CT_\lambda$ are line subbundles. Thus they may be viewed as a reduction of $\CT$ to $\bB\subset\bG$ in this neighbourhood. Since $\CT$ is trivialized in the formal neighbourhood of $\infty\in C$, we obtain a map from this neighbourhood to the flag manifold $\bX$. b) The value of the above map at $\infty\in C$ equals $\bB_-\in\bX$. We have the following cartesian diagram: $$ \begin{CD} \CG\CZ^\alpha_\eta @>\bi>> \fZ^\alpha \\ @V{\bp}VV @V{\bp}VV \\ \oCG_\eta @>{\bi}>> \fM \end{CD} $$ {\bf Proposition.} For a perverse sheaf $\CF\in\CP(\oCG_\eta,\bI)$ the sheaf $\IC(\fZ^\alpha)\otimes\bp^*\bi_*\CF[-\dim\ufM^\eta]$ is supported on $\CG\CZ^\alpha_\eta$ and is perverse. {\em Proof.} Similar to the proof of the Proposition ~\ref{j}. $\Box$ \subsubsection{Remark} \label{none} Let us denote the embedding of $\fZ^\alpha$ into $\fQ^\alpha$ by $s$. One can easily check that $\IC(\fZ^\alpha)=s^*\IC(\fQ^\alpha)[-\dim\bX]$. \subsection{} \label{isomor} Recall the notations of ~\ref{sig}. {\bf Proposition.} a) $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ (resp. $\bq:\ \CG\CZ_\eta^\alpha\lra\CZ^{\eta+\alpha}$) is proper; b) Restriction of $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ (resp. $\bq:\ \CG\CZ_\eta^\alpha\lra\CZ^{\eta+\alpha}$) to $\qp^{\eta+\alpha} \subset\CQ^{\eta+\alpha}$ (resp. to $\zp^{\eta+\alpha}\subset\CZ^{\eta+\alpha}$) is an isomorphism. {\em Proof.} a) is evident. b) It suffices to consider the case of $\bq:\ \CG\CQ^\alpha_\eta\lra\CQ^{\eta+\alpha}$. Let $(\fL_\lambda)_{\lambda\in X^+}\in\CQ^{\eta+\alpha}$. Then $(\CU_\lambda,\fL_\lambda')_{\lambda\in X^+}\in\bq^{-1} (\fL_\lambda)_{\lambda\in X^+}$ iff for any $\lambda\in X^+$ we have $\fL_\lambda'=\fL_\lambda(\langle\eta,\lambda\rangle0)$, and $\CU_\lambda\supset\CV_\lambda(-\langle\eta,\lambda\rangle0)+\fL_\lambda'$. Consider $\fL_\lambda=\CV_\lambda^{\bN_-}(-\langle\eta+\alpha, \lambda\rangle\infty)$, so that $\varphi=(\fL_\lambda)_{\lambda\in X^+}\in\qp^{\eta+\alpha}$. Then $(\CU_\lambda,\fL_\lambda')_{\lambda\in X^+}\in\bq^{-1} (\fL_\lambda)_{\lambda\in X^+}$ iff for any $\lambda\in X^+$ we have $\fL_\lambda'=\fL_\lambda(\langle\eta,\lambda\rangle0)= \CV_\lambda^{\bN_-}(\langle\eta,\lambda\rangle0-\langle\eta+ \alpha,\lambda\rangle\infty)$, and $\CU_\lambda\supset\CV_\lambda(-\langle\eta,\lambda\rangle0)+\fL_\lambda'= \CV_\lambda(-\langle\eta,\lambda\rangle0)+ \CV_\lambda^{\bN_-}(\langle\eta,\lambda\rangle0)$. In other words, $(\CU_\lambda)_{\lambda\in X^+}$ lies in the intersection of $\oCG_\eta$ with the semiinfinite orbit $T_\eta$ (see ~\ref{Iwasawa}). This intersection consists exactly of one point (see ~\cite{mv}). Thus $\bq^{-1}(\varphi)$ consists of one point. Recall the cartesian diagram ~\ref{cartes}. We have $\fs(\qp^{\eta+\alpha})\subset\dbQ^0$ (notations of ~\ref{QQ}), in particular, $\fs(\varphi)\in\dbQ^0$. Since the map $\bq:\ \CG\bQ_\eta\lra \bQ$ is $\bG(\CO)$-equivariant, and its fiber over $\fs(\varphi)$ consists of one point, we conclude that $\bq$ is isomorphism over the $\bG(\CO)$-orbit $\dbQ^0$. Since $\qp^{\eta+\alpha}=\fs^{-1}(\dbQ^0)$, applying the cartesian diagram ~\ref{cartes}, we deduce that $\bq$ is isomorphism over $\qp^{\eta+\alpha}.\ \Box$ \subsection{Proof of the Proposition ~\ref{GM}} \label{later} We are interested in the stalk of $\IC(\fQ^\alpha)$ at a point $(\CT,(\fL_\lambda)_{\lambda\in X^+})\in\ofQ^\gamma\times C^{\beta-\gamma}_\Gamma\subset\fQ^\alpha$. Suppose that the isomorphism class of $\bG$-torsor $\CT$ equals $\eta\in Y^+$, i.e. $\CT\in\fM_\eta$. The stalk in question evidently does not depend on a choice of $\CT\in\fM_\eta$ and the defect $D\in C^{\beta-\gamma}_\Gamma$. In particular, we may (and will) suppose that $\CT\in\bi(\CG_\eta)$, and $D\in (C-0)^{\beta-\gamma}_\Gamma$. Then one can see easily that the stalk in question is isomorphic, up to a shift, to the stalk of Goresky-MacPherson sheaf $\IC(\CG\CQ^\alpha_\eta)$ at the point $(\CT,(\fL_\lambda)_{\lambda\in X^+})\in\CG\CQ^\alpha_\eta$. On the other hand, according to the Proposition ~\ref{isomor} b), the latter stalk is isomorphic to the stalk of $\IC(\CQ^{\eta+\alpha})$ at the point $\bq(\CT,(\fL_\lambda)_{\lambda\in X^+})$. This point has the same defect $D$. Applying ~\ref{GMQ} we complete the proof of the Proposition ~\ref{GM}. $\Box$ \subsection{} \label{semismall} Recall that a map $\pi:\ \CX\lra\CY$ is called {\em dimensionally semismall} if the following condition holds: let $\CY_m$ be the set of all points $y\in\CY$ such that $\dim(\pi^{-1}y)\geq m$, then for $m>0$ we have $\codim_\CY\CY_m\geq2m$. Let us define $\CX_m=\pi^{-1}\CY_m$. Then we can formulate an equivalent condition of semismallness as follows: for any $m\geq0$ we have $\codim_\CX\CX_m\geq m$. Suppose $\CX$ (resp. $\CY$) is equipped with a stratification $\fS$ (resp. $\fT$), and $\pi$ is stratified with respect to $\fS$ and $\fT$. Then $\pi$ is called {\em stratified semismall} (see ~\cite{mv}) if $\pi$ is proper, and the restriction $\pi|_S$ to any stratum in $\fS$ is dimensionally semismall. In this case $\pi_*=\pi_!$ takes perverse sheaves on $\CX$ smooth along $\fS$ to perverse sheaves on $\CY$ smooth along $\fT$ (see {\em loc. cit.}). \subsection{} \label{q} Recall the fine Schubert stratifications of $\CQ^{\eta+\alpha}$ (resp. $\CG\CQ_\eta^\alpha$) introduced in ~\ref{fineS Q} (resp. ~\ref{fineS GQ}). The map $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ is stratified with respect to these stratifications. {\bf Proposition.} $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ is stratified semismall. {\em Proof} will use a few Lemmas. \subsubsection{Lemma} \label{lemma1} $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ is dimensionally semismall. {\em Proof.} Recall that we have $\CQ^{\eta+\alpha}=\sqcup_{\beta\leq \eta+\alpha}\qp^\beta$. It is enough to prove that for $(\fL_\lambda)_{\lambda\in X^+}\in\qp^\beta$ we have $\dim\bq^{-1}(\fL_\lambda)_{\lambda\in X^+}\leq |\eta+\alpha-\beta|$. Let us start with the case $\fL_\lambda= \CV_\lambda^{\bN_-}(\langle\beta-\eta-\alpha,\lambda\rangle0- \langle\beta,\lambda\rangle\infty)$. Then, like in the Proposition ~\ref{isomor}, we have $\bq^{-1}(\fL_\lambda)_{\lambda\in X^+}=\oCG_\eta\cap \overline{T}_{\beta-\alpha}$, and according to ~\cite{mv}, we have $\dim(\oCG_\eta\cap \overline{T}_{\beta-\alpha})\leq |\eta+\alpha-\beta|$. Now $\qp^\beta$ is stratified by the defect: $\qp^\beta= \sqcup^{\beta\geq\gamma\geq0}_{\Gamma\in\fP(\beta-\gamma)} \qc^\gamma\times (C-0)^{\beta-\gamma}_\Gamma$, and $\bq$ is evidently stratified with respect to this stratification. The point $(\CV_\lambda^{\bN_-}(\langle\beta-\eta-\alpha,\lambda\rangle0-\langle\beta, \lambda\rangle\infty))_{\lambda\in X^+}$ lies in the smallest (closed) stratum $\gamma=0,\Gamma=\{\{\beta\}\}$. Since the dimension of preimage is a lower semicontinuous function on $\qp^\beta$, we conclude that for any point $(\fL_\lambda)_{\lambda\in X^+}\in\qp^\beta$ we have $\dim\bq^{-1}(\fL_\lambda)_{\lambda\in X^+}\leq|\eta+\alpha-\beta|$. $\Box$ \subsubsection{} \label{lemma2} Recall the fine stratification of $\CG\CQ^\alpha_\eta$ (resp. of $\CQ^{\eta+\alpha}$) introduced in ~\ref{fine GQ} (resp. in ~\ref{fineS Q}). {\em Lemma.} $\bq:\ \CG\CQ_\eta^\alpha\lra\CQ^{\eta+\alpha}$ is stratified semismall with respect to fine stratifications. {\em Proof.} We consider a fine stratum $\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma\subset \CG\CQ^\alpha_\eta$ (see ~\ref{fine GQ}). Temporarily we will write $\bq^\alpha_\eta$ for $\bq$ to stress its dependence on $\eta$ and $\alpha$. The restriction of $\bq^\alpha_\eta$ to the stratum $\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma$ decomposes into the following composition of morphisms: $$\ooGQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma \stackrel{a\times\id}{\hookrightarrow} \CG\CQ_\chi^\gamma\times(\BP^1-0)^{\beta-\gamma}_\Gamma \stackrel{\bq^\gamma_\chi\times\id}{\lra} \CQ^{\chi+\gamma}\times(\BP^1-0)^{\beta-\gamma}_\Gamma \stackrel{b}{\hookrightarrow} \CQ^{\eta+\alpha}$$ Here $a$ is the open inclusion; and $b(\fL,D')= \fL((\beta-\alpha+\chi-\eta)0-D')$. Now $\bq^\gamma_\chi\times\id$ is semismall according to the Lemma ~\ref{lemma1}. This completes the proof of the Lemma. $\Box$ \subsubsection{Lemma} \label{lemma3} The restriction of $\bq$ to the fine Schubert stratum $\ooGQ^\alpha_{w,\eta}\subset\CG\CQ^\alpha_\eta$ is dimensionally semismall for any $w\in\CW_f/\CW_\eta$. {\em Proof.} Let $\bK\subset\bI\subset\bG(\CO)$ denote the first congruence subgroup formed by the loops $g(z)\in\bG(\CO)$ such that $g(0)=1$. The point $\eta(z)\in\CG_{e,\eta}\subset\CG_\eta$ was introduced in ~\ref{Iwahori}. For a positive integer $m$ the subset $(\CG\CQ^\alpha_\eta)_m$ (with respect to $\bq:\ \CG\CQ^\alpha_\eta\lra\CQ^{\eta+\alpha}$) was introduced in ~\ref{semismall}. {\em Claim 1.} $\codim_{\ooGQ_\eta^\alpha}(\ooGQ_\eta^\alpha)_m= \codim_{\obp^{-1}(g\cdot\bK\cdot\eta(z))}[\obp^{-1}(g\cdot\bK\cdot\eta(z)) \cap(\ooGQ^\alpha_\eta)_m]$ for any $m\geq0$ and $g\in\bG$. In effect, due to the $\bG$-equivariance of $\obp$ and $\bq$, the RHS does not depend on a choice of $g\in\bG$, so it suffices to consider $g=e$. The stabilizer of $\eta(z)$ in $\bG$ is nothing else than the parabolic subgroup $\bP(I_\eta)$ introduced in ~\ref{Iwahori}. We have $$\CG_\eta=\bG\times_{\bP(I_\eta)}[\bK\cdot\eta(z)];\ \ooGQ_\eta^\alpha=\bG\times_{\bP(I_\eta)}[\obp^{-1}(\bK\cdot\eta(z))];\ (\ooGQ_\eta^\alpha)_m=\bG\times_{\bP(I_\eta)} [\obp^{-1}(\bK\cdot\eta(z))\cap(\ooGQ^\alpha_\eta)_m]$$ The Claim follows. {\em Claim 2.} $\codim_{\obp^{-1}(g\cdot\bK\cdot\eta(z))} [\obp^{-1}(g\cdot\bK\cdot\eta(z))\cap(\ooGQ^\alpha_\eta)_m]= \codim_{\ooGQ^\alpha_{w,\eta}} [\ooGQ^\alpha_{w,\eta}\cap(\ooGQ^\alpha_\eta)_m]$ for any $m\geq0,\ g\in\bG$, and $w\in\CW_f/\CW_\eta$. In effect, let us choose $g$ in the normalizer of $\bH$ representing $w$. Let us denote by $\bP_w$ the intersection $\bP(I_\eta)\cap g\bP(I_\eta)g^{-1}$. Then we have $$\ooGQ^\alpha_{w,\eta}=\bP(I_\eta)\times_{\bP_w} [\obp^{-1}(g\cdot\bK\cdot\eta(z))];\ \ooGQ^\alpha_{w,\eta}\cap(\ooGQ^\alpha_\eta)_m=\bP(I_\eta)\times_{\bP_w} [\obp^{-1}(g\cdot\bK\cdot\eta(z))\cap(\ooGQ^\alpha_\eta)_m]$$ The Claim follows. Comparing the two Claims we obtain $$\codim_{\ooGQ^\alpha_{w,\eta}}(\ooGQ^\alpha_{w,\eta})_m\geq \codim_{\ooGQ^\alpha_{w,\eta}}[\ooGQ^\alpha_{w,\eta}\cap(\ooGQ^\alpha_\eta)_m]= \codim_{\ooGQ^\alpha_\eta}(\ooGQ^\alpha_\eta)_m\geq m$$ The last inequality holds by the virtue of the Lemma ~\ref{lemma1}. This completes the proof of the Lemma. $\Box$ \subsubsection{} Now we are ready to finish the proof of the Proposition. It remains to show that the restriction of $\bq$ to any fine Schubert stratum is dimensionally semismall. It follows from the Lemma ~\ref{lemma3} in the same way as the Lemma ~\ref{lemma2} followed from the Lemma ~\ref{lemma1} (``twisting by defect''). This completes the proof of the Proposition. $\Box$ \subsection{Corollary} \label{exact} The functor $\bq_*=\bq_!$ takes perverse sheaves on $\CG\CQ_\eta^\alpha$ smooth along the fine Schubert stratification to perverse sheaves on $\CQ^{\eta+\alpha}$ smooth along the fine Schubert stratification. $\Box$ \subsection{} A few remarks are in order. \subsubsection{Remark} Recall the fine Schubert stratification of the local convolution diagram introduced in ~\ref{lcd} and ~\ref{fine local}. The arguments used in the proof of the Proposition ~\ref{q} along with the Lemma ~\ref{codime} show that the map $\bq:\ \CG\bQ_\eta\lra\bQ$ is stratified semismall with respect to the fine Schubert stratification. \subsubsection{Remark} The same arguments as in the proof of Lemma ~\ref{lemma3} show that the convolution $\CA*\CB$ of perverse sheaves on $\CG$ (see ~\cite{lus}, or ~\cite{g}, ~\cite{mv}) is perverse if $\CB$ is $\bG(\CO)$-equivariant. In the particular case $\CA\in\CP(\CG,\bI)$ this was also proved by G.Lusztig in ~\cite{l5} using calculations in the affine Hecke algebra. \subsection{Theorem} \label{tough} Let $\eta\in Y^+,\alpha\in\BN[I]$. Consider the following diagram: $$ \begin{CD} \oCG_\eta @<{\bp}<< \CG\CQ^\alpha_\eta @>{\bq}>> \CQ^{\eta+\alpha} \\ @V{\bi}V{\subset}V @V{\bi}V{\subset}V @. \\ \fM @<\bp<< \fQ^\alpha @. {} \end{CD} $$ For a perverse sheaf $\CF$ in $\CP(\oCG_\eta,\bI)$, the sheaf $\bc^\alpha_\CQ(\CF):=\bq_*(\IC(\fQ^\alpha)\otimes\bp^*\CF)[-\dim\ufM^\eta]$ on $\CQ^{\eta+\alpha}$ is perverse and smooth along the fine Schubert stratification. {\em Proof.} By the virtue of ~\ref{+} we know that $\IC(\fQ^\alpha)\otimes\bp^*\CF[-\dim\ufM^\eta]$ is a perverse sheaf on $\CG\CQ^\alpha_\eta$. In order to apply the Corollary ~\ref{exact} we have to check that $\IC(\fQ^\alpha)\otimes\bp^*\CF[-\dim\ufM^\eta]$ is smooth along the fine Schubert stratification. The sheaf $\bp^*\CF$ is evidently smooth along the fine Schubert stratification. The sheaf $\IC(\fQ^\alpha)$ is constant along the stratification by defect (see ~\ref{GM}), hence $\bi^*\IC(\fQ^\alpha)$ is smooth along the fine Schubert stratification. This completes the proof of the Theorem. $\Box$ \subsection{Conjecture} Let $\eta\in Y^+,\alpha\in Y$ be such that $\eta+\alpha\in\BN[I]$. Consider the following diagram: $$ \begin{CD} \oCG_\eta @<{\bp}<< \CG\CQ^\alpha_\eta @>{\bq}>> \CQ^{\eta+\alpha} \\ @V{\bi}V{\subset}V @V{\bi}V{\subset}V @. \\ \fM @<\bp<< \fQ^\alpha @. {} \end{CD} $$ For a perverse sheaf $\CF$ in $\CP(\oCG_\eta,\bI)$, the sheaf $\bq_*(\IC(\fQ^\alpha)\otimes\bp^*\CF)[-\dim\ufM^\eta]$ on $\CQ^{\eta+\alpha}$ is perverse and smooth along the fine Schubert stratification. \subsection{Corollary} \label{bunk} Let $\eta\in Y^+,\alpha\gg0$. Consider the following diagram: $$ \begin{CD} \oCG_\eta @<{\bp}<< {\ddot{\CG\CZ}}^\alpha_\eta @>\bq>> \ddZ^{\eta+\alpha} \\ @V{\bi}V{\subset}V @V{\bi}V{\subset}V @. \\ \fM @<\bp<< \fZ^\alpha @. {} \end{CD} $$ (notations of ~\ref{unwilling} and ~\ref{z}). For a perverse sheaf $\CF\in\CP(\oCG_\eta,\bI)$, the sheaf $\bc^\alpha_\CZ(\CF):=\bq_*(\IC(\fZ^\alpha)\otimes\bp^*\CF)[-\dim\ufM^\eta]$ on $\ddZ^{\eta+\alpha}$ is perverse and smooth along the fine Schubert stratification. {\em Proof.} Let us denote by $s$ the locally closed embedding $\ddZ^{\eta+\alpha}\stackrel{s}{\hookrightarrow}\CQ^{\eta+\alpha}$. Also, temporarily, let us denote the maps $\bp$ and $\bq$ from the diagram ~\ref{tough} (resp. ~\ref{bunk}) by $\bp_\CQ$ and $\bq^\CQ$ (resp. $\bp_\CZ$ and $\bq^\CZ$) to stress their difference. Then we have $\bq^\CZ_*(\IC(\fZ^\alpha)\otimes\bp_\CZ^*\CF)[-\dim\ufM^\eta]= s^*\bq^\CQ_*(\IC(\fQ^\alpha)\otimes\bp_\CQ^*\CF)[-\dim\ufM^\eta-\dim\bX]$ (cf. ~\ref{none}). We also have $\codim_{\CQ^{\eta+\alpha}}\ddZ^{\eta+\alpha}=\dim\bX$, and the fine Schubert strata in $\ddZ^{\eta+\alpha}$ are intersections of $\ddZ^{\eta+\alpha}$ with the fine Schubert strata in $\CQ^{\eta+\alpha}$. One can check readily that the functor $s^*[-\dim\bX]$ takes perverse sheaves on $\CQ^{\eta+\alpha}$ smooth along the fine Schubert stratification to perverse sheaves on $\ddZ^{\eta+\alpha}$ smooth along the fine Schubert stratification. The application of ~\ref{tough} completes the proof of the Corollary. $\Box$ \subsection{Conjecture} Let $\eta\in Y^+,\alpha\in Y$ be such that $\eta+\alpha\in\BN[I]$. Consider the following diagram: $$ \begin{CD} \oCG_\eta @<{\bp}<< \CG\CZ^\alpha_\eta @>{\bq}>> \CZ^{\eta+\alpha} \\ @V{\bi}V{\subset}V @V{\bi}V{\subset}V @. \\ \fM @<\bp<< \fZ^\alpha @. {} \end{CD} $$ For a perverse sheaf $\CF$ in $\CP(\oCG_\eta,\bI)$, the sheaf $\bq_*(\IC(\fZ^\alpha)\otimes\bp^*\CF)[-\dim\ufM^\eta]$ on $\CZ^{\eta+\alpha}$ is perverse and smooth along the fine Schubert stratification. \subsection{} Now we will compare $\bc_\CZ^\alpha(\CF)$ for a fixed $\CF\in\CP(\oCG_\eta,\bI)$ and various $\alpha$. Recall the notations of ~\ref{snop} and ~\ref{awful}. {\bf Proposition.} For any $\beta,\gamma\in\BN[I], \varepsilon>0,$ there is a factorization isomorphism $$\bc_\CZ^{\beta+\gamma-\eta}\CF|_{\ddZ^{\beta,\gamma}_{\Ue,\Upe}}\iso \bc_\CZ^{\beta-\eta}\CF|_{\ddZ^\beta_\Ue}\boxtimes \IC^\gamma|_{\ddZ^\gamma_\Upe}$$ {\em Proof.} Follows easily from ~\ref{idiot}. $\Box$ \subsection{} The above Proposition shows that we can organize the collection $(\bc_\CZ^{\alpha-\eta}\CF)$ for $\alpha\in\BN[I]$ into a snop $\bc_\CZ\CF$. Namely, we set the support estimate $\chi(\bc_\CZ\CF)=\eta,\ (\bc_\CZ\CF)^\alpha_\eta=\bc_\CZ^{\alpha-\eta}\CF$. This way we obtain an exact functor $\bc_\CZ:\ \CP(\CG,\bI)\lra\PS$. \section{Examples of convolution} \subsection{} \label{stalks} Let $\CF$ be a perverse sheaf in $\CP(\oCG_\eta,\bI)$. For $\chi\leq\eta, w\in\CW_f/\CW_\chi$ we have $\CG_{w,\chi}\subset\oCG_\eta$. The sheaf $\CF$ is constant along $\CG_{w,\chi}$, and we denote by $\CF_{w,\chi}$ its stalk at any point in $\CG_{w,\chi}$. {\em Lemma.} The stalk of $\IC(\fQ^\alpha)\otimes\bp^*\CF$ at any point in a fine Schubert stratum $\ooGQ^\gamma_{w,\chi}\times (\BC^*)^{\beta-\gamma}_\Gamma\subset\CG\CQ_\eta^\alpha$ (see ~\ref{fineS GQ}) equals $\CF_{w,\chi}\otimes\IC^{\alpha-\beta}_\Gamma$ (see ~\ref{GMZ}). {\em Proof.} Follows immediately from the Proposition ~\ref{GM}. $\Box$ \subsection{} \label{Satake} Let $\bG^L$ be the Langlands dual group. Its character lattice coincides with $Y$, and the dominant characters are exactly $Y^+$. For $\eta\in Y^+$ we denote by $W_\eta$ the irreducible $\bG^L$-module with the highest weight $\eta$. For $\chi\in Y$ we denote by $_{(\chi)}W_\eta$ the weight $\chi$-subspace of $W_\eta$. Let $\IC(\oCG_\eta)$ denote the Goresky-MacPherson sheaf of $\oCG_\eta$. A natural isomorphism $H^\bullet(\oCG_\eta,\IC(\oCG_\eta))\iso W_\eta$ is constructed in ~\cite{mv}. Recall that for $\chi\in Y,\ w\in\CW_f,$ the irreducible snop $\CL(w,\chi)$ was introduced in ~\ref{CL}. The following result was suggested by V.Ginzburg. {\bf Theorem.} There is a natural isomorphism of snops: $$\bc_\CZ\IC(\oCG_\eta)\iso\bigoplus_{\chi\in Y}\ _{(\chi)}W_\eta\otimes \CL(w_0,\chi)$$ {\em Proof.} It is a reformulation of the main result of ~\cite{mv}. In effect, by the Proposition ~\ref{j}b) we know that $\IC(\fQ^\alpha)\otimes\bp^*\IC(\oCG_\eta)[-\dim\ufM^\eta]= \IC(\CG\CQ_\eta^\alpha)$. So we have to prove that $\bq_*\IC(\CG\CQ_\eta^\alpha)= \bigoplus_{0\leq\beta\leq\eta+\alpha}\ _{(\beta-\alpha)}W_\eta\otimes\IC(\CQ^\beta)$. Here we make use of the filtration $\CQ^{\eta+\alpha}=\bigcup_{0\leq\beta\leq\eta+\alpha}\CQ^\beta$ subject to the stratification $\CQ^{\eta+\alpha}= \bigsqcup_{0\leq\beta\leq\eta+\alpha}\qp^\beta$ by the defect at $0\in C$. We know that $\bq$ is proper, semismall, and stratified with respect to the above stratification. By the Decomposition Theorem (see ~\cite{bbd}), we have {\em a priori} $\bq_*\IC(\CG\CQ_\eta^\alpha)=\bigoplus_{0\leq\beta\leq\eta+\alpha} L_\beta\otimes\IC(\CQ^\beta)$ for some vector spaces $L_\beta$. To identify $L_\beta$ with $_{(\beta-\alpha)}W_\eta$ it suffices to compute the stalks at $\phi=(\fL_\lambda)_{\lambda\in X^+}\in\qp^\beta$ where $\fL_\lambda= \CV_\lambda^{\bN_-}(\langle\beta-\eta-\alpha,\lambda\rangle\cdot0- \langle D,\lambda\rangle)$ for some $D\in(\BP^1-0)^\beta_\Gamma$. As in the proof of ~\ref{lemma1} we have $\bq^{-1}(\phi)=\oCG_\eta\cap\overline{T}_{\beta-\alpha}= \bigsqcup_{\gamma\geq0} (\oCG_\eta\cap T_{\beta-\alpha+\gamma})$. According to the Lemma ~\ref{stalks}, we have $\IC(\CG\CQ_\eta^\alpha)|_{\oCG_\eta\cap T_{\beta-\alpha+\gamma}}= \IC(\oCG_\eta)|_{\oCG_\eta\cap T_{\beta-\alpha+\gamma}}\otimes \IC^\gamma_\Gamma$. According to ~\cite{mv}, we have $H^\bullet_c(\oCG_\eta\cap T_{\beta-\alpha+\gamma},\IC(\oCG_\eta))=\ _{(\beta-\alpha+\gamma)}W_\eta$. Due to the parity vanishing (see {\em loc. cit.} and ~\ref{parity}), the spectral sequence computing $H^\bullet(\oCG_\eta\cap\overline{T}_{\beta-\alpha},\IC(\CG\CQ_\eta^\alpha))$ collapses and gives $H^\bullet(\oCG_\eta\cap\overline{T}_{\beta-\alpha},\IC(\CG\CQ_\eta^\alpha))= \bigoplus_{0\leq\gamma\leq\eta+\alpha-\beta}\ _{(\beta-\alpha+\gamma)}W_\eta\otimes\IC^0_{\{\{\gamma\}\}}= \bigoplus_{0\leq\gamma\leq\eta+\alpha-\beta}\ _{(\beta-\alpha+\gamma)}W_\eta\otimes\IC(\CQ^{\beta+\gamma})_\phi$. This completes the proof of the Theorem. $\Box$
1997-08-11T19:46:39
9707
alg-geom/9707018
en
https://arxiv.org/abs/alg-geom/9707018
[ "alg-geom", "math.AC", "math.AG", "q-alg" ]
alg-geom/9707018
James M. Turner
James M. Turner
On Simplicial Commutative Rings with Vanishing Andr\'e-Quillen Homology
14 pages, LaTeX2e
null
null
null
null
We propose a generalization of a conjecture of D. Quillen, on the vanishing of Andr\'e-Quillen homology, to simplicial commutative rings. This conjecture characterizes a notion of local complete intersection, extended to the simplicial setting, under a suitable hypothesis on the local characteristic. Further, under the condition of finite-type homology, we then prove the conjecture in the case of a simplicial commutative algebra augmented over a field of non-zero characteristic. As a consequence, we obtain a proof of Quillen's conjecture for a Noetherian commutative algebra - again augmented over a field of non-zero characteristic.
[ { "version": "v1", "created": "Thu, 31 Jul 1997 17:38:39 GMT" }, { "version": "v2", "created": "Mon, 11 Aug 1997 17:46:38 GMT" } ]
2008-02-03T00:00:00
[ [ "Turner", "James M.", "" ] ]
alg-geom
\section{Introduction} In \cite{And} and \cite{Qui2}, M. Andr\'e and D. Quillen constructed the notion of a homology\\ $D_*(A|R;M)$ for a commutative algebra $A$, over a ring $R$, and an $A$-module $M$. It was then conjectured (see Section 5 of \cite{Qui2}) that, under suitable conditions on $R$ and $A$, the vanishing of the homology in sufficiently high degrees determines $A$ as a local complete intersection. In particular, for local rings, the conjecture takes the following form. \begin{conjecture}\label{con1.1} Let $R$ be a (Noetherian, commutative) local ring with residue field $\Bbb{F}$, and let $D_s(\Bbb{F}|R) = D_s(\Bbb{F}|R;\Bbb{F}), \ s \geq 0$. Then the following are equivalent: \begin{itemize} \item[(1)] $D_s(\Bbb{F}|R) = 0, \quad s >> 0$; \item[(2)] $D_s(\Bbb{F}|R) = 0, \quad s \geq 3$; \item[(3)] $R$ is a complete intersection. \end{itemize} \end{conjecture} In this form, \ref{con1.1} was proven by L. Avramov, in outline form, in \cite{Avr}, and, in much greater generality, in \cite{Avr2}. Recall that a local ring $R$ is a complete intersection if its $I$-adic completion $\hat{R}$ is a quotient of a complete regular ring by an ideal generated by a regular sequence. From this description and the properties of Andr\'e-Quillen homology, the implications (3) $\Rightarrow$ (2) $\Rightarrow$ (1) in \ref{con1.1} are immediate. The objective of this paper is to extend a version of \ref{con1.1} for simplicial local rings in an effort to bring the full power of simplicial homotopy theory to bear on this type of problem and thereby obtain a different proof of \ref{con1.1} closer in spirit to the topological results of J.-P. Serre in \cite{Serre} and Y. Umeda in \cite{Ume}. In \cite{Qui3}, D. Quillen gave a construction of Andr\'e-Quillen homology $D_*(A|B;M)$ where $B$ is a simplicial commutative ring, $A$ a simplicial commutative $B$-algebra, and $M$ a simplicial $A$-module. Let ${\mathcal R}_{\Bbb{F}}$ be the category of (commutative) local rings, with residue field $\Bbb{F}$, and $s {\mathcal R}_{\Bbb{F}}$ the category of simplicial objects over ${\mathcal R}_{\Bbb{F}}$. It follows from \cite{Qui1} that $s {\mathcal R}_{\Bbb{F}}$ has three classes of maps, called weak equivalences, fibrations, and cofibrations, giving it the structure of a closed simplicial model category. Using this structure, we say that a simplicial local ring $R$ is an {\em $n$-extension} if there is a cofibration sequence $$ \Sigma^{n-1} S_0 \rightarrow R \rightarrow S_{1}, $$ in the homotopy category $Ho(s {\mathcal R}_{\Bbb{F}})$, such that $S_{0}$ is polynomial in ${\mathcal R}_{\Bbb{F}}$ and $\hat{S_{1}} \cong \Sigma^{n}\bar{S}_{1}$ in $Ho(s {\mathcal R}_{\Bbb{F}})$. Here $\Sigma$ denotes the suspension in $Ho(s {\mathcal R}_{\Bbb{F}})$. (See $\S$I.2 and $\S$I.3 of \cite{Qui1} for the theory of suspension and cofibration sequences in homotopical algebra.) \begin{definition}\label{def1.2} Let $R$ be an object of $s {\mathcal R}_{\Bbb{F}}$. Then: \begin{itemize} \item[(1)] $R$ is {\bf regular} if $R$ is a 1-extension with $\bar{S}_{1}$ smooth in ${\mathcal R}_{\Bbb{F}}$. \item[(2)] $R$ is a {\bf complete intersection} if it is a 1-extension with $\bar{S}_1$ regular in $s {\mathcal R}_{\Bbb{F}}$. \item[(3)] $R$ is {\bf $Q$-bounded} if $(Q \pi_* R)_s = 0$ for $s >> 0$, and {\em bounded} if $\pi_s R = 0$ for $s \gg 0$. \item[(4)] The {\bf simplicial dimension} of $R$ is the integer $$ s \cdot \dim R = \max\{s| \, D_s(\Bbb{F}|R) \neq 0\}. $$ We then say that $R$ has {\bf finite simplicial dimension} if $s \cdot \dim R < \infty$. \item[(5)] $R$ is said to have {\bf finite-type homology} provided each $D_q(\Bbb{F}|R)$ is a finite dimensional $\Bbb{F}$-vector space. \item[(6)] If $R$ has both finite-type homology and finite simplicial dimension, we call $R$ {\bf finite}. \end{itemize} Given a simplicial commutative ring $R$, then $R$ is said to be {\em locally} of any one of (1) -- (6) provided $R_{\wp}$ is such, for each simplicial prime ideal $\wp$ in $R$. Given a simplicial prime ideal $\wp$ in $R$, we denote by $\Bbb{F}(\wp)$ the residue field of $R_{\wp}$ and we say that $R$ is {\em locally of non-zero characteristic} provided char$\Bbb{F}(\wp)\neq 0$ for all such $\wp$. \end{definition} We can now state our proposed simplicial generalization of \con{con1.1}. \begin{vanishingconjecture}\label{vancon1.3} Let $R$ be a locally finite simplicial commutative ring which is locally of non-zero characteristic. Then $R$ is a locally complete intersection if and only if $R$ is locally $Q$-bounded. \end{vanishingconjecture} In the rational case, while a complete intersection may be both $Q$-bounded and of finite simplicial dimension, the converse is not true. See the note following \prop{prop3.7} regarding counter-examples. To demonstrate the validity of \ref{vancon1.3}, we consider the subcategory, ${\mathcal A}_{\Bbb{F}}$, of ${\mathcal R}_{\Bbb{F}}$, consisting of augmented $\Bbb{F}$-algebras, i.e., unitary $\Bbb{F}$-algebras $A$ together with a fixed $\Bbb{F}$-algebra map $A \rightarrow \Bbb{F}$, called the augmentation of $A$. In this paper, we give evidence for \vancon{vancon1.3} by proving: \begin{theorem}\label{thm1.5} Suppose $A$ is a finite simplicial augmented commutative $\Bbb{F}$-algebra with char$\Bbb{F} > 0$. Then $A$ is bounded if and only if $A$ is a complete intersection. \end{theorem} \begin{corollary}\label{cor1.5} Let $A$ be an augmented Noetherian commutative $\Bbb{F}$-algebra, $char\Bbb{F} > 0$. Then $A$ is a complete intersection, as a local algebra, if and only if $A$ has finite simplicial dimension. \end{corollary} {\em Proof.} By IV.55 of \cite{And}, $H^{Q}_{*}(A)$ is of finite-type. Thus, by \thm{thm1.5}, $A$ has finite simplicial dimension if and only if $A$ is a complete intersection, as a simplicial algebra, if and only if $A$ is a complete intersection, as a local algebra, by the classical implication of (2) $\Rightarrow$ (3) in \con{con1.1} (see Proposition 26 of \cite{And}). \hfill $\Box$ \bigskip \subsection*{Organization of this paper} In this section, we review the closed simplicial model category structure for $s {\mathcal A}_{\Bbb{F}}$ and the construction and properties of homotopy and Andr\'e-Quillen homology. In Section 3, we describe the notion Postnikov envelopes for objects of $s {\mathcal A}_{\Bbb{F}}$ and explore its properties. In Section 4, we study the homotopy of n-extensions. Finally, in Section 5, we introduce and study the notion of a Poincar\'e series for a simplicial algebra, obtaining just enough information to prove \thm{thm1.5}. \subsection*{Acknowledgements} The author would like to thank Haynes Miller, for suggesting this project along with the direction it should take, and Jean Lannes for many useful directions as well as for making his stay in France worthwhile. Most of the work on this project was done while the author was visiting the Institut des Hautes \'Etudes Scientifique and the Ecole Polytechnique. He would like to thank them for their hospitality and use of their facilities during his stay. Finally, the author would like to thank Julie Riddleburger for putting this paper into LaTeX form. \section{Homotopy Theory of Simplicial Augmented\\ Commutative Algebras} \setcounter{equation}{0} We now review the closed simplicial model category structure for $s {\mathcal A}_{\Bbb{F}}$. We will assume the reader is familiar with the general theory of homotopical algebra given in \cite{Qui1}. We call a map $f: \ A \rightarrow B$ in $s {\mathcal A}_{\Bbb{F}}$ a \begin{itemize} \item[(i)] weak equivalence ($\stackrel{\sim}{\rightarrow}$) $\Leftarrow\!\!\Rightarrow \pi_* f$ is an isomorphism; \item[(ii)] fibration ($\rightarrow\!\!\!\!\rightarrow$) $\Leftarrow\!\!\Rightarrow f$ surjects in positive degrees; \item[(iii)] cofibration($\hookrightarrow$) $\Leftarrow \!\!\Rightarrow f$ is a retract of an almost free map. \end{itemize} Here a map $f: \ A \rightarrow B$ in $s {\mathcal A}_{\Bbb{F}}$ is {\em almost free} if there is an almost simplicial $\Bbb{F}$-vector space (no $d_0$) $V$ (see \cite{Goe1}) together with a map of almost simplicial $\Bbb{F}$-vector spaces $V \rightarrow IB$ such that the induced map $A \otimes S(V) \stackrel{\cong}{\longrightarrow} B$ is an isomorphism of almost simplicial algebras. Here $S$ is the symmetric algebra functor. Now given a finite simplicial set $K$ and a simplicial algebra $A$, define $A \wedge K$ and $A^K$ by $$ (A \wedge K)_n = \bigotimes_{K_{n}} A_n $$ and $$ (A^K)_n = \prod_{K_{n}} A_n. $$ Here the tensor product $\otimes $ is the coproduct in $s {\mathcal A}_{\Bbb{F}}$. The product in $s {\mathcal A}_{\Bbb{F}}$ is defined as $\Lambda \times_{\Bbb{F}} \Gamma$, for $\Lambda, \Gamma$ in $s {\mathcal A}_{\Bbb{F}}$, so that the diagram $$ \begin{array}{ccc} \Lambda \times_{\Bbb{F}}\Gamma & \longrightarrow & \Gamma \\[2mm] \downarrow & & \hspace*{10pt} \downarrow \epsilon \\[2mm] \Lambda & \longrightarrow & \Bbb{F} \\[-3mm] & \epsilon & \end{array} $$ is a pullback of simplicial vector spaces. \begin{theorem}\label{thm2.1} (\cite{Qui1}, \cite{Mil}, and \cite{Goe1}) With these definitions, $s {\mathcal A}_{\Bbb{F}}$ is a closed simplicial model category. \end{theorem} Given a simplicial vector space $V$, define its normalized chain complex $NV$ by \begin{equation} N_nV = V_n/(\mbox{Im} s_0 + \cdots + \mbox{Im} s_n) \end{equation} and $\partial: \ N_n V \rightarrow N_{n-1}V$ is $\partial = \sum^n_{i=0} (-1)^id_i$. The homotopy groups $\pi_* V$ of $V$ is defined as $$ \pi_n V = H_n(NV), \quad n \geq 0. $$ Thus for $A$ in $s {\mathcal A}_{\Bbb{F}}$ we define $\pi_* A$ as above. The Eilenberg-Zilber theorem (see \cite{Mac}) shows that the algebra structure on $A$ induces an algebra structure on $\pi_*A$. If we let ${\mathcal V}$ be the category of $\Bbb{F}$-vector spaces, then there is an adjoint pair $$ S: \ {\mathcal V} \Leftarrow\!\!\Rightarrow {\mathcal A}_{\Bbb{F}}: \, I, $$ where $I$ is the augmentation ideal function and $S$ is the symmetric algebra functor. For an object $V$ in ${\mathcal V}$ and $n \geq 0$, let $K(V,n)$ be the associated Eilenberg-MacLane object in $s {\mathcal V}$ so that $$ \pi_s K(V,n) = \left\{\begin{array}{ll} V & s = n; \\[2mm] 0 & s \neq n. \end{array}\right. $$ Let $S(V,n) = S(K(V,n))$, which is an object of $s {\mathcal A}_{\Bbb{F}}$. For $A$ in ${\mathcal A}_{\Bbb{F}}$, the indecomposable functor $QA = I(A)/I^2(A)$ which is an object of ${\mathcal V}$. Furthermore, we have an adjoint pair $$ Q: \ {\mathcal A}_{\Bbb{F}} \Leftarrow \!\!\Rightarrow {\mathcal V}: \ (-)_+ $$ where $V_+$, for $V$ in ${\mathcal V}$, is the simplicial algebra $V \oplus \Bbb{F}$ where $$ (v,r) \cdot (w,s) = (sv+rw,rs) $$ for $(v,r), (w,s) \in V \oplus \Bbb{F}$. $(-)_+, Q$ provides an equivalence between ${\mathcal V}$ and the category of abelian group objects in ${\mathcal A}_{\Bbb{F}}$. For $A$ in $s {\mathcal A}_{\Bbb{F}}$, we define its Andr\'e-Quillen homology, as per \cite{Goe1} and \cite{Goe2}, by $$ H^Q_s(A) = \pi_s QX, \quad s \geq 0, $$ where we choose a factorization $$ \Bbb{F} \hookrightarrow X \stackrel{\sim\hspace*{5pt}}{\rightarrow \!\!\!\!\! \rightarrow} A $$ of the unit $\Bbb{F} \rightarrow A$ as a cofibration and a trivial fibration. This definition is independent of the choice of factorization as any two are homotopic over $A$ (note that every object of $s {\mathcal A}_{\Bbb{F}}$ is fibrant). It is known (see, for example, \cite{Mil}) that $$ H^Q_s(A) = D_s(A|\Bbb{F};\Bbb{F}). $$ From the transitivity sequence, one can easily check that $D_0(\Bbb{F}|A) = 0$, and $D_{s+1}(\Bbb{F}|A)\\ \cong H^Q_s(A)$ for all $s \geq 0$. Now, as shown in \cite{Goe2}, $$ \pi_n A = [S(n),A], $$ where $S(n) = S(\Bbb{F},n)$ and $[\quad,\quad]$ denotes the morphisms in $Ho(s {\mathcal A}_{\Bbb{F}})$. Thus the primary operational structure for the homotopy groups in $s {\mathcal A}_{\Bbb{F}}$ is determined by $\pi_* S(V_0)$ for any $V_0$ in $s {\mathcal V}$. By Dold's theorem \cite{Dold} there is a triple ${\mathcal S}$ on graded vector spaces so that \begin{equation} \pi_* S(V) \cong {\mathcal S}(\pi_* V) \end{equation} encoding this structure. If char$\Bbb{F} = 0$, ${\mathcal S}$ is the free skew symmetric functor and, if char$\Bbb{F} > 0$, ${\mathcal S}$ is the free divided power algebra on the underlying vector space of a certain free algebra constructed from the input (see, for example, \cite{Bou} and \cite{Goe1}). Now recall that maps $A \stackrel{f}{\rightarrow} B \stackrel{g}{\rightarrow} C$ is a {\em cofibration sequence} in $Ho(s {\mathcal A}_{\Bbb{F}})$ if $f$ is isomorphic to a cofibration $X \stackrel{u}{\rightarrow} Y$, of cofibrant objects, with cofibre $Y \stackrel{v}{\rightarrow} Z$ isomorphic to $g$. Thus given any map $f: \ A \rightarrow B$ in $s {\mathcal A}_{\Bbb{F}}$ there is a cofibration sequence $A \stackrel{f}{\rightarrow} B \rightarrow M(f)$ in $Ho(s {\mathcal A}_{\Bbb{F}})$ formed by factoring $\Bbb{F} \rightarrow A$ into $\Bbb{F} \hookrightarrow \bar{A} \stackrel{\sim}{\rightarrow\!\!\!\!\rightarrow} A$, form the diagram $$ \begin{array}{ccc} \bar{A} & \hookrightarrow & X \\[2mm] s \begin{picture}(1,1)\put(1,1) {$\downarrow$} \put(1,5) {$\downarrow$}\end{picture} \hspace*{10pt} && \begin{picture}(1,1)\put(1,1) {$\downarrow$} \put(1,5) {$\downarrow$}\end{picture} \hspace*{10pt} \wr \\[2mm] A & \stackrel{f}{\rightarrow} & B \end{array} $$ and then let $M(f) = X \otimes_{\bar{A}}\Bbb{F}$, which is cofibrant. As an example, the {\em suspension} $\Sigma A$ of an object $A$ in $s {\mathcal A}_{\Bbb{F}}$ by $M(\epsilon)$, where $\epsilon: \ A \rightarrow \Bbb{F}$ is the augmentation. Finally, recall that the {\it completion} $\hat{A}$ of a simplicial augmented algebra $A$ is defined as $$ \hat{A} = lim_{t} A/I^{t}. $$ If $f: A \to B$ is a map of simplicial algebras, we denote by $\hat{f}: \hat{A} \to \hat{B}$ the induced map of completions. We can now summarize methods for computing homotopy and Andr\'e-Quillen homology that we will need for this paper. \begin{proposition}\label{prop2.4} \begin{itemize} \item[(1)] If $f: \ A \stackrel{\sim}{\rightarrow} B$ is a weak equivalence in $s {\mathcal A}_{\Bbb{F}}$, then $H^Q_*(f): \ H^Q_*(A) \stackrel{\cong}{\rightarrow} H^Q_*(B)$ is an isomorphism. \item[(2)] for any $A$ in $s {\mathcal A}_{\Bbb{F}}$ there is a spectral sequence $$ E^1_{s,t} = {\mathcal S}_{s} (H^Q_*(A)) \Rightarrow \pi_t \hat{A}, $$ called Quillen's Fundamental spectral sequence (see \cite{Qui2} and \cite{Qui3}), which converges when $H^{Q}_{0}(A)=0$. \item[(3)] There is a Hurewicz homomorphism $h: \ I\pi_* A \rightarrow H^Q_*(A)$ such that if $A$ is connected and $H^Q_s (A) = 0$ for $s<n$ then $A$ is $n$-connected and $h: \ \pi_n A \stackrel{\cong}{\rightarrow} H^Q_n(A)$ is an isomorphism. \end{itemize} \end{proposition} {\em Proof.} (1) is a standard result. See, for example, \cite{Qui2} or \cite{Goe2}. For (2), see chapter IV of \cite{Goe1} and \cite{Tur}. Finally, (3) is in \cite{Goe1}. \hfill $\Box$ \bigskip The following is a selection of results from \cite{Tur}. \begin{proposition}\label{compprop} Let $A$ and $B$ be in $s {\mathcal A}_{\Bbb{F}}$. Then \begin{itemize} \item[(1)] if $f:A \to B$ is an $H^{Q}_{*}$-isomorphism then $\hat{f}: \hat{A} \to \hat{B}$ is a weak equivalence, \item[(2)] if $A$ is connected then $\pi_{*}\hat{A} \cong \pi_{*}A$, \item[(3)] $H^{Q}_{*}(\hat{A}) \cong H^{Q}_{*}(A)$, and \item[(4)] $Q\pi_{*}\hat{A} \cong Q\pi_{*}A$. \end{itemize} \end{proposition} {\em Remark.} If $H^{Q}_{0}(A) = 0$ then \prop{compprop} (4) follows from a Quillen fundamental spectral sequence argument. This is due to the fact that while this spectral sequence doesn't directly converge to $\pi_{*}A$ it does allow, under the above condition on $H^{Q}_{0}$, sufficient information to be extracted about the indecomposables (see \cite{Tur} for further details). This case is sufficient for our needs. \bigskip \begin{proposition}\label{prop2.5} Let $A \stackrel{f}{\rightarrow} B \stackrel{g}{\rightarrow} C$ be a cofibration sequence in $Ho(s {\mathcal A}_{\Bbb{F}})$. Then: \begin{itemize} \item[(1)] There is a long exact sequence $$ \begin{array}{l} \cdots \rightarrow H^Q_{s+1} (C) \stackrel{\partial}{\rightarrow} H^Q_s(A) \stackrel{H^Q_*(f)}{\rightarrow} H^Q_s(B) \\[3mm] \hspace*{20pt} \stackrel{H^Q_*(g)}{\rightarrow} H^Q_s(C) \stackrel{\partial}{\rightarrow} H^Q_{s-1}(C) \rightarrow \cdots \end{array} $$ \item[(2)] There is a first quadrant spectral sequence of algebras $$ E^2_{s,t} = Tor^{\pi_*A}_{s} (\pi_* B, \Bbb{F})_t \Rightarrow \pi_{s+t} C, $$ which we refer to as the Eilenberg-Moore spectral sequence. \item[(3)] If $A$ is connected, there is a first quadrant spectral sequence of algebras $$ E^2_{s,t} = \pi_s (C \otimes \pi_t A) \Rightarrow \pi_{s+t} B, $$ which we refer to as the Serre spectral sequence. \item[(4)] If $A$ is connected and $C$ is $n$-connected, then there is a homomorphism $\tau: \ \pi_{n+1} C \rightarrow \pi_n A$, called the {\em transgression}, such that the diagram $$ \begin{array}{ccccccc} \pi_{n+1}B & \stackrel{\pi_* f}{\rightarrow} & \pi_{n+1}C & \stackrel{\tau}{\rightarrow} & \pi_nA & \stackrel{\pi_* f}{\rightarrow} & \pi_n B \\[1mm] h \downarrow \hspace*{8pt} && h \downarrow \hspace*{8pt} && \hspace*{8pt} \downarrow h && \hspace*{8pt} \downarrow h \\[1mm] H^Q_{n+1}B & \stackrel{H^Q_*f}{\rightarrow} & H^Q_{n+1}(C) & \stackrel{\partial}{\rightarrow} & H^Q_n(A) & \stackrel{H^Q_*f}{\rightarrow} & H^Q_n B \end{array} $$ commutes and the top sequence is exact. \end{itemize} \end{proposition} {\em Proof.} (1) is just the transitivity sequence for $H^Q_*$. See \cite{Goe1}. (2) is the spectral sequence of Theorem 6(b) in $\S$II.6 of \cite{Qui1}. See also \cite{Goe1}. By Theorem 6(d) in $\S$II.6 of \cite{Qui1}, there is a $1^{st}$-quadrant spectral sequence $$ E^2_{*,*} = \pi_* (B \otimes_A \pi_* A) \Rightarrow \pi_* B, $$ where $\pi_*A$ is an $A$-module via the augmentation $A \rightarrow \pi_0 A$. Here we can assume our cofibration sequence is a cofibration with cofibre $C$. Since $A$ is connected, then $B \otimes_A \pi_* A \cong C \otimes \pi_* A$. The algebra structure follows from the construction of the spectral sequence and the fact that $A \stackrel{f}{\rightarrow} B$ is a map of simplicial algebras. This gives us (3). For (4), since $A$ is connected and $C$ is $n$-connected, then in the Serre spectral sequence $$ d^{n+1}: \ \pi_{n+1} C \cong E^{n+1}_{n+1,0} \rightarrow E^{n+1}_{0,n} \cong \pi_n A, $$ which we propose is our desired map $\tau$. From this same spectral sequence, we have $\pi_s A \cong \pi_s B$, $s < n$, and, using methods modified from the next section, we can assume that $N_s IC = 0$ for $s \leq n$ and $N_{n+1}B \rightarrow\!\!\!\!\rightarrow N_{n+1}C$ is surjective. Since $E^{1}_{n+1,0} = N_{n+1}C$ and $E^2_{n+1,0} \cong E^{n+1}_{n+1,0} \cong \pi_{n+1}C$, then $d^{n+1}$ is constructable in precisely the same way as the boundary map in homological algebra. Since we can assume cofibrancy of our objects under consideration, then the diagram $$ \begin{array}{ccccccc} H_{n+1}(NB) & \rightarrow & H_{n+1}(NC) & \stackrel{d^{n+1}}{\rightarrow} & H_n(NA) & \rightarrow & H_n(NB) \\ \downarrow && \downarrow && \downarrow && \downarrow \\ H_{n+1}(NQB) & \rightarrow & H_{n+1}(NQC) & \stackrel{\partial}{\rightarrow} & H_n(NQA) & \rightarrow & H_n(NQB) \end{array} $$ commutes by naturality. The result follows. \hfill $\Box$ \section{Postnikov Envelopes} \setcounter{equation}{0} In this section, we construct and determine some properties of a useful tool for studying simplicial algebras. First, we recall the following standard result which will be useful for us (see section II.4 of \cite{Qui1}). \begin{lemma}\label{lma3.4} Let $V$ and $W$ be simplicial vector spaces. Then the map $$ [V,W] \rightarrow \mbox{Hom}_{{\mathcal V}_{*}}(\pi_* V, \pi_* W) $$ is an isomorphism. \end{lemma} \begin{proposition}\label{prop3.5} Let $A$ in $s {\mathcal A}_{\Bbb{F}}$. Then \begin{itemize} \item[(1)] There is a map of simplicial algebras $$ f_{0}:S(H^{Q}_{0}(A),0)\to A $$ which induces an isomorphism on $H^{Q}_{0}$. \item[(2)] Suppose $A$ is $(n-1)$-connected for $n \geq 1$. Then there exists a map in $s {\mathcal A}_{\Bbb{F}}$, $$ f_n: \ S(H^Q_nA,n) \rightarrow A, $$ which is an isomorphism on $\pi_n$ and $H^Q_n$. \end{itemize} \end{proposition} {\em Proof.} (1) Let $\iota: H^{Q}_{0}(A)\to I\pi_{0}A$ be a choice of splitting for the surjection $I\pi_{0}A \to H^{Q}_{0}(A)$. By \lma{lma3.4}, $f$ can be chosen to be the adjoint of the map of simplicial vector spaces $K(H^{Q}_{0}(A),0)\to IA$ induced by $\iota$. By the transitivity sequence, $H^{Q}_{0}(A(1))=0$ so the fundamental spectral sequence for $A(1)$ converges, by \prop{prop2.4} (2), so $A(1)$ is connected. (2) By the Hurewicz theorem, \prop{prop2.4} (3), the map $h: \ \pi_nA \rightarrow H^Q_n A$ is an isomorphism. Now the adjoint functors $$ S: \ s {\mathcal V} \Leftarrow\!\!\Rightarrow s {\mathcal A}_{\Bbb{F}}: \ I $$ induce an adjoint pair $$ S: \ Ho(s {\mathcal V}) \Leftarrow\!\!\Rightarrow Ho(s {\mathcal A}_{\Bbb{F}}): \ I. $$ Thus we have isomorphisms \begin{eqnarray*} [S(H^Q_nA,n),A] & \cong & [K(H^Q_nA,n),IA] \\[2mm] & \cong & \mbox{Hom}_{{\mathcal V}}(H^Q_n A, \pi_n IA), \end{eqnarray*} using \lma{lma3.4}. Choosing $f_n$ to correspond to the inverse of $h$ gives the result. \hfill $\Box$ \bigskip Now given $A$ we form the {\em Postnikov envelopes} as the sequence of cofibrations $$ A(1) \stackrel{j_{2}}{\hookrightarrow} A(2) \stackrel{j_{3}}{\hookrightarrow} \cdots \stackrel{j_{n}}{\hookrightarrow} A(n) \stackrel{j_{n+1}}{\hookrightarrow} \cdots $$ with the following properties: \begin{itemize} \item[(1)] $A(1) = \widehat{M(f_{0})}$, \item[(2)] for each $n\geq 1$, $A(n)$ is a $(n-1)$-connected and for $s \geq n$, $$ H^Q_s A(n) \cong H^Q_sA. $$ \item[(3)] There is a cofibration sequence $$ S(H^Q_{n} A, n) \rightarrow A(n) \stackrel{j_{n+1}}{\rightarrow} A(n+1). $$ \end{itemize} The existence of a Postnikov envelopes follows easily from \prop{compprop}, \prop{prop3.5}, and \begin{lemma}\label{lma3.6} If $A$ is $(n-1)$-connected, for $n\geq 1$, then the cofibre $M(f_n)$ of $f_n: \ S(H^Q_nA,n) \rightarrow A$ is $n$-connected and satisfies $H^Q_sM(f_n) \cong H^Q_sA$ for $s > n$. \end{lemma} {\em Proof.} This follows from \ref{prop3.5} and the transitivity sequence $$ H^Q_{s+1}M(f_n) \rightarrow H^Q_s S(H^Q_n A,n) \rightarrow H^Q_s A \rightarrow H^Q_s M(f_n). $$ \hfill $\Box$ \bigskip {\bf Note:} We have been implicitly using the computation $$ H^Q_s S(V,n) = \pi_s QS(V,n) = \pi_s K(V,n) = V $$ for $s = n$ and 0 otherwise. The converse holds as well. \begin{proposition}\label{prop3.7} Let $A$ be connected in $s {\mathcal A}_{\Bbb{F}}$ and suppose $H^Q_s A = 0, \ s \neq n > 0$. Then $A \cong S(H^Q_n A,n)$ in $Ho(s {\mathcal A}_{\Bbb{F}})$. \end{proposition} {\em Proof.} Since $A$ is connected, then $A$ is $(n-1)$-connected by the Hurewicz theorem. By \ref{prop3.5}, $f_n: \ S(H^Q_nA,n) \rightarrow A$ is an $H^Q_n$-isomorphism and hence a weak equivalence by \ref{prop2.4}(2). \hfill $\Box$ \bigskip {\bf Note.} From this proposition, if char$\Bbb{F}$=0 then $S(V,n)$ has simplicial dimension $n$ and $\pi_{*}S(V,n)$ is free skew-commutative on a basis of $V$ concentrated in degree $n$. Thus $S(V,n)$ is $Q$-bounded, for any $n$, showing that \vancon{vancon1.3} fails in the zero characteristic case. \bigskip \section{The Homotopy and Homology of $n$-Extensions} \setcounter{equation}{0} Call an object $A$ in $s {\mathcal A}_{\Bbb{F}}$ a {\em simple $n$-extension} if $A$ is an $n$-extension in $s {\mathcal A}_{\Bbb{F}}$ with $\bar{S}_1 = S(V_1,0)$, $V_1$ in ${\mathcal V}$. Also, for this section and the next, we define the {\em simplicial dimension} of $A$ to be $$ s \cdot \dim A = \max\{s| \, H^{Q}_{s}(A) \neq 0\} $$ We now proceed to prove: \begin{theorem}\label{thm1.4} Let $A$ be in $s {\mathcal A}_{\Bbb{F}}$. Then: \begin{itemize} \item[(1)] If $A$ is a connected simple $n$-extension for $n \geq 2$, then, in $Ho(s {\mathcal A}_{\Bbb{F}})$, we have $$ A \cong S(H^Q_{n-1}(A),n-1) \otimes S(H^Q_n(A),n). $$ \item[(2)] $A$ is a complete intersection if and only if $A$ is a simple 1-extension. \item[(3)] If $A$ is a complete intersection then $H^Q_s(A) = 0$ for $s \geq 2$ and if $H^{Q}_{*}(A)$ is of finite-type then $A$ is bounded. \item[(4)] If $H^{Q}_{0}(A)=0$ and $H^{Q}_{*}(A)$ is of finite-type then $\pi_{*}(\hat{A})$ is of finite-type. \item[(5)] The Postnikov envelope $A(1)$ has the following properties: \begin{itemize} \item[(a)] If $A$ has finite simplicial dimension, then so does $A(1)$; \item[(b)] If $H^{Q}_{0}(A)$ is finite and $\pi_{*}A$ is bounded then $\pi_{*}A(1)$ is $Q$-bounded. \item[(c)] If $H^{Q}_{*}(A)$ is of finite-type then $H^{Q}_{*}(A(1))$ is also of finite-type. \end{itemize} \end{itemize} \end{theorem} We begin with \begin{lemma}\label{lma4.1} Let $A$ in $s {\mathcal A}_{\Bbb{F}}$ be a connected simple $n$-extension for $n \geq 2$. Then $A$ is an $n$-extension of the form $$ S(H^Q_{n-1} A,n-1) \rightarrow A \rightarrow S(H^Q_n A,n). $$ \end{lemma} {\em Proof.} Let $V_0, V_1$ be vector spaces so that there is a cofibration sequence $$ S(V_0, n-1) \rightarrow A \rightarrow S(V,n). $$ Then the transitivity sequence tells us that $H^Q_s A = 0$, $s \neq n, \ n-1$ and there is an exact sequence $$ 0 \rightarrow H^Q_nA \rightarrow V_1 \rightarrow V_0 \rightarrow H^Q_{n-1}A \rightarrow 0. $$ Thus $A$ is $n-2$ connected and Postnikov tower gives us a cofibration sequence $$ S(H^Q_{n-1}A, n-1) \rightarrow A \rightarrow A(n-1) = S(H^Q_n A,n). \eqno\Box $$ \bigskip {\em Proof of \thm{thm1.4} (1).} By \lma{lma4.1}, there is a cofibration sequence $$ S(H^Q_{n-1},A,n-1) \stackrel{i}{\rightarrow} A \stackrel{j}{\rightarrow} S(H^Q_nA,n), $$ where we can assume $A$ is cofibrant, $i$ is a cofibration, and $j$ is the cofibre. Consider the commuting diagram $$ \begin{array}{ccc} [S(H^Q_nA,n),A] & \stackrel{j_*}{\longrightarrow} & [S(H^Q_nA,n),S(H^Q_nA,n)] \\[1mm] \cong \downarrow \hspace*{10pt} && \hspace*{10pt} \downarrow \cong \\[1mm] [K(H^Q_n A,n),IA] && [K(H^Q_n A,n), IS(H^Q_n A,n)] \\[1mm] \cong \downarrow \hspace*{10pt} && \hspace*{10pt} \downarrow \cong \\[1mm] \mbox{Hom}(H^Q_nA,I \pi_n A) & \stackrel{h_*}{\longrightarrow} & \mbox{Hom}(H^Q_n A, H^Q_n A) \end{array} $$ Then $j$ will split, up to homotopy, if we can show that $h: \ \pi_n A \rightarrow H^Q_n A$ is onto. By \prop{prop2.5}(4), there is a commutative diagram $$ \begin{array}{ccccccc} \pi_n A & \stackrel{\pi_* j}{\rightarrow} & \pi_n S(H^Q_n A,n) & \stackrel{\tau}{\rightarrow} & \pi_{n-1} S(H^Q_{n-1},A, n-1) & \stackrel{\pi_* i}{\rightarrow} & \pi_{n-1} A \\[2mm] h \downarrow \hspace*{10pt} && \cong \downarrow \hspace*{10pt} && \hspace*{10pt} \downarrow \cong && \hspace*{10pt} \downarrow \cong \\[2mm] H^Q_n A & \stackrel{\cong}{\rightarrow} & H^Q_n A & \stackrel{\partial = 0}{\rightarrow} & H^Q_{n-1}A & \stackrel{\cong}{\rightarrow} & H^Q_{n-1}A \end{array} $$ with the rows exact. Thus $\pi_n j$ is onto and, hence, $h: \ \pi_n A \rightarrow H^Q_n A$ is onto. \hfill $\Box$ \bigskip \begin{lemma}\label{lma4.3} Suppose $A$ in $s {\mathcal A}_{\Bbb{F}}$ is regular. Then $S(H^{Q}_{0}(A),0) \cong \hat{A}$ in $Ho(s {\mathcal A}_{\Bbb{F}})$. \end{lemma} {\em Proof.} By the standard transitivity sequence for $D_*$ applied to $\Bbb{F} \rightarrow A \rightarrow \Bbb{F}$, $D_0(\Bbb{F}|A) = 0$ and $D_{s+1}(\Bbb{F}|A) \cong H^Q_s(A)$, so since $A$ is regular, then $H^Q_s(A) = 0$, $s > 0$. Thus $f_{0}$ is an $H^{Q}_{*}$-isomorphism and so $\hat{f_{0}}$ is a weak equivalence by \prop{compprop}. \hfill $\Box$ \bigskip {\em Proof of \thm{thm1.4} (2).} If $A$ is a complete intersection then it is a 1-extension of the form $$ S_{0}\rightarrow A \rightarrow S_{1} $$ with $S_{0}$ polynomial and $\bar{S}_{1}$ regular as simplicial augmented algebra. By \lma{lma4.3}, A is thus a simple 1-extension. The converse is clear. \hfill $\Box$ \bigskip {\em Proof of \thm{thm1.4} (3).} If $A$ is a complete intersection, then $H^Q_s(A) = 0, \ s \geq 2$ follows (2) and the transitivity sequence. Consider now the Eilenberg-Moore spectral sequence $$ E^{2}_{s,t}=Tor^{S(H^{Q}_{0}(A))}_{s}(\pi_{*}A,\Bbb{F})_{t}\Longrightarrow \pi_{s+t}M(f_{0}) $$ which is a first quadrant homology-type spectral sequence of algebras. Since $H^{Q}_{0}(A)$ is finite, $S(H^{Q}_{0}(A))$ has finite flat dimension and since, by \prop{compprop}, $Q\pi_{*}M(f_{0}) = Q\pi_{*}A(1) = Q\pi_{*}S(H^{Q}_{1}(A),1)$ is finite concentrated in degree 1 then $\pi_{*}M(f_{0})$ is bounded and we can conclude, by an induction on $dim_{\Bbb{F}}H^{Q}_{0}(A)$, that $\pi_{*}A$ is bounded. \hfill $\Box$ \bigskip {\em Example.} Suppose an augmented commutative $\Bbb{F}$-algebra $B$ is a complete intersection. Then there is a complete regular algebra $\Gamma$ and an ideal $I$, generated by a regular sequence, so that $\Gamma/I \cong \hat{B}$. As we saw, $\Gamma \cong S(V_0)$, so the condition of regularity on $I$ is equivalent to there being a {\em projective extension}, that is, (see \cite{Goe1}) an extension $$ \Bbb{F} \rightarrow S(V_1) \stackrel{i}{\rightarrow} S(V_0) \rightarrow \hat{B} \rightarrow \Bbb{F}, $$ so that $i$ makes $S(V_0)$ into a projective $S(V_1)$-module. In $Ho(s {\mathcal A}_{\Bbb{F}})$, $\hat{B}$ is equivalent $M(i)$ and so there is a cofibration sequence of the form $$ S(V_0,0) \rightarrow \hat{B} \rightarrow S(V_1,1). $$ Thus, $\hat{B}$, and hence $B$, is a complete intersection as a simplicial algebra. \hfill $\Box$ \bigskip {\em Proof of \thm{thm1.4} (4).} Since $H^{Q}_{0}(A)=0$, the fundamental spectral sequence $$ E^1_{s,t} = {\mathcal S}_{s} (H^Q_*(A))_{t} \Rightarrow \pi_t \hat{A}, $$ converges. From the known structure of ${\mathcal S}$ (see e.g. \cite{Bou}), if $V$ is a finite-dimensional vector space then each ${\mathcal S}_{s}(V)_{t}$ is finite and ${\mathcal S}_{s}(V)_{t}=0, s \gg 0$ for each fixed t. The result follows. \hfill $\Box$ \bigskip {\em Proof of \thm{thm1.4} (5).} First, (a) is immediate from the transitivity sequence. For (b), $V = H^{Q}_{0}(A)$ is finite and the Eilenberg-Moore spectral sequence has the form $$ E^{2}_{s,t}=Tor^{S(V)}_{s}(\pi_{*}A,\Bbb{F})_{t}\Longrightarrow \pi_{s+t}M(f_{0}) $$ Since $S(V)$ has finite flat dimension and $\pi_{*}A$ is a graded $S(V)$-module then $$ Tor^{S(V)}_{s}(\pi_{*}A,\Bbb{F})_{t} = Tor^{S(V)}_{s}(\pi_{t}A,\Bbb{F}) $$ vanishes for $s \gg 0$ and vanishes for $t \gg 0$ if $\pi_{*}A$ is bounded. We conclude $\pi_{*}M(f_{0})$ is bounded and hence $\pi_{*}A(1)$ is $Q$-bounded, by \prop{compprop}. Finally, for (c), \lma{lma3.6} tells us that $H^{Q}_{s}(A)=H^{Q}_{s}(A(1))$ for $s \geq 1$. Thus if $H^{Q}_{*}(A)$ is of finite-type then $H^{Q}_{*}(A(1))$ is of finite-type. \hfill $\Box$ \bigskip \section{The Poincar\'e Series of a Simplicial Algebra} \setcounter{equation}{0} For this section, we assume char$\Bbb{F} = p > 0$. Let $A$ be a connected simplicial augmented commutative $\Bbb{F}$-algebra such that $\pi_{*}A$ is of finite-type. We define its {\em Poincar\'e series} by $$ \vartheta(A,t) = \sum_{n\geq 0}(dim_{\Bbb{F}}\pi_{n}A)t^{n}. $$ If $V$ is a finite-dimensional vector space and $n>0$ we write $$ \vartheta(V,n,t) = \vartheta(S(V,n),t). $$ Given power series $f(t) = \sum a_{i}t^{i}$ and $g(t) = \sum b_{i}t^{i}$ we define the relation $f(t) \leq g(t)$ provided $a_{i}\leq b_{i}$ for each $i\geq 0$. \begin{lemma}\label{poiprop} Given a cofibration sequence $$ A \rightarrow B \rightarrow C $$ of connected objects in ${\mathcal A}_{\Bbb{F}}$ with finite-type homotopy groups, then $$ \vartheta(B,t) \leq \vartheta(A,t)\vartheta(C,t) $$ which is an equality if the sequence is split. \end{lemma} {\em Proof.} From the Serre spectral sequence $$ E^{2}_{s,t}=\pi_{s}(C\otimes\pi_{t}A) \Longrightarrow \pi_{s+t}B $$ we have $$ \vartheta(A,t)\vartheta(C,t) = \sum_{n}(\sum_{s+t=n}dim_{\Bbb{F}}E^{2}_{s,t})t^{n} \geq \vartheta(B,t). $$ If the cofibration sequence is split then the spectral sequence collapses, giving an equality. \hfill $\Box$ \bigskip If $\Pi$ is a finitely-generated abelian group and $n>0$ let $$ \vartheta(\Pi,n,t) = \sum_{s}(dim_{\Bbb{F}}H_{s}(K(\Pi,n);\Bbb{F}))t^{s}. $$ \begin{lemma}\label{lmapoi} Let $V$ be a finite-dimensional vector space and $\Pi$ a free abelian group of the same dimension. Then for any $n>0$ $$ \vartheta(V,n,t) = \vartheta(\Pi,n,t). $$ \end{lemma} {\em Proof.} As shown in \cite{Car}, there is a weak equivalence of simplicial vector spaces $$ S(V,n) \rightarrow \Bbb{F}[K(\Pi,n)] $$ which gives us the desired result. \hfill $\Box$ \bigskip \begin{proposition}\label{phiprop} Given a finite-dimensional vector space $V$ and any $n>0$ the Poin-\\car\'e series $\vartheta(V,n,t)$ converges in the open unit disc. \end{proposition} {\em Proof.} This follows from \lma{lmapoi} and the results of J.P. Serre in \cite{Serre} and Y. Umeda in \cite{Ume}. \hfill $\Box$ \bigskip Now given two power series $f(t)$ and $g(t)$ we say $f(t) \sim g(t)$ provided $lim_{t\to \infty }f(t)/g(t)\\ = 1$. Given a Poincar\'e series $\vartheta(V,n,t)$, for a finite-dimensional $\Bbb{F}$-vector space $V$ and $n>0$, let $$ \varphi(V,n,t) = log_{p}\vartheta(V,n,1-p^{-t}). $$ \begin{proposition}\label{poieq} For $V$ an $\Bbb{F}$-vector space of finite dimension $q$ and $n>0$ then $\varphi(V,n,t)$ converges on the real line and $$ \varphi(V,n,t)\sim qt^{n-1}/(n-1)!. $$ \end{proposition} {\em Proof.} This follows from \lma{lmapoi} and Th\'eor\`eme 9b in \cite{Serre}, for char$\Bbb{F}=2$, and its generalization in \cite{Ume}. \hfill $\Box$ \bigskip A major step in proving \thm{thm1.5} will be accomplished with \begin{theorem}\label{connprop} Let $A$ be a connected finite simplicial augmented commutative $\Bbb{F}$-algebra. Then if $A$ is $Q$-bounded we have $A \cong S(H^{Q}_{1}(A),1)$ in $Ho(s {\mathcal A}_{\Bbb{F}})$. \end{theorem} {\em Proof.} By \thm{thm1.4} (4) and \prop{compprop}, $\pi_{*}A$ is of finite-type and hence, as it is also $Q$-bounded, bounded as well. Let $n = s \cdot \dim A$. We must show that $n = 1$. Consider the Postnikov envelope $$ S(H^{Q}_{s-1}(A),s-1) \rightarrow A(s-1) \rightarrow A(s) $$ for each s. From the theory of cofibration sequences (see section I.3 of \cite{Qui1}) the above sequence extends to a cofibration sequence $$ A(s-1) \rightarrow A(s) \rightarrow S(H^{Q}_{s-1}(A),s). $$ Thus, by \lma{poiprop}, we have $$ \vartheta(A(s),t) \leq \vartheta(A(s-1),t)\vartheta(H^{Q}_{s-1}(A),s,t). $$ Starting at $s=n-1$ and iterating this relation, we arrive at the inequality $$ \vartheta(A(n-1),t) \leq \vartheta(A,t)\prod_{s=1}^{n-2} \vartheta(H^{Q}_{s}(A),s+1). $$ Now, $A(n) \cong S(H^{Q}_{n}(A),n)$ by \prop{prop3.7}, but, by \thm{thm1.4} (1) and \lma{poiprop}, we have $$ \vartheta(A(n-1),t) = \vartheta(H^{Q}_{n-1}(A),n-1,t)\vartheta(H^{Q}_{n}(A),n,t). $$ Since $\pi_{*}(A)$ is of finite-type and bounded then there exists a $D>p$ such that $\vartheta(A,t) \leq D$, in the open unit disc. Combining, we have $$ \vartheta(H^{Q}_{n-1}(A),n-1,t)\vartheta(H^{Q}_{n}(A),n,t) \leq D\prod_{s=1}^{n-2}\vartheta(H^{Q}_{s}(A),s+1). $$ Applying a change of variables and $log_{p}$ to the above inequality, we get $$ \varphi(H^{Q}_{n-1}(A),n-1,t)+\varphi(H^{Q}_{n}(A),n,t) \leq d+\sum_{s=1}^{n-2}\varphi(H^{Q}_{s}(A),s+1). $$ By \prop{poieq}, there is a polynomial $f(t)$ of degree $n-2$, non-negative integer $a$, and positive integers $b$ and $d$ such that $$ at^{n-2}+bt^{n-1} \leq d+f(t) $$ which is clearly false for $n>1$. Thus $n=1$. The rest of the proof follows from \prop{prop3.7}. \hfill $\Box$ \bigskip {\em Proof of \thm{thm1.5}.} The ``if'' part is \thm{thm1.4} (3). We thus concentrate on the ``only if'' part. We are given a finite simplicial augmented commutative $\Bbb{F}$-algebra $A$ such that $\pi_{*}A$ is bounded. By \thm{thm1.4} (4) and (5), $A(1)$ is connected, $Q$-bounded, finite, and $\pi_{*}A(1)$ is of finite-type. Thus $A(1)$ is bounded as well and we conclude $A(1) \cong S(H^{Q}_{1}(A), 1)$ by \thm{connprop}. \hfill $\Box$ \bigskip
1997-07-24T10:02:49
9707
alg-geom/9707015
en
https://arxiv.org/abs/alg-geom/9707015
[ "alg-geom", "math.AG" ]
alg-geom/9707015
Arnaud Beauville
Arnaud Beauville
Fano contact manifolds and nilpotent orbits
Plain TeX. Postscript file available at http://www..dmi.ens.fr/users/beauvill/
null
null
null
null
A contact structure on a complex manifold M is a corank 1 subbundle F of T(M) such that the bilinear form on F with values in the quotient line bundle L=T(M)/F deduced from the Lie bracket of vector fields is everywhere non-degenerate. In this paper we consider the case where M is a Fano manifold; this implies that L is ample. If g is a simple Lie algebra, the unique closed orbit in P(g) (for the adjoint action) is a Fano contact manifold; it is conjectured that every Fano contact manifold is obtained in this way. A positive answer would imply an analogous result for compact quaternion-Kahler manifolds with positive scalar curvature, a longstanding question in Riemannian geometry. In this paper we solve the conjecture under the additional assumptions that the group of contact automorphisms of M is reductive, and that the image of the rational map M - - -> P(H0(M,L)*) associated to L has maximum dimension. The proof relies on the properties of the nilpotent orbits in a semi-simple Lie algebra, in particular on the work of R. Brylinski and B. Kostant.
[ { "version": "v1", "created": "Thu, 24 Jul 1997 08:02:40 GMT" } ]
2008-02-03T00:00:00
[ [ "Beauville", "Arnaud", "" ] ]
alg-geom
\global\def\currenvir{subsection{\global\defth}\smallskip{subsection} \global\advance\prmno by1 \par\hskip 1truecm\relax{ (\number\secno.\number\prmno) }} \def\global\def\currenvir{subsection{\global\defth}\smallskip{subsection} \global\advance\prmno by1 {(\number\secno.\number\prmno)\ }} \def\proclaim#1{\global\advance\prmno by 1 {\bf #1 \the\secno.\the\prmno$.-$ }} \long\def\th#1 \enonce#2\endth{% \medbreak\proclaim{#1}{\it #2}\global\defth}\smallskip{th}\smallskip} \def\rem#1{\global\advance\prmno by 1 {\it #1} \the\secno.\the\prmno$.-$ } \def\isinlabellist#1\of#2{\notfoundtrue% {\def\given{#1}% \def\\##1{\def\next{##1}% \lop\next\to\za\lop\next\to\zb% \ifx\za\given{\zb\global\notfoundfalse}\fi}#2}% \ifnotfound{\immediate\write16% {Warning - [Page \the\pageno] {#1} No reference found}}% \fi}% \def\ref#1{\ifx\empty{\immediate\write16 {Warning - No references found at all.}} \else{\isinlabellist{#1}\of}\fi} \def\newlabel#1#2{\rightappenditem{\\{#1}\\{#2}}\to} \def{} \def\label#1{% \def\given{th}% \ifx\giventh}\smallskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \def\given{section}% \ifx\giventh}\smallskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno}}}\fi% \def\given{subsection}% \ifx\giventh}\smallskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \ignorespaces} \newwrite\lbl \def\openout\lbl=\jobname.lbl{\openout\lbl=\jobname.lbl} \def\closeout\lbl{\closeout\lbl} \newread\testfile \def\lookatfile#1{\openin\testfile=\jobname.#1 \ifeof\testfile{\immediate\openout\nameuse{#1}\jobname.#1 \write\nameuse{#1}{} \immediate\closeout\nameuse{#1}}\fi% \immediate\closein\testfile}% \def\newlabel{main}{0.1{\newlabel{main}{0.1} \newlabel{split}{1.1} \newlabel{equi}{1.2} \newlabel{symplectization}{1.4} \newlabel{diagramme}{1.5} \newlabel{momentmap}{1.6} \newlabel{openorbit}{1.7} \newlabel{ample}{1.8} \newlabel{koki}{2.1} \newlabel{nilorb}{2.2} \newlabel{semi-simple}{2.4} \newlabel{simple}{2.5} \newlabel{Omin}{2.6} \newlabel{unique}{2.8} \newlabel{center}{3.1} \newlabel{codim}{3.2} \newlabel{finite}{3.3} \newlabel{H3}{3.4} \newlabel{summary}{3.5} \newlabel{converse}{3.6} \newlabel{dyn}{4.1} \newlabel{gpn}{4.2} \newlabel{noyau}{4.3} \newlabel{key}{4.4} \newlabel{classif}{5.1} \newlabel{closmooth}{5.2} \newlabel{list}{6.2} \newlabel{coverings}{6.3} \newlabel{Galois}{6.4} \newlabel{arg}{6.5} \newlabel{sp}{6.7} \newlabel{nonsimple}{6.8}} \magnification 1250 \pretolerance=500 \tolerance=1000 \brokenpenalty=5000 \mathcode`A="7041 \mathcode`B="7042 \mathcode`C="7043 \mathcode`D="7044 \mathcode`E="7045 \mathcode`F="7046 \mathcode`G="7047 \mathcode`I="7049\mathchardef\H="7048 \mathcode`J="704A \mathcode`K="704B \mathcode`L="704C \mathcode`M="704D \mathchardef\N="704E \mathcode`O="704F \mathcode`P="7050 \mathchardef\Q="7051 \mathcode`R="7052 \mathcode`S="7053 \mathcode`T="7054 \mathchardef\U="7055 \mathcode`V="7056 \mathcode`W="7057 \mathchardef\Y="7059 \mathcode`Z="705A\mathchardef\X="7058 \def\spacedmath#1{\def\packedmath##1${\bgroup\mathsurround =0pt##1\egroup$} \mathsurround#1 \everymath={\packedmath}\everydisplay={\mathsurround=0pt}} \def\nospacedmath{\mathsurround=0pt \everymath={}\everydisplay={} } \spacedmath{2pt} \def\qfl#1{\buildrel {#1}\over {\longrightarrow}} \def\phfl#1#2{\baselineskip=0truept{\baselineskip=0pt \lineskip=10truept\lineskiplimit=1truept}\nospacedmath\smash {\mathop{\hbox to 8truemm{\rightarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\hfl#1#2{\baselineskip=0truept{\baselineskip=0truept \lineskip=10truept\lineskiplimit=1truept}\nospacedmath\smash{\mathop{\hbox to 12truemm{\rightarrowfill}}\limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\diagram#1{\def\baselineskip=0truept{\baselineskip=0truept \lineskip=10truept\lineskiplimit=1truept} \matrix{#1}} \def\vfl#1#2{\llap{$\scriptstyle#1$}\left\downarrow\vbox to 6truemm{}\right.\rlap{$\scriptstyle#2$}} \font\eightrm=cmr8 \font\sixrm=cmr6 \def\note#1#2{\footnote{\parindent 0.4cm$^#1$}{\vtop{\eightrm\baselineskip12pt\hsize15.5truecm\noindent #2}} \parindent 0cm} \def\lhook\joinrel\mathrel{\longrightarrow}{\lhook\joinrel\mathrel{\longrightarrow}} \def\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}}{\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}}} \def\gdir_#1^#2{\mathrel{\mathop{\scriptstyle\bigoplus}\limits_{#1}^{#2}}} \def\sdir_#1^#2{\mathrel{\mathop{\kern0pt\oplus}\limits_{#1}^{#2}}} \def\pprod_#1^#2{\raise 2pt \hbox{$\mathrel{\scriptstyle\mathop{\kern0pt\prod}\limits_{#1}^{#2}}$}} \def\pc#1{\tenrm#1\sevenrm} \def\up#1{\raise 1ex\hbox{\sevenrm#1}} \def\kern-1.5pt -{\kern-1.5pt -} \def\cqfd{\kern 2.5truemm\unskip\penalty 500\vrule height 4pt depth 0pt width 4pt\medbreak} \def\vrule height 5pt depth 0pt width 5pt{\vrule height 5pt depth 0pt width 5pt} \def\virg{\raise .4ex\hbox{,}} \def\decale#1{\smallbreak\hskip 28pt\llap{#1}\kern 5pt} \defn\up{o}\kern 2pt{n\up{o}\kern 2pt} \def\par\hskip 1truecm\relax{\par\hskip 1truecm\relax} \def\par\hskip 0.5truecm\relax{\par\hskip 0.5truecm\relax} \def\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}{\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}} \def\kern 1pt{\scriptstyle\circ}\kern 1pt{\kern 1pt{\scriptstyle\circ}\kern 1pt} \def\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}}{\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}}} \def\mathop{\rm End}\nolimits{\mathop{\rm End}\nolimits} \def\mathop{\rm Hom}\nolimits{\mathop{\rm Hom}\nolimits} \def\mathop{\rm Aut}\nolimits{\mathop{\rm Aut}\nolimits} \def\mathop{\rm Im}\nolimits{\mathop{\rm Im}\nolimits} \def\mathop{\rm Ker}\nolimits{\mathop{\rm Ker}\nolimits} \def\mathop{\rm Coker}{\mathop{\rm Coker}} \def\mathop{\rm det}\nolimits{\mathop{\rm det}\nolimits} \def\mathop{\rm Pic}\nolimits{\mathop{\rm Pic}\nolimits} \def\mathop{\rm Div}\nolimits{\mathop{\rm Div}\nolimits} \def\mathop{\rm dim}\nolimits{\mathop{\rm dim}\nolimits} \def\mathop{\rm Card}\nolimits{\mathop{\rm Card}\nolimits} \def\mathop{\rm Tr}\nolimits{\mathop{\rm Tr}\nolimits} \def\mathop{\rm rk\,}\nolimits{\mathop{\rm rk\,}\nolimits} \def\mathop{\rm div\,}\nolimits{\mathop{\rm div\,}\nolimits} \def\mathop{\rm Ad}\nolimits{\mathop{\rm Ad}\nolimits} \def\mathop{\rm Res}\nolimits{\mathop{\rm Res}\nolimits} \def\mathop{\rm Lie}\nolimits{\mathop{\rm Lie}\nolimits} \def\mathop{\oalign{lim\cr\hidewidth$\longrightarrow $\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longrightarrow $\hidewidth\cr}}} \def\mathop{\oalign{lim\cr\hidewidth$\longleftarrow $\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longleftarrow $\hidewidth\cr}}} \def\mathop{\rm Ann}\nolimits{\mathop{\rm Ann}\nolimits} \def\mathop{\rm ad}\nolimits{\mathop{\rm ad}\nolimits} \def\mathop{\rm Ad}\nolimits{\mathop{\rm Ad}\nolimits} \def\mathop{\rm codim}\nolimits{\mathop{\rm codim}\nolimits} \def\mathop{\rm Spec}\nolimits{\mathop{\rm Spec}\nolimits} \font\san=cmssdc10 \def\hbox{\san \char3}{\hbox{\san \char3}} \def\hbox{\san \char83}{\hbox{\san \char83}} \def\hbox{\san \char88}{\hbox{\san \char88}} \def^{\scriptscriptstyle\times }{^{\scriptscriptstyle\times }} \input amssym.def \input amssym \catcode`\@=11 \mathchardef\dabar@"0\msafam@39 \def\longdash#1{\mathop{\dabar@\dabar@\dabar@\dabar@ \dabar@\mathchar"0\msafam@4B}\limits^{\scriptstyle#1}} \catcode`\@=12 \vsize = 25truecm \hsize = 16truecm \voffset = -.5truecm \parindent=0cm \baselineskip15pt \overfullrule=0pt \newlabel{main}{0.1 \centerline{\bf Fano contact manifolds and nilpotent orbits} \smallskip \smallskip \centerline{Arnaud {\pc BEAUVILLE\note{1}{Partially supported by the European HCM project ``Algebraic Geometry in Europe" (AGE).}}} \vskip1truecm {\bf Introduction} \smallskip \par\hskip 1truecm\relax A {\it contact structure} on a complex manifold $M$ is a corank $1$ subbundle $F\i T_M$ such that the bilinear form on $F$ with values in the quotient line bundle $L=T_M/F$ deduced from the Lie bracket on $T_M$ is everywhere non-degenerate. This implies that the dimension of $M$ is odd, say $\mathop{\rm dim}\nolimits M=2n+1$, and that the canonical bundle $K_M$ is isomorphic to $L^{-n-1}$. In this paper we will consider the case where $M$ is compact and $L$ is {\it ample}, that is, $M$ is a {\it Fano manifold}. \par\hskip 1truecm\relax This turns out to be a strong restriction on the manifold $M$; the only examples known so far are obtained as follows (see Prop.\ \ref{Omin} and \ref{nilorb} below). Let ${\goth g}$ be a simple complex Lie algebra; the adjoint group acting on ${\bf P}({\goth g})$ has exactly one closed orbit ${\bf P}{\cal O}_{\rm min}$, which is the projectivization of the {\it minimal nilpotent orbit} ${\cal O}_{\rm min}\i {\goth g}$. The Kostant-Kirillov symplectic structure on ${\cal O}_{\rm min}$ defines a contact structure on ${\bf P}{\cal O}_{\rm min}$. \par\hskip 1truecm\relax It is generally conjectured that {\it every Fano contact manifold is obtained in this way}. This problem is motivated by Riemannian geometry, more precisely by the study of compact {\it quaternion-K\"ahler} manifolds. I~will say only a few words here, referring for instance to [L-S], [L] and the bibliography therein for a more complete treatment. A quaternion-K\"ahler manifold $\Q$ is a Riemannian manifold with holonomy ${\it Sp}(n){\it Sp}(1)$. It carries a natural $S^2$\kern-1.5pt - bundle $M\rightarrow \Q$, the {\it twistor space}, which turns out to be a complex contact manifold; moreover if $\Q$ is compact and its scalar curvature is positive, $M$ is a Fano contact manifold. The only known examples of positive quaternion-K\"ahler manifolds are certain symmetric spaces associated to each compact simple Lie group, the so-called ``Wolf spaces"; thanks to the work of LeBrun and Salamon, a positive answer to the above conjecture would imply that every compact quaternion-K\"ahler manifold with positive scalar curvature is isometric to a Wolf space. \par\hskip 1truecm\relax Our result is the following: \th Theorem \enonce Let $M$ be a Fano contact manifold, satisfying the following conditions: \par\hskip 0.5truecm\relax{\rm (H1)} The rational map $\varphi^{}_L:M\dasharrow {\bf P}(\H^0(M,L)^*)$ associated to the line bundle $L$ is generically finite {\rm (}that is, $\mathop{\rm dim}\nolimits \varphi^{} _L(M)=\mathop{\rm dim}\nolimits M${\rm );} \par\hskip 0.5truecm\relax{\rm (H2)} The group $G$ of contact automorphisms of $M$ is reductive. \par\hskip 1truecm\relax Then the Lie algebra ${\goth g}$ of $G$ is simple, and $M$ is isomorphic to the minimal orbit ${\bf P}{\cal O}_{\rm min}\i{\bf P}({\goth g})$. \endth\label{main} \par\hskip 1truecm\relax While hypothesis (H1) is rather strong, (H2) is harmless from the point of view of Riemannian geometry: by the results of [L], it always holds for the twistor spaces of positive quaternion-K\"ahler manifolds. \par\hskip 1truecm\relax We will get an apparently stronger result, namely that $M$ and ${\bf P}{\cal O}_{\rm min}$ are isomorphic as {\it contact} complex manifolds. It is however a general fact that whenever two compact simply-connected contact manifolds are isomorphic, the isomorphism can be chosen compatible with the contact structures ([L], Prop.\ 2.3). \par\hskip 1truecm\relax The strategy of the proof is as follows. Using some elementary symplectic geometry, the map $\varphi^{}_L$ can be viewed as a ``contact moment map" $M\rightarrow {\bf P}({\goth g})$. Then (H1) implies that $G$ has an open orbit in $M$, whose image by $\varphi^{}_L$ is a nilpotent orbit ${\bf P}{\cal O}\i {\bf P}({\goth g})$. We are thus led to classify finite $G$\kern-1.5pt - equivariant coverings $M\rightarrow \overline{{\bf P}{\cal O}}$, where $M$ is smooth. Examples of such coverings appear in [B-K], with $M$ being the minimal orbit in ${\bf P}({\goth g}')$ for some simple Lie algebra ${\goth g}'$ containing ${\goth g}$; our key result is that all possible examples arise essentially in this way. Theorem \ref{main} follows then easily. \vskip1truecm \section {Contact geometry} \par\hskip 1truecm\relax Let $M$ be a complex contact projective manifold. Recall that the contact structure is given by an exact sequence $$0\rightarrow F\longrightarrow T_M\qfl{\theta }L\rightarrow 0\ ,$$such that the (${\cal O}_M$\kern-1.5pt - bilinear) alternate form $(X,Y)\mapsto \theta ([X,Y])$ on $F$ is everywhere non-degenerate. Alternatively the contact structure can be described by the twisted 1-form $\theta\in \H^0(M,\Omega^1_M\otimes L)$, the {\it contact form}. \par\hskip 1truecm\relax We denote by $G$ the neutral component of the group of automorphisms of $M$ preserving $F$. This is an algebraic group, whose Lie algebra ${\goth g}$ consists of the vector fields $X\in \H^0(M, T_M)$ such that $[X,F]\i F$. The following result is well-known (see e.g.\ [L]): \th Proposition \enonce The map $\H^0(\theta ):\H^0(M,T_M)\rightarrow \H^0(M, L)$ maps ${\goth g}$ isomorphically onto $\H^0(M,L)$. \endth\label{split} {\it Proof}: Let us first prove the decomposition $\H^0(M,T_M)=\H^0(M,F)\oplus {\goth g}$. Let $X\in \H^0(M,T_M)$. The map $U\mapsto \theta([X,U])$ from $F$ to $L$ is ${\cal O}_M$\kern-1.5pt - linear, hence there exists a unique vector field $X'$ in $F$ such that $\theta([X,U])=\theta([X',U])$ for all $U$ in $F$. This means that $[X-X',U]$ belongs to $F$, that is that $X-X'$ belongs to ${\goth g}$. Writing $X=X'+(X-X')$ provides the required direct sum decomposition. \par\hskip 1truecm\relax Let ${\cal L}\i T_M$ be the subsheaf of infinitesimal contact transformations. Applying the above result to each open subset of $M$ we get $T_M=F\oplus {\cal L}$, so that $\theta$ induces a (${\bf C}$\kern-1.5pt - linear) isomorphism of ${\cal L}$ onto $L$. Our statement follows by taking global sections.\cqfd \smallskip \global\def\currenvir{subsection\label{equi} For each $g\in G$ the automorphism $T(g)$ of $T_M$ induces an automorphism of $L$ above $g$; in other words, the line bundle $L$ has a canonical $G$\kern-1.5pt - linearization. In particular the group $G$ acts on $\H^0(M,L)$; the isomorphism $\theta :{\goth g}\rightarrow \H^0(M,L)$ is $G$\kern-1.5pt - equivariant with respect to this action and the adjoint action on ${\goth g}$. Also the rational map $\varphi^{} _L:M\dasharrow {\bf P}(\H^0(M,L)^*)$ associated to the line bundle $L$ is $G$\kern-1.5pt - equivariant. \global\def\currenvir{subsection Let $L^{\scriptscriptstyle\times }$ be the principal ${\bf C}^*$\kern-1.5pt - bundle associated to the {\it dual} line bundle $L^*$ -- that is the complement of the zero section in $L^*$, on which ${\bf C}^*$ acts by homotheties. We will say that a $p$\kern-1.5pt - form $\omega $ on $L^{\scriptscriptstyle\times }$ is ${\bf C^*}$\kern-1.5pt - {\it equivariant} if $\lambda ^*\omega =\lambda \omega $ for every $\lambda \in{\bf C}^*$. \par\hskip 1truecm\relax We have a canonical linear form $\tau :p^*L\rightarrow {\cal O}_{L^*}$, which is bijective on $L^{\scriptscriptstyle\times }$: if $s$ is a local section of $L$ on $M$, the function $\tau (p^*s)$ maps a point $(m,\xi )$ of $L^*$ $(\xi \in L(m)^*)$ to $\langle s(m),\xi \rangle$. We use $\tau $ to trivialize $p^*L$ on $L^{\scriptscriptstyle\times }$. We can therefore consider $p^*\theta $ as a $1$\kern-1.5pt - form on $L^{\scriptscriptstyle\times }$; it is ${\bf C}^*$\kern-1.5pt - equivariant. The following lemma is classical (see for instance [A], App.\ 4 E, or [L], p. 425): \th Lemma \enonce The $2$\kern-1.5pt - form $d(p^*\theta ) $ is a symplectic structure on $L^{\scriptscriptstyle\times }$. Conversely, any ${\bf C}^*$\kern-1.5pt - equivariant symplectic $2$\kern-1.5pt - form on $L^{\scriptscriptstyle\times }$ is of the form $d(p^*\theta)$, where $\theta$ is a contact form on $M$, which is uniquely determined.\cqfd \endth\label{symplectization} \smallskip \global\def\currenvir{subsection\label{diagramme} To each point $(m,\xi )$ of $L^*$ $(m\in M$, $\xi \in L(m)^*)$, we associate the linear form $\mu_L(m,\xi ) $ on $\H^0(M,L)$ defined by $\langle\mu_L(m,\xi ),s\rangle=\langle s(m),\xi \rangle$ for each $s\in \H^0(M,L)$. This gives a morphism $\mu _L:L^*\rightarrow \H^0(M,L)^*$ which is ${\bf C}^*$\kern-1.5pt - equi\-va\-riant and induces on the projectivizations the rational map $\varphi^{} _L:M \dasharrow {\bf P}(\H^0(M,L)^*)$. Using the isomorphism $\theta:{\goth g}\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}} \H^0(M,L)$ (Prop.\ \ref{split}), we get a commutative $G$\kern-1.5pt - equi\-variant diagram \def\dia#1{\def\baselineskip=0truept{\baselineskip=0truept} \offinterlineskip \matrix{#1}} \def\vrule height 12pt depth 5pt width 0pt\vrule{\vrule height 2pt depth 0pt width 0.4pt} \def\vrule height 12pt depth 5pt width 0pt{\vrule height 2pt depth 1pt width 0pt} $$\dia{ {L^{\scriptscriptstyle\times }} &\kern-5pt \hfl{\mu }{}\kern-5pt & {\goth g}^*&\cr \vrule height 2pt depth 1pt width 0pt\cr \vrule height 12pt depth 5pt width 0pt\vrule & & \vrule height 12pt depth 5pt width 0pt\vrule&\cr \vrule height 12pt depth 5pt width 0pt\vrule & &\cr \vrule height 12pt depth 5pt width 0pt\vrule & & \vrule height 12pt depth 5pt width 0pt\vrule\cr\vrule height 12pt depth 5pt width 0pt\vrule & &\cr \llap{$\scriptstyle p\ $} \vrule height 12pt depth 5pt width 0pt\vrule & & \vrule height 12pt depth 5pt width 0pt\vrule&\cr \vrule height 12pt depth 5pt width 0pt\vrule & &\cr \downarrow & &\downarrow\cr M & \kern-8pt\longdash{\varphi }{} \kern-8pt& {\bf P}({\goth g}^*)&\kern-5pt . }$$ \par\hskip 1truecm\relax As we have seen in (\ref{equi}), the action of $G$ on $M$ lifts to an action on $L^{\scriptscriptstyle\times }$, which is linear on the fibres; similarly any field $X\in{\goth g}$ lifts to a vector field $\widetilde{X}$ on $L^{\scriptscriptstyle\times }$ which projects to $X$ on $M$. \th Proposition \enonce $\mu $ is a moment map for the action of $G$ on the symplectic manifold $L^{\scriptscriptstyle\times }$. \endth\label{momentmap} {\it Proof}: This means by definition that for each $X\in{\goth g}$, the vector field $\widetilde{X}$ is the Hamiltonian vector field associated to the function $\langle\mu ,X\rangle$ on $L^{\scriptscriptstyle\times }$. To prove this, we first observe that since the $1$\kern-1.5pt - form $\eta =p^*\theta $ is preserved by $G$, its Lie derivative $L_{\widetilde{X}}\eta $ vanishes for each $X\in{\goth g}$. By the Cartan homotopy formula, this implies $i(\widetilde{X})\,d\eta = -d\langle\eta ,\widetilde{X}\rangle$. But we have $\langle\eta ,\widetilde{X}\rangle = \tau (p^*\theta (X) )=\langle\mu ,X\rangle$, thus $i(\widetilde{X})\,d\eta = -d\langle\mu ,X\rangle$, which proves our claim.\cqfd \smallskip \par\hskip 1truecm\relax The classical computation of the differential of the moment map gives: \th Proposition \enonce Let $m\in M$, and $\xi$ a point of $L^{\scriptscriptstyle\times }$ above $m$. The following conditions are equivalent: \par\hskip 0.5truecm\relax{\rm (i)} $\varphi$ is defined at $m$ and its differential $T_m(\varphi)$ is injective; \par\hskip 0.5truecm\relax{\rm (ii)} the $G$\kern-1.5pt - orbit of $\xi$ is open in $L^{\scriptscriptstyle\times }$; \par\hskip 0.5truecm\relax{\rm (iii)} the $G$\kern-1.5pt - orbit of $m$ is open in $M$ and $\xi$ is conjugate under $G$ to $\ell \xi$ for every $\ell \in{\bf C}^*$. \endth\label{openorbit} {\it Proof}: Since $\mu$ is ${\bf C}^*$\kern-1.5pt - equivariant, condition (i) is equivalent to: \par\hskip 1truecm\relax ${\rm (i')}$ $\mu(\xi)\not=0$ {\it and} $ T_\xi(\mu)$ {\it is injective}. Let $\omega$ be the symplectic 2-form on $L^{\scriptscriptstyle\times }$; for $v\in T_\xi(L^{\scriptscriptstyle\times })$ and $X\in{\goth g}$, the formula $i(\widetilde{X})\,\omega = -d\langle\mu ,X\rangle$ (\ref{momentmap}) gives $$\langle T_\xi(\mu)\cdot v\,,\,X\rangle = -\langle i(\widetilde{X})\omega_\xi\,,\,v \rangle = \omega_\xi(v,\widetilde{X}(\xi))\ , $$ so that the kernel of $T_\xi(\mu)$ is the orthogonal of $T_\xi(G\cdot \xi)$ in $T_\xi(L^{\scriptscriptstyle\times })$ (with respect to $\omega_\xi$). This gives the equivalence of ${\rm (i')}$ and (ii); since the action of $G$ commutes with the homotheties, (ii) is equivalent to (iii).\cqfd \smallskip \th Corollary \enonce {\rm a)} If $L$ is very ample, $M$ is homogeneous. \par\hskip 1truecm\relax {\rm b)} If $\varphi$ is generically finite, $M$ contains an open $G$\kern-1.5pt - orbit. \endth \label{ample} {\it Proof}: Under the hypothesis of a), each point of $M$ has an open orbit, thus necessarily equal to $M$. The hypothesis of b) implies that $\varphi$ is an immersion at a general point of $M$.\cqfd \smallskip \par\hskip 1truecm\relax Cor.\ \ref{ample} a) has also been obtained by J. Wisniewski (private communication). \vskip1truecm \section {Coadjoint orbits} \global\def\currenvir{subsection\label{koki} Let ${\goth g}$ be a Lie algebra; the adjoint group $G$ acts on the dual ${\goth g}^*$ of ${\goth g}$ through the coadjoint representation. Recall that each coadjoint orbit ${\cal O}$ carries a canonical $G$\kern-1.5pt - invariant symplectic structure $\Omega $, the {\it Kostant-Kirillov} structure: for $\xi \in{\cal O}$, the tangent space $T_\xi ({\cal O})$ is canonically isomorphic to ${\goth g}/{\goth z}^{}_\xi $, where ${\goth z}^{}_\xi=\mathop{\rm Ker}\nolimits(\xi\kern 1pt{\scriptstyle\circ}\kern 1pt\mathop{\rm ad}\nolimits)$ is the annihilator of $\xi $ in ${\goth g}$; the 2-form $\Omega _\xi $ is induced by the alternate form $(X,Y)\mapsto \xi ([X,Y])$ on ${\goth g}$. The following result shows that whenever ${\cal O}$ is invariant under homotheties, its image ${\bf P}{\cal O}$ in ${\bf P}({\goth g}^*)$ carries a natural contact structure: \th Proposition \enonce Let ${\goth g}$ be a Lie algebra, $G$ its adjoint group, $\xi $ a nonzero linear form on ${\goth g}$, ${\cal O}$ its coadjoint orbit in ${\goth g}^*$, ${\bf P}{\cal O}$ the image of ${\cal O}$ in ${\bf P}({\goth g}^*)$. The following conditions are equivalent: \par\hskip 0.5truecm\relax{\rm (i)} ${\bf P}{\cal O}$ is odd-dimensional; \par\hskip 0.5truecm\relax{\rm (ii)} the orbit ${\cal O}\i{\goth g}^*$ is invariant by homotheties; \par\hskip 0.5truecm\relax{\rm (iii)} for each $\ell \in{\bf C}^*$, $\ell \, \xi $ is $G$\kern-1.5pt - conjugate to $\xi$; \par\hskip 0.5truecm\relax{\rm (iv)} there exists $H\in {\goth g}$ such that $\xi \kern 1pt{\scriptstyle\circ}\kern 1pt \mathop{\rm ad}\nolimits(H)=\xi$; \par\hskip 0.5truecm\relax{\rm (v)} the annihilator ${\goth z}^{}_\xi $ of $\xi $ in ${\goth g}$ is contained in $\mathop{\rm Ker}\nolimits\xi $. \par\hskip 1truecm\relax When these conditions are satisfied, the Kostant-Kirillov symplectic structure on ${\cal O}$ comes from a $G$\kern-1.5pt - invariant contact structure on ${\bf P}{\cal O}$. \endth\label{nilorb} \par\hskip 0.5truecm\relax ${\rm (i)\Leftrightarrow (iii)}$: Let $Z_\xi$ be the stabilizer of $\xi $ in $G$, and $Z_{[\xi ]}$ the stabilizer of the image $[\xi ]$ of $\xi$ in ${\bf P}({\goth g}^*)$. The action of $Z_{[\xi]} $ on the line $[\xi]$ defines a homomorphism $\ell :Z_{[\xi ]}\rightarrow {\bf C}^*$, and we have an exact sequence $$0\rightarrow Z_\xi \longrightarrow Z_{[\xi ]}\qfl{\ell }{\bf C}^*\ . $$Since the orbit ${\cal O}$ is even-dimensional, (i) is equivalent to $\mathop{\rm dim}\nolimits Z_{[\xi ]}=\mathop{\rm dim}\nolimits Z_\xi +1$, that is to the surjectivity of $\ell $, which is nothing but condition (iii). \par\hskip 0.5truecm\relax ${\rm (ii)\Leftrightarrow (iii)}$: Clear. \par\hskip 0.5truecm\relax ${\rm (iii)\Leftrightarrow (iv)}$: The Lie algebra ${\goth z}^{}_{[\xi]} $ of $Z_{[\xi]} $ consists of the elements $H$ of ${\goth g}$ such that $\xi \kern 1pt{\scriptstyle\circ}\kern 1pt \mathop{\rm ad}\nolimits(H)=\lambda \xi $ for some $\lambda=\lambda (H) \in{\bf C}$. The homomorphism $\lambda :{\goth z}^{}_{[\xi]}\rightarrow {\bf C}$ thus defined is the Lie derivative of $\ell $, so the surjectivity of $\ell $ is equi\-valent to the surjectivity of $\lambda $, that is to (iv). \par\hskip 0.5truecm\relax ${\rm (iv)\Leftrightarrow (v)}$: The linear map $u:H\mapsto \xi \kern 1pt{\scriptstyle\circ}\kern 1pt \mathop{\rm ad}\nolimits(H)$ of ${\goth g}$ into ${\goth g}^*$ is antisymmetric, hence $\mathop{\rm Im}\nolimits u = (\mathop{\rm Ker}\nolimits u)^\perp$. But (iv) is equivalent to $\xi \in\mathop{\rm Im}\nolimits u$ and (v) to $\xi \in (\mathop{\rm Ker}\nolimits u)^\perp$. \smallskip \par\hskip 1truecm\relax Finally when ${\cal O}$ is invariant by homotheties, the Kostant-Kirillov 2-form on ${\cal O}$ is ${\bf C}^*$\kern-1.5pt - equivariant, and therefore comes from a $G$\kern-1.5pt - invariant contact structure on ${\bf P}{\cal O}$ (lemma \ref{symplectization}).\cqfd \smallskip \rem{Remark} Assume that the equivalent conditions of Prop.\ \ref{nilorb} hold; the contact structure on ${\bf P}{\cal O}$ can be described explicitely as follows. Let $\psi\in{\cal O}$; the tangent space $T_{[\psi]}({\bf P}{\cal O})$ is canonically isomorphic to ${\goth g}/{\goth z}_{[\psi]}$. Observe that ${\goth z}_{[\psi]}$ {\it is contained in} $\mathop{\rm Ker}\nolimits \psi$: each element $Z$ of ${\goth z}_{[\psi]}$ satisfies $\psi\kern 1pt{\scriptstyle\circ}\kern 1pt \mathop{\rm ad}\nolimits(Z)=\lambda \psi$ for some $\lambda\in{\bf C}$; if $\lambda=0$ we have $\psi(Z)=0$ by (v) above, while if $\lambda\not=0$ we have $\psi(Z)=\lambda^{-1} \psi([Z,Z])=0$. Then the contact structure $F\i T_{{\bf P}{\cal O}}$ is defined by $F_{[\psi]}= (\mathop{\rm Ker}\nolimits \psi)/{\goth z}_{[\psi]} $. \medskip \global\def\currenvir{subsection\label{semi-simple} Suppose that the Lie algebra ${\goth g}$ is {\it semi-simple}. Using the Killing form we identify the $G$\kern-1.5pt - module ${\goth g}^*$ to ${\goth g}$ endowed with the adjoint action. The element $\xi $ corresponds to a nonzero element $N$ of ${\goth g}$. Conditions (iii) to (v) read: \par\hskip 0.5truecm\relax${\rm (iii')}$ for each $\ell \in{\bf C}^*$, $\ell N $ is $G$\kern-1.5pt - conjugate to $N$; \par\hskip 0.5truecm\relax${\rm (iv')}$ there exists $H\in {\goth g}$ such that $[H,N]=N $; \par\hskip 0.5truecm\relax${\rm (v')}$ the centralizer ${\goth z}^{}_N $ of $N $ in ${\goth g}$ is orthogonal to $N$. \par\hskip 1truecm\relax They are equivalent to $N$ being {\it nilpotent}: ${\rm (iii')}$ implies $\mathop{\rm Tr}\nolimits \rho (N)^p=0$ for any representation $\rho $ of ${\goth g}$ and any $p\ge 1$; conversely, if $N$ is nilpotent, ${\rm (iv')}$ holds by the Jacobson-Morozov theorem. \global\def\currenvir{subsection\label{simple} Let ${\goth h}$ be a Cartan subalgebra of ${\goth g}$, $R=R({\goth g},{\goth h})$ the root system of ${\goth g}$ relative to ${\goth h}$. We have a direct sum decomposition $\displaystyle {\goth g}={\goth h}\,\oplus \gdir_{\alpha\in R}^{}{\goth g}^\alpha$. A nonzero vector $X_\alpha\in {\goth g}^\alpha$ is called a {\it root vector} (relative to $\alpha $). \par\hskip 1truecm\relax If ${\goth g}$ is simple, the Weyl group acts transitively on the set of roots with a given length, and the corresponding root vectors are conjugate. This defines the (nilpotent) orbits ${\cal O}_{min}$ of a long root vector and ${\cal O}_{short}$ of a short root vector; these orbits coincide if and only if all roots have the same length (types $A_l,D_l,E_l$). \th Proposition \enonce Let ${\goth g}$ be a simple complex Lie algebra. There exists exactly one closed orbit in ${\bf P}({\goth g})$ {\rm (}for the adjoint action{\rm ),} namely the orbit ${\bf P}{\cal O}_{min}$ of a long root vector. Every orbit contains ${\bf P}{\cal O}_{min}$ in its closure. \endth\label{Omin} {\it Proof}: Let $N$ be a nonzero element of ${\goth g}$. The orbit of $[N]$ in ${\bf P}({\goth g})$ is closed if and only if ${\goth z}^{}_{[N]}$ contains a Borel subalgebra ${\goth b}$, so that there exists a linear form $\lambda$ on ${\goth b}$ such that $\mathop{\rm ad}\nolimits(X)\cdot N=\lambda(X)N$ for all $X\in{\goth b}$. This means that $N$ is a highest weight vector for the adjoint representation; since ${\goth g}$ is simple, the adjoint representation is irreducible, and its highest weight vector is $X_\theta $, where $\theta $ is the highest root with respect to the basis of $R({\goth g},{\goth h})$ such that ${\goth b}={\goth h}\,\oplus\gdir_{\alpha\ge 0}^{}{\goth g}^\alpha$. We conclude that the orbit ${\bf P}{\cal O}_{min}$ of $X_\theta$ is the unique closed orbit in ${\bf P}({\goth g})$.\cqfd \medskip \rem{Examples} For the classical case, we get the following Fano contact manifolds: \par\hskip 0.5truecm\relax$A_l$: the projectivized cotangent bundle ${\bf P}T^*({\bf P}^l)$; \par\hskip 0.5truecm\relax$B_l,D_l$: the Grassmannian ${\bf G}_{iso}(2,V)$ of isotropic 2-planes in a quadratic vector space $V$, of dimension $2l+1$ and $2l$ respectively; \par\hskip 0.5truecm\relax$C_l$: the projective space ${\bf P}^{2l-1}$. \par\hskip 1truecm\relax For the type $G_2$ we get a Fano 5-fold of index 3 which appears in the work of Mukai [Mu]. The other exceptional Lie algebras give rise to Fano contact manifolds of dimension 15, 21, 33 and 57. \smallskip \rem{Remark}\label{unique} It follows from [L], Cor.\ 3.2, or from a direct computation, that if ${\goth g}$ is not of type $C_l$ the manifold ${\bf P}{\cal O}_{min}$ admits a {\it unique} contact structure; in all cases, the contact structure we have defined is the unique $G$\kern-1.5pt - invariant contact structure. \vskip1truecm \section{First consequences of (H1) and (H2)} \global\def\currenvir{subsection\label{center} From now on we assume that $\varphi$ {\it is generically finite} (or equivalently, $\mathop{\rm dim}\nolimits \varphi (M)=\mathop{\rm dim}\nolimits M$). By Cor.\ \ref{ample}, this implies that $G$ has an open orbit $M^{\rm o}$ in $M$. Since $\varphi $ is $G$\kern-1.5pt - equivariant, it is everywhere defined on $M^{\rm o}$; the image $\varphi (M^{\rm o})$ is an orbit ${\bf P}{\cal O}$ of $G$ in ${\bf P} ({\goth g}^*)$, and the induced map $\varphi ^{\rm o}:M^{\rm o}\rightarrow {\bf P}{\cal O}$ is a finite \'etale covering. \par\hskip 1truecm\relax Let us mention at once an immediate consequence: if a connected normal subgroup of $G$ fixes a point $[\xi ]\in{\bf P}{\cal O}$, it acts trivially on $M^{\rm o}$, hence on $M$; it follows that {\it the stabilizer ${\goth z}^{ }_{[\xi ]}$ of $[\xi ]$ in ${\goth g}$ contains no nonzero ideal of} ${\goth g}$. In particular, {\it the center of ${\goth g}$ is trivial}. \th Lemma \enonce Assume that the character group of $G$ is trivial, and $\mathop{\rm Pic}\nolimits(M)={\bf Z}$. Then $M\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}} M^{\rm o}$ has codimension $\ge 2$ in $M$. \endth\label{codim} {\it Proof}: Let $m$ be a point of $M^{\rm o}$, $[\xi ]$ its image in ${\bf P}({\goth g}^*)$. The stabilizer $Z_m$ of $m$ in $G$ is a subgroup of finite index in the stabilizer $Z_{[\xi ]}$ of $[\xi ]$. Since $M^{\rm o}$ and therefore ${\bf P}{\cal O}$ are odd-dimensional, the equivalent conditions of Prop.\ \ref{nilorb} are satisfied; hence the homomorphism $\ell :Z_{[\xi ]}\rightarrow {\bf C}^*$ deduced from the action of $Z_{[\xi ]}$ on the line $[\xi ]$ is surjective, and so is its restriction to $Z_m$. \par\hskip 1truecm\relax Recall that the group $\mathop{\rm Pic}\nolimits^G(M^{\rm o})$ of $G$\kern-1.5pt - linearized line bundles on $M^{\rm o}\cong G/Z_m$ is canonically isomorphic to the character group $\hbox{\san \char88}(Z_m)$. On the other hand, the hypothesis on $G$ ensures that the forgetful map $\mathop{\rm Pic}\nolimits^G(M^{\rm o})\rightarrow \mathop{\rm Pic}\nolimits(M^{\rm o})$ is injective ([M], Ch.\ 1, Prop.\ 1.4). Since we have found a surjective character of $Z_m$, it follows that $\mathop{\rm Pic}\nolimits(M^{\rm o})$ contains an infinite cyclic group. \par\hskip 1truecm\relax Let $(D_i)_{i\in I}$ be the family of one-codimensional components of $M\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}} M^{\rm o}$. We have an exact sequence $${\bf Z}^I\,\hfl{(D_i)}{}\, \mathop{\rm Pic}\nolimits(M)\longrightarrow \mathop{\rm Pic}\nolimits(M^{\rm o})\rightarrow 0\ .$$ Since $\mathop{\rm Pic}\nolimits(M)={\bf Z}$ and each $D_i$ has a nonzero class in $\mathop{\rm Pic}\nolimits(M)$, the only possibility is $I=\varnothing$.\cqfd \th Lemma \enonce Let $M$ be a normal projective variety, $L$ an ample line bundle on $M$, $\varphi :M\dasharrow {\bf P}^r$ the associated rational map, $\N\i{\bf P}^r$ its image. Assume that there are open subsets $M^{\rm o}\i M$ and $\N^{\rm o}\i \N$, whose complements have codimension $\ge 2$, such that $\varphi $ is defined everywhere on $M^{\rm o}$, $\varphi (M^{\rm o})=\N^{\rm o}$ and the induced morphism $\varphi^{\rm o}:M^{\rm o}\rightarrow \N^{\rm o}$ is finite. Then $\varphi $ is everywhere defined and finite. \endth\label{finite} {\it Proof}: Replacing $\N$ by its normalization we may assume that $\N$ is normal; then the restriction maps $\H^0(\N,{\cal O}_{\N}(n))\rightarrow \H^0(\N^{\rm o},{\cal O}_{\N}(n))$ and $\H^0(M,L^n)\rightarrow \H^0(M^{\rm o},L^n)$ are bijective. Let $CM=\mathop{\rm Spec}\nolimits \sdir_{n\ge 0}^{}\H^0(M,L^n)$ and $C\N=\mathop{\rm Spec}\nolimits \sdir_{n\ge 0}^{}\H^0(\N,{\cal O}_{\N}(n))$ be the cones over $M$ and $\N$ respectively associated to the line bundles $L$ and ${\cal O}_{\N}(1)$. The homomorphism $(\varphi^{\rm o})^*$ induces a finite morphism $C\varphi :CM\rightarrow C\N$, which is ${\bf C}^*$\kern-1.5pt - equivariant. The inverse image of the vertex of $C\N$ under $C\varphi $ is finite and stable under ${\bf C}^*$, hence reduced to the vertex of $CM$. Therefore $C\varphi $ induces a finite morphism $M\rightarrow \N$ which extends $\varphi ^{\rm o}$.\cqfd \smallskip \global\def\currenvir{subsection\label{H3} Let us now assume that ${\goth g}$ is reductive (this is our hypothesis (H2)). By (\ref{center}) this actually implies that ${\goth g}$ is {\it semi-simple}. We will always identify ${\goth g}^*$ with ${\goth g}$ using the Killing form. We also make a third hypothesis: \par\hskip 1truecm\relax (H3) $\mathop{\rm Pic}\nolimits(M)={\bf Z}$. \par This is innocuous because Theorem \ref{main} is known to be true when $b_2\ge 2$, as a consequence of a theorem of Wisniewski (see [L-S], cor.\ 4.2). \th Proposition \enonce Under the hypotheses {\rm (H1)} to {\rm (H3)}, the map $\varphi:M\rightarrow {\bf P}({\goth g})$ is a finite morphism onto the closure of a nilpotent orbit ${\bf P}{\cal O}$. $M$ has only finitely many orbits; each orbit is a finite \'etale covering of a nilpotent orbit in ${\bf P}({\goth g})$. \endth\label{summary} {\it Proof}: Since $G$ is semi-simple, the hypotheses of lemma \ref{codim} hold. We have already seen that the orbit ${\cal O}$ is ${\bf C}^*$\kern-1.5pt - invariant, hence nilpotent (\ref{semi-simple}). Therefore $\overline{{\bf P}{\cal O}}$ is a finite union of nilpotent orbits in ${\bf P}({\goth g})$. Since such an orbit is odd-dimensional, the codimension of ${\bf P}{\cal O}$ in $\overline{{\bf P} {\cal O}}$ is $\ge 2$, so we can apply lemma \ref{finite}; the Proposition follows.\cqfd \smallskip \rem{Remark}\label{converse} Conversely, suppose given a compact manifold $M$ with an action of $G$ and a finite surjective $G$\kern-1.5pt - equivariant morphism $\varphi :M\rightarrow \overline{{\bf P}{\cal O}}$ onto the closure of a nilpotent orbit in ${\bf P}({\goth g})$. Then $M$ {\it is a Fano contact manifold}. Indeed, let $M^{\rm o}=\varphi^{-1} ({\bf P}{\cal O})$, and $L=\varphi^*{\cal O}(1)$. The contact structure of ${\bf P}{\cal O}$ pulls back to a contact structure $\theta^{\rm o}\in \H^0(M^{\rm o},\Omega^1_{M^{\rm o}}\otimes L)$, which extends to a contact structure $\theta\in\H^0(M,\Omega^1_{M}\otimes L)$ because $M\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}} M^{\rm o}$ has codimension $\ge 2$. Since $L$ is ample, $M$ is a Fano contact manifold. \par\hskip 1truecm\relax We have thus reduced our problem to a question about nilpotent orbits of semi-simple Lie algebras, which we will study in the next sections. \vskip1truecm\section{Nilpotent orbits} \global\def\currenvir{subsection\label{dyn} At this point we need to recall Dynkin's classification of nilpotent orbits in a semi-simple Lie algebra ${\goth g}$ (a general reference for the material in this section is [C-M]). We fix a nilpotent element $N_0$ of ${\goth g}$, and denote by ${\cal O}$ its orbit in ${\goth g}$ (under the adjoint action). \par\hskip 1truecm\relax By the Jacobson-Morozov theorem, there exist elements $H$ and $N_1$ in ${\goth g}$ satisfying $$[H,N_0]=2N_0\qquad [H,N_1]=-2N_1 \qquad [N_1,N_0]=H\ ,$$ so that the subspace of ${\goth g}$ spanned by $N_0,N_1,H$ is a Lie subalgebra isomorphic to ${\goth s}{\goth l}_2$. As a ${\goth s}{\goth l}_2$\kern-1.5pt - module, ${\goth g}$ is then isomorphic to a direct sum of simple modules $\hbox{\san \char83}^kV$, where $V$ is the standard 2-dimensional representation. It follows easily that: \par\hskip 1truecm\relax (\ref{dyn}.{\it a}) there is a direct sum decomposition ${\goth g}=\sdir_{i\in{\bf Z}}^{}{\goth g}(i)$, where ${\goth g}(i)$ is the subspace of elements $ X\in {\goth g}$ with $[H,X]=iX$. \par\hskip 1truecm\relax (\ref{dyn}.{\it b}) Put ${\goth p}=\sdir_{i\ge 0}^{}{\goth g}(i)$, ${\goth n}=\sdir_{i\ge 2}^{}{\goth g}(i)$. Then ${\goth p}$ is a parabolic subalgebra of ${\goth g}$; ${\goth n}$ is a unipotent ideal in ${\goth p}$. The map $\mathop{\rm ad}\nolimits(N_0):{\goth p}\rightarrow {\goth n}$ is surjective. \par\hskip 1truecm\relax (\ref{dyn}.{\it c}) Let ${\goth h}$ be a Cartan subalgebra of ${\goth g}$ containing $H$. There exists a basis $B$ of the root system $R({\goth g},{\goth h})$ such that $\alpha (H)\in\{0,1,2\}$ for each $\alpha \in B$. The {\it weighted Dynkin diagram} of $N_0$ is obtained by labelling each node $\alpha \in B$ of the Dynkin diagram of ${\goth g}$ with the number $\alpha (H)\in\{0,1,2\}$. It depends only on the orbit ${\cal O}$ of $N_0$; two different nilpotent orbits give rise to different weighted diagrams. \medskip \defG\times ^P{\kern-1.5pt\goth n}{G\times ^P{\kern-1.5pt\goth n}} \def{\goth g}\times ^{\goth p}{\kern-1.5pt\goth n}{{\goth g}\times ^{\goth p}{\kern-1.5pt\goth n}} \def{\goth z}^{}_N{{\goth z}^{}_N} \global\def\currenvir{subsection\label{gpn} Let $P$ be the parabolic subgroup of $G$ with Lie algebra ${\goth p}$. We denote by $G\times ^P{\kern-1.5pt\goth n}$ the quotient of $G\times {\goth n}$ by $P$ acting by $p\cdot (g,N)=(gp^{-1},\mathop{\rm Ad}\nolimits(p)N)$; in other words, $G\times ^P{\kern-1.5pt\goth n}$ is the $G$\kern-1.5pt - homogeneous vector bundle on $G/P$ associated to the adjoint action of $P$ on ${\goth n}$. For $g\in G$, $N\in {\goth n}$, we denote by $(g,N)\dot{}$ the image of $(g,N)$ in $G\times ^P{\kern-1.5pt\goth n}$; the tangent space to $G\times ^P{\kern-1.5pt\goth n}$ at $(g,N)\dot{}$ is canonically isomorphic to the quotient of ${\goth g}\times {\goth n}$ by the subspace of elements $({\it P},[N,{\it P}])$ with ${\it P}\in{\goth p}$. \par\hskip 1truecm\relax The orbit $G\cdot(1,N_0)\dot{}$ is open in $G\times ^P{\kern-1.5pt\goth n}$. Since the stabilizer in $G$ of $(1,N_0)\dot{}$ is $Z_{N_0}$, there is a unique $G$\kern-1.5pt - equivariant isomorphism ${\cal O}\mathrel{\mathop{\kern 0pt\longrightarrow }\limits^{\sim}} G\cdot(1,N_0)\dot{}$ mapping $N_0$ onto $(1,N_0)\dot{}$. We will identify ${\cal O}$ to the open orbit of $G\times ^P{\kern-1.5pt\goth n}$ through this isomorphism. \th Lemma \enonce The Kostant-Kirillov symplectic $2$\kern-1.5pt - form on ${\cal O}$ extends to a $G$\kern-1.5pt - invariant $2$\kern-1.5pt - form $\omega$ on $G\times ^P{\kern-1.5pt\goth n}$. Let $(g,N)\dot{}\inG\times ^P{\kern-1.5pt\goth n}$; the kernel of $\omega_{(g,N)\dot{}} $ consists of the images of the elements $(X,[N,X])$, with $X\in{\goth n}^\perp=\sdir_{i\ge -1}^{}{\goth g}(i)$ and $[N,X]\in{\goth n}$. \endth\label{noyau} {\it Proof}: Consider the alternate bilinear form on ${\goth g}\times {\goth n}$ defined by $$((X,Q),(X',Q')) \mapsto (N\,|\,[X,X'])+(X\,|\,Q')-(X'\,|\,Q)\ .$$ Its kernel consists of pairs $(X,Q)$ with $X\in {\goth n}^\perp$ and $Q=[N,X]$; in particular, it contains the elements $({\it P},[N,{\it P}])$ for ${\it P}\in {\goth p}$, so that our form factors through $T_{(g,N)\dot{}}(G\times ^P{\kern-1.5pt\goth n})$ and defines a $G$\kern-1.5pt - invariant 2-form $\omega $ on $G\times ^P{\kern-1.5pt\goth n}$. \par\hskip 1truecm\relax The isomorphism ${\cal O}\rightarrow G\cdot(1,N_0)\dot{}$ induces on the tangent spaces the isomorphism ${\goth g}/{\goth z}^{}_{N_0}\rightarrow T_{(1,N_0)\dot{}}(G\times ^P{\kern-1.5pt\goth n})$ which maps the class of $X\in{\goth g}$ to the class of $(X,0)$. Through this isomorphism, $\omega_{(1,N_0)\dot{}}$ corresponds to the alternate form $(X,X')\mapsto (N_0\,|\,[X,X'])$, that is to the Kostant-Kirillov $2$\kern-1.5pt - form $\omega _0$ at $N_0$. Since $\omega$ and $\omega_0$ are $G$\kern-1.5pt - invariant, the restriction of $\omega$ to ${\cal O}$ is equal to $\omega_0$.\cqfd \smallskip \par\hskip 1truecm\relax The following lemma will be the key technical ingredient for our proof of the main theorem. We put ${\goth g}^{\scriptscriptstyle\times }={\goth g}\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}\{0\}$, ${\goth n}^{\scriptscriptstyle\times }={\goth n}\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}\{0\}$. \th Lemma \enonce Let $N\in{\goth n}$. Let $\overline{\cal O}$ be the closure of ${\cal O}$ in ${\goth g}^{\scriptscriptstyle\times }$. Assume that the normalization $\widetilde{{\cal O}}$ of $\overline{\cal O}$ is smooth above $N$. Then the centralizer ${\goth z}^{}_N$ is contained in ${\goth n}^\perp$. \endth\label{key} {\it Proof}: Consider the morphism $G\times ^P{\kern-1.5pt\goth n}^{\scriptscriptstyle\times } \rightarrow {\goth g}^{\scriptscriptstyle\times }$ which maps $(g,N)\dot{}$ to $\mathop{\rm Ad}\nolimits(g)N$. Its image is the closure $\overline{\cal O}$ of ${\cal O}$ in ${\goth g}^{\scriptscriptstyle\times }$; since $G\times ^P{\kern-1.5pt\goth n}$ is smooth, it factors through $\widetilde{\cal O}$. The induced morphism $\pi :G\times ^P{\kern-1.5pt\goth n}^{\scriptscriptstyle\times } \rightarrow \widetilde{\cal O}$ is proper and birational: it induces the identity on the open orbit ${\cal O}\iG\times ^P{\kern-1.5pt\goth n}^{\scriptscriptstyle\times }$. \par\hskip 1truecm\relax Since the complement of ${\cal O}$ in $\widetilde{{\cal O}}$ has codimension $\ge 2$, the symplectic 2-form on ${\cal O}$ extends to a 2-form $\varpi$ on the smooth part $\widetilde{{\cal O}}_{sm}$ of $\widetilde{{\cal O}}$; the pull-back of $\varpi$ to $\pi ^{-1}(\widetilde{{\cal O}}_{sm})\iG\times ^P{\kern-1.5pt\goth n} $ coincides with the restriction of $\omega$. It follows that every tangent vector at the point $x=(1,N)\dot{}$ of $G\times ^P{\kern-1.5pt\goth n}^{\scriptscriptstyle\times }$ killed by $T_x(\pi)$ belongs to the kernel of $\omega _x $. Since the orbit of $x$ under $Z^{\rm o}_N$ maps to a point in $\widetilde{{\cal O}}$, the vectors $({\it Z},0)$ with ${\it Z}\in{\goth z}^{}_N$ must belong to the kernel of $\omega_x $; in view of Lemma \ref{noyau}, this means that ${\goth z}^{}_N$ is contained in ${\goth n}^\perp$.\cqfd \vskip1truecm \section{The birational case} \par\hskip 1truecm\relax In this section we will prove Theorem \ref{main} in the simpler case when the map $\varphi^{}_L$ is assumed to be birational. We start with a technical lemma about Lie algebras; we keep the notation of (\ref{dyn}). \th Lemma \enonce Assume that $N_0$ is not contained in a proper ideal of ${\goth g}$, and that for every nonzero elements $N\in{\goth g}(2)$ and $Q\in{\goth g}(-2)$ the bracket $[N,Q]$ is nonzero. Then ${\goth g}$ is simple, and either ${\cal O}$ is the minimal orbit, or ${\goth g}$ is of type $G_2$ and ${\cal O}$ is the orbit of a short root vector. \endth\label{classif} {\it Proof}: Assume first that ${\goth g}$ is a product of two nonzero semi-simple Lie algebras ${\goth g}'$ and ${\goth g}''$. Write $N_0=(N'_0,N''_0)$, $H=(H',H'')$, $N_1=(N'_1,N''_1)$; the hypothesis on $N_0$ ensures that $N'_0$ and $N''_0$ (and therefore also $H',H'',N'_1,N''_1$) are nonzero. We have $N'_1\in {\goth g}(-2)$, $N''_0\in{\goth g}(2)$ and $[N'_1,N''_0]=0$, contrary to the hypothesis. Thus ${\goth g}$ is simple. \par\hskip 1truecm\relax For any nonzero $N\in{\goth g}(2)$, we have ${\goth z}^{}_N\cap{\goth g}(-2)=(0)$; by [C-M], 3.4.17, this implies that $N$ is conjugate to $N_0$. There exists a root $\alpha $ with $\alpha (H)=2$ (the corresponding root vectors span ${\goth g}(2)$); therefore $N_0$ is conjugate to $X_\alpha $. \par\hskip 1truecm\relax Assume that ${\goth g}$ is of type $B_l,C_l$ or $F_4$, and that $\alpha $ is a short root. According to [C-M] the weighted Dynkin diagram of $X_\alpha $ is one of the following: \def\kern-12pt\raise2pt\vbox{\hrule width .94truecm{\kern-12pt\raise2pt\vbox{\hrule width .94truecm} \kern-11.5pt} \def\kern-12pt\Longrightarrow\!\!=\!=\kern-12pt{\kern-12pt\Longrightarrow\!\!=\!=\kern-12pt} \def\kern-12pt =\!=\!\!\Longleftarrow\kern-12pt{\kern-12pt =\!=\!\!\Longleftarrow\kern-12pt} \def\diaram#1{\def\baselineskip=0truept{\baselineskip=0truept \lineskip=4truept\lineskiplimit=1truept} \matrix{#1}} \def\scriptstyle 2{\scriptstyle 2} \def\scriptstyle 1{\scriptstyle 1} \def\scriptstyle 0{\scriptstyle 0} \vskip-10pt$$\diaram{\baselineskip4pt \scriptstyle 2 && \scriptstyle 0 & & \scriptstyle 0 & & \scriptstyle 0 && \scriptstyle 0\cr \circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ & \kern-2pt\cdots\cdots \kern-2pt& \circ & \kern-12pt\Longrightarrow\!\!=\!=\kern-12pt &\circ}$$\vskip-20pt $$\diaram{\baselineskip4pt \scriptstyle 0 && \scriptstyle 1 & & \scriptstyle 0 & & \scriptstyle 0 && \scriptstyle 0\cr \circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ & \kern-2pt\cdots\cdots \kern-2pt& \circ & \kern-12pt =\!=\!\!\Longleftarrow\kern-12pt &\circ}$$\vskip-20pt $$\diaram{\baselineskip4pt \scriptstyle 0 && \scriptstyle 0 & & \scriptstyle 0 & & \scriptstyle 1 \cr \circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\Longrightarrow\!\!=\!=\kern-12pt &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ}$ In each case the highest root $\theta $ satisfies $\theta (H)=2$, hence $X_\theta $ should be conjugate to $X_\alpha $ -- a contradiction. Therefore either $\alpha $ is a long root, or ${\goth g}$ is of type $G_2$.\cqfd \medskip \th Proposition \enonce Let ${\cal O}$ be a nilpotent orbit in ${\goth g}$ and $\overline{\cal O}$ its closure in ${\goth g}^{\scriptscriptstyle\times }$. Assume that ${\cal O}$ is not contained in a proper ideal of ${\goth g}$, and that the normalization of $\overline{\cal O}$ is smooth. Then ${\goth g}$ is simple, and either ${\cal O}$ is the minimal nilpotent orbit, or ${\goth g}$ is of type $G_2$ and ${\cal O}$ is the orbit of a short root vector. \endth\label{closmooth} \par\hskip 1truecm\relax In the first case $\overline{{\cal O}}$ is equal to ${\cal O}$, hence smooth. In the second case $\overline{\cal O}$ is not normal, and its normalization is isomorphic to the minimal nilpotent orbit in ${\goth s}{\goth o}(7)$ [L-Sm]. \smallskip {\it Proof}: By Lemma \ref{key}, we have ${\goth z}^{}_N\i{\goth n}^\perp$ for each nonzero element $N$ of ${\goth n}$. Taking $N$ in ${\goth g}(2)$, we see that the hypotheses of Lemma \ref{classif} are satisfied, hence the result.\cqfd \smallskip \th Corollary \enonce Let $M$ be a Fano contact manifold, such that \par\hskip 1truecm\relax {\rm (i)} the rational map $\varphi:M\dasharrow {\bf P}({\goth g})$ is generically injective; \par\hskip 1truecm\relax {\rm (ii)} the group $G$ of contact automorphisms of $M$ is reductive.\par Then $\varphi $ induces an isomorphism of $M$ onto the minimal nilpotent orbit in ${\bf P}({\goth g})$. \endth {\it Proof}: Consider the commutative diagram (\ref{diagramme}) $$\diagram{ L^{\scriptscriptstyle\times } & \hfl{\mu }{} & {\goth g}^{\scriptscriptstyle\times }&\cr \vfl{}{} & & \vfl{}{}& \cr M & \hfl{\varphi }{} & {\bf P}({\goth g})&\kern-10pt. }$$ By Prop.\ \ref{summary} $\varphi $ is a finite birational morphism onto the closure of a nilpotent orbit ${\bf P}{\cal O}$ in ${\bf P}({\goth g})$; since the diagram is cartesian, $\mu $ is finite and birational onto $\overline{\cal O}$, hence realizes $L^{\scriptscriptstyle\times }$ as the normalization of $\overline{{\cal O}}$. Since the image $\overline{{\bf P}{\cal O}}$ of $\varphi $ spans ${\bf P}({\goth g})$, ${\cal O}$ cannot be contained in any proper subspace of ${\goth g}$. By Prop.\ \ref{closmooth}, this implies either that ${\cal O}$ is a minimal orbit, or that ${\goth g}$ is of type $G_2$ and ${\cal O}$ is the orbit of a short root vector; in that case $M$ is isomorphic to ${\bf P}{\cal O}'$, where ${\cal O}'$ is the minimal orbit in ${\goth s}{\goth o}(7)$, and this isomorphism preserves the contact structures (remark \ref{unique}). But then ${\goth g}$ contains ${\goth s}{\goth o}(7)$, a contradiction.\cqfd \vskip1truecm \section{The general case} \global\def\currenvir{subsection As explained in Remark \ref{converse}, we want to classify finite $G$\kern-1.5pt - equivariant surjective morphisms $\varphi:M\rightarrow \overline{{\bf P}{\cal O}}$, where $M$ is smooth and ${\cal O}\i{\goth g}$ is a nilpotent orbit; such a morphism will be called for short a $G$\kern-1.5pt - {\it covering} of $\overline{{\bf P}{\cal O}}$. Examples of $G$\kern-1.5pt - coverings appear in the classification of ``shared orbit pairs" [B-K], associated to certain pairs ${\goth g}\i {\goth g}'$ of simple Lie algebras: the manifold $M$ is the minimal orbit ${\bf P}{\cal O}'_{min}$ for ${\goth g}'$, while the orbit ${\cal O}\i{\goth g}$ is given in the list below. Brylinski and Kostant find the following cases: \def\vrule height 12pt depth 5pt width 0pt{\vrule height 12pt depth 5pt width 0pt} \def\vrule height 15pt depth 15pt width 0pt{\vrule height 15pt depth 15pt width 0pt} \def\vrule height 12pt depth 5pt width 0pt\vrule{\vrule height 12pt depth 5pt width 0pt\vrule} \def\noalign{\hrule}{\noalign{\hrule}} \def\hfill}\def\hq{\hfill\quad{\hfill}\def\hq{\hfill\quad} $$\vcenter{\offinterlineskip \halign{\vrule height 12pt depth 5pt width 0pt\vrule\hq#\hq& \vrule height 12pt depth 5pt width 0pt\vrule\hq#\hq& \vrule height 12pt depth 5pt width 0pt\vrule\hq#\hq& \vrule height 12pt depth 5pt width 0pt\vrule\hq#\hq\vrule height 12pt depth 5pt width 0pt\vrule\cr\noalign{\hrule} \vrule height 15pt depth 15pt width 0pt ${\goth g}$ & ${\goth g}'$ & ${\cal O}$ & $\deg\varphi$ \cr\noalign{\hrule} $A_2$ & $G_2$ & ${\cal O}_{(3)}$ & 3 \cr\noalign{\hrule} $B_l$ & $D_{l+1}$ & ${\cal O}_{(3,1,\ldots,1)}$ & 2 \cr\noalign{\hrule} $B_4$ & $F_4$ & ${\cal O}_{(2,2,2,2,1)}$ & 2 \cr\noalign{\hrule} $C_l$ & $A_{2l-1}$ & ${\cal O}_{(2,2,1,\ldots,1)}$ & 2 \cr\noalign{\hrule} $D_l$ & $B_{l}$ & ${\cal O}_{(3,1,\ldots,1)}$ & 2 \cr\noalign{\hrule} $D_4$ & $F_4$ & ${\cal O}_{(3,2,2,1)}$ & 4 \cr\noalign{\hrule} $F_4$ & $E_6$ & ${\cal O}_{short}$ & 2 \cr\noalign{\hrule} $G_2$ & $B_3$ & ${\cal O}_{short}$ & 1 \cr\noalign{\hrule} $G_2$ & $D_4$ & ${\cal O}_{sub}$ & 6 \cr\noalign{\hrule} }}\leqno{\global\def\currenvir{subsection}$$ \label{list} \par\hskip 1truecm\relax The notation for the orbit ${\cal O}$ requires some explanation: in the classical cases, ${\goth g}$ is viewed as an algebra of matrices via the standard representation; then ${\cal O}_{(d_1,\ldots,d_k)}$ denote the conjugacy class of a matrix in ${\goth g}$ with Jordan type $(d_1,\ldots,d_k)$. As in (\ref{simple}), ${\cal O}_{short}$ is the orbit of a short root vector. Finally ${\cal O}_{sub}$ is the so-called subregular orbit, that is the unique codimension 2 orbit in the nilpotent cone. \th Proposition \enonce Let $G$ be a simple complex Lie group acting on a manifold $M$, ${\goth g}$ the Lie algebra of $G$, ${\cal O}\i{\goth g}$ a nilpotent orbit, $\varphi:M\rightarrow \overline{{\bf P}{\cal O}}$ a finite $G$\kern-1.5pt - equivariant surjective morphism. Then either ${\cal O}={\cal O}_{min}$ and $\varphi$ is an isomorphism, or $\varphi $ is {\rm (}up to isomorphism{\rm )} one of the $G$\kern-1.5pt - coverings appearing in the list $(\ref{list})$. \endth\label{coverings} {\it Proof}:\global\def\currenvir{subsection\label{Galois} Let $M^{\rm o}$ be the open $G$\kern-1.5pt - orbit in $M$; let $m$ be a point of $M^{\rm o}$, $\H^{\rm o}$ its stabilizer in $G$ and $\H$ the stabilizer of $\varphi(m)$. Since $M$ is Fano, $M$ and therefore $M^{\rm o}$ are simply connected; this implies that $\H^{\rm o}$ is the neutral component of $\H$. So the covering $M^{\rm o}\rightarrow {\bf P}{\cal O}$ is a Galois covering, with Galois group $\Gamma := \H/\H^{\rm o}$. Since $M={\rm Proj}\sdir_{n\ge 0}^{}\H^0(M^{\rm o},L^n)$, the action of $\Gamma $ on $M^{\rm o}$ extends to an action on $M$, which commutes with the $G$\kern-1.5pt - action. \par\hskip 1truecm\relax Observe that the $G$\kern-1.5pt - covering $M\rightarrow \overline{{\bf P}{\cal O}}$ is uniquely determined by ${\cal O}$: the open $G$\kern-1.5pt - orbit $M^{\rm o}\i M$ is the simply-connected covering of ${\bf P}{\cal O}$, and $M$ is the integral closure of $\overline{{\bf P}{\cal O}}$ in $M^{\rm o}$. Thus our task is to prove that only the orbits listed in (\ref{list}) can occur. \global\def\currenvir{subsection\label{arg} We will prove this by induction on the dimension of ${\cal O}$, the case ${\cal O}={\cal O}_{min}$ being clear in view of (\ref{Galois}). By Prop.\ \ref{closmooth} we can assume $\deg(\varphi)>1$. Let $\gamma\in\Gamma $, and let $F$ be a component of the fixed locus of $\gamma$. Then $F$ is a closed submanifold of $M$, stable under $G$; the map $\varphi$ induces a $G$\kern-1.5pt - covering $F\rightarrow \overline{{\bf P}{\cal O}}_F$ for some orbit ${\cal O}_F\i\overline{\cal O}$. By the induction hypothesis, $F$ is isomorphic to the minimal orbit ${\bf P}{\cal O}'_{min}$ for some simple Lie algebra ${\goth g}'$ containing ${\goth g}$; either ${\goth g}'={\goth g}$, or the pair $({\goth g},{\goth g}')$ is one of the pairs appearing in the list (\ref{list}). \par\hskip 1truecm\relax Let us say for short that an orbit ${\cal O}'\i\overline{\cal O}$ is {\it ramified} if $\varphi^{-1} ({\bf P}{\cal O}')$ is contained in the fixed locus of some nontrivial element of $\Gamma $. Let ${\cal O}'\i\overline{\cal O}$ an orbit which is not ramified; since $\varphi$ induces an isomorphism of $M/\Gamma $ onto the normalization $\widetilde{{\bf P}{\cal O}}$ of $\overline{{\bf P}{\cal O}}$, we have: \par\hskip 0.5truecm\relax (\ref{arg}.{\it a}) $\widetilde{{\bf P}{\cal O}}$ is smooth along ${\bf P}{\cal O}'$; in particular, the centralizer of any element of ${\cal O}'\cap{\goth n}$ is contained in ${\goth n}^\perp$ (lemma \ref{key}). \par\hskip 0.5truecm\relax (\ref{arg}.{\it b}) Any nonzero element $N\in {\cal O}'\cap{\goth g}(2)$ satisfies ${\goth z}^{}_N \cap{\goth g}(-2)=(0)$, hence is conjugate to $N_0$ by [C-M], 3.4.17; therefore if ${\cal O}'\cap{\goth g}(2)\not=(0)$, then ${\cal O}'={\cal O}$. \par\hskip 0.5truecm\relax (\ref{arg}.{\it c}) Assume that $\overline{\cal O}$ is normal along ${\cal O}'$. Then $\varphi$ is \'etale above ${\bf P}{\cal O}'$, so that $T_m(\varphi)$ is injective at each point $m$ of $\varphi^{-1} ({\bf P}{\cal O}')$. But this implies that $m$ belongs to the open orbit $M^{\rm o}$ (Prop.\ \ref{openorbit}), hence ${\cal O}'={\cal O}$ again. \par\hskip 0.5truecm\relax (\ref{arg}.{\it d}) Assume that the Galois group $\Gamma $ is cyclic of prime order, and that $\overline{\cal O}$ is normal. Let $M^{\Gamma}$ be the fixed locus of $\Gamma $ in $M$. Then $\varphi $ induces an isomorphism of $M^{\Gamma }$ onto its image; in particular, $\varphi (M^\Gamma )$ is smooth. By Prop.\ \ref{closmooth}, this implies that the only ramified orbit is ${\cal O}_{min}$, so by (\ref{arg}.{\it c}) we have $\overline{\cal O}={\cal O}\cup {\cal O}_{min}$. \smallskip \global\def\currenvir{subsection Now we examine which orbits ${\cal O}\i{\goth g}$ may occur. We order the nilpotent orbits by the relation ``${\cal O}'\le {\cal O}$ iff ${\cal O}'\i\overline{{\cal O}}$". Given the Lie algebra ${\goth g}$, the possible ramified orbits are those contained in the closure of the orbit ${\cal O}$ in (\ref{list}). Using the above arguments we will show that only one more orbit is allowed: its boundary must contain only ramified orbits. This gives us for each Lie algebra ${\goth g}$ a small list of orbits, among which we may eliminate those which are simply connected; we will show that the remaining ones are those which appear in the list (\ref{list}). \medskip {\it Type $A_l\ (l\ge 4)$}\vglue0pt \par\hskip 1truecm\relax All orbit closures in case $A_l$ are normal [K-P1], so by (\ref{arg}.{\it c}) there is only one orbit which is not ramified. There is no shared orbit pair, so the only ramified orbit is the minimal one. The next orbit in the partial ordering is ${\cal O}_{(2,2,1,\ldots)}$, which is simply-connected [C-M, p.\ 92].\smallskip {\it Type $A_3$} \par\hskip 1truecm\relax The possible ramified orbits are ${\cal O}_{min}$ and ${\cal O}_{(2,2)}$; the next orbit in the partial ordering is ${\cal O}_{(3,1)}$. The orbit ${\cal O}_{(2,2)}$ gives rise to case $(D_3,B_3)$ of (\ref{list}); ${\cal O}_{(3,1)}$ is simply-connected. \smallskip {\it Type $A_2$} \par\hskip 1truecm\relax There are only two orbits, ${\cal O}_{min}$ and the principal orbit ${\cal O}_{(3)}$, which gives rise to case $(A_2,G_2)$ of (\ref{list}). \medskip \par\hskip 1truecm\relax For the types $B_l,C_l \hbox{ or }D_l$, most orbit closures are normal, with the following exceptions [K-P2]: \par\hskip 0.5truecm\relax a) There may exist an orbit ${\cal O}$ whose closure is non-normal along a codimension 2 orbit ${\cal O}'$, but whose normalization is singular along ${\cal O}'$. In this case by (\ref{arg}.{\it a}) ${\cal O}'$ is ramified; \par\hskip 0.5truecm\relax b) When ${\goth g}$ is of type $D_l$, there are orbits (corresponding to the so-called ``very even" classes) whose closure is not known to be normal. However these orbit closures have a boundary component of codimension 2 along which they are normal, so that (\ref{arg}.{\it c}) still applies.\smallskip {\it Type $B_l$ and $D_l$, $l\ge 5$} \par\hskip 1truecm\relax The Lie algebra ${\goth g}$ is ${\goth s}{\goth o}(n)$ $(n\ge 10)$. The possible ramified orbits are ${\cal O}_{min}$ and ${\cal O}_{(3,1,\ldots)}$; the only possible next orbit is ${\cal O}_{(2,2,2,2,1,\ldots)}$ (${\cal O}_{(3,2,2,1,\ldots)}$ is excluded because its closure contains ${\cal O}_{(2,2,2,2,1,\ldots)}$ which is not ramified). The orbit ${\cal O}_{(3,1,\ldots)}$ gives rise to cases $(B_l,D_{l+1} )$ and $(D_l,B_l)$; ${\cal O}_{(2,2,2,2,1,\ldots)}$ is simply-connected (\hbox{[C-M]}, p.\ 92). \smallskip {\it Type $B_4$} \par\hskip 1truecm\relax The configuration of orbits is the same as above, but here the orbit ${\cal O}_{(2,2,2,2,1)}$ can be ramified. Therefore the next orbit ${\cal O}_{(3,2,2,1,1)}$ might occur. However its fundamental group is ${\bf Z}/2$, and its closure is normal \hbox{[K-P2]}, so we deduce from (\ref{arg}.{\it d}) that this orbit does not occur. \par\hskip 1truecm\relax The orbit ${\cal O}_{(2,2,2,2,1)}$ is no longer simply-connected; it gives rise to case $(B_4,F_4)$ in (\ref{list}). \smallskip {\it Type $B_3$}\vglue0pt \par\hskip 1truecm\relax Again the orbit ${\cal O}_{(3,2,2)}$ can occur a priori; the same argument as for $B_4$ applies. \smallskip {\it Type $D_4$} \par\hskip 1truecm\relax The possible ramified orbits are ${\cal O}_{min}$, the three orbits next to ${\cal O}_{min}$ in the partial ordering (namely ${\cal O}_{(3,1,\ldots)}$ and the two orbits ${\cal O}_{(2,2,2,2)}$), and ${\cal O}_{(3,2,2,1)}$; the next orbit is ${\cal O}_{(3,3,1,1)}$. \par\hskip 1truecm\relax The three orbits next to ${\cal O}_{min}$ have the same weighted Dynkin diagram up to automorphisms, and are therefore isomorphic; they give the case $(D_4,B_4)$. The orbit ${\cal O}_{(3,2,2,1)}$ gives the case $(D_4,F_4)$. Finally ${\cal O}_{(3,3,1,1)}$ has fundamental group ${\bf Z}/2$ and normal closure [K-P2], so is excluded by (\ref{arg}.{\it d}). \medskip {\it Type} $C_l$ $(l\ge 2)$ \par\hskip 1truecm\relax The possible ramified orbits are ${\cal O}_{min}$ and ${\cal O}_{(2,2,1,\ldots)}$; the next orbit is ${\cal O}_{(2,2,2,1,\ldots)}$ if $l\ge 3$, and ${\cal O}_{(4)}$ if $l=2$. This orbit has fundamental group ${\bf Z}/2$ and is normal [K-P2], so it is excluded again by (\ref{arg}.{\it d}). The orbit ${\cal O}_{(2,2,1,\ldots)}$ gives the case $(C_l,A_{2l-1})$. \medskip {\it Type} $E_l$ \par\hskip 1truecm\relax The only possible ramified orbit is the minimal one. If ${\cal O}$ is not reduced to ${\cal O}_{min}$ it contains the next orbit ${\cal O}_1$ in the partial ordering, which is the orbit of $X_\lambda +X_\mu $, where $\lambda $ and $\mu $ are two orthogonal roots. By (\ref{arg}.{\it a}) the centralizer of an element of ${\cal O}_1\cap{\goth n}$ is contained in ${\goth n}^\perp$. \par\hskip 1truecm\relax Let $\sigma $ be the sum of the simple roots, and $\alpha ,\beta ,\gamma $ the simple roots corresponding to the three ends of the Dynkin graph. Then $\sigma ,\,\sigma -\alpha ,\,\sigma -\beta ,\,\sigma -\gamma $ are roots ([B], \S 1, n\up{o}\kern 2pt 6, cor.\ 3 of prop.\ 19) and $\sigma -\alpha $ and $\sigma -\beta $ are orthogonal; the element $N=X_{\sigma -\alpha }+X_{\sigma -\beta }$ satisfies $[N\,,\,X_{\gamma -\sigma }]=0$. Let $s=\sigma (H)$ and $m=\max \{\alpha (H),\beta (H),\gamma (H)\}$. If $s-m\ge 2$ we have $N\in {\goth n}$ and $X_{\gamma -\sigma }\notin {\goth n}^\perp$, a contradiction. \par\hskip 1truecm\relax Suppose $s=2$ and $\alpha (H)=\beta (H)=0$. Then $N$ belongs to ${\goth g}(2)$, which by (\ref{arg}.{\it b}) implies ${\cal O}={\cal O}_1$; this is excluded because ${\cal O}_1$ is simply-connected ([C-M], pp.\ 129, 130, 132). \par\hskip 1truecm\relax Looking at the list of possible weighted Dynkin diagrams in {\it loc.\ cit.}\ and eli\-mi\-nating the simply-connected orbits, the above constraints leave us with only one possible case, the weighted Dynkin diagram \def\scriptstyle 2{\scriptstyle 2} \def\scriptstyle 1{\scriptstyle 1} \def\scriptstyle 0{\scriptstyle 0} \def\vrule height 3pt depth 0pt width 0pt{\vrule height 3pt depth 0pt width 0pt} $$\dia{ \scriptstyle 1 && \scriptstyle 0 & & \scriptstyle 0 & & \scriptstyle 0 && \scriptstyle 0 && \scriptstyle 0 && \scriptstyle 1 \cr \vrule height 3pt depth 0pt width 0pt \cr \circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ \cr && && \vrule height10pt depth4pt width 0.4pt &&&&&&&&\cr && && \circ &&&&&&&&\cr\vrule height 3pt depth 0pt width 0pt\cr && && \scriptstyle 0 &&&&&&&&\cr}$$ for $E_8$. In that case one finds easily two orthogonal roots $\lambda $ and $\mu $ with $\lambda (H)=\mu (H)=2$, for instance (with the notation of [B], planche VII) $\lambda ={1\over 2}\sum_i\varepsilon _i$ and $\mu =\varepsilon_8-\varepsilon _7$; we conclude again by (\ref{arg}.{\it b}) that ${\cal O}={\cal O}_1$. \medskip {\it Type $F_4$}\vglue0pt \par\hskip 1truecm\relax The orbits which can be ramified are ${\cal O}_{min}$ and ${\cal O}_{short}$. If ${\cal O}$ is bigger than ${\cal O}_{short}$, it contains the orbit ${\cal O}_1$ next to ${\cal O}_{short}$; this is the orbit of $X_\alpha +X_\beta $, where $\alpha $ and $\beta $ are two orthogonal roots of distinct lengths. Let $$\diaram{ {\scriptstyle l_1} && {\scriptstyle l_2} & & {\scriptstyle l_3} & & {\scriptstyle l_4} \cr \circ & \kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ &\kern-12pt\Longrightarrow\!\!=\!=\kern-12pt &\circ &\kern-12pt\raise2pt\vbox{\hrule width .94truecm &\circ}$$ be the weighted Dynkin diagram of ${\cal O}$. Assume first $l_1+l_2+l_3\ge 2$. Using the notation of [B], planche VIII, let $$\nospacedmath\displaylines{ \alpha =\varepsilon _2=\alpha _1+\alpha _2+\alpha _3\ ,\quad \beta =\varepsilon _1-\varepsilon _4=\alpha _1+2\alpha _2+2\alpha _3+2\alpha _4\ ,\cr \gamma =\varepsilon _1+\varepsilon _4=\alpha _1+2\alpha _2+4\alpha _3+2\alpha _4\ .}$$ We have $[X_\alpha +X_\beta \,,\,X_{-\gamma }]=0$, $X_\alpha +X_\beta\in{\goth n}$ and $X_{-\gamma }\notin {\goth n}^\perp$, contradicting (\ref{arg}.{\it a}). \par\hskip 1truecm\relax A glance at the tables ([C-M], p.\ 128) shows that the nilpotent orbits with $l_1+l_2+l_3\le 1$ are simply-connected, with the exception of ${\cal O}_{short}$; the latter gives the case $(F_4,E_6)$. \medskip {\it Type} $G_2$ \par\hskip 1truecm\relax The only orbit which is not simply-connected is the subregular orbit ([C-M], p.~128), which gives rise to case $(G_2,D_4)$.\cqfd \medskip \rem{Example}\label{sp} Let us give an example of a $G$\kern-1.5pt - covering when ${\goth g}$ is not simple. Let ${\bf n}=(n_1,\ldots,n_k)$ be a sequence of positive integers; for each $i$, let ${\goth g}_i$ be the Lie algebra ${\goth s}{\goth p}(2n_i)$, and $V_i\ (\cong {\bf C}^{2n_i})$ its standard representation. Then ${\goth g}_i$ can be identified with $\hbox{\san \char83}^2V_i$; the minimal nilpotent orbit ${\cal O}_i\i{\goth g}_i$ is then identified with the cone of rank one tensors, so that we have a 2-to-1 map $\mu _i:V_i\rightarrow \overline{\cal O}_i={\cal O}_i\cup\{0\}$ mapping a vector $v$ to $v^2$. We put ${\goth g}=\pprod_{i}^{}{\goth g}_i$, ${\cal O}=\pprod_i^{}{\cal O}_i$, $M={\bf P}(V)$ with $V=\ \sdir_i^{}V_i$. The maps $\mu _i$ define a $G$\kern-1.5pt - covering $\varphi_{\bf n} :{\bf P}(V)\rightarrow \overline{{\bf P}{\cal O}}$, of degree $2^{k-1}$. Note that $M$ is a minimal orbit in ${\bf P}({\goth g}')$, with ${\goth g}'={\goth s}{\goth p}(V)$. \th Proposition \enonce Assume that ${\goth g}$ is a product of simple Lie algebras ${\goth g}_1,\ldots,{\goth g}_k$ $(k>1)$. Let $\varphi:M\rightarrow \overline{{\bf P}{\cal O}}$ be a $G$\kern-1.5pt - covering. Then there exists a sequence ${\bf n}=(n_1,\ldots,n_k)$ of positive integers such that $\varphi $ is isomorphic to the $G$\kern-1.5pt - covering $\varphi _{\bf n}$ of example $\ref{sp}$. In particular, ${\goth g}_i$ is isomorphic to ${\goth s}{\goth p}(2n_i)$ for each $i$, the orbit ${\cal O}$ is the product of the minimal orbits ${\cal O}_i\i{\goth g}_i$, and $M$ is isomorphic to ${\bf P}^{2n-1}$ with $n=\sum n_i$. \endth\label{nonsimple} {\it Proof}: The orbit ${\cal O}$ is a product of nontrivial orbits ${\cal O}_i\i{\goth g}_i$. Let ${\cal O}_i^{sc}$ be the simply-connected covering of ${\cal O}_i$, and $\overline{{\cal O}_i^{sc}}$ the integral closure of $\overline{\cal O}_i$ in ${\cal O}_i^{sc}$ (contrary to an earlier notation, we denote by $\overline{\cal O}_i$ the closure of ${\cal O}_i$ {\it in} ${\goth g}$). The action of $G\times {\bf C}^*$ on ${\cal O}_i$ extends to an action on ${\cal O}_i^{sc}$ and $\overline{{\cal O}_i^{sc}}$. There is only one point $o^{}_i$ of $\overline{{\cal O}_i^{sc}}$ above $0\in{\goth g}$; the open subset $\overline{{\cal O}_i^{sc}}\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}\{o_i\}$ is a principal ${\bf C}^*$\kern-1.5pt - bundle over a variety $M_i$ which admits a finite $G$\kern-1.5pt - equivariant morphism onto $\overline{{\bf P}{\cal O}}_i$. \par\hskip 1truecm\relax Let $M'=(\pprod_i^{} \overline{{\cal O}_i^{sc}})^{\scriptscriptstyle\times }/{\bf C}^*$, where the superscript $^{\scriptscriptstyle\times }$ means that we take out the point $(o^{}_1,\ldots,o^{}_k)$. This is a normal variety, with a finite morphism onto $\overline{{\bf P}{\cal O}}$; the open subset $(\pprod_i^{}{\cal O}_i^{sc})/{\bf C}^*$ is simply-connected and its complement has codimension $\ge 2$. This implies that $M'$ is isomorphic to $M$. \par\hskip 1truecm\relax Since $M$ is smooth, it follows that each $\overline{{\cal O}_i^{sc}}$ must be smooth. This implies first of all that $ {\cal O}_i^{sc}$ is smooth, hence by Prop.\ \ref{closmooth} and \ref{coverings} isomorphic to the minimal orbit ${\bf P}{\cal O}'_i$ for some simple Lie algebra ${\goth g}'_i$ containing ${\goth g}_i$. Then ${\cal O}_i^{sc}$ is the simply-connected cover of ${\cal O}'_i$, and $\overline{{\cal O}_i^{sc}}$ is its integral closure in $\overline{\cal O}_i$. Since $\overline{{\cal O}_i^{sc}}$ is smooth, this happens if and only if ${\goth g}_i={\goth g}'_i\cong {\goth s}{\goth p}(2n_i)$ for some integer $n_i\ge 1$ ([B-K], thm.\ 4.6); then ${\cal O}_i={\cal O}'_i$ by Prop.\ \ref{coverings}, so we are in the situation of example \ref{sp}.\cqfd\medskip \par\hskip 1truecm\relax The above results imply directly Theorem \ref{main}, in a slightly more precise form: \th Theorem \enonce Let $M$ be a Fano contact manifold, satifying the conditions $(\H1)$ and $(\H2)$ of Theorem $\ref{main}$. Then the Lie algebra ${\goth g}$ of infinitesimal contact transformations of $M$ is simple, and the canonical map $\varphi:M\rightarrow {\bf P}({\goth g})$ induces an isomorphism of $M$ onto the minimal orbit ${\bf P}{\cal O}_{min}\i {\bf P}({\goth g})$. \endth {\it Proof}: By (\ref{H3}), we can assume that $M$ satisfies also (H3); then $\varphi$ induces a $G$\kern-1.5pt - covering $M\rightarrow \overline{{\bf P}{\cal O}}$ onto the closure of some nilpotent orbit in ${\bf P}({\goth g})$ (Prop.\ \ref{summary}). By Prop.\ \ref{coverings} and \ref{nonsimple}, $M$ is isomorphic to the minimal orbit in ${\bf P}({\goth g}')$ for some simple Lie algebra ${\goth g}'$ containing ${\goth g}$; moreover if $\varphi$ is not an embedding, ${\goth g}'$ contains strictly ${\goth g}$, which is impossible since ${\goth g}'$ is an algebra of infinitesimal contact transformations of $M$ (see remark \ref{unique}). Therefore $\varphi$ is an embedding and ${\goth g}'={\goth g}$.\cqfd \vskip2cm \centerline{ REFERENCES} \vglue15pt\baselineskip13.4pt \def\num#1{\smallskip \item{\hbox to\parindent{\enskip [#1]\hfill}}} \parindent=1.3cm \num{A} V.\ {\pc ARNOLD}: {\sl Mathematical methods of classical mechanics}. Graduate Texts in Math.\ {\bf 60}; Springer-Verlag, New York-Heidelberg (1978). \num{B} N.\ {\pc BOURBAKI}: {\sl Groupes et alg\`ebres de Lie}, Chap.\ VI. Hermann, Paris (1968). \num{B-K} R.\ {\pc BRYLINSKI}, B.\ {\pc KOSTANT}: {\sl Nilpotent orbits, normality and Hamiltonian group actions}. J.\ of the A.M.S.\ {\bf 7}, 269-298 (1994). \num{C-M} D.\ {\pc COLLINGWOOD}, W.\ {\pc MC}{\pc GOVERN}: {\sl Nilpotent orbits in semi-simple Lie algebras}. Van Nostrand Reinhold Co., New York (1993). \num{K-P1} H.\ {\pc KRAFT}, C.\ {\pc PROCESI}: {\sl Closures of conjugacy classes of matrices are normal}. Invent.\ math.\ {\bf 53}, 227-247 (1979). \num{K-P2} H.\ {\pc KRAFT}, C.\ {\pc PROCESI}: {\sl On the geometry of the conjugacy classes in classical groups}. Comment.\ Math.\ Helvetici {\bf 57}, 539-602 (1982). \num{L} C.\ {\pc LE}{\pc BRUN}: {\sl Fano manifolds, contact structures, and quaternionic geometry}. Int.\ J.\ of Math. {\bf 6}, 419-437 (1995). \num{L-S} C.\ {\pc LE}{\pc BRUN}, S.\ {\pc SALAMON}: {\sl Strong rigidity of quaternion-K\"ahler manifolds}. Invent.\ math.\ {\bf 118}, 109-132 (1994). \num{L-Sm} T.\ {\pc LEVASSEUR}, S.\ {\pc SMITH}: {\sl Primitive ideals and nilpotent orbits in type $G_2$}. J.\ of Algebra {\bf 114}, 81-105 (1988). \num{M} D.\ {\pc MUMFORD}, J.\ {\pc FOGARTY}: {\sl Geometric invariant theory}. 2\up{nd} edition. Springer-Verlag, New York-Heidelberg (1982). \smallskip \num{Mu} S.\ {\pc MUKAI}: {\sl Biregular classification of Fano $3$\kern-1.5pt - folds and Fano manifolds of coindex $3$}. Proc.\ Nat.\ Acad.\ Sci.\ USA {\bf 86}, 3000-3002 (1989). \vskip1cm \def\pc#1{\eightrm#1\sixrm} \hfill\vtop{\eightrm\hbox to 5cm{\hfill Arnaud {\pc BEAUVILLE}\hfill} \hbox to 5cm{\hfill DMI -- \'Ecole Normale Sup\'erieure\hfill} \hbox to 5cm{\hfill (URA 762 du CNRS)\hfill} \hbox to 5cm{\hfill 45 rue d'Ulm\hfill} \hbox to 5cm{\hfill F-75230 {\pc PARIS} Cedex 05\hfill}} \end
1998-02-23T16:31:47
9707
alg-geom/9707004
en
https://arxiv.org/abs/alg-geom/9707004
[ "alg-geom", "hep-th", "math.AG" ]
alg-geom/9707004
Robert Friedman
Robert Friedman, John W. Morgan, Edward Witten
Principal G-bundles over elliptic curves
AMS-TeX, 20 pages, amsppt style; minor errors corrected
Math.Res.Lett.5:97-118,1998
null
null
null
Let $G$ be a simple and simply connected complex Lie group. We discuss the moduli space of holomorphic semistable principal $G$ bundles over an elliptic curve $E$. In particular we give a new proof of a theorem of Looijenga and Bernshtein-Shvartsman, that the moduli space is a weighted projective space. The method of proof is to study the deformations of certain unstable bundles coming from special maximal parabolic subgroups of $G$. We also discuss the associated automorphism sheaves and universal bundles, as well as the relation between various universal bundles and spectral covers.
[ { "version": "v1", "created": "Mon, 7 Jul 1997 17:43:09 GMT" }, { "version": "v2", "created": "Mon, 23 Feb 1998 15:31:46 GMT" } ]
2010-04-07T00:00:00
[ [ "Friedman", "Robert", "" ], [ "Morgan", "John W.", "" ], [ "Witten", "Edward", "" ] ]
alg-geom
\section{1. Introduction.} Let $E$ be an elliptic curve with origin $p_0$, and let $G$ be a complex simple algebraic group. For simplicity, we shall only consider the case where $G$ is simply connected, although all of the methods discussed below can be extended to the case of a general group $G$. The goal of this note is to announce some results concerning the moduli of principal holomorphic $G$-bundles over $E$. Detailed proofs, as well as a more thorough discussion of the case where $E$ is allowed to be singular or to vary in families and of the connections with del Pezzo surfaces, elliptic $K3$ surfaces, and Calabi-Yau manifolds which are elliptic or $K3$ fibrations, will appear elsewhere. Grothendieck \cite{21} considered principal holomorphic $G$-bundles over $\Pee ^1$, and showed that it was always possible to reduce the structure group to a Cartan subgroup, i\.e\. to a maximal (algebraic) torus in $G$. Atiyah \cite{1} classified all holomorphic vector bundles over an elliptic curve (in other words, the cases $G=SL(n, \Cee)$ or $G =PGL(n, \Cee)$), without however considering the problem of trying to construct a moduli space or find a universal bundle. In \cite{16}, \cite{17}, and \cite{18}, this problem is studied in the rank two case with a view toward constructing relative moduli spaces in families. This approach has been generalized to arbitrary rank in \cite{20}. A great deal of work has been done on the moduli spaces and stacks of $G$-bundles over a curve of genus at least two, partly motivated by the study of conformal blocks and the Verlinde formulas, by very many authors, e\.g\. \cite{5}, \cite{15}. A basic method here is to relate the moduli stack to an appropriate loop group. Related constructions in the case of genus one have been carried out by Baranovsky-Ginsburg \cite{4}, based on unpublished work of Looijenga (see for example \cite{13}). They relate semistable $G$-bundles to conjugacy classes in a corresponding affine Kac-Moody group. Recently Br\"uchert \cite{9} has constructed a Steinberg-type cross-section for the adjoint quotient of the affine Kac-Moody group whose image lies in the set of regular elements, and this construction leads to a moduli space for semistable $G$-bundles which is equivalent to the one we construct in Section 4 below. (We are indebted to Slodowy for calling our attention to the work of Br\"uchert and sketching an argument for the equivalence of the approach described above with the one we give in this paper.) Finally, many of the results in this note, along with applications to physics, are discussed in \cite{19}. The contents of this note are as follows. We will be concerned with the classification of semistable $G$-bundles. As is typical in invariant theory or moduli problems, the classification will be up to a coarser equivalence than isomorphism, which is usually called S-equivalence and will be defined more precisely in Section 2. In Section 2, we describe the moduli space of semistable $G$-bundles over $E$ via flat connections for the maximal compact subgroup $K$ of $G$, or equivalently via conjugacy classes of representations $\rho\: \pi _1(E) \to K$. Such bundles, which for a simply connected group $G$ are exactly the bundles whose structure group reduces to a Cartan subgroup, have an automorphism group which is as large as possible in a certain sense within a fixed S-equivalence class. The main result here is a theorem due to Looijenga and Bernshtein-Shvartsman which describes this moduli space as a weighted projective space. At the end of the section, we connect this description, in the case where $G=E_6, E_7, E_8$, with the moduli space of del Pezzo surfaces of degree $3,2,1$ respectively and with the deformation theory of simple elliptic singularities. In Section 3, we describe regular $G$-bundles, which by contrast with flat bundles have automorphism groups whose dimensions are as small as possible within a fixed S-equivalence class. The generic $G$-bundle is both flat and regular. However at special points of the moduli space we can choose either a unique flat representative or a unique regular representative, and it is the regular representatives which fit together to give holomorphic families. In Section 4, we show how special unstable bundles over certain maximal parabolic subgroups can be used to give another description of the moduli space in terms of regular bundles and obtain a new proof of the theorem of Looijenga and Bernshtein-Shvartsman. Finally, in the last section we discuss the existence of universal bundles and give a brief description of how our construction can be twisted with the help of a certain spectral cover. \section{2. Split semistable bundles.} We fix notation for the rest of this paper. As before, $E$ denotes an elliptic curve with origin $p_0$. Let $G$ be a simple and simply connected complex Lie group of rank $r$, and let $\xi \to E$ be a holomorphic principal $G$-bundle over $E$. The following definition differs from that given in Ramanathan \cite{32}, but is equivalent to it. \definition{Definition 2.1} The principal bundle $\xi \to E$ is {\sl semistable\/} if the associated vector bundle $\ad \xi$ is a semistable vector bundle. The principal bundle $\xi \to E$ is {\sl unstable\/} if it is not semistable. \enddefinition Note that, if $\xi$ is stable in the sense of \cite{32}, it is still possible for the vector bundle $\ad \xi$ to be strictly semistable. However, in our case ($G$ simply connected), there are essentially no properly stable bundles over $E$, and so the above definition will suffice for our purposes. If $\xi$ is an unstable bundle, the structure group of $\xi$ reduces canonically to a parabolic subgroup $P$ of $G$, the {\sl Harder-Narasimhan parabolic\/} associated to $\xi$ (see for example \cite{31} or \cite{2}, pp\. 589--590). The canonical reduction holds over a general base curve. In the case of a base curve $E$ of genus one, it is easy to see that the structure group further reduces to a Levi factor of $P$. Recall the following standard terminology: a {\sl family\/} of principal $G$-bundles over $E$ parametrized by a complex space (or scheme) $S$ is a principal $G$-bundle $\Xi$ over $E\times S$. The family $\Xi$ is a family of {\sl semistable\/} principal $G$-bundles over $E$ if $\Xi|E\times \{s\}=\Xi _s$ is semistable for all $s\in S$. Finally, let $\xi$ and $\xi'$ be two semistable bundles over $E$. We say that $\xi$ and $\xi'$ are {\sl S-equivalent\/} if there exists a family of semistable bundles $\Xi$ parametrized by an irreducible $S$ and a point $s\in S$ such that, for $t\neq s$, $\Xi |E \times \{t\} \cong \xi$ and $\Xi |E \times \{s\} \cong \xi'$. More generally, we let S-equivalence be the equivalence relation generated by the above relation. The following holds only under our assumption that $G$ is simply connected. \proposition{2.2} Let $\xi$ be a semistable principal $G$-bundle, and suppose that the rank of $G$ is $r$. Then $h^0(E; \ad \xi) \geq r$. Equivalently, $\dim \Aut _G\xi \geq r$, where $\dim \Aut _G\xi$ denotes the group of global automorphisms of $\xi$ \rom(as a $G$-bundle\rom). \endstatement \definition{Definition 2.3} Let $\xi$ be a semistable principal $G$-bundle. We call $\xi$ {\sl regular\/} if $h^0(E; \ad \xi) = r$, or equivalently if $\dim \Aut _G(\xi) = r$. We call $\xi$ {\sl split\/} if its structure group reduces to a Cartan subgroup of $G$, i\.e\. a maximal (algebraic) torus. \enddefinition It is easy to check that split bundles have the following closure property: if there exists a family of semistable bundles $\Xi$ parametrized by an irreducible $S$ and a point $s\in S$ such that, for $t\neq s$, the bundles $\Xi |E \times \{t\}$ are split and all isomorphic to each other, then $\Xi |E \times \{s\}$ is isomorphic to $\Xi |E \times \{t\}$, $t\neq s$, and thus it is split as well. In general, however, the condition of being split is neither open nor closed. On the other hand, by the upper semicontinuity theorem, regularity is an open condition: if $\Xi$ is a family of semistable bundles parametrized by $S$ and $\Xi |E \times \{s\}$ is regular, then $\Xi |E \times \{t\}$ is regular for all $t$ in an open neighborhood of $s$. To describe the set of split bundles, we introduce flat bundles on the compact group. Let $K$ be a maximal compact subgroup of $G$. Then $K$ is a compact, simple and simply connected Lie group. If $\frak k$ is the Lie algebra of $K$ and $\frak g$ is the Lie algebra of $G$, then $\frak g$ is the complexification of $\frak k$. Given a representation $\rho\: \pi _1(E) \cong \Zee \oplus \Zee \to K$, we can form the associated principal $K$-bundle $(\tilde E \times K)/\pi _1(E) \to E$, where $\pi _1(E)$ acts on $\tilde E$, the universal cover of $E$, in the usual way, and on $K$ via $\rho$. We shall call such a $K$-bundle a {\sl flat\/} $K$-bundle. Using the inclusion $K\subset G$, we can also view a flat $K$-bundle as a $G$-bundle, and we shall also incorrectly refer to the induced $G$-bundle as a flat $K$-bundle. We will need the following version of the theorem of Narasimhan-Seshadri \cite{29} and Ramanathan \cite{32} (see also Atiyah-Bott \cite{2} and Donaldson \cite{12}): \theorem{2.4} Let $\xi\to E$ be a semistable principal $G$-bundle. Then there is a flat $K$-bundle $S$-equivalent to $\xi$, and it is unique up to isomorphism of flat $K$-bundles. More precisely, there is a family of semistable principal $G$-bundles $\Xi$ over $E\times \Cee$, such that, for $t\neq 0$, $\xi_t=\Xi|E\times \{t\}\cong \xi$, and such that $\xi _0= \Xi|E\times \{0\}$ is the $G$-bundle associated to a flat $K$-bundle via the inclusion $K\subset G$. Finally, two flat $K$-bundles are isomorphic as $G$-bundles if and only if they are isomorphic as $K$-bundles. \endstatement We note that Theorem 2.4 also holds for a non-simply connected group. The special feature of simply connected groups which we need to describe the moduli space of flat $K$-bundles is contained in the following result of Borel \cite{7} (see also \cite{22} for the analogous algebraic result, due to Springer and Steinberg): \theorem{2.5} Let $K$ be a compact, simple, and simply connected Lie group, and let $r_1$ and $r_2$ be two commuting elements of $K$. Then there exists a maximal torus $T$ in $K$ with $r_1, r_2\in T$. \endstatement Since $\pi _1(E) \cong \Zee\oplus \Zee$, to give a representation $\rho\: \pi _1(E) \to K$ is to give two commuting elements $r_1, r_2\in K$. Thus a flat $K$-bundle reduces to a $T$-bundle. In particular, we see that for a simply connected group $G$, every $G$-bundle associated to a flat bundle is split, and conversely. On the other hand, if $G$ is not simply connected, every split bundle lifts to the universal cover $\tilde G$ of $G$, so that a $G$-bundle which does not lift to $\tilde G$ cannot be split. Thus the correct notion for unliftable bundles is that of a flat bundle. Returning to the case of a simply connected group $G$, let $T$ be a maximal torus in the compact group $K$. One checks that two homomorphisms from $\pi _1(E)$ to $T$ are conjugate by an element of $K$ if and only if they are conjugate by an element of the normalizer of $T$ in $K$. Thus we have: \theorem{2.6} There is a natural bijection from the set of flat $K$-bundles up to isomorphism, or equivalently the set of semistable $G$-bundles up to S-equivalence, to the set $\Hom (\pi _1(E), T)/W$, where $W$ is the Weyl group of $K$, acting in the usual way on the maximal torus $T$. \endstatement Fix a maximal torus $T$ in $K$. If $\Lambda = \pi _1(T)$, then $T\cong U(1) \otimes _{\Zee}\Lambda$. Moreover, since $K$ is simply connected, if $\frak t_\Ar$ denotes the real Lie algebra of $T$, then $\Lambda\subset \frak t_\Ar$ is the lattice generated by the coroots $\alpha \spcheck$, where $\alpha \in \frak t_\Ar^*$ is a root. Now given a homomorphism $\rho\: \pi _1(E) \cong \Zee\oplus \Zee\to K$, the image of $\rho$ is generated by two commuting elements of $K$ and so, after conjugation, lies in $T$. The set of flat $T$-bundles is naturally $$\Hom (\pi _1(E), T) = \Hom (\pi _1(E),U(1) \otimes _{\Zee}\Lambda ) \cong \Hom (\pi _1(E), U(1))\otimes _\Zee \Lambda.$$ Now $\Hom (\pi _1(E), U(1))$ is the set of flat line bundles on $E$, and is naturally identified with $\Pic ^0E$. Since we have fixed a base point $p_0\in E$, we can further identify $\Pic ^0E$ with $E$. Thus the space of flat $T$-bundles is naturally $E\otimes _\Zee \Lambda$. On the other hand, as we are classifying not flat $T$-bundles but flat $K$-bundles, we must take the quotient of $E\otimes _\Zee \Lambda$ by the action of the Weyl group $W$ of $G$ acting on $E\otimes _\Zee \Lambda$ via the natural action of $W$ on $\Lambda$. We have thus described the coarse moduli space of semistable $G$-bundles over $E$ as $(E\otimes _\Zee \Lambda)/W$. A different proof of this result has been given by Laszlo \cite{25}. The varieties $(E\otimes _\Zee \Lambda)/W$ have been studied by Looijenga \cite{27} and Bernshtein-Shvartsman \cite{6}, who proved the following theorem: \theorem{2.7} Let $E$ be an elliptic curve and let $\Lambda$ be the coroot lattice of a simple root system $R$ with Weyl group $W$. Then $(E\otimes _\Zee \Lambda)/W$ is a weighted projective space $WP(g_0, \dots, g_r)$, where the weights $g_i$ are given as follows: $g_0 = 1$, and the remaining roots $g_i$ are found by choosing a set of simple roots $\alpha _1, \dots , \alpha _r$, and then writing the coroot $\tilde \alpha \spcheck$ dual to the highest root $\tilde \alpha$ as a linear combination $\sum _ig_i\alpha _i\spcheck$ of the coroots dual to the simple roots. In case $R$ is simply laced, we can identify the dual coroot $\alpha \spcheck \in R\spcheck$ to $\alpha$ with $\alpha$, and consequently the $g_i$ are the coefficients of $\tilde \alpha$ in terms of the basis $\alpha _1, \dots , \alpha _r$. \endstatement The proof of \cite{27} and \cite{6} makes use of formal theta functions for a complexified affine Weyl group. We shall outline a different proof of (2.7) below. Since it will be important to motivate the construction of Section 4, let us give Looijenga's reason for studying the space $(E\otimes _\Zee\Lambda)/W$. Let $(X, x_0)$ be the germ of a simple elliptic singularity whose minimal resolution has a single exceptional component which is a smooth elliptic curve $E$ with self-intersection $-3, -2$, or $-1$. These are exactly the simple elliptic singularities which can be realized as hypersurface singularities in $(\Cee ^3, 0)$, and we shall refer to them as being of type $\tilde E_6, \tilde E_7, \tilde E_8$ respectively. These singularities are weighted cones over $E$ corresponding to a line bundle $L$ on $E$ of degree $3$, $2$, or $1$, and thus have a $\Cee ^*$-action. Moreover $\Cee ^*$ also acts on the tangent space to the deformations of $(X, x_0)$. The zero weight directions (in other words those directions fixed by the $\Cee^*$-action) correspond to deforming $(X, x_0)$ in an equisingular family by deforming $E$. The remaining weights are negative, and deformations in the neagative weight space correspond to deforming $(X, x_0)$ to a rational double point (RDP) singularity or smoothing it. The local action of $\Cee ^*$ on the negative weight deformations may be globalized, and the quotient corresponding to the singularity $\tilde E_r$ is a weighted projective space $WP(g_0, \dots, g_r)$, where the weights $g_i$ are those defined above for the root system $E_r$. On the other hand, by the general theory of negative weight deformations of singularities with $\Cee ^*$-actions, and in particular by work of Pinkham \cite{30}, Looijenga \cite{26}, and later M\'erindol \cite{28}, the points of this weighted projective space parametrize triples $(\bar S, D, \varphi)$, where $\bar S$ is a generalized del Pezzo surface of degree $9-r$ (i\.e\., $\bar S$ has at worst rational double point singularities and the inverse of the dualizing sheaf $K_{\bar S}$ is ample on $\bar S$, with $K_{\bar S}^2 = 9-r$), $D \in |-K_{\bar S}|$ is a smooth divisor, not passing through the singularities of $\bar S$, and $\varphi$ is an isomorphism from $D$ to the fixed elliptic curve $E$ such that $\varphi ^*L = N_{D/\bar S}$. The moduli of such triples $(\bar S, D, \varphi)$ can be described directly in terms of the defining equations for $\bar S$ and can also be checked directly to be a weighted projective space with the correct weights. (Similar but slightly more involved arguments also handle the case of degree $4$ and $5$, in which case the singularity is a codimension two complete intersection, in the case of degree $4$, and the corresponding root system is $D_5$, or a Pfaffian singularity in case the degree is $5$, and the root system is $A_4$.) Now an elementary Torelli-type theorem shows that the pair $(\bar S, D)$ (ignoring the extra structure of $\varphi$) is determined by the homomorphism $\psi_0\: H^2_0(S; \Zee) \to D$, where $S$ is the minimal resolution of $\bar S$ and $H^2_0(S; \Zee)$ is the orthogonal complement of $[K_S]$ in $H^2(S; \Zee)$, given as follows: represent a class $\lambda\in H^2_0(S; \Zee)$ by a holomorphic line bundle $L$ on $S$ such that $\deg ( L|D) = 0$, and define $\psi_0 (\lambda)$ to be the element $L|D\in \Pic ^0D \cong D$. But $H^2_0(S; \Zee)$ is isomorphic to the root lattice for the corresponding root system $E_r$, and this isomorphism is well-defined up to the action of the Weyl group. The choice of the isomorphism $\varphi$ enables one to extend the map $\psi_0$ to a map $\psi\: H^2(S;\Zee)/\Zee[D] \to E$, essentially because on the fixed curve $E$ we can choose a $(9-r)^{\text{th}}$ root of the line bundle $L$, and conversely the choice of such a root fixes an isomorphism from $D$ to $E$ which lines up $L$ with $N_{D/\bar S}$. Now $H^2(S;\Zee)/\Zee[D]$ is dual to the coroot lattice $\Lambda$ of the root system $E_r$, and $\psi$ defines an element of $E\otimes _\Zee\Lambda$, well-defined modulo the action of $W$. In this way, we have identified $WP(g_0, \dots, g_r)$ with $(E\otimes _\Zee\Lambda)/W$. Let $\bar S$ be the result of contracting all of the curves on $S$ not meeting $D$. Thus $\bar S$ has certain rational double point (RDP) singularities. Under the identification of the moduli space of pairs $(S,D)$ with the set of $\psi\: \Lambda \spcheck \to E$, it is not difficult to show that the RDP singularities on $S$ correspond to homomorphisms $\psi$ such that there is a sub-root lattice $\Lambda ' \subseteq \Ker \psi$. In fact, the maximal such lattice $\Lambda '$ describes the type of the RDP singularities on $\bar S$. Here the main point is to show, by a Riemann-Roch argument, that if $\gamma \in \Ker \psi$ with $\gamma ^2 = -2$, then $\pm \gamma$ is represented by an effective curve on $S$ disjoint from $D$, and thus gives a singular point on the surface obtained by contracting all such curves. In this way, there is a link between subgroups of $E_r$, $r = 6,7,8$, and singularities of the corresponding del Pezzo surfaces. \section{3. Regular bundles.} Recall that, for a simply connected group $G$, the bundle $\xi$ is {\sl regular\/} if $h^0(E; \ad \xi)$ is equal to the rank of $G$. We begin by giving a detailed description of the set of regular bundles in case $G$ is one of the classical groups. At the end of the section we shall outline the general structure of regular bundles. Let us give a preliminary definition: \definition{Definition 3.1} Let $I_n$ be the vector bundle of rank $n$ and trivial determinant on $E$ defined inductively as follows: $I_1=\scrO_E$, and $I_n$ is the unique nonsplit extension of $I_{n-1}$ by $\scrO_E$. More generally, if $\lambda$ is a line bundle on $E$ of degree zero, we define $I_n(\lambda) = I_n\otimes \lambda$. \enddefinition An easy argument shows that the algebra $\Hom (I_n, I_n)$ is isomorphic to $\Cee[t]/(t^n)$, and in particular it is a commutative unipotent $\Cee$-algebra of dimension $n$. If $V$ is an arbitrary semistable vector bundle of degree zero over $E$ and $\lambda$ is a line bundle of degree zero over $E$, let $V_\lambda \subseteq V$ be the sum of all of the subbundles of $V$ which are filtered by a sequence of subbundles whose successive quotients are isomorphic to $\lambda$. An easy argument shows that $V_\lambda$ itself is the maximal such subbundle with this property and that $V = \bigoplus _\lambda V_\lambda$. A straightforward induction classifies the possible $V_\lambda$ as a direct sum $\bigoplus _jI_{k_j}(\lambda)$. From this, it is easy to check: \proposition{3.2} Let $V$ be a semistable vector bundle over $E$ with trivial determinant, i\.e\. $V$ is a principal $SL(n)$-bundle over $E$. If $V\cong \bigoplus _{i=1}^rI_{d_i}(\lambda _i)$, where the $\lambda _i$ are line bundles on $E$ of degree zero, such that $\lambda _1^{d_1} \otimes \cdots \otimes \lambda _r^{d_r} =\scrO_E$ and $\sum _id_i = n$, then $V$ is regular if and only if $\lambda _i\neq \lambda _j$ for all $i\neq j$. \endstatement To deal with the case of the symplectic or orthogonal group, the main point is to decide when a bundle $V$ carries a nondegenerate alternating or symmetric form. The crucial case is that of $I_n$. In this case, we have the following: \proposition{3.3} There exists a nondegenerate alternating pairing on $I_n$ if and only if $n$ is even. There exists a nondegenerate symmetric pairing on $I_n$ if and only if $n$ is odd. In both cases, every two such nondegenerate pairings on $I_n$ are conjugate under the action of $\Aut I_n$. \endstatement With this said, we can describe the regular symplectic bundles. It is simplest to describe them via the standard representation: \proposition{3.4} Let $V$ be a vector bundle of rank $2n$ over $E$ with a nondegenerate alternating form, and suppose that the dimension of the group of symplectic automorphisms of $V$ is $n$. Then there exist positive integers $d_i$ and nonnegative integers $a_j$, $0\leq j\leq 3$, with $\sum _id_i + \sum _ja_j = n$, such that $V$ is isomorphic to $$\bigoplus _i\left(I_{d_i}(\lambda _i) \oplus I_{d_i}(\lambda _i^{-1})\right)\oplus I_{2a_0} \oplus I_{2a_1}(\eta _1) \oplus I_{2a_2}(\eta _2) \oplus I_{2a_3}(\eta _3),$$ where the $\lambda _i$ are line bundles of degree zero, not of order two, such that, for all $i\neq j$, $\lambda _i\neq \lambda _j^{\pm1}$, and $\eta_1, \eta _2, \eta _3$ are the three distinct line bundles of order two on $E$. Conversely, suppose that $V$ is a vector bundle as given above. Then $V$ has a nondegenerate alternating form, all such forms have a group of symplectic automorphisms of dimension exactly $n$, and every two nondegenerate alternating forms on $V$ are equivalent under the action of $\Aut V$. \endstatement In particular, we see that a regular symplectic bundle is always a regular bundle in the sense of $SL(2n)$-bundles. For $SO(2n)$ and $SO(2n+1)$, the situation is a little more complicated for two reasons. First, we shall only consider those bundles which can be lifted to $Spin (2n)$ or $Spin (2n+1)$, but shall not describe here the actual choice of a lifting. Secondly, because of (3.3), it turns out that a regular $SO(n)$-bundle does not always give a regular $SL(n)$-bundle. \proposition{3.5} Let $V$ be a vector bundle of rank $2n$ over $E$ with a nondegenerate symmetric form, and suppose that the dimension of the group of orthogonal automorphisms of $V$ is $n$. Finally suppose that $V$ can be lifted to a principal $Spin (2n)$-bundle. Then $V$ is isomorphic to $$\bigoplus _i\left(I_{d_i}(\lambda _i) \oplus I_{d_i}(\lambda _i^{-1})\right)\oplus \bigoplus _j\left(I_{2a_j+1}(\eta _j)\oplus\eta _j\right)$$ where the $\lambda _i$ are line bundles of degree zero, not of order two, such that, for all $i\neq j$, $\lambda _i\neq \lambda _j^{\pm1}$, $\eta _0 = \scrO_E, \eta_1, \eta _2, \eta _3$ are the four distinct line bundles of order two on $E$, and the second sum is over some subset \rom(possibly empty\rom) of $\{0,1,2,3\}$. Conversely, every such vector bundle $V$ has a nondegenerate symmetric form, all such forms have a group of orthogonal automorphisms of dimension exactly $n$, and every two nondegenerate symmetric forms on $V$ are equivalent under the action of $\Aut V$. \endstatement Here the symmetric form on $I_{2a_0+1}\oplus\scrO_E$ consists of the orthogonal direct sum of the nondegenerate form on the factor $I_{2a_0+1}$ given by (3.3), together with the obvious form on $\scrO_E$, and similarly for the summands $I_{2a_i+1}(\eta _i)\oplus \eta _i$. Moreover, not all of the summands $I_{2a_j+1}(\eta _j)\oplus\eta _j$ need be present in $V$. We remark that, if a vector bundle $\bigoplus _jI_{d_j}(\lambda _j)$ is isomorphic to its dual, and the sum of all the factors where $\lambda _j = \eta _i$ for some $i$ has odd rank, then the same must be true for all of the $\eta _i$. Thus, if the automorphism group of $V$ is to be as small as possible, then either $V$ is as described in (3.5) or $V$ is of the form $$\bigoplus _i\left(I_{d_i}(\lambda _i) \oplus I_{d_i}(\lambda _i^{-1})\right)\oplus I_{2a_0+1}\oplus I_{2a_1+1}(\eta _1)\oplus I_{2a_2+1}(\eta_2)\oplus I_{2a_3+1}(\eta _3).$$ But in this last case $V$ does not lift to a $Spin (2n)$-bundle. The case of $SO(2n+1)$, which we shall not state explicitly, is completely analogous, except that the summand $I_{2a_0+1}\oplus\scrO_E$ is replaced by the odd rank summand $I_{2a_0+1}$, which must always be present. We return now to the study of regular bundles over a general group $G$. \proposition{3.6} Let $\xi$ be a semistable principal $G$-bundle over $E$. Then the structure group of $\xi$ reduces to an abelian subgroup of $G$. If furthermore $\xi$ is regular, the structure group of $\xi$ reduces to an abelian subgroup of $\Aut _G\xi$, which naturally sits inside $G$ up to conjugation. \endstatement In fact, one can take the structure group of $\xi$ to be of the following form. Let $\xi _0$ be the split bundle S-equivalent to $\xi$, corresponding to the representation $\rho\: \pi _1(E) \to T\subset K$. Let $T_0$ be the image of $\rho$. Then there exists a subgroup $U$ of $G$ commuting with $T_0$, which is either trivial or a $1$-parameter commutative unipotent subgroup, such that the structure group of $\xi$ reduces to $T_0U$. We now describe the set of bundles which are simultaneously regular and split. If $\xi$ is split, then $\xi$ corresponds to a point of $(E\otimes_{\Zee} \Lambda)/W$. After lifting this point to an element $\mu$ of $E\otimes_{\Zee} \Lambda$, we see that we can describe $\ad \xi$ as follows. A root $\alpha$ defines a homomorphism $\Lambda \to \Zee$, and thus a homomorphism $E\otimes _{\Zee}\Lambda \to E \cong \Pic ^0E$. Denote the image of $\mu$ in $E$ by $\alpha(\mu)$ and the corresponding line bundle by $\lambda_{\alpha(\mu)}$. Then, as vector bundles, $$\ad \xi \cong \scrO_E^r \oplus \bigoplus _\alpha\lambda_{\alpha(\mu)}.$$ Hence $\xi$ is regular if and only if, for every root $\alpha$, $\alpha (\mu) \neq 0$. In particular, there is a nonempty Zariski open subset of $(E\otimes_{\Zee} \Lambda)/W$ such that all of the corresponding split bundles are regular. In fact, on this open subset, S-equivalence is the same as isomorphism. At the other extreme, we can consider bundles which are S-equivalent to the trivial bundle. The split representative for the S-equivalence class corresponds to the image of $0\in E\otimes_{\Zee} \Lambda$ in $(E\otimes_{\Zee} \Lambda)/W$, which has the unique preimage $0\in E\otimes_{\Zee} \Lambda$. To describe the regular representative, or more precisely its adjoint bundle, we first recall the definition of the {\sl Casimir weights\/} $d_1, \dots, d_r$ of a root system $R$. These can be defined to be the numbers $m_i +1$, where the $m_i$ are the exponents of $R$ (cf\. \cite{8}, V (6.2)), and they are also the degrees of a set of homogeneous generators for the invariants of the symmetric algebra of the vector space corresponding to the root system $R$ under the action of the Weyl group. To describe a regular bundle S-equivalent to the trivial bundle, we shall describe its adjoint bundle. (Here an $(\ad G)$-bundle has in general finitely many liftings to a $G$-bundle, but exactly one of these will turn out to be S-equivalent to the trivial bundle.) \proposition{3.7} There is a unique regular $G$-bundle $\xi$ S-equivalent to the trivial bundle. As vector bundles over $E$, $$\ad \xi \cong \bigoplus _iI_{2d_i-1},$$ where the $d_i$ are the Casimir weights of the root system of $G$. \endstatement The bundle $\ad \xi$ can be seen to be an ($\ad G$)-bundle as follows: start with the bundle $I_3 = \Sym ^2I_2$. It is an $SL(2)$ bundle which descends to an $SO(3)$-bundle. Now there is a ``maximal" embedding of $SO(3)$ in $G$, unique up to conjugation. Thus there is a representation $\rho$ of $SO(3)$ on the Lie algebra $\frak g$. Under this representation $\frak g$ decomposes as a direct sum $$\frak g = \bigoplus _i\Sym ^{2d_i-2}(\Cee ^2),$$ where we view $\Cee ^2$ as the standard representation of $SL(2)$, and thus its odd symmetric powers give representations of $SO(3)$. In particular, the $G$-bundle induced by $\rho$ gives rise to the ($\ad G$)-bundle described above. We can generalize the above picture for the trivial bundle to an arbitrary bundle. Let $\xi$ be an arbitrary semistable $G$-bundle and let $\xi _0$ be the unique split bundle S-equivalent to $\xi$. Then $\Aut \xi _0$ is up to isogeny a product of $N$ factors $G_i$, where each factor $G_i$ is either simple or isomorphic to $\Cee^*$. Let $\mu \in E\otimes_{\Zee} \Lambda$ be a representative for the class of $\xi _0$. The Lie algebra $H^0(E;\ad \xi _0)$ of $\Aut \xi _0$ is identified with $$\frak h \oplus \bigoplus _{ \alpha(\mu) = 0}\frak g^\alpha,$$ where $\frak g^\alpha$ is the root space corresponding to the root $\alpha$ (and thus in particular the rank of this reductive Lie algebra is $r$). We then have: \proposition{3.8} With notation as above, let $\xi_{\text{reg}}$ be a regular semistable bundle S-equivalent to $\xi_0$. Let $r_i$ be the rank of $G_i$, where by definition $r_i =1$ if $G_i\cong \Cee^*$, and let $d_{ij}, 1\leq j\leq r_i$ be the Casimir weights of $G_i$, where we set $d_{i1} = 1$ if $G_i\cong \Cee ^*$. Then the maximal subbundle of $\ad \xi_{\text{reg}}$ which is filtered by subbundles whose successive quotients are $\scrO_E$ is $$\left(\ad \xi_{\text{reg}}\right)_{\scrO_E} = \bigoplus _{i=0}^N\bigoplus _{j=1}^{r_i}I_{2d_{ij}-1}.$$ \endstatement From this, it is possible in principle to give a complete description of $\ad \xi_{\text{reg}}$. As a consequence of Proposition 3.6, one can show: \proposition{3.9} Let $\xi$ be a semistable principal $G$-bundle. Then $\xi$ is S-equivalent to a unique regular semistable bundle and to a unique split bundle. \endstatement There are thus two canonical representatives for every S-equivalence class, depending on whether we choose the regular or the split representative. For an open dense subset of bundles, these two representatives will in fact coincide. As should be clear from Section 2, the split representatives arise most naturally from the point of view of flat connections. However, if we try to find a universal holomorphic $G$-bundle, then we must work instead with regular bundles. In fact, even working locally, it is not possible to fit the split bundles together into a universal bundle, even for $SL(n)$. Finally, we make some comments about the automorphism group of a regular bundle. \proposition{3.10} Let $\xi$ be a regular semistable $G$-bundle. If $\Aut _G(\xi)$ is the automorphism group of $\xi$ and $(\Aut_G(\xi))^0$ is the component of $\Aut _G(\xi)$ containing the identity, then $(\Aut _G(\xi))^0$ is abelian. Moreover, $\Aut _G(\xi)$ is itself abelian if and only if $\xi$ corresponds to a smooth point of the moduli space of S-equivalence classes of semistable $G$-bundles. \endstatement In fact, a careful analysis of the root systems involved shows that the singular locus of the moduli space corresponding to $\Zee/d\Zee$-isotropy is smooth and irreducible, of dimension equal to the number of $i$ such that $d|g_i$, in the notation of Theorem 2.7. Of course, this statement also follows directly from Theorem 2.7. \section{4. The parabolic construction.} In this section, we describe a method of constructing families of regular semi\-stable $G$-bundles. The motivation is as follows: we seek to find an analogue for bundles of the singularities picture outlined above in Section 2. That is, we seek to find a mildly ``singular" (in other words, unstable) $G$-bundle $\xi _0$ together with a $\Cee ^*$-action on its deformation space, such that the weighted projective space corresponding to the quotient of the negative weight deformations of $\xi _0$ by $\Cee^*$ is both the weighted projective space $WP(g_0, \dots, g_r)$ and is the coarse moduli space of semistable $G$-bundles modulo S-equivalence. (Actually, with our conventions the action of $\Cee^*$ will be by positive weights.) It will also turn out that the points of the weighted projective space parametrize regular $G$-bundles, as opposed to split bundles, and will thus enable us to find locally a universal $G$-bundle away from the orbits where $\Cee ^*$ does not act freely. In fact, in many cases we can use this construction to produce a global universal $G$-bundle. To pursue this idea further, we have seen that unstable $G$-bundles over $E$ reduce to a parabolic subgroup of $E$, and further to a Levi factor $L$. Conversely, fix a maximal parabolic subgroup $P$ of $G$ and a Levi decomposition $P = LU$, where $U$ is the unipotent radical of $P$ and $L$ is the reductive or Levi factor. Then $U$ is normal, all Levi factors are conjugate in $P$, and the quotient homomorphism $P \to L$ is well-defined. The group $L$ is never semisimple; in fact, since $P$ is a maximal parabolic, the connected component of the center of $L$ is $\Cee ^*$. The maximal parabolic subgroup $P$ has a canonical character $\chi\: P \to \Cee ^*$ (the unique primitive dominant character), which is induced from a character $L\to \Cee$. Using this character, we can define the determinant line bundle of a principal $L$-bundle over $E$. Fix an $L$-bundle $\eta$, such that $\det \eta$ has negative degree. The induced $G$-bundle $\xi _0$ is unstable, because $\xi_0$ also reduces to the opposite parabolic to $P$, and the determinant line bundle for the primitive dominant character of the opposite parabolic has positive degree. Consider the set of all $P$-bundles $\xi$ such that the associated $L$-bundle (via the homomorphism $P\to L$) is $\eta$. It is straightforward to classify all such bundles: the group $L$ acts by conjugation on $U$, and the $L$-bundle $\eta$ and the action of $L$ on $U$ define a sheaf of unipotent groups $\underline{U}(\eta)$ on $E$, which is in general nonabelian. The set of all isomorphism classes of $P$-bundles $\xi$ which reduce to $\eta$ may then be identified with the cohomology set $H^1(E; \underline{U}(\eta))$. The $\Cee^*$ in the center of $L$ then acts on $H^1(E; \underline{U}(\eta))$. Cohomology sets similar to $H^1(E; \underline{U}(\eta))$, arising from the $H^1$ of a sheaf of unipotent groups over a base curve, have been studied in a different context by Babbitt and Varadarajan \cite{3}, following ideas of Deligne, as well as by Faltings \cite{14}. Using similar ideas, one can show that $H^1(E; \underline{U}(\eta))$ has a (non-canonical) linear structure and that $\Cee ^*$ acts linearly in this structure with positive weights (following certain standard conventions), so that the quotient is isomorphic to a weighted projective space. In the case of $SL(n)$, it is easy to make these ideas explicit. The maximal parabolic subgroups of $SL(n)$ correspond to filtrations $\{0\} \subset \Cee ^d \subset \Cee ^n$, where $0< d < n$. For each such $d$, there is a unique stable bundle $W_d$ over $E$ of rank $d$ such that $\det W_d = \scrO_E(p_0)$. The unstable bundle which we consider is then $W_d^* \oplus W_{n-d}$, and it has a nontrivial $\Cee^*$-action which acts trivially on $\det (W_d^* \oplus W_{n-d})$. In this case, a straightforward argument shows: \theorem{4.1} Let $V$ be a regular semistable vector bundle of rank $n$. Then there is an exact sequence $$0 \to W_d^* \to V \to W_{n-d} \to 0. $$ Moreover, the automorphism group of $V$ acts transitively on the set of subbundles of $V$ isomorphic to $W_d^*$ whose quotients are isomorphic to $W_{n-d}$. Finally, if $V$ is a nonsplit extension of $W_{n-d}$ by $W_d^*$, then $V$ is in fact a regular semistable vector bundle. \endstatement We note that in this case the parabolic subgroup in question is $$P = \left\{\,\pmatrix A&B\\O&D\endpmatrix: A\in GL(d), D\in GL(n-d), \det A\cdot \det D = 1\,\right\},$$ the Levi factor of $P$ is given by $$L= \left\{\,\pmatrix A&O\\O&D\endpmatrix: A\in GL(d), D\in GL(n-d), \det A\cdot \det D = 1\,\right\},$$ and the unipotent radical $U$ of $P$, which in this case is abelian, is given by $$U= \left\{\,\pmatrix I&B\\O&I\endpmatrix: B \text{ is a $d\times (n-d)$ matrix}\,\right\}.$$ It is easy then to identify $H^1(E; \underline{U}(\eta))$ with the usual sheaf cohomology group $H^1(E; W_{n-d}^*\otimes W_d^*)$ and the $\Cee ^*$-action with the usual one, up to a factor. In this way, the moduli space of regular semistable vector bundles over $E$ of rank $n$ and trivial determinant is identified with $\Pee ^{n-1}$, a fact which could also be established by spectral cover methods \cite{20}. The full tangent space to the deformations of the unstable bundle $W_d^* \oplus W_{n-d}$ keeping the determinant trivial is $H^1(E; \ad(W_d^* \oplus W_{n-d}))$. This group contains the subgroup $H^1(E; W_{n-d}^*\otimes W_d^*)$ which is tangent to the set of extensions described above. The one remaining direction has $\Cee^*$-weight zero, which correponds to moving the point $p_0$ on $E$ and which should be viewed as a one parameter family of locally trivial deformations. In the case of $SL(n)$, or equivalently the root system $A_{n-1}$, every maximal parabolic subgroup has an abelian unipotent radical and there is an appropriate construction from any such subgroup giving the moduli space of regular semistable $G$-bundles. In all other cases, we have the following: \theorem{4.2} Let $G$ be a complex, simple, and simply connected group, not of type $A_n$. Then there exists a unique maximal parabolic subgroup $P$ of $G$, up to conjugation, such that, if $L$ is the Levi factor of $P$, then there exists an $L$-bundle $\eta$ with the following properties: \roster \item"{(i)}" The connected component of the automorphism group of $\eta$ as an $L$-bundle is $\Cee ^*$. \item"{(ii)}" The line bundle $\det \eta$ has negative degree, and so the $G$-bundle $\xi _0$ induced by $\eta$ is unstable. \item"{(iii)}" If $U$ is the unipotent radical of $P$, then the nonabelian cohomology set $H^1(E; \underline{U}(\eta))$ has the structure of affine $(r+1)$-dimensional space. \item"{(iv)}" There exists a linear structure on $H^1(E; \underline{U}(\eta))$ for which the natural copy of $\Cee ^*\subseteq \Aut_G\xi _0$ acts linearly, fixing the trivial element, and with negative weights. The stabilizer of every nontrivial element of $H^1(E; \underline{U}(\eta))$ is finite, and the quotient $(H^1(E; \underline{U}(\eta)) - \{0\})/\Cee^*$ is a weighted projective space $WP(g_0, \dots, g_r)$. \item"{(v)}" If $\xi$ is a $P$-bundle over $E$ corresponding to an element of $H^1(E; \underline{U}(\eta)) - 0$, then $\xi$ is a regular semistable bundle. \endroster In all cases, the bundle $\eta$ with the above properties is uniquely specified by requiring that $\det \eta = \scrO_E(-p_0)$. \endstatement In fact, (iv) and (v) are a consequence of the other properties. If we do not specify that $\det \eta = \scrO_E(-p_0)$, then it is still the case that $\det\eta$ must have degree $-1$, and so $\eta$ is specified up to translation on $E$. We note that all of the weights are equal, in other words the weighted projective space is an ordinary projective space, exactly in the cases $A_n$ and $C_n$, in other words for the groups $SL(n+1)$ and $Sp(2n)$. In all other cases, for a simply connected group $G$, the weighted projective space will in fact have singularities. To describe the maximal parabolic subgroups which arise in Theorem 4.2, note first that maximal parabolic subgroups of $G$, up to conjugation, are in one-to-one correspondence with the vertices of the Dynkin diagram of the corresponding root system. In case $G$ is $D_n$ or $E_6, E_7, E_8$, the maximal parabolic subgroup in Theorem 4.2 corresponds to the unique trivalent vertex of the Dynkin diagram. In the remaining cases, the vertex in question is the unique vertex meeting the multiple edge which is the long root. (Such vertices will be trivalent in an appropriate sense except for the case $C_n$.) Let us describe the construction explicitly for the remaining classical groups. The simplest case after $A_n$ is the case of $Sp(2n)$, in other words $C_n$. In this case the parabolic in question corresponds to those elements of $Sp(2n)$ which preserve a totally isotropic subspace of dimension $n$. Thus $$P = \left\{\,\pmatrix T&B\\O&{}^tT^{-1}\endpmatrix: T\in GL(n), T^{-1}B = {}^t(T^{-1}B)\,\right\},$$ the Levi factor of $P$ is given by $$L= \left\{\,\pmatrix T&O\\O&{}^tT^{-1}\endpmatrix: T\in GL(n) \,\right\},$$ and the unipotent radical $U$ of $P$, which in this case is also abelian, is given by $$U= \left\{\,\pmatrix I&B\\O&I\endpmatrix: {}^tB = B\,\right\}.$$ The unstable symplectic bundle corresponding to $\eta$ is the bundle $W_n^*\oplus W_n$, with the first factor embedded as a totally isotropic subbundle and the second as its dual. It is easy then to identify $H^1(E; \underline{U}(\eta))$ with the usual sheaf cohomology $H^1(E; \Sym ^2W_n^*)$. Here $\Cee^*$ acts with constant weight, so that the quotient is an ordinary (smooth) $\Pee^{n-1}$. Next we consider $Spin(2n)$, although here it will be more convenient to work in $SO(2n)$. The natural analogue of the construction for the symplectic group would lead to the unstable bundle $W_n^*\oplus W_n$, together with the symmetric nondegenerate form for which $W_n^*$ is isotropic and which identifies the dual of $W_n^*$ with the complementary $W_n$. Such orthogonal bundles do not lift to $Spin(2n)$, although this construction does identify all of the regular semistable $SO(2n)$-bundles with $w_2\neq 0$ with the projective space on $H^1(E; \bigwedge ^2W_n^*)$, which is a $\Pee ^{n-2}$. For liftable $SO(2n)$-bundles, we use the parabolic subgroup corresponding to the trivalent vertex, which is the subgroup of $g\in SO(2n)$ preserving an isotropic subspace of rank $n-2$. In this case the unipotent radical is nonabelian. The bundle $\eta$, viewed as an unstable $SO(2n)$-bundle $\xi _0$, is the bundle $$\xi _0 = W_{n-2}^* \oplus Q_4 \oplus W_{n-2},$$ where $W_{n-2}^*$ is an isotropic subspace, $Q_4$ is the $SO(4)$-bundle $\scrO_E\oplus \eta _1\oplus \eta _2 \oplus \eta _3$, in the notation of Section 3, with a diagonal nondegenerate symmetric pairing, and $Q_4$ is orthogonal to the direct sum $W_{n-2}^* \oplus W_{n-2}$. More invariantly $Q_4 = Hom (W_2, W_2)$ with the quadratic form given by the trace. Note that neither $Q_4$ nor $W_{n-2}^* \oplus W_{n-2}$ lifts to a $Spin$-bundle, and hence the direct sum is liftable. In this case, the nonabelian cohomology set $H^1(E; \underline{U}(\eta))$ (for $Spin(2n)$) has a weight $1$ piece given by $H^1(Q_4\otimes W_{n-2}^*)$, of rank $4$, and a weight $2$ piece given by $H^1(E; \bigwedge ^2W_{n-2}^*)$, which has rank $n-3$. Similar results hold for $Spin (2n+1)$, by replacing $Q_4$ by $$Q_3 = \eta _1\oplus \eta _2 \oplus \eta _3 = \ad W_2.$$ Returning to the general case, let us show that the weighted projective space $WP(g_0, \dots, g_r)$ arising from the parabolic construction can be naturally identified with $(E\otimes _\Zee\Lambda)/W$, thus giving a new proof of Looijenga's theorem. One first shows that there exists a universal $G$-bundle over $E\times H^1(E; \underline{U}(\eta))$ in the appropriate sense. By general properties, there is a $\Cee^*$-equivariant map from $H^1(E; \underline{U}(\eta))-0$ to the moduli space of semistable $G$-bundles, in other words to $(E\otimes _\Zee\Lambda)/W$. \theorem{4.3} The induced map $WP(g_0, \dots, g_r) \to (E\otimes _\Zee\Lambda)/W$ is an isomorphism. \endstatement The essential point of the proof is to compare the determinant line bundles on the two sides, and then to use the elementary fact that a degree one morphism from a weighted projective space to a normal variety is an isomorphism. On the weighted projective side, the determinant line bundle is always Cartier, and in fact it is the line bundle $K_{WP^r}^{2}$. On the other side, it is easy to calculate the preimage of the determinant line bundle in $E\otimes _\Zee\Lambda$. At least in the case of a simply laced root system $R$, the fact that the degree of the morphism in question is one then follows from the fact that the order of the Weyl group is $r!(g_1\cdots g_r)\det R$ \cite{8}. The parabolic construction also leads to a proof of the existence of universal bundles in certain cases. For a fixed $G$, we denote by $\Cal M_E=\Cal M_E(G)$ the moduli space of regular semistable $G$-bundles over $E$ and by $\Cal M_E^0$ the smooth locus of $\Cal M_E$. \theorem{4.4} If $G= SL(n)$, let $P_d$ be the maximal parabolic subgroup of $SL(n)$ stabilizing the flag $\{0\} \subset \Cee^d\subset \Cee^n$, and if $G\neq SL(n)$, let $P$ be the maximal parabolic subgroup of $G$ described in Theorem \rom{4.2}. Let $n_P$ be the positive integer defined as follows\rom: \roster \item"{(i)}" If $G=SL(n)$ and $P=P_d$, then $n_{P_d} = n/\gcd(d,n)$. \item"{(ii)}" If $G$ is of type $C_n$, $B_n$ with $n$ even, or $D_n$ with $n$ odd, then $n_P = 2$. \item"{(iii)}" In all other cases, $n_P =1$. \endroster Let $\bar G$ be the quotient of $G$ by the unique subgroup of the center of $G$ of order $n_P$. Then the universal $G$-bundle over $E\times H^1(E; \underline{U}(\eta))$ descends to a universal $\bar G$-bundle $\overline{\Xi}$ on $E\times \Cal M_E^0$. \endstatement Let us mention the analogous results for families of elliptic curves over a base $B$. Let $\pi \: Z \to B$ be a flat family, all of whose fibers are smooth elliptic curves or more generally irreducible curves of arithmetic genus one (i\.e\. smooth, nodal, or cuspidal curves). Let $\sigma$ be a section of $\pi$ meeting each fiber in a smooth point. Associated to $Z$ is the line bundle $L$ on $B$ defined by $L^{-1} = R^1\pi _*\scrO_Z$, which can be identified with $\scrO_Z(\sigma)|\sigma$ under the isomorphism $\sigma \to B$ induced by $\pi$. We want to describe the parabolic construction along the family $Z$. To do so, recall that we have the weights $g_i$ of (2.8), which we assume ordered so that $g_i \leq g_{i+1}$. Recall also that we have defined the Casimir weights $d_1, \dots, d_r$ of a root system $R$ in Section 3. We order the $d_i$ by increasing size, except in the case of $D_n$, where we order the $d_i$ by: $2, 4,n,6, 8, \dots, 2n-2$. Our result in families can then be somewhat loosely stated as follows: \theorem{4.5} Suppose that $G\neq E_8$. The parabolic construction then globalizes over $Z$ to give a bundle of nonabelian cohomology groups over $B$. This bundle is a bundle of affine spaces with a $\Cee^*$-action which is isomorphic to the vector bundle $$\scrO_B \oplus L^{-d_1} \oplus \cdots \oplus L^{-d_r}.$$ Via this isomorphism $\Cee ^*$ acts diagonally on the line bundles in the direct sum, by the weight $g_i$ on the factor $ L^{-d_i}$ \rom(and with weight $g_0 = 1$ on the factor $\scrO_B$\rom). The associated bundle of weighted projective spaces is then a universal relative moduli space for $G$-bundles which are regular and semistable on every fiber. \endstatement A result closely related to Theorem 4.5 was established by Wirthm\"uller \cite{33}, who also noted the exceptional status of $E_8$. We note that, from our point of view, in the case of $E_8$ there is a family of weighted projective spaces over the open subset $B'$ of $B$ over which the fibers of $\pi$ are either smooth or nodal. However, this family is not the quotient of a vector bundle minus its zero section by $\Cee^*$ acting diagonally. Furthermore, the construction degenerates in an essential way at the cuspidal curves. A similar phenomenon appears if we try to classify generalized del Pezzo surfaces of degree one with an appropriate hyperplane section. \section{5. Automorphism sheaves and spectral covers.} In this section, we fix $G$ and denote by $\Cal M_E(G)=\Cal M_E$ the moduli space of regular semistable $G$-bundles over $E$. Likewise, given an elliptic fibration with a section $\pi \: Z \to B$ whose fibers are smooth elliptic curves or nodal or cuspidal cubics (except in the case $G=E_8$ where we will not allow cuspidal fibers), we have a relative moduli space $\Cal M_{Z/B}= \Cal M_{Z/B}(G)$. Thus in all cases $\Cal M_{Z/B}$ is a bundle of weighted projective spaces. Because of fixed points for the $\Cee ^*$ action, the universal $G$-bundle over $E\times H^1(E; \underline{U}(\eta))$ does not descend to a universal $G$-bundle over $E\times \Cal M_E$, even locally, near the singular points of $\Cal M_E$, and a similar statement holds in families. However, let $\Cal M_E^0$ denote the smooth locus of $\Cal M_E$, and similarly for $\Cal M_{Z/B}^0$. Then locally in either the classical or \'etale topology there exists a universal bundle $\Xi$ over $E\times \Cal M_E^0$, and similarly for $Z\times _B\Cal M_{Z/B}^0$. As we have seen in Theorem 4.4, there also exists a $\bar G$-bundle $\overline{\Xi}$ over $E\times \Cal M_E^0$, where $\bar G$ is a quotient of $G$ by a subgroup of the center of order at most two. In particular, a universal adjoint bundle always exists. In this section, we describe the issues of the existence and uniqueness of a global universal bundle over $E\times \Cal M_E^0$ or $Z\times _B\Cal M_{Z/B}^0$. There are other questions closely related to these. Given a family $\pi \: Z \to B$ as above, suppose that $\Xi$ is a $G$-bundle over $Z$ such that $\Xi |\pi ^{-1}(b)$ is a semistable bundle for all $b$ for which $\pi ^{-1}(b)$ is smooth. Then $\Xi$ defines a section of $\Cal M_{Z/B}$ over the open subset of $B$ consisting of such $b$. At the singular points of $\Cal M_{Z/B}$, the section is locally liftable to the affine bundle of cohomology groups over $B$. Conversely, a locally liftable section defines local $G$-bundles over $\pi ^{-1}(U)$ for all sufficiently small open sets $U$ of $B$ (in the classical or \'etale topology). Note that the parabolic construction extends over the singular fibers of $\pi$ (except for cuspidal fibers in case $G=E_8$), dictating the correct definition of regular semistable $G$-bundles for a singular fiber. When does a locally liftable section of $\Cal M_{Z/B}$ actually determine a $G$-bundle over $Z$? More generally, how can we describe the set (possibly empty) of all bundles corresponding to a given section? For simplicity, we shall assume that the section does not pass through the singular points of $\Cal M_{Z/B}$. Thus, if there existed a relative universal bundle over $Z\times _B\Cal M_{Z/B}^0$, we could simply pull this bundle back by the section to obtain a bundle over $B$. While a relative universal bundle does not usually exist, there are many cases where a section does indeed determine a $G$-bundle. However, our answers are complete only in the cases $G = SL(n), Sp(2n)$. Working for the moment with a single curve $E$, over an open subset of $\Cal M_E^0$ where there exists a local universal bundle $\Xi$, there is an associated group scheme $\underline{\Aut}(\Xi)$. Because the associated automorphism groups are abelian on $\Cal M_E^0$, as follows from (3.10), these local group schemes piece together to give an abelian group scheme over $\Cal M_E^0$, whose associated sheaf of sections will be denoted $\Cal A$. In the usual way, the obstruction to finding a global universal principal $G$-bundle over $E\times \Cal M_E^0$ lies in $H^2(\Cal M_E^0; \Cal A)$, and if this obstruction is zero, then the set of all such principal bundles is a principal homogeneous space over $H^1(\Cal M_E^0; \Cal A)$. More generally, given an elliptic fibration $\pi\: Z \to B$ as above, we can fit together the automorphism group schemes of local universal bundles to find an abelian group scheme over $\Cal M_{Z/B}^0$ whose fiber over every point $b\in B$ is the group scheme constructed above. Let $\Cal A_B$ denote the sheaf of sections of this group scheme. Given a section $s$ of $\Cal M_{Z/B}^0 \to B$, we can pull back the the above group scheme to obtain a group scheme over $B$, whose sheaf of sections we denote by $\Cal A_B(s)$. Just as in the case of a single smooth elliptic curve, the obstruction to finding a $G$-bundle over $Z$ corresponding to the section $s$ lies in $H^2(B; \Cal A_B(s))$, and if this obstruction is zero, then the set of all such bundles is a principal homogeneous space over $H^1(B; \Cal A_B(s))$. Let us describe the sheaf $\Cal A$ in the case of $SL(n)$ and a fixed elliptic curve $E$ in more detail. For each integer $d$, $1\leq d \leq n-1$, one can construct a universal extension $\Cal E_d$ over $E\times \Pee ^{n-1}$, viewing $\Pee^{n-1}$ as $\Ext ^1(W_{n-d}, W_d^*)$, which fits into an exact sequence $$0 \to \pi _1^*W_d^* \otimes \pi _2^*\scrO_{\Pee ^{n-1}}(1) \to \Cal E_d \to \pi _1^*W_{n-d} \to 0.$$ Clearly, $\det \Cal E_d$ has trivial restriction to each slice $E\times \{s\}$ but is not in fact trivial. On the other hand, since the restriction of $\Cal E_d$ to every fiber is regular and semistable, $\pi _2{}_*Hom (\Cal E_d, \Cal E_d)$ is a sheaf of locally free commutative $\Cee$-algebras over $\Pee^{n-1}$ of rank $n$, and thus corresponds to a finite morphism $\nu \: T \to \Pee ^{n-1}$ of degree $n$, which we shall call the {\sl spectral cover\/} of $\Pee ^{n-1}$. It is straightforward to identify the base $\Pee ^{n-1}$ with the complete linear system $|np_0|$ and the cover $T$ with the incidence correspondence in $\Pee ^{n-1}\times E$, in other words $$T = \left\{\, (\sum _{i=1}^ne_i, e): \sum _{i=1}^ne_i \in |np_0|, e = e_i {\text{ for some $i$}}\,\right\}.$$ Thus $T$ is smooth, and it has the structure of a $\Pee ^{n-2}$-bundle over $E$ such that the $\Pee ^{n-2}$ fibers are mapped to hyperplanes in $\Pee^{n-1}$ under $\nu$. Another way to describe $T$ is as follows: let $\Lambda\cong \Zee ^{n-1}$ as the sublattice of $\Zee ^n$ of vectors whose sum is zero, acted on by the Weyl group $\frak S_n$, so that $\Pee ^{n-1} = |np_0| = (E\otimes \Lambda)/\frak S_n$. Let $W_0 =\frak S_{n-1}\subset \frak S_n$ be the stabilizer of the vector $e_n \in \Zee ^n$. Then $T = (E\otimes \Lambda)/W_0$. A standard argument shows that, if $\Cal V$ is a vector bundle over $E\times \Pee^{n-1}$ whose restriction to every slice is isomorphic to the corresponding restriction of $\Cal E_d$, then $\pi _2{}_*Hom (\Cal V, \Cal E_d)$ is locally free of rank one over $\pi _2{}_*Hom (\Cal E_d, \Cal E_d) = \nu _*\scrO_T$, and thus corresponds to a line bundle on $T$, and conversely every line bundle on $T$ defines a vector bundle $\Cal V$ with the above property. It is helpful to compare this situation with the one usually encountered in algebraic geometry, where we try to make a moduli space of simple vector bundles and then the only choice is to twist by the pullback of a line bundle from the moduli space factor. From this it follows that, in the case of $SL(n)$, the automorphism sheaf $\Cal A$ is given by the kernel of the norm homomorphism $\nu_*\scrO_T^* \to \scrO_{\Pee ^{n-1}}^*$. Hence there is an exact sequence $$0 \to H^1(\Pee^{n-1}; \Cal A) \to \Pic T \to \Pic \Pee ^{n-1} \to H^2(\Pee^{n-1}; \Cal A) \to H^2(\scrO_T^*)\to 0.$$ Thus, $H^1(\Pee^{n-1}; \Cal A) \cong \Zee\times E$ for $n>2$ and $H^1(\Pee^1; \Cal A) \cong E$, and $H^2(\Pee^{n-1}; \Cal A) \cong H^3(T; \Zee)$. There is also an analogue of the above exact sequence where we take \'etale cohomology. In this case, $H^1(\Pee^{n-1}; \Cal A)$ is unchanged and $H^2(\Pee^{n-1}; \Cal A) \cong H_{\text{\'et}}^2(T; \Bbb G_m)$, which is a torsion group. The obstruction to gluing together local families (in either the classical or \'etale topology) of $SL(n)$-bundles to make a global $SL(n)$-bundle over $E\times \Pee^{n-1}$ lives in $H^2(\Pee^{n-1}; \Cal A)$, and in case the obstruction is zero the set of all such bundles in then a principal homogeneous space over $H^1(\Pee^{n-1}; \Cal A)$. In our case, a direct construction using the pushforward of appropriate line bundles on $E\times T$ shows that the obstruction in $H^2(\Pee^{n-1}; \Cal A)$ vanishes. Thus the family of universal $SL(n)$-bundles $\Cal V$ over $E\times \Pee^{n-1}$ is parametrized by $\Zee\times E$ for $n>2$ and by $E$ if $n=2$. In case we consider the corresponding situation in families $Z\to B$, then there exist mod $2$ obstructions to finding a principal $SL(n)$-bundle over the entire family, and these obstructions are not in general zero. On the other hand, there always exists a universal $GL(n)$-bundle $V$ such that $V|\pi ^{-1}(b)$ has trivial determinant for all $b$, so that $\det V$ is pulled back from $B$. See \cite{20} for more detail in the case of vector bundles. Similar explicit constructions can be carried out for the symplectic group. Let $\Lambda = \Zee ^n$, and let the Weyl group $W = \frak S_n \ltimes (\Zee/2\Zee)^n$ act on $\Lambda$, where the symmetric group acts by permuting the basis elements and $(\Zee/2\Zee)^n$ acts by sign changes. Then $(E\otimes \Lambda)/W = \Sym ^n\Pee ^1 = \Pee ^n$. Let $W_0 = \frak S_{n-1} \ltimes (\Zee/2\Zee)^{n-1}$ be the subgroup of $W$ fixing the last basis vector, and set $T^{\text{sp}} = (E\otimes \Lambda)/W_0 = \Pee ^{n-1} \times T$. The group $W_0$ is a subgroup of index two in the larger group $W_1 = \frak S_{n-1} \ltimes (\Zee/2\Zee)^n$, and there is an induced involution $\iota$ on $T^{\text{sp}}$, with quotient $T^{\text{sp}}/\iota = S = \Pee ^{n-1} \times \Pee ^1$. We then have: \proposition{5.1} For the symplectic group $Sp(2n)$, the automorphism sheaf $\Cal A^{\text{sp}}$ over $\Cal M_E(Sp(2n)) \cong \Pee^n$ is given by $$\Cal A^{\text{sp}} = \{\, f\in \nu_*\scrO_{T^{\text{sp}}}^*: \iota ^*f= f^{-1}\,\}.$$ \endproclaim Using (5.1) one can show that there is a universal bundle over $E\times \Cal M_E(Sp(2n))$ as well, and that the set of all universal bundles is parametrized by $E$. Thus we have constructed universal bundles over $E\times \Cal M_E$ in the two cases where the moduli space is smooth. It then follows from Theorem 4.4 that a universal bundle exists over $\Cal M_E^0(G)$ in all cases, with the possible exception of $G= Spin (4n+1)$ and $G = Spin (4n+2)$. We return to the case of a general $G$ and analyze the structure of the sheaf $\Cal A$ over $\Cal M_E^0$. Since $\Cal A$ is the sheaf of sections of an abelian algebraic group scheme, there is the exponential map $\exp$ from the corresponding sheaf of Lie algebras $\operatorname{Lie}\Cal A$ to $\Cal A$. The kernel of $\exp$ is a constructible sheaf, which we denote by $\underline{\Lambda}$, and the image of $\exp$ is the sheaf $\Cal A^0$ which, locally, consists of all sections of $\underline{\Aut}(\Xi)$ passing through the identity component of every fiber. First we note that $\Cal A = \Cal A^0$ on the Zariski open subset $U$ of $\Cal M^0$ consisting of split bundles, where the fiber over $x\in U$ of the group scheme corresponding to $\Cal A$ is $(\Cee ^*)^r$ and is connected. If the root system for $G$ is simply laced, we can say more: \proposition{5.2} Suppose that $G$ is simply laced. If $G\neq SL(2)$, then the set $$\{\, \xi \in \Cal M^0: \Cal A_\xi \neq \Cal A^0_\xi\,\}$$ has codimension at least two in $\Cal M^0$. \endproclaim As a consequence, in the relative setting, for $G$ simply laced, if $\dim B = 1$ and $G \neq SL(2)$, then for a generic section $s$ of $\Cal M_{Z/B}^0$, we can always assume that $\Cal A_B(s) = \Cal A_B^0(s)$. The above proposition does not hold if $G$ is not simply laced; for example, it fails for $Sp(2n)$. Next we turn to $\Cal A^0 =\operatorname{Lie}\Cal A /\underline{\Lambda}$. Note that, in case there is a universal bundle $\Xi$ over $E\times \Cal M^0$, then $\operatorname{Lie}\Cal A = R^0p_2{}_*(\ad \Xi)$ is dual to $R^1p_2{}_*(\ad \Xi)$, which is the tangent bundle to $\Cal M^0$. Thus $\operatorname{Lie}\Cal A = \Omega ^1_{\Cal M^0}$ is the cotangent bundle. In fact, this statement always holds, since a universal bundle exists locally and the automorphism sheaf is abelian. Another way to describe the cotangent bundle is as follows: let $\frak h$ be the Lie algebra of a Cartan subgroup of $G$. Then the Weyl group acts on $E\otimes \Lambda$ and on the trivial vector bundle $\scrO_{E\otimes \Lambda}\otimes _\Cee\frak h$, and the sheaf of $W$-invariant sections is a coherent sheaf over $(E\otimes \Lambda)/W= \Cal M$ whose restriction to $\Cal M^0$ is locally free, and in fact is $\Omega ^1_{\Cal M^0}$. The constructible sheaf $\underline{\Lambda}$ can be described as follows. Let $U$ be the open subset of $\Cal M^0$ over which the map $E\otimes \Lambda \to \Cal M$ is unramified, and let $i\: U \to \Cal M^0$ be the inclusion. Then the action of $W$ on $\Lambda$ gives a locally constant sheaf $\underline{\Lambda} _0$ on $U$, and $\underline{\Lambda} = i_*\underline{\Lambda}_0$. The map $\Lambda \to \frak h$ induces an inclusion $\underline{\Lambda} \to \left(\scrO_{E\otimes \Lambda}\otimes _\Cee\frak h\right)^W$, and this is the same as the inclusion $\underline{\Lambda} \to \operatorname{Lie}\Cal A$. This picture is related to the general theory of spectral covers of \cite{24} and \cite{10} (as has also been noted by Donagi in \cite{11}). Suppose that $\varpi$ is an element of $\frak h$ such that $W\cdot\varpi$ spans $\frak h$ over $\Cee$. In the typical application, $\varpi$ is (the dual of) a minuscule weight, if such exist. Let $W_0$ be the stabilizer of $\varpi$. If we set $T = (E\otimes \Lambda)/W_0$, then there is a surjection $\nu \: T \to \Cal M$. By pure algebra, $$\nu _*\scrO_T = \left(\scrO_{E\otimes \Lambda}\otimes_\Cee \Cee[W/W_0]\right)^W.$$ On the other hand, there is a surjection $\Cee[W/W_0] \to \frak h$ whose kernel consists of the relations in the orbit $W\cdot \varpi$. Correspondingly, there is a surjection $$\left(\scrO_{E\otimes \Lambda}\otimes_\Cee \Cee[W/W_0]\right)^W \to \left(\scrO_{E\otimes \Lambda}\otimes_\Cee \frak h\right)^W.$$ In particular $H^1(\Cal M;\operatorname{Lie} \Cal A)$ is a quotient of $H^1(\Cal M; \nu_*\scrO_T)$. Now suppose that we are in the relative case of an elliptic fibration $\pi \: Z\to B$. There is then a relative universal moduli space $\Cal M_{Z/B}$ (with the usual care in the case of $E_8$). The covers $T \to \Cal M$ defined over every smooth fiber extend to a finite morphism $\Cal T_{Z/B} \to \Cal M_{Z/B}$. A section $s$ of the map $\Cal M^0_{Z/B} \to B$ defines a finite cover $C_s$ of $B$, which we will call the {\sl spectral cover\/} in this case. Of course, $C_s$ need not be smooth or even reduced. In case $\dim B = 1$ and $s$ is generic, the above discussion identifies the connected components of $H^1(B;\Cal A_B(s))$ with an abelian variety which is a quotient of the Jacobian $J(C_s)$, and which is called the {\sl Prym-Tyurin variety\/} of the spectral cover. A straightforward dimension argument shows: \proposition{5.3} Suppose that $\dim B = 1$ and that $\Cal A_B(s)_b = \Cal A_B^0(s)_b$ for at least one point $b\in B$. Then $H^2(B; \Cal A_B(s)) = 0$. In other words, there exists a universal $G$-bundle over $B$ corresponding to the section $s$. \endstatement If however $\Cal A_B(s)_b \neq \Cal A_B^0(s)_b$ for all $b\in B$, then it is possible for there not to exist a universal $G$-bundle over $B$ corresponding to $s$, even when $G= SL(2)$. For $\dim B$ arbitrary, the possible obstructions in the case of $SL(n)$ are analyzed in detail in \cite{20}. Let us work out the twisting group $H^1(B;\Cal A_B(s))$ explicitly in the simplest cases $G= SL(n), Sp(2n)$, with $\dim B$ arbitrary: \proposition{5.4} Suppose that $G = SL(n)$. Let $C_s \to B$ be the spectral cover defined above. Then $$H^1(B;\Cal A_B(s)) = \Ker \{\, \operatorname{Norm}\: \Pic (C_s) \to \Pic B\,\}.$$ If $G = Sp(2n)$, let $C_s$ be the corresponding degree $2n$ cover of $B$, let $\iota\: C_s\to C_s$ be the induced involution, and let $f\: C_s\to D_s$ be the degree two quotient of $C_s$ by $\iota$. Then $$H^1(B;\Cal A_B(s)) = \Ker \{\, \operatorname{Norm}\: \Pic (C_s) \to \Pic D_s\,\}.$$ \endstatement Thus, in case $B$ is a curve, $H^1(B;\Cal A_B(s))$ is the generalized Prym variety of the cover $C_s\to D_s$. Similar results hold for the remaining classical groups $Spin (2n)$ and $Spin (2n+1)$. On the other hand, suppose that $G= E_6, E_7, E_8$, that $\dim B = 1$, and that the section $s$ is generic. In this case, there is an associated fibration of del Pezzo surfaces $p\: Y \to B$, where $Y$ is a smooth threefold. Moreover $Z$ is included as a smooth divisor on $Y$ so that $p|Z = \pi \: Z \to B$. Let $J^3(Y)$ denote the intermediate Jacobian of $Y$. There is an induced morphism $J^3(Y) \to J(B)$, where $J(B)$ is the ordinary Jacobian of $B$ coming from the homomorphism $H^*(Y) \to H^*(Z) \to H^{*-2}(B)$. Denote the kernel of the morphism $J^3(Y) \to J(B)$ by $J^3(Y/B)$. Finally set $$H^{2,2}_0(Y; \Zee) = \left.\Ker \{\, H^4(Y; \Zee) \to H^2(B; \Zee)\,\}\right/\Zee\cdot [Y_t],$$ where $Y_t$ is a general fiber of $p$. In general $H^{2,2}_0(Y; \Zee)$ is a finite group. We then obtain the following theorem, first proved by Kanev \cite{24} in the case $B =\Pee^1$ via the Abel-Jacobi homomorphism: \proposition{5.5} In the above situation, there is an exact sequence $$0\to J^3(Y/B) \to H^1(\Cal A_B(s)) \to H^{2,2}_0(Y; \Zee) \to 0.$$ \endstatement We note that we can interpret $H^1(\Cal A_B(s))$ as a relative Deligne cohomology group. \Refs \ref \no 1\by M. Atiyah \paper Vector bundles over an elliptic curve \jour Proc. London Math. Soc. \vol 7\yr 1957 \pages 414--452\endref \ref \no 2\by M. Atiyah and R. Bott \paper The Yang-Mills equations over Riemann surfaces \jour Phil. Trans. Roy. Soc. London A\vol 308\yr 1982\pages 523--615\endref \ref \no 3\by D.G. Babbitt and V.S. Varadarajan \paper Local moduli for meromorphic differential equations \jour Ast\'erisque \vol 169--170 \yr 1989 \pages 1--217 \endref \ref \no 4\by V. Baranovsky and V. Ginzburg \paper Conjugacy classes in loop groups and $G$-bundles on elliptic curves \jour Internat. Math. Res. Notices \vol 15 \yr 1996 \pages 733--751 \endref \ref \no 5\by A. Beauville and Y. Laszlo \paper Conformal blocks and generalized theta functions \jour Comm. Math. Phys. \vol 164 \yr 1994 \pages 385--419 \endref \ref \no 6 \by I.N. Bernshtein and O.V. Shvartsman \paper Chevalley's theorem for complex crystallographic Coxeter groups \jour Funct. Anal. Appl. \vol 12 \yr 1978 \pages 308--310 \endref \ref \no 7 \by A. Borel \paper Sous-groupes commutatifs et torsion des groupes de Lie compactes \jour T\^ohoku Math. Jour. \vol 13 \yr 1961 \pages 216--240 \endref \ref \no 8\by N. Bourbaki \book Groupes et Alg\`ebres de Lie \bookinfo Chap. 4, 5, et 6 \publ Masson \publaddr Paris \yr 1981 \endref \ref \no 9\by G. Br\"uchert \paper Trace class elements and cross-sections in Kac-Moody groups\paperinfo PhD. thesis, Hamburg \yr 1995 \endref \ref \no 10\by R. Donagi \paper Spectral covers \inbook Current topics in complex algebraic geometry (Berkeley 1992/93) \bookinfo MSRI Publications \vol 28 \publ Cambridge University Press \publaddr Cambridge \yr 1995 \endref \ref \no 11\bysame \paper Principal bundles on elliptic fibrations \jour Asian J. Math. \vol 1 \yr 1997 \pages 214--223\endref \ref \no 12 \by S.K. Donaldson \paper A new proof of a theorem of Narasimhan and Seshadri \jour J. Differential Geometry \vol 18 \yr 1983 \pages 269--277 \endref \ref \no 13\by P. Etingof and I. Frenkel \paper Central extensions of current algebras in two dimensions \jour Comm. Math. Phys. \vol 165 \yr 1994\pages 429--444\endref \ref \no 14 \by G. Faltings \paper Stable $G$-bundles and projective connections \jour J. Algebraic Geometry \vol 2\yr 1993 \pages 507--568 \endref \ref \no 15 \by G. Faltings \paper A proof for the Verlinde formula \jour J. Algebraic Geometry \vol 3\yr 1994 \pages 347--374 \endref \ref \no 16\by R. Friedman \paper Rank two vector bundles over regular elliptic surfaces \jour Inventiones Math. \vol 96 \yr 1989 \pages 283--332 \endref \ref \no 17 \bysame \paper Vector bundles and $SO(3)$-invariants for elliptic surfaces \jour J. Amer. Math. Soc. \vol 8 \yr 1995 \pages 29--139 \endref \ref \no 18\by R. Friedman and J.W. Morgan \book Smooth Four-Manifolds and Complex Surfaces \bookinfo Ergebnisse der Mathematik und ihrer Grenzgebiete 3. Folge \vol 27 \publ Springer-Verlag \publaddr Berlin Heidelberg New York \yr 1994 \endref \ref \no 19 \by R. Friedman, J.W. Morgan and E. Witten \paper Vector bundles and $F$ theory \jour Commun. Math. Phys. \vol 187 \yr 1997 \pages 679--743 \endref \ref \no 20 \bysame \paper Vector bundles over elliptic fibrations \paperinfo alg-geom/9709029 \endref \ref \no 21 \by A. Grothendieck \paper Sur la classification des fibr\'es holomorphes sur la sph\`ere de Riemann \jour Amer. J. Math \vol 79 \yr 1956 \pages 121--138 \endref \ref \no 22 \by J. Humphreys \book Conjugacy Classes in Semisimple Algebraic Groups \bookinfo Mathematical Surveys and Monographs \vol 43\publ Amer. Math. Soc. \publaddr Providence \yr 1995 \endref \ref \no 23\by V. Kanev \paper Intermediate Jacobians and Chow groups of threefolds with a pencil of del Pezzo surfaces \jour Annali Mat. pura appl. \vol 154 \yr 1989 \pages 13--48 \endref \ref \no 24\bysame \paper Spectral covers and Prym-Tjurin varieties I\inbook Abelian Varieties \bookinfo Proceedings of the International Conference (Egloffstein 1993) \eds W. Barth, K. Hulek, and H. Lange \publ de Gruy\-ter \publaddr Berlin-New York \yr 1995 \endref \ref \no 25\by Y. Laszlo \paperinfo preprint \endref \ref \no 26\by E. Looijenga \paper On the semi-universal deformation of a simple elliptic hypersurface singularity II: the discriminant\jour Topology \vol 17 \yr 1975 \pages 23--40 \endref \ref \no 27\bysame \paper Root systems and elliptic curves \jour Invent. Math. \vol 38 \yr 1976 \pages 17--32 \endref \ref \no 28\by J.-Y. M\'erindol \paper Les singularit\'es simples elliptiques, leurs d\'eformations, les surfaces de del Pezzo et les transformations quadratiques \jour Ann. Scient. E.N.S. \vol 15 \yr 1982 \pages 17--44 \endref \ref \no 29\by M. S. Narasimhan and C. S. Seshadri \paper Stable and unitary bundles on a compact Riemann surface \jour Annals of Math. \vol 82 \yr 1965 \pages 540--567 \endref \ref \no 30\by H. Pinkham \paper Simple elliptic singularities, del Pezzo surfaces, and Cremona transformations \inbook Several Complex Variables \bookinfo Proc. Symp. Pure Math. \vol 30 \publ Amer. Math. Soc. \publaddr Providence \yr 1977 \pages 69--71 \endref \ref \no 31\by S. Ramanan and A. Ramanathan\paper Some remarks on the instability flag \jour T\^ohoku Math. Jour. \vol 36 \yr 1984 \pages 269--291 \endref \ref \no 32\by A. Ramanathan\paper Stable principal bundles on a compact Riemann surface \jour Math. Ann. \vol 213 \yr 1975 \pages 129--152 \endref \ref \no 33\by K. Wirthm\"uller \paper Root systems and Jacobi forms\jour Compositio Math. \vol 82 \yr 1992 \pages 293--354 \endref \endRefs \enddocument \end
1997-10-27T07:41:28
9707
alg-geom/9707003
en
https://arxiv.org/abs/alg-geom/9707003
[ "alg-geom", "math.AG" ]
alg-geom/9707003
Hosono Shinobu
S.Hosono
GKZ Systems, Gr\"obner Fans and Moduli Spaces of Calabi-Yau Hypersurfaces
28 pages, LaTeX. (The statements in Prop.4.7 and Claim 5.8 are clarified.)
null
null
null
null
We present a detailed analysis of the GKZ(Gel'fand, Kapranov and Zelevinski) hypergeometric systems in the context of mirror symmetry of Calabi-Yau hypersurfaces in toric varieties. As an application we will derive a concise formula for the prepotential about large complex structure limits. (Talk given at the Taniguchi Symposium ``Topological Field Theory, Primitive Forms and Related Topics'' December, 1996)
[ { "version": "v1", "created": "Thu, 3 Jul 1997 09:57:30 GMT" }, { "version": "v2", "created": "Mon, 27 Oct 1997 06:41:28 GMT" } ]
2008-02-03T00:00:00
[ [ "Hosono", "S.", "" ] ]
alg-geom
\section{1. Introduction} Mirror symmetry of Calabi-Yau manifolds has been playing a central role in revealing non-perturbative aspects of the type II string vacua, i.e., the moduli spaces for a family of Calabi-Yau manifolds. Since the success in determining the quantum geometry on the IIA moduli space made by Candelas et al\cite{CdGP} in 1991, there have been many progresses and a lot of communications between physics and mathematics on this topics\cite{GY}. In this article, we will be concerned with the mirror symmetry of Calabi-Yau hypersurfaces in toric varieties. In this case, the mirror symmetry may be traced to a rather combinatorial properties of the reflexive polytopes which determines the ambient toric varieties due to ref.\cite{Bat1}. Furthermore since the period integrals of Calabi-Yau hypersurfaces turn out to satisfy the hypergeometric differential equation, ${\cal A}$-hypergeometric system, introduced by Gel'fand, Kapranov and Zelevinski (GKZ), we can study in great detail the moduli spaces of Calabi-Yau hypersurfaces. Based on the analysis of GKZ-hypergeometric system in our context, we will derive a closed formula for the prepotential, which defines the special K\"ahler geometry on the moduli spaces. In section 2, we will review the mirror symmetry of Calabi-Yau hypersurfaces in toric varieties. This is meant to fix our notations as well as to introduce the mirror symmetry due to Batyrev. In section 3, we will introduce GKZ-hypergeometric system (${\Delta^*}$-hypergeometric system) as an infinite set of differential equations satisfied by period integrals and summarize basic results following \cite{GKZ1}. We also define the extended ${\Delta^*}$-hypergeometric system incorporating the automorphisms of the toric varieties. We will remark that the ${\Delta^*}$-hypergeometric system in our context is resonant in general. In section 4, we will review basic properties of the toric ideal and the Gr\"obner fan as an equivalence classes of the term orders in the toric ideal. We will use the Gr\"obner fan to compactify the space of the variables in the ${\Delta^*}$-hypergeometric system, and propose it as a natural compactification of the corresponding family of Calabi-Yau hypersurfaces. In section 5, we will prove general existence of the so-called large complex structure limits, at which the monodromy becomes maximally unipotent\cite{Mor}. We will also present a general formula for the local solutions about these points. In the final section, we will derive a closed formula for the prepotential, which is valid about a large complex structure limit for arbitrary Calabi-Yau hypersurfaces in toric varieties. Our formula determines the special K\"ahler geometry about a large complex structure limit as well as the quantum corrected Yukawa coupling. Claim 5.8, Claim 5.11, and Claim 6.8 in the last two sections are meant to state those results that are verified in explicit calculations by many examples without general proofs. All the results except Prop.6.7 for the prepotential in the final section have already reported in refs.\cite{HKTY1}\cite{HKTY2} \cite{HLY1}\cite{HLY2}. \vspace{0.3cm}\noindent {\bf Acknowledgments.} This article is based on the joint works with A. Klemm, B.H. Lian, S. Theisen and S.-T. Yau. The author would like to thank them for their enjoyable collaboration. He is also grateful to express his thanks to the organizers of the Taniguchi symposium, as well as the Taniguchi Foundation, where he had valuable discussions with many participants. He also express his thanks to L. Borisov and the referee for their valuable comments on the first version of this article. \section{2. Mirror Symmetry of Calabi-Yau Hypersurfaces} In this section, we will summarize mirror symmetry of Calabi-Yau hypersurfaces in toric varieties due to Batyrev. We refer the paper\cite{Bat1} for details. Let $M\cong {\bf Z}^d$ be a lattice of rank $d$ and $N$ be its dual. We denote the dual pairing $M\times N \rightarrow {\bf Z}$ by $ \langle , \rangle $. A (convex) polytope ${\Delta}$ is a convex hull of a finite set of points in $M_{{\Bbb R}}:=M\otimes{\Bbb R}$. In the following, we assume ${\Delta}$ contains the origin in its interior. The polar dual ${\Delta^*}\subset N_{{\Bbb R}}$ is defined by \begin{equation} {\Delta^*}=\{ x \in N_{\Bbb R} \; \vert\; \langle x,y \rangle \geq -1 ,\; y\in {\Delta} \;\} \;\;. \label{eqn: Ds} \end{equation} \vspace{0.3cm}\noindent {\bf Definition 2.1.} A polytope ${\Delta}$ is called reflexive if it is a convex hull of a finite set of integral points in $M_{\Bbb R}$ and contains only the origin in its interior. \vspace{0.3cm}\noindent {\bf Proposition 2.2.} When a polytope ${\Delta}$ is reflexive, its dual ${\Delta^*}$ is also reflexive. \vspace{0.2cm} Since $({\Delta^*})^*={\Delta}$, reflexive polytopes come with a pair $({\Delta},{\Delta^*})$. The following descriptions about ${\Delta^*}$ with $N$ equally apply to ${\Delta}$ with $M$ by symmetry. \vspace{0.3cm}\noindent {\bf Definition 2.3.} A maximal triangulation $T_o$ of ${\Delta^*}$ is a simplicial decomposition of ${\Delta^*}$ with properties; 1) every $d$-simplex contains the origin as its vertex, 2) 0-simplices consist of all integral points of ${\Delta^*}$. \vspace{0.2cm} For a maximal triangulation $T_o$ of ${\Delta^*}$, we consider a complete fan $\Sigma({\Delta^*},T_o)$ over the triangulation $T_o$ in $N_{\Bbb R}$. Associated to the data $(\Sigma({\Delta^*},T_o),$ $N)$ we consider a toric variety ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ \cite{Oda}\cite{Ful}. Due to the property that ${\Delta^*}$ is reflexive, we have \vspace{0.3cm}\noindent {\bf Proposition 2.4.} (Prop.2.2.19 in \cite{Bat1}) {\it ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ is a projective variety for at least one maximal triangulation with its anti-canonical class $-K=\displaystyle{ \sum_{\rho \in N\cap {\Delta^*} \setminus \{ 0\} } } D_\rho$ ample. } \vspace{0.3cm} \noindent {\bf Note.} In \cite{Bat1}, the maximal triangulations with the property in this proposition are called {\it projective}. In case of $d\leq 4$, we can observe widely that every maximal triangulation is projective. More generally we observe that every triangulation of a reflexive polytope is {\it regular} which generalize projective(, see right after eq.(\ref{eqn:convex}) for the definition). For a restricted class of reflexive polytopes (the type I or II in the following classification), it has been proved (Th.4.10 in \cite{HLY2}) that every nonsingular maximal triangulation is projective, see also Remark after Th.2.5. In the following, we will write a projective toric variety by ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ choosing a projective maximal triangulation $T_o$ of ${\Delta^*}$. \vspace{0.2cm} Let us fix a basis $\{ n_1,\cdots,n_d\}$ of $N$ and denote its dual basis by $\{m_1,\cdots,m_d\}$. With respect to this basis, we denote the coordinate ring of the torus $T_N:={\rm Hom}\,_{\bf Z}(M,{\Bbb C}^*) \subset {\Bbb P}_{\Sigma({\Delta^*},T_o)}$ by ${\Bbb C}[Y_1^{\pm1},\cdots,Y_d^{\pm1}]$ with the generators $Y_k={\rm{\bf e}}(m_k):T_N\rightarrow {\Bbb C}^*$ defined by {\bf e}$(m_k)(t)=t(m_k)$. Consider a Laurent polynomial $f_{\Delta} = \sum_{\nu\in {\Delta}\cap M} c_{\nu} Y^\nu$ with complex coefficients $c_\nu$. We denote by $X_{\Delta}$ the Zariski closure of the zero locus $(f_{\Delta}=0)$ in ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ for generic $c_\nu$'s. Similarly, we consider a projective toric variety ${\Bbb P}_{\Sigma({\Delta},T_o)}$ associated to a projective maximal triangulation $T_o$ of ${\Delta}$, and denote the coordinate ring of $T_M:={\rm Hom}\,_{\bf Z}(N,{\Bbb C}^*) \subset {\Bbb P}_{\Sigma({\Delta},T_o)}$ by ${\Bbb C}[X_1^{\pm1},\cdots,X_d^{\pm1}]$ with $X_k={\rm{\bf e}}(n_k)$. \vspace{0.3cm}\noindent {\bf Theorem 2.5.} (Th.4.2.2, Corollary 4.2.3, Th.4.4.3 in \cite{Bat1}) {\it Let $({\Delta},{\Delta^*})$ be a pair of reflexive polytopes in dimensions $d\leq4$ (, in $M_{\Bbb R}$ and $N_{\Bbb R}$, respectively). Then; 1) generic hypersurfaces $X_{\Delta}\subset {\Bbb P}_{\Sigma({\Delta^*},T_o)}$ and $X_{\Delta^*} \subset {\Bbb P}_{\Sigma({\Delta},T_o)}$ define smooth Calabi-Yau manifolds, 2) these two hypersurfaces are mirror symmetric in their Hodge numbers, i.e., $h^{1,1}(X_{\Delta})=h^{d-2,1}(X_{\Delta^*}) , \;$ $ h^{d-2,1}(X_{\Delta})= h^{1,1}(X_{\Delta^*})$. } \vspace{0.3cm} \noindent {\bf Remark.} Depending on the toric data of the reflexive polytopes, the ambient spaces have (Gorenstein) singularities (Prop.2.2.2 in \cite{Bat1}) in general. We call a maximal triangulation is {\it nonsingular} if its simplices of maximal dimensions consists of unit simplices, i.e., simplices with unit volume. It is easy to deduce that the toric variety is nonsingular if the maximal triangulation is so. Now we classify the reflexive polytopes into the following three types: \par\noindent $\bullet$ type I; the polytope has no integral point in the interior of all codimension one faces, and has a nonsingular maximal triangulation, \par\noindent $\bullet$ type II; the polytope has at least one integral point in the interior of some codimension one face, and has a nonsingular maximal triangulation, \par\noindent $\bullet$ type III; the polytope does not have a nonsingular maximal triangulation. In the following, we always consider a nonsingular maximal triangulation $T_o$ for the polytopes of type I and II. Then the toric varieties ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ are projective and nonsingular for both the polytopes ${\Delta^*}$ of type I and II (Th.4.10 in \cite{HLY2}), however we distinguish these two because of the difference in the {\it root system} for their dual polytopes ${\Delta}$; \begin{equation} \begin{array}{rcl} R({\Delta},M)=\{ \alpha \in {\Delta^*}\cap N \;\vert\; && \makebox[-1em]{} \exists m_\alpha \in {\Delta}\cap M \; s.t. \langle m_\alpha, \alpha \rangle =-1 \; \\ &&\makebox[-1em]{} {\rm and}\; \langle m , \alpha \rangle \geq 0 \;(m\not= m_\alpha \in {\Delta}\cap M) \;\; \} \\ \end{array} \label{eqn: rootD} \end{equation} The root system determines the automorphisms of the toric variety ${\Bbb P}_{\Sigma({\Delta},T_o)}$ infinitesimally due to the following result, which we will utilize in the next section; \vspace{0.3cm}\noindent {\bf Proposition 2.6.} (Prop.3.13 in \cite{Oda}) {\it For a nonsingular toric variety ${\Bbb P}_{\Sigma({\Delta},T_o)}$, we have a direct sum decomposition via the root system $R({\Delta},M)$; \begin{equation} \begin{array}{rcl} {\rm Lie}({\rm Aut}\,({\Bbb P}_{\Sigma({\Delta},T_o)})) &=& H^0({\Bbb P}_{\Sigma({\Delta},T_o)},\Theta_{{\Bbb P}_{\Sigma({\Delta},T_o)}}) \\ &=& {\rm Lie}(T_M) \oplus \( \oplus_{\alpha\in R({\Delta},M)} {\Bbb C} {\rm{\bf e}}(\alpha)\delta_{m_{\alpha}}\) \\ \end{array} \label{eqn: lie} \end{equation} where $\delta_m \; (m \in M)$ is the derivation on $T_M$ defined by $\delta_m {\rm{\bf e}}(n):= \langle m,n \rangle {\rm{\bf e}}(n)$. } \newpage \section{3. Resonance in GKZ Hypergeometric System} We consider a family of Calabi-Yau hypersurfaces $X_{{\Delta^*}}(a) \subset {\Bbb P}_{\Sigma({\Delta},T_o)}$ varying the coefficients $a_{\ns{{}}}$ in the defining equation $f_{\Delta^*}(a)=\sum_{\ns{{}}\in {\Delta^*}\cap N} a_{\ns{}} X^{\ns{}}$. By this polynomial deformation, we describe the complex structure deformation of $X_{\Delta^*}$. This deformation space is mapped to that of (complexified) K\"ahler class of $X_{\Delta}\subset {\Bbb P}_{\Sigma({\Delta^*},T_o)}$ under the mirror symmetry. According to the local Torelli theorem \cite{BG}, we can introduce a local coordinate on the moduli space in terms of period integrals. In case of hypersurfaces in toric varieties, we have one canonical period integral \cite{Bat2}\cite{BC} \begin{equation} \Pi(a)={1\over (2\pi i)^d} \int_{C_0} {1\over f_{\Delta^*}(X,a)} \prod_{i=1}^{d} {d X_i \over X_i} \;\;, \label{eqn: Pi} \end{equation} with the cycle $C_0=\{ \vert X_1\vert=\cdots=\vert X_d\vert=1\}$ in $T_M$. Here we study the differential equation satisfied by (\ref{eqn: Pi}). \newsubsec{(3-1) Extended GKZ hypergeometric system } Let ${\cal A}=\{ \bar\chi_0,\cdots,\bar\chi_p\}$ be a finite set of integral points in $\{1\}\times{\Bbb R}^n\subset{\Bbb R}^{n+1}$. We assume the vectors $\bar\chi_0,\cdots,\bar\chi_p$ span ${\Bbb R}^{n+1}$. \vspace{0.3cm}\noindent {\bf Definition 3.1.} Consider the lattice of relations among the set ${\cal A}$, \begin{equation} L=\{ \; (l_0,\cdots,l_p)\in {\bf Z}^{p+1}\; \vert\; \sum_{i=0}^p l_i\bar\chi_{i,j}=0 \;, (j=1,\cdots,n+1) \;\} \;, \label{eqn: LA} \end{equation} where $\bar\chi_{i,j}$ represents the $j$-th component of the vector $\bar\chi_i$. {\it ${\cal A}$-hypergeometric system with exponent $\beta \in {\Bbb C}^{n+1}$ } is a system of differential equations for a complex function $\Psi(a)$ on ${\Bbb C}^{\cal A}$; \begin{eqnarray} {\cal D}_l \Psi(a)&=&\{ \prod_{l_i>0}\({\partial \; \over \partial a_i}\)^{l_i} - \prod_{l_i<0}\({\partial \; \over \partial a_i}\)^{-l_i} \} \Psi(a)=0 \;\; (l \in L ) \\ {\cal Z}\Psi(a)&=&\{ \sum_{i=0}^p \bar\chi_i a_i {\partial \; \over \partial a_i} -\beta \} \Psi(a) =0 \;\; . \label{eqn: gkz} \end{eqnarray} \vspace{0.3cm}\noindent {\bf Proposition 3.2.} (\cite{Bat2}) {\it The period integral (\ref{eqn: Pi}) satisfies the ${\cal A}$-hyper-\break geometric system with ${\cal A}=\{1\}\times ({\Delta^*}\cap N)$ and $\beta=(-1)\times\vec 0$. We call this hypergeometric system as ${\Delta^*}$-hypergeometric system. } \vspace{0.2cm} By direct evaluation of the action of ${\cal D}_l$ and ${\cal Z}$ on the period integral (\ref{eqn: Pi}), we obtain this proposition. Here we consider the meaning of the linear operator ${\cal Z}$. The first component of this vector operator is exactly the Euler operator, and just says that the period integral has homogeneous degree $-1$ as a function of $a_i$'s. For the other components, it is easy to deduce that these come from the invariance of the period integral under the torus actions, which act infinitesimally on the coordinate $X_k={\rm{\bf e}}(n_k)$ by $\delta_m X_k = \langle m, n_k \rangle X_k$. It is now clear that these actions should be considered for all elements in ${\rm Lie}({\rm Aut}\, ({\Bbb P}_{\Sigma({\Delta},T_o)}))$. Then we may write the invariance of the period integral under the infinitesimal action of $\xi \in {\rm Lie}({\rm Aut}\, ({\Bbb P}_{\Sigma({\Delta},T_o)}))$ by a linear differential operator ${\cal Z}_\xi$ acting on $\Pi(a)$ through \begin{equation} {\cal Z}_{\xi} \Pi(a)= \int_{C_0} \xi \( {1\over f_{\Delta^*}(X,a)} \) \prod_{i=1}^{d} {d X_i \over X_i} = 0 \;\;. \label{eqn: Zxi} \end{equation} For explicit forms of the operators ${\cal Z}_\xi$, we refer to the examples given in p.541 of \cite{HLY1}. \vspace{0.3cm}\noindent {\bf Proposition 3.3.} ((2.13) in \cite{HLY1}) {\it The period integral $\Pi(a)$ satisfies \begin{equation} \begin{array}{crl} && {\cal D}_l \Pi(a) =0 \; (l\in L) \;,\; \\ && \\ && {\cal Z}_E \Pi(a)=0 \;,\; {\cal Z}_\xi \Pi(a)= 0 \;\; (\xi \in {\rm Lie}({\rm Aut}\, ({\Bbb P}_{\Sigma({\Delta},T_o)}))), \\ \end{array} \label{eqn: extGKZ} \end{equation} where we denote the Euler operator by ${\cal Z}_E=\sum_{i=0}^p a_i{\partial\; \over \partial a_i} +1$. } \vspace{0.2cm} We call this system as {\it extended GKZ-hypergeometric system} or {\it extended ${\Delta^*}$-hypergeometric system}. By Prop.2.6, it is clear that this extended system reduces to the GKZ system if the polytope ${\Delta^*}$ is of type I. In the following, we take an approach to study mainly the ${\Delta^*}$-hypergeometric system because the set of the solutions of the extended ${\Delta^*}$-hypergeometric system can be found in that of the ${\Delta^*}$-hypergeometric system. \newsubsec{(3-2) Convergent series solutions } Here we summarize general results in \cite{GKZ1} about the convergent series solution of the ${\cal A}$-hypergeometric system with exponent $\beta$. This is to introduce the notion of the secondary fan as well as to fix our conventions and notations. Since our interest is in the period integrals, we assume ${\cal A}=\{1\}\times ({\Delta^*}\cap N)$ and $\beta=(-1)\times \vec 0$. Hereafter we write the integral points explicitly by ${\Delta^*}\cap N=\{\ns{0},\cdots,\ns{p}\} \;(\ns{0}\equiv\vec 0)$ and $\bn{i}:=1\times\ns{i} \; (i=0,\cdots,p)$. We start with a formal solution of the ${\cal A}$-hypergeometric system with exponent $\beta$ given by \begin{equation} \Pi(a,\gamma)=\sum_{l\in L} {1 \over \prod_{0\leq i\leq p} \Gamma(l_i+\gamma_i+1)}a^{l+\gamma} \;\;, \label{eqn: fsol} \end{equation} where $\beta=\sum_i \gamma_i \bn{i}$. Now define an affine space $\Phi(\beta):=\{ \gamma\in {\Bbb R}^{p+1}\vert \beta=\sum \gamma_i\bn{i} \}$. A subset $I\subset \{ 0,\cdots,p\}$ is called a base if $\bn{I}:=\{\bn{i}\vert i\in I\}$ form a basis of ${\Bbb R}^{d+1}$. Given a base $I$, we may solve $\sum_{j\in I}\gamma_j \bn{j} = \beta -\sum_{j\not\in I}\gamma_j\bn{j}$ for $\gamma_j \;(j\in I)$ and define $\Phi_{{\bf Z}}(\beta,I)=\{ \gamma\in \Phi(\beta) \;\vert\; \gamma_j \in {\bf Z} \;(j\not\in I) \}$. We choose an integral basis $A=\{ \l1, \cdots, \l{p-d} \}$ of the lattice $L$, and define $\Phi^A_{{\bf Z}}(\beta,I)=\{ \gamma\in \Phi_{{\bf Z}}(\beta,I) \;\vert\; \gamma_j=\sum_{k=1}^{p-d} \lambda_k \l{k}_j \; (0\leq \lambda_k<1, \; j\not\in I) \}$. Then $\Phi^A_{{\bf Z}}(\beta,I)$ provides a set of representatives of the quotient $\Phi_{\bf Z}(\beta,I)/L$ and kills the invariance $\gamma \rightarrow \gamma+v \; (v\in L)$ in the formal solution (\ref{eqn: fsol}). \vspace{0.3cm}\noindent {\bf Definition 3.4.} For a base $I$, define a cone in $L_{\Bbb R}=L\otimes{\Bbb R}$ by ${\cal K}({\cal A},I)=\{ l \in L_{\Bbb R} \;\vert\; l_i\geq0\;(i\not\in I)\}$. A ${\bf Z}$-basis $A\subset L$ is said {\it compatible} with a base $I$ if it generates a cone that contains ${\cal K}({\cal A},I)$. \vspace{0.3cm}\noindent {\bf Theorem 3.5.} (Prop.1 in \cite{GKZ1}) {\it Fix a base $I$ and choose a ${\bf Z}$-basis $A=\{\l1,\cdots,\l{p-d}\}$ compatible with it. Then the formal solution (\ref{eqn: fsol}) takes the form $\Pi(a,\gamma)=a^\gamma \sum_{m\in {\bf Z}^{p-d}_{\geq0}}c_m(\gamma) x^m$ for each $\gamma\in \Phi_{\bf Z}^A(\beta,I)$ with $x_k=a^{\l{k}}$. This powerseries converges for sufficiently small $\vert x_k \vert$. } \vspace{0.3cm}\noindent {\bf Remark.} The coefficient $c_m(\gamma)$ is given explicitly by $$ c_m(\gamma)={1\over \prod_{i=0}^p \Gamma(\sum_k m_k\l{k}_i + \gamma_i+1)} \;. $$ For some index $i$ of the base $I$ in the above theorem it can happen that $\sum m_k\l{k}_i+\gamma_i+1$ is non-positive for all $m \in {\bf Z}_{\geq0}^{p-d}$ , which means we have the trivial solution $\Pi(a,\gamma)\equiv0$. In this case, we multiply an infinite number $\Gamma(\gamma_i+1)$ to obtain nonzero powerseries, i.e., $ {\Gamma(\gamma_i+1) \over \Gamma(\sum m_k\l{k}_i+\gamma_i+1)} := \lim_{\varepsilon\rightarrow0} {\Gamma(\gamma_i+1+\varepsilon) \over \Gamma(\sum m_k\l{k}_i+\gamma_i+1+\varepsilon)} . $ \vspace{0.3cm}\noindent {\bf Definition 3.6.} Consider $P={\rm Conv.}\( \{ 0, \bn{0},\cdots,\bn{p} \}\)$ in ${\Bbb R}^{d+1}$. A collection of bases $T=\{I\}$ is called a triangulation of $P$ if $\cup_{I\in T} \langle \bn{I} \rangle =P$ for simplices $ \langle \bn{I} \rangle ={\rm Conv.}\(\{0\}\cup\bn{I}\)$ $(I\in T)$, and $ \langle \bn{I_1} \rangle \cap \langle \bn{I_2} \rangle \;(I_1,I_2\in T)$ is a lower dimensional common face. \vspace{0.3cm}\noindent {\bf Note.} Since $(d+1)$-simplices in $P$ are in one-to-one correspondence to $d$-simplices in ${\Delta^*}$, we identify a triangulation of $T$ of $P$ with its corresponding triangulation of ${\Delta^*}$. We call a triangulation $T$ of $P$ is maximal if it corresponds to a maximal triangulation of ${\Delta^*}$(Def.2.3). \vspace{0.2cm} For a base $I$ and a point $\eta\in {\Bbb R}^{p+1}$, we consider a linear function $h_{I,\eta}$ on ${\Bbb R}^{p+1}$ by $h_{I,\eta}(\bn{i})=\eta_i \;(i\in I)$. We define ${\cal C}({\cal A},I)=\{ \eta\in {\Bbb R}^{p+1} \;\vert\; h_{I,\eta}\;(\bn{i}) \leq \eta_i \;(i\not\in I)\}$ and ${\cal C}({\cal A},T):=\cap_{I\in T}{\cal C}({\cal A},I)$ for a triangulation $T$. Then it is easy to see that ${\cal C}({\cal A},T)$ consists of $\eta \in {\Bbb R}^{p+1}$ for which we have a convex piecewise linear function $h_{T,\eta}$ on $T$ determined by $h_{T,\eta}\vert_{ \langle \bn{I} \rangle }= h_{I,\eta} \; (I\in T)$ and satisfies $h_{T,\eta}(\bn{i})\leq \eta_i$ for $\bn{i}$ not a vertex of $T$, i.e., \begin{equation} \begin{array}{crl} {\cal C}({\cal A},T)=\{ \eta \in {\Bbb R}^{p+1}\; \vert \;\; && \makebox[-2em]{} h_{I_1,\eta}(v)\leq h_{I_2,\eta}(v) \;\; (v\in \langle \bn{I_2} \rangle , \; I_1,I_2\in T), \\ && \makebox[-2em]{} h_{T,\eta}(\bn{i})\leq \eta_i \;\; (\bn{i} \;\,{\rm is} \;\, {\rm not}\;\, {\rm a} \;\, {\rm vertex}\;\,{\rm of}\; T)\; \}\\ \end{array} \label{eqn:convex} \end{equation} A triangulation is called {\it regular} if ${\cal C}({\cal A},T)$ has interior points. We say a ${\bf Z}$-basis $A\subset L$ is compatible with a triangulation $T$ if it is compatible with all bases $I$ in $T$. \vspace{0.3cm}\noindent {\bf Proposition 3.7.} (Prop.5 in \cite{GKZ1}) {\it For every regular triangulation $T$, there exists (infinitely many) ${\bf Z}$-basis of $L$ compatible with $T$. } \vspace{0.2cm} The exponent $\beta$ is called {\it $T$-nonresonant} if the set $\Phi^A_{\bf Z}(\beta,I)\; (I\in T)$ are pairwise disjoint. We set $\Phi^A_{\bf Z}(\beta,T):=\cup_{I\in T}\Phi^A_{\bf Z}(\beta,I)$. We normalize the volume of the standard $(d+1)$-simplex to $1$. \vspace{0.3cm}\noindent {\bf Theorem 3.8.} (Th.3 in \cite{GKZ1}) {\it For a regular triangulation $T$ of the polytope $P$, and a ${\bf Z}$-basis $A=\{\l{1},\cdots,\l{p-d}\}$ of $L$ compatible with $T$, there are integral powerseries in the variables $x_k=a^{\l{k}}$ for $a^{-\gamma}\Pi(a,\gamma) \;(\gamma\in \Phi^A_{\bf Z}(\beta,T))$, which converge for sufficiently small $\vert x_k\vert$. If the exponents $\beta$ is $T$-nonresonant, these series constitute $vol(P)$ linearly independent solutions. } \vspace{0.3cm}\noindent {\bf Remark.} In our case of ${\Delta^*}$-hypergeometric system with $\beta=(-1)\times\vec0$, we have one special element $\gamma=(-1,0,\cdots,0)$ in the set $\Phi_{\bf Z}(\beta,I)$ for any base $I$. If the polytope ${\Delta^*}$ is of type I or II in our classification, a nonsingular maximal triangulation $T$ of $P$ consists of those bases $I$ for which $\vert \det (\bar\nu_{j,i}^*)_{1\leq i\leq d+1,\; j\in I}\vert =1$. Because of this unimodularity, we have \begin{equation} \Phi_{\bf Z}(\beta,I)=(-1,0,\cdots,0)+L \;\; , \label{eqn: phiL} \end{equation} and $\Phi_{\bf Z}^A(\beta,I)=\{ (-1,0,\cdots,0)\}$ for every base $I$ of the maximal triangulation and any ${\bf Z}$-basis $A$ compatible with it. Thus we encounter a ``maximally $T$-resonant'' situation. \vspace{0.3cm}\noindent {\bf Definition 3.9.} For a regular triangulation $T$ and a ${\bf Z}$-basis $A=\{ \l{1},\cdots,$ $\l{p-d}\}$ compatible with $T$, we define a power series $w_0(x,\rho; A)=a_0\Pi(a,\gamma)$ (with $\gamma=\sum_{k=1}^{p-d}\rho_k\l{k}+(-1,0,\cdots,0)$) by \begin{equation} w_0(x,\rho;A)=\sum_{m\in {\bf Z}_{\geq0}^{p-d} } { \Gamma(-\sum_k (m_k+\rho_k)\l{k}_0+1) \over \prod_{1\leq i\leq p} \Gamma(\sum_k(m_k+\rho_k)\l{k}_i+1) } x^{m+\rho} \;\;, \label{eqn: wnot} \end{equation} where $x_k=(-1)^{\l{k}_0}a^{\l{k}}$. \vspace{0.3cm}\noindent {\bf Remark.} Here we have applied our recepie of multiplying the constant $\Gamma(\gamma_0+1)$ to the formal solution $\Pi(a,\gamma)$. We adopt this definition because for a maximall triangulation $T_o$, we encounter the situation $\Pi(a,\gamma)\equiv 0$, namely, $\sum m_k \l{k}_0 +\gamma_0+1 \in {\bf Z}_{\leq0} \; (m\in {\bf Z}^{p-n}_{\geq0})$ for a ${\bf Z}$-basis $A=\{\l{1},\cdots,\l{p-n}\}$ compatible with $T_o$. (See Prop.4.8, Prop.4.9 and Th.4.10 in \cite{HLY2}). In general a basis $A$ compatible with a regular triangulation $T$ contains both the bases vectors $\l{k}$ with positive 0-th component and nonpositive 0-th component. Taking a value $\rho_0 \in {\bf Z}^{p-n}_{\geq0}$ under this situation cause infinity for some $m$ in the numerator of the coefficients of (\ref{eqn: wnot}). In this case we understand in our definition (\ref{eqn: wnot}) to take a limit $\rho \rightarrow \rho_0$ in a generic way. (If we encounter infinities under this limit in some coefficient, we go back to the original definition (\ref{eqn: fsol}). For a maximall triangulation $T_o$, we observe that this limit exists for all coefficients in (\ref{eqn: wnot}), see Claim5.8.) \newsubsec{(3-3) Secondary fan} It is known that the set ${\cal C}({\cal A},T)$ is a closed polyhedral cone and that these cones cover ${\Bbb R}^{p+1}$ when we vary the triangulations. Thus the set of these cones and their lower dimensional faces all together define a complete, polyhedral fan ${\cal F}({\cal A})$ called the {\it secondary fan}\cite{BFS}\cite{OP}. Let $\overline M = {\bf Z} \times M$ and $\overline N = {\bf Z} \times N$. We extend naturally the pairing $ \langle , \rangle $ to that of $\overline M$ and $\overline N$. Consider the lattice $\overline {\cal M}:= \oplus_{\ns{} \in {\Delta^*}\cap N} {\bf Z} \en{}\;(=\oplus_{i=0}^p {\bf Z}\en{i})$ and its dual $\overline {\cal N}={\rm Hom}\,_{\bf Z}(\overline {\cal M},{\bf Z})$. Then we have the following exact sequences \begin{equation} \begin{array}{crl} && \makebox[2.5cm]{} 0 \smash{\mathop{\longrightarrow}\limits^{}} \; \overline M \; \smash{\mathop{\longrightarrow}\limits^{\quad {\bf A}\quad}} \; \overline{\cal M} \; \smash{\mathop{\longrightarrow}\limits^{\quad {\bf B}\quad }} \; \Xi(\overline M) \; \smash{\mathop{\longrightarrow}\limits^{}} 0 \;\;, \\ && 0 \smash{\mathop{\longleftarrow}\limits^{}} \; {\rm Coker}{\bf A}^* \; \smash{\mathop{\longleftarrow}\limits^{}} \; \overline N \; \smash{\mathop{\longleftarrow}\limits^{\quad {\bf A}^* \quad}} \; \overline{\cal N} \; \smash{\mathop{\longleftarrow}\limits^{\quad {\bf B}^* \quad}} \; \Xi(\overline N) \; \smash{\mathop{\longleftarrow}\limits^{}} 0 \;\;, \\ \end{array} \label{eqn: exacts} \end{equation} where ${\bf A}(\bar m)=\sum_{i=0}^p \langle \bar m,\bn{i} \rangle \en{i} \;\; (\bar m\in \overline M)$ and ${\bf B}$ is the quotient. The dual is given by ${\bf A}^*(\mu)=\sum_{i=0}^p \mu(\en{i})\bn{i} \;\; (\mu\in \overline N)$. The pair $\{ {\cal B}, \Xi(\overline M) \}$ with ${\cal B}:=\{ {\bf B}(\en{0}), \cdots, {\bf B}(\en{p}) \}$ is called {\it Gale transform} of a pair $\{ {\cal A}, $ $\overline N \}$. Under this general setting, let us consider a polyhedral cone in $\Xi(\overline M)_{\Bbb R}$ \begin{equation} {\cal C}'({\cal A},T)=\cap_{I\in T} \( \sum_{i\not\in I} {\Bbb R}_{\geq0}{\bf B}(\en{i}) \)\;\;. \end{equation} \vspace{0.3cm}\noindent {\bf Proposition 3.10.}(Lemma 4.2 in \cite{BFS}) {\it The map ${\bf B}$ induces the following decomposition \begin{equation} {\cal C}({\cal A},T)={\rm Ker}({\bf B})\oplus {\cal C}'({\cal A},T) \;\;. \label{eqn:decomp} \end{equation} } \vspace{0.3cm} By definition, the cone ${\cal C}'({\cal A},T)$ is strongly convex. Using the above decomposition, we redefine the secondary fan to be \begin{equation} {\cal F}({\cal A})=\{ {\cal C}'({\cal A},T) \vert \; T: {\rm regular} \;\; {\rm triangulation} \} . \end{equation} Now the secondary fan consists of strongly convex, polyhedral cones. If the polytope ${\Delta^*}$ is of type I or II, then the quotient $\Xi(\overline M)$ is torsion free and thus $({\cal F}({\cal A}),\Xi(\overline M))$ defines a toric variety. Even in the case of type III, we may consider the corresponding toric variety by simply projecting out the torsion part of $\Xi(\overline M)$. We will use this toric variety for the compactification of the moduli space in the next section. Now let us note that $\Xi(\overline N)\cong {\rm Ker}({\bf A}^*)$ is identified with the lattice $L$, and thus ${\cal K}({\cal A},T) \subset \Xi(\overline N)_{\Bbb R}$. By definition of ${\cal C}'({\cal A},T)$, we may deduce \begin{equation} {\cal K}({\cal A},T)={\cal C}'({\cal A},T)^\vee \;\;. \label{eqn: Cdual} \end{equation} Since for a regular triangulation $T$, ${\cal C}'({\cal A},T)$ is a strongly convex polyhedral cone with interior points, the dual cone ${\cal K}({\cal A},T)$ is also strongly convex polyhedral cone. Therefore we see that there are infinitely many ${\bf Z}$-basis of the lattice $L$ compatible with $T$ (Prop.3.7). \section{4. Toric Ideal and Gr\"obner Fan} In this section we will reduce the infinite set of operators ${\cal D}_l\; (l\in L)$ in our ${\Delta^*}$-hypergeometric system to a finite set. This will be related to the compactification problem of the moduli spaces. \newsubsec{(4-1) Toric ideal and Gr\"obner fan} We write the operators ${\cal D}_l$ in (6) simply by ${\cal D}_l=({\partial\;\over\partial a})^{l_+} - ({\partial\;\over\partial a})^{l_-}$ with $l=l_+-l_-$. Keeping this form in mind we define {\it toric ideal} in a polynomial ring: \vspace{0.3cm}\noindent {\bf Definition 4.1.} Consider a polynomial ring ${\Bbb C}[y]:={\Bbb C}[y_0,\cdots,y_p]$. Toric ideal ${\cal I}_{\cal A}$ is defined to be an ideal generated by {\it binomials} $y^{l_+}-y^{l_-} \; (l\in L)$, \begin{equation} {\cal I}_{\cal A}= \langle \; y^{l_+}-y^{l_-} \;\vert\; l\in L \; \rangle \;. \end{equation} \vspace{0.3cm} Toric ideal has been extensively studied in ref.\cite{Stu1}\cite{GKZ2}. Here we summarize relevant results for our purpose. A {\it term order} (monomial ordering) on ${\Bbb C}[y]$ is a total order $\prec$ on the set of monomials $\{ y^\alpha \;\vert\; \alpha\in {\bf Z}_{\geq0}^{p+1}\}$ satisfying, 1) $y^\alpha \prec y^\beta$ implies $y^{\alpha+\gamma} \prec y^{\beta+\gamma}$ and 2) $1$ is the unique minimal element. When we have an ideal ${\cal I} \subset {\Bbb C}[y]$ and fix a term order, we can speak of the leading term $LT_\prec(f)$ for every non-zero polynomial in ${\cal I}$. Then we define {\it initial ideal} of ${\cal I}$ by \begin{equation} \langle LT_\prec ({\cal I}) \rangle = \langle LT_\prec (f) \;\vert\; f\in {\cal I}, f\not=0 \rangle \;. \end{equation} A finite set ${\cal G}\subset {\cal I}$ is called {\it Gr\"obner basis} with respect to a term order $\prec$ if it generates the initial ideal; \begin{equation} \langle LT_\prec ({\cal I}) \rangle = \langle LT_\prec(g) \;\vert\; g \in {\cal G} \rangle \;. \label{eqn: groebner} \end{equation} \vspace{0.3cm}\noindent {\bf Theorem 4.2.} (Th.1.2 in \cite{Stu2}) {\it For every ideal ${\cal I}\subset {\Bbb C}[y]$, there are only finitely many distinct initial ideals. } \vspace{0.3cm} We consider representing the term orders by weight vectors ${\omega} = (w_0,\cdots,$ ${\omega}_p)\in {\Bbb R}^{p+1}$. For a polynomial $f=\sum_\alpha c_\alpha y^\alpha$, we define its {\it leading terms} $LT_{\omega}(f)$ to be a sum of terms $c_\alpha y^\alpha$ whose weight $t_{\omega}(y^\alpha) :={\omega}_0\alpha_0+\cdots+{\omega}_p\alpha_p$ is maximal. It is obvious that if the components of ${\omega} \in {\Bbb R}_{\geq0}^{p+1}$ are rationally independent, the weight determines a term order on ${\Bbb C}[y]$. When we fix an ideal ${\cal I}\subset {\Bbb C}[y]$, we may relax the condition for the weight ${\omega}$ to be a term order; we say a weight ${\omega}\in {\Bbb R}^{p+1}$ defines a term order of ${\cal I}$ if $ \langle LT_{\omega}({\cal I}) \rangle = \langle LT_\prec({\cal I}) \rangle $ for some term order $\prec$. The following proposition provides a 'converse' statement, \vspace{0.3cm}\noindent {\bf Proposition 4.3.} (Prop.1.11 in \cite{Stu2}) {\it For any term order $\prec$, there exists a weight ${\omega}\in {\Bbb R}_{\geq0}^{p+1}$ such that $ \langle LT_w({\cal I}) \rangle = \langle LT_\prec({\cal I}) \rangle $. } \vspace{0.3cm} Now {\it Gr\"obner region} is defined to be a set \begin{equation} GR({\cal I})=\{ {\omega}\in {\Bbb R}^{p+1} \vert \langle LT_{\omega}({\cal I}) \rangle = \langle LT_{{\omega}'}({\cal I}) \rangle \; {\rm for} \;{\rm some} \; {\omega}'\in {\Bbb R}_{\geq0}^{p+1} \} \; . \end{equation} \vspace{0.3cm}\noindent {\bf Proposition 4.4.} (Prop.1.12 in \cite{Stu2}) {\it If an ideal ${\cal I}\in {\Bbb C}[y]$ is a homogeneous ideal with some grading $deg(y_i)=d_i >0$, then $GR({\cal I})={\Bbb R}^{p+1}$. } \vspace{0.3cm} Since the toric ideal ${\cal I}_{\cal A}$ is homogeneous ideal with $deg(y_0)=\cdots=$ $deg(y_p)$ $=1$, we see $GR({\cal I}_{\cal A})={\Bbb R}^{p+1}$. For a term order ${\omega}$ of ${\cal I}_{\cal A}$, we define \begin{equation} {\cal C}({\cal I}_{\cal A},{\omega})=\{ {\omega}' \in {\Bbb R}^{p+1} \;\vert\; \langle LT_{\omega}({\cal I}_{\cal A}) \rangle = \langle LT_{{\omega}'}({\cal I}_{\cal A}) \rangle \;\} \;. \label{eq:Cw} \end{equation} It is known that this set constitutes an open, convex, polyhedral cone in ${\Bbb R}^{p+1}$ (Prop.2.1 in \cite{Stu1}). In the following, we mean by ${\cal C}({\cal I}_{\cal A}, {\omega})$ the closure of the set (\ref{eq:Cw}). Then due to Th.4.2 and Prop.4.4, the collection $\{ {\cal C}({\cal I}_{\cal A}, {\omega}) \}$ is finite and defines a complete polyhedral fan ${\cal F}({\cal I}_{\cal A})$ in ${\Bbb R}^{p+1}$, called the {\it Gr\"obner fan}. \newsubsec{(4-2) Indicial ideal and compactification of ${\rm Hom}\,_{\bf Z}(L,{\Bbb C}^*)$ } In the previous section, we called a triangulation $T$ of the polytope $P$ regular if the cone ${\cal C}({\cal A},T)$ has interior points. Here we characterize the regular triangulation in a geometrical way. To this aim let us first consider a polytope $P_{\omega}:={\rm Conv.}\(\{ {\omega}_0\times\ns{0},\cdots, {\omega}_p\times\ns{p}\} \)$ in ${\Bbb R}^{d+1}$ for a weight ${\omega} \in {\Bbb R}^{p+1}$. If we project a polytope $P_{\omega}$ to $1\times {\Bbb R}^d$, then we have the polytope $1\times {\Delta^*}$. Thus we may regard the weight ${\omega}$ giving a hight to each vertex of $1\times{\Delta^*}$. For generic weight ${\omega}$, the 'lower' faces of the polytope $P_{\omega}$ consist of simplices and define, under the projection, a simplicial decomposition of ${\Delta^*}$ and thus induce a triangulation $T_{\omega}$. The regular triangulation of the polytope $P$ is a triangulation $T_{\omega}$ obtained for some weight ${\omega}$ in this way (see Def.5.3 of \cite{Zie} for more details). It is not difficult to see the relation of the polytope $P_{\omega}$ to the piecewise linear function $h_{T,\eta}$ in (\ref{eqn:convex}) with $\eta={\omega}$. Given a (regular) triangulation $T$ of the polytope $P$, the {\it Stanley-Reisner ideal} $SR_T$ in ${\Bbb C}[y]$ is defined to be the ideal generated by all monomials $y_{i_1}\cdots y_{i_k}$ for which the vertices $\bn{i_1},\cdots,\bn{i_k}$ do not make a simplex in $T$. The following theorem is due to Sturmfels: \vspace{0.3cm}\noindent {\bf Theorem 4.5.} (Thm. 3.1 in \cite{Stu1}) {\it If a weight ${\omega}$ defines a term order of the toric ideal ${\cal I}_{\cal A}$, then it induces a regular triangulation $T_{\omega}$. Moreover the Stanley-Reisner ideal $SR_{T_{\omega}}$ is equal to the radical of the initial ideal $ \langle LT_{\omega}({\cal I}_{\cal A}) \rangle $. } \vspace{0.3cm} As an immediate corollary to this theorem, we see that the Gr\"obner fan is a refinement of the fan $\{ {\cal C}({\cal A},T_{\omega}) \}$. Since the cone ${\cal C}({\cal A},T_{\omega})$ decomposes according to (\ref{eqn:decomp}), we have similar decomposition of ${\cal C}({\cal I}_{\cal A},{\omega})$ to ${\cal C}'({\cal I}_{\cal A},{\omega})$. In the following we call the collection $\{{\cal C}'({\cal I}_{\cal A},{\omega})\}$ as the Gr\"obner fan ${\cal F}({\cal I}_{\cal A})$. Now we determine a finite set of operators ${\cal D}_l$ which characterize the power series $w_0(x,\rho;A)$ for each regular triangulation and a ${\bf Z}$-basis $A$ compatible with it. This provides us a way to analyze our resonant GKZ hypergeometric system. Let us consider a term order ${\omega}$ of ${\cal I}_{\cal A}$. According to Th.4.5, the term oder ${\omega}$ determines a regular triangulation $T_{\omega}$ and also a cone ${\cal C}'({\cal I}_{\cal A},{\omega}) \subset {\cal C}'({\cal A},T_{\omega})$. If the cone ${\cal C}'({\cal I}_{\cal A},{\omega})$ is simplicial and regular, i.e., the integral generators of its one-dimensional boundary cones generate the lattice points ${\cal C}'({\cal I}_{\cal A},{\omega}) \cap \Xi(\overline M)$, we simply make its dual cone ${\cal C}'({\cal I}_{\cal A},{\omega})^\vee$ and take the integral generators of this cone as a canonical ${\bf Z}$-basis $A$ of $L$ which is compatible with $T_{\omega}$. If not, we subdivide the cone ${\cal C}'({\cal I}_{\cal A},{\omega})$ into simplicial, regular cones and reduce the problem to the former case. More generally, we may take a ${\bf Z}$-basis $A_\tau=\{ \l{1}_\tau,\cdots,\l{p-d}_\tau \}$ of $L$ compatible with $T_{\omega}$ considering any simplicial, regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$ and making its dual $\tau^\vee$. Associated to ${\omega}$, we have a Gr\"obner basis ${\cal B}_{\omega} \subset {\cal I}_{\cal A}$. By B\"uchberger's algorithms to construct the (reduced) Gr\"obner basis, we see that every generator $g \in {\cal B}_{\omega}$ is a binomial of the form $y^{l_+}-y^{l_-}$ with some $l \in L$. In the following, we assume ${\cal B}_{\omega}$ to be the reduced Gr\"obner basis which is determined uniquely for a term order ${\omega}$ (, see Chapter 2 of \cite{CLO} for the properties of the reduced Gr\"obner basis). Translating this to differential operator, we write the Gr\"obner basis ${\cal B}_{\omega}=\{ {\cal D}_{l_1}, \cdots, {\cal D}_{l_s} \} \; (1\leq s < \infty)$. Now, for each generator, we define \begin{equation} J_l(\rho;A_\tau):= a_0 x_\tau^{-\rho} a^{l_\pm}\(\da{}\)^{l_\pm} a_0^{-1}x_\tau^{\rho} \;\;, \label{eqn: Jl} \end{equation} where the choice in $l_\pm$ is made respectively by ${\omega}\cdot l_+ - {\omega}\cdot l_- >0 \;\; (<0) $. (The factor $a_0$ originate from the definition $w_0(x,\rho;A):= a_0 \Pi(a,\gamma)$ in Def.3.9.) \vspace{0.3cm}\noindent {\bf Definition 4.6.} For a term order ${\omega}$ of ${\cal I}_{\cal A}$ and an arbitrary regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$, we define, through the Gr\"obner basis ${\cal B}_{\omega}=\{{\cal D}_{l_1},\cdots, {\cal D}_{l_s}\}$, an {\it indicial ideal} in ${\Bbb C}[\rho_1,\cdots,\rho_{p-d}]$; \begin{equation} Ind_{\omega}(\tau)= \langle J_{l_1}(\rho,A_\tau),\cdots,J_{l_s}(\rho,A_\tau) \rangle \;\;. \end{equation} Similarly to the indicial equations of the differential equations of Fuchs type, we also consider the {\it indicial equations} for our ${\Delta^*}$-hypergeometric system as algebraic equations for $\rho$ coming from the leading terms of the operators $D_l \;(l\in L)$. (Note that the leading term of an operator $D_l$ varies in general when a term order ${\omega}$ varies. Here we consider for the indicial equations all possible leading terms when ${\omega}$ varies inside $\tau$.) \vspace{0.3cm}\noindent {\bf Proposition 4.7.} {\it In the notation above, the indicial ideal $Ind_{\omega}(\tau)$ coincides with the ideal generated by the indicial equations for the indices $\rho$ of the powerseries $w_0(x_\tau,\rho;A_\tau)$. } \noindent {\bf (Proof)} Consider an operator ${\cal D}_l \in {\cal B}_{\omega}$. If ${\omega}\cdot l_+ - {\omega}\cdot l_- >0$, we multiply $a^{l_+}$ to obtain \begin{equation} a^{l_+}{\cal D}_l =a^{l_+}\(\da{}\)^{l_+} -a^{l_+-l_-}a^{l_-}\(\da{}\)^{l_-} \; . \end{equation} Since the initial ideal $ \langle LT_{\omega}({\cal I}_{\cal A}) \rangle $ and thus the reduced Gr\"obner basis ${\cal B}_{\omega}$ does not change for ${\omega}$ in the interior of $\tau$, $Int(\tau)$, we have ${\omega}\cdot(l_+-l_-)>0$ for all ${\omega}\in Int(\tau)$, i.e., $l_+-l_-\in \tau^\vee\cap L$. Since we have chosen the ${\bf Z}$-basis $A_\tau=\{ \l{1}_\tau,\cdots,\l{p-d}_\tau\}$ so that it generates all integral points in $\tau^\vee \cap L$, $a^{l_+-l_-}$ is a monomial of $x_\tau$, which vanish in the limit $x_\tau \rightarrow 0$. The same argument applies to the case ${\omega}\cdot l_+ - {\omega}\cdot l_- <0$. Therefore the indicial equations arising from the operators $D_l \in {\cal B}_{\omega}$ exactly coincide with the generators of the indicial ideal (\ref{eqn: Jl}). For general operators $D_l \; (l\in L)$ , depending on the weight ${\omega} \in Int(\tau)$, we have two possible leading terms. However for both of them, owing to the defining property of the Gr\"obner basis, we have $LT_{\omega}(D_l)=\(\da{}\)^\mu LT_{\omega}(D_{l_k})$ for some $k$ and $\mu$. Multiplying a monomial $a^{\mu+l_{k\pm}}$, we obtain \begin{equation} a_0x_\tau^{-\rho} a^{\mu+l_{k\pm}} LT_{\omega}(D_l) a_0^{-1} x_\tau^{\rho} =F(\rho) \, J_{l_k}(\rho;A_\tau) \;\;, \end{equation} with some polynomial $F(\rho)$. Thus we see all polynomial relations of $\rho$ comming from the leading terms are in $Ind_{\omega}(\tau)$. \par Conversely, since all generators of the ideal $Ind_{\omega}(\tau)$ give the indicial equations related to ${\cal B}_{\omega}$, the ideal $Ind_{\omega}(\tau)$ is contained in the other. Therefore the two ideals are the same. \hfill $\Box$ \vspace{0.3cm} Now based on Prop.4.7, we may claim the following; \vspace{0.3cm}\noindent {\bf Proposition 4.8.} {\it Consider a compact toric variety ${\Bbb P}_{{\cal F}({\cal I}_{\cal A})}$ associated to the Gr\"obner fan $({\cal F}({\cal I}_{\cal A}), \Xi(\overline M))$. Then for any resolution ${\Bbb P}_{\tilde{\cal F}({\cal I}_{\cal A})} \rightarrow {\Bbb P}_{{\cal F}({\cal I}_{\cal A})}$ associated to a refinement $(\tilde{\cal F}({\cal I}_{\cal A}), \Xi(\overline M)) \rightarrow ({\cal F}({\cal I}_{\cal A}), \Xi(\overline M))$, we have integral powerseries of the form $w_0(x_\tau,\rho;A_\tau)$ $(\rho\in V(Ind_{\omega}(\tau)))$ at each boundary point given by the normal crossing toric divisors, namely at the origin of $Hom_{s.g.}(\tau^\vee\cap L,{\Bbb C})$. We will call this compactification Gr\"obner compactification. } \vspace{0.3cm}\noindent {\bf Remark.} Since Prop.4.7 provides us only a necessary condition for the indices $\rho$ to give a powerseries solution $w_0(x,\rho;A_\tau)$, we do not claim by Prop.4.8, although we expect, that all $\rho\in V(Ind_{\omega}(\tau))$ form the powerseries solutions of our ${\Delta^*}$-hypergeometric system. \newsubsec{(4-3) Resonance of ${\Delta^*}$-hypergeometric system} When the polytope ${\Delta^*}$ is of type I or II, we have seen in the Remark right after Th.3.8 that the ${\Delta^*}$-hypergeometric system becomes ``maximally resonant'' for a maximal triangulation $T_o$. Here we study this resonance in detail restricting our attention to the polytopes of type I or II. We also comment about the case of type III. We call a collection of vertices ${\cal P}=\{\bn{i_1},\cdots,\bn{i_a}\}$ {\it primitive} if ${\cal P}$ does not form a simplex in $T_o$ but ${\cal P}\setminus \{ \bn{i_s}\}$ does for any $\bn{i_s}\in {\cal P}$. By definition of the Stanley-Reisner ideal, it is easy to deduce that the monomials that corresponds to primitive collections generate the ideal $SR_{T_o}$. Let us denote by $\Sigma(1\times{\Delta^*},T_o)$ the fan in $\overline N_{\Bbb R}$ that is naturally associated to the triangulation $T_o$ of $P$. Since the volumes of all $d+1$ simplices in $T_o$ are unimodular for the polytope ${\Delta^*}$ of type I or II, the fan $\Sigma(1\times{\Delta^*},T_o)$ consists of regular cones. Therefore if we have a primitive collection ${\cal P}=\{\bn{i_1},\cdots,\bn{i_a}\}$, we obtain \begin{equation} \bn{i_1}+\cdots+\bn{i_a}=\sum_k c_k \bn{j_k} \;\; (c_k \in {\bf Z}_{\geq0}) \label{eqn: prim} \end{equation} where $\{\bn{j_k} \vert c_k\not= 0\}$ generates a cone that contains the vector in the left hand side. Writing (\ref{eqn: prim}) as $\bn{i_1}+\cdots+\bn{i_a}-\sum c_k \bn{j_k}=0$, we read the corresponding {\it primitive relation} $l({\cal P}) \in L$. \vspace{0.3cm}\noindent {\bf Lemma 4.9.} {\it Every primitive collection of a maximal triangulation $T_o$ does not contain the point $\bn{0}=1\times \vec 0$. } \par \noindent {\bf (Proof)} Suppose a primitive collection is given by ${\cal P}=\{ \bn{0}, \bn{i_1}, \cdots,\bn{i_a} \} $ $(1\leq i_1,\cdots,i_a \leq p)$. Since it is primitive, the simplex $ \langle \bn{i_1},\cdots,\bn{i_a} \rangle $ must be a simplex in the triangulation $T_o$, which means that this simplex is a face of some maximal dimensional simplex in $T_o$. Since $T_o$ is a maximal triangulation in which every maximal dimensional simplex contains the vertex $\bn{0}$, we see the simplex $ \langle \bn{0},\bn{i_1},\cdots,\bn{i_a} \rangle $ must be a simplex in $T_o$, which is a contradiction. \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Proposition 4.10.} {\it For a term order ${\omega}$ such that $T_{\omega}$ is a maximal triangulation, the initial ideal $ \langle LT_{\omega}({\cal I}_{\cal A}) \rangle $ is radical and $ \langle LT_{\omega}({\cal I}_{\cal A}) \rangle =SR_{T_{\omega}}$. } \par\noindent {\bf (Proof)} Consider the primitive collections for the triangulation $T_{\omega}$, which generate the Stanley-Reisner ideal $SR_{T_{\omega}}$. Write a primitive collection ${\cal P}=\{ \bn{i_1},\cdots,\bn{i_a}\}$ and the corresponding primitive relation as $l({\cal P})$ considering the relation (\ref{eqn: prim}). For a term order ${\omega}$, the regular triangulation $T_{\omega}$ is induced from the lower faces of the polytope $P_{\omega}={\rm Conv.}(\,\{\tilde\ns{0}, \cdots ,\tilde\ns{p} \;\vert\; \tilde\ns{k}=$ ${\omega}_k\times\ns{k}\;(k=0,\cdots,p)\}\,)$. Then the convex hull ${\rm Conv.}(\{\tilde\ns{i}\;\vert\; \bn{i}\in{\cal P}\})$ is not a simplex that corresponds to a lower face of $P_{\omega}$. Therefore we have a ``height'' inequality $( \tilde\ns{i_1}+\cdots+\tilde\ns{i_a} )_1 > (\sum c_k \tilde\ns{j_k})_1$, namely, \begin{equation} {\omega}_{i_1}+\cdots+{\omega}_{i_a} > \sum_k c_k {\omega}_{j_k} \;\;. \label{eqn: height} \end{equation} This means that $LT_{\omega}(y^{l({\cal P})_+}-y^{l({\cal P})_-})= y_{i_1}\cdots y_{i_a}$, which is one of the generators of the ideal $SR_{T_{\omega}}$. Since this argument applies to all primitive collections, we conclude $SR_{T_{\omega}}\subset \langle LT_{\omega}({\cal I}_{\cal A}) \rangle $. Since the opposite inclusion follows from $SR_{T_{\omega}}=\sqrt{ \langle LT_w({\cal I}_{\cal A}) \rangle }$ (Th.4.5), we conclude $SR_{T_{\omega}}= \langle LT_w({\cal I}_{\cal A}) \rangle $, which proves the initial ideal is radical. \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Corollary 4.11.} {\it Under the hypothesis in the previous proposition, the set of all possible primitive collections $\{ {\cal P}_1,\cdots,{\cal P}_s \}$ of $T_{\omega}$ determines the Gr\"obner basis by ${\cal B}_{\omega}=\{ {\cal D}_{l({\cal P}_1)}, \cdots, {\cal D}_{l({\cal P}_s)} \}$. And the indicial ideal $Ind_{\omega}(\tau)$ is homogeneous for an arbitrary regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$. } \par\noindent {\bf (Proof)} By definition, $SR_{T_{\omega}}$ is generated by the monomials corresponding to primitive collections. From the argument in the proof of Prop.4.10, we know $SR_{T_{\omega}}= \langle LT_{\omega}( {\cal D}_{l({\cal P}_1)}),$ $\cdots, LT_{\omega}( {\cal D}_{l({\cal P}_s)}) \rangle $. This combined with $ \langle LT_{\omega}({\cal I}_{\cal A}) \rangle =SR_{T_{\omega}}$ establishes that ${\cal B}_{\omega}$ is the Gr\"obner basis. {}For the rest, we note that any primitive collection ${\cal P}= \{ \bn{i_1},\cdots,\bn{i_a}\}$ does not contain $\bn0$ according to Lemma 4.9. Now we have \begin{equation} J_{l({\cal P})}(\rho;A_\tau)=a_0x_\tau^{-\rho} a^{l({\cal P})_+} \(\da{}\)^{l({\cal P})_+} a_0^{-1}x_\tau^\rho = x_\tau^{-\rho} \ta{{i_1}}\cdots\ta{{i_a}} x_\tau^\rho , \end{equation} where $ \ta{{}} =a \da{} $, which proves that the generator $J_{l({\cal P})}(\rho;A_\tau)$ is homogeneous in $\rho$. \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Remark.} If we combine a general result that the GKZ system is holonomic \cite{GKZ1}, i.e., its solution space is finite dimensional, with our Corol. 4.11, we may conclude that the zero is the only solution for the indices $\rho$. This is the maximal $T$-resonance in our approach. We will give following \cite{HLY2} an independent proof about this in the next section. As we remarked before, our ${\Delta^*}$-hypergeometric system for the polytope ${\Delta^*}$ of type III does not share this property. Here we can explain the difference. We first note that the primitive collections generate the Stanley-Reisner ideal and has the property in Lemma 4.9 irrespective to the type of polytopes. The only change in the above arguments is in the definition of the primitive relation. Namely, since all cones are not regular in type III case, for some primitive collection the equation (\ref{eqn: prim}) should be replaced by \begin{equation} \lambda_{i_1}\bn{i_1}+ \cdots + \lambda_{i_a} \bn{i_a} = \sum_k c_k \bn{j_k} \;\;, \label{eqn: sing} \end{equation} with some positive integers $\lambda_{i_1},\cdots,\lambda_{i_a}$ not all equal to one. Accordingly the leading term $LT_{\omega}(y^{l({\cal P})_+}-y^{l({\cal P})_-})$ will be replaced by $(y_{i_1})^{\lambda_{i_1}}\cdots (y_{i_a})^{\lambda_{i_a}}$. This indicates that the initial ideal $LT_{\omega}({\cal I}_{\cal A})$ is no longer radical, and therefore the generators $J_l(\rho;A_\tau)$ become inhomogeneous. When translating the monomial $y^{l({\cal P})_+}$ to the differential operator $a^{l({\cal P})_+} \( \da{} \)^{l({\cal P})_+}$, each $\lambda_i$-fold degeneration to zero 'splits' to simple zeros. Thus every index does not degenerate to zero, although we still have a simple zero. \section{5. Large compex Structure Limit} Here we will study in detail the maximal resonance of the ${\Delta^*}$-hypergeo-metric system. We will identify this resonance with the large complex structure limit (LCSL), i.e., a celebrated boundary point in the moduli space of Calabi-Yau manifolds\cite{Mor}. \newsubsec{(5-1) Maximal degeneration} In this subsection, we will restrict our arguments to the polytopes of type I or II. In these two cases, we have a nonsingular projective toric variety ${\Bbb P}_{\Sigma( {\Delta^*},T_o)}$ for a maximal triangulation. We focus on the Chow ring of this toric variety. The Chow ring of a variety is a free abelian group generated by irreducible closed subvarieties, modulo rational equivalence, which is endowed with the ring structure via the intersection products. In case of non-singular compact toric varieties ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$, it has a simple description in terms of the (toric) divisors; \vspace{0.3cm}\noindent {\bf Proposition 5.1.} (sect.3.3 of \cite{Oda}, sect.5.2 of \cite{Ful}) {\it The Chow ring $A^*({\Bbb P}_{\Sigma({\Delta^*},T_o)})$ is isomorphic to the cohomology ring $H^{2*}({\Bbb P}_{\Sigma({\Delta^*},T_o)},{\bf Z})$ and is given by \begin{equation} A^*({\Bbb P}_{\Sigma({\Delta^*},T_o)})={\bf Z}[D_0,\cdots,D_p]/(SR_{T_o}+\bar R), \end{equation} where $D_k \; (k>0)$ represents the toric-divisor determined by the integral point $\ns{k}$. $SR_{T_o}$ is the Stanley-Reisner ideal and $\bar R$ is the ideal generated by linear relations $\sum_{k=0}^p \langle 1\times u, \bn{k} \rangle D_{k} =0 \;\;(u\in M)\;. $ } \vspace{0.3cm}\noindent {\bf Note.} Owing to lemma 4.9, we can take the generators of $SR_{T_o}$ that do not contain $D_0$. Therefore the generator $D_0$ plays only a dummy role, although it makes sense as a divisor if we consider a toric variety defined by the (non-complete) fan $\Sigma(1\times{\Delta^*},T_o) \subset \overline N_{\Bbb R}$. \vspace{0.3cm} Now consider a term order ${\omega}$ of the toric ideal ${\cal I}_{\cal A}$ and denote the Gr\"obner basis by ${\cal B}_{\omega}=\{ {\cal D}_{l_1},\cdots,{\cal D}_{l_s}\}$. We define \begin{equation} I_l(\ta{{}}):=a_0 a^{l_\pm}\(\da{{}} \)^{l_\pm} a_0^{-1} \end{equation} for each $LT_{\omega}({\cal D}_l)=(\da{{}})^{l_\pm}$ in a similar way to $J_l(\rho;A_\tau)$. Obviously these two are related by $J_l(\rho;A_\tau)=x_\tau^{-\rho}I_l(\ta{{}} )x_\tau^{\rho}$. We consider the following ideals in ${\Bbb C}[\ta0,\cdots,\ta{p}]$, \begin{equation} I_{\omega}:= \langle I_{l_1}(\ta{{}} ),\cdots, I_{l_s}(\ta{{}} ) \rangle \;,\; \bar R_a:= \langle \sum_{i=0}^p \langle 1\times u, \bn{i} \rangle \ta{i} \;\vert\; u\in M \rangle \;\;. \end{equation} \vspace{0.3cm}\noindent {\bf Proposition 5.2.} {\it For a term order ${\omega}$ of ${\cal I}_{\cal A}$ and an arbitrary regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$, we have \begin{equation} {\Bbb C}[\rho]/Ind_{\omega}(\tau) \cong {\Bbb C}[\ta{{}}]/(I_{\omega}+\bar R_a) \;\;. \end{equation} } \par \noindent {\bf (Proof)} When we take the ${\bf Z}$-basis $A_\tau=\{ \l1_\tau, \cdots,\l{p-d}_\tau \}$, we have $\ta{i}=\sum_{k=1}^{p-d} (\l{k}_\tau)_i \theta_{x^{(k)}_\tau}$. Then the homomorphism $\phi : {\Bbb C}[\ta{{}}]\rightarrow {\Bbb C}[\theta_{x_\tau}]\cong{\Bbb C}[\rho]$ induced by this relation is surjective, since rank$(L)=p-d$, and satisfies ${\rm Ker}\, \phi = \bar R_a $ and $\phi(I_{\omega})= Ind_{\omega} (\tau)$. This proves the assertion. \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Proposition 5.3.} {\it Consider a term order ${\omega}$ of ${\cal I}_{\cal A}$ with $T_{\omega}$ a maximal triangulation $T_o$. Then for any regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$, the variety associated to the indicial ideal $Ind_{\omega}(\tau)$ consists only one point, i.e., \begin{equation} V(Ind_{\omega}(\tau))=\{0\} \;\;. \end{equation} } \par \noindent {\bf (Proof)} By Corol. 4.11, we know the indicial ideal is homogeneous for a term order ${\omega}$ of the given property. Moreover the ideal $I_{\omega}$ coincides with the Stanley-Reisner ideal $SR_{T_o}$. Therefore we have \begin{equation} {\Bbb C}[\rho]/Ind_{\omega}(\tau) \cong {\Bbb C}[\ta{{}}]/(I_{\omega}+\bar R_a) \cong A^*({\Bbb P}_{\Sigma({\Delta^*},T_o)})\otimes{\Bbb C} \;\;, \end{equation} which is finite dimensional. Since the ideal is homogeneous, the claim follows. \hfill $\Box$ We write our series (\ref{eqn: wnot}) for generic $\rho$ by $w_0(x,\rho;A)= \sum_{m\in {\bf Z}^{p-n}_{\geq 0}} c(m+\rho)x^{m+\rho}$. As remarked after Def.3.9, the value $\rho=0$ might cause the infinity in the numerator for some coefficient $c(m+\rho)$. One way to treat this infinity problem is to take the limit $\rho\rightarrow 0$ as discussed there and we will come back to this recepie in Claim 5.8. Here following \cite{HLY2}, we introduce the series \begin{equation} w_0(x_\tau,0;A_\tau)_{\geq0}:= \sum_{m\in{\bf Z}^{p-n}_{\geq0},\; -\sum_k m_k\l{k}_\tau\geq0} c(m) x^m \;\;. \label{eqn: wnotpositive} \end{equation} \vspace{0.3cm}\noindent {\bf Theorem 5.4.} (Th.5.2 in \cite{HLY2}) {\it For a term order ${\omega}$ with $T_{\omega}$ a maximal triangulation and any regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$, the series $w_0(x_\tau,0; A_\tau)_{\geq0}$ is the only powerseries solution of the ${\Delta^*}$-hypergeometric system about the origin of $U_\tau=Hom_{s.g.}(\tau^\vee\cap L,{\Bbb C})$. } \vspace{0.3cm} To prove this theorem, we prepare the following lemma; \vspace{0.3cm}\noindent {\bf Lemma 5.5.} {\it Consider a subset $S\not=\{\phi\}$ that is contained in $\tau^\vee\cap L$. There exist an element $\delta\in S$ and a simplicial, regular cone $C_\delta^\vee \subset L_{\Bbb R}$ such that $C_\delta^\vee$ contains both the subset $S-\delta$ and the cone $\tau^\vee$. } \par\noindent {\bf (Proof)} Consider a hyperplane $H(v;z_0)$ with a normal vector $v\in \tau$ and passing through a point $z_0$ in $\tau^\vee$. When we consider a parallel transport $H(v,t z_0)$ $(t\geq0)$ of the hyperplane, we may find the minimal $t_0$ such that $H(v, t_0 z_0)\cap S \not=\{\phi\}$ while $H(v,t z_0)\cap S =\{\phi\} \;\; (t <t_0)$. Changing the normal vector $v$ slightly, if necessary, we may assume the intersection $H(v, t_0 z_0)\cap S$ occurs at a point $\delta$. Now for this $\delta$, we see that the union $U:=(\tau^\vee\cap L) \cup (S-\delta) \setminus\{0\}$ is contained in the half space $H_>(v,0)$. Therefore the normal cone to the set $U$ at the origin is strongly convex, polyhedral cone. Since a strongly convex, polyhedral cone can be inside a simplicial, regular cone, the assertion follows. \hfill $\Box$ \noindent {\bf (Proof of Th.5.4.)} To prove the theorem, we write the series $w_0(x_\tau,0;$ $A_\tau)$ in terms of $a_0,\cdots,a_p$ by \begin{equation} w_0(a,0,\tau)=\sum_{l \in \tau^\vee\cap L} c_l a^l \;\; , \end{equation} with $c_0=1$. Now suppose we have two different series of this form. Then the difference of the two may be written by $r(a,0,S)=\sum_{l \in S} d_l a^l $ with a subset $S\subset \tau^\vee \cap L\setminus \{0\}$. Using the result in the lemma 5.5, we may write this series via nonzero $\delta$ as \begin{equation} r(a,0,C_\delta)=a^\delta \sum_{l \in C_\delta^\vee \cap L} d_{l+\delta}a^{l} \;\;, \label{eqn: delta} \end{equation} or $r(x_\tau,0,A_{C_\delta})=x_\tau^{\rho(\delta)} \sum_{n \in {\bf Z}^{p-d}_{\geq 0}} d(n) x_\tau^n $ with $\rho(\delta)\not=0 ,d(0)\not=0$ and $C_\delta \subset \tau$. This is a contradiction to Prop.5.3. \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Remark.} By direct evaluation of the period integral (\ref{eqn: Pi})\cite{Bat2}, we can verify that $a_0\Pi(a)$ exactly coincides with the powerseries in Th.5.4 when expressed in terms of the ${\bf Z}$-basis $A_\tau$ (Prop.5.15 \cite{HLY2}). \newsubsec{(5-2) All solutions about maximal degeneration points} Here we determine other solutions about maximal degeneration points, all of which contains logarithmic singularities. As in the previous subsection, our arguments are restricted to the polytopes of type I or II. Let us note that the first degree elements of the Chow ring, $A^1({\Bbb P}_{\Sigma({\Delta^*},T_o)})$, describe the Picard group of the toric variety ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ and may be expressed by \begin{equation} A^1({\Bbb P}_{\Sigma({\Delta^*},T_o)})={\bf Z} D_0\oplus\cdots\oplus{\bf Z} D_p / \bar R \;\cong\; \Xi(\overline M) \;\;. \end{equation} From this we see a dual pairing between the Picard group and the lattice $L\cong \Xi(\overline N)$; \begin{equation} A^1({\Bbb P}_{\Sigma({\Delta^*},T_o)})\times L \rightarrow {\bf Z} \;\;. \label{eqn: dualpL} \end{equation} \vspace{0.3cm}\noindent {\bf Definition 5.6.} For a ${\bf Z}$-basis $A_\tau=\{ \l{1}_\tau, \cdots, \l{p-d}_\tau \}$ of $L$ determined from a term order ${\omega}$ with $T_{\omega}$ equal to a maximal triangulation $T_o$, we denote its dual by $A_\tau^\vee=\{ J_{\tau,1},\cdots,J_{\tau,p-d} \}$ or simply by $\{ J_1\cdots,J_{p-d}\}$ when its dependence on $\tau$ is obvious. \vspace{0.3cm}\noindent {\bf Note.} By construction, the basis $A_\tau^\vee$ consists of the integral generators of the simplicial, regular cone $\tau$ contained in ${\cal C}'({\cal I}_{\cal A},{\omega})$ ($= {\cal C}'({\cal A},T_o)$ by Prop.4.10). ${\cal C}'({\cal A},T_o)$ consists of convex functions on $T_o$ which may be identified with the convex functions on the fan $\Sigma({\Delta^*},T_o)$. Since the set of all convex functions on the fan $\Sigma({\Delta^*},T_o)$ determines the closure of the K\"ahler cone of ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ (see Corol.2.15 in \cite{Oda}), the bases $J_{\tau,1},\cdots, J_{\tau,p-d}$ generate a simplicial, regular cone contained in this closure of the K\"ahler cone. \vspace{0.3cm}{\noindent} {\bf Definition 5.7.} For the powerseries $w_0(x_\tau,\rho;A_\tau)= \sum_{n\in{\bf Z}_{\geq0}^{p-d}} c(n+\rho) x_\tau^{n+\rho}$ in (\ref{eqn: wnot}), we define \begin{equation} w_0(x_\tau,J;A_\tau):= \sum_{n\in{\bf Z}_{\geq0}^{p-d}} c\(n+{J \over 2\pi i}\) x_\tau^{n+{J\over 2\pi i}} \;\;, \label{eqn: wJ} \end{equation} as the Taylor series expansion of $w_0(x_\tau,\rho;A_\tau)$ about $\rho=0$ followed by the substitution $\rho={J\over 2\pi i}$, where $J$'s are defined in Def.5.6. \vspace{0.3cm} In refs.\cite{HKTY1}\cite{HKTY2}\cite{HLY1}\cite{HLY2}, it is widely verified \vspace{0.3cm}\noindent {\bf Claim 5.8.} {\it The expansion (\ref{eqn: wJ}) exists as an element in $A^*({\Bbb P}_{\Sigma({\Delta^*},T_o)})\otimes {\Bbb C}\{x_\tau\}[\log x_\tau]$, and the coefficient series constitute a complete set of the local solutions about the maximal degeneration points. Especially the limit $w_0(x_\tau,\rho;A_\tau)\vert_{\rho\rightarrow0}$ coincides with $w_0(x_\tau,0; A_\tau)_{\geq0}$ in Th.5.4. } \vspace{0.3cm}\noindent {\bf Remark.} We comment about the case of the polytopes of type III. In this case, since the toric variety ${\Bbb P}_{\Sigma({\Delta^*},T_o)}$ is singular, the Chow ring should be considered over ${\Bbb Q}$. Under this modification the expansion (\ref{eqn: wJ}) makes sense in $A^*({\Bbb P}_{\Sigma({\Delta^*},T_o)})_{\Bbb Q} \otimes {\Bbb C}\{x_\tau\}[\log x_\tau]$. Then the coefficient series should be in a subspace of the whole solution space of the ${\Delta^*}$-hypergeometric system. More precisely, as we see in Remark after Corol. 4.11, the initial ideal $LT_{\omega}({\cal I}_{\cal A})$ is no longer radical but we have a strict inclusion $LT_{\omega}({\cal I}_{\cal A}) \subset \sqrt{LT_{\omega}({\cal I}_{\cal A})}$. As discussed there, we have $LT_{\omega}({\cal D}_l)=(\da{i_1})^{\lambda_{i_1}} \cdots (\da{i_a})^{\lambda_{i_a}}$ for some element of the Gr\"obner basis ${\cal B}_{\omega}=\{ {\cal D}_{l_1}, \cdots, {\cal D}_{l_s} \}$. If we define ${\rm rad}(LT_{\omega}({\cal D}_l)) :=\da{i_1}\cdots\da{i_a}$, then the radical may be expressed by $\sqrt{LT_{\omega}({\cal I}_{\cal A})}= \langle {\rm rad}(LT_w({\cal D}_{l_1})), \cdots, {\rm rad}(LT_w({\cal D}_{l_s})) \rangle $. Correspondingly, if we define $\tilde I_{\omega}(\ta{{}}):= a_0 a_{i_1}\cdots a_{i_a} {\rm rad}(LT_{\omega}({\cal D}_l)) a_0^{-1}$, we naturally come to the ``radical'' of the indicial ideal $\tilde{Ind}_{\omega}(\tau):= \langle \tilde J_{l_1}(\rho;A_\tau), \cdots, \tilde J_{l_s}(\rho;A_\tau) \rangle $ with $\tilde J_l(\rho;A_\tau):=x_\tau^{-\rho}\tilde I_{\omega}(\ta{{}}) x_\tau^{\rho}$. By definition, we have strict inclusions $Ind_{\omega}(\tau)$ $\subset \tilde{Ind}_{\omega}(\tau)$ and $V(Ind_{\omega}(\tau)) \supset V(\tilde{Ind}_{\omega}(\tau))$. As is clear now, our Prop.5.2 and Prop.5.3 apply to the ``radical'' $\tilde{Ind}_{\omega}(\tau)$ under the replacements $I_{\omega}$ by $\tilde I_{\omega}$ and the Chow ring by that over ${\Bbb Q}$. The expansion (\ref{eqn: wJ}) gives all logarithmic solutions which arise from the degeneration $V(\tilde{Ind}_{\omega}(\tau))$ $=\{ 0\}$. \newsubsec{(5-3) LCSL of Calabi-Yau hypersurfaces} So far we have been concerned with the ${\Delta^*}$-hypergeometric system. Since the period integral (\ref{eqn: Pi}) of Calabi-Yau hypersurface $X_{\Delta^*}$ satisfies the (extended) ${\Delta^*}$-hypergeometric system, a complete set of the period integrals of $X_{\Delta^*}$ should be found in the set of solutions of the ${\Delta^*}$-hypergeometric system. We will find that the expansion (\ref{eqn: wJ}) contains the period integrals in a natural way from the mirror symmetry. Before going into this topic, we need to discuss about the compactification of the moduli space ${\cal M}(X_{\Delta^*}(a))$ of the polynomial deformation of the Calabi-Yau hypersurface $X_{\Delta^*}$. Through a detailed analysis of the local solutions of the ${\Delta^*}$-hypergeometric system, we have arrived at a natural compactification, the Gr\"obner compactification ${\Bbb P}_{{\cal F}({\cal I}_{\cal A})}$ in Prop.4.8. Now it is natural to adopt this compactification as that of the moduli space ${\cal M}(X_{\Delta^*}(a))$. However, one problem arises when the hypersurface (, precisely its ambient space,) has non-trivial automorphisms. We need to mod out the space ${\Bbb P}_{{\cal F}({\cal I}_{\cal A})}$ by the induced actions from the automorphisms, whose infinitesimal forms are described in (\ref{eqn: extGKZ}). Here to avoid getting involved in the problems related to the actions of the automorphisms, we take a ``gauge choice'' that sets to zero all polynomial deformations corresponding to integral points on codimension-one faces of ${\Delta^*}$. Note that, in view of Prop.2.6, the degree of the freedom associated to the non-trivial automorphisms would be fixed by this gauge choice. In the following, we use the subscript $s$ (, $s$ of simply-minded!,) to indicate this naive choice of the ``gauge''; for example $\Delta_s^*$-hypergeometric system, the toric ideal ${\cal I}_{{\cal A}_s}$ etc. Note that all the polytopes of type II will be treated as the polytopes of type III under this prescription. \vspace{0.3cm}\noindent {\bf Definition 5.9.} As an compactification of ${\cal M}(X_{\Delta^*}(a))$, we define \begin{equation} \overline{\cal M}(X_{{\Delta^*}}(a))={\Bbb P}_{{\cal F}({\cal I}_{{\cal A}_s})} \;\;. \end{equation} \vspace{0.3cm} Now we consider the toric part of the Chow ring of the Calabi-Yau hypersurface $X_{{\Delta}_s}$, which comes from the ambient space by restriction. Since we have $[X_{{\Delta}_s}]=D_1+\cdots+D_p$ for the divisor of the Calabi-Yau hypersurface, the restriction may be attained by the quotient as follow; \vspace{0.3cm}\noindent {\bf Definition 5.10.} \begin{equation} A^*(X_{{\Delta}_s})_{toric}= A^*({\Bbb P}_{\Sigma(\Delta_s^*,T_o)})_{\Bbb Q}/ Ann(D_1+\cdots+D_p) \;\;, \label{eqn: chowX} \end{equation} where $Ann(x)$ is defined by $Ann(x)=\{ \; y\in {\cal R} \;\vert\; x\,y=0 \;\}$ for a ring ${\cal R}$. \vspace{0.3cm}\noindent {\bf Claim 5.11.} {\it Period integrals about a LCSL of Calabi-Yau hypersurface $X_{\Delta_s^*}$ are extracted from the series $w_0(x_\tau,J;A_\tau)$ (\ref{eqn: wJ}) expanded in $A^*(X_{{\Delta}_s})_{toric}\otimes {\Bbb C}\{x_\tau\}[\log x_\tau]$. } \vspace{0.3cm}\noindent {\bf Remark.} In general, the period integrals of Calabi-Yau hypersurfaces satisfy the differential equations of Fuchs type, so-called the Picard-Fuchs equations \cite{PF}. Picard-Fuchs equations determines the period integrals as its solutions. Our Claim 5.11 says that our ${\Delta^*}$-hypergeometric system is reducible in general and contains the Picard-Fuchs equation as a component of it. In refs.\cite{HKTY1}\cite{HLY1}, it is verified in several examples of all types of the polytopes that Picard-Fuchs equations are derived from the (extended) ${\Delta^*}$-hypergeometric system after a factorization of the operator $\ta{1}+\cdots+\ta{p}$ from the left, which we may identify with the quotient by $Ann(D_1+\cdots+D_p)$ in (\ref{eqn: chowX}). \section{6. Prepotential} In this section, we study so-called the prepotential \cite{Saito} near a LCSL in detail. Under the mirror map, a LCSL is mapped to a large radius limit in which the instanton corrections to the prepotential are suppressed exponentially, and has important applications to the enumerative geometry. Also the prepotential determines the {\it special K\"ahler geometry} on the moduli space $\overline {\cal M}(X_{\Delta_s^*})$ and that on the complexified K\"ahler moduli space of the mirror $X_{{\Delta}_s}$. In this section, we fix a term order ${\omega}$ for which $T_{\omega}$ is a maximal triangulation of $\Delta_s^*$ and take the ${\bf Z}$-basis $A_\tau$ choosing a regular cone $\tau$ in ${\cal C}'({\cal I}_{{\cal A}_s},{\omega})$. Based on Claim 5.11, we expand the series $w_0(x_\tau,J;A_\tau)$ defined in (\ref{eqn: wJ}) (see also \cite{Sti}) as; \begin{equation} w_0(x_\tau,J;A_\tau)= w^{(0)}(x_\tau,J)+w^{(1)}(x_\tau,J)+{1\over2!}w^{(2)}(x_\tau,J) +{1\over3!}w^{(3)}(x_\tau,J) \;, \label{eqn: wJexp} \end{equation} where the superscripts indicate the degree in the Chow ring $A^*(X_{{\Delta}_s})_{toric}\otimes {\Bbb C}\{x_\tau\}[\log x_\tau]$. \vspace{0.3cm}\noindent {\bf Definition 6.1.} The {\it special coordinate} $(t_1,\cdots, t_{p-d})$ of the special K\"ahler geometry is defined by the ratios of the period integrals; \begin{equation} t{\hskip-0.02cm\cdot\hskip-0.02cm} J = {w^{(1)}(x_\tau,J) \over w^{(0)}(x_\tau,J) } \;\;, \label{eqn: special} \end{equation} where we abuse the letters $J_1, \cdots, J_{p-d}$ to represent the images of the $J$'s under the quotient (\ref{eqn: chowX}). The inverse series of this relation will be called the {\it mirror map}. \vspace{0.3cm}\noindent {\bf Note.} Since $w^{(1)}(x_\tau,J)$ is linear in log$x_\tau$, the mirror map takes the form $x_\tau(q):=x_\tau(q_1,\cdots,q_{p-d})$ with $q_k:={\rm e}^{2\pi i t_k}$. It is easy to see that $x_\tau^{(k)}(q) =q_k(1+{\cal O}(q))$. \vspace{0.3cm}\noindent {\bf Definition 6.2.} We define the prepotential in the special coordinate by \begin{equation} F(t):=\int_{X_{{\Delta}_s}} {\cal F}(x_\tau(q), J) \;\;, \label{eqn: Ft} \end{equation} with the {\it invariant} density \begin{equation} {\cal F}(x_\tau(q), J)={1\over2} \({1\over w^{(0)}}\)^2 \{w^{(0)}(-{1\over3!}w^{(3)}-{c_2(X_{{\Delta}_s})\over 12}w^{(1)})+ w^{(1)}({1\over2!}w^{(2)}) \} \;. \label{eqn: Fd} \end{equation} The integration symbol $\int_{X_{{\Delta}_s}} :=\int_{{\Bbb P}_{\Sigma({\Delta_s^*},T_o)}} [X_{{\Delta}_s}]$ is meant to take the coefficient of the 'volume form' in the Chow ring $A^*({\Bbb P}_{\Sigma(\Delta_s^*,T_o)})_{\Bbb Q}$ normalized by $\int_{{\Bbb P}_{\Sigma({\Delta_s^*},T_o)}} [X_{{\Delta}_s}]c({\Bbb P}_{\Sigma(\Delta_s^*,T_o)})/(1+[X_{{\Delta}_s}])=\chi(X_{{\Delta}_s})$. (For the normalization when $\chi(X_{{\Delta}_s})=0$, see ref.\cite{HLY1}.) \vspace{0.3cm}\noindent {\bf Note.} It would be instructive to summarize the general description\cite{Str} of the special K\"ahler geometry on the complex structure moduli space of Calabi-Yau threefolds. Let us denote the holomorphic 3-form of a family of Calabi-Yau threefolds $W_\psi$ by $\Omega(\psi)$. We take a symplectic basis $\{A_a,B_b\}$ $(a,b=0,\cdots,h^{2,1}(W))$ of $H_3(W,{\bf Z})$ and construct the period integrals $z_a(\psi)=\int_{A_a} \Omega(\psi)$ and ${\cal G}_b(\psi)=\int_{B_b}\Omega(\psi)$. Then the holomorphic 3-form may be written by $\Omega(\psi)=\sum_a z_a(\psi) \alpha_a + \sum_b {\cal G}_b(\psi)\beta_b$ in terms of the dual bases $\alpha_a$ and $\beta_b$ in $H^3(W,{\bf Z})$. Locally we can introduce on the moduli space a K\"ahler metric, so-called the Weil-Peterson metric \cite{Tian}, through the K\"ahler potential $K(\psi,\bar\psi)=-\log i\int_M \Omega(\psi) \wedge \bar\Omega(\psi)$ $= -\log i\sum_a (z_a(\psi) \overline{ {\cal G}_a(\psi)}$ $- {\cal G}_a(\psi) \overline{ z_a(\psi)} )$. It is shown in ref.\cite{Str} that the prepotential $F(\psi)=$ ${1\over 2}\sum_a z_a(\psi){\cal G}_a(\psi)$ describes the potential $K(\psi,\bar\psi)$ by \begin{equation} K(\psi,\bar\psi)= -{\rm log}i\sum_a \( z_a(\psi)\overline{{\partial F(\psi)\over \partial z_a}} -\overline{z_a(\psi)} {\partial F(\psi) \over \partial z_a} \) \;, \end{equation} and defines the {\it special K\"ahler geometry} on the moduli space. Our definition (\ref{eqn: Fd}) of the prepotential, up to the prefactor $(w^{(0)})^{-2}$ which makes the prepotential invariant under $\Omega(\psi)\mapsto f(\psi)\Omega(\psi)$, implicitly contains a claim that $(w^{(0)},w^{(1)},{1\over2!}w^{(2)},-{1\over3!}w^{(3)}- {c_2(X_{{\Delta}_s}) \over 12} w^{(1)})$ form the period integrals for a symplectic basis of $H_3(X_{\Delta_s^*},{\bf Z})$. Several evidences for this claim are reported in \cite{HLY3}. In the following, we will restrict our attention to the form of the prepotential near a LCSL assuming its application to the enumerative geometry (the instanton counting). \vspace{0.3cm} Now consider the following expansion in the Chow ring associated to the series $w_0(x_\tau,0,A_\tau)=\sum_{n\in {\bf Z}_{\geq0}^{p-d}}c(n)x_\tau^n$; \begin{equation} \sum_{n\in {\bf Z}^{p-d}_{\geq0}}c\(n+{J\over 2\pi i}\)x_\tau^n= \tilde w^{(0)}(x_\tau,J)+\tilde w^{(1)}(x_\tau,J)+ {1\over2!}\tilde w^{(2)}(x_\tau,J) +{1\over3!}\tilde w^{(3)}(x_\tau,J) \;. \label{eqn: wJexpp} \end{equation} \vspace{0.3cm}\noindent {\bf Lemma 6.3.} {\it The two definitions of the series (\ref{eqn: wJexp}) and (\ref{eqn: wJexpp}) are related by \begin{equation} \begin{array}{crl} w^{(0)}&=&\tilde w^{(0)} \;, \\ w^{(1)}&=&(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J) w^{(0)} + \tilde w^{(1)} \;,\\ w^{(2)}&=&(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J)^2 w^{(0)} + 2(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J)\tilde w^{(1)} + \tilde w^{(2)}\;,\\ w^{(3)}&=&(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J)^3 w^{(0)} + 3(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J)^2\tilde w^{(1)} + 3(\log x{\hskip-0.02cm\cdot\hskip-0.02cm} \hat J) \tilde w^{(2)} + \tilde w^{(3)}\;,\\ \end{array} \end{equation} where we have introduced an abbreviation $\log x{\hskip-0.02cm\cdot\hskip-0.02cm}\hat J= \sum_{a=1}^{p-d} (\log x_a){1\over 2\pi i}J_a$. } \vspace{0.3cm}\noindent {\bf Lemma 6.4.} {\it The series $\tilde w^{(d)}(x_\tau,J) \;\;(d=1,2,3)$ in (\ref{eqn: wJexpp}) have the form \begin{equation} \begin{array}{crl} \tilde w^{(1)}(x_\tau,J)&=&\sum_n c(n) \Psi^{(1)}(n) x_\tau^n \;\;, \\ \tilde w^{(2)}(x_\tau,J)&=&\sum_n c(n) \{(\Psi^{(1)}(n))^2+\Psi^{(2)}(n)\}x_\tau^n \;\;, \\ \tilde w^{(3)}(x_\tau,J)&=&\sum_n c(n) \{(\Psi^{(1)}(n))^3+ 3\Psi^{(1)}(n)\Psi^{(2)}(n)+\Psi^{(3)}(n)\} x_\tau^n \;\;, \\ \end{array} \label{eqn: wPsi} \end{equation} where $\Psi^{(k)}(n)$'s are elements in the Chow ring of degree $k$ defined by \begin{equation} \begin{array}{crl} &&\Psi^{(1)}(n) =-(\hat J \cdt l_0)\psi(1-n \cdt l_0)-\sum_{i=1}^{p} (\hat J \cdt l_i) \psi(1+n \cdt l_i) \;,\\ &&\Psi^{(2)}(n)=(\hat J \cdt l_0)^2\psi'(1-n \cdt l_0)- \sum_{i=1}^{p}(\hat J \cdt l_i)^2\psi'(1+n \cdt l_i) \;, \\ &&\Psi^{(3)}(n)=-(\hat J \cdt l_0)^3\psi''(1-n \cdt l_0)- \sum_{i=1}^{p}(\hat J \cdt l_i)^3\psi''(1+n \cdt l_i) \;, \\ \end{array} \end{equation} with $\hat J{\hskip-0.02cm\cdot\hskip-0.02cm} l_k = \sum_{a=1}^{p-d} {J_a \over 2\pi i} l^{(a)}_k \; (k=0,\cdots, p)$, $\psi(z)={d \; \over dz}\log\Gamma(z)$ and the derivatives of $\psi(z)$. } \vspace{0.3cm}\noindent {\bf Lemma 6.5.} {\it \begin{equation} \Psi^{(1)}(0)=0 \;\;,\;\; \Psi^{(2)}(0)=-{c_2(X_{{\Delta}_s})\over 12} \;\;,\;\; \Psi^{(3)}(0)=-{6\zeta(3) \over (2\pi i)^3} c_3(X_{{\Delta}_s}) \;\;. \end{equation} } \par \noindent {\bf (Proof)} These constant terms originate from those of the $\psi$-functions; $\psi(1)=-\gamma \;,\; \psi'(1)={\pi^2 \over 6} \;,\; \psi''(1)=-2 \zeta(3)$. These values of the $\psi$-functions combined with the adjunction formula for the total Chern class, with $D_i=J{\hskip-0.02cm\cdot\hskip-0.02cm} l_i$ under the rational equivalence in the Chow ring, \begin{equation} c(X_{{\Delta}_s})={\prod_{i=1}^p(1+D_i) \over 1+[X_{{\Delta}_s}] } ={\prod_{i=1}^p(1+J{\hskip-0.02cm\cdot\hskip-0.02cm} l_i) \over 1-J{\hskip-0.02cm\cdot\hskip-0.02cm} l_0} \;, \end{equation} result in our claim for the leading terms. (Note that $c_1(X_{{\Delta}_s})=0$ for $\Psi^{(1)}(0)$.) \hfill $\Box$ \vspace{0.3cm}\noindent {\bf Remark.} We can subtract these constant terms $\Psi^{(k)}(0)$ in a systematic way modifying the expansion (\ref{eqn: wJexpp}) slightly as follows; \begin{equation} \sum_{n\in {\bf Z}^{p-d}_{\geq 0}} {c(n+{J\over2\pi i}) \over c({J\over2\pi i})} x_\tau^n = w^{(0)}(x_\tau)+\tilde w_r^{(1)}(x_\tau,J)+ {1\over 2}\tilde w_r^{(2)}(x_\tau,J) + {1\over 3!} \tilde w_r^{(3)}(x_\tau,J) \;. \end{equation} This is because this change of normalization in the series $w_0$ simply results in the replacement $\Psi^{(k)}(n)$ with $\Psi^{(k)}_r(n):= \Psi^{(k)}(n)-\Psi^{(k)}(0)$ in (\ref{eqn: wPsi}). \vspace{0.3cm} Now it is immediate from Lemmas 6.4 and 6.5 to obtain \vspace{0.3cm}\noindent {\bf Lemma 6.6.} \begin{eqnarray*} \tilde w^{(1)}&=&\tilde w^{(1)}_r \;\;,\;\; \tilde w^{(2)}=-{c_2(X_{{\Delta}_s}) \over 12}w^{(0)} + \tilde w_r^{(2)} \;,\\ \tilde w^{(3)}&=&-{6\zeta(3) \over (2\pi i)^3} c_3(X_{{\Delta}_s})w^{(0)} -{c_2(X_{{\Delta}_s}) \over 4} \tilde w_r^{(1)} + \tilde w_r^{(3)} . \\ \end{eqnarray*} \vspace{-2.5cm} \begin{equation} \label{eqn: wr} \end{equation} \par\vspace{0.8cm}\noindent \vspace{0.3cm} Now using the results in Lemmas 6.3-6.6, we may arrive at our final form of the prepotential, see also \cite{Sti}, modulo the kernel of the integration $\int_{X_{{\Delta}_s}}$ in Def.6.2; \vspace{0.3cm}\noindent {\bf Proposition 6.7.} {\it The invariant form of the prepotential ${\cal F}(x,J)$ may be expressed by \begin{eqnarray} {\cal F}(t,J) &=&{1\over6}(t{\hskip-0.02cm\cdot\hskip-0.02cm} J)^3-{c_2(X_{{\Delta}_s}) \over 24}(t{\hskip-0.02cm\cdot\hskip-0.02cm} J)+ {\zeta(3) \over 2(2\pi i)^3}c_3(X_{{\Delta}_s}) \nonumber \\ && \quad\quad -{1\over 2}\log \left( \sum_{n\in {\bf Z}_{\geq0}^{p-d}} { c(n+{J\over 2\pi i}) \over c({J\over 2\pi i}) } x_\tau^n \right) \;\;, \nonumber \\ \end{eqnarray} with the mirror map $x_\tau=x_\tau(q)$. } \vspace{0.3cm}\noindent {\bf Claim 6.8.} {\it Three times derivatives of the prepotential give the instanton corrected Yukawa couplings; \begin{eqnarray} &&K_{t_a t_b t_c}(t)={\partial^3 \;\; \over \partial t_a \partial t_b \partial t_c} F(t) \nonumber \\ &&= \int_{X_{{\Delta}_s}} J_aJ_bJ_c + \sum_{ \Gamma\in H_2(X_{{\Delta}_s},{\bf Z}) \atop \Gamma\not=0 } (\Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} J_a)(\Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} J_b) (\Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} J_c) N(\Gamma) {{\rm e}^{2\pi i \Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} (t{\hskip-0.02cm\cdot\hskip-0.02cm} J)} \over 1- {\rm e}^{2\pi i \Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} (t{\hskip-0.02cm\cdot\hskip-0.02cm} J)}} \;, \nonumber \\ \end{eqnarray} where $\Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} J:= \int_\Gamma J$ and $N(\Gamma)$ counts the number of the rational curves of class $\Gamma$ on the Calabi-Yau manifolds $X_{{\Delta}_s}$.} \vspace{0.3cm}\noindent {\bf Note.} Since the mirror map has the $q$-expansion $x_\tau^{(k)}(q)=q_k(1+{\cal O}(q))$, it is immediate to deduce that the number of lines $N(\Gamma)$ in $X_{{\Delta}_s}$ is counted by \begin{equation} N(\Gamma)=\int_{X_{{\Delta}_s}} -{1\over2}{c((\Gamma{\hskip-0.02cm\cdot\hskip-0.02cm} J)+J) \over c(J)} \;\;. \end{equation} We see that the famous number $2785$ for the quintic in ${\Bbb P}^4$ \cite{CdGP} is counted by this formula as \begin{equation} N(1)=-{1\over2}\int_{{\Bbb P}^4} 5J {(5+5J)(4+5J)(3+5J)(2+5J)(1+5J) \over (1+5J)^5}\;\;. \end{equation} The invariant form of the prepotential ${\cal F}(x,J)$ may have significant applications to extracting the predicted numbers of the rational curves $N(\Gamma)$. In a recent work \cite{HSS}, this form has been utilized efficiently to verify that the numbers $N(\Gamma)$ of a certain Calabi-Yau manifold (Schoen's Calabi-Yau manifold) are related to the modular forms, the theta function of the $E_8$ lattice and Dedekind's eta function. \def\thebibliography#1{\vskip 1.2pc{\centerline {\bf References}}\vskip 4pt \list {[\arabic{enumi}]}{\settowidth\labelwidth{[#1]}\leftmargin\labelwidth \advance\leftmargin\labelsep \usecounter{enumi}} \def\hskip .11em plus .33em minus .07em{\hskip .11em plus .33em minus .07em} \sloppy\clubpenalty4000\widowpenalty4000 \sfcode`\.=1000\relax} \let\endthebibliography=\endlist
1997-07-31T12:26:47
9707
alg-geom/9707017
en
https://arxiv.org/abs/alg-geom/9707017
[ "alg-geom", "math.AG" ]
alg-geom/9707017
Andre. Hirschowitz
A. Hirschowitz and S. Ramanan
New evidence for Green's conjecture on syzygies of canonical curves
Tex-type: LaTeX
null
null
null
null
We prove that two weakened forms of Green's conjectures for canonical curves are equivalent when the genus $g$ is odd.
[ { "version": "v1", "created": "Thu, 31 Jul 1997 11:27:37 GMT" } ]
2008-02-03T00:00:00
[ [ "Hirschowitz", "A.", "" ], [ "Ramanan", "S.", "" ] ]
alg-geom
\section{Introduction} Some twelve years ago, Mark Green [G] made a few conjectures regarding the behaviour of syzygies of a curve $C$ imbedded in $\mbox{$I\!\!P$}^n$ by a complete linear system. The so-called {\it generic} Green conjecture on canonical curves pertains to this question when the linear system is the canonical one and the curve is generic in the moduli, and predicts what are the numbers of syzygies in that case. Green and Lazarsfeld [GL] have observed that curves with nonmaximal Clifford index have extra syzygies and we will call {\it specific} Green conjecture on canonical curves the stronger prediction that the curves which have the numbers of syzygies expected for generic curves are precisely those with maximal Clifford index ($[(g-1)/2]$). (As a matter of fact, the full Green conjecture on canonical curves relates more closely the Clifford number with the existence of extra syzygies.) Many attempts have been made to settle this question, and some nice results have been obtained ([Sch][V]). In this note, we work over an algebraically closed field of arbitrary characteristic and prove that, as stated above, the generic and specific Green conjectures for canonical curves are equivalent at least when the genus $g$ is odd. Let $C$ be a curve canonically imbedded in $\mbox{$I\!\!P$}^{g-1}$ with ideal sheaf ${\cal I}_C$. We denote by $Q$ the universal quotient on $\mbox{$I\!\!P$}^{g-1}$, so that $Q(1)$ is the tangent bundle, and by $Q_C$ its restriction to $C$. It is generally known (see [P-R] for example) that extra syzygies appear when, for some $i\leq [(g-1)/2]$, the natural map $\Lambda ^i (\Gamma (C,Q_C)) \rightarrow \Gamma (C, \Lambda ^iQ_C)$ is not surjective. It is easy to see that the relevant quotient of $\Gamma (C, \Lambda ^iQ_C)$ by $\Lambda ^i (\Gamma (C,Q_C))$ is isomorphic to $\Gamma (\Lambda^{i+1}Q \otimes {\cal I}_C(1))$ (cf 2.1). \begin{thm} Let $g = 2k-1 \geq 5$ be an odd integer. If the generic curve $C$ of genus $g$ has the expected number of syzygies (i.e. $\Gamma (\Lambda ^{k} Q \otimes {\cal I}_C(1))=0$), then so does any curve of genus $g$ with maximal Clifford index, namely $k-1.$ \end{thm} To prove this, we compute a virtual (divisor) class $v$ for the locus (in the moduli) of curves $C$ for which the cohomology group $\Gamma (\Lambda^{k} Q \otimes {\cal I}_C(1))$ does not vanish. Once $v$ is computed, we compare it with the class $c$ of the locus of $k$-gonal curves (these are curves with non-maximal Clifford index in our case), which, thanks to Harris and Mumford [HM], is already known, and we find that $v=(k-1)c$. We conclude by proving that the generic $k$-gonal curve has at least $k-1$ extra syzygies, which implies that the $k$-gonal locus occurs with multiplicity $k-1$ in $v$, leaving no room for another component. Our proof gives another consequence of the generic Green's conjecture, namely that the number of extra syzygies (more precisely $h^0(\Lambda^{k} Q \otimes {\cal I}_C(1))$) is exactly $k-1$ for any $k$-gonal curve $C$ in the smooth part of the $k$-gonal locus. Finally, we observe that our argument fails completely in the case of even genus, where the expected codimension of our jump locus is no more one. \section { Preliminaries on syzygies} \label {prel} We collect here a few useful remarks on syzygies. \begin{prop} Let $S$ be a linearly normal subscheme of $\mbox{$I\!\!P$}^n $ (i.e. $\Gamma (\mbox{$I\!\!P$}^n,{\cal O}(1)) \to \Gamma (S, {\cal O}(1))$ is an isomorphism) with ideal sheaf ${\cal I}_S$. Then the cokernel of $$ \Lambda^ i\Gamma (\mbox{$I\!\!P$}^n,Q) = \Gamma (\mbox{$I\!\!P$}^n, \Lambda ^i Q)\rightarrow \Gamma (S,\Lambda^i Q_S) $$ is canonically isomorphic to $$\Gamma (\mbox{$I\!\!P$}^n,\Lambda^{i+1} Q\otimes {\cal I}_S(1)).$$ \end{prop} \noindent{\bf Proof.} Consider the following exact sequence of sheaves on $P(V) = \mbox{$I\!\!P$}^n$: $$ 0\rightarrow {\cal O}(-1) \rightarrow V_P \rightarrow Q \rightarrow 0 $$ By taking the exterior $(i+1)$-th power and tensoring with ${\cal O}(1)$, we get the exact sequence: $$0\rightarrow \Lambda^{i}Q \rightarrow \Lambda^{i+1} V_P (1)\rightarrow \Lambda^ {i+1}Q (1) \rightarrow 0. $$ This exact sequence of vector bundles remains exact on tensorisation by ${\cal I}_S$ as well as ${\cal O}_S$. Thus we get the commutative diagram $$\begin{array}{ccccccccc} & & 0 & & 0 & & 0 & &\\ & & \downarrow & & \downarrow & & \downarrow & &\\ 0 & \rightarrow & \Lambda^i Q \otimes {\cal I}_S & \rightarrow & \Lambda^{i+1} V_P \otimes {\cal I}_S(1) & \rightarrow & \Lambda^{i+1} Q \otimes {\cal I}_S(1) & \rightarrow & 0 \\ & & \downarrow & & \downarrow & & \downarrow & &\\ 0 & \rightarrow & \Lambda^i Q & \rightarrow & \Lambda^{i+1} V_P (1) & \rightarrow & \Lambda^{i+1} Q(1) & \rightarrow & 0\\ & & \downarrow & & \downarrow & & \downarrow & &\\ 0 & \rightarrow & \Lambda^i Q_S & \rightarrow & \Lambda^{i+1} V_S(1) & \rightarrow & \Lambda^{i+1} Q_S(1) & \rightarrow & 0 \\ & & \downarrow & & \downarrow & & \downarrow & &\\ & & 0 & & 0 & & 0 & & \end{array} $$ Now apply the section functor $\Gamma $ : the middle row remains exact. Thus we may apply the snake lemma to the two lower rows. This yields the desired isomorphism because under our assumption, $\Gamma (\Lambda ^{i+1}V_P(1)) \to \Gamma (\Lambda ^{i+1}V_S(1))$ is an isomorphism as well. $\hfill\square$ \begin{rem} Thus we will think of $\Gamma (\Lambda ^j Q \otimes {\cal I}_C(1))$ as the space of extra syzygies. From this point of view, extra syzygies behave in a monotonic way with respect to the degree $j$ and the subvariety $C$: a) If $C \subset S$ are two subvarieties of $\mbox{$I\!\!P$}^{g-1}$, then $h^0 (\Lambda ^j Q \otimes {\cal I}_C(1))\geq h^0 (\Lambda ^j Q \otimes {\cal I}_S(1))$. We will estimate syzygies of our canonical curves by using a scroll $S$ containing them. b) If $i < j$, then $h^0 (\Lambda ^j Q \otimes {\cal I}_C(1))\geq h^0 (\Lambda ^i Q \otimes {\cal I}_C(1)).$ The above proposition is applicable with our canonical curve: $S = C$. Also, the Clifford index of the generic curve of genus $g$, is well-known to be $[(g-1)/2]$. Finally, if $ i < j$ then $\Gamma (\Lambda ^j Q \otimes {\cal I}_C(1))=0$ implies $\Gamma (\Lambda ^i Q \otimes {\cal I}_C(1))=0$ so that we have an equivalent formulation of the specific conjecture of Green: \noindent{\bf Specific Green's conjecture.} {\it Let $C$ be a canonically imbedded curve with maximum Clifford index $[(g - 1)/2]$. Then $\Gamma (\Lambda ^j Q \otimes {\cal I}_C(1))$ is zero for $j = [(g+1)/2]$.} \end {rem} We will use in Section \ref {scroll} the following semi-continuity statement: \begin{prop} \label {scs} Let $p: W \rightarrow T$ be a smooth family of projective varieties parametrized by the spectrum $T$ of a discrete valuation ring. We suppose that $W$ is endowed with a line bundle ${\cal L}$, that $h^0(W_t, {\cal L}_t)$ is constant and that for each point $t \in T$, ${\cal L}_t$ is generated by global sections. This yields a $T$-morphism $m$ from $W$ to $\mbox{$I\!\!P$}(p_*{\cal L})$. We denote by $I_t$ the ideal sheaf of $m(W_t)$ and by $Q_t$ the tautological quotient bundle on the fibre $\mbox{$I\!\!P$}_t=\mbox{$I\!\!P$}(H^0(W_t, {\cal L}_t))$. Then for any $i$, the dimension $h^0 (\mbox{$I\!\!P$}_t, \Lambda^iQ_t \otimes I_t(1))$ is upper-semi-continuous. \end{prop} \noindent {Proof.} By properness of the Hilbert scheme, there exists a $T$-flat subscheme $\bar W$ of $\mbox{$I\!\!P$}(p_*{\cal L})$ with the property that its fibre over the general point $t_1$ of $T$ is $m(W_{t_1})$. By continuity, its special fibre $\bar W_{t_0}$ contains $m(W_{t_0})$ (indeed, they are equal, but we don't need this). Thus, by inclusion, we have $$h^0 (\mbox{$I\!\!P$}_{t_0}, \Lambda^iQ_{t_0} \otimes I_{t_0}(1)) \geq h^0 (\mbox{$I\!\!P$}_{t_0}, \Lambda^iQ_{t_0} \otimes I_{\bar W_{t_0}}(1)),$$ and by semi-continuity, $$h^0 (\mbox{$I\!\!P$}_{t_0}, \Lambda^iQ_{t_0} \otimes I_{\bar W_{t_0}}(1)) \geq h^0 (\mbox{$I\!\!P$}_{t_1}, \Lambda^iQ_{t_1} \otimes I_{\bar W_{t_1}}(1)),$$ which altogether prove our claim.$\hfill \square$ \section{The syzygy locus in the case of odd genus} \label {S3} In this section, we write ${\cal M}={\cal M}^{o}_g$ for the open subvariety of ${\cal M}_g$ consisting of points that represent isomorphism classes of smooth curves with trivial automorphism group. What we need to know of ${\cal M}$ is that an effective divisor on it which is rationally equivalent to zero is indeed zero: this is for instance because ${\cal M}$ has a projective compactification with two-codimensional boundary (cf e.g. [A]). Let $x$ be a point in ${\cal M}$ and $C$ the corresponding curve. We consider the canonical imbedding of $C$ in $\mbox{$I\!\!P$}^{g-1}$ ($C$ is not hyperelliptic); we denote by ${\cal I}_C$ the ideal sheaf of $C$ and by $Q$ the tautological quotient bundle of rank $g-1$ on $\mbox{$I\!\!P$}^{g-1}$. Finally, we denote by $S_{g}$ the locus in ${\cal M}$ of (points corresponding to) curves $C$ satisfying $\Gamma (\Lambda^{k} Q \otimes {\cal I}_C(1)) \neq 0.$ As a jump locus, $S_g$ has a natural Cartier divisor structure (see e.g. the proof of the next proposition) and we compare its class in the Picard group of ${\cal M}$ with the class $c$ of the $k$-gonal locus (cf [HM]). \begin{prop} Let $g = 2k-1\geq 5$ be an odd integer such that the generic curve $C$ of genus $g$ satisfies $\Gamma (\Lambda^{k} Q \otimes {\cal I}_C(1))=0.$ Then, in the Picard group of ${\cal M}$, the rational class $v$ of $S_{g}$ is $(k-1)c$. \end{prop} \noindent{\bf Proof.} There exists a universal curve ${\cal C}$ over ${\cal M}$, that is to say a smooth variety $\cal C$ and a smooth projective morphism $\pi :\cal C\to \cal M$ such that for any $x\in \cal M$ the fibre of $\pi $ over $x$ is the curve of genus $g$ whose isomorphism class is given by the point $x$. Let $\omega = \omega_{\pi }$ be the cotangent bundle along the fibres and $E$ its direct image on ${\cal M}$ by $\pi $. Then $\pi$ factors through the natural canonical imbedding of ${\cal C}$ in the projective bundle $p:\mbox{$I\!\!P$} = \mbox{$I\!\!P$} (E) \to {\cal M}$. Let ${\cal I}$ be the ideal sheaf of ${\cal C}$ in $\mbox{$I\!\!P$}$. The relatively ample (hyperplane) line bundle along the fibres of $\mbox{$I\!\!P$} $ will be denoted as usual by $ {\cal O}_p(1)$. Finally we write again $Q$ for the vector bundle on $\mbox{$I\!\!P$} $ given by the exact sequence $$ 0\rightarrow {\cal O}_p(-1) \rightarrow p ^*(E)^* \rightarrow Q \rightarrow 0. $$ Observe that $p _*(\Lambda ^l Q(1))$ is a vector bundle of rank ${g\choose l}g - {g\choose l - 1}$. Similarly, on each fibre ${\cal C}_x$, $\Lambda ^l Q_{{\cal C}_x}\otimes \omega_{{\cal C}_x}$ is semi-stable (cf [PR]) of slope $2l + 2g - 2$, thus non-special, and $p _*(\Lambda ^l Q(1)\otimes {\cal O}_{\cal C})$ is also a vector bundle, of rank ${g-1\choose l}(2l + g - 1)$, for each $l>0.$ Substituting $k$ for $l$, the restriction from $\mbox{$I\!\!P$}$ to the universal curve yields a morphism $r$ from $p _*(\Lambda ^k Q(1))$ to $p _*(\Lambda ^k Q(1)\otimes {\cal O}_{\cal C})$, and our assumption means that this morphism is injective (at the generic point). We observe that the two vector bundles have the same rank, namely ${2k - 2\choose k }(4k - 2) = {2k - 1\choose k - 1}(2k - 2)$. Thus the above map defines a (degeneracy) divisor in ${\cal M}$ and this is $S_g$, by definition. Its (virtual) rational class is $v=c_1(p _*(\Lambda ^k Q(1)\otimes {\cal O}_{\cal C}))-c_1(p _*(\Lambda ^k Q(1))).$ We will compute this class in $Pic({\cal M})$ as a multiple of $\lambda= c_1(E)$. We will start with the following computation in the appropriate Grothendieck group $K$. Let $t$ be an indeterminate and for any vector bundle $V$, let $\lambda _t(V)$ denote the element $\sum t^i \Lambda ^i(V)$ in $K[[t]]$. This extends to a homomorphism of $K$ into the multiplicative group consisting of power series with constant term 1 in $K[[t]]$, and this map is still denoted by $\lambda _t$. Consider now $x = p _!(\lambda _t(Q).{\cal I}(1)) = p _!(\lambda _t(Q).({\cal O}_p(1) - {\cal O}_{\cal C}(1)))$. Substitute $Q = p ^*(E^*) - {\cal O}_p(-1)$. Then we obtain \begin{eqnarray*} x & = & p _!( \frac {\lambda _t(p ^*(E^*))}{\lambda _t({\cal O}(-1))} ({\cal O}_p(1) - {\cal O}_{\cal C}(1))\cr & = & \lambda _t(E^*) p _! \left ( \frac {{\cal O}_p(1) - {\cal O}_{\cal C}(1)}{1 + t{\cal O}_p(-1)}\right )\cr & = & {\lambda }_t(E^*) \sum _{j = 0} ^{j = \infty } (-1)^j t^j (p _!({\cal O}_p(1-j)) - p _!({\cal O}_{\cal C}(1-j))). \end{eqnarray*} Our class $v$ is the coefficient of $t^k$ in the first Chern class of $-x$. Observe that $p _! ({\cal O}_p(1 - j)) = 0$, whenever $2 \leq j \leq g$. Also we have $p _!({\cal O}) = 1$ and $p _!({\cal O}_p(1)) = E. $ On the other hand, $p _!({\cal O}_{\cal C}(1 - j))$ can be seen to be $ E^* - 1$ for $j = 0$ and to be $1 - E$ for $j = 1$. The first Chern class of the direct images for $j\geq 2$ can be computed by the Grothendieck-Riemann-Roch theorem to be $(1 - 6(1 - j) + (1-j)^2) \lambda $ (see [M]). The first Chern class of $\lambda _t(E^*)$ is clearly equal to $-t(1 + t)^{g - 1}\lambda $. Also the rank of $\lambda _t(E)$ is $(1 + t)^g$, while the rank of $p _!({\cal O}_{\cal C}(n))$ is $(g - 1)(2n - 1)$. Thus $v$ is equal to $N\lambda $ where $-N$ is the coefficient of $t^k$ in $$ (1 + t)^g\{ 1 - \sum ^{i = \infty}_{i = 0} (-1)^i (1 + 6i + 6i^2)t^i \} - t(1 + t)^{g - 1}\{ g - t - (g - 1)\sum _{ i = 0}^{i = \infty } (- 1)^i t^i (1 - 2i)\}. $$ On the one hand, we have $(1 + t)^g (1 - \sum _{i = 0} ^{i = \infty } (-1)^i(1 - 6i + 6i^2)t^i)$ $= (1 + t)^g ( 1 - \sum _{i = 0} ^{i = \infty }(-1)^i(6(i + 1)(i + 2) - 24(i + 1) + 13)t^i )$ $=(1 + t)^g ( 1 - {12\over (1 + t)^3} + {24\over (1 + t)^2} - {13\over 1 + t})$ $=(1 + t)^{g - 3}( (1+t)^3 - {13}(1 + t)^2 + {24} (1 + t) - 12)$ $=t(1 + t)^{g - 3} (t^2 - 10t +1),$ and on the other, $t(1 + t)^{g - 1} ( g - t - (g-1)\sum _{i = 0} ^{i = \infty } (-1)^i( 3 - 2(i + 1))t^i)$\\ $=t(1 + t)^{g - 1}(g - t - (g - 1)({3\over 1 + t} - {2\over (1 + t)^2}))$\\ $=t(1 + t)^{g - 3}((1 + t)^2(g - t) - (g - 1)3(1 + t) - 2(g - 1))$\\ $=t(1 + t)^{g - 3}(-t^3 + (g-2)t^2 + (-g + 2)t + 1).$ This leads to the determination of $N$ to be the coefficient of $t^k$ in $$t^2(1 +t)^{2k - 4}(-t^2 +(2k - 4)t - (2k - 13)),$$ namely $$- {2k - 4\choose k - 4} + (2k - 4){2k - 4\choose k - 3} - (2k - 13){2k - 4\choose k - 2}$$ and this simplifies to $$ 6(k +1)(k - 1){(2k - 4)!\over (k-2)!k!}.$$ Now Harris and Mumford [HM] have studied the locus of $k$-gonal curves in ${\cal M}$ and have shown that this variety is a divisor whose class is $6(k+1){(2k - 4)!\over (k-2)!k!}\lambda $, which proves our claim. $\hfill \square$ \section{Syzygies of scrolls} \label {scroll} Extra syzygies of $k$-gonal curves arise because they lie on scrolls. So we start with estimating some syzygies of scrolls. \begin{prop} Let $W$ be a vector bundle on $\mbox{$I\!\!P$} ^1$ of rank $k-1$ and degree $k$. We suppose $W$ to be globally generated. We denote by $I_W$ the ideal of the image of the natural morphism from $\mbox{$I\!\!P$} (W)$ into $\mbox{$I\!\!P$} \Gamma (W)$ and by $Q$ the tautological quotient bundle on this projective space. Then the dimension $h^0 (\mbox{$I\!\!P$} \Gamma (W), \Lambda ^{k } Q\otimes I_W(1))$ is at least $k - 1$. \end{prop} \noindent {\bf Proof.} By \ref {scs}, we may suppose that $W$ is generic namely $W= {\cal O}(1)^{\oplus {k-2}} \oplus {\cal O}(2)$. In this case, the natural morphism $\mbox{$I\!\!P$}(W) \to \mbox{$I\!\!P$} \Gamma (W)$ is an imbedding. We will use freely the identification (2.1). Consider $X = \mbox{$I\!\!P$} ^1 \times \mbox{$I\!\!P$} ^1$ and the variety $Y = \mbox{$I\!\!P$}^1 \times \mbox{$I\!\!P$} (W)$. Let us denote as usual by $p_1$ and $p_2$ the two projections (in both cases) and by $\pi $ the fibration $\mbox{$I\!\!P$} (W)\to \mbox{$I\!\!P$} ^1$, as well as the morphism $Y\to X$ given by $I \times \pi $. Let $\Delta $ be the diagonal divisor in $X$ and $D$ its inverse image in $Y$. Let $Q$ be the universal quotient bundle on $\mbox{$I\!\!P$} \Gamma (W)$ and its restriction to $\mbox{$I\!\!P$} (W)$. Now consider on $Y$ the bundle homomorphism $p_1^*(W)^* \to p_2^*(Q)$ obtained as the composition of the pull back by $p_1$ of the natural inclusion $W^* \to \Gamma (W)^*\otimes {\cal O}$ and the pull-back by $p_2$ of the tautological map $\Gamma (W)^* \otimes {\cal O} \to Q$. This homomorphism is injective as a sheaf morphism but has one-dimensional kernel on the fibres over points of $D$. Thus we obtain an inclusion of ${\cal L}:= p_1^*(\Lambda ^{k - 1}W^*)\otimes {\cal O}(D)$ into $p_2^*(\Lambda ^{k - 1}Q)$. Note that ${\cal O}(D)$ is isomorphic to $ p_1^*{\cal O}(1) \otimes p_2^* \pi ^*{\cal O}(1)$ so that ${\cal L}$ is isomorphic to $p_1^*({\cal O}(-k + 1))\otimes p_2^*({\pi ^*(\cal O}(1)))$. Taking direct image by $p_1$ we get a homomorphism of ${\cal O}(-k + 1) \otimes \Gamma (\mbox{$I\!\!P$} ^1, {\cal O}(1))$ into $\Gamma (\Lambda ^{k - 1}Q)$. This fits in the following commutative diagram $$ \begin{array}{ccc} {\cal O}(-k)& \rightarrow & \Lambda ^{k - 1}\Gamma (Q)\otimes {\cal O}\\ \downarrow & & \downarrow \\ {\cal O}(-k + 1)\otimes \Gamma (\mbox{$I\!\!P$} ^1, {\cal O}(1)) & \rightarrow & \Gamma (\Lambda ^{k - 1}Q) \otimes {\cal O} \\ \downarrow & & \downarrow \\ {\cal O}(-k + 2)&\rightarrow & coker\Lambda ^{k - 1}\Gamma (Q)\otimes {\cal O}\to \Gamma (\Lambda ^{k - 1}Q)\otimes {\cal O}. \end{array} $$ We wish to make two remarks here. Firstly the lower horizontal arrow is nonzero. In fact, for any point $x$ of $\mbox{$I\!\!P$} ^1$, the middle horizontal arrow gives a two-dimensional space of sections of $\Lambda ^{k - 1}Q$. This is obtained as follows. Consider the sheaf inclusion of the trivial subbundle $W^*_x$ in $\Gamma (W)^*$ on $\mbox{$I\!\!P$}(W)$ and compose it with the natural homomorphism of the trivial bundle with fibre $\Gamma (W)^*$ into $Q$. Take the $(k - 1)$-th exterior power of this map. This becomes an inclusion of ${\cal O}(x) = {\cal O}(1)$ in $\Lambda ^{k - 1}Q$. Thus at the $\Gamma $-level this gives the two-dimensional space of sections required. Clearly the sections of the trivial bundle $\Lambda ^{k -1}(W_x^*)$ give a one-dimensional subspace of this. This is the top horizontal arrow in our diagram. Conversely let $s$ be an element of $\Lambda ^{k - 1}\Gamma (Q) = \Lambda ^{k - 1}(\Gamma W)^*$, then its exterior product with any element of $W_x^*$ gives an element of $\Lambda ^k\Gamma (Q) = \Lambda ^k (\Gamma W)^*$. If $s$ is actually a section of the sub-bundle generated by $W_x$, then this exterior product should be zero at the generic point and hence 0. This implies that $s$ belongs to $\Lambda ^{k - 1}(W_x)^* = {\cal O}(-k)$. Secondly, since all our constructions are canonical and $W$ is a homogeneous bundle, it follows that the lower horizontal arrow is $SL(2)$-equivariant. Now the proposition is a consequence of the following claim: If ${\cal O}(-n)$ admits a non-zero map into a trivial bundle, which is equivariant for the natural $SL(2)$-actions, then the rank of the trivial bundle is at least $n + 1$. To prove it, use the dual map of the trivial bundle into ${\cal O}(n)$ and use the fact that $\Gamma ({\cal O}(n))$ is an irreducible $SL(2)$-module. This implies that the induced map at the $\Gamma $-level, which is nonzero by assumption, is actually injective. $\hfill \square$ \section { Extra syzygies of gonal curves} \label {S5} We say that a curve of genus $g = 2k - 1$ is $k$-gonal if it carries a line bundle $L$ of degree $k$ whose linear system has no base points and thus yields a $k$-sheeted morphism $\pi $ onto $\mbox{$I\!\!P$}^1$. In this paragraph, we prove that $k$-gonal curves of genus $2k-1$ have at least $k-1$ extra syzygies. \begin{prop} Let $C$ be a nonhyperelliptic $k$-gonal curve of genus $2k - 1$, with $L$ the special line bundle of degree $k$ and $Q_C$ the restriction of the tautological quotient bundle on $\mbox{$I\!\!P$}^{g -1}$ to the canonically imbedded curve $C$. Then the dimension of $H^0(C, \Lambda ^{k - 1}Q_C)$ is at least ${g\choose k - 1} + k - 1$. \end{prop} \noindent {\bf Proof.} Consider the direct image $V$ of the canonical line bundle $K$ of $C$ by $\pi $. The so-called trace map gives a homomorphism of $V$ onto $K_{\mbox{$I\!\!P$} ^1} = {\cal O}(-2)$. Let $W$ be its kernel. Thus we have an exact sequence $$ 0 \rightarrow W \rightarrow V \rightarrow {\cal O}(-2)\rightarrow 0 $$ Since ${\cal O}(-2)$ has no nonzero sections it follows that $\Gamma (W) = \Gamma (V) = \Gamma (C,K)$. Moreover, the kernel of the evaluation map $\Gamma (C,K) \to W_p$ at any point $p\in \mbox{$I\!\!P$} ^1$ is simply the set of sections vanishing on $\pi ^{-1}(p)$, that is to say $s\Gamma (K\otimes L^{-1})$ where $s$ is a nonzero section of $L$ vanishing on this fibre. On computing the dimension of this space to be $k$ by Riemann-Roch, we find that the evaluation map from $\Gamma (C,K)_{\mbox{$I\!\!P$} ^1}$ to $V$ is actually onto $W$. Thus, $W$ is generated by global sections and we get a morphism of $\mbox{$I\!\!P$} (W)$ into $\mbox{$I\!\!P$} \Gamma (C,K)$. Finally the pull-back of $V$ to $C$, namely $\pi ^*\pi _*(K)$ comes with a natural homomorphism onto $K$. Indeed the natural surjection of $\Gamma (K)$ onto $K$ factors through this map, which can be thought of as `evaluation along fibres'. Thus we have a morphism from $C$ to $\mbox{$I\!\!P$} (W)$ the composition of which with the above mentioned morphism from $\mbox{$I\!\!P$} (W) \to \mbox{$I\!\!P$} \Gamma(K)$ to $C$ is the canonical imbedding. Thus our claim follows from Section \ref {scroll}. $\hfill \square$ \section{Proof of the theorem} In this section, we give the proof of our theorem. We start with the \begin{lem} Let $S$ be a smooth variety and $E$ and $F$ two vector bundles of the same rank $n$. Let $f:E\to F$ be a homomorphism which is generically an isomorphism, and $D$ a subvariety of codimension 1 in $S$ on which $f$ has kernel of rank $\geq r$, then the degeneracy divisor of $f$ contains $D$ as a component of multiplicity at least $r$. \end{lem} \noindent {\bf Proof.} Note that the question is local and localising at the generic point of $D$, we may assume that $S$ is a discrete valuation ring with maximal ideal $\gotm M$ and that $f$ is a square matrix of nonzero determinant. Then by a proper choice of basis we may assume $f$ to be diagonal of the form $\delta_{i,j}t^{m_i}, 0 \leq i,j \leq n$, where $t$ is a generating parameter. Our assumption ensures $m_i > 0$ for at least $r$ indices. Then clearly $det(f)$ is in $\gotm M^r$. Since the degeneracy locus is defined by $det (f)$, this proves our assertion. $\hfill \square$ Before turning to the proof, we state again our \begin{thm} Let $g = 2k-1 \geq 5$ be an odd integer. If the generic curve $C$ of genus $g$ has no extra syzygies (i.e. $\Gamma (\Lambda ^{k} Q \otimes {\cal I}_C(1))=0$), then so does any curve of genus $g$ with maximal Clifford index, namely $k-1.$ \end{thm} \noindent{\bf Proof.} We have shown (see Section 3) that the syzygy divisor $S_g$ is the degeneracy locus of a homomorphism of a vector bundle into another of same rank, and (see Section 5) that at the generic $k$-gonal curve this homomorphism has kernel of dimension at least $k - 1$. Thanks to the previous lemma, this implies that the locus of $k$-gonal curves is contained in $S_g$ with multiplicity at least $k - 1$. By our computation in Section 3, the residual divisor has rational class zero, thus is the zero divisor (this is what we need to know about ${\cal M}$). Thus in ${\cal M}$, curves with extra syzygies are in the $k$-gonal divisor. Now even around curves with automorphisms, we can see by going to a covering where a universal curve exists, that the locus of curves with extra syzygies is a divisor. Since the locus of curves with automorphisms is of codimension at least two in ${\cal M}$, we get that even in ${\cal M}_g$, curves with extra syzygies are in the $k$-gonal divisor, thus have nonmaximal Clifford index. This proves our theorem.$\hfill \square$ \begin {rem} We may even conclude that curves with nonmaximal Clifford index (which have extra syzygies by [GL]) are all in the $k$-gonal divisor. Note that this result is true (without our assumption on the generic curve), cf [ELMS]. \end {rem}
1998-07-20T11:38:20
9707
alg-geom/9707008
en
https://arxiv.org/abs/alg-geom/9707008
[ "alg-geom", "math.AG" ]
alg-geom/9707008
Pelham Wilson
P. M. H. Wilson
Flops, Type III contractions and Gromov-Witten invariants on Calabi-Yau threefolds
20 pages, latex2e, minor changes to previous version
null
null
null
null
We investigate Gromov-Witten invariants associated to exceptional classes for primitive birational contractions on a Calabi-Yau threefold X. It was observed in a previous paper that these invariants are locally defined, in that they can be calculated from knowledge of an open neighbourhood of the exceptional locus of the contraction; in this paper, we make this explicit. For Type I contractions (i.e. only finitely many exceptional curves), a method is given for calculating the Gromov-Witten invariants, and these in turn yield explicit expressions for the changes in the cubic form $D^3$ and the linear form $D.c_2$ under the corresponding flop. For Type III contractions (when a divisor E is contracted to a smooth curve C of singularities), there are only two relevant Gromov-Witten numbers n(1) and n(2). Here n(2) is the number (suitably defined) of simple pseudo-holomorphic rational curves representing the class of a fibre of E over C, and n(1) the number of simple curves representing half this class. Explicit formulae for n(1) and n(2) are given (n(1) in terms of the singular fibres of E over C and n(2)=2g(C)-2). An easy proof of these formulae is provided when g(C)>0. The main part of the paper then gives a proof valid in general (including the case g(C)=0).
[ { "version": "v1", "created": "Wed, 9 Jul 1997 16:46:00 GMT" }, { "version": "v2", "created": "Tue, 14 Oct 1997 16:58:47 GMT" }, { "version": "v3", "created": "Mon, 16 Mar 1998 17:44:10 GMT" }, { "version": "v4", "created": "Mon, 20 Jul 1998 09:38:18 GMT" } ]
2008-02-03T00:00:00
[ [ "Wilson", "P. M. H.", "" ] ]
alg-geom
\section{Introduction} In this paper, we investigate Gromov--Witten invariants associated to exceptional classes for primitive birational contractions on a Calabi--Yau threefold $X$. As already remarked in \cite{18}, these invariants are locally defined, in that they can be calculated from knowledge of an open neighbourhood of the exceptional locus of the contraction; intuitively, they are the numbers of rational curves in such a neighbourhood. In \S\ref{sec1}, we make this explicit in the case of Type~I contractions, where the exceptional locus is by definition a finite set of rational curves. Associated to the contraction, we have a flop; we deduce furthermore in Proposition~\ref{prop_1.1} that the changes to the basic invariants (the cubic form on $H^2(X,\mathbb Z)$ given by cup product, and the linear form given by cup product with the second Chern class $c_2$) under the flop are explicitly determined by the Gromov--Witten invariants associated to the exceptional classes. The main results of this paper concern the Gromov--Witten invariants associated to classes of curves contracted under a Type~III primitive contraction. Recall \cite{17} that a primitive contraction $\varphi\colon X\to\Xbar$ is of Type~III if it contracts down an irreducible divisor $E$ to a curve of singularities $C$. For $X$ a smooth Calabi--Yau threefold, such contractions were studied in \cite{18}; in particular, it was shown there that the curve $C$ is smooth and that $E$ is a conic bundle over $C$. We denote by $2\eta\in H_2(X,\mathbb Z)/\operatorname{Tors}$ the numerical class of a fibre of $E$ over $C$. In the case when $E$ is a $\mathbb P ^1$-bundle over $C$, this may in fact be a primitive class, and so the notation is at slight variance with that adopted in \S\ref{sec1}, where $\eta$ is assumed to be the primitive class. In the case when the class of a fibre is not primitive (for instance, when $E$ is not a $\mathbb P ^1$-bundle over $C$), the primitive class contracted by $\varphi$ will be $\eta$. We denote the Gromov--Witten numbers associated to $\eta$ and $2\eta$ by $n_1$ and $n_2$, with the convention that $n_1=0$ if $2\eta$ is the primitive class. The above conventions have been adopted so as to achieve consistency of notation for all Type~III contractions. If the genus $g$ of the curve $C$ is strictly positive, under a general holomorphic deformation of the complex structure on $X$, the divisor $E$ disappears leaving only finitely many of its fibres, and (except in the case of elliptic quasiruled surfaces, where all the Gromov--Witten invariants vanish) we have a Type~I contraction. The results of \S\ref{sec1} may then be applied to deduce the Gromov--Witten invariants associated to the classes $m\eta$ for $m>0$. These are all determined by the Gromov--Witten numbers $n_1$ and $n_2$, and explicit formulas for $n_1$ and $n_2$ are given in Proposition~\ref{prop_2.3}; in particular $n_2=2g-2$. The formulas for $n_1$ and $n_2$ remain valid also for $g=0$, although the slick proof given in Proposition~\ref{prop_2.3} for the case $g>0$ no longer works. The formula for $n_1$ is proved for all values of $g(C)$ by local deformation arguments in Theorem~\ref{thm_2.5}. Verifying that $n_2=-2$ in the case when $g(C)=0$ is rather more difficult, and involves the technical machinery of moduli spaces of stable pseudo\-holomorphic maps and the virtual neighbourhood method, as used in \cite{2,9} in order to construct Gromov--Witten invariants for general symplectic manifolds. In particular, we shall need a cobordism result from \cite{13}, which we show in Theorem~\ref{thm_3.1} applies directly in the case where no singular fibre of $E$ is a double line. The general case may be reduced to this one by making a suitable almost complex small deformation of complex structure. In \S\ref{sec4}, we give an application of our calculations. In \cite{18}, it was shown that if $X_1$, $X_2$ are Calabi--Yau threefolds which are symplectic deformations of each other (and general in their complex moduli), then their K\"ahler cones are the same. Now we can deduce (Corollary~\ref{cor4.1}) that corresponding codimension one faces of these cones have the same contraction type. The author thanks Yongbin Ruan for the benefit of conversations concerning material in \S\ref{sec3} and his preprint \cite{13}. \section{Flops and Gromov--Witten invariants}\label{sec1} If $X$ is a smooth Calabi--Yau threefold with K\"ahler cone $\mathcal K$, then the nef cone $\overline{\mathcal K}$ is locally rational polyhedral away from the cubic cone \[ W^*=\bigl\{ D\in \HR 2 X \ ; \ D^3=0\bigr\}; \] moreover, the codimension one faces of $\overline{\mathcal K}$ (not contained in $W^*$) correspond to primitive birational contractions $\varphi\colon X\to\Xbar$ of one of three different types \cite{17}. In the numbering of \cite{17}, Type~I contractions are those where only a finite number of curves (in fact $\mathbb P^1$s) are contracted. The singular threefold $\overline X$ then has a finite number of cDV singularities. Whenever one has such a small contraction on $X$, there is a flop of $X$ to a different birational model $X'$, also admitting a birational contraction to $\overline X$; moreover, identifying $\HR 2 {X'}$ with $\HR 2 {X}$, the nef cone of $X'$ intersects the nef cone of $X$ along the codimension one face which defines the contraction to $\overline X$ \cite{6, 7}. It is well known \cite{7} that $X'$ is smooth, projective and has the same Hodge numbers as $X$, but that the finer invariants, such as the cubic form on $\HZ 2 X$ given by cup product, and the linear form on $\HZ 2 X$ given by cup product with $c_2(X)=p_1(X)$, will in general change. Recall that, when $X$ is simply connected, these two forms along with $\HZ 3 X$ determine the diffeomorphism class of $X$ up to finitely many possibilities \cite{14}, and that if furthermore $H_2(X,\mathbb Z)$ is torsion free, this information determines the diffeomorphism class precisely \cite{16}. When the contraction $\varphi\colon X\to\Xbar$, corresponding to such a {\em flopping face} of $\overline{\mathcal K}$, contracts only isolated $\mathbb P^1$s with normal bundle $(-1,-1)$ (that is, $\overline X$ has only simple nodes as singularities), then it is a standard calculation to see how the above cubic and linear forms (namely the cup product $\mu\colon \HZ 2 X\to\mathbb Z$, and the form $c_2\colon \HZ 2 X\to\mathbb Z$) change on passing to $X'$ under the flop. Since any flop is an isomorphism in codimension one, we have natural identifications \[ \HR 2 {X'}\cong \operatorname{Pic}_{\mathbb R}(X')\cong \operatorname{Pic}_{\mathbb R}(X)\cong \HR 2 X. \] If we are in the case where the exceptional curves $C_1,\dots,C_N$ are isolated $\mathbb P^1$s with normal bundle $(-1,-1)$, and if we denote by $D'$ the divisor on $X'$ corresponding to $D$ on $X$, then \[ (D')^3=D^3-\sum(D\cdot C_i)^3 \quad \text{and}\quad c_2(X')\cdot D'=c_2(X)\cdot D+2\sum D\cdot C_i \ . \] This is an easy verification -- see for instance \cite{1}. \begin{prop}\label{prop_1.1} Suppose that $X$ is a smooth Calabi--Yau threefold, and $\varphi\colon X\to\Xbar$ is any Type~I contraction, with $X'$ denoting the flopped Calabi--Yau threefold. The cubic and linear forms $(D')^3$ and $D'\cdot c_2(X')$ on $X'$ are then explicitly determined by the cubic and linear forms $D^3$ and $D\cdot c_2(X)$ on $X$, and the $3$-point Gromov--Witten invariants $\Phi_A $ on $X$, for $A\in H_2(X,\mathbb Z)$ ranging over classes which vanish on the flopping face. \end{prop} \begin{rem} This is essentially the statement from physics that the A-model 3-point correlation function on $\mathcal K(X)$ may be analytically continued to give the A-model 3-point correlation function on $\mathcal K(X')$. \end{rem} \begin{pf} We use the ideas from \cite{18}; in particular, we know that on a suitable open neighbourhood of the exceptional locus of $\varphi$, there exists a small holomorphic deformation of the complex structure for which the exceptional locus splits up into disjoint $(-1,-1)$-curves (\cite{18}, Proposition~1.1). Let $A\in H_2(X,\mathbb Z)$ be a class with $\varphi_* A=0$. The argument from \cite{18}, Section~1 then shows how the Gromov--Witten invariants $\Phi_A(D,D,D)$ can be calculated from local information. Having fixed a K\"ahler form $\omega$ on $X$, a small deformation of the holomorphic structure on a neighbourhood of the exceptional locus may be patched together in a $C^{\infty}$ way with the original complex structure to yield an almost complex structure tamed by $\omega$, and the Gromov--Witten invariants can then be calculated in this almost complex structure. The Gromov Compactness Theorem is used in this argument to justify the fact that all of the pseudo\-holomorphic rational curves representing the class $A$ have images which are $(-1,-1)$-curves in the deformed local holomorphic structure. Here we also implicitly use the Aspinwall--Morrison formula for the contribution to Gromov--Witten invariants from multiple covers of infinitesimally rigid $\mathbb P^1$s, now proved mathematically by Voisin \cite{15}. So if $n(B)$ denotes the number of $(-1,-1)$-curves representing a class given $B$, then \[ \Phi_A(D,D,D)=(D\cdot A)^3 \sum_{kB=A} n(B)/k^3, \] where the sum is taken over all integers $k>0$ and classes $B\in H_2(X,\mathbb Z)$ such that $kB=A$. So if $H_2(X,\mathbb Z)$ is torsion free and $A$ is the primitive class vanishing on the flopping face, this says that \[ \Phi_{mA}(D,D,D)=(D\cdot A)^3 \sum_{d|m} n(dA)d^3. \] Recall that the Gromov--Witten invariants used here are the ones (denoted $\widetilde\Phi$ in \cite{12}) which count marked parametrized curves satisfying a perturbed pseudo\-holomorphicity condition. Knowledge of the numbers $n(A)$ for the classes A with $\varphi_* A=0$ determines the Gromov--Witten invariants $\Phi_{A}$ for classes A with $\varphi_*A=0$, and vice-versa. If we can now show that the local contributions to $(D')^3$ and $D'\cdot c_2(X')$ are well-defined and invariant under the holomorphic deformations of complex structure we have made locally, then the obvious formulas for them will hold. Let $\eta\in H_2(X,\mathbb Z)/\operatorname{Tors}$ be the primitive class with $\varphi_* \eta=0$ and $n_d$ denote the total number of $(-1,-1)$-curves on the deformation which have numerical class $d\eta$; the $n_d$ are therefore nonnegative integers (cf.\ \cite{10}, Remark~7.3.6). Then \begin{align} (D')^3&=D^3-(D\cdot\eta)^3\sum_{d>0}n_dd^3, \tag{2.1.1}\\ D'\cdot c_2(X')&=D\cdot c_2(X)+2(D\cdot\eta)\sum_{d>0}n_dd. \tag{2.1.2} \end{align} To justify the premise in the first sentence of the paragraph, the basic result needed is that of local conservation of number, as stated in \cite{3}, Theorem~10.2. For calculating the change in $D^3$ for instance, let $X$ now denote the neighbourhood of the exceptional locus of $\varphi$ and $\pi\colon \mathcal X\to B$ the small deformation under which the exceptional locus splits up into $(-1,-1)$-curves. So we have a regular embedding (of codimension six) \[ \renewcommand{\arraystretch}{1.3} \begin{matrix} \mathcal X &\hookrightarrow & \mathcal X \times \mathcal X \times \mathcal X &=\mathcal Y \\ \downarrow && \downarrow \\ B & \kern1.2em=\kern-1.2em & B \end{matrix} \] In order to calculate the triple products $D_1'\cdot D_2'\cdot D_3'$ from $D_1\cdot D_2\cdot D_3$ and the numbers $n_d$, we may assume {\em wlog} that the $D_i$ are very ample, and so in particular we get effective divisors $\mathcal D_1$, $\mathcal D_2$ and $\mathcal D_3$ on $\mathcal X /B$. Applying \cite{8}, Theorem~11.10, we can flop in the family $\mathcal X\to B$, hence obtaining a deformation $\mathcal X'\to B$ of the flopped neighbourhood $X'$. We wish to calculate the local contribution to $D_1'\cdot D_2'\cdot D_3'$; with the notation as in \cite{3}, Theorem~10.2, we have a fibre square \[ \renewcommand{\arraystretch}{1.3} \begin{matrix} \mathcal W & \longrightarrow & \mathcal D_1' \times \mathcal D_2'\times \mathcal D_3' \\ \downarrow && \downarrow \\ \mathcal X' &\longrightarrow & \mathcal X' \times \mathcal X'\times \mathcal X' \end{matrix} \] with $\operatorname{Supp}(\mathcal W)=\bigcap \operatorname{Supp}(\mathcal D_i')$. Furthermore, we may assume that the divisors $\mathcal D_i $ were chosen so that $\mathcal D_1\cap\mathcal D_2\cap\mathcal D_3$ has no points in $\mathcal X$, and so in particular $\mathcal W$ is proper over $B$. Letting $D_i'(t)$ denote the restriction of $\mathcal D_i'$ to the fibre $X_t'$, we therefore have a well-defined local contribution to $D_1'(t)\cdot D_2'(t)\cdot D_3'(t)$ (concentrated on the flopping locus of $X_t'$), which is moreover independent of $t\in B$. Thus by making the local calculation as in (7.4) of \cite{1}, we deduce that \[ D_1\cdot D_2\cdot D_3-D_1'\cdot D_2'\cdot D_3' =(D_1\cdot \eta)(D_2\cdot \eta)(D_3\cdot \eta) \sum_{d>0} n_d d^3 \] as required. The proof for $c_2\cdot D$ is similar. Here we consider the graph $\widetilde X \subset X \times X'$ of the flop, with $\pi_1\colon \widetilde X\to X $ and $\pi_2\colon \widetilde X\to X' $ denoting the two projections, and $E \subset \widetilde X$ the exceptional divisor for both $\pi_1$ and $\pi_2$. Then $\pi_2^*(T_{X'})\rest{\widetilde X \setminus E}=\pi_1^*(T_X)\rest{\widetilde X \setminus E} $, and so in particular $c_2(\pi_2^* T_{X'})-c_2(\pi_1^*T_X)$ is represented by a 1-cycle $Z$ on $E$. Suppose {\em wlog} that $D$ is very ample, and that $D'$ denotes the corresponding divisor on $X'$. Set $\pi_1^* D=\widetilde D$ and $ \pi_2^* D'=\widetilde D +F$, with $F$ supported on $E$. Then $c_2(X')\cdot D'=c_2(\pi_2^* T_{X'})\cdot(\widetilde D+F)$. Hence \[ c_2(X')\cdot D'-c_2(X)\cdot D=c_2(\pi_2^* T_{X'})\cdot F+Z\cdot \widetilde D=c_2((\pi_2^* T_{X'})\rest F)+(Z\cdot \widetilde D)_E \] where the right-hand side is purely local. Note the slight abuse of notation here that $F$ denotes also the fixed {\em scheme} for the linear system $ |\pi_2^* D' |$. Now taking $X$ to be a local neighbourhood of the flopping locus, and taking a small deformation $\mathcal X\to B$ as before, we obtain families $\mathcal X'$, $\widetilde \mathcal X $, $\mathcal D$, $\mathcal E$, $\mathcal F$ and $\mathcal Z$ over B (corresponding to $X'$, $\widetilde X$, $D$, $E$, $F$ and $Z$). For ease of notation, we shall use $\pi_1$ and $\pi_2$ also for the morphisms of families $\widetilde \mathcal X\to \mathcal X$, respectively $\widetilde \mathcal X\to \mathcal X'$. Applying \cite{3}, Theorem~10.2 to the family of vector bundles $(\pi_2^* T_{\mathcal X'/B})\rest{\mathcal F}$ on the scheme $\mathcal F$ over $B$ yields that $c_2((\pi_2^* T_{\mathcal X'/B})\rest{F_t})$ is independent of $t\in B$. Noting that $\widetilde\mathcal D \hookrightarrow \widetilde\mathcal X$ is a regular embedding, we apply the same theorem to the fibre square \[ \renewcommand{\arraystretch}{1.3} \begin{matrix} \widetilde\mathcal D \times_{\widetilde \mathcal X} \mathcal E & \longrightarrow & \mathcal E \\ \downarrow && \downarrow \\ \widetilde\mathcal D &\longrightarrow & \widetilde\mathcal X \end{matrix} \] and the cycle $\mathcal Z$ on $\mathcal E$. This yields that $(Z_t\cdot\widetilde D_t)_{E_t}$ on $E_t$ is independent of $t\in B$, where by definition \[ Z_t=c_2(\pi_2^* T_{\mathcal X'/B})\rest{X_t}-c_2(\pi_1^* T_{\mathcal X /B}) \rest{X_t}. \] Thus the local contribution to $D'(t)\cdot c_2(X_t')$ is well-defined and independent of $t$, and so we need only make the local calculation for generic $t$ (where the exceptional locus of the flop consists of disjoint $(-1,-1)$-curves). This calculation may be found in \cite{1}, (7.4). \end{pf} \begin{spec} There are reasons for believing that only the numbers $n_1$ and $n_2$ are nonzero, and hence that the Gromov--Witten invariants associated to classes $m\eta$ for $m>2$ all arise from multiple covers. If this speculation is true, then the changes under flopping to the cubic form and the linear form would be determined by these two integers, and conversely. \end{spec} \section{Type~III contractions and Gromov--Witten invariants}\label{sec2} The main results of this paper concern the Gromov--Witten invariants associated to classes of curves contracted under a Type~III primitive contraction. Recall \cite{17} that a primitive contraction $\varphi\colon X\to\Xbar$ is of Type~III if it contracts down an irreducible divisor $E$ to a curve of singularities $C$. For $X$ a smooth Calabi--Yau threefold, such contractions were studied in \cite{18}; in particular, it was shown there that the curve $C$ is smooth and that $E$ is a conic bundle over $C$. We denote by $2\eta\in H_2(X,\mathbb Z)/\operatorname{Tors}$ the numerical class of a fibre of $E$ over $C$. As explained in the Introduction, we denote by $n_1$ and $n_2$ the Gromov--Witten numbers associated to the classes $\eta$ and $2\eta$, where $n_1=0$ if $E$ is a $\mathbb P^1$-bundle over $C$. If the generic fibre of $E$ over $C$ is reducible (consisting of two lines, each with class $\eta$), then, except in two cases, it follows from the arguments of \cite{18}, \S4 that, by making a global holomorphic deformation of the complex structure, we may reduce down to the case where the generic fibre of $E$ over $C$ is irreducible. The two exceptional cases are: \begin{enumerate} \renewcommand{\labelenumi}{(\alph{enumi})} \item $g(C)=1$ and $E$ has no double fibres. \item $g(C)=0$ and $E$ has two double fibres. \end{enumerate} However, Case~(a) is an {\em elliptic quasi-ruled} surface in the terminology of \cite{18}, and hence disappears completely under a generic global holomorphic deformation. In particular, we know that all the Gromov--Witten invariants $\Phi_A$ are zero, for $A\in H_2(X,\mathbb Z)$ having numerical class $m\eta$ for any $m>0$. In Case~(b), $E$ is a nonnormal generalized del Pezzo surface $\overline\F_{3;2}$ of degree 7 (see \cite{18}). As argued there however, we may make a holomorphic deformation in a neighbourhood of $E$ so that $E$ deforms to a {\em smooth} del Pezzo surface of degree 7, and where the class $\eta$ is then represented by either of two `lines' on the del Pezzo surface (which are $(-1,-1)$-curves on the threefold); hence $n_1=2$. In fact, the smooth del Pezzo surface is fibred over $\mathbb P^1$ with one singular (line pair) fibre. The arguments we give below may be applied locally (more precisely with the global almost complex stucture obtained by suitably patching the local small holomorphic deformation on an open neighbourhood of $E$ with the original complex structure), and the Gromov--Witten invariants may be calculated as if the original contraction $\varphi$ had contracted such a smooth del Pezzo surface of degree 7. In particular, $n_1=2$ comes from the two components of the singular fibre (Theorem~\ref{thm_2.5}), and $n_2=-2$ is proved in \S\ref{sec3} (see also Remark~\ref{rem_2.4}). Let us therefore assume that the generic fibre of $E$ over $C$ is irreducible, and so in particular $E\to C$ is obtained from a $\mathbb P^1$-bundle over $C$ by means of blowups and blowdowns. Moreover $E$ itself is a conic bundle over $C$, and so its singular fibres are either line pairs or double lines. \begin{lem} In the above notation, $E$ has only singularities on the singular fibres of the map $E\to C$. When the singular fibre is a line pair, we have an $\rA{n}$ singularity at the point where the two components meet (we include here the possibility $n=0$ when the point is a smooth point of $E$). When the singular fibre is a double line, we have a $\rD{n}$ singularity on the fibre (here we need to include the case $n=2$, where we in fact have two $\rA1$ singularities, and $n=3$, where we have an $\rA3$ singularity). \end{lem} \begin{pf} The proof is obvious, once the correct statement has been found. The statement of this result in \cite{17} omits (for fibre a double line) the cases $\rD{n}$ for $n>2$. \end{pf} \begin{lem}\label{lem_2.2} Suppose that $E\to C$ as above has $a_r$ fibres which are line pairs with an $\rA{r}$ singularity and $b_s$ fibres which are double lines with a $\rD{s}$ singularity (for $r\ge0$ and $s\ge2$), then \[ K_E^2=8(1-g)-\sum_{r\ge 0} a_r (r+1)-\sum_{s\ge 2} b_s s, \] where $g$ denotes the genus of $C$. \end{lem} This enables us to give a slick calculation of the Gromov--Witten invariants when the base curve has genus $g>0$. In this case, it was shown in \cite{17} that for a generic deformation of $X$, only finitely many fibres from $E$ deform, and hence the Type~III contraction deforms to a Type~I contraction. Thus Gromov--Witten numbers $n_1$ and $n_2$ may be defined as in Section 1, and are nonnegative integers. \begin{prop}\label{prop_2.3} When $g>0$, we have \[ n_1=2\sum_{r\ge 0} a_r (r+1)+2\sum_{s\ge 2} b_s s \quad \text{and} \quad n_2=2g-2. \] \end{prop} \begin{pf} We take a generic 1-parameter deformation of $X$, for which the Type~III contraction deforms to a Type~I contraction. We therefore have a diagram \[ \renewcommand{\arraystretch}{1.3} \begin{matrix} \mathcal X &\longrightarrow &\overline\mathcal X \\ \downarrow && \downarrow \\ \Delta &=& \Delta \end{matrix} \] where $\Delta\subset\mathbb C$ denotes a small disc. Since the singular locus of $\overline\mathcal X $ consists only of curves of cDV singularities, we may again apply \cite{8}, Theorem~11.10 to deduce the existence of a (smooth) flopped fourfold $\mathcal X'\to \overline\mathcal X$. The induced family $\mathcal X'\to\Delta$ is given generically by flopping the fibres, and at $t=0$ it is easily checked that $X_0'\cong X_0$; this operation is often called an {\em elementary transformation} on the family. Identifying the groups $H^2 (X_t,\mathbb Z)\cong H^2 (X_t',\mathbb Z)$ as before, this has the effect (at $t=0$) of sending $E$ to $-E$ (cf.\ the discussion in \cite{5}, \S3.3). So if $E'$ denotes the class in $H^2(X_t',\mathbb Z)$ corresponding to the class $E$ in $H^2(X_t,\mathbb Z)$, we have $(E')^3=-E^3$. For $t\ne 0$, we just have a flop, and so $(E')^3$ can be calculated from equation (2.1.1), namely $(E')^3=E^3+n_1+8n_2$. Therefore, using Lemma~\ref{lem_2.2} \[ n_1+8n_2=-2E^3=16(g-1)+2\sum_{r\ge0}a_r(r+1)+2\sum_{s\ge2}b_ss. \] Similarly, we have $c_2(X')\cdot E'=- c_2(X)\cdot E$, and so from equation (2.1.2) it follows that $2n_1+4n_2=2 c_2\cdot E$. An easy calculation of the right-hand side then provides the second equation \[ 2n_1+4n_2=8(g-1)+4\sum_{r\ge0}a_r(r+1)+4\sum_{s\ge2}b_ss. \] Solving for $n_1$ and $n_2$ from these two equations gives the desired result. \end{pf} \begin{rem}\label{rem_2.4} This result remains true even when $g=0$, although the slick proof given above is no longer valid. The formula for $n_1$ is checked in Theorem~\ref{thm_2.5} by local deformation arguments (for which the genus $g$ is irrelevant), showing that the contribution to $n_1$ from a line pair fibre with $\rA{r}$ singularity is $2(r+1)$, and from a double line fibre with $\rD{s}$ singularity is $2s$. Let $A\in H_2(X,\mathbb Z)$ denote the class of a fibre of $E\to C$. Observe that any pseudo\-holomorphic curve representing the numerical class $\eta$ will be a component of a singular fibre of $E\to C$. Moreover, the components $l$ of a singular fibre represent the same class in $H_2(X,\mathbb Z)$, and so in particular twice this class is $A$. Thus the Aspinwall--Morrison formula (as proved in \cite{15}) yields the contribution to the Gromov--Witten invariants $\Phi_A (D,D,D)$ from double covers, purely in terms of $n_1$ and $D\cdot A$. The difference may be regarded as the contribution to $\Phi_A (D,D,D)$ from simple maps, and taking this to be $n_2 (D\cdot A)^3$ determines the number $n_2$ (in \S\ref{sec3}, we shall see how $n_2$ may be determined directly from the moduli space of simple stable holomorphic maps). If $g>0$, the above argument shows that this is in agreement with our previous definition, and yields moreover the equality $n_2=2g-2$. The fact that $n_2=-2$ when $g=0$ requires a rather more subtle argument involving technical machinery -- see Theorem~\ref{thm_3.1}. I remark that the value $n_2=-2$ is needed in physics, and that there is also a physics argument justifying it (see \cite{4}, \S5.2 and \cite{5}, \S3.3) -- essentially, it comes down to a statement about the A-model 3-point correlation functions. In \S\ref{sec3} below, we give a rigorous mathematical proof of the assertion. \end{rem} \begin{thm}\label{thm_2.5} The formula for $n_1$ in Proposition~\ref{prop_2.3} is valid irrespective of the value of the genus $g=g(C)$. \end{thm} \begin{pf} By making a holomorphic deformation of the complex structure on an open neighbourhood $U$ in $X$ of the singular fibre $Z$ of $E\to C$, we may calculate the contribution to $n_1$ from that singular fibre -- see \cite{18}, (4.1). The deformation of complex structure is obtained as in \cite{18} by considering the one dimensional family of Du Val singularities in $\overline X$, and deforming this family locally in a suitable neighbourhood $\overline U$ of the dissident point. Our assumption is that the family $\overline U\to\Delta$ has just an $\rA1$ singularity on $\overline U_t$ for $t\ne 0$, and we may assume also that $\overline U\to \Delta$ is a good representative (in the sense explained in \cite{18}). The open neighbourhood $U$ is then the blowup of $\overline U$ in the smooth curve of Du Val singularities (\cite{18}, p.\ 569). The contribution to $n_1$ may be calculated locally, and will not change when we make small holomorphic deformations of the complex structure on $U$, which in turn corresponds to making small deformations to the family $\overline U\to\Delta$. First we consider the case where the singular fibre $Z$ is a line pair -- from this, it will follow that the dissident singularity on $\overline U$ is a $\rcA{n}$ singularity with $n>1$, and that $\overline U$ has a local analytic equation of the form \[ x^2+y^2+z^{n+1}+tg(x,y,z,t)=0 \] in $\mathbb C^3\times\Delta$ (here $t$ is a local coordinate on $\Delta$, and $x=y=z=0$ the curve $C$ of singularities). For $t\ne 0$, we have an $\rA1$ surface singularity, which implies that $g$ must contain a term of the form $t^r z^2$ for some $r\ge 0$. By an appropriate analytic change of coordinates, we may then assume that $\overline U$ has a local analytic equation of the form \[ x^2+y^2+z^{n+1}+t^{r+1}z^2+t h(x,y,z,t)=0, \] where $h$ consists of terms which are at least cubic in $x,y,z$. By making a small deformation of the family $\overline U\to \Delta$, we may reduce to the case $n=2$, that is, $\overline U$ having local equation $x^2+y^2+z^3+t^{r+1}z^2+th=0$. At this stage, we could in fact also drop the term $th$ (an easy check using the versal deformation family of an $\rA2$ singularity), but this will not be needed. We now make a further small deformation to get $\overline U_{\varepsilon} \subset \mathbb C^3 \times \Delta $ given by a polynomial \[ x^2+y^2+z^3+t^{r+1} z^2+\varepsilon z^2+th=x^2+y^2+z^2 (z+t^{r+1}+\varepsilon)+th \ . \] This then has $r+1$ values of $t$ for which the singularity is an $\rA2$ singularity -- for other values of $t$, it is an $\rA1$ singularity. If we blow up the singular locus of $\overline U_{\varepsilon}$, we therefore obtain a smooth exceptional divisor for which $r+1$ of the fibres over $\Delta$ are line pairs. By the argument of \cite{18}, (4.1), this splitting of the singular fibre into $r+1$ line pair singular fibres of the simplest type can be achieved by a local holomorphic deformation on a suitable open neighbourhood of the fibre in the original threefold $X$. It is however clear that a line pair coming from a dissident $\rcA2$ singularity of the above type contributes precisely two to the Gromov--Witten number $n_1$ -- one for each line in the fibre. In terms of equations, we have a local equation for $\overline X$ of the form $x^2+y^2+z^3+w z^2=0$; deforming this to say $x^2+y^2+z^3+w z^2+\varepsilon w=0$, we get two simple nodes, and hence two disjoint $(-1,-1)$-curves on the resolution. The argument of \cite{18}, (4.1) shows that the Gromov--Witten number $n_1$ may be calculated purely from these local contributions, and so the total contribution to $n_1$ from the line pair singular fibre of $E$ with $\rA{r}$ singularity is indeed $2(r+1)$, as claimed. For the case of the singular fibre $Z$ of $E$ being a double line, the dissident singularity must be $\rcE6$, $\rcE7$, $\rcE8$, or $\rcD{n}$ for $n\ge4$. Thus $\overline U$ has a local analytic equation of the form $f(x,y,z)+tg(x,y,z,t)$ in $\mathbb C^3\times\Delta$ for $f$ a polynomial of the appropriate type ($t$ a local coordinate on $\Delta$, and $x=y=z=0$ the curve of singularities). To simplify matters, we may deform $f$ to a polynomial defining a $\rD4$ singularity, and hence make a small deformation of the family to one in which the dissident singularity is of type $\rcD4$. We then have a local analytic equation of the form \[ x^2+y^2 z+z^3+t g(x,y,z,t)=0. \] For $t\ne 0$, we have an $\rA1$ singularity, and so the terms of $g$ must be at least quadratic in $x,y,z$. Moreover, by changing the $x$-coordinate, we may take the equation to be of the form \[ x^2+y^2 z+z^3+t^a y^2+t^b yz+t^c z^2+t h(x,y,z,t)=0, \] with $a,b,c$ positive, and where the terms of $h$ are at least cubic in $x,y,z$. The fact that the blowup $U$ of $\overline U$ in $C$ is smooth is easily checked to imply that $a=1$. Since \[ ty^2+2 t^b yz=t(y+t^{b-1}z)^2-t^{2b-1}z^2, \] we have an obvious change of $y$-coordinate which brings the equation into the form \[ x^2+y^2 z+z^3 +t y^2+t^r z^2+t h_1 (x,y,z,t)=0, \] where $r=\operatorname{min} \{ c, 2b-1 \}$ and $h_1$ has the same property as $h$. When we blow up $\overline U$ along the curve $x=y=z=0$, we obtain an exceptional locus $E$ with a double fibre over $t=0$, on which we have a $\rD{r+1}$ singularity (including the case $r=1$ of two $\rA1$ singularities, and $r=2$ of an $\rA3$ singularity). Moreover, this was also true of our original family, since the small deformation of $f$ we made did not affect the local equation of the exceptional locus. Moreover, by adding a term $\varepsilon_1 y^2+\varepsilon_2 z^2$, we may deform our previous equation to one of the form \[ x^2+y^2(z+t+\varepsilon_1)+z^2(z+t^r+\varepsilon_2)+th_1(x,y,z,t)=0. \] When $ t+\varepsilon_1=0$, we have an $\rA3$ singularity, and when $t^r+\varepsilon_2=0$, an $\rA2$ singularity. Moreover, when we blow up the singular locus of this deformed family, the resulting exceptional divisor is smooth and has line pair fibres for these $r+1$ values of $t$. Thus, as seen above, the contribution to $n_1$ from the original singular fibre (a double line with a $\rD{r+1}$ singularity) is $2(r+1)$ as claimed. \end{pf} \section{Calculation of $n_2$ for Type~III contractions}\label{sec3} Let $\varphi\colon X\to\Xbar $ be a Type~III contraction on a Calabi--Yau threefold $X$, which contracts a divisor $E$ to a (smooth) curve $C$ of genus $g$. When $g>0$, it was proved in Proposition~\ref{prop_2.3} that the Gromov--Witten number $n_2$ (defined for arbitrary genus via Remark~\ref{rem_2.4}) is $2g-2$. The purpose of this Section is to extend this result to include the case $g=0$ ($C$ is isomorphic to $\mathbb P^1$), and to prove $n_2=2g-2$ in general. Arguing as in \cite{18}, it is clear that the desired result is a local one, depending only on a neighbourhood of the exceptional divisor $E$. As remarked in \S\ref{sec2}, we may then always reduce down to the case that the generic fibre of $E\to C$ is irreducible. If all the fibres of $E\to C$ are smooth (so $E$ is a $\mathbb P^1$-bundle over $C$), the fact that $n_2=2g-2$ was proved in Proposition 5.7 of \cite{11}, using a cobordism argument. This latter result was extended by Ruan in \cite{13}, Proposition~2.10, using the theory of moduli spaces of stable maps and the virtual neighbourhood technique (cf.~\cite{2,9}). If the singular fibres of $E\to C$ are line pairs, Ruan's result applies directly. We prove below that the linearized Cauchy--Riemann operator has constant corank for the stable (unmarked) rational curves given by the fibres of $E$ over $C$, and hence by Ruan's result that there is an obstruction bundle $\mathcal H$ on $C$, with $n_2$ determined by the Euler class of $\mathcal H$. By Dolbeault cohomology, there is a natural identification of $\mathcal H$ with the cotangent bundle $T_C^*$ on $C$, and hence the formula for $n_2$ follows. We note however that for Ruan's result to hold, we do not need an integrable almost complex structure on $X$. Provided we have a natural identification between the cokernel of the linearized Cauchy--Riemann operator and the cotangent space at the corresponding point of $C$, we can still deduce that $n_2=2g-2$. In the general case of a Type~III contraction which has double fibres, we show below that we can make a small local deformation of the almost complex structure on $X$ so that $E$ deforms to a family of pseudo\-holomorphic rational curves over $C$ with at worst line pair singular fibres, and for which Ruan's method applies. \begin{thm}\label{thm_3.1} For any Type~III contraction $\varphi\colon X\to\Xbar $, the Gromov--Witten number $n_2=2g-2$. \end{thm} \begin{pf} We saw above that we may assume that the generic fibre of $E\to C$ is irreducible. Furthermore, we initially assume also that the singular fibres are all line pairs, and later reduce the general case to this one. We let $J$ denote the almost complex structure on $X$, which we know is integrable (at least in a neighbourhood of $E$), and tamed by a symplectic form $\omega$. Let $A\in H_2(X,\mathbb Z)$ be the class of a fibre of $E\to C$. Adopting the notation from \cite{13}, we consider the moduli space $\overline{\mathcal M}_A(X,J)=\overline{\mathcal M}_A(X,0,0,J)$ of stable unmarked rational holomorphic maps, a compactification of the space of (rigidified) pseudo\-holomorphic maps $\mathbb C\mathbb P^1\to X$, representing the class $A$. The theory of stable maps, as explained in Section 3 of \cite{13}, goes through for unmarked stable maps, by taking each component of the domain as a bubble component, and adding marked points (in addition to the double points) as in \cite{13} in order to stabilize the components (thus taking a local slice of the automorphism group). In the case that all the singular fibres of $E\to C$ are line pairs, $\overline{\mathcal M}_A (X,J)$ has two components, one corresponding to simple maps and the other to double covers. It is now a simple application of Gromov compactness to see that these two components are disjoint, since a sequence of double cover maps cannot converge to a simple map. A similar argument will show that for all almost complex structures $J_t$ in some neighbourhood of $J=J_0$, the moduli space $\overline{\mathcal M}_A (X,J_t)$ will consist of two disjoint components, one corresponding to the simple maps and the other to the double covers. Since any stable unmarked rational holomorphic map must be an embedding, it is clear that the component $\overline{\mathcal M}'_A (X,J)$ corresponding to the simple maps can be identified naturally with the smooth base curve $C$, and that for all almost complex structures in some neighbourhood of $J=J_0$, the moduli space $\overline{\mathcal M}'_A (X,J_t)$ of simple unmarked stable holomorphic maps is compact. The Gromov--Witten invariant $n_2$ that we seek can then be defined via Ruan's virtual neighbourhood invariant $\mu_{\mathcal S}$, and may be evaluated on $(X,J)$ by using \cite{13}, Proposition~2.10. Let us now go into more details of this. We consider $C^{\infty}$ stable maps $f\in\overline B_A (X)=\overline B_A (X,0,0)$ in the sense of \cite{13}, Definition 3.1, where Ruan shows later in the same Section that the naturally stratified space $\overline B_A (X)$ satisfies a property which he calls {\em virtual neighbourhood technique admissable} or {\em VNA}, and as he says, for the purposes of the virtual neighbourhood construction, behaves as if it were a Banach $V$-manifold. Since any simple marked holomorphic stable map $f$ in $\overline{\mathcal M}'_A (X,J)$ is forced to be an embedding, we may restrict our attention to $C^{\infty}$ stable maps whose domain $\Sigma$ comprises at most two $\mathbb P^1$s. We stratify $\overline B_A (X)$ according to the combinatorial type $D$ of the domain $\Sigma$. Thus any $f\in \overline{\mathcal M}'_A (X,J)$ belongs to one of two strata of $\overline B_A (X)$. In general, for $k$-pointed $C^{\infty}$ stable maps of genus $g$, Ruan shows that for any given combinatorial type $D$, the substratum $ B_D (X,g,k)$ is a Hausdorff Frechet V-manifold (\cite{13}, Proposition 3.6). As mentioned above, he needs to add extra marked points in order to stabilize the nonstable components of the domain $\Sigma$, thus taking a local slice of the action of the automorphism group on the unstable marked components of $\Sigma$. Moreover, the tangent space $T_f B_D (X,g,k)$ is identified with $\Omega^0 (f^* T_X)$, as defined in his equation \cite{13}, (3.29). The tangent space $T_f \overline B_A (X,g,k)$ can then be defined as $T_f B_D (X,g,k) \times \mathbb C_f $, where $\mathbb C_f$ is the space of gluing parameters (see \cite{13}, equation before (3.67)). In our case, however, things are a bit simpler. Given $f\in \overline{\mathcal M}'_A (X,J)$ with domain $\Sigma$ consisting of two $\mathbb P^1$s, the tangent space $T_f \overline B_A (X)$ is of the form $\Omega^0 (f^* T_X) \times \mathbb C$, and we have a neighbourhood $\widetilde U_f$ of $f$ in $\overline B_A (X)$ defined by \cite{13}, (3.43), consisting of stable maps $\overline f^{v,w}$ parametrized locally by \[ \bigl\{w\in\Omega^0(f^*T_X)\ ;\ \|w\|_{C^1}<\varepsilon'\bigr\} \] (corresponding to deformations within the stratum $B_D(X)$), and by $v\in\mathbb C_f^\varepsilon $ (an $\varepsilon$-ball in $\mathbb C_f=\mathbb C$ giving the gluing parameter at the double point). This then corresponds to the above decomposition of $T_f\overline B_A (X)$ into two factors. On the first factor, the linearization $D_f\overline\partial_J $ of the Cauchy--Riemann operator restricts to \[ \overline\partial_{J,f}\colon\Omega^0(f^*T_X)\to\Omega^{0,1}(f^*T_X) \] in the notation of \cite{13}. The index of this operator may be calculated using Riemann--Roch on each component of $\Sigma$ (cf.~the proof of Lemma 3.16 in \cite{13}, suitably modified to take account of the extra marked points), and is seen to be $-2$. Let us now consider the stable maps $f^v=f^{v,0}$ for $v\in\mathbb C_f^\varepsilon\setminus\{0\}$. These are stable maps $\mathbb C\mathbb P^1\to X$ which differ from $f$ only in small discs around the double point, and in this sense are approximately holomorphic. Set $v=r e^{i\theta}$; then the gluing to get $f^v\colon \Sigma^v\to X$ is only performed in discs around the double point of radius $2r^2 /\rho$ in the two components ($\rho$ a suitable constant). It can then be checked for any $2<p<4$ that $\| \overline\partial_J (f^v)\|_{L^p_1} \le Cr^{4/p}$ (see \cite{13} Lemma 3.23, and \cite{10} Lemma A.4.3), from which it follows that the linearization \[ L_A=D_f \overline\partial_J \] of the Cauchy--Riemann operator should be taken as zero on the factor $\mathbb C_f$ in $T_f \overline B_A (X)$. Thus we deduce that the index of $L_A$ is zero, and that $\operatorname{coker} L_A$ is same as the cokernel of $\overline\partial_{J,f}\colon \Omega^0 (f^* T_X)\to \Omega^{0,1} (f^* T_X)$, which by Dolbeault cohomology may be identified as \[ H^1(f^*T_X)=H^1(Z,T_X{}\rest Z), \] where $Z$ is the fibre of $E\to C$ (over a point $x\in C$) corresponding to the image of $f$. We note that these are exactly the same results as are obtained in the smooth case, when $\Sigma$ consists of a single $\mathbb P^1$. Here, we need to add three marked points to stabilize $\Sigma$, and Riemann--Roch then gives immediately that the index of $L_A$ is zero. Observe that $Z$ is a complete intersection in $X$, and so for our purposes is as good as a smooth curve. Via the obvious exact sequence, $H^1(T_X{}\rest Z)$ may be naturally identified with $H^1(N_{Z/X})$, which in turn may be naturally identified with $H^0 (N_{Z/X})^*$ (since $K_Z=\bigwedge^2 N_{Z/X}$, we have a perfect pairing $H^0 (N_{Z/X}) \times H^1(N_{Z/X})\to H^1(K_Z)\cong \mathbb C$). Observing that $H^0 (N_{Z/X})=H^0 (\mathcal O_Z\oplus\mathcal O_Z (E))\cong \mathbb C$, we know that $\operatorname{coker} L_A $ has complex dimension one and is naturally identified with $T^*_{C,x}$, the dual of the tangent space at $x$ to the Hilbert scheme component $C$. This we have seen is true for all $f\in\overline{\mathcal M}'_A(X,J)$. We now apply \cite{13}, Proposition~2.10, (2) to our set-up, where $C=\overline{\mathcal M}'_A (X,J)=\mathcal M_{\mathcal S}=\mathcal S^{-1}(0)$ for $\mathcal S$ the Cauchy--Riemann section of $\overline \mathcal F_A(X)$ (as constructed in \cite{13}, \S3) over a suitable neighbourhood of $\mathcal M_{\mathcal S}$ in $\overline B_A(X)$. The above calculations verify that the conditions of Proposition~2.10, (2) are satisfied, with $\ind(L_A)=0$, $\operatorname{dim}(\operatorname{coker} L_A)=2$ and $\operatorname{dim} (\mathcal M_{\mathcal S})=2$. Moreover, we deduce that the obstruction bundle $\mathcal H$ on $\mathcal M_{\mathcal S}$ is just the cotangent bundle $T_C^*$ on $C$. The Gromov--Witten number $n_2$ may then be defined to be $\mu_{\mathcal S}(1)$. It follows from the basic Theorem~4.2 from \cite{13} that this is independent of any choice of tamed almost complex structure and is a symplectic deformation invariant. Thus by considering a small deformation of the almost complex structure and using \cite{13}, Proposition~2.10, (1), it is the invariant $n_2$ that we seek. Applying Ruan's crucial Proposition~2.10, (2), the invariant can be expressed as \[ \mu_{\mathcal S}(1)=\int_{\mathcal M'_A (X,J)} e(T_C^*), \] from which it follows that $n_2=2g-2$ as claimed. The general case (where $E\to C$ also has double fibres) can now be reduced to the case considered above. Suppose we have a point $Q\in C$ for which the corresponding fibre is a double line. We choose an open disc $\Delta \subset C$ with centre $Q$, and a neighbourhood $U$ of $Z$ in $X$, with $U$ fibred over $\Delta$, the fibre $U_0$ over $Q$ containing the fibre $Z$. Letting $\overline U\to \Delta$ denote the image of $U$ under $\varphi$, a family of surface Du Val singularities, we make a small deformation $\overline\mathcal U\to \Delta'$ of $\overline U$, as in the proof of Theorem~\ref{thm_2.5} of this paper, and in this way obtain a holomorphic deformation $\mathcal U\to \Delta'$ of $U$ under which $E_0=E\rest{\Delta}$ deforms to a family of surfaces $E_t$ ($t\in \Delta'$), all fibred over $\Delta$, and with at worst line pair singular fibres for $t\ne 0$. Considering $\overline\mathcal U\to \Delta \times \Delta'$ as a two parameter deformation of the surface singularity $ \overline U_0$, we may take a good representative and apply Ehresmann's fibration theorem (with boundary) to the corresponding resolution $\mathcal U\to\Delta\times\Delta'$ (cf.\ \cite{18}, proof of Lemma 4.1). In this way, we may assume that $\mathcal U\to \Delta \times \Delta'$ is differentiably trivial over the base. In particular, the family $\mathcal U\to\Delta'$ is also differentiably trivial, and hence determines a holomorphic deformation of the complex structure on a fixed neighbourhood $U$ of $Z$, where $U\to \Delta$ is also differentiably trivial. We perform this procedure for each singular fibre $Z_1,\dots,Z_N $ of $E\to C$, obtaining, for each $i$, an open neighbourhood $U_i$ of $Z_i$ fibred over $\Delta_i \subset C$, and a holomorphic complex structure $J_i$ on $U_i$ with the properties explained above (of course, if $Z_i$ is a line pair, we may take $J_i$ to be the original complex structure $J$). Let $\frac{1}{2}\overline\Delta_i$ denote the closed subdisc of $\Delta_i$ with half the radius, $C^*=C\setminus\bigcup_{i=1}^N\frac{1}{2}\overline\Delta_i$, and $E^*=E\rest{C^*}\to C^*$ the corresponding open subset of $E$. We then take a tubular neighbourhood $U^*\to C^*$ of $E^*\to C^*$, equipped with the original complex structure $J$. By taking deformations to be sufficiently small and shrinking radii of tubular neighbourhoods if necessary, all these different complex structures may be patched together in a $C^{\infty}$ way (tamed by the symplectic form) over the overlaps in $C$. In this manner, we obtain an open neighbourhood $W$ of $E$ in $X$, and a tamed almost complex structure $J'$ on $W$, which is a small deformation of the original complex structure $J$ and which satisfies the following properties: \begin{enumerate} \renewcommand{\labelenumi}{(\alph{enumi})} \item Each singular fibre $Z_i$ of $E\to C$ has an open neighbourhood $U_i \subset W$ fibred over $\Delta_i \subset C$ with $J'$ inducing an integrable complex structure on each fibre (thus $U_i\to \Delta_i$ is a $C^{\infty}$ family of holomorphic surface neighbourhoods). \item The almost complex structure $J'$ is integrable in a smaller neighbourhood $U'_i \subset U_i$ of each singular fibre, with the corresponding family $U'_i\to \Delta'_i$ being holomorphic. \item On the complement of $\bigcup U_i$ in $W$, the almost complex structure $J'$ coincides with the original complex structure $J$. \item $E$ deforms to a $C^{\infty}$ family of pseudo\-holomorphic rational curves $E'\to C$ in $(W,J')$, with generic fibre $\mathbb C\mathbb P^1$ and the only singular fibres being line pairs. Moreover, we may assume that any such singular fibre is contained in one of the above open sets $U'_i$. \end{enumerate} Of course, we may now patch $J'$ on $W$ with the original complex structure $J$ on $X$ to get a global tamed almost complex structure on $X$, which we shall also denote by $J'$. Provided we have taken our deformations sufficiently small, the standard argument via Gromov compactness ensures that any pseudo\-holomorphic stable map representing the class $A$ has image contained in a fibre of $E'\to C$. The theory of \cite{13} applies equally well to almost complex structures, and hence to our almost complex manifold $X'$ with complex structure $J'$. Clearly, all the calculations remain unchanged for stable maps whose image (a fibre of $E'\to C$) has a neighbourhood on which $J'$ is integrable, and in particular this includes all the singular fibres. Suppose therefore that $f\colon \mathbb C\mathbb P^1\to X'$ is a pseudo\-holomorphic rational curve whose image $Z$ is contained in an overlap $U_i \setminus U'_i$ (where $J'$ may be nonintegrable). The linearized Cauchy--Riemann operator $L_A$ still has index zero, since by the argument of \cite{10}, p.~24, the calculation via Riemann--Roch continues to give the correct value. We therefore need to show that $\operatorname{coker} L_A$ is still identified naturally as $T^*_{C,x}$, and hence that the obstruction bundle is $\mathcal H=T^*_C$ as before. Setting $U=U_i$ and $\Delta=\Delta_i$, we know that $U\to \Delta$ is locally (around the image $Z$ of $f$) a $C^{\infty}$ family of holomorphic surface neighbourhoods. Moreover, the linearized Cauchy--Riemann operator $L_A = D_f\colon C^{\infty} (f^*T_U)\to \Omega^{0,1} (f^*T_U)$ fits into the following commutative diagram (with exact rows) \[ \renewcommand{\arraystretch}{1.3} \begin{matrix} 0&\to&C^{\infty}(f^*T_{U/\Delta})&\to &C^{\infty}(f^*T_U)&\to&C^{\infty}(g^*T_{\Delta})&\to&0 \\ &&\Bigm\downarrow\overline\partial_{f}&&\Bigm\downarrow D_f&&\Bigm\downarrow&& \\ 0&\to&\Omega^{0,1}(f^*T_{U/\Delta})&\to &\Omega^{0,1}(f^*T_U)&\to&\Omega^{0,1}(g^*T_{\Delta})&\to&0 \end{matrix} \] where $g$ is the constant map on $\mathbb C\mathbb P^1$ with image the point $x\in \Delta$, and where the fibre of $E'$ over $x$ is $Z$. Let us denote by $U_x$ the corresponding holomorphic surface neighbourhood, the fibre of $U$ over $x$. The cokernel of \[ \overline\partial_{f}\colon C^{\infty}(f^*T_{U/\Delta})\to\Omega^{0,1}(f^*T_{U/\Delta}) \] is then naturally identified via Dolbeault cohomology with $H^1(T_{U_x}{}\rest Z)\cong H^1(N_{Z/U_x})$. This latter space is in turn naturally identified with $H^1(N_f)\cong H^0(N_f)^*\cong T^*_{C,x}$. I claim now that $J'$ may be found as above for which $\operatorname{coker} L_A$ has the correct dimension (namely real dimension two) for all fibres of $E'\to C$. Since $L_A$ has index zero and $\ker L_A$ has dimension at least two, we need to show that that the dimension of $\operatorname{coker} L_A$ is not more than two. This follows by a Gromov compactness argument. Suppose that the dimension is too big for some fibre of $E'\to C$, however close we take $J'$ to $J$. We can then find sequences of almost complex structures $J'_\nu$ (with the properties (a)--(d) described above) converging to $J=J_0$, and pseudo\-holomorphic rational curves $f_\nu\colon\mathbb C\mathbb P^1\to (X, J'_\nu)$ at which $\operatorname{coker} L_A$ has real dimension $>2$. By construction, the image of such a map is not contained in any $U'_i$ (since $J'_\nu$ would then be integrable on some neighbourhood of the image, and then we know that $\operatorname{coker} L_A$ has the correct dimension). Thus the image of $f_\nu$ has nontrivial intersection with the compact set $X \setminus \bigcup U'_i$. By Gromov compactness, the $f_\nu$ may be assumed to converge to a pseudo\-holomorphic rational curve on $(X,J)$ whose image is not contained in any $U'_i$. This is therefore just an embedding $f\colon\mathbb C\mathbb P^1\to (X, J)$ of some smooth fibre of $E\to C$, at which we know that $\operatorname{coker} L_A$ has real dimension precisely two; this then is a contradiction. A similar argument, via Gromov compactness, then yields the fact that $J'$ may be found as above such that the linear map $\operatorname{coker} (\overline\partial_{f})\to \operatorname{coker} (D_f)$ is an isomorphism for all smooth fibres of $E'\to C$, since this is true for all the smooth fibres of $E\to C$ on $(X,J)$. For such a $J'$, we deduce that $\operatorname{coker} L_A$ is naturally identified with $T^*_{C,x}$ for all fibres, and hence the obstruction bundle identified as $T^*_C$. The previous argument may then be applied directly with the almost complex structure $J'$, showing that the symplectic invariant $n_2$ is $2g-2$ in general. The proof of Theorem~\ref{thm_3.1} is now complete. \end{pf} \section{An application to symplectic deformations of Calabi--Yaus} \label{sec4} If $X$ is a Calabi--Yau threefold which is general in moduli, we know that any codimension one face of its nef cone $\overline{\mathcal K} (X)$ (not contained in the cubic cone $W^*$) corresponds to a primitive birational contraction $\varphi\colon X\to\Xbar $ of Type~I, II or $\text{III}_0$, where Type~$\text{III}_0$ denotes a Type~III contraction for which the genus of the curve $C$ of singularities on $\overline X$ is zero. In \cite{18}, we studied Calabi--Yau threefolds which are symplectic deformations of each other. One of the results proved there (Theorem~2) said that if $X_1$ and $X_2$ are Calabi--Yau threefolds, general in their complex moduli, which are symplectic deformations of each other, then their K\"ahler cones are the same. The proof of this essentially came down to showing that certain Gromov--Witten invariants associated to exceptional classes were nonzero. Using the much more precise information obtained in this paper, we are able to make a stronger statement. \begin{cor}\label{cor4.1} With the notation as above, any codimension one face (not contained in $W^*$) of $\overline{\mathcal K} (X_1)=\overline{\mathcal K} (X_2)$ has the same contraction type (Type~I, II or $\text{\sl III}_0$) on $X_1$ as on $X_2$. \end{cor} \begin{pf} The fact that Type~II faces correspond is easy, since for $D$ in the interior of such a face, the quadratic form $q(L)=D\cdot L^2$ is degenerate, which is not the case for $D$ in the interior of a Type~I or Type~$\text{III}_0$ face. Stating it another way, if we consider the Hessian form associated to the topological cubic form $\mu$, then $h$ is a form of degree $\rho=b_2$ which has a linear factor corresponding to each Type~II face. Thus the condition that a face is of Type~II is topologically determined. The result will therefore follow if we can show that a face of the nef cone which is Type~I for one of the Calabi--Yau threefolds is not of Type~$\text{III}_0$ for the other. However, for a Type~I face, we saw in \S\ref{sec1} that $n_d$ is always nonnegative; for a Type~$\text{III}_0$ face, we proved in Theorem~\ref{thm_3.1} that $n_2=-2$. Since Gromov--Witten invariants are invariant under symplectic deformations, the result is proved. \end{pf} \begin{rem} It is still an open question whether there exist examples of Calabi--Yau threefolds $X_1$ and $X_2$ which are symplectic deformations of each other but not in the same algebraic family. \end{rem}
1997-07-09T19:51:21
9707
alg-geom/9707009
en
https://arxiv.org/abs/alg-geom/9707009
[ "alg-geom", "hep-th", "math.AG", "math.QA", "q-alg" ]
alg-geom/9707009
Martin Markl
Martin Markl
Simplex, associahedron, and cyclohedron
42 pages, LaTeX, article 12pt + leqno style
null
null
null
null
The aim of the paper is to give an `elementary' introduction to the theory of modules over operads and discuss three prominent examples of these objects - simplex, associahedron (= the Stasheff polyhedron) and cyclohedron (= the compactification of the space of configurations of points on the circle). Keywords: (right) module over an operad, module associated to a cyclic operad, Koszul module over an operad.
[ { "version": "v1", "created": "Wed, 9 Jul 1997 17:55:00 GMT" } ]
2008-02-03T00:00:00
[ [ "Markl", "Martin", "" ] ]
alg-geom
\section{\@startsection {section}{1}{\z@ }% {-3.5ex plus -1ex minus -.2ex}{2.3ex plus .2ex}{\bf }} \def\thebibliography#1{% \section *{References.\@mkboth {REFERENCES}{REFERENCES}}% \list {[\arabic {enumi}]}{\settowidth \labelwidth {[#1]}% \leftmargin \labelwidth \advance \leftmargin \labelsep % \usecounter {enumi}} \def \newblock % {\hskip .11em plus .33em minus -.07em} \sloppy \clubpenalty 4000% \widowpenalty 4000 \sfcode`\.=1000\relax} \def\@maketitle{% \newpage \null \vskip 2em \begin{center}{\Large\bf \@title \par } \vskip 1.5em {\large \lineskip .5em \begin {tabular}[t]{c}\@author \end{tabular}\par } \vskip .8em {June 20, 1994} \end{center}\par \vskip 1.5em} \def\footnote{\@ifnextchar [{\@xfootnote }{\stepcounter {\@mpfn }% \begingroup \let \protect \noexpand \xdef \@thefnmark {\hskip-3mm}% \endgroup \@footnotetext }} \def\@makefnmark{} \def\abstract{% \if@twocolumn \section *{Abstract} \else \small\quotation\noindent{\bf Abstract.}\fi} \def\fnum@figure{{\bf \figurename {} \thefigure }} \catcode`\@=13 \def\hspace*{\fill{\hspace*{\fill} \mbox{\hphantom{mm}\rule{0.25cm}{0.25cm}}\\} \def{\cal B}{{\cal B}} \def{\cal I}{{\cal I}} \def{\bf R}{{\bf R}} \def{{\cal B}C}{{{\cal B}C}} \def\rada#1#2{#1,\ldots,#2} \def{{\cal I}C}{{{\cal I}C}} \def{\cal C}{{\cal C}} \def{\cal P}{{\cal P}} \def\coll#1{\{{#1}(n)\}_{n\geq 1}} \def\colla#1{\{{#1}_n\}_{n\geq 1}} \def{\cal E}{{\cal E}} \def\Hom#1#2{{\rm Hom}(#1,#2)} \def\mbox{$1\!\!1$}{\mbox{$1\!\!1$}} \def\mbox{$\rule{.15mm}{1.9mm}\hskip-.75mm\times$}{{\mbox{$|\!\!\times$}}} \def{\overline {\cal B}}{{\overline {\cal B}}} \def{\overline K}{{\overline K}} \def{\overline {\cal A}}{{\overline {\cal A}}} \def{A($\infty$)}{{A($\infty$)}} \def{\overline {{\cal B}C}}{{\overline {{\cal B}C}}} \def{\overline W}{{\overline W}} \def{\cal M}{{\cal M}} \def\mbox{\large$\land$}{\mbox{\large$\land$}} \def\mbox{$\rule{.15mm}{1.9mm}\hskip-.75mm\times$}{\mbox{$\rule{.15mm}{1.9mm}\hskip-.75mm\times$}} \def\squeeze{\times\hskip-1.5mm \cdot \hskip-1mm% \cdot\hskip-1mm\cdot\hskip-1.5mm\times} \def{\bf Z}{{\bf Z}} \def{\cal F}{{\cal F}} \def{\cal T}{{\cal T}} \def{\tt T}{{\tt T}} \def\cyclsum{ {\raisebox{-.4mm}{\Large $\circ$}% \hskip-4.3mm\sum} } \def\znamenko#1{{(-1)^{#1}\cdot}} \def{\cal A}{{\cal A}} \def{\Omega}{{\Omega}} \def{\it Ass}{{\it Ass}} \def{\it Cycl}{{\it Cycl}} \def{\it UAss}{{\it UAss}} \def{\it UPoiss}{{\it UPoiss}} \def{\it Poiss}{{\it Poiss}} \def{\it Comm}{{\it Comm}} \def{\it UComm}{{\it UComm}} \def{\it U}{\cal P}{{\it U}{\cal P}} \def{\bf k}{{\bf k}} \def\prez#1#2{\langle #1;#2\rangle} \def{\bf s}{{\bf s}} \def{\rm Span}{{\rm Span}} \def\prezmod#1#2#3{{\langle #1;#2;#3\rangle}} \def{\cal M}{{\cal M}} \def\uparrow\!{\uparrow\!} \def{\bf s\hskip0mm}{{\bf s\hskip0mm}} \def{\cal Q}{{\cal Q}} \def{\partial_{\Cob}}{{\partial_{{\Omega}}}} \def{\rm sgn}{{\rm sgn}} \def{\rm sgn\hskip1mm}{{\rm sgn\hskip1mm}} \def{\sf F}{{\sf F}} \def{\varphi}{{\varphi}} \def{\underline {\sf F}}{{\underline {\sf F}}} \def{\buildrel \circ \over {{\sf F}}}{{\buildrel \circ \over {{\sf F}}}} \def{\otimes}{{\otimes}} \def{\bf E}{{\bf E}} \def\skel#1#2{{#1}^{(\leq #2)}} \def{\cal D}{{\cal D}} \def{\overline {\cal D}}{{\overline {\cal D}}} \def{\overline \Delta}{{\overline \Delta}} \def\dual#1{{{#1}^*}} \def\dualI#1{{(#1)^*}} \def\set#1{{\{#1\}}} \def{\bf Z}{{\bf Z}} \def\tttr#1#2{{\tt T}_{{#1},\ldots,{#2}}} \count88=1 \def\odstintro{{\vskip2mm\noindent% {\bf I.\the\count88.} \global\advance\count88 by 1}} \def{\rm Ker}{{\rm Ker}} \def\oDelta{\stackrel{\mbox{\scriptsize o}}% {\Delta}\hskip-1mm\rule{0mm}{2mm}} \def\downarrow\!{\downarrow\!} \long\def\comment#1\endcomment{{}} \begin{document} \pagestyle{myheadings} \bibliographystyle{plain} \baselineskip20pt plus 2pt minus 1pt \parskip3pt plus 1pt minus .5pt \begin{center} {\Large \bf Simplex, associahedron, and cyclohedron} \end{center} \begin{center} {\large Martin Markl} \end{center} \footnote{\noindent{\bf Mathematics Subject Classification:} 57P99} \footnote{Supported by a Fulbright grant, by the grant AV \v CR \#1019507 and by the grant GA \v CR \#201/96/0310} \section*{Introduction.} \vskip2mm \odstintro The paper deals with three types of convex polyhedra. The most classical is the $n$-dimensional {\em simplex\/} $\Delta^n$~\cite[\S10.1]{switzer:75}, the basic ingredient of simplicial topology and perhaps one of the most important mathematical objects at all~\cite{may:67}. Another polyhedron is the Stasheff polytope $K_n$, also called the {\em associahedron\/}, the basic tool for the study of homotopy associative Hopf spaces~\cite[page~277]{stasheff:TAMS63}. The last type is the polyhedron $W_n$, defined as the Axelrod-Singer compactification of the configuration space of $n$ points on the circle, and introduced by R.~Bott and C.~Taubes~\cite[page~5249]{bott-taubes:JMP94} in connection with the study of nonperturbative link invariants, recently dubbed by J.~Stasheff the {\em cyclohedron\/}~\cite{stasheff:from-ops}. \odstintro The crucial property of the collection $K =\{K_n\}_{n\geq 1}$ is that it forms a cellular {\em operad\/}~\cite[page~278]{stasheff:TAMS63}. J.~Stasheff observed in~\cite{stasheff:from-ops} that the collection $W = \{W_n\}_{n\geq 1}$ is a right {\em module\/}, in the sense of~\cite[page~1476]{markl:zebrulka}, over the operad $K$. In Theorem~\ref{vicko} we prove that also the collection $\Delta = \{\Delta^n\}_{n\geq 0}$ is a natural right module over the operad ${\it Ass}$ for associative algebras. \odstintro Operads were introduced to encode varieties of algebras. We show that, in the same spirit, also modules over operads describe varieties of some objects. We call these objects {\em traces\/} (Definition~\ref{el}), since they naturally generalize traces on associative algebras. For a so-called {\em cyclic\/} operad ${\cal P}$~\cite[Definition~2.1]{getzler-kapranov:cyclic} we construct a natural ${\cal P}$-module $M_{{\cal P}}$, the module {\em associated\/} to the operad ${\cal P}$ (Definition~\ref{Turmo}). We show that $M_{\cal P}$-traces are exactly {\em invariant bilinear forms\/} in the sense of E.~Getzler and M.M.~Kapranov~\cite[Definition~4.1]{getzler-kapranov:cyclic}. \odstintro Algebras over the cellular chain operad $CC_*(K)$ of the associahedron are {A($\infty$)}-algebras introduced by J.~Stasheff in~\cite[page~294]{stasheff:TAMS63}. They can be understood as algebras with the usual associativity condition \[ (ab)c = a(bc) \] satisfied only up to a system of coherent homotopies. In Proposition~\ref{nuzky} we show that the traces over the cellular chain complex $CC_*(W)$ of the cyclohedron are {\em homotopy traces\/} on {A($\infty$)}-algebras, for which the usual condition \[ T(ab) = T(ba) \] is satisfied only up to a system of coherent homotopies. The traces over the cellular chain complex $CC_*(\Delta)$ of the simplex are described in Theorem~\ref{vicko}. \odstintro The cellular chain complex $CC_*(K)$ of the associahedron has a very effective description -- it is the {\em operadic bar construction\/} on the operad ${\it Ass}$ for associative algebras~\cite[Example~4.1]{markl:zebrulka}. We introduce the bar construction on a {\em module\/} over an operad (this definition was independently made by V.~Ginzburg and A.A.~Voronov in~\cite{ginzburg-voronov}) and show that the cellular chain complex $CC_*(W)$ of the cyclohedron is the bar construction on the ${\it Ass}$-module ${\it Cycl}$, which describes traces (ordinary, not homotopy) on associative algebras (Theorem~\ref{ucpavka}). A fully algebraic description of $CC_*(\Delta)$ is given in Theorem~\ref{resiz}. \odstintro V.Ginzburg and M.M.~Kapranov~\cite[Definition~4.1.3]{ginzburg-kapranov:DMJ94} introduced so-called {\em Koszul operads\/}, with all expected nice properties, and the related notion of the {\em Koszul dual\/} of a {\em quadratic operad\/}~\cite[\S2.1.9]{ginzburg-kapranov:DMJ94}. We introduce analogous notions for {\em modules\/} over operads, i.e.~we introduce {\em quadratic modules\/}, their {\em Koszul (quadratic) duals\/} and the property of {\em Koszulness\/} for these modules. These definitions were again independently made by V.~Ginzburg and A.A.~Voronov in~\cite{ginzburg-voronov}. \odstintro The operad ${\it Ass}$ for associative algebras is Koszul~\cite[Corollary~4.2.7]{ginzburg-kapranov:DMJ94}. Since the operadic bar construction on ${\it Ass}$ is the cellular chain complex of the associahedron, the Koszulness of ${\it Ass}$ follows from the acyclicity of $K$, which in turn follows from the fact that it is a convex polyhedron. In a similar manner, we show in Theorem~\ref{Amphora1} that the module ${\it Cycl}$ describing traces on associative algebras is Koszul, as a consequence of the convexity of the cyclohedron $W$. A more general argument is to observe that ${\it Cycl}$ is the module associated to the cyclic operad ${\it Ass}$ (Example~\ref{kacirek}) and then apply Theorem~\ref{Katalogizacni} saying that a module associated to a Koszul operad is Koszul. \odstintro We show in Lemma~\ref{whoop} that, for each module over an operad, there exists a {\em spectral sequence\/}, converging to the homology of the bar construction. We also prove in Proposition~\ref{myska} that for modules over Koszul operads this spectral sequence collapses. Our spectral sequence, applied to an ${\it Ass}$-module ${\it Cycl}$, carries a strong geometrical message -- the initial term is the cellular chain complex of the cyclohedron, while the next term is the cellular chain complex of the simplex. If we interpret the cyclohedron as the compactification of the simplex constructed by a sequence of blow-ups~\cite[page~5249]{bott-taubes:JMP94}, then the spectral sequence describes the inverse process -- `deblowing-up' the cyclohedron back to the simplex, see Section~\ref{22}. \odstintro {\em Some further suggestions.\/} Consider the following `standard situation' closely related to the topological quantum field theory. Let $C(S^m)$ be the Axelrod-Singer compactification~\cite[Section~5]{axelrod-singer:preprint} of the configuration space of distinct points in the sphere $S^m$, and ${\sf F}_m$ the compactification of the moduli space of configurations of distinct points in ${\bf R}^m$~\cite[\S3.2]{getzler-jones:preprint}. It is known that ${\sf F}_n$ is a topological operad~\cite[\S3.2]{getzler-jones:preprint} and that this operad acts on the right module $C(S^m)$~\cite[Theorem~5.2]{markl:cf} (to be precise, if $m \not= 1,3,7$, the sphere $S^m$ is not parallelizable and we need a suitable framed versions of the objects above). The homology operad $H_*({\sf F}_m)$ describes a form of graded Poisson algebras~\cite[Theorem~3.1]{getzler-jones:preprint} (or Batalin-Vilkovisky algebras, in the framed case)~\cite[Section~4]{G2}, and it is not difficult to see that $H_*(C(S^n))$ is the module associated to the cyclic operad $H_*({\sf F}_m)$ in the sense of our Definition~\ref{Turmo}. Our paper deals with the above situation for $n=1$, while all the machinery cries for an application to a general situation. Another suggestion for further research is the following. E.~Getzler and M.M.~Kapranov introduced in~\cite[Definition~5.2]{getzler-kapranov:cyclic} the cyclic homology of an algebra over a cyclic operad as the (left, nonabelian) derived functor of the universal invariant bilinear form functor $\lambda({\cal P},-)$. We propose to study, for a (noncyclic) operad ${\cal P}$ and a ${\cal P}$-module $M$, the derived functor of the universal $M$-{\em trace\/} as a natural generalization of the cyclic homology. The cyclic homology will be then a special case for ${\cal P}$ cyclic and $M = M_{\cal P}$. There are two ways to read the paper -- either as an exposition of the properties of the associahedron, cyclohedron and simplex, with some generalizations, or as a paper on general theory of modules over operads, with a special attention paid to the three examples above. \noindent {\em Acknowledgment:\/} I would like to express my thanks to Jim Stasheff for numerous discussions and hospitality during my stay at the University of North Carolina. Also the communication with Sasha Voronov, who was working independently on~\cite{ginzburg-voronov}, was very useful. I am also very grateful to Steve Shnider and the referee for careful reading the manuscript and many useful remarks. \section*{Plan of the paper:} \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{1968}: Associahedron and the cyclohedron as a truncation of the simplex.\/} We recall the convex realization of the associahedron as a truncation of the simplex, due to S.~Shnider and S.~Sternberg, and construct a similar realization of the cyclohedron. \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{bolehlav}: Cyclohedron as a module over the associahedron.\/} We recall (right) modules over operads and introduce traces as algebraic objects described by these modules. We introduce the module ${\it Cycl}$ for traces on associative algebras. We prove that the cyclohedron is a module over the associahedron and describe the corresponding traces. \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{hrnicek1}: Simplex as a module over the operad for associative algebras.\/} We show that the simplex is a module over the operad ${\it Ass}$ for associative algebras. We prove that the associated cellular chain complex is free and describe the corresponding traces. \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{cervena-tuzka}: Quadratic operads and modules; modules associated to cyclic operads.\/} We present a class of operads and modules having a particularly easy description. We recall cyclic operads and introduce the module associated to a cyclic operad. \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{penezenka}: Cyclohedron as the cobar construction.\/} We introduce the cobar construction on a module over an operad. We define quadratic Koszul modules. We show that the cellular chain complex of the cyclohedron is the cobar construction on the module ${\it Cycl}$ and deduce from this fact that ${\it Cycl}$ is Koszul. \noindent \hangindent=5mm \hangafter=1 {\em Section~\ref{22}: Cyclohedron as a compactification of the simplex.\/} We view the cyclohedron as a compactification of the simplex, constructed as a sequence of blow-ups. We show that the spectral sequence related to the cobar construction `deflates' the cyclohedron back to the simplex. \noindent \hangindent=5mm \hangafter=1 {\em Appendix: Traces versus invariant bilinear forms.\/} We show that traces over the module associated to a cyclic operad are exactly invariant bilinear forms of E.~Getzler and M.M.~Kapranov. \section{Associahedron and the cyclohedron as a truncation of the simplex} \label{1968} Let ${\cal B}(n)$ denote the set of all meaningful bracketings of $n$ independent variables $\rada1n$. The {\em associahedron\/} $K_n$ is a convex $(n-2)$-dimensional polyhedron whose faces are indexed by elements of ${\cal B}(n)$. To be more precise, ${\cal B}(n)$ is a poset (= partially-ordered set) ordered by saying that $b' \prec b''$ if $b''$ is obtained from $b'$ by removing one or more pair of brackets. Then $K_n$ is a convex polyhedron whose poset of faces is (isomorphic to) ${\cal B}(n)$. See Figure~\ref{k3andk4} for $K_3$ and $K_4$. A nice picture of $K_5$ can be found in~\cite[page~151]{markl-stasheff:JofAlg94}. \begin{figure}[hb] \begin{center} \unitlength 1.70mm \thicklines \begin{picture}(54.33,25.50) \put(32.33,12.50){\line(1,1){10.00}} \put(42.33,22.50){\line(1,-1){10.00}} \put(52.33,12.50){\line(-1,-2){5.00}} \put(47.33,2.50){\line(-1,0){10.00}} \put(37.33,2.50){\line(-1,2){5.00}} \put(42.33,22.50){\makebox(0,0)[cc]{$\bullet$}} \put(52.33,12.50){\makebox(0,0)[cc]{$\bullet$}} \put(47.33,2.50){\makebox(0,0)[cc]{$\bullet$}} \put(37.33,2.50){\makebox(0,0)[cc]{$\bullet$}} \put(32.33,12.50){\makebox(0,0)[cc]{$\bullet$}} \put(30.33,12.50){\makebox(0,0)[rc]{$((12)3)4$}} \put(54.33,12.50){\makebox(0,0)[lc]{$1(2(34))$}} \put(42.33,25.50){\makebox(0,0)[cc]{$(12)(34)$}} \put(35.33,0.50){\makebox(0,0)[cc]{$(1(23))4$}} \put(49.33,0.50){\makebox(0,0)[cc]{$1((23)4)$}} \put(37.33,19.50){\makebox(0,0)[rc]{$(12)34$}} \put(47.33,19.50){\makebox(0,0)[lc]{$12(34)$}} \put(33.33,6.50){\makebox(0,0)[rc]{$(123)4$}} \put(51.33,6.50){\makebox(0,0)[lc]{$1(234)$}} \put(42.33,4.50){\makebox(0,0)[cc]{$1(23)4$}} \put(42.33,11.50){\makebox(0,0)[cc]{$1234$}} \put(-5.00,12.50){\line(1,0){15.00}} \put(-5.00,12.50){\makebox(0,0)[cc]{$\bullet$}} \put(10.00,12.50){\makebox(0,0)[cc]{$\bullet$}} \put(2.00,15.50){\makebox(0,0)[cc]{$123$}} \put(-7.00,17.50){\makebox(0,0)[cc]{$(12)3$}} \put(12.00,17.50){\makebox(0,0)[cc]{$1(23)$}} \end{picture} \end{center} \caption{$K_3$ (left) and $K_4$ (right).\label{k3andk4}} \end{figure} We recall a very cute `linear convex realization' of $K_n$ as a truncation of the $(n-2)$-dimensional simplex, due to S.~Shnider and S.~Sternberg~\cite{shnider-sternberg:book}. Our exposition follows the corrected version given in~\cite[Appendix~B]{stasheff:from-ops}. We need an alternative description of the poset ${\cal B}(n)$. Let $P(n)$ denote the set of all proper subintervals of the interval $[1,n-1] = \{\rada1{n-1}\}$. Two intervals $I,J \in P(n)$ are called {\em compatible\/}, if $I\cup J$ is not an interval properly containing both $I$ and $J$, i.e.~if either $J\subset I$, or $I\subset J$, or $I\cup J$ is not an interval. Let ${\cal I}(n)$ be the set of all subsets $\iota$ of $P(n)$ such that $I$ and $J$ are compatible for any $I,J \in \iota$. The poset structure on ${\cal I}(n)$ is given by the set inclusion: $\iota \preceq \kappa$ if $\kappa\subset \iota$. \begin{lemma} \label{prim} (Shnider-Sternberg) The posets ${\cal B}(n)$ and ${\cal I}(n)$ are isomorphic. \end{lemma} \noindent {\bf Proof.} For $I = [i,j]\in P(n)$, let $b(I)$ be the bracketing $1\cdots(i \cdots j+1)\cdots n$. This correspondence is easily seen to induce a poset isomorphism ${\cal I}(n) \cong {\cal B}(n)$.\hspace*{\fill \noindent Define the function $c: P(n)\to {\bf R}_{>0}$ by $c(I):= 3^{\#I}$, for $I \in P(n)$. Let $K_n \subset {\bf R}^{n-1}$ be the convex polytope defined by \[ K_n = \left\{(t_1,\ldots,t_{n-1})\in {\bf R}^{n-1};\ \sum_{k=1}^{n-1}t_k = c([1,n-1]),\ \sum_{k\in I}t_k \geq c(I),\ I\in P(n)\right\}. \] Denote also, for $I\in P(n)$, by $P_I$ the hyperplane \[ P_I := \left\{(t_1,\ldots,t_{n-1})\in {\bf R}^{n-1};\ \sum_{k\in I}t_k =c(I)\right\}. \] The proof of the following proposition is given in~\cite[Appendix~B]{stasheff:from-ops}. \begin{proposition}(Shnider-Sternberg) \label{Skoda} The polytope $K_n$ has nonempty interior in the $(n-2)$-dimensional hyperplane $\{(t_1,\ldots,t_{n-1})\in {\bf R}^{n-1};\ \sum_{k=1}^{n-1}t_k = 3^{n-1}\}$. The intersection \[ K_n \cap \bigcap\{P_{I},\ I \in \iota\} \] defines a nonempty $(n-\#I-2)$-dimensional face of $K_n$ for any $\iota \in {\cal I}(n)$. All faces of $K_n$ are obtained in this way. \end{proposition} If we denote, for $\iota \in {\cal I}(n)$, by $P_{\iota}$ the intersection $\bigcap\{P_{I},\ I \in \iota\}$, then the above proposition immediately implies that the correspondence $\iota \mapsto K_n \cap P_\iota$ defines an isomorphism of the poset ${\cal I}(n)$ and the poset of faces of $K_n$. This is the promised convex realization of $K_n$. The case $n=4$ is illustrated on Figure~\ref{realK4}. \begin{figure}[hbtp] \begin{center} \unitlength 0.95mm \linethickness{0.4pt} \begin{picture}(85.00,100.33) \thinlines \put(0.00,4.33){\vector(1,0){100}} \put(0.00,4.33){\vector(0,1){100.00}} \put(-5.00,89.33){\line(1,-1){87.89}} \put(-3.00,64.33){\line(1,0){65.00}} \put(60.00,1.33){\line(0,1){60.11}} \put(10.00,99.33){\line(0,-1){97.00}} \put(-3.00,14.33){\line(1,0){88.00}} \put(-2.00,84.33){\line(1,0){4.00}} \put(-5.00,84.33){\makebox(0,0)[rc]{$24$}} \put(-5.00,64.33){\makebox(0,0)[rc]{$18$}} \put(-5.00,14.33){\makebox(0,0)[rc]{$3$}} \put(60.00,-0.67){\makebox(0,0)[ct]{$18$}} \put(80.00,-0.67){\makebox(0,0)[ct]{$24$}} \put(10.00,-0.67){\makebox(0,0)[ct]{$3$}} \put(2.00,98.33){\makebox(0,0)[lb]{$t_3$}} \put(43.00,47.33){\makebox(0,0)[lc]{$P_{\{[2]\}}$}} \put(38.00,67.33){\makebox(0,0)[cb]{$P_{\{[1,2]\}}$}} \put(64.00,46.33){\makebox(0,0)[lc]{$P_{\{[2,3]\}}$}} \put(13.00,95.33){\makebox(0,0)[lc]{$P_{\{[1]\}}$}} \put(25.00,17.33){\makebox(0,0)[cb]{$P_{\{[3]\}}$}} \put(80.00,1.33){\line(0,1){6.11}} \thicklines \put(10.00,14.33){\line(1,0){50.00}} \put(60.00,14.33){\line(0,1){10.00}} \put(60.00,24.33){\line(-1,1){40.00}} \put(20.00,64.33){\line(-1,0){10.00}} \put(10.00,64.33){\line(0,-1){50.00}} \put(10.00,64.33){\makebox(0,0)[cc]{$\bullet$}} \put(20.00,64.33){\makebox(0,0)[cc]{$\bullet$}} \put(10.00,14.33){\makebox(0,0)[cc]{$\bullet$}} \put(60.00,14.33){\makebox(0,0)[cc]{$\bullet$}} \put(60.00,24.33){\makebox(0,0)[cc]{$\bullet$}} \put(12.17,61.50){\makebox(0,0)[cc]{$\cdot$}} \put(16.67,60.50){\makebox(0,0)[cc]{$\cdot$}} \put(18.83,56.00){\makebox(0,0)[cc]{$\cdot$}} \put(12.17,44.83){\makebox(0,0)[cc]{$\cdot$}} \put(12.17,36.50){\makebox(0,0)[cc]{$\cdot$}} \put(16.67,52.16){\makebox(0,0)[cc]{$\cdot$}} \put(16.67,43.83){\makebox(0,0)[cc]{$\cdot$}} \put(16.67,35.50){\makebox(0,0)[cc]{$\cdot$}} \put(13.67,49.00){\makebox(0,0)[cc]{$\cdot$}} \put(18.83,47.66){\makebox(0,0)[cc]{$\cdot$}} \put(18.83,39.33){\makebox(0,0)[cc]{$\cdot$}} \put(12.33,24.66){\makebox(0,0)[cc]{$\cdot$}} \put(16.83,23.66){\makebox(0,0)[cc]{$\cdot$}} \put(13.83,20.50){\makebox(0,0)[cc]{$\cdot$}} \put(19.00,27.50){\makebox(0,0)[cc]{$\cdot$}} \put(19.00,19.16){\makebox(0,0)[cc]{$\cdot$}} \put(25.83,43.83){\makebox(0,0)[cc]{$\cdot$}} \put(25.83,35.50){\makebox(0,0)[cc]{$\cdot$}} \put(22.83,49.00){\makebox(0,0)[cc]{$\cdot$}} \put(22.83,40.66){\makebox(0,0)[cc]{$\cdot$}} \put(22.83,32.33){\makebox(0,0)[cc]{$\cdot$}} \put(28.00,47.66){\makebox(0,0)[cc]{$\cdot$}} \put(28.00,39.33){\makebox(0,0)[cc]{$\cdot$}} \put(26.00,23.66){\makebox(0,0)[cc]{$\cdot$}} \put(28.17,27.50){\makebox(0,0)[cc]{$\cdot$}} \put(39.83,27.50){\makebox(0,0)[cc]{$\cdot$}} \put(32.83,27.33){\makebox(0,0)[cc]{$\cdot$}} \put(39.83,19.16){\makebox(0,0)[cc]{$\cdot$}} \put(36.83,32.66){\makebox(0,0)[cc]{$\cdot$}} \put(43.83,24.50){\makebox(0,0)[cc]{$\cdot$}} \put(29.83,32.50){\makebox(0,0)[cc]{$\cdot$}} \put(36.83,24.33){\makebox(0,0)[cc]{$\cdot$}} \put(36.83,16.00){\makebox(0,0)[cc]{$\cdot$}} \put(35.00,39.50){\makebox(0,0)[cc]{$\cdot$}} \put(35.17,19.33){\makebox(0,0)[cc]{$\cdot$}} \put(52.33,19.33){\makebox(0,0)[cc]{$\cdot$}} \put(49.33,24.50){\makebox(0,0)[cc]{$\cdot$}} \put(49.33,16.16){\makebox(0,0)[cc]{$\cdot$}} \put(54.50,23.16){\makebox(0,0)[cc]{$\cdot$}} \put(22.00,57.33){\makebox(0,0)[cc]{$\cdot$}} \put(30.83,49.00){\makebox(0,0)[cc]{$\cdot$}} \put(39.67,40.67){\makebox(0,0)[cc]{$\cdot$}} \end{picture} \end{center} \caption{$(t_1,t_3)$-projection of the convex realization of $K_4$.\label{realK4}} \end{figure} As observed in~\cite[Appendix~B]{stasheff:from-ops}, the above construction works also for other choices of the function $c : P(n)\to {\bf R}_{>0}$ provided it is admissible in the sense that \[ c(I)+c(J)< c(I\cup J), \mbox{ if $I\cup J$ properly contains both $I$ and $J$.} \] Let us proceed to the definition of the cyclohedron $W_n$. As the associahedron, it will be a convex polyhedron characterized by the poset ${{\cal B}C}(n)$ indexing its faces. Consider again $n$ independent formal variables, labeled by natural numbers $1,\ldots,n$. The elements of ${{\cal B}C}(n)$ will be equivalence classes represented by bracketing of the chain $\rada{\sigma(1)}{\sigma(n)}$, where $\sigma \in \Sigma_n$ is a {\em cyclic\/} permutation. In contrast to the case of ${\cal B}(n)$, we allow also the bracketing which embraces all elements. Thus, for instance, $(3(12))$ represents an element of ${{\cal B}C}(n)$. The equivalence relation is given as follows. Let $\sigma\in \Sigma_n$ be a cyclic permutation, let $b'$ be a bracketing of $\sigma(1)\cdots\sigma(s)$ and $b''$ a bracketing of $\sigma(s+1)\cdots\sigma(n)$, for some $1\leq s\leq n$. Then we identify $b'b''$ to $b''b'$. Thus, for example, $3(12)= (12)3$ in ${{\cal B}C}(3)$ (but $(3(12))\not= ((12)3)$). The partial order on ${{\cal B}C}(n)$ is defined, as for ${\cal B}(n)$, by deleting pairs of brackets. Each element $b$ of ${{\cal B}C}(n)$ can be uniquely represented by a symbol, obtained from a representative of $b$ by forcing the indeterminates into the natural order. We call such symbols {\em cyclic bracketings\/}. The formal definition will be obvious from the following example. \begin{example}{\rm\ The poset ${{\cal B}C}(2)$ contains three elements, $(12)$, $(21)$ and $12$, where $(21)$ is represented by the cyclic bracketing $1)(2$. The poset structure is depicted by the interval, see Figure~\ref{W2}. \begin{figure}[hbtp] \begin{center} \unitlength=1mm \begin{picture}(53.00,15.00)(10,10) \thicklines \put(20.00,10.00){\line(1,0){30.00}} \put(20.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(50.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(35.00,15.00){\makebox(0,0)[cc]{$12$}} \put(17.00,15.00){\makebox(0,0)[cc]{$(12)$}} \put(53.00,15.00){\makebox(0,0)[cc]{$1)(2$}} \end{picture} \end{center} \caption{$W_2$.\label{W2}} \end{figure} \noindent Below are listed elements of the poset ${{\cal B}C}(3)$: \[ \begin{array}[b]{|c|c||c|c|} \hline \mbox{elements of ${{\cal B}C}(3)$}&\mbox{cyclic bracketings}&\mbox{cont.}&\mbox{\hskip11mm cont.\hskip11mm} \\ \hline \hline (1(23))&(1(23))&(231)&1)(23 \\ ((12)3)&((12)3)&2(31)=(31)2&1)2(3 \\ ((23)1)&1)((23)&(312)&12)(3 \\ (2(31))&1))(2(3&(12)3=3(12)&(12)3 \\ ((31)2)&1)2)((3&(123)&(123) \\ (3(12))&(12))(3&123=231=312&123 \\ 1(23)=(23)1&1(23)&&\\ \hline \end{array} \] The poset structure of ${{\cal B}C}(3)$ is depicted on Figure~\ref{W3}. }\end{example} \begin{figure}[hbtp] \begin{center} \unitlength2mm \begin{picture}(62.00,35.00)(10,3) \thicklines \put(20.00,20.00){\line(1,1){10.00}} \put(30.00,30.00){\line(1,0){20.00}} \put(50.00,30.00){\line(1,-1){10.00}} \put(60.00,20.00){\line(-1,-1){10.00}} \put(50.00,10.00){\line(-1,0){20.00}} \put(30.00,10.00){\line(-1,1){10.00}} \put(20.00,20.00){\makebox(0,0)[cc]{$\bullet$}} \put(30.00,30.00){\makebox(0,0)[cc]{$\bullet$}} \put(50.00,30.00){\makebox(0,0)[cc]{$\bullet$}} \put(60.00,20.00){\makebox(0,0)[cc]{$\bullet$}} \put(50.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(30.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(28.00,35.00){\makebox(0,0)[cc]{$((12)3)$}} \put(52.00,35.00){\makebox(0,0)[cc]{$(1(23))$}} \put(62.00,20.00){\makebox(0,0)[lc]{$1)((23)$}} \put(52.00,5.00){\makebox(0,0)[cc]{$1))(2(3$}} \put(28.00,5.00){\makebox(0,0)[cc]{$1)2)((3$}} \put(18.00,20.00){\makebox(0,0)[rc]{$(12))(3$}} \put(40.00,35.00){\makebox(0,0)[cc]{$(123)$}} \put(40.00,5.00){\makebox(0,0)[cc]{$1)2(3$}} \put(27.00,25.00){\makebox(0,0)[lc]{$(12)3$}} \put(27.00,15.00){\makebox(0,0)[lc]{$12)(3$}} \put(53.00,25.00){\makebox(0,0)[rc]{$1(23)$}} \put(53.00,15.00){\makebox(0,0)[rc]{$1)(23$}} \put(40.00,20.00){\makebox(0,0)[cc]{$123$}} \end{picture} \end{center} \caption{$W_3$.\label{W3}} \end{figure} The structure of ${{\cal B}C}(4)$ is indicated on Figure~\ref{W4}. The picture is already rather complicated, so we labeled only the vertices (= the minimal elements of ${{\cal B}C}(4)$). The label of an arbitrary face can be easily found -- it is the least upper bound of all vertices of the face. For example, the pentagon on the top of $W_4$ is labeled by $123)(4$, the front hexagon is labeled by $(12)34$, etc. \begin{figure}[hbtp] \begin{center} \unitlength=0.67mm \begin{picture}(200.11,131.11) \thicklines \put(20.00,10.00){\line(1,0){160.05}} \put(0.00,20.00){\line(1,1){89.87}} \put(20.00,10.00){\line(-2,1){19.95}} \put(89.87,109.90){\line(-3,4){9.74}} \put(180.00,10.00){\line(2,1){19.10}} \put(0.00,20.00){\line(0,1){23.00}} \put(200.00,20.00){\line(0,1){23.00}} \put(0.00,43.00){\line(1,1){80.00}} \put(100.11,131.11){\line(5,-2){20.00}} \put(110.00,110.00){\line(3,4){9.89}} \put(120.00,123.00){\line(1,-1){80.11}} \put(80.00,123.00){\line(5,2){20.11}} \put(110.00,110.00){\line(1,-1){90.00}} \put(110.11,110.00){\line(0,0){0.00}} \put(90.00,110.00){\line(1,0){20.11}} \thinlines \put(170.00,25.00){\line(1,2){5.00}} \put(30.00,25.00){\line(-1,2){5.00}} \put(30.00,25.00){\line(3,1){60.00}} \put(180.00,10.00){\line(-2,3){10.00}} \put(85.11,55.11){\line(-3,-1){60.00}} \put(90.00,45.00){\line(1,0){19.89}} \put(200.00,43.00){\line(-3,-1){15.05}} \put(175.00,35.00){\line(3,1){6.05}} \put(109.89,45.11){\line(1,2){4.89}} \put(100.00,65.11){\line(3,-2){15.11}} \put(90.00,45.11){\line(-1,2){4.89}} \put(85.00,55.00){\line(3,2){15.00}} \put(20.00,10.00){\line(2,3){10.00}} \put(114.78,55.11){\line(3,-1){60.22}} \put(0.00,43.00){\line(3,-1){14.96}} \put(25.00,35.00){\line(-3,1){5.00}} \put(170.00,25.00){\line(-3,1){60.11}} \put(100.00,65.00){\line(0,1){43}} \put(100.00,131.00){\line(0,-1){17}} \put(175.00,35.00){\makebox(0,0)[cc]{$\bullet$}} \put(200.00,43.00){\makebox(0,0)[cc]{$\bullet$}} \put(200.00,20.00){\makebox(0,0)[cc]{$\bullet$}} \put(180.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(170.00,25.00){\makebox(0,0)[cc]{$\bullet$}} \put(90.00,45.00){\makebox(0,0)[cc]{$\bullet$}} \put(85.00,55.00){\makebox(0,0)[cc]{$\bullet$}} \put(100.00,65.00){\makebox(0,0)[cc]{$\bullet$}} \put(115.00,55.00){\makebox(0,0)[cc]{$\bullet$}} \put(110.00,45.00){\makebox(0,0)[cc]{$\bullet$}} \put(80.00,123.00){\makebox(0,0)[cc]{$\bullet$}} \put(100.00,131.00){\makebox(0,0)[cc]{$\bullet$}} \put(120.00,123.00){\makebox(0,0)[cc]{$\bullet$}} \put(90.00,110.00){\makebox(0,0)[cc]{$\bullet$}} \put(110.00,110.00){\makebox(0,0)[cc]{$\bullet$}} \put(0.00,43.00){\makebox(0,0)[cc]{$\bullet$}} \put(0.00,20.00){\makebox(0,0)[cc]{$\bullet$}} \put(20.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(25.00,35.00){\makebox(0,0)[cc]{$\bullet$}} \put(30.00,25.00){\makebox(0,0)[cc]{$\bullet$}} \put(0.00,46.00){\makebox(0,0)[rb]{\scriptsize $1)2)\!)\!(3(\!(4$}} \put(0.00,17.00){\makebox(0,0)[rt]{\scriptsize $(12)\!)\!)\!(3(4$}} \put(18.00,7.00){\makebox(0,0)[ct]{\scriptsize $(12)\!)\!(\!(34)$}} \put(32.00,23.00){\makebox(0,0)[lc]{\scriptsize $1)2)\!(\!(\!(34)$}} \put(28.00,34.00){\makebox(0,0)[lc]{\scriptsize $1)\!)2)\!(\!(3(4$}} \put(88.00,38.00){\makebox(0,0)[cc]{\scriptsize $1)\!)\!(2(\!(34)$}} \put(112.00,38.00){\makebox(0,0)[cc]{\scriptsize $1)\!(\!(\!(2(34)\!)$}} \put(83.00,58.00){\makebox(0,0)[rc]{\scriptsize $1)\!)\!)\!(2(3(4$}} \put(117.00,58.00){\makebox(0,0)[lc]{\scriptsize $1)\!(\!(\!(23)4)$}} \put(103.00,67.00){\makebox(0,0)[lb]{\scriptsize $1)\!)\!(\!(23)\!(4$}} \put(78.00,126.00){\makebox(0,0)[rb]{\scriptsize $1)2)3)\!(\!(\!(4$}} \put(122.00,126.00){\makebox(0,0)[lb]{\scriptsize $(1(23)\!)\!(4$}} \put(100.00,135.00){\makebox(0,0)[cb]{\scriptsize $1)\!(23)\!)\!(\!(4$}} \put(112.00,110.00){\makebox(0,0)[lc]{\scriptsize $(\!(\!12\!)3)\!)\!(\!4$}} \put(88.00,110.00){\makebox(0,0)[rc]{\scriptsize $(\!12\!)\!)3)\!(\!(\!4$}} \put(180.00,5.00){\makebox(0,0)[lc]{\scriptsize $(\!(12)\!(34)\!)$}} \put(200.00,15.00){\makebox(0,0)[lc]{\scriptsize $(\!(\!(12)3)4)$}} \put(200.00,47.00){\makebox(0,0)[lc]{\scriptsize $(\!(1(23)\!)4)$}} \put(166.00,22.00){\makebox(0,0)[rc]{\scriptsize $(1(2(34)\!)\!)$}} \put(172.00,34.00){\makebox(0,0)[rc]{\scriptsize $1(\!(23)4)\!)$}} \end{picture} \end{center} \caption{$W_4$.\label{W4}} \end{figure} We construct, mimicking the approach of Shnider and Sternberg, a convex realization of the poset ${{\cal B}C}(n)$. First some terminology. By a {\em cyclic subinterval\/} of $[1,n]$ we mean either a `normal' subinterval $[i,j]$, $1\leq i\leq j \leq n$, representing the subset $\{\rada ij\}$ of $\{\rada 1n\}$, or the symbol $i][j$, $1\leq i <j \leq n$, representing $\{\rada 1i\}\cup \{\rada jn\}$. We will always suppose that the corresponding sets are proper subsets of $\{\rada 1n\}$, i.e.~we exclude the intervals $[1,n]$ and $i][i+1$, for $1\leq i <n$. Let us denote by $PC(n)$ the set of all cyclic subintervals in the above sense. We denote by ${{\cal I}C}(n)$ the set of all subsets of $PC(n)$ consisting of {\em nested\/} subintervals, meaning that, for $I,J \in \iota \in {{\cal I}C}(n)$, either $I\subset J$ or $J \subset I$. Again, ${{\cal I}C}(n)$ is a poset, the order being induced by the inclusion. We have the following analog of Lemma~\ref{prim}. \begin{lemma} The posets ${{\cal B}C}(n)$ and ${{\cal I}C}(n)$ are isomorphic. \end{lemma} \noindent {\bf Proof.} Define, for $I \in PC(n)$, the cyclic bracketing $b(I)\in {{\cal B}C}(n)$ by \[ b(I)=\left\{ \begin{array}{ll} 1\cdots(i\cdots j+1)\cdots n, &\mbox{ for $I = [i,j],\ j< n$,} \\ 1)\cdots(i\cdots n, &\mbox{ for $I = [i,n]$, and} \\ 1\cdots i+1)\cdots(j\cdots n, &\mbox{ for $I = i][j$.} \end{array} \right. \] This correspondence induces the desired poset isomorphism.\hspace*{\fill Our convex realization of ${{\cal B}C}(n)$, whose possibility was predicted in~\cite[Appendix~B]{stasheff:from-ops}), is defined as follows. Let $W_n \subset {\bf R}^n$ be the convex polyhedron \[ W_n = \left\{(t_1,\ldots,t_n)\in {\bf R}^n;\ \sum_{k=1}^nt_k = c([1,n]),\ \sum_{k\in I}t_k \geq c(I),\ I\in PC(n)\right\}. \] For $I\in P(n)$, let $P_I$ be the hyperplane \[ P_I := \left\{(t_1,\ldots,t_n)\in {\bf R}^n;\ \sum_{k\in I}t_k =c(I)\right\}. \] The proof of the following proposition is a straightforward modification of the proof of Proposition~\ref{Skoda} as given in~\cite[Appendix~B]{stasheff:from-ops}). \begin{proposition} The polytope $W_n$ has nonempty interior in the $(n-1)$-dimensional hyperplane $\{(t_1,\ldots,t_n)\in {\bf R}^n;\ \sum_{k=1}^n t_k = c([1,n])\}$. The intersection \[ W_n \cap \bigcap\{P_{I},\ I \in \iota\} \] defines a nonempty $(n-\#I-1)$-dimensional face of $W_n$ for any $\iota\in {{\cal I}C}(n)$ and all faces of $W_n$ are obtained in this way. \end{proposition} For $\iota \in {{\cal I}C}(n)$, let $P_{\iota}$ be the intersection $\bigcap\{P_{I},\ I \in \iota\}$. Then the correspondence $\iota \mapsto W_n \cap P_\iota$ defines an isomorphism between the poset ${{\cal I}C}(n)$ and the poset of faces of the polytope $W_n$. This is our convex realization of $W_n$. The convex realization of $W_3$ is shown on Figure~\ref{realW3}. \begin{figure}[hbtp] \begin{center} \unitlength 0.95mm \linethickness{0.4pt} \begin{picture}(100.00,100.00)(0,-7) \thinlines\put(-5,80.00){\makebox(0,0)[rc]{$24$}} \put(-5.00,60.00){\makebox(0,0)[rc]{$18$}} \put(-5.00,10.00){\makebox(0,0)[rc]{$3$}} \put(60.00,-5.00){\makebox(0,0)[ct]{$18$}} \put(80.00,-5.00){\makebox(0,0)[ct]{$24$}} \put(10.00,-5.00){\makebox(0,0)[ct]{$3$}} \put(0.00,0.00){\vector(1,0){100.00}} \put(0.00,0.00){\vector(0,1){100.00}} \put(-5.00,85.00){\line(1,-1){87.89}} \put(-3.00,60.00){\line(1,0){65.00}} \put(60.00,-3.00){\line(0,1){60.11}} \put(10.00,95.00){\line(0,-1){98.00}} \put(-3.00,10.00){\line(1,0){88.00}} \put(-2.00,80.00){\line(1,0){4.00}} \put(88.00,2.00){\makebox(0,0)[cb]{$t_1$}} \put(2.00,94.00){\makebox(0,0)[lb]{$t_3$}} \put(43.00,43.00){\makebox(0,0)[lc]{$P_{\{[2]\}}$}} \put(38.00,63.00){\makebox(0,0)[cb]{$P_{\{[1,2]\}}$}} \put(64.00,42.00){\makebox(0,0)[lc]{$P_{\{[2,3]\}}$}} \put(13.00,91.00){\makebox(0,0)[lc]{$P_{\{[1]\}}$}} \put(25.00,13.00){\makebox(0,0)[cb]{$P_{\{[3]\}}$}} \put(80.00,-3.00){\line(0,1){6.11}} \put(-5.00,55.00){\line(1,-1){57.00}} \put(22.00,33.00){\makebox(0,0)[lc]{$P_{\{1][3\}}$}} \thicklines \put(60.00,10.00){\line(0,1){10.00}} \put(60.00,20.00){\line(-1,1){40.00}} \put(20.00,60.00){\line(-1,0){10.00}} \put(10.00,60.00){\makebox(0,0)[cc]{$\bullet$}} \put(20.00,60.00){\makebox(0,0)[cc]{$\bullet$}} \put(60.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(60.00,20.00){\makebox(0,0)[cc]{$\bullet$}} \put(10.00,60.00){\line(0,-1){20.00}} \put(10.00,40.00){\line(1,-1){30.00}} \put(40.00,10.00){\line(1,0){20.00}} \put(10.00,40.00){\makebox(0,0)[cc]{$\bullet$}} \put(40.00,10.00){\makebox(0,0)[cc]{$\bullet$}} \put(13.33,56.67){\makebox(0,0)[cc]{.}} \put(19.17,53.67){\makebox(0,0)[cc]{.}} \put(14.50,51.33){\makebox(0,0)[cc]{.}} \put(22.00,48.50){\makebox(0,0)[cc]{.}} \put(26.17,45.17){\makebox(0,0)[cc]{.}} \put(45.17,22.67){\makebox(0,0)[cc]{.}} \put(32.00,42.17){\makebox(0,0)[cc]{.}} \put(51.00,19.67){\makebox(0,0)[cc]{.}} \put(27.33,39.83){\makebox(0,0)[cc]{.}} \put(46.33,17.33){\makebox(0,0)[cc]{.}} \put(34.83,37.00){\makebox(0,0)[cc]{.}} \put(53.83,14.50){\makebox(0,0)[cc]{.}} \put(35.83,32.67){\makebox(0,0)[cc]{.}} \put(41.67,29.67){\makebox(0,0)[cc]{.}} \put(37.00,27.33){\makebox(0,0)[cc]{.}} \put(44.50,24.50){\makebox(0,0)[cc]{.}} \put(40.00,27.00){\makebox(0,0)[cc]{.}} \put(41.00,22.67){\makebox(0,0)[cc]{.}} \put(42.17,17.33){\makebox(0,0)[cc]{.}} \put(18.17,43.50){\makebox(0,0)[cc]{.}} \put(13.50,41.17){\makebox(0,0)[cc]{.}} \put(21.00,38.33){\makebox(0,0)[cc]{.}} \end{picture} \end{center} \caption{$(t_1,t_3)$-projection of the convex realization of $W_3$.\label{realW3}} \end{figure} \begin{observation}{\rm\ The cyclohedron $W_n$ has $n(n-1)$ codimension-one faces, represented by the bracketings \begin{equation} \label{csa} b_{k,n} := (\rada{\sigma(1)}{\sigma(k)})\rada{\sigma(k+1)}{\sigma(n)},\ 1 < k\leq n, \end{equation} where $\sigma \in \Sigma_n$ is a cyclic permutation. The face represented by the bracketing $b_{k,n}$ is isomorphic to the product $W_{n-k+1}\times K_k$. For example, $W_4$ depicted on Figure~\ref{W4}, has \begin{itemize} \item[-] $4$ hexagonal faces, corresponding to $(12)34$, $(23)41$, $(34)12$ and $(41)23$, isomorphic to $W_3 \times K_2 = W_3 \times \mbox{point}$, \item[-] 4 square faces, corresponding to $(123)4$, $(234)1$, $(341)2$ and $(412)3$, isomorphic to $W_2 \times K_3$, and \item[-] 4 pentagonal faces, corresponding to $(1234)$, $(2341)$, $(3412)$ and $(4123)$, isomorphic to $W_1 \times K_4 = \mbox{point} \times K_4$. \end{itemize} This was observed by J.~Stasheff who realized that this is a strong motivation for the existence of a module structure which we will discuss in the following Section~\ref{bolehlav}. }\end{observation} \begin{observation}{\rm\ It is clear from our constructions that the cyclohedron $W_n$ is a truncation of the associahedron $K_{n+1}$, for $n\geq 1$. This is, of course, a trivial statement -- any convex polyhedron is a truncation of an arbitrary other convex polyhedron of the same dimension, so we must be more precise. For any $n\geq 1$ there is an obvious map $P(n+1)\hookrightarrow PC(n)$ which decomposes $PC(n)$ as \[ PC(n)= P(n+1) \sqcup E(n), \] where $E(n)$ is the subset of `exotic' cyclic intervals of the form $i][j$, $1\leq i<j\leq n$. The polyhedron $W_n$ is then the truncation of $K_{n+1}$ by hyperplanes $P_I$ indexed by the `exotic' intervals $I\in E(n)$. Compare Figures~\ref{realK4} and~\ref{realW3} for $n=3$. We do not know whether this observation has any deeper meaning. }\end{observation} \begin{observation}{\rm \label{sdff} Choose, for each $t=\rada 1n$, a point $P_t$ in the interior of the codimension one face of $W_n$ corresponding to $(\rada tn,\rada 1{t-1})$. The convex hull of the set $\set{\rada{P_1}{P_n}}$ is a simplex, closely related to the `deblowing up' of $W_n$ described in Section~\ref{22}. We will use this simplex to introduce an orientation of $W_n$. }\end{observation} \section{Cyclohedron as a module over the associahedron} \label{bolehlav} We believe that there is no need to give a detailed definition of an operad. Recall only that an {\em operad\/} (in a symmetric monoidal category ${\cal C}= ({\cal C},\times)$) is a sequence ${\cal P} = \{{\cal P}(n); n\geq 1\}$ of objects of ${\cal C}$ together with morphisms \[ \gamma=\gamma_{m_1,\ldots,m_l}:{\cal P}(l)\times {\cal P}(m_1)\times\cdots\times{\cal P}(m_l) \longrightarrow {\cal P}(m_1+\cdots+m_l), \] given for any $l,m_1,\ldots,m_l \geq 1$, satisfying the usual axioms~\cite[Definition~3.12]{may:1972}. If not stated otherwise, we assume our operads to be {\em symmetric\/}, i.e.~we assume that each ${\cal P}(n)$ has a right action of the symmetric group $\Sigma_n$, $n\geq 2$, which has again to satisfy some axioms~\cite[Definition~1.1]{may:1972}. We frequently write $p(\rada{p_1}{p_l})$ instead of $\gamma(p,\rada{p_1}{p_l})$. One comment concerning the action of the symmetric group is in order here. Our convention is determined by the conventional choice of the multiplication in the symmetric group. We accepted the standard one with $\sigma \cdot \tau$ meaning $\sigma(\tau)$, i.e.~the permutation (= a map) $\tau$ followed by $\sigma$. Then ${\cal P}(n)$ must be a {\em right\/} $\Sigma_n$-module, which is the convention used in the original definition of P.~May quoted above. Recall that, for any object $V\in {\cal C}$, there exist the so-called {\em endomorphism operad\/} ${\cal E}_V = \coll{{\cal E}_V}$ with ${\cal E}_V(n):= \Hom {V^{\times n}}V$. If ${\cal P}$ is an operad in ${\cal C}$, then a {\em${\cal P}$-algebra structure on $V$\/} is an operad map $a: {\cal P} \to {\cal E}_V$. \begin{example}{\rm\ \label{hrnicek} The collection ${\cal B} = \coll {{\cal B}}$ introduced in Section~\ref{1968} has a structure of a (nonsymmetric) operad in the category of posets. The composition $\gamma(b;\rada{b_1}{b_l})$ is, for $b\in {\cal B}(l)$ and $b_i \in {\cal B}(m_i)$, defined as the bracketing $b(\rada{b_1}{b_l})$ obtained by inserting $b_i$ at the $i$-th position in $b$, $1\leq i\leq l$. We believe that it is clear what we mean by this. For example \begin{eqnarray*} \gamma_{{\cal B}}(12;1(23),12)& =& 1(23)45, \\ \gamma_{{\cal B}}((12)3;(12)3,12,(12)(24)) & =& ((12)345)(67)(89),\ \mbox{etc.} \end{eqnarray*} A classical result of J.~Stasheff~\cite[page~278]{stasheff:TAMS63} says that the collection of the associahedra $K = \{K_n\}_{n\geq 1}$ has a cellular (nonsymmetric) operad structure which induces on the collection ${\cal B} = \coll {{\cal B}}$ of its faces the operad structure of Example~\ref{hrnicek}. As a consequence, cellular chains on $K$ form an operad ${\cal A} = \coll {\cal A}$, ${\cal A}(n)= CC_*(K_n)$, in the category of differential graded vector spaces. Algebras over the operad ${\cal A}$ are {A($\infty$)}-algebras~\cite[page~294]{stasheff:TAMS63}. }\end{example} Probably the most effective way to describe the operad ${\cal A}$ is to say that ${\cal A} = {\Omega}(\dualI{{\bf s} {\it Ass}})$, the operadic cobar construction on the dual cooperad $\dualI {{\bf s} {\it Ass}}$, where ${\bf s} {\it Ass}$ is the suspension of the operad for associative algebras. This is the same, since the operad ${\it Ass}$ is Koszul, as to say that ${\cal A}$ is the {\em minimal model\/} of the associative operad ${\it Ass}$. All this is explained in~\cite{markl:zebrulka}. We are going to make a similar analysis for the cyclohedron. \begin{definition} \label{Amphora} A (right) module over an operad ${\cal P}$ is a collection $M = \coll M$ such that each $M(n)$ is, for $n\geq 1$, a $\Sigma_n$-module, together with morphisms \begin{equation} \label{dymka} \nu= \nu_{\rada{m_1}{m_l}}: M(l) \times {\cal P}(m_1)\times \cdots \times {\cal P}(m_l) \longrightarrow M(m_1+\cdots+ m_l), \end{equation} given for any $l, m_1,\ldots,m_l \geq 1$. The structure maps $\nu_{\rada{m_1}{m_l}}$ must satisfy the axioms which are obtained by replacing, in the May's definition of an operad~\cite[Definition~3.12]{may:1972}, the first occurrence of ${\cal P}$ by $M$, see~\cite[Definition~1.3]{markl:zebrulka} for details. \end{definition} \begin{remark}{\rm\ Observe the resemblance of the above definition to the definition of an operad. This is due to the fact that right modules over an operad are special cases of general modules, which are abelian groups object in a certain comma category of operads, see the discussion in~\cite[page~1476]{markl:zebrulka}. }\end{remark} \begin{remark}{\rm\ \label{mince} The structure map $\nu$ is, as in the case of operads, determined by the system of `comp' maps \begin{equation} \label{comp} \circ_i :M(m) \otimes {\cal P}(n) \to M(m+n-1),\ m,n \geq 1,\ 1\leq i\leq m, \end{equation} defined by $\circ_i (x,p) := \nu(x;1,\ldots,1,p,1,\ldots,1)$ ($p$ at the $i$-th position) which have to satisfy certain axioms~\cite[Formula~(1)]{markl:zebrulka}. }\end{remark} \begin{example}{\rm\ An operad ${\cal P}$ is a module over itself. Very important nontrivial examples are provided by the Axelrod-Singer compactification of configuration spaces of points in a manifold. The result is a module over the operad of `local' configurations, see~\cite{markl:cf}. There is the following analog of the endomorphism operad. Let $A,W$ be objects of the category ${\cal C}$. Then the collection ${\cal E}_{A,W}= \coll{{\cal E}_{A,W}}$ with ${\cal E}_{A,W}(n):= \Hom{A^{\times n}}W$ is a module over the endomorphism operad ${\cal E}_A$, the module structure being given by the obvious composition of maps, as in the case of the endomorphism operad. A ${\cal P}$-algebra structure $a: {\cal P} \to {\cal E}_A$ on $A$ induces a ${\cal P}$-module structure on ${\cal E}_{A,W}$. }\end{example} We are going now to introduce objects described, in the similar sense as algebras are described by operads, by {\em modules\/} over operads. We will call them, from the reasons which will be explained in Example~\ref{why-traces}, {\em traces\/} over algebras. \begin{definition} \label{el} Let $M$ be a ${\cal P}$-module and let $A$ be a ${\cal P}$-algebra. An $M$-trace over $A$ is a map $t : M \to {\cal E}_{A,W}$ of ${\cal P}$-modules, where ${\cal E}_{A,W}$ has the ${\cal P}$-module structure induced from the ${\cal P}$-algebra structure on $A$. \end{definition} \begin{example}{\rm\ Rather dull examples of traces are given by taking $M= {\cal P}$. For example, an ${\it Ass}$-trace over an associative algebra is (given by) a bilinear map $B : A \times A \to W$ such that $B(ab,c) = B(a,bc)$, $a,b,c\in A$, i.e.~by an (not necessary symmetric) invariant bilinear form. }\end{example} We will need the following notation. Let, for permutations $\sigma \in \Sigma_l$ and $\sigma_i\in \Sigma_{m_i}$, $1\leq i\leq l$, \begin{equation} \label{myska1} \sigma(\rada{\sigma_1}{\sigma_l})\in \Sigma_{m_1+\cdots+ m_l} \end{equation} denote the permutation $\sigma(\rada{m_1}{m_l})\cdot (\sigma_1 \oplus \cdots \oplus \sigma_l)$, where the meaning of $\sigma_1 \oplus \cdots \oplus \sigma_l$ is clear and $\sigma(\rada{m_1}{m_l})$ permutes the blocks of $\rada{m_1}{m_l}$-elements via $\sigma$. This defines a map $\Sigma_l \times \Sigma_{m_1}\squeeze \Sigma_{m_l} \to \Sigma_{m_1+\cdots +m_l}$. \begin{example}{\rm\ \label{tuzka} More interesting example of a trace can be constructed as follows. Take again the (symmetric) operad {\it Ass}\ for associative algebras. Recall that ${\it Ass}(n) = {\bf k}[\Sigma_n]$, the group ring of the symmetric group over the ground field ${\bf k}$. The operad structure map $\gamma = \gamma_{{\it Ass}}$ is defined by $\gamma(\sigma;\rada{\sigma_1}{\sigma_l}) = \sigma(\rada{\sigma_1}{\sigma_l})$, where $\sigma(\rada{\sigma_1}{\sigma_l})$ has the same meaning as in~(\ref{myska1}). The group of cyclic permutations ${\bf Z}_n = {\bf Z}/n{\bf Z}$ acts from the left on $\Sigma_n$. The group $\Sigma_{n-1}$ is imbedded in $\Sigma_n$ as permutations leaving $1$ fixed. This embedding is a cross-section to the ${\bf Z}_n$-action, thus we can identify $\Sigma_{n-1}$ as a {\em set\/} to the coset space ${\bf Z}_n \backslash \Sigma_n$. The projection $\pi_n : \Sigma_n \to {\bf Z}_n \backslash \Sigma_n \cong \Sigma_{n-1}$ then induce on ${\bf k}[\Sigma_{n-1}]$ a structure of a right $\Sigma_n$-module. Define the collection ${\it Cycl} = \coll{{\it Cycl}}$ by ${\it Cycl}(n):= {\bf k}[\Sigma_{n-1}]$, $n\geq 1$. The system of maps $\{\pi_n: \Sigma_n \to \Sigma_{n-1}\}_{n\geq 1}$ is the projection $\pi : {\it Ass} \to {\it Cycl}$ of collections. \begin{lemma} The projection $\pi : {\it Ass} \to {\it Cycl}$ induces on ${\it Cycl}$ the structure of a module over the operad ${\it Ass}$. \end{lemma} \noindent {\bf Proof.} The structure maps $\nu = \nu_{{\it Cycl}}$ are determined, for $\sigma \in \Sigma_l$ and $\sigma_i \in \Sigma_{m_i}$, $1\leq i \leq l$, by \[ \nu(\pi(\sigma);\rada{\sigma_1}{\sigma_l}) := \gamma_{{\it Ass}}(\sigma;\rada{\sigma_1}{\sigma_l}), \] where $\gamma_{{\it Ass}}(\sigma;\rada{\sigma_1}{\sigma_l})= \sigma(\rada{\sigma_1}{\sigma_l})$. The proof is then finished by an easy verification that, if $\sigma' \equiv \sigma''$ mod ${\bf Z}_l$, then \[ \sigma'(\rada{\sigma_1}{\sigma_l}) \equiv \sigma''(\rada{\sigma_1}{\sigma_l}) \mbox{ mod } {\bf Z}_{m_1+\cdots +m_l}, \] which we leave to the reader.\hspace*{\fill }\end{example} \begin{example}{\rm\ \label{why-traces} A ${\it Cycl}$-trace over an associative algebra $A$ is (characterized by) a map $T: A \to W$ such that \begin{equation} \label{nabla} T(ab) = \znamenko{|a|\cdot |b|}T(ba),\ a,b\in A, \end{equation} i.e.~$T$ is a trace in the usual sense. We postpone the verification of this statement to Example~\ref{nabijecka}. }\end{example} In the rest of this section we show that the collection $W := \colla W$ of the cyclohedra is a natural cellular (right) module over the cellular operad $K = \colla K$ and describe $W$-traces on an A($\infty$)- (= $K$)-algebra. It is convenient to consider harmless symmetrizations. Recall that ${\cal B}(n)$ was the poset of all bracketings of $\rada 1n$. Take instead be the poset ${\overline {\cal B}}(n)$ of all bracketings of $\rada{\sigma(1)}{\sigma(n)}$, $\sigma \in \Sigma_n$. Obviously ${\overline {\cal B}}(n) = \Sigma_n \times {\cal B}(n)$ and ${\overline {\cal B}}(n)$ is the poset of faces of the $n!$-connected polyhedron ${\overline K}_n := \Sigma_n\times K_n$. The collection ${\overline K} := \colla {\overline K}$ is a (symmetric) cellular operad and the corresponding operad of cellular chains ${\overline {\cal A}} := CC_*({\overline K})$ is the (symmetric) operad for {A($\infty$)}-algebras. There is a similar symmetrization of the cyclohedron. We introduced ${{\cal B}C}(n)$ as the poset of equivalence classes of bracketings of $\rada {\sigma(1)}{\sigma(n)}$ with a {\em cyclic\/} permutation $\sigma \in \Sigma_n$. If we admit all permutations, we obtain the poset ${\overline {{\cal B}C}}(n)= \Sigma_{n-1}\times {{\cal B}C}(n)$ whose realization is the $(n-1)!$-connected polyhedron ${\overline W}_n := \Sigma_{n-1} \times W_n$. \begin{lemma} \label{paska} The collection ${\overline {{\cal B}C}} := \coll{{\overline {{\cal B}C}}}$ is a natural module over the operad ${\overline {\cal B}}:= \coll{{\overline {\cal B}}}$ in the symmetric monoidal category of posets. \end{lemma} \noindent {\bf Proof.} The easiest way to define the module structure is the following. Let $b$ be a bracketing of $\rada{1}{l}$ representing an element $[b]\in {{\cal B}C}(l)\subset {\overline {{\cal B}C}}(l)$. Let $b_i \in {\cal B}(m_i)\subset {\overline {\cal B}}(m_i)$ be, for $1\leq i \leq l$ , a bracketing of $\rada{1}{m_i}$. Then we define $\nu_{{\overline {{\cal B}C}}}(b;\rada{b_1}{b_l}) \in {\overline {{\cal B}C}}(m)$, $m = m_1+\cdots m_l$, to be the element represented by the composite (in the same sense as in Example~\ref{hrnicek}) bracketing $b(\rada{b_1}{b_l})$ of $m$. The set ${\overline {{\cal B}C}}(l)$ is $\Sigma_l$-generated by elements of the same form as $b$, i.e.~by elements represented by a bracketing of the `unpermuted' string $\rada1l$, and the same is true also for ${\overline {\cal B}}(m_i)$, $1\leq i\leq l$. Thus the formula for the composition of arbitrary elements is dictated by the equivariance of the module composition map. We leave the verification that this definition is correct to the reader. \hspace*{\fill \begin{theorem} \label{sirky} The collection ${\overline W} := \colla{{\overline W}}$ carries a structure of a module over the operad ${\overline K} := \colla{{\overline K}}$ in the category of cellular complexes which induces, on the level of the poset of faces, the structure of Lemma~\ref{paska}. The homology collection $H_*({\overline W})= \{H_*({\overline W}_n)\}_{n\geq 1}$ coincides, as an ${\it Ass} = H_*({\overline K})$-module, to the module ${\it Cycl}$ introduced in Example~\ref{why-traces}. \end{theorem} \noindent {\bf Proof.} The proof is a modification of the proof of the existence of an operad structure on the associahedron, given by J.~Stasheff in~\cite[page~278]{stasheff:TAMS63}. By Remark~\ref{mince}, the ${\overline K}$-module structure on ${\overline W}$ is given by the `comp' maps \[ \circ_i :{\overline W}_m \times {\overline K}_n \to {\overline W}_{m+n-1},\ m,n\geq 1,\ 1\leq i\leq m. \] As a matter of fact, in our case it is enough to specify \begin{equation} \label{vrsek} \circ_1 :W_m \times K_n \to W_{m+n-1}, \end{equation} the remaining `comp' maps are determined by the equivariance. We define $\circ_1$ of~(\ref{vrsek}) to be the identification of the product $W_m \times K_n$ to the face of $W_{m+n-1}$ indexed by the bracketing $b_{n,n+m-1}$ of~(\ref{csa}). The second part is immediate.\hspace*{\fill \begin{observation}{\rm\ The ${\overline K}$-module structure on the `symmetrized' cyclohedron ${\overline W}$ restricts to the right action of the nonsymmetric operad $K$ on the `nonsymmetric' cyclohedron $W$. This is the structure observed by J.~Stasheff in~\cite[Section~4]{stasheff:from-ops}. Another argument for the existence of the module structure of Theorem~\ref{sirky} is the interpretation of the cyclohedron to the compactification of a configuration space, see Section~\ref{22}. }\end{observation} Let us consider the ${\overline {\cal A}}$-module ${\cal M}:= CC_*({\overline W})$ of cellular chains on the cyclohedron. To describe traces over ${\cal M}$, it is convenient to accept the following notation. For graded indeterminates $\rada{a_1}{a_n}$ and a permutation $\sigma\in \Sigma_n$, the {\em Koszul sign\/} $\epsilon(\sigma)=\epsilon(\sigma;\rada{a_1}{a_n})$ is defined by \[ a_1\land\dots\land a_n = \epsilon(\sigma;a_1,\dots,a_n) \cdot a_{\sigma(1)}\land \dots \land a_{\sigma(n)}, \] which has to be satisfied in the free graded commutative algebra $\mbox{\large$\land$}(\rada{a_1}{a_n})$. Denote also \[ \chi(\sigma)=\chi(\sigma;\rada{a_1}{a_n}) := {\rm sgn}(\sigma)\cdot \epsilon(\sigma;\rada{a_1}{a_n}). \] For an expression $X(\rada {a_1}{a_n})$ in indeterminates $\rada {a_1}{a_n}$, let the {\em cyclic sum\/} \begin{equation} \label{cyclsum} \cyclsum X(\rada{a_1}{a_n}) := \sum_\sigma \chi(\sigma) X(\rada {a_{\sigma(1)}}{a_{\sigma(n)}}) \end{equation} be the summation over all cyclic permutations. A convincing example of the use of this convention is the (graded) Jacobi identity written as \[ \cyclsum [a_1,[a_2,a_3]] = 0. \] The following proposition, whose proof we postpone after Theorem~\ref{ucpavka}, describes ${\cal M}$-traces over {A($\infty$)}-algebras. \begin{proposition} \label{nuzky} Let $A = (A; m_1=\partial,m_2,m_3,\ldots)$ be an {A($\infty$)}-algebra~(\cite[page~294]{stasheff:TAMS63}, but we use the sign convention of~\cite[\S1.4]{markl:JPAA92}). Then an ${\cal M}$-trace is given by a differential graded vector space $W = (W,\delta)$, $\deg{\delta} = -1$, and a system $T_n :A^{\otimes n}\to W$ of degree-$(n-1)$ linear maps, $n\geq 1$, such that, for all $\rada{a_1}{a_n} \in A$, \begin{itemize} \item[(i)] $T_n(\rada{a_1}{a_n})= \chi(\sigma)\cdot T_n(\rada{a_{\sigma(1)}}{a_{\sigma(n)}})$ for all cyclic permutations $\sigma \in \Sigma_n$, and \item[(ii)] for all $n\geq 1$, \begin{equation} \label{Ax} \delta T_n(\rada{a_1}{a_n}) = \cyclsum \sum_{1\leq k\leq n} \znamenko{k+n} T_{n-k+1}(m_k(\rada{a_1}{a_{k}}), \rada{a_{k+1}}{a_n}). \end{equation} \end{itemize} We call such objects homotopy traces over an {A($\infty$)}-algebra $A$. \end{proposition} \noindent Let us write down the axiom~(\ref{Ax}) explicitly for small $n$. For $n=1$ it gives \[ \partial T_1(a) = T_1 (\delta(a)),\ a\in A, \] which means that $T_1$ is a homomorphism of differential graded spaces $(A,\partial)$ and $(W,\delta)$. For $n=2$ it becomes \[ \delta T_2(a,b) +T_2(\partial a,b) -\znamenko{|a|\cdot |b|} T_2(\partial b,a) = T_1(m_2(a,b)) -\znamenko{|a|\cdot |b|} T_1(m_2(b,a)),\ a,b\in A, \] i.e.~$T_1$ is a trace for the `multiplication' $m_2$ up to a homotopy $T_2$. For higher $n$'s, the axiom~(\ref{Ax}) represents `coherence conditions' for the homotopy $T_2$. An important special case is when $A$ is an ordinary associative algebra, that is the only nontrivial structure map is $m_2$, which is an associative multiplication $\cdot$. The axioms for the corresponding trace are obtained by putting, in Proposition~\ref{nuzky}, $m_k=0$ for $k\geq 3$. We also substitute $(-1)^\frac{n(n-1)}{2}T_n$ for $T_n$, to get rid of the overall sign $(-1)^n$. A homotopy trace is then a system $\{T_n:A^{\otimes n}\to W \}_{n\geq 1}$ of degree-$(n-1)$ linear maps, satisfying~\ref{nuzky}(i) and \begin{eqnarray} \delta T_1(a)\!\! &=&\!\! 0\nonumber \\ \delta T_2 (a,b)\!\! &=& \!\! T_1(a\cdot b)-\znamenko{|a|\cdot |b|} T_1(b\cdot a)\nonumber \\ \delta T_3 (a,b,c)\!\! &=&\!\!T_2(a\cdot b,c) +\znamenko{|a|\cdot(|b|+|c|)} T_2(b\cdot c,a) +\znamenko{|c|\cdot(|a|+|b|)} T_2(c\cdot a,b)\nonumber \\ &\vdots&\nonumber \\ \label{Blatter} \hskip5mm \delta T_n(\rada {a_1}{a_n})\!\!&=& \!\! \cyclsum T_{n-1}(a_1\cdot a_2,\rada{a_3}{a_n}),\ n\geq 4. \end{eqnarray} Equation~(\ref{Blatter}) describes a trace over a certain ${\it Ass}$-module ${\overline {\cal D}}$, closely related to the simplex. The following Section~\ref{hrnicek1} is devoted to the study of this module. \section{Simplex as a module over the operad for associative algebras} \label{hrnicek1} Let $\Delta_n$ be the standard $(n-1)$-dimensional simplex (!observe that the conventional notation for our $\Delta_n$ is $\Delta^{n-1}$!). An explicit description of $\Delta_n$ is the following. Denote, for $1\leq i\leq n$, by $e_i$ the point $(\rada 00,1,\rada 00)\in {\bf R}^n$ ($1$ at the $i$-th position). Then $\Delta_n \subset {\bf R}^n$ is the convex hull of the set $\{\rada{e_1}{e_n}\}$. Figure~\ref{simplex} shows $\Delta_n$ for $n=3$. \begin{figure}[hbtp] \begin{center} \unitlength 1.20mm \begin{picture}(34.44,46.16) \unitlength 1.6mm \thinlines \put(13.83,13.16){\line(-5,-3){23.83}} \put(10.61,13.28){\line(5,-3){23.83}} \put(12.33,10.33){\line(0,1){25.83}} \put(12.33,27.16){\makebox(0,0)[cc]{$\bullet$}} \put(-4.00,2.33){\makebox(0,0)[cc]{$\bullet$}} \put(28.83,2.33){\makebox(0,0)[cc]{$\bullet$}} \put(-4.06,-0.62){\makebox(0,0)[cc]{$\{1\}$}} \put(28.83,-0.78){\makebox(0,0)[cc]{$\{2\}$}} \put(9.33,29.66){\makebox(0,0)[cc]{$\{3\}$}} \put(1.66,17.66){\makebox(0,0)[cc]{$\{31\}$}} \put(22.66,18.16){\makebox(0,0)[cc]{$\{23\}$}} \put(10.83,-1.28){\makebox(0,0)[cc]{$\{12\}$}} \put(10.50,6.33){\makebox(0,0)[cc]{$\{123\}$}} \put(12.33,27.33){\line(-1,0){1.50}} \put(10.83,27.33){\line(1,0){3.00}} \put(16.17,13.16){\makebox(0,0)[cc]{$\cdot$}} \put(3.83,11.00){\makebox(0,0)[cc]{$\cdot$}} \put(13.00,8.83){\makebox(0,0)[cc]{$\cdot$}} \put(7.33,11.50){\makebox(0,0)[cc]{$\cdot$}} \put(11.33,16.66){\makebox(0,0)[cc]{$\cdot$}} \put(8.17,12.33){\makebox(0,0)[cc]{$\cdot$}} \put(6.67,15.83){\makebox(0,0)[cc]{$\cdot$}} \put(20.50,9.33){\makebox(0,0)[cc]{$\cdot$}} \put(21.50,5.83){\makebox(0,0)[cc]{$\cdot$}} \put(17.33,5.00){\makebox(0,0)[cc]{$\cdot$}} \put(10.50,18.00){\makebox(0,0)[cc]{$\cdot$}} \put(14.00,18.50){\makebox(0,0)[cc]{$\cdot$}} \put(0.83,6.16){\makebox(0,0)[cc]{$\cdot$}} \put(-5.22,3.50){\line(1,-1){2.44}} \put(27.56,0.94){\line(1,1){3.00}} \put(2.56,3.83){\makebox(0,0)[cc]{$\cdot$}} \put(17.67,16.17){\makebox(0,0)[cc]{$\cdot$}} \put(11.00,22.94){\makebox(0,0)[cc]{$\cdot$}} \thicklines \put(12.33,27.33){\line(-2,-3){16.67}} \put(-4.34,2.33){\line(1,0){33.33}} \put(29.00,2.33){\line(-2,3){16.78}} \end{picture} \end{center} \caption{$\Delta_3$.\label{simplex}} \end{figure} There is a classical correspondence between the poset of subsets of $\{\rada1n\}$ and the poset of faces of $\Delta_n$ given by \[ \mbox{subset $S$ of $\{\rada1n\}$} \longleftrightarrow \mbox{convex hull of the set $\{e_i\}_{i\in S}\subset {\bf R}^n$.} \] See~\cite[\S10.1]{switzer:75} for details. Let $\Delta := \{\Delta_n\}_{\geq 1}$. In fact, it is more convenient to consider the symmetrized version ${\overline \Delta} := \{{\overline \Delta}_n\}_{\geq 1}$, where ${\overline \Delta}_n$ is the disjoint union of $(n-1)!$ copies of $\Delta_n$, indexed by cyclic orders of its vertices. This means that the poset of faces of ${\overline \Delta}_n$ consists of elements of the form \begin{equation} \label{bookshop} \set{\rada{i_1}{i_l}}\times [\sigma], \end{equation} where $\set{\rada{i_1}{i_l}}$ is a subset of $\set{\rada 1n}$, $1\leq l\leq n$, and $[\sigma]$ is an equivalence class from the left coset ${\bf Z}_n \backslash \Sigma_n$. We define the right action of $\Sigma_n$ by \[ (\set{\rada{i_1}{i_l}}\times [\sigma])\cdot \rho := \set{\rada{\rho^{-1}(i_1)}{\rho^{-1}(i_l)}}\times [\sigma\rho]. \] Let $e = \set{\rada{i_1}{i_l}}\times [\sigma]$ be a face (= cell) of ${\overline \Delta}(n)$ as in~(\ref{bookshop}). An orientation of $e$ is given by choosing an order of elements of $\set{\rada{i_1}{i_l}}$. Two such orders induce the same orientation if and only if they differ by a permutation of signature $+1$. Thus the cellular cell complex ${\overline {\cal D}}(n) := CC_*({\overline \Delta}(n))$ is a vector space with the basis \begin{equation} \label{Y} \langle\rada{i_1}{i_l}\rangle\times [\sigma] \end{equation} where $\langle\rada{i_1}{i_l}\rangle$ denotes the cell $\set{\rada{i_1}{i_l}}\times [\sigma]$ with the orientation induced by the order $i_1< \cdots < i_l$. The right action of $\Sigma_n$ is given by \[ (\langle\rada{i_1}{i_l}\rangle\times [\sigma])\cdot \rho := (\langle\rada{\rho^{-1}(i_1)}{\rho^{-1}(i_l)}\rangle)\times [\sigma\rho]. \] \begin{theorem} \label{vicko} The collection ${\overline \Delta} := \{{\overline \Delta}_n\}_{\geq 1}$ of cell complexes has a natural structure of a (right) module over the operad ${\it Ass}$ for associative algebras. The traces over the cellular chain complex ${\overline {\cal D}} := CC_*({\overline \Delta})$ are the objects described by~(\ref{Blatter}). \end{theorem} \noindent {\bf Proof.} We observed in Remark~\ref{mince} that the action is determined by a system of `comp' maps \[ \circ_i : {\overline \Delta}_n \otimes {\it Ass}(m) \to {\overline \Delta}_{m+n-1},\ n,m\geq 1. \] Since ${\overline \Delta}_n$ is $\Sigma_n$-generated by $\Delta_n = \Delta_n \times [\mbox{$1\!\!1$}_n] \subset {\overline \Delta}_n$ (the copy corresponding to the `normal' cyclic order $(\rada 1n)$) and ${\it Ass}(m) = {\bf k}[\Sigma_n]$ is $\Sigma_m$-generated by the identity permutation $\mbox{$1\!\!1$}_m \in \Sigma_m$, it is enough to specify $\circ_i(t,x)$ for $t\in \Delta_n$ and $x = \mbox{$1\!\!1$}_m$. We define $\circ_i(-,\mbox{$1\!\!1$}_m): \Delta_n \to \Delta_{m+n-1}$ to be the unique simplicial map such that \[ \circ_i(\set j \times [\mbox{$1\!\!1$}_n], [\mbox{$1\!\!1$}_m]) := \left\{ \begin{array}{ll} \set j,& \mbox{ for $1 \leq j \leq i$, and} \\ \set{j+m-1},& \mbox{ for $i < j \leq n$.} \end{array} \right. \] In other words, $\circ_i(-,\mbox{$1\!\!1$}_m)$ is the canonical inclusion $\Delta_n \hookrightarrow \Delta_{m+n-1}$, identifying $\Delta_n$ to the $(n-1)$-dimensional face of $\Delta_m$ corresponding to the subset $\{\rada 1i,\rada{i+m}{n+m-1}\}$. The induced map of the cellular chain complex satisfies \begin{equation} \label{napoleon} \circ_i(\langle \rada 1n \rangle \times [\mbox{$1\!\!1$}_n], [\mbox{$1\!\!1$}_m])= \langle \rada 1i,\rada{i+m}{n+m-1}\rangle \times [\mbox{$1\!\!1$}_{m+n-1}]. \end{equation} It is a straightforward verification to prove that this really defines an ${\it Ass}$-action. The second part will be proved after we formulate Theorem~\ref{resiz}.\hspace*{\fill We are going to give an algebraic characterization of the ${\it Ass}$-module ${\overline {\cal D}}$. To do this, we need some more or less standard notions, which we will also find useful later. From now on, if not stated otherwise, the underlying symmetric monoidal category will be the category of (differential) graded vector spaces. For any collection $E= \coll E$ there exists the {\em free operad\/} ${\cal F}(E)$ on $E$~\cite[page~226]{ginzburg-kapranov:DMJ94}. The operad ${\cal F}(E)$ has the following very explicit description in terms of trees. Denote by ${\cal T}$ the set of (labeled rooted) trees and by ${\cal T}_n$ the subset of ${\cal T}$ consisting of trees having $n$ input edges. Let $E({\tt T})$ denote, for ${\tt T} \in {\cal T}$, the set of `multilinear' colorings of the vertices of ${\tt T}$ by the elements of $E$ such that a vertex with $k$ input edges is colored by an element of $E(k)$. The free operad ${\cal F}(E)$ on $E$ may be then described as \begin{equation} \label{fax} {\cal F}(E)(n):= \bigoplus_{{\tt T}\in{\cal T}_n}E({\tt T}) \end{equation} with the operad structure on ${\cal F}(E)$ given by the operation of `grafting' trees. We will in fact always assume that $E(1)= 0$, thus we consider in~(\ref{fax}) only trees whose all vertices are at least binary, i.e.\ they have at least two incoming edges. The details may be found in~\cite{ginzburg-kapranov:DMJ94,getzler-jones:preprint}. \begin{example} \label{zajicek_usacek}{\rm\ The set ${\cal T}_2$ has only one element (Figure~\ref{t2andt3}) and ${\cal F}(E)(2)= E(2)$. The set ${\cal T}_3$ has four elements (see again Figure~\ref{t2andt3}) \begin{figure}[hbtp] \begin{center} \unitlength 0.80mm \thicklines \begin{picture}(157.62,22.95) \put(7.62,22.95){\line(0,-1){10.00}} \put(7.62,12.95){\line(-1,-1){10.00}} \put(7.62,12.95){\line(1,-1){10.00}} \put(57.62,22.95){\line(0,-1){10.00}} \put(57.62,12.95){\line(-1,-1){10.00}} \put(57.62,12.95){\line(1,-1){10.00}} \put(52.62,7.95){\line(1,-1){5.00}} \put(7.62,12.95){\makebox(0,0)[cc]{$\bullet$}} \put(52.62,7.95){\makebox(0,0)[cc]{$\bullet$}} \put(57.62,12.95){\makebox(0,0)[cc]{$\bullet$}} \put(-2.38,-2.05){\makebox(0,0)[cc]{$1$}} \put(17.62,-2.05){\makebox(0,0)[cc]{$2$}} \put(47.62,-2.05){\makebox(0,0)[cc]{$1$}} \put(57.62,-2.05){\makebox(0,0)[cc]{$2$}} \put(67.62,-2.05){\makebox(0,0)[cc]{$3$}} \put(87.62,22.95){\line(0,-1){10.00}} \put(87.62,12.95){\line(-1,-1){10.00}} \put(87.62,12.95){\line(1,-1){10.00}} \put(82.62,7.95){\line(1,-1){5.00}} \put(82.62,7.95){\makebox(0,0)[cc]{$\bullet$}} \put(87.62,12.95){\makebox(0,0)[cc]{$\bullet$}} \put(77.62,-2.05){\makebox(0,0)[cc]{$2$}} \put(87.62,-2.05){\makebox(0,0)[cc]{$3$}} \put(97.62,-2.05){\makebox(0,0)[cc]{$1$}} \put(117.62,22.95){\line(0,-1){10.00}} \put(117.62,12.95){\line(-1,-1){10.00}} \put(117.62,12.95){\line(1,-1){10.00}} \put(112.62,7.95){\line(1,-1){5.00}} \put(112.62,7.95){\makebox(0,0)[cc]{$\bullet$}} \put(117.62,12.95){\makebox(0,0)[cc]{$\bullet$}} \put(107.62,-2.05){\makebox(0,0)[cc]{$3$}} \put(117.62,-2.05){\makebox(0,0)[cc]{$1$}} \put(127.62,-2.05){\makebox(0,0)[cc]{$2$}} \put(147.62,22.95){\line(0,-1){10.00}} \put(147.62,12.95){\line(-1,-1){10.00}} \put(147.62,12.95){\line(1,-1){10.00}} \put(147.62,12.95){\line(0,-1){10.00}} \put(147.62,12.95){\makebox(0,0)[cc]{$\bullet$}} \put(137.62,-2.05){\makebox(0,0)[cc]{$1$}} \put(147.62,-2.05){\makebox(0,0)[cc]{$2$}} \put(157.62,-2.05){\makebox(0,0)[cc]{$3$}} \end{picture} \end{center} \caption{The sets ${\cal T}_2$ (left) and ${\cal T}_3$ (right).\label{t2andt3}} \end{figure} and ${\cal F}(E)(3)$ consists of three copies of $E(2)\otimes E(2)$ which corresponds to the three binary trees in ${\cal T}_3$ and one copy of $E(3)$ corresponding to the corolla (= the tree with one vertex). Compare also~\cite[Figure~7]{ginzburg-kapranov:DMJ94}. }\end{example} In the same manner, for each operad ${\cal P}$ and for each collection $X = \coll X$ there exists the {\em free (right) ${\cal P}$-module\/} generated by the collection $X$, which we denote $X\circ {\cal P}$. An explicit description is~\cite[page~312]{markl:dl} \begin{equation} \label{cajicek} (X\circ {\cal P})(m) = \def.7{.7} \bigoplus \left( \mbox{Ind}^{\Sigma_m}_{\Sigma_{m_1} \times \cdots \times \Sigma_{m_l}} (X(l)\otimes {\cal P}(m_1)\otimes \cdots \otimes {\cal P}(m_l)) \right)_{\Sigma_l}, \end{equation} where the summation is taken over all $m_1+\cdots +m_l =n$, $l\geq 1$. On the right-hand side, $\mbox{Ind}^{\Sigma_m}_{\Sigma_{k_1} \times \cdots \times \Sigma_{k_l}}(-)$ denotes the induced representation and $(-)_{\Sigma_l}$ the quotient under the obvious action of $\Sigma_l$. The term $X(l)\otimes {\cal P}(m_1)\otimes \cdots \otimes {\cal P}(m_l)$ on the right-hand side of~(\ref{cajicek}) can be interpreted as colorings of the tree ${\tt T}_{m_1,\ldots,m_l}$ from Figure~\ref{thetree} such that the output vertex is colored by an element of $X(l)$ and the remaining vertices by elements of ${\cal P}$. \begin{figure}[hbtp] \begin{center} \unitlength 1.00mm \thicklines \begin{picture}(59.69,22.89) \put(6.69,5.39){\line(-6,-5){6.83}} \put(6.69,5.56){\line(-1,-2){2.75}} \put(6.69,5.23){\line(1,-1){5.33}} \put(17.02,5.73){\line(-2,-5){2.20}} \put(17.02,5.73){\line(1,-1){5.83}} \put(53.02,6.39){\line(-6,-5){7.50}} \put(52.85,6.56){\line(-2,-5){2.73}} \put(52.69,6.73){\line(1,-1){7.00}} \put(31.69,20.39){\makebox(0,0)[cc]{$\bullet$}} \put(6.52,5.39){\makebox(0,0)[cc]{$\bullet$}} \put(17.02,5.73){\makebox(0,0)[cc]{$\bullet$}} \put(52.85,6.56){\makebox(0,0)[cc]{$\bullet$}} \put(34.69,12.73){\makebox(0,0)[cc]{$\cdots$}} \put(7.35,-0.44){\makebox(0,0)[cc]{$\cdots$}} \put(19.52,-0.61){\makebox(0,0)[cc]{$\cdots$}} \put(54.19,-0.61){\makebox(0,0)[cc]{$\cdots$}} \put(32.71,22.89){\makebox(0,0)[lb]{$l$ inputs}} \put(4.85,7.23){\makebox(0,0)[rb]{$m_1$ inputs}} \put(19.85,6.06){\makebox(0,0)[lb]{$m_2$ inputs}} \put(55.52,8.56){\makebox(0,0)[lb]{$m_l$ inputs}} \put(6.59,5.41){\line(0,1){15.19}} \put(52.90,6.41){\line(0,1){14.18}} \put(52.90,20.59){\line(-1,0){46.32}} \put(27.50,20.62){\line(0,-1){9.30}} \put(17.19,5.53){\line(-1,-6){0.92}} \put(17.02,20.56){\line(0,-1){15.17}} \end{picture} \end{center} \caption{The tree ${\tt T}_{m_1,\ldots,m_l}$. The output vertex is symbolized as a `rake' with no output edge, to underline its distinguished character. \label{thetree}} \end{figure} \begin{example} \label{haficek} {\rm\ We have $(X\circ {\cal P})(1) = X(1)$, corresponding to the tree ${\tt T}_1$ on Figure~\ref{odrazka}. \begin{figure}[hbtp] \begin{center} \unitlength 1.00mm \thicklines \begin{picture}(142.67,43.00) \put(131.67,15.00){\line(-1,-1){10.00}} \put(131.67,15.33){\line(1,-1){10.33}} \put(12.34,35.00){\makebox(0,0)[cc]{$\bullet$}} \put(72.01,34.38){\makebox(0,0)[cc]{$\bullet$}} \put(131.67,34.33){\makebox(0,0)[cc]{$\bullet$}} \put(12.34,15.00){\makebox(0,0)[cc]{$\bullet$}} \put(59.29,15.33){\makebox(0,0)[cc]{$\bullet$}} \put(83.34,15.67){\makebox(0,0)[cc]{$\bullet$}} \put(131.67,15.00){\makebox(0,0)[cc]{$\bullet$}} \put(-0.00,42.33){\makebox(0,0)[cc]{${\tt T}_1:$}} \put(57.67,42.66){\makebox(0,0)[cc]{${\tt T}_{1,1}:$}} \put(118.67,43.00){\makebox(0,0)[cc]{${\tt T}_2:$}} \put(16.34,36.33){\makebox(0,0)[lc]{$X(1)$}} \put(16.00,17.33){\makebox(0,0)[lc]{${\cal P}(1)={\bf k}$}} \put(74.36,38.21){\makebox(0,0)[lc]{$X(2)$}} \put(62.67,12.33){\makebox(0,0)[lc]{${\cal P}(1)={\bf k}$}} \put(87.01,19.33){\makebox(0,0)[lc]{${\cal P}(1)={\bf k}$}} \put(135.33,37.00){\makebox(0,0)[lc]{$X(1)$}} \put(135.33,18.00){\makebox(0,0)[lc]{${\cal P}(2)$}} \put(12.34,0.00){\makebox(0,0)[cc]{$1$}} \put(59.34,0.00){\makebox(0,0)[cc]{$1$}} \put(83.34,0.33){\makebox(0,0)[cc]{$2$}} \put(122.00,0.00){\makebox(0,0)[cc]{$1$}} \put(142.67,0.00){\makebox(0,0)[cc]{$2$}} \put(12.34,35.00){\line(0,-1){29.00}} \put(59.34,34.33){\line(1,0){24.00}} \put(83.34,34.33){\line(0,-1){28.00}} \put(59.34,34.33){\line(0,-1){27.67}} \put(131.67,34.33){\line(0,-1){19.33}} \end{picture} \end{center} \caption{$(X\circ {\cal P})(1)$ (left) and $(X\circ {\cal P})(2)$ (right).\label{odrazka}} \end{figure} The vector space $(X\circ {\cal P})(2)$ consists of a copy of $X(2)$ corresponding to ${\tt T}_{1,1}$ and a copy of $X(1)\otimes {\cal P}(2)$ corresponding to ${\tt T}_2$, see again Figure~\ref{odrazka}. Note that we still assume that ${\cal P}(1)= {\bf k}$. }\end{example} For a graded vector space $V=\bigoplus_p V_p$ let $\uparrow\! V$ (resp. $\downarrow\! V$) be the {\em suspension\/} (resp. the {\em desuspension\/}) of $V$, i.e. the graded vector space defined by $(\uparrow\! V)_p := V_{p-1}$ (resp. $(\downarrow\! V)_p := V_{p+1}$). We have the obvious natural maps $\uparrow : V \to \uparrow\! V$ and $\downarrow: V\to \downarrow\! V$. For a collection $E$, the {\em suspension\/} ${\bf s\hskip0mm} E$ is the collection with \begin{equation} \label{sign-factor} ({\bf s\hskip0mm} E)(n):= {\rm sgn}\ \otimes \uparrow^{n-1}E(n), \end{equation} $n\geq 1$, where $\uparrow^{n-1}$ is the $(n-1)$-fold suspension introduced above and ${\rm sgn}$ is the signum representation of $\Sigma_n$ on ${\bf k}$. The reason why we need the signum factor is that we intend to apply the suspension to operads and modules over operads. Without this factor, the composition induced on the suspension will not be equivariant, compare also~\cite[page~8]{getzler-jones:preprint}. There is an obvious similar notion of the {\em desuspension\/} ${\bf s\hskip0mm}^{-1} E$ of the collection $E$. Let us come back to the promised algebraic characterization of the ${\it Ass}$-module ${\overline {\cal D}}$. Consider the free ${\it Ass}$-module ${\bf s\hskip0mm}{\it Cycl} \circ {\it Ass}$ generated by the suspension of the collection ${\it Cycl}$ introduced in Example~\ref{tuzka}. It will be useful to have an explicit description of the elements of ${\bf s\hskip0mm} {\it Cycl} \circ {\it Ass}$. Consider the free graded right $\Sigma_n$-module $H(n)$ generated by the trees ${\tt T}_{m_1,\ldots,m_l}$ introduced in Figure~\ref{thetree}, with $m_1+\cdots+m_l=n$. The grading is given by $\deg({\tt T}_{m_1,\ldots,m_l}):= l-1$ Thus $H(n)$ is, by definition, the graded vector space with the basis \[ \left\{ {\tt T}_{m_1,\ldots,m_l}\times \sigma,\ \sigma \in \Sigma_n,\ 1\leq l\leq n,\ m_1+\cdots+m_l=n \right\},\ \deg({\tt T}_{m_1,\ldots,m_l}\times \sigma)= l-1. \] A neat graphical presentation of the symbol ${\tt T}_{m_1,\ldots,m_l}\times \sigma$ is the tree ${\tt T}_{m_1,\ldots,m_l}$ with the inputs labeled by $\rada{\sigma^{-1}(1)}{\sigma^{-1}(n)}$. For example, ${{\tt T}_{2,1}}\times (123)$ can be depicted as \begin{center} \unitlength 1.20mm \thicklines \begin{picture}(10,13)(53,23) \put(52.63,33.53){\line(1,0){8.67}} \put(61.30,33.53){\line(0,-1){4.33}} \put(61.30,29.19){\line(-1,-1){4.00}} \put(61.30,29.19){\line(1,-1){4.00}} \put(52.63,33.53){\line(0,-1){8.00}} \put(52.66,22.95){\makebox(0,0)[cc]{2}} \put(57.19,22.95){\makebox(0,0)[cc]{3}} \put(57.19,33.53){\makebox(0,0)[cc]{$\bullet$}} \put(61.35,28.95){\makebox(0,0)[cc]{$\bullet$}} \put(65.21,22.95){\makebox(0,0)[cc]{1}} \end{picture} \end{center} Define now the left action of the group ${\bf Z}_l$, considered as the group of cyclic permutations of order $l$, on $H(n)$ as follows. Recall that, for $\zeta \in {\bf Z}_l \subset \Sigma_l$, we denoted by $\zeta(\rada{m_1}{m_l}) \in \Sigma_n$ the permutation which permutes the blocks of $\rada{m_1}{m_l}$-elements via $\zeta$. Then we put \begin{equation} \label{pejska_Mikinka} \zeta \left( \tttr{m_1}{m_l} \times \sigma \right) := {\rm sgn}(\zeta) \cdot \tttr{m_{\zeta^{-1}(1)}}{m_{\zeta^{-1}(l)}} \times \zeta(\rada{m_1}{m_l})\cdot\sigma. \end{equation} The following lemma is an easy consequence of definitions and formula~(\ref{cajicek}). \begin{lemma} \label{MGD} The graded vector space $({\bf s\hskip0mm}{\it Cycl} \circ {\it Ass})(n)$ can be identified with the graded vector space with the basis given by equivalence classes of symbols \begin{equation} \label{3M} \tttr{m_1}{m_l} \times \sigma \in H(n),\ \deg({\tt T}_{m_1,\ldots,m_l})= l-1, \end{equation} modulo the left action of the group ${\bf Z}_l$ defined in~(\ref{pejska_Mikinka}). Under this identification, the right action of $\Sigma_n$ on the equivalence class $[\tttr{m_1}{m_l} \times \sigma]$ is described as \[ [\tttr{m_1}{m_l} \times \sigma]\cdot \rho := {\rm sgn}(\sigma)\cdot [\tttr{m_1}{m_l} \times \sigma\rho]. \] \end{lemma} Let $\xi_n\in {\it Cycl}(n)$ be the generator represented by the identical permutation $\mbox{$1\!\!1$}_n\in \Sigma_n$ and let $\alpha_2 = \mbox{$1\!\!1$}_2 \in {\it Ass}(2) = {\bf k}[\Sigma_2]$. Define the differential $\partial$ on ${\bf s\hskip0mm}{\it Cycl} \circ {\it Ass}$ by \begin{eqnarray} \label{guma} \partial (\uparrow\!^{n-1} \xi_n) &:=& -\sum_{\sigma} {\rm sgn}(\sigma)\cdot \nu(\uparrow\!^{n-2} \xi_{n-1};\alpha_2,\rada 11)\cdot \sigma \\ \nonumber &=&-\cyclsum \nu(\uparrow\!^{n-2}\xi_{n-1};\alpha_2,\rada 11).\ \mbox{ /the cyclic sum notation of~(\ref{cyclsum})/} \end{eqnarray} Using the identification of Lemma~\ref{MGD}, this could be also written as \[ \partial ({{\tt T}}_{\underbrace% {\mbox{\scriptsize $1,\ldots,1$}}_{n\times}} \times [\mbox{$1\!\!1$}_n])= -\cyclsum ({{\tt T}}_{2,\underbrace% {\mbox{\scriptsize $1,\ldots,1$}}_{n-1\times}}\times[\mbox{$1\!\!1$}_n] ), \] or, in a diagrammatic shorthand, \vskip2mm \begin{center} \unitlength 1.30mm \thicklines \begin{picture}(69.67,5.67) \put(10.00,5.67){\line(0,0){0.00}} \put(10.00,5.67){\line(1,0){15.00}} \put(25.00,5.67){\line(0,-1){4.00}} \put(10.00,1.67){\line(0,1){4.00}} \put(13.00,1.67){\line(0,1){4.00}} \put(8.33,3.67){\makebox(0,0)[rc]{$\partial($}} \put(26.83,3.67){\makebox(0,0)[lc]{$)$}} \put(30.67,3.67){\makebox(0,0)[cc]{$=$}} \put(17.67,5.67){\makebox(0,0)[cc]{$\bullet$}} \put(10.00,-0.16){\makebox(0,0)[cc]{$1$}} \put(13.00,-0.16){\makebox(0,0)[cc]{$2$}} \put(16.00,-0.16){\makebox(0,0)[cc]{$3$}} \put(25.00,-0.16){\makebox(0,0)[cc]{$n$}} \put(20.33,2.17){\makebox(0,0)[cc]{$\cdots$}} \put(16.00,1.67){\line(0,1){4.00}} \put(16.00,5.67){\line(0,-1){4.00}} \put(-10,0){ \put(54.67,5.67){\line(1,0){15.00}} \put(69.67,5.67){\line(0,-1){4.00}} \put(54.67,5.67){\line(0,-1){1.00}} \put(59.00,5.67){\line(0,-1){4.00}} \put(49.67,3.67){\makebox(0,0)[rc]{$-\ \displaystyle\cyclsum$}} \put(64.50,2.34){\makebox(0,0)[cc]{$\cdots$}} \put(54.67,4.59){\line(-2,-5){1.20}} \put(54.67,4.50){\line(1,-3){0.94}} \put(55.61,1.67){\line(0,0){0.06}} \put(53.17,0.00){\makebox(0,0)[cc]{$1$}} \put(56.17,0.00){\makebox(0,0)[cc]{$2$}} \put(59.17,0.00){\makebox(0,0)[cc]{$3$}} \put(69.50,0.00){\makebox(0,0)[cc]{$n$}} \put(62.33,5.67){\makebox(0,0)[cc]{$\bullet$}} } \end{picture} \end{center} Since the elements $\{ \uparrow\!^{n-1} \xi_n\}_{n\geq 1}$ generate ${\bf s\hskip0mm}{\it Cycl} \circ {\it Ass}$, formula~(\ref{guma}) is enough to determine the differential $\partial$. We leave to the reader to verify that the definition is correct and that $\partial^2=0$. \begin{theorem} \label{resiz} The ${\it Ass}$-module ${\overline {\cal D}} = CC_*({\overline \Delta})$ is isomorphic to the free differential ${\it Ass}$-module $({\bf s\hskip0mm}{\it Cycl} \circ {\it Ass},\partial)$ constructed above. \end{theorem} \noindent {\bf Proof.} Any differential ${\it Ass}$-module map $\omega: ({\bf s\hskip0mm}{\it Cycl} \circ {\it Ass},\partial) \to ({\overline {\cal D}},\partial_{{\overline {\cal D}}})$ is determined by the values $\omega(\uparrow\!^{n-1}\xi_n)$, $n\geq 1$. We define $\omega(\uparrow\!^{n-1}\xi_n):= e_n$, where $e_n \in {\overline {\cal D}}(n)$ is the top $(n-1)$-dimensional oriented cell $\langle \rada 1n\rangle$. We shall verify that the above defined map $\omega$ commutes with the differentials, \[ \omega(\partial \uparrow\!^{n-1}\xi_n) = -\omega(\cyclsum \nu(\uparrow\!^{n-2}\xi_{n-1},\alpha_2,\rada11)) = \partial e_n. \] Because $\omega$ is a module homomorphism, the above equation can be rewritten as \begin{equation} \label{hvezdicka} \cyclsum\nu_{{\overline {\cal D}}}(e_{n-1};\alpha_2,\rada11) = -\partial e_n. \end{equation} The standard formula for the boundary of $\langle \rada 1n \rangle$~\cite[\S10.1]{switzer:75} says that \begin{equation} \label{ja} \partial e_n = \partial \langle\rada 1n\rangle = \sum_{1\leq i \leq n} \znamenko{i+1} \langle \rada 1{i-1},\rada {i+1}n\rangle \end{equation} while the defining formula~(\ref{napoleon}) for the ${\it Ass}$-module action on ${\overline \Delta}$ gives \[ \nu_{{\overline {\cal D}}}(e_{n-1};\alpha_2,\rada11) = \langle1,\rada3n\rangle. \] Now it is enough to observe that \[ \cyclsum\langle1,\rada3n\rangle= - \sum_{1\leq i \leq n}\znamenko{i+1}\langle\rada1{i-1},\rada{i+1}n\rangle \] which, together with~(\ref{ja}), gives~(\ref{hvezdicka}). It remains to prove that $\omega$ is an isomorphism. To this end, we give an explicit formula for the map $\omega$. Let $\tttr{m_1}{m_l} \times \sigma \in H(n)$ be as in Lemma~\ref{MGD}. The numbers $\rada{m_1}{m_l}$ determine a sequence $\rada{i_1}{i_l}$ by $i_s:= m_1+\cdots m_{s-1}+1$, $1\leq s\leq l$. Consider a map $\varphi: H(n)\to {\overline {\cal D}}(n)$ defined by \[ \varphi(\tttr{m_1}{m_l} \times \sigma) := \langle\rada{\sigma^{-1}(i_1)}{\sigma^{-1}(i_l)}\rangle \times [\sigma], \] where we denoted elements of ${\overline {\cal D}}(n)$ (= cells of ${\overline \Delta}(n)$) as in~(\ref{Y}). It is immediate to see that $\varphi$ is an $\Sigma_n$-equivariant epimorphism. For the left action of $\zeta \in {\bf Z}_l$ we have \begin{eqnarray*} \varphi(\zeta(\tttr{m_1}{m_l} \times \sigma) &=& {\rm sgn}(\zeta)\!\cdot\! \varphi( \tttr{m_{\zeta^{-1}(1)}}% {m_{\zeta^{-1}(l)}} \times \zeta(\rada{m_1}{m_l})\sigma) \\ &=& {\rm sgn}(\zeta)\!\cdot\! \langle \rada{\sigma^{-1}(i_{\zeta(1)})}{\sigma^{-1}(i_{\zeta(l)})} \rangle \times [\sigma] = \langle\rada{\sigma^{-1}(i_1)}{\sigma^{-1}(i_l)} \rangle\times [\sigma], \end{eqnarray*} which shows that $\varphi(x) = \varphi(\zeta y)$. On the other hand, a moment's reflection show that $\varphi(x) = \varphi(y)$, for $x,y\in H(n)$, implies the existence of some $\zeta \in {\bf Z}_l$ such that $x = \zeta y$. Thus the map $\varphi$ induces an equivariant isomorphism $({\bf s\hskip0mm}{\it Cycl} \circ {\it Ass})(n) = {\bf Z}_l \backslash H(n) \cong {\overline {\cal D}}(n)$, which is exactly our map $\omega$. The nature of the map $\omega$ is illustrated on Figure~\ref{peniz}.\hspace*{\fill \begin{figure}[hbtp] \begin{center} \unitlength 1.20mm \thicklines \begin{picture}(91.08,94.33) \put(-11.20,94.32){\makebox(0,0)[rc]{\fbox{$n=1$:}}} \put(-1.37,94.15){\makebox(0,0)[cc]{$\bullet$}} \put(-1.37,86.65){\makebox(0,0)[cc]{$1$}} \put(-11.20,77.48){\makebox(0,0)[rc]{\fbox{$n=2$:}}} \put(12.30,77.48){\line(1,0){5.00}} \put(17.30,77.48){\line(0,-1){5.00}} \put(12.13,77.48){\line(0,-1){5.00}} \put(14.72,77.48){\makebox(0,0)[cc]{$\bullet$}} \put(-3.87,77.48){\line(1,0){5.00}} \put(1.13,77.48){\line(0,-1){5.00}} \put(-4.04,77.48){\line(0,-1){5.00}} \put(-1.45,77.48){\makebox(0,0)[cc]{$\bullet$}} \put(-4.04,69.65){\makebox(0,0)[cc]{$1$}} \put(1.13,69.65){\makebox(0,0)[cc]{$2$}} \put(12.13,69.65){\makebox(0,0)[cc]{$2$}} \put(17.30,69.65){\makebox(0,0)[cc]{$1$}} \put(7.97,77.48){\makebox(0,0)[cc]{$=-$}} \put(20.79,77.48){\makebox(0,0)[lc]{$=\langle12\rangle,$}} \put(42.46,78.32){\makebox(0,0)[cc]{$\bullet$}} \put(42.46,74.65){\makebox(0,0)[cc]{$\bullet$}} \put(45.84,76.48){\makebox(0,0)[lc]{$= \langle1\rangle,$}} \put(61.76,78.32){\makebox(0,0)[cc]{$\bullet$}} \put(61.63,74.65){\makebox(0,0)[cc]{$\bullet$}} \put(65.01,76.48){\makebox(0,0)[lc]{$=-\langle2\rangle$}} \put(40.13,68.30){\makebox(0,0)[cc]{$1$}} \put(44.96,68.30){\makebox(0,0)[cc]{$2$}} \put(59.30,68.30){\makebox(0,0)[cc]{$2$}} \put(64.30,68.30){\makebox(0,0)[cc]{$1$}} \put(-11.20,56.52){\makebox(0,0)[rc]{\fbox{$n=3$:}}} \put(-6.37,56.69){\line(1,0){10.00}} \put(3.63,56.69){\line(0,-1){5.17}} \put(-6.37,56.69){\line(0,-1){5.17}} \put(-6.37,51.52){\line(0,0){0.00}} \put(-1.37,56.69){\makebox(0,0)[cc]{$\bullet$}} \put(11.13,56.69){\line(1,0){10.00}} \put(28.63,56.69){\line(1,0){10.00}} \put(21.13,56.69){\line(0,-1){5.17}} \put(38.63,56.69){\line(0,-1){5.17}} \put(11.13,56.69){\line(0,-1){5.17}} \put(28.63,56.69){\line(0,-1){5.17}} \put(11.13,51.52){\line(0,0){0.00}} \put(28.63,51.52){\line(0,0){0.00}} \put(16.13,56.69){\makebox(0,0)[cc]{$\bullet$}} \put(33.63,56.69){\makebox(0,0)[cc]{$\bullet$}} \put(7.46,57.35){\makebox(0,0)[cc]{$=$}} \put(25.13,57.52){\makebox(0,0)[cc]{$=$}} \put(42.80,57.52){\makebox(0,0)[lc]{$=\langle123\rangle$}} \put(-6.37,49.02){\makebox(0,0)[cc]{$1$}} \put(-1.37,49.02){\makebox(0,0)[cc]{$2$}} \put(3.63,49.02){\makebox(0,0)[cc]{$3$}} \put(11.13,49.02){\makebox(0,0)[cc]{$2$}} \put(16.13,49.02){\makebox(0,0)[cc]{$3$}} \put(21.13,49.02){\makebox(0,0)[cc]{$1$}} \put(28.63,49.02){\makebox(0,0)[cc]{$3$}} \put(33.63,49.02){\makebox(0,0)[cc]{$1$}} \put(38.63,49.02){\makebox(0,0)[cc]{$2$}} \put(-6.37,33.53){\line(1,0){8.67}} \put(2.30,33.53){\line(0,-1){4.33}} \put(2.30,29.19){\line(-1,-1){4.00}} \put(2.30,29.19){\line(1,-1){4.00}} \put(-6.37,33.53){\line(0,-1){8.00}} \put(2.30,28.86){\makebox(0,0)[cc]{$\bullet$}} \put(27.96,33.53){\line(-1,0){8.67}} \put(19.29,33.53){\line(0,-1){4.33}} \put(19.29,29.19){\line(1,-1){4.00}} \put(19.29,29.19){\line(-1,-1){4.00}} \put(27.96,33.53){\line(0,-1){8.00}} \put(19.29,28.86){\makebox(0,0)[cc]{$\bullet$}} \put(11.96,33.53){\makebox(0,0)[cc]{$=-$}} \put(52.63,33.53){\line(1,0){8.67}} \put(61.30,33.53){\line(0,-1){4.33}} \put(61.30,29.19){\line(-1,-1){4.00}} \put(61.30,29.19){\line(1,-1){4.00}} \put(52.63,33.53){\line(0,-1){8.00}} \put(61.30,28.86){\makebox(0,0)[cc]{$\bullet$}} \put(86.96,33.53){\line(-1,0){8.67}} \put(78.29,33.53){\line(0,-1){4.33}} \put(78.29,29.19){\line(1,-1){4.00}} \put(78.29,29.19){\line(-1,-1){4.00}} \put(86.96,33.53){\line(0,-1){8.00}} \put(78.29,28.86){\makebox(0,0)[cc]{$\bullet$}} \put(70.96,33.53){\makebox(0,0)[cc]{$=-$}} \put(-6.37,12.19){\line(1,0){8.67}} \put(2.30,12.19){\line(0,-1){4.33}} \put(2.30,7.86){\line(-1,-1){4.00}} \put(2.30,7.86){\line(1,-1){4.00}} \put(-6.37,12.19){\line(0,-1){8.00}} \put(2.30,7.52){\makebox(0,0)[cc]{$\bullet$}} \put(27.96,12.19){\line(-1,0){8.67}} \put(19.29,12.19){\line(0,-1){4.33}} \put(19.29,7.86){\line(1,-1){4.00}} \put(19.29,7.86){\line(-1,-1){4.00}} \put(27.96,12.19){\line(0,-1){8.00}} \put(19.29,7.52){\makebox(0,0)[cc]{$\bullet$}} \put(11.96,12.19){\makebox(0,0)[cc]{$=-$}} \put(42.40,74.85){\line(-1,-2){2.32}} \put(42.40,74.85){\line(1,-2){2.32}} \put(61.67,74.85){\line(-2,-5){1.91}} \put(61.67,74.70){\line(3,-5){2.78}} \put(64.45,70.07){\line(0,0){0.06}} \put(-1.37,94.33){\line(0,-1){5.33}} \put(42.50,74.73){\line(0,1){3.73}} \put(61.70,74.73){\line(0,1){3.73}} \put(-1.37,56.73){\line(0,-1){5.33}} \put(16.10,56.73){\line(0,-1){5.33}} \put(33.56,56.73){\line(0,-1){5.33}} \put(-1.90,33.53){\makebox(0,0)[cc]{$\bullet$}} \put(23.63,33.53){\makebox(0,0)[cc]{$\bullet$}} \put(57.03,33.53){\makebox(0,0)[cc]{$\bullet$}} \put(82.69,33.53){\makebox(0,0)[cc]{$\bullet$}} \put(-1.90,12.20){\makebox(0,0)[cc]{$\bullet$}} \put(23.63,12.20){\makebox(0,0)[cc]{$\bullet$}} \put(52.66,22.95){\makebox(0,0)[cc]{2}} \put(57.19,22.95){\makebox(0,0)[cc]{3}} \put(65.21,22.95){\makebox(0,0)[cc]{1}} \put(74.34,22.95){\makebox(0,0)[cc]{3}} \put(82.36,22.95){\makebox(0,0)[cc]{1}} \put(86.89,22.95){\makebox(0,0)[cc]{2}} \put(91.08,33.58){\makebox(0,0)[lc]{$= \langle23\rangle$}} \put(-6.44,23.12){\makebox(0,0)[cc]{1}} \put(-1.73,23.12){\makebox(0,0)[cc]{2}} \put(6.29,23.12){\makebox(0,0)[cc]{3}} \put(15.24,23.12){\makebox(0,0)[cc]{2}} \put(23.26,23.12){\makebox(0,0)[cc]{3}} \put(27.97,23.12){\makebox(0,0)[cc]{1}} \put(-6.44,2.03){\makebox(0,0)[cc]{3}} \put(-1.73,2.03){\makebox(0,0)[cc]{1}} \put(6.29,2.03){\makebox(0,0)[cc]{2}} \put(15.24,2.03){\makebox(0,0)[cc]{1}} \put(23.44,2.03){\makebox(0,0)[cc]{2}} \put(27.97,2.03){\makebox(0,0)[cc]{3}} \put(30.93,33.76){\makebox(0,0)[lc]{$=\langle12\rangle$,}} \put(30.93,12.14){\makebox(0,0)[lc]{$=-\langle13\rangle$}} \put(6.50,94.33){\makebox(0,0)[cc]{$= \langle1\rangle$}} \end{picture} \end{center} \caption{ A representation of oriented faces of $\Delta_1$, $\Delta_2$ and $\Delta_3$ by equivalence classes of labeled planar trees, representing elements of $H(1)$, $H(2)$ and $H(3)$. \label{peniz}} \end{figure} A ${\overline {\cal D}}$-trace $t:{\overline {\cal D}} \to {\cal E}_{A,V}$ is, under the identification of Theorem~\ref{resiz}, given by the values $T_n := t(\uparrow\!^{n-1}\xi_n)$, $n\geq 1$. The axiom~(\ref{Blatter}) then reflects~(\ref{hvezdicka}). The relation between the axiom~(\ref{Blatter}) and the geometry of the simplex is visualized on Figure~\ref{delta}. \begin{figure}[hbtp] \begin{center} \unitlength 1.00mm \linethickness{0.4pt} \begin{picture}(74.16,52.17) \put(-40.33,49.67){\makebox(0,0)[lc]{$\delta T_1 (a) = 0$:}} \put(-13.00,49.67){\makebox(0,0)[rc]{$\delta($}} \put(-10.50,49.67){\makebox(0,0)[cc]{$\bullet$}} \put(-10.50,52.17){\makebox(0,0)[cc]{$a$}} \put(-8.00,49.67){\makebox(0,0)[lc]{$)=0$}} \put(-40.33,39.67){\makebox(0,0)[lc] {$\delta T_2 (a,b) = T_1(a\cdot b)-\znamenko{|a|\cdot |b|}T_1(b\cdot a):$}} \put(37.34,39.67){\makebox(0,0)[lc]{$\delta($}} \put(43.00,39.67){\makebox(0,0)[cc]{$\bullet$}} \put(53.00,39.67){\makebox(0,0)[cc]{$\bullet$}} \put(55.50,39.67){\makebox(0,0)[lc]{$)=$}} \put(64.66,39.67){\makebox(0,0)[cc]{$\bullet$}} \put(64.66,43.17){\makebox(0,0)[cc]{$ab$}} \put(69.16,39.84){\makebox(0,0)[cc]{$-$}} \put(74.16,39.67){\makebox(0,0)[cc]{$\bullet$}} \put(74.16,43.17){\makebox(0,0)[cc]{$ba$}} \put(43.00,39.67){\vector(1,0){10.00}} \put(-40.50,26.34){\makebox(0,0)[lc] {$\delta T_3 (a,b,c) = T_2(a\cdot b,c)+\znamenko{|a|\cdot(|b|+|c|)} T_2(b\cdot c,a)+\znamenko{|c|\cdot(|a|+|b|)} T_2(c\cdot a,b)$:}} \put(-40.50,8.67){\makebox(0,0)[lc]{$\delta\left(\rule{0mm}{10mm}\right.$}} \put(-22.65,14.50){\vector(-3,-4){8.13}} \put(-14.53,3.67){\vector(-3,4){8.13}} \put(-30.80,3.67){\vector(1,0){16.25}} \put(-7.17,8.84){\makebox(0,0)[cc]{$\left.\rule{0mm}{10mm}\right)=$}} \put(4.49,8.84){\vector(1,0){14.00}} \put(21.99,8.84){\makebox(0,0)[cc]{$+$}} \put(42.49,8.84){\makebox(0,0)[cc]{$+$}} \put(24.99,8.84){\vector(1,0){14.00}} \put(45.49,8.84){\vector(1,0){14.00}} \put(-26.45,10.67){\makebox(0,0)[rb]{$(ca)b$}} \put(-18.61,10.34){\makebox(0,0)[lb]{$(bc)a$}} \put(-22.78,2.84){\makebox(0,0)[ct]{$(ab)c$}} \put(11.50,11.50){\makebox(0,0)[cb]{$(ab)c$}} \put(32.00,11.67){\makebox(0,0)[cb]{$(bc)a$}} \put(52.00,11.67){\makebox(0,0)[cb]{$(ca)b$}} \put(42.99,43.17){\makebox(0,0)[cc]{$ab$}} \put(53.16,43.00){\makebox(0,0)[cc]{$ba$}} \end{picture} \end{center} \caption{ Relation of the axiom~(9) to the geometry of the simplex. \label{delta} } \end{figure} \section{Quadratic operads and modules; modules associated to cyclic operads} \label{cervena-tuzka} Each operad ${\cal P}$ can be presented as a quotient ${\cal F}(E)/I$, for a collection $E$ and an `operadic' ideal $I$. The operad ${\cal P}$ is {\em quadratic\/}~\cite[page~228]{ginzburg-kapranov:DMJ94} if it has a presentation such that the collection $E$ is concentrated in degree $2$, $E = E(2)$, and the ideal $I$ is generated by a subspace $R \subset {\cal F}(E)(3)$. In this case we write ${\cal P} = \prez ER$. \begin{example}{\rm\ \label{genius} Quadratic operads are omnipresent. Just recall that ${\it Ass} = \prez ER$ for $E = E(2) = {\bf k}[\Sigma_2]$, the regular representation of $\Sigma_2$. Choosing a generator $\mu \in E$, we can write $E = {\rm Span}(\mu,\mu S_{21})$, where $S_{21}\in \Sigma_2$ is the transposition. Then $R\subset {\cal F}(E)(3)$ is the $\Sigma_3$-subspace generated by $\mu(1,\mu)- \mu(\mu,1)$. If we think of $\mu$ as corresponding to a multiplication, then the generator of $R$ expresses the associativity. Sometimes we simplify the notation and write ${\it Ass}= \prez \mu{\mu(1,\mu)- \mu(\mu,1)}$. }\end{example} Similarly, for each ${\cal P}$-module $M$ there exists a collection $X$ such that $M = (X\circ {\cal P})/J$ for some right submodule $J\subset X\circ {\cal P}$. The following definition was introduced independently also in~\cite{ginzburg-voronov}. \begin{definition} \label{celenka} The module $M$ is called quadratic, if, in the above presentation, the collection $X$ is concentrated in degree $1$, $X= X(1)$, and the right submodule $J$ is generated by a subspace $G \subset (X \circ {\cal P})(2)$. In this case we write $M = \prezmod X{{\cal P}}G$. \end{definition} \begin{example}{\rm\ \label{nabijecka} Let $X=X(1)$ be generated by one element $g$; $X(1)= {\rm Span}(g)$. Let $G \subset (X\circ {\it Ass})(2)$ be defined as $G = {\rm Span}(g(\mu)(1- S_{21}))$, where $\mu$ and $S_{21}\in \Sigma_2$ has the same meaning as in Example~\ref{genius}. Then it is not difficult to see that ${\it Cycl} = \prezmod X{{\it Ass}}G$ or, in a more explicit notation, \begin{equation} \label{Spitfire} {\it Cycl} = \prezmod g{{\it Ass}}{g(\mu)- g(\mu)S_{21}}. \end{equation} Now we can give the characterization of ${\it Cycl}$ traces promised in Example~\ref{why-traces}. Since the ${\it Ass}$-module ${\it Cycl}$ is generated by $g\in {\it Cycl}(1)$, any ${\it Cycl}$-trace $t: {\it Cycl} \to {\cal E}_{A,W}$ is determined by the image $T := t(g)\in {\cal E}_{A,V}(1)= \Hom AW$. The symmetry~(\ref{nabla}) then follows from the condition $t(g(\mu)(1- S_{21})) = 0$. }\end{example} In fact, we show that ${\it Cycl}$ is a very special case of a module over the {\em cyclic\/} operad ${\it Ass}$. Cyclic operads were introduced by E.~Getzler and M.M.~Kapranov~\cite[Section~2]{getzler-kapranov:cyclic}. The definition we are going to recall is based on the following convention. We interpret the symmetric group $\Sigma_{n+1}$ as the group of permutations of the set $\{\rada 0n\}$ and $\Sigma_n$ as the subgroup of $\Sigma_{n+1}$ consisting of permutations $\sigma \in \Sigma_{n+1}$ with $\sigma(0)= 0$. If $\tau_n \in \Sigma_{n+1}$ is the cycle $(\rada 0n)$, then $\tau_n$ and $\Sigma_n$ generate $\Sigma_{n+1}$. \begin{definition} \label{tabal} Cyclic operad is an ordinary operad ${\cal P}$ such that the right $\Sigma_n$-action on ${\cal P}(n)$ extends, for $n\geq 1$, to an action of $\Sigma_{n+1}$ such that \begin{itemize} \item[(i)] $\tau_1(1)= 1$, where $1\in {\cal P}(1)$ is the unit and \item[(ii)] for $p\in {\cal P}(m)$ and $q\in {\cal P}(n)$, \[ \gamma(p;q,\rada 11)\cdot \tau_{m+n-1}= \znamenko{|p|\cdot|q|} \gamma(q\cdot\tau_n;\rada11,p\cdot\tau_m). \] \end{itemize} \end{definition} An intuitive feeling for the action of $\tau_n$ is suggested by Figure~\ref{feel}. \begin{figure}[hbtp] \begin{center} \unitlength 1.00mm \thinlines \begin{picture}(60.84,24.84) \put(49.67,16.00){\line(-1,-1){10.17}} \put(39.50,5.83){\line(1,0){20.00}} \put(59.50,5.83){\line(-1,1){10.00}} \put(49.50,15.83){\line(0,1){7.33}} \put(41.33,5.83){\line(0,-1){4.83}} \put(44.83,5.83){\line(0,-1){4.83}} \put(48.00,5.83){\line(0,-1){4.83}} \put(57.33,5.83){\line(0,-1){4.83}} \put(53.00,3.33){\makebox(0,0)[cc]{$\cdots$}} \put(9.83,16.00){\line(-1,-1){10.17}} \put(-0.33,5.83){\line(1,0){20.00}} \put(19.67,5.83){\line(-1,1){10.00}} \put(9.67,15.83){\line(0,1){7.33}} \put(1.50,5.83){\line(0,-1){4.83}} \put(5.00,5.83){\line(0,-1){4.83}} \put(8.17,5.83){\line(0,-1){4.83}} \put(17.50,5.83){\line(0,-1){4.83}} \put(13.17,3.33){\makebox(0,0)[cc]{$\cdots$}} \put(39.08,0.91){\oval(4.50,3.17)[b]} \put(36.83,1.00){\line(0,1){22.33}} \put(55.17,23.09){\oval(11.33,3.50)[t]} \put(60.83,23.67){\line(0,-1){23.00}} \put(9.67,10.33){\makebox(0,0)[cc]{$p$}} \put(49.50,10.33){\makebox(0,0)[cc]{$p\tau_n$}} \put(27.00,10.33){\makebox(0,0)[cc]{$\longmapsto$}} \put(1.50,-2.67){\makebox(0,0)[cc]{$1$}} \put(5.00,-2.67){\makebox(0,0)[cc]{$2$}} \put(8.17,-2.67){\makebox(0,0)[cc]{$3$}} \put(17.50,-2.67){\makebox(0,0)[cc]{$n$}} \put(44.83,-2.67){\makebox(0,0)[cc]{$1$}} \put(48.00,-2.67){\makebox(0,0)[cc]{$2$}} \put(60.83,-2.67){\makebox(0,0)[cc]{$n$}} \end{picture} \end{center} \caption{A `visualization' of the action of $\tau_n$. The element $\tau_n$ turns $p\in {\cal P}(n)$, represented as a `thing' with $n$ inputs and one output, a bit so that the first input becomes the output and the output becomes the last input of $p\cdot \tau_n$.\label{feel}} \end{figure} Let ${\cal P} = \prez ER$ be a quadratic operad. Thus $E(2)$ is a $\Sigma_2= {\bf Z}_2$ space and the homomorphism ${\rm sgn}: \Sigma_3\to {\bf Z}_2$ equips $E(2)$ with a $\Sigma_3$-action which induces on ${\cal F}(E)$ a cyclic operad structure. We say, according to~\cite[\S3.2]{getzler-kapranov:cyclic}, that ${\cal P}$ is {\em cyclic quadratic\/}, if the subspace $R \subset {\cal F}(E)(3)$ is $\Sigma_4$-invariant. In this case the operad ${\cal P}$ carries a natural cyclic structure induced from the cyclic structure of ${\cal F}(E)$. An example of a cyclic quadratic operad is the operad ${\it Ass}$, see~\cite[Proposition~2.4]{getzler-kapranov:cyclic} for a very explicit description of the cyclic structure, and also for other examples of cyclic operads. Let ${\cal P} = ({\cal P},\gamma)$ be a cyclic operad in the sense of Definition~\ref{tabal}. Define the ${\cal P}$-module $M_{{\cal P}}$ as follows. As a collection, $M_{{\cal P}}(n+1)= {\cal P}(n)$, for $n\geq 0$. The structure maps are given by \begin{eqnarray} \label{eqv} \nu(x;p_0,\rada 11)&:=&\znamenko{|p_0|\cdot |x|} \gamma(p_0\cdot\tau_{m_0};\rada 11,x), \mbox{ and} \\ \nonumber \nu(x;1,\rada{p_1}{p_n}) &:=& \gamma(x;\rada{p_1}{p_n}), \end{eqnarray} for $x\in M_{{\cal P}}(n+1)={\cal P}(n)$, $p_i\in {\cal P}(m_i)$, $0\leq i\leq n$. These conditions, along with the module axioms (Definition~\ref{Amphora}), imply that \[ \nu(x;\rada{p_0}{p_n})= \znamenko{|p_0|\cdot |x|} \gamma(p_0\cdot\tau_{m_0};\rada11, \nu(x;1,\rada{p_1}{p_n})). \] The verification of module axioms of $M_{{\cal P}}$ is routine. \begin{definition} \label{Turmo} We call the module $M_{\cal P}$ the module associated to the cyclic operad ${\cal P}$. \end{definition} We will also consider {\em unital\/} operads which describe algebras {\em with unit\/}. Unital operad is an operad such that ${\cal P}(0)$ is nonempty, generated by an element $\vartheta$, encoding the unit ${\bf k} \to A$ in the corresponding algebra $A$. The element $\vartheta$ determines, for $n\geq 1$ and $1\leq i\leq n$, the `degeneracy' maps $s_i : {\cal P}(n)\to {\cal P}(n-1)$ by $s_i(p):= \gamma(p;\rada 11,\vartheta,\rada 11)$ ($\vartheta$ at the $i$-th position). These maps satisfy certain commutation relations~\cite[page~278, Proposition~3]{stasheff:TAMS63} which follow from the axioms of an operad. For us, the most important is the relation $s_1(p)= s_2(p S_{21})$ for $p\in {\cal P}(2)$, which follows from the equivariance of structure maps. This, together with a natural requirement that $s_1 = s_2$ on ${\cal P}(2)$, gives that the maps $s_1=s_2$ are $\Sigma_2$-equivariant on ${\cal P}(2)$. We denote both maps $s_1$ and $s_2$ (which are the same) by $s$. \begin{example} \label{vypocetni} {\rm\ There is the operad ${\it UAss}$ for associative algebras with unit, which is the same as the operad ${\it Ass}$ except that ${\it UAss}(0)= {\rm Span}(\vartheta)$. The map $s :{\it Ass}(2) = {\bf k}[\Sigma_2]\to {\bf k}$ is the standard augmentation of the group ring ${\bf k}[\Sigma_2]$. Another example is the operad ${\it UComm}$ for unital commutative algebras. It has, for $n\geq 1$, ${\it UComm}(n)= {\it Comm}(n)= \mbox{$1\!\!1$}$ (the trivial one-dimensional representation), and ${\it UComm}(0)= {\rm Span}(\vartheta)$. The map $s : {\it UComm}(2)= {\bf k} \to {\bf k}$ is the identity. A more complicated example is the operad ${\it UPoiss}$ for Poisson algebras with unit. Here by a Poisson algebra with unit we mean an ordinary Poisson algebra~\cite[Example~3.3]{markl:dl} $P = (P, \cdot, [-,-])$ with a distinguished element $1\in P$ which is a two-sided unit for the commutative multiplication $\cdot$, while $[x,1]= [1,x]= 0$, for all $x\in P$. The operad ${\it UPoiss}$ coincides with the operad ${\it Poiss}$ for Poisson algebras (which is very explicitly described in~\cite{fox-markl:ContM97}), except that ${\it UPoiss}(0) ={\rm Span}(\vartheta)$. The component ${\it UPoiss}(2)$ is the direct sum $\mbox{$1\!\!1$} \oplus {\rm sgn}$, with the trivial one-dimensional representation $\mbox{$1\!\!1$}$ concentrated in degree zero, and the signum representation ${\rm sgn}$ in degree 1. The map $s : {\it UPoiss}(2)\to {\bf k} = {\it UPoiss}(1)$ is the projection on the zero-dimensional component. }\end{example} We saw that a natural way to construct cyclic operads was to take a quadratic operad ${\cal P} = \prez ER$ for which the relations $R$ were invariant under the natural $\Sigma_4$-action; the operad ${\cal P}$ had then a natural cyclic structure. There is a similar approach to unital operads. So, let ${\cal P} = \prez ER$ be a quadratic operad and suppose we are given an epimorphism $s :E \to {\bf k}$. This will be a model for the two degeneracy maps $s_1 = s_2 : {\cal P}(2)= E \to {\cal P}(1)= {\bf k}$. These two maps induce degeneracy maps on the free operad ${\cal F}(E)$ satisfying the correct commutation relations. The following definition expresses the conditions assuring that this structure preserves the relations $R$. \begin{definition} \label{virgo} Let ${\cal P} = \prez ER$ be a quadratic (resp.~cyclic quadratic) operad. Suppose that we are given an epimorphism $s : E \to {\bf k}$ such that \begin{itemize} \item[(i)] $s$ is invariant under the $\Sigma_2$ (resp.~$\Sigma_3$, in the cyclic case) action, and \item[(ii)] the induced maps $s_1,s_2,s_3: {\cal F}(E)(3)\to {\cal F}(E)(2)$ send the subspace $R\subset {\cal F}(E)(3)$ to zero. \end{itemize} Then the collection ${\it U}{\cal P}$, defined by ${\it U}{\cal P}(n):= {\cal P}(n)$ for $n\geq 1$ and ${\it U}{\cal P}(0)= {\bf k}$, has a natural structure of a unital operad. We call operads of this form quadratic unital (resp.~cyclic quadratic unital) operads. \end{definition} All the three operads ${\it UAss}$, ${\it UComm}$ and ${\it UPoiss}$ from Example~\ref{vypocetni} are cyclic quadratic unital operads in the sense of Definition~\ref{virgo}. \begin{proposition} \label{zabacek} Let ${\cal P} = \prez ER$ be a cyclic unital quadratic operad in the sense of Definition~\ref{virgo}. Then the associated module $M_{{\it U}{\cal P}}$ is quadratic, \[ M_{{\it U}{\cal P}} = \prezmod{{\rm Span}(\vartheta)}{{\cal P}}{{\rm Ker}(s):E \to {\bf k}}. \] \end{proposition} \noindent {\bf Proof.} Let $X = X(1):= {\rm Span}(g)$ and consider the map $\psi : X\circ {\cal P} \to M_{{\it U}{\cal P}}$ defined by $p(g) := \vartheta \in M_{{\it U}{\cal P}}(1)= {\it U}{\cal P}(0)$. Because clearly $X\circ {\cal P} = {\cal P}$, $\psi$ is, by~(\ref{eqv}), given as \[ (X\circ {\cal P})(n)= {\cal P}(n)\ni p \longmapsto \nu(\vartheta;p) = \gamma(p\cdot\tau_n; \rada 11,\vartheta)\in M_{{\it U}{\cal P}}(n) = {\it U}{\cal P}(n-1). \] Thus $\psi$ is a sequence $\psi(n):{\cal P}(n)\mapsto {\it U}{\cal P}(n-1)$ given by $\psi(n)(p):= \gamma(p\cdot\tau_n; \rada 11,\vartheta)$. These maps are epimorphisms, because $E$ generates ${\cal P}$ and $s:E \to {\bf k}$ (= the composition with $\vartheta$) is epi, by assumption. On the other hand, $\psi(2):{\cal P}(2)=E \to {\it U}{\cal P}(1)={\bf k}$ is exactly the map $s$, thus the submodule of $X\circ {\cal P}$ generated by ${\rm Ker}(s)$ is certainly contained in the kernel of $\psi$. A moment's reflections shows that the whole kernel of $\psi$ is generated by ${\rm Ker}(s)$.\hspace*{\fill \begin{example}{\rm\ \label{kacirek} If ${\cal P} = {\it UAss}$ is the operad for unital associative algebras from Example~\ref{vypocetni}, then ${\rm Ker}(s:E= {\bf k}[\Sigma_2] \to {\bf k})= {\rm sgn}$. So, by Proposition~\ref{zabacek}, $M_{{\it UAss}}= \prezmod{\vartheta}{{\it Ass}}{{\rm sgn}}$, thus $M_{{\it UAss}} = {\it Cycl}$, by~(\ref{Spitfire}). For the operad ${\it UComm}$ the kernel of $s$ is trivial, hence \[ M_{{\it UComm}} = {\rm Span}(\vartheta)\circ {\it Comm} \cong {\it Comm}. \] In other words, $M_{{\it UComm}}$ is the operad ${\it Comm}$ considered as a right module over itself. We recommend to the reader to make the similar discussion for the operad ${\it UPoiss}$. }\end{example} \section{Cyclohedron as the cobar construction} \label{penezenka} For a graded vector space $V$, let $\dual V$ denote the graded dual of $V$, i.e.\ the graded vector space $\bigoplus_p(\dual V)_p$ with $(\dual V)_p ={\rm Hom}^{-p}(V,{{\bf k}})={\rm Hom}(V_{p},{{\bf k}})$, the space of linear maps from $V_{p}$ to ${\bf k}$. If $V$ is a right $\Sigma_n$-module, then we equip $\dual V$ with the transposed $\Sigma_n$-action. We believe that there is no real risk of confusion of the ${}^*$ indicating the dual with the star indicating the degree. Recall also~\cite[\S3.5.1]{ginzburg-kapranov:DMJ94} that a {\em cooperad\/} is a collection ${\cal Q}=\coll {{\cal Q}}$ together with a system of maps \[ \omega = \omega_{\rada{m_1}{m_l}}: {\cal Q}(m_1+\cdots+m_l) \to {\cal Q}(l)\otimes {\cal Q}(m_1)\otimes \cdots \otimes {\cal Q}(m_l), \] which satisfy the axioms which are exactly the duals of the axioms for an operad. A typical example of a cooperad is the dual $\dual{\cal P}$ of an operad ${\cal P}$, i.e.~the collection $\dual{\cal P} = \{\dualI{{\cal P}(n)}\}_{n\geq 1}$ with the cooperad structure defined by the dualization of the structure maps of ${\cal P}$; here an obvious finite type assumption is necessary, but it will be always satisfied in the paper and we will make no explicit comments about it. Observe that if ${\cal P}$ is an operad, then both the suspension ${\bf s\hskip0mm} {\cal P}$ and the desuspension ${\bf s\hskip0mm}^{-1} {\cal P}$ introduced in Section~\ref{hrnicek1}, with the sign convention of~(\ref{sign-factor}), have a natural operad structure induced from the operad structure on ${\cal P}$. The structure maps of a cooperad ${\cal Q}$ determine (and are determined by) a map ${\overline \nu}: {\cal Q}\to {\cal F}({\cal Q})$ of collections. Composing this map with the (de)suspensions gives a degree -1 map $\downarrow\! {\cal Q} \to {\cal F}(\downarrow\! {\cal Q})$ which uniquely (because of the freeness of ${\cal F}(\downarrow\! {\cal Q})$) extends to a degree -1 derivation ${\partial_{\Cob}}$ of the operad ${\cal F}(\downarrow\! {\cal Q})$ which satisfies, as a consequence of the axioms of a cooperad, ${\partial_{\Cob}} \circ{\partial_{\Cob}} =0$. The differential graded operad \[ {\Omega}({\cal Q}) := ({\cal F}(\downarrow\! {\cal Q}),{\partial_{\Cob}}) \] is called the {\em cobar construction\/} on the cooperad ${\cal Q}$~\cite[\S3.2.12]{ginzburg-kapranov:DMJ94}. Let ${\cal P} = \prez ER$ be a quadratic operad. Take the dual $\dual E$ of $E$ and let $R^{\perp}$ be the annihilator of the space $R\subset {\cal F}(E)(3)$ in ${\cal F}(\dual E \otimes {\rm sgn})(3)$. The quadratic operad ${\cal P}^! := \prez {\dual E \otimes {\rm sgn}}{R^{\perp}}$ is, according to~\cite[\S2.1.9]{ginzburg-kapranov:DMJ94}, called the {\em Koszul\/} (or {\em quadratic\/}) {\em dual\/} of the operad ${\cal P}$. We always have a map $\downarrow\! \dualI{{\bf s\hskip0mm} {\cal P}} \to {\cal P}^!$ of collections defined as the composition \[ \downarrow\! \dualI{{\bf s\hskip0mm} {\cal P}}\stackrel{\rm proj.}{\longrightarrow} (\downarrow\! \dualI{{\bf s\hskip0mm} {\cal P})(2)} = \dual E\otimes {\rm sgn} = {\cal P}^!(2) \hookrightarrow {\cal P}^! \] which extends, by the freeness of ${\cal F}(\downarrow\! \dualI{{\bf s\hskip0mm} {\cal P}})$, to a differential graded operad map \begin{equation} \label{budicek} \pi : {\Omega}(\dualI{{\bf s\hskip0mm} {\cal P}}) = ({\cal F}(\downarrow\! \dualI{{\bf s\hskip0mm} {\cal P}}),{\partial_{\Cob}}) \to ({\cal P}^!,\partial=0). \end{equation} The quadratic operad ${\cal P}$ is called {\em Koszul\/}~\cite[Definition~4.1.3]{ginzburg-kapranov:DMJ94} if the map in~(\ref{budicek}) is a homology isomorphism. In the rest of the paper, the space of generators $E=E(2)$ of a quadratic operad will always be ungraded, concentrated in degree zero. Then the components of both the operad ${\cal P}$ and its dual ${\cal P}^!$ are concentrated in degree zero as well, and the Koszulness implies that the complex $({\Omega}(\dualI{{\bf s\hskip0mm} {\cal P}})(n), {\partial_{\Cob}})$ is acyclic in positive dimensions, for all $n$. On the other hand, as shown in~\cite[Theorem~4.1.13]{ginzburg-kapranov:DMJ94}, this acyclicity condition implies the Koszulness of ${\cal P}$. \begin{example}{\rm\ \label{yhr} The operad ${\it Ass}$ is well-known to be Koszul self-dual, ${\it Ass} = {\it Ass}^!$~\cite[Theorem~2.1.11]{ginzburg-kapranov:DMJ94}, and Koszul~\cite[Corrolary~4.2.7]{ginzburg-kapranov:DMJ94}. We have already remarked that the operad ${\overline {\cal A}}$ of cellular chains on the (symmetrized) associahedron ${\overline K}$ coincides, as a differential graded operad, with the cobar construction ${\Omega}(\dualI{{\bf s\hskip0mm}{\it Ass}})$. Let us give an explicit illustration of this statement. The $n$-th piece $(\downarrow\! \dualI{{\bf s\hskip0mm} {\it Ass}})(n)$ of the collection $(\downarrow\! \dualI{{\bf s\hskip0mm} {\it Ass}})$ is isomorphic to one copy of the regular representation ${\bf k}[\Sigma_n]$ concentrated in degree $(n-2)$. The isomorphism is not unique, but we may choose, for example, $\lambda : {\bf k}[\Sigma_n] \to \downarrow\! ({\bf s\hskip0mm} {\it Ass})^*(n)$ given by $\lambda(\sigma)(\uparrow\!^{n-2}\rho) := {\rm sgn}(\sigma)\cdot \delta_{\sigma,\rho}$, where $\sigma, \rho \in \Sigma_n$ and the meaning of the `Kronecker delta' $\delta_{\sigma,\rho}$ is clear. Formula~(\ref{fax}) then describes ${\cal F}(\downarrow\! \dualI{{\bf s\hskip0mm} {\it Ass}})(n)$ as the vector space spanned by the set of all {\em planar\/} (rooted, labeled) $n$-trees, having at least binary vertices. The identification of these trees with the bracketings of $\rada{\sigma(1)}{\sigma(n)}$, $\sigma \in \Sigma_n$, i.e.~with the elements of the set ${\overline {\cal B}}(n)$, is classical -- see~\cite[\S1.4]{boardman-vogt:73}; two examples are shown on Figure~\ref{hreben}. \begin{figure}[hbtp] \begin{center} \unitlength 0.70mm \thicklines \begin{picture}(174.84,50.34) \put(40.00,50.00){\line(0,-1){10.00}} \put(40.00,40.00){\line(-1,-1){30.00}} \put(40.00,39.83){\line(1,-1){30.33}} \put(30.00,30.00){\line(1,-1){20.17}} \put(29.83,30.00){\makebox(0,0)[cc]{$\bullet$}} \put(40.00,39.67){\makebox(0,0)[cc]{$\bullet$}} \put(16.74,16.91){\makebox(0,0)[cc]{$\bullet$}} \put(43.26,16.57){\makebox(0,0)[cc]{$\bullet$}} \put(10.00,5.00){\makebox(0,0)[cc]{$4$}} \put(23.67,5.00){\makebox(0,0)[cc]{$1$}} \put(36.33,4.83){\makebox(0,0)[cc]{$3$}} \put(50.17,5.00){\makebox(0,0)[cc]{$5$}} \put(70.33,5.00){\makebox(0,0)[cc]{$2$}} \put(7.67,31.33){\makebox(0,0)[rc]{$((41)(35))2:$}} \put(108.67,10.00){\line(1,1){30.17}} \put(138.84,40.17){\line(0,1){10.17}} \put(138.84,40.17){\line(-1,-3){7.89}} \put(131.00,16.67){\line(-6,-5){7.67}} \put(130.84,16.67){\line(6,-5){7.50}} \put(138.84,40.00){\line(2,-3){15.56}} \put(154.00,16.67){\line(-1,-1){6.67}} \put(154.17,17.17){\line(6,-5){7.83}} \put(138.84,40.00){\line(6,-5){36.00}} \put(138.84,39.83){\makebox(0,0)[cc]{$\bullet$}} \put(130.84,16.50){\makebox(0,0)[cc]{$\bullet$}} \put(154.17,17.00){\makebox(0,0)[cc]{$\bullet$}} \put(108.50,5.00){\makebox(0,0)[cc]{$1$}} \put(123.50,5.00){\makebox(0,0)[cc]{$6$}} \put(138.17,5.00){\makebox(0,0)[cc]{$2$}} \put(147.50,4.83){\makebox(0,0)[cc]{$3$}} \put(161.84,4.83){\makebox(0,0)[cc]{$4$}} \put(174.67,4.83){\makebox(0,0)[cc]{$5$}} \put(108.33,31.34){\makebox(0,0)[rc]{$1(62)(34)5:$}} \put(43.24,16.56){\line(-1,-1){6.97}} \put(16.74,16.74){\line(1,-1){6.97}} \end{picture} \end{center} \caption{Two examples of an identification of planar trees with elements of ${\cal B}$.\label{hreben}} \end{figure} The fact that the cellular differential coincides with ${\partial_{\Cob}}$ is a routine combinatorics. See~\cite[Example~4.1]{markl:zebrulka} for details. The identification above shows that the operad ${\it Ass}$ is Koszul. More precisely, we know that $H_*({\overline K})= H_0({\overline K})= {\it Ass}$, because ${\overline K}(n)$ is the union of convex polyhedra indexed by the elements of the symmetric group. On the other hand, due to the identification above, $H_*({\Omega}(\dualI{{\bf s\hskip0mm} {\cal P}}))= H_*({\overline K})$, thus the map $\pi$ of~(\ref{budicek}) is a homology isomorphism. }\end{example} A {\em comodule\/} over a cooperad ${\cal Q}$ is a collection $N = \coll N$ together with structure maps \[ \kappa = \kappa_{\rada{m_1}{m_l}}: N(m_1+\cdots+m_l) \to N(l)\otimes {\cal Q}(m_1)\otimes \cdots \otimes {\cal Q}(m_l), \] satisfying axioms dual to the axioms of a module over an operad. An example is the dual $\dual {\cal M}$ of a ${\cal P}$-module ${\cal M}$, which is a natural comodule over the cooperad $\dual{\cal P}$. Let $N$ be a ${\cal Q}$-comodule. As in the case of cooperads, the structure maps induce a degree -1 differential ${\partial_{\Cob}}$ on the free module $N\circ {\cal F}(\downarrow\! {\cal Q})$. The right differential graded ${\Omega}({\cal Q})$-module \[ {\Omega}(N;{\cal Q}) :=(N\circ {\cal F}(\downarrow\! {\cal Q}),{\partial_{\Cob}}) \] is called the (right) {\em cobar construction\/} on the ${\cal Q}$-comodule $N$. Suppose that $M = \prezmod X{{\cal P}}G$ is a quadratic ${\cal P}$-module, in the sense of Definition~\ref{celenka}, over a quadratic operad ${\cal P} = \prez ER$. Take the dual $\dual X$ and let $G^{\perp}$ be the annihilator of $G$ in $(\dual X\circ (\dual E \otimes {\rm sgn}))(2)$. Then the quadratic ${\cal P}^!$-module $M^! := \prezmod {\dual X}{{\cal P}^!}{G^{\perp}}$ is called the {\em Koszul\/} (or {\em quadratic\/}) {\em dual\/} of $M$. The above definitions were independently made in~\cite{ginzburg-voronov}. \begin{proposition} The ${\it Ass}$-module ${\it Cycl}$ is Koszul self-dual, ${\it Cycl}^! = {\it Cycl}$. \end{proposition} \noindent {\bf Proof.} Under the notation of Example~\ref{nabijecka}, the $\Sigma_2$-space $(X\circ {\it Ass})(2)$ is the direct sum $\mbox{$1\!\!1$} \oplus {\rm sgn}$ of the trivial and signum representations, while clearly $R= {\rm sgn}$, generated by $g(\mu) - g(\mu)S_{21}$. Then $R^\perp$ is easily seen to be ${\rm sgn}$ and the proposition follows.\hspace*{\fill Observe that, for a ${\cal P}$-module $M = \coll M$, the collection ${\bf s\hskip0mm} M$ has an induced ${\bf s\hskip0mm} {\cal P}$-module structure. For a quadratic ${\cal P}$-module $\prezmod X{{\cal P}}G$ over a quadratic operad ${\cal P} = \prez ER$ we have, as in~(\ref{budicek}), the map \begin{equation} \label{Opusem} \pi :{\Omega}(\dualI {{\bf s\hskip0mm} M},\dualI {{\bf s\hskip0mm} {\cal P}}) \to (M^!,\partial = 0), \end{equation} induced by the composition \[ \dualI {{\bf s\hskip0mm} M} \stackrel{\rm proj.}{\longrightarrow} \dualI{{\bf s\hskip0mm} M}(1) = \dual X = M^!(1) \hookrightarrow M^!. \] The following definition was independently made in~\cite{ginzburg-voronov}. \begin{definition} \label{telefon} A quadratic module $M$ over a quadratic operad ${\cal P}$ is called Koszul if the map $\pi$ in~(\ref{Opusem}) is a homology isomorphism. \end{definition} \begin{theorem} \label{Amphora1} The module ${\it Cycl}$ is Koszul. \end{theorem} The theorem will follow from Theorem~\ref{ucpavka} which explicitly identifies the cobar construction ${\Omega}(\dualI{{\bf s\hskip0mm}{\it Cycl}}, \dualI{{\bf s\hskip0mm}{\it Ass}})$ to the cellular chain complex ${\cal M}$ of the symmetrized cyclohedron ${\overline W}$. Thus the Koszulness of ${\it Cycl}$ is, as in the case of the operad ${\it Ass}$, a consequence of the fact that the cyclohedron is a convex polyhedron. At the end of the section we formulate a more general statement which also implies Theorem~\ref{Amphora1}. \begin{theorem} \label{ucpavka} The cobar construction ${\Omega}(\dualI{{\bf s\hskip0mm}{\it Cycl}}, \dualI{{\bf s\hskip0mm} {\it Ass}})$ is isomorphic to the cellular chain complex ${\cal M} = CC_*({\overline W})$ of the cyclohedron ${\overline W}$. \end{theorem} \noindent {\bf Proof.} As we observed in Example~\ref{yhr}, the collection $\downarrow\!\dualI{{\bf s\hskip0mm} {\it Ass}}(n)$ consists of one copy of the regular representation ${\bf k}[\Sigma_n]$ concentrated in degree $(n-2)$. Similarly, the collection $\dualI{{\bf s\hskip0mm} {\it Cycl}}$ is isomorphic to the collection ${\bf s\hskip0mm} {\it Cycl}$. We are going to give an explicit description of the space $(\dualI{{\bf s\hskip0mm}{\it Cycl}}\ \circ \downarrow\!\dualI{{\bf s\hskip0mm} {\it Ass}})(n)$, similar to that of Lemma~\ref{MGD}. Suppose that $T_i$ is, for each $1\leq i \leq l$, a planar $m_i$-tree, whose all vertices are at least binary. Let $R(\rada{T_1}{T_l})$ be the tree obtained by grafting the trees $\rada{T_1}{T_l}$ at the inputs of the `$l$-rake,' or, pictorially: \begin{center} \def\SetFigFont#1#2#3{{}} \def\smash#1{{#1}} \setlength{\unitlength}{0.0062500in}% \begin{picture}(370,213)(20,440) \thicklines \put(150,540){\line( 0, 1){63}} \put(150,603){\line( 1, 0){300}} \put(450,603){\line( 0,-1){63}} \put(150,540){\line(-1,-2){ 30}} \put(120,480){\line( 1, 0){ 60}} \put(180,480){\line(-1, 2){ 30}} \put(123,480){\line( 0,-1){ 21}} \put(138,480){\line( 0,-1){ 21}} \put(171,480){\line( 0,-1){ 21}} \put(240,603){\line( 0,-1){63}} \put(240,540){\line(-1,-2){ 30}} \put(210,480){\line( 1, 0){ 60}} \put(270,480){\line(-1, 2){ 30}} \put(450,540){\line(-1,-2){ 30}} \put(420,480){\line( 1, 0){ 60}} \put(480,480){\line(-1, 2){ 30}} \put(261,480){\line( 0,-1){ 21}} \put(471,480){\line( 0,-1){ 21}} \put(426,480){\line( 0,-1){ 21}} \put(441,480){\line( 0,-1){ 21}} \put(216,480){\line( 0,-1){ 21}} \put(231,480){\line( 0,-1){ 21}} \put(153,462){\makebox(0,0)[b]{ \smash{ \SetFigFont{12}{14.4}{rm}$\cdot\!\!\cdot\!\!\cdot$ } }} \put(330,568){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$\cdots$}}} \put(246,462){\makebox(0,0)[b]% {\smash{\SetFigFont{12}{14.4}{rm}$\cdot\!\!\cdot\!\!\cdot$}}} \put(456,462){\makebox(0,0)[b]% {\smash{\SetFigFont{12}{14.4}{rm}$\cdot\!\!\cdot\!\!\cdot$}}} \put(300,597){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$\bullet$}}} \put(150,537){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$\bullet$}}} \put(154,492){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$T_1$}}} \put(240,537){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$\bullet$}}} \put(244,492){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$T_2$}}} \put(450,537){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$\bullet$}}} \put(454,492){\makebox(0,0)[b]{\smash{\SetFigFont{12}{14.4}{rm}$T_l$}}} \put(10,540 ){\makebox(0,0)[b]% {\smash{\SetFigFont{12}{14.4}{rm}$R(\rada{T_1}{T_l}) =$}}} \end{picture} \end{center} Let $J(n)$ be the free graded right $\Sigma_n$-module generated by the symbols $R(\rada{T_1}{T_l})$, $m_1+\cdots+ m_l = n$, with the degree defined by \[ \deg(R(\rada{T_1}{T_l}))= n-1-\sum_{i=1}^l \#{\rm vert}(T_i), \] where $ \#{\rm vert}(T_i)$ is the number of vertices of the tree $T_i$. For $\zeta \in {\bf Z}_l$ we put (see~(\ref{pejska_Mikinka}) for the notation) \begin{equation} \label{monce1} \zeta(R(\rada{T_1}{T_l}) \times \sigma) :={\rm sgn}(\zeta) \cdot R(\rada{T_{\eta^{-1}(1)}}{T_{\eta^{-1}(l)}}) \times \zeta(\rada{m_1}{m_l}) \sigma. \end{equation} Then the graded vector space $(\dualI{{\bf s\hskip0mm}{\it Cycl}}\ \circ \downarrow\!\dualI{{\bf s\hskip0mm} {\it Ass}})(n)$ is spanned by equivalence classes of elements \[ R(\rada{T_1}{T_l}) \times \sigma \in J(n), \] modulo the left action of the group ${\bf Z}_l$ introduced in~(\ref{monce1}). The right action of the group $\Sigma_n$ is given by \[ [R(\rada{T_1}{T_l}) \times \sigma]\cdot \rho := [R(\rada{T_1}{T_l}) \times \sigma\rho]. \] We may symbolize the element $R(\rada{T_1}{T_l}) \times \sigma \in J(n)$ as the tree $R(\rada{T_1}{T_l})$ with the inputs labeled by $\rada{\sigma^{-1}(1)}{\sigma^{-1}(l)}$. There is an almost obvious one-to-one correspondence between these labeled planar $n$-trees and cyclic bracketings of $n$ indeterminates from ${\overline {{\cal B}C}}(n)$. This becomes absolutely clear after looking at Figure~\ref{master}. \begin{figure}[hbtp] \begin{center} \unitlength 1.20mm \thicklines \begin{picture}(91.08,167.33) \put(-11.20,167.32){\makebox(0,0)[rc]{\fbox{$n=1$:}}} \put(-1.37,167.15){\makebox(0,0)[cc]{$\bullet$}} \put(-1.37,159.65){\makebox(0,0)[cc]{$1$}} \put(-11.20,150.48){\makebox(0,0)[rc]{\fbox{$n=2$:}}} \put(12.30,150.48){\line(1,0){5.00}} \put(17.30,150.48){\line(0,-1){5.00}} \put(12.13,150.48){\line(0,-1){5.00}} \put(14.72,150.48){\makebox(0,0)[cc]{$\bullet$}} \put(-3.87,150.48){\line(1,0){5.00}} \put(1.13,150.48){\line(0,-1){5.00}} \put(-4.04,150.48){\line(0,-1){5.00}} \put(-1.45,150.48){\makebox(0,0)[cc]{$\bullet$}} \put(-4.04,142.65){\makebox(0,0)[cc]{$1$}} \put(1.13,142.65){\makebox(0,0)[cc]{$2$}} \put(12.13,142.65){\makebox(0,0)[cc]{$2$}} \put(17.30,142.65){\makebox(0,0)[cc]{$1$}} \put(7.97,150.48){\makebox(0,0)[cc]{$=-$}} \put(20.79,150.48){\makebox(0,0)[lc]{$=12,$}} \put(42.46,151.32){\makebox(0,0)[cc]{$\bullet$}} \put(42.46,147.65){\makebox(0,0)[cc]{$\bullet$}} \put(45.84,149.48){\makebox(0,0)[lc]{$= (12),$}} \put(61.76,151.32){\makebox(0,0)[cc]{$\bullet$}} \put(61.63,147.65){\makebox(0,0)[cc]{$\bullet$}} \put(65.01,149.48){\makebox(0,0)[lc]{$= (21)$}} \put(40.13,141.30){\makebox(0,0)[cc]{$1$}} \put(44.96,141.30){\makebox(0,0)[cc]{$2$}} \put(59.30,141.30){\makebox(0,0)[cc]{$2$}} \put(64.30,141.30){\makebox(0,0)[cc]{$1$}} \put(-11.20,129.52){\makebox(0,0)[rc]{\fbox{$n=3$:}}} \put(-6.37,129.69){\line(1,0){10.00}} \put(3.63,129.69){\line(0,-1){5.17}} \put(-6.37,129.69){\line(0,-1){5.17}} \put(-6.37,124.52){\line(0,0){0.00}} \put(-1.37,129.69){\makebox(0,0)[cc]{$\bullet$}} \put(11.13,129.69){\line(1,0){10.00}} \put(28.63,129.69){\line(1,0){10.00}} \put(21.13,129.69){\line(0,-1){5.17}} \put(38.63,129.69){\line(0,-1){5.17}} \put(11.13,129.69){\line(0,-1){5.17}} \put(28.63,129.69){\line(0,-1){5.17}} \put(11.13,124.52){\line(0,0){0.00}} \put(28.63,124.52){\line(0,0){0.00}} \put(16.13,129.69){\makebox(0,0)[cc]{$\bullet$}} \put(33.63,129.69){\makebox(0,0)[cc]{$\bullet$}} \put(7.46,130.35){\makebox(0,0)[cc]{$=$}} \put(25.13,130.52){\makebox(0,0)[cc]{$=$}} \put(42.80,130.52){\makebox(0,0)[cc]{$=$}} \put(48.30,130.52){\makebox(0,0)[cc]{$123$}} \put(-6.37,122.02){\makebox(0,0)[cc]{$1$}} \put(-1.37,122.02){\makebox(0,0)[cc]{$2$}} \put(3.63,122.02){\makebox(0,0)[cc]{$3$}} \put(11.13,122.02){\makebox(0,0)[cc]{$2$}} \put(16.13,122.02){\makebox(0,0)[cc]{$3$}} \put(21.13,122.02){\makebox(0,0)[cc]{$1$}} \put(28.63,122.02){\makebox(0,0)[cc]{$3$}} \put(33.63,122.02){\makebox(0,0)[cc]{$1$}} \put(38.63,122.02){\makebox(0,0)[cc]{$2$}} \put(-1.37,61.85){\line(-1,-1){6.17}} \put(-1.37,61.85){\line(1,-1){6.00}} \put(-1.37,61.69){\makebox(0,0)[cc]{$\bullet$}} \put(-1.37,65.19){\makebox(0,0)[cc]{$\bullet$}} \put(25.96,61.85){\line(-1,-1){6.17}} \put(53.30,61.85){\line(-1,-1){6.17}} \put(25.96,61.85){\line(1,-1){6.00}} \put(53.30,61.85){\line(1,-1){6.00}} \put(25.96,61.69){\makebox(0,0)[cc]{$\bullet$}} \put(53.30,61.69){\makebox(0,0)[cc]{$\bullet$}} \put(25.96,65.19){\makebox(0,0)[cc]{$\bullet$}} \put(53.30,65.19){\makebox(0,0)[cc]{$\bullet$}} \put(-7.54,52.02){\makebox(0,0)[cc]{$1$}} \put(-1.37,52.02){\makebox(0,0)[cc]{$2$}} \put(4.63,52.02){\makebox(0,0)[cc]{$3$}} \put(8.96,62.19){\makebox(0,0)[cc]{$=(123),$}} \put(19.80,52.19){\makebox(0,0)[cc]{$3$}} \put(25.96,52.19){\makebox(0,0)[cc]{$1$}} \put(31.96,52.19){\makebox(0,0)[cc]{$2$}} \put(36.96,62.35){\makebox(0,0)[cc]{$=(312)$,}} \put(47.13,52.19){\makebox(0,0)[cc]{$1$}} \put(53.30,52.19){\makebox(0,0)[cc]{$2$}} \put(59.30,52.19){\makebox(0,0)[cc]{$3$}} \put(64.13,62.19){\makebox(0,0)[cc]{$=(123)$}} \put(-1.61,36.99){\line(-1,-1){6.31}} \put(-1.61,41.25){\makebox(0,0)[cc]{$\bullet$}} \put(-1.61,36.61){\makebox(0,0)[cc]{$\bullet$}} \put(26.22,36.99){\line(-1,-1){6.31}} \put(54.04,36.99){\line(-1,-1){6.31}} \put(26.22,41.25){\makebox(0,0)[cc]{$\bullet$}} \put(54.04,41.25){\makebox(0,0)[cc]{$\bullet$}} \put(26.22,36.61){\makebox(0,0)[cc]{$\bullet$}} \put(54.04,36.61){\makebox(0,0)[cc]{$\bullet$}} \put(-8.10,27.90){\makebox(0,0)[cc]{$1$}} \put(5.81,27.90){\makebox(0,0)[cc]{$3$}} \put(-6.37,106.53){\line(1,0){8.67}} \put(2.30,106.53){\line(0,-1){4.33}} \put(2.30,102.19){\line(-1,-1){4.00}} \put(2.30,102.19){\line(1,-1){4.00}} \put(-6.37,106.53){\line(0,-1){8.00}} \put(2.30,101.86){\makebox(0,0)[cc]{$\bullet$}} \put(27.96,106.53){\line(-1,0){8.67}} \put(19.29,106.53){\line(0,-1){4.33}} \put(19.29,102.19){\line(1,-1){4.00}} \put(19.29,102.19){\line(-1,-1){4.00}} \put(27.96,106.53){\line(0,-1){8.00}} \put(19.29,101.86){\makebox(0,0)[cc]{$\bullet$}} \put(11.96,106.53){\makebox(0,0)[cc]{$=-$}} \put(52.63,106.53){\line(1,0){8.67}} \put(61.30,106.53){\line(0,-1){4.33}} \put(61.30,102.19){\line(-1,-1){4.00}} \put(61.30,102.19){\line(1,-1){4.00}} \put(52.63,106.53){\line(0,-1){8.00}} \put(61.30,101.86){\makebox(0,0)[cc]{$\bullet$}} \put(86.96,106.53){\line(-1,0){8.67}} \put(78.29,106.53){\line(0,-1){4.33}} \put(78.29,102.19){\line(1,-1){4.00}} \put(78.29,102.19){\line(-1,-1){4.00}} \put(86.96,106.53){\line(0,-1){8.00}} \put(78.29,101.86){\makebox(0,0)[cc]{$\bullet$}} \put(70.96,106.53){\makebox(0,0)[cc]{$=-$}} \put(-6.37,85.19){\line(1,0){8.67}} \put(2.30,85.19){\line(0,-1){4.33}} \put(2.30,80.86){\line(-1,-1){4.00}} \put(2.30,80.86){\line(1,-1){4.00}} \put(-6.37,85.19){\line(0,-1){8.00}} \put(2.30,80.52){\makebox(0,0)[cc]{$\bullet$}} \put(27.96,85.19){\line(-1,0){8.67}} \put(19.29,85.19){\line(0,-1){4.33}} \put(19.29,80.86){\line(1,-1){4.00}} \put(19.29,80.86){\line(-1,-1){4.00}} \put(27.96,85.19){\line(0,-1){8.00}} \put(19.29,80.52){\makebox(0,0)[cc]{$\bullet$}} \put(11.96,85.19){\makebox(0,0)[cc]{$=-$}} \put(42.40,147.85){\line(-1,-2){2.32}} \put(42.40,147.85){\line(1,-2){2.32}} \put(61.67,147.85){\line(-2,-5){1.91}} \put(61.67,147.70){\line(3,-5){2.78}} \put(64.45,143.07){\line(0,0){0.06}} \put(-1.59,36.88){\line(1,-1){6.50}} \put(26.24,36.88){\line(1,-1){6.50}} \put(54.08,36.88){\line(1,-1){6.50}} \put(0.58,34.55){\line(-1,-1){3.67}} \put(0.58,34.55){\makebox(0,0)[cc]{$\bullet$}} \put(28.41,34.55){\line(-1,-1){3.67}} \put(56.24,34.55){\line(-1,-1){3.67}} \put(28.41,34.55){\makebox(0,0)[cc]{$\bullet$}} \put(56.24,34.55){\makebox(0,0)[cc]{$\bullet$}} \put(-1.46,17.82){\line(1,-1){6.31}} \put(-1.46,22.08){\makebox(0,0)[cc]{$\bullet$}} \put(-1.46,17.44){\makebox(0,0)[cc]{$\bullet$}} \put(-1.48,17.71){\line(-1,-1){6.50}} \put(-3.65,15.38){\line(1,-1){3.67}} \put(-3.65,15.38){\makebox(0,0)[cc]{$\bullet$}} \put(26.37,17.82){\line(1,-1){6.31}} \put(54.20,17.82){\line(1,-1){6.31}} \put(26.37,22.08){\makebox(0,0)[cc]{$\bullet$}} \put(54.20,22.08){\makebox(0,0)[cc]{$\bullet$}} \put(26.37,17.44){\makebox(0,0)[cc]{$\bullet$}} \put(54.20,17.44){\makebox(0,0)[cc]{$\bullet$}} \put(26.35,17.71){\line(-1,-1){6.50}} \put(54.18,17.71){\line(-1,-1){6.50}} \put(24.18,15.38){\line(1,-1){3.67}} \put(52.01,15.38){\line(1,-1){3.67}} \put(24.18,15.38){\makebox(0,0)[cc]{$\bullet$}} \put(52.01,15.38){\makebox(0,0)[cc]{$\bullet$}} \put(-1.37,167.33){\line(0,-1){5.33}} \put(42.50,147.73){\line(0,1){3.73}} \put(61.70,147.73){\line(0,1){3.73}} \put(-1.37,129.73){\line(0,-1){5.33}} \put(16.10,129.73){\line(0,-1){5.33}} \put(33.56,129.73){\line(0,-1){5.33}} \put(-1.90,106.53){\makebox(0,0)[cc]{$\bullet$}} \put(23.63,106.53){\makebox(0,0)[cc]{$\bullet$}} \put(57.03,106.53){\makebox(0,0)[cc]{$\bullet$}} \put(82.69,106.53){\makebox(0,0)[cc]{$\bullet$}} \put(-1.90,85.20){\makebox(0,0)[cc]{$\bullet$}} \put(23.63,85.20){\makebox(0,0)[cc]{$\bullet$}} \put(-1.37,65.33){\line(0,-1){3.60}} \put(25.83,62.00){\line(0,1){3.20}} \put(53.30,61.87){\line(0,1){3.33}} \put(-1.37,61.73){\line(0,-1){6.13}} \put(25.83,61.73){\line(0,-1){6.13}} \put(53.30,61.73){\line(0,-1){6.27}} \put(-1.64,41.33){\line(0,-1){4.40}} \put(-1.50,22.13){\line(0,-1){4.53}} \put(26.23,41.33){\line(0,-1){4.40}} \put(54.10,41.33){\line(0,-1){4.40}} \put(26.36,22.13){\line(0,-1){4.53}} \put(54.23,22.13){\line(0,-1){4.53}} \put(52.66,95.95){\makebox(0,0)[cc]{2}} \put(57.19,95.95){\makebox(0,0)[cc]{3}} \put(65.21,95.95){\makebox(0,0)[cc]{1}} \put(74.34,95.95){\makebox(0,0)[cc]{3}} \put(82.36,95.95){\makebox(0,0)[cc]{1}} \put(86.89,95.95){\makebox(0,0)[cc]{2}} \put(91.08,106.58){\makebox(0,0)[lc]{= 2(31)}} \put(-6.44,96.12){\makebox(0,0)[cc]{1}} \put(-1.73,96.12){\makebox(0,0)[cc]{2}} \put(6.29,96.12){\makebox(0,0)[cc]{3}} \put(15.24,96.12){\makebox(0,0)[cc]{2}} \put(23.26,96.12){\makebox(0,0)[cc]{3}} \put(27.97,96.12){\makebox(0,0)[cc]{1}} \put(-6.44,75.03){\makebox(0,0)[cc]{3}} \put(-1.73,75.03){\makebox(0,0)[cc]{1}} \put(6.29,75.03){\makebox(0,0)[cc]{2}} \put(15.24,75.03){\makebox(0,0)[cc]{1}} \put(23.44,75.03){\makebox(0,0)[cc]{2}} \put(27.97,75.03){\makebox(0,0)[cc]{3}} \put(30.93,106.76){\makebox(0,0)[lc]{$=1(23)$,}} \put(30.93,85.14){\makebox(0,0)[lc]{$=3(12)$}} \put(-3.13,28.13){\makebox(0,0)[cc]{$2$}} \put(19.89,28.13){\makebox(0,0)[cc]{$2$}} \put(24.77,28.13){\makebox(0,0)[cc]{$3$}} \put(32.79,28.13){\makebox(0,0)[cc]{$1$}} \put(47.78,28.31){\makebox(0,0)[cc]{$3$}} \put(52.49,28.31){\makebox(0,0)[cc]{$1$}} \put(60.51,28.31){\makebox(0,0)[cc]{$2$}} \put(3.67,38.24){\makebox(0,0)[lc]{$=(1(23))$,}} \put(30.87,38.42){\makebox(0,0)[lc]{$=(2(31))$,}} \put(58.94,38.59){\makebox(0,0)[lc]{$=(3(12))$}} \put(3.72,19.67){\makebox(0,0)[lc]{$=(1(23))$,}} \put(30.88,19.83){\makebox(0,0)[lc]{$=(2(31))$,}} \put(58.88,20.00){\makebox(0,0)[lc]{$=((31)2)$}} \put(-8.10,8.40){\makebox(0,0)[cc]{$1$}} \put(5.81,8.40){\makebox(0,0)[cc]{$3$}} \put(0.04,8.63){\makebox(0,0)[cc]{$2$}} \put(19.89,8.63){\makebox(0,0)[cc]{$2$}} \put(27.94,8.63){\makebox(0,0)[cc]{$3$}} \put(32.79,8.63){\makebox(0,0)[cc]{$1$}} \put(47.78,8.81){\makebox(0,0)[cc]{$3$}} \put(55.66,8.81){\makebox(0,0)[cc]{$1$}} \put(60.51,8.81){\makebox(0,0)[cc]{$2$}} \put(6.50,167.33){\makebox(0,0)[cc]{$= (1)$}} \end{picture} \end{center} \caption{A representation of elements of ${{\cal B}C}(1)$, ${{\cal B}C}(2)$ and ${{\cal B}C}(3)$ by equivalence classes of planar trees.\label{master}} \end{figure} To be more formal, the isomorphism ${\varphi} : (\dualI{{\bf s\hskip0mm}{\it Cycl}}\ \circ \downarrow\!\dualI{{\bf s\hskip0mm} {\it Ass}}) \to {\cal M}$ is defined by ${\varphi}(\uparrow\!^{n-1} \xi_n) := f_n$, where $f_n = (\rada 1n)$ is the top $n$-dimensional cell of the cyclohedron $W_n$ and $\xi_n$ the generator of ${\it Cycl}(n)$ represented by $\mbox{$1\!\!1$}_n \in \Sigma_n$. We must specify also the orientation of $f_n$. In Observation~\ref{sdff} we constructed points $\rada {P_1}{P_n}$ spanning a simplex $\Delta_{f_n}\in f_n$. We orient $f_n$ coherently with the orientation of $\Delta_{f_n}\in f_n$ induced by the order $P_1 < \cdots < P_n$ of its vertices. The cobar differential $\partial_{{\Omega}}$ is given by \begin{equation} \label{tuti} \partial_{{\Omega}}(\uparrow\!^{n-1} \xi_n) = \cyclsum \sum_{1\leq k\leq n}\znamenko{n+k} \nu(\uparrow\!^{n-k}\xi_{n-k+1};\alpha_k,\rada11), \end{equation} where $\alpha_k= \mbox{$1\!\!1$}_k\in {\bf k}[\Sigma_k]= {\it Ass}$ is the generator, $k\geq 1$. We shall compare now~(\ref{tuti}) to the geometric boundary of the top-dimensional cell $f_n$ of the cyclohedron $W_n$. This can be done exactly as in the proof of Theorem~\ref{resiz} and we leave it to the reader.\hspace*{\fill The {\em proof of Theorem~\ref{nuzky}\/} is now immediate. An ${\cal M}$-trace $T: {\cal M}\to {\cal E}_{A,W}$ is determined by a system $\{T_n := t(\uparrow\!^{n-1}\xi_n): A^{\otimes n}\to W\}_{n \geq 1}$. The axiom~(\ref{Ax}) then reflects~(\ref{tuti}). We finish this section by the following theorem whose proof, based on a straightforward but involved spectral sequence argument, we omit. \begin{theorem} \label{Katalogizacni} Let $M_{{\it U}{\cal P}}$ be a module associated to a cyclic unital quadratic operad ${\cal P}$. Then $M_{U{\cal P}}$ is Koszul if and only if ${\cal P}$ is. \end{theorem} Because ${\it Cycl} = M_{{\it UAss}}$ (Example~\ref{kacirek}) and the operad ${\it Ass}$ is Koszul~\cite[Corollary~4.2.7]{ginzburg-kapranov:DMJ94}, Theorem~\ref{Katalogizacni} gives an alternative proof of Theorem~\ref{Amphora1}. \section{Cyclohedron as a compactification of the simplex} \label{22} For a compact Riemannian manifold $V$, let $C^0_n(V) =\{(\rada{v_1}{v_n});\ v_i\not= v_j\}$ be the configuration space of $n$ distinct points in $V$. Axelrod and Singer constructed in~\cite{axelrod-singer:preprint} a compactification $C_n(V)$ of this space, by adding to $C^0_n(V)$ the blow-ups along the diagonals. The space $C_n(V)$ is a manifold with corners, whose open part (= top-dimensional stratum) is $C^0_n(V)$. There exists a similar compactification of the {\em moduli space\/} ${\buildrel \circ \over {{\sf F}}}_m(n)$ of configurations of $n$ distinct points in the $m$-dimensional Euclidean plane ${\bf R}^m$ modulo the action of the affine group, described by Getzler and Jones in~\cite[\S3.2]{getzler-jones:preprint} and denoted by ${\sf F}_m(n)$. The authors of~\cite{getzler-jones:preprint} also observed that the collection ${\sf F}_m := \{{\sf F}_m(n)\}_{n\geq 1}$ has a natural structure of a topological operad. In~\cite{markl:cf} we proved that \begin{theorem} \label{1725} If $V$ is an $m$-dimensional parallelizable Riemannian manifold, then the collection $C(V) := \{C_n(V)\}_{n\geq 1}$ forms a right module over the operad ${\sf F}_m$ in the category of manifolds with corners. \end{theorem} There is also a `framed' version of the above theorem for manifolds which are not parallelizable, but we will not need it. Take $V=S^1$. Then it is immediately seen that the space $C_n^0(V)$ has $(n-1)!$ components indexed by cyclic orders of $n$ points on the circle. Each of these components is isomorphic to $\oDelta^n \times S^1$, where $\oDelta^n$ is the open $n$-dimensional simplex. It is `well-known' (see Remark~\ref{beuo} below) that the compactification $C_n(S^1)$ is isomorphic to ${\overline W}_n \times S^1$, the product of the symmetrized cyclohedron with the circle~\cite[page~5249]{bott-taubes:JMP94}. Similarly, ${\buildrel \circ \over {{\sf F}}}_1(n)$ is easily seen to have $n!$ components indexed by orders of the set of $n$ points on the line, each component being isomorphic to $\oDelta^{n-2}$. Again, it is `well-known' that the compactification ${\sf F}_1(n)$ is the (symmetrized) associahedron ${\overline K}_n$~\cite[3.2(1)]{getzler-jones:preprint}. This assumed, our statement (Theorem~\ref{sirky}) about the existence of a ${\overline K}$-module structure on the cyclohedron follows from Theorem~\ref{1725} applied on $C(S^1)= {\overline W} \times S^1$ (the extra factor $S^1$ plays no r\^ole). \begin{remark}{\rm \label{beuo} The Axelrod-Singer compactification $C_n(S^1)$ is a manifold with corners, constructed by a very explicit sequence of blow-ups. We do not know any `universal' characterization of this space. Thus to prove that $C_n(S^1) \cong {\overline W}_n \times S^1$ would require an explicit construction of an isomorphism of two manifolds with corners, which is certainly not a tempting challenge. But a reflection on the structure of these two object `proves' the isomorphism `beyond any doubts', which is the opinion shared by many authors. The same remark applies also to the isomorphism ${\sf F}_1(n)\cong {\overline K}_n$. }\end{remark} \begin{remark}{\rm It follows from general properties of manifold-with-corners that both $K_n$ and $W_n$ are truncations of a simplex~\cite[Proposition~6.1]{markl:cf}, but this existence statement says nothing about an explicit linear convex realization of Section~\ref{1968}. }\end{remark} As we observed above, the cyclohedron $W_n$ can be viewed as the simplex $\Delta^n$, some faces of whose were blown-up. In the rest of this section we show that a very natural spectral sequence related to the cobar construction can be interpreted as an inverse process -- `deblowing-up' of the cyclohedron back to the closed simplex. For a collection $X$ and $p\geq 1$, let $\skel Xp\subset X$ be the subcollection defined by $\skel Xp(n) = X(n)$ for $n\leq p$ and $\skel Xp(n) = 0$ otherwise. Let us consider, for a comodule $N$ over an cooperad ${\cal Q}$ and for a natural $n$, the subspace \[ F_p(n) := (\skel N{p+1} \circ {\cal F}(\downarrow\! {\cal Q}))(n) \subset (N \circ {\cal F}(\downarrow\! {\cal Q}))(n). \] It is easily seen that $F_p(n)$ is ${\partial_{\Cob}}$-invariant, thus $\{F_p(n)\}_{p\geq 0}$ is an increasing filtration of the $n$-th piece ${\Omega}(N,{\cal Q})(n)$ of the cobar construction ${\Omega}(N,{\cal Q}) = (N \circ {\cal F}(\downarrow\! {\cal Q}), {\partial_{\Cob}})$. Let ${\bf E}(n) = (E^r_{pq}(n), d^r)$ be the corresponding spectral sequence. The following lemma is an easy exercise. \begin{lemma} \label{whoop} The spectral sequence ${\bf E}(n) = (E^r_{pq}(n), d^r)$ constructed above converges to $H_*({\Omega}(N,{\cal Q})(n))$. The first term $E^1$ is described as \[ E^1_{pq}(n)= (N(p+1)\circ H_*({\Omega}({\cal Q})))(n)_{p+q}, \] the space of elements of degree $p+q$ in the $n$-th piece of the free $H_*({\Omega}({\cal Q}))$-module on the $(p+1)$-th piece of the collection $N$. \end{lemma} If the cooperad ${\cal Q}$ and the comodule $N$ are Koszul, the spectral sequence above collapses at the 1st term, which has a very explicit description. Since we did not formulate the Koszulness for cooperads and comodules (though the definition is an exact dual), we suppose from now on that $N = \dualI{{\bf s\hskip0mm} M}$ and ${\cal Q} = \dualI{{\bf s\hskip0mm} {\cal P}}$, for a module $M$ over an operad ${\cal P}$. We also suppose that $M$ and ${\cal P}$ are not graded, i.e.~ that both $M(n)$ and ${\cal P}(n)$ are concentrated in degree $0$, $n\geq 1$. \begin{proposition} \label{myska} If the operad ${\cal P}$ is Koszul, then $E^1_{pq}(n)= 0$ for $q\geq 1$, while \[ E^1_{p0}(n)= (\dualI{{\bf s\hskip0mm} M}(p+1) \circ {\cal P}^!)(n), \] and the spectral sequence collapses at $E^1$. The module $M$ is Koszul if and only if the complex \[ 0 \stackrel{d^1}{\longleftarrow} E^1_{00}(n) \stackrel{d^1}{\longleftarrow} E^1_{10}(n) \stackrel{d^1}{\longleftarrow} E^1_{20}(n) \longleftarrow \cdots \] is acyclic in positive dimensions, for all $n\geq 1$. \end{proposition} \noindent {\bf Proof.} If ${\cal P}$ is Koszul, then $H_*({\Omega}({\cal P}))= H_0({\Omega}({\cal P}))={\cal P}^!$, by definition. Thus, by Lemma~\ref{whoop}, $E^1_{pq}(n)= (\dualI{ {\bf s\hskip0mm} M}(p+1)\circ {\cal P}^!)(n)_{p+q}$ which may be nonzero only for $q=0$, because $\dualI{ {\bf s\hskip0mm} M}(p+1)$ is concentrated in degree $p$. The collapsing is obvious from degree reasons. The second part of the statement follows immediately from the definition of the Koszulness of a module (Definition~\ref{telefon}).\hspace*{\fill Our spectral sequence has, for ${\cal P} = {\it Ass}$ and $M ={\it Cycl}$, a beautiful geometric meaning. The initial term ${\bf E}^0 = (E^0_{pq},d^0)$ is the cobar construction ${\Omega}(\dualI{{\bf s\hskip0mm}{\it Cycl}}, \dualI{{\bf s\hskip0mm}{\it Ass}})$ which is isomorphic, by Theorem~\ref{ucpavka}, to the cellular chain complex of the cyclohedron ${\overline W}_n$, while ${\bf E}^1 = (E^1_{pq},d^1) = ({\bf s\hskip0mm}{\it Cycl} \circ {\it Ass},\partial)$ is isomorphic, by Theorem~\ref{resiz}, to the cellular chain complex of the simplex ${\overline \Delta}_n$. The passage from ${\bf E}^0$ to ${\bf E}^1$ can be interpreted as the `deblowing-up' of the cyclohedron back to the simplex. This process is visualized on Figure~\ref{deblow}. \begin{figure}[hbtp] \begin{center} \unitlength 1.50mm \thicklines \begin{picture}(91.66,52.08) \put(0.50,37.00){\line(1,1){10.00}} \put(10.72,47.41){\line(1,0){20.00}} \put(30.72,47.08){\line(1,-1){10.00}} \put(40.39,37.41){\line(-1,-1){10.00}} \put(30.72,27.08){\line(-1,0){20.00}} \put(10.86,27.21){\line(-1,1){10.00}} \put(0.72,37.08){\makebox(0,0)[cc]{$\bullet$}} \put(10.72,47.08){\makebox(0,0)[cc]{$\bullet$}} \put(30.72,47.08){\makebox(0,0)[cc]{$\bullet$}} \put(40.72,37.08){\makebox(0,0)[cc]{$\bullet$}} \put(30.72,27.08){\makebox(0,0)[cc]{$\bullet$}} \put(10.72,27.08){\makebox(0,0)[cc]{$\bullet$}} \put(8.72,52.08){\makebox(0,0)[cc]{$((12)3)$}} \put(32.72,52.08){\makebox(0,0)[cc]{$(1(23))$}} \put(42.72,37.08){\makebox(0,0)[lc]{$1)((23)$}} \put(32.72,22.08){\makebox(0,0)[cc]{$1))(2(3$}} \put(8.72,22.08){\makebox(0,0)[cc]{$1)2)((3$}} \put(-1.28,37.08){\makebox(0,0)[rc]{$(12))(3$}} \put(20.72,52.08){\makebox(0,0)[cc]{$(123)$}} \put(20.72,22.08){\makebox(0,0)[cc]{$1)2(3$}} \put(7.72,42.08){\makebox(0,0)[lc]{$(12)3$}} \put(7.72,32.08){\makebox(0,0)[lc]{$12)(3$}} \put(33.72,42.08){\makebox(0,0)[rc]{$1(23)$}} \put(33.72,32.08){\makebox(0,0)[rc]{$1)(23$}} \put(20.72,37.08){\makebox(0,0)[cc]{$123$}} \put(55.83,4.33){\line(5,6){16.67}} \put(72.50,24.33){\line(5,-6){16.81}} \put(89.30,4.16){\line(-1,0){33.64}} \put(10.72,46.75){\line(1,0){20.00}} \put(40.86,36.94){\line(-1,-1){10.00}} \put(10.39,26.75){\line(-1,1){10.00}} \put(72.33,24.11){\makebox(0,0)[cc]{$\bullet$}} \put(55.66,4.11){\makebox(0,0)[cc]{$\bullet$}} \put(89.44,4.11){\makebox(0,0)[cc]{$\bullet$}} \put(53.00,1.89){\makebox(0,0)[cc]{$\{3\}$}} \put(91.66,2.11){\makebox(0,0)[cc]{$\{2\}$}} \put(72.55,27.44){\makebox(0,0)[cc]{$\{1\}$}} \put(63.25,16.77){\makebox(0,0)[rb]{$\{13\}$}} \put(81.44,17.00){\makebox(0,0)[lb]{$\{12\}$}} \put(72.33,1.44){\makebox(0,0)[ct]{$\{23\}$}} \put(72.33,12.33){\makebox(0,0)[cc]{$\{123\}$}} \put(44.84,29.43){\line(6,-5){9.02}} \put(44.20,28.74){\line(6,-5){9.01}} \put(53.93,21.00){\line(-1,0){2.61}} \put(53.93,21.00){\line(0,1){2.16}} \end{picture} \end{center} \caption{$\Delta_3$ as deblowing-up of $W_3$. The faces of $W_3$ which are contracted by $d^0$ to a vertex are indicated by double lines, $(123)$ is contracted to $\set 1$, $1)(23$ to $\set 2$ and $12)(3$ to $\set 3$.\label{deblow}} \end{figure} \section*{Appendix: Traces versus invariant bilinear forms.} Let us recall the following notion of~\cite[Definition~4.1]{getzler-kapranov:cyclic}. If ${\cal P}$ is a cyclic operad and $A$ a ${\cal P}$-algebra, then a bilinear form $B: A\otimes A\to W$ with values in a vector space $W$ is called {\em invariant\/} if, for all $n\geq 0$, the map $B_n:{\cal P}(n)\otimes A^{\otimes (n+1)} \to W$ defined by the formula \begin{equation} \label{plus} B_n(p{\otimes} x_0 {\otimes} x_1 {\otimes} \cdots {\otimes} x_n):= \znamenko {|x_0|\cdot |p|}B(x_0,p(\rada {x_1}{x_n})), \end{equation} is invariant under the action of the symmetric group $\Sigma_{n+1}$ on ${\cal P}(n){\otimes} A^{\otimes (n+1)}$. \vskip2mm \noindent {\bf Proposition A.1.} {\it Let ${\cal P}$ be a cyclic operad and let $M_{{\cal P}}$ be the associated module introduced in Definition~\ref{Turmo}. Let $A$ be a ${\cal P}$-algebra. Then there exists a 1-1 correspondence between $M_{{\cal P}}$-traces on the ${\cal P}$-algebra $A$ in the sense of Definition~\ref{el}, and invariant bilinear forms on $A$.} \vskip2mm \noindent {\bf Proof.} Let $t : M_{\cal P} \to {\cal E}_{A,W}$ be an $M_{\cal P}$-trace. Because $M_{\cal P}(n+1)= {\cal P}(n)$, the trace is represented by a system $\{t_n :{\cal P}(n-1)\to \Hom{A^{\otimes n}}W\}_{n\geq 2}$ of linear maps. We claim that $B := t_2(1): A\otimes A \to W$, where $1\in {\cal P}(1)$ is the unit, is an invariant bilinear form. To see it, observe that~(\ref{plus}) can be rewritten as \[ B_n(p\otimes x_0 \otimes \cdots \otimes x_n) = \nu_{{\cal E}_{A,W}}(t_2(1);1,p)(\rada{x_0}{x_n}), \] while \[ \nu_{{\cal E}_{A,W}}(t_2(1);1,p) = \nu_{{\cal E}_{A,W}}(\nu_{M_{\cal P}}(1;1,p)) = t_{n+1}(p \cdot \tau_n), \] thus \begin{equation} \label{xplus} B_n(p\otimes x_0 \otimes \cdots \otimes x_n) = t_{n+1}(p \cdot \tau_n)(\rada{x_0}{x_n}) \end{equation} and the equivariance of $B_n$ follows from the equivariance of $t_{n+1}$. On the other hand, if $B$ is an invariant bilinear form, then~(\ref{xplus}) defines a trace.\hspace*{\fill
1997-07-13T19:55:18
9707
alg-geom/9707012
en
https://arxiv.org/abs/alg-geom/9707012
[ "alg-geom", "math.AG" ]
alg-geom/9707012
Kalle Karu
Dan Abramovich and Kalle Karu
Weak semistable reduction in characteristic 0
AMS-LaTeX, 22 pages
null
null
null
null
Let X->B be a morphism of varieties in characteristic zero. Semistable reduction has been proved for dim(B)=1 (Kempf, Knudsen, Mumford, Saint-Donat), dim(X)=dim(B)-1 (de Jong) and dim(X)=dim(B)+2 (Alexeev, Kollar, Shepherd-Barron). In this paper we consider the general case. First we define what we mean by a semistable morphism in terms of toroidal embeddings. Then we reduce the varieties to toroidal embeddings and solve a slightly weaker version of semistable reduction. We also state the full semistable reduction problem in terms of combinatorics of the associated polyhedral complexes.
[ { "version": "v1", "created": "Sun, 13 Jul 1997 17:55:10 GMT" } ]
2008-02-03T00:00:00
[ [ "Abramovich", "Dan", "" ], [ "Karu", "Kalle", "" ] ]
alg-geom
\section{INTRODUCTION} Regretfully, we work over an algebraically closed field $k$ of characteristic 0. \subsection{The problem} Roughly speaking, the semistable reduction problem we address here asks for the following: \begin{quote} Let $X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective morphism of complex projective varieties with geometrically integral generic fiber. Find a generically finite proper surjective morphism (that is, an alteration) $B_1\rightarrow} \newcommand{\dar}{\downarrow B$, and a proper birational morphism (that is, a modification) $Y\rightarrow} \newcommand{\dar}{\downarrow X\times_B B_1$, such that the morphism $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ is nice. \end{quote} Of course, one needs to decide what a ``nice morphism'' means. The question was posed, among other places, in the introduction of \cite{te}, p. vii. It can be viewed as a natural extension of Hironaka's theorem on resolution of singularities, which is in a sense ``the general fiber'' of semistable reduction. \subsection{Brief history} The case { $\dim B=1, \dim X=2$ is very old, see \cite{aw}. When $\dim B=1,$ semistable reduction was obtained in \cite{te}, in the best possible sense: $Y$ is nonsingular, and all the fibers are reduced, strict divisors of normal crossings. Using a result of Kawamata on ramified covers (see \cite{kawamata}, theorem 17), one can obtain semistable reduction ``in codimension 1'' over a base of arbitrary dimension. Below, we will refer to the result of Kawamata as {\bf ``Kawamata's trick''.} We will discuss it in detail in section \ref{reduced-fibers}. The case where $\dim X = \dim B + 1$ has recently been proven by de Jong \cite{dj}. Here one shows that any family of curves can be made into a family of nodal curves, which are indeed as ``nice'' as one may expect. Using recent difficult results of Alexeev, Koll\'ar and Shepherd-Barron (see \cite{alex}, \cite{alex2}), one obtains a version of the case $\dim X = \dim B + 2$. Here each fiber is a semi-log-canonical surface. Up until recently, not much has been known about the case $\dim X >\dim B+2$. Often one finds remarks of the following flavor: ``since we do not have a semistable reduction result over a base of higher dimension, we will work around it in the following technical manner...''. \subsection{Definition of semistable families} We give here a description of the best possible kind of morphisms we have in mind. Let $f:X\to B$ be a flat morphism of nonsingular projective varieties with connected fibers. Somewhat informally, we say that $f$ is {\bf semistable} if for each point $x\in X$ with $f(x) = b$ there is a choice of formal coordinates $\hat{B}_b = {\operatorname{Spec\ }}\ k[[t_i]]$ and $\hat{X}_x = {\operatorname{Spec\ }}\ k[[x_j]]$, such that $f$ is given by: $$t_i = \prod_{j=l_{i-1}+1}^{l_i} x_j. $$ Here $0 = l_0 < l_1 \cdots < l_m \leq n$, where $n=\dim X$ and $m = \dim B$. To be more precise, we give things a more global structure using the notion of a toroidal morphism. At the same time we describe a slightly weaker condition which will appear below: \begin{dfn} The morphism $f:X\to B$ above is called {\bf weakly semistable} if \begin{enumerate} \item the varieties $X$ and $B$ admit toroidal structures $U_X\subset X$ and $U_B\subset B$, with $U_X=f^{-1}U_B$; \item with this structure, the morphism $f$ is toroidal; \item the morphism $f$ is equidimensional; \item all the fibers of the morphism $f$ are reduced; and \item $B$ is nonsingular. \end{enumerate} If also $X$ is nonsingular, we say that the morphism $f:X\to B$ is {\bf semistable}. \end{dfn} \subsection{The ultimate goal} The result one would really like to have is: \begin{conj}\label{conj-semistable} Let $X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective morphism of complex projective varieties with geometrically integral generic fiber. There is a projective alteration $B_1\rightarrow} \newcommand{\dar}{\downarrow B$, and a projective modification $Y\rightarrow} \newcommand{\dar}{\downarrow X\times_B B_1$, such that $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ is semistable. \end{conj} Na\"{\i}vely one might hope to have each fiber isomorphic to a divisor of normal crossings. But already in the case of a 2-parameter family of surfaces $t_1 = x_1x_2;\, t_2 = x_3 x_4$, this is impossible. It seems that the definition above is the best one can hope for. \subsection{A. J. de Jong's results} In \cite{dj}, Johan de Jong shows, among many other results, that if one allows $Y\rightarrow} \newcommand{\dar}{\downarrow X\times_B B_1$ to be an alteration instead of a modification, one can make $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ very nice indeed: $Y$ is nonsingular, $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ is semistable as in the definition above, and moreover it can be written as a composition of nodal curve fibrations $Y=Y_0\to Y_1\to \cdots \to Y_k=B_1$. De Jong's methods and ideas will serve as a starting point for investigating the semistable reduction conjecture. \subsection{Our main result} The main result of this paper is the following: \begin{th}[Weak semistable reduction]\label{th-weak-semistable-reduction} Let $X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective morphism of complex projective varieties with geometrically integral generic fiber. There exist an alteration $B_1\rightarrow} \newcommand{\dar}{\downarrow B$ and a modification $Y\rightarrow} \newcommand{\dar}{\downarrow X\times_B B_1$, such that $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ is weakly semistable. \end{th} With a little more work we will get $X$ to have only quotient singularities. There are many cases (such as when $f$ is a family of surfaces) where we can actually prove the semistable reduction conjecture. These will be pursued elsewhere. Hopefully, by the time this paper achieves its final form the conjecture will be fully proven. \subsection{Mild morphisms} A few words are in order about the significance of our result. Note that the property of a morphism being semistable is far from being stable under base changes. One may ask, what remains from semistability after at least {\em dominant} base changes? Here is a suggestion: \begin{dfn} We define a morphism $X\rightarrow} \newcommand{\dar}{\downarrow B$ as above to be {\bf mild}, if for any dominant $B_1\rightarrow} \newcommand{\dar}{\downarrow B$ where $B_1$ has at most rational Gorenstein singularities, we have that $X\times_B B_1$ has at most rational Gorenstein singularities as well. \end{dfn} Mild morphisms arise naturally in moduli theory. Indeed, mild families of curves are precisely nodal families; families of Gorenstein semi-log-canonical surfaces mentioned above are mild. For a discussion of why mild morphisms are useful, see \cite{fibered}. In fact, the paper \cite{fibered} would have been much simplified, had mild reduction been available. Already in the case $\dim B=1$, mild reduction is a much easier task than semistable reduction. Indeed, lemma 2 on page 103 of \cite{te}, and the discussion there, already give mild reduction in this case. The delicate combinatorics of chapter III of \cite{te} is not used for this purpose. It will be shown (see section \ref{mildness}) that weakly semistable morphisms are indeed mild. \subsection{Structure of the proof} After the introduction, section \ref{toroidal-morphisms} will be devoted to a general discussion of toroidal morphisms. The proof itself will begin with section \ref{toroidal-reduction}. Semistable reduction has at least two flavors: first, the fibers of the morphism $Y\rightarrow} \newcommand{\dar}{\downarrow B_1$ should have nice local defining equations. Second, the family should have nice algebraic properties. We will perform a number of reduction steps, incrementally improving one or the other of these flavors. \subsubsection{Toroidal reduction} In the first step, carried out in section \ref{toroidal-reduction}, we will show that any morphism can be modified to a toroidal morphism. The construction is inspired by the inductive procedure of \cite{dj}, and follows closely the proofs in \cite{aj}. Just as in \cite{aj}, the construction we give is very non-canonical. Even when the generic fiber of $X\to B$ is smooth, it will be blown up during the construction. One hopes that methods such as those of \cite{bm} or \cite{villa} could be adapted to this situation and give a more canonical procedure. It is tempting to state the following conjecture. \begin{conj} Let $X\to B$ be a morphism as in the theorem. Let $U\subset B$ be an open set over which $X$ is toroidal, and let $\Sigma = B \setmin U$. There exists modifications $X'\to X$ and $B'\to B$, each of which is the composition of a sequence of blowings up with smooth centers lying over $\Sigma$, and a lifting $X'\to B'$ which is toroidal. \end{conj} It should be noted, that in view of recent results of Morelli \cite{m} and W{\l}odarczyk \cite{w}, this conjecture implies the strong blow-up - blow-down conjecture. \subsubsection{Improving the toroidal morphism} In sections \ref{remove-horizontal} and \ref{equidimensional} we perform a couple of simple reduction steps to improve our situation. {Let $f:(U_X \subset X) \rightarrow} \newcommand{\dar}{\downarrow(U_B\subset B)$ be any toroidal morphism, with $B$ nonsingular. By the results of \cite{te}, we can find a toroidal resolution of singularities $X'\rightarrow} \newcommand{\dar}{\downarrow X$. Let $f':X'\rightarrow} \newcommand{\dar}{\downarrow B$ be the resulting projection. We first show that now $f^{-1} U_B\subset X'$ is also a toroidal embedding, which is easier to handle: there are no horizontal divisors. For convenience, we replace $X\rightarrow} \newcommand{\dar}{\downarrow B$ by the new morphism. We remark that one can proceed a fair distance without removing these horizontal divisors, and, we believe, the results one can obtain are of interest (e.g., the inductive structure of de Jong can be preserved), but this would make the present paper much more cumbersome, so we delay that investigation to a future occasion. Now, our morphism $X\rightarrow} \newcommand{\dar}{\downarrow B$ is not necessarily equidimensional. We repair this by an appropriate decomposition of the associated conical polyhedral complexes $\Delta_X$ and $\Delta_B$. We make sure that, after the modification, the base remains nonsingular, and then the morphism is automatically flat. \subsubsection{Kawamata's trick and reduced fibers.} We start section \ref{reduced-fibers} with a discussion of Kawamata's trick and its relation with toroidal morphisms in some detail. Then we use Kawamata's trick to find a finite base change, after which all the fibers are reduced. This finishes the proof of the main theorem, since the resulting morphism is weakly semistable. A variant of Kawamata's trick for global ``index 1 covers'' is discussed in section \ref{sec-cartier}. \subsubsection{Mild reduction.} We begin section \ref{mildness} by checking that the resulting fibers are Gorenstein. Using a base change and descent argument, and the fact that toroidal singularities are always rational, we then prove that the resulting family is mild. \subsubsection{Combinatorial restatement} In section \ref{combinatorial} the semistable reduction conjecture is restated purely in combinatorial terms. We end the paper with a discussion of the problems one encounters when trying to go from weak semistable reduction to semistable reduction. \subsection{Acknowledgments} We would like to thank O. Gabber, A.J. de Jong, H. King and K. Matsuki for helpful and inspiring discussions. \subsection{Terminology} A {\bf modification} is a proper birational morphism of irreducible varieties. An {\bf alteration} $a:B_1\rightarrow} \newcommand{\dar}{\downarrow B$ is a proper, surjective, generically finite morphism of irreducible varieties, see \cite[2.20]{dj}. The alteration $a$ is a {\bf Galois alteration} if there is a finite group $G\subset {\operatorname{Aut}}_B (B_1)$ such that the associated morphism $B_1/G\rightarrow} \newcommand{\dar}{\downarrow B$ is birational, compare \cite[5.3]{dj2}. \section{Toroidal morphisms}\label{toroidal-morphisms} We have collected in this section some notations and preliminaries about toric varieties, toroidal embeddings, and their morphisms (see \cite{te} for details). \begin{rem} Our approach here is based on the formalism of \cite{te}. A different approach, using {\em logarithmic structures}, was developed by K. Kato, see \cite{kato}, \cite{kato1}. It is our belief, that the approach via logarithmic structures should eventually prevail - it provides us with a flexible category, in which toroidal embeddings ($=$ logarithmically regular schemes) and toroidal morphisms ($=$ logarithmically smooth(!) morphisms) play a special role. Some of our statements below are rendered almost trivial with Kato's formalism, e.g. Lemmas \ref{lem-tor-composition} and \ref{lem-tor-prod}. The reason we decided to stick with the formalism of \cite{te} is, that the theory of logarithmic structures is not yet in stable form (see the many flavors of such structures introduced in Kato's papers), and, more importantly, it has not yet gained widespread acceptance as a basic formalism. It might have turned away some readers (especially those combinatorially inclined) had we used the theory of logarithmic structures throughout. It is also worth noting, that Kato's notion of a fan, although it has a nice structural morphism, is much less amenable to combinatorial manipulation than the polyhedral complexes of \cite{te}. \end{rem} \subsection{Toric varieties} Given a lattice $N\cong{\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}^n$, its dual $M={\operatorname{Hom }}(N,{\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}})$, a strictly convex rational polyhedral cone $\sigma\subset N_{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}} = N\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ with its dual $\sigma^\vee = \{m\in M_{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}} | m(u)\geq 0$ for all $u\in \sigma\}$, we define the {\bf affine toric variety} $X_\sigma = {\operatorname{Spec\ }} S[\sigma]$ where $S[\sigma]$ is the semigroup algebra of $\sigma^\vee\cap M$ over the ground field. If more than one toric variety is considered, we use a subscript: $N_\sigma$, $M_\sigma$. We denote by $\sigma^{(1)}$ the 1-dimensional edges of $\sigma$. The indivisible points $v$ in $\sigma^{(1)}\cap N$ are called the {\bf primitive points} of $\sigma$. The variety $X_\sigma$ is nonsingular if and only if the primitive points of $\sigma$ form a part of a basis of $N$. In that case we say that $\sigma$ is nonsingular. The toric variety $X_\sigma$ contains an n-dimensional algebraic torus $T=\bfg^n_m$ as an open dense subset, and the action of $T$ on itself extends to an action on $X_\sigma$. Thus, $X_\sigma$ is a disjoint union of orbits of this action. There is a one-to-one correspondence between the orbits and the faces of $\sigma$. In particular, 1-dimensional faces ${\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}_{+}v_i$ correspond to codimension~1 orbits $\bfo_{v_i}$. A {\bf toric morphism} $f:X_\sigma\rightarrow} \newcommand{\dar}{\downarrow X_\tau$ is a dominant equivariant morphism of toric varieties defined by a linear map $f_\Delta: (N_\sigma,\sigma) \rightarrow} \newcommand{\dar}{\downarrow (N_\tau,\tau)$. We use the same notation for the scalar extension $f_\Delta: N_\sigma\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}} \rightarrow} \newcommand{\dar}{\downarrow N_\tau\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$. \subsection{Toroidal embeddings} Given a normal variety $X$ and an open subset $U_X\subset X$, the embedding $U_X\subset X$ is called {\bf toroidal} if for every closed point $x\in X$ there exist a toric variety $X_\sigma$, a point $s\in X_\sigma$, and an isomorphism of complete local algebras \[ \hat{{\cal{O}}}_{X,x} \cong \hat{{\cal{O}}}_{X_{\sigma},s} \] so that the ideal of $X\setmin U_X$ corresponds to the ideal of $X_{\sigma}\setmin T$. Such a pair $(X_\sigma, s)$ is called a local model at $x\in X$. By restricting $X_{\sigma}$ if necessary, we can assume that the orbit of $s$ is the unique closed orbit in $X_{\sigma}$. \begin{dfn}\label{def-toroidal-map} A dominant morphism ${f}:(U_X\subset X)\rightarrow} \newcommand{\dar}{\downarrow(U_B\subset B)$ of toroidal embeddings is called {\bf toroidal} if for every closed point $x\in X$ there exist local models $(X_{\sigma},s)$ at $x$, $(X_{\tau},t)$ at ${f}(x)$ and a toric morphism $g: X_{\sigma}\rightarrow} \newcommand{\dar}{\downarrow X_{\tau}$ so that the following diagram commutes \[ \begin{CD} \hat{{\cal{O}}}_{X,x} @<{\cong}<< \hat{{\cal{O}}}_{X_{\sigma},s}\\ @A{\hat{f}^{*}}AA @AA{\hat{g}^{*}}A\\ \hat{{\cal{O}}}_{B,f(x)} @<{\cong}<< \hat{{\cal{O}}}_{X_{\tau},t} \end{CD} \] where $\hat{f}^{*}$ and $\hat{g}^{*}$ are the ring homomorphisms induced by $f$ and $g$. \end{dfn} \subsection{Cones and polyhedral complexes} Let $X\setmin U_X = \cup_{i\in I}E_i$ where $E_i$ are irreducible and have codimension~1. We will assume that all the $E_i$ are normal, that is, $U_X\subset X$ is a {\bf toroidal embedding without self-intersection} (also known as a {\bf strict} toroidal embedding). In that case, we can use the irreducible components of $\cap_{i\in J} E_i$ for all $J\subset I$ to define a stratification of $X$ (these components are the closures of strata). Closures of strata formally corresponds to closures of orbits in local models. Since a toric morphism maps orbits to orbits, a toroidal morphism maps strata to strata. Let Y be a stratum in $X$, which is by definition an open set in an irreducible component of $\cap_{i\in J} E_i$ for some $J\subset I$. The star of $Y$ is the union of strata in whose closure $Y$ lies (each of these corresponds to some $K\subset J\subset I$). To the stratum $Y$ we associate \begin{enumerate} \item $M^Y$: -- the group of Cartier divisors in ${\operatorname{Star}}(Y)$ supported in ${\operatorname{Star}}(Y)\setmin U_X$ \item $N^Y$: -- ${\operatorname{Hom }}(M^Y,{\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}})$ \item $M_{+}^Y \subset M^Y$: -- effective Cartier divisors \item $\sigma^Y\subset N^Y_{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$: -- the dual of $M_{+}^Y$ \end{enumerate} It is shown in \cite{te} (Corollary 1, page 61) that if $(X_{\sigma},s)$ is a local model at $x\in X$ in the stratum $Y$, then \begin{eqnarray*} M^Y &\cong& M_\sigma/\sigma^{\bot} \\ \sigma^Y &\cong& \sigma \end{eqnarray*} The cones $\sigma^Y$ glue together to form a polyhedral complex $\Delta_X = (|\Delta_X|,\{\sigma^Y\},\{M^Y\})$, where the lattices $M^Y$ form an integral structure on $\Delta_X$. Equivalently, instead of $M^Y$ we may give the lattices $N^Y$ and embeddings $\sigma^Y\hookrightarrow N^Y_{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$. We also denote $M_{\sigma^Y}=M^Y$, $N_{\sigma^Y}=N^Y$. Then, comparing to the lattices of local models, \begin{eqnarray*} M_{\sigma^Y} &\cong& M_\sigma/\sigma^{\bot} \\ N_{\sigma^Y} &\cong& N_\sigma\cap \mbox{Span}(\sigma) \\ \end{eqnarray*} If $N_{\sigma^Y}\neq N_\sigma$ we can choose a splitting of $N_\sigma$ so that at a point $x\in Y$ the local model is in the form $(X_{\sigma'}\times\bfg_m^l,(s',1))$ where $(N_{\sigma'},\sigma')\cong (N_{\sigma^Y},\sigma^Y)$. \begin{lem} A toroidal morphism ${f}:X\rightarrow} \newcommand{\dar}{\downarrow B$ induces a morphism $f_\Delta:\Delta_X \rightarrow} \newcommand{\dar}{\downarrow \Delta_B$, which for each cone $\sigma^Y$ is the restriction of $g_\Delta: (\sigma,N_\sigma)\rightarrow} \newcommand{\dar}{\downarrow(\tau,N_\tau)$ where $(X_{\sigma},s)$, $(X_{\tau},t)$ are local models at $x\in Y\subset X$, $f(x)\in Z\subset B$, and $g_\Delta$ is the linear map determined by the toric morphism $g: X_{\sigma}\rightarrow} \newcommand{\dar}{\downarrow X_{\tau}$ in Definition~\ref{def-toroidal-map}. \end{lem} {\bf Proof.} To see that the maps $f_\Delta$ defined for different cones $\sigma^Y$ agree on the overlaps it suffices to notice that the dual morphism $f_\Delta^\vee:M_{\tau^Z}\rightarrow} \newcommand{\dar}{\downarrow M_{\sigma^Y}$ is defined by pulling back a Cartier divisor and restricting it to ${\operatorname{Star}}(Y)\setmin U_X$. Since the pullback is defined independently of the stratum, we see that $f_\Delta$ is well defined. \qed \begin{rem} Note that the polyhedral morphism $f_\Delta:\Delta_X \rightarrow} \newcommand{\dar}{\downarrow \Delta_B$ is well defined even if $f$ is not toroidal, as long as $f(U_X)\subset U_B$. \end{rem} \begin{lem}\label{lem-tor-composition} If $e:X\rightarrow} \newcommand{\dar}{\downarrow Y$ and $f:Y\rightarrow} \newcommand{\dar}{\downarrow Z$ are toroidal morphisms, then $f\circ e:X\rightarrow} \newcommand{\dar}{\downarrow Z$ is also toroidal. \end{lem} \begin{rem} This lemma is a triviality if one uses logarithmic structures. \end{rem} {\bf Proof.} Let $x\in X$, $y=e(x)\in Y$, $z=f(y)\in Z$, and choose local models at $x$, $y$ and $z$ as in Definition~\ref{def-toroidal-map}. Consider the tower \[ \begin{CD} \hat{{\cal{O}}}_{X,x} @<{\cong}<< \hat{{\cal{O}}}_{X_{\sigma},s}\\ @A{\hat{e}^{*}}AA @AA{\hat{g}^{*}}A\\ \hat{{\cal{O}}}_{Y,y} @<{\cong}<< \hat{{\cal{O}}}_{X_{\tau_1},t_1}\\ @A{id}AA @AA{\alpha}A\\ \hat{{\cal{O}}}_{Y,y} @<{\cong}<< \hat{{\cal{O}}}_{X_{\tau_2},t_2}\\ @A{\hat{f}^{*}}AA @AA{\hat{h}^{*}}A\\ \hat{{\cal{O}}}_{Z,z} @<{\cong}<< \hat{{\cal{O}}}_{X_{\rho},r} \end{CD} \] where the upper and lower squares commute by the definition of toroidal morphism, and where $\alpha$ is defined by tracing the other three sides of the middle square. Then the middle square also commutes. Since $\tau_1$ and $\tau_2$ are isomorphic, we can take $\tau_1=\tau_2=\tau$ and $X_{\tau_1}= X_{\tau_2}= X_{\tau}$. The map $\alpha$, of course, need not be the identity. Let the coordinate rings of the tori in $X_{{\sigma}}$, $X_{{\tau}}$ and $X_{{\rho}}$ be $k[x_1,x_1^{-1},\ldots,x_n,x_n^{-1}]$, $k[y_1,y_1^{-1},\ldots,y_m,y_m^{-1}]$ and $k[z_1,z_1^{-1},\ldots,z_l,z_l^{-1}]$, respectively, so that the toric morphisms $g$ and $h$ are defined by \[ y_i \mapsto \prod_{j=1}^{n} x_j^{a_{ij}}, \hspace{.5in} z_i \mapsto \prod_{j=1}^{m} y_j^{b_{ij}} \] for some $a_{ij}, b_{ij} \in {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}$. The maps in the third line of the tower identify the group of T-invariant Cartier divisors in $X_{{\tau_2}}$ with the group of Cartier divisors in $Y$ supported in $Y\setmin U_Y$ and passing through $y$. The maps in the second line of the tower identify the latter group with the group of T-invariant Cartier divisors in $X_{{\tau_1}}$. Hence $\alpha$ induces a group homomorphism between T-invariant Cartier divisors of $X_{{\tau_2}}$ and $X_{{\tau_1}}$. Since T-invariant Cartier divisors in $X_{{\tau}}$ are given by products of $y_i$, then (using the same letter $y_i$ for the image of $y_i$ in the completed local ring) $\alpha$ maps \[ y_i \mapsto u_i \prod_{j=1}^{m} y_j^{c_{ij}} \] where $u_i$ are units in $\hat{{\cal{O}}}_{X_{{\tau}},t_1}$, and $c_{ij}\in {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}$. The composition $\hat{g}^{*}\circ\alpha\circ\hat{h}^{*}$ is then defined by \[ z_i \mapsto v_i \prod_{j=1}^{m} x_i^{d_{ij}} \] where the matrix with entries $d_{ij}$ is the product of the matrices with entries $a_{ij}$, $c_{ij}$ and $b_{ij}$, and where $v_i=\hat{g}^{*}(\prod_j u^{c_{ij}})$ are units in $\hat{{\cal{O}}}_{X_{{\sigma}},s}$. The matrix $(d_{ij})$ is equivalent to a matrix in the form \[ \begin{pmatrix} D \\ 0\end{pmatrix} \] where D is an $l\times l$ diagonal matrix with diagonal entries $d_1,\ldots,d_l \in {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}$. Hence we can change the coordinate functions $x_i$ and $z_i$ of the respective tori so that $\hat{g}^{*}\circ\alpha\circ\hat{h}^{*}$ maps \[ z_i \mapsto v_i x_i^{d_i}\] If we now set \[ \tilde{x_i} = v_i^{\frac{1}{d_i}} x_i \] then in these new coordinates the composition $\hat{g}^{*}\circ\alpha\circ\hat{h}^{*}$ is induced by a toric morphism defined by \[ z_i \mapsto x_i^{d_i} \] and thus $f\circ e:X\rightarrow} \newcommand{\dar}{\downarrow Z$ is toroidal. \qed Given a toroidal embedding $U_X\subset X$ with polyhedral complex $\Delta_X$, and a subdivision $\Delta'_X$ of $\Delta_X$, one constructs (see \cite{te}) a new toroidal embedding $U_{X'}\subset X'$ with polyhedral complex $\Delta'_X$, and a birational toroidal morphism $f': X'\rightarrow} \newcommand{\dar}{\downarrow X$ such that the induced map of the polyhedral complexes $ \Delta'_X \rightarrow} \newcommand{\dar}{\downarrow \Delta_X$ is the given subdivision. If $Y$ is a stratum in $X$ corresponding to the cone $\sigma^Y \in \Delta_X$, and if $\sigma'\subset \sigma$ is a cone in the subdivision, define \[ V_{\sigma'} = {\operatorname{Spec\ }}_{{\operatorname{Star}}(Y)} \sum_{D\in {\sigma'}^{\vee} \cap M^Y} {\cal O}} \def\nor{{\rm nor}} \def\question{{\bf (?)}_{{\operatorname{Star}}(Y)}(-D), \] where the sum is taken inside the field of rational functions of ${\operatorname{Star}}(Y)$. Then $X'$ is formed by gluing together the open sets $V_{\sigma'}$. A subdivision $\Delta_X'$ of $\Delta_X$ is called {\bf projective} if there exists a continuous function $\psi:|\Delta_X|\rightarrow} \newcommand{\dar}{\downarrow {\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$, taking rational values on $\sigma\cap N_\sigma$, which is convex and piecewise linear on each cone $\sigma\in\Delta_X$, and the largest pieces in $\sigma$ where $\psi$ is linear are the cones of the subdivision. Such $\psi$ is called a {\bf good function} (or {\em lifting function}, or {\em order function}), for the subdivision $\Delta_X'$ of $\Delta_X$. A projective subdivision corresponds to a projective modification $X'\rightarrow} \newcommand{\dar}{\downarrow X$. \begin{lem}\label{lem-lifting} Let $f:X\to B$ be a toroidal morphism, $f_\Delta:\Delta_X\to \Delta_B$ the associated morphism of polyhedral complexes. Let $X'\to X$ and $B'\to B$ be toroidal modifications, with associated subdivisions $\Delta_{X'}$ and $\Delta_{B'}$. Then $f$ lifts to a morphism $f': X' \to B'$ if and only if for each cone $\sigma' \in \Delta_{X'}$, there exists a cone $\tau' \in \Delta_{B'}$ such that $f_\Delta(\sigma') \subset \tau'$. \end{lem} {\bf Proof.} Let $\sigma'\subset \sigma^Y \in \Delta_X$ and $\tau'\subset\tau^Z\in\Delta_B$ be cones in the subdivisions such that $f_\Delta(\sigma') \subset \tau'$. The homomorphism ${\cal{O}}_{{\operatorname{Star}}(Z)} \rightarrow} \newcommand{\dar}{\downarrow {\cal{O}}_{{\operatorname{Star}}(Y)}$ extends to \[ \sum_{D\in {\tau'}^{\vee} \cap M^Z} {\cal O}} \def\nor{{\rm nor}} \def\question{{\bf (?)}_{{\operatorname{Star}}(Z)}(-D) \rightarrow} \newcommand{\dar}{\downarrow \sum_{E\in {\sigma'}^{\vee} \cap M^Y} {\cal O}} \def\nor{{\rm nor}} \def\question{{\bf (?)}_{{\operatorname{Star}}(Y)}(-E) \] because $f^{\vee}_{\Delta} ({\tau'}^\vee) \subset {\sigma'}^{\vee}$. This shows that the rational map $f'$ is a morphism on each $V_{\sigma'}$. Since $V_{\sigma'}$ cover $X'$, $f'$ is a morphism. Conversely, suppose $f'$ is a morphism. Let $x'\in X', f'(x')=b'\in B'$. Since the three morphisms $X'\rightarrow} \newcommand{\dar}{\downarrow X$, $X\rightarrow} \newcommand{\dar}{\downarrow B$ and $B'\rightarrow} \newcommand{\dar}{\downarrow B$ are toroidal, we have a toric morphism $g: X_{\sigma'}\rightarrow} \newcommand{\dar}{\downarrow X_{\tau'}$ where $(X_{\sigma'},s')$ and $(X_{\tau'},t')$ are local models at $x'$ and $b'$. Thus $g_{\Delta}(\sigma') \subset \tau'$. But $g_{\Delta}|_{\sigma'} = f_{\Delta}|_{\sigma'}$, hence $f_{\Delta}(\sigma') \subset \tau'$. Since this is true for $x'$ in any stratum of $X$, we get that for any cone $\sigma' \in \Delta_{X'}$, $f_\Delta(\sigma') \subset \tau'$ for some $\tau' \in \Delta_{B'}$. \qed \section{Toroidal reduction}\label{toroidal-reduction} \subsection{Statement of result} The purpose of this section is to modify any family of varieties into a toroidal morphism. \begin{th}\label{th-toroidal-reduction} Let $f:X\to B$ be a projective, surjective morphism with geometrically integral generic fiber, and assume $B$ integral. Let $Z\subset X$ be a proper closed subscheme. There exist a diagram as follows: $$\begin{array}{lclcl} U_X & \subset & X' &\stackrel{m_X}{\to} &X \\ \dar & & \dar f' & & \dar f\\ U_B & \subset & B' & \stackrel{m_B}{\to} &B \end{array} $$ such that $m_B$ and $m_X$ are modifications, $X'$ and $B'$ are nonsingular, the inclusions on the left are toroidal embeddings, and such that \begin{enumerate} \item $f'$ is toroidal. \item Let $Z' = m_X^{-1}Z$. Then $Z'$ is a strict normal crossings divisor, and $Z'\subset X'\setmin U_{X'}$. \end{enumerate} \end{th} \subsection{To begin the proof,} we proceed by induction on the relative dimension of $f$. If the relative dimension of $f$ is 0, let $m_X:X'\to X$ be a resolution of singularities such that $Z'=m_X^{-1}Z$ is a strict normal crossings divisor, let $B'=X'$ and $m_B = f\circ m_X$, and $f'=id$ the identity. Assume we have proven the result for morphisms of relative dimension $n-1$, and consider the case $\operatorname{rel.dim } f = n$. \subsection{Preliminary reduction steps} First, we may replace $X$ by its normalization, therefore we may assume $X$ normal, and by blowing up $Z$ in $X$ we may assume $Z$ a Cartier divisor. Let $\eta\in B$ be the generic point of $B$. By the projectivity assumption we have $X\subset {\Bbb{P}}} \newcommand{\bfc}{{\Bbb{C}}^N_B$ for some $N$. Choosing a generic projection ${\Bbb{P}}} \newcommand{\bfc}{{\Bbb{C}}^N_\eta \das {\Bbb{P}}} \newcommand{\bfc}{{\Bbb{C}}^{n-1}_\eta$ we get a rational map $X_\eta \das {\Bbb{P}}} \newcommand{\bfc}{{\Bbb{C}}^{n-1}_\eta$. Replacing $X$ by the closure of the graph of this map, we may assume that we have a morphism $g:X \to {\Bbb{P}}} \newcommand{\bfc}{{\Bbb{C}}^{n-1}_B = P$. \subsection{Semistable reduction of a family of curves} By \cite{dj2}, Theorem 2.4, we have a diagram as follows: $$\begin{array}{lcl} X_1 & \stackrel{\alpha}{\rightarrow} \newcommand{\dar}{\downarrow} & X \\ \dar g_1 & & \dar g \\ P_1 & \stackrel{a}{\rightarrow} \newcommand{\dar}{\downarrow} & P \\ & & \dar \\ & & B \end{array} $$ and a finite group $G\subset {\operatorname{Aut}}_PP_1$, with the following properties: \begin{enumerate} \item The morphism $a:P_1\to P$ is a Galois alteration with Galois group $G$. \item The action of $G$ lifts to ${\operatorname{Aut}}_XX_1$, and $\alpha:X_1\to X$ is a Galois alteration with Galois group $G$. \item There are $n$ disjoint sections $\sigma_i:P_1\rightarrow} \newcommand{\dar}{\downarrow X_1$ such that the strict altered transform $Z_1\subset X_1$ of $Z$ is the union of their images, and $G$ permutes the sections $\sigma_i$. \item The morphism $g_1: X_1 \to P_1$ is a nodal family of curves, and $\sigma_i(P_1)$ is disjoint from $\operatorname{Sing}g_1$. \end{enumerate} We may replace $X$, $P$ and $Z$ by $X_1/G$ and $P_1/G$, and $\alpha^{-1}Z/G$. Note that $\alpha^{-1}Z/G$ is not necessarily equal to the union of the images of $\sigma_i$, but the complement lies over a proper closed subset in $P_1$. \subsection{Using the inductive hypothesis} Let $\Delta\subset P$ be the union of the loci over which $Z,P_1$ or $X_1$ are not smooth. We apply the inductive assumption to $\Delta\subset P \to B$, and obtain a diagram as follows: $$\begin{array}{lclcl} U_P & \hookrightarrow & P' &\stackrel{m}{\to} &P \\ \dar & & \dar & & \dar\\ U_B & \hookrightarrow & B' &\to &B \end{array} $$ Such that $P', B'$ are nonsingular, $P' \to P$ and $B' \to B$ are modifications, the left square is a toroidal morphism, and $m^{-1}\Delta$ is a divisor of strict normal crossings contained in $P' \setmin U_P$. We may again replace $P, B$ by $P', B'$, and further we may replace $X,X_1, P_1,Z$ and $ \sigma_i$ by their pullback to $P'$. In particular $P\to B$ has a toroidal structure, and $P_1\to P$ is unramified over $U_P$. By Abhyankar's lemma, since $P_1$ is normal, it inherits a toroidal structure given by $U_{P_1} = m^{-1}U_P$ as well, so that $P_1\to P$ is a toroidal finite morphism. \subsection{Conclusion of proof} Now $X_1\to P_1$ is a nodal family which is smooth over $U_P$, therefore it as well inherits a toroidal structure $U_{X_1}\subset U_X$, where $U_{X_1} = ({g_1}^{-1} U_{P_1}) \setmin (\cup\sigma_i(P_1))$; e.g. local equations around a node are of the form $ uv = f(t)$, where $f(t)$ is a monomial on $P_1$. Notice that $\alpha^{-1} Z $ is a divisor contained in $U_{X_1}$ (see \cite{aj}, 1.3). In this situation we can apply the procedure of \cite{aj}, section 1.4 to make the group $G$ act toroidally on $X_1$: first we blow up the scheme $\operatorname{Sing} g_1$ to separate the branches of the nodes. Then we are in the situation of Proposition 1.8 of \cite{aj}, namely there is a canonical $G$-equivariant blowup $d:\tilde{X_1} \to X_1$ such that $G$ acts strictly toroidally on $b^{-1}U_{X_1}\subset\tilde{X_1}$. Let ${X'} =\tilde{X_1}/G$, then $ {X'}\to B $ inherits a toroidal structure and ${X'}\to X$ is birational; moreover, ${Z'}\subset {X'}$ is a divisor contained in ${X'}\setmin U_{{X'}}$. Applying toroidal resolution of singularities, the induction step is proven. \qed \section{Removing horizontal divisors}\label{remove-horizontal} We may now replace $X\to B$ by $X'\to B'$, and thus we may assume that the morphism $f$ is toroidal. Our goal in this section is to arrive at a situation where $f^{-1} U_B = U_X$. The rough idea is, that a morphism between nonsingular toroidal embeddings $f:X\to B$ is locally given by monomials $t_i = x_1^{k_1} \cdots x_r^{k_r}$, in which the variables defining horizontal divisors cannot appear, so these divisors are unnecessary in the toroidal description. We make this precise by a simple translation argument. \begin{prp} Let $U_X\subset X$ and $U_B \subset B$ be nonsingular toroidal embeddings and $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ a surjective toroidal morphism. Then, denoting $U_X'= f^{-1}(U_B) \supset U_X$, we have that $U_X'\subset X$ is a toroidal embedding, and $f: (U_X'\subset X)\rightarrow} \newcommand{\dar}{\downarrow (U_B\subset B)$ is a toroidal morphism. \end{prp} {\bf Proof.} Since $f$ maps $U_X$ into $U_B$, $f^{-1}(B\setmin U_B)$ as a set is a union of divisors supported in $X\setmin U_X$. In local models these divisors are all T-invariant. Consider local models $(X_{\sigma},s)$ at $x$, $(X_{\tau},t)$ at $f(x)$, and the toric morphism $g:X_\sigma\rightarrow} \newcommand{\dar}{\downarrow X_\tau$. We may assume that $v_1,\ldots,v_n$ is a basis of $N_\sigma$ and $\sigma$ is generated by $v_1,\ldots,v_k$. Then $X_\sigma \cong {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^k\times\bfg_m^{n-k}$, and we may take $s=(0,1)$. Let the closures of the orbits corresponding to $v_1,\ldots,v_j$ be the horizontal divisors. That means, $g_\Delta(v_i)=0$ for $i=1,\ldots,j$, and $g$ factors through the projection: \[ g: X_\sigma \cong {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^j\times{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^{k-j}\times\bfg^{n-k}_m \rightarrow} \newcommand{\dar}{\downarrow {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^{k-j}\times\bfg^{n-k}_m \to X_\tau \] Now take $s'=(1,0,1)\subset {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^j\times{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^{k-j}\times{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^{n-k}$. From the factorization of $g$ we see that $g(s')=g(s)$. Translation by $(1,0,0)$ gives an isomorphism of the local rings at $s$ and $s'$ so that the ideals of the T-invariant divisors passing trough $s'$ corresponds to the ideal of the vertical T-invariant divisors passing through $s$. Thus, $(X_\sigma,s')$ is a local model for the embedding $U_X'\subset X$ at $x\in X$ and $g:(X_\sigma,s')\rightarrow} \newcommand{\dar}{\downarrow(X_\tau,t)$ is the toric morphism of the local models representing $f$. \qed \section{Making the morphism equidimensional}\label{equidimensional} The goal of this section is to perform modifications on $B$ and $X$, after which the morphism becomes equidimensional. First a lemma which characterizes equidimensional toroidal morphisms: \begin{lem} Let $f:X\to B$ be a surjective toroidal morphism, $f_\Delta: \Delta_X\to\Delta_B$ the associated morphism of polyhedral complexes. Then $f$ is equidimensional if and only if for any cone $\sigma \in \Delta_X$, we have $f_\Delta(\sigma) \in \Delta_B$. That is, the image of a cone of $\Delta_X$ is a cone of $\Delta_B$. \end{lem} {\bf Proof.} Computing the dimension of a local ring commutes with taking the completion. Thus, it suffices to consider local models. The generic fiber of a toric morphism $f: X_\sigma \rightarrow} \newcommand{\dar}{\downarrow X_\tau$ has dimension ${\operatorname{Rank}}(N_\sigma) - {\operatorname{Rank}}(N_\tau)$. Now $f$ maps a $k$-dimensional orbit corresponding to a $({\operatorname{Rank}}(N_\sigma)-k)$-dimensional face $\sigma'$ of $\sigma$ onto some $l$-dimensional orbit corresponding to a $({\operatorname{Rank}}(N_\tau)-l)$-dimensional face $\tau'$ of $\tau$. Thus $f$ is equidimensional if and only if \[ ({\operatorname{Rank}}(N_\sigma)-k) - ({\operatorname{Rank}}(N_\tau)-l) \leq {\operatorname{Rank}}(N_\sigma) - {\operatorname{Rank}}(N_\tau), \] that means $l \leq k$. Every $k$-dimensional cone maps to an $l$-dimensional cone for $l\leq k$ if and only if the image of every cone is a cone. \qed \begin{rem} In case $\tau$ is simplicial the condition of the theorem is equivalent to the statement that all 1-dimensional faces of $\sigma$ map to 0 or 1-dimensional faces of $\tau$. \end{rem} The following lemma is a slight generalization of the toric Chow's lemma (\cite{danilov} 6.9.2 page 119). \begin{lem} Given a polyhedral complex $\Delta$ and a subdivision $\Delta'$ of $\Delta$, there exists a projective subdivision $\Delta''$ of $\Delta$ which refines $\Delta'$. \end{lem} {\bf Proof.} First, we show that it suffices to find a ``good'' function $\psi$ (see Section~\ref{toroidal-morphisms}) on each cone $\sigma\in\Delta$. Indeed, if for each $\sigma\in\Delta$ we have found a good function $\psi_\sigma: \sigma\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$, then a good function on $|\Delta|$ is the sum \[ \psi = \sum_{\sigma\in\Delta} \bar{\psi}_\sigma \] where $\bar{\psi}_\sigma: |\Delta|\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ is a good extension of $\psi_\sigma$ constructed as follows. To extend $\psi_\sigma$ to a cone $\tau\in\Delta$ we proceed by induction on the dimension of $\tau$ ({\em cf.} \cite{te} Lemma~1, page~33). If $\dim \tau = 1$ and $\tau$ is not a face of $\sigma$, define $\bar{\psi}_\sigma|_\tau\equiv 0$. Now assume that $\dim {\tau} > 1$ and $\bar{\psi}_\sigma$ is defined on $\partial\tau$. Choose a point $x$ in the relative interior of $\tau$ and define \[ \bar{\psi}_\sigma (\lambda x + (1-\lambda) y) = \lambda C + (1-\lambda) \bar{\psi}_\sigma(y), \hspace{0.3in} y\in\partial\tau, 0\leq\lambda\leq 1.\] For big $C\in{\Bbb{Q}}} \newcommand{\bff}{{\Bbb{F}}$ the extension $\bar{\psi}_\sigma|_\tau$ is convex. Now let $\sigma\in\Delta$ with $\dim {\sigma}=n$. For every $n-1$ dimensional cone $\tau\in\Delta'$ in the subdivision of $\sigma$ choose a linear rational function $l_\tau: \sigma\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ such that $\tau$ is the subset of $\sigma$ defined by $l_\tau=0$. Then the sum \[ \psi_\sigma(x) = -\sum_{\tau}|l_\tau(x)| \] is a good function on $\sigma$. \qed \begin{prp} Let $U_X\subset X$ and $U_B \subset B$ be toroidal embeddings with polyhedral complexes $\Delta_X$ and $\Delta_B$ respectively, and assume that $B$ is nonsingular. Let $ f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective toroidal morphism. Then there exist projective subdivisions $\Delta_X'$ of $\Delta_X$ and $\Delta_B'$ of $\Delta_B$ with $\Delta_B'$ nonsingular, such that the induced map $f':X'\rightarrow} \newcommand{\dar}{\downarrow B'$ is an equidimensional toroidal morphism. If $f^{-1}(U_B)=U_X$ then $(f')^{-1}(U_{B'})=U_{X'}$. \end{prp} {\bf Proof.} There exists a subdivision of $\Delta_B$ ``induced'' by $\Delta_X$. For $x\in\tau\in\Delta_B$ let $S_x$ be the set of cones $\sigma\in\Delta_X$ such that $\sigma\cap f_\Delta^{-1}(x)\neq\{0\}$. Since $f_\Delta$ is surjective, $S_x\neq \emptyset$. The relation $x\sim y \Leftrightarrow S_x = S_y$ for $x,y\in\tau$ is clearly an equivalence, hence it defines a partition of $\Delta_B$. The equivalence class of $x$ is \[ \bigcap_{\sigma\in S_x} f_\Delta(\sigma) \] which is a convex rational polyhedral subcone of $\tau$. Thus the partition defines a subdivision $\Delta_B^0$ of $\Delta_B$ such that $f_\Delta(\sigma)$ for any cone $\sigma\in\Delta_X$ is a union of cones in $\Delta_B^0$. By the previous lemma there exists a refinement $\Delta_B^1$ of $\Delta_B^0$ which is a projective subdivision of $\Delta_B$. Finally, we let $\Delta_B'$ be a nonsingular projective subdivision of $\Delta_B^1$. For $f_\Delta$ to map cones of $\Delta_{X}$ to cones of $\Delta_B'$, the complex $\Delta_{X}$ has to be subdivided. Since the subdivision $\Delta_B'$ of $\Delta_{B}$ is projective, there exists a good function $\psi$ on $|\Delta_B'|=|\Delta_B|$. The piecewise linear function $\psi\circ f_\Delta$ defines a projective subdivision $\Delta_X'$ of $\Delta_{X}$ whose cones map into cones of $\Delta_B'$. If $\sigma \in \Delta_X'$ then $f_\Delta(\sigma)$ is a union of faces of some cone $\tau\in\Delta_B$. Since $f_\Delta$ is linear on $\sigma$, $f_\Delta(\sigma)$ is convex in $\tau$, hence $f_\Delta(\sigma)$ is a face of $\tau$. If we assume from the beginning that $f^{-1}(U_B)=U_X$, that means $f_\Delta^{-1}(0)\cap |\Delta_X| = 0$, then clearly the same is true for any subdivision $|\Delta_X'| = |\Delta_X|$, hence $(f')^{-1}(U_{B'})=U_{X'}$. \qed \section{Kawamata's trick and reduced fibers}\label{reduced-fibers} \subsection{Statement of result} The goal in this section is to find a finite base change, after which all the fibers in the resulting morphism are reduced. The base change we perform will not necessarily be toroidal, but the morphism after base change will still be toroidal. \begin{prp}\label{prop-red} Let $U_X\subset X$ and $U_B \subset B$ be projective toroidal embeddings, and assume that $B$ is nonsingular. Let $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective equidimensional toroidal morphism with $f^{-1}(U_B)=U_X$. Then there exists a finite surjective morphism $p: B'\rightarrow} \newcommand{\dar}{\downarrow B$ so that, denoting by $X'$ the normalization of $X\times_B B'$, we have that $B'$ and $X'$ are toroidal embeddings, the projection $f': X'\rightarrow} \newcommand{\dar}{\downarrow B'$ is an equidimensional toroidal morphism with reduced fibers, and $(f')^{-1}(U_{B'})=U_{X'}$. \end{prp} The construction of $X'$ and $B'$ is more explicitly given in Proposition~\ref{prop-compl}, where the polyhedral complexes of $X'$ and $B'$ are also described. \subsection{The toric pictures} To start, we characterize equidimensional toroidal morphisms with reduced fibers in terms of polyhedral complexes. \begin{lem} Let $f:X\to B$ be an equidimensional toroidal morphism, $f_\Delta:\Delta_X\to \Delta_B$ the associated morphism of polyhedral complexes. Then $f$ has reduced fibers if and only if for any cone $\sigma \in \Delta_X$, with image $\tau\in \Delta_B$, we have $f_\Delta(N_\sigma\cap\sigma) = N_\tau\cap\tau$. That is, the image of the lattice in any cone of $X$ is the lattice in the image cone. \end{lem} {\bf Proof.} It suffices to consider the toric morphism of local models $f: X_\sigma \rightarrow} \newcommand{\dar}{\downarrow X_\tau$ and the fiber over a point $t\in X_\tau$ lying in the closed orbit of $X_\tau$. If the orbit of $t$ is not $\{t\}$ then $X_\tau$ is a product $X_\tau = X_{\tau'}\times \bfg_m^q$ for some $q>0$. Without loss of generality we may then replace $X_\tau$ by $X_{\tau'}$, and replace $f$ by $p\circ f$ where $p: X_\tau\rightarrow} \newcommand{\dar}{\downarrow X_{\tau'}$ is the projection. Indeed, the fiber $(p\circ f)^{-1}(p(t))$ is isomorphic to $f^{-1}(t)\times \bfg_m^q$, and $p_\Delta$ gives an isomorphism $p_\Delta: N_\tau\cap\tau \cong N_{\tau'}\cap\tau'$. Thus we may assume that $\{t\}$ is the unique closed orbit of $X_\tau$. The ideal of $f^{-1}(t)$ is generated by $k[f_\Delta^{\vee}(\tau^\vee\cap(M_\tau\setmin \{0\}))] \subset k[\sigma^\vee\cap M_\sigma]$. The fiber is reduced if and only if the image $f_\Delta^{\vee}(\tau^\vee\cap(M_\tau\setmin \{0\}))$ is saturated in $\sigma^\vee\cap M_\sigma$. This happens if and only if $f_\Delta(N_\sigma\cap\sigma) = N_\tau\cap\tau$. \qed When $X_\tau$ is nonsingular, the condition of the lemma is that primitive points of $\sigma$ map to primitive points of $\tau$. \begin{lem}\label{lem-tor-prod} Let $X_\sigma\rightarrow} \newcommand{\dar}{\downarrow X_\tau$ be a toric morphism with $X_\tau$ nonsingular. Let $X_{\tau'}$ be a toric variety given by $\tau'=\tau$ and $N_{\tau'}\subset N_\tau$ a sublattice of finite index. Then every irreducible component of the normalization of $X_\sigma \times_{X_\tau} X_{\tau'}$ is a toric variety $X_{\sigma'}$ given by the cone $\sigma'=\sigma$ and integral lattice $N_{\sigma'} = N_\sigma \cap f_\Delta^{-1}(N_{\tau'})$. \end{lem} {\bf Proof.} The ring of regular functions of $X_{\tau'} \times_{X_\tau} X_\sigma$ is \[ {\cal{O}}_{X_{\tau'}}\otimes_{{\cal{O}}_{X_\tau}} {\cal{O}}_{X_\sigma} = k[(\tau')^{\vee}\cap M_{\tau'}] \otimes_{k[\tau^{\vee}\cap M_\tau]} k[\sigma^{\vee}\cap M_\sigma] = k[\pi] \] where $\pi$ is the pushout of $j: \tau^{\vee}\cap M_\tau \to (\tau')^{\vee}\cap M_{\tau'} $ and $f^\vee_\Delta: \tau^{\vee}\cap M_\tau \to \sigma^{\vee}\cap M_\sigma$: \[ \pi = ((\tau')^{\vee}\cap M_{\tau'}) \times (\sigma^{\vee}\cap M_\sigma) /\sim.\] Here $\sim$ is the equivalence relation generated by: $(v_1,w_1)\sim (v_2,w_2)$ whenever there exists $u\in \tau^{\vee}\cap M_\tau$ such that $ (v_1,w_1) = (v_2,w_2) \pm (u,-f^\vee_\Delta(u)). $ Let $M$ be the abelian group \[ M = M_{\tau'}\times M_\sigma /((u,-f^\vee_\Delta(u))) | u\in \tau^{\vee}\cap M_\tau). \] We will show that the semigroup homomorphism $\iota: \pi \rightarrow} \newcommand{\dar}{\downarrow M$ is injective. Suppose that $\iota(v_1,w_1) = \iota(v_2,w_2)$, where $v_1,v_2 \in (\tau')^{\vee}\cap M_{\tau'} $ and $w_1,w_2 \in \sigma^{\vee}\cap M_\sigma$. Say $ (v_1,w_1) = (v_2,w_2) + (u,-f^\vee_\Delta(u))$ for some $u\in M_\tau$. Let $t_1,\ldots,t_m \in \tau^\vee$ be a basis of $M_\tau$, so that in this basis $\tau^\vee$ is given by the inequalities $t_i\geq 0$, for $i=1,\ldots,k$. Write $u=\sum_i \alpha_it_1$ with $\alpha_i\in {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}$. Collecting positive and negative terms, we get $u = u_1-u_2$ with $u_1,u_2\in \tau^{\vee}\cap M_\tau.$ Writing $v_1,v_2$ in terms of $t_i$ and expanding the equation $v_1=v_2+u_1-u_2$, we see that $v_2-u_2 \in \tau^{\vee}\cap M_{\tau'}$. Let $(v_3,w_3) = (v_2-u_2,w_2+f^\vee_\Delta(u_2))$. Clearly $$(v_1,w_1)\sim (v_3,w_3)\sim (v_2,w_2)$$ and $\iota$ is injective. The integral closure of $k[\pi]$ in $k[M]$ is the semigroup algebra $k[\tilde{\pi}]$ where $\tilde{\pi}$ is the saturation of $\pi$ in $M$. Write $M=F\oplus T$ where $F$ is the free part and $T$ is torsion. Then $T\subset\tilde{\pi}$ because $m T=0$ for some $m>0$. For any $f+t \in \tilde{\pi}$ where $f\in F$ and $t\in T$, we have $-t\in T\subset \tilde{\pi}$, hence $f\in \tilde{\pi}$. Thus \[ \tilde{\pi} \cong (\tilde{\pi}\cap F) \oplus T.\] If $t_i$ are generators of $T$ of order $m_i$, and if $x_i$ is the image of $t_i$ in $k[\tilde{\pi}]$, then \[ k[\tilde{\pi}] \cong k[\tilde{\pi}\cap F][\ldots,x_i,\ldots] / (x_i ^{m_i}=1) \] Thus the normalization of ${\operatorname{Spec\ }} {k[\tilde{\pi}]}$ has $|T|$ components with $x_i = \zeta_i$ where $\zeta_i$ are $m_i$'th roots of unity, each component isomorphic to ${\operatorname{Spec\ }} {k[\tilde{\pi}\cap F]}$. Next we show that $F$ can be embedded in $M_\sigma\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ so that $M_\sigma\subset F$. The image of the homomorphism $\phi: M\rightarrow} \newcommand{\dar}{\downarrow M_\sigma\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ defined by $\phi(v,w)=f_\Delta^\vee(v)+w$ for $v\in M_{\tau'}, w\in M_\sigma$ contains $M_\sigma$. Since $M$ and $M_\sigma$ have the same rank, $\phi$ embeds $F$ in $M_\sigma\otimes{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ as a lattice of full rank containing $M_\sigma$. As $f_\Delta^\vee$ takes $\tau^\vee$ into $\sigma^\vee$, we see that $\phi$ maps $\tilde{\pi}\cap F$ into $\sigma^\vee$ so that the image contains $\sigma^{\vee}\cap M_\sigma$. Therefore, ${\operatorname{Spec\ }} {k[\tilde{\pi}\cap F]}$ is a toric variety defined by the cone $\sigma'=\sigma$ and integral lattice $N_{\sigma'} = F^\vee$. To determine $F^{\vee}$, first, we have $M_\sigma\subset F$, hence $F^{\vee}\subset N_\sigma$; second, $f_\Delta^{\vee}(M_{\tau'})\subset F$ implies that $f_\Delta(F^{\vee})\subset N_{\tau'}$. Conversely, $N_\sigma\cap f_\Delta^{-1}(N_{\tau'}) \subset F^{\vee}$, so the two are equal. \qed To get a toric morphism with reduced fibers, one can take a base change $X_{\tau'} \rightarrow} \newcommand{\dar}{\downarrow X_\tau$, where $\tau'=\tau$ and $N_{\tau'}\subset N_{\tau}$ is a sublattice of finite index. By Lemma~\ref{lem-tor-prod} every component $X_{\sigma'}$ of the normalization of $X_\sigma \times_{X_\tau} X_{\tau'}$ is a toric variety defined by the cone $\sigma' = \sigma$ and integral lattice $N_{\sigma'} = N_\sigma \cap f_\Delta^{-1}(N_{\tau'})$. By a judicious choice of $N_{\tau'}$ the fibers of $X_{\sigma'}\rightarrow} \newcommand{\dar}{\downarrow X_{\tau'}$ are reduced. \subsection{Kawamata's covering} To perform a similar base change in the toroidal case we need a toroidal morphism $B'\rightarrow} \newcommand{\dar}{\downarrow B$ which ramifies over a divisor $D$ with a certain index. \begin{dfn} \label{def-cyclic-cover} Let $L$ be a Cartier divisor on $B$, and let $D\in|m L|$. Choose a rational section $s_L$ of $ {\cal{O}}_B(L)$ defining $L$ and a regular section $s_D$ of ${\cal{O}}_B(mL)$ defining $D$. Consider the rational function $\phi = s_D/s_L^m$ on $B$. Then the field $K(B)(\sqrt[m]{\phi})$ depends only on $D$ and ${\cal{O}}(L)$. The normalization of $B$ in $K(B)(\sqrt[m]{\phi})$ is called {\bf the cyclic cover ramified along $D$ with index $m$. } \end{dfn} \begin{rem} Another way to define the cyclic cover, is as the normalization of $${\cal S}pec_B{\operatorname{Sym}}^\bullet\bigl({\cal{O}}_B(L)^\vee \bigr)/(\{f- s(f)\}_f) $$ where we view $s$ as a morphism $s:{\cal{O}}_B(kL)\to{\cal{O}}_B((k+m)L)$. \end{rem} When $B$ is nonsingular and $D$ a divisor of normal crossings so that $s_D=x_1\cdots x_l$ for some local parameters $x_1,\ldots,x_l$, then the cyclic cover has a local equation \[ z^m = x_1\cdots x_l. \] It is nonsingular if and only if $l=1$. Let $U_B\subset B$ be a nonsingular projective (strict) toroidal embedding. Then $B\setmin U_B = \sum D_i$ is a strict divisor of normal crossings. Consider the data $(D_i, m_i)$ where $D_i$ are the irreducible components of $B\setmin U_B$ and $m_i$ are positive integers, $i=1,\ldots,m$. A {\it Kawamata covering package} consists of $(D_i, m_i, H_{ij})$ with the following properties: \begin{enumerate} \item $H_{ij}$ are effective reduced nonsingular divisors on $B$, for $i=1,\ldots,m$, $j=1,\ldots,\dim B$. \item $\sum_i D_i + \sum_{i,j} H_{ij}$ is a reduced divisor of normal crossings (in particular, $H_{ij}$ are distinct). \item $D_i+ H_{ij} \in |m_i L_i|$ for some Cartier divisor $L_i$ for all $i,j$. \end{enumerate} To find $H_{ij}$, let $M$ be an ample divisor. Take a multiple of $M$ if necessary so that $m_i M-D_i$ is very ample for all $i$, and choose $H_{ij}$ general members in $|m_i M-D_i|$. Now let $B_{ij}$ be the cyclic cover ramified along $D_i+H_{ij}$ with index $m_i$; let $B'$ be the normalization of \[ B_{1,1} \times_B \ldots \times_B B_{m,{\operatorname{dim}}{B}}\] and let $p:B'\rightarrow} \newcommand{\dar}{\downarrow B$ be the projection. \begin{lem} [Kawamata] The variety $B'$ is nonsingular, ramified with index $m_i$ along $D_i+H_{ij}$. The reduction of the inverse image $p^*(\sum_i D_i + \sum_{i,j} H_{ij})_{\red}$ is a divisor of normal crossings. \end{lem} {\bf Proof.} Let $x_i$ be local equations of $D_i$, and let $y_{ij}$ be local equations of $H_{ij}$ at $b\in B$. Then $B_{ij}$ is locally given by the equation \[ z_{ij}^{m_i} = x_i y_{ij} \] It suffices to prove that the normalization of $\times_j B_{ij}$ is nonsingular for all $i$. If $b\in D_i$ then since $\bigcup_j H_{ij} \cap D_i = \emptyset$, say $b\notin H_{i,0}$, and $y_{i,0}$ is a unit in ${\cal{O}}_{B,b}$. The normalization of the product $\times_j B_{ij}$ is locally defined by the equations \begin{eqnarray*} z_{i,1}^{m_i} &=& x_i \\ (\frac{z_{i,j}}{z_{i,1}})^{m_i} &=& y_{i,j} \hspace{.5in} j=2,\ldots \dim B. \end{eqnarray*} Since $x_i, y_{i,2},\ldots,y_{i,\dim B}$ are either units or local parameters in ${\cal{O}}_{B,b}$, it follows that the normalization of the product $\times_j B_{ij}$ is nonsingular at $b$. A similar situation happens when $b\notin D_i$, since then $x_i$ is a unit in ${\cal{O}}_{B,b}$. Thus, $B'$ is nonsingular and ramified with index $m_i$ along $D_i$ and $H_{ij}$. Replacing $B$ by $D_i$ or $H_{ij}$, we get that $p^*(\sum_i D_i + \sum_{i,j} H_{ij})_{\red}$ is a divisor of normal crossings. \qed Let $\tilde{U}_B = B\setmin (\bigcup_{i,j} H_{ij}\cup D_i)$, and $\tilde{U}_{B'} = p^{-1} (\tilde{U}_B)$. Both $\tilde{U}_B \subset B$ and $ \tilde{U}_{B'} \subset B'$ are toroidal embeddings because $B, B'$ are nonsingular and the divisors $B\setmin\tilde{U}_B$, $B'\setmin\tilde{U}_{B'}$ cross normally. From the construction and local equations we also see that $p$ is a toroidal morphism with respect to this structure. If $b'\in B'$, $b=p(b')\in B$, then there exist local parameters $\{x_l\}$ at $b$ and $\{z_l\}$ at $b'$ such that $p$ is defined by $x_l = z_l^{a_l}$ where $a_l = m_i$ if the divisor defined by $x_l$ is $D_i$ or $H_{ij}$, otherwise $a_l =1$. We denote the polyhedral complexes of $\tilde{U}_B \subset B$ and $ \tilde{U}_{B'} \subset B'$ by $\tilde{\Delta}_B$ and $\tilde{\Delta}_{B'}$, respectively. So, every cone $\tau' \in \tilde{\Delta}_{B'}$ is mapped homeomorphically to a cone $\tau \in \tilde{\Delta}_{B}$ by $p_\Delta$. If $N_\tau$ has basis $u_1,\ldots,u_m$ then $N_{\tau'}$ has basis $a_1 u_1, \ldots, a_m u_m$ where $a_l$ are as above. \subsection{The toroidal picture} We need to see how adding the divisors $H_{ij}$ to the toroidal structure of $B$ affects the toroidal structure of $X$. \begin{lem} Let $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be toroidal, $B$ nonsingular, H a generic hyperplane section of $B$. Let $x\in X$, $b=f(x)\in H$. Then \begin{itemize} \item[(i)] there exist local models $(X_\sigma = X_\sigma'\times \bfg_m, z\times 1)$ at $x$ and $(X_\tau = X_\tau'\times \bfg_m, y\times 1)$ at $b$ such that $H$ corresponds to the divisor $X_\tau'\times \{1\}$ and the morphism $f$ is a product of toroidal morphisms \[ f=g \times id: X_\sigma'\times \bfg_m \rightarrow} \newcommand{\dar}{\downarrow X_\tau'\times \bfg_m \] \item[(ii)] $U_B\setmin H\subset B$, $U_X\setmin f^{-1}(H)\subset X$ are toroidal embeddings and $f$ is a toroidal morphism of these embeddings. \end{itemize} \end{lem} {\bf Proof.} Let $(X_{\sigma},s)$ and $(X_{\tau},t)$ be local models at $x\in X$ and $b=f(x)\in B$, respectively, and let $f$ also denote the toric morphism of the local models defined by \[ f_\Delta: (N_\sigma,\sigma)\rightarrow} \newcommand{\dar}{\downarrow(N_\tau,\tau) \] Clearly $U_B\setmin H\subset B$ is toroidal and we may assume that it has a local model $(X_{<\tau,v>},t')$, where $<\tau,v>$ is the cone spanned by $\tau$ and some indivisible $v\in N_{\tau}$, and where $t'$ lies in the unique closed orbit of $X_{<\tau,v>}$. Write $N_{\tau,2}$ for the saturated sublattice of $N_{\tau}$ generated by $v$, and choose a splitting \begin{eqnarray*} N_{\tau} &\cong& N_{\tau,1} \oplus N_{\tau,2} \\ X_{\tau} &\cong& X_{\tau}' \times \bfg_m \\ X_{<\tau,v>} &\cong& X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \end{eqnarray*} Since $f$ is dominant, $\tilde{N}_{\tau,2} = f_\Delta(f_\Delta^{-1}( N_{\tau,2}))$ is a sublattice of finite index of $N_{\tau,2}$. The inclusion $\tilde{N}_{\tau,2} \subset N_{\tau,2}$ corresponds to an \'etale cover of $\bfg_m$. Since the completed local rings are isomorphic, we may replace the local model by a local model in the \'etale cover and assume that $f_\Delta^{-1}(N_{\tau,2})$ surjects onto $N_{\tau,2}$. Let $N_{\sigma,1}=f_\Delta^{-1}(N_{\tau,1}) \subset N_{\sigma}$ and let $N_{\sigma,2} \subset N_{\sigma}$ be generated by some $u\in N_{\sigma}$ such that $f_\Delta(u)=v$. For any $w\in N_{\sigma}$, $f_\Delta(w) = v_1+v_2 \in N_{\tau,1} \oplus N_{\tau,2}$, there exists $m\in {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}$ so that $f_\Delta(m u)=v_2$. Hence $w-m u \in N_{\sigma,1}$, and $N_{\sigma}=N_{\sigma,1}+N_{\sigma,2}$. Since $N_{\sigma,1}\cap N_{\sigma,2} = \{0\}$ the sum is direct, \begin{eqnarray*} N_{\sigma} &\cong& N_{\sigma,1} \oplus N_{\sigma,2} \\ X_{\sigma} &\cong& X_{\sigma}' \times \bfg_m \end{eqnarray*} and $f_\Delta$ maps $N_{\sigma,1}$ to $N_{\tau,1}$ and $N_{\sigma,2}$ to $N_{\tau,2}$. Thus the toric morphism $f$ is a product $f=g\times h$ where $g:X_{\sigma}'\rightarrow} \newcommand{\dar}{\downarrow X_{\tau}'$ and $h=id:\bfg_m \rightarrow} \newcommand{\dar}{\downarrow \bfg_m$ are the toric morphisms induced by the restriction of $f_\Delta$ to $N_{\sigma,1}$ and $N_{\sigma,2}$, respectively. Since $f_\Delta$ maps $<\sigma,u>$ to $<\tau,v>$, $f$ extends to a toric morphism \[ f: X_{<\sigma,v>} \cong X_{\sigma}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 @>{g\times id}>> X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \cong X_{<\tau,v>}\] For $t=(y,1)\in X_{\tau}' \times \bfg_m \subset X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$ the complete local rings $\hat{{\cal{O}}}_{X_{\tau}' \times \bfg_m,(y,1)}$ and $\hat{{\cal{O}}}_{X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1,(y,0)}$ are isomorphic via translation. The closures of codimension 1 orbits through $(y,0)$, are those through $(y,1)$ plus $X_{\tau}' \times \{0\}$ which formally corresponds to $H$. Similarly, for $s=(z,1)\in X_{\sigma}' \times \bfg_m \subset X_{\sigma}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$ the ring $\hat{{\cal{O}}}_{X_{\sigma}' \times \bfg_m,(z,1)}$ is isomorphic to $\hat{{\cal{O}}}_{X_{\sigma}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1,(z,0)}$ via translation, and the closures of codimension 1 orbits through $(z,0)$, are those through $(z,1)$ plus $X_{\sigma}' \times \{0\} = f^{*}(X_{\tau}' \times \{0\})$ which formally corresponds to $f^{*}(H)$. Thus $(X_{<\sigma,u>},(z,0))$ and $(X_{<\tau,v>},(y,0))$ are local models at $x\in X$ and $b\in B$, respectively, and the morphism \[ f: (U_X\setmin f^{-1}(H)\subset X)\rightarrow} \newcommand{\dar}{\downarrow(U_B\setmin H\subset B)\] is toroidal. \qed It follows from the lemma that $f:(\tilde{U}_B\subset B) \rightarrow} \newcommand{\dar}{\downarrow (\tilde{U}_X\subset X)$ is toroidal, where $\tilde{U}_X = f^{-1}(\tilde{U}_B)$. By Lemma~\ref{lem-tor-prod}, the normalization $X'$ of $X\times_B B'$ is toroidal. Let $f': X'\rightarrow} \newcommand{\dar}{\downarrow B'$, $p':X'\rightarrow} \newcommand{\dar}{\downarrow X$ be the (toroidal) projections, and let $\tilde{\Delta}_X$, $\tilde{\Delta}_{X'}$ be the polyhedral complexes of $X, X'$. \[ \begin{CD} X' @>{p'}>> X\\ @V{f'}VV @VV{f}V\\ B' @>{p}>> B \end{CD} \] Then $p_\Delta'$ maps a cone $\sigma' \in \tilde{\Delta}_{X'}$ homeomorphically to a cone $\sigma \in \tilde{\Delta}_{X}$; the integral lattice of $\sigma'$ can then be identified with a sublattice $N_{\sigma'} \subset N_\sigma$. The following lemma shows that the added divisors $p^{-1}(H_{ij})$ and $(p\circ f')^{-1}(H_{ij})$ can be removed from the toroidal structures of $B'$ and $X'$ so that $f'$ (but not $p$ or $p'$) remains toroidal. \begin{lem} $p^{-1}(U_B) \subset B'$ and $(p\circ f')^{-1}(U_B) \subset X'$ are toroidal embeddings and $f'$ is a toroidal morphism of these embeddings. \end{lem} We show how to remove one irreducible divisor $H=H_{ij}$ for some $i,j$. Since the question is local, choose local models $X_{\sigma}$, $X_{\tau}$ and $X_{\tau'}$ of $X$, $B$, and $B'$ so that both $f$ and $p$ are products \[ f: X_{\sigma} \cong X_{\sigma}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 @>{g\times id}>> X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \cong X_{\tau}\] \[ p: X_{\tau'} \cong X_{\rho}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 @>{q\times r}>> X_{\tau}'\times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \cong X_{\tau} \] where all morphisms are toric and $H$ corresponds to $X_{\tau}' \times \{0\}$. Since $p$ ramifies along $H$, the morphism $r:{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$ has degree $m\geq 2$. The fiber product is then \begin{eqnarray*} X'' & = & (X_{\sigma}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1)\times_{X_{\tau}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1} (X_{\tau'}' \times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1) \\ & \cong & (X_{\sigma}' \times_{X_{\tau}'} X_{\tau'}')\times ({\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \times_{{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1} {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1) \\ & \cong & (X_{\sigma}' \times_{X_{\tau}'} X_{\tau'}')\times {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1 \end{eqnarray*} and the projection $X''\rightarrow} \newcommand{\dar}{\downarrow X_{\tau'}$ is the product of the projection $X_{\sigma}' \times_{X_{\tau}'} X_{\tau'}' \rightarrow} \newcommand{\dar}{\downarrow X_{\tau'}'$ and $id: {\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$. The divisor $(p\circ f')^{*}(H)$ in $X''$ is $X_{\sigma}' \times_{X_{\tau}'} X_{\tau'}' \times \{0\}$ and by the same translation argument as above we can remove $p^{*}(H)$ and $(p\circ f')^{*}(H)$ from the toroidal structures of $B'$ and $X''$, respectively. \qed Let $\Delta_{X'}, \Delta_{B'}$ be the polyhedral complexes of the embeddings $p^{-1}(U_B) \subset B'$ and $(p\circ f')^{-1}(U_B) \subset X'$. Removing the divisors $p^{-1}(H_{ij})$ and $(p\circ f')^{-1}(H_{ij})$ means removing the corresponding edges (and everything attached to them) from the polyhedral complexes $\tilde{\Delta}_{B'}$ and $\tilde{\Delta}_{X'}$. As $p_\Delta$ and $p_\Delta'$ map cones of $\tilde{\Delta}_{B'}$ and $\tilde{\Delta}_{X'}$ homeomorphically to cones of $\tilde{\Delta}_{B}$ and $\tilde{\Delta}_{X}$, restrictions of $p_\Delta$ and $p_\Delta'$ map cones of ${\Delta}_{B'}$ and ${\Delta}_{X'}$ homeomorphically to cones of ${\Delta}_{B}$ and ${\Delta}_{X}$. We summarize the previous constructions in the following proposition. \begin{prp}\label{prop-compl} With the assumptions of Proposition~\ref{prop-red}, let $u_i$ be the primitive points of $\Delta_B$, and let $m_i>0$ for $i=1,\ldots,l$. There exists a finite covering $p:B'\rightarrow} \newcommand{\dar}{\downarrow B$ so that, if $X'$ is the normalization of $X\times_B B'$ and $f':X'\rightarrow} \newcommand{\dar}{\downarrow B'$, $p':X'\rightarrow} \newcommand{\dar}{\downarrow X$ the two projections, we have \begin{itemize} \item[(i)] $U_{X'}\subset X'$ and $U_{B'}\subset B'$ are toroidal embeddings with polyhedral complexes $\Delta_{X'}$ and $\Delta_{B'}$; moreover, $f'$ is a toroidal morphism of these embeddings. \item[(ii)] There exist morphisms $$p_\Delta:\Delta_{B'}\rightarrow} \newcommand{\dar}{\downarrow \Delta_{B}$$ and $$p_\Delta': \Delta_{X'}\rightarrow} \newcommand{\dar}{\downarrow \Delta_{X},$$ such that \begin{list}{$\circ$}{} \item for any cone $\tau'\in \Delta_{B'}$, the morphism $p_\Delta$ maps $\tau'$ isomorphically to $\tau$, and identifies $N_{\tau'}$ with the sublattice of $N_{\tau}$ generated by $m_i u_i$; \item for any cone $\sigma'\in \Delta_{X'}$, $p_\Delta'$ maps $\sigma'$ isomorphically to $\sigma$, and identifies $N_{\sigma'}$ with the sublattice $N_\sigma \cap f_\Delta^{-1} (N_{\tau'})$ of $N_{\sigma}$ \item The following diagrams commute: \begin{minipage}{2.5in} \[ \begin{CD} \sigma' @>{\cong}>p_\Delta'> \sigma\\ @V{f_\Delta'}VV @VV{f_\Delta}V\\ \tau' @>{\cong}>p_\Delta> \tau \end{CD} \] \end{minipage} \begin{minipage}{2.5in} \[ \begin{CD} N_{\sigma'} @>{\subset}>p_\Delta'> N_\sigma\\ @V{f_\Delta'}VV @VV{f_\Delta}V\\ N_{\tau'} @>{\subset}>p_\Delta> N_{\tau} \end{CD} \] \end{minipage} \\ \end{list} \end{itemize} \end{prp} \qed {\bf Proof of \ref{prop-red}} Let $\Delta_X$, $\Delta_B$ be the polyhedral complexes of $X$, $B$, and $f_\Delta: \Delta_X\rightarrow} \newcommand{\dar}{\downarrow\Delta_B$. By equidimensionality, $f_\Delta$ maps $\sigma^{(1)}$ to $\tau^{(1)}$. If $u_i$ are the primitive points of $\Delta_B$, let $v_{ij}$ be the primitive points of $\Delta_X$ such that $f_\Delta(v_{ij}) = m_{ij} u_i$ for some $m_{ij} >0$. Set $m_i = {\operatorname{lcm}}_j \{m_{ij}\}$. Then for all $i,j$, some multiple of $v_{ij}$ maps to $m_i u_i$. Now we use the covering data $(D_i, m_i)$ to define the toroidal morphism $f':X'\rightarrow} \newcommand{\dar}{\downarrow B'$ as in the previous proposition. Let $\sigma' \in \Delta_{X'}$, $f_\Delta'(\sigma') = \tau' \in \Delta_{B'}$. There exist $\sigma\in\Delta_X$, $\tau\in\Delta_B$ such that $\sigma'\cong\sigma$, $\tau'\cong\tau$. The integral lattice $N_{\tau'}$ is generated by $m_i u_i$ with $u_i$ lying on the edges of $\tau$, and $N_{\sigma'}=N_\sigma \cap f_\Delta^{-1} (N_{\tau'})$. For any edge of $\sigma'$ spanned by $v_{ij}$, some multiple $a_{ij} v_{ij}$ maps to $m_i u_i$. Hence $a_{ij} v_{ij} \in N_{\sigma'}$ is primitive and maps to the primitive point $m_i u_i$. This proves that $f'$ has reduced fibers. \qed \section{Mildness of the morphism}\label{mildness} Recall that a normal variety $X$ is said to have Gorenstein singularities if it is Cohen-Macaulay and has an invertible dualizing sheaf $\omega_X$. It is said to have rational singularities, if for a resolution of singularities $r:X'\to X$ we have $r_*\omega_{X'} = \omega_X$. Every toric variety is Cohen-Macaulay with rational singularities (\cite{te} Theorem~14, p. 52). The dualizing sheaf of an affine toric variety $X_\tau$ is the coherent sheaf $\omega_{X_\tau} = {\cal{O}}(-\sum D_i)$ where $D_i$ are closures of the codimension 1 orbits. This sheaf is invertible if and only if there exists an element of $M$ (namely a linear function on $\tau$ taking integer values on $N$) such that $\psi(v) = -1$ for all $v\in \tau^{(1)}$. \begin{lem} Let $X_\sigma$ and $X_\tau$ be affine toric varieties with $X_\tau$ nonsingular. Let $f: X_\sigma \rightarrow} \newcommand{\dar}{\downarrow X_\tau$ be an equidimensional toric morphism without horizontal divisors, having only reduced fibers. Then $X_\sigma$ has rational Gorenstein singularities. \end{lem} {\bf Proof.} Since $X_\tau$ is nonsingular, it has rational Gorenstein singularities. Let $\psi:\tau\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ be the linear interpolation of $\psi(u_i)=-1$ for every primitive point $u_i$ of $\tau$. Then $\psi\circ f_\Delta:\sigma\rightarrow} \newcommand{\dar}{\downarrow{\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}$ is the required function. It is linear because $\psi$ and $f_\Delta$ are; since $f$ is equidimensional and has reduced fibers, primitive points in $\sigma$ map to primitive points in $\tau$, hence $\psi\circ f_\Delta$ takes the value $-1$ on every primitive point of $\sigma$. \qed \begin{lem}\label{lem-rational-Gorenstein} Let $U_X\subset X$ and $U_B \subset B$ be toroidal embeddings, and assume that B is nonsingular. Let $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be an equidimensional toroidal morphism, without horizontal divisors, and with reduced fibers. Then $X$ has rational Gorenstein singularities. \end{lem} {\bf Proof.} Having rational Gorenstein singularities is a local analytic property. Since all local models have rational Gorenstein singularities, so does $X$. \qed To show that the morphism is mild, we need to look at the situation after dominant base changes. We first look at the cases where the base change is relatively nice: \begin{lem}\label{lem-log-smoothness} Let $f:X\to B$ be as above. Let $g:B'\to B$ be a dominant morphism, where $B'$ a nonsingular variety, and assume that $ g^{-1} (B\setmin U_B)$ is a normal crossings divisor. Let $X'$ be the pullback of $X$ to $B'$. Then $X'\to B'$ admits a toroidal structure relative to $ g^{-1}U_B \subset B'$. \end{lem} {\bf Proof.} We use the formalism of logarithmic structures. By \cite{kato1}, \S 8.1, the morphism $X\to B$ is logarithmically smooth. Also $ g^{-1}U_B \subset B'$ endows $B'$ with a logarithmically regular structure. Moreover $g:B'\to B$ is a morphism of logarithmic schemes. The variety $X'$ thus inherits a logarithmic structure. The morphism $X'\to B'$ clearly satisfies the formal lifting property for logarithmic smoothness (\cite{kato1}, \S 8.1,(i)). It is left to show that $X'$ satisfies condition (S) (\cite{kato1} \S 1.5). Indeed, since $f$ is equidimensional, the fibers of $f$ are reduced and $B'$ is normal, it follows that $X'$ is regular in codimension 1. Since $f$ is a Gorenstein morphism and $B$ is Gorenstein, we have that $X'$ is Gorenstein, in particular it is Cohen Macaulay. It then follows that $X'$ is normal. Combining this with the assumption that $f^*M_B$ is saturated in $M_X$, we have that the monoids giving the logarithmic charts on $X'$ are integral and saturated. Altogether, we have that $X'$ satisfies condition (S). Therefore $f':X'\to B'$ is logarithmically smooth, and thus toroidal. \begin{lem}\label{lem-descent} Let $f:X\to B$ be a flat Gorenstein morphism and assume that $B$ has rational Gorenstein singularities. Assume there is a modification $r:B'\to B$ with rational Gorenstein singularities, such that $X'=X\times_B B'$ has rational singularities. Then $X$ has rational singularities as well. \end{lem} {\bf Proof.} Consider the diagram $$\begin{array}{rcl} X' &\stackrel{h}{\to} & X \\ f' \downarrow & & \downarrow f\\ B' &\stackrel{r}{\to} & B. \end{array}$$ By base change we have $h^*\omega_f = \omega_{f'}$. By assumption we have $\omega_B = r_*\omega_{B'}$. The flatness of $f$ implies $f^*r_*\omega_{B'} = h_*{f'}^*\omega_{B'}$. Therefore we have \begin{eqnarray*} h_*\omega_{X'} & = & h_*(\omega_{f'} \otimes {f'}^*\omega_{B'} )\\ & = & h_*(h^*\omega_f \otimes {f'}^*\omega_{B'}) \\ & = & \omega_f \otimes h_*{f'}^*\omega_{B'} \\ & = & \omega_f \otimes f^*r_*\omega_{B'} \\ & = & \omega_f \otimes f^*\omega_B \\ & = & \omega_X. \end{eqnarray*} Thus $X$ has rational singularities. \qed \begin{prp} Let $U_X\subset X$ and $U_B \subset B$ be toroidal embeddings and assume that B is nonsingular. Let $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be an equidimensional morphism with reduced fibers, which is toroidal such that $U_X=f^{-1}U_B$. Then $f$ is mild. \end{prp} {\bf Proof.} Let $B_1\to B$ be a dominant morphism such that $B_1$ has rational Gorenstein singularities. We need to show that $X_1 = X\times_BB_1$ has rational Gorenstein singularities. By lemma \ref{lem-rational-Gorenstein} it has Gorenstein singularities. Pick a resolution of singularities $B'\to B_1$ such that the inverse image of $B\setmin U_B$ in $B'$ is a divisor with normal crossings. By lemma \ref{lem-log-smoothness} we have that $X' = X\times_BB'$ is toroidal, therefore $X'$ has rational singularities. By lemma \ref{lem-descent} we have that $X_1$ has rational singularities as well, which is what we needed. \qed \section{The Cartier covering}\label{sec-cartier} In section \ref{combinatorial} below we will translate the semistable reduction conjecture in purely combinatorial terms. In order to maximize the flexibility of the combinatorial operations, we need to generalize Kawamata's trick slightly to accommodate cases where $B$ has quotient singularities. In such a situation we need a finite cover of $B$ which is nonsingular, and such that the resulting polyhedral complex is easily described. The following statement will suffice for this purpose. \begin{prp}\label{prop-cartier} Let $U_X\subset X$ and $U_B \subset B$ be projective toroidal embeddings, and assume that $B$ has only quotient singularities. Let $f: X\rightarrow} \newcommand{\dar}{\downarrow B$ be a surjective mild toroidal morphism, satisfying $f^{-1}(U_B)=U_X$. Then there exists a finite surjective morphism $p: B'\rightarrow} \newcommand{\dar}{\downarrow B$ so that, denoting $X'=X\times_B B'$, $U_{B'}=p^{-1}U_B$ and $U_{X'}= X\times_B U_{B'}$ we have \begin{enumerate} \item $B'$ and $X'$ are toroidal embeddings with polyhedral complexes $\Delta_{X'}$ and $\Delta_{B'}$, with $B'$ nonsingular; \item the projection $f': X'\rightarrow} \newcommand{\dar}{\downarrow B'$ is a mild toroidal morphism; \item There exist morphisms $$p_\Delta:\Delta_{B'}\rightarrow} \newcommand{\dar}{\downarrow \Delta_{B}$$ and $$p_\Delta': \Delta_{X'}\rightarrow} \newcommand{\dar}{\downarrow \Delta_{X},$$ such that \begin{list}{$\circ$}{} \item for any cone $\tau'\in \Delta_{B'}$, the morphism $p_\Delta$ maps $\tau'$ isomorphically to $\tau$, and identifies $N_{\tau'}$ with the sublattice of $N_{\tau}$ generated by the primitive vectors of $\tau$; \item for any cone $\sigma'\in \Delta_{X'}$, $p_\Delta'$ identifies $N_{\sigma'}$ with the sublattice $N_\sigma \cap f_\Delta^{-1} (N_{\tau'})$ of $N_{\sigma}$, and maps $\sigma'$ isomorphically to $\sigma$. \item The following diagrams commute: \begin{minipage}{2.5in} \[ \begin{CD} \sigma' @>{\cong}>p_\Delta'> \sigma\\ @V{f_\Delta'}VV @VV{f_\Delta}V\\ \tau' @>{\cong}>p_\Delta> \tau \end{CD} \] \end{minipage} \begin{minipage}{2.5in} \[ \begin{CD} N_{\sigma'} @>{\subset}>p_\Delta'> N_\sigma\\ @V{f_\Delta'}VV @VV{f_\Delta}V\\ N_{\tau'} @>{\subset}>p_\Delta> N_{\tau} \end{CD} \] \end{minipage} \\ \end{list} \end{enumerate} \end{prp} {\bf Proof.} We use a construction analogous to Kawamata's covering. First, let $B$ be a variety, $E$ an effective Weil divisor on $B$. We make the following assumptions: \begin{enumerate} \item as a scheme, $E$ is integral and normal; \item the divisor $ mE = D$ is a Cartier; \item There is a Cartier divisor $L$ on $B$ such that ${\cal{O}}(D) = {\cal{O}}(mL)$. \end{enumerate} As in Definition \ref{def-cyclic-cover}, we choose a function $\phi$ with $\operatorname{div}(\phi) = D-mL$, and define the cyclic cover $p:B'\to B$ by taking $m$-th root of $\phi$. The point is, that the ${\Bbb{Q}}} \newcommand{\bff}{{\Bbb{F}}$-Cartier divisor $p^*E$ is in fact Cartier. Let us see what happens in the toric case. \begin{lem} Let $X_\tau$ be an affine toric variety with quotient singularities and $E$ a toric divisor corresponding to a primitive vector $v$. Write $\tau={\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}_+v\times \tau_1$. Choose $m\in \bfn$ such that $D = mE$ is Cartier. Then ${\cal{O}}(mD)$ is trivial, and the corresponding cyclic cover is $X_{\tau'}$, where $|\tau'| = \tau$ and $N_{\tau'} = {\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}} v\times (N_\tau\cap\tau_1)$. \end{lem} This lemma is easy and left to the reader. If we iterate the lemma with all the toric divisors, we obtain a nonsingular covering, whose lattice is generated by the primitive vectors of $\tau$. Assume that in addition we have a mild morphism $X_\sigma\to X_\tau$. Lemma \ref{lem-tor-prod} works word-for-word in this case. The only thing which needs to be changed in the proof is, that since $f_\Delta(N_\sigma) = N_\tau$, the semigroup homomorphism $i:\pi\to M$ is still injective. (This is not a serious business - we could replace the product by its reduction anyway.) In order to define a Kawamata covering package, we need the following Bertini type lemma (this is a special case of the stratified Bertini Theorem) . \begin{lem} Let $U_B\subset B$ be a toroidal embedding. Let $H_0$ be a very ample divisor on $B$ and let $H$ be a general element of $|H_0|$. Let $U_B' = U_B\setmin H$. We have \begin{enumerate} \item $U_B'\subset B$ is a toroidal embedding; \item let $b\in H$ and assume that $b$ lies on a stratum $Y\subset B$ of the toroidal embedding $U_B\subset B$. Then $H\cup Y$ is nonsingular at $b$; \item there is a local model $X_\rho \times \bfg_m^k$ for $U_B\subset B$ at $b$, where $Y$ corresponds to the factor $\bfg_m^k$, and $H$ corresponds to $X_\rho \times \bfg_m^{k-1}\times\{1\}$. \item there is a local model $X_\rho \times \bfg_m^{k-1}\times{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$ for $U_B'\subset B$ at $b$, where $Y$ corresponds to $\bfg_m^k\times{\Bbb{A}}} \newcommand{\bfn}{{\Bbb{N}}^1$, and $H$ corresponds to $X_\rho \times \bfg_m^{k-1}\times\{0\}$. \end{enumerate} \end{lem} {\bf Proof.} Since $B$ is a stratified space, it has a local product structure $X_\rho\times Y$ at $b$. We need to adapt this structure to the divisor $H$ as in the lemma. Applying Bertini's theorem to the closure of the stratum $Y$, we have that $H\cup Y$ is nonsingular. Pick a regular function $u$ near $b$ defining $H$ on a neighborhood of $b$. Then the restriction $u_Y$ of $u$ to $Y$ is a regular parameter. We can find a regular system of parameters on $Y$ at $b$ including $u_Y$ and lift them to $B$. This lifting gives the desired adaptation of the product structure. \qed Using this lemma, the proof of Proposition \ref{prop-compl} works word-for-word here, and yields Proposition \ref{prop-cartier}. \qed \section{Towards semistable reduction}\label{combinatorial} \subsection{Combinatorial semistable reduction} We are going to restate the semistable reduction conjecture (Conjecture \ref{conj-semistable}) in purely combinatorial terms. First let us define semistability combinatorially. \begin{dfn}\label{dfn-comb-semi} Let $\Delta_X$ and $\Delta_B$ be rational conical polyhedral complexes, and let $f_\Delta:\Delta_X\to \Delta_B$ be a surjective polyhedral map. We say that $f_\Delta$ is {\bf weakly semistable} if the following conditions hold. \begin{enumerate} \item We have $f_\Delta^{-1}(0) = \{0\}$. \item For any cone $\tau\in \Delta_X$, we have $f_\Delta(\tau)\in \Delta_B$. \item\label{item-reduced} For any cone $\tau\in \Delta_X$, we have $f_\Delta(N_\tau) = N_{f_\Delta(\tau)}$. \item $\Delta_B$ is simplicial with index 1. \end{enumerate} If also $\Delta_X$ is simplicial with index 1, we say that $f_\Delta$ is {\bf semistable}. \end{dfn} We now define the operations we are allowed to perform. \begin{dfn} Let $\Delta$ be a rational conical polyhedral complex. A {\bf lattice alteration} $\Delta_1\to \Delta$ consists of an integral structure induced by a consistent choice of sublattices $N^1_\sigma\subset N_\sigma$ for each cone $\sigma\subset \Delta$. An {\bf alteration} $\Delta_1\to \Delta$ is a composition $\Delta_1\to \Delta'\to \Delta$ of a lattice alteration $\Delta_1\to \Delta'$ with a subdivision $\Delta'\to \Delta$. The alteration is {\bf projective} if the corresponding subdivision $\Delta'\to \Delta$ is projective. \end{dfn} Note that the composition of alterations is an alteration. Be warned that the factorization of a polyhedral alteration above is not analogous to the Stein factorization in algebraic geometry - it has the opposite order. \begin{dfn} \begin{enumerate} \item Let $\Delta_X\to \Delta_B$ be a polyhedral map of rational conical polyhedral complexes. Let $\Delta_B'\to \Delta_B$ be a subdivision. The induced subdivision $\Delta_X'\to \Delta_X$ is the minimal subdivision admitting a polyhedral map to $\Delta_B'$. The cones of $\Delta_X'$ are of the form $\sigma\cap f_\Delta^{-1}(\tau')$ for $\sigma\in \Delta_X$ and $\tau'\in \Delta_B'$. \item Let $\Delta_X\to \Delta_B$ be a polyhedral map of rational conical polyhedral complexes. Let $\Delta_B^1\to \Delta_B$ be a lattice alteration. The induced lattice alteration $\Delta_X^1\to \Delta_X$ is the minimal lattice alteration mapping to $\Delta_B^1$. The lattices of $\Delta_X^1$ are of the form $N_\sigma\cap f_\Delta^{-1}(N_{\tau^1})$ for $\sigma\in \Delta_X$ and $\tau^1\in \Delta_B^1$. \item Let $\Delta_X\to \Delta_B$ be a polyhedral map of rational conical polyhedral complexes. Let $\Delta_B^1\to \Delta_B'\to \Delta_B$ be an alteration, factored as a lattice alteration followed by a subdivision. The induced alteration $\Delta_X^1\to \Delta_X$ is the induced lattice alteration $\Delta_X^1\to \Delta_X'$ of the induced subdivision $\Delta_X'\to \Delta_X$. \end{enumerate}\end{dfn} Note that an alteration induced by a projective alteration is projective. We are now ready to state our conjecture. \begin{conj}\label{conj-comb-semistable} Let $f_\Delta:\Delta_X\to \Delta_B$ be a polyhedral map of rational conical polyhedral complexes, and assume for simplicity that $f_\Delta^{-1}(0) = \{0\}$. Then there exists a projective alteration $\Delta_B^1\to \Delta_B$, with induced alteration $\Delta_X^1\to \Delta_X$, and a projective subdivision $\Delta_Y\to \Delta_X^1$, such that $\Delta_Y \to \Delta_B^1$ is semistable. \end{conj} The conjectures are tied together by the following proposition: \begin{prp} Conjecture \ref{conj-comb-semistable} implies Conjecture \ref{conj-semistable}. \end{prp} {\bf Proof.} Let $X\to B$ be a surjective morphism of complex projective varieties with geometrically integral generic fiber. By Theorem \ref{th-toroidal-reduction} we may assume the morphism toroidal. Let $f_\Delta:\Delta_X \to \Delta_B$ be the associated polyhedral map. By theorem \ref{th-weak-semistable-reduction} we may assume $f_\Delta$ is weakly semistable. We now invoque Conjecture \ref{conj-comb-semistable}. Let $\Delta_B^1\to \Delta_B'\to\Delta_B$ be a projective alteration, $\Delta_X^1\to \Delta_X$ the induced alteration, and $\Delta_Y\to \Delta_X^1$ a projective subdivision such that $\Delta_Y \to \Delta_B^1$ is semistable. The subdivision $\Delta_B'\to\Delta_B$ and the induced subdivision $\Delta_X'\to\Delta_X$ give rise to modifications $B'\to B$ and $X'\to X$ with a lifting $X'\to B'$ (see lemma \ref{lem-lifting}). Let $\Delta_B^0\to \Delta_B'$ be the lattice alteration given by the sublattice generated by the primitive vectors, and $\Delta_X^0\to \Delta_X'$ the induced lattice alteration. Proposition \ref{prop-cartier} gives rise to an alteration $\widetilde{B_0} \to B'$, with pullback $\widetilde{X_0} \to X'$, and a polyhedral map $\Delta_{\widetilde{B_0}}\to \Delta_B^0$ which is an isomorphism on each cone. The same holds for $\Delta_{\widetilde{X_0}}\to \Delta_X^0$. The lattice alteration $\Delta_B^1\to \Delta_B'$, being of index 1, factors through $\Delta_B^0$. Proposition \ref{prop-compl} gives an alteration $\widetilde{B_1}\to\widetilde{B_0}$ with pullback $\widetilde{X_1}\to\widetilde{X_0}$ and a polyhedral map $\Delta_{\widetilde{B_1}}\to \Delta_B^1$ which is an isomorphism on each cone, and the same holds for $\Delta_{\widetilde{X_1}}\to \Delta_X^1$. The subdivision $\Delta_Y\to \Delta_X^1$ induces a subdivision $\widetilde{\Delta_{Y}}\to\Delta_{\widetilde{X_1}}$, giving rise to a toroidal modification $\widetilde{Y}\to \widetilde{X_1}$. Since $\Delta_Y\to \Delta_B^1$ is semistable, we have that $\widetilde{\Delta_{Y}}\to\Delta_{\widetilde{B_1}}$ is semistable as well, and therefore $\widetilde{Y}\to\widetilde{B_1}$ is semistable, as required. \qed \subsection{How far can we push the results?} In \cite{ar}, it is shown that if $\Delta_0\subset \Delta$ is a subcomplex, and $\Delta_0'\to \Delta_0$ is a {\em projective} triangulation, then there is a projective triangulation $\Delta'\to \Delta$ extending $\Delta_0'$, without new edges: ${\Delta'}^{(1)} = {\Delta}^{(1)} \cup {\Delta_0'}^{(1)}$. Applying the main theorem of \cite{te}, Chapter III, it is shown in \cite{ar} that given a weakly semistable $X\to B$, one can locally find a finite map $B_1\to B$ and a modification $X_2\to X\times_BB_1$ such that $X_2\to B_1$ is semistable in codimension 1. If we apply Kawamata's trick, then clearly this can be done globally. Moreover, $X_2$ has only quotient singularities. A weakly semistable morphism is said to be {\bf almost semistable} if it is semistable in codimension 1 and moreover $X$ has quotient singularities. Thus, in theorem \ref{th-weak-semistable-reduction} we may replace ``weakly semistable'' by ``almost semistable''. An analogous definition can be made on the polyhedral side. It is important to note that an almost semistable morphism is not necessarily semistable. It is easy to give a polyhedral example: let $\tau = ({\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}_+)^2$ be endowed with the standard lattice ${\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}^2$, and let $\sigma = ({\Bbb{R}}} \newcommand{\bfo}{{\Bbb{O}}_+)^4$ be given the lattice generated by ${\Bbb{Z}}} \newcommand{\bfg}{{\Bbb{G}}^4$ and the vector $w=(1/2,1/2,1/2,1/2)$. We have a polyhedral map $\sigma\to \tau$ given by $(a,b,c,d)\mapsto (a+b, c+d)$. It is easy to see that this is almost semistable, but not semistable, since $\sigma$ has index 2. Needless to say, a corresponding toroidal example can be easily constructed as well. The example we just gave is easy to amend. Indeed, if we subdivide $\tau$ at its barycenter $(1,1)$, take the induced subdivision of $\sigma$, then its star subdivision centered at $w$, and extend this to a triangulation using \cite{ar}, we obtain a semistable map. This can be extended to families of surfaces in general - the main observation (see \cite{wang} for the case $\dim(B)=1$) is that one can use Pick's theorem and subdivide, with no need for additional lattice alteration. We plan to pursue this elsewhere. One last remark: the second author has shown, that in order to prove semistable reduction, it is sufficient to produce $B_1$ and $Y$ such that $Y\to B_1$ satisfies all but condition \ref{item-reduced} of the requirements for semistability in Definition \ref{dfn-comb-semi}. Again, this will be pursued elsewhere.
1997-07-10T11:37:16
9707
alg-geom/9707011
en
https://arxiv.org/abs/alg-geom/9707011
[ "alg-geom", "math.AG" ]
alg-geom/9707011
Dionisi Carla
Carla Dionisi
The tangent space at a special symplectic instanton bundle on P^{2n+1}
Latex, 11 pages, to appear in Annali di Matematica
null
null
null
null
Let $MI_{Simp,P^{2n+1}}(k)$ be the moduli space of stable symplectic instanton bundles on $P^{2n+1}$ with second Chern class $c_2=k$ (it is a closed subscheme of the moduli space $MI_{P^{2n+1}}(k)$), We prove that the dimension of its Zariski tangent space at a special (symplectic) instanton bundle is $2k(5n-1)+4n^2-10n+3, k\geq 2$. It follows that special symplectic instanton bundles are smooth points for $ k \leq 3 $
[ { "version": "v1", "created": "Thu, 10 Jul 1997 09:43:34 GMT" } ]
2008-02-03T00:00:00
[ [ "Dionisi", "Carla", "" ] ]
alg-geom
\section*{Introduction} Symplectic instanton bundles on $\mbox{{I \hspace{-,18cm}P}}^{2n+1}$ are holomorphic bundles of rank 2n (see \cite{A} ,\cite{MS} and \cite{OS}) that correspond to the self-dual solutions of Yang-Mills equations on $\mbox{{I \hspace{-,18cm}P}}^n(\mbox{{I \hspace{-.18cm}H}})$. \ They are given by some monads (see section 2 for precise definitions) and their only topological invariant is $c_2=k$.\\ At present the dimension of their moduli space $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$ is not known except in the cases n=1, where the dimension is $8k-3$ (see \cite{HN}), and in few other cases corresponding to small values of k.\\ $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$ is a closed subscheme of $MI_{\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$ and this last scheme parametrizes stable instanton bundles with structural group $GL(2n)$.\\ The class of special instanton bundles was introduced in \cite{ST}. \\ Let $E \in MI_{\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$ be a special symplectic instanton bundle. The tangent dimension $h^1(End(E))$ was computed in \cite{OT} and it is equal to $4(3n-1)k+(2n-5)(2n-1)$ .\\ The Zariski tangent space of $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$ at $E$ is isomorphic to $H^1(S^2E)$ and in this paper we prove that \begin{equation} \label{zero} h^2(S^2E)=\left(\begin{array}{c} k-2 \\ 2 \end{array} \right) \cdot \left(\begin{array}{c} 2n-1 \\ 2 \end{array} \right) \qquad \forall \ k \geq 2 \end{equation} By the Hirzebruch-Riemann-Roch formula , since $h^0(S^2E)=0$ and $h^i(S^2E)=0 \ \forall\ i\geq3$, it follows that: \begin{center} $\chi(S^2E)=h^2(S^2E)-h^1(S^2E)=2n^2+n+\frac{1}{2}\left[ k^2\left( \begin{array}{c} 2n-1 \\ 2 \end{array} \right) -k(10n^2-5n-1) \right] $ \end{center} and \begin{theorem} \label{acca1} Let $E$ be a special symplectic instanton bundle.Then $$h^1(S^2E)=2k(5n-1)+4n^2-10n+3 \qquad ,k\geq 2 $$ \end{theorem} (for $n=1$ it is well known that \ $ h^1(S^2E)=8k-3$ ).\\ Now, since by the Kuranishi map $H^2(S^2E)$ is the space of obstructions to the smoothness at $E$ of $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$, we obtain \begin{cor} $\forall k\geq 2$ the dimension of any irreducible component of $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(k)$, containing a special symplectic instanton bundle is bounded by the value \\ $2k(5n-1)+4n^2-10n+3 \quad $ ( linear in k ) \end{cor} \begin{cor} $\forall n$ \ $MI_{\mbox{Simp},\mbox{{I \hspace{-,18cm}P}}^{2n+1}}(3)$ is smooth at a special instanton bundle $E$,and the dimension of any irreducible component containing $E$ is $ 4n^2+20n-3 $. \end{cor} The main remark of this paper is that it is easier to compute $H^2(S^2E)$ and $H^2(\stackrel{2}{\wedge} E)$ together as SL(2)-modules (although this second cohomology space has a geometrical meaning only for orthogonal bundles) than to compute $H^2(S^2E)$ alone. \section{Preliminaries} Throughout this paper $\mbox{{\itshape I \hspace{-.38cm} K}}$ denotes an algebraically closed field of characteristic zero. $U$ denotes a 2-dimensional $\mbox{{\itshape I \hspace{-.38cm} K}}$ vector space \((U=<s,t>)\), \ $S_n=S^nU$ its n-th symmetric power $\ms{0.2}(S_n=<s^n, s^{n-1}t,\ldots, t^n>$) , $V_n=U\otimes S_n $ $\ms{0.2}(V_n=<s\otimes s^n, s\otimes s^{n-1}t,\ldots s\otimes t, \ldots t\otimes t^n>)$ \ and $\mbox{{I \hspace{-,18cm}P}}^{2n+1}=\mbox{{I \hspace{-,18cm}P}}(V_n)$. \begin{defin} \label{def 1.1} A vector bundle $E$ on $\mbox{{I \hspace{-,18cm}P}}^{2n+1}$ of rank $2n$ is called an {\bf instanton bundle of quantum number $k$} if: \begin{itemize} \pagebreak \item $E$ has Chern polinomial $c_t(E) \ = \ (1-t^2)^{-k}$; \item $E(q)$ has natural cohomology in the range $-(2n+1) \ \leq \ q \ \leq 0$, that is $H^i(E(q)) \ \neq \ 0$ for at most one $i = i(q)$. \end{itemize} \end{defin} By \cite{OS},\cite{AO1},the Definition ~\ref{def 1.1} is equivalent to : \\ i)$E$ is the cohomology bundle of a monad: \[ 0 \rightarrow O(-1)^k \rightarrow \Omega^1(1)^k \rightarrow O^{2n(k-1)} \rightarrow 0 \] or ii) $E$ is the cohomology bundle of a monad: \[ 0 \rightarrow O(-1)^k \fr{A} O^{2n+2k} \fr{B^t} O(1)^k \rightarrow 0 \] (where, after we have fixed a coordinate system, A and B can be identified with matrices in the space $Mat(k,2n+2k,S_1)$) \begin{defin} An instanton bundle $E$ is called {\bf symplectic} if there is an isomorphism $\varphi:E \rightarrow E^ {\vee}$ satisfying $\varphi = -\varphi^{\vee}$. \end{defin} \begin{defin} An instanton bundle is called {\bf special} if it arises from a monad where the morfism $B^t$ is defined in some system of homogeneous coordinates $x_0, \cdots x_n, y_0 \cdots y_n$ on $\mbox{{I \hspace{-,18cm}P}}^{2n+1}$ by the trasposed of the matrix: \[ B = \left( \begin{array}{cccccccccccc} x_0 & \cdots & x_n & 0 & \cdots & 0 & y_0 & \cdots & y_n & 0 & \cdots & 0 \\ 0 & x_0 & \cdots & x_n & 0 & \cdots & 0 & y_0 & \cdots & y_n & 0 & \cdots \\ \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots & \cdots\\ \cdots & 0 & x_0 & \cdots & x_n & 0 & \cdots & 0 & y_0 & \cdots & y_n & 0 \\ 0 & \cdots & 0 & x_0 & \cdots & x_n & 0 & \cdots & 0 & y_0 & \cdots & y_n \end{array} \right) \] \\ \end{defin} The following lemma is well known (and easy to prove) \pagebreak \begin{lemma} $$H^{0} (O(1)) \cong V^{\vee}$$ $$H^{0}(\Omega^1(2)) \cong \stackrel{2}{\wedge} V^{\vee}$$ $$H^i (\mbox{{I \hspace{-,18cm}P}}^n, S^2 \Omega^1(1)) = \left \{ \begin{array}{ll} 0 & \mbox{se } i\neq 1 \\ \stackrel{2}{\wedge} V^{\vee} & \mbox{se } i= 1 \end{array} \right.$$ \end{lemma} \section{Existence of a special symplectic instanton bundle} There is a natural exact sequence of GL(U)-equivariant maps for any $k , n\geq 1$ (Clebsch-Gordan sequence): \begin{equation} \label{uno_3} 0 \rightarrow \stackrel{2}{\wedge} U \otimes S_{k-1} \otimes V_{n-1} \fr{\beta} S_k \otimes V_n \fr{\mu} V_{k+n} \rightarrow 0 \end{equation} where $\mu$ is the multiplication map and $\beta$ is defined by $(s \wedge t) \otimes f \otimes g \rightarrow (sf \otimes tg - tf \otimes sg)$ We can define (see \cite{OT}) the morphism \\ \centerline{ $\ac{b}: S_{k-1}^{\vee} \otimes \Omega^1(1) \rightarrow \stackrel{2}{\wedge} U^{\vee} \otimes S_{k-2}^{\vee} \otimes V_{n-1}^{\vee} \otimes O$ } and it is induced the complex \begin{equation} \label{cinque_3} A \otimes O(-1) \fr{\ac{a}} S^{\vee}_{k-1} \otimes \Omega^1(1) \fr{\ac{b}} \stackrel{2}{\wedge} U^{\vee} \otimes S^{\vee}_{k-2} \otimes V_{n-1}^{\vee} \otimes O \end{equation} where A is a k-dimensional subspace of $S^{\vee}_{2n+k-1} \otimes \stackrel{2}{\wedge} U^{\vee}$ such that ~\refeq{cinque_3} is a monad and the cohomology bundle $E$ is a special symplectic instanton bundle. It was proved in \cite{OT} that $$ H^2(EndE)\cong Ker(\Phi^{\vee})^{\vee}$$ where \centerline{ $\Phi^{\vee}: S_{k-2}^{\otimes 2} \otimes V_{n-1}^{\otimes 2} \rightarrow S_{k-1}^{\otimes 2} \otimes \stackrel{2}{\wedge} V_{n}$ } and there is an isomorphism of SL(2)-representations $$ \varepsilon : S^{\vee}_{k-3} \otimes S^{\vee}_{k-3} \otimes S^2V^{\vee}_{n-2} \rightarrow Ker(\Phi^{\vee})$$ \section{How to identify $H^2(S^2E)$ and $H^2(\stackrel{2}{\wedge} E)$} \begin{prop} Let $E$ be special symplectic instanton bundle ,cohomology of monad ~\refeq{cinque_3} and $N=Ker \ac{b}$.Then \begin{description} \item[(i)] $H^2 (S^2E) \cong H^2(S^2N) $ \item[(ii)] $H^2 (\stackrel{2}{\wedge} E) \cong H^2(\stackrel{2}{\wedge} N) $ \end{description} \end{prop} \begin{proof} \\ We denote $B:= S_{k-1}^{\vee}$ \qquad and \qquad $C:=\stackrel{2}{\wedge} U^{\vee} \otimes S_{k-2}^{\vee} \otimes V_{n-1}^{\vee} $ \\ The result follows from the two exact sequences given by monad ~\refeq{cinque_3} : \begin{equation} \label{cinque_4} 0 \rightarrow N \rightarrow B \otimes \Omega^1 (1) \rightarrow C \otimes O \rightarrow 0 \end{equation} \begin{equation} 0 \rightarrow A \otimes O(-1) \rightarrow N \rightarrow E \rightarrow 0 \end{equation} In fact, by performing the second symetric and alternating power of sequence ~\refeq{cinque_4}, we have \centerline{ \begin{minipage}{2in} \begin{tabbing} $ 0 \rightarrow S^2 N \rightarrow $\=$\tilde{A}$\=$\rightarrow B \otimes C \otimes \Omega^1(1) \rightarrow \stackrel{2}{\wedge} C \otimes O \rightarrow 0$\\ \> \>$\searrow$\=\hspace{0.5 cm}$\nearrow$\=\\ \> \> \> $ M^1$\\ \> \>$\nearrow$\hspace{0.5 cm}$\searrow$\\ \> 0 \> \> \> 0 \end{tabbing} \end{minipage} } \begin{equation} \label{DUE} \end{equation} where $\tilde{A}:= S^2(B \otimes \Omega^1(1)) = (S^2 B \otimes S^2(\Omega^1(1))) \oplus ( \stackrel{2}{\wedge} B \otimes \Omega^2(2))$ \\ and \\ \centerline{ \begin{minipage}{2in} \begin{tabbing} $ 0 \rightarrow \stackrel{2}{\wedge} N \rightarrow $\=$\overline{A} $\=$\ \rightarrow B \otimes C \otimes \Omega^1(1) \rightarrow O \otimes S^2 C \rightarrow 0$\\ \> \>$\searrow$\=\hspace{0.5 cm}$\nearrow$\=\\ \> \> \> $ M$\\ \> \>$\nearrow$\hspace{0.5 cm}$\searrow$\\ \> 0 \> \> \> 0 \end{tabbing} \end{minipage} } \begin{equation} \label{TRE} \end{equation} where $\overline{A}:= \stackrel{2}{\wedge}(B \otimes \Omega^1(1)) = (\stackrel{2}{\wedge} B \otimes S^2(\Omega^1(1))) \oplus ( S^2 B \otimes \Omega^2(2))$ \\ \end{proof} \subsection{Identifying $ H^2(S^2 N)$ and $H^2 ( \stackrel{2}{\wedge} N) $} i) \label{par43} Diagram ~\refeq{DUE} gives the following two exact sequences:\\ \begin{minipage}{5 in} \begin{equation}\label{quattro4} O \rightarrow H^0(M^1) \rightarrow H^1(S^2N) \rightarrow H^1(\ac{A}) \rightarrow H^1(M^1) \rightarrow H^2(S^2(N)) \rightarrow H^2(\ac{A}) \rightarrow \cdots \end{equation} \begin{tabbing} $O \rightarrow H^0(M^1) \rightarrow B \otimes C \otimes H^0$\=$(\Omega^1(1))\rightarrow \stackrel{2}{\wedge} C\rightarrow H^1(M^1) \rightarrow B \otimes C \otimes H^1$\=$(\Omega^1(1))\rightarrow \cdots$\\ \>$\parallel$\> $\parallel$\\ \>$0 $ \> $0$ \end{tabbing} \begin{equation}\label{cinque4} \end{equation} \end{minipage} Sequence \refeq{cinque4} implies: $H^0(M^1) = 0$ \quad and \quad $H^1(M^1) \cong \stackrel{2}{\wedge} C$ Then, by using the two formulas: $H^1(\tilde{A}) = (S^2 B \otimes H^1(S^2 \Omega^1(1))) \oplus ( \stackrel{2}{\wedge} B \otimes H^1(\Omega^2(2)) =S^2 B \otimes \stackrel{2}{\wedge} V^{\vee}$ and:\ $H^2(\tilde{A})= (S^2 B \otimes H^2(S^2 \Omega^1(1))) \oplus ( \stackrel{2}{\wedge} B \otimes H^2(\Omega^2(2)) =0$ sequence ~\refeq{quattro4} becomes: \[0 \rightarrow H^1(S^2N) \rightarrow H^1(\ac{A}) \rightarrow H^1(M^1) \rightarrow H^2(S^2(N)) \rightarrow 0\] i.e.\[0 \rightarrow H^1(S^2 N) \rightarrow S^2 B \otimes \stackrel{2}{\wedge} V^{\vee} \fr{\ac{\Phi}} \stackrel{2}{\wedge} C \rightarrow H^2(S^2N) \rightarrow 0 \] \[ \Longrightarrow \qquad H^2 ( S^2 N ) \cong \mbox{Coker}(\ac{\Phi}) = (\mbox{Ker}(\ac{\Phi}^{\vee}))^{\vee} \] Then: \[H^2 ( S^2 N) ^{\vee} = \mbox{Ker} \left[ \stackrel{2}{\wedge} (S_{k-2} \otimes V_{n-1} ) \fr{\ac{\Phi}^{\vee}} S^2(S_{k-1}) \otimes \stackrel{2}{\wedge} V_n\right]\] ii) Diagram ~\refeq{TRE} gives the following two exact sequences: \begin{minipage}{5in} \begin{equation}\label{sei4} O \rightarrow H^0(M) \rightarrow H^1(\stackrel{2}{\wedge} N) \rightarrow H^1(\overline{A}) \rightarrow H^1(M) \rightarrow H^2(\stackrel{2}{\wedge} N) \rightarrow H^2(\overline{A}) \rightarrow \cdots \end{equation} \begin{tabbing} $O \rightarrow H^0(M) \rightarrow B \otimes C \otimes H^0$\=$(\Omega^1(1))\rightarrow S^2 C $\=$\otimes H^0(O) \rightarrow H^1(M) \rightarrow 0 \rightarrow \cdots$\\ \>$\parallel$\> $\parallel$\\ \>$0 $ \> $S^2 C$ \end{tabbing} \begin{equation}\label{sette4} \end{equation} \end{minipage}\\ and, from sequence ~\refeq{sette4}, we get \\ $H^0(M) = 0$ \quad and \quad $H^1(M) \simeq S^2 C$ \\ Then, since :\\ $H^1(\overline{A})=(H^1(S^2(\Omega^1(1))\otimes \stackrel{2}{\wedge} B) \oplus (S^2 B \otimes H^1(\Omega^2(2)))= \stackrel{2}{\wedge} B \otimes \stackrel{2}{\wedge} V^{\vee}$ \\ and \qquad $H^2(\overline A) = 0$\\ sequence ~\refeq{sei4} becomes : \begin{minipage}{5 in} \begin{tabbing} $O \rightarrow H^0$\=$(M) \rightarrow H^1(\stackrel{2}{\wedge} N)$\=$ \rightarrow H^1(\overline{A}) \rightarrow H^1(M) \rightarrow H^2(\stackrel{2}{\wedge} N) \rightarrow 0$\\ \>$\parallel$\\ \>0 \end{tabbing} \end{minipage} i.e. \qquad $0 \rightarrow H^1(\stackrel{2}{\wedge} N) \rightarrow \stackrel{2}{\wedge} B \otimes \stackrel{2}{\wedge} V^{\vee} \fr{\overline{\Phi}} S^2 C \rightarrow H^2(\stackrel{2}{\wedge} N ) \rightarrow 0 $ \\ \[ \Longrightarrow \qquad H^2 ( \stackrel{2}{\wedge} N ) \cong \mbox{Coker}(\overline{\Phi}) = (\mbox{Ker}(\overline{\Phi}^{\vee}))^{\vee} \] Then we obtain : \[(H^2(\stackrel{2}{\wedge} N))^{\vee} = \mbox{Ker} \left[ S^2(S_{k-2}\otimes V_{n-1})\fr{\overline{\Phi}^{\vee}}\stackrel{2}{\wedge} S_{k-1}\otimes \stackrel{2}{\wedge} V_n \right]\] \subsection{\bf Identifying $H^2(S^2 E)$} We have \qquad \qquad $H^2(S^2E)^{\vee} \cong \mbox{Ker} \ \ac{\Phi}^{\vee}$ \\ where \qquad $\ac{\Phi}^{\vee} : \stackrel{2}{\wedge}(S_{k-2}\otimes V_{n-1}) \rightarrow S^2 S_{k-1} \otimes \stackrel{2}{\wedge} V_n$ \\ is explicitly given by \begin{tabbing} $\ac{\Phi}^{\vee}((g \otimes v)\wedge (g^1 \otimes v^1)) =$\=$sg \cdot sg^1 \otimes (tv \wedge tv^1)-sg\cdot tg^1 \otimes (tv \wedge sv^1)+$\\ \>$-tg\cdot sg^1 \otimes (sv \wedge tv^1)+tg\cdot tg^1 \otimes (sv \wedge sv^1) $ \\ \end{tabbing} i.e. \qquad $ \ac{\Phi}^{\vee} = \ac{p} \circ (\stackrel{2}{\wedge}\beta)$ \\ where \qquad $\beta : \ \ \stackrel{2}{\wedge} U \otimes S_{k-2} \otimes V_{n-1} \rightarrow S_{k-1} \otimes V_n \; \; \; $ is such that \[ (s \wedge t) \otimes (g \otimes v) \mapsto (sg \otimes tv) - (tg \otimes sv) \] and \begin{tabbing} $\ac{p} \ : \ $\=$\stackrel{2}{\wedge}($\=$S_{k-1} \otimes V_n) \rightarrow S^2 S_{k-1} \otimes \stackrel{2}{\wedge} V_n$ \\ \>\>$\|$ \\ \>$(\stackrel{2}{\wedge} S_{k-1} \otimes S^2V_n) \oplus (S^2S_{k-1} \otimes \stackrel{2}{\wedge} V_n)$\\ \end{tabbing} is such that $$ (f \otimes u) \wedge (f' \otimes u^1) \mapsto f\cdot f' \otimes u \wedge u^1 \mbox{.} $$ Now, we consider the $SL(2)$-equivariant morphism:\\ \[ \ac{\varepsilon}^1 \ : \ \stackrel{2}{\wedge} (S_{k-3} \otimes V_{n-2}) \rightarrow \stackrel{2}{\wedge}(S_{k-2} \otimes V_{n-1}) \] where, up to the order of factors, the map $\ac{\varepsilon}^1 := \beta^1 \wedge \beta^1$ and $\beta^1:S_{k-3} \otimes V_{n-2} \rightarrow S_{k-2} \otimes V_{n-1}$ is defined as $\beta$. Hence, $\ac{\varepsilon}^1 $ is injective. Finally, we define \[ \ac{\varepsilon} \ : \ \stackrel{2}{\wedge} S_{k-3} \otimes S^2V_{n-2} \rightarrow \stackrel{2}{\wedge}(S_{k-2} \otimes V_{n-1}) \] as $\ac{\varepsilon} = \ac{\varepsilon}^1 \circ \ac{i}$, \ where \begin{tabbing} $\ac{i} \ :\ $\=$ \stackrel{2}{\wedge} S_{k-3} \otimes S^2V_{n-2} \rightarrow \stackrel{2}{\wedge}(S_{k-3} \otimes V_{n-2}) \ \ $ such that \\ \> $f \wedge f' \otimes u\cdot u^1 \longmapsto (f \otimes u) \wedge (f' \otimes u^1) + (f \otimes u^1) \wedge (f' \otimes u)$\\ \end{tabbing} is an injective map. Then, also $\ac{\varepsilon}$ is injective. \begin{lem} \label{lemma442} Im $ \ac{\varepsilon} \subset \mbox{Ker} \ \ac{\Phi}^{\vee}$ \end{lem} \begin{proof} Straightforward computation. \end{proof} \subsection {\bf Identifying $H^2(\stackrel{2}{\wedge} E)$} We have $$H^2(\stackrel{2}{\wedge} E)^{\vee} \cong \mbox{Ker} \ \ol{\Phi}^{\vee}$$ where \qquad $ \ol{\Phi}^{\vee} \ : \ S^2(S_{k-2}\otimes V_{n-1}) \rightarrow \stackrel{2}{\wedge} S_{k-1} \otimes \stackrel{2}{\wedge} V_n $ \qquad is explicity given by \begin{tabbing} $\ol{\Phi}^{\vee}((g \otimes v)\cdot (g^1 \otimes v^1)) =$\=$sg \wedge sg^1 \otimes (tv \wedge tv^1) - sg \wedge tg^1 \otimes(tv \wedge sv^1)$\\ \>$-tg \wedge sg^1 \otimes (sv \wedge tv^1)+(tg \wedge tg^1) \otimes (sv \wedge sv^1)$ \end{tabbing} i.e. \qquad $\ol{\Phi}^{\vee} = \ol{p} \circ(S^2 \beta)$ \qquad where \pagebreak \begin{tabbing} $\ol{p} \ : \ $\=$S^2($\=$S_{k-1}\otimes V_n) \rightarrow \stackrel{2}{\wedge} S_{k-1} \otimes \stackrel{2}{\wedge} V_n$\\ \>\>$\|$\\ \>$(\stackrel{2}{\wedge} S_{k-1}\otimes \stackrel{2}{\wedge} V_n) \oplus (S^2 S_{k-1} \otimes S^2 V_n)$ \end{tabbing} is such that \[ \ol{p}((f \otimes u) \cdot (f' \otimes u^1)) = f \wedge f' \otimes u \wedge u^1 \] We consider the $SL(2)$-equivariant morphism: \[ \ol{\varepsilon}^1 \ :\ S^2(S_{k-3} \otimes V_{n-2}) \rightarrow S^2 (S_{k-2} \otimes V_{n-1}) \] such that: \begin{tabbing} $\ol{\varepsilon}^1((f \otimes u) \cdot (f' \otimes u^1)) =$\=$(sf \otimes tu) \cdot(sf' \otimes tu^1)-(sf \otimes su) \cdot (tf' \otimes tu^1)+$\\ \>$-(tf \otimes tu) \cdot (sf' \otimes su^1) + (sf \otimes tu) \cdot (sf' \otimes tu^1)$ \end{tabbing} ($\ol{\varepsilon}^1 = S^2 \beta^1 $ hence $ \ol{\varepsilon}^1$ is injective). Finally, we define \[ \ol{\varepsilon} \ : \ S^2 S_{k-3} \otimes S^2 V_{n-2} \rightarrow S^2 (S_{k-2} \otimes V_{n-1}) \] as $\ol{\varepsilon} = \ol{\varepsilon}^1 \circ \ol{i}$ \quad where \begin{tabbing} $\ol{i}\ : \ $\=$ S^2 S_{k-3} \otimes S^2V_{n-2} \rightarrow S^2 (S_{k-3} \otimes V_{n-2})$ \ such that \\ \>$f \cdot f' \otimes uu^1 \mapsto (f \otimes u)(f' \otimes u^1) + (f \otimes u^1)(f' \otimes u)$ \end{tabbing} is an injective map. Then, also $\ol{\varepsilon}$ is injective \begin{lem} \label{lemma443} Im $ \ol{\varepsilon} \subset \mbox{Ker} \ \ol{\Phi}^{\vee}$ \end{lem} \begin{proof} Straightforward computation. \end{proof} \begin{theorem} For any special symplectic instanton bundle $E$ \[ H^2(S^2E)\ \simeq\ \stackrel{2}{\wedge} (S_{k-3})^{\vee} \otimes S^2(V_{n-2})^{\vee}\] \end{theorem} \begin{proof} By lemma ~\ref{lemma442} and ~\ref{lemma443} we have the following diagram with exact rows and columns: \begin{minipage}{4in} \begin{tabbing} $0 \rightarrow H^2(\stackrel{2}{\wedge}$\=$ N)^{\vee} \rightarrow S^2(S_{k-2} \otimes$\=$V_{n-1}) \fr{\ol{\Phi}^{\vee}} \stackrel{2}{\wedge} S_{k-1} \otimes$\=$\stackrel{2}{\wedge} V_n \rightarrow H^1(\stackrel{2}{\wedge}$\=$N)^{\vee}\rightarrow 0$ \kill \> $0$ \> $0$ \> $0$ \> $0$\\ \>$\downarrow$ \> $\downarrow$ \> $\downarrow$\> $\downarrow$\\ $0 \rightarrow H^2(\stackrel{2}{\wedge} N)^{\vee} \rightarrow S^2(S_{k-2} \otimes V_{n-1}) \fr{\ol{\Phi}^{\vee}} \stackrel{2}{\wedge} S_{k-1} \otimes \stackrel{2}{\wedge} V_n \rightarrow H^1(\stackrel{2}{\wedge} N)^{\vee}\rightarrow 0$ \\ \>$\downarrow$ \> $\downarrow$ \> $\downarrow$\> $\downarrow$\\ $0 \rightarrow H^2(N \otimes N)^{\vee} \rightarrow S_{k-2} ^{\otimes 2} \otimes V_{n-1}^{\otimes 2} \fr{\Phi^{\vee}} S_{k-1}^{\otimes 2} \otimes \stackrel{2}{\wedge} V_n \rightarrow H^1(N \otimes N)^{\vee}\rightarrow 0$ \\ \>$\downarrow$ \> $\downarrow$ \> $\downarrow$\> $\downarrow$\\ $0 \rightarrow H^2(S^2 N)^{\vee} \rightarrow \stackrel{2}{\wedge} (S_{k-2} \otimes V_{n-1}) \fr{\ac{\Phi}^{\vee}} S^2 S_{k-1} \otimes \stackrel{2}{\wedge} V_n \rightarrow H^1(S^2 N)^{\vee}\rightarrow 0$ \\ \>$\downarrow$ \> $\downarrow$ \> $\downarrow$\> $\downarrow$\\ \> $0$ \> $0$ \> $0$ \> $0$\\ \end{tabbing} \end{minipage} \\ It was shown in \cite{OT} that: $H^2(EndE) \simeq{Ker}\ \Phi^{\vee} = H^2(N \otimes N)^{\vee} \simeq S_{k-3}^{\otimes 2} \otimes S^2 V_{n-2}$ We have proved that there are two injective maps: \[ \ac{\varepsilon}\ : \ \stackrel{2}{\wedge} (S_{k-3}) \otimes S^2V_{n-2} \rightarrow \mbox{Ker} \ \ac{\Phi}^{\vee} \simeq H^2(S^2N)^{\vee} \simeq H^2(S^2E)^{\vee} \] \[ \ol{\varepsilon}\ : \ S^2(S_{k-3}) \otimes S^2 V_{n-2} \rightarrow \mbox{Ker} \ \ol{\Phi}^{\vee} \simeq H^2(\stackrel{2}{\wedge} N)^{\vee} \simeq H^2(\stackrel{2}{\wedge} E)^{\vee} \] Then, we can consider the following diagram: \begin{tabbing} $0 \rightarrow S^2 S_{k-3} $\=$ \otimes S^2 V_{n-2} \rightarrow S_{k-3}^{\otimes 2} \otimes $\=$ S^2V_{n-2} \rightarrow \stackrel{2}{\wedge} S_{k-3}$\=$\otimes V_{n-2}$\kill \> $0$ \> $0$ \> $0$ \\ \> $\downarrow$ \> $\downarrow$ \> $\downarrow$ \\ $0 \rightarrow S^2 S_{k-3} \otimes S^2 V_{n-2} \rightarrow S_{k-3}^{\otimes 2} \otimes S^2V_{n-2} \rightarrow \stackrel{2}{\wedge} S_{k-3} \otimes S^2V_{n-2} \rightarrow 0 $\\ \> $\downarrow \ol{\varepsilon}$ \> $\downarrow \varepsilon$ \> $\downarrow \ac{\varepsilon}$ \\ \ \ \ \ \ \ \ \ \ $0\rightarrow $ \> $H^2(\stackrel{2}{\wedge} E)^{\vee} \rightarrow H^2(EndE)^{\vee} \rightarrow H^2(S^2E)^{\vee} \rightarrow 0 $ \\ \>\> $\downarrow$\\ \>\> $0$\\ \end{tabbing} and by the {\bf Snake-Lemma} there is the exact sequence : \begin{tabbing} $0 \rightarrow$ Ker \ \= $\ol{\varepsilon} \rightarrow $ Ker\ \= $\varepsilon \rightarrow $ Ker \= $\ac{\varepsilon} \rightarrow \mbox{Coker}\ \ol{\varepsilon} \rightarrow \mbox{Coker}\ $\=$ \varepsilon \rightarrow \mbox{Coker}\ \ac{\varepsilon} \rightarrow 0$\\ \> $\|$ \> $\|$ \> $\|$ \> $\|$ \\ \> $0$ \> $0$ \> $0$ \> $0$\\ \end{tabbing} $ \Rightarrow \mbox{Coker} \ \ol{\varepsilon} = 0 \Rightarrow \ol{\varepsilon}\ \mbox{is an isomorphism} \ \Rightarrow \ac{\varepsilon}\ \mbox{is an isomorphism}. $ \\ Thus: \[H^2(S^2E)^{\vee}\ \cong\ \stackrel{2}{\wedge} (S_{k-3}) \otimes S^2(V_{n-2})\] i.e. \qquad \ $H^2(S^2E)\ \simeq\ \stackrel{2}{\wedge} (S_{k-3})^{\vee} \otimes S^2(V_{n-2})^{\vee}$ \quad as we wanted.\\ \end{proof} \begin{oss} By this theorem formula ~\ref{zero} and theorem ~\ref{acca1} are easily proved. \end{oss}
1997-07-09T16:28:38
9707
alg-geom/9707005
en
https://arxiv.org/abs/alg-geom/9707005
[ "alg-geom", "math.AG" ]
alg-geom/9707005
Jesper Funch Thomsen
Jesper Funch Thomsen
Irreducibility of \bar{M}_{0,n}(G/P,\beta)
8 pages, AmsLaTeX
null
null
null
null
Let G be a linear algebraic group, P be a parabolic subgroup of G and \beta be a cycle of dimension 1 in the Chow group of the quotient G/P. Using geometric arguments and Borel's fixed point theorem, we prove that the moduli space \bar{M}_{0,n}(G/P, \beta) of n-pointed genus 0 stable maps representing \beta is irreducible.
[ { "version": "v1", "created": "Wed, 9 Jul 1997 14:28:33 GMT" } ]
2008-02-03T00:00:00
[ [ "Thomsen", "Jesper Funch", "" ] ]
alg-geom
\section{Introduction} Let $G$ be a complex connected linear algebraic group, $P$ be a parabolic subgroup of $G$ and $\beta \in A_1(G/P)$ be a 1-cycle class in the Chow group of G/P. An $n$-pointed genus $0$ stable map into $G/P$ representing the class $\beta$, consists of data $( \mu : C \rightarrow X ; p_1, \dots , p_n )$, where $C$ is a connected, at most nodal, complex projective curve of arithmetic genus $0$, and $\mu$ is a complex morphism such that $\mu_*[C] = \beta$ in $A_1(G/P)$. In addition $p_i$, $i=1, \dots, n$ denote $n$ nonsingular marked points on $C$ such that every component of $C$, which by $\mu$ maps to a point, has at least 3 points which is either nodal or among the marked points (this we will refer to as every component of $C$ being stable). The set of $n$-pointed genus $0$ stable maps into $G/P$ representing the class $\beta$, is parameterized by a coarse moduli space $\overline{M}_{0,n}(G/P,\beta)$. In general it is known that $\overline{M}_{0,n}(G/P,\beta)$ is a normal complex projective scheme with finite quotient singularities. In this paper we will prove that $\overline{M}_{0,n}(G/P,\beta)$ is irreducible. It should also be noted that we in addition will prove that the boundary divisors in $\overline{M}_{0,n}(G/P,\beta)$ , usually denoted by $D(A,B,\beta_1, \beta_2)$ ($\beta = \beta_1 + \beta_2$, $A \cup B$ a partition of $\{ 1, \dots , n \}$) , are irreducible. After this work was carried out we learned that B. Kim and R. Pandharipande \cite{KimPan} had proven the same results, and even proved connectedness of the corresponding moduli spaces in higher genus. Our methods however differs in many ways. For example in this paper we consider the action of a Borel subgroup of $G$ on $\overline{M}_{0,n}(G/P,\beta)$, while Kim and Pandharipande mainly concentrate on maximal torus action. Another important difference is that we in this presentation proceed by induction on $\beta$. This means that the question of $\overline{M}_{0,n}(G/P,\beta)$ being irreducible, can be reduced to simple cases. This work was carried out while I took part in the program ``Enumerative geometry and its interaction with theoretical physics'' at the Mittag-Leffler Institute. I would like to use this opportunity to thank the Mittag-Leffler Institute for creating a stimulating atmosphere. Thanks are also due to Niels Lauritzen and S{\o}ren Have Hansen for useful discussions concerning the generalization from full flag varieties to patial flag varieties. \section{Summary on $\overline{M}_{0,n}(G/P,\beta)$} \label{summary} In this section we will summarize the properties of the coarse moduli space $\overline{M}_{0,n}(G/P,\beta)$ which we will make use of. The notes on quantum cohomology by W. Fulton and R. Pandharipande \cite{FulPan} will serve as our main reference. As mentioned in the introduction the moduli space $\overline{M}_{0,n}(G/P,\beta)$ parameterizes $n$-pointed genus $0$ stable maps into $G/P$ representing the class $\beta$. By definition $\beta$ is effective if it is represented by some $n$-pointed genus $0$ stable map. In the following we will only consider values of $n$ and $\beta$ where $\overline{M}_{0,n}(G/P,\beta)$ is non-empty. This means $\beta$ must be effective and $n \geq 0$, and if $\beta = 0$ we must have $n \geq 3$. The moduli space $\overline{M}_{0,n}(G/P,\beta)$ is known to be a normal projective scheme (see \cite{FulPan}). This implies that $\overline{M}_{0,n}(G/P,\beta)$ splits up into a finite disjoint union of its components. This we will use several times. \subsection{Contraction morphism} On $\overline{M}_{0,n+1}(G/P,\beta)$ we have a contraction morphism $$ \overline{M}_{0,n+1}(G/P,\beta) \rightarrow \overline{M}_{0,n}(G/P,\beta)$$ which ``forget'' the $(n+1)$'th marked point. The contraction morphisms value on a closed point in $\overline{M}_{0,n+1}(G/P,\beta)$, represented by $(\mu : C \rightarrow G/P; p_1, \dots , p_{n+1})$, is the point in $\overline{M}_{0,n}(G/P,\beta)$ represented by $( \mu^{\circ} : C^{\circ} \rightarrow G/P; p_1 , \dots , p_n)$, where $C^{\circ}$ denote $C$ with the unstable components collapsed, and $\mu^{\circ}$ is the map induced from $\mu$. From the construction of $\overline{M}_{0,n}(G/P,\beta)$ it follows, that the contraction map is a surjective map with connected fibres. \subsection{Evaluation map} For each element $a \in \{1 , \dots n \} $ we have an evaluation map $$ \delta_a :\overline{M}_{0,n}(G/P,\beta) \rightarrow G/P.$$ Its value on a closed point in $\overline{M}_{0,n}(G/P,\beta)$ corresponding to $(\mu :C \rightarrow G/P ; p_1, \dots , p_n)$ is defined to be $\mu(p_a)$. \subsection{Boundary} By a boundary point in $\overline{M}_{0,n}(G/P,\beta)$ we will mean a point which correspond to a reducible curve. Let $ A \cup B =\{1, \dots n \}$ be a partition of $\{ 1, \dots n \} $ in disjoint sets, and let $\beta_1, \beta_2 \in A_1(X)$ be effective classes such that $\beta = \beta_1 + \beta_2$. We will only consider the cases when $\beta_1 \neq 0$ (resp. $\beta_2 \neq 0$) or $|A| \geq 2$ (resp. $|B| \geq 2$). With these conditions on $\beta_1, \beta_2, A$ and $B$ we let $ D(A,B,\beta_1, \beta_2)$ denote the set of elements in $\overline{M}_{0,n}(G/P,\beta)$ where the corresponding curve $C$ is of the following form : \begin{itemize} \item $C$ is the union of (at most nodal) curves $C_A$ and $C_B$ meeting in a point. \item The markings of $A$ and $B$ lie on $C_A$ and $C_B$ respectively. \item $C_A$ and $C_B$ represent the classes $\beta_1$ and $\beta_2$ respectively. \end{itemize} Notice here that our restrictions on $A$,$B$,$\beta_1$ and $\beta_2$ is the stability conditions on $C_A$ and $C_B$. \noindent It is clear that every boundary element lies in at least one of these $D(A,B,\beta_1,\beta_2)$. The sets $D(A,B,\beta_1, \beta_2)$ are in fact closed, and we will regard them as subschemes of $\overline{M}_{0,n}(G/P,\beta)$ by giving them the reduced scheme structure. Closely related to $D(A,B,\beta_1, \beta_2)$ is the scheme $M(A,B, \beta_1, \beta_2)$ defined by the fibre square $$ \begin{CD} M(A,B,\beta_1, \beta_2) @>p_2>> \overline{M}_{0,A \cup \{ \centerdot \} }(G/P,\beta_1) \times \overline{M}_{0,B \cup \{ \centerdot \} }(G/P,\beta_2) \\ @Vp_1VV @VV\delta_{\centerdot}^A \times \delta_{\centerdot}^BV \\ G/P @>\Delta >> G/P \times G/P \end {CD} $$ Here $\Delta$ is the diagonal embedding and $\delta_{\centerdot}^A$ and $\delta_{\centerdot}^B$ denotes the evaluation maps with respect to the point $ \{ \centerdot \}$. In \cite{FulPan} it is proved that $M(A,B,\beta_1,\beta_2)$ is a normal projective variety and that we have a canonical map $$ M(A,B,\beta_1,\beta_2) \longrightarrow D(A,B,\beta_1, \beta_2). $$ This map is clearly surjective. As $M(A,B,\beta_1,\beta_2)$ is a closed subscheme of $\overline{M}_{0,A \cup \{ \centerdot \} }(G/P,\beta_1) \times \overline{M}_{0,B \cup \{ \centerdot \} }(G/P,\beta_2)$ we can regard the closed points of $M(A,B,\beta_1, \beta_2)$ as elements of the form $(z_1,z_2)$, where $z_1 \in \overline{M}_{0,A \cup \{ \centerdot \} }(G/P,\beta_1)$ and $z_2 \in \overline{M}_{0,B \cup \{ \centerdot \} }(G/P,\beta_2)$. The image of $(z_1,z_2)$ in $D(A,B,\beta_1,\beta_2)$ will then be denoted by $z_1 \sqcup z_2$. Given $z_1 \in \overline{M}_{0,A \cup \{ \centerdot \} }(G/P,\beta_1)$, $z_2 \in \overline{M}_{0,B \cup \{ \centerdot \} \cup \{ * \}}(G/P,\beta_2)$ and $z_3 \in \overline{M}_{0,C \cup \{ * \} }(G/P,\beta_3)$ , with $\delta_{\centerdot}(z_1) = \delta_{\centerdot}(z_2)$ and $\delta_*(z_2) = \delta_*(z_3)$, we then have the identity $(z_1 \sqcup z_2) \sqcup z_3 = z_1 \sqcup (z_2 \sqcup z_3)$ inside $ \overline{M}_{0,A \cup B \cup C}(G/P,\beta_1+ \beta_2 +\beta_3)$. \subsection{G-action} As mentioned in the introduction we have a $G$-action $$G \times \overline{M}_{0,n}(G/P,\beta) \rightarrow \overline{M}_{0,n}(G/P,\beta).$$ On closed points we can describe the action in the following way. Let $x \in \overline{M}_{0,n}(G/P,\beta)$ be a closed point corresponding to the data $(\mu : C \rightarrow G/P ; p_1 , \dots , p_n )$, and let $g$ be a closed point in $G$. Then $g \cdot x$ is the point in $\overline{M}_{0,n}(G/P,\beta)$ corresponding to $(\mu_g : C \rightarrow G/P ; p_1 , \dots , p_n )$, where $\mu_g = (g \cdot) \circ \mu$. Here $g \cdot$ denotes multiplication with $g$ on $G/P$. \subsection{Special cases} The following special cases of our main result follows from the construction and formal properties of our moduli spaces. \noindent $\beta = 0$ : Here the moduli space $\overline{M}_{0,n}(G/P,\beta)$ is canonical isomorphic to $\overline{M}_{0,n} \times G/P$, where $\overline{M}_{0,n}$ denote the moduli space of stable $n$-pointed curves of genus $0$. As $\overline{M}_{0,n}$ is known to be irreducible \cite{Knudsen} we get that $\overline{M}_{0,n}(G/P,0)$ is irreducible. \noindent $G/P = \P^1$ : The irreducibility of $\overline{M}_{0,n}(\P^1,d)$ follows from the construction of the moduli space in \cite{FulPan}. First of all $\overline{M}_{0,0}(\P^1, 1) \cong \operatorname{Spec} ({\Bbb C})$ so we may assume that $(n,d) \neq (0,1)$. With this assumption $\overline{M}_{0,n}(\P^1, d)$ is the quotient of a variety $M$ by a finite group. Now $M$ is glued together by the moduli spaces $\overline{M}_{0,n}(\P^1,d,\overline{t})$ of $\overline{t}$-maps spaces (here $\overline{t}=(t_0,t_1)$ is a basis of $\O_{\P^1}(1)$). See section 3 in \cite{FulPan} for a definition of $\overline{M}_{0,n}(\P^1, d, \overline{t})$. The moduli spaces $\overline{M}_{0,n}(\P^1, d, \overline{t})$ are irreducible (in fact they are ${\Bbb C}^*$-bundles over an open subscheme of $\overline{M}_{0,m}$ for a suitable $m$). This follows from the proof of Proposition 3.3 in \cite{FulPan}. It is furthermore clear that $\overline{M}_{0,n}(\P^1, d, \overline{t})$, and $\overline{M}_{0,n}(\P^1, d, \overline{t}')$ intersect non-trivially for different choices of bases $\overline{t}$ and $\overline{t}'$. This imply that $\overline{M}_{0,n}(\P^1,d)$ is connected, and as it is locally normal it must be irreducible. \section{Flag varieties} \label{flag} In this section we will give a short review on flag varieties. Main references will be \cite{Springer}, \cite{Demazure} and \cite{Kock}. In \cite{Springer} one can find the general theory on the structure of linear algebraic groups. The Chow group of $G/B$, where $B$ is a Borel subgroup, can be found in \cite{Demazure}. From this one easily recovers the Chow group for a general flag variety $G/P$ (e.g. \cite{Kock} Section 1). \subsection{Schubert varieties.} Let $G$ be a complex connected linear algebraic group and $P$ be a parabolic subgroup of $G$. As we will only be interested in the quotient $G/P$, we may assume that $G$ is semisimple. Fix a maximal torus $T$ and a Borel subgroup $B$ such that $$ T \subseteq B \subseteq P \subseteq G. $$ Let $W$ (resp. $R$) denote the Weyl group (resp. roots) associated to $T$ and let $R^+$ denote the positive roots with respect to $B$. Let further $D \subseteq R^+$ denote the simple roots. Given $\alpha \in R$ we let $s_{\alpha} \in W$ denote the corresponding reflection. From general theory on algebraic groups we know that $P$ is associated to a unique subset $I \subseteq D$, such that $P = B W_I B$, where $W_I$ is the subgroup of $W$ generated by the reflections $s_{\alpha}$ with $\alpha \in I$. The flag variety $G/P$ is then the disjoint union of a finite number of $B$-invariant subsets $C(w) = B w P / P$ with $w \in W^I$, where $$W^I = \{ w \in W | w \alpha \in R^+ \text{ for all } \alpha \in I \}.$$ Each $C(w)$, $w \in W^I$ is isomorphic to ${\Bbb A}^{l(w)}$. Here $l(w)$ denotes the length of a shortest expression of $w$ as a product of simple reflections $s_{\alpha}$, $\alpha \in D$. The closures of $C(w)$, $w \in W$, inside $G/P$ is called the generalized Schubert varieties. We will denote them by $X_w$, $w \in W$, respectively. In case $l(w)=1$ we have $X_w \cong \P^{\,1}$. \subsection{Chow group.} The Chow group $A_*(G/P)$ is freely generated. As a basis we can pick $[X_w]$, $w \in W^I$. In \cite{Kock} it is proved that this basis is orthogonal. Using that positive classes intersect in positive classes on G/P (Cor. 12.2 in \cite{FulInt}), we conclude that a class in $A_*(G/P)$ is positive (or zero) if and only if it is of the form $$ \sum _{w \in W^I} a_w [X_w] \text{ with } a_w \geq 0.$$ \subsection{Effective classes.} Let $\beta \in A_1(G/P)$. From above it is clear that $\beta$ can only be effective (in the sense of Section \ref{summary}), if $\beta$ is a positive linear combination of $[X_{s_{\alpha}}]$ with $\alpha \in D \cap W^I = D \setminus I$. Noticing that $X_{s_{\alpha}} \cong \P^{\,1}$, $\alpha \in D \setminus I$, implies the inverse, that is, a positive linear combination of $[ {X_{s_{\alpha}}}]$, $\alpha \in D \setminus I$ is effective. Using the above we can introduce a partial ordering on the set of effective classes in $A_1(G/P)$. \begin{defn} Let $\beta_1$ and $\beta_2$ be effective classes. If there exist an effective class $\beta_3$ such that $\beta_2 = \beta_1 + \beta_3$ we write $\beta_1 \prec \beta_2$. If $\beta$ is an effective class with the property $$\beta' \prec \beta \Rightarrow \beta' = 0 \text{ or } \beta' = \beta$$ we say that $\beta$ is irreducible. An effective class $\beta$ is reducible if it is not irreducible. \end{defn} Notice that a non-zero effective class $\beta$ is irreducible if and only if $\beta = [X_{s_{\alpha}}]$ for some $\alpha \in D \setminus I$. In the proof of the irreducibility of $\overline{M}_{0,n}(G/P,\beta)$ we will use induction on $\beta$ with respect to this ordering. This is possible because given an effective class $\beta \in A_1(G/P)$, there is only finitely many other effective classes $\beta'$ with $\beta' \prec \beta$. \subsection{Summary.} We are ready to summarize what will be important for us \begin{itemize} \item The set of effective classes in $A_1(G/P)$ has a ${\Bbb Z}_{\geq 0}$-basis represented by $B$-invariant closed subvarieties $X_{s_{\alpha}}$, $\alpha \in D \setminus I$, of $G/P$. \item The subsets $X_{s_{\alpha}}$, $\alpha \in D \setminus I$ are the only $B$-invariant irreducible 1-dimensional closed subsets of $G/P$. \item $X_{s_{\alpha}} \cong \P^{\,1}$, $\alpha \in D \setminus I$. \end{itemize} \section{Boundary of $\overline{M}_{0,n}(X,\beta)$} \label{boundary} In this section we begin the proof of our main result. Remember that our convention is that whenever we write $\overline{M}_{0,n}(G/P,\beta)$, $D(A,B,\beta_1,\beta_2)$ or $M(A,B,\beta_1,\beta_2)$, we assume that these are well defined and non-empty. From now on we will assume that $G$, a semisimple linear algebraic group, and a parabolic subgroup $P$ have been fixed. We let $X$ denote $G/P$. We will need to know when $D(A,B,\beta_1,\beta_2)$ is irreducible and for this purpose we have the following proposition. \begin{prop} \label{divisor} Suppose that $\overline{M}_{0,A \cup \{ \centerdot \} }(X,\beta_1)$ and $\overline{M}_{0,B \cup \{ \centerdot \} }(X,\beta_2)$ are irreducible. Then the scheme $M(A,B,\beta_1,\beta_2)$ is also irreducible. In particular $D(A,B,\beta_1,\beta_2)$ will be irreducible in this case. \begin{pf} As $M(A,B,\beta_1,\beta_2)$ is a normal scheme it splits up into a disjoint union of irreducible components $C_1,C_2, \dots, C_l$. Our task is to show that $l=1$. Consider the natural map $\pi : G \rightarrow G/P$. Locally (in the Zariski topology) this map has a section (\cite{Jantzen} p.183) , i.e. there exists an open cover $ \{U_i \}_{i \in I}$ of $X$ (we assume $U_i \neq \emptyset$) and morphisms $s_i : U_i \rightarrow G$ such that $\pi \circ s_i$ is the identity map. By pulling back the covering $ \{U_i \}_{i \in I}$ of $X$, by the evaluation maps $\delta_{\centerdot}^A$ and $\delta_{\centerdot}^B$, we get open coverings $ \{ V_i^A \}_{i \in I} $ and $ \{ V_i^B \}_{i \in I} $ of $\overline{M}_{0,A \cup \{ \centerdot \} }(X,\beta_1)$ and $\overline{M}_{0,B \cup \{ \centerdot \} }(X,\beta_2)$ respectively. Finally an open cover $ \{W_i \}_{i \in I}$ of $M(A,B,\beta_1,\beta_2)$ is obtained by setting $W_i = p_1^{-1}(U_i)= p_2^{-1}(V_i^A \times V_i^B)$. We claim $$\forall i,j \in I : W_i \cap W_j \neq \emptyset.$$ To see this consider $U_i,U_j \subseteq X$. As $X$ is irreducible there exists a closed point $x \in U_i \cap U_j$. Using that $G$ acts transitively on $X$ we can choose elements $z_1 \in \overline{M}_{0,A \cup \{ \centerdot \} }(X,\beta_1)$ and $z_2 \in \overline{M}_{0,B \cup \{ \centerdot \} }(X,\beta_2)$, with $\delta_{\centerdot}^A(z_1) = \delta_{\centerdot}^B(z_2)=x$. With these choices it is clear that $(z_1,z_2)$ correspond to a point in $W_i \cap W_j$. Next we want to show that $W_i$ is irreducible. For this consider the map $$ \begin{array}{cccc} \psi_i : & V_i^A \times V_i^B & \rightarrow & \overline{M}_{0,A \cup \{ \centerdot \} }(X,\beta_1) \times \overline{M}_{0,B \cup \{ \centerdot \} }(X,\beta_2) \vspace{2ex}\\ & (z_1,z_2) & \mapsto & (z_1, ((s_i \circ \delta_{\centerdot}^A)(z_1)) ((s_i \circ \delta_{\centerdot}^B)(z_2))^{-1} z_2) \end{array} $$ where we use the group action of $G$ on $\overline{M}_{0,B \cup \{ \centerdot \} }(X,\beta_2)$. By definition $\psi_i$ factors through $W_i$. We therefore have an induced map $$\psi_i^{'} : V_i^A \times V_i^B \rightarrow W_i. $$ Clearly $\psi_i^{'} \circ p_2$ is the identity map. This implies that $\psi_i^{'}$ is surjective, and as $V_i^A \times V_i^B$ is irreducible, we get that $W_i$ is irreducible. At last we notice that as $W_i$ is irreducible it must be contained in one of the components $C_1,C_2, \dots ,C_l$ of $M(A,B,\beta_1, \beta_2)$. On the other hand the $W_i$'s intersect non-trivially so all of them must be contained in the same component. But $ \{ W_i \} _{i \in I }$ was an open cover of $M(A,B,\beta_1,\beta_2)$. We conclude that $l = 1$, as desired. Being a surjective image of $M(A,B,\beta_1,\beta_2)$ this implies that $D(A,B,\beta_1,\beta_2)$ is also irreducible. \end{pf} \end{prop} \section{ Properties of the components of $\overline{M}_{0,n} (X,\beta)$} In this section we study the behaviour of the components of $\overline{M}_{0,n}(X,\beta)$. Let $K_1, K_2, \dots, K_l$ denote the components of $\overline{M}_{0,n}(X,\beta)$. As $ \overline{M}_{0,n}(X,\beta)$ is normal, the $K_i$'s are disjoint. Remember that we had a group action of $G$ on $\overline{M}_{0,n}(X,\beta)$ which was introduced in Section \ref{summary}. We claim \begin{lem} \label{inv} Let $K$ be a component of $\overline{M}_{0,n}(X,\beta)$. Then $K$ is invariant under the group action of $G$ on $\overline{M}_{0,n} (X,\beta)$. \begin{pf} Let $\eta : G \times \overline{M}_{0,n}(X,\beta) \rightarrow \overline{M}_{0,n}(X,\beta)$ denote the group action, and consider the image $\eta(G \times K)$ of $G \times K$. As $G \times K$ is irreducible $\eta(G \times K)$ will also be irreducible. This means that $\eta(G \times K)$ is contained in a component, say $K_1$, of $\overline{M}_{0,n}(X,\beta)$. On the other hand $\eta( \{ e \} \times K) \subseteq K$ (here $e$ denotes the identity element in $G$) so we conclude that $K=K_1$. \end {pf} \end{lem} The next lemma concerns the boundary of components of $\overline{M}_{0,0} (X,\beta)$ when $\beta$ is reducible. \begin{lem} \label{bound} Let $\beta$ be a reducible effective class and $K$ be a component of $\overline{M}_{0,0}(X,\beta)$. Then there exist boundary elements in $K$. \begin{pf} Assume $K$ do not have boundary elements. Then by definition of boundary points, each element in $K$ would correspond to an irreducible curve. Using Lemma \ref{inv} we have an induced $B$-action on $K$. As $K$ is projective, and $B$ is a connected solvable linear algebraic group, we can use Borel's fixed point Theorem (see \cite{Springer} p.159) to conclude that this action has a fixed point. This means that there exist $z \in K$ such that $b z = z$, for all $b \in B$. Let $\P^{\,1} \stackrel{\mu}{\rightarrow} X$ be the stable curve with its morphism to $X$ which correspond to $z$. By definition of the group action of $B$ on $z$ we conclude that $\mu(\P^{\,1})$ must be a $B$-invariant subset of $X$. On the other hand $\mu (\P^{\,1})$ is closed, irreducible and of dimension 1. By the properties stated in Section \ref{flag} $\mu(\P^{\,1})$ must be equal to a 1-dimensional Schubert variety $X_{s_{\alpha}}$ of $X$. From this we conclude that $\mu_*[\P^{\,1}] = m [X_{s_{\alpha}}]$, where $m$ is a positive integer. As $\beta$ is reducible $m \geqq 2$. The closed embedding $i : X_{s_{\alpha}} \rightarrow X$ induces a map $i_* : \overline{M}_{0,0}(X_{s_{\alpha}}, m [X_{s_{\alpha}}]) \rightarrow \overline{M}_{0,0}(X,\beta)$, where an element $(C \stackrel{f} {\rightarrow} X_{s_{\alpha}}) \in \overline{M}_{0,0} (X_{s_{\alpha}}, m [X_{s_{\alpha}}])$ goes to $i_*(C \stackrel{f} {\rightarrow} X_{s_{\alpha}}) = (C\stackrel{i \circ f}{\rightarrow} X)$. As $X_{s_{\alpha}}$ is isomorphic to $\P^{\,1}$ we know that $ \overline{M}_{0,0}(X_{s_{\alpha}}, m [{X}_{s_{\alpha}}])$ is irreducible. On the other hand $z$ is in the image of $i_*$ so we conclude that $i_*( \overline{M}_{0,0}(X_{s_{\alpha}}, m [X_{s_{\alpha}}])) \subseteq K$. But a boundary element in $\overline{M}_{0,0}(X_{s_{\alpha}}, m [ X_{s_{\alpha}}])$ is easy to construct by hand (as $m \geq 2$), which gives us the desired contradiction. \end{pf} \end{lem} The following will also be useful. \begin{lem} \label{hit} Let $\beta \in A_1(X)$ be a reducible effective class and suppose that $\overline{M}_{0,0}(X,\beta')$ is irreducible for $\beta' \prec \beta$. Furthermore let $K$ be a component of $\overline{M}_{0,0}(X,\beta)$. Then there exists a non-zero irreducible class $\beta'$, with $\beta -\beta'$ effective, such that $D(\emptyset, \emptyset, \beta', \beta - \beta') \cap K \neq \emptyset$. \begin{pf} By Lemma \ref{bound} we can choose a boundary point $z \in K$. There exists effective classes $\beta_1$ and $\beta_2$ such that $z \in D(\emptyset, \emptyset, \beta_1, \beta_2)$. We may assume that $\beta_1$ is reducible. Choose an effective non-zero irreducible class $\beta'$ and an effective class $\beta''$ such that $\beta_1 = \beta' + \beta''$. Choose also $z_1 \in \overline{M}_{0, \{ Q_1 \} }(X,\beta')$, $z_2 \in \overline{M}_{0, \{ Q_1 \} \cup \{ Q_2 \} }(X,\beta'')$ and $z_3 \in \overline{M}_{0, \{ Q_2 \} }(X,\beta_2)$, such that $\delta_{Q_1}(z_1) = \delta_{Q_1}(z_2)$ and $\delta_{Q_2}(z_2) = \delta_{Q_2}(z_3)$. Then $z_1 \sqcup z_2 \in \overline{M}_{0, \{ Q_2 \} }(X,\beta_1)$ from which we conclude $(z_1 \sqcup z_2) \sqcup z_3 \in D(\emptyset, \emptyset, \beta_1, \beta_2)$. On the other hand $z_2 \sqcup z_3 \in \overline{M}_{0, \{ Q_1 \} }(X,\beta - \beta')$ by which we conclude $z_1 \sqcup (z_2 \sqcup z_3) \in D(\emptyset, \emptyset, \beta', \beta - \beta')$. Using Proposition \ref{divisor} we know that $ D(\emptyset, \emptyset, \beta_1, \beta_2)$ is irreducible and as $z \in D(\emptyset, \emptyset,\beta_1, \beta_2) \cap K$, we must have $ D(\emptyset, \emptyset,\beta_1, \beta_2) \subseteq K$, in particular $(z_1 \sqcup z_2) \sqcup z_3 \in K$. On the other hand $$ (z_1 \sqcup z_2) \sqcup z_3 = z_1 \sqcup ( z_2 \sqcup z_3) \in D(\emptyset, \emptyset, \beta', \beta - \beta').$$ This proves the lemma. \end{pf} \end{lem} \section{Irreducibility of $\overline{M}_{0,n}(X,\beta)$} In this section we will prove that the moduli spaces $\overline{M}_{0,n}(X,\beta)$ are irreducible. First we notice that for $\beta \neq 0$ we can restrict our attention to a fixed $n$. \begin{lem} \label{n-inv} Let $n_1,n_2 \geqq 0$ be integers and $\beta \in A_1(X) \setminus \{ 0 \}$ be an effective class. Then $\overline{M}_{0,n_1}(X,\beta)$ is irreducible if and only if $\overline{M}_{0,n_2}(X,\beta)$ is irreducible. \begin{pf} It is enough to consider the case $n_2=n+1$ and $n_1=n$ for a positive integer $n$. The contraction morphism $f : \overline{M}_{0,n+1}(X,\beta) \rightarrow \overline{M}_{0,n}(X,\beta)$ which forgets the $(n+1)$'th point is a surjective map with connected fibres. Let $K_1,K_2, \dots, K_s$ (resp. $C_1,C_2, \dots, C_t$) be the components of $\overline{M}_{0,n+1}(X,\beta)$ (resp. $\overline{M}_{0,n}(X,\beta)$). As the components are mutually disjoint and $f$ is surjective we must have $s \geqq t$. Let us now restrict our attention to one of the components of $\overline{M}_{0,n}(X,\beta)$, say $C_1$. Assume that $K_1,K_2, \dots, K_r$ ($r \leqq s$) are the components which by $f$ maps to $C_1$. It will be enough to show that $r=1$. Assume $r \geqq 2$. As $$ C_1 = \bigcup_{i=1}^{r} f(K_i) $$ and as $C_1$ is irreducible, at least one of the components $K_1 ,K_2, \dots, K_r$ maps surjectively onto $C_1$. So there must exist a point $x$ in $C_1$ which is in the image of at least 2 of the components in $\overline{M}_{0,n+1}(X,\beta)$. But then the fibre of $f$ over $x$ is not connected, which is a contradiction. \end{pf} \end{lem} The idea in proving the irreducibility of $\overline{M}_{0,n}(X,\beta)$ is to use induction on the class $\beta \in A_1(X)$. By this we mean that we will prove that $\overline{M}_{0,n}(X,\beta)$ is irreducible assuming the same condition is true for $\beta' \prec \beta$. The first step in the induction procedure will be to show that $\overline{M}_{0,0}(X,\beta)$ is irreducible, when $\beta$ is a non-zero irreducible class. \begin{lem} \label{nonred} Let $\beta$ be a non-zero irreducible class. Then $\overline{M}_{0,0}(X,\beta)$ is irreducible. \begin{pf} As $\beta$ is a non-zero irreducible class, $\beta$ must be the class of a 1-dimensional Schubert variety $X_{s_{\alpha}}$. Let $K$ be a component of $\overline{M}_{0,0}(X,\beta)$. As in the proof of Lemma \ref{bound} we have a $B$-action on $K$, which by Borel's fixed point theorem is forced to have a fixed point. Let $x \in K$ be a fixed point. As $\beta$ is irreducible $x$ must correspond to an irreducible curve, i.e. $x$ correspond to a map of the form $ \P^{\,1} \stackrel{\mu}{\rightarrow} X$. The image $\mu(\P^{\,1})$ is a closed 1-dimensional $B$-invariant irreducible subset of $X$. As by assumption $\mu_*[\P^{\,1}] = [\overline {X}_{s_{\alpha}}]$, we conclude that $\mu(\P^{\,1}) = X_{s_{\alpha}}$. Now $X_{s_{\alpha}}$ is isomorphic to $\P^{\,1}$, so $\mu$ must be an isomorphism onto its image. But clearly every map $ \P^{\,1} \stackrel{f}{\rightarrow} X$ with $f(\P^{\,1})= X_{s_{\alpha}}$, which is an isomorphism onto its image, represent the same point in $\overline{M}_{0,0}(X,\beta)$. Above we have shown that this point belongs to every component of $\overline{M}_{0,0}(X,\beta)$. Using that the components of $\overline{M}_{0,0}(X,\beta)$ are disjoint the lemma follows. \end{pf} \end{lem} Now we are ready for the general case. \begin{thm} \label{thm} Let $\beta \in A_1(X)$ be an effective class and $X = G/P$ be a flag variety. Then $\overline{M}_{0,n}(X,\beta)$ is irreducible for every positive integer n. \begin{pf} The case $\beta = 0 $ is trivial as noted in Section \ref{summary}. By Lemma \ref{n-inv} we may therefore assume that $n = 0$. As remarked above we will proceed by induction. Assume that the theorem has been proven for $\beta '$ with $\beta ' \prec \beta$. Referring to Lemma \ref{nonred} we may assume that $\beta$ is reducible. Write $\beta = \sum_{i=1}^{m} \beta_i$ as a sum of non-zero irreducible effective classes $\beta_i$. Then $m \geq 2$. We divide into 2 cases. Assume first that $m=2$. So $\beta = \beta' + \beta''$, where $\beta'$ and $\beta''$ are effective irreducible classes. In this case every boundary element lie in $D(\emptyset , \emptyset, \beta', \beta'')$, which we by induction know is irreducible (Proposition \ref{divisor}). On the other hand do every component of $\overline{M}_{0,0} (X,\beta)$ contain a boundary point (by Lemma \ref{bound}). Using that the components of $\overline{M}_{0,0}(X,\beta)$ are disjoint, the theorem follows in this case. Assume therefore that $m \geq 3$. For each $i=1, \dots,m$ choose $z_i \in \overline{M}_{0,\{ Q_i \} }(X,\beta_i)$, a point such that $\delta_{Q_i} (z_i)= eP$, where $\delta_{Q_i}$ is the evaluation map onto $X$. Let $Q = \{ Q_1, Q_2, \dots, Q_m\}$, and choose a point $z_0 \in \overline{M}_{0,Q}(X,0)$ corresponding to a curve $C \cong \P^{\,1}$ and a map $\mu : C \rightarrow X$ such that $\mu(C) = eP$. Define $$ z = z_0 \sqcup (\sqcup_{i=1}^{m} z_i) \in \overline{M}_{0,0}(X,\beta).$$ Then clearly $z \in D(\emptyset , \emptyset, \beta_i, \beta -\beta_i)$ for all $i$. Let $K$ be the component of $\overline{M}_{0,0}(X,\beta)$ which contains $z$. By the induction hypothesis and Proposition 1, $D(\emptyset , \emptyset, \beta_i, \beta - \beta_i)$ is irreducible for all $i$, which implies $D(\emptyset , \emptyset, \beta_i, \beta -\beta_i) \subseteq K$ for all $i$. On the other hand, by Lemma \ref{hit}, every component of $\overline{M}_{0,0}(X,\beta)$ will intersect at least one of the sets $D(\emptyset , \emptyset, \beta_i, \beta - \beta_i)$. Using, and now for the last time, that the components of $\overline{M}_{0,0}(X,\beta)$ are disjoint, the theorem follows. \end{pf} \end{thm} \begin{cor} Let $X=G/P$ be a flag variety. Then the boundary divisors $D(A,B,\beta_1,\beta_2)$ of $\overline{M}_{0,n}(X,\beta)$ are irreducible. \begin{pf} Use Proposition \ref{divisor} and Theorem \ref{thm}. \end{pf} \end{cor} \bibliographystyle{amsplain} \ifx\undefined\leavevmode\hbox to3em{\hrulefill}\, \newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,} \fi
1997-12-19T22:43:55
9712
alg-geom/9712022
en
https://arxiv.org/abs/alg-geom/9712022
[ "alg-geom", "math.AG" ]
alg-geom/9712022
Alexander Polishchuk
Alexander Polishchuk
Poisson structures and birational morphisms associated with bundles on elliptic curves
20 pages, AMSLatex
null
null
null
null
In this paper we define a Poisson structure on some moduli spaces related to principal G-bundles on elliptic curves, the simplest example being the moduli space of stable pairs: a vector bundle and its global section. We also study birational morphisms between projective spaces appearing as such moduli spaces.
[ { "version": "v1", "created": "Fri, 19 Dec 1997 21:43:54 GMT" } ]
2007-05-23T00:00:00
[ [ "Polishchuk", "Alexander", "" ] ]
alg-geom
\section{Stable triples} Let us recall the definition of stable triples from \cite{BG}. Let $T=(E_1,E_2,\Phi)$ be a triple consisting of two vector bundles $E_1$ and $E_2$ on $X$ and a homomorphism $\Phi:E_2\rightarrow E_1$. For a real parameter $\sigma$ the $\sigma$-degree of $T$ is defined as follows: $$\deg_{\sigma}(T)=\deg(E_1)+\deg(E_2)+\sigma\cdot\operatorname{rk}(E_2).$$ Now the $\sigma$-slope of $T$ is defined by the formula $$\mu_{\sigma}(T)=\frac{\deg_{\sigma}(T)}{\operatorname{rk}(E_1)+\operatorname{rk}(E_2)}.$$ Note that if $L$ is a line bundle then we can define a tensor of a triple $T$ with $L$ naturally, so that one has $\mu_{\sigma}(T\otimes L)=\mu_{\sigma}(T)+\deg L$. The triple $T$ is called $\sigma$-stable if for every non-zero proper subtriple $T'\subset T$ one has $\mu_{\sigma}(T')<\mu_{\sigma}(T)$. Sometimes it is convinient to introduce another stability parameter $\tau=\mu_{\sigma}(T)$. The category of triples $T=(E_1,E_2,\Phi)$ is equivalent to the category of extensions \begin{equation}\label{equivext} 0\rightarrow p^*E_1\rightarrow F\rightarrow p^*E_2(2)\ra0 \end{equation} on $X\times\P^1$ where $p:X\times\P^1\rightarrow X$ is the projection. Indeed, the space of such extensions is $\operatorname{Ext}^1_{X\times\P^1}(p^*E_2(2),p^*E_1)\simeq\operatorname{Hom}_X(E_2,E_1)$. This extension has a unique $\operatorname{SL}_2$-equivariant structure and as shown in \cite{BG} the $\sigma$-stability condition on $T$ is equivalent to the $\operatorname{SL}_2$-equivariant stability of $F$ with respect to some polarization on $X\times\P^1$ depending on $\sigma$. Let us denote by ${\cal M}_{\sigma}={\cal M}_{\sigma}(d_1,d_2,r_1,r_2)$ the moduli space of $\sigma$-stable triples $T=(E_1,E_2,\Phi)$ on $X$ with $\deg(E_i)=d_i$, $\operatorname{rk} E_i=r_i$. When using another stability parameter $\tau$ we will denote the same moduli space by ${\cal M}_{\tau}$. This moduli space can be constructed using geometric invariant theory as in \cite{Be}. We claim that in the case of elliptic curve all these moduli spaces are smooth. \begin{lem} The moduli space ${\cal M}_{\sigma}$ is smooth. \end{lem} \noindent {\it Proof}. According to \cite{BG} we have to show that $H^2(X\times\P^1,\underline{\operatorname{End}} F)^{\operatorname{SL}_2}=0$ for the $\operatorname{SL}_2$-equivariant vector bundle $F$ associated with a $\sigma$-stable triple. Consider the exact sequence $$0\rightarrow K\rightarrow \underline{\operatorname{End}} F\rightarrow\underline{\operatorname{Hom}}(p^*E_1,p^*E_2(2))\rightarrow 0.$$ Since the direct image $Rp_*$ of the last term will have no $\operatorname{SL}_2$-invariant part we have $H^*(X\times\P^1,\underline{\operatorname{End}} F)^{\operatorname{SL}_2}\simeq H^*(X\times\P^1,K)$. Now $K$ sits in the exact triangle $$\underline{\operatorname{Hom}}(p^*E_2(2),p^*E_1))\rightarrow K\rightarrow\underline{\operatorname{End}} p^*E_1\oplus \underline{\operatorname{End}} p^*E_2\rightarrow\underline{\operatorname{Hom}}(p^*E_2(2),p^*E_1)[1]\rightarrow\ldots$$ It follows that equivariant direct image of $K$ with respect to the projection $p$ is quasi-isomorphic to the complex \begin{equation}\label{tangcomp} C^{\cdot}:\underline{\operatorname{End}} E_1\oplus\underline{\operatorname{End}} E_2\stackrel{d}{\rightarrow} \underline{\operatorname{Hom}}(E_2,E_1) \end{equation} concentrated in degrees $0$ and $1$, where $d(A,B)=A\Phi-\Phi B$. We have the exact sequence of cohomologies $$H^1(X,\underline{\operatorname{End}}E_1\oplus\underline{\operatorname{End}} E_2)\rightarrow H^1(X,\underline{\operatorname{Hom}}(E_2,E_1))\rightarrow H^2(X, C^{\cdot})\rightarrow 0.$$ Now by Serre duality we have $H^1(X,\underline{\operatorname{Hom}}(E_2,E_1))^*\simeq H^0(X,\underline{\operatorname{Hom}}(E_1,E_2))$. According to Lemma 4.4 of \cite{BG} this space is zero unless $\Phi$ is an isomorphism. In the latter case the first arrow in the above exact sequence is surjective, so in either case we get $H^2(X\times\P^1,K)=H^2(X, C^{\cdot})=0$. \qed\vspace{3mm} The proof of this lemma also shows that the tangent space to ${\cal M}_{\sigma}$ at a triple $T$ is identified with the hypercohomology space $H^1(X,C^{\cdot})$ where $C^{\cdot}$ is the complex (\ref{tangcomp}). This can also be shown directly considering infinitesemal deformations of the first order for triples. \section{Poisson structure} \label{mainsec} Let us fix a trivialization $\omega_X\simeq\O_X$ of the canonical bundle of $X$. Then we can define a Poisson structure on the moduli space of triples ${\cal M}_{\sigma}$. As we have seen above the tangent space to ${\cal M}_{\sigma}$ at a triple $T$ is identified with $H^1(X,C)$ where $C$ is the complex (\ref{tangcomp}). By Serre duality the cotangent space is isomorphic to $H^0(X,C^*)=H^1(X,C^*[-1])$, where the complex $C^*[-1]=((C^1)^*\stackrel{-d^*}{\rightarrow}(C^0)^*)$ is concentrated in degrees $0$ and $1$. Using the natural autoduality of $\operatorname{End} E_i$ the complex $C^*[-1]$ can be identified with $$\underline{\operatorname{Hom}}(E_1,E_2)\stackrel{-d^*}\rightarrow \underline{\operatorname{End}}E_1\oplus\underline{\operatorname{End}}E_2$$ where $-d^*(\Psi)=(-\Phi\Psi,\Psi\Phi)$. Now let us consider the morphism of complexes $\phi:C^*[-1]\rightarrow C$ with components $\phi_1=0$ and \begin{equation}\label{phi0} \phi_0:\underline{\operatorname{Hom}}(E_1,E_2)\rightarrow \underline{\operatorname{End}}E_1\oplus\underline{\operatorname{End}}E_2: \Psi\mapsto (\Phi\Psi,\Psi\Phi). \end{equation} Since $d\circ\phi_0=0$, we have indeed the morphism of complexes. Therefore, we can take the induced map on hypercohomologies $$H_T=\phi_*:H^1(X,C^*[-1])\rightarrow H^1(X,C).$$ Note that we get a map from the cotangent space to the tangent space of ${\cal M}_{\sigma}$ at $T$. This construction easily globalizes to give a morphism $H$ from the cotangent bundle to the tangent bundle of ${\cal M}_{\sigma}$. \begin{thm} $H$ defines a Poisson structure on ${\cal M}_{\sigma}$. \end{thm} \noindent {\it Proof} . Let us check that $H^*=-H$. First of all we claim that $\phi^*[-1]=\phi$ in the homotopy category of complexes. Indeed, by definition $\phi^*[-1]$ has components $(\phi^*[-1])_0=0$ and $$(\phi^*[-1])_1:\underline{\operatorname{End}}E_1\oplus\underline{\operatorname{End}}E_2\rightarrow \underline{\operatorname{Hom}}(E_2,E_1):(A,B)\mapsto A\Phi+\Phi B.$$ Now let us consider the map $$h:\underline{\operatorname{End}}E_1\oplus\underline{\operatorname{End}}E_2\rightarrow\underline{\operatorname{End}}E_1\oplus \underline{\operatorname{End}}E_2:(A,B)\mapsto (-A,B).$$ Immediate check shows that $h$ provides a homotopy from $\phi^*[-1]$ to $\phi$. Now the skew-commutativity of $H$ follows immediately from the skew-commutativity of the natural pairing $H^1(C)\otimes H^1(C')\rightarrow H^2(C\otimes C')$ that comes from the minus sign in the commutativity constraint for the tensor product of complexes. The Jacoby identity will be proven in section \ref{gener} in more general situation. \qed\vspace{3mm} We can interpret the Poisson bivector $H$ in terms of $\operatorname{SL}_2$-equivariant bundles on $X\times\P^1$ as follows. Let $F$ be the extension (\ref{equivext}) associated with a triple $T$. Then the tangent space to ${\cal M}_{\sigma}$ at $T$ is identified with $H^1(X\times\P^1, \underline{\operatorname{End}} F)^{\operatorname{SL}_2}$, hence by Serre duality the cotangent space is identified with $H^1(X\times\P^1, \underline{\operatorname{End}}(F)(-2))$. Now we claim that $H_T$ is induced by some canonical morphism $$\a:\underline{\operatorname{End}}F(-2)\rightarrow\underline{\operatorname{End}}F$$ on $X\times\P^1$. Namely, let $\underline{\operatorname{End}}(F,p^*E_1)$ be the kernel of the natural projection $\underline{\operatorname{End}}F\rightarrow\underline{\operatorname{Hom}}(p^*E_1,p^*E_2(2))$. The dual morphism to the embedding gives a morphism $\underline{\operatorname{End}}F\rightarrow\underline{\operatorname{End}}(F,p^*E_1)^*$. Thus, to construct $\a$ it is sufficient to construct a morphism $$\widetilde{\a}:\underline{\operatorname{End}}(F,p^*E_1)^*(-2)\rightarrow\underline{\operatorname{End}}(F,p^*E_1).$$ Now the bundle $\underline{\operatorname{End}}(F,p^*E_1)$ sits in the following exact triple $$0\rightarrow\underline{\operatorname{Hom}}(p^*E_2(2),p^*E_1)\rightarrow \underline{\operatorname{End}}(F,p^*E_1) \rightarrow\underline{\operatorname{End}}(p^*E_1)\oplus\underline{\operatorname{End}}(p^*E_2)\rightarrow 0.$$ We have the morphism to the last term of this triple $$p^*\phi_0:\underline{\operatorname{Hom}}(p^*E_1,p^*E_2)\rightarrow\underline{\operatorname{End}}(p^*E_1)\oplus \underline{\operatorname{End}}(p^*E_2)$$ where $\phi_0$ is defined in (\ref{phi0}). It is easy to check that $p^*\phi_0$ lifts uniquely to a morphism $$\underline{\operatorname{Hom}}(p^*E_1,p^*E_2)\rightarrow \underline{\operatorname{End}}(F,p^*E_1).$$ Now we define $\widetilde{\a}$ to be the composition of the latter morphism with the natural projection $\underline{\operatorname{End}}(F,p^*E_1)^*(-2)\rightarrow\underline{\operatorname{Hom}}(p^*E_1,p^*E_2)$. One has the natural morphism $\det:{\cal M}_{\sigma}\rightarrow\operatorname{Pic}(X)^2$ associating to a triple $(E_1,E_2,\Phi)$ the pair of line bundles $(\det E_1,\det E_2)$. We claim that $\det$ is a Casimir morphism, i.e. preimage of any local function downstairs is a Casimir function upstairs (that is a function having zero Poisson bracket with any other function). Indeed, the cotangent map to $\det$ is just the natural embedding $$i:H^0(X,\O_X)^2\rightarrow H^1(X, C^{\cdot})$$ which factors through $H^0(X,C^1)=H^0(X,\underline{\operatorname{End}} E_1)\oplus H^0(X,\underline{\operatorname{End}} E_2)$. On the other hand, $H$ factors through the map $H^1(X, C^{\cdot})\rightarrow H^1(X,C^0)$, hence the image of $i$ is killed by $H$. In particular, the Poisson bracket on ${\cal M}_{\sigma}$ induces Poisson brackets on the fibers of the morphism $\det$. These fibers can be identified with moduli spaces ${\cal M}_{\sigma}(L_1,L_2,r_1,r_2)$ of triples with fixed determinants $\det E_i\simeq L_i$. Tensoring with a fixed line bundle $L$ gives a Poisson isomorphism ${\cal M}_{\sigma}(L_1,L_2,r_1,r_2)\simeq {\cal M}_{\sigma}(L_1\otimes L^{\otimes r_1}, L_2\otimes L^{\otimes r_2},r_1,r_2)$. An automorphism $\phi:X\rightarrow X$ induces an isomorphism of moduli spaces ${\cal M}_{\sigma}(L_1,L_2,r_1,r_2)\rightarrow{\cal M}_{\sigma}(\phi^*L_1, \phi^*L_2,r_1,r_2)$ compatible with Poisson structures. \section{Special cases} For any bundle $E_1$ and a subbundle $E_2\subset E_1$ let us denote by $\underline{\operatorname{End}}(E_1,E_2)$ the sheaf of local homomorphisms of $E_1$ preserving $E_2$. In other words, this is the kernel of the natural projection $\underline{\operatorname{End}}(E_1)\rightarrow\underline{\operatorname{Hom}}(E_2,E_1/E_2)$. \begin{lem}\label{subbun} Let $T=(E_1,E_2,\Phi)$ be a $\sigma$-stable triple such that $\Phi:E_2\rightarrow E_1$ is an embedding of $E_2$ as a subbundle. Then the tangent space to ${\cal M}_{\sigma}$ at $T$ can be identified with $H^1(X,\underline{\operatorname{End}}(E_1,E_2))$. \end{lem} \noindent {\it Proof}. The natural embedding $\underline{\operatorname{End}}(E_1,E_2)\rightarrow\underline{\operatorname{End}}(E_1)\oplus\underline{\operatorname{End}}(E_2)$ induces the quasi-isomorphism $\underline{\operatorname{End}}(E_1,E_2)\rightarrow C^{\cdot}$. Hence the assertion. \qed\vspace{3mm} Under the identification of this lemma our Poisson structure at the triple $T$ for which $\Phi$ is an embedding of a subbundle can be described as follows. From the exact triple $$0\rightarrow\underline{\operatorname{Hom}}(E_1/E_2,E_2)\rightarrow\underline{\operatorname{End}}E_1 \rightarrow\underline{\operatorname{End}}(E_1,E_2)^*\rightarrow 0$$ we get a boundary homomorphism $$H^0(X,\underline{\operatorname{End}}(E_1,E_2)^*)\rightarrow H^1(X,\underline{\operatorname{Hom}}(E_1/E_2,E_2)).$$ Composing it with the natural morphism $$H^1(X,\underline{\operatorname{Hom}}(E_1/E_2,E_2))\rightarrow H^1(X,\underline{\operatorname{End}}(E_1,E_2))$$ we get a morphism $$H^1(X,\underline{\operatorname{End}}(E_1,E_2))^*\simeq H^0(X,\underline{\operatorname{End}}(E_1,E_2)^*)\rightarrow H^1(X,\underline{\operatorname{End}}(E_1,E_2)),$$ which coincides with $H_T$ under identification of Lemma \ref{subbun}. Equivalently, we may start with the natural morphism $$H^0(X,\underline{\operatorname{End}}(E_1,E_2)^*)\rightarrow H^0(X,\underline{\operatorname{Hom}}(E_2,E_1/E_2))$$ and compose it with the boundary homomorphism $$H^0(X,\underline{\operatorname{Hom}}(E_2,E_1/E_2))\rightarrow H^1(X,\underline{\operatorname{End}}(E_1,E_2))$$ coming from the exact triple \begin{equation}\label{imptriple} 0\rightarrow\underline{\operatorname{End}}(E_1,E_2)\rightarrow\underline{\operatorname{End}} E_1\rightarrow \underline{\operatorname{Hom}}(E_2, E_1/E_2)\rightarrow 0. \end{equation} The equivalence of this description with the previous one follows immediately from the commutative diagram \begin{equation} \begin{array}{ccccc} \underline{\operatorname{Hom}}(E_1/E_2,E_2) &\lrar{} &\underline{\operatorname{End}}(E_1,E_2) &\lrar{} &\underline{\operatorname{End}} E_2\oplus\underline{\operatorname{End}} E_1/E_2 \\ \ldar{\operatorname{id}} & & \ldar{} & & \ldar{} \\ \underline{\operatorname{Hom}}(E_1/E_2,E_2) &\lrar{} &\underline{\operatorname{End}} E_1 &\lrar{} &\underline{\operatorname{End}}(E_1,E_2)^* \\ &&\ldar{} & & \ldar{}\\ &&\underline{\operatorname{Hom}}(E_2,E_1/E_2) &\lrar{\operatorname{id}}& \underline{\operatorname{Hom}}(E_2,E_1/E_2) \end{array} \end{equation} It is easy to check using the above descriptions that the restriction of our Poisson bracket to the space of triples for which $\Phi$ is an embedding of a subbundle is a particular case of the Poisson bracket on moduli of principal bundles over parabolic subgroups defined by Feigin and Odesskii in \cite{FO}. \begin{lem}\label{simple} Keep the assumptions of Lemma \ref{subbun}. Then we have an exact triple $$0\rightarrow \operatorname{End} E_2\oplus\operatorname{End} (E_1/E_2)\rightarrow\ker H_T\rightarrow \operatorname{End} E_1/\operatorname{End}(E_1,E_2)\rightarrow 0.$$ If $\Phi(E_2)$ is a direct summand of $E_1$ then $H$ vanishes at $T$. Otherwise, the dimension of $\ker H_T$ is minimal (and equals to $2$) if and only if the bundles $E_1$, $E_2$ and $E_1/E_2$ are simple. \end{lem} \noindent {\it Proof} . Considering the second of the above descriptions of $H_T$ we see immediately that there is an exact sequence \begin{equation}\label{triplepf} 0\rightarrow \operatorname{End} E_2\oplus \operatorname{End} (E_1/E_2)\rightarrow\ker H_T\rightarrow \ker (H^0(X,\underline{\operatorname{Hom}}(E_2,E_1/E_2))\rightarrow H^1(\underline{\operatorname{End}}(E_1,E_2)). \end{equation} Using (\ref{imptriple}) the last term can be identified with $\operatorname{End} E_1/ \operatorname{End}(E_1,E_2)$. Moreover, the last arrow in (\ref{triplepf}) is surjective since we have a natural map $\operatorname{End} E_1\rightarrow\ker H_T$ coming from the morphism $\underline{\operatorname{End}} E_1\rightarrow\underline{\operatorname{End}}(E_1,E_2)^*$, and the composition of this map with the last arrow of (\ref{triplepf}) is just the canonical projection to $\operatorname{End} E_1/\operatorname{End}(E_1,E_2)$. If $\Phi(E_2)$ is a direct summand in $E_1$ then the boundary homomorphism used in the definition of $H_T$ is zero, hence $H_T=0$. Otherwise, $\dim \ker H_T=2$ if and only if $E_2$ and $E_1/E_2$ are simple and all endomorphisms of $E_1$ preserve $\Phi(E_2)$. We claim this can happen only when $E_1$ is also simple. Indeed, let $A:E_1\rightarrow E_1$ be any endomorphism. By assumption $A$ preserves $\Phi(E_2)$. Adding a constant to $A$ we may assume that $A|_{\Phi(E_2)}=0$. Then it induces a map $E_1/E_2\rightarrow E_2$. However, $\sigma$-stability of our triple implies by Lemma 4.4 of \cite{BG} that $\operatorname{Hom}(E_1,E_2)=0$ since $\Phi$ is not an isomorphism in our situation. It follows that $\operatorname{Hom}(E_1/E_2,E_2)=0$, hence, $A=0$. Thus, $\operatorname{End} E_1={\Bbb C}$. \qed\vspace{3mm} We are mainly interested in the case when $E_2=\O_X$, $\det E_1$ is fixed. In terms of parameter $\tau$ the stability condition on $\Phi:\O_X\rightarrow E_1$ is that $\mu(E'_1)<\tau$ for every proper non-zero subbundle $E'_1\subset E_1$ and $\mu(E_1/E'_1)>\tau$ for every proper subbundle $E'_1\subset E_1$ such that $\Phi\in H^0(X,E'_1)$. Now let $E$ be a stable bundle on $X$ of degree $d$ and rank $r$ (in particular, $gcd(d,r)=1$). Set $\tau=\mu(E)=\frac{d}{r}$ and consider the moduli space ${\cal M}_{\tau}(\det E,\O_X,r+1,1)$. It is easy to see that the stability condition on such a triple $\Phi:E_2=\O_X\rightarrow E_1$ is equivalent to the condition that $\Phi$ is nowhere vanishing section and the quotient $E_1/\Phi(\O_X)$ is a stable bundle. Moreover, since there exists a unique stable bundle of rank $r$ and determinant $\det E$ it follows that $E_1/\Phi(\O_X)\simeq E$. Thus, we can identify the moduli space of such triples with the projective space $\P\operatorname{Ext}^1(E,\O_X)$. If $gcd(d,r+1)=1$ then generic extension of $E$ by $\O_X$ is stable. Hence, according to Lemma \ref{simple} in this case the Poisson bracket on $\P\operatorname{Ext}^1(E,\O_X)$ is symplectic at general point. Let $t_x:X\rightarrow X$ be the translation by $x\in X$. Then by functoriality we have a natural Poisson isomorphism $$\P\operatorname{Ext}^1(E,\O_X)\widetilde{\rightarrow} \P\operatorname{Ext}^1(t_x^*E,\O_X).$$ Note that as $x$ varies $t_x^*E$ runs through all stable bundles of given degree $d$ and rank $r$. Let $K\subset X$ be the finite subgroup of order $d^2$ consisting of $x$ such that $t_x^*\det E\simeq\det E$. Then for $x\in K$ one has $t_x^*E\simeq E$, therefore, $K$ acts on $\P\operatorname{Ext}^1(E,\O_X)$ by linear transformations preserving the Poisson structure. Another special moduli space associated with a fixed stable bundle $E$ is ${\cal M}_{\tau}(\det E,\O_X,\operatorname{rk} E, 1)$ where $\tau=\mu(E)$. Then the condition on a triple just means that $E_1$ is stable, hence isomorphic to $E$, and $\Phi$ is an arbitrary non-zero section. Therefore, this moduli space can be identified with $\P H^0(X,E)$. The Poisson bracket in this case can be described as follows. The tangent space $T_{[s]}$ to the line generated by $s\in H^0(X,E)$ is identified with $\operatorname{coker}(H^0(X,\O_X)\stackrel{s}{\rightarrow} H^0(X,E))$. Thus the cotangent space is $T^*_{[s]}=\ker((H^1(X,E^*)\rightarrow H^1(X,\O_X))$. Let $D\subset X$ be the divisor of zeroes of $s$, so that $s:\O_X\rightarrow E$ factors as $\O_X\rightarrow\O_X(D)\rightarrow E$ where $\O_X(D)$ is embedded as a subbundle into $E$. Then we have the natural map \begin{equation}\label{map} H^1(X,E^*)\rightarrow H^1(X,E^*(D))\rightarrow H^1(X,\underline{\operatorname{End}}(E,\O_X(D))). \end{equation} The latter space sits in the exact sequence $$0\rightarrow H^0(X,E(-D)/\O_X)\rightarrow H^1(X,\underline{\operatorname{End}}(E,\O_X(D)))\rightarrow H^1(X,\underline{\operatorname{End}} E) \simeq H^1(X,\O_X).$$ It follows that (\ref{map}) induces a map $T^*_{[s]}\rightarrow H^0(X,E(-D)/\O_X)$. Furthermore, it is easy to check that its composition with the boundary homomorphism $H^0(X,E(-D)/\O_X)\rightarrow H^1(X,\O_X)$ is zero, hence, we get a map $$T^*_{[s]}\rightarrow H^0(X,E(-D))/H^0(X,\O_X).$$ Now the latter space is naturally embedded into $T_{[s]}$ and the composition with this embedding gives our Poisson structure at $[s]\in\P H^0(X,E)$. \begin{lem} Let $H_{[s]}:T_{[s]}^*\rightarrow T_{[s]}$ be the above Poisson structure on $\P H^0(X,E)$. If $\operatorname{rk} E=1$ then $H_{[s]}=0$. Otherwise, one has an exact sequence $$0\rightarrow H^1(X,\operatorname{ad}(E/\O_X(D)))\rightarrow\operatorname{coker} H_{[s]}\rightarrow H^0(D,E|_D)\rightarrow 0$$ where $D$ is the zero divisor of $s$. \end{lem} \noindent {\it Proof} . Let us denote by $V\subset T_{[s]}$ the subspace $H^0(X,E(-D))/H^0(X,\O_X)$. Since the image of $H_{[s]}$ is contained in $V$ we have the exact sequence $$0\rightarrow\operatorname{coker}(T^*_{[s]}\rightarrow V)\rightarrow\operatorname{coker} H_{[s]}\rightarrow T_{[s]}/V\rightarrow 0.$$ Since $E(-D)$ is stable of positive slope, it follows that $H^1(X, E(-D))=0$, hence we have an isomorphism $$T_{[s]}/V\simeq H^0(X,E)/H^0(X,E(-D))\simeq H^0(D,E|_D).$$ Now the assertion follows easily from the exact sequence $$T_{[s]}^*\rightarrow H^1(X,\underline{\operatorname{End}}(E,\O(D)))\rightarrow H^1(X,\underline{\operatorname{End}}(E/\O_X))\oplus H^1(X,\O_X)\rightarrow 0.$$ \qed\vspace{3mm} More generally, we can consider the moduli space ${\cal M}_{\tau}(L,\O_X,r,1)$ where $L$ is a fixed line bundle of degree $d$, $\tau=\frac{d}{r}+\epsilon$ where $\epsilon>0$ is sufficiently small. Then we get the moduli space of pairs $s:\O_X\rightarrow E$ where $E$ is a semistable bundle with determinant $L$, $\operatorname{rk} E=r$, $s$ is a section which doesn't belong to any destabilizing subbundle of $E$. We have a Casimir morphism from this moduli space to the moduli stack of semistable bundles, so the fibers inherit the Poisson structure. In particular, if we take the semistable bundle $E=(E_0)^{\oplus k}$ where $E_0$ is a stable bundle, then the corresponding fiber is identified with the Grassmannian $G(k,H^0(E_0))$ of $k$-dimensional subspaces in $H^0(E_0)$, so we get some family of Poisson structures on the Grassmannians. \section{Fourier transforms} Let $m:X\times X\rightarrow X$ be the group law on $X$, $x_0\in X$ be the neutral element. Let $${\cal P}=m^*\O_X(x_0)\otimes p_1^*\O_X(-x_0)\otimes p_2^*\O_X(-x_0)$$ be the Poincar\'e line bundle on $X\times X$ inducing an isomorphism of $X$ with the dual elliptic curve. We denote by ${\cal F}$ the corresponding Fourier-Mukai transform which is an autoequivalence of the the derived category ${\cal D}^b(X)$ of coherent sheaves on $X$ given by $${\cal F}(E)=Rp_{2*}(Lp_1^*E\sideset{^L}{}{\otimes}{\cal P}).$$ One has ${\cal F}\circ{\cal F}\simeq (-\operatorname{id}_X)^*[-1]$ (see \cite{Mukai}). It is easy to see that for every $E\in{\cal D}^b(X)$ one has $\operatorname{rk}{\cal F}(E)=\deg E$, $\deg{\cal F}(E)=-\operatorname{rk} E$. It follows that if $T:{\cal D}^b(X)\rightarrow{\cal D}^b(X)$ is a composition of some sequence of Fourier transforms and tensorings with line bundles, then the vector $v(T(E))=(\operatorname{rk} T(E), \deg T(E))$ is obtained from the vector $v(E)=(\operatorname{rk} E, \deg E)$ by applying some matrix $A\in\operatorname{SL}_2({\Bbb Z})$. Furthermore, one can lift the natural action of $\operatorname{SL}_2({\Bbb Z})$ on vectors $(\operatorname{rk}, \deg)$ to the action of a central extension of $\operatorname{SL}_2({\Bbb Z})$ by ${\Bbb Z}$ on ${\cal D}^b(X)$. More precisely, we can consider the standard presentation of $\operatorname{SL}_2({\Bbb Z})$ by generators $S=\left( \matrix 0 & 1\\ -1 & 0 \endmatrix \right)$ and $R=\left( \matrix 1 & 0\\ 1 & 1 \endmatrix \right)$ subject to relations $$S^2=(RS)^3,\ S^4=1.$$ Then the central extension in question is the group $\widetilde{\operatorname{SL}}_2({\Bbb Z})$ generated by $S$ and $R$ with the only relation $S^2=(RS)^3$. The action of this group on ${\cal D}^b(X)$ is the following: $S$ acts as the Fourier-Mukai transform while $R$ acts as tensoring with $\O_X(x_0)$ (see \cite{Mukai}). We will consider the action of $\widetilde{\operatorname{SL}}_2({\Bbb Z})$ on morphisms of stable bundles. For this it will be useful to know the orbits of $\operatorname{SL}_2({\Bbb Z})$ on pairs of primitive vectors in ${\Bbb Z}^2$. First of all, for a pair of vectors $v_1=(r_1,d_1)$, $v_2=(r_2,d_2)$ such that $gcd(r_i,d_i)=1$ for $i=1,2$ the determinant $\det(v_1,v_2)\in{\Bbb Z}$ is invariant of $\operatorname{SL}_2({\Bbb Z})$. We consider only pairs $v_1,v_2$ with $\det(v_1,v_2)\neq 0$. For such pairs there is a second $\operatorname{SL}_2({\Bbb Z})$-invariant $\a(v_1,v_2)\in({\Bbb Z}/\det(v_1,v_2))^*$ defined from the condition $$v_1\equiv\a(v_1,v_2)v_2\operatorname {mod}\det(v_1,v_2){\Bbb Z}^2$$ It is easy to see that the $\operatorname{SL}_2({\Bbb Z})$-orbit of such $(v_1,v_2)$ consists of all pairs with the same $\det$ and $\a$. Henceforward, we restrict ourself to considering stable bundles on $X$ with determinant isomorphic to $\O_X(nx_0)$ for some $n$. The reason is that the group $\widetilde{\operatorname{SL}}_2({\Bbb Z})$ preserves the set $S_{x_0}$ of objects of the form $E[k]$ where $k\in{\Bbb Z}$, $E$ is either a stable bundle with determinant $\O_X(nx_0)$ for some $n$, or $\O_{x_0}$. An element of $S_{x_0}$ is determined by its degree and rank uniquely up to a shift. It follows that the group $\widetilde{\operatorname{SL}}_2({\Bbb Z})$ acts transitively on $S_{x_0}$. Furthermore, an element $T\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ is completely determined by its action on a pair of elements of $S_{x_0}$ which are not isomorphic up to shift. The first immediate consequence of the action of $\widetilde{\operatorname{SL}}_2({\Bbb Z})$ is that for stable bundles $E_1$ and $E_2$ such that $\det E_1\simeq\det E_2\simeq\O_X(dx_0)$ and $\operatorname{rk} E_1\equiv \operatorname{rk} E_2\operatorname {mod}(d)$ there is a canonical isomorphism $$\P\operatorname{Ext}^1(E_1,\O_X)\simeq\P\operatorname{Ext}^1(E_2,\O_X).$$ Indeed, under these conditions there is a unique element $T\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ such that $T(\O_X)\simeq\O_X$ and $T(E_1)\simeq E_2$. Considering the action of $T$ on morphisms from $E_1$ to $\O_X[1]$ we get the above isomorphism. Now for every stable bundle $E$ with $\det E\simeq\O_X(dx_0)$ where $d>1$ we can find an element $T\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ such that $T(E)\simeq \O_X$. Then $T(\O_X)=E'[n]$ for some stable bundle $E'$ and some $n\in{\Bbb Z}$. Since $\operatorname{Hom}(\O_X,E)\neq 0$ we should have $\operatorname{Hom}(T(\O_X), T(E))\neq 0$, hence $n=0$ or $-1$. Consider first the case $n=-1$. Then one has $\det E'\simeq \O_X(dx_0)$ and $$r\cdot r'\equiv -1\operatorname {mod}(d)$$ where $r=\operatorname{rk} E$, $r'=\operatorname{rk} E'$ (this is deduced comparing invariants $\a$ for the pair of vectors $(v(\O_X),v(E))$ and its image under $T$). Conversely, for every $E'$ satisfying these conditions there exists a unique element $T\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ sending $E$ to $\O_X$ and $\O_X$ to $E'[-1]$. The transformation $T$ induces an isomorphism \begin{equation}\label{linisom} T_*:\P H^0(E)\widetilde{\rightarrow}\P\operatorname{Ext}^1(E',\O_X). \end{equation} (an isomorphism of this kind with $r=1$, $r'=d-1$ was constructed in \cite{FMW} by a different method.) Note that in the previous section we identified both sides of the isomorphism (\ref{linisom}) with some special moduli spaces of pairs, in particular, they carry natural Poisson structures. \begin{prop} The isomorphism $T_*$ is compatible with Poisson structures. \end{prop} \noindent {\it Proof} . Let $s:\O_X\rightarrow E$ be a non-zero section, $\O_X\rightarrow\widetilde{E}\rightarrow E'$ be the corresponding extension of $E'$ by $\O_X$ with class $T_*(s)\in\operatorname{Ext}^1(E',\O_X)$. It suffices to prove that $T_*$ preserves Poisson structure over a non-empty open subset, hence we can assume that $E/\O_X$ has no torsion. By Serre duality the cotangent space $T^*_{[s]}\P H^0(E)$ can be identified with $$\ker(\operatorname{Ext}^1(E,\O_X)\rightarrow H^1(\O_X))\simeq \operatorname{Ext}^1(E/\O_X,\O_X)/{\Bbb C}\cdot e$$ where $e\in\operatorname{Ext}^1(E/\O_X, \O_X)$ is the class of the extension $\O_X\rightarrow E\rightarrow E/\O_X$. Under this identification the Poisson bracket on $\P H^0(E)$ at the point $[s]$ is induced by the natural morphism $$H_{[s]}:\operatorname{Ext}^1(E/\O_X,\O_X)/{\Bbb C}\cdot e\rightarrow T_{[s]}\P H^0(E)$$ which comes from the identification of $T_{[s]}\P H^0(E)$ with $\ker(H^1(\underline{\operatorname{End}}(E,\O_X))\rightarrow H^1(\O)^2)$ and from the natural morphism $$\operatorname{Ext}^1(E/\O_X,\O_X)=H^1(\underline{\operatorname{Hom}}(E/\O_X,\O_X))\rightarrow H^1(\underline{\operatorname{End}}(E,\O_X)).$$ In other words, we have a morphism from a neighborhood of $[e]$ in the space of extensions $\P\operatorname{Ext}^1(E/\O_X,\O_X)$ to $\P H^0(E)$ (since in the neighborhood of $[e]$ such an extension is necessarily isomorphic to $E$), and $H_{[s]}$ is just the tangent map to this morphism at the point $[e]$. Similarly, the Poisson bracket on $\P\operatorname{Ext}^1(E',\O_X)$ at the point $[T_*s]$ can be identified with the tangent map $$\operatorname{Hom}(\widetilde{E},E')/{\Bbb C}\cdot f\rightarrow T_{[T_*(s)]}\P\operatorname{Ext}^1(E',\O_X)$$ to the local morphism from $\P\operatorname{Hom}(\widetilde{E},E')$ to $\P\operatorname{Ext}^1(E',\O_X)$ at the point $[f]$ where $f:\widetilde{E}\rightarrow E'$ is the canonical morphism. Here we use the natural identification of $\operatorname{Hom}(\widetilde{E},E')/{\Bbb C}\simeq\ker(H^0(E')\rightarrow H^1(\O_X))$ with the cotangent space to $\P\operatorname{Ext}^1(E',\O_X)$ at $[T_*(s)]$. Now we have the following commutative square of local morphisms in the neighborhood of points induced by $s$: \begin{equation} \begin{array}{ccc} \P\operatorname{Ext}^1(E/\O_X,\O_X) & \lrar{} & \P H^0(E)\\ \ldar{T_*} && \ldar{T_*}\\ \P\operatorname{Hom}(\widetilde{E},E') &\lrar{} & \P\operatorname{Ext}^1(E',\O_X) \end{array} \end{equation} Considering the corresponding commutative square of tangent maps we get the compatibility of $T_*$ with Poisson brackets. \qed\vspace{3mm} For some other choice of autoequivalence $T:{\cal D}^b(X)\rightarrow{\cal D}^b(X)$ sending $E$ to $\O_X$ one has $T(\O_X)\simeq (E'')^*$ for a stable bundle $E''$ of degree $\deg E''=\deg E=d$ and rank $r''$ satisfying the congruence $$r''\cdot r\equiv 1\operatorname {mod}(d)$$ where $r=\operatorname{rk} E$. In this case we get an isomorphism $$\P H^0(E)\simeq \P H^0(E'').$$ We claim that it is also compatible with the natural Poisson structures on both sides. Indeed, this is proven exactly as in the above proposition using the following commutative diagram of local morphisms: \begin{equation} \begin{array}{ccc} \P\operatorname{Ext}^1(E/\O_X,\O_X) & \lrar{} & \P H^0(E)\\ \ldar{T_*} && \ldar{T_*}\\ \P\operatorname{Hom}(E'',E''/\O_X) &\lrar{} & \P H^0(E'') \end{array} \end{equation} Combining these isomorphisms we also get Poisson isomorphisms between $\P\operatorname{Ext}^1(E_1,\O_X)$ and $\P\operatorname{Ext}^1(E_2,\O_X)$ for stable bundles $E_1$ and $E_2$ of the same degree $d$ and ranks $r_1$ and $r_2$ satisfying $r_1 r_2\equiv 1\operatorname {mod} (d)$. We denote by ${\cal M}(d,r)$ the projective space $\P\operatorname{Ext}^1(E,\O_X)$ where $E$ is a stable bundle with determinant $\O_X(dx_0)$ and rank $r$ (in particular, $gcd(d,r)=1$), considered as a Poisson variety. Then the above results show that ${\cal M}(d,r)$ depends only on $d$ and on the residue $r\in({\Bbb Z}/d{\Bbb Z})^*$, furthermore, we have \begin{equation}\label{fourisom} {\cal M}(d,r)\simeq{\cal M}(d,r^{-1}). \end{equation} Also for every stable bundle $E$ of degree $d>0$ and rank $r$ we have an isomorphism of Poisson varieties $$\P H^0(E)\simeq{\cal M}(d,-r^{-1}).$$ The Poisson isomorphism (\ref{fourisom}) is the classical limit of the following isomorphism of Sklyanin algebras $$Q_{d,r}(x)\simeq Q_{d,r'}(x)$$ for every $\tau\in X$, where $rr'\equiv 1\operatorname {mod}(d)$ (cf. \cite{FOf}). To see this isomorphism let us make the substitutions $i=r'j'$, $j=r'i'$, $n=r'(n'+i'-j')$ in the defining relation (\ref{relation}) of $Q_{d,r}(x)$. Then using the relation $\theta_{-i}(-x)=a\cdot b^i\cdot\theta_i(x)$ (where $a$ and $b$ are some non-zero constants, $b^d=1$) we can rewrite the quadratic relations in the form \begin{equation}\label{relation2} \sum_{n'\in{\Bbb Z}/d{\Bbb Z}}\frac{\theta_{j'-i'-(r'-1)n'}(0)} {\theta_{r'n'}(x)\theta_{j'-i'-n'}(-x)} t_{j'-n'}t_{i'+n'}=0. \end{equation} Now it is obvious that the map $t_i\mapsto t_{r'i}$ defines an isomorphism from $Q_{d,r}(x)$ to $Q_{d,r'}(x)$ as required. \section{Birational transformations} If one changes the stability parameter $\tau$ the moduli spaces ${\cal M}_{\tau}(L_1,L_2,r_1,r_2)$ undergo some birational transformations, see \cite{T}, \cite{BDW}. Clearly, these birational transformations are compatible with the Poisson structures on their domain of definition. In particular, considering moduli of pairs $\O_X\rightarrow E$ where $\deg E=d$, $\operatorname{rk} E=r$ are such that $gcd (r,d)=1$ and $gdc(r+1,d)=1$ we get a Poisson birational transformation from ${\cal M}(d,r)$ to ${\cal M}(d,-(r+1)^{-1})\simeq\P H^0(E)$ where $E$ is a stable bundle of degree $d$ and rank $r+1$. Let us denote by $R_d\subset{\Bbb Z}/d{\Bbb Z}$ the set of residues $r$ such that $r\in({\Bbb Z}/d{\Bbb Z})^*$ and $r+1\in({\Bbb Z}/d{\Bbb Z})^*$. The map $\phi:r\mapsto -(r+1)^{-1}$ preserves $R_d$ and satisfies $\phi^3=\operatorname{id}$. On the other hand, the involution $\b:R_d\rightarrow R_d: r\mapsto r^{-1}$ also preserves $R_d$ and we have $\phi\b=\b\phi^{-1}$. It follows that $\b$ and $\phi$ generate the action of the symmetric group $S_3$ on $R_d$. Recall that in the previous section we defined an isomorphism ${\cal M}(d,r)\rightarrow{\cal M}(d,r^{-1})={\cal M}(d,\b(r))$. \begin{thm}\label{s3} The birational morphisms ${\cal M}(d,r)\rightarrow{\cal M}(d,\phi(r))$ and ${\cal M}(d,r)\rightarrow{\cal M}(d,\b(r))$ defined above extend to an action of $S_3$ on $\sqcup_{r\in R_d}{\cal M}(d,r)$. \end{thm} \noindent {\it Proof} . For every residue $r\in R_d$ let us denote by $E_r$ a stable bundle with determinant $\O_X(dx_0)$ and rank $\operatorname{rk} E\equiv r\operatorname {mod}(d)$ such that $0<\operatorname{rk} E<d$. Let us check the relation $\phi^3=\operatorname{id}$. For this we have to show that the corresponding composition of birational transformations $${\cal M}(d,r)\rightarrow{\cal M}(d,\phi(r))\rightarrow{\cal M}(d,\phi^2(r))\rightarrow{\cal M}(d,r)$$ is the identity. By definition the first arrow is the composition of the map that associates to a generic morphism $f:E_r\rightarrow\O_X[1]$ the corresponding morphism $\operatorname{Cone}(f)[-1]:\O_X\rightarrow E_{r+1}$ (considered up to constant) with the autoequivalence $T_r\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ which sends $E_{r+1}$ to $\O_X[1]$ and $\O_X$ to $E_{\phi(r)}$. It follows that the above triple composition amounts to applying the construction $\operatorname{Cone}(\cdot)[-1]$ thrice (note that in our situation this construction is functorial) and applying the functor $T_{\phi^2(r)}T_{\phi(r)}T_r$. The triple composition of $\operatorname{Cone}(\cdot)[-1]$ is isomoprhic to the shift $\operatorname{id}[-2]$, while $T_{\phi^2(r)}T_{\phi(r)}T_r=\operatorname{id}[2]$, hence the assertion. It remains to check the relation between our birational transformations corresponding to the relation $\phi\b=\b\phi^{-1}$. This amount to checking the following relation between contravariant functors from $D^b(X)$ to itself: $$DT_{\b\phi(r)}DU_{\phi(r)}T_r=U_r$$ where $D:{\cal D}^b(X)^{op}\rightarrow{\cal D}^b(X)$ is the duality functor $E\mapsto R\underline{\operatorname{Hom}}(E,\O_X)$, $U_r\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ is the unique element sending $E_r$ to $\O_X$ and $\O_X$ to $E_{\b(r)}^*$. Note that for every $T\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ we have $DTD\in\widetilde{\operatorname{SL}}_2({\Bbb Z})$ (this follows from the compatibility between the Fourier-Mukai transform and duality, cf. \cite{Mukai}), moreover the corresponding involution on $\operatorname{SL}_2({\Bbb Z})$ is just the conjugation by the matrix $\left(\matrix 1 & 0 \\ 0 & -1 \endmatrix \right)$. Using this one can check the above identity up to shift. Finally, since both parts send $E_r$ to $\O_X$ the identity follows. \qed\vspace{3mm} It follows from the above theorem that for every $\sigma\in S_3$ and $r\in R_d$ such that $\sigma(r)=r$ we get a birational automorphism $f_{\sigma}$ of ${\cal M}(d,r)$. Since $\b$ acts as an isomorphism on our moduli spaces there are essentially two different cases to consider: the residue $r\in R_d$ is fixed $\phi$ or by $\phi\b$. The fixed points of $\phi$ are residues $r$ satisfying $$r^2+r+1\equiv 0\operatorname {mod}(d).$$ In this case we get a Poisson birational automorphism $f_{\phi}$ of ${\cal M}(d,r)$ such that $f_{\phi}^3=\operatorname{id}$. The map $\phi\b$ has the only fixed point $r=-2\operatorname {mod}(d)$ (provided that $d$ is odd), so we get a Poisson birational involution $f_{\phi\b}$ of ${\cal M}(d,d-2)$. Let us describe these birational automorphisms more explicitly. Consider first the case when $r^2+r+1\equiv 0\operatorname {mod}(d)$, i.e. $\phi(r)=r$. Using the notation of the proof of the above theorem we have $T_r(E_{r+1})=\O_X[1]$, $T_r(\O_X)=E_r$, and $T_r(E_r)= E_{r+1}[1]$. Now let us consider the closed subvariety of $$Z\subset \P\operatorname{Hom}(E_{r+1},E_r)\times\P\operatorname{Ext}^1(E_r,\O_X)\times \P\operatorname{Hom}(\O_X,E_{r+1})$$ consisting of triples of lines $([v_1],[v_2],[v_3])$ such that all three pairwise compositions $v_2\circ v_1$, $v_3\circ v_2$ and $v_1\circ v_3$ are zeroes. It is easy to see that any of three projections of $Z$ to the projective spaces are birational. In fact, general point of $Z$ corresponds to the exact triangle $$\O_X\stackrel{v_3}{\rightarrow} E_{r+1}\stackrel{v_1}{\rightarrow} E_r \stackrel{v_2}{\rightarrow}\O_X[1].$$ In particular, $Z$ is birational to ${\cal M}(d,r)=\P\operatorname{Ext}^1(E_r,\O_X)$. On the other hand, the functor $T_r$ gives rise to the automorphism of $Z$ given by $$\Phi:(v_1,v_2,v_3)\mapsto (T_r(v_2)[-1],T_r(v_3),T_r(v_1)[-1]).$$ Now $\Phi$ induces our birational automorphism of ${\cal M}(d,r)$ with cube equal to the identity. In the simplest non-trivial case $d=7$, $r=2$ the functor we use has form $T_r:A\mapsto {\cal F}(A(-2x_0))(3x_0)[1]$. In this case one can describe $Z$ as the double blow-up of ${\cal M}(7,2)$. For this it is more convenient to use the isomorphism of ${\cal M}(7,2)$ with $\P H^0(E_3)$. Then we have a natural embedding $S^2 C\rightarrow\P H^0(E_3): D\mapsto H^0(E_3(-D))$ where $D$ is an effective divisor of degree 2 on $X$. We also define the 4-dimensional variety $V\subset \P H^0(E_3)$ containing $S^2 C$ as the union of chords of $S^2 C\subset \P H^0(E_3)$ connecting $D_1$ and $D_2$ in $S^2 C$ such that $D_1\cap D_2\neq\emptyset$. Then our variety $Z$ is obtained by first blowing-up of $\P H^0(E_3)$ along $S^2 C$, and then blowing up the proper transform of $V$. The automorphism $\Phi$ cyclically permutes the following three divisors on $Z$: two exceptional divisors and the proper transform of the chord variety of $S^2 C$ (which is a hypersurface in $\P H^0(E_3)$). In the case $r\equiv -2\operatorname {mod}(d)$ (where $d$ is odd) the birational autoequivalence of ${\cal M}(d,d-2)$ is described as follows. Again using the notation from the proof of Theorem \ref{s3} we have $U_{d-1}(E_{d-1})=\O_X$, $U_{d-1}(\O_X)=E_{d-1}^*$ and $U_{d-1}(E_{d-2})=E_{d-2}^*[1]$. Now let us consider the subvariety $$Y\subset \P\operatorname{Hom}(E_{d-1},E_{d-2})\times\P\operatorname{Ext}^1(E_{d-2},\O_X)$$ consisting of pairs $([v_1],[v_2])$ such that the composition $v_2v_1\in\operatorname{Ext}^1(E_{d-1},\O_X)$ is zero. Then both projections of $Y$ to projective spaces are birational. On the other hand, the functor $U_{d-1}$ induces an involution of $Y$ sending $(v_1,v_2)$ to $(U_{d-1}(v_2)^*[1], U_{d-1}(v_1)^*[1])$, hence our birational involution of ${\cal M}(d,d-2)=\P\operatorname{Ext}^1(E_{d-2},\O_X)$. \section{Generalization}\label{gener} In this section we consider a generalization of the main construction of section \ref{mainsec} to the case of principal bundles with other structural groups than $\operatorname{GL}$. For this note that the datum of a triple $(E_1,E_2,\Phi)$ with $\operatorname{rk} E_i=r_i$ for $i=1,2$ is equivalent to that of a pair $(P,s)$ where $P$ is a principal bundle with structure group $\operatorname{GL}_{r_1}\times\operatorname{GL}_{r_2}$ and a section $s\in V(P)$ of the vector bundle $V(P)$ associated with the natural representation of $\operatorname{GL}_{r_1}\times\operatorname{GL}_{r_2}$ on the space of $r_1\times r_2$-matrices. To generalize this let us consider a general reductive group $G$ and its representation $V$. Then one can consider the moduli stack ${\cal M}_{G,V}$ of pairs $(P,s)$ where $P$ is a principal $G$-bundle, $s\in H^0(X,V(P))$ is a global section of the corresponding vector bundle associated to $V$ and $P$. Let $\gg$ be the Lie algebra of $G$. Assume that we are given a symmetric invariant tensor $t\in S^2(\gg)^{\gg}$. Then $t$ induces a morphism of $\gg$-modules $t_*:S^2(V)\rightarrow S^2(V)$ as follows. Let $t=\sum_i x_i\otimes y_i$, then $$t_*(v\otimes v)=\sum (x_i\cdot v)\otimes (y_i\cdot v)$$ where $\cdot$ denotes the action of $\gg$ on $V$. Assume that $t_*=0$. Then fixing a trivialization $\omega_X\simeq\O_X$ we can construct a Poisson bracket on the smooth locus of ${\cal M}$ as follows. The tangent space to ${\cal M}$ at a point $(P,s)$ can be identified with the hypercohomology space $H^1(X,C^\cdot)$ where $C^{\cdot}$ is the complex $\gg(P)\stackrel{d}{\rightarrow} V(P)$ concentrated in degrees $0$ and $1$, where $\gg(P)$ is the vector bundle associated with $P$ and the adjoint representation, the map $d$ is induced by the Lie action of $\gg(P)$ on $V(P)$: $d(A)=A\cdot s$. Hence, the contangent space can be identified with $H^1(C^*[-1])$. Now we can construct the morphism of complexes $\phi:C^*[-1]\rightarrow C$ as before setting $\phi_1=0$ and $\phi_0:V^*(P)\rightarrow\gg(P)$ to be the composition of $d^*$ with the map $\gg^*(P)\rightarrow\gg(P)$ induced by $t$. We claim that our condition on $t$ and $V$ implies that $d\circ\phi_0=0$ and that the obtained morphism $H$ from the cotangent space of ${\cal M}$ to the tangent space is skew-symmetric. Indeed, essentially we have to check that for every $v\in V$ the following composition is zero: $$V^*\stackrel{d_v^*}{\rightarrow}\gg^*\stackrel{t}{\rightarrow}\gg \stackrel{d_v}{\rightarrow} V$$ where $d_v(A)=A\cdot v$. This is equivalent to the condition $t_*(v\otimes v)=0$. Now the skew-symmetry follows as before: the homotopy beteen $\phi$ and $\phi^*[-1]$ is constructed using the map $\gg^*(P)\rightarrow\gg(P)$ induced by $t$. \begin{thm} The above construction defines a Poisson bracket on the smooth locus of ${\cal M}$. \end{thm} \noindent {\it Proof} . We have to check the Jacoby identity for our bracket. We will use the approach similar to that of \cite{Bo}, \cite{Bo2}. The Jacoby identity can be rewritten in terms of the morphism $H:T_{{\cal M}}^*\rightarrow T_{{\cal M}}$ as follows: \begin{equation}\label{Jacobi} H(\omega_1)\cdot\langle H(\omega_2),\omega_3\rangle-\langle [H(\omega_1),H(\omega_2)], \omega_3\rangle+ cp(1,2,3)=0 \end{equation} where $\omega_i\in T_{{\cal M}}^*$ are local 1-forms on ${\cal M}$, $[\cdot,\cdot]$ is the commutator of vector fields, $cp(1,2,3)$ indicates terms obtained by cyclic permutation of 1,2 and 3 from the first two terms. Working over an affine \'etale open $U\rightarrow{\cal M}$ we can represent every 1-form $\omega\in T_{{\cal M}}^*(U)$ by a Cech cocycle $(\phi_{ij},\psi_i)$ for some open covering $\{ U_i\}$ of $U\times X$, where $\phi_{ij}\in \Gamma(U_i\cap U_j,V^*(P))$, $\psi_i\in\Gamma(U_i,\gg^*(P))$ are such that $-d^*\phi_{ij}=\psi_j-\psi_i$ over $U_i\cap U_j$, $P$ is the universal $G$-bundle on ${\cal M}$. Similarly, every vector field $v\in T_{{\cal M}}(U)$ can be represented by a Cech cocycle $(\a_{ij},\nu_i)$, where $\a_{ij}\in\Gamma(U_{ij},\gg(P))$, $\nu_i\in\Gamma(U_i, V(P))$ are such that $d\a_{ij}=\nu_j-\nu_i$. In terms of these representatives the pairing between $T^*_{{\cal M}}$ and $T_{{\cal M}}$ takes form \begin{equation} \langle (\a_{ij},\nu_i),(\phi_{ij},\psi_i)\rangle= \operatorname{Tr}(\langle\a_{ij},\psi_j\rangle+\langle\phi_{ij},\nu_i\rangle), \end{equation} where $\operatorname{Tr}:H^1(U\times X,\O_{U\times X})\rightarrow H^0(U,\O_U)$ is the morphism induced by the trivialization of $\omega_X$. The map $H$ sends a 1-form $(\phi_{ij},\psi_i)$ to the vector field represented by the cocycle $(t\circ d^*\phi_{ij},0)$ where $t$ is considered as a map $\gg^*\rightarrow\gg$. Since $d^*\phi_{ij}=\psi_i-\psi_j$ we have $$H(\phi_{ij},\psi_i)=(0,d\circ t(\psi_i))\operatorname {mod}(\operatorname{im}(\delta))$$ where $\delta$ is the differential in the Cech complex of $C^{\cdot}$. Note that since $d\circ t\circ d^*=0$, we have $d\circ t(\psi_i)=d\circ t(\psi_j)$ over $U_i\cap U_j$, hence we obtain the global section $d\circ t(\psi_{\cdot})\in \Gamma(U\times X, V(P))$. It follows that \begin{equation} \langle H(\phi_{ij},\psi_i),(\phi'_{ij},\psi'_i)\rangle= \operatorname{Tr}(\langle\phi'_{ij},d\circ t(\psi_i)\rangle)= \operatorname{Tr}(\langle\psi'_i-\psi'_j, t(\psi_i)\rangle. \end{equation} Let us consider the relative Atiyah extension for the universal bundle $P$: $$0\rightarrow\gg(P)\rightarrow\AA(P)\rightarrow p^*T_{{\cal M}}\rightarrow 0.$$ where $p:{\cal M}\times X\rightarrow{\cal M}$ is the projection, $\AA(P)$ is the bundle of relative infinitesemal symmetries of $P$. Then for sufficiently fine covering $\{ U_i\}$ a Cech cocycle $(\a_{ij},\nu_i)$ representing a local vector field on ${\cal M}$ can be written as follows: $\a_{ij}=D_j-D_i$, $\nu_i=D_i(s)$ where $D_i\in\Gamma(U_i,\AA(P))$, $s\in V(P)$ is the universal section, the symbol of $D_i$ is equal to the restriction of a given vector field to $U_i$. In particular, for a vector field represented by a cocycle $(0,\nu)$ where $\nu\in\Gamma(U\times X, V(P))$ we have $\nu=D(s)$ for some $D\in\Gamma(U\times X,\AA(P))$. After these remarks we can start proving (\ref{Jacobi}). Let us denote by $(\phi^h_{ij},\psi^h_i)$ Cech cocycles representing 1-forms $\omega_h$ for $h=1,2,3$. Let $D^h\in\Gamma(U\times X,\AA(P))$ be infinitesemal symmetries corresponding to $H(\omega_h)$ so that $D^h(s)=d\circ t(\psi_i)=t(\psi_i)(s)$ over $U_i$. Then we have $$H(\omega_1)\cdot\langle H(\omega_2),\omega_3\rangle= D^1\cdot\operatorname{Tr}(\langle\psi^3_i-\psi^3_j, t(\psi^2_i)\rangle)= \operatorname{Tr}(\langle D^1(\psi^3_i-\psi^3_j), t(\psi^2_i)\rangle+ \langle \psi^3_i-\psi^3_j, D^1(t(\psi^2_i))\rangle).$$ On the other hand, it is easy to compute the commutator: $$[H(\omega_1),H(\omega_2)]=(0,D^1D^2(s)-D^2D^1(s))= (0,D^1(t(\psi^2_i)(s))-D^2(t(\psi^1_i)(s))).$$ Hence, we have \begin{align*} &\langle [H(\omega_1),H(\omega_2)], \omega_3\rangle= \operatorname{Tr}\langle\phi^3_{ij}, D^1(t(\psi^2_i)(s))-D^2(t(\psi^1_i)(s))\rangle=\\ &\operatorname{Tr}\langle\phi^3_{ij},D^1(t(\psi^2_i))(s)-D^2(t(\psi^1_i))(s) +t(\psi^2_i)(D^1(s))-t(\psi^1_i)(D^2(s))\rangle= \\ &\operatorname{Tr}\langle\psi^3_i-\psi^3_j,D^1(t(\psi^2_i))-D^2(t(\psi^1_i))- [t(\psi^1_i),t(\psi^2_i)]\rangle. \end{align*} It follows that \begin{align*} &H(\omega_1)\cdot\langle H(\omega_2),\omega_3\rangle- \langle [H(\omega_1),H(\omega_2)], \omega_3\rangle=\\ &\operatorname{Tr}(\langle D^1(\psi^3_i-\psi^3_j), t(\psi^2_i)\rangle+ \langle\psi^3_i-\psi^3_j,t(D^2(\psi^1_i))+ [t(\psi^1_i),t(\psi^2_i)]\rangle=\\ & \operatorname{Tr}(\langle D^1(\psi^3_i-\psi^3_j), t(\psi^2_i)\rangle+ \langle D^2(\psi^1_i), t(\psi^3_i-\psi^3_j)\rangle+ \langle\psi^3_i-\psi^3_j,[t(\psi^1_i),t(\psi^2_i)]\rangle) \end{align*} Since this is equal to $D^1\cdot\operatorname{Tr}\langle\phi^3_{ij},dt(\psi^2_{\cdot})\rangle- \operatorname{Tr}\langle\phi^3_{ij},[D^1,D^2](s)\rangle$ which is skew-symmetric in $i,j$, we can also skew-symmetrize in $i,j$ the expression obtained above. Then after adding terms obtained by cyclic permutation of $1$, $2$ and $3$ we obtain the trace of the following expression $$\langle\psi^3_i-\psi^3_j, [t(\psi^1_i),t(\psi^2_i)]+[t(\psi^1_j),t(\psi^2_j)]\rangle+ cp(1,2,3).$$ Up to a coboundary this is equal to $$\langle\psi^3_i,\psi^1_j,\psi^2_j\rangle_t- \langle\psi^3_j,\psi^1_i,\psi^2_i\rangle_t+ cp(1,2,3),$$ where we denote $\langle x,y,z\rangle_t=\langle x, [t(x),t(y)]\rangle$. From the fact that $t\in (S^2\gg)^*$ one can deduce easily that $\langle \cdot,\cdot,\cdot\rangle_t$ is $\gg$-invariant and skew-symmetric. This implies the following identity: $$\langle\psi^3_i,\psi^1_j,\psi^2_j\rangle_t- \langle\psi^3_j,\psi^1_i,\psi^2_i\rangle_t+ cp(1,2,3) =\langle\psi^1_i-\psi^1_j,\psi^2_i-\psi^2_j,\psi^3_i- \psi^3_j\rangle_t= -\langle d^*\phi^1_{ij},d^*\phi^2_{ij},d^*\phi^3_{ij}\rangle_t.$$ It remains to notice that $\langle d^*\phi^1, d^*\phi^2, d^*\phi^3\rangle_t=0$ for any $\phi^1,\phi^2,\phi^3\in V^*(P)$. Indeed, we have to show that for any triple of elements $\varphi_1,\varphi_2,\varphi_3\in V^*$ and any $v\in V$ one has $$\langle d^*_v\varphi_1, d^*_v\varphi_2, d^*_v\varphi_3\rangle_t=0$$ where $d^*_v\varphi_h\in\gg^*$ is defined by $d^*_v\varphi_h(x)=\varphi_h(x\cdot v)$, $h=1,2,3$. Let $t=\sum_i x_i\otimes y_i$. Then \begin{align*} & \langle d^*_v\varphi_1,d^*_v\varphi_2,d^*_v\varphi_3\rangle_t= \langle d^*\varphi_1, \sum_{i,j}[d^*\varphi_2(y_i)x_i,d^*\varphi_3(y_j)x_j]\rangle=\\ &\sum_{i,j}\varphi_1([x_i,x_j]\cdot v)\varphi_2(y_i\cdot v) \varphi_3(y_j\cdot v). \end{align*} Now $$\sum_{i,j} [x_i,x_j]\cdot v\otimes y_i\cdot v\otimes y_j\cdot v= \sum_{i,j}x_ix_jv\otimes y_iv\otimes y_jv- \sum_{i,j}x_jx_iv\otimes y_iv\otimes y_jv=0$$ since $\sum_i x_iv\otimes y_iv=0$. \qed\vspace{3mm} Notice that the condition $t_*=0$ is usually not satisfied when $\gg$ is simple. However, for example if $S^2(V)$ is irreducible and $\gg$ is simple then we necessarily have $t_*=\lambda\cdot\operatorname{id}$ for some scalar $\lambda$. It follows that we can replace $G$ by its product $G\times{\Bbb G}_m$ with one-dimensional torus, $t$ by its sum with the appropriate multiple of the square of the generator of $\operatorname{Lie}({\Bbb G}_m)$, so that for the new tensor $t'$ and the same representation $V$ (on which ${\Bbb G}_m$ acts via identity character) the condition $t'_*=0$ will be satisfied. In the case $G=\operatorname{GL}_{r_1}\times \operatorname{GL}_{r_2}$ and $V=\operatorname{Mat}(r_1,r_2)$ the tensor $t$ is equal to $(t_1, -t_2)$ where $t_1=\sum E_{ij}\otimes E_{ji}$ is the standard symmetric invariant tensor for ${\frak gl}_{r_1}$, $t_2$ is the similar tensor for ${\frak gl}_{r_2}$. An interesting case is $G=\operatorname{GSp}_{2r}$, the group of invertible matrices preserving the symplectic from up to a scalar. In this case we can take $V$ to be the standard representation of $G$ of rank $2r$, then $S^2(V)$ is isomorphic to the adjoint representation of $\Sp_{2r}$ on which $\operatorname{GSp}_{2r}$ acts naturally, and one can easily find a non-zero invariant tensor $t\in (S^2\gg)^{\gg}$ with $t_*=0$ (in fact, such $t$ is unique up to a constant). The corresponding moduli stack is the stack of the following data: a vector bundle $E$ together with a symplectic form $$E\otimes E\rightarrow L$$ inducing an isomorphism $E\simeq E^*\otimes L$, and a section $s:\O_X\rightarrow E$. In particular, taking $E$ to be a fixed bundle with a symplectic form as above, we can consider sometimes the appropriate quotient space of $\P H^0(E)$ by the group of $\operatorname{GSp}$-automorphisms of $E$ as a Poisson substack in the above stack. More precisely, we can define the stability condition for such pairs $(E, s)$ depending on a parameter $\tau$: the only difference with the case of $\operatorname{GL}$ is that one should consider totally isotropic subbundles of $E$. Then for $\tau=\mu(E)+\epsilon$ we have the Casimir morphism from such moduli space to the stack of semistable $\operatorname{GSp}$-bundles, hence its fibers inherit the Poisson structure. For example, if $E_0$ is a stable bundle of degree 2 then there is a natural $\operatorname{GSp}_4$-structure on $E=E_0\oplus E_0$ such that both summands are totally isotropic. Then the $\tau$-stability condition (with $\tau=\mu(E_0)+\epsilon)$ allows only sections $s\in H^0(E)= H^0(E_0)\oplus H^0(E_0)$ with non-zero projections to both summands. Hence, the space of such sections up to the action of symplectic automorphisms of $E$ is $S^2\P H^0(E)$, so we get a Poisson structure on the latter variety.
1999-02-08T22:55:15
9712
alg-geom/9712028
en
https://arxiv.org/abs/alg-geom/9712028
[ "alg-geom", "math.AG" ]
alg-geom/9712028
Victor Vinnikov
Joseph A. Ball (Virginia Tech) and Victor Vinnikov (Weizmann Institute)
Zero-pole interpolation for matrix meromorphic functions on a compact Riemann surface, and a matrix Fay trisecant identity
AMS-LaTeX, 37 pages; final version (several references added and some misprints corrected), to appear in Amer. J. of Math
null
null
null
null
This paper presents a new approach to constructing a meromorphic bundle map between flat vector bundles over a compact Riemann surface having a prescribed Weil divisor (i.e., having prescribed zeros and poles with directional as well as multiplicity information included in the vector case). This new formalism unifies the earlier approach of Ball-Clancey (in the setting of trivial bundles over an abstract Riemann surface) with an earlier approach of the authors (where the Riemann surface was assumed to be the normalizing Riemann surface for an algebraic curve embedded in ${\bold C}^2$ with determinantal representation, and the vector bundles were assumed to be presented as the kernels of linear matrix pencils). The main tool is a version of the Cauchy kernel appropriate for flat vector bundles over the Riemann surface. Our formula for the interpolating bundle map (in the special case of a single zero and a single pole) can be viewed as a generalization of the Fay trisecant identity from the usual line bundle case to the vector bundle case in terms of Cauchy kernels. In particular we obtain a new proof of the Fay trisecant identity.
[ { "version": "v1", "created": "Tue, 23 Dec 1997 23:04:31 GMT" }, { "version": "v2", "created": "Mon, 8 Feb 1999 21:55:15 GMT" } ]
2007-05-23T00:00:00
[ [ "Ball", "Joseph A.", "", "Virginia Tech" ], [ "Vinnikov", "Victor", "", "Weizmann\n Institute" ] ]
alg-geom
\section{Introduction} \label{S:intro} The following zero-pole interpolation problem is one of the main objects of study in the recent monograph \cite{bgr}. We state here the simplest case where all zeros and poles are assumed to be simple and disjoint. {\sl Given a finite collection $\lambda^1, \dots, \lambda^{n_0}, \mu^1, \dots, \mu^{n_\infty}$ of distinct points in the complex plane ${\bold C}$, nonzero column vectors $u_1, \dots, u_{n_\infty} \in {\bold C}^{r \times 1}$ and nonzero row vectors $x_1, \dots, x_{n_0} \in {\bold C}^{1 \times r}$, find (if possible) an $r \times r$ matrix function $T(z)$ having value equal to the identity matrix $I$ at infinity such that (1) $T(z)$ is analytic on $({\bold C} \cup \{\infty\}) \backslash \{\mu^1, \dots, \mu^{n_\infty}\}$ and $T(z)^{-1}$ has analytic continuation to $({\bold C} \cup \{\infty \}) \backslash \{\lambda^1, \dots, \lambda^{n_0} \}$, (2) for $i=1, \dots, n_0$, $T(\lambda^i)$ has rank $r-1$ and $x_i T(\lambda^i) = 0$, and (3) for $j=1, \dots, n_\infty$, $T(\mu_j)^{-1}$ (i.e., the analytic continuation of $T(z)^{-1}$ to $z=\mu^j$) has rank $r-1$ and $T(\mu^j)^{-1} u_j = 0$.} A complete solution, along with numerous applications to problems in factorization, matrix interpolation and $H_\infty$-control, is given in \cite{bgr}. The solution (for this simple case) is as follows. {\sl A solution exists if and only if the $n_0 \times n_\infty$ matrix \begin{equation} \Gamma = [\Gamma_{ij}]\text{ with } \Gamma_{ij} = \dfrac{x_i u_j}{\mu^j - \lambda^i} \label{Gamma} \end{equation} is square and invertible. In this case the unique solution is given by \begin{equation} T(z) = I + \sum_{j=1}^{n_\infty} u_j (z-\mu^j)^{-1} c_j \label{genus0solution} \end{equation} where $c= \begin{bmatrix} c_1 & \dots & c_{n_\infty} \end{bmatrix} ^T$ is given by} \begin{equation} c= \Gamma^{-1} \begin{bmatrix} x_1^T & \dots & x_{n_0}^T \end{bmatrix}^T. \label{scalarcoefficients} \end{equation} The solution in \cite{bgr} uses system theory ideas, especially the state space similarity theorem specifying the level of uniqueness for two realizations of the same rational matrix function as the transfer function of a linear system. Later work (see \cite{bgrak}) handles the nonregular case (where $\det T(z)$ vanishes identically and the nature of the zero structure must be enlarged) by elementary linear algebra, without recourse to the state space similarity theorem. There have now appeared two seemingly distinct generalizations of this result to higher genus. In \cite{bc}, the problem is posed to construct a (global, single-valued) meromorphic matrix function on the compact Riemann surface $X$ satisfying conditions as in (1), (2) and (3) above. A matrix analogous to $\Gamma$ appears, but the solution criterion is not as simple; nevertheless, an explicit formula analogous to \eqref{genus0solution} and \eqref{scalarcoefficients} was found for the solution when it exists. The approach in \cite{bc} can be seen as an analogue of that in \cite{bgrak} (i.e., system theory ideas are avoided and a simple ansatz is used to reduce the problem to an analysis of a linear system of equations). The paper \cite{hip}, on the other hand, while formulating a more general problem (involving bundle maps between certain types of flat vector bundles rather than global meromorphic matrix functions) in an abstract setting, works primarily in a more concrete setting, where the Riemann surface $X$ is taken to be the normalizing Riemann surface for an algebraic curve $C$ having a determinantal representation \[ C = \{ (\lambda_1, \lambda_2) \in {\bold C}^2 \colon \det (\lambda_1 \sigma_2 - \lambda_2 \sigma_1 + \gamma) = 0 \} \] (where $\sigma_1, \sigma_2, \gamma$ are $M \times M$ matrices and $\lambda = (\lambda_1, \lambda_2)$ are affine coordinates), and the input and output bundles $E$ and $\widetilde{E}$ are assumed to have kernel representations, e.g., \[ E(\lambda) = \{ v \in {\bold C}^M \colon (\lambda_1 \sigma_2 - \lambda_2 \sigma_1 + \gamma) v = 0\}. \] In this setting, a non-metric version of the several-variable system theory connected with the model theory for commuting operators due to Livsic (i.e., a version with all Hilbert space inner products dropped) applies, any meromorphic bundle map satisfying appropriate conditions at the points at infinity can be realized as the transfer function, or the joint characteristic function, of a Livsic-Kravitsky 2D system, and the zero-pole bundle-map interpolation problem can be solved using the state-space similarity theorem for this setting in a manner completely parallel to that of \cite{bgr} for the genus 0 case. (For a recent systematic treatment of the Livsic theory, we refer to \cite{lkmv} and \cite{vinsurvey}). In this solution there is a matrix $\Gamma$ analogous to the $\Gamma$ in \eqref{Gamma} along with an explicit formula for the solution (when it exists) as in \eqref{genus0solution}. In this setting, one specifies an output bundle $\widetilde{E}$ having a kernel bundle representation as well as the zero-pole interpolation data. The invertibility of $\Gamma$ is then equivalent to the existence of an input bundle $E$ also having a kernel representation together with a bundle map $T \colon E \to \widetilde{E}$ meeting the zero-pole interpolation conditions. The purpose of this paper is to synthesize these two approaches. We obtain a generalization of the approach of \cite{bc} which handles the vector bundle problem, and clarify the solution criterion as well as the role of the invertibility of $\Gamma$ in this abstract setting. To obtain an analogue of the basic ansatz in \cite{bc} used for the form of the solution for the general bundle case, we need a version of the Cauchy kernel $(z,w) \to \frac{1}{z-w}$ for sections of a flat vector bundle $\chi$ satisfying $h^0(\chi \otimes \Delta) = 0$ where $\Delta$ is a line bundle of differentials of order 1/2 (a theta characteristic or a spin structure). This object was introduced in \cite{hip} (see also \cite{AV2}, \cite{vin1} for the line bundle case) but a proof of its existence for the vector bundle case relied on the theory of determinantal representations of algebraic curves and of kernel representations of bundles over such curves. Here we give a simple, direct existence proof using only some cohomology theory of vector bundles and the Riemann-Roch theorem. A similar proof of the same result for the line bundle case is given in \cite{raina1} and \cite{raina2}; the general case is also handled in \cite{newfay} but by using completely different techniques (involving the theory of the Green's function for the heat equation over $X$). Various other forms of the Cauchy kernel for a Riemann surface have appeared earlier in the literature, in particular in connection with the Riemann-Hilbert problem (see \cite{Rodin}, \cite{Zverovich}). However, these are developed within the framework of meromorphic differentials whereas our Cauchy kernel is defined as a multiplicative meromorphic differential of order $1/2$. The use of half-order differentials has the advantage that no extraneous poles are introduced in the Cauchy kernel. We mention that the paper \cite{AV} applies our Cauchy kernel to the study of indefinite Hardy spaces on a finite bordered Riemann surface. Taking the constant term in the Laurent expansion of the Cauchy kernel around the diagonal allows us to define a certain flat connection on the flat bundle $\chi$. In the concrete setting of the determinantal representations, this connection was already introduced in \cite{hip}. This flat connection is determined canonically up to a choice of a bundle $\Delta$ of half-order differentials. We also make explicit the mappings between the concrete and abstract settings; in this way we are able to see explicitly the equivalence between the solution in \cite{hip} and the solution in \cite{bc}. The main ingredient is the explicit formula for the determinantal representation of an algebraic curve with a given kernel bundle in terms of the Cauchy kernel of the bundle. We also specialize the results to the line bundle case. For this case, both the solution to the zero-pole interpolation problem and the Cauchy kernel can be expressed explicitly in terms of theta functions (see \cite{oldfay}, \cite{mumford} and \cite{farkaskra} for background material on theta functions). When this is done, the equality between these two forms of the solution of the interpolation problem leads to a new proof of the trisecant identity due to Fay (see \cite{oldfay} Corollary 2.19 or \cite{mumford} Volume II page 3.214). In the general vector bundle case, the formula for the solution of the interpolation problem in terms of the Cauchy kernels (in the case of a single zero and a single pole---see \eqref{3.12a}) can be viewed as a matrix version of the Fay trisecant identity. We mention that in the genus 1 case one can obtain an explicit formula for the Cauchy kernel for the general case of flat vector bundles (see \cite{bcv}). We close the introduction by mentioning three other possible further applications of our Cauchy kernel. First of all, as will be shown in Section \ref{S:detrep}, the vector bundle $\chi$ is completely determined by the values $K(\chi; x^i,x^j)$ of the Cauchy kernel at a certain finite collection of points $x^1, \dots, x^m$ (forming a line section in a birational planar embedding of $X$). This suggests that these values can be used as affine coordinates for the bundle $\chi$ in the corresponding moduli space of semistable bundles on the complement of the generalized theta divisor. A very similar construction for line bundles on hyperelliptic curves is due to Jacobi (see \cite{Jacobi}) and has been given a modern treatment by Mumford (see Volume II of \cite{mumford}). Secondly, in the line bundle case consideration of the Fay trisecant identity when some of the points come together leads to very interesting identities, showing in particular that the theta function satisfies the KP equations (see \cite{oldfay}, Volume II of \cite{mumford} and \cite{Shiota}). An interesting line of research is to consider similar limiting versions of our matrix Fay trisecant identity \eqref{3.12a}. A related problem is to find the relation between the Cauchy kernel and the matrix Baker-Akhiezer function of Krichever and Novikov \cite{KrNo} whose definition involves the so called Tjurin parameters \cite{Tju1,Tju2} of the vector bundle. Thirdly, the absence of explicit formulas for the Cauchy kernel makes it interesting to try to find formulas for the Cauchy kernel of one vector bundle in terms of the Cauchy kernel for another. In particular it would be interesting to find how the Cauchy kernel behaves under pullback and direct image. The paper is organized as follows. Section \ref{S:intro} is this introduction. Section \ref{S:cauchyker} develops the Cauchy kernel for a flat line bundle. Section \ref{S:absint} then formulates and solves the zero-pole interpolation problem in the abstract setting. Section \ref{S:linebundle} obtains explicit formulas for all the results in the line bundle case and obtains the new proof of the Fay trisecant identity. Section \ref{S:detrep} explains how to use Cauchy kernels to obtain a canonical map from the abstract to the concrete setting. Finally, Section \ref{S:conint} explains the connections with the concrete interpolation problem solved in \cite{hip}. \section{The Cauchy kernel for a flat vector bundle} \label{S:cauchyker} We assume that we are given a compact Riemann surface $X$. Let $\Delta$ be a line bundle of differentials of order $\frac{1}{2}$ on $X$, i.e., a line bundle satisfying $\Delta \otimes \Delta \cong K$, where $K$ is the canonical line bundle (i.e., the line bundle with local holomorphic sections equal to local holomorphic differentials on $X$). Note that since $\deg(K) = 2g-2$, $\deg(\Delta) = g-1$ where $g$ is the genus of $X$. In addition we assume that we are given a holomorphic complex vector bundle $\chi$ of degree $0$ (and of rank $r$, say) over $X$ such that \[ h^0(\chi \otimes \Delta) = 0, \] i.e., $\chi \otimes \Delta$ has no nonzero global holomorphic sections. The condition implies that $\chi$ is necessarily semistable, and means that the (equivalence class of) $\chi$ lies on the complement of the generalized theta divisor in the moduli space of semistable vector bundles of rank $r$ and degree $0$ on $X$ (see \cite{newfay}, \cite{Seshadri} and \cite{Drezet}). It also follows immediately from Weil's criterion for flatness \cite{Gunning} that $\chi$ is actually a flat vector bundle (see \cite{hip} page 275 for details). By definition, since $\chi$ is flat, sections $h$ of $\chi$ have the property that they lift to ${\bold C}^r$-vector functions $\widetilde{h}$ defined on the universal cover $\widetilde X$ of $X$ such that \[ h(R \widetilde{p}) = \chi(R) h(\widetilde{p}) \] for all $\widetilde{p} \in \widetilde{X}$ where $R$ is any element of the group of deck transformations Deck($\widetilde{X}/X)$ $ \cong \pi_1(X)$ and where $R \to \chi(R) \in GL(r)$ is a (constant) factor of automorphy associated with the bundle $\chi$. (We somewhat abuse the notation denoting a constant factor of automorphy and the corresponding flat vector bundle by the same letter.) The main object of this section is to define an object $K(\chi; \cdot, \cdot)$ associated with any such bundle $\chi$ which we shall call the {\it Cauchy kernel} for the bundle $\chi$. Let $M$ denote the Cartesian product $M=X \times X$ and let $\pi_1 \colon M \to X$ be the projection map onto the first coordinate and $\pi_2 \colon M \to X$ the projection onto the second coordinate. The defining property of $K(\chi; \cdot, \cdot)$ is that $K(\chi; \cdot, \cdot)$ be a meromorphic mapping of the vector bundles $\pi_2^* \chi$ and $\pi_1^* \chi \otimes \pi_1^*\Delta \otimes \pi_2^* \Delta$ on $M$ which is holomorphic outside of the diagonal ${\cal D} = \{(p,p) \in M \colon p \in X\}$, where it has a simple pole with residue $I_r$. More precisely, the latter condition means the following: for any $\widetilde{p}_0 \in \widetilde{X}$ and any local parameter $t$ on $X$, if we let $\sqrt{dt}$ be the corresponding local holomorphic frame for $\Delta$ lifted to a neighborhood of $\widetilde{p}_0$ on $\widetilde{X}$, then near $(\widetilde{p}_0, \widetilde{p}_0) \in \widetilde{X} \times \widetilde{X}$ the lift of $K(\chi; \cdot, \cdot)$ to $\widetilde{X} \times \widetilde{X}$ has the form \[ \frac{ K(\chi; \widetilde{p},\widetilde{q}) }{ \sqrt{dt}(\widetilde{p}) \sqrt{dt}(\widetilde{q}) } = \frac{1}{ t(\widetilde{p}) - t(\widetilde{q}) } \left[ I_r + O\left(\sqrt{ |t(\widetilde{p})|^2 + |t(\widetilde{q})|^2 } \right) \right]. \] Thus, if $e$ is in the fiber of $\chi$ at a point $q$ on $X$, and $t$ is a local parameter of $X$ centered at $q$, then $K(\chi; \cdot, q) \frac{e}{\sqrt{dt}(q)}$ is a meromorphic section of $\chi \otimes \Delta$ that has a single simple pole at $q$, with a residue (in terms of the local parameter $t$) equal to $e \sqrt{dt}(q)$. Note that since $\chi \otimes \Delta$ has no nontrivial global holomorphic sections, such a meromorphic section is unique whenever it exists. Note that when $X$ is the Riemann sphere and $\chi$ is (necessarily) trivial, then \[K(\chi;p,q) = \frac{I_r}{t(p) - t(q)} \sqrt{dt}(p) \sqrt{dt}(q) \] where $t$ is the standard coordinate on the complex plane, i.e., we get the usual Cauchy kernel. Existence was shown in \cite{hip} by exhibiting an explicit formula for $K(\chi; \cdot, \cdot)$; the construction involved using a representation of $X$ as the normalized Riemann surface for an algebraic curve $C$ embedded in ${\bold P}^2$ and representing the bundle $E = \chi \otimes \Delta \otimes {\cal O}(1)$ as the kernel bundle associated with a determinantal representation of the curve $C$ \begin{gather} C = \{[\mu_0,\mu_1,\mu_2] \in {\bold P}^2 \colon \det (\mu_1\sigma_2 - \mu_2 \sigma_1 + \mu_0 \gamma) = 0\} \notag \\ E(\mu) = \ker (\mu_1\sigma_2 - \mu_2 \sigma_1 + \mu_0 \gamma). \notag \end{gather} When the rank $r$ of the vector bundle $\chi$ is 1, one can get an explicit formula (in terms of the Abel-Jacobi map and classical theta functions on the Jacobian variety of $X$) for the Cauchy kernel (see \cite{hip}). Details of this formula will be reviewed in Section \ref{S:linebundle} of this paper, where other special aspects of the line bundle case will also be discussed. Our purpose in this section is to give an alternative proof of the existence of such a Cauchy kernel $K(\chi; \cdot, \cdot)$ for a flat vector bundle $\chi$ by a direct, simple, more abstract argument (without relying on representing $X$ as the normalizing Riemann surface for a curve $C$ having a determinantal representation as in \cite{hip}). This is the content of the following theorem. \begin{theorem} \label{T:cauchyker} Let $\chi$ be a flat vector bundle over the Riemann surface $X$ with $h^0(\chi \otimes \Delta) = 0$ as above. Then the Cauchy kernel $K(\chi; \cdot, \cdot)$ exists, i.e., there is a unique meromorphic mapping of the vector bundles $\pi_2^* \chi$ and $\pi_1^* \chi \otimes \pi_1^*\Delta \otimes \pi_2^* \Delta$ on $M=X \times X$ which is holomorphic outside of the diagonal ${\cal D} = \{(p,p) \in M \colon p \in X\}$, where it has a simple pole with residue $I_r$. \end{theorem} \begin{pf} Note that we have the following exact sequence of vector bundles over $M=X \times X$: \begin{equation} \label{exseq1} 0 \to {\cal O}(-{\cal D}) \to {\cal O} \to {\cal O}|_{\cal D} \to 0. \end{equation} For ease of notation, define vector bundles $F$ and $W$ over $M$ and $V$ and $K$ over $X$ by \begin{gather} F= \pi_1^* \chi \otimes \pi_1^* \Delta \otimes \pi_2^* \chi^\vee \otimes \pi_2^* \Delta \notag \\ W = F \otimes {\cal O}({\cal D}) \notag \\ V = \chi \otimes \Delta \notag \\ K = \text{ the canonical line bundle on } X \notag \end{gather} where $\chi^\vee$ is the dual bundle of $\chi$. Tensoring the exact sequence \eqref{exseq1} with $W$ gives us \begin{equation} \label{exseq2} 0 \to F \to W \to W \otimes {\cal O}|_{\cal D} \to 0. \end{equation} The map of taking the residue along the diagonal defines a linear mapping $$ {\cal R} \colon H^0(M,W) \rightarrow H^0(X, End\ V). $$ Our goal is to show that there exists a unique element $K(\chi; \cdot, \cdot)$ of $H^0(M,W)$ so that ${\cal R}(K(\chi; \cdot, \cdot)) = I_V$. Note that the bundle $W \otimes {\cal O}|_{\cal D}$ can be identified with the bundle $End\ V$ of endomorphisms of $V$. Moreover the residue mapping ${\cal R}$ is exactly the mapping from $H^0(M,W)$ into $H^0(X, End\ V)$ induced by the mapping $W \rightarrow \left. W \otimes {\cal O}\right|_{\cal D}$ in \eqref{exseq2}. The vector bundle exact sequence \eqref{exseq2} induces (see the Basic Fact on page 40 of \cite{gh}) the exact cohomology sequence \begin{align} \notag 0 \to H^0(M,F) & \to H^0(M,W) \to H^0(X, End\ V) \to \\ \to H^1(M, F) & \to \cdots. \label{exseq3} \end{align} Next we argue that (i) $h^0(F) = 0$ and (ii) $h^1(F) = 0$. The statement (i) follows easily from our assumption that $h^0(V) = 0$. As for statement (ii) it follows from the Kunneth formulas (see page 58 of \cite{gh}) that \begin{align} H^1(M,F) = & H^1(M, \pi_1^*(V) \otimes \pi_2^*( V^\vee \otimes K)) \notag \\ \cong & \left(H^0(M, \pi_1^*(V)) \otimes H^1(M, \pi_2^*(V^\vee \otimes K)) \right) \\ & \oplus \left( H^1(M, \pi_1^*(V)) \otimes H^0(M, \pi_2^*(V^\vee \otimes K) ) \right). \label{kunneth} \end{align} By our assumption that $h^0(V)=0$ it follows that the first term on the right hand side of \eqref{kunneth} is 0. Since $h^0(V) = 0 $ we also have $h^0(V^\vee \otimes K) = 0$ as well, by the Riemann-Roch Theorem for vector bundles on an algebraic curve (see \cite{Gunning}) and the assumption that deg $V=r(g-1)$. Hence the second term in \eqref{kunneth} is zero as well. This verifies the desired fact (ii). Hence the exact sequence \eqref{exseq3} collapses to \begin{equation} \label{exseq4} 0 \to H^0(M,W) \overset{{\cal R}}{\to} H^0(X,End\ V) \to 0. \end{equation} It follows that ${\cal R} \colon H^0(M,W) \to H^0(X,End\ V)$ is an isomorphism and the Theorem follows. \end{pf} Let $\widetilde{p}_0$ be an arbitrary point of $\widetilde{X}$, and let $t$ be a local coordinate for $\widetilde{X}$ near $\widetilde{p}_0$. Then by definition the Cauchy kernel $K(\chi; \cdot, \cdot)$ is such that \[ \left(t(\widetilde p) - t(\widetilde q) \right) \frac{K(\chi; \widetilde p, \widetilde q)}{\sqrt{dt}(\widetilde p) \sqrt{dt}(\widetilde q)} \] is analytic in $(\widetilde p, \widetilde q)$ near $(\widetilde p_0, \widetilde p_0)$ with value at $(\widetilde p_0, \widetilde p_0)$ equal to $I_r$; hence, for $(\widetilde p, \widetilde q)$ close to $(\widetilde p_0, \widetilde p_0)$, $K(\chi; \cdot, \cdot)$ has a representation of the form \begin{multline} \label{cauchyexp} \frac{K(\chi; \widetilde p, \widetilde q)}{\sqrt{dt}(\widetilde p) \sqrt{dt}(\widetilde q)} \\ = \frac{1}{t(\widetilde p) - t(\widetilde q)} \left[ I_r + \frac{A_\ell}{dt}(\widetilde p_0) t(\widetilde p) + \frac{A}{dt}(\widetilde p_0) t(\widetilde q) + O\left( |t(\widetilde p)|^2 + |t(\widetilde q)|^2 \right) \right] \end{multline} for appropriate holomorphic matrix $K$-valued coefficients $A(\widetilde p_0)$ and $A_\ell(\widetilde p_0)$. An important point for us is the following result. \begin{lemma} \label{L:cauchyexp} Let $A_\ell$ and $A$ be the linear coefficients appearing in the Laurent expansion of the Cauchy kernel $K(\chi; \cdot, \cdot)$ as in \eqref{cauchyexp}. Then \[A(\widetilde p_0) + A_\ell(\widetilde p_0) = 0 \] for all $\widetilde p_0 \in \widetilde X$. \end{lemma} \begin{pf} Let $\widetilde{p}_0$ be an arbitrary point of $\widetilde X$ and let $t$ be a local coordinate for $\widetilde X$ near $\widetilde p_0$. For $\widetilde p \in X$ near $\widetilde p_0$, define $$ f(\widetilde p) = (t(\widetilde p) - t(\widetilde q)) \cdot \left. \frac{K(\chi; \widetilde p, \widetilde q)}{\sqrt{dt}(\widetilde p) \sqrt{dt}(\widetilde q)} \right|_{\widetilde p = \widetilde q}. $$ From \eqref{cauchyexp} we see that $$ f(\widetilde p) = I_r + \left[ \frac{A_\ell}{dt}(\widetilde p_0) + \frac{A}{dt}(\widetilde p_0)\right] t(\widetilde p) + O(|t(\widetilde p)|^2). $$ In particular, \begin{equation} \label{der1} \left. \frac{df}{dt}(\widetilde p) \right|_{\widetilde p = \widetilde p_0} = \frac{A_\ell}{dt}(\widetilde p_0) + \frac{A}{dt}(\widetilde p_0). \end{equation} On the other hand, we can use $t'(\widetilde p') = t(\widetilde p') - t(\widetilde p)$ as a local coordinate for the variable $\widetilde p'$ near the point $\widetilde p \in X$. From \eqref{cauchyexp} again we have \begin{multline} [(t(\widetilde p') - t(\widetilde p)) - (t(\widetilde q') - t(\widetilde p))] \frac{K(\chi; \widetilde p', \widetilde q')}{\sqrt{dt}(\widetilde p') \sqrt{dt}(\widetilde q')} \\ \notag = I_r + \frac{A_\ell}{dt}(\widetilde p) (t(\widetilde p') - t(\widetilde p)) + \frac{A}{dt}(\widetilde p) (t(\widetilde q') - t(\widetilde p)) + O(|t(\widetilde p') - t(\widetilde p)|^2 + |t(\widetilde q') - t(\widetilde p)|^2). \end{multline} Evaluation of both sides of this equation at $\widetilde p' = \widetilde q' = \widetilde p$ yields $f(\widetilde p) = I_r$ from which we get \begin{equation} \label{der2} \frac{df}{dt}(\widetilde p) = 0 \end{equation} for all $\widetilde p \in X$. Comparison of \eqref{der1} and \eqref{der2} now gives $A(\widetilde p_0) + A_\ell(\widetilde p_0) = 0$ as asserted. \end{pf} We can use these coefficients $A(\widetilde p)$ and $A_\ell(\widetilde q)$ defined by \eqref{cauchyexp} to define connections $\nabla_\chi$ on $\chi$ and $\nabla^*_\chi$ on $\chi^\vee$ according to the formulas \begin{align} \nabla_\chi y &= A y + dy \notag \\ \nabla^*_\chi x &= A_\ell^T x + dx. \label{connection} \end{align} for local holomorphic sections $y$ of $\chi$ and $x$ of $\chi^\vee$. The result $A_\ell + A = 0$ from Lemma \ref{L:cauchyexp} is equivalent to the fact that $\nabla_\chi$ and $\nabla_\chi^*$ are {\it dual to each other}, i.e. \[ d(x^Ty) = x^T (\nabla_\chi y) + (\nabla_\chi^* x)^T y \] for local holomorphic sections $y$ of $\chi$ and $x$ of $\chi^\vee$, where $(y, x) \to x^Ty$ is the pairing between $\chi$ and $\chi^\vee$. Moreover, from the formula for $\nabla_\chi$ we see that the connection matrix associated with $\nabla_\chi$ is a matrix of holomorphic $(1,0)$-forms. Hence, $\nabla_\chi$ is {\it compatible with the complex structure of} $X$ and moreover, since we are in complex dimension 1, the connection $\nabla_\chi$ is {\it flat}, i.e., $\nabla_\chi$ has {\it zero curvature} (see Section~5 of Chapter~0 of \cite{gh} for all relevant definitions). The existence of such a flat connection on $\chi$ in turn implies that $\chi$ itself is a flat vector bundle (see \cite{mst} pages 294--295). In general there are many choices of distinct flat connections on a flat vector bundle; our construction via the Cauchy kernel provides a canonical choice of such a flat connection (up to a choice of a bundle $\Delta$ of half-order differentials). An explicit formula for $\nabla$ in the line bundle case is given in Section \ref{S:linebundle}. {\bf Remark:} The proof of Theorem \ref{T:cauchyker} used only the fact that $\deg\chi = 0$ and did not use the flatness of $\chi$. Since the existence of a flat connection (compatible with the complex structure) implies that the bundle is flat, our construction gives a direct proof of the flatness of $\chi$ independent of Weil's theorem. \section{The abstract interpolation problem} \label{S:absint} In this section we consider as given two flat vector bundles $\chi$ and $\widetilde \chi$ over the Riemann surface $X$ for which both $h^0(\chi \otimes \Delta)=0$ and $h^0(\widetilde \chi \otimes \Delta)=0$, where again, $\Delta$ is a line bundle of half-order differentials over $X$. We are interested in studying pole-zero interpolation conditions imposed on a bundle map of $\chi$ to $\widetilde \chi$. The data for the interpolation problem is as follows. We assume that we are given $n_\infty$ distinct points $\mu^1, \dots, \mu^{n_\infty}$ (the prescribed poles) together with $n_0$ distinct points $\lambda^1, \dots, \lambda^{n_0}$ (the prescribed zeros). For each fixed index $j$ ($j=1, \dots, n_\infty$) we specify a linearly independent set $\{u_{j1}, \dots, u_{j,s_j}\}$ of $s_j$ vectors in the fiber $\widetilde \chi(\mu^j)$ of $\widetilde \chi$ over $\mu^j$ (the prescribed pole vectors) and for each fixed index $i$ ($i=1, \dots, n_0$) we specify a linearly independent set $\{x_{i1}, \dots, x_{i,t_i}\}$ of $t_i$ vectors in the fiber $\widetilde \chi^\vee(\lambda^i)$ of the dual bundle $\widetilde \chi^\vee$ of $\widetilde \chi$ (the prescribed null vectors). Also, for each pair of indices ($i,j$) for which $\lambda^i = \mu^j=:\xi^{ij}$, we specify a collection $\{\rho_{ij,\alpha\beta} \colon 1 \le \alpha \le t_i, 1 \le \beta \le s_j\}$ of numbers that depend on the choice of the local parameter at the point $\xi^{ij}$. The {\it Abstract Interpolation Problem} (ABSINT) which we study in this section is the following: {\it determine if there exists a bundle map $T \colon \chi \to \widetilde \chi$ with transpose $T^\vee \colon \widetilde \chi^\vee \to \chi^\vee$ such that: \begin{enumerate} \item[(i)] $T$ has poles only at the points $\{\mu^1, \dots, \mu^{n_\infty}\}$; for each $j=1, \dots, n_\infty$, the pole of $T$ at $\mu^j$ is simple, and the residue $R_j = \text{ Res }_{p=\mu^j} T \colon \chi(\mu^j) \to \widetilde \chi(\mu^j)$ of $T$ at $\mu^j$ is such that $\{u_{j1}, \dots, u_{j,s_j}\}$ spans the image space $\text{im } R_j$ of $R_j$. \item[(ii)] The bundle map $(T^\vee)^{-1} \colon \chi^\vee \to \widetilde \chi^\vee$ has poles only at $\{\lambda^1, \dots, \lambda^{n_0}\}$; for each $i=1, \dots, n_0$, the pole of $(T^\vee)^{-1}$ at $\lambda^i$ is simple and the residue $\widehat R_i = \text{ Res}_{p=\lambda^i}(T^\vee)^{-1} \colon \chi^\vee(\lambda^i) \to \widetilde \chi^\vee (\lambda^i)$ of $(T^\vee)^{-1}$ at $\lambda^i$ is such that $\{x_{i1}, \dots, x_{i,t_i}\}$ spans the image space $\text{ im } \widehat R_i$ of $\widehat R_i$. \item[(iii)] For each pair of indices ($i,j$) for which $\lambda^i = \mu^j =: \xi^{ij}$, and for $\alpha = 1, \dots, t_i$, let $x_{i \alpha}(p)$ be a local holomorphic section of $\widetilde \chi^\vee$ with $x_{i \alpha}(\xi^{ij}) = x_{i \alpha}$ such that $T^\vee(p)x_{i \alpha}(p)$ has analytic continuation to $p = \xi^{ij}$ with value at $p = \xi^{ij}$ equal to 0. Then \[ (\nabla^*_{\widetilde \chi} x_{i \alpha}(\xi^{ij}))^T u_{j \beta} = \rho_{ij, \alpha \beta} \] for $\beta = 1, \dots , s_j$. \end{enumerate} When such a bundle map $T$ exists, give an explicit formula for the construction of $T$.} In order for solutions to exist, the compatibility condition \begin{equation} \label{comp} x_{i \alpha}u_{j \beta} = 0 \text{ whenever } \lambda^i = \mu^j. \end{equation} must hold. This follows from the requirement that the meromorphic section $$ x_{i \alpha}(p)T(p) $$ be analytic at the point $p=\xi^{ij}:= \lambda^i = \mu^j$. Hence we shall always assume that our data collection \begin{equation} \label{dataset} {\boldsymbol \omega} = \{(x_{i\alpha},\lambda^i), (u_{j\beta}, \mu^j), \rho_{ij, \alpha \beta} \} \end{equation} also satisfies this compatibility condition. It will be convenient to work with an alternate form of the interpolation condition (iii) in (ABSINT). Suppose that $u(p) = T(p) \varphi(p)$, where $\varphi$ is a local holomorphic section of $\chi$ near $\xi^{ij}$ chosen so that $\text{Res}_{p=\xi^{ij}} u(p) = u_{j \beta}$ with respect to the local coordinate $t^{ij}$ centered at $\xi^{ij}$. Then \begin{align} \left(x_{i \alpha}(p) \right)^T\left( t^{ij}(p) u(p) \right) & = t^{ij}(p) \cdot \left(x_{i \alpha}(p)\right)^T T(p) \varphi(p) \\ & = \left( t^{ij}(p) T^\vee(p) x_{i \alpha}(p)\right)^T \varphi(p) \end{align} has a double order zero at $p=\xi^{ij}$, and hence \[ d\left( x_{i \alpha}(p)^T (t^{ij}(p) u(p) ) \right)|_{p=\xi^{ij}} = 0. \] Since $\nabla_{\widetilde{\chi}}^*$ and $\nabla_{\widetilde{\chi}}$ are dual connections as a consequence of Lemma \ref{L:cauchyexp}, we therefore have \[ \left( \nabla_{\widetilde{\chi}}^* x_{i \alpha}(\xi^{ij})\right)^T u_{j \beta} + x_{i \alpha}^T \nabla_{\widetilde{\chi}} (t^{ij}(p) u(p))|_{ p= \xi^{ij}} = 0. \] Thus the interpolation condition in part (iii) of (ABSINT) can be expressed alternatively as \begin{equation} \label{coupledint} x_{i \alpha}^T \nabla_{\widetilde{\chi}}(t^{ij}(p) u(p))|_{p = \xi^{ij}} = - \rho_{ij, \alpha \beta} \end{equation} where $u(p) = T(p) \varphi(p)$ for a local holomorphic section $\varphi$ of $\chi$ near $\xi^{ij}$ such that $\text{Res}_{p=\xi^{ij}}u(p) = u_{j \beta}$. In the scalar case ($r=1$), the compatibility condition \eqref{comp} can never be satisfied in a nontrivial way and hence the third set of interpolation conditions is absent under our assumptions; this corresponds to the fact that a scalar meromorphic function cannot have a zero and a pole at the same point $\xi^{ij}$. Moreover, for the case $r=1$, necessarily $t_i=1$ for all $i$ and $s_j=1$ for all $j$. In the case where both $\chi$ and $\widetilde{\chi}$ are trivial (or more generally if we use coordinates with respect to a local holomorphic frame for $\widetilde{\chi}$ near $\mu^j$ or $\widetilde{\chi}^\vee$ near $\lambda^i$), there is no loss of generality in taking $x_i:=x_{i1}=1$ and $u_j:=u_{j1}=1$ for all $i$ and $j$. Thus the only remaining relevant data are the zeros $\lambda^1, \dots \lambda^{n_0}$ and the poles $\mu^1, \dots, \mu^{n_\infty}$ (all assumed here to be distinct). As is standard in algebraic geometry, the formal sum \[ {\boldsymbol \lambda} - {\boldsymbol \mu}:= \lambda^1 + \dots + \lambda^{n_0} - \mu^1 - \dots - \mu^{n_\infty} \] is said to be a {\it divisor} on $X$. If $f$ is a meromorphic function, the associated principal divisor $(f)$ is defined to be the formal sum $ p^1 + \dots + p^{n_0} - q^1 - \dots - q^{n_\infty}$ where the $p^i$'s are the zeros of $f$ and the $q^j$'s are the poles of $f$ (with repetitions according to respective multiplicities). Associated with any divisor ${\boldsymbol \lambda} - {\boldsymbol \mu}$ as above is the vector bundle ${\cal O}({\boldsymbol \mu} - {\boldsymbol \lambda})$ whose holomorphic sections can be identified with global meromorphic functions $h$ such that \[ (h) \ge {\boldsymbol \lambda} - {\boldsymbol \mu}, \] i.e., such that the zeros of $h$ include the points $\lambda^1, \dots, \lambda^{n_0}$ (all with multiplicity at least 1) and the poles of $h$ are a subset of $\mu^1, \dots, \mu^{n_\infty}$ (all with multiplicity at most 1). It is convenient for us to introduce matrix analogues of these ideas. Let $\boldsymbol \omega$ be an interpolation data set as in \eqref{dataset}. Let us introduce the notation \begin{equation} \label{poledata} ({\boldsymbol \mu}, {\bold u}) = \{ (\mu^j, u_{j \beta}) \colon 1 \le j \le n_\infty, 1 \le \beta \le s_j \} \end{equation} for the pole part of $\boldsymbol \omega$. We let ${\cal M}(\widetilde{\chi}\otimes \Delta)$ be the sheaf of meromorphic sections of $\widetilde{\chi} \otimes \Delta$. For $U$ an open subset of $X$ we define \begin{align} \notag {\cal O}(\widetilde{\chi} \otimes \Delta)({\boldsymbol \mu}, {\bold u}) (U) = & \{u \in {\cal M}(\widetilde{\chi} \otimes \Delta ) (U) \colon u \text{ has poles only at } \mu^j, \\ & u_{-1}:=\text{Res}_{p=\mu^j} \in \text{span }\{u_{j \beta} \colon 1 \le \beta \le s_j\} \} \notag \end{align} and \begin{align} \notag {\cal O}(\widetilde{\chi}\otimes \Delta)({\boldsymbol \omega})(U) = & \{u \in {\cal O}(\widetilde{\chi} \otimes \Delta) ({\boldsymbol \mu}, {\bold u})(U) \colon \text{ if } \lambda^i \ne \mu^j, \text{ then } x_{i \alpha}u(\lambda^i) = 0; \\ \notag & \text{if } \lambda^i = \mu^j =:\xi^{ij}, \text{ there is a local holomorphic section } x_{i \alpha}(p) \text{ of } \widetilde{\chi}^\vee \\ \notag & \text{such that (i) } x_{i \alpha}(\xi^{ij}) = x_{i \alpha}, \text{ (ii) } x_{i \alpha}(p)^T \frac{u}{ \sqrt{ dt^{ij} } }(p) \text{ has analytic} \\ \notag & \text{continuation to } p = \xi^{ij} \text{with value } 0 \text{ there, and} \\ \notag & \text{ (iii) } x_{i \alpha}^T \nabla_{\chi}\left(t^{ij} \frac{u}{ \sqrt{dt^{ij}} }\right) = - \sum_{\beta = 1}^{s_j}\rho_{ij,\alpha \beta} c_\beta \frac{1}{ \sqrt{dt^{ij}} (\xi^{ij})} \\ \notag & \text{if Res}_{p=\mu^j} u = \sum_{\beta=1}^{s_j} u_{j \beta} c_\beta.\} \end{align} It is obvious that ${\cal O}(\widetilde{\chi} \otimes \Delta)({\boldsymbol \mu}, {\bold u})$ and ${\cal O}(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$ are locally free sheaves of rank $r$, and we denote the corresponding rank $r$ vector bundles by $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \mu}, {\bold u})$ and $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. It is also obvious that $T$ is a solution of (ABSINT) if and only if $T$ is an isomorphism from $\chi \otimes \Delta$ to $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. The solution of the zero-pole interpolation problem introduced at the beginning of this section is as follows. \begin{theorem} \label{T:absint} Define a $n_0 \times n_\infty$ block matrix $\Gamma = [\Gamma_{ij}]$ ($1 \le i \le n_0$, $1 \le j \le n_\infty$) where the block entry $\Gamma_{ij}$ in turn is a $t_i \times s_j$ matrix $\Gamma_{ij} = [\Gamma_{ij, \alpha \beta}]$ ($1 \le \alpha \le t_i$, $1 \le \beta \le s_j$) with matrix entries $\Gamma_{ij, \alpha \beta}$ given by \begin{equation} \label{defGamma} \Gamma_{ij, \alpha \beta} = \begin{cases} - x_{i \alpha}^T K(\widetilde{\chi};\lambda^i, \mu^j) u_{j \beta}, &\text{if } \lambda^i \ne \mu^j; \\ -\rho_{ij, \alpha \beta} & \text{if } \lambda^i = \mu^j. \end{cases} \end{equation} In addition we introduce the block matrices \begin{gather} \notag {\bold u}_i = \begin{bmatrix} u_{i1} & \dots & u_{i s_i} \end{bmatrix}, \quad {\bold x}_j = \begin{bmatrix} x_{j1} & \dots & x_{j t_i} \end{bmatrix}, \\ K_{{\boldsymbol \mu}, {\bold u}}(p) = \begin{bmatrix} K(\widetilde{\chi}; p, \mu^1) {\bold u}_1 & \dots & K(\widetilde{\chi}; p, \mu^{n_\infty}) {\bold u}_{n_\infty} \end{bmatrix}, \\ K^{{\bold x},{\boldsymbol \lambda}}(q) = \begin{bmatrix} {\bold x}^T_1 K(\widetilde{\chi}; \lambda^1, q) \\ \vdots \\ {\bold x}^T_{n_0} K(\widetilde{\chi}; \lambda^{n_0}, q)\end{bmatrix}. \end{gather} Let $q$ be a point of $X$ disjoint from all the interpolation nodes $\lambda^1, \dots, \lambda^{n_0}$, $\mu^1, \dots, \mu^{n_\infty}$ and let $Q$ be an invertible linear map of the fiber space $\chi(q)$ to the fiber space $\widetilde{\chi}(q)$. Then the abstract interpolation problem (ABSINT) has a solution $T$ with value $Q$ at the point $q$ if and only if the matrix $\Gamma$ is square and invertible and \begin{equation} \label{residues} [K(\widetilde{\chi}; p^i,q) + K_{{\boldsymbol\mu}, {\bold u}}(p^i) \Gamma^{-1} K^{{\bold x},{\boldsymbol \lambda}}(q)] Q (\text{Res}_{p^i}\ K(\chi; \cdot , q)^{-1}) = 0 \end{equation} at each pole $p^i$ of $K(\chi; \cdot, q)^{-1}$. In this case the unique solution $T$ of the interpolation problem (ABSINT) with value $Q$ at $q$ is given by \begin{equation} \label{solution} T(p) = [K(\widetilde{\chi}; p,q) + K_{{\boldsymbol \mu}, {\bold u}}(p) \Gamma^{-1} K^{{\bold x},{\boldsymbol \lambda}}(q)] Q K(\chi; p,q)^{-1} \end{equation} with inverse given by \begin{equation} \label{inversesolution} T^{-1}(p) = K(\chi;p,q)^{-1} T^{-1}(q)[K(\widetilde{\chi};q,p) + K_{{\boldsymbol \mu},{\bold u}}(q) \Gamma^{-1} K^{ {\bold x}, {\boldsymbol \lambda} }(p)]. \end{equation} \end{theorem} Two special cases of formula \eqref{solution} deserve to be mentioned. The first is the case where $n_0=n_\infty = 1$, $t_1=s_1=1$ and $\lambda^1 \ne \mu^1$. If we set $x =x_1$, $\lambda = \lambda^1$, $\mu = \mu^1$ and $u = u_1$, then $\Gamma = -x K(\widetilde{\chi}; \lambda, \mu) u$ is just a number and the formula \eqref{solution} becomes \begin{equation} \label{3.12a} T(p) K(\chi; p,q) T(q)^{-1} = \dfrac{ K(\widetilde{\chi}; p,q) - K(\widetilde{\chi}; p,\mu) u x^T K(\widetilde{\chi}; \lambda,q)}{x^T K(\widetilde{\chi}; \lambda, \mu) u}. \end{equation} In the line bundle case, the identity \eqref{3.12a} reduces to the Fay trisecant identity and will be discussed in Section \ref{S:linebundle}. The second special case of interest is the case where the given zero and pole vectors at each interpolation node span the whole fiber space. In this case there is an explicit multiplicative formula for the interpolant $T$ in terms of the prime form $E(p,q)$; this will be discussed in detail at the end of Section \ref{S:linebundle}. A second version of the abstract interpolation problem (ABSINT) has the same form (i), (ii) and (ii) as (ABSINT), but with the input bundle $\chi$ left also as an unknown to be found, subject to the proviso that it also be flat and have $h^0(\chi \otimes \Delta)=0$. This version of the problem was studied in \cite{hip} in a more concrete setting where $X$ is the normalizing Riemann surface for an algebraic curve $C$ embedded in ${\bold P}^2$ having a maximal rank $r$ determinantal representation; we will discuss the connections of this setup with ours in Sections 5 and 6. At this time we also state the solution to the modified (ABSINT). \begin{theorem} \label{T:bvabsint} Let $\boldsymbol \omega$ be a data set for (ABSINT) as above and form the matrix $\Gamma$ as in \eqref{defGamma}. Then there exists a flat bundle $\chi$ with $h^0(\chi \otimes \Delta) = 0$ and a meromorphic bundle map $T \colon \chi \to \widetilde{\chi}$ satisfying the interpolation conditions (i), (ii) and (iii) of (ABSINT) if and only if $\Gamma$ is square and invertible. \end{theorem} To prove Theorem \ref{T:absint} we need some preliminary lemmas. \begin{lemma} \label{L1:absint} A global meromorphic section of $\widetilde{\chi}\otimes \Delta$ is in ${\cal O}(\widetilde{\chi} \otimes \Delta) ({\boldsymbol \mu},{\bold u})(X)$ if and only if $h$ has the form \[ h= \sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{s_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} c_{j \beta} \] for some scalars $c_{j \beta}$. \end{lemma} \begin{pf} Suppose $h \in {\cal O}(\widetilde{\chi} \otimes \Delta) ( {\boldsymbol \mu}, {\bold u})(X)$. Choose scalars $c_{j \beta}$ so that \[ \text{Res}_{p = \mu^j} h(p) = \sum_{\beta=1}^{s_j} u_{j \beta} c_{j \beta} \] and set \[ \widehat{h}(p) = \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} c_{j \beta}. \] Then $\widehat{h} \in {\cal M}(\widetilde{\chi} \otimes \Delta)(X)$ and $h - \widehat{h} \in {\cal O}(\widetilde{\chi} \otimes \Delta)(X)$. Thus $h = \widehat{h}$ since ${\cal O}(\widetilde{\chi} \otimes \Delta)(X) = H^0(X, \widetilde{\chi} \otimes \Delta) = 0$ by our standing assumptions on $\widetilde{\chi}$. The converse direction follows easily from the defining properties of the Cauchy kernel $K(\widetilde{\chi}; \cdot, \cdot)$. \end{pf} \begin{lemma} \label{L2:absint} The map \[ [[c_{j \beta}]_{1 \le \beta \le s_j}]_{1 \le j \le n_\infty} \to h(p)=\sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{t_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} c_{j \beta} \] establishes a one-to-one correspondence between $\ker \Gamma$ and ${\cal O}(\widetilde{\chi} \otimes \Delta)( {\boldsymbol \omega})(X)$. In particular \[ \dim \ker \Gamma = h^0( (\widetilde{\chi} \otimes \Delta) ({\boldsymbol \omega}) ). \] \end{lemma} \begin{pf} Suppose first that $h \in {\cal O}(\widetilde{\chi} \otimes \Delta)( {\boldsymbol \omega})(X)$. In particular $h \in {\cal O}(\widetilde{\chi} \otimes \Delta) ({\boldsymbol \mu}, {\bold u})(X)$, so by Lemma \ref{L1:absint} there exists a collection of complex numbers $\{c_{j \beta}\}_{1\le j \le n_\infty, 1 \le \beta \le s_j}$ so that \[ h(p) = \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{t_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} c_{j \beta}. \] We next see what conditions the other requirements on $h$ for admission to the class ${\cal O}(\widetilde{\chi} \otimes \Delta) ({\boldsymbol \omega})(X)$ impose on the scalars $c_{j \beta}$. If $i$ is any index for which $\lambda^i \ne \mu^j$ for all $j$, we must have \begin{align} \notag 0 = x_{i \alpha}h(\lambda^i) & = \sum_{j=1}^{n_\infty} \sum_{\beta = 1} ^{t_j} x_{i \alpha} K(\widetilde{\chi}; \lambda^i, \mu^j) u_{j \beta} c_{j \beta} \\ \label{kerGamma1} &= - \sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{t_j} \Gamma_{ij, \alpha, \beta} c_{j \beta}. \end{align} If, on the other hand, $i$ is an index such that $\lambda^i = \mu^j =: \xi^{ij}$ for some index $j$, then we write the Laurent series expansion for $\frac{h}{\sqrt{dt^{ij}}}$ near $\xi^{ij}$ (where $t^{ij}(p)$ is a local coordinate for $X$ centered at $\xi^{ij}$) \[ \frac{h}{ \sqrt{dt^{ij}} }= \left[ \frac{h}{\sqrt{dt^{ij}}}\right]_{-1} \frac{1}{t^{ij}} + \left[\frac{h}{\sqrt{dt^{ij}}}\right]_0 + O(|t^{ij}(p)|). \] We compute \begin{equation} \label{res} \left[\frac{h}{\sqrt{dt^{ij}}}\right]_{-1} = \text{Res}_{p=\xi^{ij}}\ \frac{h}{\sqrt{t^{ij}}}(p) =\sum_{\beta=1}^{s_j} u_{j \beta} \sqrt{dt^{ij}}(\xi^{ij}) c_{j \beta} \end{equation} and hence \begin{equation} \label{rescoeff} \text{Res}_{p=\xi^{ij}} \frac{h}{\sqrt{dt^{ij}}}(p) = \sum_{\beta = 1}^{s_j} u_{j \beta} {\bold c}_{j \beta} \end{equation} where ${\bold c}_{j \beta} = \sqrt{dt^{ij}}(\xi^{ij}) c_{j \beta}$. Moreover,, we have \begin{align} \notag \left[ \frac{h}{\sqrt{dt^{ij}}}\right]_0 = & \sum_{k \ne j} \sum_{\beta = 1} ^{s_k} \frac{K(\widetilde{\chi}; \xi^{ij}, \mu^k)}{\sqrt{dt^{ij}}(\xi^{ij})} u_{k \beta}c_{k \beta} \\ \label{0coeff} &+ \sum_{\beta=1}^{s_j} \frac{A_{\ell}(\xi^{ij})}{dt(\xi^{ij})} \sqrt{dt^{ij}}(\xi^{ij}) u_{j \beta} c_{j \beta}. \end{align} We next compute \begin{gather} \notag x_{i \alpha}^T ( A(\xi^{ij})\left[\frac{h}{\sqrt{dt^{ij}}}\right]_{-1} + \left[ \frac{h}{ \sqrt{dt^{ij}} } \right]_0 dt^{ij}(\xi^{ij}) ) = \sum_{\beta=1}^{s_j} x^T_{i \alpha} A(\xi^{ij}) u_{j \beta} \sqrt{dt^{ij}}(\xi^{ij}) c_{j \beta} \\ \notag +\sum_{k \ne j} \sum_{\beta=1}^{s_j} x^T_{i \alpha} \frac{ K(\widetilde{\chi}; \xi^{ij}, \mu^k) }{ \sqrt{dt^{ij}}(\xi^{ij}) } u_{j \beta} c_{j \beta} dt^{ij}(\xi^{ij}) + \sum_{\beta=1}^{s_j} x_{i \alpha} \frac{ A_\ell(\xi^{ij}) }{ \sqrt{dt^{ij}}(\xi^{ij}) } u_{j \beta} c_{j \beta} dt^{ij}(\xi^{ij}) \\ \notag = \left\{ \sum_{\beta=1}^{s_j} x_{i \alpha}^T (A(\xi^{ij}) + A_\ell(\xi^{ij})) u_{j \beta} c_{j \beta} + \sum_{k \ne j} \sum_{\beta=1}^{s_k} x_{i \alpha}^T K(\widetilde{\chi}; \xi^{ij}, \mu^k) u_{j \beta} c_{j \beta} \right\} \sqrt{dt^{ij}}(\xi^{ij}). \end{gather} By Lemma \ref{L:cauchyexp} the first term in the braces vanishes. Combining this fact with the formula \eqref{rescoeff} for the coefficients ${\bold c}_{j \beta}$, we see that the interpolation condition \[ x_{i \alpha}^T \nabla_{\widetilde{\chi}}\left( t^{ij} \frac{h}{\sqrt{dt^{ij}}}\right) = - \sum_{\beta=1}^{s_j} \rho_{ij, \alpha \beta} {\bold c}_{j \beta} \] becomes \[ \left\{ \sum_{k \ne j} \sum_{\beta = 1}^{s_k} x_{i \alpha}^T K(\widetilde{\chi}; \xi^{ij}, \mu^k) u_{j \beta} c_{j \beta} \right\} \sqrt{dt^{ij}}(\xi^{ij}) = - \left( \sum_{\beta = 1} ^{s_j} \rho_{ij, \alpha \beta} c_{j \beta} \right) \sqrt{dt^{ij}}(\xi^{ij}). \] After canceling off $\sqrt{dt^{ij}}(\xi^{ij})$ and recalling that $\xi^{ij} = \lambda^i$ we see that this can be rewritten as \begin{equation} \label{kerGamma2} \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_j} \Gamma_{ij, \alpha, \beta} c_{j \beta} = 0 \end{equation} and this equation holds for all pairs of indices $(i, \alpha)$ such that $\lambda^i = \mu^j$ for some $j$. Combining \eqref{kerGamma2} and \eqref{kerGamma1} we see that the column vector $[[c_{j \beta}]_{1 \le \beta \le s_j}]_{1\le j \le n_\infty}]$ is in $\ker \Gamma$ as claimed. Conversely, if $[[c_{j \beta}]_{1 \le \beta \le s_j}]_{1 \le j \le n_\infty}$ is in $\ker \Gamma$ and we set \[ h(p) = \sum_{j=1}^{n_\infty} \sum_{\beta =1}^{s_j} K(\widetilde{\chi}; p, \mu^j) c_{j \beta}, \] then one can verify that $h \in {\cal O}(\widetilde{\chi} \otimes \Delta) ({\boldsymbol \omega})(X)$ by reversing the steps of the above argument. \end{pf} The analogue of Lemma \ref{L1:absint} at the level of bundle endomorphisms is the following. \begin{lemma} \label{L1':absint} Suppose that $T$ is a holomorphic bundle map from $\chi \otimes \Delta$ to $ (\widetilde{\chi} \otimes \Delta) ( {\boldsymbol \mu}, {\bold u})$ such that $T(q) = Q \colon \chi(q) \to \widetilde{\chi}(q)$. Then there exists a unique choice of operators $\widehat{x}_{j \beta} \colon \widetilde{\chi}(q) \to {\bold C}$ such that \[ T(p) = \left[ K(\widetilde{\chi}; p, q) + \sum_{j=1}^{n_\infty} \sum_{\beta = 1} ^{s_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} \widehat{x}_{j \beta} \right] Q K(\chi;p,q)^{-1}. \] \end{lemma} \begin{pf} Choose operators $\widehat{x}_{j \beta} \colon \widetilde{\chi}(q) \to {\bold C}$ so that \[ \text{Res}_{\mu^j} T(\cdot) K(\chi; \mu^j,q)Q^{-1} = \sum_{ \beta}^{s_j} u_{j \beta} \widehat{x}_{j \beta}. \] Set \[ \widehat{T}(p) = \left[ K(\widetilde{\chi}; p,q) + \sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{s_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} \widehat{x}_{j \beta} \right] Q(K(\chi, p,q)^{-1}. \] Then, for any vector $v \in \chi(q)$ we have \begin{gather} \notag T(\cdot) K(\chi; \cdot, q)v - \widehat{T}(\cdot) K(\chi; \cdot, q) v \\ \notag =T(\cdot) K(\chi; \cdot, q) v - \left[ K(\widetilde{\chi}; \cdot, q)Q +\sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_j} K(\widetilde{\chi}; \cdot, \mu^j) u_{j \beta} \widehat{x}_{j \beta}Q \right] v \\ \end{gather} is an element of ${\cal O}(\widetilde{\chi} \otimes \Delta)(X)$. By our standing assumption that $h^0(\widetilde{\chi} \otimes \Delta) = 0$, we conclude that $(T-\widehat{T})(\cdot) K(\chi; \cdot, q)v=0$ for all $v \in \chi(q)$. This is enough to force $T=\widehat{T}$, and the lemma follows. \end{pf} \begin{pf*}{Proof of Theorem \ref{T:absint}} We first argue that necessarily $\Gamma$ is square and invertible if a solution $T$ to the interpolation problem (ABSINT) exists. To do this, we show first that $\ker \Gamma = \{0\}$ and secondly, that $\Gamma$ is square. To see that $\ker \Gamma =\{0\}$, we proceed as follows. If $T$ is a solution of (ABSINT), then multiplication by $T$ induces a biholomorphic bundle map between $\chi \otimes \Delta$ and $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. By assumption, $h^0(\chi \otimes \Delta) = 0$. Hence we also have $h^0( (\widetilde{\chi} \otimes \Delta) ({\boldsymbol \omega})) = 0$. Now it follows from Lemma \ref{L2:absint} that $\ker \Gamma = \{0\}$. Next we argue that $\Gamma$ is square. Since $\chi$ and $\widetilde{\chi}$ by assumption are both flat, both $\chi$ and $\widetilde{\chi}$ have degree 0, as do $\det \chi$ and $\det \widetilde{\chi}$. Then \[ \text{deg}(\det T) = \text{deg}(\det \widetilde{\chi}) - \text{ deg} (\det \chi) = 0. \] If $T$ is a solution of (ABSINT), then the total number of zeros $n_0(T)$ of $T$ (counted with multiplicities as appropriate for meromorphic matrix functions---see Chapter 3 of \cite{bgr}) is equal to $\sum_{i=1}^{n_0} t_i =:N_0$ which is the number of rows of $\Gamma$, while the total number of poles $n_\infty(T)$ (again counted with multiplicities) is equal to $ \sum_{j=1}^{n_\infty} s_j =:N_\infty$ which is equal to the number of columns of $\Gamma$. In general we have $\text{deg}(\det T) = n_0(T) - n_\infty(T)$. Hence the equality $\text{deg}(\det T) = 0$ for $T$ a solution of (ABSINT) implies that $\Gamma$ is square. Combining this with the result of the previous paragraph, we see that $\Gamma$ is invertible as well. If $T$ is a solution of (ABSINT), then in particular $T$ satisfies the hypotheses of Lemma \ref{L1':absint} and hence $T$ has the form \begin{equation} \label{ansatz} T(p) = \left[ K(\widetilde{\chi};p,q) + \sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{s_j} K(\widetilde{\chi}; p, \mu^j) u_{j \beta} \widehat{x}_{j \beta} \right] Q K(\chi, p,q)^{-1} \end{equation} for appropriate operators $\widehat{x}_{j \beta} \colon \widetilde{\chi}(q) \to {\bold C}$. We now find what additional restrictions on $\widehat{x}_{j \beta}$ are forced by the zero and coupled zero-pole interpolation conditions (ii) and (iii) in (ABSINT). Suppose that $i$ is an index for which $\lambda^i \ne \mu^j$ for any $j$. Then the zero interpolation condition $x_{i \alpha}^T T(\lambda^i) =0$ forces, for all $\alpha$ between $1$ and $t_i$, \[ x_{i \alpha}^T K(\widetilde{\chi}; \lambda^i,q) Q K(\chi; \lambda^i, q)^{-1} + \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_i}x_{i \alpha}^T K(\widetilde{\chi}; \lambda^i, \mu^j) u_{j \beta} \widehat{x}_{j \beta} Q K(\chi; \lambda^i,q)^{-1} = 0. \] Recalling the definition of $\Gamma_{ij, \alpha \beta}$, we can rewrite this as \begin{equation} \label{Gamma1} \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_j} \Gamma_{ij, \alpha \beta} \widehat{x}_{j \beta} = x_{i \alpha} K(\widetilde{\chi}, \lambda^i, q) \end{equation} for all index pairs $(i,\alpha)$ such that $\lambda^i \ne \mu^j$ for any $j$. We next consider an index $i$ for which $\lambda^i = \mu^j:= \xi^{ij}$ for some $j$. Let $x_{i \alpha}(p)$ be a local holomorphic section of $\widetilde{\chi}^\vee$ as in the third set of interpolation conditions. Let $\varphi(p)$ be the meromorphic local section of $\chi$ given by \[ \varphi(p) = \frac{K(\chi; p,q)}{\sqrt{dt^{ij}}(p)} e \] for a vector $e \in \chi(q)$ where $t^{ij}(p)$ is a local coordinate on $X$ centered at $\xi^{ij}$, and let $u(p)$ be the local meromorphic section of $\widetilde{\chi}$ given by $u(p) = T(p) \varphi(p)$. From \eqref{ansatz} we have then \[ u(p) = \frac{K(\widetilde{\chi};p,q)}{\sqrt{dt^{ij}}(p)} Q e + \sum_{j=1}^{n_\infty} \sum_{\beta = 1}^{s_j} \frac{K(\widetilde{\chi};p,\mu^j)}{\sqrt{dt^{ij}}(p)} u_{j \beta} \widehat{x}_{j \beta} Q e. \] Then the coefficients $[u]_{-1}$ and $[u]_0$ in the Laurent expansion of $u(p)$ centered at $\xi^{ij}$ with respect to local coordinate $t^{ij}$ are given by \begin{align} \label{-1Laurent} [u]_{-1} &= \sum_{\beta = 1}^{s_j} u_{j \beta} \sqrt{dt^{ij}}(\xi^{ij}) \widehat{x}_{j \beta} Q e, \\ \label{0Laurent} [u]_0 &= \frac{K(\widetilde{\chi}; \xi^{ij},q)}{\sqrt{dt^{ij}}(\xi^{ij})} Q e + \sum_{k \ne j} \sum_{\beta = 1}^{s_k} \frac{K(\widetilde{\chi}; \xi^{ij}, \mu^k)}{\sqrt{dt^{ij}}(\xi^{ij})} u_{k \beta} \widehat{x}_{k \beta} Q e \\ \notag & + \sum_{\beta=1}^{t_j} \frac{A_\ell(\xi^{ij})}{dt(\xi^{ij})} \sqrt{dt}(\xi^{ij}) u_{j \beta} \widehat{x}_{j \beta} Q e \end{align} so the alternate coupled interpolation condition (iii) given by \eqref{coupledint} implies that \[ x_{i \alpha}^T \left( A(\xi^{ij}) [u]_{-1} + [u]_0 dt^{ij}(\xi^{ij}) \right) = - \sum_{\beta = 1}^{s_j} \rho_{ij, \alpha \beta} \widehat{x}_{j \beta} Q e \cdot \sqrt{dt^{ij}}(\xi^{ij}). \] Substitution of the expressions \eqref{-1Laurent} and \eqref{0Laurent} for $[u]_{-1}$ and $[u]_0$ gives \begin{align} \notag \sum_{\beta=1}^{s_j} x_{i \alpha}^T A(\xi^{ij}) u_{j \beta} \widehat{x}_{j \beta} Q e \sqrt{dt^{ij}}(\xi^{ij}) & + x_{i \alpha}^T \frac{K(\widetilde{\chi}; \xi^{ij}, q)}{\sqrt{dt^{ij}}(\xi^{ij})} Q e dt^{ij}(\xi^{ij}) \\ \notag + \sum_{k \ne j} \sum_{\beta = 1}^{s_k} x_{i \alpha}^T \frac{K(\widetilde{\chi}; \xi^{ij}, \mu^k)}{\sqrt{dt^{ij}}(\xi^{ij})} u_{k \beta} \widehat{x}_{k \beta} Q e \cdot dt^{ij}(\xi^{ij}) & + \sum_{\beta = 1}^{t_j} x_{i \alpha}^T \frac{A_\ell(\xi^{ij})}{\sqrt{dt^{ij}}(\xi^{ij})} u_{j \beta} \widehat{x}_{j \beta} Q e \cdot dt^{ij}(\xi^{ij}) \\ \notag = &- \sum_{\beta = 1}^{s_j} \rho_{ij, \alpha \beta} \widehat{x}_{j \beta} Q e \cdot \sqrt{dt^{ij}}(\xi^{ij}). \end{align} By using the result of Lemma \ref{L:cauchyexp} and recalling the definition \eqref{defGamma} of $\Gamma_{ij, \alpha \beta}$ we see that this expression collapses to \[ \sum_{k=1}^{n_\infty} \sum_{\beta = 1}^{s_k} \Gamma_{ik, \alpha \beta} \widehat{x}_{k \beta} Q e = x_{i \alpha}^T K(\widetilde{\chi}; \xi^{ij},q) Qe. \] Since this must hold for all $e \in \chi(q)$, we arrive at the operator equation \begin{equation} \label{Gamma2} \sum_{k=1}^{n_\infty} \sum_{\beta=1}^{s_k} \Gamma_{ik, \alpha \beta} \widehat{x}_{k \beta} = x_{i \alpha}^T K(\widetilde{\chi}; \xi^{ij},q) \end{equation} which must hold for all index pairs $(i, \alpha)$ for which $\lambda^i = \mu^j$ for some $j$. Combining \eqref{Gamma1} and \eqref{Gamma2} gives us \[ \Gamma \widehat{{\bold x}} = K^{{\boldsymbol \lambda}, {\bold x}}(q) \] where we have set $\widehat{{\bold x}}$ equal to the column vector $[[ \widehat{x}_{j \beta}]_{ 1 \le \beta \le s_j}]_{1 \le j \le n_\infty}$. Plugging this value into \eqref{ansatz} leaves us with the formula \eqref{solution} for the solution $T$. This also establishes the uniqueness of the solution of (ABSINT) whenever it exists. Since $T$ is analytic at the points $p^1, \dots, p^n$ where $K(\chi; \cdot, q)^{-1}$ has poles, necessarily the residue conditions \eqref{residues} must hold as well. The necessity and uniqueness parts of the theorem are now established. Conversely, assume that $\Gamma$ is invertible and that the residue conditions \eqref{residues} hold. We define $T(p)$ by the formula \eqref{solution}. Then $T$ is a meromorphic bundle map of $\chi$ and $\widetilde{\chi}$ with only simple poles which occur at most at the points $\mu^1, \dots, \mu^{n_\infty}$ with \[ \text{im Res}_{\mu^i} T(\cdot) \subset \text{span }\{u_{j \alpha} \colon 1 \le \alpha \le s_j\} \] for $j=1, \dots, n_\infty$. Since $T$ has the form \eqref{ansatz} with operators $\widehat{x}_{j \beta}$ ($1 \le j \le n_\infty$, $1 \le \beta \le s_j$) satisfying \eqref{Gamma1} and \eqref{Gamma2}, we see that also $T$ satisfies the interpolation conditions (ii) and (iii) in (ABSINT) as well. Hence, the number of poles $n_\infty(T)$ of $T$ (counting multiplicities for a meromorphic matrix function as in Chapter 3 of \cite{bgr}) is at most $\sum_{j=1}^{n_\infty} s_j =: N_\infty$ and the number of zeros $n_0(T)$ of $T$ (counting multiplicities) is at least $\sum_{i=1}^{n_\infty} t_i =: N_0$. As $T$ is a bundle map of the flat bundles $\chi$ and $\widetilde{\chi}$, we know that $n_0(T) = n_\infty(T)$. On the other hand, since $\Gamma$ is square we have $N_0=N_\infty$. From the chain of inequalities \[ N_0 \le n_0(T) = n_\infty(T) \le N_\infty \] combined with the equality $N_0 = N_\infty$, we get that $n_0(T) = N_0$ and $n_\infty(T) = N_\infty$. This implies that necessarily \[ \text{im Res}_{p = \mu^j} T(p) = \text{span}\{u_{j1}, \dots, u_{j s_j} \} \] and \[ {\text{im Res}}_{p=\lambda^i}(T^{\vee})^{-1}(p) = \text{span}\{x_{i1}, \dots, x_{i t_i}\} \] and that $T(p)$ is analytic and invertible at every point $p \in X$ outside of $\mu^1, \dots, \mu^{n_\infty}$, $\lambda^1, \dots, \lambda^{n_0}$. This verifies that $T$ is a bona fide solution of the interpolation problem (ABSINT). It remains only to verify the formula \eqref{inversesolution} for the inverse of $T$. To see this, we note that $(T^{-1})^T$ is also the solution of an interpolation problem of the type (ABSINT), namely the one with data set $[{\boldsymbol \omega}]^\vee$ given by \begin{enumerate} \item $( {\boldsymbol \lambda},{\bold x})$ in place of $( {\boldsymbol \mu}, {\bold u})$, \item $({\boldsymbol \mu}, {\bold u})$ in place of $({\boldsymbol \lambda}, {\bold x})$, and \item $-\rho_{ji, \beta \alpha}$ in place of $\rho_{ij, \alpha \beta}$. \end{enumerate} The matrix $\Gamma$ associated with this interpolation problem turns out to be exactly $-\Gamma^T$ where $\Gamma$ is as in \eqref{defGamma}. Hence by Theorem \ref{T:absint} the bundle map $(T^{-1})^T$ must be given by \begin{equation} \label{Tinvtr} (T^{-1})^T = [K(\widetilde{\chi}^\vee;p,q) - K_{{\boldsymbol \lambda}, {\bold x}}(p) (\Gamma^{-1})^T K^{{\bold u}, {\boldsymbol \mu}}(q)] (T(q)^{-1})^T K(\chi^\vee; p,q)^{-1}. \end{equation} By the uniqueness property of Cauchy kernels, it is easy to see that \[ K(\chi^\vee; p,q)^T = - K(\chi; q,p). \] Hence, taking transpose on both sides of \eqref{Tinvtr} gives \begin{align} \notag T^{-1}(p) = & -K(\chi;q,p)^{-1} T(q)^{-1}[-K(\widetilde{\chi}; q,p) - K_{ {\boldsymbol \mu}, {\bold u}}(q) \Gamma^{-1} K^{{\bold x}, {\boldsymbol \lambda} }(p) ] \\ \notag = & K(\chi;q,p)^{-1} T(q)^{-1}[-K(\widetilde{\chi}; q,p) + K_{{\boldsymbol \mu}, {\bold u}}(q) \Gamma^{-1} K^{{\bold x},{\boldsymbol \lambda}}(p) ] \end{align} and the formula \eqref{inversesolution} follows. \end{pf*} {\bf Remark:} In case $\widetilde{\chi} = \chi =:\chi_0$ are both taken to be the trivial bundle of rank $r$, the Cauchy kernel $K(\chi_0; \cdot, \cdot)$ has the scalar form $k_0(\cdot, \cdot) I_r$ where $k_0(\cdot, \cdot)$ is the Cauchy kernel for the trivial line bundle over $X$. In this case the ansatz \eqref{ansatz} simplifies to \[ T(p) = Q + \sum_{j=1}^{n_\infty} \sum_{\beta=1}^{s_j} f_{\mu^j}(p) u_{j \beta} \widehat{x}^\prime_{j \beta} \] where we have set \[f_{\mu}(p) = \frac{k_0(p,\mu)}{k_0(p,q)} \] and the row vectors $\widehat{x}^\prime_{j \beta} = \widehat{x}_{j \beta} Q$ are now taken to be the unknowns. Note that $k_0(\cdot, q)$ is a half-order differential with divisor of degree $g-1$ and a pole at $q$; if we assume that the zeros are distinct, this divisor has the form $p^1 + \dots + p^g - q$ for distinct points $p^1, \dots, p^g,q \in X$. If the image of the divisor $p^1 + \dots + p^g$ under the Abel-Jacobi map is not on the classical theta divisor in the Jacobian (i.e. if $(p^1 + \dots + p^g$ is a {\it non-special} divisor), then there are no nonzero constant meromorphic functions with only poles equal to at most simple poles at the points $p^1, \dots, p^g$; this corresponds to our assumption that $h^0(\chi_0 \otimes \Delta)=h^0(\Delta)=0$. Furthermore, in this case, the global scalar meromorphic function $f_\mu(\cdot)$ on $X$ (for $\mu$ a point of $X$ disjoint from $p^1, \dots, p^g,q$) is uniquely determined (up to a nonzero scalar multiple) by the condition that it have a pole at $\mu$ and that its divisor $(f_\mu)$ satisfy \[ (f_\mu) \ge q - \mu - p^1 - \dots - p^g. \] In this way our results and analysis on the (ABSINT) problem reduce to the work in \cite{bc} for the trivial bundle case. Notice that $\chi_0$ can be replaced here by $\xi\otimes\chi_0$ for any line bundle $\xi$ of degree $0$ satisfying $h^0(\xi \otimes \Delta)=0$, replacing the Cauchy kernel $k_0(\cdot,\cdot)$ for the trivial line bundle by the Cauchy kernel for $\xi$; this corresponds to letting $p^1 + \dots + p^g$ be any non-special effective divisor of degree $g$. \vspace{.1in} One remaining piece of business in this section is the proof of Theorem \ref{T:bvabsint}. The problem of identifying the unknown input bundle in a more explicit fashion will be addressed in Section \ref{S:conint}. \begin{pf*}{Proof of Theorem \ref{T:bvabsint}} If there exists such an input bundle $\chi$ and meromorphic bundle map $T$, then $T$ implements a biholomorphic bundle map between $\chi \otimes \Delta$ and $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. Since $h^0(\chi \otimes \Delta) = 0$, it then follows from Lemma \ref{L2:absint} that $\Gamma$ is injective. Since deg$(\chi \otimes \Delta) = r(g-1)$, it must be the case that deg$(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})=r(g-1)$ as well. This means that $\Gamma$ is square, and hence invertible. Conversely, suppose that $\Gamma$ is square and invertible. Define a bundle $\chi$ so that $\chi \otimes \Delta \cong (\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. Since $\Gamma$ is square, it follows that \[ \text{deg}\left( (\widetilde{\chi}\otimes \Delta)({\boldsymbol \omega})\right) = \text{deg}(\widetilde{\chi} \otimes \Delta) = r(g-1), \] and hence $\text{deg}(\chi \otimes \Delta) = r(g-1)$. Since $\ker \Gamma = \{0\}$, we know by Lemma \ref{L2:absint} that $h^0( (\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega}))=0$; thus $h^0(\chi \otimes \Delta) = 0$. It follows from these two facts as in the proof of Theorem 3.1 in \cite{hip} that $\chi$ is flat. Let now $S \colon \chi \otimes \Delta \to (\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$ be an implementation of the holomorphic bundle isomorphism between $\chi \otimes \Delta$ and $(\widetilde{\chi} \otimes \Delta)({\boldsymbol \omega})$. Define $T \colon \chi \to \widetilde{\chi}$ so that $S = T \otimes I_{{\cal O}(\Delta)}$. Then $T$ is a meromorphic bundle map from $\chi$ to $\widetilde{\chi}$ which solves the interpolation problem (ABSINT). \end{pf*} \section{The line bundle case and Fay's identity} \label{S:linebundle} In this section we specialize the work of the preceding sections to the line bundle case. We shall need here some basic facts concerning the Jacobian variety, the Abel-Jacobi map and associated theta functions (theta function, theta functions with characteristics and prime form) for the Riemann surface $X$. The review here is quite sketchy; for complete details the reader should consult \cite{oldfay}, \cite{farkaskra} or \cite{mumford}. When the rank $r$ of the vector bundle $\chi$ is 1, one can get an explicit formula for $K(\chi; \cdot, \cdot)$ in terms of the Abel-Jacobi map for the surface $X$ and various variants of the classical theta function associated with the Jacobian variety of $X$ (see \cite{hip}). Specifically, in this case we may assume that $\chi$ is a flat unitary line bundle with factor of automorphy (also called $\chi$) given by $\chi(A_j) = \exp (- 2 \pi i a_j)$, $\chi(B_j) = \exp (2 \pi i b_j)$, $j=1, \dots, g$, where $A_1, \dots, A_g, B_1, \dots, B_g$ form a canonical integral homology basis on $X$. Let $\Omega$ be the corresponding period matrix, let $J(X)= {\bold C}^g / {\bold Z}^g + \Omega {\bold Z}^g$ be the Jacobian variety of $X$ and let $\phi \colon X \to J(X)$ be the Abel-Jacobi map. As is standard, we extend $\phi$ by linearity to any divisor on $X$, and, using the correspondence between linear equivalence classes of divisors and isomorphism classes of line bundles, we consider $\phi$ to be defined on any line bundle on $X$ as well. One can verify that then $\phi(\chi) = z$ where $z = \Omega a + b$ and $a,b \in {\bold R}^g$ have respective coordinates $a_j,b_j$. Then the explicit formula for the Cauchy kernel (as given in \cite{hip}) is the following. The verification is straightforward, once one has in hand the properties and factors of automorphy for the various objects involved. \begin{theorem} \label{T:scalarCauchykernel} For the case where $\chi$ is a flat unitary line bundle as above, the Cauchy kernel as defined in Section \ref{S:cauchyker} is given explicitly by \begin{equation} K(\chi;p,q) = \frac{ \theta \begin{bmatrix} a \\ b \end{bmatrix} (\phi(q) - \phi(p)) } {\theta \begin{bmatrix} a \\ b \end{bmatrix} (0) E(q,p)}. \label{lineCauchyker} \end{equation} \end{theorem} In the statement of Theorem \ref{T:scalarCauchykernel} $\theta \begin{bmatrix} a \\ b \end{bmatrix} ( \cdot)$ is the associated theta function with characteristics $\begin{bmatrix} a \\ b \end{bmatrix}$, $E(\cdot, \cdot)$ is the prime form on $X \times X$, and we assume the line bundle $\Delta$ of differentials of order $\frac{1}{2}$ has been chosen so that $\phi(\Delta) = - \kappa$, where $\kappa \in J(X)$ is Riemann's constant (see \cite{oldfay} and \cite{mumford}). Note that a consequence of Riemann's theorem is that $\theta(z) \ne 0$ if and only if $h^0(\chi \otimes \Delta) = 0$, and hence $\theta \begin{bmatrix} a \\ b \end{bmatrix} (0) \ne 0$ in \eqref{lineCauchyker} and the formula makes sense. No such explicit formula is known at present for the higher rank case except in genus 1 (see \cite{bcv}). In the line bundle case one can also give an explicit formula for the canonical connections $\nabla_\chi$, $\nabla_\chi^*$ associated with the flat unitary line bundle $\chi$. This is the content of the following Proposition. \begin{proposition} \label{P:connection} For the case where $\chi$ is a flat unitary line bundle with normalizations as above, then the canonical connections $\nabla_\chi$ and $\nabla_\chi^*$ are given by \begin{align} \nabla_\chi y = & \left[ \sum_{j=1}^g \frac{\partial}{\partial z_j} \log \theta \begin{bmatrix} a \\ b \end{bmatrix} (0) \ \omega_j(p) \right] y + dy \\ = & \left[ \sum_{j=1}^g [2 \pi i a_j + \frac{\partial}{\partial z_j} \log \theta(z) ] \omega_j(p) \right] y + dy,\\ \nabla_\chi x = & - \left[ \sum_{j=1}^g \frac{\partial}{\partial z_j} \log \theta \begin{bmatrix} a \\ b \end{bmatrix} (0) \ \omega_j(p) \right] x + dx \\ = & - \left[ \sum_{j=1}^g [2 \pi i a_j + \frac{\partial}{\partial z_j} \log \theta(z) ] \omega_j(p) \right] x + dx. \end{align} \end{proposition} \begin{pf} This follows directly by comparing the general expansion for the Cauchy kernel $$ \frac{K(\chi; p,p_0)}{\sqrt{dt}(p) \sqrt{dt}(p_0)} = \frac{1}{t(p) - t(p_0)} \left[ I_r + \frac{A_\ell}{dt}(p_0) t(p) + O(|t(p)|^2) \right] $$ on the one hand and substituting the expansion of the theta function $$ \theta \begin{bmatrix} a \\ b \end{bmatrix} \left(\phi(p_0) - \phi(p)\right) = \theta \begin{bmatrix} a \\ b \end{bmatrix} (0) - \sum_{j=1}^g \frac{\partial \theta \begin{bmatrix} a \\ b \end{bmatrix}} {\partial z_j} ( 0 ) \frac{\omega_j}{dt} \left( p_0 \right) t(p) + O(|t(p)|^2) $$ and the expansion of the prime form (see Corollary 2.5 in \cite{oldfay}) $$ E(p_0,p) = t(p) + O(|t(p)|^3) $$ into \eqref{lineCauchyker}. \end{pf} {\bf Remark.} From the formula for $\nabla_\chi$ and $\nabla_\chi^*$ it follows that the coefficients $A(p)$ and $A_\ell(p)$ are independent of the choice of homology bases (i.e., marking) on the Riemann surface $X$ as long as the bundle $\Delta$ of half-order differentials defined by $\phi(\Delta) = \kappa$ remains the same, since the unitary flat representative for a flat line bundle is unique. It is an amusing exercise to verify this independence directly by using the transformation law for theta functions (see \cite{mumford} and \cite{Igusa}). We next specialize the work of Section \ref{S:absint} to the scalar (or line bundle) case, where $\chi$ and $\widetilde{\chi}$ are flat unitary line bundles. As explained in Section \ref{S:absint}, necessarily the multiplicities $s_j$ and $t_i$ are all 1 and without loss of generality we may take $u_{j1}=1$, $x_{i1}=1$ for all $i$ and $j$. Then the compatibility condition \eqref{comp} forces the third interpolation condition to be absent. The data of the problem consists simply of the set of $n_\infty + n_0$ distinct points $\mu^1, \dots, \mu^{n_\infty}, \lambda^1, \dots, \lambda^{n_0}$ together with the flat unitary line bundles $\chi$ and $\widetilde{\chi}$. The problem then is to produce a bundle map $T \colon \chi \to \widetilde{\chi}$ with divisor equal to ${\boldsymbol{\lambda}} - {\boldsymbol{\mu}}$ (where we have set ${\boldsymbol{\lambda}} = \lambda^1 + \dots + \lambda^{n_0}$ and ${\boldsymbol \mu} = \mu^1 + \dots + \mu^{n_\infty}$). If we view the bundles in terms of factors of automorphy, we can view $T$ simply as a multivalued function on $X$ having divisor equal to ${\boldsymbol \lambda} - {\boldsymbol \mu}$ and factor of automorphy $\chi_T$ given by \[ \chi_T(A_j) = e^{-2 \pi i a_j}, \ \chi_T(B_j) = e^{2 \pi i b_j} \text{ for } j=1, \dots, g \] where $\phi(\widetilde{\chi}) - \phi(\chi) = \Omega a + b$ (where we have set $a= \begin{bmatrix} a_1 & \dots & a_g \end{bmatrix}^T$ and $b= \begin{bmatrix} b_1 & \dots & b_g \end{bmatrix}^T$). In the genus zero case where $X= \bold{C} \cup \{\infty\}$ is the Riemann sphere, any flat unitary line bundle is trivial and the problem is to produce a global meromorphic function with divisor equal to ${\boldsymbol \lambda} - {\boldsymbol \mu}$. Trivially a solution exists if and only if $n_0=n_\infty$ and then the unique solution with value 1 at infinity is given in the multiplicative form \begin{equation} T(z) = \dfrac{\prod_{i=1}^{n_0} (z-\lambda^i)}{\prod_{j=1}^{n_\infty} (z-\mu^j)}. \label{g0prod} \end{equation} or in the partial fraction form \begin{equation} T(z) = 1 + \sum_{j=1}^{n_\infty} c_j (z-\mu^j)^{-1} \label{g0partialfrac} \end{equation} where $c^T= \begin{bmatrix} c_1 & \dots & c_{n_\infty} \end{bmatrix}$ is the unique solution of the linear system of equations $S c = \begin{bmatrix} 1 & \dots & 1 \end{bmatrix}^T$ with $S$ equal to the Sylvester matrix \[ S = [S_{ij}] \text{ with } S_{ij} = \frac{1}{\mu^j - \lambda^i}, \] or, in other words, \begin{equation} c = S^{-1} \begin{bmatrix} 1 & \dots & 1 \end{bmatrix}^T. \label{g0partialfraccoef} \end{equation} It is possible to evaluate the vector $c$ explicitly from \eqref{g0partialfraccoef} once one knows the entries of $S^{-1}$ explicitly. This in turn can be done once one knows an explicit formula for the determinant of a Sylvester matrix $S$, since the cofactor matrices are again of the same form. In this way one can verify directly the equivalence of the two formulas \eqref{g0prod} and \eqref{g0partialfrac}. For details on the algebra of this computation, we refer to Theorem 4.3.2 of \cite{bgr}. We shall see that an analogous pair of formulas holds for the solution of the abstract interpolation problem (ABSINT) for the higher genus case (for the line bundle setting). This is the content of the next Theorem. \begin{theorem} \label{T:lineabsint} Consider the problem (ABSINT) for the case where $\chi$ and $\widetilde{\chi}$ are flat unitary line bundles and given data set equal to $\boldsymbol{\lambda} - \boldsymbol{\mu} = \lambda^1 + \dots + \lambda^{n_0} - \mu^1 - \dots - \mu^{n_\infty}$ as above. Then a solution exists if and only if \begin{equation} n_0 = n_\infty \text{ and } \phi(\widetilde{\chi}) - \phi(\chi) = \phi({\boldsymbol \lambda}) - \phi(\boldsymbol \mu). \label{necessity} \end{equation} In this case, if $q$ is a point of $X$ disjoint from the set $\{\lambda^1, \dots, \lambda^{n_0}, \mu^1, \dots, \mu^{n_\infty}\}$ of prescribed zeros and poles and $Q$ is any invertible fiber map from $\chi(q)$ onto $\widetilde{\chi}(q)$, then a solution of (ABSINT) having value $Q$ at $q$ is given in multiplicative form as \begin{equation} T(p) = \dfrac{ \prod_{i=1}^{n_0} E(p,\lambda^i)/E(q,\lambda^i)}{\prod_{j=1}^{n_\infty} E(p,\mu^j)/E(q,\mu^j)} \exp(-2 \pi i a^T (\phi(p) - \phi(q)) Q \label{prod} \end{equation} where $a^T = \begin{bmatrix} a_1 & \dots & a_g \end{bmatrix}$ and $\phi({\boldsymbol{\lambda}}) - \phi({\boldsymbol{\mu}}) = \Omega a + b$ with $a,b \in {\bold R}^g$. If $\chi \otimes \Delta$ and $\widetilde{\chi} \otimes \Delta$ have no nontrivial holomorphic sections, then the solution $T$ with $T(q)=Q$ is unique and alternatively is given by the partial fraction formula \begin{align} T(p)=& \left\{ \dfrac{\theta[\widetilde{z}](\phi(q) - \phi(p))}{\theta[\widetilde{z}](0) E(q,p)} + \sum_{j=1}^{n_\infty} \sum_{k=1}^{n_0} \dfrac{\theta[\widetilde{z}](\phi(\mu^j) - \phi(p))}{\theta[\widetilde{z}](0)E(\mu^j,p)} \cdot [\Gamma^{-1}]_{jk} \cdot \dfrac{\theta[\widetilde{z}](\phi(q)- \phi(\lambda^k))}{\theta[\widetilde{z}](0) E(q,\lambda^k) } \right\} \notag \\ & \times Q \cdot \dfrac{\theta[z](0) E(q,p)}{\theta[z](\phi(q) - \phi(p))} \label{partialfrac} \end{align} where the $n_0 \times n_\infty$ matrix $\Gamma$ is given by \[ \Gamma = [\Gamma_{ij}] \text{ with } \Gamma_{ij} =- \dfrac{\theta[\widetilde{z}](\phi(\mu^j) - \phi(\lambda^i))}{\theta[\widetilde{z}](0) E(\mu^j, \lambda^i)}. \] and where we have set \[ z = \phi(\chi), \ \widetilde{z} = \phi(\widetilde{\chi}). \] Here we write $\theta[z](\lambda)$ rather than $\theta\begin{bmatrix} a \\ b \end{bmatrix}(\lambda)$ if $z = \Omega a + b \in \bold{C}^g$ with $a, b \in {\bold R}^g$. \end{theorem} \begin{pf} If such a bundle map exists, then the bundle $\chi \otimes {\cal O}(\boldsymbol{\lambda} - \boldsymbol{\mu})$ and $\widetilde{\chi}$ are holomorphically equivalent. In particular, the divisor $\boldsymbol{\lambda} - \boldsymbol{\mu}$ must have degree 0 since both $\chi$ and $\widetilde{\chi}$ are flat bundles. The equality $\phi(\widetilde{\chi}) = \phi(\chi) + \phi(\boldsymbol{\lambda}) - \phi(\boldsymbol{\mu})$ then follows from the correspondence between flat bundles and linear equivalence classes of divisors mentioned above. This verifies the necessity condition \eqref{necessity}. Conversely, assume that \eqref{necessity} holds and define $T$ by the right-hand side of \eqref{prod}. That the zero-pole divisor of $T$ is $\boldsymbol{\lambda} - \boldsymbol{\mu}$ follows directly from the fact that the divisor of the prime form $(p,q) \to E(p,q)$ is the diagonal $\{(p,p) \colon p \in X\} \subset X \times X$. One can next check from the known period relations of $E$ that the right-hand side of \eqref{prod} has the factor of automorphy $\chi_T$ \[ \chi_T(A_j) = \exp (-2 \pi i a_j), \ \chi_T(B_j) = \exp (2 \pi i b_j) \text{ for } j=1, \dots, g \] where $a_1, \dots, a_g, b_1, \dots, b_g$ are respective components of $a,b \in {\bold R}^g$ chosen so that $\Omega a + b = \phi(\boldsymbol{\lambda}) - \phi(\boldsymbol{\mu})$. The second condition in \eqref{necessity} now guarantees that $T(\cdot)$ so defined is a bundle map from $\chi$ into $\widetilde{\chi}$. The uniqueness assertion is a consequence of Lemma \ref{L1':absint}. The alternative formula \eqref{partialfrac} is simply a rewriting of the formula \eqref{solution} from Theorem \ref{T:absint} specialized to the line bundle case, where we have substituted the explicit formula \eqref{lineCauchyker} for the Cauchy kernel from Theorem \ref{T:scalarCauchykernel}. \end{pf} Note that part of the content of Theorem \ref{T:lineabsint} is that the matrix $\Gamma$ is invertible whenever $\chi$ and $\widetilde{\chi}$ have no nontrivial holomorphic sections and (ABSINT) has a solution. It is of interest to specialize Theorem \ref{T:lineabsint} to the case of one prescribed zero and pole $\lambda - \mu = \lambda^1 - \mu^1$. When this is done we obtain the following result. \begin{theorem} \label{T:redlineabsint} If $\chi$ and $\widetilde{\chi}$ are two flat unitary line bundles such that neither $\chi \otimes \Delta$ nor $\widetilde{\chi} \otimes \Delta$ have nontrivial holomorphic sections and $\lambda$ and $\mu$ are two distinct points of $X$, then the unique meromorphic bundle map from $\chi$ to $\widetilde{\chi}$ with zero-pole divisor equal to $\lambda - \mu$ and value $Q \neq 0$ at the point $q \in X$ is given by either \begin{equation} \label{specialprod} T(p) = \dfrac{E(p,\lambda)}{E(p, \mu)} \dfrac{E(q,\mu)}{E(q,\lambda)} \exp(-2 \pi i a^T(\phi(p) - \phi(q)) Q \end{equation} or \begin{align} T(p) =& \exp(-2 \pi i a^T(\phi(p) - \phi(q))) \left\{ \dfrac{\theta(z + \phi(\lambda) - \phi(\mu) + \phi(q) - \phi(p)) \theta(z) }{\theta(z + \phi(\lambda) - \phi(\mu)) \theta(z + \phi(q) - \phi(p))} \right. \notag \\ & \left. - \dfrac{\theta(z + \phi(\lambda) - \phi(p)) \theta(z+ \phi(q) - \phi(\mu)) E(\mu, \lambda) E(q,p)} {\theta(z+\phi(\lambda) - \phi(\mu)) \theta(z + \phi(q) - \phi(p)) E(\mu,p) E(q,\lambda)} \right\} Q. \label{specialpartialfrac} \end{align} \end{theorem} \begin{pf} The starting point of course is formula \eqref{prod} and \eqref{partialfrac} specialized to the case $\boldsymbol{\lambda} - \boldsymbol{\mu} = \lambda - \mu$. The formula \eqref{specialprod} is an immediate consequence of \eqref{prod}. Derivation of \eqref{specialpartialfrac} requires a little bit of algebra. We use the definition \[ \theta[z](\lambda) = \exp( \pi i a^T \Omega a + 2 \pi i a (\lambda + b)) \theta(\lambda + z) \] (where $z = \Omega a + b$ with $a, b \in {\bold R}^g$) to express all theta functions with characteristic $\theta[z](\cdot)$ in terms of the theta function itself $\theta(\cdot)$. When this is plugged into \eqref{partialfrac} and little bit of algebra is used to collect the exponential factor (noting that $a = a_{\widetilde{z}} - a_z$ if $\widetilde{z} - z (=\phi(\lambda) - \phi(\mu)) = \Omega a + b$ and $\widetilde{z} = \Omega a_{\widetilde{z}} + b_{\widetilde{z}}$, $z = \Omega a_z + b_z$ with $a$, $b$, $a_{\widetilde{z}}$, $b_{\widetilde{z}}$, $a_z$, $b_z$ in ${\bold R}^g$), we get \begin{align} T(p) = & \exp(- 2 \pi i a^T(\phi(p) - \phi(q) ) ) \left\{ \dfrac{ \theta(z + \phi(\lambda) - \phi(\mu) + \phi(q) - \phi(p))}{\theta(z + \phi(\lambda) - \phi(\mu)) E(q,p)} \right. \notag \\ & - \dfrac{\theta(z + \phi(\lambda) - \phi(p))}{\theta(z+\phi(\lambda) - \phi(\mu)) E(\mu,p)} \cdot \dfrac{ \theta(z + \phi(\lambda) - \phi(\mu) E(\mu, \lambda)}{\theta(z)} \notag \\ & \left. \cdot \dfrac{\theta(z+\phi(q) - \phi(\mu))} {\theta(z + \phi(\lambda) - \phi(\mu)) E(q, \lambda)} \right\} \cdot Q \dfrac{\theta(z) E(q,p)}{\theta(z+ \phi(q) - \phi(p))}. \notag \end{align} The formula \eqref{specialpartialfrac} now follows by simple algebraic manipulation. \end{pf} As a Corollary we obtain a version of Fay's Trisecant Identity (see \cite{oldfay} formula (45) page 34 or \cite{mumford} Volume II page 3.214). \begin{corollary} \label{C:faytrisecant} For $X$ a compact Riemann surface, $\phi$ its Abel-Jacobi map, $\theta(\lambda)$ and $E(p,q)$ its associated respective theta function and prime form, $p,q,\lambda, \mu$ points of $X$ and $z \in {\bold C}^g$, the following identity holds: \begin{gather} \theta(z+\phi(\lambda) - \phi(\mu)) \theta(z+\phi(q) - \phi(p)) E(p,\lambda) E(q,\mu) \notag \\ + \theta(z+\phi(\lambda) - \phi(p)) \theta(z + \phi(q) - \phi(\mu)) E(\lambda, \mu) E(q,p) \notag \\ = \theta(z + \phi(\lambda) - \phi(\mu) + \phi(q) - \phi(p)) \theta(z) E(p,\mu) E(q,\lambda). \label{faytrisecant} \end{gather} \end{corollary} \begin{pf} From the identity of the two expressions \eqref{specialprod} and \eqref{specialpartialfrac} for $T(p)$ in Theorem \ref{T:redlineabsint}, we have equality of the following two expressions for $\exp(2 \pi i a^T(\phi(p) - \phi(q))) T(p) Q^{-1}$: \begin{align} \dfrac{E(p,\lambda)}{E(p,\mu)} \dfrac{E(q,\mu)}{E(q,\lambda)} = & \dfrac{\theta(z + \phi(\lambda) - \phi(\mu) + \phi(q) - \phi(p)) \theta(z) }{\theta(z + \phi(\lambda) - \phi(\mu)) \theta(z + \phi(q) - \phi(p))} \notag \\ & - \dfrac{\theta(z + \phi(\lambda) - \phi(p)) \theta(z+ \phi(q) - \phi(\mu)) E(\mu, \lambda) E(q,p)} {\theta(z+\phi(\lambda) - \phi(\mu)) \theta(z + \phi(q) - \phi(p)) E(\mu,p) E(q,\lambda)} \notag \end{align} Multiplication of both sides by $\theta(z+\phi(\lambda)-\phi(\mu)) \theta(z + \phi(q) - \phi(p)) E(\mu,p) E(q, \lambda)$ along with a liberal use of the general identity $E(p,q) = -E(q,p)$ along with some algebra now leads to Fay's identity \eqref{faytrisecant} as desired. \end{pf} The identity \eqref{faytrisecant} is actually a special case of a more general identity (see Corollary 2.19 in \cite{oldfay}) which gives an explicit expression for the determinant of a matrix $M$ of the form \[ M = [M_{ij}] \text{ where } M_{ij} = \dfrac{\theta(z + \phi(\mu^j) - \phi(\lambda^i))}{E(\mu^j, \lambda^i)}. \] Since the cofactor matrices of such a matrix are of the same form, one can then compute explicitly (in terms of theta functions and prime forms) the entries of the inverse of the matrix $\Gamma$ appearing in Theorem \ref{T:lineabsint}. In this way one can verify by direct computation the identity of the two expressions \eqref{prod} and \eqref{partialfrac} for $T(p)$ in Theorem \ref{T:lineabsint}. This then is a canonical higher genus generalization of Theorem 4.3.2 in \cite{bgr}. Note that this proof of Fay's identity arises from equating a multiplicative formula for the solution of a zero-pole interpolation problem to a partial-fraction expression for the same solution. Formula \eqref{solution} in Theorem \ref{T:absint} gives an analogue of the partial fraction expression for the solution of a zero-pole interpolation problem for a vector bundle endomorphism. Formula \eqref{solution}, giving a connection between the Cauchy kernel $K(\chi; p,q)$ and $K(\widetilde \chi; p,q)$, can be viewed as a matrix-valued version of the Fay trisecant identity. There is one case in higher rank when a multiplicative representation does exist, namely the case of full rank zero-pole interpolation (where the given pole vectors $\{u_{j \beta} \colon 1 \le \beta \le s_j = r\}$ span the fiber space $\widetilde{\chi}(\mu^j)$ and the given null vectors $\{ x_{i \alpha} \colon 1 \le \alpha \le t_i =r\}$ span the fiber space $\widetilde{\chi}^\vee(\lambda^i)$) for each $i$ and $j$. Then $T(p)$ is again given by \eqref{prod} where $Q$ now is a product of a scalar from the fiber of ${\cal O}({\boldsymbol\lambda} - {\boldsymbol \mu})(q)$ and a value at $q$ of an automorphism of $\widetilde \chi$. Without loss of generality we may assume that $\{x_{i \alpha} \colon 1 \le \alpha \le r\}$ and $\{u_{j \beta} \colon 1 \le \beta \le r\}$ consist of the standard basis vectors for each $i$ and $j$. Then in the partial fraction expansion \eqref{solution} of $T(p)$ we have that $\Gamma$ has the block matrix form \begin{equation} \label{Gammafull} \Gamma = -[K(\widetilde{\chi}; \lambda^i, \mu^j)]_{i=1, \dots, n_0; j=1, \dots, n_\infty}, \end{equation} and that $K_{ {\boldsymbol \mu},{\bold u}}(p)$ and $K^{\bold x, {\boldsymbol \lambda}}(q)$ are block row and column matrices respectively \begin{gather} K_{{\boldsymbol \mu}, {\bold u}}(p) = \begin{bmatrix} K(\widetilde{\chi}; p, \mu^1) & \dots & K(\widetilde{\chi}; p, \mu^{n_\infty}) \end{bmatrix}, \\ K^{{\bold x},{\boldsymbol \lambda}}(q) = \begin{bmatrix} K(\widetilde{\chi}; \lambda^1, q) \\ \vdots \\ K(\widetilde{\chi}; \lambda^{n_0}, q)\end{bmatrix}. \end{gather} Equating this multiplicative formula to the partial fraction expansion leads to the same result as in formula (2.16) in \cite{newfay}. \section{Determinantal representations of algebraic curves and kernel bundles via Cauchy kernels} \label{S:detrep} In \cite{hip} zero-pole interpolation problems of the sort discussed here were studied in a more concrete setting of vector bundles over an algebraic curve embedded in projective space with fiber space given as the kernel of a two-variable matrix pencil. In this section we make the connections between that setting and the abstract compact Riemann surface setting of Section~\ref{S:absint} of this paper explicit. As we shall see, the link between the two settings is provided by the Cauchy kernels introduced in Section~\ref{S:cauchyker}. We first review the setting from \cite{hip}. Suppose that we are given three $M \times M$ matrices $\sigma_1, \sigma_2, \gamma$ and let $U_0(z) = U_0(z_1, z_2)$ be the two-variable linear matrix pencil \[ U_0(z) = z_1 \sigma_2 - z_2 \sigma_1 + \gamma, \quad z=(z_1, z_2). \] We will also often consider the homogenization $U(\mu)$ (where $\mu = [\mu_0,\mu_1,\mu_2]$ are projective coordinates in ${\bold P}^2$) given by \[ U(\mu) = \mu_0 U_0(\frac{\mu_1}{\mu_0}, \frac{\mu_2}{\mu_0}) =\mu_1 \sigma_2 - \mu_2 \sigma_1 + \mu_0 \gamma. \] Although $\det U(\mu)$ is not well-defined as a function of the projective variable $\mu=[\mu_0,\mu_1,\mu_2]$, nevertheless its zero set is well-defined and defines a curve $C \subset {\bold P}^2$ by \[ C=\{\mu=[\mu_0,\mu_1,\mu_2] \in {\bold P}^2\colon \det U(\mu)=0\}. \] We shall assume that $U(\mu)$ defines a maximal irreducible determinantal representation of rank $r$; this means that $\det U(\mu) = F(\mu)^r$ where $F$ is an irreducible homogeneous polynomial of degree $m$ (so $M=rm$), and that $\ker U(\mu) = r$ for all smooth points $\mu$ of $C$, i.e., points $\mu^0$ where at least one of $\frac{\partial F}{\partial \mu_0}(\mu^0)$, $\frac{\partial F}{\partial \mu_1}(\mu^0)$, and $\frac{\partial F}{\partial \mu_2}(\mu^0)$ is not zero. In case of a singular point $\mu^0$, we assume that $\dim \ker U(\mu^0)$ is as large as possible, namely $sr$ where $s$ is the multiplicity of $\mu^0$. Under these conditions $E(\mu) = \ker U(\mu)$ lifts to a vector bundle $E$ of rank $r$ over the normalizing Riemann surface $X$ of $C$; note that the bundle $E$ is realized concretely as a rank $r$ subbundle of the trivial bundle of rank $M$ over $X$. The normalizing Riemann surface $X$ is a Riemann surface such that there is a holomorphic mapping $\pi \colon X \to {\bold P}^2$ whose image equals $C$ such that $\pi$ is a one-to-one immersion on the inverse image of smooth points of $C$; we call $\pi \colon X \rightarrow C$ a birational embedding of $X$ in ${\bold P}^2$. For more details, see \cite{hip}. As in \cite{hip}, we shall assume for simplicity that all the singular points of $C$ are nodes (i.e., $\pi^{-1}(q) = \{p^1, p^2 \}$ where $p^1$ and $p^2$ are distinct points on $X$ with neighborhoods $U_1$ and $U_2$ such that $\pi$ is an immersion at both $p^1$ and $p^2$ and the analytic arcs $\pi(U_1) $ and $\pi(U_2)$ intersect transversally at $q$). We also assume that the line at infinity $\{\mu_0 = 0\}$ is nowhere tangent to $C$. The holomorphic vector bundle $E_\ell$ which is dual to $E$ can be realized concretely as a subbundle of the trivial rank $M$ bundle over $X$ (with fibers now written as row vectors) via \[ E_\ell(\mu) = \ker_\ell U(\mu) =\{x \in {\bold C}^{1 \times M} \colon x U(\mu) = 0\}. \] A concrete pairing between $E_\ell \otimes {\cal O}(1) \otimes \Delta$ and $E \otimes {\cal O}(1) \otimes \Delta$ is given by \begin{equation} \label{pairing} \{u_\ell, u\} = \frac{u_\ell}{\mu_0} \frac{\xi_1\sigma_1 + \xi_2 \sigma_2}{\xi_1\ d\lambda_1 + \xi_2\ d\lambda_2} \frac{u}{\mu_0}. \end{equation} Here $u_\ell$ and $u$ are local holomorphic sections of $E_\ell \otimes {\cal O}(1) \otimes \Delta$ and $E\otimes {\cal O}(1) \otimes \Delta$ respectively, $\lambda_1$ and $\lambda_2$ are meromorphic functions on $X$ given by $\lambda_1 = z_1 \circ \pi$ and $\lambda_2 = z_2 \circ \pi$, and $\xi_1,\xi_2$ are arbitrary (not both zero) complex parameters. If $E$ and $E_\ell$ are right and left kernel bundles determined by a rank $r$ maximal determinantal representation $U(\mu)$ of a curve $C$ as above, then it can be shown that necessarily $E \otimes {\cal O}(1) \otimes \Delta$ is isomorphic to a flat bundle $\chi$ over $X$ with the property that $h^0(\chi \otimes \Delta) = 0$, and that $E_\ell \otimes {\cal O}(1) \otimes \Delta$ is isomorphic to the dual $\chi^\vee$ of $\chi$. This isomorphism of $E \otimes {\cal O}(1) \otimes \Delta$ with $\chi$ is implemented explicitly by a {\it matrix of normalized sections} $u^\times(p)$. Explicitly, $u^\times$ is an $M \times r$ matrix whose columns are meromorphic sections of the pullback of $E \otimes \Delta$ to the universal cover $\widetilde{X}$ of $X$ such that: \begin{enumerate} \item[1.] $\dfrac{1}{\sqrt{dt}(R \widetilde{p})} u^\times(R \widetilde{p}) = \frac{1}{\sqrt{dt}(\widetilde{p})}u^\times(\widetilde{p}) \chi^{-1}(R)$ for all $\widetilde{p} \in \widetilde{X}$ and all $R \in \text{Deck}(\widetilde{X}/X)$ $ \cong \pi_1(X)$, where $t$ is a local parameter on $X$ and $\sqrt{dt}$ is the corresponding local holomorphic frame for $\Delta$ lifted to the neighborhoods of $\widetilde{p}$ and $R\widetilde{p}$ on $\widetilde{X}$. \item[2.] Each column of $u^\times$ has first order poles at (the points of $\widetilde{X}$ over) the points of $C$ at infinity, and is holomorphic everywhere else. \item[3.] For each $p \in X$, the columns of $u^\times(\widetilde{p})$ form a basis for the fiber $(E \otimes \Delta)(p)$, where $\widetilde{p} \in \widetilde{X}$ is over $p$ (if $p$ is a point of $C$ at infinity we have first to multiply $u^\times$ by a local parameter centered at $p$). \end{enumerate} Simply speaking, $u^\times$ consists of a multiplicative $\Delta$-valued meromorphic frame for $E$, normalized to have poles exactly at the points of $C$ at infinity. An isomorphism $\chi \to E \otimes {\cal O}(1) \otimes \Delta$ is now given explicitly by $y \to \mu_0 u^\times y$ where $y$ is a local holomorphic section of $\chi$. An $r \times M$ matrix of normalized sections $u_\ell^\times$ of $E_\ell$, whose rows are meromorphic sections of the pullback of $E_\ell \otimes \Delta$ to the universal covering $\widetilde{X}$ of $X$, is defined similarly, with item (1) replaced by \begin{enumerate} \item[$1_\ell .$] $\dfrac{1}{\sqrt{dt}} (R \widetilde{p}) u_\ell^\times(R \widetilde p) = \frac{1}{\sqrt{dt}(\widetilde{p})} \chi(R) u_\ell^\times(\widetilde{p})$ for all $\widetilde{p} \in \widetilde{X}$ and all $R \in \text{Deck}(\widetilde{X}\backslash X) \cong \pi_1(X)$, where $t$ and $\sqrt{dt}$ are as before. \end{enumerate} An isomorphism $\chi^\vee \to E_\ell \otimes {\cal O}(1) \otimes \Delta$ is given explicitly by $x \to \mu_0 x^T u_\ell^\times$, where $x$ is a local holomorphic section of $\chi^\vee$. Given $u^\times$, the {\it dual } matrix of normalized section $u_\ell^\times$ is determined uniquely by \[ u_\ell^\times \frac{\xi_1 \sigma_1 + \xi_2 \sigma_2}{\xi_1\ d\lambda_1 + \xi_2\ d\lambda_2} u^\times = I_r \] (where $I_r$ is the $r \times r$ identity matrix), so that under the isomorphisms $\chi \cong E \otimes {\cal O}(1) \otimes \Delta$ and $\chi^\vee \cong E_\ell \otimes {\cal O}(1) \otimes \Delta$ the natural duality pairing between $\chi^\vee$ and $\chi$ equals the pairing \eqref{pairing}. If we now define $K(\chi; \widetilde{p}, \widetilde{q})$ by \begin{equation} \label{cauchypairing} K(\chi; \widetilde{p}, \widetilde{q}) = u_\ell^\times(\widetilde{p}) \frac{\xi_1 \sigma_1 + \xi_2 \sigma_2}{\xi_1(\lambda_1(\widetilde{p}) - \lambda_1(\widetilde{q})) + \xi_2 (\lambda_2(\widetilde{p}) - \lambda_2(\widetilde{q}))} u^\times(\widetilde{q}), \end{equation} then $K$ has all the properties of the Cauchy kernel as defined in Section \ref{S:cauchyker}. This method of constructing the Cauchy kernel, via a dual pair of normalized sections of the kernel bundles associated with a maximal determinantal representation of an algebraic curve $C$ embedded in ${\bold P}^2$ which has the Riemann surface $X$ as its normalizing surface, was presented in \cite{hip}. Here we wish to make explicit the reverse path. We start with a compact Riemann surface $X$ and a flat holomorphic vector bundle $\chi$ over $X$ for which $h^0(\chi \otimes \Delta) = 0$. We assume as given the associated Cauchy kernel as developed in Section \ref{S:cauchyker}. We then construct a birational embedding of $X$ into ${\bold P}^2$ with image equal to the curve $C$ together with a rank $r$ maximal determinantal representation of $C$ in such a way that we recover the Cauchy kernel $K(\chi; \cdot, \cdot)$ from a dual pair of normalized sections for the associated left and right kernel bundles associated with this determinantal representation of $C$, as in \eqref{cauchypairing}. We first need some preparations. Let $\chi$ be a flat vector bundle over the Riemann surface $X$ such that $h^0(\chi \otimes \Delta)=0$. In addition choose two scalar meromorphic functions $\lambda_1$, $\lambda_2$ on $X$ such that ${\cal M}(X) = {\bold C}(\lambda_1, \lambda_2)$, i.e., rational functions in $\lambda_1,\lambda_2$ generate the whole field of (scalar) meromorphic functions on $X$. Assume that all poles of $\lambda_1$ and $\lambda_2$ are simple, and denote the set of poles by $x^1, \dots, x^m \in X$. Define complex numbers $c_{ik}$ ($1 \le i \le m$, $k=1,2$) by \[ c_{ik} = -\text{Res}_{p=x^i}\lambda_k(p) \] where the residue is with respect to some fixed local coordinate $t^i = t^i(p)$ centered at $p=x^i$. On occasion we shall also need the next coefficient $-d_{ik}$ in the Laurent expansion of $\lambda_k$ at $x^\i$: \[ \lambda_k(p) = -\frac{c_{ik}}{t^i} - d_{ik} + O(|t^i|). \] Define $M \times M$ matrices (where $M = mr$) $\sigma_1, \sigma_2,\gamma$ by \begin{equation} \label{pencilcoef} \sigma_1 = \underset{1\le i \le m}{\text{diag.}} (c_{i1} I_r), \quad \sigma_2 = \underset{1\le i \le m}{\text{diag.}}(c_{i2}I_r) \quad \gamma = [\gamma_{ij}]_{i,j=1,\dots,m} \end{equation} where \[ \gamma_{ij} = \begin{cases} d_{i1}c_{i2} -d_{i2} c_{i1}, & i=j \\ (c_{i1} c_{j2}- c_{j1}c_{i2}) \dfrac{K(\chi; x^i, x^j)}{dt^j(x^j)}, & i \ne j. \end{cases} \] Also define \begin{equation} \label{normsec} u^\times(p) = \begin{bmatrix} K(\chi; x^1, p) \\ \vdots \\ K(\chi; x^m,p) \end{bmatrix}, \quad u_\ell^\times(p) = -\begin{bmatrix} K(\chi; p, x^1) & \dots & K(\chi; p, x^m) \end{bmatrix} \end{equation} Then we have the following result. \begin{theorem} \label{T:detrep} Let $\chi$ be a flat vector bundle over the Riemann surface $X$ such that $h^0(\chi \times \Delta) = 0$ with associated Cauchy kernel $K(\chi; \cdot, \cdot)$ and use a pair of meromorphic functions $\lambda_1(p),\lambda_2(p)$ on $X$ which generate the field ${\cal M}(X)$ of meromorphic functions on $X$ to define matrices $\sigma_1$, $\sigma_2$ and $\gamma$ as in \eqref{pencilcoef}. Then: (i) The map $\pi_0 \colon X \to {\bold C}^2$ given by \[ \pi_0(p) = (\lambda_1(p), \lambda_2(p)) \] maps $X \backslash \{x^1, \dots, x^m\}$ onto the affine part $C_0$ of an algebraic curve $C \subset {\bold P}^2$ and extends to a birational embedding $\pi \colon X \rightarrow C$ of $X$ in ${\bold P}^2$. The defining irreducible homogeneous polynomial $F(\mu_0, \mu_1, \mu_2)$ of $C$ is such that $ \det (\mu_1 \sigma_2 - \mu_2 \sigma_1 + \mu_0 \gamma) = F(\mu_0, \mu_1, \mu_2)^r$. (ii) Denote by $E$ the kernel bundle over $C$ given in affine coordinates by \[ E(\lambda) = \ker (\lambda_1 \sigma_2 - \lambda_2 \sigma_1 + \gamma),\quad \lambda= (\lambda_1, \lambda_2). \] Then $\chi \cong E \otimes {\cal O}(1) \otimes \Delta$ with $u^\times$ and $u_\ell^\times$ given by \eqref{normsec} equal to the dual matrices of normalized sections of $E$ and $E_\ell$. \end{theorem} We shall prove Theorem \ref{T:detrep} under the assumption that all the singular points of $C$ are nodes. \begin{pf} We define the curve $C$ as the compactification in projective space of the image of the map $\pi$ in the statement of the theorem \[ C_0 = \{(\lambda_1(p), \lambda_2(p))\colon p \in X\}. \] Then $X$ is the normalizing Riemann surface for the curve $C$ and the degree of $C$ is equal to the number of intersections with the line at infinity, namely, deg $C=m$. Let $f(z_1,z_2)=0$ be an irreducible polynomial of degree $m$ such that $f(z_1,z_2)=0$ is the defining equation for $C$ (in affine coordinates). Thus $f(\lambda_1(p),\lambda_2(p))=0$ for all $p \in X$. We must show that $z_1\sigma_2 - z_2 \sigma_1 + \gamma$ is a maximal determinantal representation of $f(z_1,z_2)^r=0$, that $E \otimes {\cal O}(1) \otimes \Delta \cong \chi$ and that $u^\times$ and $u_\ell^\times$ are dual matrices of normalized sections of $E$ and $E_\ell$. The first step is to prove the identities \begin{align} (\lambda_1(p) \sigma_2 -\lambda_2(p)\sigma_1+\gamma)u^\times(p) & =0 \label{identity1} \\ u_\ell^\times(p) (\lambda_1(p) \sigma_2 - \lambda_2(p) \sigma_1 + \gamma) & = 0 \label{identity2} \\ \frac{u_\ell^\times(p) (\xi_1 \sigma_1 + \xi_2 \sigma_2 ) u^\times(p)} {\xi_1 \ d\lambda_1(p) + \xi_2\ d\lambda_2(p)} & = 1. \label{identity3} \end{align} To prove \eqref{identity1}, set \[ h(p) = (\lambda_1(p) \sigma_2 - \lambda_2(p) \sigma_1 + \gamma) u^\times(p). \] Note that $h(p)$ is a meromorphic section of $\chi \otimes \Delta$. To check that $h=0$ it suffices to check that $h$ has no poles (since $h^0(\chi \otimes \Delta)=0$). The only possible poles in the formula for $h$ occur at the points $x^1, \dots, x^m$ and these are at most double poles. For $\alpha = 1, \dots, m$, let us write down the Laurent expansion for $h$ near $x^\alpha$ as \[ h(p) = [h]^{\alpha,-2}(t^\alpha)^{-2} + [h]^{\alpha, -1} (t^\alpha)^{-1} + [\text{analytic at $x^\alpha$]}. \] We must show that $[h]^{\alpha, -2}=0$ and $[h]^{\alpha, -1} = 0$ for $1 \le \alpha \le m$. Each of $[h]^{\alpha, -2}$ and $[h]^{\alpha, -1}$ in turn is a block $m \times 1$ column matrix: $[h]^{\alpha, -2} = \left[ [h]^{\alpha, -2}_i\right]$ and $[h]^{\alpha, -1} = \left[ [h]^{\alpha, -1}_i \right]$ with $i = 1, \dots m$. We compute \begin{align} \notag [h]^{\alpha, -2}_i =& \sum_{j=1}^m (-c_{\alpha 1} c_{i2} \delta_{ij} + c_{\alpha 2} c_{i 1} \delta_{ij}) \delta_{j \alpha}\cdot (- dt^\alpha(x^\alpha))\\ \notag =& ( c_{\alpha 1} c_{\alpha 2} - c_{\alpha 2} c_{\alpha 1}) dt^\alpha(x^\alpha) = 0 \end{align} where $\delta_{ij}$ is the Kronecker delta. Similarly, \begin{align} \notag [h]^{\alpha, -1}_i =& \sum_{j,\ j\ne \alpha} (-c_{\alpha 1} c_{i 2} \delta_{ij} +c_{\alpha 2} c_{i 1} \delta_{ij}) K(\chi, x^j, x^\alpha) \\ \label{hresidue} & + \sum_j^m \{-d_{\alpha 1} c_{i 2} \delta_{ij} + d_{\alpha 2} c_{i 1} \delta_{ij} + \gamma_{ij} \} \delta_{j \alpha}(- dt^\alpha(x^\alpha)). \end{align} For $i = \alpha$ \eqref{hresidue} becomes \[ (-d_{\alpha 1} c_{\alpha 2} + d_{\alpha 2} c_{\alpha 1} + d_{\alpha 1} c_{\alpha 2} - c_{\alpha 1} d_{\alpha 2})(- dt^\alpha(x^\alpha))= 0 \] while, for $i \ne \alpha$, \eqref{hresidue} becomes \[ (-c_{\alpha 1} c_{i2} + c_{\alpha 2} c_{i 1}) K(\chi; x^i, x^\alpha) + (c_{i1} c_{\alpha 2} - c_{\alpha 1} c_{i 2}) \frac{K(\chi; x^i, x^\alpha)}{dt^\alpha(x^\alpha)}(- dt^\alpha(x^\alpha)) = 0. \] Thus $h=0$ and \eqref{identity1} follows. By a similar calculation of Laurent series coefficients one can verify \eqref{identity2}. To verify \eqref{identity3}, by a standard lemma (see Proposition 2.3 in \cite{hip}), it suffices to show that \[ \frac{u^\times_\ell(p) \sigma_k u^\times(p)}{d\lambda_k(p)} = 1 \text{ for } k=1,2. \] The numerator of the expression on the left is given by \begin{gather} \notag \begin{bmatrix} K(\chi; p,x^1) & \dots & K(\chi; p,x^m) \end{bmatrix} \begin{bmatrix} -c_{ik} & & \\ & \ddots & \\ & & - c_{mk} \end{bmatrix} \begin{bmatrix} K(\chi; x^1,p) \\ \vdots \\ K(\chi; x^m, p) \end{bmatrix} \\ \notag = -\sum_{i=1}^m K(\chi; p, x^i) [\text{Res}_{p=x^i} \lambda_k(p)] K(\chi; x^i,p) \end{gather} and hence \[ \frac{u^\times_\ell(p) \sigma_k u^\times(p)}{d\lambda_k(p)} = - \frac{\sum_{i=1}^m K(\chi; p, x^i) [\text{Res}_{p=x^i}\lambda_k(p)] K(\chi; x^i,p)} {d\lambda_k(p)}. \] The double pole at each $x^i$ in the numerator is cancelled by a double pole at each $x^i$ in the denominator. Note also that the product of two half-order differentials in the numerator is cancelled by the differential in the denominator. The resulting quotient is a well-defined holomorphic section of $Hom(\chi,\chi)$ which has the value $I$ at $x^1, \dots, x^m$, and hence must equal $I$ at all $p \in X$. Equation \eqref{identity3} now follows. Denote by $d(z_1,z_2)$ the polynomial $d(z_1,z_2)=\det (z_1\sigma_2-z_2\sigma_1 + \gamma)$. By \eqref{identity1} or \eqref{identity2} we know that $d(\lambda_1(p),\lambda_2(p))=0$ identically in $p \in X$. Since $f$ is by assumption the irreducible defining polynomial for $X$, it follows that $f(z_1,z_2) | d(z_1,z_2)$, and hence, \begin{equation} \label{fact} d(z_1,z_2) = f(z_1,z_2)^s g(z_1,z_2) \end{equation} for some positive integer $s$ and some polynomial $g$ relatively prime with respect to $f$. By inspection we see that $d$ has degree equal to $M=mr$, while, as already mentioned, $f$ has degree equal to $m$. From the factorization we see that deg $d \ge s(\text{deg } f)$, or $r \ge s$. On the other hand, at a smooth point $(z_1,z_2) \in C$, we know that $\dim E(z) \ge r$ since the $r$ linearly independent columns of $u^\times(p)$ are in $\ker (\lambda_1(p) \sigma_2 - \lambda_2(p) \sigma_1 + \gamma)$. But also, from the factorization \eqref{fact} and an inductive argument working with the minors the matrix pencil $z_1 \sigma_2 - z_2 \sigma_1 + \gamma$ (see the proof of Theorem 3.2 in \cite{hip}) one can show that $\dim E(z) \le s$. From $r \ge s$ and $r \le \dim E(z) \le s$ we conclude that $r=s$. From \eqref{fact} and degree counting we conclude that the polynomial $g$ is a constant, and without loss of generality, $d=f^r$. Thus $z_1\sigma_2 - z_2 \sigma_1 + \gamma$ is a maximal determinantal representation of $f(z_1,z_2)^r=0$ as asserted, except that we still must check that for any node $q$ on $C$ the columns of $u^\times(p^1)$ and $u^\times(p^2)$ are linearly independent, where $\pi^{-1}(q) = \{p^1, p^2\}$. It then follows from \eqref{identity1}, \eqref{identity2} and \eqref{identity3} that $u^\times(p)$ and $u^\times_\ell(p)$ form the associated dual pair of normalized cross-sections for $E$ and $E_\ell$ respectively. We first claim if $L$ is a straight line nowhere tangent to $C$ and $y^1, \dots, y^m$ are the preimages on $X$ of the points of intersection of $C$ with $L$, then the block matrix $[K(\chi; x^i, y^j]_{i,j = 1, \dots, m}$ is invertible. This follows immediately from the discussion of the full rank zero-pole interpolation problem at the end of Section 4, as we now show. Since the divisor $x^1 + \dots + x^m - y^1 - \dots - y^m$ is equivalent to $0$, the input bundle in the full rank zero-pole interpolation problem with zeros $x^1, \dots, x^m$ and poles $y^1, \dots, y^m$ and output bundle $\chi$ is again (isomorphic to) $\chi$. Since $h^0(\chi \otimes \Delta) = 0$, the matrix $\Gamma$ \eqref{Gammafull} is invertible. Next, by taking any line $L$ through the node $q$ which is nowhere tangent to $C$, we see that the columns of $u^\times(p^1)$ and $u^\times(p^2)$ are columns in a $M \times M$ invertible matrix (namely, the associated matrix $\Gamma$), and hence are linearly independent. \end{pf} We remark that the same proof works when the singularities of $C$ are any ordinary singular points, or more generally, are such that a singular point of multiplicity $s$ has $s$ distint preimages on $X$. \section{The concrete interpolation problem for meromorphic bundle maps between kernel bundles of determinantal representations of an algebraic curve} \label{S:conint} In the paper \cite{hip} the following problem was considered. We are given an irreducible algebraic curve $C$ in ${\bold P}^2$ together with its normalizing compact Riemann surface $X$ and the normalization map $\pi \colon X \to C$. We assume that the defining polynomial for $C$ is an irreducible polynomial $f$ of degree $m$ (in affine coordinates). For simplicity we assume again that the only singularities of $C$ are nodes and that $C$ intersects the line at infinity in $m$ distinct smooth points. We suppose in addition that $f^r$ has a maximal determinantal representation \[ f^r(z_1,z_2) = \det (z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma}) \] where $\sigma_1$, $\sigma_2$ and $\widetilde{\gamma}$ are $M \times M$ matrices ($M=mr$), with which is associated the kernel bundle $\widetilde{E}$ of rank $r$ over $C \backslash C_{sing}$ ($C_{sing}$ is the set of the singular points of $C$) with fibers (over affine points) given by \begin{equation} \label{outputbundle} \widetilde{E}(z) = \ker (z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma}). \end{equation} As explained in Section \ref{S:detrep}, we may consider the pullback of $\widetilde{E}$ to $X \backslash \pi^{-1}(C_{sing})$ as extended to a rank $r$ vector bundle over all of $X$. We also have the left kernel bundle $\widetilde{E}_\ell$ where \begin{equation} \label{dualoutputbundle} \widetilde{E}_\ell(z) = \ker_\ell (z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma}) \end{equation} with pullback under $\pi$ also extendable to a rank $r$ vector bundle defined over all of $X$. These bundles, or more precisely their twists $\widetilde{E} \otimes {\cal O}(1) \otimes \Delta$ and $\widetilde{E}_\ell \otimes {\cal O}(1) \otimes \Delta$, have the canonical pairing \eqref{pairing} with each other, as explained in Section \ref{S:detrep}. The data for the concrete interpolation (CONINT) problem to be considered in this section consists of: \begin{enumerate} \item[(D1)] $n_\infty$ distinct smooth, finite points $\mu^1 = (\mu^1_1, \mu^1_2), \dots, \mu^{n_\infty} = (\mu^{n_\infty}_1, \mu^{n_\infty}_2)$ of $C$ (the preassigned poles), \item[(D2)] for each $j=1, \dots, n_\infty$, a linearly independent set $\{\varphi_{j1}, \dots, \varphi_{j,s_j} \}$ of $s_j$ vectors in the fiber $\widetilde{E}(\mu^j)$ (the preassigned pole vectors), \item[(D3)] $n_0$ distinct smooth, finite points $\lambda^1 = (\lambda^1_1, \lambda^1_2), \dots, \lambda^{n_0}=(\lambda^{n_0}_1, \lambda^{n_0}_2)$ of $C$ (the preassigned zeros), \item[(D4)] for each $i=1, \dots, n_0$, a linearly independent set $\{\psi_{i1}, \dots, \psi_{i,t_i}\}$ of $t_i$ vectors in the fiber $\widetilde{E}_\ell(\lambda^i)$ (the preassigned null vectors), and \item[(D5)] for each pair of indices $(i,j)$ for which $\lambda^i = \mu^j =: \xi^{ij}$, a choice of a local coordinate $t^{ij}$ on $X$ centered at $\xi^{ij}$ and a collection of numbers $\{\rho_{ij, \alpha \beta} \colon 1 \le \alpha \le t_i, 1 \le \beta \le s_j\}$ (the preassigned coupling numbers with respect to the chosen local coordinate). \end{enumerate} The interpolation problem then is to find an $M \times M$ matrix $\gamma$ defining a maximal determinantal representation of $f^r$ \begin{equation} \label{detrepforfr} f^r(z_1,z_2) = \det (z_1 \sigma_2 - z_2 \sigma_1 + \gamma) \end{equation} giving the kernel bundle $E$ over $X$ with fiber over a smooth finite point $z \in C$ given by \begin{equation} \label{inputbundle} E(z) = \ker (z_1 \sigma_2 - z_2 \sigma_1 + \gamma) \end{equation} and the left kernel bundle $E_\ell$ given by \begin{equation} \label{dualinputbundle} E_\ell(z) = \ker_\ell (z_1 \sigma_2 - z_2 \sigma_1 + \gamma) \end{equation} together with meromorphic bundle maps \[ S \colon E \to \widetilde{E},\quad S_\ell \colon \widetilde{E}_\ell \to E_\ell \] (where we write bundle maps on left kernel bundles as acting from the right), where $S \otimes I_{{\cal O}(1) \otimes \Delta}$ and $S_\ell \otimes I_{{\cal O}(1) \otimes \Delta}$ are transposes of each other with respect to the pairing \eqref{pairing}, so that $S$ (and $S_\ell$) act as the identity operator $I$ on the corresponding fibers at the points at infinity, and the following set of interpolation conditions is satisfied: \begin{enumerate} \item[(I1)] $S$ has poles only at $\mu^1, \dots, \mu^{n_\infty}$; for each $j=1, \dots, n_\infty$, the pole of $S$ at $\mu^j$ is simple, and the vectors $\{\varphi_{j1}, \dots, \varphi_{j,n_\infty}\}$ span the image space of the residue $R_j \colon E(\mu^j) \to \widetilde{E}(\mu^j)$ of $S$ at $\mu^j$. \item[(I2)] The bundle map $S_\ell^{-1} \colon E_\ell \to \widetilde{E}_\ell$ has poles only at the points $\{\lambda^1, \dots, \lambda^{n_0}\}$; for each $i=1, \dots, n_0$, the pole of $S_\ell^{-1}$ at $\lambda^i$ is simple and the vectors $\{\psi_{i1} \dots, \psi_{i t_i}\}$ span the image space of the residue $\widehat{R}_i \colon E_\ell(\lambda^i) \to \widetilde{E}_\ell(\lambda^i)$ of $S_\ell^{-1}$ at $\lambda^i$. \item[(I3)] For each pair of indices $(i,j)$ where $\lambda^i = \mu^j =: \xi^{ij}$, and for $\alpha = 1, \dots, t_i$, let $\psi_{i \alpha}(p)$ be a local holomorphic section of $\widetilde{E}_\ell$ near $\xi^{ij}$ with \[ \psi_{i \alpha}(t^{ij}) = \psi_{i \alpha} + \psi_{i \alpha 1} t^{ij} + o(t^{ij}) \] such that $\psi_{i \alpha} S_\ell(p)$ has analytic continuation to $p = \xi^{ij}$ with value there equal to $0$. Then, for any choice of complex parameters $\xi_1$ and $\xi_2$ \[ \psi_{i \alpha 1} \frac{\xi_1 \sigma_1 + \xi_2 \sigma_2} {\xi_1 \lambda_1^{\prime}(\xi^{ij}) + \xi_2 \lambda_2^{\prime}(\xi^{ij})}\varphi_{j \beta} - \psi_{i \alpha} (\xi_1 \sigma_1 + \xi_2 \sigma_2) \varphi_{j \beta} \frac{\xi_1 \lambda_1^{\prime \prime}(\xi^{ij}) + \xi_2 \lambda_2^{\prime \prime}(\xi^{ij})} {2(\xi_1 \lambda_1^{\prime}(\xi^{ij}) + \xi_2 \lambda_2^{\prime}(\xi^{ij}))^2} = \rho_{ij, \alpha \beta}. \] Here ${}^{\prime} = \dfrac{d}{dt^{ij}}$. \end{enumerate} It can be shown that a necessary consistency condition on the data set for the problem to have a solution is that \begin{equation} \label{ZP} \psi_{i \alpha} (\xi_1 \sigma_1 + \xi_2 \sigma_2) \varphi_{j \beta} = 0. \end{equation} Given the data set (D1)--(D5) we form the $n_0 \times n_\infty$ block matrix $\Gamma^0 = [\Gamma^0_{ij}]$ ($1 \le i \le n_0, 1 \le j \le n_\infty$) where $\Gamma^0_{ij}=[\Gamma^0_{ij,\alpha \beta}]$ ($1 \le \alpha \le t_i$, $1 \le \beta \le s_j$) in turn is the $t_i \times s_j$ matrix with entries given by \begin{equation} \label{Gamma0} \Gamma^0_{ij, \alpha \beta} = \begin{cases} \psi_{i \alpha} \frac{\xi_1 \sigma_1 + \xi_2 \sigma_2} {\xi_1(\mu^j_1 - \lambda^i_1) + \xi_2(\mu^j_2 - \lambda^i_2)} \varphi_{j \beta} & \text{if } \lambda^i \ne \mu^j \\ -\rho_{ij, \alpha \beta} & \text{if } \lambda^i = \mu^j. \end{cases} \end{equation} Additional matrices which we shall need are \begin{align} A_1 = \begin{bmatrix} \mu^1_1I_{s_1} & & \\ & \ddots & \\ & & \mu^{n_\infty}_1 I_{s_{n_\infty}} \end{bmatrix}, & \quad A_2 = \begin{bmatrix} \mu^1_2I_{s_1} & & \\ & \ddots & \\ & & \mu^{n_\infty}_2 I_{s_{n_\infty}} \end{bmatrix}, \notag \\ Z_1=\begin{bmatrix}\lambda^1_1 I_{t_1} & & \\ & \ddots & \\ & & \lambda^{n_0}_1 I_{t_{n_0}} \end{bmatrix}, & \quad Z_2=\begin{bmatrix}\lambda^1_2 I_{t_1} & & \\ & \ddots & \\ & & \lambda^{n_0}_2 I_{t_{n_0}} \end{bmatrix}, \notag \\ \varphi_j = \begin{bmatrix} \varphi_{j1} & \dots & \varphi_{j s_j} \end{bmatrix}, & \quad \varphi = \begin{bmatrix} \varphi_1 & \dots & \varphi_{n_\infty} \end{bmatrix}, \notag \\ \psi_i = \begin{bmatrix} \psi_{i1} \\ \vdots \\ \psi_{i t_i} \end{bmatrix}, & \quad \psi = \begin{bmatrix} \psi_1 \\ \vdots \\ \psi_{n_0} \end{bmatrix}. \label{matrices} \end{align} The solution of the concrete interpolation problem (CONINT) obtained in \cite{hip} is as follows. \begin{theorem} \label{T:conint} (See Theorem 4.1 of \cite{hip}.) Assume that we are given a curve $C$ with defining irreducible polynomial $f$, a maximal determinantal representation for $f^r$ as in \eqref{detrepforfr} together with associated kernel bundle $\widetilde{E}$ \eqref{outputbundle} and left kernel bundle $\widetilde{E}_\ell$ \eqref{dualoutputbundle} and a data set (D1)--(D5) for the interpolation problem (I1)--(I3). Then the interpolation problem has a solution if and only if the interpolation data satisfy the compatibility conditions \eqref{ZP} at the overlapping zeros and poles, and the matrix $\Gamma^0$ given by \eqref{Gamma0} is square and invertible. In this case the unique solution of the interpolation problem (I1)--(I3) is given by \begin{equation} \label{gamma} \gamma = \widetilde{\gamma} - \sigma_1 \varphi (\Gamma^0)^{-1} \psi \sigma_2 + \sigma_2 \varphi (\Gamma^0)^{-1} \psi \sigma_1 \end{equation} with associated kernel bundle $E$ \eqref{inputbundle} and left kernel bundle $E_\ell$ \eqref{dualinputbundle}, with $S(z)$ given by \begin{equation} \label{formulaforS} S(z) = [I + \varphi (\xi_1(z_1I - A_1) + \xi_2(z_2 I -A_2))^{-1} (\Gamma^0)^{-1} \psi (\xi_1 \sigma_1 + \xi_2 \sigma_2)]|_{E(z)} \end{equation} and with $S_\ell^{-1}(z)$ given (as a right multiplication operator) by \begin{equation} \label{formulaforSell} S_\ell^{-1}(z) = [I-(\xi_1 \sigma_1 + \xi_2 \sigma_2) \varphi (\Gamma^0)^{-1} \left(\xi_1 (z_1 I - Z_1) + \xi_2( z_2 I - Z_2) \right)^{-1} \psi] |_{E_\ell(z)}. \end{equation} Here the matrices $A_1,A_2,Z_1,Z_2,\psi,\varphi$ are as in \eqref{matrices}. \end{theorem} The main goal of this section is to use the machinery developed in Section \ref{S:detrep} to make explicit the connections between Theorem \ref{T:conint} and Theorem \ref{T:absint} of Section \ref{S:absint}. Suppose therefore that $\chi$ and $\widetilde{\chi}$ are two flat bundles over the Riemann surface $X$ with $h^0(\chi \otimes \Delta)=0=h^0(\widetilde{\chi} \otimes \Delta)$. We use a fixed pair $(\lambda_1(p), \lambda_2(p))$ of meromorphic functions on $X$ generating ${\cal M}(X)$ to produce a map $\pi \colon X \to C$. The respective Cauchy kernels $K(\chi; \cdot, \cdot)$ and $K(\widetilde{\chi}; \cdot, \cdot)$ generate corresponding maximal determinantal representations $z_1 \widetilde{\sigma}_2 - z_2 \widetilde{\sigma}_1 + \widetilde{\gamma}$ and $z_1 \sigma_2^{\prime} - z_2 \sigma_1^{\prime} + \gamma^{\prime}$ for $f(z_1,z_2)^r=0$ as in Theorem \ref{T:detrep}. Note however that the formulas for $\sigma_1$ and $\sigma_2$ in \eqref{pencilcoef} depend only on the choice of embedding functions $\lambda_1(p)$ and $\lambda_2(p)$, and not on the particular flat bundle $\chi$; hence we may and shall write simply $\sigma_i$ in place of $\widetilde{\sigma}_i$ and $\sigma_i^{\prime}$ for $i=1,2$. We also have associated dual pairs of matrices of normalized sections: $\widetilde{u}^\times, \widetilde{u}_\ell^\times$ for $\widetilde{E}= \ker(z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma})$ and $\widetilde{E}_\ell = \ker_\ell (z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma})$ respectively, and $u^{\times \prime}, u^{\times \prime}_\ell$ for $E= \ker (z_1 \sigma_2 - z_2 \sigma_1 + \gamma^{\prime})$ and $E_\ell = \ker_\ell (z_1 \sigma_2 - z_2 \sigma_1 + \gamma^{\prime})$. Since $\widetilde{u}^\times$ implements an isomorphism between $\widetilde{\chi}$ and $\widetilde{E} \otimes {\cal O}(1) \otimes \Delta$ and $u^{\times \prime}$ implements an isomorphism between $\chi$ and $E^\prime \otimes {\cal O}(1) \otimes \Delta$, any meromorphic bundle map $T \colon \chi \to \widetilde{\chi}$ induces a meromorphic bundle map $S \colon E^\prime \to \widetilde{E}$ determined by \[ S(p) u^{\times \prime}(p) = \widetilde{u}^\times(p) T(p). \] However, in the solution of (CONINT) from \cite{hip} stated in Theorem \ref{T:conint}, the solution $S$ is normalized to act as the identity operator over the points of $C$ at infinity. In order for the map $S$ constructed as above from the abstract bundle map $T \colon \chi \to \widetilde{\chi}$ to achieve this normalization, we must make an adjustment \begin{equation} \label{adjustment} \alpha (z_1 \sigma_2 - z_2 \sigma_1 + \gamma^\prime) \beta = z_1 \sigma_2 - z_2 \sigma_1 + \gamma \end{equation} on the input determinantal representation, where $\alpha, \beta \in GL(M^{mr}, {\bold C})$. If $u^\times, u^\times_\ell$ is the dual pair of normalized sections for $E=\ker (z_1 \sigma_2 - z_2 \sigma_1 + \gamma)$ and $E_\ell = \ker_\ell (z_1 \sigma_2 - z_2 \sigma_1 + \gamma)$, then \[ u^\times(p) = \beta^{-1} u^{\times \prime}(p), \quad u^\times_\ell(p) = u^{\times \prime}_\ell(p) \alpha^{-1} \] and we seek to solve instead the equation \[ S(p) u^\times(p) = \widetilde{u}^\times(p) T(p), \] or equivalently \begin{equation} \label{intertwining} S(p) \beta^{-1} u^{\times \prime}(p) = \widetilde{u}^\times(p) T(p) \end{equation} for $S,\alpha,\beta$ subject to the proviso that $S(x^i)= I_{E(x^i)=\widetilde{E}(x^i)}$ for $i=1, \dots, m$. Since the columns of $u^{\times \prime}(p)$ and $\widetilde{u}^\times(p)$ (after multiplication by a local parameter on $X$ at $p$) simply form a standard basis in ${\bold C}^M$ when evaluated at $x^1, \dots, x^m$, we see that we should take \[ \beta = \begin{bmatrix} T(x^1) & & \\ & \ddots & \\ & & T(x^m) \end{bmatrix}^{-1}. \] In order to guarantee $\alpha \sigma_k \beta = \sigma_k$ for $k=1,2$ as required in \eqref{adjustment} we then take \[ \alpha = \begin{bmatrix} T(x^1) & & \\ & \ddots & \\ & & T(x^m) \end{bmatrix}. \] Thus $\gamma = \alpha \gamma^{\prime} \beta$ is given by $\gamma = [\gamma_{ij}]_{i,j=1,\dots, m}$ with \begin{equation} \label{inputgamma} \gamma_{ij} = \begin{cases} d_{i1}c_{i2} - c_{i1} d_{i2} & \text{if } i=j \\ (c_{i1}c_{j2} - c_{j1}c_{i2})T(x^i)K(\chi;x^i,x^j)T(x^j)^{-1} & \text{if } i \ne j. \end{cases} \end{equation} We remark that the ``adjustment'' of $\gamma$ by the values of a bundle map at the points of $X$ over the points of $C$ at infinity plays a central role in the construction of triangular models for commuting nonselfadjoint operators; see \cite{vin1} and \cite{lkmv}, Chapter 12. We now suppose that we are given an abstract interpolation data set $\omega$ as in \eqref{dataset} for an Abstract Interpolation Problem (ABSINT) as in Section \ref{S:absint} with output bundle $\widetilde{\chi}$, and let $y \to \widetilde{u}^\times y$ and $x \to x^T \widetilde{u}^\times_\ell$ be the associated bundle isomorphisms from $\widetilde{\chi}$ to $\widetilde{E} \otimes {\cal O}(1) \otimes \Delta$ and from $\widetilde{\chi}^\vee$ to $\widetilde{E}_\ell \otimes {\cal O}(1) \otimes \Delta$, where $\widetilde{E}$ and $\widetilde{E}_\ell$ are the right and left kernel bundles respectively associated with the maximal determinantal representation $z_1 \sigma_2 - z_2 \sigma_1 + \widetilde{\gamma}$ (with $\sigma_1,\sigma_2,\widetilde{\gamma}$ given by \eqref{pencilcoef}, and with $\widetilde{u}^\times$ and $\widetilde{u}^\times_\ell$ given by \eqref{normsec}, all with $\widetilde{\chi}$ in place of $\chi$). We assume that the pair of meromorphic functions $\lambda_1(p)$ and $\lambda_2(p)$ is chosen in such a way that the set of poles $x^1, \dots, x^m$ is disjoint from the preassigned poles $\mu^1, \dots, \mu^{n_\infty}$ and the set of preassigned zeros $\lambda^1, \dots, \lambda^{n_0}$. Define a data set $\omega_0$ for a (CONINT) problem as follows: \begin{enumerate} \item The preassigned poles consist of the points $\pi(\mu^1) = (\mu^1_1, \mu^1_2), \dots, \pi(\mu^{n_\infty}) = (\mu^{n_\infty}_1, \mu^{n_\infty}_2)$ with associated pole vectors $\varphi_{j \beta} \in \widetilde{E}(\pi(\mu^j))$ given by $\varphi_{j \beta} = \widetilde{u}^\times(\mu^j) u_{j \beta}$ for $j=1, \dots, n_\infty$ and $\beta = 1, \dots, s_j$. \item The preassigned zeros consist of the points $\pi(\lambda^1) = (\lambda^1_1, \lambda^1_2), \dots, \pi(\lambda^{n_0}) = ( \lambda^{n_0}_1, \lambda^{n_0}_2)$ with associated sets of null vectors $\psi_{i \alpha} \in \widetilde{E}_\ell(\pi(\lambda))$ given by $\psi_{i \alpha} = x^T_{i \alpha} \widetilde{u}^\times_\ell(\lambda^i)$ for $i=1, \dots, n_0$ and $\alpha=1, \dots, t_i$. \item For those pairs of indices $(i,j)$ for which $z^i=w^j=:\xi^{ij}$ we take the associated coupling numbers $\rho_{ij, \alpha \beta}$ to be the same as those specified for the (ABSINT) problem. \end{enumerate} With this choice of data set, the reader can check that the matrices $\varphi$ and $\psi$ as defined in \eqref{matrices} reduce to \[ \varphi = \begin{bmatrix} K_{\boldsymbol \mu, {\bold u}}(x^1) \\ \vdots \\ K_{\boldsymbol \mu,{\bold u}}(x^m) \end{bmatrix}, \quad \psi = -\begin{bmatrix} K^{{\bold x}, \boldsymbol \lambda}(x^1) & \dots & K^{{\bold x}, \boldsymbol \lambda}(x^m) \end{bmatrix} \] where the notation $K_{\boldsymbol \mu,{\bold u}}(p)$ and $K^{{\bold x}, \boldsymbol \lambda}(p)$ is as in the statement of Theorem \ref{T:absint}. It turns out that the (ABSINT) problem with data set $\omega$ is equivalent to the (CONINT) problem with data set $\omega_0$ under the identifications $\pi \colon X \to C$ and $\widetilde{u}^\times: \widetilde{\chi} \to \widetilde{E} \otimes {\cal O}(1) \otimes \Delta$ and $\widetilde{u}^\times_\ell: \widetilde{\chi}^\vee \to \widetilde{E}_\ell \otimes {\cal O}(1) \otimes \Delta$ sketched above, in that a solution $T$ of (ABSINT) corresponds to a solution $S$ of (CONINT) under the correspondence (including the normalization at the points over infinity) between abstract bundle maps $T$ and concrete bundle maps $S$ discussed above. (For the first two interpolation conditions, this observation is rather transparent. For the third interpolation condition (I3), this requires the relation between the interpolation condition (I3) for the (CONINT) problem with a flat connection on the bundle $\widetilde{E}_\ell \otimes {\cal O}(1) \otimes \Delta$ and the correspondence of this connection with the coefficient $A_\ell(p)$ in the Laurent expansion of the Cauchy kernel $K(\widetilde{\chi}; \cdot, \cdot)$ along the diagonal; this is explained in Section 3.2 of \cite{hip}.) Hence the formula for the solution $T$ of (ABSINT) in \eqref{solution} must correspond to the formula for the solution $S$ of (CONINT) in \eqref{formulaforS} under the correspondence \eqref{intertwining}. The point of the next result is to verify this directly; in addition we see that the matrix $\Gamma$ appearing in Theorem \ref{T:absint} is identical to the matrix $\Gamma^0$ appearing in Theorem \ref{T:conint}. \begin{theorem} \label{T:comparison} Let $\omega$ be the data set for an (ABSINT) problem with $\omega_0$ the corresponding data set for a (CONINT) problem. Then $\Gamma = \Gamma_0$. Furthermore, if $\Gamma$ is invertible and $\chi$ is the input bundle for which (ABSINT) has a solution, then a solution $T$ of (ABSINT) is related to the unique solution $S$ of (CONINT) having value identity on the fibers over the points at infinity according to the intertwining condition \eqref{intertwining}. \end{theorem} \begin{pf} The fact that $\Gamma = \Gamma_0$ is a simple consequence of the definitions and of the formula \eqref{cauchypairing} expressing the Cauchy kernel $K(\widetilde{\chi}; \cdot, \cdot)$ in terms of a dual pair $\widetilde{u}^\times, \widetilde{u}^\times_\ell$ of normalized sections of $\widetilde{E}$ and $\widetilde{E}_\ell$. It remains to verify the intertwining relation \eqref{intertwining} \[ S(p) \beta^{-1} u^{\times \prime}(p) = \widetilde{u}^\times(p) T(p) \] where $S$ is given by \eqref{formulaforS}, $T$ by \eqref{solution} and $\beta^{-1} = \underset{1\le i\le m}{\text{diag.}} \{T(x^i)\}$. We compute \begin{align} \notag [S(p) \beta^{-1} u^{\times \prime}(p)]_i = & \sum_{j=1}^m S_{ij}(p) T(x^j) K(\chi; x^j,p) \\ \notag = T(x^i) K(\chi; x^i,p) & - K_{\boldsymbol \mu,{\bold u}}(x^i) \cdot \underset{i'}{\text{diag.}} \{( \xi_1 \lambda_1(p)+\xi_2 \lambda_2(p) - \xi_1 \mu^{i^\prime}_1 -\mu^{i^\prime}_2)^{-1} I_{s_{i^\prime}} \} \cdot \\ & \cdot \Gamma^{-1} \cdot \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K^{{\bold x}, \boldsymbol \lambda}( \lambda^j) T(x^j) K(\chi; x^j,p). \label{Sside1} \end{align} From \eqref{solution} with $x^i$ in place of $p$ and $p$ in place of $q$ (and hence $T(p)$ in place of $Q$) we see that \[ T(x^i) K(\chi; x^i,p)=[K(\widetilde{\chi};x^i,p)+ K_{\boldsymbol \mu,{\bold u}}(x^i) \Gamma^{-1} K^{{\bold x}, \boldsymbol \lambda}(p)] T(p). \] We use this identity both in the form indicated and with $x^j$ in place of $x^i$ to convert \eqref{Sside1} to \begin{align} \notag [S(p) \beta^{-1} u^{\times \prime}(p)]_i = & [K(\widetilde{\chi};x^i,p) +K_{\boldsymbol \mu,{\bold u}}(x^i) \Gamma^{-1} K^{{\bold x},\boldsymbol \lambda}(p)] T(p) - \\ \notag - K_{\boldsymbol \mu,{\bold u}}(x^i) \cdot & \underset{i^\prime}{\text{diag.}} \{(\xi_1 \lambda_1(p) + \xi_2 \lambda_2(p) - \xi_1 \mu^{i^\prime}_1 - \xi_2 \mu^{i^\prime}_2)^{-1}I_{s_{i^\prime}}\} \cdot \\ \label{Sside2} \cdot \Gamma^{-1} \cdot \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) & K^{{\bold x},\boldsymbol \lambda}(x^j) [K(\widetilde{\chi};x^j,p) +K_{\boldsymbol \mu,{\bold u}}(x^j) \Gamma^{-1} K^{{\bold x},\boldsymbol \lambda}(p)] T(p). \end{align} Next we use the general identity (see also \cite{AV}) \begin{gather} \notag \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K(\widetilde{\chi};p,x^j) K(\widetilde{\chi};x^j,q) = \\ \label{collection} =(\xi_1 \lambda_1(q) + \xi_2 \lambda_2(q) - \xi_1 \lambda_1(p) - \xi_2 \lambda_2(p)) K(\widetilde{\chi}; p,q) \end{gather} which is valid for all distinct points $p,q$ in $X$ which are disjoint from $x^1, \dots, x^m$. To prove this ``collection formula'' \eqref{collection}, consider each side as a function of $p$ with $q$ fixed. Since $h^0(\widetilde{\chi} \otimes \Delta) = 0$, it suffices to show that the local principal part in the Laurent series expansion at each pole of each side matches with the local principal part of the other side. One can check that the only possible poles are all simple and occur at $x^1, \dots, x^m$ with residue of each side at $x^i$ equal to the common value $-(\xi_1c_{i1} + \xi_2 c_{i2}) dt^i(x^i) K(\widetilde{\chi}; x^i,q)$. Immediate consequences of the identity \eqref{collection} which are important for our context here are: \begin{align} \notag \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) & K^{{\bold x}, \boldsymbol \lambda}(\lambda^i) K(\widetilde{\chi};x^j,p) = \\ \label{consequence1} = & \underset{i^\prime}{\text{diag.}} \{ (\xi_1 \lambda_1(p) + \xi_2 \lambda_2(p) - \xi_1 \lambda^{i^\prime}_1 - \xi_2 \lambda^{i^\prime}_2)I_{t_{i^\prime}}\} \cdot K^{{\bold x},\boldsymbol\lambda}(p), \end{align} and, if $(i^\prime, j^\prime)$ is a pair of indices for which $\lambda^{i^\prime} \ne \mu^{j^\prime}$ then the $(i^\prime, j^\prime)$-matrix entry of $\sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K^{{\bold x}, \boldsymbol \lambda}(x^j) K_{\boldsymbol \mu,{\bold u}}(x^j)$ is given by \begin{gather} \notag [ \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K^{{\bold x}, \boldsymbol\lambda}(x^j) K_{\boldsymbol \mu,{\bold u}}(x^j)]_{i^\prime, j^\prime} = \\ \notag = (\xi_1 \mu^{j^\prime}_1 + \xi_2 \mu^{j^\prime}_2 - \xi_1 \lambda^{i^\prime}_1 - \xi_2 \lambda^{i^\prime}_2) {\bold x}_{i^\prime} K(\widetilde{\chi}; \lambda^{i^\prime}, \mu^{j^\prime}) {\bold u}_{j^\prime} \\ \notag = - (\xi_1 \mu^{j^\prime}_1 + \xi_2 \mu^{j^\prime}_2 - \xi_1 \lambda^{i^\prime}_1 - \xi_2 \lambda^{i^\prime}_2) \Gamma_{i^\prime, j^\prime}. \end{gather} Hence in matrix form we have \begin{gather} \notag \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K^{{\bold x}, \boldsymbol \lambda}(x^j) K_{\boldsymbol\mu,{\bold u}}(x^j) = \\ \label{consequence2} = \underset{i^\prime}{\text{diag.}} \{(\xi_1 \lambda^{i^\prime}_1 + \xi_2 \lambda^{i^\prime}_2)I_{t_{i^\prime}} \} \cdot \Gamma - \Gamma \cdot \underset{j^\prime}{\text{diag.}}\{(\xi_1 \mu^{j^\prime}_1 + \xi_2 \mu^{j^\prime}_2) I_{s_{j^\prime}} \}. \end{gather} In the case where $p=q$, the collection formula \eqref{collection} takes the limiting form \begin{gather} \notag \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K(\widetilde{\chi};p,x^j) K(\widetilde{\chi};x^j,p) = \\ \label{degencollection} -(\xi_1 \lambda^\prime_1(p) + \xi_2 \lambda^\prime_2(p)) \ dt(p) \end{gather} where ${}^\prime = \dfrac{d}{dt}$ where $t$ is a local coordinate centered at $p$. An application of this degenerate collection formula \eqref{degencollection} gives, for $\lambda^{i^\prime} = \mu^{j^\prime} =: \xi^{i^\prime j^\prime}$, \begin{gather} \notag [ \sum_{j=1}^m (\xi_1 c_{j1} + \xi_2 c_{j2}) K^{{\bold x}, \boldsymbol \lambda}(x^j) K_{\boldsymbol \mu,{\bold u}}(x^j)]_{i^\prime, j^\prime} = \\ \notag = -(\xi_1 \lambda^\prime_1(\xi^{i^\prime j^\prime}) + \xi_2 \lambda^\prime_2(\xi^{i^\prime j^\prime}) )\ dt(\xi^{i^\prime j^\prime}) \ {\bold x}^T_{i^\prime} {\bold u}_{j^\prime} = 0 \end{gather} where we used the compatibility condition \eqref{comp} for the last step. We conclude that \eqref{consequence2} continues to be valid even in the case where $\lambda^{i^\prime} = \mu^{j^\prime}$. Making the substitutions \eqref{consequence1} and \eqref{consequence2} in \eqref{Sside2}, we obtain \begin{gather} \notag [S(p) \beta^{-1} u^{\times \prime}(p)]_i = [K(\widetilde{\chi};x^i,p)+ K_{\boldsymbol \mu,{\bold u}}(x^i) \Gamma^{-1} K^{{\bold x},\boldsymbol \lambda}(p)] T(p) - \\ \notag - K_{\boldsymbol\mu, {\bold u}}(x^i) \cdot \underset{i'}{\text{diag.}} \{( \xi_1 \lambda_1(p)+\xi_2 \lambda_2(p) - \xi_1 \mu^{i^\prime}_1 -\xi_2 \mu^{i^\prime}_2)^{-1} I_{s_{i^\prime}} \} \cdot \\ \notag \cdot \Gamma^{-1} \cdot \left( \underset{i^\prime}{\text{diag.}}\{(\xi_1 \lambda_1(p) + \xi_2\lambda_2(p) - \xi_1 \lambda^{i^\prime}_1 -\xi_2 \lambda^{i^\prime}_2) I_{t_{i^\prime}}\} K^{{\bold x}, \boldsymbol \lambda}(p) + \right. \\ \notag + \left[ \underset{i^\prime}{\text{diag.}}\{ (\xi_1 \lambda^{i^\prime}_1 + \xi_2 \lambda^{i^\prime}_2) I_{t_{i^\prime}} \} \Gamma \right. - \Gamma \underset{j^\prime}{\text{diag.}} \{ ( \xi_1 \mu^{j^\prime}_1 + \xi_2 \mu^{j^\prime}_2) I_{s_{j^\prime}}\} \left. \right] \Gamma^{-1} K^{{\bold x}, \boldsymbol\lambda}(p) \left. \right) = \\ \notag = K(\widetilde{\chi};x^i,p) T(p) + K_{\boldsymbol\mu, {\bold u}}(x^i) \Gamma^{-1} K^{{\bold x}, \boldsymbol\lambda}(p) T(p) - \\ \notag - K_{\boldsymbol \mu, {\bold u}}(x^i) \cdot \underset{i^\prime}{\text{diag.}} \{ (\xi_1 \lambda_1(p) + \xi_2 \lambda_2(p) - \xi_1 \mu^{i^\prime}_1 - \xi_2 \mu^{i^\prime}_2)^{-1} I_{s_{i^\prime}} \} \cdot \\ \notag \cdot \left( (\xi_1 \lambda_1(p) + \xi_2 \lambda_2(p)) \Gamma^{-1} K^{{\bold x}, \boldsymbol \lambda}(p) \right. - \Gamma^{-1} \underset{i^\prime}{\text{diag.}} \{( \xi_1 \lambda^{i^\prime}_1 + \xi_2 \lambda^{i^\prime}_2) I_{t_{i^\prime}} \} K^{{\bold x}, \boldsymbol \lambda}(p) + \\ \notag + \Gamma^{-1} \underset{i^\prime}{\text{diag.}} \{ (\xi_1 \lambda^{i^\prime}_1 + \xi_2 \lambda^{i^\prime}_2) I_{t_{i^\prime}} \} K^{{\bold x}, \boldsymbol \lambda}(p) \left. - \underset{j^\prime}{\text{diag.}} \{(\xi_1 \mu^{j^\prime}_1 + \xi_2 \mu^{j^\prime}_2) I_{s_{j^\prime}} \} \Gamma^{-1} K^{{\bold x}, \boldsymbol\lambda}(p) \right) \\ \notag = K(\widetilde{\chi}; x^i,p) T(p) \end{gather} and the intertwining relation \eqref{intertwining} follows. \end{pf} {\bf REMARK:} Note that the results of this section in principle give a means of computing explicitly the unknown input bundle $\chi$ appearing in Theorem \ref{T:bvabsint} in Section \ref{S:absint}. Indeed, given a data set $\omega$ for an (ABSINT) problem, we can convert it to a data set $\omega_0$ for a (CONINT) problem, as explained in the discussion preceding the statement of Theorem \ref{T:comparison}. In the context of the (CONINT) problem, Theorem \ref{T:conint} solves the problem of identifying the input bundle. Namely, the input bundle $E$ is defined to be $E=\ker (z_1\sigma_2 - z_2 \sigma_1 + \gamma)$ where $\sigma_1, \sigma_2 $ and $\widetilde{\gamma}$ are given by \eqref{pencilcoef} and $\gamma$ is given by \eqref{gamma}. The vector bundle $\chi$ is then determined up to biholomorphic equivalence by the condition \[ \chi \cong E \otimes {\cal O}(1) \otimes \Delta. \] Moreover, as explained in \cite{hip}, it is possible to construct a matrix of normalized sections $u^\times$ for $E$ from working with the minors of the matrix pencil $z_1 \sigma_2 - z_2 \sigma_1 + \gamma$. Such a matrix of normalized sections $u^\times$ in turn implements concretely the biholomorphic equivalence between $\chi$ and $E \otimes {\cal O}(1) \otimes \Delta$.
1998-12-07T01:29:44
9712
alg-geom/9712031
en
https://arxiv.org/abs/alg-geom/9712031
[ "alg-geom", "math.AG" ]
alg-geom/9712031
James S. Milne
J.S. Milne
Descent for Shimura Varieties
6 pages
Michigan Math. J. 46 (1999), no. 1, 203--208
null
null
null
We verify that the descent maps provided by Langlands's Conjugacy Conjecture do satisfy the continuity condition necessary for them to be effective. Thus Langlands's conjecture does imply the existence of canonical models. This replaces an earlier version of the paper --- the proof in this version is simpler, and the exposition more detailed.
[ { "version": "v1", "created": "Tue, 30 Dec 1997 00:59:16 GMT" }, { "version": "v2", "created": "Mon, 7 Dec 1998 00:29:43 GMT" } ]
2021-01-19T00:00:00
[ [ "Milne", "J. S.", "" ] ]
alg-geom
\subsubsection{Notations and Conventions} A variety over a field $k$ is a geometrically reduced scheme of finite type over $\Spec k$ (\emph{not} necessarily irreducible). For a variety $V$ over a field $k$ and a homomorphism $\sigma \colon k\rightarrow k^{\prime }$, $% \sigma V$ is the variety over $k^{\prime }$ obtained by base change. The ring of finite ad\`{e}les for $\mathbb{Q}$ is denoted by $\mathbb{A}_{f}{}$. \section{Descent of Varieties.} In this section, $\Omega$ is an algebraically closed field of characteristic zero. For a field $L\subset\Omega$, $A(\Omega/L)$ denotes the group of automorphisms of $\Omega$ fixing the elements of $L$. Let $V$ be a variety over $\Omega$, and let $k$ be a subfield of $\Omega$. A family $(f_{\sigma})_{\sigma\in A(\Omega/k)}$ of isomorphisms $f_{\sigma }\colon\sigma V\rightarrow V$ will be called a \emph{descent system\/} if $% f_{\sigma\tau}=f_{\sigma}\circ\sigma f_{\tau}$ for all $\sigma,\tau\in A(\Omega/k)$. We say that a model $(V_{0},f\colon V_{0,\Omega}\rightarrow V)$ of $V$ over $k$ \emph{splits} $(f_{\sigma})$ if $f_{\sigma}=f\circ(\sigma f)^{-1}$ for all $\sigma\in A(\Omega/k)$, and that a descent system is \emph{% effective\/} if it is split by some model over $k$. The next theorem restates results of Weil 1956. \begin{theorem} Assume that $\Omega $ has infinite transcendence degree over $k$. A descent system $(f_{\sigma })$$_{\sigma \in A(\Omega /k)}$ on a quasiprojective variety $V$ over $\Omega $ is effective if, for some subfield $L$ of $\Omega $ finitely generated over $k$, the descent system $(f_{\sigma })_{\sigma \in A(\Omega /L)}$ is effective. \end{theorem} \begin{proof} Let $k^{\prime}$ be the algebraic closure of $k$ in $L$ --- then $k^{\prime}$ is a finite extension of $k$ and $L$ is a regular extension of $k^{\prime}$. Let $(V_{t^{\prime}},f^{\prime}\colon V_{t^{\prime},\Omega}\rightarrow V)$ be the model of $V$ over $L$ splitting $(f_{\sigma})_{\sigma\in A(\Omega/L)}$% . Let $t\colon L\rightarrow k_{t}$ be a $k^{\prime}$-isomorphism from $L$ onto a subfield $k_{t}$ of $\Omega$ linearly disjoint from $L$ over $% k^{\prime}$, and let $V_{t}=V_{t^{\prime}}\otimes_{L,t}k_{t}$. Zorn's Lemma allows us to extend $t$ to an automorphism $\tau$ of $\Omega$ over $% k^{\prime}$. The isomorphism \begin{equation*} f_{t,t^{\prime}}\colon V_{t^{\prime},\Omega}\overset{f^{\prime}}{\rightarrow }V\overset{f_{\tau}^{-1}}{\rightarrow}\tau V\overset{(\tau f^{\prime})^{-1}}{% \rightarrow}V_{t,\Omega} \end{equation*} is independent of the choice of $\tau$, is defined over $L\cdot k_{t}$, and satisfies the hypothesis of Weil 1956, Theorem 6, which gives a model $(W,f)$ of $V$ over $k^{\prime}$ splitting $(f_{\sigma})_{\sigma\in A(\Omega /k^{\prime})}$. For $\sigma \in A(\Omega /k)$, $g_{\sigma }\overset{\text{df}}{=}f_{\sigma }\circ \sigma f\colon \sigma W_{\Omega }\rightarrow V$ depends only on $% \sigma |k^{\prime }$. For $k$-homomorphisms $\sigma ,\tau \colon k^{\prime }\rightarrow \Omega $, define $f_{\tau ,\sigma }=g_{\tau }^{-1}\circ g_{\sigma }\colon \sigma W\rightarrow \tau W$. Then $f_{\tau ,\sigma }$ is defined over the Galois closure of $k^{\prime }$ in $\Omega $ and the family $(f_{\tau ,\sigma })$ satisfies the hypotheses of Weil 1956, Theorem 3, which gives a model of $V$ over $k$ splitting $(f_{\sigma })_{\sigma \in A(\Omega /k)}.$ \end{proof} \begin{corollary} Let $\Omega$, $k$, and $V$ be as in the theorem, and let $% (f_{\sigma})_{\sigma\in A(\Omega/k)}$ be a descent system on $V$. If there is a finite set $\Sigma$ of points in $V(\Omega)$ such that \begin{enumerate} \item any automorphism of $V$ fixing all $P\in \Sigma $ is the identity map, and \item there exists a subfield $L$ of $\Omega$ finitely generated over $k$ such that $f_{\sigma}(\sigma P)=P$ for all $P\in\Sigma$ and all $\sigma\in A(\Omega/L)$, \end{enumerate} then $(f_{\sigma})_{\sigma\in A(\Omega/k)}$ is effective. \end{corollary} \begin{proof} After possibly replacing the $L$ in (b) with a larger finitely generated extension of $k$, we may suppose that $V$ has a model $(W,f)$ over $L$ for which the points of $\Sigma$ are rational, i.e., such that for each $% P\in\Sigma$, $P=f(P^{\prime})$ for some $P^{\prime}\in W(L)$. Now, for each $% \sigma\in A(\Omega/L)$, $f_{\sigma}$ and $f\circ\sigma f^{-1}$ are both isomorphisms $\sigma V\rightarrow V$ sending $\sigma P$ to $P$, and so hypothesis (a) implies they are equal. Hence $(f_{\sigma})_{\sigma\in A(\Omega/L)}$ is effective, and the theorem applies. \end{proof} \begin{remark} \begin{enumerate} \item It is easy to construct noneffective descent systems. For example, take $\Omega$ to be the algebraic closure of $k$, and let $V$ be a variety $% k $. A one-cocycle $h\colon A(\Omega/k)\rightarrow\Aut(V_{\Omega})$ can be regarded as a descent system --- identify $h_{\sigma}$ with a map $\sigma V_{\Omega }=V_{\Omega}\rightarrow V_{\Omega}$. If $h$ is not continuous, for example, if it is a homomorphism into $\Aut(V)$ whose kernel is not open, then the descent system will not be effective. \item An example (Dieudonn\'{e} 1964, p 131) shows that the hypothesis that $V$ be quasiprojective in (1.1) is necessary unless the model $V_{0}$ is allowed to be an algebraic space in the sense of M. Artin. \item Theorem 1.1 and its corollary replace Lemma 3.23 of Milne 1994, which omits the continuity conditions. \end{enumerate} \end{remark} \subsection{Application to moduli problems.} Suppose we have a contravariant functor $\mathcal{M}$ from the category of algebraic varieties over $\Omega $ to the category of sets, and equivalence relations $\sim $ on each of the sets $\mathcal{M}(T)$ compatible with morphisms. The pair $(\mathcal{M},\sim )$ is then called a \emph{moduli problem} over $\Omega $. A $t\in T(\Omega )$ defines a map \begin{equation*} m\mapsto m_{t}\overset{\text{df}}{=}t^{\ast }m\colon \mathcal{M}% (T)\rightarrow \mathcal{M}(\Omega ). \end{equation*} \noindent A \emph{solution to the moduli problem }is a variety $V$ over $% \Omega$ together with an isomorphism $\alpha\colon\mathcal{M}(\Omega )/\!\!\sim\rightarrow V(\Omega)$ such that: \begin{enumerate} \item for all varieties $T$ over $\Omega$ and all $m\in\mathcal{M}(T)$, the map $t\mapsto\alpha(m_{t})\colon T(\Omega)\rightarrow V(\Omega)$ is regular (i.e., defined by a morphism $T\rightarrow V$ of $\Omega$-varieties); \item for any variety $W$ over $\Omega$ and map $\beta\colon\mathcal{M}% (\Omega)/\!\!\sim\rightarrow W(\Omega)$ satisfying the condition (a), the map $P\mapsto\beta(\alpha^{-1}(P))\colon V(\Omega)\rightarrow W(\Omega)$ is regular. \end{enumerate} \noindent Clearly, a solution to a moduli problem is unique up to a unique isomorphism when it exists. Let $(\mathcal{M},\sim )$ be a moduli problem over $\Omega $, and let $k$ be a subfield $\Omega $. For $\sigma \in A(\Omega /k)$, define $^{\sigma }% \mathcal{M}$ to be the functor sending an $\Omega $-variety $T$ to $\mathcal{% M}(\sigma ^{-1}T)$. We say that $(\mathcal{M},\sim )$ is \emph{rational over} $k$ if there is given a family $(g_{\sigma })_{\sigma \in A(\Omega /k)}$ of isomorphisms $g_{\sigma }\colon ^{\sigma }\mathcal{M}\rightarrow \mathcal{M}$, compatible with $\sim$, such that $g_{\sigma \tau }=g_{\sigma }\circ \sigma g_{\tau }$ for all $% \sigma $, $\tau \in A(\Omega /k)$ --- the last equation means that $% g_{\sigma \tau }(T)=g_{\sigma }(T)\circ g_{\tau }(\sigma ^{-1}T)$ for all varieties $T$. Note that $^{\sigma }\mathcal{M}(\Omega )=\mathcal{M}(\Omega ) $, and that the rule $\sigma m=g_{\sigma }(m)$ defines an action of $% A(\Omega /k)$ on $\mathcal{M}(\Omega )$. A \emph{solution to a moduli problem% } $(\mathcal{M},\sim ,(g_{\sigma }))$ \emph{rational over} $k$ is a variety $% V_{0}$ over $k$ together with an isomorphism $\alpha \colon \mathcal{M}% (\Omega )/\!\!\sim \rightarrow V_{0}(\Omega )$ such that \begin{enumerate} \item $(V_{0,\Omega},\alpha)$ is a solution to the moduli problem $(% \mathcal{M},\sim)$ over $\Omega$, and \item $\alpha $ commutes with the actions of $A(\Omega /k)$ on $\mathcal{M}% (\Omega )$ and $V_{0}(\Omega )$. \end{enumerate} \noindent Again, $(V_{0},\alpha)$ is uniquely determined up to a unique isomorphism (over $k$) when it exists. \begin{theorem} Assume that $\Omega $ has infinite transcendence degree over $k$. Let $(% \mathcal{M},\sim ,(g_{\sigma }))$ be a moduli problem rational over $k$ for which $(\mathcal{M},\sim )$ has a solution $(V,\alpha )$ over $\Omega $. Then $(\mathcal{M},\sim ,(g_{\sigma }))$ has a solution over $k$ if there exists a finite subset $\Sigma \subset \mathcal{M}(\Omega )$ such that \begin{enumerate} \item any automorphism of $V$ fixing $\alpha(P)$ for all $P\in\Sigma$ is the identity map, and \item there exists a subfield $L$ of $\Omega $ finitely generated over $k$ such that $g_{\sigma }(P)\sim P$ for all $P\in \Sigma $ and all $\sigma \in A(\Omega /L)$. \end{enumerate} \end{theorem} \begin{proof} The family $(g_{\sigma})$ defines a descent system on $V$, which Corollary 1.2 shows to be effective. \end{proof} \section{Descent of Shimura Varieties.} In this section, all fields will be subfields of $\mathbb{C}$. For a subfield $E$ of $\mathbb{C}$, $E^{\text{ab}}$ denotes the composite of all the finite abelian extensions of $E$ in $\mathbb{C}$. Let $(G,X)$ be a pair satisfying the axioms (2.1.1.1--2.1.1.3) of Deligne 1979 to define a Shimura variety, and let $\Sh(G,X)$ be the corresponding Shimura variety over $\mathbb{C}$. We regard $\Sh(G,X)$ as a pro-variety endowed with a continuous action of $G(\mathbb{A}_{f})$ --- in particular (ibid. 2.7.1) this means that $\Sh(G,X)$ is a projective system of varieties $(\Sh_{K}(G,X))$ indexed by the compact open subgroups $K$ of $G(\mathbb{A}% _{f})$. Let $[x,a]$ $=([x,a]_{K})_{K}$ denote the point in $\Sh(G,X)(\mathbb{% C})$ defined by a pair $(x,a)\in X\times G(\mathbb{A}_{f})$, and let $E(G,X)$ be the reflex field of $(G,X)$. For a special point $x\in X$, let $% E(x)\supset E(G,X)$ be the reflex field for $x$ and let \begin{equation*} r_{x}\colon \Gal(E(x)^{{ \text{ab} }}/E(x))\rightarrow T(\mathbb{A}_{f})/T(\mathbb{Q}% )^{-} \end{equation*} be the reciprocity map defined in Milne 1992, p164 (inverse to that in Deligne 1979, 2.2.3). Here $T$ is a subtorus of $G$ such that $\text{Im}% (h_{x})\subset T_{\mathbb{R}}$ and $T(\mathbb{Q})^{-}$ is the closure of $T(% \mathbb{Q})$ in $T(\mathbb{A}{}_{f})$. A\emph{\ model\/} of $\Sh(G,X)$ over a field $k$ is a pro-variety $S$ over $k$ endowed with an action of $G(% \mathbb{A}_{f})$ and a $G(\mathbb{A}_{f})$-equivariant isomorphism $f\colon S_{\mathbb{C}}\rightarrow \Sh(G,X)$. A model of $\Sh(G,X)$ over $E(G,X)$ is \emph{canonical } if, for each special point $x\in X$ and $a\in G(\mathbb{A}% _{f})$, $[x,a]$ is rational over $E(x)^{\text{ab}}$ and $\sigma \in \Gal% (E(x)^{\text{ab}}/E(x))$ acts on $[x,a]$ according\footnote{% More precisely, the condition for $(S,f)$ to be canonical is the following: if $P\in S(\mathbb{C})$ corresponds under $f$ to $[x,a]$, then $\sigma P$ corresponds under $f$ to $[x,r_{x}(\sigma )\cdot a].$} to the rule: \begin{equation*} \sigma \lbrack x,a]=[x,r_{x}(\sigma )\cdot a]. \end{equation*} Let $k$ be a field containing $E(G,X)$. A \emph{descent system} for $\Sh(G,X) $ over $k$ is a family of isomorphisms \begin{equation*} (f_{\sigma }\colon \sigma \Sh(G,X)\rightarrow \Sh(G,X))_{\sigma \in A(% \mathbb{C}/k)} \end{equation*} such that, \begin{enumerate} \item for all $\sigma,\tau\in A(\mathbb{C}/k)$, $f_{\sigma\tau}=f_{\sigma }\circ\sigma f_{\tau}$, and \item for all $\sigma\in A(\mathbb{C}/k) $, $f_{\sigma}$ is equivariant for the actions of $G(\mathbb{A}_{f})$ on $\Sh(G,X)$ and $\sigma\Sh(G,X)$. \end{enumerate} \noindent We say that a model $(S,f)$ of $\Sh(G,X)$ over $k$ \emph{splits }$% (f_{\sigma})$ if $f_{\sigma}=f\circ\sigma f^{-1}$ for all $\sigma\in A(\Omega/k)$, and that a descent system if \emph{effective} if it is split by some model over $k$. A descent system $(f_{\sigma})$ for $\Sh(G,X)$ over $% E(G,X)$ is \emph{canonical} if \begin{equation*} f_{\sigma}(\sigma\lbrack x,a])=[x,r_{x}(\sigma|E(x)^{{ \text{ab} }})\cdot a] \end{equation*} \noindent whenever $x$ is a special point of $X$, $\sigma\in A(\mathbb{C}% /E(x))$, and $a\in G(\mathbb{A}_{f})$. \begin{remark} \begin{enumerate} \item For a Shimura variety $\Sh(G,X)$, there exists at most one canonical descent system for $\Sh(G,X)$ over $E(G,X)$. (Apply Deligne 1971, 5.1, 5.2.) \item Let $(S,f)$ be a model of $\Sh(G,X)$ over $E(G,X)$, and let $% f_{\sigma }=f\circ(\sigma f)^{-1}$. Then $(f_{\sigma})_{\sigma\in A(\mathbb{C% }/k)}$ is a descent system for $\Sh(G,X)$, and $(f_{\sigma})$ is canonical if and only if $(S,f)$ is canonical. \item Suppose $\Sh(G,X)$ has a canonical descent system $% (f_{\sigma})_{\sigma\in A(\mathbb{C}/E(G,X))}$; then $\Sh(G,X)$ has a canonical model if and only if $(f_{\sigma})$ is effective. (Follows from (a) and (b).) \item A descent system $(f_{\sigma})_{\sigma\in A(\mathbb{C}/k)}$ on $\Sh% (G,X)$ defines for each compact open subgroup $K$ of $G(\mathbb{A}_{f})$ a descent system $(f_{\sigma,K})_{\sigma\in A(\mathbb{C}/k)}$ on the variety $% \Sh_{K}(G,X)$ (in the sense of \S1). If $(f_{\sigma})$ is effective, then so also is $(f_{\sigma,K})$ for all $K$; conversely, if $(f_{\sigma,K})_{\sigma% \in A(\mathbb{C}/k)}$ is effective (in the sense of \S1) for all sufficiently small $K$, then $(f_{\sigma})_{\sigma\in A(\mathbb{C}/k)}$ is effective (in the sense of this section). \end{enumerate} \end{remark} \begin{lemma} The automorphism group of the quotient of a bounded symmetric domain by a neat arithmetic group is finite. \end{lemma} \begin{proof} According to Mumford 1977, Proposition 4.2, such a quotient is an algebraic variety of logarithmic general type, which implies that its automorphism group is finite (Iitaka 1982, 11.12). Alternatively, one sees easily that the automorphism group of the quotient of a bounded symmetric domain $D$ by a neat arithmetic subgroup $\Gamma$ is $% N/\Gamma$ where $N$ is the normalizer of $\Gamma$ in $\Aut(D)$. Now $N$ is countable and closed (because $\Gamma$ is closed), and hence is discrete (Baire category theorem). Because the quotient of $\Aut(D)$ by $\Gamma$ has finite measure, this implies that $\Gamma$ has finite index in $N$. Cf. Margulis 1991, II 6.3. \end{proof} \begin{theorem} Every canonical descent system on a Shimura variety is effective. \end{theorem} \begin{proof} Let $(f_{\sigma})_{\sigma\in A(\mathbb{C}/E(G,X))}$ be a canonical descent system for the Shimura variety $\Sh(G,X)$. Let $K$ be a compact open subgroup of $G(\mathbb{A}_{f})$, chosen so small that the connected components of $\Sh_{K}(G,X)$ are quotients of bounded symmetric domains by \emph{neat} arithmetic groups. Let $x$ be a special point of $X$. According to Deligne 1971, 5.2, the set $\Sigma=\{[x,a]_{K}\mid a\in G(\mathbb{A}% _{f})\}$ is Zariski dense in $\Sh_{K}(G,X)$. Because the automorphism group of $\Sh _{K}(G,X)$ is finite, there is a finite subset $\Sigma_{f}$ of $\Sigma$ such that any automorphism $\alpha$ of $\Sh_{K}(G,X)$ fixing each $P\in\Sigma_{f}$ is the identity map. The rule \begin{equation*} \sigma \ast \lbrack x,a]_{K}=[x,r_{x}(\sigma )\cdot a]_{K} \end{equation*} defines an action of $\Gal(E(x)^{\text{ab}}/E(x))$ on $\Sigma $ for which the stabilizer of each point of $\Sigma $ is open. Therefore, there exists a finite abelian extension $L$ of $E(x)$ such that $\sigma \ast P=P$ for all $% P\in \Sigma _{f}$ and all $\sigma \in \Gal(E(x)^{\text{ab}}/L)$. Now, because $(f_{\sigma})_{\sigma\in A(\Omega/E(G,X))}$ is canonical, $% f_{\sigma,K}(\sigma P)=P$ for all $P\in\Sigma_{f}$ and all $\sigma\in A(% \mathbb{C}{}/L)$, and we may apply Corollary 1.2 to conclude that $% (f_{\sigma,K})_{\sigma\in A(\mathbb{C}/E(G,X))}$ is effective. As this holds for all sufficiently small $K$, $(f_{\sigma})_{\sigma\in A(\mathbb{C}% /E(G,X))}$ is effective. \end{proof} \begin{remark} \begin{enumerate} \item If Langlands's Conjugacy Conjecture (Langlands 1979, p232, 233) is true for a Shimura variety $\Sh(G,X)$, then $\Sh(G,X)$ has a canonical descent system (ibid. \S 6; also Milne and Shih 1982, \S 7). \item Langlands's Conjugacy Conjecture is true for all Shimura varieties (Milne 1983). Hence canonical models exist for all Shimura varieties. \end{enumerate} \end{remark} Another proof, based on different ideas, that the descent maps given by Langlands's conjecture are effective can be found in Moonen 1998. (I thank the referee for this reference.) \section*{References} Deligne, P., Travaux de Shimura, in S\'{e}minaire Bourbaki, 23\`{e}me ann\'{e}e (1970/71), Exp. No. 389, 123--165. Lecture Notes in Math., 244, Springer, Berlin, 1971. Deligne, P., Vari\'{e}t\'{e}s de Shimura: Interpr\'{e}tation modulaire, et techniques de construction de mod\`{e}les canoniques, Proc. Symp. Pure Math. 33 Part 2, pp. 247--290, 1979. Dieudonn\'{e}, J., Fondements de la G\'{e}om\'{e}trie Alg\'{e}brique Moderne, Presse de l'Universit\'{e} de Montr\'{e}al, 1964. Iitaka, S., Algebraic Geometry, Springer, Heidelberg, 1982. Langlands, R., Automorphic representations, Shimura varieties, and motives, Ein M\"{a}rchen, Proc. Symp. Pure Math. 33 Part 2, pp. 205--246, 1979. Margulis, G.A., Discrete subgroups of semisimple Lie groups, Springer, Heidelberg, 1991. Milne, J.S., The action of an automorphism of $\mathbb{C}$ on a Shimura variety and its special points, Prog. in Math., vol. 35, Birkh\"{a}user, Boston, pp. 239--265, 1983. Milne, J.S., The points on a Shimura variety modulo a prime of good reduction, in The Zeta Function of Picard Modular Surfaces (Langlands and Ramakrishnan, eds), Les Publications CRM, Montr\'{e}al, pp. 153--255, 1992. Milne, J. S., Shimura varieties and motives, in Motives (Seattle, WA, 1991), 447--523, Proc. Sympos. Pure Math., Part 2, Amer. Math. Soc., Providence, RI, 1994. Milne, J.S. and Shih, K-y., Conjugates of Shimura varieties, in Hodge Cycles, Motives, and Shimura Varieties, Lecture Notes in Math., vol. 900, Springer, Heidelberg, pp. 280--356, 1982. Moonen, B., Models of Shimura varieties in mixed characteristics, in Galois Representations in Arithmetic Geometry (A.J. Scholl and R.L Taylor, editors), Cambridge University Press, pp. 267--350, 1998. Mumford, D., Hirzebruch's proportionality theorem in the non-compact case, Invent. Math. 42, 239--272, 1977. Weil, A., The field of definition of a variety, Amer. J. Math 78, pp. 509--524, 1956. \end{document}
1997-12-16T01:39:01
9712
alg-geom/9712009
en
https://arxiv.org/abs/alg-geom/9712009
[ "alg-geom", "hep-th", "math.AG", "math.DG", "q-alg" ]
alg-geom/9712009
Andrei Okounkov
Spencer Bloch and Andrei Okounkov
The Character of the Infinite Wedge Representation
Latex, 57 pages; typos in bibliography corrected
null
null
null
null
We study the character of the infinite wedge projective representation of the algebra of differential operators on the circle. We prove quasi-modularity of this character and also compute certain generating functions for traces of differential operators which we call correlation functions. These correlation functions are sums of determinants built from genus 1 theta functions and their derivatives.
[ { "version": "v1", "created": "Tue, 9 Dec 1997 20:55:53 GMT" }, { "version": "v2", "created": "Tue, 16 Dec 1997 00:39:01 GMT" } ]
2007-05-23T00:00:00
[ [ "Bloch", "Spencer", "" ], [ "Okounkov", "Andrei", "" ] ]
alg-geom
\section{Introduction}\label{sec:intro} The purpose of this paper is to study the character associated to a certain basic projective representation, the infinite wedge representation \cite{KP}, of the Lie algebra ${\mathcal D}$ of differential operators on the circle. Concretely, we write \begin{equation}\label{01} D_n := \left(t\frac{d}{dt}\right)^n,\quad n\ge 0 \end{equation} acting as differential operators on the ring $R={\Bbb C}[t,t^{-1}]$ of Laurent polynomials. (Here $D_0 = 1$ is the identity operator. We occasionally write $D=D_1$.) ${\mathcal D}$ has a vector space basis $t^mD_n,\ m\in {\Bbb Z},\ n\ge 0$. The Witt algebra of derivations of $R,\ {\mathcal W}=\text{\rm Der}(R)\subset {\mathcal D}$ is spanned by $t^mD,\ m\in{\Bbb Z}$. These algebras are graded, with $t^mD_n$ having weight $m$. One has a central extension \cite{KR} $0\to {\Bbb C}\cdot c \to {\widehat{\mathcal D}} \to {\mathcal D} \to 0$ inducing the Virasoro central extension $\widehat{{\mathcal W}}$ of ${\mathcal W}$, and these central extensions are also graded, with $c$ having degree $0$. Certain graded highest weight representations of $\widehat{{\mathcal W}}$ arise in connection with conformal field theory and have been of considerable interest to physicists. These representations have the property that for a suitable choice $\widetilde{D}\in \widehat{{\mathcal W}}$ lifting $D$, the character $$\text{\rm Trace}(q^{\widetilde{D}}) $$ is well-defined (i.e. the action of $\widetilde{D}$ is semi-simple with finite eigenspaces) and is the $q$-expansion of a modular form \cite{ZHU}. Notice that a different choice of lifting of $D$ to $\widehat{{\mathcal W}}$ multiplies the character by $q^a$ for some constant $a$. In particular, there is at most one lifting for which the character is modular. For representations of ${\widehat{\mathcal D}}$ one may consider the Abelian subalgebra \begin{equation}\label{02} {\mathcal H} := {\Bbb C} D_0\oplus{\Bbb C} D_1\oplus{\Bbb C} D_2\oplus\ldots\subset {\mathcal D}. \end{equation} Suppose the action of ${\mathcal H}$ is semi-simple with finite dimensional simultaneous eigenspaces. If we choose liftings $\widetilde{D}_n \in {\widehat{\mathcal D}}$ of $D_n$ we may define the character \begin{equation}\label{03} \Omega(q_0,q_1,\ldots) := \text{\rm Trace}\left(q_0^{\widetilde{D}_0}q_1^{\widetilde{D}_1}q_2^{\widetilde{D}_2}\cdots \right)\,. \end{equation} It is also frequently convenient to write $q_r = e^{2\pi i\tau_r}$ so the character becomes (with abuse of notation) \begin{equation}\Omega(\tau_0,\tau_1,\ldots) := \text{\rm Trace}\left( \exp\left(2\pi i\sum_{r\ge 0}\widetilde{D}_r\tau_r\right)\right)\,. \end{equation} Assuming the eigenspaces of $D_1$ are themselves finite dimensional, we may specialize the $\tau_r \mapsto 0$ for $r\ne 1$ and develop $\Omega(\tau_0,\tau_1,\ldots)$ in a formal Taylor series expansion in $\tau_0,\tau_2,\ldots$ with coefficients functions in $\tau_1$: \begin{equation}\label{05} \Omega(\tau_0,\tau_1,\ldots) = \sum_A\omega_A(\tau_1)\tau^A/A!\,. \end{equation} Here $A=(a_0,a_2,a_3,\ldots)$ with almost all $a_j=0$, and $\tau^A/A!$ is multi-index notation. Finally, $\widetilde{D}_0$ is central in ${\widehat{\mathcal D}}$ so its eigenspaces are stable under the ${\widehat{\mathcal D}}$ action. We write \begin{equation}\label{06}V(q_1,q_2,\ldots) = \text{\rm Coeff. of $q_0^0$ in }\Omega(q_0,q_1,\ldots) \end{equation} for the character of the $\widetilde{D}_0=0$ eigenspace. Again, by abuse of notation, we also write $V(\tau_1,\tau_2,\ldots)$ and we expand in a series \begin{equation}\label{07} V = \sum_{B=(b_2,b_3,\ldots)} v_B(\tau_1)\tau^B/B!\,, \end{equation} where, by construction, \begin{equation}\label{07a} \frac{v_B(\tau_1)}{(2\pi i)^{b_2+b_3+\dots}}= \text{\rm Trace}\big|_{\widetilde{D}_0=0} \left( q_1^{\widetilde{D}_1} \prod_{k=2}^\infty \left(\widetilde{D}_k\right)^{b_k} \right)\,. \end{equation} Here the trace is taken in the $\widetilde{D}_0=0$ subspace. In the case of the infinite wedge representation, these characters have the following shape \begin{multline}\label{08} \Omega(q_0,q_1,q_2,\ldots) = \\ q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots \prod_{r\ge 0} (1+q_0q_1^{r+{\frac{1}{2}}}q_2^{(r+{\frac{1}{2}})^2}\cdots)(1+q_0^{-1}q_1^{r+{\frac{1}{2}}}q_2^{-(r+{\frac{1}{2}})^2}\cdots) \end{multline} \begin{multline}\label{09}V(q_1,q_2,\ldots) = q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots \\ \times\sum_{(m|n)}q_1^{\sum_j (m_j+{\frac{1}{2}})+(n_j+{\frac{1}{2}})}q_2^{\sum_j (m_j+{\frac{1}{2}})^2-(n_j+{\frac{1}{2}})^2} q_3^{\sum_j (m_j+{\frac{1}{2}})^3+(n_j+{\frac{1}{2}})^3}\cdots \end{multline} Here \begin{equation}\label{010}\xi(s) := \sum_{n\ge 1}\left(n-{\textstyle \frac{1}{2}}\right)^{-s} = (2^s-1)\zeta(s). \end{equation} Also, the sums are over all \begin{gather}\label{011}(m|n) = (m_1,m_2,\ldots,m_a|n_1,\ldots,n_a) \\ m_a>m_{a-1}>\cdots >m_1\ge 0,\quad n_a>n_{a-1}>\cdots>n_1\ge 0\notag \,. \end{gather} Such data are {\it Frobenius coordinates} for partitions $\lambda$ (cf. \cite{MacD}, p. 3.) In particular, the exponents in \eqref{09} are functions on the set of partitions. We define \begin{equation}\label{012}p_r(\lambda) := \sum_j \left(m_j+{\textstyle \frac{1}{2}}\right)^r+ (-1)^{r+1}\left(n_j+{\textstyle \frac{1}{2}}\right)^r \,. \end{equation} Thus $p_0(\lambda)=0$ and $p_1(\lambda) = |\lambda|$ is the number being partitioned. The formulas \eqref{08} and \eqref{09} appear (without the anomaly terms) in \cite{FKRW} and \cite{AFMO}. The anomaly terms depend on a choice of liftings of the $D_n$ to ${\widehat{\mathcal D}}$. For representations of the Virasoro algebra, this lifting is determined by modularity. The correct analog, quasimodularity, for representations of ${\widehat{\mathcal D}}$ was suggested by Dijkgraaf \cite{D} based on an interpretation of $V(\tau_1,\tau_2,0,0,\ldots)$ as a generating function counting covers of a fixed elliptic curve E with given degree and genus. From this point of view, a mirror symmetry argument suggested an interpretation in terms of (nonholomorphic) modular forms on the dual moduli space. More precisely, note that in \eqref{05} and \eqref{07} the variable $\tau_1$ plays a distinguished role. The analogue of modularity involves the behavior of $\Omega$ or $V$ under the transformations \begin{equation}\label{12} \tau_1 \mapsto \frac{a\tau_1+b}{c\tau_1+d},\qquad \tau_j \mapsto \frac{\tau_j}{(c\tau_j+d)^{j+1}},\ j\ne 1\,, \end{equation} for $\begin{pmatrix}a & b \\ c & d \end{pmatrix}\in \Gamma$, where $\Gamma\subset \text{\rm SL}_2({\Bbb Z})$ is some subgroup of finite index. Given $C=(c_0,c_2,c_3,\ldots)$ (resp. $C=(c_2,c_3,\ldots)$) with almost all $c_i=0$ define the weight of $C$ \begin{equation}\label{0013} \text{\rm wt}(C) := \sum (i+1)c_i \,. \end{equation} To say that the $\omega_A(\tau_1)$ (resp. the $v_B(\tau_1)$) were modular of weight $\text{\rm wt}(A)+w$ (resp. $\text{\rm wt}(B)+w$) for some $w$ would imply that $\Omega$ (resp. $V$) was modular of weight $w$ under the transformation \eqref{12}. This doesn't happen in the examples we know. Instead, the $\omega_A$ and $v_B$ are {\it quasimodular} of these weights. The notion of quasimodular form is developed in \cite{KZ}, where a rigorous proof of quasimodularity for the $\omega_A$ (resp. $v_B$) is given for $A=(a_0,a_2,0,0,\ldots)$ (resp. $B=(b_2,0,0,\ldots)$). By definition, the ring of quasimodular forms is the graded algebra generated over the ring of modular forms by the Eisenstein series of weight $2$ $$ G_2(q) := -\frac{B_2}{4}+\sum_{n=1}^\infty\sum_{d|n}dq^n. $$ Unlike the ring of modular forms, this ring is closed under the derivation $q\frac{d}{dq}$. One can canonically associate to a quasimodular form a certain nonholomorphic ({\it almost holomorphic} in the terminology of op.\ cit.) modular form. \begin{defn}\label{defn01} A series $F = \sum_A f_A(\tau_1)\tau^A/A!$ for $A=(a_0,a_2,\ldots)$ is quasimodular of weight $w$ if each $f_A(\tau_1)$ is quasimodular of weight $w+\text{\rm \rm wt}(A)$. \end{defn} \begin{thm}\label{th02}$\Omega(\tau_0,\tau_1,\ldots)$ (resp. $V(\tau_1,\tau_2,\ldots)$) is quasimodular of weight $0$ (resp. weight $-{\frac{1}{2}}$). \end{thm} Although the anomaly factors $q_{2r+1}^{-\xi(-2r-1)}$ are uniquely determined by quasimodularity, they can also be understood solely in representation-theoretic terms. Roughly speaking, if one tries to construct the infinite wedge representation as a representation of ${\mathcal D}$ (rather than ${\widehat{\mathcal D}}$) one is lead to meaningless infinite constants of the form $\sum_{n\ge 1} \left(n-{\frac{1}{2}}\right)^r,\ r\ge 1$. Regularizing these sums via analytic continuation of $\xi(s)$ and using the fact that $\xi(-2r)=0$ for $r\ge 0$ leads to the stated values. The analogy between \eqref{08} and the triple product formulas for the genus 1 theta functions continues with the following elliptic transformation formula for $\Omega(\tau_0,\tau_1,\ldots)$. \begin{thm}\label{th03} Define the transformation $T$ by $$ T(\tau_j) = \tau_j -\binom{j+1}{1}\tau_{j+1}+\binom{j+2}{2}\tau_{j+2} -\ldots\,, $$ and set $q'_i=\exp(2\pi i T(\tau_i))$. Then we have $$ \Omega(q_0',q_1',q_2',\ldots) = q_0q_1^{-{\frac{1}{2}}}q_2^{+\frac{1}{3}}q_3^{-\frac{1}{4}}\cdots\Omega(q_0,q_1,\ldots) $$ \end{thm} The above results constitute a pleasant but perhaps not terribly surprising generalization of the work in \cite{D} and \cite{KZ}. The character $V$ possesses some other hidden structure and relation to the genus 1 theta functions, which seems to us quite surprising and new. The idea is to replace the bulky generating function \eqref{09} for the quantities \eqref{07a} by other generating functions, where the auxiliary variables are attached not to the exponents $b_i$ in \eqref{07a} but to the indices $k$ of $\widetilde{D}_k$. These new generating functions admit a neat evaluation in terms of genus 1 theta functions. We call these generating functions $n$-point \emph{correlation functions} because both by their definition (as averages of product of $n$ generating series) and by their analytic structure (determinants built from theta functions and their derivatives) these functions closely resemble correlation functions in QFT. However, we were not able to find any precise connection and in the present text we work with these functions using solely the classical methods of analysis. Concretely, the definition of these $n$-point functions is the following. Let $f(\lambda)$ be a function on partitions, and define \begin{equation}\label{015} \langle f\rangle_q \ := \sum_\lambda f(\lambda)q^{|\lambda|}\Big/ \sum_\lambda q^{|\lambda|}\,. \end{equation} That is, $\langle f\rangle_q$ is the expectation of the function $f$ provided the the probability of each partition $\lambda$ is proportional to $q^{|\lambda|}$. Then, the equations \eqref{07a}, \eqref{09} and \eqref{011} can be restated as follows: \begin{equation}\label{015a} \text{\rm Trace}\big|_{\widetilde{D}_0=0} \left( q_1^{\widetilde{D}_1} \prod_{k=2}^\infty \left(\widetilde{D}_k\right)^{b_k} \right)=\eta(q_1) \left\langle \prod_{k=2}^\infty \left(p_k(\lambda)-\xi(-k)\right)^{b_k} \right\rangle_{q_1}\,. \end{equation} One checks (see section \ref{sec:prepart}) that \begin{equation}\label{016a} p_k(\lambda)-\xi(-k)=\left.\left(t\frac{d}{dt}\right)^k \left(\sum_{i=1}^\infty t^{\lambda_i-i+1/2} - \frac1{\log t}\right) \right|_{t=1}\,. \end{equation} By definition, set for $n=1,2,3,\dots$ \begin{equation}\label{016} F(t_1,\dots,t_n;q)= \left\langle \prod_{k=1}^n \left(\sum_{i=1}^\infty t_k^{\lambda_i-i+{\frac{1}{2}}}\right) \right\rangle_{q} \,, \end{equation} that is, $F(t_1,\dots,t_n;q)$ is the expectation of the product of $n$ generating series for the quantities \eqref{016a}. By linearity, all quantities \eqref{015a} satisfying $b_2+b_3+\dots\le m$ for some $m$ are encoded in the functions \eqref{016} with $n\le m$. \begin{defn}\label{defn02} The functions \eqref{016} are called $n$-point correlation functions. \end{defn} The equation \eqref{016a} can be restated as follows. Write $t_k = e^{u_k}$ and define a differential operator \begin{equation}\label{017a} \delta(u) := \frac1{u} +(2\pi i)^{-1}\sum_{r=1}^\infty \frac{\partial}{\partial\tau_r}u^r/r!\,. \end{equation} Then \begin{equation}\label{017} \Big<\prod_{k=1}^n \Big(\sum_{i=1}^\infty t_k^{\lambda_i-i+{\frac{1}{2}}}\Big)\Big>_{q_1} = \eta(q_1)\delta(u_1)\circ\ldots\circ\delta(u_n)V|_{\tau_2=\tau_3=\dotsc = 0}. \end{equation} This is also proved in section \ref{sec:prepart}. The form of the singular term $1/u$ and in \eqref{017a} and the corresponding term in \eqref{016a} very much depends on our choice of anomaly factors in the character. Write $q=q_1$, write $F(t_1,\dotsc,t_n)$ for the $n$-point function, and write $\Theta(t)=\Theta(t;q)$ for the following theta function: $$ \Theta(t) := \eta(q)^{-3}\sum_{n\in {\Bbb Z}} (-1)^nq^{\frac{(n+{\frac{1}{2}})^2}{2}}x^{n+{\frac{1}{2}}}= (q)^{-2}_\infty(x^{\frac{1}{2}} - x^{-{\frac{1}{2}}})(qx)_\infty(q/x)_\infty\,. $$ Let $\Theta^{(p)}(t) = (t\frac{d}{dt})^p\Theta(t)$. Then our main result on the $n$-point function is following: \begin{thm}\label{thm04} \begin{equation}\label{018} \vspace{-2 \jot} F(t_1,\dots,t_n)= \sum_{\sigma\in\mathfrak{S}(n)}\, \frac {\displaystyle \det\left( \frac{\displaystyle \Theta^{(j-i+1)}(t_{\sigma(1)}\cdots t_{\sigma(n-j)})}{\displaystyle (j-i+1)!} \right)} {\displaystyle \Theta(t_{\sigma(1)})\,\Theta(t_{\sigma(1)} t_{\sigma(2)}) \dots \Theta(t_{\sigma(1)}\cdots t_{\sigma(n)})} \,\,. \end{equation} Here $\sigma$ runs through all permutations $\mathfrak{S}(n)$ of $\{1,\dotsc,n\}$, the matrices in the numerator have size $n\times n$, and we define $1/(-n)!=0$ if $n\ge 1$. \end{thm} For the $1$-point function, this is simply $$\Big<\sum_{i=1}^\infty t^{\lambda_i-i+{\frac{1}{2}}}\Big>_q = \frac{1}{\Theta(t)}\,, $$ because $$ \Theta'(1)=1\,. $$ For $n=3$ the equation \eqref{018} becomes: \begin{multline} F(t_1,t_2,t_3)=\\ \frac1{\displaystyle\Theta(t_1 t_2 t_3)} \sum_{\sigma\in\mathfrak{S}(3)} \det \left( \begin{array}{rrc} \frac{\displaystyle\Theta'\left(t_{\sigma(1)} t_{\sigma(2)}\right)} {\displaystyle \Theta\phantom{{}'}\left(t_{\sigma(1)} t_{\sigma(2)}\right)}& \frac{\displaystyle 1}{\displaystyle 2} \, \frac{\displaystyle\Theta''\left(t_{\sigma(1)}\right)} {\displaystyle\Theta\phantom{{}''}\left(t_{\sigma(1)}\right)}& \frac{\displaystyle\Theta'''(1)}{\displaystyle 3!}\\ {\displaystyle 1}& \frac{\displaystyle\Theta'\phantom{{}'}\left(t_{\sigma(1)}\right)} {\displaystyle\Theta\phantom{{}''}\left(t_{\sigma(1)}\right)}& 0\\ & {\displaystyle 1}& 1 \end{array} \right) \end{multline} The traces \eqref{015a} can be computed from \eqref{018} using repeated differentiation, L'Hospital's rule and formulas for the derivatives $\Theta^{(p)}(1;q)$ in terms of the Eisenstein series: \begin{align*} \Theta^{(3)}(1;q)&= - 6 G_2(q)\,,\\ \Theta^{(5)}(1;q)&= - 10 G_4(q)+ 60 G_2(q)^2\,, \\ \Theta^{(7)}(1;q)&= - 14 G_6(q)+420 G_4(q) G_2(q)- 840 G_2(q)^3\,, \end{align*} and so on, see \eqref{th'}. Note that all even order derivatives of the odd function $\Theta(x)$ vanish at $x=1$. Sections \ref{sec:results}-\ref{sec:pf} are devoted to the proof of \eqref{018}. An essential ingredient of this proof are the following $q$-difference equations for the functions $F(t_1,\dots,t_n)$ which is established in section 8: \begin{thm}\label{thm05} For $n=1,2,\dots$ we have \begin{multline*} F(qt_1,t_2,\dots,t_n)=-q^{1/2} t_1 \dots t_n \times \\ \sum_{s=0}^{n-1} (-1)^s \sum_{1<i_1<\dots < i_s \le n} F(t_1 t_{i_1} t_{i_2} \cdots t_{i_s}, \dots,\widehat{\,t_{i_1}}, \dots,\widehat{\,t_{i_s}}, \dots) \,. \end{multline*} \end{thm} These equations are analogs of elliptic transformations for the $n$-point functions $F(t_1,\dots,t_n)$\,. Here is another point of view about these things. Let ${\mathcal A}$ be an associative algebra with $1$, and let $\rho : {\mathcal A} \to \text{\rm End}(F)$ be a representation of the associated Lie algebra (with commutator bracket). We do {\it not} assume $\rho$ compatible with the associative algebra structures. Let $D\in {\mathcal A}$ be given and assume $\rho(D)$ is semisimple with finite dimensional eigenspaces. Let $F_a \subset F$ be the eigenspace of $\rho(D)$ with eigenvalue $a$. Since $[D, D^n]=0$, $\rho(D^n)$ stabilizes $F_a$, and it makes sense to consider for a variable $u$ the expression \begin{equation}\label{019}f_\rho(u,z):= \chi_\rho(z)^{-1}\sum_a\Big(\text{\rm Tr}_{F_a}\sum_{n\ge 0} \rho(D^n)u^n/n!\Big)e^{az}\,. \end{equation} where $\chi_\rho(x) := \text{\rm Tr}e^{x\rho(D)}$ is the trace of $\rho$. If in fact $\rho(D^n)=\rho(D)^n$ for all $n$, we find \begin{equation}f_\rho(u,z) = \chi_\rho(z)^{-1}\sum_a \text{\rm Tr}_{F_a}e^{at}e^{az}=\chi_\rho(u+z)/\chi_\rho(z) \,. \end{equation} More generally, letting $n=(n_1,\ldots,n_r)$ run through $r$-tuples of non-negative integers and replacing $\rho(D^n)$ with $\prod_i \rho(D^{n_i})$ and $u^n/n!$ with $\prod u_i^{n_i}/n_i!$ we may define $f_\rho(u_1,\ldots,u_r;z)$. When $\rho$ is compatible with associative multiplication, we have $f_\rho = \chi_\rho(u_1+\ldots+u_r+z)/\chi_\rho(z)$. When $\rho$ is a projective representation, $\rho(D^n)$ is well defined only up to adding $\alpha_n\cdot \text{\rm Id}$. Such a modification replaces $f_\rho(u,z)$ with $f_\rho(u,z)+(\sum_n\alpha_nu^n/n!)$. If ${\mathcal H}$ is spanned by the powers of $D$ as in \eqref{02} we find writing $\chi_\rho(\tau_0,\tau_1,\tau_2,\ldots)$ for the full character that \begin{multline}\label{020} f_\rho(u_1,\ldots,u_r;\tau_1) = \\ \chi_\rho(\tau_1)^{-1}\sum_{n_1,\ldots,n_r\ge 0}\left(\prod_{i=1}^r\frac{\partial}{\partial\tau_{n_i}}\right) \chi_\rho(\tau_0,\tau_1,\tau_2,\ldots) |_{\tau_0=\tau_2=\ldots=0}\prod_i u_i^{n_i}/n_i! \,. \end{multline} For the representation \eqref{06}, this is \eqref{017}. Finally, section \ref{sec:example} proves quasimodularity and computes the $n$-point function for a representation of a subalgebra of ${\mathcal D}$ studied in \cite{B}. The first author would like to acknowledge considerable inspiration from conversations and correspondence with V. Kac and D. Zagier. It was Zagier who suggested that these characters might be quasimodular. \section{The Infinite Wedge Representation}\label{sec:rep} Material in this section is quite well known. References are \cite{K}, \cite{K2}, \cite{PS}, \cite{FKRW}. We work throughout with vector spaces, Lie algebras, and representations of Lie algebras over the field ${\Bbb C}$. Recall that a (projective) representation of a Lie algebra ${\mathcal L}$ is a Lie algebra homomorphism ${\mathcal L} \to \text{\rm End}(F)$; (resp. ${\mathcal L} \to \text{\rm End}(F)/{\Bbb C}\cdot\text{\rm Id}$) for a vector space $F$. A projective representation gives rise to a central extension of Lie algebras via pullback: \begin{equation}\begin{CD}0 @>>> {\Bbb C}\cdot c @>>> \widetilde{{\mathcal L}} @>>> {\mathcal L} @>>> 0 \\ @. @VVV @VVV @VVV @. \\ 0 @>>>{\Bbb C}\cdot\text{\rm Id} @>>> \text{\rm End}(F) @>>> \text{\rm End}(F)/{\Bbb C}\cdot\text{\rm Id} @>>> 0\,. \end{CD} \end{equation} Let $V$ be a vector space with basis $v_n,\ n\in {\Bbb Z}$. Define the finite matrices \begin{equation}{\mathcal A}_0 := \bigoplus_{i,j\in {\Bbb Z}} {\Bbb C} E_{ij}\subset \text{\rm End}(V)\,. \end{equation} where $E_{ij}(v_i)=v_j$ and $E_{ij}(v_k)=0;\ k\ne i$. Let ${\mathcal A}\supset{\mathcal A}_0$ be the larger space of all matrices supported in a bounded strip. An infinite matrix $M=\sum_{\text{\rm infinite}}a_{ij}E_{ij}$ lies in $A$ if and only if there exists a constant $c$ such that $a_{ij}=0$ if $|i-j|>c$. For example, ${\rm Id}\in{\mathcal A}-{\mathcal A}_0$. Suppose given a representation $\alpha_0$ of ${\mathcal A}_0$ which extends to a projective representation $\alpha$ of ${\mathcal A}$: \begin{equation}\label{13} \begin{array}{ccc} {\mathcal A}_0 & \stackrel{\iota}{\hookrightarrow} & \makebox[.8in][l]{${\mathcal A}$} \\ \rule[-4mm]{0cm}{9mm}\makebox[.1in][l]{$\downarrow\! \alpha_0$} & \makebox[.1in]{$\swarrow\!\sigma$} & \makebox[.7in][l]{$\downarrow\! \alpha$} \\ \text{\rm End}(F) & \twoheadrightarrow & \text{\rm End}(F)/{\Bbb C}\cdot{\rm Id}\,. \end{array} \end{equation} Let $\sigma$ be a lifting of $\alpha$ which is a map of vector spaces, not necessarily compatible with Lie algebra structures and not necessarily extending $\alpha_0$. Assume, however, that $\sigma$ is continuous in the sense that for any $f\in F$ and any $x=\sum_{\text{\rm infinite}}x_{ij}\iota(E_{ij})\in {\mathcal A}$ there exists $N=N(x,f)$ such that $x_{ij}\sigma\iota(E_{ij})(f)=0$ for $|i|+|j|>N$, and $$\sigma(x)(f) = \sum_{|i|+|j|\le N}x_{ij}\sigma\iota(E_{ij})(f)\,. $$ Notice that such a $\sigma$ is unique up to a finite modification. More precisely, if $\sigma'$ is another map with the same property, $\sigma\iota(E_{ij})=\sigma'\iota(E_{ij})$ for almost all $i,j$. We obtain in this way a map \begin{equation}\label{14} \epsilon := \alpha_0-\sigma\circ\iota : {\mathcal A}_0 \to {\Bbb C}\cdot{\rm Id}\,. \end{equation} The functional $\epsilon$ is uniquely determined (up to a finite modification) by $\alpha_0,\alpha$. \begin{defn}\label{reg} Let ${\mathcal S}$ be a collection of sequences of complex numbers $c=\{c_{ij}\}_{i,j\in {\Bbb Z}}$. A regularization scheme for ${\mathcal S}$ is a map $r:{\mathcal S} \to {\Bbb C}$ such that if $\{c_{ij}\}, \{c_{ij}'\}\in{\mathcal S}$ differ for only finitely many pairs $i,j$, then $$r(\{c_{ij}\}) = r(\{c_{ij}'\})+\sum_{i,j}(c_{ij}-c_{ij}'). $$ (We will usually assume ${\mathcal S}$ is saturated in the sense that if $c\in {\mathcal S}$ and $c'$ differs from $c$ for only finitely many $i,j$ then $c'\in {\mathcal S}$.) \end{defn} Assume now we have $\alpha, \alpha_0$, and that there exists a $\sigma$ as in \eqref{13} above. Let $\widehat{{\mathcal A}}$ be the corresponding central extension of ${\mathcal A}$. Let ${\mathcal S},r$ be as in definition (\ref{reg}). Suppose given $x=\sum x_{ij}\iota(E_{ij})\in {\mathcal A}$, such that $\{x_{ij}\epsilon(E_{ij})\}\in {\mathcal S}$, where $\epsilon$ is defined in \eqref{14}. Then $$\sigma(x) + r(\{x_{ij}\epsilon(E_{ij})\})\cdot{\rm Id} \in \text{\rm End}(F) $$ is well-defined independent of the choice of $\sigma$. Thus we have a lifting $\hat x\in \widehat{{\mathcal A}}$ depending on the regularization scheme $r$ but not on the choice of lifting $\sigma$. \begin{exam}\label{ex12} For the infinite wedge representation, \begin{equation}\epsilon(E_{ij}) = \begin{cases} 1 & i=j\le 0\,, \\ 0 & \text{\rm else\,.} \end{cases} \end{equation} The elements $x$ we want to lift have the form $x=\sum_{m\in{\Bbb Z}}\left(m-{\textstyle \frac{1}{2}}\right)^kE_{mm}$ for $k=0,1,2,\ldots$, so the sequences we need to regularize are $$\left\{\left(m-{\textstyle \frac{1}{2}}\right)^k\right\}_{m\le 0}\,. $$ The natural way to do this is by analytic continuation of the function \begin{equation}\label{16}\xi(s) = \sum_{n\ge 1} \left(n-{\textstyle \frac{1}{2}}\right)^{-s} \end{equation} We define $r(\{(m-{\frac{1}{2}})^k\}_{m\le 0})=(-1)^k\xi(-k)$\,. \end{exam} The infinite wedge representation $F:=\Lambda^\infty V$ is the vector space with basis \begin{equation}v_I = v_{i_m}\wedge v_{i_{m-1}}\wedge v_{i_{m-2}}\wedge\ldots\,, \end{equation} where $i_{k-1}<i_k$ and $i_k = k$ for $k\ll 0$. By definition $m$ is the {\it charge} of $v_I$ and $$\text{\rm energy}(v_I) := \sum_{k=m}^{k=-\infty}(i_k-k)\,. $$ For example, the elements \begin{equation} |m\rangle := v_m\wedge v_{m-1}\wedge v_{m-2}\wedge\ldots \end{equation} have charge $m$ and energy $0$. The wedge here is formal, but we can use the expected alternating linearity to make sense for $A\in {\mathcal A}_0$ of expressions like \begin{equation}\label{19} Av_I := \sum_{k=m}^{k=-\infty}v_{i_m}\wedge\ldots\wedge Av_{i_k}\wedge v_{i_{k-1}}\wedge \ldots \quad . \end{equation} We obtain a representation $\alpha_0 : {\mathcal A}_0 \to \text{\rm End}(F)$. To extend $\alpha_0$ to ${\mathcal A}$, we would have to define expressions like \begin{equation}\label{110} \sum_{p\in {\Bbb Z}}a_pE_{p,p+r}(v_I)\,. \end{equation} When $r\ne 0$ it is easy to see that $E_{p,p+r}v_I = 0$ for $|p|>>0$, so \eqref{110} makes sense. Similarly $E_{p,p}v_I=0$ for $p>>0$. Define $\sigma : {\mathcal A}_0 \to \text{\rm End}(F)$ by \begin{equation} \sigma(E_{ij}) = \begin{cases}\alpha_0(E_{ij}) & i\ne j \text{\rm or }i=j>0 \\ \alpha_0(E_{ii})-{\rm Id} & i\le 0\end{cases} \end{equation} It is straightforward to check that for any $r$ and any $I$ there exists an $N$ such that $\sigma(E_{p,p+r})(v_I)=0$ for $|p|\ge N$, so $\sigma$ extends to a map ${\mathcal A}\to\text{\rm End}(F)$ as in \eqref{13}. Define $\alpha$ to be the composition $${\mathcal A}\stackrel{\sigma}{\to}\text{\rm End}(F)\to \text{\rm End}(F)/{\Bbb C}\cdot{\rm Id}\,. $$ \begin{lem} The map $\alpha$ is a projective representation of ${\mathcal A}$. \end{lem} \begin{proof} The assertion is there exists a bilinear map $a: {\mathcal A}\times {\mathcal A} \to {\Bbb C}$ with $$[\sigma(x),\sigma(y)] = \sigma([x,y]) + a(x,y){\rm Id}. $$ One defines $$a(E_{ij},E_{k\ell}) = \begin{cases}1 & i=\ell\le 0 \text{\rm\ and }j=k\ge 1\,, \\ -1 & i=\ell\ge 1 \text{\rm\ and } j=k\le 0 \,,\\ 0 & \text{\rm else}\,.\end{cases} $$ Because of the conditions on the signs of the indices, $a$ extends to ${\mathcal A}\times{\mathcal A}$ as desired. \end{proof} Given $u\in {\Bbb C}$, we have an action as differential operators of ${\mathcal D}$ on ${\Bbb C}[t,t^{-1}]t^u$. Identifying this space with $V$ via $t^{n+u} \mapsto v_n$, we get a mapping of associative algebras (and hence a fortiori of Lie algebras) \begin{gather*} \delta_u : {\mathcal D} \to {\mathcal A}\subset \text{\rm End}(V)\,, \\ \delta_u(t^pD^r) = \sum_{n\in {\Bbb Z}}(n+u)^rE_{n,n+p} \notag\,. \end{gather*} We define a projective representation \begin{gather*}\rho_u := \alpha\circ\delta_u : {\mathcal D} \to \text{\rm End}(F)/{\Bbb C}\cdot{\rm Id}\,, \\ \rho := \rho_{-{\frac{1}{2}}}\notag\,. \end{gather*} We will customarily view $\rho$ as a representation on a central extension, $\rho : {\widehat{\mathcal D}} \to \text{\rm End}(F)$. Our objective now is to compute the character of the representation $\rho_u$. Define operators $\psi^*_{-r},\ r\ge 0$ and $\psi_{-r},\ r>0$ on $F$ by $$\psi_{-r}(v_I) := v_r\wedge v_I;\quad \psi^*_{-r}(\ldots\wedge v_{-r}\wedge\ldots) = \ldots\wedge\widehat{\rule{0cm}{3mm}v_{-r}}\wedge\ldots\quad. $$ A basis for $F$ can be written \begin{multline*} \psi_{-i_1}\cdots\psi_{-i_a}\psi^*_{-j_1}\cdots\psi^*_{-j_b}|0\rangle\,,\\ 0<i_1<i_2<\ldots<i_a,\ 0\le j_1<j_2<\ldots<j_b\,. \end{multline*} We have \begin{multline*}\sigma\circ\delta_u(D_r)\Big(\psi_{-i_1}\cdots\psi_{-i_a}\psi^*_{-j_1} \cdots\psi^*_{-j_b}|0\rangle\Big) = \\ \sigma\begin{pmatrix}\ddots &&\\ &(n+u)^r &\\ &&\ddots\end{pmatrix}\Big(\psi_{-i_1}\cdots \psi_{-i_a}\psi^*_{-j_1} \cdots\psi^*_{-j_b}|0\rangle\Big) = \\ \Big(\sum_{\ell=1}^a (i_\ell+u)^r - \sum_{m=1}^b (-j_m+u)^r\Big)\Big(\psi_{-i_1}\cdots \psi_{-i_a}\psi^*_{-j_1} \cdots\psi^*_{-j_b}|0\rangle\Big). \end{multline*} For example, writing $q_r = \exp(2\pi i\tau_r)$ we get \begin{gather*}\exp\Big(2\pi i\sum_{r\ge 0}\tau_r\cdot\sigma\circ\delta_u(D_r)\Big)(\psi_{-n}|0\rangle) = q_0q_1^{n+u}q_2^{(n+u)^2}\cdots \,,\\ \exp\Big(2\pi i\sum_{r\ge 0}\tau_r\cdot\sigma\circ\delta_u(D_r)\Big)(\psi^*_{-n}|0\rangle) = q_0^{-1}q_1^{n+u}q_2^{-(n+u)^2}\cdots \,.\notag \end{gather*} On all of $F$ we find \begin{multline*}\text{\rm Tr}\exp\Big(2\pi i\sum_{r\ge 0}\tau_r\cdot\sigma\circ\delta_u(D_r)\Big) = \\ \prod_{n\ge 0}\big(1+q_0q_1^{n+u+1}q_2^{(n+u+1)^2}\cdots\big)\big(1+q_0^{-1}q_1^{n-u}q_2^{-(n-u)^2}\cdots\big). \end{multline*} We now specialize to the case $u=-{\frac{1}{2}}$. The reason why this particular value of $u$ yields a quasimodular character is not clear, but it may have to do with the fact that for $u=p+{\frac{1}{2}}\in {\Bbb Z}+{\frac{1}{2}}$ the space $V={\Bbb C}[t,t^{-1}]t^u$ has a nondegenerate, symmetric bilinear form $$(t^{n+u},t^{m+u}) := \text{\rm res}_{t=0} t^{n+m+2u-1}dt. $$ Further, if we consider the involution $$\sigma : {\mathcal D} \to {\mathcal D};\quad \sigma(t^aD^b) := -t^a(-D-a)^b, $$ we find $$\sigma^2 = I;\quad \sigma[x,y] = [\sigma(x),\sigma(y)];\ (x(a),b)+(a,\sigma(x)(b)) = 0. $$ With reference to example (\ref{ex12}), we note that $\xi(s) = (2^s-1)\zeta(s)$ vanishes at $s=0,-2,-4,\ldots$ We therefore define the character of the infinite wedge representation of ${\mathcal D}$ to be \begin{multline}\label{1_18}\Omega(q_0,q_1,\ldots) = \Omega(\tau_0,\tau_1,\ldots) = \\ q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots\prod_{n\ge 0}\big(1+q_0q_1^{n+{\frac{1}{2}}}q_2^{(n+{\frac{1}{2}})^2}\cdots\big)\big(1+q_0^{-1}q_1^{n+{\frac{1}{2}}} q_2^{-(n+{\frac{1}{2}})^2}\cdots\big)\,. \end{multline} Note that $\delta_u(D_0)=\sum_{i\in {\Bbb Z}} E_{ii}$ is the charge operator \begin{multline*}\delta_u(D_0)\psi_{-i_1}\cdots \psi_{-i_a}\psi^*_{-j_1} \cdots\psi^*_{-j_b}|0\rangle = \\ (a-b)\psi_{-i_1}\cdots\psi_{-i_a}\psi^*_{-j_1}\cdots\psi^*_{-j_b}|0\rangle. \end{multline*} Operators in ${\mathcal D}$ preserve the charge, so the charge eigenspaces are stable under $\rho_u$. For $u=-{\frac{1}{2}}$, the characters of the subrepresentation of charge $n$ is the coefficient of $q_0^n$ in \eqref{1_18}. Of particular interest is the charge $0$ part. Using \eqref{09} and \eqref{012}, this can be written \begin{multline}\label{120} V(q_1,q_2,\ldots) = V(\tau_1,\tau_2,\ldots) := \\ q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots\{\text{\rm Coeff. of $q_0^0$ in $\Omega(q_0,q_1,\ldots)$}\}= \\ q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots\sum_\lambda q_1^{p_1(\lambda)}q_2^{p_2(\lambda)}\cdots\,. \end{multline} Here the sum is over all partitions $\lambda$. \section{Elliptic transformation of the character $\Omega$}\label{sec:transf} In this section we want to expose some interesting analogies between $\Omega(q_0,q_1,q_2,\ldots)$ \eqref{1_18} and the classical genus $1$ theta function, which we will write \begin{equation}\theta(q_0,q_1) := \sum_{n\in{\Bbb Z}}q_0^nq_1^{n^2/2}. \end{equation} The triple product formula (cf. \cite{MUM}, p. 70) implies \begin{multline}\label{22}\theta(q_0,q_1) = \prod_{m\ge 1}(1-q_1^m)\prod_{n\ge 0}(1+q_0q_1^{n+{\frac{1}{2}}}) (1+q_0^{-1}q_1^{n+{\frac{1}{2}}}) = \\ \eta(q_1)\Omega(q_0,q_1,1,1,\ldots). \end{multline} Here $\eta(q_1)=q_1^{\frac{1}{24}}\prod_{m\ge 1}(1-q_1^m)$ is the classical eta function, and we have used the fact that the first anomaly factor in $\Omega$ is given by $$q_1^{-\xi(-1)}=q_1^{-(2^{-1}-1)\zeta(-1)} = q_1^{-\frac{1}{24}}. $$ We want to generalize to $\Omega$ the elliptic transformation law \begin{equation}\theta(q_0q_1^{-1},q_1) = q_0q_1^{-{\frac{1}{2}}}\theta(q_0,q_1). \end{equation} For this, we define infinite matrices \begin{equation} U:= \begin{pmatrix}0 & -1 & 0 & 0 & \hdots \\ 0 & 0 & -2 & 0 & \hdots \\ 0 & 0 & 0 & -3 & \hdots \\ \vdots & \vdots & \vdots & \vdots \end{pmatrix} \end{equation} and \begin{equation} T:= \exp(U) = \begin{pmatrix}1 & -1 & 1 & -1 & 1 &\hdots \\\vspace{4pt} 0 & 1 & -\binom{2}{1} & \binom{3}{2} & -\binom{4}{3} & \hdots \\ 0 & 0 & 1 & -\binom{3}{1} & \binom{4}{2} & \hdots \\ \vdots & \vdots & \vdots & \vdots & \vdots \end{pmatrix} \quad. \end{equation} Define for $j\ge 0$ \begin{gather*}\tau_j' := T(\tau_j) = \tau_j -\binom{j+1}{1}\tau_{j+1}+\binom{j+2}{2}\tau_{j+2} -\ldots\,, \\ q_j' := T(q_j) = q_jq_{j+1}^{-\binom{j+1}{1}}q_{j+2}^{\binom{j+2}{2}}\cdots\,. \notag \end{gather*} \begin{thm}\label{thm21} $\Omega(q_0',q_1',q_2',\ldots) = q_0q_1^{-{\frac{1}{2}}}q_2^{+\frac{1}{3}}q_3^{-\frac{1}{4}}\cdots\Omega(q_0,q_1,\ldots)$. \end{thm} \begin{lem}\label{lem22} For $n\ge 2$ we have $$\sum_{i=1}^{n-1}(-1)^i\binom{n}{i}\xi(i-n) = \frac{(-1)^{n+1}}{n+1}+\frac{(-1)^n}{2^n}\,. $$ \end{lem} \begin{proof}[Proof of lemma] Recall one has Bernoulli numbers $B_n$, $n\ge 0$ satisfying $B_0=1,\ B_1=-{\frac{1}{2}},\ \zeta(1-n)=-\frac{B_n}{n}$ for $n\ge 2$. Substituting and using the identity $\frac{n+1}{j+1}\binom{n}{j}= \binom{n+1}{j+1}$, the desired formula becomes \begin{equation}\label{27}\sum_{k=1}^n(-1)^{k-1}\binom{n+1}{k}(1-2^{1-k})B_k \stackrel{?}{=} \frac{n+1}{2^n} - 1 \,. \end{equation} Consider the Bernoulli polynomials $$B_N(x) := \sum_{k=0}^N\binom{N}{k}B_kx^{N-k} \,, $$ which may be defined by the generating function $$\sum_{N=0}^\infty B_N(x)t^N/N! = \frac{te^{xt}}{e^t-1}\,. $$ As a consequence we get \begin{equation}\label{28}B_N(x+1)-B_N(x) = Nx^{N-1} \,. \end{equation} Summing for $-1\ge x\ge -p$ we get for $N=n+1$ and $p\ge 1$ the equivalent identities \begin{gather}(n+1)\sum_{\ell=-1}^{-p}\ell^n+B_{n+1}(-p)-B_{n+1}(0)=0 \,,\notag \\ (n+1)\sum_{\ell=-1}^{-p}\ell^n+(-p)^{n+1}+\sum_{k=1}^n B_k \binom{n+1}{k}(-p)^{n+1-k} = 0\,. \label{29} \end{gather} (We are grateful to V. Kac for suggesting \eqref{29}.) It is straightforward to deduce \eqref{27} from \eqref{29} taking $p=1,2$\,. \end{proof} \begin{proof}[Proof of theorem] The transformation $T$ satisfies \begin{gather*}T(q_0)T(q_1)^sT(q_2)^{s^2}\cdots = q_0q_1^{s-1}q_2^{(s-1)^2}\cdots\,, \\ T(q_0)^{-1}T(q_1)^sT(q_2)^{-s^2}\cdots = q_0^{-1}q_1^{s+1}q_2^{-(s+1)^2}\cdots\,. \notag \end{gather*} Note the second identity follows from the first, replacing $s$ by $-s$ and inverting. The first identity is left for the reader. It follows that \begin{multline}\label{211}\frac{\prod_{n\ge 0}\big(1+T(q_0)T(q_1)^{n+{\frac{1}{2}}}\cdots\big)\big(1+T(q_0)^{-1}T(q_1)^{(n+{\frac{1}{2}})} T(q_2)^{-(n+{\frac{1}{2}})^2}\cdots\big)}{\prod_{n\ge 0}\big(1+q_0q_1^{n+{\frac{1}{2}}}\cdots\big)\big(1+q_0^{-1}q_1^{(n+{\frac{1}{2}})} q_2^{-(n+{\frac{1}{2}})^2}\cdots\big)} \\ = \frac{1+q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{4}}q_3^{-\frac{1}{8}} \cdots}{1+q_0^{-1}q_1^{{\frac{1}{2}}}q_2^{-\frac{1}{4}} \cdots} = q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{4}}\cdots\,. \end{multline} As a consequence of the lemma we have \begin{equation}\label{212} \frac{T(q_1)^{-\xi(-1)}T(q_3)^{-\xi(-3)}\cdots}{q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots} = q_2^{\frac{1}{3}-\frac{1}{4}}q_3^{-\frac{1}{4}+\frac{1}{8}}q_4^{\frac{1}{5}-\frac{1}{16}} \cdots \,. \end{equation} The proof follows by combining \eqref{211} and \eqref{212}. \end{proof} Recall we have defined $V(q_1,q_2,\ldots)$ to be the coefficient of $q_0^0$ in $\Omega(q_0,q_1,\ldots)$. \begin{thm}\label{thm23} We have the following series expansion for $\Omega$: $$\Omega(q_0,q_1,q_2,\ldots) = \sum_{n=-\infty}^{n=\infty}V(T^{-n}(q_1),T^{-n}(q_2),\ldots) q_0^nq_1^{n^2/2}q_2^{n^3/3}\cdots . $$ \end{thm} \begin{proof} First note \begin{equation}T(q_0)^nT(q_1)^{n^2/2}T(q_2)^{n^3/3}\cdots = q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{3}}\cdots q_0^{n-1}q_1^{(n-1)^2/2}\cdots. \end{equation} Write $\Omega = \sum_{-\infty}^\infty V_n(q_1,\ldots)q_0^nq_1^{n^2/2}\cdots$. Then \begin{multline*}q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{3}}\cdots\sum_nV_{n-1}(q)q_0^{n-1}q_1^{(n-1)^2/2}\cdots = \\ q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{3}}\cdots\Omega(q) = \Omega(T(q)) = \\ \sum V_n(T(q))T(q_0)^nT(q_1)^{n^2/2}\cdots = \\ q_0q_1^{-{\frac{1}{2}}}q_2^{\frac{1}{3}}\cdots\sum_nV_{n}(T(q))q_0^{n-1}q_1^{(n-1)^2/2}\cdots. \end{multline*} Taking coefficients of $q_0^{n-1}$ we get $V_n(T(q))=V_{n-1}(q)$. The formula follows since $V_0=V$. \end{proof} \begin{exam} Consider the formula in the theorem with $q_j\mapsto 1,\ j\ge 2$. We have $$T(q)|_{q_j\mapsto 1,j\ge 2} = q_1 $$ and by \eqref{012} $$V(q_1,1,1,\ldots) = q_1^{-\frac{1}{24}}\sum_\lambda q_1^{p_1(\lambda)} = q_1^{-\frac{1}{24}}\prod_{m\ge 1}(1-q_1^m)^{-1}=\eta(q_1)^{-1}. $$ The assertion of the theorem is then \begin{equation}\eta(q_1)^{-1}\sum_{n\in {\Bbb Z}} q_0^nq_1^{n^2/2} = q_1^{\frac{-1}{24}}\prod_{m\ge 0}(1+q_0q_1^{m+{\frac{1}{2}}})(1+q_0^{-1}q_1^{m+{\frac{1}{2}}}). \end{equation} Multiplying through by $\eta(q_1)$ yields the triple product formula \eqref{22}. \end{exam} \section{Quasimodular forms}\label{sec:quasi} In this section, we recall the theory of quasimodular forms as developed in \cite{KZ}. All results are due to Kaneko and Zagier and are recalled here solely for the convenience of the reader. We fix a subgroup of finite index $\Gamma\subset\Gamma_1 := \text{\rm SL}(2,{\Bbb Z})$. A holomorphic modular form of weight $k$ (for $\Gamma$) is a holomorphic function $f(\tau)$ on the upper half-plane ${\mathcal H}=\{\tau=x+iy\ |\ y>0\}$ satisfying \begin{equation}\label{31}f\left(\frac{a\tau+b}{c\tau+d}\right) = (c\tau+d)^kf(\tau),\quad \begin{pmatrix}a & b \\ c & d\end{pmatrix}\in\Gamma\,. \end{equation} Note for some $\ell\ge 1$ we have $(\begin{smallmatrix}1 & \ell \\ 0 & 1\end{smallmatrix})\in\Gamma$, so $f$ may be expanded in a Fourier series \begin{equation}\label{32}f(\tau) = \sum a_n\exp(2\pi in\tau/\ell) = \sum a_n q^{n/\ell}\,. \end{equation} We assume that $f$ is holomorphic at $i\infty$, i.e. $a_n=0$ for $n<0$ and the Fourier series converges for $|q|<1$. The holomorphic modular forms constitute a graded ring $$M_*(\Gamma) := \oplus M_k(\Gamma) $$ graded by the weight $k$. An {\it almost holomorphic} function on ${\mathcal H}$ will be a function \begin{equation} F(\tau) = \sum_{m=0}^N f_m(\tau)y^{-m} \end{equation} on ${\mathcal H}$ where $y=\text{\rm Im}(\tau)$ and each $f_m$ has a Fourier expansion as in \eqref{32}. An almost holomorphic modular form of weight $k$ is an almost holomorphic function $f(\tau)$ satisfying the weight $k$ modularity property \eqref{31}. \begin{exam} The classical Eisenstein series ($B_k$ Bernoulli number as in Lemma (\ref{lem22})) \begin{equation}\label{34}G_k(q):=\frac{-B_k}{2k}+\sum_{n=1}^\infty\Big(\sum_{d|n} d^{k-1}\Big)q^n,\quad k=2,4,6,\ldots \end{equation} is modular of weight $k$ for $\Gamma=\Gamma_1$ and $k\ge 4$. On the other hand, $G_2$ satisfies the transformation \begin{equation}\label{35} G_2\left(\frac{a\tau+b}{c\tau+d}\right) = (c\tau+d)^2G_2(\tau)-\frac{c(c\tau+d)}{4\pi i} \,. \end{equation} Notice, however, that \begin{equation}\label{36} \left[\text{\rm Im}\left(\frac{a\tau+b}{c\tau+d}\right)\right]^{-1} -(c\tau+d)^2y^{-1} = -2ic(c\tau+d) \,. \end{equation} It follows that the function \begin{equation}\label{37} G_2^*(\tau) := G_2(\tau)+\frac{y^{-1}}{8\pi} \end{equation} is an almost holomorphic modular form of weight $2$ for $\Gamma_1$. \end{exam} \begin{defn}\label{defn32} A quasimodular form of weight $k$ is a holomorphic function $f(\tau)$ on ${\mathcal H}$ such that there exists an almost holomorphic modular form $F = \sum_{m=0}^N f_my^{-m}$ of weight $k$ with $f_0=f$. \end{defn} \begin{exam} $G_2$ is quasimodular of weight $2$. Indeed, one can take $F=G_2^*$. \end{exam} Let us write $AHM_*(\Gamma)$ (resp. $QM_*(\Gamma)$) for the graded ring of almost holomorphic (resp. quasimodular) forms. (Kaneko and Zagier write $\widehat{M}$ and $\widetilde{M}$, but this makes it difficult to remember which is which.) \begin{prop}\label{prop34} The assignment $F=\sum_{j=0}^My^{-j}f_j\mapsto f_0$ defines an isomorphism of graded rings, $AHM_*(\Gamma)\cong QM_*(\Gamma)$. \end{prop} \begin{proof} Note this is well defined, i.e. $\sum_{j=0}^My^{-j}f_j(\tau) \equiv 0$ for holomorphic $f_j$ if and only if all the $f_j$ are zero. To see this one can e.g.\ apply the differential operator $\frac{iM}{2}y^{-1}+\frac{d}{d\bar\tau}$ and argue by induction on $M$. The map in question is surjective by definition, so it suffices to show injectivity. Suppose for some $r\ge 1$ and $f_r \ne 0$ that $F=y^{-r}f_r+\ldots+y^{-M}f_M$ is almost holomorphic modular of weight $k$. Let $A=(\begin{smallmatrix}a & b \\ c & d\end{smallmatrix}) \in\Gamma$ and write $j=c\tau+d$. Using \eqref{36} we get \begin{multline*} (j^2y^{-1}-2icj)^rf_r(A\tau)+\ldots +(j^2y^{-1}-2icj)^Mf_M(A\tau) = \\ j^ky^{-r}f_r(\tau)+\ldots+j^ky^{-M}f_M(\tau). \end{multline*} Now identify coefficients of powers of $y^{-1}$ \begin{align}\label{39} f_M(A\tau) &= j^{k-2M}f_M(\tau) \,,\\ f_{M-1}(A\tau)-2icj\binom{M}{1}f_M(A\tau) &= j^{k-2M+2}f_{M-1}(\tau) \notag \,,\\ f_{M-2}(A\tau)-2icj\binom{M-1}{1}f_{M-1}(A\tau) & +(2icj)^2\binom{M}{2}f_M(A\tau) \notag\\ &= j^{k-2M+4}f_{M-2}(\tau)\notag\,, \\ \makebox[5cm][c]{$\vdots$} & \makebox[3cm][c]{$\vdots$} \notag \displaybreak[0]\\ f_r(A\tau)-2icj\binom{r+1}{1}f_{r+1}(A\tau) \qquad\quad& \notag \\ +(2icj)^2\binom{r+2}{2}f_{r+2}(A\tau) -\ldots &= j^{k-2r}f_r(\tau)\notag \,. \end{align} The terms not involving $y$ give \begin{equation}(-2icj)^rf_r(A\tau)+\ldots+(-2cij)^Mf_M(A\tau) = 0. \label{310} \end{equation} Solve \eqref{39} recursively: \begin{align*}f_M(A\tau) &= j^{k-2M}f_M(\tau) \,,\\ f_{M-1}(A\tau) &= 2ic\binom{M}{1}j^{k-2M+1}f_M(\tau)+j^{k-2M+2}f_{M-1}(\tau)\notag \,,\\ f_{M-2}(A\tau) &= \ldots + j^{k-2M+4}f_{M-2}(\tau) \notag \,, \\ \makebox[2cm][c]{$\vdots$} & \makebox[5cm][c]{$\vdots$} \notag \\ f_r(A\tau) & = \ldots + j^{k-r}f_r(\tau)\,. \notag \end{align*} Finally, substituting in \eqref{310} yields \begin{multline}\label{312} (c\tau+d)^{k-r}c^rf_r(\tau) +\alpha_{r+1}(c\tau+d)^{k-r-1}c^{r+1}f_r(\tau) +\ldots \\ + \alpha_M (c\tau+d)^{k-M}c^Mf_M(\tau) = 0. \end{multline} Here the $\alpha_j$ are constants independent of $c$ and $d$. Varying $A\in \Gamma$ yields a contradiction. Indeed, the map $\Gamma\to{\Bbb C}^2,\ A\mapsto (c,d)$ has Zariski dense image, so in the above identity, $c$ and $d$ can be taken to be independent variables. The coefficient of $c^rd^{k-r}$ is nontrivial. \end{proof} \begin{prop}\begin{enumerate}\item[(i)] $M_*(\Gamma)\subset QM_*(\Gamma)$. \item[(ii)] $QM_*(\Gamma) = M_*(\Gamma)\otimes{\Bbb C}[G_2]$. \item[(iii)] $QM_*(\Gamma)$ is stable under the operator (of degree $2$) $D:= \frac{d}{d\tau}$. We have $$QM_k(\Gamma) = \oplus_{0\le i\le k/2}D^iM_{k-2i}(\Gamma)\oplus{\Bbb C}\cdot D^{k/2-1}G_2. $$ \end{enumerate} \end{prop} \begin{proof} (i) is clear. To prove (ii), we claim first that the map \begin{equation}\label{313}M_*(\Gamma)\otimes{\Bbb C}[G_2^*] \to AHM_*(\Gamma) \end{equation} is an isomorphism, where $G_2^*$ is as in \eqref{37}. Indeed, for $F = f_0+\ldots+y^{-M}f_M$ almost holomorphic modular of weight $k$, it follows from the first line of \eqref{39} that $f_M$ is holomorphic modular of weight $k-2M$. We have $$F - f_M\cdot (8\pi G_2^*)^M = g_0+\ldots+y^{-M+1}g_{M-1} $$ is almost holomorphic modular of weight $k$. Surjectivity of \eqref{313} follows by induction on $M$. Injectivity is straightforward, keeping track of powers of $y^{-1}$. Assertion (ii) now follows from proposition (\ref{prop34}). Finally, (iii) is left for the reader. \end{proof} \section{Quasimodularity for characters $\Omega$ and $V$}\label{sec:qmov} The purpose of this section is to prove \begin{thm} The series $\Omega(\tau_0,\tau_1,\ldots)$ and $V(\tau_1,\tau_2,\ldots)$ are quasimodular of weights $0$ and $-{\frac{1}{2}}$ respectively. \end{thm} We focus first on $\Omega$. Recall (definition (\ref{defn32})) a series $F(\tau_0,\tau_1,\tau_2,\ldots)$ is said to be quasimodular of weight $k$ if it can be expanded in a formal Taylor series $$\sum_{J=(j_0,j_2,j_3,\ldots)} B_J(\tau_1)\tau^J/J! $$ with $B_J(\tau_1)$ quasimodular of weight \begin{equation}\label{4wt} k+{\rm wt}(J):= k+j_0+3j_2+4j_3+\ldots. \end{equation} Since we have not specified a group $\Gamma$, there will be no harm e.g. in replacing $\tau_1$ by $2\tau_1$. Define $$\Phi(q)=\Omega(-q_0,q_1,q_2,\ldots).$$ One has $\Omega(q) = \Phi(q_0^2,q_1^2,q_2^2,\ldots)/\Phi(q)$, so, since the space of quasimodular forms is a ring, it will suffice to prove $\Phi(q)$ is quasimodular. We have $$\Phi(1,x_1,1,1,\ldots) = x_1^{-\frac{1}{24}}\prod_{r\ge 0}(1-x_1^{r+\frac{1}{2}})^2 = \big(\eta(x^{1/2})/\eta(x)\big)^2, $$ which is modular of weight $0$. It therefore suffices to show $$G :=\log\big(\Phi(x)/\Phi(1,x_1,1,1,\ldots)\big) $$ is quasimodular. We have (with $\xi(s)$ as in \eqref{16}) \begin{multline}\label{41} G = -2\pi i\big(\sum_{n\ge 1}\xi(-2n-1)\tau_{2n+1}\big)+ \\ \sum_{r\ge 0}\big[\log(1-x_0x_1^{n+\frac{1}{2}}x_2^{(n+\frac{1}{2})^2}\cdots)+ \log(1-x_0^{-1}x_1^{n+\frac{1}{2}}x_2^{-(n+\frac{1}{2})^2}\cdots)- \\ 2\log(1-x_1^{n+\frac{1}{2}})\big]\,. \end{multline} We expand the final sum in \eqref{41} \begin{multline*}B:= \\ -\sum_{n\ge 0\,,\ l\ge 1}\frac{x_1^{(n+{\frac{1}{2}})l}}{l} \left(\exp\left(2\pi il\left(\tau_0+\left(n+{\textstyle \frac{1}{2}}\right)^2\tau_2+\left(n+{\textstyle \frac{1}{2}}\right)^3\tau_3+\ldots\right)\right)\right. \\ +\left. \exp\left(2\pi il\left(-\tau_0-\left(n+{\textstyle \frac{1}{2}}\right)^2\tau_2+\left(n+{\textstyle \frac{1}{2}}\right)^3\tau_3-\ldots\right)\right) - 2\right). \end{multline*} Define coefficients $\alpha_m(J)$, where $J=(j_0,j_2,j_3,\ldots)$ and $\ j_k \ge 0$, by \begin{equation}\frac{\partial^J}{\partial \tau^J}B|_{\tau_0=\tau_2=\ldots=0} = \sum_{m=1}^\infty\alpha_m(J)x_1^{m/2}\,. \end{equation} Note $\alpha_m(0)=0$. For $J\ne 0$, write $|J|=j_0+j_2+\ldots$. We get \begin{equation}\alpha_m(J) = -(2\pi i)^{|J|}\sum_{n\ge 0\,,\ l\ge 1}l^{|J|-1}(1+(-1)^{{\rm wt}(J)})\left(n+{\textstyle \frac{1}{2}}\right)^{{\rm wt}(J)-|J|}x_1^{(n+{\frac{1}{2}})l}. \end{equation} Thus, $\alpha_m(J)=0$ for ${\rm wt}(J)$ odd. For even weight, we find \begin{multline}\label{45} \frac{\partial^J}{\partial \tau^J}B|_{\tau_0=\tau_2=\ldots=0} = -2(2\pi i)^{|J|}2^{|J|-{\rm wt}(J)}\sum_{m\ge 1}m^{|J|-1}x_1^{m/2}\sum_{d|m;\ d \text{\rm odd}}d^{{\rm wt}(J)-2|J|+1} \\ = -2(2\pi i)^{|J|}2^{|J|-{\rm wt}(J)}\sum_{m\ge 1}m^{{\rm wt}(J)-|J|}x_1^{m/2}\sum_{l|m;\ \frac{m}{l} \text{\rm odd}}l^{2|J|-{\rm wt}(J)-1} \,. \end{multline} The Eisenstein series of weight $k$ and level $1$ was given in \eqref{34}. As in \cite{KZ}, we work with the level two Eisenstein series \begin{gather*}F_k^{(1)}(q) := G_k(q^{{\frac{1}{2}}}) - G_k(q) = \sum_{n=1}^\infty\Big(\sum_{d|n,2\not\,\mid d}(n/d)^{k-1}\Big)q^{n/2}\,, \\ F_k^{(2)} := G_k(q^{{\frac{1}{2}}}) - 2^{k-1}G_k(q) = (1-2^{k-1})\zeta(1-k)/2+ \sum_{n=1}^\infty\Big(\sum_{d|n,2\not\,\mid d}d^{k-1}\Big)q^{n/2}. \end{gather*} Suppose first that the inequality $k(J) := {\rm wt}(J)-2|J|+2>1$ holds. It follows from the first equality in \eqref{45} that \begin{equation}\label{48} \frac{\partial^J}{\partial \tau^J}B|_{\tau_0=\tau_2=\ldots=0} = -4\pi i(2^{-k(J)+1})(\partial/\partial\tau_1)^{|J|-1}(F_{k(J)}^{(2)}(x_1)-F_{k(J)}^{(2)}(0))\,. \end{equation} Note $k(J)$ is even, so the remaining possibility is $k(J) \le 0$. In this case, the second equality in \eqref{45} implies \begin{equation}\frac{\partial^J}{\partial \tau^J}B|_{\tau_0=\tau_2=\ldots=0} = -2(2\pi i)^{2-k(J)}(\partial/\partial\tau_1)^{{\rm wt}(J)-|J|}(F_{2-k(J)}^{(1)}(x_1))\,. \end{equation} Since the $F^{(i)}_k$ are quasimodular, it follows that the $\frac{\partial^J}{\partial \tau^J}B|_{\tau_0=\tau_2=\ldots=0}$ are quasimodular except possibly (because of the constant term on the right) in in the case $k(J)\ge 2;\ |J|=1$ where the constant term is not correct. This only happens if $J=(0,\ldots,0,1,0,\ldots)$ where $j=1$ for some $j\ge 3$ odd and all the other entries are zero. In this case $k(J)=j+1$ and it follows from \eqref{48} and the identity $\xi(s)=(2^s-1)\zeta(s)$ that \begin{multline*}\frac{\partial}{\partial\tau_j} B|_{\tau=0} = (-4\pi i)2^{-j}\big(F^{(2)}_{j+1}(x_1)+(2^j-1)\zeta(-j)/2\big)= \\ (-4\pi i)2^{-j}F^{(2)}_{j+1}(x_1)+2\pi i\xi(-j)\,. \end{multline*} The last constant term exactly cancels the $-2\pi i\xi(-j)$ which appears in the first sum on the right in \eqref{41}. This completes the proof of quasimodularity for $\Omega$. It remains to prove quasi-modularity for $V(x_1,\ldots)$. The proof parallels that in \cite{KZ}. We begin by expanding \begin{gather} V(\tau_1,\tau_2,\ldots) = \sum_{K = (k_2,k_3,\ldots)}A_K(\tau_1)\tau^K/K!\,, \\ V(T^{-n}(\tau)) = \sum_{m=0}^\infty\sum_{K}(\partial/\partial\tau_1)^mA_K(\tau_1) (T^{-n}(\tau_1)-\tau_1)^m(m!)^{-1}(T^{-n}(\tau))^K/K!\,, \\ \label{413} \Omega(\tau) = \sum_{n=-\infty}^\infty\sum_{J=(j_0,j_2,j_3,\ldots)}\sum_{m=0}^\infty\sum_{K} (\partial/\partial\tau_1)^mA_K(\tau_1)(T^{-n}(\tau_1)-\tau_1)^m(m!)^{-1}\times \\ \times (T^{-n}(\tau))^K(K!)^{-1}a_Jn^{\text{\rm wt}(J)}\tau^J(J!)^{-1}x_1^{n^2/2}\,. \notag \end{gather} Here $\sum_Ja_Jn^{\text{\rm wt}(J)}\tau^J(J!)^{-1}$ is the power series expansion of $x_0^nx_2^{n^3/3}x_3^{n^4/4}\cdots$, with $a_J$ independent of $n$. One verifies the expansion \begin{equation}\label{414} T^{-n}(\tau_j) = \tau_j+\binom{j+1}{1}n\tau_{j+1}+\binom{j+2}{2}n^2\tau_{j+2}+\ldots\quad . \end{equation} Fix $m\ge 0$ and $K=(k_2,k_3,\ldots)$. Let $P(\tau)=\tau_2^{p_2}\tau_3^{p_3}\ldots$ and $Q_i(\tau)=\tau_i^{q_{ii}}\tau_{i+1}^{q_{i,i+1}}\ldots;\ i\ge 2$ be monomials with $\deg(P)=\sum p_j = m$ and $\deg(Q_i)=k_i$. Define the weights by ${\rm wt}(P) = \sum (j+1)p_j$ and ${\rm wt}(Q_i) = \sum_{j\ge i} (j+1)q_{ij}$. As a consequence of \eqref{414}, $P$ appears in $(T^{-n}(\tau_1)-\tau_1)^m(m!)^{-1}$ with a coefficient $c_{P,m}n^{{\rm wt}(P)-2m}$, where $c_{P,m}$ is constant independent of $n$. Similarly, $Q:=\prod Q_i$ appears in $(T^{-n}(\tau))^K(K!)^{-1}$ with coefficient $c_{\{Q_i\},K}n^{\sum_i {\rm wt}(Q_i)-(i+1)k_i}$. Now fix $J=(j_0,j_2,\ldots)$. We obtain a contribution to the coefficient of the monomial $P(\tau)Q(\tau)\tau^J$ in \eqref{413} of the form \begin{equation}\label{415} c_{J,K,m,\{Q_i\},P}\frac{\partial^m}{\partial\tau_1^m}A_K(\tau_1) \sum_{n=-\infty}^\infty n^{{\rm wt}(P)-2m+\sum_i ({\rm wt}(Q_i)-(i+1)k_i)+{\rm wt}(J)}x_1^{n^2/2}. \end{equation} If the exponent of $n$ is odd, this cancels. Assume this exponent equals $2r$ for $r\ge 0$ an integer. Then \eqref{415} can be rewritten \begin{equation}\label{416} b_{J,K,m,\{Q_i\},P}\frac{\partial^m}{\partial\tau_1^m}A_K(\tau_1) \frac{\partial^r}{\partial\tau_1^r}\theta_{00}(\tau_1)\,, \end{equation} where $\theta_{00}(\tau_1)=\sum \exp(\pi in^2\tau_1)$ is a modular form of weight $1/2$. Since differentiation preserves quasi-modularity and increases weight by $2$, it follows that, assuming $A_K(\tau_1)$ is quasimodular of weight ${\rm wt}(K)-{\frac{1}{2}}$, the expression in \eqref{416} is quasi-modular of weight ${\rm wt}(P)+\sum{\rm wt}(Q_i)+{\rm wt}(J)$. We will prove $A_K(\tau_1)$ is quasimodular of weight ${\rm wt}(K)-1/2$ by induction on ${\rm wt}(K)$. Write \begin{multline*}\theta_{00}(x) = \sum_{-\infty}^\infty x^{n^2/2}= \prod_{m\ge 1}(1-x^m)\prod_{m\ge 0}(1+x^{m+{\frac{1}{2}}})^2 = \\ \eta(x)\Omega(1,x,1,1\ldots)\,. \end{multline*} Substituting $x_j=1;\ j\ne 1$ in Theorem (\ref{thm23}) yields \begin{gather*}\Omega(1,x_1,1,\ldots) = V(x_1,1,\ldots)\sum x_1^{n^2/2} = V(x_1,1,\ldots)\theta_{00}(x_1) \,,\notag \\ V(x_1,1,\ldots) = \eta(x_1)^{-1} \,. \end{gather*} This is modular of weight $-1/2$ as desired. Now fix an index set $M=(m_2,m_3,\ldots)$ with ${\rm wt}(M)>0$ and assume $A_K(\tau_1)$ is quasimodular of weight ${\rm wt}(K)-1/2$ for all $K$ with ${\rm wt}(K)<{\rm wt}(M)$. Consider the coefficient $B_M(\tau_1)$ of $\tau^M$ in the expansion of $\Omega(\tau)$ \eqref{413}. We know that $B_M$ is quasimodular of weight ${\rm wt}(M)$. It is a sum of terms \eqref{416}, of which all but one involve $A_K$ with ${\rm wt}(K)<{\rm wt}(M)$ and are quasimodular of weight ${\rm wt}(M)$ by our inductive hypothesis. The one remaining term, which appears with coefficient $(M!)^{-1}$, is $A_M(\tau_1)\theta_{00}(\tau_1)$ (take $J=0=m,\ P=1,\ Q_i = \tau_i^{m_i}$.) It follows that $A_M(\tau_1)$ is quasimodular of weight ${\rm wt}(M)-1/2$ as desired. \section{Preliminaries on partitions}\label{sec:prepart} Let $\lambda = \lambda_1\ge \lambda_2\ge\ldots\ge\lambda_\ell >0=\lambda_{\ell+1}=\ldots$ be a partition. $\ell=\ell(\lambda)$ is the length of the partition, and $|\lambda| = \sum\lambda_i$ is the number being partitioned. Let $f(\lambda)$ be a function on the set of partitions. Assuming $f$ does not grow too rapidly with $|\lambda|$ we may consider for $|q|\ll 1$ the ratio \begin{equation}\label{51} \langle f\rangle_q\ := \frac{\sum_\lambda f(\lambda)q^{|\lambda|}}{\sum_\lambda q^{|\lambda|}} = (q)_\infty \sum_\lambda f(\lambda)q^{|\lambda|}. \end{equation} Here, we write for $q$ given and $n\le \infty$, \begin{equation}\label{52}(a)_n = (1-a)(1-aq)(1-aq^2)\cdots (1-aq^n). \end{equation} Fix an integer $n\ge 1$ and variables $t_1,\ldots,t_n$. We will be interested in the following ``$n$-point correlation function'' \begin{equation}\label{53} F(t_1,\ldots,t_n) := \Big<\prod_{k=1}^n\Big(\sum_{i=1}^\infty t_k^{\lambda_i-i+{\frac{1}{2}}}\Big)\Big>_q. \end{equation} To understand the relation between $F$ and the character $V(q)$, we take a small detour, beginning with some basic ideas from the theory of partitions. The {\it diagram} of a partition $\lambda$ is simply the array of dots (or squares) with $\lambda_1$ dots in the first row, $\lambda_2$ dots in the second row, etc. Here is $4,4,3,2,1$. $$\begin{array}{cccc} \bullet & \bullet & \bullet & \bullet \\ \bullet & \bullet & \bullet & \bullet \\ \bullet & \bullet & \bullet \\ \bullet & \bullet \\ \bullet \end{array} $$ The transposed partition $\lambda'$ is obtained by flipping along the diagonal $$\begin{array}{ccccc} \bullet & \bullet & \bullet & \bullet & \bullet\\ \bullet & \bullet & \bullet & \bullet \\ \bullet & \bullet & \bullet \\ \bullet & \bullet \end{array} $$ For $\lambda_i\ge i$ let $m_i=\lambda_i-i$ be the number of dots to the right of the $i$-th point on the diagonal, and let $n_i$ be the number of points below it. (so $n_i(\lambda)=m_i(\lambda')$.) Note $m_1>m_2>\ldots \ge 0$ and similarly for the $n_i$. The $m_i$ and $n_i$ are called Frobenius coordinates of $\lambda$ and written $(m_i,\ldots,m_p|n_1,\ldots,n_p)$. For example, Frobenius coordinates for $4,4,3,2,1$ are $(3,2,0|4,2,0)$. Frobenius coordinates are related to the $\lambda_i$ via the following generating function identity. \begin{lem}\label{lem51} Let $\lambda$ be a partition of length $\ell$ with Frobenius coordinates\newline $(m_i,\ldots,m_p|n_1,\ldots,n_p)$. Then $$\sum_i (t^{\lambda_i-i+{\frac{1}{2}}}-t^{-i+{\frac{1}{2}}}) = \sum_k (t^{m_k+{\frac{1}{2}}}-t^{-(n_k+{\frac{1}{2}})}). $$ \end{lem} \begin{proof} It is clear that the $m_k$ correspond to the nonnegative $\lambda_i-i$. Thus it will suffice to identify the terms with negative exponents on the two sides. This follows from the identity of sets of numbers $$\{\lambda_{p+1}-(p+1),\dots,\lambda_{\ell}-\ell,-n_1-1,\dots,-n_p-1\} = \{-\ell,\dots,-2,-1\},, $$ which, in turn, follows by comparing the negative elements from the identity $$ \{\lambda_1-1,\dots,\lambda_{\ell}-\ell,-\lambda'_1,\dots,-\lambda'_k + k-1\}=\{-\ell,\dots, k-1\}\,, \forall k \ge \lambda_1\,, $$ established in \cite{MacD}, Chapter I, (1.7). \end{proof} Substituting $t=\exp(u)$ and expanding in the previous lemma yields identities \begin{multline}\label{54} \sum_i \left(\lambda_i-i+{\textstyle \frac{1}{2}}\right)^r +(-1)^{r+1} \left(i-{\textstyle \frac{1}{2}}\right)^r = \\ \sum_j \left(m_j+{\textstyle \frac{1}{2}}\right)^r + (-1)^{r+1}\left(n_j+{\textstyle \frac{1}{2}}\right)^r =: p_r(\lambda). \end{multline} For example, $|\lambda| = p_1(\lambda)$. Recall we have already encountered the $p_r(\lambda)$ in the expansion for the character $V(q)$: \begin{equation} V(q_1,q_2,q_3,\ldots) = q_1^{-\xi(-1)}q_3^{-\xi(-3)}\cdots \sum_\lambda \prod_{r\ge 1}q_r^{p_r(\lambda)}. \end{equation} To see the connection more specifically, write $q_n = \exp(2\pi i \tau_n)$ and consider the differential operator \begin{equation}\label{56} \delta = \delta(u) := u^{-1}+(2\pi i)^{-1}\sum_{n=1}^\infty \frac{u^n}{n!}\frac{\partial}{\partial \tau_n}. \end{equation} \begin{prop} We have with $F$ as in \eqref{53} and $t=\exp(u)$ $$F(t) = \eta(q_1)\delta V|_{\tau_2=\tau_3=\ldots =0}. $$ \end{prop} \begin{proof} Consider the Bernoulli polynomials, defined by \begin{equation}\label{57} \frac{u\exp(xu)}{\exp(u)-1} = \sum_{k=0}^\infty B_k(x)u^k/k!;\quad B_0(x) = 1,\ B_1(x) = x-{\frac{1}{2}} \end{equation} One has (\cite{hida}, theorem 1, p. 43) \begin{equation}\label{58} \xi(1-m) = \frac{-B_m({\frac{1}{2}})}{m}. \end{equation} Combining these identities yields \begin{equation}\label{59} -\sum_{n=1}^\infty \xi(-n)u^n/n! = \frac{1}{t^{\frac{1}{2}} - t^{-{\frac{1}{2}}}}-u^{-1} = \sum_{i=1}^\infty t^{-i+{\frac{1}{2}}} -u^{-1}. \end{equation} On the other hand, using lemma \ref{lem51} and the identity \begin{equation}\label{510} \eta(q_1)^{-1} = q_1^{-\xi(-1)}\sum_\lambda q_1^{|\lambda|}, \end{equation} one computes \begin{multline*}\eta(q_1)\delta V|_{\tau_2=\ldots = 0} = u^{-1} -\sum \xi(-n)u^n/n! + \\ + \Big(\sum_\lambda q_1^{|\lambda|}\Big)^{-1}\sum_\lambda q_1^{|\lambda|}\sum_{i=1}^\infty (t^{\lambda_i - i+{\frac{1}{2}}} - t^{-i+{\frac{1}{2}}} ) = \\ = u^{-1}+\Big(\sum_\lambda q_1^{|\lambda|}\Big)^{-1}\sum_\lambda q_1^{|\lambda|}\sum_{i=1}^\infty t^{\lambda_i - i+{\frac{1}{2}}} - u^{-1} = \\ =\Big<\sum_{i=1}^\infty t^{\lambda_i - i+{\frac{1}{2}}}\Big>_{q_1} = F(t). \end{multline*} \end{proof} More generally, one finds \begin{equation}\label{512} \eta(q_1)\delta(u_1)\circ\ldots\circ\delta(u_n)V|_{\tau_2=\ldots=0} = \Big<\prod_{k=1}^n\Big(\sum_{i=1}^\infty t_k^{\lambda_i-i+{\frac{1}{2}}}\Big)\Big>_{q_1} = F(t_1,\dotsc,t_n). \end{equation} This is also equivalent to the formula \eqref{016a} from the introduction. \section{The formula for correlation functions}\label{sec:results} In this section we give a detailed statement of our results on the generating function $F(t_1\dotsc,t_n)$ defined in \eqref{53}. To simplify notation, we write $\Theta(x)=\Theta(x;q)$ for the following theta function \begin{align}\label{61} \Theta(x) = \T_{11}(x;q)&=\eta^{-3}(q)\sum_{n\in{\Bbb Z}} (-1)^n q^{\frac{(n+1/2)^2}2} x^{n+1/2} \\ &=(q)_\infty^{-2} (x^{1/2}-x^{-1/2}) (q x)_\infty (q/x)_\infty \,. \end{align} The function $\Theta(x)$ is odd, \begin{equation*} \Theta(x^{-1})=-\Theta(x), \end{equation*} and satisfies the difference equation \begin{equation}\label{62} \Theta(q^m x)=(-1)^m q^{-m^2/2} x^{-m} \Theta(x)\,, \quad m\in{\Bbb Z} \,. \end{equation} Define \begin{equation} \Theta^{(k)}(x):= \left(x\d x\right)^k \T_{11}(x;q)\, ,\ \Theta' = \Theta^{(1)},\ \text{\rm etc.} \end{equation} where $x\d x$ is the natural invariant vector field on the group ${\Bbb C}^* = {\Bbb C}\setminus 0$. We shall prove the following \begin{thm}\label{thm61} \begin{equation}\label{64} \vspace{-3 \jot} F(t_1,\dots,t_n)= \sum_{\sigma\in\mathfrak{S}(n)}\, \frac {\displaystyle \det\left( \frac{\displaystyle \Theta^{(j-i+1)}(t_{\sigma(1)}\cdots t_{\sigma(n-j)})}{\displaystyle (j-i+1)!} \right)_{i,j=1}^n} {\displaystyle \Theta(t_{\sigma(1)})\, \Theta(t_{\sigma(1)} t_{\sigma(2)}) \dots \Theta(t_{\sigma(1)}\cdots t_{\sigma(n)})} \end{equation} Here $\sigma$ runs through all permutations $\mathfrak{S}(n)$ of $\{1,\dotsc,n\}$,the matrices in the numerator have size $n\times n$, and we define $1/(-n)!=0$ if $n\ge 1$. \end{thm} In particular, since $\Theta'(1)=1$ and $\Theta''(1)=0$, we have \begin{align*} F(t_1) &=\frac{1}{\Theta(t_1)}\,,\notag\\ F(t_1,t_2) &=\frac{1}{\Theta(t_1 t_2)} \left( \frac{\Theta'(t_1)}{\Theta(t_1)} + \frac{\Theta'(t_2)}{\Theta(t_2)} \right)\,. \notag \end{align*} For $n=3$ the formula \eqref{64} looks as follows: \begin{multline}\label{66} F(t_1,t_2,t_3)=\\ \frac1{\displaystyle\Theta(t_1 t_2 t_3)} \sum_{\sigma\in\mathfrak{S}(3)} \det \left( \begin{array}{rrc} \frac{\displaystyle\Theta'\left(t_{\sigma(1)} t_{\sigma(2)}\right)} {\displaystyle \Theta\phantom{{}'}\left(t_{\sigma(1)} t_{\sigma(2)}\right)}& \frac{\displaystyle 1}{\displaystyle 2} \, \frac{\displaystyle\Theta''\left(t_{\sigma(1)}\right)} {\displaystyle\Theta\phantom{{}''}\left(t_{\sigma(1)}\right)}& \frac{\displaystyle\Theta'''(1)}{\displaystyle 3!}\\ {\displaystyle 1}& \frac{\displaystyle\Theta'\phantom{{}'}\left(t_{\sigma(1)}\right)} {\displaystyle\Theta\phantom{{}''}\left(t_{\sigma(1)}\right)}& 0\\ & {\displaystyle 1}& \Theta'(1) \end{array} \right)\\ = \frac{1}{\Theta(t_1 t_2 t_3)} \left( \sum_{1\le i \ne j \le 3} \frac{\Theta'(t_i)}{\Theta(t_i)} \frac{\Theta'(t_i t_j)}{\Theta(t_i t_j)} - \sum_{i=1}^3 \frac{\Theta''(t_i)}{\Theta(t_i)} + \Theta'''(1) \right) \,. \end{multline} The determinants in \eqref{64}, having only one non-zero diagonal below the main diagonal, have a nice combinatorial expansion, see the formula \eqref{78} below. One has the following simple \begin{lem}\label{lem61a} \begin{equation}\label{th'} \Theta^{(2m+1)}(1;q)=(2m+1)! \sum_{k_1+2k_2+\dots=m} \frac{(-2)^{k_1+k_2+\dots}} {k_1!\, k_2!\, \cdots} \prod_i \left(\frac{G_{2i}(q)}{(2i)!}\right)^{k_i} \,. \end{equation} \end{lem} Observe that the sum in \eqref{th'} is over all partitions $1^{k_1} 2^{k_2} 3^{k_3} \dots$ of the number $m$ and $k_1+k_2+\dots$ is the length of such a partition. In particular, \begin{align*} \Theta^{(3)}(1;q)&= - 6 G_2(q)\,,\\ \Theta^{(5)}(1;q)&= - 10 G_4(q)+ 60 G_2(q)^2\,, \\ \Theta^{(7)}(1;q)&= - 14 G_6(q)+420 G_4(q) G_2(q)- 840 G_2(q)^3\,. \end{align*} This lemma will be proved below. \begin{remark} The equality in Theorem \ref{thm61} may look like an equality of multivalued functions but in fact the only ambiguity is the factor $$ \sqrt{\,t_1\cdots t_n} $$ which appears on both sides of of the theorem and, therefore, can be ignored. \end{remark} \begin{remark} Alternatively, one may consider the $(n+1)\times (n+1)$ matrix \begin{equation}\label{67}\left(\begin{array}{ccccc} \Theta(t_1\cdots t_n) & \Theta'(t_1\cdots t_{n-1}) &\hdots & \frac{1}{n!}\Theta^{(n)}(1) \\ 0 & \Theta(t_1\cdots t_{n-1}) & \hdots & \frac{1}{(n-1)!}\Theta^{(n-1)}(1) \\ \vdots & \vdots & \hdots & \vdots \\ 0 & 0 & \hdots & \Theta'(1) \\ 0 & 0 & \hdots & 0 \end{array}\right) \end{equation} (Note the bottom right hand entry is $\Theta(1)=0$.) This matrix has rank $n$, so there will be a unique column vector $$\begin{pmatrix}v_n(\td{n}) \\ v_{n-1}(\td{n-1}) \\ \vdots \\ v_1(t_1) \\ 1\end{pmatrix} $$ which is killed by multiplication on the left by the matrix \eqref{67}. Then $$F(\td{r}) = \sum_{\sigma} v_r(t_{\sigma(1)},\dotsc,t_{\sigma(r)}). $$ \end{remark} \begin{proof}[Proof of Lemma \ref{lem61a}] Let $f(a)$ be some function of a variable $a$ and let $b$ be another variable. We have: \begin{align*} e^{f(a+b)-f(a)}&=\exp\left(\sum_{m=1}^\infty \frac{f^{(m)}(a)}{m!} b^m\right)\\ &=\prod_{m=1}^\infty \exp\left(\frac{f^{(m)}(a)}{m!} b^m\right)\\ &=\prod_{m=1}^\infty \sum_{k=0}^\infty \left(\frac{f^{(m)}(a)}{m!} \right)^k \frac{b^{mk}}{k!}\\ &=\sum_{s=0}^\infty \sum_{k_1+2k_2+3k_3+\dots=s} \frac{b^s}{k_1!\, k_2!\, k_3!\cdots} \prod_{i=1}^\infty\left(\frac{f^{(i)}(a)}{i!}\right)^{k_i}\,. \end{align*} In other words, we have \begin{multline}\label{67a} e^{-f(a)} \frac{d^s}{d a^s} e^{f(a)} = \\ s! \sum_{k_1+2k_2+3k_3+\dots=s} \frac1{k_1!\, k_2!\, k_3!\cdots} \left(\frac{f'}{1!}\right)^{k_1} \left(\frac{f''}{2!}\right)^{k_2} \left(\frac{f'''}{3!}\right)^{k_3} \cdots \,. \end{multline} We wish to apply the above formula to the triple product formula for the theta function and use the fact that $$ \left.\left(x\frac{d}{dx}\right)^{m} \log\left((qx)_\infty (q/x)_\infty\right) \right|_{x=1}= \begin{cases} -2 G_{m}(q) - B_{m}/m\,, &\text{$m$ is even}\\ 0\,, &\text{$m$ is odd} \,. \end{cases} $$ The derivative $\Theta^{(2m+1)}(1;q)$ is quasi-modular of weight $2m$. It is, therefore, sufficient, to compute the term $$ \left. (2m+1) \left(x\frac{d}{dx} (x^{1/2}-x^{-1/2}) \right) \left(x\frac{d}{dx}\right)^{2m} (qx)_\infty (q/x)_\infty \right|_{x=1}\,, $$ which gives the contribution of the maximal weight. Using \eqref{67a} we obtain that the above term equals the RHS of \eqref{th'} modulo terms of lower weight. This establishes \eqref{th'}. \end{proof} Before diving into the proof of Theorem \ref{thm61} in the next section, we give as a warmup the proof for the $1$-point function. \begin{thm}\label{thm64} $$\Big<\sum_i t^{\lambda_i -i+{\frac{1}{2}}}\Big>_q = \frac{1}{\Theta(t)}. $$ \end{thm} \begin{lem}\label{lem65} $$\Big<t^{\lambda_k}\Big>_q = \frac{(q)_{\infty}}{(1-q)\cdots (1-q^{k-1})(1-q^kt)(1-q^{k+1}t)\cdots } $$ \end{lem} \begin{proof} The notation is as in \eqref{52}. Recall in section (\ref{sec:prepart}) we defined the transpose $\lambda'$ of a partition $\lambda$. The lemma is straightforward once one observes that $$\Big<t^{\lambda_k}\Big>_q = \Big<t^{\lambda_k'}\Big>_q = \Big<t^{\#\{i|\lambda_i\ge k\}}\Big>_q. $$ \end{proof} Recall the following Heine's $q$-analog of the Gauss ${}_2F_1$-summation. Given any $a$, $b$, $c$ satisfying $|c|<|ab|$ we have \begin{equation}\label{69} \sum_{n=0}^\infty \frac{(a)_n (b)_n}{(c)_n (q)_n} \left(\frac{c}{ab}\right)^n = \frac{(c/a)_\infty (c/b)_\infty}{(c)_\infty (c/ab)_\infty}\,, \end{equation} see for example \cite{GR}, Section 1.5. Recall that the symbol $(a)_n$ was defined in \eqref{52}. \begin{proof}[Proof of theorem \ref{thm64}] \begin{multline*} \Big<\sum_i t^{\lambda_i -i+{\frac{1}{2}}}\Big>_q = \sum_{k=1}^\infty t^{{\frac{1}{2}} -k}\frac{(q^k)_\infty}{(q^kt)_\infty} = \\ = t^{-{\frac{1}{2}}}\frac{(q)_\infty}{(qt)_\infty}\sum_{r=0}^\infty t^{-r}\frac{(qt)_i}{(q)_i} = \frac{(q)_\infty^2}{(t^{\frac{1}{2}} - t^{-{\frac{1}{2}}})(qt)_\infty(q/t)_\infty}= \frac{1}{\Theta(t)}. \end{multline*} \end{proof} \section{Beginning of the proof of Theorem \ref{thm61}}\label{sec:not} Our strategy for proving theorem \ref{thm61} will be based on the following simple implication $$ \Bigg( \begin{matrix} \text{\rm $f(x)$ is holomorphic on ${\Bbb C}\setminus 0$ and }\\ f(qx)=f(x) \end{matrix} \Bigg) \Rightarrow \Bigg( \text{\rm $f(x)$ is a constant} \Bigg)\quad . $$ We shall show that both sides of theorem \ref{thm61} satisfy the same difference equation and have the same singularities. We now introduce some useful notation. Given indices $ 1< i_1<\dots<i_n \,. $ set \begin{align}\label{71} \ex{i_1,\dots,i_r}{t_1,\dots,t_r} :&= \left\langle \prod_{k=1}^n t_k^{\lambda_{i_k}-i_k+1/2} \right\rangle_q \\ &= \prod_{k=1}^n t_k^{1/2-i_k} \left\langle \prod_{k=1}^n t_k^{\lambda'_{i_k}} \right\rangle_q \notag \\ &= (q)_\infty \frac {t_1^{1/2-i_1} \dots t_r^{1/2-i_r}} {(q)_{i_1-1} (q^{i_1} t_1)_{i_2-i_1} (q^{i_2} t_1 t_2)_{i_3-i_2} \dots (q^{i_r} t_1 \cdots t_r)_\infty} \,,\notag \end{align} The notation $(a)_n$ is as in \eqref{52}, and the last equality in \eqref{71} is a straightforward generalization of lemma (\ref{lem65}). For brevity, write $$ \ex{i}{t} :=\ex{i_1,\dots,i_r}{t_1,\dots,t_r} \,. $$ Set \begin{align} H(t_1,\dots,t_n)&:=\sum_{i_1<\dots<i_n} \ex{i}{t} \,, \label{72} \\ G(t_1,\dots,t_n)&:=\sum_{\sigma\in\mathfrak{S}(n)} H(t_{\sigma(1)},\dotsc,t_{\sigma(n)}) \,. \label{73} \end{align} By $\P n$ denote the set of all partitions of the set $\{1,\dots,n\}$. An element $\pi$ of $\P n$ $$ \pi=\{\pi_1,\dots,\pi_\ell\} \in \P n\,, \quad \pi_i\subset\{1,\dots,n\}\,, $$ is by definition an unordered collection of subsets of $\{1,\dots,n\}$ such that \begin{align*} &\pi_i \cap \pi_j = \emptyset\,, \quad i\ne j\,, \\ &\pi_1 \cup \dots \cup \pi_\ell = \{1,\dots,n\} \,. \end{align*} Note the difference between partitions of the \emph{set} $\{1,\dots,n\}$ and partitions of the \emph{number} $n$. The number $\ell$ is called the \emph{length} of $\pi$ and is denoted by $\ell(\pi)$. The subsets $\pi_i$ are called the \emph{blocks} of $\pi$. Given a partition $\pi\in\P n$ set $$ G^\pi(t_1,\dots,t_n):=G\left(\prod_{k\in\pi_1} t_k\,, \dots , \prod_{k\in\pi_{\ell(\pi)}} t_k \right) \,. $$ For example, if $n=3$ and $\pi=\{\{1,2\},\{3\}\}$ then $$ G^\pi(t_1,t_2,t_3)=G(t_1 t_2, t_3) \,. $$ With this notation we have \begin{equation}\label{76} F(t_1,\dots,t_n)=\sum_{\pi\in\P n} G^\pi(t) \,. \end{equation} Set also \begin{equation}\label{77} F^\pi(t_1,\dots,t_n):= F\left(\prod_{k\in\pi_1} t_k\,, \dots , \prod_{k\in\pi_{\ell(\pi)}} t_k \right)\,, \quad \pi\in\P n \,. \end{equation} We shall also need the set ${\Bbb G} n$ of all {\it compositions} of the set $\{1,\dots,n\}$. By definition, the set ${\Bbb G} n$ consists of all {\it ordered} collections $$ \gamma=(\gamma_1,\dots,\gamma_\ell)\,, \quad \gamma_i\subset\{1,\dots,n\}\,, $$ such that \begin{align*} &\gamma_i \cap \gamma_j = \emptyset\,, \quad i\ne j\,, \\ &\gamma_1 \cup \dots \cup \gamma_\ell = \{1,\dots,n\} \,. \end{align*} The number $\ell$ is called again the {\it length} of $\gamma$ and is denoted by $\ell(\g)$. The set ${\Bbb G} n$ is in a natural one-to-one correspondence with the set ${\Upsilon}_n$ of all partial flags $$ \emptyset={\upsilon}_1\varsubsetneq{\upsilon}_2\varsubsetneq\dots\varsubsetneq{\upsilon}_\ell \varsubsetneq{\upsilon}_{\ell+1}=\{1,\dots,n\} $$ in the set $\{1,\dots,n\}$, namely, we set \begin{equation}\label{77a} {\upsilon}(\gamma)_k=\gamma_1 \cup \dots \cup \gamma_{k-1}\,. \end{equation} For any partition $\pi$ or composition $\gamma$ of an $n$-element set of length $\ell$ define its sign as $$ (-1)^\pi = (-1)^\gamma = (-1)^{n+\ell}\,. $$ Then, in particular, the sign of a permutation equals the sign of the corresponding partition into disjoint cycles. Denote by $ U(t_1,\dots,t_n) $ the RHS of \eqref{64}, and set \begin{multline*} T(t_1,\dots,t_n):= \Theta(t_1 \dotsc t_n) U(t_1,\dots,t_n) \\ =\sum_{\sigma\in\mathfrak{S}(n)} \frac 1{\Theta(t_{\sigma(1)})\Theta(t_{\sigma(1)} t_{\sigma(2)}) \dotsc \Theta(t_{\sigma(1)}\cdots t_{\sigma(n-1)})}\times \notag \\ \times \det\left( \frac{\Theta^{(j-i+1)}(t_{\sigma(1)}\cdots t_{\sigma(n-j)})}{(j-i+1)!} \right)\,. \notag \end{multline*} Expanding the determinants one obtains \begin{equation}\label{78} T(t_1,\dots,t_n)=\sum_{\gamma\in{\Bbb G} n} (-1)^{\gamma} \Theta^{(\#\gamma_1)}(1) \, \prod_{k=2}^{\ell(\g)} \frac{\Theta^{(\#\gamma_k)}\left(\prod_{i\in{\upsilon}(\gamma)_{k}} t_i\right)} {\Theta\left(\prod_{i\in{\upsilon}(\gamma)_{k}} t_i\right)}\,, \end{equation} where we used the notation \eqref{77a} for the partial sums of a composition $\gamma$. Define the functions $T^\pi(t)$, $\pi\in\P n$, as in \eqref{77}. We will use two subsets of $\P n$. By $\Po n$ denote the set of all partitions $\pi$ such that $\pi$ has at most one block of cardinality $>1$ which, in addition, contains the number 1. By $\Pt n$ denote the set of all partitions $\pi$ that have $\{1\}$ as a block. We have, for example, \begin{gather*} \{\{1,2\},\{3\}\}\in\Po 3\,,\\ \{\{1\},\{2,3\}\}\in\Pt 3\,. \end{gather*} Denote by $$ \text{Atom}({\{1,\dots,n\}})=\{\{1\},\dotsc,\{n\}\} \in \P n $$ the partition into 1-element subsets. It is clear that $$ \Po n \cap \Pt n = \text{Atom}({\{1,\dots,n\}})\,. $$ Define the subsets $\Go n,\Gt n \subset {\Bbb G} n$ in the same way as for partitions of $\{1,\dots,n\}$. We shall need the following identities \begin{lem} \begin{gather}\label{710} \sum_{\pi\in\P n} (-1)^{\pi} \ell(\pi) ! = \sum_{\gamma\in{\Bbb G} n} (-1)^{\gamma} = 1 \\ \label{711} \sum_{\pi\in\P n} (-1)^{\ell(\pi)} (\ell(\pi)-1) ! = 0 \,. \end{gather} \end{lem} \begin{proof} We shall prove \eqref{710}. The proof of \eqref{711} is similar. Define $f: \prod_n \to \prod_{n-1}$ in the obvious way, by simply omitting $n$ from a partition. We argue by induction on $n$. Note $$\# f^{-1}(\pi) = \ell(\pi)+1. $$ More precisely, $f^{-1}(\pi)$ contains $\ell(\pi)$ partitions of length $\ell(\pi)$ and one partition of length $\ell(\pi)+1$. We compute \begin{multline*} \sum_{\pi\in\prod_n} (-1)^{n+\ell(\pi)}\ell(\pi)! = \\ \sum_{\pi\in\prod_{n-1}} (-1)^n\Big((-1)^{\ell(\pi)} \ell(\pi) \ell(\pi)!+(-1)^{\ell(\pi)+1}(\ell(\pi)+1)!\Big) = \\ \sum_{\pi\in\prod_{n-1}} (-1)^{n-1+\ell(\pi)}\ell(\pi)! = 1. \end{multline*} \end{proof} \section{Difference equations for the correlation functions}\label{sec:difF} The aim of this section is the following \begin{thm}\label{thm81} $$ F(q t_1, t_2,\dots,t_n)= -q^{1/2} t_1\cdots t_n \left( \sum_{\pi\in\Po n} (-1)^{\pi} F^\pi (t_1,\dots,t_n) \right) \,. $$ \end{thm} Recall that, by definition, \begin{multline*} \sum_{\pi\in\Po n} (-1)^{\pi} F^\pi (t_1,\dots,t_n)=\\ \sum_{s=0}^{n-1} (-1)^s \sum_{1<i_1<\dots < i_s \le n} F(t_1 t_{i_1} t_{i_2} \cdots t_{i_s}, \dots,\widehat{\,t_{i_1}}, \dots,\widehat{\,t_{i_s}}, \dots) \,.\notag \end{multline*} We begin with some auxiliary propositions. \begin{prop}\label{prop82} For any $k=1,\dots,n$ \begin{align*} H(t_1,\dots,qt_k,\dots,t_n)= &-q^{1/2} t_1 \cdots t_n H(t_1,\dots,t_n) \notag\\ &+ \frac{q t_k}{1-q t_k} H(t_1,\dots,t_{k-1}, qt_k t_{k+1}, t_{k+2},\dots,t_n) \\ &- \frac{1}{1-q t_k} H(t_1,\dots,t_{k-2}, qt_{k-1} t_{k}, t_{k+1},\dots,t_n) \,.\notag \end{align*} Here if $k=1$ (or $k=n$) the third (second) summand should be omitted. \end{prop} \begin{cor}\label{cor83} \begin{align} G(q t_1, t_2,\dots,t_n)=& -q^{1/2} t_1\cdots t_n G(t_1,\dots,t_n) \label{82} \\ &-\sum_{k=2}^n G(q t_1 t_k,t_2,\dots,t_{k-1},t_{k+1},\dots, t_n)\notag \\ =& -q^{1/2} t_1\cdots t_n \left( \sum_{\pi\in\Po n} (-1)^{\pi} (n-\ell(\pi))! \, \, G^\pi (t) \right) \,. \label{83} \end{align} \end{cor} \begin{lem}\label{lem84} Let $u$ and $v$ be two variables. For any integers $a < b$ \begin{equation} \sum_{a < i < b} \frac{(q v)^{1/2-i}} {(q^a u)_{i-a} (q^i u v)_{b-i}} = \frac{1}{1-q v} \left( \frac{(q v)^{3/2-b}}{(q^a u)_{b-a-1}} - \frac{(q v)^{1/2-a}}{(q^{a+1} u v )_{b-a-1}} \right) \,. \end{equation} In particular, for $a=1$ and $u=1$ we obtain \begin{equation} \sum_{1 \le i < b} \frac{(q v)^{1/2-i}} {(q)_{i-1} (q^i v)_{b-i}} = \frac{1}{1-q v} \frac{(q v)^{3/2-b}}{(q)_{b-2}} \,. \end{equation} \end{lem} \begin{proof} Follows from the following particular case of the $q$-Gauss summation formula \eqref{69}: for any $c$ and $d$ one has \begin{equation} \sum_{i=c}^{\infty} (q v)^{1/2 - i} \frac{(q^d u v)_{i-c}}{(q^d u)_{i-c}}= -(q v)^{3/2-c} \frac{1-q^{d-1} u}{1- q v} \,. \end{equation} \end{proof} In the rest of the section we shall write $a\ll b$ if $a+1<b$. \begin{lem}\label{lem85} We have \begin{align} &H(t_1,\dots,qt_k,\dots,t_n)= \notag \\ &\quad= \sum_{i_1<\dots < i_{k} \ll i_{k+1} < \dots < i_n} q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k}) t_{k+1} \cdots t_n \ex{i_1,\dots,i_n}{t_1,\dots,t_n}\label{88} \\ &\quad= \sum_{i_1<\dots < i_{k-1} \ll i_{k} < \dots < i_n} q^{3/2-i_k} (1-q^{i_k-1} t_1 \cdots t_{k-1}) t_k \cdots t_n \ex{i_1,\dots,i_n}{t_1,\dots,t_n} \label{89} \end{align} \end{lem} \begin{proof} Recall that by definition $$ H(t_1,\dots,qt_k,\dots,t_n) = \sum_{i_1<\dots<i_n} \ex{i_1,\dots,i_n}{t_1,\dots,qt_k,\dots,t_n} \,. $$ We have the two following identities: \begin{align} &\ex{i_1,\dots,i_n}{t_1,\dots,qt_k,\dots,t_n} = \notag \\ &\qquad= q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k}) t_{k+1} \cdots t_n \ex{i_1,\dots,i_k,i_{k+1}+1,\dots,i_n+1}{t_1,\dots,t_n} \label{810} \\ &\qquad= q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k-1}) t_k \cdots t_n \ex{i_1,\dots,i_{k-1},i_{k}+1,\dots,i_n+1}{t_1,\dots,t_n} \label{811} \end{align} Replacing in \eqref{810} the indices $i_{k+1}+1,\dots,i_n+1$ by $i_{k+1},\dots,i_n$ we obtain \eqref{88}. Replacing in \eqref{811} the indices $i_{k}+1,\dots,i_n+1$ by $i_{k},\dots,i_n$ we obtain \eqref{89}. \end{proof} \begin{proof}[Proof of proposition \eqref{prop82}] Consider the general case $1 < k < n$. The two other cases $k=1$ and $k=n$ are similar. Rewrite the sum \eqref{88} as follows \begin{multline}\label{812} \sum_{i_1<\dots < i_{k} \ll i_{k+1} < \dots < i_n} q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k}) t_{k+1} \cdots t_n \ex{i_1,\dots,i_n}{t_1,\dots,t_n} \\ = \sum_{i_1< \dots < i_n} (\dots) - \sum_{\stackrel{i_1<\dots < i_{k} \ll i_{k+2} < \dots < i_n}{ i_{k+1}=i_k+1}} (\dots)\,, \end{multline} and denote the two summands in the RHS of \eqref{812} by $\s1$ and $\s2$. The formula \eqref{88} together with the following identity \begin{multline*} q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k}) t_{k+1} \cdots t_n \ex{i_1,\dots,i_k,i_k+1,i_{k+2},\dots,i_n}{t_1,\dots,t_n} = \\ q^{1/2-i_k} (1-q^{i_k} t_1 \cdots t_{k+1}) t_{k+2} \cdots t_n \ex{i_1,\dots,i_{k-1},i_k,&i_{k+2},\dots,i_n}{t_1,\dots,t_{k-1},t_k t_{k+1},&t_{k+2}, \dots,t_n} \end{multline*} implies that $$ \s2=H(t_1,\dots,q t_k t_{k+1}, \dots, t_n) \,. $$ Set \begin{align} \s3:&=\s1 + q^{1/2} t_1\cdots t_n H(t_1,\dots,t_n) \notag \\ &=t_{k+1}\cdots t_n \sum_{i_1<\dots<i_n} q^{1/2-i_k} {\ex{i}{t} }\ \,. \label{814} \end{align} Using Lemma \ref{lem84} we can evaluate one of the nested sums in \eqref{814} as follows: \begin{multline*} \sum_{i_{k-1} < i_k < i_{k+1}} \frac {(qt_k)^{1/2-i_k}} {(q^{i_{k-1}} t_1 \cdots t_{k-1})_{i_k-i_{k-1}} (q^{i_{k}} t_1 \cdots t_{k})_{i_{k+1}-i_{k}}} \\= \frac1{1-q t_k} \left( \frac{(qt_k)^{3/2-i_{k+1}}} {(q^{i_{k-1}} t_1 \cdots t_{k-1})_{i_{k+1}-i_{k-1}-1}} - \frac{(qt_k)^{1/2-i_{k-1}}} {(q^{i_{k-1}+1} t_1 \cdots t_{k})_{i_{k+1}-i_{k-1}-1}} \right) \,. \end{multline*} Then sum $\s3$ becomes the following difference \begin{align*} &\frac{t_k \cdots t_n}{1-q t_k} \sum_{\dots<i_{k-1} \ll i_{k+1} < \dots} q^{3/2-i_{k+1}} (1- q^{i_{k+1}-1}t_1\cdots t_{k-1}) \ex{\dots,i_{k-1},i_{k+1},\dots}{ \dots,t_{k-1}, t_k t_{k+1},\dots} \\ &-\frac{t_{k+1} \cdots t_n}{1-q t_k} \sum_{\dots<i_{k-1} \ll i_{k+1} < \dots} q^{1/2-i_{k-1}} (1- q^{i_{k-1}}t_1\cdots t_{k}) \ex{\dots,i_{k-1},i_{k+1},\dots}{ \dots,t_{k-1}t_k, t_{k+1}, \dots} \,. \end{align*} By Lemma \ref{lem85} this yields \begin{multline*} \s3= \frac1{1-qt_k} H(t_1,\dots,t_{k-1},q t_k t_{k+1}, \dots, t_n) \\ - \frac1{1-qt_k} H(t_1,\dots,q t_{k-1}t_k, t_{k+1}, \dots, t_n) \,. \end{multline*} Thus, \begin{align*} H(t_1,\dots,qt_k,\dots,t_n)=& \s3 - q^{1/2} t_1 \cdots t_n H(t_1,\dots,t_n) -\s2 \\ =&-q^{1/2} t_1 \cdots t_n H(t_1,\dots,t_n)\\ &+ \frac{q t_k}{1-q t_k} H(t_1,\dots,t_{k-1}, qt_k t_{k+1}, t_{k+2},\dots,t_n) \\ &- \frac{1}{1-q t_k} H(t_1,\dots,t_{k-2}, qt_{k-1} t_{k}, t_{k+1},\dots,t_n) \,. \end{align*} This concludes the proof. \end{proof} \begin{proof}[Proof of corollary \ref{cor83}] We will prove \eqref{82}; equation \eqref{83} follows easily from \eqref{82}. Substitute the formula from proposition \ref{prop82} into \eqref{73} and consider the coefficient of a summand $$ H(\dots, q t_1 t_j, \dots)\,, \quad j=2,\dots,n \,. $$ Such a summand arises from the expansion of $$ H(\dots, q t_1, t_j, \dots)\quad\text{\rm\ and} \quad H(\dots, t_j, q t_1, \dots) \,. $$ The sum of the coefficients equals $$ \frac{q t_1}{1- q t_1} - \frac{1}{1- q t_1} = -1 \,. $$ This proves \eqref{82}. \end{proof} \begin{prop}\label{prop86} \begin{equation}\label{815} F(q t_1, t_2,\dots,t_n)= -q^{1/2} t_1\cdots t_n \left( \sum_{\pi\in\Pt n} G^\pi(t) \right) \,. \end{equation} \end{prop} \begin{proof} Substitute \eqref{83} into \eqref{76}. Let $c_\pi$ denote the coefficient of the summand $G^\pi(t)$ in this expansion. It is clear that $$ c_\pi = -q^{1/2} t_1\cdots t_n \,, \pi \in \Pt n \,. $$ Suppose that $\pi \notin \Pt n$ and show that $c_\pi=0$. Since $F(q t_1, t_2,\dots,t_n)$ is symmetric in variables $t_2,\dots,t_n$ we can assume that $$ \pi=\{\{1,2,\dots,m\},\pi_2,\dots,\pi_l\}\,, \quad 1<m\le n\,, $$ where $\{\pi_2,\dots,\pi_l\}$ is a partition of the set $\{m+1,\dots,n\}$. Let $\sigma$ be a partition of $\{1,\dots,m\}$ $$ \sigma=\{\sigma_1,\dots,\sigma_s\} \in \P m \,. $$ Set $$ \pi\land\sigma=\{\sigma_1,\dots,\sigma_s,\pi_2,\dots,\pi_l\} \in \P n \,. $$ It is easy to see that the term $G^\pi(t)$ arises precisely from the expansion of the terms of the form $$ G^{\pi\land\sigma} (qt_1,t_2,\dots,t_n)\,, \quad \sigma\in\P m \,. $$ By \eqref{83} it arises with the coefficient $$ q^{1/2} t_1\cdots t_n (-1)^{\ell(\sigma)} (\ell(\sigma)-1)! \,. $$ By \eqref{711}, the sum of these coefficients over all $\sigma\in\P m$ equals zero. This concludes the proof. \end{proof} \begin{proof}[Proof of theorem \ref{thm81}] We will show that the RHS in theorem \ref{thm81} equals the RHS of \eqref{815}. Substitute \eqref{76} into RHS of theorem \ref{thm81} and let $c_\pi$ denote the coefficient of the summand $G^\pi(t)$ in this expansion. Again, it is clear that $$ c_\pi = -q^{1/2} t_1\cdots t_n \,, \pi \in \Pt n \,. $$ and we have to show that $$ c_\pi = 0 \,, \pi \notin \Pt n \,. $$ Again, we can suppose that $$ \pi=\{\{1,2,\dots,m\},\pi_2,\dots,\pi_l\}\,, \quad 1<m\le n\,, $$ where $\{\pi_2,\dots,\pi_l\}$ is a partition of the set $\{m+1,\dots,n\}$. For any set $S$ denote by $\text{Atom}(S)$ the partition of $S$ into one-element blocks. The summand $G^\pi(t)$ arises in the RHS of from the expansion of terms $$ F^{\left\{\{S\},\text{Atom}(\{1,\dots,n\}\setminus S)\right\}}(t)\,, \quad 1\in S \subset \{1,\dots,m\}\,, $$ and therefore the coefficient $c_\pi$ equals \begin{align*} c_\pi&=-q^{1/2} t_1\cdots t_n \sum_{1\in S \subset \{1,\dots,m\}} (-1)^{|S|+1} \\ &=-q^{1/2} t_1\cdots t_n \sum_{k=0}^{m-1} (-1)^k \binom{m-1}k \\ &=0 \,. \end{align*} This concludes the proof. \end{proof} \section{Singularities of correlation functions}\label{sec:sing} In this section we consider the singularities of the function $F(t_1,\dots,t_n)$. The series \eqref{76} converges uniformly on compact subsets of the set $$ \left\{|t_i|>1\,, i=1,\dots,n\right\}\setminus \bigcup_{S\subset\{1,\dots,n\}} \bigcup_{m=1}^\infty \left\{q^m \prod_{k\in S} t_k =1 \right\} $$ and has simple poles on the divisors $$ \left\{q^m \prod_{k\in S} t_k =1 \right\}\,, \quad S\subset\{1,\dots,n\} \,. $$ The difference equation from theorem \ref{thm81} gives a meromorphic continuation of the function $F(t)$ onto (a double cover of) the domain $({\Bbb C}\setminus 0)^n$. By symmetry, it suffices to consider the divisors \begin{equation}\label{91} q^m t_1 t_2 \dots t_k = 1\,, \quad k\le n,\quad m=1,2,\dots \,. \end{equation} We shall prove the following \begin{thm}\label{thm91} We have \begin{equation}\label{92} F(t_1,\dots,t_n)=(-1)^{m} \frac{q^{m^2/2} m^{k-1}}{q^m t_1\dots t_k-1} \frac{F(t_{k+1},\dots,t_n)}{(t_{k+1}\dots t_n)^m}+ \dots \,, \end{equation} where dots stand for terms regular at the divisor \eqref{91} and we assume that $$ F(t_{k+1},\dots,t_n)=1\,,\quad k=n\,. $$ \end{thm} A curious property of \eqref{92} is that the residue \begin{equation}\label{93} \operatorname{Res}_{\{q^m t_1 t_2 \dots t_k = 1\}} F(t_1,\dots,t_n) \end{equation} does not depend on the variables $t_1,\dots,t_k$. Introduce the following notation. Given a function $f(i_1,\dots,i_n)$ set $$ \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} f(i_1,\dots,i_n) = \sum_{1\le i_1 \le \dots \le i_n \le m} \frac{f(i_1,\dots,i_n)} {\# \operatorname{Stab}_{S(n)} \{i_1,\dots,i_n\}} \,. $$ If the function $f$ is symmetric, we have \begin{equation}\label{94} \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} f(i_1,\dots,i_n) = \frac 1{n!} \sum_{i_1,\dots,i_n=1}^m f(i_1,\dots,i_n) \,. \end{equation} With this notation we have $$ F(t)= \sum_{\stackrel{\text{\rm over all permutations}}{ \text{\rm of $t_1,\dots,t_n$}}} \,\, \widetilde{\sum_{1\le i_1 \le \dots \le i_n}} \ex{i}{t} \,. $$ It is clear that the evaluation of the residue \eqref{93} boils down to the evaluation of the following sum \begin{multline}\label{95} \sum_{\stackrel{\text{\rm permutations}}{\text{\rm of $t_1,\dots,t_k$}}} \\ \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} \frac{t_1^{1/2-i_1} \dots t_k^{1/2-i_k}} {(q)_{i_1-1} (q^{i_1} t_1)_{i_2-i_1} (q^{i_2} t_1 t_2)_{i_3-i_2} \dots (q^{i_k} t_1 \dots t_k)_{m-i_k}} \,, \end{multline} where the variables $t_1,\dots,t_k$ are subject to constraint $$ q^m t_1 \cdots t_k =1 \,. $$ This evaluation follows from a curious rational function identity which shall now be established. Set $$ \cbm {i_1,\dots,i_k}{t_1,\dots,t_k}:= \frac1 {(q)_{i_1-1} (q^{i_1} t_1)_{i_2-i_1} (q^{i_2} t_1 t_2)_{i_3-i_2} \dots (q^{i_k} t_1 \dots t_k)_{m-i_k}} \,. $$ Then we have the following \begin{thm}\label{thm92} Let $m,k\in{\Bbb N}$ and suppose that $$ q^m t_1 t_2 \dots t_k = 1\,. $$ Then \begin{equation}\label{96} \scy\quad \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} \cbm {i_1,\dots,i_k}{t_1,\dots,t_k} = \frac{m^{k-1}}{(k-1)!} \,. \end{equation} In particular, this sum is independent of variables $q,t_1,\dots,t_k$. \end{thm} \begin{proof}[Proof of Theorem \ref{thm92}] Let us use $q,t_1,\dots,t_{k-1}$ as independent variables. First, show that LHS of \eqref{96} does not depend on $t_1,\dots,t_{k-1}$. Since it is a rational function bounded on infinity it suffices to show that it does not have any poles. The only poles it can possibly have are simple poles on divisors $$ t_s t_{s+1} \dots t_{r-1} t_r = q^{-l}\,, \quad s\le r\le k-1,\, 1\le l \le m \,. $$ By symmetry, it suffices to consider the pole on \begin{equation}\label{97} t_1 \dots t_r = q^{-l}\,, \quad \quad r\le k-1,\, 1\le l \le m \,. \end{equation} This pole arises from two type of summands in \eqref{96}. First, it arises from the summands \begin{equation}\label{98} \cbm {i_1,\dots,i_k}{t_1,\dots,t_k}\,, \quad i_r \le l < i_{r+1}\,, \end{equation} and it also arises from summands \begin{equation}\label{99} \cbm {i_1,\dots,i_k}{t_{r+1},\dots,t_k,t_1,\dots,t_r}\,, \quad i_{k-r} \le m-l < i_{k-r+1}\,. \end{equation} We shall match each summand of the form \eqref{98} to a summand of the form \eqref{99} in such a way that the poles in each pair will cancel out. Namely, it is easy to see that the sum \begin{equation}\label{910} \cbm {i_1,\dots,i_k}{t_1,\dots,t_k} \,+\, \cbm {i_{r+1}-l,\dots,i_k-l,i_1+m-l,\dots,i_r+m-l} {t_{r+1},\dots,t_k,t_1,\dots,t_r}\,, \end{equation} where $$ 1\le i_1\le \dots \le i_r \le l < i_{r+1}\le \dots \le i_k \le m\,, $$ is regular on the divisor \eqref{97}; notice that the inequalities $$ 1\le i_1 \le \dots \le i_r < i_{r+1} \le \dots \le i_k \le m $$ imply that $$ \#\operatorname{Stab}\{i_1,\dots,i_k\}= \#\operatorname{Stab}\{i_{r+1}-l,\dots,i_k-l,i_1+m-l,\dots,i_r+m-l\}\,. $$ This proves that the LHS of \eqref{96} is regular on \eqref{97}. Therefore, the LHS of \eqref{96} does not depend on $t_1,\dots,t_k$ and we can substitute $$ t_1=t_2=\dots=t_{k-1}=1\,,\quad t_k=q^{-m} \,. $$ Then the LHS of \eqref{96} becomes \begin{equation}\label{911} \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} \,\, \sum_{r=1}^k \frac1{(q)_{i_r} (q^{i_r-m})_{m-i_r}} \,. \end{equation} Since the inner sum in \eqref{911} is symmetric in $i_1,\dots,i_k$ we can use \eqref{94} to obtain \begin{align*} &\frac1{k!} \sum_{i_1,\dots,i_k=1}^m \sum_{r=1}^k \frac1{(q)_{i_r} (q^{i_r-m})_{m-i_r}} \notag\\ &= \frac{m^{k-1}}{(k-1)!} \sum_{i=1}^m \frac1{(q)_{i} (q^{i-m})_{m-i}} \notag\\ &= \frac{m^{k-1}}{(k-1)!} \frac1{(q^{1-m})_{m-1}} \sum_{i=1}^m \frac{(q^{1-m})_{i-1}}{(q)_{i}} \notag\\ &=\frac{m^{k-1}}{(k-1)!} \,,\notag \end{align*} where the last equality follows from the q-binomial theorem. This concludes the proof. \end{proof} Replacing in \eqref{96} the variables $q,t_1,\dots,t_k$ by their reciprocals one obtains the following \begin{cor}\label{cor93} Suppose that $q^m t_1 t_2 \dots t_k = 1$. Then the sum \begin{equation}\label{912} \scy \widetilde{\sum_{1\le i_1 \le \dots \le i_k \le m}} \frac{t_1^{1/2-i_1} \dots t_k^{1/2-i_k}} {(q)_{i_1-1} (q^{i_1} t_1)_{i_2-i_1} (q^{i_2} t_1 t_2)_{i_3-i_2} \dots (q^{i_k} t_1 \dots t_k)_{m-i_k}} \end{equation} equals $$ (-1)^{m-1} \frac{q^{m^2/2} m^{k-1}}{(k-1)!}\,, $$ for all $m$ and $k$. \end{cor} This corollary gives the evaluation of the sum \eqref{95} and this proves \eqref{92} \section{Difference equations for the RHS in \eqref{64}}\label{sec:dif} In this section we show that the functions $U(t_1,\dots,t_n)$ satisfy the very same difference equations as the functions $F(t)$ do. Since we shall use only the difference equation \eqref{62} for the theta function $\Theta(x)$, let us consider following general situation. Let $t_0$ be an auxiliary variable. We shall eventually let $$ t_0\to 1 \,. $$ Modify temporarily the definition \eqref{77a} as follows \begin{equation}\label{101} {\upsilon}(\gamma)_k=\{0\}\cup \gamma_1 \cup \dots \cup \gamma_{k-1}\,. \end{equation} Suppose that a function $$ r(x;m)\,, \quad m=0,1,2,\dots\,, $$ satisfies the two following properties: \begin{equation}\label{102} r(x;0)=1 \end{equation} identically and \begin{equation}\label{103} r(qx;m)=\sum_{i=0}^m (-1)^i \binom mi\, r(x;m-i) \,. \end{equation} Consider the function \begin{equation}\label{104} R(t_1,\dots,t_n\,|\, t_0)=\sum_{\gamma\in{\Bbb G} n} (-1)^{\gamma} \prod_{k=1}^{\ell(\g)} r\left({{\textstyle \prod}}_{i\in{\upsilon}(\gamma)_{k}} t_i \,; \#\gamma_k \right) \,, \end{equation} where we use the modified definition \eqref{101}. We want to show that \begin{thm}\label{thm101} Suppose a function $r(x;m)$ satisfies \eqref{102} and \eqref{103}. Then \begin{equation}\label{105} R(qt_1,t_2,\dots,t_n\,|\, t_0)= \sum_{\pi\in\Po n} (-1)^{\pi} R^\pi(t_1,\dots,t_n\,|\, t_0) \,, \end{equation} where the function $R(t_1,\dots,t_n\,|\, t_0)$ was defined in \eqref{104} \end{thm} Here, as usual, \begin{equation}\label{106} R^\pi(t_1,\dots,t_n\,|\, t_0):= R\left(\left.\prod_{k\in\pi_1} t_k\,, \dots , \prod_{k\in\pi_{\ell(\pi)}} t_k \,\right|\, t_0 \right)\,, \quad \pi\in\P n \,. \end{equation} \begin{proof} Given a composition $$ \gamma=(\gamma_1,\dots,\gamma_l)\,, $$ let $m=m(\gamma)$ denote the number of the block that contains 1. Substitute \eqref{103} into \eqref{104}. We obtain \begin{equation}\label{107} R(qt_1,t_2,\dots,t_n\,|\, t_0)= \sum_{\gamma\in{\Bbb G} n} \sum_{s_{m+1}=0}^{\#\gamma_{m+1}} \dots \sum_{s_l=0}^{\#\gamma_{l}} \sqb {\gamma_{m+1},\dots,\gamma_l} {s_{m+1},\dots,s_l}\,, \end{equation} where \begin{equation}\label{108} \sqb {\gamma_{m+1},\dots,\gamma_l} {s_{m+1},\dots,s_l} \end{equation} stands for the following product \begin{multline*} (-1)^{n+\ell(\g)+s_{m+1}+\dots+s_{l}} \prod_{k=1}^{m} r \left({{\textstyle \prod}}_{i\in{\upsilon}(\gamma)_{k}} t_i\,; \#\gamma_k \right)\times \\ \prod_{k=m+1}^{l} \binom {\#\gamma_i}{s_i} r \left({{\textstyle \prod}}_{i\in{\upsilon}(\gamma)_{k}} t_i\,; \#\gamma_k - s_i \right) \,. \end{multline*} Let us divide all summands \eqref{108} into 3 following types according to occurrence of certain patters in the sequence $s_{m+1},\dots,s_l$. The summands of the form \begin{equation}\label{109} \sqb {\gamma_{m+1},&\dots,&\gamma_k,&\dots} {\#\gamma_{m+1},&\dots,&\#\gamma_k,&0,\dots,0}\,, \quad m+1\le k \le l \,, \end{equation} will be called type I summands. If \eqref{108} is not of type I then let $k$, $m+1\le k \le l$, be the minimal number such that $$ \big(0 < s_k < \#\gamma_k\big)\quad \text{\rm\ or}\quad \big(s_k=0\text{\rm\ and }s_{k+1}=\#\gamma_{k+1}\big) \,. $$ We shall say that \eqref{108} is of type II (type III) if the first (second) parenthesis contains a true statement. We shall show that the the type II summands cancel with the type III summand while the type I summands produce the RHS of \eqref{105}. The cancelation of the type II and type III summands follows from the following identity. Suppose that $$ 0 < s_k < \#\gamma_k\,. $$ Then \begin{multline}\label{1010} \sqb{\dots,\gamma_k,\dots} {\dots,s_k,\dots} + \\ \sum_{\stackrel{\delta \subset \gamma_k}{ \#\delta=s_k}} \left[\begin{matrix} \gamma_1,\dots,\gamma_m\\{}\end{matrix} \left|\begin{matrix} \dots,&\gamma_k\setminus\delta,&\delta,&\dots \\ \dots,&0,&\#\delta,&\dots \end{matrix} \right. \right] = 0 \,. \end{multline} To see \eqref{1010} notice that all $\binom {\gamma_k}{s_k}$ summands in the sum over subsets $\delta\subset \gamma_k$ are equal and proportional to the first summand in \eqref{1010}. Now consider a type I summand \eqref{109}. Set \begin{equation}\label{1011} \delta=\gamma_{m+1}\cup \dots \cup \gamma_k \,. \end{equation} We want to fix a subset $$ \delta\subset \{2,\dots,n\} $$ and compute the sum of all type I summands \eqref{109} satisfying \eqref{1011}. Let us consider the nontrivial case $$ \delta\ne\emptyset \,. $$ First sum over all $$ (\gamma_{m+1}, \dots, \gamma_k)\in\Gamma(\delta)\,, $$ where $\Gamma(\delta)$ stands for the set of all compositions of the set $\delta$. We have \begin{multline}\label{1012} \sqb {\gamma_{m+1},&\dots,&\gamma_k,&\gamma_{k+1},\dots} {\#\gamma_{m+1},&\dots,&\#\gamma_k,&0,\dots,0}=\\ (-1)^{k-m+1} \sqb {\delta &\gamma_{k+1},\dots} {\#\delta,&0,\dots,0} \,. \end{multline} Since \eqref{1012} depends only on the parity of number of parts in the composition \eqref{1011} we can use \eqref{710} to obtain \begin{multline*} \sum_{ (\gamma_{m+1},\dots,\gamma_k)\in\Gamma(\delta)} \sqb {\gamma_{m+1},&\dots,&\gamma_k,&\gamma_{k+1},\dots} {\#\gamma_{m+1},&\dots,&\#\gamma_k,&0,\dots,0} = \\ (-1)^{\#\delta+1} \sqb {\delta, &\gamma_{k+1},\dots} {\#\delta,&0,\dots,0} \,. \end{multline*} Now sum over the remaining blocks $$ (\gamma_1,\dots,\gamma_{m},\gamma_{k+1},\dots,\gamma_l)\,. $$ It follows from the definition \eqref{104}, \eqref{106} that \begin{multline*} (-1)^{\#\delta+1} \sum_{ (\gamma_1,\dots,\gamma_{m},\gamma_{k+1},\dots,\gamma_l)} \sqb {\delta, &\gamma_{k+1},\dots} {\#\delta,&0,\dots,0} \\= (-1)^{n+\ell(\sigma)} R^{\sigma} (t\,|\, t_0) \,, \end{multline*} where $\sigma=\sigma(\delta)$ is the following element of $\Po n$ $$ \sigma=\{{1\cup \delta}, \text{Atom}(\{2,\dots,n\}\setminus \delta)\} \,. $$ Recall that $\text{Atom}(\{2,\dots,n\}\setminus \delta)$ denotes the partition of the set $\{2,\dots,n\}\setminus \delta$ into 1-element blocks. Thus, the sum of all type I summands in \eqref{107} equals the RHS of \eqref{105}. This concludes the proof of the theorem. \end{proof} \begin{cor}\label{cor102} \begin{equation}\label{1014} T(qt_1,t_2,\dots,t_n)= \sum_{\pi\in\Po n} (-1)^{n+\ell(\pi)} T^\pi(t_1,\dots,t_n) \,. \end{equation} \end{cor} \begin{proof} Take $$ r(x;m)=\frac{\Theta^{(m)}(x)}{\Theta(x)} \,. $$ By virtue of \eqref{62} this function satisfies \eqref{103} and it obviously satisfies \eqref{102}. We have $$ T(t_1,\dots,t_n)=\operatorname{Res}_{\,t_0=1} R(t_1,\dots,t_n\,|\, t_0) \,. $$ Taking the residue in \eqref{105} we obtain \eqref{1014}. \end{proof} \begin{cor}\label{cor103} \begin{equation}\label{1015} {U(qt_1,t_2,\dots,t_n)} = - q^{1/2} t_1 \cdots t_n \left( \sum_{\pi\in\Po n} (-1)^{n+\ell(\pi)} {U^\pi(t_1,\dots,t_n)} \right) \,. \end{equation} \end{cor} \section{Singularities of the RHS in \eqref{64}}\label{sec:Using} In this section we shall prove that $U(t_1,\dots,t_n)$ has exactly the same singularities as $F(t_1,\dots,t_n)$. \begin{thm}\label{thm111} For $k=1,\dots,n$ we have \begin{equation}\label{111} U(t_1,\dots,t_n)=(-1)^{m} \frac{q^{m^2/2} m^{k-1}}{q^m t_1\cdots t_k-1} \frac{U(t_{k+1},\dots,t_n)}{(t_{k+1}\cdots t_n)^m}+ \dots \,, \end{equation} where dots stand for terms regular at the divisor $$ q^m t_1\cdots t_k=1 $$ and we assume that $$ U(t_{k+1},\dots,t_n)=1\,,\quad k=n\,. $$ \end{thm} Let us again point out that \eqref{111} implies that the residue $$ \operatorname{Res}_{q^m t_1\dots t_k=1} U(t_1,\dots,t_n) $$ is independent of $t_1,\cdots,t_k$. In the proof we shall use a curious identity good for any odd smooth function which we shall state as a separate Theorem \ref{thm113}. It is clear that since the function $U(t)$ satisfies the very same difference equation as $F(t)$ does it suffices to consider the case $$ m=0 \,. $$ That is, we have to show that \begin{equation}\label{112} U(t_1,\dots,t_n)=\frac1{t_1-1} U(t_2,\dots,t_n)+ \dots\,, \quad t_1\to 1 \,, \end{equation} and that \begin{equation}\label{113} \text{\rm $U(t)$ is regular on $\{t_1\cdots t_k=1\}$ for $1<k\le n$ } \end{equation} It follows from the very definition of the function $U(t)$ (look, for example, at the formula \eqref{66}) that \begin{align*} U(t_1,\dots,t_n)&=U(t_1,\dots,t_k) U(t_{k+1}\dots t_n) + \dots \\ &=\frac{T(t_1,\dots,t_k)}{t_1 \cdots t_k -1} U(t_{k+1}\dots t_n) + \dots\,, \end{align*} where dots stand for terms regular on the divisor $$ t_1\cdots t_k=1 \,. $$ Since $$ T(1)=1 \,. $$ the Theorem 4.1 will follow from the following \begin{thm}\label{thm112} We have \begin{equation}\label{114} T(t_1,\dots,t_n)\Big|_{t_1\cdots t_n=1}=0 \end{equation} provided $n>1$. \end{thm} \begin{proof} Till the end of the proof the variables $t_1,\dots,t_n$ will be always subject to constraint \begin{equation}\label{115} t_1\cdots t_n=1\,. \end{equation} Induct on $n$. Since $$ T(t_1,t_2)= \frac{\Theta'(t_1)}{\Theta(t_1)} + \frac{\Theta'(t_2)}{\Theta(t_2)} $$ the case $n=2$ is clear. First, show that \begin{equation}\label{116} \text{\rm $T(t_1,\dots,t_n)$ is a constant.} \end{equation} By the difference equation \eqref{1014} and the induction hypothesis $$ T(q t_1, t_2,\dots,t_n)=T(t_1,\dots,t_n)+(-1)^{n+1} \,. $$ Similarly $$ T(q t_1, t_2,\dots,t_n)=T(q t_1, q^{-1}t_2,\dots,t_n)+(-1)^{n+1} \,. $$ It follows that $$ T(q t_1, q^{-1}t_2,\dots,t_n)=T(t_1,\dots,t_n) \,. $$ Hence, by virtue of the strategy enunciated at the beginning of section \ref{sec:not}, the claim \eqref{116} will follow from the following claim \begin{equation}\label{117} \text{\rm $T(t_1,\dots,t_n)$ is regular}\,, \end{equation} which will now be established. By symmetry it suffices to show that $T(t_1,\dots,t_n)$ is regular on the divisor \begin{equation}\label{118} t_1 \dots t_k = 1\,, \quad 1\le k < n-2 \,. \end{equation} Again, from the definition \eqref{78} it is clear that \begin{multline}\label{119} T(t_1,\dots,t_n)=\\ \left(\frac1{\Theta(t_1\dots t_k)} + \frac1{\Theta(t_{k+1}\dots t_n)} \right)T(t_1\dots t_k)T(t_{k+1}\dots t_n) + \dots \end{multline} where dots stand for a function regular on \eqref{118}. By \eqref{115} the sum in parentheses in \eqref{119} vanishes and this proves \eqref{116} and \eqref{117}. Thus, what is left is to show that LHS of \eqref{114} vanishes at some point. This follows from the general identity established below in Theorem \ref{thm113} where one has to substitute $$ f(x)=\Theta(x) \,. $$ This concludes the proof of the theorem. \end{proof} Let $f(x)$ be an odd function $$ f(x^{-1})=-f(x) \,. $$ Set $$ \p{k}(x):= \left(x\d x\right)^k f(x)\,, k\in{\Bbb N} \,. $$ Consider the following function \begin{equation}\label{1110} \Phi(t_1,\dots,t_n):=\sum_{\gamma\in{\Bbb G} n} \Psi_\gamma (t) \,, \end{equation} where $$ \Psi_\gamma(t)=(-1)^{\ell(\g)} \p{\#\gamma_1}(1)\, \prod_{i=2}^{\ell(\g)} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} \,. $$ Note that $\Psi_\gamma(t)=0$ if $\#\gamma_1$ is even. We want to prove the following \begin{thm}\label{thm113} Suppose that $n>1$ and the variables $t_1,\dots,t_n$ are subject to constraint $$ t_1 \cdots t_n = 1 \,. $$ Then for any odd function $f(x)$ with a simple zero at $x=1$ $$ f(1)=0\,, \quad f'(1)\ne 0 $$ we have \begin{equation}\label{1111} \Phi(t_1,\dots,t_n)\to 0\,, \quad t_1 \to 1 \,, \end{equation} where the function $\Phi(t)$ was defined in \eqref{1110}. \end{thm} \begin{proof} Let us divide all summands in \eqref{1110} into 3 following types: \begin{enumerate} \item $\gamma_{\ell(\g)}=\{1\}$\,, \item $\gamma_1=\{1\}$\,, \item others. \end{enumerate} Note that all type 3 summands are regular at $t_1=1$. Let $\gamma=(\gamma_1,\dots,\gamma_{\ell-1},\{1\})$ be a type 1 composition. Then the composition $$ \gamma'=(\{1\},\gamma_1,\dots,\gamma_{\ell-1}) $$ is of type 2. Consider the sum \begin{equation}\label{1112} \Psi_\gamma (t) + \Psi_{\gamma'} (t) \,. \end{equation} There are two possible cases: $\#\gamma_1$ is even and $\#\gamma_1$ is odd. If $$ \text{\rm $\#\gamma_1$ is even} $$ then the first summand in \eqref{1112} is zero and the second one is regular at $t_1=1$. In this case we obtain \begin{equation}\label{1114} \lim_{t_1\to 1} \left(\Psi_\gamma (t) + \Psi_{\gamma'} (t) \right) = (-1)^{\ell} \p{\#\gamma_1+1}(1)\,\, \prod_{i=2}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} \,. \end{equation} It is easy to see that \eqref{1114} will exactly cancel with the type 3 summand corresponding to the composition \begin{equation}\label{1115} (\{1\}\cup\gamma_1,\dots,\gamma_{\ell-1}) \,. \end{equation} Note that if $\#\gamma_1$ is odd then the contribution of the composition \eqref{1115} is zero. Now consider the sum \eqref{1112} in the case $$ \text{\rm $\#\gamma_1$ is odd}\,. $$ We have \begin{alignat*}{2} \Psi_\gamma(t)&= (-1)^{\ell+1} \p{\#\gamma_1}(1)\,\frac{f'(t_1)}{f(t_1)} &&\prod_{i=2}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} \\ \Psi_{\gamma'}(t)&= (-1)^{\ell} f'(1) \,\frac{\p{\#\gamma_1}(t_1)}{f(t_1)} &&\prod_{i=2}^{\ell-1} \frac{ \p{\#\gamma_i}\left(t_1\cdot {\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left(t_1 \cdot{\tprod}_{j\in\ups(\g)_i}t_j\right)} \end{alignat*} Observe that \begin{equation}\label{1117} f'(1) \,\, \frac{\p{\#\gamma_1}(t_1)}{f(t_1)} - \p{\#\gamma_1}(1)\,\, \frac{f'(t_1)}{f(t_1)} \to 0\,, \quad t_1 \to 1 \end{equation} because \eqref{1117} is regular and odd. Therefore \begin{multline*} \lim_{t_1\to 1} \left(\Psi_\gamma (t) + \Psi_{\gamma'} (t) \right) = \\ (-1)^\ell \p{\#\gamma_1}(1)\,\, \lim_{t_1\to 1} \frac{f'(t_1)}{f(t_1)} \left( \prod_{i=2}^{\ell-1} \frac{ \p{\#\gamma_i}\left(t_1\cdot {\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left(t_1 \cdot{\tprod}_{j\in\ups(\g)_i}t_j\right)} - \right.\\ \left. \prod_{i=2}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)}\right) \,. \end{multline*} By L'Hospital rule this limit equals \begin{multline*} (-1)^\ell \p{\#\gamma_1}(1) \left( \sum_{k=2}^{\ell-1} \left( \frac {\p{\#\gamma_k+1}\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} {f\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} - \right. \right. \\ - \left. \left. \frac {\p{\#\gamma_k}\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)f'\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} {f\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)^2} \right) \prod_{\stackrel{i=2}{i\ne k}}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} \right) \,. \end{multline*} It is easy to see that the summand $$ (-1)^\ell \p{\#\gamma_1}(1) \,\, \frac {\p{\#\gamma_k+1}\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} {f\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} \prod_{\stackrel{i=2}{i\ne k}}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} $$ cancels with the contribution of the type 3 composition \begin{equation}\label{1118} (\gamma_1,\dots, \{1\}\cup \gamma_k ,\dots, \gamma_{\ell-1}) \,, \quad 2\le k \le l-1\,. \end{equation} Similarly the summand \begin{multline*} (-1)^{\ell+1} \p{\#\gamma_1}(1) \,\, \frac {\p{\#\gamma_k}\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)f'\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)} {f\left({{\textstyle \prod}}_{j\in{\upsilon}(\gamma)_k} t_j\right)^2} \times \\ \prod_{\stackrel{i=2}{i\ne k}}^{\ell-1} \frac{ \p{\#\gamma_i}\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} {f\left({\tprod}_{j\in\ups(\g)_i}t_j\right)} \,. \end{multline*} cancels with the contribution of the type 3 composition \begin{equation}\label{1119} (\gamma_1,\dots,\{1\},\gamma_k,\dots,\gamma_{\ell-1}) \,, \quad 2\le k \le l-1 \,. \end{equation} It is clear that the compositions of the form \eqref{1115}, \eqref{1118}, and \eqref{1119} exhaust the set of type 3 compositions. This concludes the proof. \end{proof} \section{Conclusion of the proof of Theorem \ref{thm61}}\label{sec:pf} Induct on $n$. Suppose that $n=1$. By Theorems \ref{thm81} and \ref{thm91} the function $$ \Theta(t_1) F(t_1) $$ is holomorphic on ${\Bbb C}\setminus 0$, invariant under the transformation $$ t_1 \mapsto q t_1 $$ and equal to $1$ for $t_1=1$. It follows that $$ F(t_1)=\frac{1}{\Theta(t_1)} \,. $$ Suppose that $n>1$ and consider the function \begin{equation}\label{121} \Theta(t_1\cdots t_n) F(t_1,\dots,t_n) - T(t_1,\dots,t_n) \,. \end{equation} By induction hypothesis, we have $$ \Theta(t_1\cdots t_n) F^\pi (t_1,\dots,t_n) - T^\pi(t_1,\dots,t_n)\,, $$ provided $$ \ell(\pi) < n \,. $$ Therefore by Theorem 1.1 and Corollary 3.2 the function is invariant under the transformation $$ t_1 \mapsto q t_1\,. $$ By symmetry it is invariant under all transformations $$ t_i \mapsto q t_i\,, \quad i=1,\dots,n \,. $$ By Theorems \ref{thm91} and \ref{thm111} the function \eqref{121} is holomorphic and vanishes if $$ t_1 \cdots t_n = 0 \,. $$ Therefore \eqref{121} equals zero. This concludes the proof. \setcounter{section}{12} \section{Another Example}\label{sec:example} Another representation of a subalgebra of the algebra $ {\mathcal D}$ of differential operators on ${\Bbb C}[t,t^{-1}]$ consisting of differential operators which are skew-adjoint in a suitable sense, was studied in \cite{B}. This algebra contains Virasoro and also the odd powers $D,D^3,D^5,\ldots$ of $D = t\frac{d}{dt}$. The $D^{2n+1}$ act semisimple with finite eigenspaces, and the resulting character is \begin{equation}\label{131}\Psi(\tau_1,\tau_3,\ldots) = q_1^{\zeta(-1)/2}q_3^{\zeta(-3)/2}\cdots \prod_{n=1}^\infty (1-q_1^nq_3^{n^3}\cdots)^{-1} \end{equation} In this section, we will show that the Taylor expansion for $\Psi$ is quasimodular of weight -1/2 (cf. \eqref{4wt}), and we will calculate the $n$-point function \begin{multline}\label{132}{\mathcal F}_n(\tau_1,z_1,\dotsc,z_n) := \\ \sum_{k_1,\dotsc,k_n\ge 1} \frac{\partial^n}{\partial\tau_{2k_1-1}\cdots\partial\tau_{2k_n-1}}(\Psi)|_{\tau_3 = \cdots = 0} \frac{z_1^{2k_1-1}\cdots z_n^{2k_n-1}}{(2k_1-1)!\cdots (2k_n-1)!}. \end{multline} As before, we write $$q_r = \exp(2\pi i \tau_r);\quad r\ge 1. $$ We compute \begin{gather*} \frac{1}{2\pi i}\frac{\partial}{\partial \tau_{2j-1}} \log\Psi = \zeta(1-2j)/2 + \sum_{m,n=1}^\infty m^{2j-1}q_1^{nm}q_3^{nm^3}\cdots \\ \frac{1}{(2\pi i)^r}\frac{\partial^r}{\partial \tau_{2j_1-1}\cdots\partial\tau_{2j_r-1}} \log\Psi = \sum_{m,n=1}^\infty m^{2(\sum_k j_k-r)-1}(nm)^{r-1}q_1^{nm}q_3^{nm^3}\cdots \notag \end{gather*} Note these expressions depend only on $r$ and $\sum_{k=1}^r j_k$. Define \begin{gather}\label{134} h_{r, 2(j_1+\ldots+j_r)}(\tau_1,\tau_3,\ldots) := \frac{1}{(2\pi i)^r}\frac{\partial^r}{\partial \tau_{2j_1-1}\cdots\partial\tau_{2j_r-1}} \log\Psi,\\ g_{r, 2(j_1+\ldots+j_r)}(\tau_1) := h_{r, 2(j_1+\ldots+j_r)}(\tau_1,0,0,\ldots) = \frac{\partial^{r-1}}{\partial\tau_1^{r-1}}G_{2(j_1+\ldots+j_r-r+1)}(\tau_1) \label{135} \end{gather} where $G_{2p}(\tau)$ is the Eisenstein series of weight $2p$ as in \eqref{34}. Note $g_{r,s}$ is quasimodular of weight $2a$. For $A = (a_3,a_5,\ldots)$ write ${\rm wt}(A) = 4a_3+6a_5+\ldots$ (compare \eqref{4wt}). The coefficient of $\tau^A/A!$ in the Taylor expansion for $\log\Psi$ is $g_{r,{\rm wt} A}(\tau_1)$. Also \begin{equation}\label{136} \Psi|_{\tau_3=\cdots = 0} = \eta^{-1}(\tau_1) \end{equation} Thus, by \eqref{4wt} we conclude that $\log(\Psi\eta)$ is quasimodular of weight $0$. Exponentiating yields \begin{prop}\label{prop131} $\Psi$ is quasimodular of weight $-1/2$. \end{prop} We now consider the $n$-point function \eqref{132} \begin{lem} Let ${\mathcal S} = \{j_1,\dotsc,j_n\}$ be a set of positive integers. Let $\Pi({\mathcal S})$ denote the set of all partitions $\mu = \{\mu_1,\dotsc,\mu_\ell\}$ of ${\mathcal S}$. For any finite set $\phi$ let $\#\phi$ denote the number of elements in $\phi$. For a subset $\phi\subset {\mathcal S}$ let $|\phi|$ be the sum over the elements. Then $$\frac{1}{(2\pi i)^n}\frac{\partial^n}{\partial \tau_{2j_1-1}\cdots\partial\tau_{2j_n-1}}\Psi = \Psi\sum_{\mu\in \Pi({\mathcal S})} \prod_{k=1}^{\#\mu}h_{\#\mu_k,2|\mu_k|} $$ \end{lem} \begin{proof}By induction on $n$. For $n=1$ this is just the logarithmic derivative. Suppose now $n\ge 2$ and the assertion holds for $n-1$. Note $$\frac{1}{2\pi i}\frac{1}{\partial\tau_{2j_n-1}}h_{n-1,2(j_1+\ldots+j_{n-1})} = h_{n,2(j_1+\ldots+j_{n})} $$ Let ${\mathcal T} = \{j_1,\dotsc,j_{n-1}\}$. We have inductively \begin{multline*}\frac{1}{(2\pi i)^n}\frac{\partial^n}{\partial \tau_{2j_1-1}\cdots\partial\tau_{2j_n-1}}\Psi = \frac{1}{2\pi i}\frac{\partial}{\partial\tau_{2j_n-1}}\Big(\Psi\sum_{\mu\in \Pi({\mathcal T})} \prod_{k=1}^{\#\mu}h_{\#\mu_k,2|\mu_k|}\Big) = \\ = \Psi h_{1,2j_n}\sum_{\mu\in \Pi({\mathcal T})} \prod_{k=1}^{\#\mu}h_{\#\mu_k,2|\mu_k|} + \Psi\sum_{\mu\in \Pi({\mathcal T})}\sum_{p=1}^{\#\mu}h_{\#\mu_p+1,2(|\mu_p|+j_n)}\prod_{k\neq p}h_{\#\mu_k,2|\mu_k|} \\ = \Psi\sum_{\mu\in\Pi({\mathcal S})}\prod_{k=1}^{\#\mu}h_{\#\mu_k,2|\mu_k|}. \end{multline*} \end{proof} Define \begin{equation}{\mathcal G}_n(\tau,z) := \sum_{r=1}^\infty \frac{\partial^{n-1}}{\partial\tau^{n-1}} G_{2r}(\tau)z^{2r+n-2}/(2r+n-2)! \end{equation} One has, with $\Theta$ as in \eqref{61} \begin{equation}2{\mathcal G}_1(\tau,2\pi iz) := \frac{1}{2\pi i}(-\frac{d}{dz}\log\Theta(z)+\frac{1}{z}) \end{equation} To see this, let $\sigma(z,\tau)$ be the elliptic sigma function as defined e.g. in \cite{L}, p. 247. One has \begin{multline*}\frac{-1}{2\pi i}\frac{d}{dz}\log\Theta = \frac{-1}{2\pi i} \frac{d}{dz}\log\sigma +2G_2(\tau)(2\pi iz) = \\ \frac{-1}{2\pi iz} +2G_2(\tau)(2\pi iz)+2G_4(\tau)(2\pi iz)^3/3!+2G_6(\tau)(2\pi iz)^5/5!+\ldots \end{multline*} Combining \eqref{132}, \eqref{135}, and \eqref{136} we get \begin{multline}\label{1310} {\mathcal F}(\tau,z_1,\dotsc,z_n) = \eta(\tau)^{-1}\sum_{k_1,\ldots,k_n\ge 1} \frac{z_1^{2k_1-1}\cdots z_n^{2k_n-1}}{(2k_1-1)!\cdots (2k_n-1)!}\times \\ \sum_{\mu\in\Pi\{k_1,\dotsc,k_n\}} \prod_{p=1}^{\#\mu} \frac{\partial^{\#\mu_p -1}}{\partial\tau^{\#\mu_p -1}}G_{2(|\mu_p|-\#\mu_p +1)}(\tau). \end{multline} For a function $f(z_1,\dotsc,z_n)$ let $\epsilon f$ denote the ``oddification'' of $f$, e.g. $$\epsilon f(z_1,z_2) = \frac{1}{4}(f(z_1,z_2)-f(-z_1,z_2)-f(z_1,-z_2)+f(-z_1,-z_2)) $$ Note $$\epsilon((z_1+\ldots+z_n)^r/r!) = \sum_{\ell_1+\ldots+\ell_n=(r+n)/2} \frac{z_1^{2\ell_1-1}\cdots z_n^{2\ell_n-1}}{(2\ell_1-1)!\cdots (2\ell_n-1)!} $$ It follows that \begin{multline*}\epsilon{\mathcal G}_n(\tau,z_1+\ldots+z_n) = \sum_{r=1}^\infty \frac{\partial^{n-1}} {\partial\tau^{n-1}} G_{2r}(\tau)\times \\ \sum_{k_1+\ldots+k_n=r+n-1}\frac{z_1^{2k_1-1}\cdots z_n^{2k_n-1}}{(2k_1-1)!\cdots (2k_n-1)!} \end{multline*} We conclude \begin{prop} The $n$-point function is given by $${\mathcal F}(\tau,z_1,\dotsc,z_n) = \eta(\tau)^{-1}\sum_{\mu\in\Pi_n}\prod_{k=1}^{k=\#\mu}\epsilon{\mathcal G}_{\#\mu_k}(\sum_{i\in \mu_k} z_i) $$ \end{prop} \begin{proof} We have an obvious identification $$\Pi_n = \Pi\{1,\dotsc,n\} = \Pi\{k_1,\dotsc,k_n\} $$ (Note $\{k_1,\dotsc,k_n\}$ is treated formally, i.e. we do not identify the $k_i$ if they are equal.) Write $\kappa=\{k_1,\dotsc,k_n\}$ and let $\mu_{p,\kappa}\subset \kappa$ be the corresponding subset. Given $\mu\in\Pi_n$ define \begin{multline*}{\mathcal F}_\mu(\tau,z_1,\dotsc,z_n) = \eta(\tau)^{-1}\sum_{k_1,\ldots,k_n\ge 1} \frac{z_1^{2k_1-1}\cdots z_n^{2k_n-1}}{(2k_1-1)!\cdots (2k_n-1)!}\times \\ \prod_{p=1}^{\#\mu} \frac{\partial^{\#\mu_p -1}}{\partial\tau^{\#\mu_p -1}}G_{2(|\mu_{p,\kappa}|-\#\mu_p +1)}(\tau). \end{multline*} so ${\mathcal F} = \sum_{\Pi_n} {\mathcal F}_\mu$. It will suffice to show \begin{equation}\label{1312}{\mathcal F}_\mu = \eta(\tau)^{-1}\prod_{p=1}^{p=\#\mu}\epsilon{\mathcal G}_{\#\mu_p}(\sum_{i\in \mu_p} z_i). \end{equation} But given $\kappa$, the term $z_1^{2k_1-1}\cdots z_n^{2k_n-1}$ occurs exactly once in \eqref{1312} and has the correct coefficient in $\tau$. \end{proof} \bibliographystyle{plain} \renewcommand\refname{References}
1997-12-03T00:49:42
9712
alg-geom/9712004
en
https://arxiv.org/abs/alg-geom/9712004
[ "alg-geom", "math.AG" ]
alg-geom/9712004
Janos Kollar
J\'anos Koll\'ar
Real Algebraic Threefolds I: Terminal Singularities
LATEX2e, 24 pages
null
null
null
null
This is the first of a series of papers studying real algebraic threefolds using the minimal model program. The main results are outlined in Part II. The present part I. contains the necessary preliminary work concerning terminal singularities. First I give standard forms for 3-dimensional terminal singularities over arbitrary fields. This is then used to develop a fairly complete topological classification over the reals.
[ { "version": "v1", "created": "Tue, 2 Dec 1997 23:49:41 GMT" } ]
2007-05-23T00:00:00
[ [ "Kollár", "János", "" ] ]
alg-geom
\section{Introduction} In real algebraic geometry, considerable attention has been paid to the study of real algebraic curves (in connection with Hilbert's 16th problem) and also to real algebraic surfaces. See \cite{Viro90}, \cite{Riesler93} and the references there. In higher dimensions one of the main avenues of investigation was initiated by \cite{Nash52}, and later developed by many others (see \cite{AK92} for some recent directions). One of these results says that every compact differentiable manifold can be realized as the set of real points of an algebraic variety. \cite{Nash52} posed the problem of obtaining similar results using a restricted class of varieties, for instance rational varieties. The aim of this series of papers is to develop the theory of minimal models for real algebraic threefolds. This approach gives very strong information about the topology of real algebraic threefolds, and it also answers the above mentioned question of \cite{Nash52}. For algebraic threefolds over $\c$, the minimal model program provides a very powerful tool. The method of the program is the following. (See \cite{Koll87} or \cite{CKM88} for introductions) Starting with a smooth projective variety $X$, we perform a series of ``elementary" birational transformations $$ X=X_0\map X_1\map \cdots \map X_n $$ until we reach a variety $X_n$ whose global structure is ``simple". In essence the minimal model program allows us to investigate many questions in two steps: first study the effect of the ``elementary" transformations and then consider the ``simple" global situation. In parctice both of these steps are frequently rather difficult. For instance, we still do not have a complete list of all possible ``elementary" steps, despite repeated attempts to obtain it. A somewhat unpleasant feature of the theory is that the varieties $X_i$ are not smooth, but have so called terminal singularities. In developing the theory of minimal models for real algebraic threefolds, we again have to understand the occurring terminal singularities. The aim of this paper is to give a classification of terminal 3-fold singularities over $\r$. Minimal models serve only as a background, the proofs depend entirely on well established methods of singularity theory. I do not even use the definition of terminal singularities! Terminal 3-fold singularities over $\c$ are completely classified. \cite{Reid85} is a very readable introduction and survey. I will take the result of this classification as my definition, since the theory over $\r$ can be most naturally developed in this setting. The classification is, in some sense, not complete. In a few cases I obtain unique normal forms (\ref{ca1-top.thm}), but in most cases this seems nearly impossible (see \cite{Markushevich85} for a special case over $\c$). My aim is to write the singularities in a form that allows one to determine their topology over $\r$. The resulting lists and algorithms are given in sections 4--5. It turns out that the normal forms of 3-fold terminal singularities are essentially the same over any field of characteristic zero. Thus in sections 2--3 I work with any subfield of $\c$. As a consequence of the classification over $\c$, we know that 3-fold terminal singularities come in two types. Some are hypersurface singularities, and the others are quotients of these hypersurface singularities by a finite cyclic group. Accordingly, the classification over any field is done in two steps. Section 2 deals with terminal hypersurface singularities. These results are mostly routine generalizations of the theory over $\c$. Quotient singularities frequently have ``twisted" forms over a subfield of $\c$. ``Twisted" forms do not appear for 3-fold terminal singularities, and so the classification ends up very similar to the one over $\c$. \begin{ack} I thank G. Mikhalkin for answering my numerous questions about real algebraic geometry. Partial financial support was provided by the NSF under grant number DMS-9622394. Most of this paper was written while I visited RIMS, Kyoto Univ. \end{ack} \section{Terminal hypersurface singularities} \begin{notation} For a field $K$ let $K[[x_1,\dots,x_n]]$ denote the ring of formal power series in $n$ variables over $K$. For $K=\r$ or $K=\c$, let $K\{x_1,\dots,x_n\}$ denote the ring of those formal power series which converge in some neighborhood of the origin. For any $F\in K\{x_1,\dots,x_n\}$ the set $(F=0)$ is a germ of a real or complex analytic set. I will refer to it as a singularity. If $F\in K[[x_1,\dots,x_n]]$ then by the singularity $(F=0)$ I mean the scheme $\spec_KK[[x_1,\dots,x_n]]/(F)$. For a power series $F$, $F_d$ denotes the degree $d$ homogeneous part. The multiplicity, denoted by $\mult_0F$, is the smallest $d$ such that $F_d\neq 0$. If we write a power series as $F_{\geq d}$ then it is assumed that its multiplicity is at least $d$. Two power series $F,G\in K[[x_1,\dots,x_n]]$ are called equivalent over $K$ if there is an automorphism of $K[[x_1,\dots,x_n]]$ given by $x_i\mapsto \phi_i(x_1,\dots,x_n)\in K[[x_1,\dots,x_n]]$ and an invertible $u(x_1,\dots,x_n)\in K[[x_1,\dots,x_n]]$ such that $$ u(x_1,\dots,x_n)G(x_1,\dots,x_n)=F(\phi_1,\dots,\phi_n). $$ Thus $F$ and $G$ are equivalent iff the corresponding singularities $(F=0)$ and $(G=0)$ are isomorphic (over $K$). We have to pay special attention to cases when $F$ and $G$ are not equivalent over $K$ but are equivalent over some larger field. For instance, $F=x_1^2+x_2^2$ and $G=x_1^2-x_2^2$ are not equivalent over $\r$ but are equivalent over $\c$. If $K=\r,\c$ and $F,G\in K\{x_1,\dots,x_n\}$ then I am mainly interested in equivalences where $u,\phi_i\in K\{x_1,\dots,x_n\}$. If $F,G\in K\{x_1,\dots,x_n\}$ have isolated critical points at the origin, then $F$ and $G$ are equivalent in $K\{x_1,\dots,x_n\}$ iff they are equivalent in $K[[x_1,\dots,x_n]]$ (cf. \cite[p.121]{AGV85}), thus we do not have to be careful about this distinction. \end{notation} \begin{defn} Let $K$ be a field of characteristic zero with algebraic closure $\bar K$. $(F(x,y,z)=0)$ is called a {\it Du Val} singularity (or a rational double point) iff over $\bar K$ it is equivalent to one of the standard forms \begin{enumerate} \item[$A_n$] \quad $x^2+y^2+z^{n+1}=0$; \item[$D_n$] \quad $x^2+y^2z+z^{n-1}=0$; \item[$E_6$] \quad $x^2+y^3+z^4=0$; \item[$E_7$] \quad $x^2+y^3+yz^3=0$; \item[$E_8$] \quad $x^2+y^3+z^5=0$. \end{enumerate} Du Val singularities have many interesting intrinsic characterizations, (cf. \cite{Durfee79, Reid85}) but I will not use this. \end{defn} The following definition introduces our basic objects of study. \begin{defn}\label{cdv.def} Let $K$ be a field of characteristic zero with algebraic closure $\bar K$. $(F(x,y,z,t)=0)$ is called a {\it compound Du Val} singularity (or {\it cDV} for short) iff over $\bar K$ it is equivalent to $$ h(x,y,z)+tf(x,y,z,t)=0 $$ where $(h=0)$ is a Du Val singularity. $(F(x,y,z,t)=0)$ is called a $cA_n$ (resp. $cD_n$ or $cE_n$) singularity if its equation can be written as above with $h$ having type $A_n$ (resp. $D_n$ or $E_n$), but it does not admit such representation with a smaller value of $n$. It is called a $cA$ (resp. $cD$ or $cE$) singularity if the value of $n$ is not specified. \end{defn} The reason we are interested in cDV singularities is the following: \begin{thm}\cite{Reid80} A 3-dimensional hypersurface singularity over $\c$ is terminal iff it is an isolated cDV singularity.\qed \end{thm} The aim of this section is to develop ``normal forms" for cDV singularities over any field $K$. This will then give ``normal forms" for 3-dimensional terminal hypersurface singularities over $K$. The proof is a rather standard application of the methods of \cite{AGV85}. \begin{say}\label{nf.meth} We repeatedly use 3 methods: \begin{enumerate} \item The Weierstrass preparation theorem. This is frequently stated only over $\c$, but it works over any field since the Weierstrass normal form is unique. \item The elimination of the $y^{n-1}$-term from the polynomial $a_ny^n+a_{n-1}y^{n-1}+\dots$ by a coordinate change $y\mapsto y-a_{n-1}/na_n$ when $a_n$ is invertible. \item Let $M_1,\dots,M_k$ be monomials in the variables $x_1,\dots,x_m$. Assume that $x_0M_1,\dots,x_0M_k$ are multiplicatively independent. Then any power series of the form $\sum M_i\cdot u_i(x_1,\dots,x_m)$ where $u_i(0)\neq 0$ for all $i$ is equivalent to $\sum M_i\cdot u_i(0)$ by a suitable coordinate change $x_i\mapsto x_i\cdot(\mbox{unit})$. \end{enumerate} \end{say} These elementary operations are sufficient to deal with the $cA$ and $cE$ cases. In the $cD$ case the following generalization of (\ref{nf.meth}.2) is needed. \begin{const}\label{nf.meth.weighted} In $K[[x_1,\dots,x_m]]$, assign positive integral weights to the variables $w(x_i)=w_i$. For a monomial set $w(\prod x_i^{c_i})=\sum c_iw_i$. Write a power series in terms of its weighted homogeneous pieces $F=F_d+F_{d+1}+\dots$. Choose $g_i\in K[[x_1,\dots,x_m]]$ such that $w(g_i)=w(x_i)+e$ for some $e>0$. Then $$ F(x_i+g_i)=F(x_i)+\sum g_i\frac{\partial F_d}{\partial x_i}+ R_{>(d+e)}(x_i). $$ Repeatedly using this for higher and higher degrees, we see that, for every $N>0$, $F$ is equivalent to a power series $F^N+R_{>N}$ where $F^N$ is a polynomial of degree $N$ and no linear combination of the monomials in $F^N$ can be written in the form $\sum g_i(\partial F_d/\partial x_i)$ as above. In the ring of formal power series this can be continued indefinitely, thus at the end we can kill all the degree $>d$ elements of the Jacobian ideal $$ \Delta(F_d):=\left(\frac{\partial F_d}{\partial x_1}, \dots , \frac{\partial F_d}{\partial x_m}\right). $$ If $F\in K\{x_1,\dots,x_m\}$ defines an isolated singularity, then by Tougeron's lemma (cf. \cite[p.121]{AGV85}), $F^N+R_{>N}$ is equivalent to $F^N$ by an analytic coordinate change for $N\gg 1$. Thus the final conclusion is the same. \end{const} \begin{prop}\label{morse.lem} Any power series $F_{\geq 2}(x_1,\dots,x_n)$ is equivalent to a power series $$ a_1x_1^2+\dots+a_kx_k^2+G_{\geq 3}(x_{k+1},\dots,x_n). $$ \end{prop} Proof. By a linear change of coordinates we can diagonalize $F_2$, thus we can assume that $F_2=a_1x_1^2+\dots+a_kx_k^2$. Repeatedly applying (\ref{nf.meth}.1) to the variables $x_1,\dots,x_k$ we reach a situation when $F$ is a quadratic polynomial in the variables $x_1,\dots,x_k$. (\ref{nf.meth}.2) can then be used to eliminate the linear terms in $x_1,\dots,x_k$.\qed \begin{thm}\label{ca.thm} Assume that $F_{\geq 1}(x,y,z,t)\in K[[x,y,z,t]]$ defines a terminal singularity of type $cA$. Then $F$ is equivalent to one of the following: \begin{enumerate} \item[$cA_0$] \quad $x=0$. \item[$cA_1$] \quad $ax^2+by^2+cz^2+dt^m=0$, where $abcd\neq 0$. \item[$cA_{>1}$] \quad $ax^2+by^2+f_{\geq 3}(z,t)=0$, where $ab\neq 0$. This has type $cA_n$ for $n=\mult_0f-1$. \end{enumerate} \end{thm} Proof. If $F_1\neq 0$ then (\ref{nf.meth}.1) gives $cA_0$. Thus assume that $F_1=0$. $F$ has type $cA$, hence $F_2$ is a quadric of rank at least 2. If the rank is 2 then (\ref{morse.lem}) gives the $cA_{>1}$ cases. Assume finally that $F_2$ has rank 3 or 4. By (\ref{morse.lem}) we can write $F$ as $ax^2+by^2+cz^2+g(t)=0$. Using (\ref{nf.meth}.3) we obtain $ax^2+by^2+cz^2+dt^m=0$. In all these cases we can multiply through by $a^{-1}$ to get a somewhat simpler form when the coefficient of $x^2$ is 1. \qed \begin{thm}\label{cd.thm} Assume that $F_{\geq 2}(x,y,z,t)\in K[[x,y,z,t]]$ defines a terminal singularity of type $cD$. Then $F$ is equivalent to one of the following: \begin{enumerate} \item[$cD_4$] \quad $x^2+f_{\geq 3}(y,z,t)$, where $f_3$ is not divisible by the square of a linear form. \item[$cD_{>4}$] \quad $x^2+y^2z+ayt^r+h_{\geq s}(z,t)$, where $a\in K$, $r\geq 3$, $s\geq 4$ and $h_s\neq 0$. This has type $cD_n$ where $n=\min\{2r, s+1\}$ if $a\neq 0$ and $n=s+1$ if $a=0$. \end{enumerate} \end{thm} Proof. $F_2$ is a rank one quadric, thus in suitable coordinates the equation becomes $ax^2+f_{\geq 3}(y,z,t)$. Here $f_3\neq 0$ is not the cube of a linear form since otherwise we would have a type $cE$ singularity. If $f_3$ is not divisible by the square of a linear form then we have case $cD_4$. If $f_3$ is divisible by the square of a linear form, then $f_3=l_1^2l_2$ for two linear forms $l_i$, and both of them are defined over $K$. We can change coordinates $l_1\mapsto y$ and $l_2\mapsto z$. At this point our power series is $x^2+y^2z+(\mbox{higher order terms})$. Assign weights $w(x)=3, w(y)=w(z)=2, w(t)=6$. The leading term is $x^2+y^2z$. Using (\ref{nf.meth.weighted}) we can eliminate all monomials which contain $y^2$ or $yz$. To see the last part, take the hyperplane section $t=\lambda z$. The term $ay\lambda^rz^r$ can be eliminated by a substitution $y\mapsto y+(a/2)\lambda^rz^{r-1}$. This creates a term $-(a/2)^2\lambda^{2r} z^{2r-1}$. The only problem could be that $h(z,\lambda z)$ has multiplicity $2r-1$ and there is cancellation. However, $h_{2r-1}(z,\lambda z)=z^{2r-1} h_{2r-1}(1,\lambda)$ is a polynomial of degree $2r-1$ in $\lambda$, thus it does not equal $-(a/2)^2\lambda^{2r} z^{2r-1}$. \qed \begin{thm}\label{ce.thm} Assume that $F_{\geq 2}(x,y,z,t)\in K[[x,y,z,t]]$ defines a terminal singularity of type $cE$. Then $F$ is equivalent to one of the following: \begin{enumerate} \item[$cE_6$] \quad $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$, where $h_4\neq 0$. \item[$cE_7$] \quad $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)$, where $g_3\neq 0$. \item[$cE_8$] \quad $x^2+y^3+yg_{\geq 4}(z,t)+h_{\geq 5}(z,t)$, where $h_5\neq 0$. \end{enumerate} \end{thm} Proof. $F_2$ is a rank one quadric by (\ref{cdv.def}), thus in suitable coordinates the equation becomes $ax^2+f_{\geq 3}(y,z,t)$. Here $f_3\neq 0$ and it is the cube of a linear form since otherwise we would have a type $cD$ singularity. (\ref{nf.meth}.1--2) gives an equation $$ ax^2+by^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t). $$ Multiply the equation by $a^3b^2$ and then make the substitutions $x\mapsto xa^{-2}b^{-1}$ and $y\mapsto ya^{-1}b^{-1}$ to get the required normal forms.\qed \section{Higher index terminal singularities} The classification of non-hypersurface terminal 3-fold singularities over $\c$ relies on the following construction: Let $\z_n$ denote the cyclic group of order $n$ and $\epsilon$ a primitive $n^{th}$ root of unity. Assume that $\z_n$ acts on $\c^4$ by $$ \sigma: (x,y,z,t)\mapsto (\epsilon^{a_x}x,\epsilon^{a_y}y,\epsilon^{a_z}z,\epsilon^{a_t}t). $$ I will use the shorter notation $\frac{1}{n}(a_x,a_y,a_z,a_t)$ to denote such an action. If $F(x,y,z,t)$ is equivariant with respect to this action, then $\z_n$ acts on the hypersurface $(F=0)$ and we can take the quotient, denoted by $(F=0)/\frac{1}{n}(a_x,a_y,a_z,a_t)$. By \cite{Reid80}, every terminal 3-fold singularity $X$ over $\c$ is of the form $(F=0)/\frac{1}{n}(a_x,a_y,a_z,a_t)$, where $F$ defines a terminal hypersurface singularity. The value of $n$ is uniquely determined by $X$, it is called the {\it index} of $X$. It is not easy to come up with a complete list of terminal 3-fold singularities, but by now the list is well understood; see \cite{Reid85} for a good survey. It turns out that most actions do not produce terminal quotients and we have only a few cases: \begin{thm}\label{hind.overC.thm}\cite{Mori85} Let $0\in X$ be a 3-fold terminal nonhypersurface singularity over $\c$. Then $0\in X$ is isomorphic to a singularity described by the following list: $$ \begin{tabular}{|c|l|c|l|l|} \hline name &\qquad equation & index & \quad action & condition\\ \hline cA/n& $xy+f(z,t)$ & $n$ & $(r,-r,1,0)$ & $(n,r)=1$\\ \hline cAx/2& $x^2+ y^2 +f_{\geq 4}(z,t)$ & $2$ & $(0,1,1,1)$&\\ \hline cAx/4 & $x^2+y^2 +f_{\geq 2}(z,t)$ & $4$ & $(1,3,1,2)$& $f_2(0,1)=0$\\ \hline cD/2&$x^2+f_{\geq 3}(y,z,t)$ & $2$ &$(1,0,1,1)$&\\ \hline cD/3&$x^2+f_{\geq 3}(y,z,t)$ &$3$ &$(0,2,1,1)$& $f_3(1,0,0)\neq 0$\\ \hline cE/2&$x^2+y^3+f_{\geq 4}(y,z,t)$ &$2$ &$(1,0,1,1)$&\\ \hline \end{tabular} $$ \end{thm} The equations have to satisfy 2 obvious conditions: \begin{enumerate} \item The equations define a terminal hypersurface singularity. \item The equations are $\z_n$-equivariant. (In fact $\z_n$-invariant, except for $cAx/4$.) \end{enumerate} If we work over a field $K$ which does not contain the $n^{th}$ roots of unity, then the action $\frac{1}{n}(a_1,\dots,a_m)$ is not defined over $K$. There is, however, another way of loking at the quotient which does make sense over any field. Any action of the cyclic group $\z_n$ on $\c^m$ defines a $\z_n$-grading $w$ of $\c[[x_1,\dots,x_m]]$ by $$ w(\prod x_i^{c_i})=a \qtq{iff} \sigma (\prod x_i^{c_i})=\epsilon^a \cdot \prod x_i^{c_i}. $$ If $F$ is $\z_n$-equivariant then $(F)\subset \c[[x_1,\dots,x_m]]$ is a homogeneous ideal, hence the grading descends to a grading of $\c[[x_1,\dots,x_m]]/(F)$. The ring of functions on the quotient $(F=0)/\frac{1}{n}(a_1,\dots,a_m)$ can be identified with the ring of grade zero elements of $\c[[x_1,\dots,x_m]]/(F)$. If $K$ is any field, $n\in \n$ and $a_i\in \z$, then we obtain a $\z_n$-grading $w=w(a_1,\dots,a_m)$ of $K[[x_1,\dots,x_m]]$ (or of $\r\{x_1,\dots,x_m\}$) by $$ w(\prod x_i^{c_i})=\sum c_ia_i \in \z_n. $$ Let $R\subset K[[x_1,\dots,x_m]]$ denote the subring of grade zero elements. Then $\spec_KR$ gives a singularity over $K$ which is denoted by $$ {\Bbb A}^m/{\textstyle \frac{1}{n}}(a_1,\dots,a_m). $$ (Especially when $K=\r$, one might be tempted to write ${\Bbb R}^m/{\textstyle \frac{1}{n}}(a_1,\dots,a_m)$ instead. However, the set of real points of ${\Bbb A}^m/{\textstyle \frac{1}{n}}(a_1,\dots,a_m)$ is not in any sense a quotient of the set $\r^n$ (cf. (\ref{2ind.quot})), so this may lead to confusion.) If $F\in K[[x_1,\dots,x_m]]$ is graded homogeneous, then $w$ gives a grading of $K[[x_1,\dots,x_m]]/(F)$. Let $R/(R\cap(F))\subset K[[x_1,\dots,x_m]]/(F)$ be the subring of grade zero elements. $\spec_K R/(R\cap(F))$ defines a singularity over $K$. By construction, $$ \spec_K R/(R\cap(F))\times_{\spec K}\spec \bar K\cong (F=0)/{\textstyle \frac{1}{n}}(a_1,\dots,a_m). $$ Thus $\spec_K R$ is a terminal singularity over $K$ iff $(F=0)/{\textstyle \frac{1}{n}}(a_1,\dots,a_m)$ is a terminal singularity over $\bar K$. Under certain conditions, every $K$-form of a quotient is obtained this way: \begin{thm}\label{hind.gen.thm} $K$ be a field of characteristic zero with algebraic closure $\bar K$. Let $\z_n$ denote the cyclic group of order $n$ and $\epsilon$ a primitive $n^{th}$ root of unity. Assume that $\z_n$ acts on $\bar K^m$ by $\sigma: (x_i)\mapsto (\epsilon^{w_i}x_i)$. Let $F\in \bar K[[x_1,\dots,x_m]]$ be equivariant with respect to this action, and assume that the fixed point set of $\sigma$ has codimension at least 2 in $(F=0)$. Assume in addition that $$ w(F)-\sum w_i\qtq{is relatively prime to} n. $$ Let $0\in X$ be a singularity over $K$ such that $$ X\times_{\spec K}\spec \bar K\cong (F=0)/ \textstyle{\frac{1}{n}(w_1,\dots,w_m)}. $$ Then there is an $F^K\in K[[x_1,\dots,x_m]]$ such that $F$ and $F^K$ are equivalent over $\bar K$ and $$ X\cong (F^K=0)/ \textstyle{\frac{1}{n}(w_1,\dots,w_m)}. $$ \end{thm} It is worthwhile to note that the condition about $n$ and $w(F)-\sum w_i $ being relatively prime is essential: \begin{exmp} Consider the quotient singularity $\c[u,v]/\frac{1}{n}(1,-1)$. It is isomorphic to $(xy-z^n=0)$ via the substitutions $x=u^n,y=v^n,z=uv$. Over $\c$ we have a Du Val singularity $A_{n-1}=(x^2+y^2+z^n=0)$. Over $\r$ we see that $(x^2-y^2-z^n=0)\cong {\Bbb A}^2/\frac{1}{n}(1,-1)$. Another $\r$-form of $A_{n-1}$ is $x^2+y^2-z^n$. This can also be obtained as a quotient, but this time we act on ${\Bbb A}^2$ by rotation with angle $2\pi/n$. Finally, if $n$ is even, then there is another $\r$-form of $A_{n-1}$ given by $(x^2+y^2+z^n=0)$. The only $\r$-point is the origin, so we do not even have a nonzero map $\r^2\to (x^2+y^2+z^n=0)$. As another example, take the 4-dimensional terminal singularity $\c^4/\frac{1}{n}(a,-a,b,-b)$ for any $(ab,n)=1$. It has another $\r$-form given as ${\Bbb A}^4/\z_n$ where we act on the first two coordinates by rotation with angle $2a\pi/n$ and on the last two coordinates by rotation with angle $2b\pi/n$. In some special cases there are further $\r$-forms. Take for instance $\c^4/\frac{1}{2}(1,1,1,1)$. This can be realized as the cone over $\c\p^3$ embedded by the quadrics to $\c\p^9$. Let $C\subset \r\p^2$ be a smooth conic. Taking symmetric powers we have $S^3C\subset S^3\r\p^2$ and $S^3H^0(\r\p^2,\o(1))$ embeds it to $\r\p^9$. If $C$ has a real point, then $S^3C\cong \r\p^3$ and we get the Veronese embedding. If $C$ has no real points then the image is a variety over $\r$ without real points. The cone over it is a real form of ${\Bbb A}^4/\frac12(1,1,1,1)$ with an isolated real point at the origin. \end{exmp} Proof of (\ref{hind.gen.thm}). Set $S=\bar K[[x_1,\dots,x_m]]/(F)$, the ring of functions on $\tilde X_{\bar K}:=(F=0)$. The $\z_n$-action defines a $\z_n$-grading $S=\sum_{i=0}^{n-1}S_i$. $S_0$, the ring of grade $0$ elements, is exactly the ring of functions on $X_{\bar K}$. Our aim is to find an algebraic way of reconstructing $S$ from $S_0$, which then hopefully generalizes to nonclosed fields. There is another summand which can be easily seen algebraically. Set $d=w(F)-\sum w_i$. Note that $$ \frac{1}{\partial F/\partial x_m}dx_1\wedge\dots\wedge dx_{m-1} $$ is a local generator of $\omega_S$ and it has weight $-d$. Thus $$ \omega_{S_0}\cong S_d\frac{1}{\partial F/\partial x_m}dx_1\wedge\dots\wedge dx_{m-1}. $$ Once $S_d$ is determined, we obtain $S_{jd}$ as follows. The multiplication map $$ S_a\otimes_{S_0} S_b\to S_{a+b}\qtq{(subscripts modulo $n$)} $$ are isomorhisms over the open set where the $\z_n$-action is free. We assumed that the complement has codimension at least 2, thus $S_{jd}\cong S_d^{[j]}$, where $ S_d^{[j]}$ denotes the double dual of $S_d^{\otimes j}$. If $d$ and $n$ are relatively prime, then we obtain every summand $S_i$ this way. In particular, $$ S=\sum_{i=0}^{n-1}S_i\cong \sum_{j=0}^{n-1}\omega_{S_0}^{[j]}. $$ Over an arbitrary field, we can thus proceed as follows. Let $\omega_X$ be the dualizing sheaf of $X$. This is also the reflexive sheaf $\o_X(K_X)$ where $K_X$ is the canonical class. Then $\omega_X^{[n]}$ is isomorphic to $\o_X$, where $n$ is the index. (We know this over $\bar K$. Isomorphism of two sheaves $F,G$ is a question about $\Hom(F,G)$ and this commutes with base field extensions.) Fix such an isomorphism $s:\omega_X^{[n]}\to \o_X$. Consider the $\o_X$-algebra $$ R(X,s):=\sum_{j=0}^{n-1}\omega_X^{[j]}, $$ where multiplication for $j+k\geq n$ is given by $$ \omega_X^{[j]}\otimes \omega_X^{[k]}\mapsto \omega_X^{[j+k]}\cong \omega_X^{[n]}\otimes \omega_X^{[j+k-n]}\stackrel{s\otimes 1}{\longrightarrow} \omega_X^{[j+k-n]}. $$ This has a $\z_n$ grading by declaring $\omega_X^{[j]}$ to have grade $j$. (Note. Two isomorphisms $s_1,s_2:\omega_X^{[r]}\to \o_X$ differ by an invertible function $h\in \o_X^*$. If $h$ is an $n^{th}$-power, then the resulting algebras $R(X,s_i)$ are isomorphic, but they need not be isomorphic otherwise. This is connected with the topological aspects observed in (\ref{2ind.quot}).) Over $\bar K$, $R(X,s)$ is isomorphic to $\o_{\tilde X}$. Thus $R(X,s)$ is a $K$-form of $\o_{\tilde X}$. In particular, $R(X,s)$ is an algebra of the form $K[[x_1,\dots,x_m]]/(F^K)$, where $F$ and $F^K$ are equivalent over $\bar K$. The grading lifts to a grading of $K[[x_1,\dots,x_m]]$ such that $F^K$ is graded homogeneous. We can choose $x_i$ to be homogeneous.\qed As a corollary, we obtain the following classification of terminal 3-fold nonhypersurface singularities over nonclosed fields: \begin{thm}\label{hind.term.thm} Let $K$ be a field of characteristic zero and $0\in X$ a 3-fold terminal nonhypersurface singularity over $K$. Then $0\in X$ is isomorphic over $K$ to a singularity described by the following list: $$ \begin{tabular}{|c|l|c|l|l|} \hline name &\qquad equation & index & \quad weights & condition\\ \hline cA/2& $ax^2+by^2+f(z,t)$ & $2$ & $(1,1,1,0)$&\\ \hline cA/n& $xy+f(z,t)$ & $n\geq 3$ & $(r,-r,1,0)$& $(n,r)=1$\\ \hline cAx/2& $ax^2+by^2 +f_{\geq 4}(z,t)$ & $2$ & $(0,1,1,1)$&\\ \hline cAx/4 & $ax^2+by^2 +f_{\geq 2}(z,t)$ & $4$ & $(1,3,1,2)$& $f_2(0,1)=0$\\ \hline cD/2&$x^2+f_{\geq 3}(y,z,t)$ & $2$ &$(1,0,1,1)$&\\ \hline cD/3&$x^2+f_{\geq 3}(y,z,t)$ &$3$ &$(0,2,1,1)$& $f_3(1,0,0)\neq 0$\\ \hline cE/2&$x^2+y^3+f_{\geq 4}(y,z,t)$ &$2$ &$(1,0,1,1)$&\\ \hline \end{tabular} $$ \end{thm} \begin{complement} The corresponding quotient singularity is terminal iff the equations satisfy 2 obvious conditions: \begin{enumerate} \item The equations define a terminal hypersurface singularity. \item The equations are graded homogeneous. \end{enumerate} With these assumptions, a terminal singularity corresponds to exactly one case on the above list. \end{complement} Proof. By looking at the list of (\ref{hind.overC.thm}), we see that the assumptions of (\ref{hind.gen.thm}) are satisfied. Hence we know that $X$ is of the form $(F^K=0)/\frac1{n}(a_x,a_y,a_z,a_t)$ where $\frac1{n}(a_x,a_y,a_z,a_t)$ is on the list of (\ref{hind.overC.thm}). Once we know a $\z_n$-grading on $K[[x,y,z,t]]$ and a graded homogeneous power series $F^K$, we can try to bring it to some normal form using the methods (\ref{nf.meth}) and (\ref{nf.meth.weighted}). They are set up in such a way that if $F^K$ is homogeneous in a $\z_n$-grading the all coordinate changes respect the grading. The proofs of (\ref{ca.thm}, \ref{cd.thm}, \ref{ce.thm}) remain unchanged. The only difference is in (\ref{morse.lem}). It is not true that a quadratic form can be diagonalized using a linear transformation which respects the $\z_n$-grading. The best one can achieve is a sum of forms in disjoint sets of variables $\sum q_i$ where each $q_i$ is either $au_i^2$ or $u_iv_i$. The latter case is necessary iff the two variables have different $\z_n$-grading. In the $cD$ and $cE$ cases the quadric has rank 1, so it can be diagonalized. In the $cA/2$ and $cAx/2$ cases every grade 0 quadric is diagonalizable. In the $cAx/4$ case $x^2,xz,y^2, z^2$ are the only grade 2 quadratic monomials. A quadratic form like this can again be diagonalized. Finally let us look at the $cA/n$-case for $n\geq 3$. The only grade 0 degree 2 monomials are $xy, t^2$ and $xz$ if $r=-1$ or $yz$ if $r=1$. We need to get a rank $\geq 2$ quadric, so $xy $ (or $xz$ if $r=-1$, $yz$ if $r=1$) must appear. In the $r=\pm 1$ case we may need to perform a linear change of variables to get the normal form $xy+f(z,t)$. \qed \section{The topology of terminal hypersurface singularities} Let $0\in X$ be a real singularity. It's real points $X(\r)$ form a topological space, which can be triangulated (cf. \cite[9.2]{BCR87}). We may assume that $0$ is a vertex of the triangulation. Then locally near $0$, $X(\r)$ is PL-homeomorphic to the cone over a simplicial complex $L=L(X(\r))$, which is called the {\it link} of $0$ in $X(\r)$. The local topology of $X(\r)$ at $0$ is thus determined by $L$. In general one needs to contemplate the dependence of $L$ on various choices made. I am mainly interested in the case when $X$ is a 3-dimensional isolated singularity. In this case $L$ is a compact surface (without boundary) and so $L$ and $X(\r)$ determine each other up to homeomorphism. The aim of this section is to classify terminal singularities over $\r$ according to their local topology. To be precise, we give a classification in the $cA$ cases and provide a procedure in the $cD$ and $cE$ cases which reduces the 3-dimensional problem to some questions about plane curve singularities. \begin{notation} $M\sim N$ denotes that $M$ and $N$ are homeomorphic. $\uplus$ denotes disjoint union. $M\uplus rN$ denotes the disjoint union of $M$ and of $r$ copies of $N$. $M_g$ denotes the unique compact, closed and orientable surface of genus $g$. \end{notation} We start with a general lemma. \begin{lem}\label{orient} Let $X$ be a smooth real hypersurface. Then $X(\r)$ is orientable. \end{lem} Proof. Let $X=(f=0)$ be a real equation where $f\in \r[x_1,\dots,x_n]$ or $f\in \r\{x_1,\dots,x_n\}$. At each point $p\in X$, $X$ divides a neighborhood of $p$ into two halves. $f$ is positive on one half and negative on the other half. Choosing a sign thus determines an orientation. \qed \begin{thm}\label{ca1-top.thm} The following table gives a complete list of 3-dimensional terminal singularities of type $cA_1$ over $\r$. In the table $n\geq 1$. Case 4, $n=1$ and case 5, $n=1$ are isomorphic. Aside from this, two singularities are isomorphic iff they correspond to the same case and the same value of $n$. $$ \begin{tabular}{|c|l|c|} \hline case & \qquad equation & $L$ \\ \hline $cA_1(1)$ & $x^2+y^2+ z^2\pm t^{2n+1}$ & $S^2$ \\ \hline $cA_1(2)$ & $x^2+y^2- z^2\pm t^{2n+1}$ & $S^2$ \\ \hline $cA_1(3)$& $x^2+y^2+z^2+ t^{2n}$ & $\emptyset$ \\ \hline $cA_1(4)$& $x^2+y^2+z^2-t^{2n}$ & $S^2\uplus S^2$ \\ \hline $cA_1(5)$& $x^2+y^2-z^2+ t^{2n}$ & $S^2\uplus S^2$ \\ \hline $cA_1(6)$& $x^2+y^2-z^2- t^{2n}$ & $S^1\times S^1$ \\ \hline \end{tabular} $$ \end{thm} Proof. The equations follow from (\ref{ca.thm}), once we note that after multiplying by $\pm 1$ we may assume that the quadratic part has at least 2 positive eigenvalues. The topology is easy to figure out. Since all the claims are special cases of the next result, I discuss them in more detail there.\qed \begin{thm}\label{can-top.thm} A 3-dimensional terminal singularity of type $cA_{>1}$ over $\r$ is equivalent to a form $$ x^2\pm y^2\pm h(z,t)\prod_{i=1}^m f_i(z,t)=0, $$ where the $f_i$ are irreducible power series (over $\r$) such that $(f_i(z,t)=0)$ changes sign on $\r^2\setminus\{0\}$ and $h(z,t)$ is positive on $\r^2\setminus\{0\}$. The following table gives a complete list of the possibilites for the topology of $X(\r)$. $$ \begin{tabular}{|l|l|c|} \hline \quad case & \qquad equation & $L(X(\r))$ \\ \hline $cA_{>1}^+(0,+)$& $x^2+y^2+h$ & $\emptyset$ \\ \hline $cA_{>1}^+(0,-)$& $x^2+y^2-h$ & $S^1\times S^1$ \\ \hline $cA_{>1}^+(m)$& $x^2+y^2\pm hf_1\cdots f_m$ & $\uplus m S^2$\\ \hline $cA_{>1}^-(0)$& $x^2-y^2\pm h$ & $S^2\uplus S^2$ \\ \hline $cA_{>1}^-(m)$& $x^2-y^2\pm hf_1\cdots f_m$ &$M_{m-1}$\\ \hline \end{tabular} $$ \end{thm} Proof. We already have the form $x^2\pm y^2+f(z,t)$ by (\ref{ca.thm}). Write $f$ as a product of irreducible power series over $\r$. Those factors which do not vanish on $\r^2\setminus\{0\}$ are multiplied together to get $h$. By writing $\pm h$ we may assume that $h$ is positive on $\r^2\setminus\{0\}$. (Since the signs of the other factors are not fixed, the sign of $h$ matters only if there are no other factors.) Let $f_i$ be the remaining factors of $f$. Assume now that we are in the $cA^+$-case: $x^2+ y^2\pm h\prod f_i$. Projection to the $(z,t)$-plane is a proper map whose fibers are as follows: \begin{enumerate} \item $S^1$ if $\pm h(z,t)\prod f_i(z,t)<0$, \item a point if $\pm h(z,t)\prod f_i(z,t)=0$, \item empty if $\pm h(z,t)\prod f_i(z,t)>0$. \end{enumerate} If $m=0$ then $X(\r)\setminus\{0\}$ is a circle bundle over either $\r^2\setminus\{0\}$ or over the empty set. The first case gives $L\sim S^1\times S^1$ by (\ref{orient}). If $m>0$, we have to describe the semi-analytic set $U:=(\prod f_i(z,t)\leq 0)\subset \r^2$. Semi-analytic sets can be triangulated (cf. \cite[9.2]{BCR87}), thus in a neighborhood of the origin, $U$ is the cone over $U\cap (z^2+t^2=\epsilon)$. Each $(f_i=0)$ is an irreducible curve germ over $\r$, thus homeomorphic to $\r^1$. So each $f_i$ has 2 roots on the circle $(z^2+t^2=\epsilon)$. Hence $U\cap (z^2+t^2=\epsilon)$ is the disjoint union of $m$ closed arcs. Therefore $L$ has $m$ connected components, each homeomorphic to $S^2$. The second possibility is the $cA^-$-case: $x^2- y^2- h\prod f_i$. (The two choices of $\pm h$ are equivalent by interchanging $x$ and $y$.) Here we project to the $(y,z,t)$-hyperplane. The fiber over a point $(y,z,t)$ is \begin{enumerate} \item 2 points if $y^2+ h(z,t)\prod f_i(z,t)>0$, \item 1 point if $y^2+ h(z,t)\prod f_i(z,t)=0$, \item empty if $y^2+ h(z,t)\prod f_i(z,t)<0$. \end{enumerate} Thus we have to determine the region $$ U:=\{y^2+ h(z,t)\prod f_i(z,t)\geq 0\}\subset (y^2+z^2+t^2=\epsilon)\sim S^2, $$ and then take its double cover to get $L$. If $m=0$ then $U=S^2$ and so $L=S^2\uplus S^2$. If $m>0$ then $h\prod f_i$ is negative on $m$ disjoint arcs in the circle $(z^2+t^2=\epsilon)$, and $y^2+ h(z,t)\prod f_i(z,t)$ is negative in contractible neighborhoods of these intervals. Thus $U=S^2\setminus(\mbox{$m$-discs})$ and so $L$ is a surface of genus $m-1$, orientable by (\ref{orient}). \qed \begin{exmp} It is quite instructive to consider the following incorrect approach to the topology of $cD$ and $cE$-type singularities. I illustrate it in the $cE_6$-case. Over $\r$, a surface singularity of type $E_6$ is $x^2+y^2\pm z^4$. In both cases, projection to the $(x,z)$-plane is a homeomorphism. Consider a $cE_6$-type point $X$. If $L(X(\r))$ has several connected components, then a suitable hyperplane intersects at least two of them. By a small perturbation we obtain an $E_6$-singularity as the intersection, thus we conclude that $L(X(\r))$ is connected. This is especially suggestive if we note that instead of a plane we could use a small perturbation of any smooth hypersurface. Unfortunately the conclusion is false, as we se in (\ref{ce6.exmp}). $L(X(\r))$ can have several components, and some of them are not seen by general hypersurface sections. These look like very ``thin" cones, as opposed to the main component which is ``thick". It would be interesting to give precise meaning to this observation and to see its significance in the study of singularities. \end{exmp} The following approach to the topology of $cD$ and $cE$-type singularities is taken from \cite[Sec.12]{AGV85}. \begin{say}[Deformation to the weighted tangent cone]{\ } Let $X:=(f(x_1,\dots,x_n)=0)$ be a hypersurface singularity. For simplicity of notation I assume that $f$ converges for $|x_i|< 1+\delta$. Assign integral weights to the variables $w(x_i)=w_i$ and write $f$ as the sum of weighted homogeneous pieces $$ f=f_d+f_{d+1}+f_{d+2}+\dots, $$ where $f_s$ is weighted homogeneous of degree $s$. For a parameter $\lambda\neq 0$ set \begin{eqnarray*} f^{\lambda}(x_1,\dots,x_n):= \lambda^{-d}f(\lambda^{w_1}x_1,\dots,\lambda^{w_n}x_n)\\ = f_d+\lambda f_{d+1}+\lambda^2f_{d+2}+\dots \end{eqnarray*} This suggests that if we define $f^0:=f_d$ then $$ X^{\lambda}:=(f^{\lambda}=0)\qtq{for} \lambda\in \r $$ is a ``nice" family of hypersurface singularities. For $\lambda\neq 0$ they are all isomorphic to $(f=0)$ and for $\lambda=0$ we obtain the weighted tangent cone $(f_d=0)$. This can be used to determine the topology of $X(\r)$ in 2 steps. First describe $X^0$ and then try to relate $X^{\lambda}$ and $X^0$ for small values of $\lambda$. Let $w$ be a common multiple of the $w_i$ and set $u_i=w/w_i$. \end{say} \begin{prop}\label{def.nc.prop} Notation as above. Assume that $X=(f=0)$ is an isolated hypersurface singularity. Then there is a $0<\lambda_0$ such that for every $0<\lambda\leq \lambda_0$ \begin{enumerate} \item $L^{\lambda}:=X^{\lambda}\cap (\sum x_i^{2u_i}=1)$ is smooth, and \item $X^{\lambda}\cap (\sum x_i^{2u_i}\leq 1)$ is homeomorphic to the cone over $L^{\lambda}$. \end{enumerate} \end{prop} Proof. The map $\r^m\to \r^+$ given by $(x_1,\dots,x_n)\mapsto\sum x_i^{2u_i}$ is proper. Thus its restriction to $X^{\lambda}$ is also proper. The proposition follows once we establish that the resulting map $$ t:X^{\lambda}\to \r $$ has no critical points with critical value in $(0,1]$ for $0<\lambda\leq \lambda_0$. The critical values of a real algebraic morphism form a semi-algebraic set (cf. \cite[9.5]{BCR87}), thus there is a $0<\mu_0$ such that $t:X^1\to \r$ has no critical values in $(0,\mu_0]$. The following diagram is commutative $$ \begin{array}{ccc} X^{\lambda} & \stackrel{x_i\mapsto \lambda^{-w_i}x_i}{\longrightarrow} & X^1\\ t\downarrow\ & & \ \downarrow t\\ \r & \stackrel{s\mapsto \lambda^{-w}s}{\longrightarrow}& \r \end{array} $$ which shows that (\ref{def.nc.prop}) holds with $\lambda_0=\mu_0^{1/w}$.\qed So far we have not done much, but the advantage of this approach is that we can view $L^{\lambda}$ as a deformation of the compact real algebraic variety $L^0$. If $L^0$ is smooth then this deformation is locally trivial differentiably. Thus we obtain: \begin{cor} Assume that $(f_d=0)$ defines an isolated singularity. Then $L^{\lambda}$ is diffeomorphic to $L^0$.\qed \end{cor} This is sufficient to describe the the topology of ``general" members of several families of terminal singularities: \begin{cor}\label{ce.top.gen} Let $X$ be a terminal singularity given by one of the of the following equations: \begin{enumerate} \item[$cD_4$] \quad $x^2+f_{\geq 3}(y,z,t)$, where $f_3=0$ has no real singular point. \item[$cE_6$] \quad $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$, where $h_4$ has no multiple real linear factor. \item[$cE_8$] \quad $x^2+y^3+yg_{\geq 4}(z,t)+h_{\geq 5}(z,t)$, where $h_5$ has no multiple real linear factor. \end{enumerate} Then: \begin{enumerate} \item[$cD_4$] \quad $L(X(\r))\sim S^2$ if $(f_3=0)\subset \r\p^2$ has one connected component and $L(X(\r))\sim S^2\uplus (S^1\times S^1)$ if $(f_3=0)$ has two connected components. \item[$cE_6$] \quad $L(X(\r))\sim S^2$. \item[$cE_8$] \quad $L(X(\r))\sim S^2$. \end{enumerate} \end{cor} Proof. We use deformation to the weighted tangent cone with (suitable integral multiples of the) weights $(1/2,1/3,1/3,1/3)$ in the $cD_4$-case, $(1/2,1/3,1/4,1/4)$ in the $cE_6$-case, and $(1/2,1/3,1/5,1/5)$ in the $cE_8$-case. The equations for $X^0$ are $x^2+f_3(y,z,t)=0$, $x^2+y^3+h_4(z,t)=0$ and $x^2+y^3+h_5(z,t)=0$. Our conditions guarantee that $X^0(\r)$ has isolated singularities, thus it is sufficient to determine $L(X^0(\r))$. In the $cE$-cases, projection to the $(x,z,t)$ hyperplane is a homeomorphism from $X^0(\r)$ to $\r^3$, thus $L(X^0(\r))\sim S^2$ In the $cD_4$-cases we project to the $(y,z,t)$-hyperplane. As in the proof of (\ref{can-top.thm}), we can get $L(X^0(\r))$ once we know the set $U\subset (y^2+z^2+t^2=1)$ where $f_3$ is nonnegative. The boundary $\partial U$ doubly covers the projective curve $(f_3=0)\subset \r\p^3$. If $(f_3=0)\subset \r\p^2$ has one connected component then it is a pseudo-line and $\partial U$ is a connected double cover, hence $U$ is a disc. If $(f_3=0)$ has two connected components, then one is a pseudo-line the other an oval. $\partial U$ has 3 connected components, and $U$ is a disc plus an annulus. Thus $L(X^0(\r))\sim S^2\uplus (S^1\times S^1)$. \qed \begin{rem} In the $cE$ cases of the above example, projection to the $(x,z,t)$ plane is a homeomorphism from $X^0(\r)$ to $\r^3$ even if $h_4$ or $h_5$ have multiple factors. In these cases, however, we can not conclude that $X(\r)$ is also homeomorphic to $\r^3$. In fact we see in (\ref{ce6.exmp}) that this is not always true. \end{rem} Similar arguments work in some of the $cD_{>4}$-cases: \begin{cor}\label{cd.top.gen} Let $X$ be a terminal singularity given by equation $x^2+y^2z+h_{\geq s}(z,t)$, where $z\not\vert h_{s}$ and $h_{s}$ has no multiple real linear factors. Let $s$ be the number of real linear factors of $h_{s}$. There are three cases: \begin{enumerate} \item $s=2r+1$ and $L(X(\r))\sim M_r\uplus rS^2$; \item $s=2r,\ h(0,1)<0$ and $L(X(\r))\sim M_r\uplus (r-1)S^2$; \item $s=2r,\ h(0,1)>0$ and $L(X(\r))\sim M_{r-1}\uplus rS^2$. \end{enumerate} \end{cor} Proof. We use deformation to the weighted tangent cone with weights $(1/2,(s-1)/2s,1/s,1/s)$. Thus we need to figure out the topology of $x^2+y^2z+h_{(n-1)}(z,t)=0$. As before, this reduces to understanding the set where $y^2z+h_{s}(z,t)\leq 0$. This can be done by projecting to the $(z,t)$ plane. Details are left to the reader. \qed \begin{say}\label{choose.weights} For many terminal singularities one can not choose weights so that the weighted tangent cone has an isolated singularity at the origin, but in all cases it is possible to choose weights so that the weighted tangent cone has 1-dimensional singular locus: $$ \begin{tabular}{|c|l|ccc|} \hline name &\quad equation is \quad $x^2+$ &$w(y)$ &$w(z)$&$w(t)$\\ \hline $cD_4$& $f_{\geq 3}(y,z,t)$ & $\frac13$&$\frac13$&$\frac13$\\ \hline $cD_{>4}(1)$& $y^2z+ayt^r+h_{\geq s}(z,t)$ & $\frac{s-1}{2s}$&$\frac1{s}$&$\frac1{s}$\\ \hline $cD_{>4}(2)$& $y^2z\pm yt^r+h_{\geq s}(z,t)$ & $\frac{r-1}{2r-1}-$\mbox{\scriptsize $\epsilon$}& $\frac1{2r-1}+$\mbox{\scriptsize $2\epsilon$} &$\frac1{2r-1}+\frac{\epsilon}{r}$\\ \hline $cE_6$ &$y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$ & $\frac13$&$\frac14$&$\frac14$\\ \hline $cE_7$ &$y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)$ & $\frac13$&$\frac29$&$\frac29$\\ \hline $cE_8$ &$y^3+yg_{\geq 4}(z,t)+h_{\geq 5}(z,t)$ & $\frac13$&$\frac15$&$\frac15$\\ \hline \end{tabular} $$ In the $cD_{>4}$ case we asume that $h_s\neq 0$ and use the first weight sequence if $a=0$ or $2r>s+1$ and the second weight sequence if $2r\leq s+1$, where $\epsilon$ is a small positive number. (We could use $\epsilon=0$ except when $2r=s-1$.) Let $(w_1,w_2,w_3,w_4)$ be integral multiples of these weights. The weighted tangent cone, and its singularities are the following: $$ \begin{tabular}{|c|l|l|} \hline name &weighted tangent cone& \qquad singularities \\ \hline $cD_4$& $x^2+f_3(y,z,t)$ &at singular pts of $(f_3=0)$\\ \hline $cD_{>4}(1)$& $x^2+y^2z+h_s(z,t)$&at multiple real factors of $zh_s$\\ \hline $cD_{>4}(2)$& $x^2+y^2z\pm yt^r$ & at the $z$-axis\\ \hline $cE_6$ &$x^2+y^3+h_4(z,t)$& at multiple real factors of $h_4$\\ \hline $cE_7$ &$x^2+y^3+yg_3(z,t)$& at real factors of $g_3$\\ \hline $cE_8$ &$x^2+y^3+h_5(z,t)$& at multiple real factors of $h_5$\\ \hline \end{tabular} $$ These equations have the form $x^2+F(y,z,t)$ and the deformation to the weighted tangent cone leaves this form invariant: $$ x^2+F^{\lambda}(y,z,t)=x^2+\lambda^{-d}F(\lambda^{w_2}y,\lambda^{w_3}z,\lambda^{w_4}t). $$ Set $$ U^{\lambda}:=\{(y,z,t)\vert F^{\lambda}(y,z,t)\leq 0\subset (y^{2u_2}+z^{2u_3}+t^{2u_4}=1)\}. $$ $U^{\lambda}$ is a semi-algebraic set and its boundary is the real algebraic curve $$ C^{\lambda}:=\{(y,z,t)\vert F^{\lambda}(y,z,t)=0\subset (y^{2u_2}+z^{2u_3}+t^{2u_4}=1)\}. $$ We have established the following: \end{say} \begin{prop} $C^{\lambda}$ is a deformation of the real algebraic curve $C^0$ inside the smooth real algebraic surface $(y^{2u_2}+z^{2u_3}+t^{2u_4}=1)$.\qed \end{prop} \begin{say} The deformations of real algebraic curves can be understood in two steps (cf. \cite{Viro90}). Put small discs around the singularities. Outside the discs all small deformations are topologically trivial and inside the discs we have a local problem involving real curve singularities. Here we have the advantage that $(y^{2u_2}+z^{2u_3}+t^{2u_4}=1)$ is an affine algebraic surface, thus we can choose the local deformations independently and they can always be patched together. Thus we can describe the possible cases for $C^{\lambda}$, and thereby the topological types of the corresponding 3-dimensional terminal singularities, if we can describe the deformations of the occurring real plane curve singularities. By looking at the equations we see that the only singularities that we have to deal with are the 2-variable versions of the Du Val singularities: \begin{enumerate} \item[$A_n$] \quad $y^2\pm z^{n+1}=0$; \item[$D_n$] \quad $y^2z\pm z^{n-1}=0$; \item[$E_6$] \quad $y^3+z^4=0$; \item[$E_7$] \quad $y^3\pm yz^3=0$; \item[$E_8$] \quad $y^3+z^5=0$. \end{enumerate} For all of these cases, a complete list of the topological types of real deformations is known \cite{Chislenko88}. The list can also be found in \cite[Figs. 16--28]{Viro90}, which contains many further examples. \end{say} \begin{exmp} Consider for instance the $cD_4$ cases. The various possibilities for $f_3$ are easy to enumerate. The most interesting is $f_3=yzt$. Here $C^0$ is the union of the 3 coordinate hyperplanes intersecting $(y^{6}+z^{6}+t^{6}=1)$. We have 6 singular points of type $u^2-v^2=0$. At each of them we can choose a deformation $u^2-v^2\pm \epsilon=0$. This gives $2^6$ possibilities. The symetries of the octahedron act on the configurations so it is easy to get a complete list. At the end we get $7$ possible topological types for $L(X(\r)$ where $X=(x^2+yzt+f_{\geq 4}(y,z,t)=0)$: $$ M_2, M_1\uplus S^2, M_1, S^2, 2S^2, 3S^2, 4S^2. $$ It truns out that these exhaust all the cases given by $cD_4$. \end{exmp} \begin{exmp}\label{ce6.exmp} Consider the $cE_6$-type points $$ x^2+y^3+yg_{\geq 3}(z,t)\pm z^2t^2+h_{\geq 5}(z,t). $$ Using the methods of (\ref{nf.meth.weighted}) these can be brought to the form $$ x^2+y^3\pm z^2t^2+ya(z)+yb(t)+c(z)+d(t). $$ The weighted tangent cone, $(x^2+y^3\pm z^2t^2=0)$ is singular along the $z$ and $t$-axes. In order to understand the sigularity type of $C^{\lambda}$, say along the positive $z$-halfaxis, set $t=\epsilon$. We get an equation $$ x^2+y^3\pm z^2\epsilon^2+ya(z)+yb(\epsilon)+c(z)+d(\epsilon). $$ $\mult_0a\geq 3$ and $\mult_0c\geq 5$, thus all the terms involving $z$ can be absorbed into $z^2$, and we obtain the equivalent form $$ x^2\pm z^2+y^3+yb(\epsilon)+d(\epsilon). $$ The cubic $y^3+yb(\epsilon)+d(\epsilon)$ has 3 real roots if $4b(\epsilon)^3+27d(\epsilon)^2<0$ and 1 real root if $4b(\epsilon)^3+27d(\epsilon)^2>0$. $C^0$ is homeomorphic to $S^1$ and it has 4 singular points (along the $z$ and $t$ halfaxes). $C^{\lambda}$ is a smooth curve which has an oval near a singular point of $C^0$ if the corresponding cubic has 3 real roots and no ovals if only 1 real root. Thus $C^{\lambda}$ has at most 5 connected components. We can also determine the location of the ovals relative to the ``main component" of $C^{\lambda}$. In deforming $z^2-y^3=0$, the oval can appear only toward the negative $y$-direction. Putting all this together, we get the following possibilities for $L(X(\r))$: $$ \begin{array}{ll} rS^2,\ 1\leq r\leq 5 &\mbox{in the $+z^2t^2$-case, and} \\ M_r,\ 0\leq r\leq 4 &\mbox{in the $-z^2t^2$-case.} \end{array} $$ \end{exmp} \section{The topology of terminal quotient singularities} Let $0\in X$ be 3-fold terminal singularity and $\pi:\tilde X\to X$ its index one cover. As we proved, $X=\tilde X/\frac{1}{n}(a_1,\dots,a_m)$ where $n$ is the index of $X$ and the $a_i$ are integers. We use this representation to determine the topology of $X$ in terms of the already known topology of $\tilde X$. The main question is to determine the real points of ${\Bbb A}^m/\frac{1}{n}(a_1,\dots,a_m)$. Let $\sigma:\c^m\to \c^m$ be the corresponding action of $1\in \z_n$. The answer depends on the parity of $n$. First we discuss the odd index cases which are easier. \begin{prop}\label{odd.quot} Assume that $n$ is odd and set $Y={\Bbb A}^m/\frac{1}{n}(a_1,\dots,a_m)$. Then the induced map $\r^n\to Y(\r)$ is a homeomorphism. \end{prop} Proof. Let $R\subset \r[[x_1,\dots,x_m]]$ denote the ring of invariant functions. A point $P\in {\Bbb A}^m$ maps to a real point of $X$ iff $f(P)\in \r$ for every $f\in R$. Let $\epsilon$ be a primitive $n^{th}$ root of unity. If $P=(p_1,\dots,p_m)$ is real then $$ \sigma^b(p_1,\dots,p_m)=(\epsilon^{ba_1}p_1,\dots,\epsilon^{ba_m}p_m) $$ is also real iff $\sigma^b(P)=P$. This shows that the quotient map $\r^n\to X(\r)$ is an injection. Let $Q\in X(\r)$ be a point. Then $\pi^{-1}(Q)\subset {\Bbb A}^m$ has an odd number of closed points over $\c$ (usually $n$ of them) and as a scheme it is defined over $\r$. Thus it has a real point, hence $\r^n\to X(\r)$ is also surjective.\qed \begin{cor}\label{oddind.termquot} Let $0\in X$ be a 3-fold terminal singularity of odd index and $\pi:\tilde X\to X$ its index one cover. Then $\pi: \tilde X(\r)\to X(\r)$ is a homeomorphism.\qed \end{cor} The even index case is more subtle. For purposes of induction we allow the case when $n$ is odd. Consider the action $\frac{1}{n}(a_1,\dots,a_m)$ on ${\Bbb A}^m$. Write $n=2^sn'$ where $n'$ is odd. Let $\eta$ be a primitive $2^{s+1}$-st root of unity and $j:\r^m\to \c^m$ the map $$ j(x_1,\dots,x_m)=(\eta^{a_1}x_1,\dots,\eta^{a_m}x_m). $$ (If $n$ is odd then $\eta=-1$, hence $j(\r^m)=\r^m$.) Write $a_i=2^ca'_i$ such that $a'_i$ is odd for some $i$. If $s>c$, let $\tau:\c^m\to \c^m$ be the $\z_2$-action $$ \tau(x_1,\dots,x_m)=((-1)^{a'_1}x_1,\dots,(-1)^{a'_m}x_m). $$ For $s=c$ let $\tau$ be the identity. Note that both $\r^n$ and $j(\r^n)$ are $\tau$-invariant. \begin{prop}\label{2ind.quot} Set $Y={\Bbb A}^m/\frac{1}{n}(a_1,\dots,a_m)$. Define $j$ and $\tau$ as above. Then $Y(\r)$ is the quotient of $\r^m\cup j(\r^m)$ by $\tau$. \end{prop} Proof. The proof is by induction on $m$ and $n$. We can assume that the action is faithful, that is $\sum b_ia_i=1$ is solvable in integers. Indeed, for non faithful actions we get the same quotient from a smaller group action. The definitions of $j$ and $\tau$ are set up such that they do not change if we change the group this way. Set $F:=\prod_i x_i^{b_i}$. By induction on $m$ we know that (\ref{2ind.quot}) holds on each coordinate hyperplane. Thus we have to deal with points $P=(p_1,\dots,p_m)$ such that each $p_i\neq 0$. Assume that $\pi(P)$ is real. Let $\epsilon$ be a primitive $(2n)^{th}$ root of unity. $p_i^n$ is real, hence $p_i=\epsilon^{c_i}\cdot(\mbox{real number})$ for some $c_i\in \z$. Thus $F(P)=\epsilon^{c}\cdot(\mbox{real number})$. $F(\sigma(P))=\epsilon^2F(P)$, hence by replacing $P$ by $\sigma^r(P)$ for some $r$ we may assume that $F(P)\in\r$ or $F(P)\in\eta\cdot\r$. Assume first that $F(P)$ is real. For each $i$ the function $F^{n-a_i}x_i$ is invariant, hence has a real value at $P$. Thus $P\in \r^n$. If $F(P)\in\eta\cdot\r$ then the same agrument shows that $p_i\in \eta^{a_i}\cdot \r$, thus $P\in j(\r^m)$. This shows that $\r^m\cup j(\r^m)\to Y(\r)$ is surjective. It is also $\tau$-invariant. Finally, if $P=(p_1,\dots,p_m)\in \r^m\cup j(\r^m)$ then $$ \sigma^s(P)=(\epsilon^{ba_1}p_1,\dots,\epsilon^{ba_m}p_m) \in \r^m\cup j(\r^m) $$ iff $\sigma^s(P)=P$ or $\sigma^s(P)=\tau(P)$. Thus $\r^m\cup j(\r^m)\to Y(\r)$ is $2:1$ for $s>c$ and $1:1$ for $s=c$. \qed \begin{defn} Let $F\in \r[[x_1,\dots,x_m]$ be a power series, homogeneous of grade $d$ under the grading $\frac{1}{n}(a_1,\dots,a_m)$. Let $\eta$ be as above. Define the {\it companion} $F^c$ of $F$ with respect to the action $\frac{1}{n}(a_1,\dots,a_m)$ by $$ F^c(x_1,\dots,x_m):=\eta^{-d}F(\eta^{a_1}x_1,\dots,\eta^{a_m}x_m). $$ Note that $F^c\in \r[[x_1,\dots,x_m]]$. \end{defn} \begin{cor}\label{2ind.termquot} Let $0\in X$ be a 3-fold terminal singularity of even index and $\pi:\tilde X\to X$ its index one cover. Then $$ L(X(\r)) \sim L(\tilde X(\r))/(\tau)\uplus L(\tilde X^c(\r))/(\tau). $$ \end{cor} Proof. We use the notation of (\ref{2ind.quot}). Let $F=0$ be the equation of $\tilde X$ and let $W:=(F=0)\cap(\r^m\cup j(\r^m))$. Then $X(\r)$ is the quotient of $W$ by $\tau$. $(F=0)\cap\r^m=\tilde X(\r)$. $(F=0)\cap j(\r^m)$ can be identified with the set of real zeros of $F(\eta^{a_1}x_1,\dots,\eta^{a_m}x_m)=0$. (The normalizing factor $\eta^{-d}$ does not change the set of zeros.) In the terminal case the group action is fixed point free on $(F=0)\setminus\{0\}$, thus $(F=0)\cap\r^m$ and $(F=0)\cap j(\r^m)$ intersect only at the origin.\qed As a byproduct we obtain the following: \begin{cor}\label{h.ind.not.isol} Let $0\in X$ be a 3-fold terminal singularity of index$>1$. Then $0\in X(\r)$ is not an isolated point. \end{cor} Proof. Let $\tilde X$ be the index 1 cover. We are done, unless $0\in \tilde X(\r)$ is an isolated point. This happens only in cases $cA/2, cAx/2$ (and maybe for $cAx/4$) where the equation of $\tilde X$ is $F=x^2+y^2+f(z,t)$ and $f(z,t)$ is positive on $\r^2\setminus\{0\}$. Let us compute $F^c$. In the $cA/2$ ase we get $-x^2-y^2+f(iz,t)$ and this has nontrivial solutions in the $(x,y,t)$-hyperplane. In the $cAx/2$ case we get $x^2-y^2+f(iz,it)$ and this has nontrivial solutions in the $(x,y)$-plane. In the $cAx/4$ case already $\tilde X$ has nontrivial $\r$-points. Indeed, here $f$ has grade 2, thus every $t$-power in it has an odd exponent. Thus $f(z,t)$ is not positive on the $t$-axis. \qed \begin{say}[Orientability of index 2 quotients] We have seen in (\ref{orient}) that every real algebraic hypersurface is orientable, and so are their quotients by odd order groups (\ref{odd.quot}). With index 2 quotients, the question of orientablilty is interesting. Consider a quotient $(F=0)/\frac12(w_1,\dots,w_m)$. Let $\sigma$ be the corresponding $\z_2$ action. We can orient $X:=(F=0)$ by choosing an orientation of $\r^m$ and at each smooth point of $X$ we choose the normal vector pointing in the direction where $F$ is positive. $\sigma$ preserves the orientation of $\r^m$ iff $\sum w_i$ is even. The parity of $w(F)$ determines the sign in $\sigma(F)=\pm F$. Thus $\sigma$ preserves the induced orientation of $X$ iff $w(F)+\sum w_i$ is even. If $w(F)+\sum w_i$ is odd, the induced orientation is not preserved. If $\sigma$ fixes a connected component of (the nonsingular part of) $X(\r)$, then the corresponding quotient is not orientable. If, however, $\sigma$ only permutes the connected components of $X(\r)$ then the quotient is still orientable. Thus we obtain: \begin{lem}\label{ind2.orient} Let $0\in X:=(F=0)/\frac12(w_1,\dots,w_m)$ be an isolated singular point, with index one cover $\tilde X$ and companion $\tilde X^c$. Then $L(X(\r))$ is nonorientable iff $w(F)+\sum w_i$ is odd and $\sigma$ fixes at least one of the connected components of $L(\tilde X(\r))$ or $L(\tilde X^c(\r))$.\qed \end{lem} \end{say} \begin{exmp}[The topology of $cA/2$ points] {\ } The simplest case is ${\Bbb A}^3/{\textstyle \frac1{2}(1,1,1)}$. The link of ${\Bbb A}^3$ is the sphere $(x^2+y^2+z^2=\epsilon^2)$. We act by the antipodal map, and the quotient is $\r\p^2$. The quotient of the purely imaginary subspace also gives real points, thus $L(X(\r))\sim 2\r\p^2$. In the $cA_{>0}/2$ cases write $$ X:=(x^2\pm y^2+f(z,t)=0)/\frac12(1,1,1,0) $$ with cover $\tilde X:=(x^2\pm y^2+f(z,t)=0)$. Only even powers of $z$ occur in $f(z,t)$, thus we can write $f(z,t)=G(z^2,t)$. As in (\ref{can-top.thm}) we factor it as $G(z^2,t)=\pm h(z,t)\prod_{i=1}^r f_i(z,t)$. The companion cover is $\tilde X^c=(x^2\pm y^2-G(-z^2,t)=0)$. We have to be careful since the product decomposition of $G$ is not preserved. A factor of $h$ may become indefinite and it can also happen that two factors become conjugate over $\r$. Thus we write $-G(-z^2,t)=\pm h'(z,t)\prod_{j=1}^{r'} f'_j(z,t)$. By (\ref{2ind.termquot}), $L(X(\r))=L(\tilde X(r))/\tau\uplus L(\tilde X^c(r))/\tau$. In all these cases (\ref{ind2.orient}) shows that if $\tau$ fixes a connected component, the quotient is not orientable. Thus if $L(\tilde X(r))$ or $L(\tilde X^c(r))$ is connected, the quotient is not orientable. This holds in all the $cA^-_{>0}(r>0)$ cases. In the $cA^-_{>0}(0)$ case the equation is $(x^2=y^2+h)$, and each halfspace of $(x\neq 0)$ contains a unique connected component. Thus $\tau$ interchanges the two connected components and the quotient is orentable. In the $cA^+_{>0}$ case $(\pm h(z,t)\prod_{i=1}^r f_i(z,t)=0)$ is negative on $r$ connected regions $P_1,\dots,P_r\subset \r^2$, and $L(\tilde X(\r))$ consist of $r$ copies of $S^2$, one for each $P_j$. The involution $\tau$ fixes the $t$-axis pointwise, thus if one of the half $t$-axes is contained in some $P_j$, then $\tau$ fixes the $S^2$ over that region. The other copies of $S^2$ are interchanged. The same holds for $\tilde X^c$. Along the $t$-axis $G(z^2,t)$ and $-G(-z^2,t)$ have opposite signs. Thus among the 4 pairs $$ \begin{array}{cc} (\mbox{positive half $t$-axis}, G(z^2,t)) & (\mbox{positive half $t$-axis}, -G(-z^2,t))\\ (\mbox{negative half $t$-axis}, G(z^2,t))& (\mbox{negative half $t$-axis}, -G(-z^2,t)) \end{array} $$ there are two where the function is negative along the half axis. We obtain the following list of possibilities. (We use the notation $K_r:=S^2\# r\r\p^2$, thus $K_2$ is the Klein bottle.) $$ \begin{tabular}{|l|l|c|} \hline \qquad $\tilde X$ & \qquad $\tilde X^c$ & $L(X(\r))$ \\ \hline $cA_0$& $cA_0$ & $\r\p^2\uplus \r\p^2$ \\ \hline $cA_{>0}^-(r>0)$& $cA_{>0}^-(r'>0)$ & $K_r\uplus K_{r'}$ \\ \hline $cA_{>0}^-(r>0)$& $cA_{>0}^-(0)$ & $K_r\uplus S^2$ \\ \hline $cA_{>0}^-(0)$& $cA_{>0}^-(0)$ & $S^2\uplus S^2$ \\ \hline $cA_{>0}^+(r>0)$& $cA_{>0}^+(r'>0)$ & $2\r\p^2\uplus {\textstyle\frac{r+r'-2}{2}}S^2$ \\ \hline $cA_{>0}^+(r>0)$& $cA_{>0}^+(0,+)$ & $2\r\p^2\uplus {\textstyle\frac{r-2}{2}}S^2$ \\ \hline $cA_{>0}^+(r>0)$& $cA_{>0}^+(0,-)$ & $K_2\uplus {\textstyle\frac{r}{2}}S^2$ \\ \hline $cA_{>0}^+(0,-)$& $cA_{>0}^+(0,+)$ & $K_2$ \\ \hline \end{tabular} $$ Thus $X(\r)$ is not orientable, except in the fourth case. This indeed occurs: Let $\tilde X:=(x^2-y^2+z^{4m}+t^{2n}=0)$ and $X:=\tilde X/{\textstyle \frac12(1,1,1,0)}$ with companion $\tilde X^c=(-x^2+y^2+z^{4m}+t^{2n}=0)$. $\tilde X^c\cong \tilde X$ and $L(\tilde X(\r))\sim S^2\uplus S^2$. $\tau$ interchanges the two copies of $S^2$. Thus $X(\r)$ is orientable and $ L(X(\r))\sim S^2\uplus S^2$. \end{exmp}
1997-12-15T18:23:18
9712
alg-geom/9712015
en
https://arxiv.org/abs/alg-geom/9712015
[ "alg-geom", "math.AG" ]
alg-geom/9712015
H. Esnault
H\'el\`ene Esnault, V. Srinivas, Eckart Viehweg
The universal regular quotient of the Chow group of 0-cycles on a singular projective variety
Latex 2e
null
null
null
null
We show the existence of a regular universal quotient as a smooth commutative algebraic group of the Chow group of 0-cycles on a projective reduced variety, and give over the field of complex numbers an analytic description of it. This generalizes the classical theory of the Albanese. The Chow group of 0-cycles is then isomorphic to this smooth algebraic group if and only if it is finite dimensional in the sense of Mumford. This generalizes the classical theorem of Roitman.
[ { "version": "v1", "created": "Mon, 15 Dec 1997 17:23:17 GMT" } ]
2007-05-23T00:00:00
[ [ "Esnault", "Hélène", "" ], [ "Srinivas", "V.", "" ], [ "Viehweg", "Eckart", "" ] ]
alg-geom
\section{Chow groups and regular homomorphisms} We begin by recalling the definition of the Chow group of 0-cycles $CH^n(X)$, as given in \cite{LW} (see also \cite{BiS}). As in \cite{BiS}, we adopt the convention that a point lying on a lower dimensional component of $X$ is deemed to be singular. Let $X_{{\rm sing}}$ denote the (closed) subset of singular points, and $X_{{\rm reg}}=X-X_{{\rm sing}}$ the complementary open set. The closure of $X_{{\rm reg}}$ is the union of the $n$-dimensional components of $X$. The group $Z^n(X)$ of 0-cycles is defined to be the free abelian group on the closed points of $X_{{\rm reg}}$. The subgroup $R^n(X)$ of cycles rationally equivalent to 0 is defined using the notion of a Cartier curve. \begin{defn}\label{ccurve} A {\em Cartier curve} is a subscheme $C\subset X$, defined over $k$, such that \begin{points} \item $C$ is pure of dimension 1 \item no component of $C$ is contained in $X_{{\rm sing}}$ \item if $x\in C\cap X_{{\rm sing}}$, then the ideal of $C$ in ${\mathcal O}_{x,X}$ is generated by a regular sequence (consisting of $n-1$ elements). \end{points} \end{defn} If $C$ is a Cartier curve on $X$, with generic points $\eta_1,\ldots,\eta_s$, and ${\mathcal O}_{S,C}$ is the semilocal ring on $C$ of the points of $S=(C\cap X_{{\rm sing}})\cup\{\eta_1,\ldots,\eta_s\}$, there is a natural map on unit groups \[\theta_{C,X}:{\mathcal O}_{S,C}^*\>>> \bigoplus_{i=0}^s{\mathcal O}_{\eta_i,C}^*.\] Define $R(C,X)={\rm image}\,\theta_{C,X}$. For $f\in R(C,X)$, define the divisor of $(f)_C$ as follows: let $C_i$ denote the maximal Cohen-Macaulay subscheme of $C$ supported on the component with generic point $\eta_i$. Then for any $x\in C_i$ the map \[{\mathcal O}_{x,C_i}\>>>{\mathcal O}_{\eta_i,C_i}={\mathcal O}_{\eta_i,C}\] is the injection of a Cohen-Macaulay local ring of dimension 1 into its total quotient ring. If $f_i$ is the component of $f$ in ${\mathcal O}_{\eta_i,C}$, then $f_i=a_x/b_x$ for some non zero-divisors $a_x,b_x\in{\mathcal O}_{x,C_i}$. Define \[(f)_C=\sum_{i=1}^s(f_i)_{C_i}=\sum_{i=1}^s\sum_{x\in C_i}(\ell({\mathcal O}_{x,C_i}/a_x{\mathcal O}_{x,C_i})-\ell({\mathcal O}_{x,C_i}/b_x{\mathcal O}_{x,C_i})) \cdot [x].\] Standard arguments imply that this is well-defined ({\it i.e.\/},\ the coefficient of $[x]$ is independent of the choice of the representation $f_i=a_x/b_x$, and vanishes for all but a finite number of $x$). Suppose $C$ is reduced. Then in the above considerations, ${\mathcal O}_{x,C_i}$ is an integral domain with quotient field ${\mathcal O}_{\eta_i,C}$. If $v_1, \ldots,v_m$ are the discrete valuations of ${\mathcal O}_{\eta_i,C}$ centered at $x$, then the multiplicity of $x$ in $(f)_{C_i}$ is \begin{equation}\label{val} \ell({\mathcal O}_{x,C_i}/a_x{\mathcal O}_{x,C_i})-\ell({\mathcal O}_{x,C_i}/b_x{\mathcal O}_{x,C_i})= \sum_{j=1}^m v_j(f_i) \end{equation} (compare \cite{Fulton}, Example~A.3.1.). In fact, let $R$ be the integral closure of $O={\mathcal O}_{x,C_i}$ in ${\mathcal O}_{\eta_i,C}$. The Chinese remainder theorem implies that \[\ell(R/a_xR) = \sum_{j=1}^m v_j(a_x),\] and similarly for $b_x$. Multiplying $a_x$ and $b_x$ by the same element of $O$ we may assume that both $a_xR$ and $b_xR$ are contained in $O$, and $$\ell(O/a_xO)+\ell(R/O) = \ell(R/a_xR) + \ell(a_xR/a_xO).$$ Since $a_x \neq 0$ the second terms on both sides are equal. \begin{defn}\label{rat-equ} Let $U \subset X_{\rm reg}$ be an open dense subscheme. $R^n(X,U)$ is defined to be the subgroup of $Z^n(U)$ generated by elements $(f)_C$ as $C$ ranges over all Cartier curves with $C\cap U$ dense in $C$, and $f\in R(C,X)$ with $(f)_C \in Z^n(U)$. For $U=X_{{\rm reg}}$ we write $R^n(X)$ instead of $R^n(X,X_{{\rm reg}})$ and define $$CH^n(X) = Z^n(X)/R^n(X).$$ Mapping a point $x \in X_{{\rm reg}}$ to its rational equivalence class defines a map $$\gamma :X_{{\rm reg}} \>>> CH^n(X).$$ If $U_1, \ldots , U_r$ denote the irreducible components of $X_{\rm reg}$ then $Z^n(X)_{\deg 0}$ and $CH^n(X)_{\deg 0}$ denote the subgroup of $Z^n(X)$ and $CH^n(X)$, respectively, of cycles $\delta$ with $\deg(\delta|_{U_i})=0$ for $i = 1, \ldots r$. \end{defn} As noted in \cite{BiS}, lemma~1.3 of \cite{LW} allows one to restrict to considering only curves $C$ such that $C\cap X_{{\rm reg}}$ has no embedded points, and any irreducible component $C'$ of $C$ which lies entirely in $X_{{\rm reg}}$ occurs in $C$ with multiplicity 1. The moving lemmas 2.2.2 and 2.2.3 of \cite{BiS} allow stronger restrictions on $C$: \begin{lemma}\label{moving} Let $A \subset X_{{\rm sing}}$ be a closed subset of dimension $\leq n-2$, and let $D \subset X$ be a closed subset of dimension $\leq n-1$. Then any element $\delta\in R^n(X)$ can be written in the form $\delta=(f)_C$ for a single (possibly reducible) Cartier curve $C$, such that \begin{enumerate} \item[(a)] $C$ is reduced \item[(b)] $C\cap A=\emptyset$ \item[(c)] $C\cap D$ is empty or consists of finitely many points. \end{enumerate} \end{lemma} \begin{cor}\label{equidim} If $U \subset X_{\rm reg}$ is an open and dense subscheme, then $$CH^n(X)=Z^n(U)/R^n(X,U) \mbox{ \ \ and \ \ }CH^n(X)_{\deg 0}=Z^n(U)_{\deg 0}/R^n(X,U).$$ \end{cor} \begin{proof} First note that the zero cycles supported on $U$ generate $CH^n(X)$ since the corresponding assertion holds true for curves. The moving lemma \ref{moving} for $D=X-U$ implies that $R^n(X)\cap Z^n(U)=R^n(X,U)$. \end{proof} \begin{rmk} Let $X^{(n)}$ denote the union of the $n$-dimensional irreducible components of $X$, and let $X^{<n}$ be the union of the lower dimensional components. Applying the corollary to $X^{(n)}$ and the open subset $U=X^{(n)}-X_{{\rm sing}}=X-X_{sing}$, we see that the natural map from $CH^n(X)$ to $CH^n(X^{(n)})$ is surjective. It seems plausible that a stronger form of lemma~\ref{moving} holds, where $A$ is allowed to be any closed subset of $X$ of codimension $\geq 2$ which is disjoint from ${\rm supp}\,(\delta)$. If this is true, then applying it to $X^{(n)}$ with $A=X^{(n)}\cap X^{<n}$, one sees that for any $\delta\in R^n(X^{(n)})\cap Z^n(X)$ there exists a reduced Cartier curve $C$ in $X^{(n)}$, disjoint from $A$, and $f \in R(C,X^{(n)})$ with $\delta=(f)_C$. Then $C$ is also a Cartier curve on $X$, and $\delta\in R^n(X)$. We deduce that $CH^n(X)\to CH^n(X^{(n)})$ is an isomorphism. We have as yet been unable to prove this. \end{rmk} \begin{rmk}\label{rmk-equi} Keeping the notation from the previous remark, we note further that for $k={\mathbb C}$, the natural maps \[H^{2n}(X,{\mathbb Z}(n))\>>> H^{2n}(X^{(n)},{\mathbb Z}(n)),\;\; D^n(X)\>>> D^n(X^{(n)}),\;\; A^n(X)\>>> A^n(X^{(n)})\] are isomorphisms, since $X^{<n}$ has constructible cohomological dimension $\leq 2(n-1)$ and coherent cohomological dimension $\leq n-1$. \end{rmk} As reflected by the notation, $R(C,X)$ depends on the pair $(C,X)$, and is not necessarily intrinsic to $C$. In fact, since we have not imposed any unit condition at singular points of $C$ which lie in $X_{{\rm reg}}$, the functions $f \in R(C,X)$ are defined on some curve $C'$, birational to $C$. \begin{defn}\label{admissible} Let $C'$ be a reduced projective curve and $\iota:C' \to X$ be a morphism. Then $(C',\iota)$ will be called {\it admissible} if $\iota: C' \to C=\iota(C')$ is birational, if $C$ is a reduced Cartier curve and if for some open neighbourhood $W$ of $X_{{\rm sing}}$ the restriction of $\iota$ to $\iota^{-1}(W)$ is a closed embedding. \end{defn} If $(C',\iota)$ is admissible one has an inclusion $R(C',C') \subset R(\iota(C),X)$ which is an equality if $\iota^{-1}(X_{{\rm reg}})$ is non-singular. \begin{lemma}\label{gysin} Let $(C',\iota)$ be admissible. Then there exists a homomorphism (of abstract groups) $$ \eta: {\rm Pic}^0(C') \cong CH^1(C')_{\deg 0} \>>> CH^n(X)_{\deg 0}$$ which maps the isomorphism class of ${\mathcal O}_{C'}(p-p')$ to $\gamma (\iota(p))-\gamma(\iota(p'))$. \end{lemma} \begin{proof} By definition ${\rm Pic}(C')= Z^1(C'_{\rm reg})/R(C',C')$ and one has a map $$\gamma \circ \iota : Z^1(C'_{\rm reg}) \>>> CH^n(X). $$ The equality (\ref{val}) shows that for $f\in R(C',C')$ the image of $(f)_{C'}$ in $CH^n(X)$ is zero. \end{proof} \begin{notations}\label{difference} Let $Y$ be a non-singular scheme with irreducible components $Y_1, \ldots ,Y_s$, let $G$ be an abstract or an algebraic group, and let $\pi: Y \to G$ a map or morphism. \begin{points} \item After choosing base points $p_i \in Y_i$ a map $$ \pi_m : S^m(Y):= S^m( \bigcup_{i=1}^s Y_i) \>>> G, $$ is defined by $\pi_m(y_1, \ldots, y_m)=\sum_{j=1}^m (\pi(y_j) - \pi(p_{\rho(j)}))$, where $\rho(j) = i$ if $y_j \in Y_i$. \item To avoid the reference to base points, we will frequently use different maps: $$ \pi^{(-)} : \Pi_Y = \bigcup_{i=1}^s Y_i\times Y_i \>>> G $$ is defined by $\pi^{(-)}(y,y') = \pi(y)-\pi(y')$, and $\pi^{(-)}_m: S^m(\Pi_Y) \to G$ is the composite $S^m(\Pi_Y) \to S^m(G) \> {\rm sum} >> G$. \end{points} \end{notations} If $G$ is an algebraic group, then the images of $\pi^{(-)}_m$ lie in the connected component of $0$. In particular for $U$ open and dense in $X_{\rm reg}$ we will frequently consider \begin{align*} &\gamma^{(-)} = \gamma_U^{(-)}: \Pi_U \>>> CH^n(X)_{\deg 0}\\ \mbox{and \ \ \ \ } &\gamma_m = \gamma_{U,m} : S^m(U) \>>> CH^n(X)_{\deg 0}. \end{align*} \begin{lemma} \label{generators} Let $G$ be a $d$-dimensional smooth connected commutative algebraic group and let $\Gamma \subset G$ be a constructible subset which generates $G$ as an abstract group. Then \begin{points} \item the image of the composite map \ $S^d(\Gamma) \>>> S^d(G) \>{\rm sum} >> G $ \ is dense \item $S^{2d}(\Gamma) \>>> S^{2d}(G) \>{\rm sum} >> G$ \ is surjective \item if $B$ is a non-singular scheme with connected components $B_1, \ldots ,B_s$ and if $\vartheta: B \to G$ is a morphism with image $\Gamma$ then the morphism $$ \vartheta_{d}^{(-)}: S^d(\Pi_B)=S^d \bigl( \bigcup_{i=1}^s B_i \times B_i \bigr) \>>> G $$ with $\vartheta_{d}^{(-)}((b_1,b_1'),\ldots (b_d,b_d')) = \sum_{i=1}^d (\vartheta(b_i)-\vartheta(b_i'))$ is surjective. \end{points} \end{lemma} \begin{proof} Let $\bar{\Gamma}_1, \ldots, \bar{\Gamma}_s$ be the irreducible components of the closure $\bar{\Gamma}$ of $\Gamma$, and let $\Gamma_i= \bar{\Gamma_i}\cap \Gamma$. It is sufficient to find non-negative integers $d_1, \ldots ,d_s$ with $\sum_{i=1}^s d_i \leq d$ such that the image of $S^{d_1}(\Gamma_1)\times \cdots \times S^{d_s}(\Gamma_s)$ is dense in $G$. To this end, we may assume that the identity of $G$ lies on each $\Gamma_i$. Let $\bar{\Gamma}_1^\nu$ be the closure of the image of $S^\nu(\Gamma_1)$ in $G$. Since $\bar{\Gamma}_1^\nu \subset \bar{\Gamma}_1^{\nu+1}$ there exists some $d_1 \leq d$ with $\bar{\Gamma}_1^{d_1} = \bar{\Gamma}_1^{d_1+1}$, and $d_1$ is minimal with this property. Hence $\bar{\Gamma}_1^{d_1} = \bar{\Gamma}_1^{2 d_1}$ and $\bar{\Gamma}_1^{d_1}$ is a subgroup of $G$ of dimension larger than or equal to $d_1$. If $s=1$, i.e. if $\bar{\Gamma}$ is irreducible, then $\bar{\Gamma}_1^{d_1}=G$. In general, replacing $G$ by $G/\bar{\Gamma}_1^{d_1}$ one obtains \ref{generators} (i) by induction on $s$. The second part is an easy consequence of (i). Let $U$ be an open dense subset of $G$, contained in $S^{d_1}(\Gamma_1)\times \cdots \times S^{d_s}(\Gamma_s)$. Given $p \in G$ the intersection of the two open sets $U$ and $p-U$ is non-empty and hence there are points $a,b \in U$ with $p-b=a$. For (iii) we may assume that for some point $b_i' \in B_i$, the image of $B_i\times\{b'_i\}$ in $\Gamma_i$ is dense, for each $i$. By (i) one finds $d_1, \ldots ,d_s$ with $\sum_{i=1}^s d_i = d$ such that the image $$ \vartheta^{(-)}_d(S^{d_1}(B_1\times \{b_1'\})\times \cdots \times S^{d_s}(B_s\times \{b_s'\})) $$ contains a subset $U$ which is open in $G$. Given $p \in G$ the intersection of $U$ and of $p+U$ is non empty and hence $p=a-b$ for two points $a$ and $b$ in $U$. Obviously $a-b$ lies in the image of $\vartheta^{(-)}_d$. \end{proof} \begin{cor}\label{generators2} Let $C'$ be a reduced curve, let $B_1, \ldots B_s$ be the connected components of $B=C'_{\rm reg}$, let $b'_j \in B_j$ be base points and let $\vartheta: B \to {\rm Pic}^0(C')$ be the morphism with $\vartheta|_{B_j}(b)={\mathcal O}_{C'}(b-b_j')$. Then there exists some open connected subscheme $W$ of $S^g(B)$, for $g=\dim_k({\rm Pic}^0(C'))$, such that $\vartheta_W := \vartheta_g|_W$ is an open embedding. \end{cor} \begin{proof} By \ref{generators} we find some $W$ with $\vartheta_W(W)$ open and $\vartheta_W$ finite over its image. On the other hand, any fibre of $\vartheta_g$ is an open subset of $\P(H^0(C',{\mathcal O}_{C'}(D)))$ for some divisor $D$ on $C'$; hence the projective spaces corresponding to points of $\vartheta_W(W)$ must be 0-dimensional. \end{proof} \begin{lemma}\label{pic} Let $G$ be a smooth commutative algebraic group, $U \subset X_{\rm reg}$ an open and dense subset, and $\pi: U \to G$ a morphism. Then the following two conditions are equivalent. \begin{enumerate} \item[(a)] There exists a homomorphism (of abstract groups) $\phi:CH^n(X)_{\deg 0} \to G$ such that $\pi^{(-)} = \phi \circ \gamma^{(-)}$ (as maps on the closed points). \item[(b)] For all admissible pairs $(C',\iota)$ with $B=(\iota^{-1}(U))_{\rm reg}$ dense in $C'$ there exists a homomorphism of algebraic groups $\psi:{\rm Pic}^0(C') \to G$ such that the diagram $$ \begin{CD} \Pi_B \>\vartheta^{(-)} >> {\rm Pic}^0(C')\\ {\mathbb V} \iota V V {\mathbb V} V \psi V \\ \Pi_U \> \pi^{(-)} >> G \end{CD} $$ commutes. Here $\vartheta: B \to {\rm Pic}(C')$ denotes the natural morphism, mapping a point $p$ to the isomorphism class of the invertible sheaf ${\mathcal O}_{C'}(p)$. \end{enumerate} Moreover, if the equivalent conditions (a) and (b) are true, the morphism $\psi$ in (b) factors as $$ \begin{TriCDV} {{\rm Pic}^0(C')}{\> \eta >>}{CH^n(X)_{\deg 0}} {\SE \psi E E }{\SW W \phi W}{G} \end{TriCDV} $$ and the image of $\phi:CH^n(X)_{\deg 0}\to G$ is contained in the connected component of the identity of $G$. \end{lemma} \begin{proof} Assume (a) and let $(C',\iota)$ be admissible and $g=\dim({\rm Pic}^0(C'))$. Choosing base points $b'_j \in B_j$, one finds by \ref{generators2} an open subscheme $W$ of $S^{g}(B)$ such that the morphism $\vartheta_W:W \to {\rm Pic}^0(C')$ is an open embedding. By \ref{gysin} one obtains a homomorphism $$ \psi:{\rm Pic}^0(C')\>\eta>>CH^n(X)_{\deg 0} \> \phi >> G $$ of abstract groups. By assumption $\pi^{(-)}$ is a morphism of schemes and the same holds true for $\pi^{(-)}\circ\iota: \Pi_B \to G.$ Thereby the restriction of $\psi$ to the open subscheme $W \subset {\rm Pic}^0(C')$ is a morphism of schemes, and being a homomorphism of abstract groups $\psi$ is a morphism of algebraic groups. Since each point of $X_{\rm reg}$ lies on some Cartier curve, the images of the connected algebraic groups ${\rm Pic}^0(C')$ generate $CH^n(X)_{\deg 0}$ and the image $\phi(CH^n(X)_{\deg 0})$ lies in the connected component of $G$, which contains the identity. The morphism $\pi^{(-)}$ induces a map $\tilde{\phi}:Z^n(U)_{\deg 0} \to G$ and it remains to verify that (b) implies that $\tilde{\phi}(R^n(X,U))=0$. By \ref{moving} each $\delta\in R^n(X,U)$ is of the form $(f)_C$ for a reduced Cartier curve $C$. There exists an admissible pair $(C',\iota)$ with $\iota(C')=C$ and with $\iota^{-1}(X_{\rm reg})$ non-singular. $(f)_C$ is the image of $(f)_{C'}$ in $Z^n(U)$ and by assumption $\iota\circ\pi^{(-)}$ factors through ${\rm Pic}^0(C')$. \end{proof} \begin{cor}\label{equ-reg} Let $\phi: CH^n(X)_{\deg 0} \to G$ be a homomorphism to a smooth commutative algebraic group $G$. Then the following conditions are equivalent: \begin{points} \item $\phi\circ\gamma^{(-)}: \Pi_{X_{\rm reg}} \>>> G$ is a morphism of schemes. \item There exists an open dense subscheme $U$ of $X_{\rm reg}$ such that $\phi\circ\gamma^{(-)}|_{\Pi_U}$ is a morphism of schemes. \item Given a base point $p_i$ on each irreducible component $U'_i$ of some open dense subscheme $U$ of $X_{\rm reg}$, the map $\pi: U \to G$ with $\pi|_{U'_i}(x) = \phi(x-p_i)$ is a morphism of schemes. \item Given any $m>0$ and base points $p_i$ on each irreducible component $U'_i$ of some open dense subscheme $U$ of $X_{\rm reg}$, $\phi\circ\gamma_m : S^m(U) \to G$ is a morphism of schemes. \end{points} \end{cor} Of course, ``$\pi$ is a morphism of schemes'' stands for ``there exists a morphism of schemes whose restriction to closed points coincides with $\pi$'', an abuse of terminology which we will repeat throughout this article. \begin{proof} Obviously (i) implies (ii). For $U \subset X_{\rm reg}$ given, the equivalence of (ii), (iii), and (iv) is an easy exercise. In fact, the morphism $\pi$ in \ref{equ-reg} (iii) is just $\phi\circ\gamma_1$. Assume that (iii) holds true for some $U$. We will show that the corresponding property holds true for $X_{\rm reg}$ itself. To this aim consider the map $\bar{\pi}: X_{\rm reg} \to G$ with $\bar{\pi}|_{U'_i}(x) = \phi(x-p_i)$ and the graph $\Gamma_{\bar{\pi}}$ of $\bar{\pi}$ in $X_{\rm reg}\times G$. By definition, $\Gamma_{\bar{\pi}} \cap U\times G$ is the graph $\Gamma_{\pi}$. Let $Z$ be the closure of $\Gamma_\pi$ in $X_{\rm reg}\times G$. $\Gamma_{\bar{\pi}}$ is contained in $Z$. In fact, given a point $x \in X_{\rm reg}$ one can find a Cartier curve $C$ through $x$ with $U\cap C$ dense in $C$ and with $B=C \cap X_{\rm reg}$ non-singular. By lemma \ref{pic} the morphism $(\pi|_{C\cap U})^{(-)} : \Pi_{C\cap U} \to G$ factors through a morphism ${\rm Pic}^0(C)\to G$ of algebraic groups and, in particular, it extends to a morphism $\Pi_{B} \to G$. Again this implies that the restriction of $\bar{\pi}$ to $B$ is a morphism, hence $\Gamma_{\bar{\pi}}\cap B\times G$ is closed and therefore contained in $Z$. By construction the morphism $p_1: Z \to X_{\rm reg}$ induced by the projection is birational and surjective. Let $V \subset X_{\rm reg}$ be the largest open subscheme with $p_1|_{p_1^{-1}(V)}$ an isomorphism. Then $\bar{\pi}|_V$ is a morphism of schemes and ${\rm codim}_{X_{\rm reg}}(X_{\rm reg}-V) \geq 2$. By theorem 1 in \cite{BLR}, 4.4, $\bar{\pi}|_V$ extends to a morphism $X_{\rm reg} \to G$. The graph of this morphism is contained in $Z$, hence it is equal to $Z$ and $\bar{\pi}$ is a morphism. \end{proof} We end this section by giving the definition of a regular homomorphism, used already in the formulation of the main theorems in the introduction. \begin{defn}\label{def-reg} Let $G$ be a smooth commutative algebraic group. A homomorphism $\phi:CH^n(X)_{\deg 0} \to G$ (of abstract groups) is called {\it a regular homomorphism}, if one of the equivalent conditions in \ref{equ-reg} holds true. \end{defn} \begin{lemma}\label{reg_sur} The image of a regular homomorphism $\phi:CH^n(X)_{\deg 0} \to G$ is a connected algebraic subgroup of $G$. \end{lemma} \begin{proof} Let $G'$ denote the Zariski closure of $\phi (CH^n(X)_{\deg 0})$. By \ref{pic}, $G'$ is connected and it is generated by the image of $\Pi_{X_{\rm reg}} \to G'$. Hence \ref{reg_sur} follows from the third part of Lemma \ref{generators}. \end{proof} \section{The cycle class map} Throughout the next three sections we will assume that the ground field $k$ is the field of complex numbers. ${\mathcal O}_X$ and $\Omega^m_{X/{\mathbb C}}$ will respectively denote the sheaves of holomorphic functions and (analytic K\"ahler) differential $m$-forms. As in the introduction consider the Deligne complex \[{\mathcal D}(n)_X=\left(0\to {\mathbb Z}_X(n)\to{\mathcal O}_X\to\Omega^{1}_{X/{\mathbb C}}\to\cdots\to \Omega^{n-1}_{X/{\mathbb C}}\to 0\right),\] and associated cohomology group $D^{n}(X)={\mathbb H}^{2n}(X,{\mathcal D}(n)_X).$ In this section we construct the cycle class homomorphism $CH^n(X) \to D^n(X)$, using Cartier curves $C$ in $X$. By the moving lemma \ref{moving} it will be sufficient to consider reduced Cartier curves $C$ in $X$. Note, however, that we do not have that $C$ is a local complete intersection in $X$, in general; this is only given to hold at points of $C\cap X_{{\rm sing}}$. This leads to a slight technical difficulty. We will need to define `Gysin' maps for Cartier curves $C$ in $X$. These are directly defined in case $C$ is a local complete intersection, and in general one has first to make a sequence of point blow ups centered in $X_{{\rm reg}}$ to reduce to this special case. Indeed, even to show that the cycle homomorphism $Z^n(X)\to D^n(X)$ respects rational equivalence, a similar procedure needs to be followed. Note that the exterior derivative yields a map of complexes ${\mathcal D}(n)_X\to \Omega^n_{X/{\mathbb C}}[-n],$ and there is an obvious map ${\mathcal D}(n)_X\to{\mathbb Z}(n)_X$. \begin{lemma}\label{local} For $x\in X_{{\rm reg}}$, there is a unique element $[x]\in H^{2n}_{\{x\}}(X,{\mathcal D}(n)_X)$ which maps to the topological cycle class of $x$ in $H^{2n}_{\{x\}}(X,{\mathbb Z}(n))$ as well as to the ``Hodge cycle class'' of $x$ in $H^n_{\{x\}}(X,\Omega^n_{X/{\mathbb C}})$. This gives rise to a well-defined cycle class homomorphism $Z^n(X)\to D^n(X)$, whose composition with $D^n(X)\to H^{2n}(X,{\mathbb Z}(n))$ is the topological cycle class homomorphism. \end{lemma} \begin{proof} The element $[x]$ exists because the topological and Hodge cycle classes both map to the de Rham cycle class of $x$ in $H^{2n}_{\{x\}}(X,{\mathbb C})={\mathbb H}^{2n}_{\{x\}}(X,\Omega^{\d}_{X/{\mathbb C}})$, by a standard local computation. See \cite{EV}, \S7, for example (though $X$ is singular, the terms in the above computation depend only on a neighbourhood of $x$ in $X$, and we have $x\in X_{{\rm reg}}$; hence \cite{EV}, \S7 is applicable). \end{proof} \begin{lemma}\label{curves1} In the above situation, if $\dim X=1$, then there is a natural quasi-isomorphism ${\mathcal D}(1)_X\cong{\mathcal O}_X^*[-1]$, yielding an identification ${\rm Pic}(X)\cong D^1(X)$ (and hence also ${\rm Pic}^0(X)\cong A^1(X)$). Under the identification, the class of a smooth point $[x]\in D^1(X)$ corresponds to the class of the invertible sheaf ${\mathcal O}_X(x)$. \end{lemma} \begin{proof} The natural quasi-isomorphism is equivalent to the exactness of the exponential sequence. The description of the class of a point $x$ as the class of the invertible sheaf ${\mathcal O}_X(x)$ is also a standard local computation. \end{proof} Now we argue as in \cite{BiS}, in order to show that the map $Z^n(X)\to D^n(X)$ factors through $CH^n(X)$. We follow the convention that the truncated de Rham complex of K\"ahler differentials $$ \Omega_{X/{\mathbb C}}^{< n}=(0\to{\mathcal O}_X\to\cdots\to\Omega^{n-1}_{X/{\mathbb C}}\to 0) $$ has ${\mathcal O}_X$ placed in degree 0; thus we have an exact sequence of complexes \[0\>>> \Omega_{X/{\mathbb C}}^{< n}[-1]\>>> {\mathcal D}(n)_X\>>> {\mathbb Z}(n)_X \>>> 0.\] \begin{lemma}\label{curves2} Let $X$ be a projective variety of dimension $n$ over ${\mathbb C}$, and $C\subset X$ be a reduced Cartier curve which is a local complete intersection in $X$. Then there is a commutative diagram \[ \begin{CD} Z^1(C) \>>> Z^n(X)\\ {\mathbb V} V V {\mathbb V} V V\\ D^1(C) \> {\rm Gysin}>> D^n(X)\\ {\mathbb V} V V {\mathbb V} V V \\ H^2(C,{\mathbb Z}(1)) \>{\rm Gysin}>> H^{2n}(X,{\mathbb Z}(n)) \end{CD} \] \end{lemma} \begin{proof} Consider the local (hyper) cohomology sheaves ${\mathcal H}^j_C({\mathcal D}(n)_X)$ of the complex ${\mathcal D}(n)_X$ with support in $C$. We claim that for any point $x\in C$, the stalks ${\mathcal H}^j_C({\mathcal D}(n)_X)_x$ vanish for $j\neq 2n-1$, unless $x$ is a singular point of $C$. Indeed, if $x\in C$ is a non-singular point (so that $x\in X_{{\rm reg}}$ as well), then there is a long exact sequence of stalks \begin{multline*} \cdots\>>>{\mathcal H}^{j-1}_C({\mathbb Z}(n)_X)_x\oplus {\mathcal H}^{j-1-n}_C(\Omega^n_{X/{\mathbb C}})_x\>>> {\mathcal H}^{j-1}_C({\mathbb C}_X)_x\>>> {\mathcal H}^j_C({\mathcal D}(n)_X)_x\\ \>>> {\mathcal H}^j_C({\mathbb Z}(n)_X)_x\oplus{\mathcal H}^{j-n}_C(\Omega^n_{X/{\mathbb C}})_x\>>> {\mathcal H}^j_C({\mathbb C}_X)_x\>>>\cdots \end{multline*} However ${\mathcal H}^i_C({\mathbb Z}(n)_X)_x={\mathcal H}^i_C({\mathbb C}_X)_x=0$ for $i \neq2n-2$, ${\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X)$ injects into ${\mathcal H}^{2n-2}_C({\mathbb C}_X)$, and ${\mathcal H}^i_C(\Omega^n_{X/{\mathbb C}})_x=0$ unless $i=n-1$, for a non-singular point $x\in C$ as above. This implies that ${\mathcal H}^j_C({\mathcal D}(n)_X)_x=0$ for $j\neq 2n-1$, for such $x$. Also ${\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)_x$ fits into an exact sequence \[0\>>>{\mathcal H}^{2n-2}_C({\mathbb C}_X/{\mathbb Z}(n)_X)_x\>>>{\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)_x\>>> {\mathcal H}^{n-1}_C(\Omega^n_{X/{\mathbb C}})_x\>>> 0,\] with ${\mathcal H}^{2n-2}_C({\mathbb C}_X/{\mathbb Z}(n)_X)_x\cong{\mathbb C}/{\mathbb Z}(1)={\mathbb C}^*$. Thus, ${\mathcal H}^j_C({\mathcal D}(n)_X)$ is supported at a finite set of points, if $j\neq 2n-1$. Hence in the local-to-global spectral sequence \[E_2^{p,q}=H^p(C,{\mathcal H}^q_C({\mathcal D}(n)_X))\Longrightarrow {\mathbb H}^{p+q}_C(X,{\mathcal D}(n)_X)\] we have $E_2^{p,q}=0$ for $p>0$, $q\neq 2n-1$. In particular, there is a well-defined injective map \[\alpha:H^1(C,{\mathcal H}^{2n-1}_C({\mathcal D}(n)_X))\>>> {\mathbb H}^{2n}_C(X,{\mathcal D}(n)_X).\] We will next construct a natural map of sheaves on $C$ \[{\mathcal O}_C^*\>>>{\mathcal H}^{2n-1}_C({\mathcal D}(n)_X).\] The desired Gysin map $D^1(C)\to D^n(X)$ is then defined to be the composition \begin{multline*} H^1(C,{\mathcal O}_C^*)\>>> H^1(C,{\mathcal H}^{2n-1}_C({\mathcal D}(n)_X))\>{\alpha}>> {\mathbb H}^{2n}_C(X,{\mathcal D}(n)_X)\>>>\\ {\mathbb H}^{2n}(X,{\mathcal D}(n)_X)=D^n(X) \end{multline*} To construct the map on sheaves ${\mathcal O}^*_C\to {\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)$, we argue locally, as follows. Let $U$ be an affine neighbourhood in $X$ of a point $x\in C$, on which the ideal of $C$ is generated by a regular sequence of functions $f_1,\ldots,f_{n-1}$, determining a morphism $f:U\to {\mathbb A}^{n-1}_{{\mathbb C}}$ such that $f^{-1}(0)=C\cap U$. Note that there are well-defined sections (of the skyscraper sheaves) \[\alpha\in \Gamma({\mathcal H}^{2n-2}_{\{0\}}({\mathbb Z}(n)_{{\mathbb A}^{n-1}_{{\mathbb C}}}))={\mathbb Z}(1),\;\;\;\;\beta\in \Gamma({\mathcal H}^{n-1}_{\{0\}}(\Omega^{n-1}_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}})),\] which have the same image $\gamma\in \Gamma({\mathcal H}^{2n-2}_{\{0\}}(\Omega_{{\mathbb A}^{n-1}_{\mathbb C}/{\mathbb C}}^{\bullet})),$ under the obvious maps, and such that $\beta$ is annihilated by the ideal of $0$ in $\Gamma({\mathcal O}_{{\mathbb A}^{n-1}_{{\mathbb C}}})$, for the natural module structure on $\Gamma({\mathcal H}^{n-1}_{\{0\}}(\Omega^{n-1}_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}}))$. In fact, these conditions uniquely determine such a pair of sections $(\alpha,\beta)$ up to sign, and there is a standard choice, with $\beta$ determined by $\dlog(z_1)\wedge\cdots\wedge\dlog(z_{n-1}),$ where $z_j$ are the coordinate functions, so that $\beta$ is the cup product of the local divisor classes \begin{equation}\label{ext} \dlog(z_j)\in \Gamma({\mathbb A}^{n-1}_{{\mathbb C}},\sext^1_{{\mathbb A}^{n-1}_{\mathbb C}}({\mathcal O}_{\{z_j=0\}}, \Omega^1_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}})) \subset \Gamma({\mathbb A}^{n-1}_{{\mathbb C}}, {\mathcal H}^1_{\{z_j=0\}}(\Omega^1_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}})). \end{equation} Hence $\gamma$ is also determined. Now consider \begin{gather*} f^*\alpha\in\Gamma(U,{\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X)), \;\;\;\; f^*\beta\in \Gamma(U,{\mathcal H}^{n-1}_C(\Omega^{n-1}_{X/{\mathbb C}})),\\ \mbox{and} \;\;\;\; f^*\gamma\in\Gamma(U,{\mathcal H}^{2n-1}_C(\Omega^{<n}_{X/{\mathbb C}})), \end{gather*} where $f^*\alpha$ and $f^*\beta$ both map to $f^*\gamma$, and $f^*\beta$ is annihilated by any section of the ideal sheaf of $C\cap U$ in $U$. Thus $f^*\alpha$ and $f^*\beta$ yield maps of sheaves \[{\mathbb Z}(1)_C\mid_U\>>> {\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X)\mid_U,\;\;\;\; {\mathcal O}_C\mid_U\>>> {\mathcal H}^{n-1}_C(\Omega^{n-1}_{X/{\mathbb C}})\mid_U,\] giving rise to a commutative diagram of sheaves $$ \begin{CD} {\mathbb Z}(1)_C\mid_U \>>> {\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X)\mid_U\\ {\mathbb V} V V {\mathbb V} V V\\ {\mathcal O}_C\mid_U \>>> {\mathcal H}^{2n-2}_C(\Omega^{<n}_{X/{\mathbb C}})\mid_U. \end{CD} $$ There is a long exact sequence of sheaves on $C$ $$ \cdots\>>>{\mathcal H}^j_C({\mathbb Z}(n)_X)\>>>{\mathcal H}^j_C(\Omega^{<n}_{X/{\mathbb C}})\>>> {\mathcal H}^{j+1}_C({\mathcal D}(n)_X)\>>>{\mathcal H}^{j+1}_C({\mathbb Z}(n)_X)\>>>\cdots $$ Hence from the exponential sequence \[0\>>> {\mathbb Z}(1)_C\>>> {\mathcal O}_C\>{\rm exp}>>{\mathcal O}_C^*\>>> 0,\] and the above commutative diagram, we deduce that there is a well-defined map of sheaves \[{\mathcal O}_C^*\mid_U\>>> {\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)\mid_U.\] We will now show that these locally defined maps patch together to give well-defined sheaf maps \begin{equation}\label{sheafmap} {\mathcal O}_C\>>> {\mathcal H}^{2n-2}_C(\Omega^{n-1}_{X/{\mathbb C}}) \mbox{ \ \ \ and \ \ \ }{\mathcal O}_C^*\>>> {\mathcal H}^{2n-1}_C({\mathcal D}(n)_X). \end{equation} To do this, it suffices to show that the classes $f^*\alpha$, $f^*\beta$ and $f^*\gamma$ defined above are in fact independent of the map $f$, {\it i.e.\/},\ of the choice of generators for the ideal of $C$ in $U$. This too can be seen ``universally''. Since the ideal sheaf of $C$ in $X$ is locally generated by a regular sequence, any two such sets of local generators for ${\mathcal I}_C$ on the affine open set $U$ differ by the operation of an element of $\GL_{n-1}({\mathcal O}_X(V))$, for some neighbourhood $V$ of $C\cap U$ in $U$. Hence it suffices to show that if $p:\GL_{n-1}({\mathbb C})\times{\mathbb C}^{n-1}\to{\mathbb C}^{n-1}$ is the projection, and $m:\GL_{n-1}({\mathbb C})\times{\mathbb C}^{n-1}\to{\mathbb C}^{n-1}$ the map given by the operation of $\GL_{n-1}({\mathbb C})$ on ${\mathbb C}^{n-1}$ by invertible linear transformations, then $p^*\alpha=m^*\alpha$, $p^*\beta=m^*\beta$, and hence also $p^*\gamma=m^*\gamma$. We leave the verification of this to the reader, as a simple application of the K\"unneth formula. Finally, note that for $U=X_{{\rm reg}}$, we have a commutative diagram with exact rows \[ \minCDarrowwidth=.8cm \begin{CD} 0\>>> {\mathbb Z}(1)_C\mid_U \>>> {\mathcal O}_C\mid_U\>>>{\mathcal O}_C^*\mid_U\>>> 0\\ \noarr {\mathbb V}\cong V V {\mathbb V} V V {\mathbb V} V V\\ 0\>>> {\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X)\mid_U\>>>{\mathcal H}^{2n-2}_C(\Omega^{<n}_{X/{\mathbb C}})\mid_U \>>> {\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)\mid_U\>>> 0 \end{CD}\] where the left vertical arrow is an isomorphism. For a smooth point $x\in C$, apply the functors ${\mathcal H}^j_{\{x\}}$ to the rows of the above diagram, and note that ${\mathcal H}^j_{\{x\}}({\mathbb Z}(1)_C)=0$ for $j\neq 2$, and ${\mathcal H}^j_{\{x\}}({\mathcal O}_C)=0$ for $j\neq 1$. We then obtain another diagram with exact rows \[ \minCDarrowwidth=.6cm \hspace*{-.4cm} \begin{CD} 0\>>> {\mathcal H}^1_{\{x\}}({\mathcal O}_C) \>>> {\mathcal H}^1_{\{x\}}({\mathcal O}_C^*) \>>> {\mathcal H}^2_{\{x\}}({\mathbb Z}(1)_C)\>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V \cong V\\ 0\>>> {\mathcal H}^1_{\{x\}}({\mathcal H}^{2n-2}_C(\Omega^{<n}_{X/{\mathbb C}})) \>>> {\mathcal H}^1_{\{x\}}({\mathcal H}^{2n-1}_C({\mathcal D}(n)_X)) \>>> {\mathcal H}^2_{\{x\}}({\mathcal H}^{2n-2}_C({\mathbb Z}(n)_X))\>>> 0 \end{CD}\] The bottom row may be identified (see \cite{Ha2}, III, Ex.~8.7, pg.~161) with the exact sequence \[0\>>> {\mathcal H}^{2n-1}_{\{x\}}(\Omega^{<n}_{X/{\mathbb C}})\>>> {\mathcal H}^{2n}_{\{x\}}({\mathcal D}(n)_X)\>>> {\mathcal H}^{2n}_{\{x\}}({\mathbb Z}(n)_X)\>>> 0.\] We claim that, under the above identification, the local cycle class of $x$ in ${\mathcal H}^2_{\{x\}}({\mathcal D}(1)_C)={\mathcal H}^1_{\{x\}}({\mathcal O}_C^*)$ maps to the corresponding local cycle class of $x$ in ${\mathcal H}^{2n}_{\{x\}}({\mathcal D}(n)_X)$. Choosing a suitable regular system of parameters on $X$ at $x$, we reduce to checking this in the special case when $x\in X$ is the origin $0\in {\mathbb C}^n$, and the curve $C$ is the $z_n$-axis, given by the vanishing of the first $n-1$ coordinates. We again leave this verification to the reader. This means that, in the commutative diagram \[ \begin{CD} H^1_{\{x\}}(C,{\mathcal O}_C^*) \>>> H^1(C,{\mathcal O}_C^*)=D^1(C)\\ {\mathbb V} V V {\mathbb V} V V\\ {\mathbb H}^{2n}_{\{x\}}(X,{\mathcal D}(n)_X) \>>> D^n(X) \end{CD}\] the cycle class of $x$ in $D^1(C)$ maps to that of $x$ in $D^n(X)$. Hence we have shown that there is a commutative diagram \[\begin{CD} Z^1(C) \>>> Z^n(X)\\ {\mathbb V} V V {\mathbb V} V V\\ D^1(C) \>{\rm Gysin}>> D^n(X) \end{CD} \] It remains to show that the Gysin map ${\rm Pic}(C)=D^1(C)\to D^n(X)$ is compatible with the topological Gysin map $H^2(C,{\mathbb Z}(1))\to H^{2n}(X,{\mathbb Z}(n))$. Since $Z^1(C)\to D^1(C)$ is surjective, the compatibility of the two Gysin maps is clear from the fact that each one maps the class of $x$ on $C$ to the corresponding class on $X$. \end{proof} \begin{rmk}\label{gys_rem} Assume that the local complete intersection curve $C$ lies in the Cohen-Macaulay locus $X_{\rm CM}$ of $X$. Then the first sheaf map in (\ref{sheafmap}) factors as \begin{equation}\label{gys_fac} {\mathcal O}_C\>>> \sext^{n-1}_X( {\mathcal O}_C,\Omega^{n-1}_{X/{\mathbb C}}) \>>> {\mathcal H}^{2n-2}_C(\Omega^{n-1}_{X/{\mathbb C}}). \end{equation} To see this, note that $$ \beta \in \Gamma({\mathbb A}^{n-1}_{{\mathbb C}},\sext^{n-1}_{{\mathbb A}^{n-1}_{\mathbb C}}({\mathcal O}_{\{0\}}, \Omega^{n-1}_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}})) \subset \Gamma({\mathbb A}^{n-1}_{{\mathbb C}}, {\mathcal H}^{n-1}_{\{0\}}(\Omega^{n-1}_{{\mathbb A}^{n-1}_{{\mathbb C}}/{\mathbb C}})) $$ as it is the product of the classes $\dlog(z_j)$ in (\ref{ext}). Further the map $U \to {\mathbb A}^{n-1}_{\mathbb C}$ is flat in a neigbourhood of $C\cap U$, as it is equidimensional. Thus $f^*\beta$ defines a class in $\Gamma(U,\sext^{n-1}_X({\mathcal O}_{C},\Omega^{n-1}_{X/{\mathbb C}}))$ mapping to $\Gamma(U,{\mathcal H}^{n-1}_{C}(\Omega^{n-1}_{X/{\mathbb C}}))$. As \begin{multline*} {\rm Ext}^{n-1}_{\GL_{n-1}({\mathbb C})\times {\mathbb C}^{n-1}}({\mathcal O}_{\GL_{n-1}({\mathbb C})\times \{0\}}, \Omega^{n-1}_{\GL_{n-1}({\mathbb C})\times {\mathbb C}^{n-1}/{\mathbb C}})\\ \subset H^{n-1}_{\GL_{n-1}({\mathbb C})\times \{0\}}(\GL_{n-1}({\mathbb C})\times {\mathbb C}^{n-1}, \Omega^{n-1}_{\GL_{n-1}({\mathbb C})\times {\mathbb C}^{n-1}/{\mathbb C}})), \end{multline*} the class $f^*\beta$ defines the factorization (\ref{gys_fac}). \end{rmk} \begin{lemma}\label{blowup} Let $X$ be projective of dimension $n$ over ${\mathbb C}$, $f:Y\to X$ the blow up of a smooth point $x\in X$. Then the natural maps $f_*:CH^n(Y)\to CH^n(X)$ and $f^*:D^n(X)\to D^n(Y)$ are isomorphisms, and there is a commutative diagram \[\begin{CD} Z^n(Y) \>>> D^n(Y)\\ {\mathbb V} f_* V V {\mathbb A} \cong A f^* A\\ Z^n(X) \>>> D^n(X) \end{CD} \] \end{lemma} \begin{proof} The isomorphism on Chow groups is easy to prove, using the fact that the exceptional divisor $E$ is a projective space (the details are in \cite{BiS}). That $f^*:D^n(X)\to D^n(Y)$ is an isomorphism is also easy to see, for the same reason, using also the exact sequence \[0\>>> f^*\Omega^1_{X/{\mathbb C}}\>>> \Omega^1_{Y/{\mathbb C}}\>>> \Omega^1_{E/{\mathbb C}}\>>> 0.\] So we need to prove that if $y\in Y$ is any smooth point, then its class in $D^n(Y)$ is the inverse image of that of $f(y)$ in $D^n(X)$. This is clear if $f(y)\neq x$. If $f(y)=x$, we may argue as follows. There is a commutative diagram with exact rows \[ \minCDarrowwidth=.6cm \hspace*{-.5cm} \begin{CD} 0 \>>> {\mathbb H}^{2n}_{\{y\}}(Y,{\mathcal D}(n)_Y)\>>> H^{2n}_{\{y\}}(Y,{\mathbb Z}(n))\oplus H^n_{\{y\}}(Y,\Omega^n_{Y/{\mathbb C}}) \>>> H^{2n}_{\{y\}}(Y,{\mathbb C}(n))\>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V V\\ 0 \>>> {\mathbb H}^{2n}_{E}(Y,{\mathcal D}(n)_Y)\>>> H^{2n}_{E}(Y,{\mathbb Z}(n))\oplus H^n_{E}(Y,\Omega^n_{Y/{\mathbb C}}) \>>> H^{2n}_{E}(Y,{\mathbb C}(n))\>>> 0\\ \noarr {\mathbb A} f^* A A {\mathbb A} f^* A A {\mathbb A} f^* A A\\ 0 \>>> {\mathbb H}^{2n}_{\{x\}}(X,{\mathcal D}(n)_X)\>>> H^{2n}_{\{x\}}(X,{\mathbb Z}(n))\oplus H^n_{\{x\}}(X,\Omega^n_{X/{\mathbb C}}) \>>> H^{2n}_{\{y\}}(X,{\mathbb C}(n))\>>> 0 \end{CD}\] Here the downward vertical arrows are the natural maps (``increase support''). It is standard that the topological local cycle classes of $x$ and $y$ have the same images in $H^{2n}_E(Y,{\mathbb Z}(n))$. Similarly, the images in $H^n_E(Y,\Omega^n_{Y/{\mathbb C}})$ of the local cycle classes of $x$ and $y$ in Hodge cohomology are also known to be equal; for example, this follows from the existence of a Gysin map $f_*:H^n_{\{y\}}(Y,\Omega^n_{Y/{\mathbb C}})\to H^n_{\{x\}}(X,\Omega^n_{X/{\mathbb C}})$, which maps the local class of $y$ to that of $x$, and which factors through $$H^n_E(Y,\Omega^n_{Y/{\mathbb C}})\>{(f^*)^{-1}}>>H^n_{\{x\}} (X,\Omega^n_{X/{\mathbb C}}).$$ Thus $f^*[x]=[y]\in {\mathbb H}^{2n}_E(Y,{\mathcal D}(n)_Y)$, and hence a similar equality is valid in $D^n(Y)$ as claimed. \end{proof} \begin{lemma}\label{equivalence} The map $Z^n(X)\to D^n(X)$ factors through $CH^n(X)$, and hence determines a homomorphism $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$. \end{lemma} \begin{proof} This is similar to the corresponding proof in \cite{BiS}. Let $C\subset X$ be a reduced Cartier curve, and $f\in R(C,X)$. Let $\pi:Y\to X$ be a composition of blow ups at smooth points so that the strict transform $\tilde{C}$ of $C$ in $Y$ satisfies $\tilde{C}_{{\rm sing}}=\tilde{C}\cap Y_{{\rm sing}}\cong C\cap X_{{\rm sing}}$. Then \[R(C,X)=R(\tilde{C},Y)=R(\tilde{C},\tilde{C}),\] and $\pi_*(f)_{\tilde{C}}=(f)_C\in Z^n(X).$ Now from lemma~\ref{curves1}, $(f)_{\tilde{C}}\mapsto 0\in D^n(Y),$ and so from lemma~\ref{blowup}, $(f)_C=\pi_*(f)_{\tilde{C}}=0\in D^n(X).$ \end{proof} \begin{cor}\label{corblowup} If $f:Y\to X$ is a composition of blow ups at smooth points, then we have a diagram \[\begin{CD} CH^n(Y) \>>> D^n(Y)\\ {\mathbb V} f_* V \cong V {\mathbb A} \cong A f^* A\\ CH^n(X) \>>> D^n(X) \end{CD} \] \end{cor} \begin{cor}\label{curves3} For any reduced Cartier curve $C\subset X$, there are commutative diagrams \[\begin{CD} Z^1(C) \>>> Z^n(X)\\ {\mathbb V} V V {\mathbb V} V V\\ {\rm Pic}(C) \>{\rm Gysin}>>CH^n(X)\\ {\mathbb V} \cong V V {\mathbb V} V V\\ D^1(C) \>{\rm Gysin}>> D^n(X)\\ {\mathbb V} V V {\mathbb V} V V \\ H^2(C,{\mathbb Z}(1)) \>{\rm Gysin}>> H^{2n}(X,{\mathbb Z}(n)) \end{CD} \mbox{ \ \ \ \ and \ \ \ \ \ } \begin{CD} {\rm Pic}^0(C) \>>> CH^n(X)_{\deg 0}\\ {\mathbb V} \cong V V {\mathbb V} V V \\ A^1(C) \>>> A^n(X) \end{CD}\] \end{cor} \begin{proof} As in the proof of lemma~\ref{equivalence}, by a compositon of blow-ups at smooth points, we reduce to the case when $C$ is a local complete intersection in $X$. Then lemma~\ref{curves2} implies the corollary. \end{proof} Considering embedded resolution of singularities one obtains from \ref{curves3} and \ref{corblowup} a second construction of the Gysin map in \ref{gysin} over ${\mathbb C}$. At the same time, it gives the compatibility of this map with the Gysin map for the Deligne cohomology, constructed in \ref{curves2}. \section{Some general properties of $A^n(X)$ over ${\mathbb C}$} It is shown in \cite{BiS} that if $X$ is projective over ${\mathbb C}$ of dimension $n$, then there is a natural surjection (which is referred to in \cite{BiS} as the {\em Abel-Jacobi map}) \[AJ^n_X:CH^n(X)_{\deg 0}\>>> J^n(X):=\frac{H^{2n-1}(X,{\mathbb C}(n))}{F^0H^{2n-1}(X,{\mathbb C}(n))+{\rm image}\,H^{2n-1}(X,{\mathbb Z}(n))},\] where by results of Deligne, $J^n(X)$ is a semi-abelian variety (since the non-zero Hodge numbers of $H^{2n-1}(X,{\mathbb Z}(n))$ lie in the set $\{(-1,0),(0,-1),(-1,-1)\}$). \begin{lemma}\label{algebraicity} Let $X$ be projective of dimension $n$ over ${\mathbb C}$. Then there is a natural surjection $\psi:A^n(X)\to J^n(X),$ whose kernel is a ${\mathbb C}$-vector space. $A^n(X)$ has a unique structure as an algebraic group such that $\psi$ is a morphism of algebraic groups, with additive kernel ({\it i.e.\/},\ with kernel isomorphic to a direct sum of copies of ${\mathbb G}_a$). \end{lemma} \begin{proof} By a result of Bloom and Herrera \cite{BH}, the natural map \[H^{2n-1}(X,{\mathbb C}(n))\>>> H^{2n-1}_{DR}(X/{\mathbb C})={\mathbb H}^{2n-1}(X,\Omega^{\d}_{X/{\mathbb C}})\] is split injective. As explained in \cite{D}~(9.3.2), if $X_{\d}\to X$ is a suitable hypercovering by a smooth proper simplicial scheme, the splitting may be given by the composition $$H^{2n-1}_{DR}(X/{\mathbb C})\to H^{2n-1}_{DR}(X_{\d}/{\mathbb C})\cong H^{2n-1}(X_{\d},{\mathbb C}(n))\cong H^{2n-1}(X,{\mathbb C}(n)). $$ {From} this description, the splitting is a map of filtered vector spaces, where $H^{2n-1}(X,{\mathbb C}(n))$ has the Hodge filtration for the mixed Hodge structure while $H^{2n-1}_{DR}(X/{\mathbb C})$ has the truncation filtration ({\it i.e.\/},\ the filtration b\^ete). Hence we obtain a commutative diagram $$ \begin{TriCDV} {H^{2n-1}(X,{\mathbb C}(n))}{\>>>}{H^{2n-1}(X,{\mathbb C})/F^0H^{2n-1}(X,{\mathbb C}(n))} {\SE E E}{\NE E {\vartheta}E} {{\mathbb H}^{2n-1}(X,\Omega^{< n}_{X/{\mathbb C}})} \end{TriCDV} $$ The map $\vartheta$ induces the map $\psi$ taking quotients modulo $H^{2n-1}(X,{\mathbb Z}(n))$. Note that by weight considerations, the natural map \[ H^{2n-1}(X,{\mathbb Z}(n))\>>> H^{2n-1}(X,{\mathbb C}(n))/F^0H^{2n-1}(X,{\mathbb C}(n)) \] has a torsion kernel. Hence the kernels of $\psi$ and $\vartheta$ are the same (and the latter is a ${\mathbb C}$-vector space). This represents $A^n(X)$ as an analytic group extension of the semi-abelian variety $J^n(X)$ by an additive group ${\mathbb G}_a^r$, for some $r$, and hence as an analytic group extension of an abelian variety by a group ${\mathbb G}_a^r\times {\mathbb G}_m^s$. As noted in \cite{D}, (10.1.3.3), for any abelian variety $A$ over ${\mathbb C}$, the isomorphism classes of analytic and algebraic groups extensions of $A$ by either ${\mathbb G}_a$ or by ${\mathbb G}_m$ coincide (as a consequence of GAGA); hence a similar property is valid for extensions by ${\mathbb G}_a^r\times {\mathbb G}_m^s$. This implies that $A^n(X)$ has a unique algebraic structure such that $\psi$ is a homomorphism of algebraic groups over ${\mathbb C}$, as claimed. \end{proof} For $X$ a curve, as in (i) of the next corollary, note that ${\rm Pic}^0(X)$ has the natural algebraic structure obtained by representing a suitable Picard functor. In particular, given an algebraic family of divisors (of degree 0) on $X$ parametrized by a variety (or scheme) $T$, the induced map $T\to{\rm Pic}^0(X)$ is automatically a morphism. On the other hand, $A^1(X)$ has the algebraic structure given by lemma~\ref{algebraicity}. Hence, a priori, the induced map $T\to A^1(X)$ obtained from such a family is only analytic, since it is essentially given by integration. From (i) of the corollary, it will follow that it is in fact algebraic. The content of (ii) of the corollary is similar. \begin{cor}\label{algcurve}\ \ \begin{points} \item If $X$ is a curve, then the natural isomorphism ${\rm Pic}^0(X)\cong A^1(X)$ is an isomorphism of algebraic groups. \item In general, if $C\subset X$ is a reduced Cartier curve, then the induced homomorphism $A^1(C)\to A^n(X)$ of corollary~\ref{curves3} is algebraic. \end{points} \end{cor} \begin{proof} (i) The identification is certainly analytic, and in both cases, when one represents the algebraic group as an extension of an abelian variety by a commutative affine group, the abelian variety in question is just ${\rm Pic}^0(\tilde{X})=J(\tilde{X})=D^1(\tilde{X})$, the Jacobian of the normalized curve $\tilde{X}$ (by which we mean the product of the Jacobians of the connected components of $\tilde{X}$). Now one argues that the identification must be algebraic as well, since one has the one-one correspondence between analytic and algebraic extensions of an abelian variety by ${\mathbb G}_a^r\times{\mathbb G}_m^s$. (ii) Let $\tilde{X} \to X^{(n)}$ be a desingularization of $X^{(n)}$ such that the proper transform $\tilde{C}$ of $C$ is the normalization of $C$. First note that one has a factorization $$ \begin{CD} A^1(C) \>>> A^n(X)\\ {\mathbb V} V V {\mathbb V} V V\\ J^1(C) \>>> J^n(X)\\ {\mathbb V} V V {\mathbb V} V V\\ A^1(\tilde{C}) \>>> A^n(\tilde{X}) \end{CD} $$ where all maps are analytic group homomorphisms, and the vertical ones are algebraic (lemma \ref{algebraicity}). Indeed the map $C \to X$ induces a morphism of mixed Hodge structures $H^1(C) \to H^{2n-1}(X)$, and therefore an analytic group homomorphim $J^1(C) \to J^n(X)$, which has to be algebraic as it is compatible with its abelian part $J^1(\tilde{C}) \to A^n(\tilde{X})$ and all analytic group homomorphisms ${\mathbb G}_m^s \to {\mathbb G}_m^{s'}$ are algebraic. Similarly, all group homomorphisms ${\mathbb G}_a^r \to {\mathbb G}_a^{r'}$ are algebraic, and therefore $A^1(C) \to A^n(X)$ is algebraic as well. \end{proof} \begin{defn}\label{deflie} For any commutative algebraic group $A$ over ${\mathbb C}$, let $\Omega(A)$ denote the dual vector space to the Lie algebra $\Lie(A)$. We may then identify $\Omega(A)$ with the vector space of (closed) translation invariant regular 1-forms on $A$. \end{defn} Our next goal is to give a description of $\Omega(A^n(X))$, generalizing the fact that for a non-singular projective variety $X$, $\Omega(A^n(X))$ is the space of holomorphic 1-forms on $X$ (since in that case, $A^n(X)$ is the Albanese variety of $X$). \begin{lemma}\label{purity} Let $X$ be projective of dimension $n$ over ${\mathbb C}$, and let $\omega_X$ denote the dualizing module of $X$ (in the sense of \cite{Ha}, Ch.~III, \S7). Let $X^{(n)}$ be the union of the $n$-dimensional components of $X$, and let $\omega_{X^{(n)}}$ denote its dualizing module. \begin{points} \item $\omega_X$ is annihilated by the ideal sheaf of $X^{(n)}$ in $X$. With its natural induced structure as an ${\mathcal O}_{X^{(n)}}$-module, $\omega_X\cong\omega_{X^{(n)}}$, and is a torsion-free ${\mathcal O}_{X^{(n)}}$-module. Hence for any coherent ${\mathcal O}_X$-module ${\mathcal F}$, the sheaf $\shom_{{\mathcal O}_X}({\mathcal F}, \omega_X)$ is also naturally an ${\mathcal O}_{X^{(n)}}$-module, which is ${\mathcal O}_{X^{(n)}}$-torsion free, and for any dense open set $U\subset X^{(n)}$, the restriction map \[{\rm Hom}_X({\mathcal F},\omega_X)\>>> {\rm Hom}_U({\mathcal F}\mid_U,\omega_X\mid_U)\] is injective. In particular, taking $U=X_{{\rm reg}}$, so that $\omega_X\mid_U=\Omega^n_{U/{\mathbb C}}$, and taking ${\mathcal F}=\Omega^{n-1}_{X/{\mathbb C}}$, we have that ${\rm Hom}_X(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$ may be identified with a ${\mathbb C}$-subspace of the vector space of holomorphic 1-forms on $X_{{\rm reg}}$ which are meromorphic on $X^{(n)}$. \item $\Omega(A^{n}(X))$ is naturally identified with the subspace of ${\rm Hom}_X(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$ consisting of {\em closed} 1-forms. \item When $n=1$, \[\Omega(A^{1}(X))=\Omega({\rm Pic}^{0}(X)))=H^{0}(X,\omega_X).\] \item Let $j:X_{\rm CM}\to X$ be the inclusion of the open subset of Cohen-Macaulay points. The natural map \[\Omega(A^{n}(X))\>>> \left(\mbox{closed 1-forms in ${\rm Hom}_X(\Omega^{n-1}_{X/{\mathbb C}},j^{m}_*j^{*}\omega_X)$}\right)\] is an isomorphism, where $j^{m}_*$ denotes the meromorphic direct image. \end{points} \end{lemma} \begin{proof} (i) We note first that $\omega_X\cong\omega_{X^{(n)}}$, and the latter is a torsion-free ${\mathcal O}_{X^{(n)}}$-module. Indeed, if we fix a projective embedding $X\hookrightarrow\P^N_{{\mathbb C}}$, then \[\omega_X=\sext_{\P^N}^{N-n}({\mathcal O}_X,\omega_{\P^N_{{\mathbb C}}}),\] and there is an analogous formula for $\omega_{X^{(n)}}$. As in \cite{Ha}, we see by Serre duality on $\P^N_{{\mathbb C}}$ that $\sext^i({\mathcal F},\omega_{\P^N_{{\mathbb C}}})=0$ for all $i\leq N-n$ for any coherent sheaf ${\mathcal F}$ supported in dimension $<n$. This gives the desired isomorphism, and implies that any local section of ${\mathcal O}_{X^{(n)}}$, which is a non zero-divisor, is also a non zero-divisor on $\sext^{N-n}_{\P^N_{{\mathbb C}}}({\mathcal O}_X,\omega_{\P^N_{{\mathbb C}}})$. This means exactly that $\omega_{X^{(n)}}$ is torsion-free. We conclude that for any coherent ${\mathcal O}_X$-module ${\mathcal F}$, the sheaf $\shom_X({\mathcal F},\omega_X)$ is a torsion-free ${\mathcal O}_{X^{(n)}}$-module as well. Applying this to ${\mathcal F}=\Omega^{n-1}_{X/{\mathbb C}}$ gives (i). (iii) is a special case of (ii). To prove (ii), first note that from the definition of $A^{n}(X)$, we have \begin{equation}\label{lie_descr} \Lie(A^{n}(X))=\coker(d:H^{n}(X,\Omega^{n-2}_{X/{\mathbb C}})\to H^{n}(X,\Omega^{n-1}_{X/{\mathbb C}})). \end{equation} {From} Serre duality for $H^{n}$ and ${\rm Hom}$, as in the definition of the dualizing sheaf in \cite{Ha}, we have an identification of the dual vector space \[H^{n}(X,\Omega^i_{X/{\mathbb C}})^{*}= {\rm Hom}_{{\mathcal O}_X}(\Omega^i_{X/{\mathbb C}},\omega_X),\] for any $i$. Thus $\Omega(A^n(X))$ is identified with the subspace of ${\rm Hom}_{{\mathcal O}_X}(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$ of elements $\varphi$ such that the composition \[\ell: H^n(X,\Omega^{n-2}_{X/{\mathbb C}})\>{d}>> H^n(X,\Omega^{n-1}_{X/{\mathbb C}}) \>{\varphi}>> H^n(X,\omega_X)\cong {\mathbb C}\] is 0. It remains to show that, identifying elements $\varphi\in {\rm Hom}_{{\mathcal O}_X}(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$ with certain holomorphic 1-forms on $X_{{\rm reg}}$, $\Omega(A^n(X))$ is just the subspace of closed 1-forms. To see this, since we may consider $\varphi$ as a meromorphic 1-form on $X$ which is holomorphic on $X_{{\rm reg}}$, we can find a coherent sheaf of ideals ${\mathcal J}$, defining the Zariski closed subset $X_{{\rm sing}}\subset X$ ({\it i.e.\/},\ the subscheme determined by ${\mathcal J}$ has $X_{{\rm sing}}$ as its underlying reduced scheme), such that \begin{points} \item $\eta\mapsto \eta\wedge\varphi$ defines an element of ${\rm Hom}_{{\mathcal O}_X}({\mathcal J}\Omega^{n-2}_{X/{\mathbb C}},\Omega^{n-1}_{X/{\mathbb C}})$ \item $\eta\mapsto \eta\wedge d\varphi$ defines an element of ${\rm Hom}_{{\mathcal O}_X}({\mathcal J}\Omega^{n-2}_{X/{\mathbb C}},\omega_X)$, where we view $\omega_X$ as a certain coherent extension of $\Omega^n_{X_{{\rm reg}}/{\mathbb C}}$ to $X$. \end{points} (Here ${\mathcal J}{\mathcal F}$ denotes ${\rm image}\,({\mathcal J}\otimes{\mathcal F}\to{\mathcal F})$, for any ideal sheaf ${\mathcal J}$ and coherent sheaf ${\mathcal F}$). Since ${\mathcal J}$ defines $X_{{\rm sing}}$ within $X$, the natural map \[H^n(X,{\mathcal J}\Omega^{n-2}_{X/{\mathbb C}})\>>> H^n(X,\Omega^{n-2}_{X/{\mathbb C}})\] is surjective, and for any $\varphi\in{\rm Hom}_{{\mathcal O}_X}(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$, the composition \[\ell_1:H^n(X,{\mathcal J}\Omega^{n-2}_{X/{\mathbb C}})\>{d}>> H^n(X,\Omega^{n-1}_{X/{\mathbb C}})\>{\varphi}>> H^n(X,\omega_X)\cong {\mathbb C}\] factors through $\ell$. Thus \[\varphi\in \Omega(A^n(X))\;\;\;\iff\;\;\; \ell_1=0.\] We have 2 other related linear functionals \[\ell_2:H^n(X,{\mathcal J}\Omega^{n-2}_{X/{\mathbb C}})\>>>{\mathbb C},\;\;\;\; \ell_3:H^n(X,{\mathcal J}\Omega^{n-2}_{X/{\mathbb C}})\>>>{\mathbb C},\] defined by \begin{gather*} \ell_2:H^n(X,{\mathcal J}\Omega^{n-2}_{X{\mathbb C}})\>{\wedge d\varphi}>> H^n(X,\omega_X)\cong {\mathbb C},\\ \ell_3: H^n(X,{\mathcal J}\Omega^{n-2}_{X/{\mathbb C}})\>{\wedge \varphi}>> H^n(X,\Omega^{n-1}_{X/{\mathbb C}})\>{d}>> H^n(X,\omega_X)\cong{\mathbb C}, \end{gather*} where in the definition of $\ell_3$, we have let $d$ also denote the composite of the exterior derivative $\Omega^{n-1}_{X/{\mathbb C}}\to\Omega^n_{X/{\mathbb C}}$ with the natural map $\Omega^n_{X/{\mathbb C}}\to\omega_X$. The formula \[d(\eta\wedge\varphi)=d\eta\wedge\varphi+(-1)^{n-2}\eta\wedge d\varphi,\] for any $n-2$ form $\eta$, implies that $\ell_3=\ell_1+(-1)^{n-2}\ell_2.$ Now by Serre duality and the ${\mathcal O}_{X^{(n)}}$-torsion freeness of $\shom_{{\mathcal O}_X}({\mathcal J}\Omega^{n-2}_{X/{\mathbb C}}, \omega_X)$ (see (i)), $\ell_2$ vanishes precisely when $d\varphi=0$ as a 2-form on $X_{reg}$. On the other hand, we claim that for any $\varphi\in {\rm Hom}_{{\mathcal O}_X}(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)$, the map $\ell_3$ constructed as above is always 0. This will imply that $\ell_1=0$ \iff $\varphi$ is a closed meromorphic 1-form. To prove that $\ell_3$ vanishes, it suffices to prove that the map \[H^n(X,\Omega^{n-1}_{X/{\mathbb C}})\>{d}>> H^n(X,\omega_X)\] vanishes. One way to understand this is to note that if $\pi:Y\to X$ is a resolution of singularities, then there is a commutative diagram \[\begin{CD} H^n(X,\Omega^{n-1}_{X/{\mathbb C}}) \>{d}>> H^n(X,\omega_X)\\ {\mathbb V} \pi^* VV {\mathbb A} A \pi_* A\\ H^n(Y,\Omega^{n-1}_{Y/{\mathbb C}}) \>{d}>> H^n(Y,\omega_Y) \end{CD} \] which reduces us to proving that $H^n(Y,\Omega^{n-1}_{Y/{\mathbb C}})\>{d}>> H^n(Y,\omega_Y)$ vanishes. This follows from Hodge theory, or alternately may be proved as in \cite{Ha2}, III, lemma~8.4. Proof of (iv):\quad We begin by recalling that since $X$ is reduced, it is Cohen-Macaulay in codimension 1, so that $Z=X-j(X_{\rm CM})$ has codimension $\geq 2$ in $X$. Let ${\mathcal I}$ denote the ideal sheaf of $Z$ in $X$. Let ${\mathcal D}_m$ be the complex of sheaves \[{\mathcal D}_m=(0\to j_!{\mathbb Z}_{X_{\rm CM}}(n)\to {\mathcal I}^{m+n-1}\by{d}{\mathcal I}^{m+n-2}\Omega^{1}_{X/{\mathbb C}}\by{d} \cdots \by{d} {\mathcal I}^{m}\Omega^{n-1}_{X/{\mathbb C}}).\] Then ${\mathcal D}_m$ is a subcomplex of ${\mathcal D}(n)_X$, whose cokernel complex consists of sheaves supported on $Z$; the 0-th term of the cokernel is ${\mathbb Z}(n)_Z$, while the other terms are coherent sheaves supported on $Z$. Since $\dim Z\leq n-2$, we see that ${\mathbb H}^{i}$ of this cokernel complex vanishes for $i\geq 2n-1$. Hence ${\mathbb H}^{2n}(X,{\mathcal D}_m)\to {\mathbb H}^{2n}(X,{\mathcal D}(n)_X)$ is an isomorphism, for all $m$. Now as in the proof of (i), one uses duality, to conclude that for all $m$, there are isomorphisms \begin{align*} {\rm Hom}(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)&\>>> {\rm Hom}({\mathcal I}^m\Omega^{n-1}_{X/{\mathbb C}},\omega_X),\\ {\rm Hom}(\Omega^{n-2}_{X/{\mathbb C}},\omega_X)&\>>> {\rm Hom}({\mathcal I}^m\Omega^{n-2}_{X/{\mathbb C}},\omega_X), \end{align*} and taking the direct limit over all $m$, we obtain (iv). \end{proof} Our next goal is the proof of proposition~\ref{basic}, which gives us another useful way to recognize elements of the vector space $\Omega(A^n(X))$. We make use of two lemmas. \begin{lemma}\label{basic1} Let $X\subset\P^N_{{\mathbb C}}$ be a reduced projective variety of dimension $n$. Then we can find a finite number of linear projections $\pi_i:X\to \P^n_{{\mathbb C}}$, each of which is a finite morphism, such that the induced sheaf map \[\bigoplus_i\pi_i^*\Omega^{n-1}_{\P^n/{\mathbb C}}\>>> \Omega^{n-1}_{X/{\mathbb C}}\] is surjective. \end{lemma} \begin{proof} For any linear projection $\pi:X\to\P^n_{{\mathbb C}}$, there is a factorization \[\pi^*\Omega^{n-1}_{\P^n_{{\mathbb C}}/{\mathbb C}}\>>> \Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathcal O}_X\>{\psi}>>\Omega^{n-1}_{X/{\mathbb C}},\] where the natural map $\psi$ is surjective. So it suffices to prove the stronger assertion that there are projections $\pi_i$ as above such that the induced sheaf map \[\bigoplus_i\pi_i^*\Omega^{n-1}_{\P^n/{\mathbb C}}\>>> \Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathcal O}_X\] is surjective. We claim that for any $x\in X$, we can find a finite set of such projections $\pi_i:X\to\P^n_{{\mathbb C}}$ such that the map of ${\mathbb C}$-vector spaces \[\bigoplus_i\pi_i^*\Omega^{n-1}_{\P^n/{\mathbb C}}\otimes{\mathbb C}(\pi_i(x))\>>> \Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathbb C}(x)\] is surjective. Indeed, the Grassmannian ${\mathbb G}_{{\mathbb C}}(n+1,N+1)$ (of $n+1$ dimensional subspaces of ${\mathbb C}^{N+1}$) parametrizes linear projections from $\P^N_{{\mathbb C}}$ to $\P^n_{{\mathbb C}}$, and it contains a dense Zariski open subset corresponding to projections which are finite morphisms on $X$. Hence the $n$-dimensional vector subspaces \[\pi^*\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}\otimes{\mathbb C}(\pi(x))\subset \Omega^1_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathbb C}(x)\] also range over a Zariski open subset of the Grassmannian of $n$-dimensional subspaces of the cotangent space of $\P^N_{{\mathbb C}}$ at $x$. In particular, we can find a finite number of them whose $(n-1)$-th exterior powers span the $(n-1)$-th exterior power of this cotangent space, namely $\Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathbb C}(x)$. Now suppose $\pi_1,\ldots,\pi_r$ are chosen finite linear projections $X\to\P^n_{{\mathbb C}}$, and that \[\bigoplus_{i=1}^r\pi_i^*\Omega^{n-1}_{\P^n/{\mathbb C}}\>>> \Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathcal O}_X\] is not surjective. We can then find a point $x\in X$ at which the cokernel is non-zero. By the above claim, we can augment the set of projections to $\pi_1,\ldots,\pi_r,\pi_{r+1},\ldots,\pi_s$ so that the cokernel of the new map \[\bigoplus_{i=1}^{r+s}\pi_i^*\Omega^{n-1}_{\P^n/{\mathbb C}}\>>> \Omega^{n-1}_{\P^N_{{\mathbb C}}/{\mathbb C}}\otimes{\mathcal O}_X\] does not have $x$ in its support. Thus the support of the cokernel has strictly decreased. Now the lemma follows by Noetherian induction. \end{proof} \begin{lemma}\label{basic2} Let ${\mathcal F}$ be a reflexive coherent sheaf on $\P^n_{{\mathbb C}}$, and $\omega$ a meromorphic section of ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}$, which is regular on some given (nonempty) Zariski open subset $W\subset\P^n_{{\mathbb C}}$. Suppose there is a non-empty open set $V$ in ${\mathbb G}_{{\mathbb C}}(2,n+1)$, the Grassmannian of lines in $\P^n_{{\mathbb C}}$, such that \begin{points} \item each line $L\in V$ meets $W$, and is disjoint from the non locally-free locus of ${\mathcal F}$ \item for each $L\in V$, the image of $\omega$ in $({\mathcal F}\otimes\Omega^1_{L/{\mathbb C}})\mid_{L\cap W}$ extends to a regular section of ${\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}$ on $L$. \end{points} Then $\omega$ extends (uniquely) to a regular section on $\P^n_{{\mathbb C}}$ of ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}$. \end{lemma} \begin{proof} Since ${\mathcal F}$ is reflexive, it is locally free outside a Zariski closed set $A$ (of codimension $\geq 3$), and any section of ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}$ defined in the complement of $A$ extends uniquely to a section on all of $\P^n_{{\mathbb C}}$. Since $\omega$ is a meromorphic section, it determines a (unique) regular section of some twist ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}(D)$, for an effective divisor $D$; there is a unique such twist $D$ which is minimal with respect to the partial order on effective divisors (determined by inclusion of subschemes). Our goal is to show that $D=0$. If $F$ is an irreducible component of ${\rm supp}\, D$ which appears in $D$ with multiplicity $r>0$, then we can find a point $x\in F$ such that \begin{points} \item $x$ is a non-singular point of $F$, and does not lie on any other component of $D$; further, ${\mathcal F}$ is locally free near $x$ \item $V$ contains a line through $x$ \item there is a regular parameter $t\in{\mathcal O}_{x,\P^n_{{\mathbb C}}}$ ({\it i.e.\/},\ $t$ is part of a regular system of parameters) such that $t$ defines the ideal of $F$ at $x$, and such that $t^r\omega$ determines a regular, non-vanishing section of ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}$ in a neighbourhood of $x$. \end{points} It then follows that for a non-empty Zariski open set of lines $L$ through $x$, we have $L\in V$, and $t^r \omega$ maps to a regular, non-vanishing section of ${\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}$ near $x$, while $\omega$ itself maps to a regular section of ${\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}$. However, $t$ vanishes at $x$. This is a contradiction. \end{proof} If $C\subset X$ is a reduced, local complete intersection Cartier curve, then in fact $C\subset X_{\rm CM}\cap X^{(n)}$ (recall that $X_{\rm CM}$ denotes the (dense) Zariski open subset of Cohen-Macaulay points of $X$). The sheaf map ${\mathcal O}_C\to {\mathcal H}^{2n-1}_C(\Omega^{n-1}_{X/{\mathbb C}})$ in (\ref{sheafmap}) induces a composite map $$\alpha_C:H^1(C,{\mathcal O}_C)\>>> H^1(C,{\mathcal H}^{2n-1}_C(\Omega^{n-1}_{X/{\mathbb C}}))\>>> H^n_C(X,\Omega^{n-1}_{X/{\mathbb C}})\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} {\mathbb H}^{2n-1}(X,\Omega^{<n}_{X/{\mathbb C}}).$$ This is just the map $\Lie({\rm Pic}^0(C))\to \Lie(A^n(X))$ on Lie algebras induced by the composition of the group homomorphisms ${\rm Pic}^0(C)\to A^1(C)$ and the Gysin map $A^1(C)\to A^n(X)$. \begin{propose}\label{basic} \ \ \begin{enumerate} \item[(a)] Let $C\subset X$ be a reduced, local complete intersection Cartier curve, and let $U\subset X_{reg}$ be a dense open subset such that $U\cap C$ is dense in $C_{reg}$. Then the dual $\alpha_C^{\vee}$ of $\alpha_C:H^1(C,{\mathcal O}_C)\to {\mathbb H}^{2n-1}(X,\Omega^{<n}_{X/{\mathbb C}})$ ({\it i.e.\/},\ of $\Lie({\rm Pic}^0(C))\to \Lie(A^n(X))$) fits into a commutative diagram $$ \begin{CD} \Omega(A^n(X)) \> \subset >> H^0(X_{\rm reg}, \Omega^1_{X_{\rm reg}/{\mathbb C}})\\ {\mathbb V} V {\alpha_C^{\vee}}V {\mathbb V} V {\rm restriction}V\\ H^0(C,\omega_C) \> \subset >> H^0(C\cap U,\Omega^1_{C\cap U/{\mathbb C}}). \end{CD} $$ (Here the right hand vertical arrow is given by restriction of 1-forms.) \item[(b)] Let $U\subset X_{{\rm reg}}$ be a dense Zariski open set, and let $\omega\in \Gamma(U,\Omega^1_{U/{\mathbb C}})$ be closed. Then $\omega\in \Omega(A^n(X))$ if and only if \begin{enumerate} \item[(i)] $\omega$ yields a meromorphic section on $X$ of $\Omega^1_{X/{\mathbb C}}$ \item[(ii)] for any reduced, local complete intersection Cartier curve $C\subset X$ such that $C\cap U$ is dense in $C$, the restriction of $\omega$ to $B=C_{reg}\cap U$ is in the image of the natural injective map \[H^0(C,\omega_C) \>>> H^0(B,\Omega^1_{B/{\mathbb C}}).\] \end{enumerate} \end{enumerate} \end{propose} \begin{proof} First we prove (a). From lemma~\ref{purity}, it suffices to prove that if $\beta_C:H^1(C,{\mathcal O}_C)\to H^n(X,\Omega^{n-1}_{X/{\mathbb C}})$ is the obvious map through which $\alpha_C$ factors, then the dual map $\beta_C^{\vee}$ fits into a commutative diagram $$ \begin{CD} H^0(X,\shom_X(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)) \> \subset >> H^0(X_{\rm reg}, \Omega^1_{X_{\rm reg}/{\mathbb C}})\\ {\mathbb V} V {\beta_C^{\vee}}V {\mathbb V} V {\rm restriction}V\\ H^0(C,\omega_C) \> \subset >> H^0(C\cap X_{\rm reg},\omega_{C\cap X_{{\rm reg}}}). \end{CD} $$ Here we have used Serre duality on $X$ and $C$ to make the identifications $$H^n(X,\Omega^{n-1}_{X/{\mathbb C}})^{\vee}= H^0(X,\shom_X(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)),$$ $$H^1(C,{\mathcal O}_C)^{\vee}=H^0(C,\omega_C).$$ Since $C$ is a reduced, local complete intersection Cartier curve in $X$ (so that $C\subset X_{\rm CM}\cap X^{(n)}$), we have the adjunction formula \[\omega_C=\shom_C(\bigwedge^{n-1}{\mathcal I}_C/{\mathcal I}_C^2,\omega_X\otimes{\mathcal O}_C).\] Hence there is a natural sheaf map \[\psi_C:\shom_X(\Omega^{n-1}_{X/{\mathbb C}},\omega_X)\>>> \shom_C(\bigwedge^{n-1}{\mathcal I}_C/{\mathcal I}_C^2,\omega_X\otimes{\mathcal O}_C)=\omega_C\] induced by restriction to $C$, and composition with the natural map \begin{align*} \bigwedge^{n-1}{\mathcal I}_C/{\mathcal I}_C^2&\>>>\Omega^{n-1}_{X/{\mathbb C}}\otimes{\mathcal O}_C,\\ f_1\wedge\cdots\wedge f_{n-1}&\longmapsto df_1\wedge\cdots\cdots df_{n-1}. \end{align*} On any open set $U\subset X_{{\rm reg}}$ with $U\cap C\subset C_{{\rm reg}}$, one verifies at once, from the explicit description, that the map $\psi_C\mid_U$ is just the restriction map on 1-forms $\Omega^1_{U/{\mathbb C}}\to\Omega^1_{C\cap U/{\mathbb C}}$. Hence the desired commutativity (which implies (a)) follows from: \begin{claim}\label{duality} $\beta^{\vee}$ is the map induced by $\psi_C$ on global sections. \end{claim} To prove the claim, first note that for the local complete intersection curve $C$ in $X_{\rm CM}$, one also has $$ \sext^{n-a}_X({\mathcal O}_C,\omega_X)= \left\{ \begin{array}{ll} \omega_C & \mbox{ \ for \ } a=1\\ 0 & \mbox{ \ for \ } a\neq 1, \end{array} \right. $$ Hence there is a Gysin map given as the composite \begin{multline*} H^1(C,\omega_C)=H^1(X,\sext^{n-1}_X({\mathcal O}_C,\omega_X)) \> \epsilon >> {\rm Ext}^n_X({\mathcal O}_C,\omega_X)\\ \>>> H^n_C(X,\omega_X) \>>> H^n(X,\omega_X) \end{multline*} where $\epsilon$ is the isomorphism resulting from the (degenerate) spectral sequence $$ E_2^{a,b-a}=H^a(X,\sext^{b-a}_X({\mathcal O}_C,\omega_X)) \Longrightarrow {\rm Ext}^b_X({\mathcal O}_C,\omega_{X}). $$ The trace map ${\rm Tr}_C:H^1(C,\omega_C) \to {\mathbb C}$ (of Serre duality on $C$) factors as $$ {\rm Tr}_C:H^1(C,\omega_C) \>{\rm Gysin}>> H^n(X,\omega_X) \> {\rm Tr}_X >> {\mathbb C} $$ (one way to verify this is to show that the composite ${\rm Tr}_X\circ{\rm Gysin}$ has the universal property of ${\rm Tr}_C$). Now the claim~\ref{duality} amounts to the assertion that the following diagram commutes: \[\begin{CD} H^1(C,{\mathcal O}_C) \>{\rm Gysin}>> H^n(X,\Omega^{n-1}_{X/{\mathbb C}})\\ {\mathbb V}{\psi_C(\varphi)} V V {\mathbb V}{\varphi} V V\\ H^1(C,\omega_C) \>{\rm Gysin}>> H^n(X,\omega_X) \end{CD}\] {From} Remark~\ref{gys_rem}, this will follow if we prove the commutativity of the diagram of ${\mathcal O}_X$-linear maps $$ \begin{CD} {\mathcal O}_C \>>> \sext^{n-1}_X({\mathcal O}_C,\Omega^{n-1}_{X/{\mathbb C}})\\ {\mathbb V} \psi_C(\varphi) V V {\mathbb V} V \varphi V\\ \omega_C \> \cong >> \sext^{n-1}_X({\mathcal O}_C,\omega_X). \end{CD} $$ As $\omega_C$ is torsion-free, it is enough to check this commutativity on a suitable open subset of the regular locus of $C$, where it is easily verified. We now show the ``if'' part of (b) (note that the other direction follows directly from (a)). By lemma~\ref{basic1}, it suffices to prove that for each finite, linear projection\\ $\pi:X\to\P^n_{{\mathbb C}}$, the meromorphic 1-form $\omega$ determines a section of $$\shom_X(\pi^*\Omega^{n-1}_{\P^n_{{\mathbb C}}/{\mathbb C}},\omega_X).$$ Since $\pi$ is a finite morphism, \[\pi_*\omega_X=\shom_{\P^n_{{\mathbb C}}}(\pi_*{\mathcal O}_X,\omega_{\P^n_{{\mathbb C}}}),\] and we have a sequence of natural identifications of sheaves \begin{align*} &\pi_*\shom_X(\pi^*\Omega^{n-1}_{\P^n_{{\mathbb C}}/{\mathbb C}},\omega_X)\cong \shom_{\P^n_{{\mathbb C}}}(\Omega^{n-1}_{\P^n_{{\mathbb C}}/{\mathbb C}}\otimes\pi_*{\mathcal O}_X, \omega_{\P^n_{{\mathbb C}}})\\ &\cong \shom_{\P^n_{{\mathbb C}}}(\pi_*{\mathcal O}_X, \shom_{\P^n_{{\mathbb C}}}(\Omega^{n-1}_{\P^n_{{\mathbb C}}/{\mathbb C}},\omega_{\P^n_{{\mathbb C}}})) \cong {\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}, \end{align*} where ${\mathcal F}=\shom_{\P^n_{{\mathbb C}}}(\pi_*{\mathcal O}_X,{\mathcal O}_{\P^n_{{\mathbb C}}})$ is a (non-zero) coherent reflexive sheaf on $\P^n_{{\mathbb C}}$. Let $W\subset\P^n_{{\mathbb C}}$ be a dense open subset such that $\pi^{-1}(W)\subset U$. Then $\omega$ determines a section of ${\mathcal F}\otimes\Omega^1_{\P^n_{{\mathbb C}}/{\mathbb C}}$ on $W$, and we want to show it extends to a global section of this sheaf. We do this by verifying that the hypotheses of lemma~\ref{basic2} are satisfied. Let $L$ be a line in $\P^n_{{\mathbb C}}$, disjoint from the non-flat locus of $\pi:X^{(n)}\to\P^n_{{\mathbb C}}$ (which is a subset of $\P^n_{{\mathbb C}}$ of codimension $\geq 2$, since $X^{(n)}$ is reduced and purely of dimension $n$). Then the scheme-theoretic inverse image of $L$ in $X^{(n)}$ is a closed, local complete intersection subscheme of $X^{(n)}$, purely of dimension 1, and which is contained in the Cohen-Macaulay locus of $X^{(n)}$ (since $X^{(n)}$ is Cohen-Macaulay precisely at all points $x\in X$ where $\pi$ is flat). If further $L$ is not contained in the branch locus of $\pi$ on $X^{(n)}$ ({\it i.e.\/},\ $\pi$ is \'etale over all but finitely many points of $L$), then $\pi^{-1}(L)=D$ is non-singular outside a finite set. Thus $D$ is a reduced, complete intersection curve in $X^{(n)}$. Further, if $D\cap X^{<n}=\emptyset$, then $D$ is a reduced local complete intersection curve in $X$, whose non-singular locus is contained in $X_{{\rm reg}}$. In particular $D$ is a reduced Cartier curve in $X$. Finally, if $L$ is not contained in the image of $X-U$, then $D$ has finite intersection with $X-U$, and hence $D\cap U$ is dense in $D$. Clearly the set of all such lines $L$ contains a non-empty open subset of the Grassmannian of lines. For a line $L$ as above, we have \begin{align*} \pi_*\omega_D&\cong \shom_{L}(\pi_*{\mathcal O}_D,\omega_L)\cong \shom_L(\pi_*{\mathcal O}_X\otimes{\mathcal O}_L,\omega_L)\\ &\cong \shom_L(\pi_*{\mathcal O}_X\otimes{\mathcal O}_L,\Omega^1_{L/{\mathbb C}}) \cong {\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}, \end{align*} since ${\mathcal F}\otimes{\mathcal O}_L\cong \shom_L(\pi_*{\mathcal O}_D,{\mathcal O}_L)$ (as $\pi$ is flat over $L$). Since we are given that the image of $\omega$ in $\Omega^1_{D\cap U/{\mathbb C}}$ extends to a global section of $\omega_D$, it follows that the corresponding section of ${\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}\mid_{L\cap W}$ extends to a global section of ${\mathcal F}\otimes\Omega^1_{L/{\mathbb C}}$. Thus we have verified the hypotheses of lemma~\ref{basic2}. \end{proof} \begin{rmk}\label{example} Two properties of $A^n(Y)$, which are true for smooth projective varieties $Y$, do not carry over to the general case: the compatibility with products, and its dimension being constant in a flat family. We give examples to illustrate these pathologies. Let $X$ and $Y$ be projective varieties of dimension $n$ and $m$, respectively, and let $r(X)$ and $r(Y)$ denote the number of irreducible components of dimensions $n$ and $m$ respectively. By \cite{D} the K\"unneth decomposition $$ H^{2(n+m)-1}(X\times Y, {\mathbb Z})/_{({\rm torsion})} = \left[H^{2n-1}(X,{\mathbb Z})^{r(Y)} \oplus H^{2m-1}(Y,{\mathbb Z})^{r(X)}\right] /_{({\rm torsion})} $$ is compatible with the Hodge structure. Thus \begin{equation}\label{product1} J^{n+m}(X\times Y) = J^n(X)^{r(Y)} \times J^m(Y)^{ r(X)} . \end{equation} For $A^{n+m}(X \times Y)$ the picture is wilder. By (\ref{lie_descr}) in the proof of \ref{purity}, we have \begin{gather}\label{product2} \Lie(A^{n+m}(X\times Y))=\\ \frac{H^n(\Omega^{n-1}_{X/{\mathbb C}})\otimes H^m(\Omega^{m}_{Y/{\mathbb C}})\oplus H^n(\Omega^{n}_{X/{\mathbb C}})\otimes H^m(\Omega^{m-1}_{Y/{\mathbb C}})}{H^n(\Omega^{n-2}_{X/{\mathbb C}})\otimes H^m(\Omega^{m}_{Y/{\mathbb C}})\oplus H^n(\Omega^{n-1}_{X/{\mathbb C}})\otimes H^m(\Omega^{m-1}_{Y/{\mathbb C}})\oplus H^n(\Omega^{n}_{X/{\mathbb C}})\otimes H^m(\Omega^{m-2}_{Y/{\mathbb C}})} \notag \end{gather} where the maps from the denominator are $$ d_X\otimes {\rm id}_Y, \ \ d_X\otimes {\rm id}_Y + (-1)^{n-1}{\rm id}_X \otimes d_Y\mbox{ \ \ and \ \ }{\rm id}_X \otimes d_Y. $$ Consider an elliptic curve $E$, the rational curve $\Gamma=(x^3-y^2z) \subset \P^2_{\mathbb C}$, with a cusp, and the union of three rational curves $ C = (xyz) \subset \P^2_{\mathbb C}$. They all are fibres of the family ${\mathcal C} \to \P=\P(H^0(\P^2,{\mathcal O}_{\P^2}(3)))$ of curves of degree three in $\P^2$. Hodge theory implies that $$ d:H^1(E,{\mathcal O}_E) \>>> H^1(E,\Omega_E^1)\mbox{ \ \ and \ \ } d:H^1(C,{\mathcal O}_C) \>>> H^1(C,\Omega_C^1)\cong {\mathbb C}^3 $$ are both zero. Using (\ref{product2}) this shows \begin{align*} &A^2(C \times E) = J^2(C \times E) = A^1(C)\times A^1(E)^{3}= {\mathbb G}_m \times E^{3}\\ &A^2(C \times C) = J^2(C \times C) = A^1(C)^{3}\times A^1(C)^{3}= {\mathbb G}_m^6. \end{align*} On the other hand, $\Gamma - (0:1:0)= {\rm Spec \,}({\mathbb C}[t^2,t^3])$ and, if $\pi: \tilde{\Gamma} \to \Gamma$ denotes the normalization, one has exact sequences \begin{gather*} 0\>>> {\mathcal O}_{\Gamma} \>>> \pi_*({\mathcal O}_{\tilde{\Gamma}}) \>>> {\mathbb C} t \>>> 0 \mbox{ \ \ and}\\ 0\>>> \Omega^1_{\Gamma/{\mathbb C}} \>>> \pi_*(\Omega^1_{\tilde{\Gamma}/{\mathbb C}}) \>>> {\mathbb C} dt \>>> 0. \hspace*{1cm} \end{gather*} Thus ${\mathbb C} t=H^1(\Gamma,{\mathcal O}_\Gamma) \> d >\cong > {\mathbb C} dt \> > \subset > H^1(\Gamma,\Omega^1_{\Gamma/{\mathbb C}}) \cong {\mathbb C}^2$ and one obtains by \ref{product2} \begin{align*} &A^2(\Gamma \times E) = {\mathbb G}_a \times E = A^1(\Gamma)\times A^1(E)\\ &A^2(\Gamma \times \Gamma) = \frac{{\mathbb C}^2 \times {\mathbb C}^2}{{\mathbb C}} = {\mathbb G}_a^3 \mbox{ \ \ whereas \ \ } A^1(\Gamma)\times A^1(\Gamma) = {\mathbb G}_a \times {\mathbb G}_a. \end{align*} In particular, a product formula as (\ref{product1}) fails for $A^n$ instead of $J^n$, and the dimension of $J^n$ and $A^n$ are not constant for the fibres ${\mathcal C} \times {\mathcal C} \to \P \times \P$. It is amusing to write down the cycle map for the last example. Writing $$ \Gamma_{\rm reg}\times \Gamma_{\rm reg} = (\Gamma -(0:0:1))\times (\Gamma-(0:0:1)) = {\rm Spec \,}({\mathbb C}[u]\otimes_{\mathbb C} {\mathbb C}[v]), $$ $\Omega(\Gamma\times\Gamma)={\rm Hom}_{\Gamma\times\Gamma}(\Omega^1_{\Gamma\times\Gamma/{\mathbb C}}, \omega_{\Gamma\times\Gamma})_{\rm cl}$ decomposes as \begin{align*} &(H^0(\Gamma,\shom(\Omega^1_{\Gamma/|C},\omega_\Gamma)) \otimes H^0(\Gamma,\omega_\Gamma) \oplus H^0(\Gamma,\omega_\Gamma) \otimes H^0(\Gamma,\shom(\Omega^1_{\Gamma/{\mathbb C}},\omega_\Gamma)))_{\rm cl}\\ &= ({\mathbb C} dv \oplus {\mathbb C} udv \oplus {\mathbb C} du \oplus {\mathbb C} vdu)_{\rm cl} = {\mathbb C} dv \oplus {\mathbb C} du \oplus {\mathbb C} (udv+vdu). \end{align*} The cycle map is \begin{align*} \Pi_{\Gamma_{\rm reg} \times \Gamma_{\rm reg}}={\mathbb A}_{\mathbb C}^2 \times {\mathbb A}_{\mathbb C}^2 & \>>> {\mathbb G}_a^3\\ ((x_1,x_2),(y_1,y_2)) & \longmapsto \left\{ \begin{array}{rll} du &\mapsto &y_1-x_1\\ dv &\mapsto &y_2-x_2\\ udv+vdu & \mapsto & y_1y_2-x_1x_2. \end{array}\right. \end{align*} \end{rmk} \section{The universal property over ${\mathbb C}$} Let $U_1,\ldots,U_r$ be the connected components of $X_{{\rm reg}}$, and for each $i$, let $p_i\in U_i$ be a base point. Let $G$ be a commutative algebraic group. By \ref{def-reg} and \ref{equ-reg} a homomorphism (of abstract groups) $\phi:CH^n(X)_{\deg 0}\to G$ is regular, if and only if $\phi\circ\gamma_m:S^m(X_{\rm reg})\to G$ is a morphism of varieties, for some $m>0$. \begin{thm}\label{regular} \ \ \begin{points} \item The homomorphism $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$ constructed in lemma~\ref{equivalence} is regular and surjective. \item The cokernel of the map $H_1(X_{\rm reg},{\mathbb Z}) \to \Lie(A^n(X))$, defined by integration of 1-forms over homology classes, is naturally isomorphic to $A^n(X)$ and the composite $(\varphi\circ\gamma)^{(-)}: \Pi_{X_{\rm reg}} \to CH^n(X)_{\deg 0}\to A^n(X)$ is given by $$ (x,y) \longmapsto \bigr\{ \omega \mapsto \int^y_x \omega \bigr\} $$ \item {\em (Universality)} $\varphi$ satisfies the following universal property: for any regular homomorphism $\phi:CH^n(X)_{\deg 0}\to G$ to a commutative algebraic group there exists a unique homomorphism $h:A^n(X)\to G$ of algebraic groups with $\phi=h\circ\varphi$. \end{points} \end{thm} \begin{proof}[Proof of (i)] It suffices to prove that $\varphi\circ\gamma_1:U=X_{\rm reg}\to A^n(X)$ is a morphism. Note that, from the definition, it is clearly analytic. Further, we have the following. \begin{enumerate} \item[(a)] the composition $U\to A^n(X)\to {\rm Alb}\,(\tilde{X})$ is a morphism, where $\tilde{X}$ is a resolution of singularities of $X^{(n)}$, since we may then regard $U$ as an open subset of $\tilde{X}$, and the map $U\to {\rm Alb}\,(\tilde{X})$ is the restriction of the Albanese mapping for $\tilde{X}$, with appropriate base-points. Here ${\rm Alb}\,(\tilde{X})$ is the product of the Albanese varities of the connected components of $\tilde{X}$, and $A^n(X)$ is an extension of ${\rm Alb}\,(\tilde{X})$ by a group ${\mathbb G}_a^r\times{\mathbb G}_m^s$, so that in particular $A^n(X)\to{\rm Alb}\,(\tilde{X})$ is a Zariski locally trivial fibre bundle. \item[(b)] For each reduced Cartier curve $C\subset X$, the composite $$C_{{\rm reg}}\>>> U\>>> A^n(X)$$ is a morphism. Indeed, for each component $B_0$ of $C_{\rm reg}$, the composition \[B_0\>>> U\>{\gamma_1}>>CH^n(X)_{\deg 0}\>{\varphi}>> A^n(X)\] agrees with \[B_0\>>> {\rm Pic}^0(C)\>>> CH^n(X)_{\deg 0}\>{\varphi}>>A^n(X)\] up to a translation, and by corollary~\ref{algcurve}, the latter is algebraic. \end{enumerate} Now we may argue as in \cite{BiS}: we are reduced to proving that if $V$ is a non-singular affine variety, a holomorphic function on $V$ which is algebraic when restricted to ``almost all'' algebraic curves in $V$, is in fact an algebraic regular function. This may be proved using Noether normalization and power series expansions for holomorphic functions on ${\mathbb C}^n$, or deduced from \cite{Si}, (1.1). Since $\Omega(A^n(X))$ is a finite dimensional subspace of 1-forms on $U$, there exist reduced local complete intersection Cartier curves $C_i \subset X$, for $i=1,\ldots,s$, such that $$ \Omega(A^n(X)) \>>> \bigoplus_{i=1}^s H^0(C_i,\omega_{C_i}) $$ is injective. Hence $$ \bigoplus_{i=1}^s {\rm Pic}^0(C_i) \> \oplus \psi_i >> A^n(X) $$ is surjective. \noindent {\it Proof of (ii) and (iii):}\quad Let $\phi:CH^n(X)_{\deg 0}\to G$ be a regular homomorphism to a commutative algebraic group $G$. By lemma \ref{pic} the image of $\phi$ is contained in the connected component of the identity of $G$. Hence we may assume without loss of generality that $G$ is connected. Now $\Omega(G)$ consists of closed, translation-invariant 1-forms. Thus if $$h=\phi\circ\gamma_1:U\>>> G,$$ then the image of $h^*:\Omega(G)\to \Gamma(U,\Omega^1_{X/{\mathbb C}})$ is contained in the subspace of closed 1-forms. We claim that in fact $h^*(\Omega(G))\subset \Omega(A^n(X)).$ This is deduced from the criterion of proposition~\ref{basic}, (b), since we know that for any reduced Cartier curve $C$ in $X$, the composition $${\rm Pic}^0(C)\>>> CH^n(X)_{\deg 0}\>{\phi}>> G$$ is a homomorphism of algebraic groups. Now we observe that if $B_0$ is any component of $C_{{\rm reg}}$, then $$ B_0 \>>> {\rm Pic}^0(C)=\Lie(A^1(C))/{\rm image}\,H_1(C_{\rm reg},{\mathbb Z}) $$ is given by integration of 1-forms in $H^0(C,\omega_C)$. Moreover the composite $$ B_0\>>> U\>{h}>> G $$ agrees with $$ B_0\>>>{\rm Pic}^0(C)\>>>G, $$ up to a translation by an element of $G$ (and elements of $\Omega(G)$ are translation invariant). Dualizing the above inclusion on 1-forms, we thus obtain a map on Lie algebras $\Lie(A^n(X))\to\Lie(G).$ This fits into a commutative diagram \[\begin{CD} H_1(U,{\mathbb Z})\>>> \Lie (A^n(X))\\ {\mathbb V} V V {\mathbb V} V V \\ H_1(G,{\mathbb Z})\>>> \Lie(G) \end{CD}\] where the horizontal arrows are given by integration of 1-forms over homology classes. Further, there is a commutative diagram \begin{equation}\label{diag} \begin{CD} U \>>> \Lie(A^n(X))/{\rm image}\,H_1(U,{\mathbb Z})\\ {\mathbb V} \gamma_1 V V {\mathbb V} V \tilde{\phi}V \\ CH^n(X)_{\deg 0}\> {\phi}>> G=\Lie(G)/{\rm image}\,H_1(G,{\mathbb Z}) \end{CD} \end{equation} where $\tilde{\phi}$ is a homomorphism of analytic groups, and where the upper horizontal arrow is given by integration of 1-forms in $\Omega(A^n(X))$. We claim that the map $H_1(U,{\mathbb Z})\to \Lie(A^n(X))=\Omega(A^n(X))^*$ factors through the (surjective) composition \[H_1(U,{\mathbb Z})\>{\cong}>> H^{2n-1}_c(U,{\mathbb Z}(n))\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} H^{2n-1}(X,{\mathbb Z}(n)),\] where $H^*_c$ denotes compactly supported cohomology, and the isomorphism is by Poincar\'e duality. Indeed, let $C\subset X^{(n)}$ be a sufficiently general reduced complete intersection curve in $X^{(n)}$. Then \[C\cap X^{<n}=\emptyset,\;\; C_{\rm sing}=C\cap X^{(n)}_{sing}=C\cap X_{{\rm sing}},\] and one has a Gysin homomorphism \[H^1(C,{\mathbb Z}(1))\to H^{2n-1}(X,{\mathbb Z}(n))\cong H^{2n-1}(X^{(n)},{\mathbb Z}(n))\] which fits into a commutative diagram with exact rows \[\begin{CD} H^0(C_{{\rm sing}},{\mathbb Z}(1)) \>>> H^1_c(C\cap U,{\mathbb Z}(1)) \>>> \hspace*{-.82cm}\to \hspace{.42cm} H^1(C,{\mathbb Z}(1))\hspace{.3cm}\\ {\mathbb V} V V {\mathbb V} V V {\mathbb V} V{\rm Gysin}V\\ H^{2n-2}(X^{(n)}_{{\rm sing}},{\mathbb Z}(n))\>>> H^{2n-1}_c(U,{\mathbb Z}(n))\>>> \hspace*{-.7cm} \to \hspace{.2cm} H^{2n-1}(X^{(n)},{\mathbb Z}(n)) \end{CD}\] The left hand vertical arrow is in fact surjective, since $H^{2n-2}(X^{(n)}_{{\rm sing}},{\mathbb Z}(n-1))$ is the free abelian group on the $(n-1)$-dimensional components of $X^{(n)}_{{\rm sing}}$, and (since $C$ is a general complete intersection) $C_{{\rm sing}}$ has non-empty intersection (which is supported at smooth points, and is transverse) with each such component of $X^{(n)}_{{\rm sing}}$. Now we note that the composite $H_1(C\cap U,{\mathbb Z})\to H_1(U,{\mathbb Z})\to \Lie(A^n(X))$ factors through the surjective composite $H_1(U,{\mathbb Z})\cong H^1_c(U,{\mathbb Z}(1))\to H^1(C,{\mathbb Z}(1))$, since $C\cap U\to U\to A^n(X)$ is compatible with a homomorphism ${\rm Pic}^0(C)\to A^n(X)$ (here ``compatible'' means that for any component $B_0$ of $C\cap U$, the composites $B_0\to U\to A^n(X)$ and $B_0\to{\rm Pic}^0(C)\to A^n(X)$ agree up to translation by an element of $A^n(X)$). Now a diagram chase implies the claim made at the beginning of the paragraph. Thus in the diagram (\ref{diag}) we see that $\Lie(A^n(X))/{\rm image}\,H_1(U,{\mathbb Z})$ is identified with \[{\Lie(A^n(X))}/{{\rm image}\,H^{2n-1}(X,{\mathbb Z}(n))}\;=A^n(X).\] Hence there is a homomorphism $\tilde{\phi}: A^n(X)\to G$, such that $\gamma_1 \circ \phi:U\to G$ factors through $A^n(X)$. Since $\gamma_1^*:\Omega(A^n(X))\to \Gamma(U,\Omega^1_{U/{\mathbb C}})$ is injective, the induced map $A^n(X)\to G$ with this property is unique, since the corresponding map on Lie algebras is uniquely determined. Since ${\rm image}\,\gamma_1$ generates $CH^n(X)_{\deg 0}$, the two homomorphisms \[\phi:CH^n(X)_{\deg 0}\>>> G,\;\; CH^n(X)_{\deg 0}\>{\varphi}>> A^n(X)\>{\tilde{\phi}}>> G\] must coincide. This proves the universal property of $\varphi$, except that we need to note that $\tilde{\phi}$ is a morphism. By lemma \ref{equ-reg}, $\phi$ induces an algebraic group homomorphism ${\rm Pic}^0(C) \to G$ for all admissible pairs $(C', \iota)$, with $C = \iota(C')$. As above, we can choose reduced complete intersection curves $C_i$, $i=1, \ldots, s$, such that $$ \bigoplus_{i=1}^s {\rm Pic}^0(C_i) \> \oplus \psi_i >> A^n(X) $$ is surjective. As $\tilde{\phi} \circ (\oplus \psi_i)$ is an algebraic group homomorphism, $\tilde{\phi}$ is an algebraic group homomorphism as well. \end{proof} \begin{rmk} Lemma~\ref{algebraicity}, combined with the Roitman Theorem proved in [BiS], imply that $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$ is an isomorphism on torsion subgroups. In other words, the Roitman Theorem is valid for $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$, over ${\mathbb C}$. This is another similarity with the Albanese mapping for a non-singular projective variety. \end{rmk} \begin{rmk} The proof of theorem \ref{regular} is close in spirit to the construction of a ``generalized Albanese variety'' in \cite{FW}. There Faltings and W\"ustholz consider a finite dimensional subspace $V \subset H^0(X_{\rm reg},\Omega^1_{X_{\rm reg}})$, containing the 1-forms with logarithmic poles on some desingularization of $X^{(n)}$, and they construct a commutative algebraic group $G_V$ together with a morphism $X_{\rm reg} \to G_V$, which is universal among the morphisms $\tau: X_{\rm reg} \to H$ to commutative algebraic groups $H$, with $\tau^*(\Omega(G)) \subset V$. \end{rmk} \section{Picard groups of Cartier curves} In the next section, we give an algebraic construction of $A^n(X)$ for a reduced projective $n$-dimensional variety $X$, defined over an algebraically closed field $k$. As in the analytic case, we will use the Picard scheme for Cartier curves in $X$ and for families of such curves. In this section, we discuss some properties of such families of curves, and the corresponding Picard schemes. In particular, we establish the technical results \ref{mu2} and \ref{mu3}, which are important steps in the algebraic construction of $A^n(X)$. Let $S$ be a non-singular variety, and let $f:{\mathcal C} \to S$ be a flat family of projective curves with reduced geometric fibres $C_s=f^{-1}(s)$. Then $$ g(C_s):= \dim_{k(s)} H^1(C_s,{\mathcal O}_{C_s}) \mbox{ \ \ \ and \ \ \ } \# C_s:= \dim_{k(s)} H^0(C_s,{\mathcal O}_{C_s}) $$ are both constant on $S$. In fact, let ${\mathcal C} \> \tilde{f} >> \tilde{S} \>\varkappa >> S$ be the Stein factorization of $f$. Since the fibres of $f$ are reduced, $\varkappa:\tilde{S}\to S$ is a finite \'etale morphism and $\# C_s=\#\varkappa^{-1}(s)$ is constant, as well as $g(C_s)= \chi(C_s,{\mathcal O}_{C_s}) - \# C_s$. Let $S' \to S$ be a finite (possibly branched) covering such that $f':{\mathcal C}' = {\mathcal C} \times_SS'\to S'$ is the disjoint union of families of curves ${\mathcal C}'_i \to S'$, for $i=1,\ldots ,s$ with connected fibres. By \cite{BLR}, 8.3, theorem 1, the relative Picard functors ${\rm Pic}_{{\mathcal C}'_i/S'}$ are represented by an algebraic space ${\rm Pic}({\mathcal C}'_i/S')$, and we define $${\rm Pic}({\mathcal C}'/S')={\rm Pic}({\mathcal C}'_i/S')\times_{S'}\cdots\times_{S"}{\rm Pic}({\mathcal C}'/S'). $$ For the smooth locus ${\mathcal C}_{\rm sm}$ of $f$ consider the $g$-th symmetric product $$ {f'}^g:S^g({\mathcal C}'_{\rm sm} /S')\>>> S' $$ over $S'$. For any open subscheme $W'\subset S^g({\mathcal C}'_{\rm sm} /S')$ there is a natural map $\vartheta_{W'} : W' \to {\rm Pic}({\mathcal C}'/S')$. By \cite{BLR}, 9.3, lemmas 5 and 6, one has the following generalization of \ref{generators2}: \begin{lemma}\label{relpic2} After replacing $S'$ by an \'etale covering, there exists an open subscheme $W'\subset S^g({\mathcal C}'_{\rm sm} /S')$ with geometrically connected fibres over $S'$, such that $\vartheta_{W'} : W' \to {\rm Pic}({\mathcal C}'/S')$ is an open embedding. \end{lemma} Recall that $X^{(n)}$ denotes the union of the $n$-dimensional irreducible components of $X$, and $X^{<n}$ is the union of the smaller dimensional components. \begin{notations}\label{linearsystem} For a very ample invertible sheaf ${\mathcal L}$ on $X^{(n)}$ we write $$ |{\mathcal L}|^{n-1} = \P(H^0(X^{(n)},{\mathcal L}))\times \cdots \times \P(H^0(X^{(n)},{\mathcal L})) \ \ \ \ (n-1)\mbox{-times} $$ and $|{\mathcal L}|^{n-1}_0$ for the open subscheme defined by $n-1$-tuples $D_1, \ldots , D_{n-1}$ of divisors such that \begin{enumerate} \item[(i)] $C = D_1 \cap \cdots\cap D_{n-1}$ is a reduced complete intersection curve in $X^{(n)}$, \item[(ii)] $C\cap X^{<n}=\emptyset$, and \item[(iii)] $X_{\rm reg} \cap C$ is non-singular and dense in $C$. \end{enumerate} Note that by (ii), $C$ is a reduced Cartier curve in $X$ which is a local complete intersection. By abuse of notation we will sometimes write $C \in |{\mathcal L}|^{n-1}$ instead of $(D_1, \ldots , D_{n-1}) \in |{\mathcal L}|^{n-1}$. \end{notations} The normalization $\pi :\tilde{C} \to C$ induces a surjection $\pi^*:{\rm Pic}^0(C) \to {\rm Pic}^0(\tilde{C})$. By \cite{BLR}, 9.2, the kernel of $\pi^*$ is the largest linear subgroup $H(C)$ of ${\rm Pic}(C)$. One has \begin{align}\label{chieq} \dim(H(C)) & = \dim({\rm Pic}^0(C)) - \dim({\rm Pic}^0(\tilde{C}))\notag\\ & = \dim_k(H^1(C,{\mathcal O}_C)) - \dim_k(H^1(\tilde{C},{\mathcal O}_{\tilde{C}}))\\ & = \chi(\tilde{C},{\mathcal O}_{\tilde{C}}) - \chi({C},{\mathcal O}_{{C}}) - ( \# \tilde{C} - \# C ),\notag \end{align} where again $\# C$ and $\# \tilde{C}$ denote the number of connected components of $C$ and $\tilde{C}$, respectively. Given a flat family of projective curves $f:{\mathcal C} \to S$ over an irreducible variety $S$ with reduced geometric fibres $C_s$, there exists a finite (possibly branched) covering $S'\to S$ and an open dense subscheme $S'_0\subset S'$ such that the normalization of ${\mathcal C}\times_S S'_0$ is smooth over $S'_0$. Hence $\# \tilde{C}_s$, and the dimension of the linear part $H(C_s)$ of ${\rm Pic}^0(C_s)$, are both constant on the image of $S'_0$. \begin{defn}\label{mu} For a reduced projective curve $C$ we define $r(C)$ to be the number of irreducible components of $C$, and $\mu(C)$ to be the dimension of the largest linear subgroup of ${\rm Pic}^0(C)$. By \cite{F}, Satz 5.2, for a very ample invertible sheaf ${\mathcal L}$ the open subscheme $|{\mathcal L}|^{n-1}_0$ is not empty. Then $r({\mathcal L})$ and $\mu({\mathcal L})$ denote the values of $r(C)$ and of $\mu(C)$ for $C \in |{\mathcal L}|^{n-1}_0$ in general position. \end{defn} By the equality (\ref{chieq}) one has: \begin{equation}\label{chieq2} \chi(\tilde{C},{\mathcal O}_{\tilde{C}}) - \chi({C},{\mathcal O}_{{C}}) \geq \mu(C) \geq \chi(\tilde{C},{\mathcal O}_{\tilde{C}}) - \chi({C},{\mathcal O}_{{C}}) - r(C) + 1. \end{equation} \begin{lemma}\label{muinequality} For a very ample invertible sheaf ${\mathcal L}$ and for a positive integer $N$, $$\mu({\mathcal L}^N) \leq N^{n-1}\cdot (\mu({\mathcal L}) + r({\mathcal L}) -1).$$ \end{lemma} \begin{proof} Given $D^{(i)}_j \in |{\mathcal L}|$, for $i=1, \ldots , N$ and $j=1, \ldots , n-1$, we write \begin{align*} I &= \{1, \ldots ,N\}^{n-1}\\ C^{(\underline{i})} &= D^{(i_1)}_1 \cap \cdots \cap D^{(i_{n-1})}_{n-1}\mbox{ \ \ for \ \ } \underline{i} = (i_1, \ldots , i_{n-1}) \in I\\ \mbox{and \ \ \ \ } C &= \bigcup_{\underline{i} \in I} C^{(\underline{i})} = \bigcap_{j=1}^{n-1} (D_j^{(1)} \cup \cdots \cup D_j^{(N)}). \end{align*} \begin{claim}\label{general} There exists a choice of the divisors $D^{(i)}_j \in |{\mathcal L}|$ such that \begin{enumerate} \item[(a)] $C^{(\underline{i})} \in |{\mathcal L}|^{n-1}_0$, \ $\mu(C^{(\underline{i})})= \mu({\mathcal L})$ \ and \ $r(C^{(\underline{i})})= r({\mathcal L})$ \vspace{.05cm} \item[(b)] $C^{(\underline{i})} \cap C^{(\underline{i}')} \cap X_{\rm sing} = \emptyset$ for $\underline{i} \neq \underline{i}'$\vspace{.05cm} \item[(c)] each point $x \in C_{\rm sing} \cap X_{\rm reg}$ lies on exactly two components $C^{(\underline{i})}$ and $C^{(\underline{i}')}$. In this case, there exists one $\nu$ with $i_j = i'_j$ for all $j\neq \nu$. Locally in $x$ the surface $$Y = D^{(i_1)}_1 \cap \cdots \cap \widehat{D^{(i_{\nu})}_\nu} \cap \cdots \cap D^{(i_{n-1})}_{n-1}\cap X_{\rm reg} $$ is nonsingular and contains $C^{(\underline{i})}$ and $C^{(\underline{i}')}$ as two smooth divisors intersecting transversally. \vspace{.05cm} \item[(d)] $C$ is a reduced complete intersection curve in $|{\mathcal L}^N|^{n-1}$. \end{enumerate} \end{claim} \begin{proof} (d) follows from (a), (b) and (c). Since $|{\mathcal L}|^{n-1}_0$ is open and dense in $|{\mathcal L}|^{n-1}$ (a) holds true for sufficiently general divisors. Counting dimensions one finds that for $\underline{i}\neq \underline{i}'$ the intersection $C^{(\underline{i})} \cap C^{(\underline{i}')}$ is either empty or consists of finitely many points. The latter can only happen, if all but one entry in $\underline{i}$ and $\underline{i}'$ are the same, and obviously one may assume that the intersection points all avoid $X_{\rm sing}$. Moreover $$ C^{(\underline{i})} \cap C^{(\underline{i}')} \cap C^{(\underline{i}'')}= \emptyset $$ for pairwise different $\underline{i},\ \underline{i}'\ \underline{i}''\in I$. Now (c) follows from the Bertini theorem \cite{F}, Satz 5.2, saying that for sufficiently general divisors $D_j^{(i)}$ \begin{gather*} Y = D^{(i_1)}_1 \cap \cdots \cap \widehat{D^{(i_{\nu})}_\nu} \cap \cdots \cap D^{(i_{n-1})}_{n-1}\cap X_{\rm reg}\\ C^{(\underline{i})}=Y \cap D_\nu^{(i_\nu)} \mbox{ \ \ \ and \ \ \ }C^{(\underline{i}')}=Y \cap D_\nu^{(i'_\nu)} \end{gather*} are non-singular and that $C^{(\underline{i})}$ and $C^{(\underline{i}')}$ meet tranversally on $Y$. \end{proof} Let ${\mathbb A}^{M+1} \subset |{\mathcal L}^N|^{n-1}$ be an affine open subspace containing the point $s_0$ which corresponds to the tuple $\{D_j^{(1)}\cup \cdots \cup D_j^{(N)}\}_{j=1,\ldots,n-1}$, and let $\P^M$ be the projective space parametrizing lines in ${\mathbb A}^{M+1}$, passing through $s_0$. There is a line $S \in \P^M$ such that \begin{points} \item the total space ${\mathcal C}$ of the restriction $$ \begin{CD} {\mathcal C} \>\tau>> X\\ {\mathbb V} f V V\\ S \> \subset >> |{\mathcal L}^N|^{n-1}. \end{CD} $$ of the universal family to $S$ is non-singular in a neighbourhood of each point $x\in C_{\rm sing} \cap X_{\rm reg}$ \item the intersection of $X_{\rm reg}$ with the general fibre of $f:{\mathcal C} \to S$ is non-singular. \end{points} In fact, using the notation from \ref{general} (c), we can choose for a point $x\in C_{\rm sing} \cap X_{\rm reg}$ a line $S$ connecting $s_0$ with a point $(D'_1, \ldots ,D'_{n-1})\in {\mathbb A}^{M+1}$, where $$ D'_j = D_j^{(1)}\cup \cdots \cup D_j^{(N)} \mbox{ \ \ for \ \ } j \neq \nu, $$ where $D'_\nu \cap Y_{\rm reg}$ is non singular, and where $x \not\in D'_\nu$. By this choice, in a neigbourhood of $x$ the restriction of the universal family ${\mathcal C}$ to $S$ is just a fibering of $Y$ over $S$. Hence the condition (i) is valid for the chosen point $x$. However, for each point $x \in C_{\rm sing} \cap X_{\rm reg}$, the condition (i) is an open condition in $\P^M$, and hence for a general line $S$, (i) hold true for all points in $C_{\rm sing} \cap X_{\rm reg}$; clearly the second condition (ii) holds as well. The family $f:{\mathcal C} \to S$ has only finitely many non-reduced fibres and outside of them $U=\tau^{-1}(X_{\rm reg})$ contains only finitely many points, which are singularities of the fibres. Replacing $S$ by an open neighbourhood of $s_0$, we may assume thereby, that for $s\neq s_0$ the fibre $C_s=f^{-1}(s)$ is reduced, that $C_s\cap X_{\rm reg}$ is non-singular and dense in $C_s$ and that $\mu(C_s)=\mu({\mathcal L}^N)$. In particular $U$ is non-singular outside of the points $C_{\rm sing}\cap X_{\rm reg}$, and by condition (i) $U$ is non singular. Moreover, $f|_U : U \to S$ is semi-stable; hence $f|_U$ is a local complete intersection morphism, smooth outside a finite subset of $U$. Let $L$ be a finite extension of the function field $k(S)$ such that the normalization of ${\mathcal C}\times_S {\rm Spec \,}(L)$ is smooth over $L$, and let $S'$ be the normalization of $S$ in $L$. Consider $$ \begin{CDS} {\mathcal C}' \> \sigma >> {\mathcal C}\times_S S' \> \eta' >> {\mathcal C}\\ & \SE E f' E {\mathbb V} pr_2 V V & {\mathbb V} V f V \\ && S' \> \eta >> S \end{CDS} $$ where $\sigma$ denotes the normalization. Since $U\times_S S' \to S'$ is a local complete intersection morphism, smooth outside a finite subset of the domain, $U\times_S S'$ is normal and $\sigma$ restricted to $U'=\sigma^{-1}(U\times_S S')$ is an isomorphism. By construction the general fibre of $f'$ is smooth and ${\mathcal C}'$ is normal. Since for all $s' \in S'$ the fibres $C'_{s'}={f'}^{-1}(s')$ of $f'$ are reduced on the open dense subvariety $U'$, they are reduced everywhere. Note also that ${\mathcal C}-U\to S$ is finite, and hence so is ${\mathcal C}'-U'\to S'$. Let $s', s'_0 \in S'$ be points, with $s_0=\eta(s'_0)$, and with $s=\eta(s')$ in general position. The inequality (\ref{chieq2}) implies that \begin{align*} \mu({\mathcal L}^N) = \mu(C_{s}) &\leq \chi(C'_{s'},{\mathcal O}_{C'_{s'}})- \chi(C_s,{\mathcal O}_{C_s}) \\ &= \chi(C'_{s'_0},{\mathcal O}_{C'_{s'_0}})-\chi(C_{s_0},{\mathcal O}_{C_{s_0}}). \end{align*} Since $C'_{s'_0}\cap U'$ is isomorphic to $C_{s_0}\cap U$ the curve $C'_{s'_o}$ is finite over and birational to $C=C_{s_0}$. Moreover, the fibres $C'_{s'_0}\cap U'$ and $C\cap U$ have the same number $\delta$ of double points. Writing ${C'}^{(\underline{i})}$ for the preimage of ${C}^{(\underline{i})}$ in $C'_{s'_0}$ one obtains \begin{align*} & \chi(C,{\mathcal O}_{C}) + \delta = \sum_{\underline{i}\in I} \chi({C}^{(\underline{i})},{\mathcal O}_{{C}^{(\underline{i})}}),\\ & \chi(C'_{s'_0},{\mathcal O}_{C'_{s'_0}})+ \delta = \sum_{\underline{i}\in I} \chi({C'}^{(\underline{i})},{\mathcal O}_{{C'}^{(\underline{i})}})\\ \mbox{and \ \ \ } &\mu({\mathcal L}^N)\leq \sum_{\underline{i}\in I}(\chi({C'}^{(\underline{i})},{\mathcal O}_{{C'}^{(\underline{i})}})- \chi({C}^{(\underline{i})},{\mathcal O}_{{C}^{(\underline{i})}}) ). \end{align*} Finally, ${C'}^{(\underline{i})}$ is finite over and birational to ${C}^{(\underline{i})}$, thus it is dominated by the normalization of ${C}^{(\underline{i})}$, and (\ref{chieq2}) implies \begin{align*} \sum_{\underline{i}\in I}(\chi({C'}^{(\underline{i})},{\mathcal O}_{{C'}^{(\underline{i})}})- \chi({C}^{(\underline{i})},{\mathcal O}_{{C}^{(\underline{i})}})) &\leq \sum_{\underline{i}\in I}(\mu({C}^{(\underline{i})}) + r({C}^{(\underline{i})}) -1)\\ &=N^{n-1}\cdot (\mu({\mathcal L}) +r({\mathcal L}) -1). \end{align*} \end{proof} Replacing ${\mathcal L}$ by its $N$-th power one obtains by lemma \ref{muinequality} ample invertible sheaves on $X^{(n)}$ with many more linearly independent sections than $\mu({\mathcal L})$. For example, if $X_1, \ldots ,X_r$ are the irreducible components of $X^{(n)}$, then $$ {\rm image}\,( H^0(X^{(n)},{\mathcal L}^N) \to H^0(X_i,{\mathcal L}^N|_{X_i}))= H^0(X_i,{\mathcal L}^N|_{X_i}), $$ for sufficiently large $N$, and its dimension i bounded below by a non-zero multiple of $N^n$, whereas by \ref{muinequality}, $\mu({\mathcal L}^N)$ is bounded above by $(\mu({\mathcal L})+r({\mathcal L})-1)\cdot N^{n-1}$. One obtains: \begin{cor}\label{mu2} There exists a very ample sheaf ${\mathcal L}$ on $X^{(n)}$ with $$ \dim_k ({\rm image}\,( H^0(X^{(n)},{\mathcal L}) \to H^0(X_i,{\mathcal L}|_{X_i}))) \geq 2\cdot \mu({\mathcal L})+r+2, $$ for $i=1, \ldots, r$. \end{cor} Over a field $k$ of positive characteristic we will need a stronger technical condition. Recall that $\Pi_{X_{\rm reg}} = \bigcup_{i=1}^r (U_i\times U_i)$, where $U_i=X_i \cap X_{\rm reg}$ are the irreducible components of $X_{\rm reg}$. \begin{ass}\label{dominant} Let $Z \subset S^d(\Pi_{X_{\rm reg}})\times S^d(\Pi_{X_{\rm reg}})\times |{\mathcal L}|^{n-1}_0$ be the incidence variety of points \begin{multline*} (((x_1,x'_1), \ldots ,(x_d,x'_d)),((x_{d+1},x'_{d+1}), \ldots ,(x_{2d},x'_{2d})), (D_1, \ldots ,D_{n-1}))\\ \in S^d(\Pi_{X_{\rm reg}})\times S^d(\Pi_{X_{\rm reg}})\times |{\mathcal L}|^{n-1}_0 \end{multline*} with $x_1, \ldots, x_{2d},x'_1, \ldots, x'_{2d} \in C=D_1 \cap \cdots \cap D_{n-1}.$ Then the projection $$ pr_{12}'=pr_{12}|_{Z}: Z \>>> S^d(\Pi_{X_{\rm reg}})\times S^d(\Pi_{X_{\rm reg}}) $$ is dominant. \end{ass} \begin{prop}\label{mu3} There exists a very ample sheaf ${\mathcal L}$ on $X^{(n)}$ which satisfies the assumption \ref{dominant}, for all $d \leq \mu({\mathcal L})$. \end{prop} \begin{proof} Let ${\mathcal I}_i$ be the ideal sheaf of $\bigcup_{j\neq i} X_j$ on $X^{(n)}$. In particular ${\mathcal I}_i|_{X_j}$ is zero, for $j\neq i$. Hence if ${\mathcal F}$ is a torsion-free coherent sheaf on $X^{(n)}$ then \begin{align*} &H^0(X_i, {\mathcal F} \otimes {\mathcal I}_i|_{X_i}/_{\rm torsion})=H^0(X^{(n)},{\mathcal F}\otimes {\mathcal I}_i/_{\rm torsion})\\ \mbox{and \ \ \ }&\bigoplus_{i=1}^r H^0(X_i, {\mathcal F} \otimes {\mathcal I}_i|_{X_i}/_{\rm torsion}) \subset H^0(X^{(n)},{\mathcal F}). \end{align*} \begin{claim}\label{mu4} There exists a very ample invertible sheaf ${\mathcal L}$ on $X^{(n)}$ such that \begin{equation}\label{mueq} \dim_k({\rm image}\,(H^0(X_i, {\mathcal L} \otimes {\mathcal I}_i|_{X_i}) \to H^0(C, {\mathcal L}|_C))) \geq 4\cdot \mu({\mathcal L}), \end{equation} for $i = 1, \ldots , r$, and for all $C \in |{\mathcal L}|^{n-1}_0$. \end{claim} \begin{proof} Given a very ample invertible sheaf ${\mathcal L}$ it suffices to find a lower bound for the dimension of the image of the composite map $$ \tau: H^0(X_i, {\mathcal L}^N \otimes {\mathcal I}_i|_{X_i}) \>>> H^0(X^{(n)},{\mathcal L}^N) \>>> H^0(C, {\mathcal L}^N|_C) $$ which is independent of $C\in |{\mathcal L}^N|^{n-1}_0$ and grows like $N^n$. If ${\mathcal J}_C$ denotes the ideal sheaf of $C$ on $X^{(n)}$, then $$ \ker(\tau)= H^0\left(X_i, {\mathcal L}^N \otimes {\mathcal J}_C \otimes {\mathcal I}_i|_{X_i}/_{({\rm torsion})}\right) \subset H^0(X^{(n)}, {\mathcal L}^N \otimes {\mathcal J}_C). $$ Hence it is sufficient to give an upper bound for $\dim(H^0(X^{(n)}, {\mathcal L}^N \otimes {\mathcal J}_C))$ by some polynomial in $N$ of degree $n-1$, independent of $C$. For $j < n$ the dimension of $H^{j}(X^{(n)}, {\mathcal L}^{-N})$ is bounded by a polynomial of degree $n-2$. In fact, $X^{(n)}$ is a subscheme of $\P^M=\P(H^0(X^{(n)}, {\mathcal L}))$ and by \cite{Ha}, III.7.1 and III.6.9 one has, for $N$ sufficiently large, \begin{align*} H^j(X^{(n)}, {\mathcal L}^{- N}) &\cong {\rm Ext}^{M-j}({\mathcal O}_{X^{(n)}}\otimes {\mathcal L}^{-N},\omega_{\P^M}) \\ &\cong H^0(\P^M,\sext^{M-j}({\mathcal O}_{X^{(n)}},\omega_{\P^M}\otimes {\mathcal O}_{\P^M}(N)))\\ &\cong H^0(\P^M,\sext^{M-j}({\mathcal O}_{X^{(n)}},\omega_{\P^M})\otimes {\mathcal O}_{\P^M}(N)). \end{align*} Since $X$ is Cohen-Macaulay outside of a subscheme $T$ of codimension $2$, the support of $\sext^{M-j}({\mathcal O}_{X^{(n)}},\omega_{\P^M})$ lies in $T$ for $M-j >M-n$. The curve $C$ being a complete intersection of divisors in $|{\mathcal L}^N|$, a resolution of the ideal sheaf ${\mathcal J}_C$ on $X^{(n)}$ is given by the Koszul complex $$ 0 \to {\mathcal L}^{-(n-1)N}=\bigwedge^{n-1}\bigl( \soplus{n-1} {\mathcal L}^{-N} \bigr) \to \ \ \cdots \ \ \to \bigwedge^{2}\bigl( \soplus{n-1} {\mathcal L}^{-N} \bigr) \to \soplus{n-1} {\mathcal L}^{-N} \to {\mathcal J}_C \to 0. $$ Therefore $\dim(H^0(X^{(n)}, {\mathcal L}^N \otimes {\mathcal J}_C))$ is bounded from above by $$\sum_{j=0}^{n-2} \dim_k \bigl( H^j\bigl(X^{(n)},{\mathcal L}^N\otimes \bigwedge^{j+1}\bigl( \soplus{n-1} {\mathcal L}^{-N} \bigr)\bigr)=\sum_{j=0}^{n-2} \dim_k \bigl( H^j\bigl(X^{(n)},{\mathcal L}^{-jN}\bigr)\bigr) {\binom{n-1}{j+1}}. $$ \end{proof} Let ${\mathcal L}$ be a very ample invertible sheaf on $X^{(n)}$ which satisfies the inequality (\ref{mueq}) in \ref{mu4}. We fix some curve $C\in |{\mathcal L}|^{n-1}_0$ and some natural number $d \leq \mu({\mathcal L})$. Each irreducible component of $S^d(\Pi_{X_{\rm reg}})\times S^d(\Pi_{X_{\rm reg}})$ is of the form $$ S_{\underline{d}}=(S^{d_1}(U_1 \times U_1) \times \cdots \times S^{d_r}(U_r \times U_r)) \times (S^{d_{r+1}}(U_1 \times U_1) \times \cdots \times S^{d_{2r}}(U_r \times U_r)), $$ for some tuple $\underline{d}=(d_1, \ldots d_{2r})$ of non-negative integers with $$ d_1 + \cdots + d_r = d_{r+1} + \cdots + d_{2r} =d. $$ Given such a tuple ${\underline{d}}$, we claim that there are (pairwise distinct) points $x_1, \ldots x_{2d},$ $x'_1 \ldots x'_{2d}$, with $x_\nu,x'_\nu \in C \cap U_i$, for $$ \sum_{j=1}^{i-1} d_j < \nu \leq \sum_{j=1}^i d_j \mbox{ \ \ \ and for \ \ \ } \sum_{j=1}^{r+i-1} d_j < \nu \leq \sum_{j=1}^{r+i} d_j, $$ such that the restriction map \begin{equation}\label{restr} H^0(X^{(n)},{\mathcal L}) \>>> \bigoplus_{i=1}^{2d} k_{x_i}\oplus k_{x'_i} \end{equation} is surjective. In fact, by the inequality (\ref{mueq}) the dimension of the image of $$ H^0(X_i, {\mathcal L} \otimes {\mathcal I}_i|_{X_i}) \>>> H^0(C\cap X_i, {\mathcal L} \otimes {\mathcal I}_i|_{C\cap X_i}) \>>> H^0(C, {\mathcal L}|_C))) $$ is at least $4\cdot \mu({\mathcal L}) \geq 4 \cdot d \geq 2 \cdot (d_i + d_{r+i})$ and for sufficiently general points $x_1, \ldots x_{2d}, x'_1 \ldots x'_{2d} \in C$ the composite $$ \bigoplus_{i=1}^rH^0(X_i,{\mathcal L}\otimes {\mathcal I}_i|_{X_i}) \subset H^0(X^{(n)},{\mathcal L}) \>>> \bigoplus_{i=1}^{2d} k_{x_i} \oplus k_{x'_i} $$ is surjective. By construction $$ w: =(((x_1,x'_1), \ldots ,(x_d,x'_d)),((x_{d+1},x'_{d+1}), \ldots ,(x_{2d},x'_{2d}))) \in S_{\underline{d}}. $$ Let $V$ denote the subspace of divisors $D \in |{\mathcal L}|$ with $$ x_1, \ldots ,x_{2d},x'_1,\ldots,x'_{2d} \in D. $$ The fibre ${pr'}^{-1}_{12}(w)$ of the morphism $pr_{12}': Z \to S^d(\Pi_{X_{\rm reg}})\times S^d(\Pi_{X_{\rm reg}})$ is the intersection of $V^{n-1}$ with $|{\mathcal L}|^{n-1}_0$. In particular, since $(w,C) \in Z$, this intersection is non-empty. If $\delta = \dim(|{\mathcal L}|)$, the surjectivity of the restriction map (\ref{restr}) implies that $\dim(V)= \delta - 4 \cdot d$ and $\dim({pr'}_{12}^{-1}(w)) = (n-1)\cdot (\delta - 4 \cdot d).$ The fibres of $pr_3':Z \to |{\mathcal L}|^{n-1}_0$ are equidimensional of dimension $4\cdot d$ and hence $Z$ is equidimensional of dimension $(n-1)\cdot \delta + 4\cdot d$. Therefore the dimension of $pr_{12}'(Z) \cap S_{\underline{d}}$ can not be smaller than $$ (n-1)\cdot \delta + 4\cdot d - (n-1)\cdot (\delta - 4 \cdot d)= n \cdot 4 \cdot d = \dim(S_{\underline{d}}). $$ \end{proof} \section{The algebraic construction of $A^n(X)$} Let $X$ be a projective variety of dimension $n$, defined over an algebraically closed field $k$. As a first step towards the construction of $A^n(X)$ we need to bound the dimension of the image of a regular homomorphism $$\phi:CH^n(X)_{\deg 0} \>>> G $$ to a smooth connected commutative algebraic group $G$. By the theorem of Chevalley and Rosenlicht (theorems 1 and 2 in \cite{BLR}, 9.2) there exists a unique smooth linear subgroup $L$ of $G$ such that $G/L=A$ is an abelian variety. In addition, $L$ is canonically isomorphic to a product of a unipotent group and a torus. Let us write $$ 0\>>> L \>>> G \> \delta >> A \>>> 0 $$ for the extension. \begin{lemma}\label{subgroups} There exists a unique smooth connected algebraic subgroup $H$ of $G$, with $\delta(H) = A$, such that every smooth connected algebraic subgroup $J$ of $G$ with $\delta(J)=A$ contains $H$. Moreover, the quotient group $G/H$ is linear. \end{lemma} \begin{proof} Given a smooth algebraic subgroup $J$ of $G$, one has the commutative diagram of exact sequences $$ \begin{CD} \noarr 0 \noarr 0 \noarr 0 \\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V V\\ 0 \>>> L\cap J \>>> J \>>> \delta(J) \>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V V\\ 0 \>>> L \>>> G \>>> A \>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V V\\ 0 \>>> L/(L\cap J) \>>> G/J \>>> A/\delta(J) \>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} V V {\mathbb V} V V\\ \noarr 0 \noarr 0 \noarr 0 \end{CD} $$ Since $A/\delta(J)$ is an abelian variety and $L/(L\cap J)$ a linear algebraic group, $\delta(J)=A$ if and only if $G/J$ is linear. Observe further, that $\delta(J) = A$ if and only if $\delta(J') = A$ for the connected component $J'$ of $J$ containing the identity. Choose $H$ to be any smooth connected algebraic subgroup of $G$ with $\delta(H)=A$ and such that $\delta(H') \neq A$ for all proper algebraic subgroups $H'$ of $H$. For $J$ as in \ref{subgroups} consider the commutative diagram of exact sequences $$ \begin{CD} 0 \>>> J\cap H \>>> G \>>> G/(J\cap H) \>>> 0\\ \noarr {\mathbb V} V V {\mathbb V} \Delta V V {\mathbb V} \iota V V\\ 0 \>>> J\oplus H \>>> G\oplus G \>>> G/J \oplus G/H \>>> 0 \end{CD} $$ where $\Delta$ is the diagonal embedding. Since $J\cap H = \Delta^{-1}(J\oplus H)$ the morphism $\iota$ is injective on closed points, and hence $G/(J\cap H)$ is a linear algebraic group. By the choice of $H$ one obtains $J\cap H = H$. \end{proof} Recall that $X$ has $n$-dimensional irreducible components $X_1, \ldots ,X_r$, whose union is denoted $X^{(n)}$, and $U_i=X_{reg}\cap X_i$. Also $X^{<n}$ is the union of the lower dimensional components of $X$. \begin{prop}\label{bound} Let ${\mathcal L}$ be a very ample invertible sheaf on $X^{(n)}$ which satisfies the assumption \ref{dominant}. Let $g=\dim_k(H^1(C,{\mathcal O}_C))$, for $C \in |{\mathcal L}|^{n-1}_0$. Let $\phi:CH^n(X)_{\deg 0} \to G$ be a surjective regular homomorphism to a smooth connected commutative algebraic group $G$. Then the induced morphism (see \ref{difference}) $$ \pi^{(-)}: S^{g+\nu\cdot \mu({\mathcal L})}(\Pi_{X_{\rm reg}}) \>>> CH^n(X)_{\deg 0} \>\phi >> G $$ is dominant, for $\nu > 0$, and surjective, for $\nu > 1$. In particular the dimension of $G$ is bounded by $2\cdot n \cdot (g+\mu({\mathcal L}))$. \end{prop} Probably the bound for the dimension of $G$ is far from being optimal. We will indicate in \ref{bound2} how to obtain $\dim (G) \leq g$ in characteristic zero, under a weaker assumption on ${\mathcal L}$. \begin{proof}[Proof of \ref{bound}] Let again $L$ be the largest smooth linear algebraic subgroup and $\delta:G \mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} A=G/L$ the projective quotient group. Recall that $|{\mathcal L}|^{n-1}_0$ denotes the set of tuples $(D_1, \ldots , D_{n-1})$ of divisors in the linear system $|{\mathcal L}|$ for which $C=D_1\cap \cdots \cap D_{n-1}$ is a reduced complete intersection curve (in $X^{(n)}$), $C\cap X^{<n}=\emptyset$, and $C\cap X_{\rm reg}$ non-singular and dense in $C$. \begin{claim}\label{surjective} There exists an open dense subscheme $S \subset |{\mathcal L}|^{n-1}_0$ such that $$ {\rm Pic}^0(C) \>\psi >> G \> \delta >> A $$ is surjective and such that the dimension of ${\rm image}\, (\psi: {\rm Pic}^0(C) \to G)$ is constant, for $C \in S$. \end{claim} \begin{proof} Returning to the notation introduced in \ref{linearsystem} let $S\subset |{\mathcal L}|^{n-1}_0$ be an open subvariety, and let $$ \begin{CD} {\mathcal C} \> \sigma >> X\\ {\mathbb V} f V V\\ S \end{CD} $$ denote the restriction of the universal complete intersection to $S$. The smooth locus of $f$ is ${\mathcal C}_{{\rm sm}}=\sigma^{-1}(X_{\rm reg})$ and ${\mathcal C}_{\rm sm}$ is dominant over $X_{\rm reg}$. Let $S' \to S$ be the finite morphism, and let $W' \subset S^g({\mathcal C}_{\rm sm}\times_S S' /S')$ be the open connected subscheeme considered in lemma \ref{relpic2}. Replacing $S$ and $S'$ by some open subschemes, and $S'$ by a branched cover if necessary, we may assume that there exists a section $\epsilon'$ of $W' \to S'$. By \ref{relpic2} \begin{align*} W' &\>>> {\rm Pic}({\mathcal C}/S)\times_S S'= {\rm Pic}({\mathcal C}\times_S S'/S')\\ w' &\longmapsto \vartheta_{W'}(w') - \vartheta_{W'}(\epsilon'(pr_2(w'))) \end{align*} is an open embedding. On the other hand, one has a morphism of schemes $$ h: W' \>>> S^g({\mathcal C}_{\rm sm}\times_S S' /S') \>>> S^g(X_{\rm reg}), $$ and the image of the connected scheme $W'$ lies in some connected component, say $S_{\underline{g}}=S^{g_1}(U_1) \times \cdots \times S^{g_r}(U_r).$ Since $\phi: CH^n(X)_{\deg 0} \to G$ is regular, the composite $$ h^{(-)}:W'\times_{S'} W' \>>> S_{\underline{g}}\times S_{\underline{g}} \> \theta >> G $$ is a morphism, where $$ \theta(\underline{x},\underline{x}')= \phi\bigl( \sum_{i=1}^g \gamma(x_i) - \sum_{i=1}^g \gamma(x'_i)\bigr) . $$ The morphism $h^{(-)}$ induces $S'$ morphisms $$ h_{S'}^{(-)}: W'\times _{S'} W' \>>> G\times S' \mbox{ \ \ and \ \ } h_{S'}^{(-)}\circ\delta:W'\times_{S'} W' \>>> A \times S'. $$ Let $W_G$ and $W_A$ be locally closed subschemes of the images $$ h_{S'}^{(-)}(W'\times_{S'} W')\mbox{ \ \ and \ \ } h_{S'}^{(-)}\circ\delta(W'\times_{S'} W') $$ respectively, dense in the closure of the images. Choosing $S'$ and $S$ small enough, one may assume that $S' \to S$ is surjective and that $W_G$ and $W_A$ are both equidimensional over $S'$. For $C\in S$ choose a point $s' \in S'$ mapping to $C \in S$ and let $W'_{s'}$ denote the fibre of $W'$ over $s'$. Then the image of $W'_{s'} \times W'_{s'}$ in ${\rm Pic}^0(C)$ is dense and thereby $\dim( \psi({\rm Pic}^0(C)))$ and $\dim(\delta(\psi({\rm Pic}^0(C)))) = d'$ are both constant on $S$. Asume that $d' < \dim(A)$. The closure of $\delta(h^{(-)}(W'_C\times W'_C))$ is the image of ${\rm Pic}^0(C)$, hence $\delta(h^{(-)}(W'_C\times W'_C))$ lies in some abelian subvariety $B$ of $A$ of dimension $d'< \dim(A)$. Since $S'$ and $W'$ are connected, and since an abelian variety $A$ does not contain non-trivial families of abelian subvarieties, $B$ is independent of the curve $C$ chosen. ${\mathcal C}_{\rm sm}$ being dominant over $X_{\rm reg}$ this implies that the image $\delta\phi(CH^n(X)_{\deg 0})$ lies in $B$, contradicting the assumptions made. \end{proof} In general, a commutative algebraic group $G$ can contain non-trivial families of subgroups and the argument used above does not extend to $G$ instead of $A$. Let $H \subset G$ be the smallest connected algebraic subgroup with $\delta(H) = A$, as constructed in \ref{subgroups}. By \ref{surjective} and by the universal property in \ref{subgroups}, for $C \in S$ the image of $\psi({\rm Pic}^0(C))$ contains $H$. By \ref{generators} the image of $S^g(\Pi_{C_{\rm reg}})$ in $G$ is $\psi({\rm Pic}^0(C))$ and hence $H$ is contained in the image of $S^g(\Pi_{X_{\rm reg}})$. In order to show that $$ \pi^{(-)}: S^{g+\mu({\mathcal L})}(\Pi_{X_{\rm reg}}) \>>> CH^n(X)_{\deg 0} \>\phi >> G $$ is dominant, it suffices to verify that the image $Y_0$ of the composite $$ \tau^{(-)}: S^{d}(\Pi_{X_{\rm reg}}) \>>> CH^n(X)_{\deg 0} \>\phi >> G \>>> G/H $$ is dense, for some $d \leq \mu({\mathcal L})$. Applying claim \ref{surjective} to $G/H$ instead of $G$ one finds a non-empty open subscheme $S\subset |{\mathcal L}|^{n-1}_0$ such that the dimension $d$ of $\psi'({\rm Pic}^0(C))$ is constant on $S$, where $\psi':{\rm Pic}^0(C) \to G/H$ is the natural map (see \ref{pic}). Since $G/H$ is a linear algebraic group, we must have $d\leq \mu({\mathcal L})$, and choosing $S$ small enough, we may assume that \begin{equation}\label{bd} d=\dim(\psi'({\rm Pic}^0(C))) \leq \mu(C) = \mu({\mathcal L}), \mbox{ \ \ \ \ for all \ $C \in S$.} \end{equation} Since $Y_0$ generates the group $G/H$, it is dense in $G/H$ if and only if its closure $Y$ is a group. By assumption the image of the incidence variety $$ Z \> pr'_{12} >>S^{d}(\Pi_{X_{\rm reg}})\times S^{d}(\Pi_{X_{\rm reg}}) \>\tau^{(-)}\times \tau^{(-)}>> Y \times Y $$ defined in \ref{dominant} contains some open dense subscheme $T$. By definition, for each $t \in T$ there exist divisors $D_1, \ldots, D_{n-1}$ with $C=D_1\cap \cdots \cap D_{n-1} \in S$ and with $$ t \in {\rm image}\,( S^{d}(\Pi_{B})\times S^{d}(\Pi_{B}) \>{\vartheta}^{(-)}\times {\vartheta}^{(-)}>> Y \times Y) $$ for $B = C\cap X_{\rm reg}$ and for the induced map $\vartheta^{(-)}$ from $\Pi_B$ to $G/H$. By \ref{generators} $\psi'({\rm Pic}^0(C))={\vartheta}^{(-)}(S^d(\Pi_B))\subset Y$. Since $\psi'({\rm Pic}^0(C))$ is an algebraic subgroup of $G/H$, the image of $t$ under the morphism $$ {\rm diff}:G/H \times G/H \>>> G/H \mbox{ \ \ \ with \ \ \ } (g,g') \longmapsto g-g' $$ is contained in $\psi'({\rm Pic}^0(C))$, hence in $Y$. Thereby $T$ is a subset of ${\rm diff}^{-1}(Y)$, and the same is true of its closure $Y \times Y$. One obtains that ${\rm diff} (Y \times Y) \subset Y$ and $Y$ is a subgroup of $G/H$. Since $Y_0$ is dense in $G/H$, by lemma \ref{generators} (ii) the image of $S^{2d}(\Pi_{X_{\rm reg}})$ is $G/H$. \end{proof} As indicated already, the proposition \ref{bound} can be improved in characteristic zero. \begin{variant}\label{bound2} Assume that ${\rm char} (k) =0$. Let ${\mathcal L}$ be a very ample invertible sheaf on $X^{(n)}$ with \begin{equation}\label{eqmu2} \dim_k ({\rm image}\,( H^0(X^{(n)},{\mathcal L}) \to H^0(X_i,{\mathcal L}|_{X_i}))) \geq 2\cdot \mu({\mathcal L}) + r + 2, \end{equation} for $i=1,\ldots ,r$. Let $G$ be a smooth connected commutative algebraic group, and let $\phi:CH^n(X)_{\deg 0} \to G$ be a surjective regular homomorphism. Then there exists an open dense subvariety $S \subset |{\mathcal L}|^{n-1}_0$ such that for each $C \in S$ the induced homomorphism (see \ref{pic}) $$ \psi: {\rm Pic}^0(C) \>>> CH^n(X)_{\deg 0} \>\phi >> G $$ is surjective. In particular the dimension of $G$ is bounded by $g=\dim_k H^1(C,{\mathcal O}_C)$. \end{variant} \begin{proof} The first part of the proof is the same as the one for \ref{bound}. In particular we may assume claim \ref{surjective} to hold true. Let $H \subset G$ be the smallest subgroup with $\delta(H) = A$, as constructed in \ref{subgroups}. By \ref{surjective} and by the universal property in \ref{subgroups}, for all $C \in S$ the image of $\psi({\rm Pic}^0(C))$ contains $H$. Hence $\psi:{\rm Pic}^0(C) \>>> G$ is surjective if and only if $$ {\rm Pic}^0(C) \> \psi >> G \>>> G/H $$ is surjective. In order to prove \ref{bound2} we may assume thereby that $G$ is linear and $A=0$. For $C \in S$, let $\gamma_B : B=C\cap X_{\rm reg} \to CH^n(X)$ denote the natural map and let $\Gamma_B$ be the image of the composite $$ \vartheta^{(-)}:\Pi_B \> \gamma_B^{(-)} >> CH^n(X)_{\deg 0} \> \phi >> G. $$ For any subset $M \subset G$ we will denote by $G(M)$ the smallest algebraic subgroup of $G$ which contains $M$. If $M$ contains a point of infinite order, then $\dim(G(M)) > 0$. In characteristic zero the converse holds true, as well. In fact, if $\dim(G(M)) > 0$ then $G(M)$ contains a subgroup isomorphic either to ${\mathbb G}_a$ or to ${\mathbb G}_m$. In characteristic zero, both contain points of infinite order. Hence if the dimension of $G(\Gamma_B) = \psi({\rm Pic}^0(C))$ is larger than zero, the constructible set $\Gamma_B$ contains a point $\alpha_1$ of infinite order and $\dim(G(\alpha_1)) > 0$. Repeating this for $G/G(\alpha_1,\ldots,\alpha_\nu)$ instead of $G$, we find recursively points $\alpha_1, \ldots, \alpha_d\in \Gamma_B$ with $G(\Gamma_B)=G(\alpha_1,\ldots, \alpha_d).$ Let us choose points $x_1, \ldots, x_d,x'_1, \ldots, x'_d \in B$ with $\alpha_j=\vartheta^{(-)}((x_j,x'_j))$, and moreover, for each component $X_i$ of $X^{(n)}$, choose a base point $q_i \in B\cap X_i$. \begin{claim}\label{dominant2} There exists a closed suscheme $Z\subset S$ such that the restriction $$ \begin{CD} {\mathcal C}'={\mathcal C}\times_S Z \> \sigma'=\sigma|_{{\mathcal C}'} >> X^{(n)}\>\subset>> X\\ {\mathbb V} f'=f|_{{\mathcal C}'} V V\\ Z \>\subset >> S\> \subset >>|{\mathcal L}|^{n-1}_0 \end{CD} $$ of the universal family satisfies: \begin{enumerate} \item[(a)] For each point $z \in Z$ the curve $C_z={f'}^{-1}(z)$ contains the points $$x_1, \ldots, x_d,x'_1, \ldots, x'_d, q_1,\ldots , q_r.$$ \item[(b)] $\sigma': {\mathcal C}' \to X^{(n)}$ is dominant. \end{enumerate} \end{claim} \begin{proof} For $$ V_i=({\rm image}\,(H^0(X^{(n)},{\mathcal L}) \to H^0(X_i,{\mathcal L}|_{X_i}))-0)/k^* \subset |({\mathcal L}|_{X_i})| $$ consider the rational map $\tilde{p}_i:|{\mathcal L}|^{n-1} \to V_i^{n-1}$. Since each $C\in S$ is a complete intersection curve, the restriction $p_i:S\to V_i^{n-1}$ of $\tilde{p}_i$ is a morphism. For $x \in X_i\cap \sigma({\mathcal C})$ the condition ``$x \in C_s$'' defines a multilinear subspace $\Delta_x^i$ of $V_i^{n-1}$ of codimension $n-1$. Let $I_i\subset\{ 1, \ldots,d\}$ denote the set of all the $\nu$ with $x_\nu, x'_\nu \in X_i$. Then the codimension of $$ \Delta^i = \Delta^i_{q_i} \cap \bigcap_{\nu \in I_i}(\Delta^i_{x_\nu} \cap \Delta^i_{x'_\nu}) $$ is at most $(n-1)\cdot(2\cdot \# I_i +1)$. Let ${\mathcal C}^i \to \Delta^i$ be the intersection on $X_i$ of the divisors in $\Delta^i \subset V_i^{n-1}$. Then the general fibre of ${\mathcal C}^i \to X_i$ has dimension at least $$\dim\Delta^i+1-\dim X=\dim V_i^{n-1}+1-n-{\rm codim}\Delta^i$$ $$\geq \dim (V_i^{n-1}) +1 -n - (n-1)\cdot(2\cdot \# I_i +1) \geq (n-1)\cdot(2\cdot(\mu({\mathcal L}) - \# I_i)+r-1). $$ Since some $C \in S$ contains all the points $x_j, x'_j$ and $q_i$, the intersection $$ Z = S\cap \bigcap_{i=1}^r p_i^{-1}(\Delta^i) $$ is non-empty. For the restriction ${\mathcal C}'$ of the universal curve ${\mathcal C}$ to $Z$ the dimension of the general fibre of $\sigma':{\mathcal C}' \to X$ over $X_i$ has dimension larger than or equal to \begin{align*} & (n-1)\cdot(2\cdot(\mu({\mathcal L}) - \# I_i)+r-1)-\sum_{j\neq i} (n-1)\cdot(2\cdot \# I_j +1)\\ &= (n-1)\cdot 2\cdot(\mu({\mathcal L}) - \sum_{j=1}^r \# I_i) =(n-1)\cdot 2\cdot(\mu({\mathcal L}) - d). \end{align*} By the inequality (\ref{bd}) the last expression is larger than or equal to $0$ and $\sigma'$ is dominant. \end{proof} Let $G(C_z)$ denote the image of ${\rm Pic}^0(C_z)$ in $G$. By the choice of $Z$ the intersection $B_z=C_z\cap X_{\rm reg}$ is non-singular and the dimension of $G(C_z)=G(\Gamma_{B_z})$ is the same as the dimension of $G(\Gamma_B)=G(\alpha_1,\ldots, \alpha_d)$. By \ref{dominant2} the points $\alpha_i = \phi(\sigma'(x_i)-\sigma'(x'_i))$ are contained in $\Gamma_{B_z}$, hence $$ G(C_z)=G(\Gamma_{B_z}) = G(\Gamma_B)=G(C) $$ for all $z \in Z$. As $\sigma'$ is dominant, $\sigma'({\mathcal C}')$ contains some $V$, open and dense in $X_{reg}$ (and hence in $X^{(n)}$). For $q\in V\cap X_i$ one finds some $z \in Z$ with $q \in C_z$. By \ref{dominant} $C_z$ contains the chosen base point $q_i$ and $$ \phi(\gamma(q_i) - \gamma(q)) \in G(C_z)=G(C). $$ By \ref{equidim} (i), the points $\gamma(q_i) - \gamma(q)$ (for $q\in V$) generate $CH^n(X)_{\deg 0}$. Since $\phi$ was assumed to be surjective, we obtain $G=G(C)$, as claimed. \end{proof} Using proposition \ref{bound} or its variant \ref{bound2} the construction of $A^n(X)$ proceeds now along the lines of Lang's construction in \cite{La} of the Albanese variety of a smooth projective variety. \begin{thm}\label{existence} There exists a smooth connected commutative algebraic group $A^n(X)$ and a surjective regular homomorphism $\varphi: CH^n(X)_{\deg 0} \to A^n(X)$ satisfying the following universal property: For any smooth commutative algebraic group $G$ and for any regular homomorphism $\phi:CH^n(X)_{\deg 0}\to G$ there exists a unique homomorphism $h:A^n(X)\to G$ of algebraic groups with $\phi=h\circ\varphi$. Moreover, if $k \subset K$ is an extension of algebraically closed fields, then $$ A^n(X\times_kK)=A^n(X)\times_kK. $$ \end{thm} \begin{proof} By lemma \ref{reg_sur} it is sufficient to consider connected groups $G$, and surjective regular homomorphisms $\phi$. By \ref{mu3} there exists a very ample invertible sheaf ${\mathcal L}$ which satisfies the assumption \ref{dominant} and we can apply \ref{bound}. (As we have seen in \ref{mu2} the inequality (\ref{eqmu2}) in \ref{bound2} holds true for some ${\mathcal L}$, and if ${\rm char}(k)=0$ we can use the variant \ref{bound2}, as well.) Let $g=\dim_k(H^1(C,{\mathcal O}_C))$, for some curve $C\in |{\mathcal L}|^{n-1}_0$ in general position. Then for all regular homomorphisms $\phi:CH^n(X)_{\deg 0}\to G$ to smooth connected commutative algebraic groups $G$ the induced morphism $$ \pi^{(-)}: S^{g+\mu({\mathcal L})}(\Pi_{X_{\rm reg}}) \>\gamma^{(-)}_{g+\mu({\mathcal L})}>> CH^n(X)_{\deg 0} \>\phi >> G $$ has a dense image in $G$. Hence for the product $\Pi$ of all the different connected components of $S^{g+\mu({\mathcal L})}(\Pi_{X_{\rm reg}})$ the induced morphism $\pi' : \Pi \to G$ is dominant and $\pi'$ induces a unique embedding of function fields $k(G) \subset k(\Pi)$. If $\phi_\nu : CH^n(X)_{\deg 0} \to G_\nu$, for $\nu = 1,2$ are two surjective regular homomorphisms to smooth connected commutative algebraic groups, then $$ \phi_3:CH^n(X)_{\deg 0} \>>> G_1\times G_2 $$ is regular. Let $G_3$ be the image of $\phi_3$. Then $\phi_\nu$ factors through the regular homomorphism $\phi_3:CH^n(X)_{\deg 0} \to G_3$ and $k(G_\nu) \subset k(G_3) \subset k(\Pi)$, for $\nu = 1,2$. Hence among the smooth connected commutative algebraic groups $G$ with a regular surjective homomorphisms from $\phi:CH^n(X)_{\deg 0}\to G$, there is one, $A^n(X)$, for which the subfield $k(A^n(X))$ is maximal in $k(\Pi)$ and $A^n(X)$ dominates all the other $G$ in a unique way. It remains to show that $A^n(X)$ satisfies base-change for algebraically closed fields. Let us write $Z_K = Z\times_kK$, for a variety $Z$ defined over $k$. We first show: \begin{claim}\label{basechange1} Let $K\supset k$ be an algebraically closed extension field of $k$. The cycle map $\varkappa^{(-)}_K:(\Pi_{X_{\rm reg}})_K \to A^n(X)_K$ factors through a surjective homomorphism $u_{K,k}:A^n(X_K) \to A^n(X)_K$ of algebraic groups. \end{claim} \begin{proof} Let $U=X_{reg}$, and let $(C', \iota)$ be an admissible pair defined over $K$, with $B =\iota^{-1}(U_K)_{{\rm reg}}$. Choose a rational function $f \in R(C',X_K)$ such that $$ {\rm div} f = \sum a_i - \sum b_i $$ for $p =(a_1,b_1,\ldots,a_m,b_m)\in S^m(\Pi_B)(K)$. Choose a smooth $k$ variety $S$ with $k(S)\subset K$, such that $C'$, $B$, $p$, $a_i$, $b_i$, $f$ come by base-change from $k(S)$ to $K$ from $${\mathcal C}' \to S, \ \ {\mathcal B} \to S, \ \ \pi: S \to S^m(\Pi_{{\mathcal B}/S}), \ \ \alpha_i, \beta_i: S \to {\mathcal B}, \ \ \varphi \in k({\mathcal B})^{\times} $$ with ${\rm div} \varphi = \sum \alpha_i - \sum \beta_i$. Since $f\in R(C',X_K)$, we can replace $S$ by a dense open subscheme, so that we can arrange that for each $s\in S(k)$, if we specialize to ${\mathcal C}'_s = {\mathcal C} _S \times s$, then $\pi(s)$ maps to zero in ${\rm Pic}^0({\mathcal C}'_s)$. As \[S^m(\Pi_{{\mathcal B}_s}) \>>> S^m (\Pi_U) \times s \>>> A^n(X) \times s\] factors through ${\rm Pic}^0({\mathcal C}'_s)$, the composite morphism $$ S \> \pi >> S^m(\Pi_{{\mathcal B}/S}) \>>> S^m(\Pi_U)\times S \>>> A^n(X)\times S $$ maps all $k$-points of $S$ to the zero section. Thus it is the zero section, and therefore $S^m(\Pi_U)_K \to A^n(X)_K$ factors through $CH^n(X_K)$, inducing $u_{K,k}$ by lemma \ref{pic}. \end{proof} Since $d_K:= \dim A^n(X_K)$ is bounded by $2n(g + \mu({\mathcal L}))$ (proposition~\ref{bound}), there is an algebraically closed field $K_1$ with $d_{K_1}=d_L$ for all $L\supset K_1$ algebraically closed. Since for any ascending chain $K_i\subset K_{i+1}$ of algebraically closed fields with $K_i \supset K_1$ one has $$ {\rm deg} \ u_{K_i,K_1} \leq {\rm deg} \ u_{K_{i+1},K_1} \leq {\rm deg} \ u_{\cup K_i, K_1}, $$ one concludes that there is an algebraically closed field $E\supset K_1$ such that $u_{E,L}$ is an isomorphism for all algebraically closed fields $L\supset E$. We will make use of the following lemma. \begin{lemma}\label{descent} Let $K$ be a field, $W$, $Y$, $Z$ be geometrically integral $K$-varieties, such that there are $K$-morphisms $\alpha:W\to Y$, $\beta:W\to Z$, such that $\alpha$ has dense image. Then: \begin{enumerate} \item[(i)] there is at most one $K$-morphism $f:Y\to Z$ such that $\beta=f\circ\alpha$ \item[(ii)] suppose that for some extension field $L$ of $K$, there is an $L$-morphism $h:Y_L\to Z_L$ such that $\beta_L=h\circ \gamma_L:W_L\to Z_L$; then there is a $K$-morphism $f:Y\to Z$ as in (i), and we have $h=f_L$. \end{enumerate} \end{lemma} \begin{proof} Let $\Gamma\subset W\times_K Z$ be the graph of $\beta$, and let $\bar{\Gamma}\subset Y\times_K Z$ be the closure of $\alpha\times 1_Z(\Gamma)$. The projection $\bar{\Gamma}\to Y$ has dense image. If there is a $K$-morphism $f:Y\to Z$ as in (i), then $\bar{\Gamma}$ must be its graph, and so there is at most one such morphism, which exists precisely when $\bar{\Gamma}\to Y$ is an isomorphism. Clearly if this is an isomorphism after base change to $L$, it is an isomorphism to begin with. \end{proof} There is a smooth $k$ variety $S$, with $k(S) \subset E$, together with a smooth commutative $S$-group scheme ${\mathcal A} \to S$ with connected fibers, such that $A^n(X_E) = {\mathcal A} \times_S {\rm Spec \,} E$. Choosing $S$ small enough one also has natural surjective $S$-morphisms $\Pi\times S \to {\mathcal A}$ and $u_{S,k}:{\mathcal A} \to A^n(X)\times S$, where $\Pi$ is the irreducible variety constructed in the first part of the proof. Let $F$ be an algebraic closure of the quotient field of $E\otimes_k E$, $p: k(S\times_k S) \hookrightarrow F$ the natural inclusion, and let $p_i^*: k(S)\hookrightarrow k(S\times_k S)$, $i=1,2$ be the inclusions defined by the two projections $p_i:S\times_k S\to S$. Set $q_i= p \circ p_i^*$, and for any $S$-scheme $T$, let $q_i^*T$ be the $F$-scheme obtained by the base change to $F$ determined by $q_i$. The surjective $S$-morphism $\Pi\times S \to {\mathcal A}$ gives rise to the surjections \[\alpha_i':q_i^* (\Pi\times_kS) = \Pi_{F} \>>> q_i^*{\mathcal A}.\] By the assumption on $E$ the two $F$-varieties $q_i^*{\mathcal A}$ are isomorphic via $$ u'= u_{F, E_1} \circ u^{-1}_{F,E_2}: q_1^*A^n(X_E) \>>> q_2^*A^n(X_E). $$ where $E_i\subset F$ are the images of the two embeddings $E\hookrightarrow E\otimes_k E\hookrightarrow F$, $x\mapsto x\otimes 1$ and $x\mapsto 1\otimes x$. By construction, $u'$ satisfies $\alpha_2'=u'\circ\alpha_1'$. Hence by lemma~\ref{descent}, applied to the extension of fields $k(S\times_kS)\hookrightarrow F$, the isomorphism $u'$ comes from an isomorphism $$u: (p_1^*{\mathcal A})_{k(S\times_k S)} \>>> (p_2^*{\mathcal A})_{k(S\times_k S)},$$ Then $u$ in fact extends uniquely to an isomorphism of groups schemes (again denoted $u$) \[u: (p_1^*{\mathcal A})_U \>>> (p_2^*{\mathcal A})_U\] over an open dense subset $U \subset S\times_k S$. Replacing $S$ by some open dense subscheme, we may assume that $p_i:U\to S$ is surjective, for $i=1,2$. Further, if $\alpha_i:\Pi\times U\to (p_i^*{\mathcal A})_U$, $i=1,2$ are the natural surjections, then $\alpha_2=u\circ \alpha_1$. The uniqueness statement in lemma~\ref{descent}~(i) similarly implies that $u$ satisfies the ``cocycle condition'' \[u_{23}\circ u_{12}=u_{13}:\pi_1^*{\mathcal A}\>>>\pi_3^*{\mathcal A}\] on the fibers over the generic point of $S\times_k S\times_k S$, and hence (by continuity) over the open dense subset \[\pi_{12}^{-1}(U)\cap\pi_{23}^{-1}(U)\cap \pi_{13}^{-1}(U)\subset S\times_k S\times_k S.\] Here $\pi_j:S\times_k S\times_k S\to S$ are the 3 projections, and $u_{ij}=\pi_{ij}^*{u}$, for the 3 projections $\pi_{ij}:S\times_k S\times_k S\to S\times_k S$. Given two points $s_i \in S(k)$, one finds a third one $s \in S(k)$ such that $(s_1,s) \in U(k)$ and $(s, s_2) \in U(k)$. The cocycle condition implies that the induced composite isomorphism $$ \theta_{s_1s_2}: {\mathcal A}|_{s_1} \> u|_{(s_1,s)} >> {\mathcal A}|_s \> u|_{(s,s_2)} >> {\mathcal A}|_{s_2}$$ does not depend on the point $s\in S(k)$ chosen. Also $u$ is compatible with the surjective morphisms $\Pi \times U \to p_i^*{\mathcal A}$. We claim that for each closed point $s\in S(k)$, the morphism $\Pi\times s\to {\mathcal A}|_s$ induces a regular homomorphism $CH^n(X)_{\deg 0}\to {\mathcal A}|_s$. Let $(C',\iota)$ be an admissible pair on $X$, defined over $k$, with $B = \iota^{-1}(U)_{{\rm reg}}$. The morphism $(\Pi_B)_E\to A^n(X_E)={\mathcal A}_E$, and the resulting morphism $S^g(\Pi_B)_E\to {\mathcal A}_E$ (with $g:=\dim{\rm Pic}^0(C')$) induces a homomorphism ${\rm Pic}^0(C')_E\to {\mathcal A}_E$, since by the defining property of $A^n(X_E)$, we have a factorization through $CH^n(X_E)_{\deg 0)}$. Since $S^g(\Pi_B)\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$}{\rm Pic}^0(C')$, lemma~\ref{descent} gives a map ${\rm Pic}^0(C') \times_k k(S) \to {\mathcal A}_{k(S)}$, compatible with the maps from $(\Pi_B)_{k(S)}$. This then induces a map ${\rm Pic}^0(C')\times S^0\to{\mathcal A}_{S^0}$ for some open dense subscheme $S^0\subset S$. Choosing a $k$-point $s_1\in S^0(k)$, we get that the map $\Pi_B\times s_1\to {\mathcal A}|_{s_1}$ is compatible with a homomorphism ${\rm Pic}^0(C')\to {\mathcal A}|_{s_1}$. The isomorphism $u$ is compatible with the morphisms $\Pi_B \times U \to p_i^*{\mathcal A}$. Hence, the isomorphisms $\theta_{ss_1}$ are compatible with the maps $\Pi_B\cong \Pi_B\times s\to {\mathcal A}|_s$ and $\Pi_B\cong \Pi_B\times s_1\to {\mathcal A}|_{s_1}$, for all $s\in S(k)$. We deduce that for any $s\in S(k)$, the map $\Pi_B\times s\to {\mathcal A}|_s$ gives rise to a compatible morphism ${\rm Pic}^0(C') \times _k s \to {\mathcal A}|_s$. This implies that there is an induced regular homomorphism $CH^n(X)_{\deg 0}\to {\mathcal A}|_s$ for each $s\in S(k)$. Hence, one obtains morphisms $v_s: A^n(X) \to {\mathcal A}|_s$, verifying $ v_t = \theta_{st} \circ v_s$ for all $s,\ t \in S(k)$. Choosing now $s \in S(k)$, we set $G= {\mathcal A}|_s$, and $ v= v_s$. The surjective morphism $\Pi\times_k S \to {\mathcal A}$ induces a surjection from $\Pi \times_k s$ onto $G$, hence $v$ is surjective. Since the composite $$ u_{S,k}\circ v : A^n(X) \>>> G \>>> A^n(X) $$ is an isomorphism, $v$ is an isomorphism. Thus $u_{S, k}: {\mathcal A} \to A^n(X)\times S$ is an isomorphism when restricted to each $t \in S(k)$, and is hence an isomorphism. By base change to $E$ one finds that $u_{E,k}:A^n(X_E) \to A^n(X)_E$ is an isomorphism. Now if $K \supset k$ is any algebraically closed field, we choose an algebraically closed field $F$ with \begin{gather*} \hspace*{1.6cm}F\supset K \supset k\mbox{ \ \ \ and \ \ \ }F\supset E\supset k, \mbox{ \ \ \ hence}\\ u_{K, k} \otimes {\rm id}_F \circ u_{F,K}= u_{F, k} = u_{E, k} \otimes {\rm id}_F \circ u_{F,E}, \end{gather*} and $u_{K,k}$ is an isomorphism as well. \end{proof} \section{Finite dimensional Chow groups of zero cycles} The definition of finite dimensionality for the Chow group of 0-cycles is a natural generalization of the definition in the non-singular (and normal) case (see \cite{M}, \cite{S}). \begin{defn}\label{fin-dim} $CH^n(X)$ is said to be {\em finite dimensional} if for some $m>0$, the map $$ \gamma_m : S^m(X_{\rm reg}) \>>> CH^n(X)_{\deg 0} $$ (introduced in \ref{difference}) is surjective. \end{defn} One can see that this is also equivalent to the statement that for some integer $m'>0$, depending only on $X$, any element of $CH^n(X)_{\deg 0}$ is represented by a 0-cycle $\sum_{i=1}^r\delta_i$, where for each $i$, the cycle $\delta_i$ is a difference of two effective 0-cycles of degree $m'$ supported in $X_i$. In the proof of the next theorem we will use the notion of a {\em regular map} $f:Z\to CH^n(X)_{\deg 0}$ from a variety $Z$. This is a map of sets such that \begin{points} \item the composition $Z\to CH^n(X)_{\deg 0}\to A^n(X)$ is a morphism \vspace{.05cm} \item there is a surjective morphism $W\to Z$ such that $$W\>>> Z\>{f}>> CH^n(X)_{\deg 0}$$ factors as $W\>{h}>>S^m(X_{\rm reg})\>{\gamma_m}>> CH^n(X)_{\deg 0},$ for some morphism $h$. \end{points} For example, let $C'$ be a reduced Cartier curve in $X$ or, more general, let $(C',\iota)$ be an admissible pair. Then the homomorphism $\eta:{\rm Pic}^0(C')\to CH^n(X)_{\deg 0}$ constructed in lemma~\ref{gysin} is regular. In fact, the first condition holds true by \ref{pic} whereas the second one follows from the dominance of $S^g(C'_{\rm reg}) \to {\rm Pic}^0(C')$, for $g=\dim_k(H^1(C',{\mathcal O}_{C'}))$. Recall that $k$ is called a universal domain, if its trancendence degree over the prime field is uncountable. \begin{thm}\label{finite} Let $X$ be a projective variety of dimension $n$ over a universal domain $k$. Then $CH^n(X)$ is finite dimensional if and only if $$ \varphi:CH^n(X)_{\deg 0}\>>> A^n(X) $$ defines an isomorphism between $CH^n(X)_{\deg 0}$ and {\em (the closed points of)} $A^n(X)$. \end{thm} \noindent{\bf Proof}\ \\ Let us write $U=X_{\rm reg}$. By lemma \ref{generators} (ii) the composite \[S^m(U)\>>> CH^n(X)_{\deg 0}\>>> A^n(X)\] is always surjective for $m=2\cdot \dim(A^n(X))$. Hence, if $CH^n(X)_{\deg 0}\to A^n(X)$ is an isomorphism, then $CH^n(X)$ is finite dimensional. So the main thrust of the theorem is the converse, that if $CH^n(X)_{\deg 0}$ is finite dimensional, then $CH^n(X)_{\deg 0}\to A^n(X)$ is an isomorphism. We imitate Roitman's proof of this result in the non-singular case, and the analogous argument for the normal case in \cite{S}; however there are extra refinements needed here, particularly in characteristic $p>0$. First, we note that by \cite{LW}, proposition 4.2, the ``graphs of rational equivalence'' \[\Gamma_{r,s}=S^r(U)\times_{CH^n(X)_{\deg 0}}S^s(U)\] decompose as a countable union of locally closed subvarieties, for each $r,s$, and over a universal domain such a decomposition is unique. This immediately implies that if $f_j:Z_j\to CH^n(X)_{\deg 0}$, $j=1,2$ are regular maps, then \[Z_1\times_{CH^n(X)_{\deg 0}}Z_2 = \{(z_1,z_2)\in Z_1\times Z_2; \ f_1(z_1)=f_2(z_2)\}\] is a countable union of locally closed subvarieties of $Z_1\times Z_2$. Now arguing as in \cite{S}, lemma~(1.3), we first see that if $G$ is a smooth connected commutative algebraic group, and $f:G\to CH^n(X)_{\deg 0}$ is any regular map which is a group homomorphism, then there is a well-defined connected component of the identity $G^0\subset \ker f$, which is a connected algebraic subgroup of $G$, and which has countable index in $\ker f$. Then the induced homomorphism $$G/G^0\>>> CH^n(X)_{\deg 0}$$ has a countable kernel. Hence, for any such homomorphism $G\to CH^n(X)_{\deg 0}$, we can define the {\em dimension of the image of $G$} to be the dimension of $G/G^0$. Next, notice that if $G_1\to CH^n(X)_{\deg 0}$ and $G_2\to CH^n(X)_{\deg 0}$ are two regular homomorphisms from smooth connected commutative algebraic groups $G_i$ such that ${\rm image}\,G_1$ is properly contained in ${\rm image}\, G_2$, then in fact \[\dim\, {\rm image}\,G_1< \dim\, {\rm image}\,G_2.\] Indeed, we may assume the maps $G_i\to CH^n(X)_{\deg 0}$ have countable kernel, so that we wish to assert that $\dim G_1<\dim G_2$. Now $G_3=G_1\times_{CH^n(X)_{\deg 0}}G_2$ is a subgroup of $G_1\times G_2$ which is a countable union of locally closed subvarieties, and hence has a connected component of the identity which is a connected algebraic group, say $H$. Then $H\to G_i$ are homomorphisms of algebraic groups with countable, hence finite kernels, such that $H\to G_1$ is surjective, and the image of $H$ in $G_2$ is a strictly smaller subgroup. Thus $\dim G_1=\dim H<\dim G_2$. Now suppose $\gamma_m$ is surjective. We claim that for any homomorphism $$ G\>>> CH^n(X)_{\deg 0} $$ as above, with countable kernel, we have $\dim G\leq \dim S^m(U).$ Indeed, \[G\times_{CH^n(X)_{\deg 0}}S^m(U)\] is a countable union of subvarieties of $G\times S^m(U)$ which projects onto $G$, and maps to $S^m(U)$ with countable fibres. Hence some irreducible component of this fibre product dominates $G$ under the projection, and maps to $S^m(U)$ with finite fibres. We now claim that we can find a finite number of reduced complete intersection curves $C_1,\ldots,C_s$ such that the induced homomorphism from $\oplus {\rm Pic}^0(C_j)$ to $CH^n(X)_{\deg 0}$ is surjective. Indeed, given a finite collection of such curves, if $$P=\bigoplus {\rm Pic}^0(C_j)\>>> CH^n(X)_{\deg 0}$$ is not surjective, we can find a curve $C$ of the same sort such that $${\rm image}\,( {\rm Pic}^0(C)\>>> CH^n(X)_{\deg 0})$$ is not contained in the image of $P$. Then the induced map $$P\times {\rm Pic}^0(C)\>>> CH^n(X)_{\deg 0}$$ has strictly larger dimensional image than that of $P$. Since the dimension of the image is bounded above by $\dim S^m(U)=mn$, this process can be repeated at most a finite number of times. So we may assume given a surjective regular homomorphism $$f: A\>>> CH^n(X)_{\deg 0}$$ with countable kernel, where $A$ is a connected smooth commutative algebraic group, and for some Cartier curves $C_1, \ldots , C_s$ a surjective homomorphism \begin{equation}\label{surjection} \bigoplus_{j=1}^s {\rm Pic}^0(C_j) \> \oplus \rho_j >> A. \end{equation} Note that the composition $h:A\to CH^n(X)_{\deg 0}\to A^n(X)$ is then a surjective homomorphism of algebraic groups. We now distinguish between the case $k={\mathbb C}$, and that of a general universal domain $k$. \begin{proof}[Proof of \ref{finite} for $k={\mathbb C}$] We first show that the surjective homomorphism $h:A\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} A^n(X)$ is an {\em isogeny}. Clearly $h$ induces an injective homomorphism $$h^*:\Omega(A^n(X))\to\Omega(A).$$ We will use proposition~\ref{basic} to show that $h^*:\Omega(A^n(X))\to\Omega(A)$ is an isomorphism. Since $h^*$ is injective, it suffices to prove that $\dim \Omega(A)\leq\dim \Omega(A^n(X))$. Consider the set $\Gamma=U\times_{CH^n(X)_{\deg 0}}A$. This is a countable union of algebraic subvarieties, and maps surjectively to $U$ under the projection. Recalling that $U=\cup_j U_j$, we can then find irreducible varieties $\Gamma_j\subset \Gamma$ such that $\Gamma_j$ dominates $U_j$ under the projection $\Gamma\to U$. Then $\pi_j:\Gamma_j\to U_j$ has countable, and hence finite, fibres. Let $d_j$ be the degree of $\pi_j$, and let $V_j\subset U_j$ be a dense open subset such that $\pi_j:\pi_j^{-1}(V_j)\to V_j$ is an \'etale covering of degree $d_j$. Let $c$ be the l.c.m. of the $d_j$, and let $c=d_jc_j$. If $q:\Gamma\to A$ is the second projection, then consider the morphism \begin{gather*} \mu:V=\bigcup_jV_j\>>> A,\\ \mu(x)=c_j\mathop{\sum}_{y\in \pi_j^{-1}(x)}q(y)\,\,\mbox{for $x\in V_j$.} \end{gather*} One verifies at once that the diagram \begin{equation}\label{diag-1} \begin{CD} \bigcup_j V_j=V \>{\mu}>> A \\ {\mathbb V} VV {\mathbb V} V f V\\ \bigcup_j U_j=U \>{c\cdot\gamma_1}>>CH^n(X)_{\deg 0} \end{CD} \end{equation} commutes. The image of $\mu(V)$ in $CH^n(X)_{\deg 0}$ generates $CH^n(X)_{\deg 0}$ as a group, since any 0-cycle on $X$ is rationally equivalent to a cycle supported on $V$. Hence the subgroup of $A$ generated by $\mu(V)$ has countable index, and is also a countable increasing union of constructible subsets, namely the images of $\mu(V)^{2m}$ under the maps \begin{align*} \sigma_m:A^{2m}&\>>>A,\;\;\;\; m\geq 1,\\ (a_1,\ldots,a_{2m})&\longmapsto a_1+\cdots+a_m-a_{m+1}-\cdots-a_{2m}. \end{align*} By dimension considerations, one of the subsets $\sigma_m(\mu(V)^{2m})$ must be dense in $A$, and then $\sigma_{2m}(\mu(V)^{4m})=A$. Hence the induced map on 1-forms $$ \Omega(A)\to H^0(V^{4m},\Omega^1_{V^{4m}/{\mathbb C}}) $$ is injective. Now the action of $\sigma_{2m}$ on 1-forms is given by \[\sigma_{2m}^*(\omega)= (\omega,\ldots,\omega,-\omega,-\omega,\cdots,-\omega).\] This means that the map on 1-forms $\Omega(A)\to H^0(V,\Omega^1_{V/{\mathbb C}})$, induced by the morphism $V\to \mu(V)\hookrightarrow A$, is injective. We claim that ${\rm image}\,\Omega(A)\subset \Omega(A^n(X))$, so that $\dim\Omega(A)\leq \dim\Omega(A^n(X))$. To see this, it suffices by proposition~\ref{basic} to show that for any reduced Cartier curve $C\subset X$ with $B=(C_{{\rm reg}})\cap V$ dense in $C$, the image of any element of $\Omega(A)$ in $H^0(B,\Omega^1_{B/{\mathbb C}})$ lies in the image of $H^0(C,\omega_C)$. Fixing base points in each component of $B$, we obtain a morphism $\vartheta:C_{{\rm reg}}\to{\rm Pic}^0(C)$. If $C_i$ is any component of $C_{{\rm reg}}$, then the two induced maps \begin{gather*} C_i\>>> {\rm Pic}^0(C)\>>> CH^n(X)_{\deg 0},\\ C_i\hookrightarrow U\>{\gamma_1}>> CH^n(X)_{\deg 0} \end{gather*} agree up to translation by a fixed element of $CH^n(X)_{\deg 0}$. Now consider the subgroup $\Gamma_C={\rm Pic}^0(C)\times_{CH^n(X)_{\deg 0}} A$. As before, this is a countable union of subvarieties of ${\rm Pic}^0(C)\times A$. Hence there is a connected algebraic subgroup $\Gamma^0_C\subset\Gamma_C$ such that $\Gamma_C/\Gamma^0_C$ is a countable group. Further, $\Gamma_C\to{\rm Pic}^0(C)$ is surjective with countable fibres. Hence $\Gamma^0_C\to{\rm Pic}^0(C)$ is an isogeny. Restricting (\ref{diag-1}) one obtains a commutative diagram \[\begin{CD} B \>{\mu}>> A\\ {\mathbb V} VV {\mathbb V} V f V\\ C_{{\rm reg}}\>{c\gamma_1}>> CH^n(X)_{\deg 0} \end{CD}\] and hence a morphism $B\to \Gamma^0_C$ such that \begin{points} \item for each component $C_i$ of $C_{{\rm reg}}$, the composite $C_i\cap B\to \Gamma^0_C\to{\rm Pic}^0(C)$ equals the restriction of the composite $C_i\to {\rm Pic}^0(C)\longby{c\cdot }{\rm Pic}^0(C)$, up to a translation (here $c\cdot$ denotes multiplication by $c$) \item $C_i\cap B\to \Gamma^0_C\to A$ agrees with $\mu$, up to a translation. \end{points} Hence, by (ii), $$ {\rm image}(\Omega(A)\>{\mu^*}>>\Gamma(B,\Omega^1_{C/{\mathbb C}})) \subset {\rm image}\left(\Omega(\Gamma^0_C)\>>> \Gamma(B,\Omega^1_{C/{\mathbb C}})\right) $$ while by (i), \[{\rm image}\,\Omega(\Gamma^0_C)={\rm image}\,\Omega({\rm Pic}^0(C))={\rm image}\,\Gamma(C,\omega_C).\] Since $C$ was arbitrary, we have verified the hypotheses of proposition~\ref{basic}. This completes the proof that the composite $h:A\to CH^n(X)_{\deg 0}\to A^n(X)$ is an isogeny. In particular, $f:A\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} CH^n(X)_{\deg 0}$ has a finite kernel. Replacing $A$ by $A/(\ker f)$, we may assume given a regular homomorphism $f:A\to CH^n(X)_{\deg 0}$ which is an {\em isomorphism} of groups. Now repeating the above arguments once more, we obtain (\ref{diag-1}) with $c=1$. By corollary~\ref{equ-reg}, this means the group isomorphism $f^{-1}:CH^n(X)_{\deg 0}\to A$ is a regular homomorphism, which must factor through $\varphi:CH^n(X)_{\deg 0}\mbox{$\,\>>>\hspace{-.5cm}\to\hspace{.15cm}$} A^n(X)$. This forces $\varphi$ to be an isomorphism of groups, as well. \end{proof} \begin{rmk}\label{roitman} Over the field of complex numbers the last part of the proof of \ref{finite} is consistent with the Roitman theorem proved in \cite{BiS}. In fact, if $$A\cong CH^n(X)_{\deg 0}\>>> A^n(X)$$ is surjective with finite kernel the generalization of Roitman's theorem implies that the composite \[A\>{\cong}>> CH^n(X)_{\deg 0}\>>> A^n(X)\>>> J^n(X)\] is an isomorphism on torsion subgroups, so that $CH^n(X)_{\deg 0}\to A^n(X)$ is an injection on torsion subgroups. Hence the isogeny $A\cong CH^n(X)_{\deg 0}\to A^n(X)$ must be an isomorphism. \end{rmk} In the algebraic case we have to modify the arguments, in particular since the lower horizontal morphism in the diagram (\ref{diag-1}) need not to be surjective in characteristic $p>0$. \begin{proof}[Proof of \ref{finite} for $k$ a universal domain] \ \\ Let us write $B$ for the kernel of $h:A \to A^n(X)$, a closed subgroup scheme of $A$, not necessarily reduced. We may replace $A$ by $A/\kappa$, for any (zero dimensional) closed subgroup scheme $\kappa$ of $B$ such that $\kappa(k)\subset\ker f$. The group $B$ acts on $U\times_{A^n(X)}A$ with quotient $U\times_{A^n(X)}A^n(X)=U$. The kernel ${\mathcal K}$ of the map $A(k) \to CH^n(X)_{\deg 0}$ consists of countably many closed points, the induced action on $U\times_{CH^n(X)_{\deg 0}}A$ is free, and the induced map on the quotient $$ (U\times_{CH^n(X)_{\deg 0}}A)/{\mathcal K} = U\times_{CH^n(X)_{\deg 0}}(A/{\mathcal K}) \>>> U $$ is a bijection on the closed points. Let $V \subset U$ be an open dense subscheme, and let $\Gamma_j$ be a locally closed irreducible subscheme of $V\times_{A^n(X)}A$, contained in $V\times_{CH^n(X)_{\deg 0}}A$, and dominant over the component $V_j=V\cap U_j$ of $V$ under the first projection. For $V$ small enough, we may assume that $\Gamma_j \to V_j$ is finite. Let $\kappa_j\subset {\mathcal K}$ be the subgroup of elements $g$ with $g(\Gamma_j) = \Gamma_j$. Then $\kappa_j$ is a finite group and $\Gamma_j/\kappa_j \to V_j$ is an isomorphism on the closed points. Replacing $\Gamma_j$ by its image in $U\times_{A^n(X)}(A/\kappa_j)$ and $A$ by $A/\kappa_j$ we may assume that $\kappa_j$ is trivial, and thereby that $\Gamma_j \to V_j$ is purely inseparable. Repeating this construction for the different components of $U$ we finally reduce to the situation, where $U$ has an open dense subscheme $V$, and where $V\times_{CH^n(X)_{\deg 0}}A$ has a closed subscheme $\Gamma$ which is purely inseparable over $V$. Assume that $\Gamma \to V$ is not an isomorphism, in particular, that the characteristic of $k$ is $p>0$. The restriction of the group action to $B\times \Gamma$ factors as $$ B \times \Gamma \> \cong >> (V\times_{A^n(X)}A)\times_V\Gamma \> pr_1 >> V\times_{A^n(X)}A $$ and the preimage $S(\Gamma)$ of $\Gamma \subset V\times_{A^n(X)}A$ is isomorphic to $\Gamma \times_V \Gamma$. Thus $S(\Gamma)$ is a subscheme of $B\times \Gamma$, supported in the zero section $\{e\}\times \Gamma$. Hence $S(\Gamma)$ is contained in the $\nu$-th infinitesimal neighbourhood $\{e\}_\nu \times \Gamma$ of the zero section, for some $\nu > 0$. The kernel $\kappa^{(\nu')}$ of the $\nu'$-th geometric Frobenius $F^{(\nu')}: B \to B^{(\nu')}$ is defined by the sheaf of ideals in ${\mathcal O}_B$, generated by the $p^{\nu'}$-th powers of the generators of the sheaf of ideals $\rm \bf m$ defining $\{e\} \subset B$. For some $\nu'>0$ it is contained in ${\rm \bf m}^\nu$ and $\{e\}_\nu$ is a subscheme of $\kappa^{(\nu')}$. Dividing $A$ by $\kappa^{(\nu')}$, we may assume that $S(\Gamma)=\Gamma\times_V\Gamma$ is isomorphic to $\Gamma$, and thereby that $\Gamma$ is isomorphic to $V$. Independent of the characteristic of $k$, we have thus reduced to the situation where $U$ has an open dense subscheme $V$, for which $$ pr_1:V\times_{CH^n(X)_{\deg 0}}A \to V $$ has a section, such that on projecting to $A$ we obtain a morphism $\mu:V\to A$ and (using the notation introduced in \ref{difference}) a commutative diagram \begin{equation}\label{diag2} \begin{CD} \Pi_V \>{\mu^{(-)}}>> A \\ {\mathbb V}\subseteq VV {\mathbb V} V f V\\ \Pi_U \>{\gamma^{(-)}}>>CH^n(X)_{\deg 0}. \end{CD} \end{equation} \begin{claim}\label{claim2} There exists a surjective homomorphism $\phi:CH^n(X)_{\deg 0} \to A$ with $\mu^{(-)}=\phi \circ \gamma^{(-)}|_{\Pi_V}$. In particular, $\phi$ is regular. \end{claim} \begin{proof} Let $(C',\iota)$ be an admissible pair with $B=(\iota^{-1}(V))_{\rm reg}$ dense in $C'$. By restriction (\ref{diag2}) gives rise to a commutative diagram \begin{equation}\label{diag3} \begin{CD} \Pi_B \>{{\mu'}^{(-)}}>> A \\ {\mathbb V}= VV {\mathbb V} V f V\\ \Pi_B \>{\gamma_B^{(-)}}>>CH^n(X)_{\deg 0} \end{CD} \end{equation} where $\mu'=\mu|_B$. By lemma \ref{pic} the lower horizontal map in the diagram (\ref{diag3}) factors as \begin{equation}\label{diag4} \begin{CDS} \Pi_B \>\gamma_B^{(-)}>> CH^n(X)_{\deg 0}\\ & \SE E \vartheta^{(-)} E {\mathbb A} \eta A A \\ && {\rm Pic}^0(C'). \end{CDS} \end{equation} Let $\Gamma_{C'}^0$ be the connected component of $\Gamma_{C'}={\rm Pic}^0({C'})\times_{CH^n(X)_{\deg 0}} A$ containing the origin. $\Gamma_{C'}/\Gamma_{C'}^0$ is a countable group and $\Gamma_{C'}^0 \to {\rm Pic}^0({C'})$ is an isogeny. Since the diagrams (\ref{diag3}) and (\ref{diag4}) are commutative, the image of $$ \Pi_B \> (\vartheta^{(-)}, {\mu'}^{(-)}) >> {\rm Pic}^0({C'})\times A $$ is contained in $\Gamma_{C'}^0$. This implies that $\Gamma_{C'}^0\to {\rm Pic}^0(C')$ must be an isomorphism. In fact, by \ref{generators2} there is an open connected subscheme $W$ of $S^g(B)$ such that the morphism $\vartheta_{W} : W \to {\rm Pic}^0(C)$ is an open embedding. On the other hand, $\vartheta_W$ factors through the isogeny $\Gamma_{C'}^0\to {\rm Pic}^0(C')$. Hence the morphism ${\mu'}^{(-)}$ in the diagram (\ref{diag3}) is the composite $$ \Pi_B \> \vartheta^{(-)} >> {\rm Pic}^0(C')\cong \Gamma_{C'}^0 \> pr_2 >> A, $$ and the condition (b) in lemma \ref{pic} holds true. Thereby the homomorphism $\phi$ in \ref{claim2} exists, and it remains to show that $\phi$ is surjective. Equivalently, it suffices to show that the image of $\phi$ generates $A$ as a group, which will follow if we show that $\mu^{-}(\Pi_V)$ generates $A$. But we know that $\gamma^{-}(\Pi_V)$ generates $CH^n(X)_{\deg 0}$, and so $\mu^{-}(\Pi_V)$ generates a subgroup of countable index in $A$. Since $k$ is a universal domain, $\mu^{-}(\Pi_V)$ generates $A$. \end{proof} By claim \ref{claim2} and by the universal property for $A^n(X)$ the regular homomorphism $\phi: CH^n(X)_{\deg 0}\to A$ factors through a homomorphism of algebraic groups $\chi:A^n(X)\to A$. Since $\phi$ is surjective, the induced morphism $\chi$ is surjective as well. Further, the composite $CH^n(X)_{\deg 0}\>{\phi}>> A\> f>> CH^n(X)_{\deg 0}$ is clearly the identity, since it is so on the image of $\Pi_V$, which is a set of generators. By the universal property of $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$, we deduce that the composite $A^n(X)\>{\chi}>> A\>h>> A^n(X)$ is the identity. Hence $\chi$ and $h$ are inverse isomorphisms of algebraic groups, and $f:A\to CH^n(X)_{\deg 0}$ and $\varphi:CH^n(X)_{\deg 0}\to A^n(X)$ are both isomorphisms (of groups) as well. \end{proof} \bibliographystyle{plain}
1997-12-11T20:58:25
9712
alg-geom/9712012
en
https://arxiv.org/abs/alg-geom/9712012
[ "alg-geom", "math.AG" ]
alg-geom/9712012
Misha Verbitsky
Misha Verbitsky
Hyperholomorphic sheaves and new examples of hyperkaehler manifolds
113 pages, v. 2.0, an error in the statement of Theorem 8.15 corrected; Mathematical Physics, 12. International Press, 1999. iv+257 pp. ISBN: 1-57146-071-3
null
null
null
http://creativecommons.org/licenses/by/3.0/
Given a compact hyperkaehler manifold $M$ and a holomorphic bundle B over $M$, we consider a Hermitian connection $\nabla$ on B which is compatible with all complex structures on $M$ induced by the hyperkaehler structure. Such a connection is unique, because it is Yang-Mills. We call the bundles admitting such connections hyperholomorphic bundles. A stable bundle is hyperholomorphic if and only if its Chern classes $c_1$, $c_2$ are SU(2)-invariant, with respect to the natural SU(2)-action on the cohomology. For several years, it was known that the moduli space of stable hyperholomorphic bundles is singular hyperkaehler. More recently, it was proven that singular hyperkaehler varieties admit a canonical hyperkaehler desingularization. In the present paper, we show that a moduli space of stable hyperholomorphic bundles is compact, given some assumptions on Chern classes of B and hyperkaehler geometry of $M$ (we also require $dim_C M>2$). Conjecturally, this leads to new examples of hyperkaehler manifolds. We develop the theory of hyperholomorphic sheaves, which are (intuitively speaking) coherent sheaves compatible with hyperkaehler structure. We show that hyperholomorphic sheaves with isolated singularities can be canonically desingularized by a blow-up. This theory is used to study degenerations of hyperholomorphic bundles.
[ { "version": "v1", "created": "Thu, 11 Dec 1997 19:58:43 GMT" }, { "version": "v2", "created": "Sun, 9 Dec 2012 22:41:59 GMT" } ]
2012-12-11T00:00:00
[ [ "Verbitsky", "Misha", "" ] ]
alg-geom
\section{Introduction} \label{_intro_Section_} \hfill For an introduction to basic results and the history of hyperk\"ahler geometry, see \cite{_Besse:Einst_Manifo_}. \hfill This Introduction is independent from the rest of this paper. \subsection{An overview} \subsubsection{Examples of hyperk\"ahler manifolds} A Riemannian manifold $M$ is called {\bf hyperk\"ahler} if the tangent bundle of $M$ is equipped with an action of quaternian algebra, and its metric is K\"ahler with respect to the complex structures $I_\iota$, for all embeddings ${\Bbb C} \stackrel{\iota}\hookrightarrow \Bbb H$. The complex structures $I_\iota$ are called {\bf induced complex structures}; the corresponding K\"ahler manifold is denoted by $(M, I_\iota)$. For a more formal definition of a hyperk\"ahler manifold, see \ref{_hyperkahler_manifold_Definition_}. The notion of a hyperk\"ahler manifold was introduced by E. Calabi (\cite{_Calabi_}). Clearly, the real dimension of $M$ is divisible by 4. For $\dim_{\Bbb R} M= 4$, there are only two classes of compact hyperk\"ahler manifolds: compact tori and K3 surfaces. Let $M$ be a complex surface and $M^{(n)}$ be its $n$-th symmetric power, $M^{(n)} = M^n/S_n$. The variety $M^{(n)}$ admits a natural desingularization $M^{[n]}$, called {\bf the Hilbert scheme of points}. The manifold $M^{[n]}$ admits a hyperk\"ahler metrics whenever the surface $M$ is compact and hyperk\"ahler (\cite{_Beauville_}). This way, Beauville constructed two series of examples of hyperk\"ahler manifolds, associated with a torus (so-called ``higher Kummer variety'') and a K3 surface. It was conjectured that all compact hyperk\"ahler manifolds $M$ with $H^1(M) =0$, $H^{2,0}(M)={\Bbb C}$ are deformationally equivalent to one of these examples. In this paper, we study the deformations of coherent sheaves over higher-dimensional hyperk\"ahler manifolds in order to construct counterexamples to this conjecture. A different approach to the construction of new examples of hyperk\"ahler manifolds is found in the recent paper of K. O'Grady, who studies the moduli of semistable bundles over a K3 surface and resolves the singularities using methods of symplectic geometry (\cite{_O'Grady_}). \subsubsection{Hyperholomorphic bundles} Let $M$ be a compact hyperk\"ahler manifold, and $I$ an induced complex structure. It is well known that the differential forms and cohomology of $M$ are equipped with a natural $SU(2)$-action (\ref{_SU(2)_commu_Laplace_Lemma_}). In \cite{_Verbitsky:Hyperholo_bundles_}, we studied the holomorphic vector bundles $F$ on $(M,I)$ which are compatible with a hyperk\"ahler structure, in the sense that any of the following conditions hold: \begin{equation}\label{_hyperho_condi_Equation_} \begin{minipage}[m]{0.8\linewidth} \begin{description} \item[(i)] The bundle $F$ admits a Hermitian connection $\nabla$ with a curvature $\Theta\in \Lambda^2(M, \operatorname{End}(F))$ which is of Hodge type (1,1) with respect to any of induced complex structures. \item[(ii)] The bundle $F$ is a direct sum of stable bundles, and its Chern classes $c_1(F)$, $c_2(F)$ are $SU(2)$-invariant. \end{description} \end{minipage} \end{equation} These conditions are equivalent (\ref{_inva_then_hyperho_Theorem_}). Moreover, the connection $\nabla$ of \eqref{_hyperho_condi_Equation_} (i) is Yang-Mills (\ref{_hyperholo_Yang--Mills_Proposition_}), and by Uhlenbeck--Yau theorem (\ref{_UY_Theorem_}), it is unique. A holomorphic vector bundle satisfying any of the conditions of \eqref{_hyperho_condi_Equation_} is called {\bf hyperholomorphic} (\cite{_Verbitsky:Hyperholo_bundles_}). Clearly, a stable deformation of a hyperholomorphic bundle is again a hyperholomorphic bundle. In \cite{_Verbitsky:Hyperholo_bundles_}, we proved that a deformation space of hyperholomorphic bundles is a singular hyperk\"ahler variety. A recent development in the theory of singular hyperk\"ahler varieties (\cite{_Verbitsky:Desingu_}, \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_}) gave a way to desingularize singular hyperk\"ahler manifolds, in a canonical way. It was proven (\ref{_desingu_Theorem_}) that a normalization of a singular hyperk\"ahler variety (taken with respect to any induced complex structure $I$) is a smooth hyperk\"ahler manifold. This suggested a possibility of constructing new examples of compact hyperk\"ahler manifolds, obtained as deformations of hyperholomorphic bundles. Two problems arise. \hfill {\bf Problem 1.} The deformation space of hyperholomorphic bundles is {\it a priori} non-compact and must be compactified. \hfill {\bf Problem 2.} The geometry of deformation spaces is notoriously hard to study. Even the dimension of a deformation space is difficult to compute, in simplest examples. How to find, for example, the dimension of the deformation space of a tangent bundle, on a Hilbert scheme of points on a K3 surface? The Betti numbers are even more difficult to compute. Therefore, there is no easy way to distinguish a deformation space of hyperholomorphic bundles from already known examples of hyperk\"ahler manifolds. \hfill In this paper, we address Problem 1. Problem 2 can be solved by studying the algebraic geometry of moduli spaces. It turns out that, for a generic deformation of a complex structure, the Hilbert scheme of points on a K3 surface has no closed complex subvarieties (\cite{_Verbitsky:Hilbert_}; see also \ref{_no_triana_subva_of_Hilb_Theorem_}). It is possible to find a 21-dimensional family of deformations of the moduli space $\operatorname{Def}(B)$ of hyperholomorphic bundles, with all fibers having complex subvarieties (\ref{_double_Fou_embedding_Lemma_}). Using this observation, it is possible to show that $\operatorname{Def}(B)$ is a new example of a hyperk\"ahler manifold. Details of this approach are given in Subsection \ref{_new_exa_F-M_checking_Subsection_}, and the complete proofs will be given in a forthcoming paper. It was proven that a Hilbert scheme of a generic K3 surface has no trianalytic subvarieties.\footnote{Trianalytic subvariety (\ref{_trianalytic_Definition_}) is a closed subset which is complex analytic with respect to any of induced complex structures.} Given a hyperk\"ahler manifold $M$ and an appropriate hyperholomorphic bundle $B$, denote the deformation space of hyperholomorphic connections on $B$ by $\operatorname{Def}(B)$. Then the moduli of complex structures on $M$ are locally embedded to a moduli of complex structures on $\operatorname{Def}(B)$ (\ref{_maps_pre_tw_curves_Claim_}). Since the dimension of the moduli of complex structures on $\operatorname{Def}(B)$ is equal to its second Betti number minus 2 (\ref{_Bogomo_etale_Theorem_}), the second Betti number of $\operatorname{Def}(B)$ is no less than the second Betti number of $M$. The Betti numbers of Beauville's examples of simple hyperk\"ahler manifolds are 23 (Hilbert scheme of points on a K3 surface) and 7 (generalized Kummer variety). Therefore, for $M$ a generic deformation of a Hilbert scheme of points on K3, $\operatorname{Def}(B)$ is either a new manifold or a generic deformation of a Hilbert scheme of points on K3. It is easy to construct trianalytic subvarieties of the varieties $\operatorname{Def}(B)$, for hyperholomorphic $B$ (see \cite{_Verbitsky:Symplectic_I_}, Appendix for details). This was the motivation of our work on trianalytic subvarieties of the Hilbert scheme of points on a K3 surface (\cite{_Verbitsky:Hilbert_}). For a generic complex structure on a hyperk\"ahler manifold, all stable bundles are hyperholomorphic (\cite{_Verbitsky:Symplectic_I_}). Nethertheless, hyperholomorphic bundles over higher-dimensional hyperk\"ahler manifolds are in short supply. In fact, the only example to work with is the tangent bundle and its tensor powers, and their Chern classes are not prime. Therefore, there is no way to insure that a deformation of a stable bundle will remain stable (like it happens, for instance, in the case of deformations of stable bundles of rank 2 with odd first Chern class over a K3 surface). Even worse, a new kind of singularities may appear which never appears for 2-dimensional base manifolds: a deformation of a stable bundle can have a singular reflexization. We study the singularities of stable coherent sheaves over hyperk\"ahler manifolds, using Yang-Mills theory for reflexive sheaves developed by S. Bando and Y.-T. Siu (\cite{_Bando_Siu_}). \subsubsection{Hyperholomorphic sheaves} A compactification of the moduli of hyperholomorphic bundles is the main purpose of this paper. We require the compactification to be singular hyperk\"ahler. A natural approach to this problem requires one to study the coherent sheaves which are compatible with a hyperk\"ahler structure, in the same sense as hyperholomorphic bundles are holomorphic bundles compatible with a hyperk\"ahler structure. Such sheaves are called {\bf hyperholomorphic sheaves} (\ref{_hyperho_shea_Definition_}). Our approach to the theory of hyperholomorphic sheaves uses the notion of admissible Yang-Mills connection on a coherent sheaf (\cite{_Bando_Siu_}). The equivalence of conditions \eqref{_hyperho_condi_Equation_} (i) and \eqref{_hyperho_condi_Equation_} (ii) is based on Uhlen\-beck--\-Yau theorem (\ref{_UY_Theorem_}), which states that evey stable bundle $F$ with $\deg c_1(F) =0$ admits a unique Yang--Mills connection, that is, a connection $\nabla$ satisfying $\Lambda\nabla^2=0$ (see Subsection \ref{_sta_bu_and_YM_Subsection_} for details). S. Bando and Y.-T. Siu developed a similar approach to the Yang--Mills theory on (possibly singular) coherent sheaves. Consider a coherent sheaf $F$ and a Hermitian metric $h$ on a locally trivial part of $F\restrict{U}$. Then $h$ is called admissible (\ref{_admi_metri_Definition_}) if the curvature $\nabla^2$ of the Hermitian connection on $F\restrict{U}$ is square-integrable, and the section $\Lambda\nabla^2\in \operatorname{End}(F\restrict{U})$ is uniformly bounded. The admissible metric is called {\bf Yang-Mills} if $\Lambda\nabla^2=0$ (see \ref{_Yang-Mills_sheaves_Definition_} for details). There exists an analogue of Uhlenbeck--Yau theorem for coherent sheaves (\ref{_UY_for_shea_Theorem_}): a stable sheaf admits a unique admissible Yang--Mills metric, and conversely, a sheaf admitting a Yang--Mills metric is a direct sum of stable sheaves with the first Chern class of zero degree. A coherent sheaf $F$ is called {\bf reflexive} if it is isomorphic to its second dual sheaf $F^{**}$. The sheaf $F^{**}$ is always reflexive, and it is called {\bf a reflexization} of $F$ (\ref{_refle_Definition_}). Applying the arguments of Bando and Siu to a reflexive coherent sheaf $F$ over a hyperk\"ahler manifold $(M, I)$, we show that the following conditions are equivalent (\ref{_hyperho_conne_exi_Theorem_}). \begin{description} \item[(i)] The sheaf $F$ is stable and its Chern classes $c_1(F)$, $c_2(F)$ are $SU(2)$-invariant \item[(ii)] $F$ admits an admissible Yang--Mills connection, and its curvature is of type (1,1) with respect to all induced complex structures. \end{description} A reflexive sheaf satisfying any of the these conditions is called {\bf reflexive stable hyperholomorphic}. An arbitrary torsion-free coherent sheaf is called {\bf stable hyperholomorphic} if its reflexization is hyperholomorphic, and its second Chern class is $SU(2)$-invariant, and {\bf semistable hyperholomorphic} if it is a successive extension of stable hyperholomorphic sheaves (see \ref{_hyperho_shea_Definition_} for details). This paper is dedicated to the study of hyperholomorphic sheaves. \subsubsection{Deformations of hyperholomorphic sheaves} By \ref{_generic_are_dense_Proposition_}, for an induced complex structure $I$ of general type, {\bf all} coherent sheaves are hyperholomorphic. However, the complex structures of general type are never algebraic, and in complex analytic situation, the moduli of coherent sheaves are, generaly speaking, non-compact. We study the flat deformations of hyperholomorphic sheaves over $(M,I)$, where $I$ is an algebraic complex structure. {\it A priori}, a flat deformation of a hyperholomorphic sheaf will be no longer hyperholomorphic. We show that for some algebraic complex structures, called {\bf $C$-restricted complex structures}, a flat deformation of a hyperholomorphic sheaf remains hyperholomorphic (\ref{_sheaf_on_C_restr_hyperho_Theorem_}). This argument is quite convoluted, and takes two sections (Sections \ref{_cohomo_hype_Section_} and \ref{_C_restri_Section_}). Further on, we study the local structure of stable reflexive hyperholomorphic sheaves with isolated singularities. We prove the Desingularization Theorem for such hyperholomorphic sheaves (\ref{_desingu_hyperho_Theorem_}). It turns out that such a sheaf can be desingularized by a single blow-up. The proof of this result is parallel to the proof of Desingularization Theorem for singular hyperk\"ahler varieties (\ref{_desingu_Theorem_}). The main idea of the desingularization of singular hyperk\"ahler varieties (\cite{_Verbitsky:DesinguII_}) is the following. Given a point $x$ on a singular hyperk\"ahler variety $M$ and an induced complex structure $I$, the complex variety $(M, I)$ admits a local ${\Bbb C}^*$-action which preserves $x$ and acts as a dilatation on the Zariski tangent space of $x$. Here we show that any stable hyperholomorphic sheaf $F$ is equivariant with respect to this ${\Bbb C}^*$-action (\ref{_Psi_equiv_hyperho_Theorem_}, \ref{_C^*_stru_on_sge_Definition_}). Then an elementary algebro-geometric argument (\ref{_desingu_C^*_equi_Proposition_}) implies that $F$ is desingularized by a blow-up. Using the desingularization of hyperholomorphic sheaves, we prove that a hyperholomorphic deformation of a hyperholomorphic bundle is again a bundle (\ref{_reflexi_defo_loca_trivi_Theorem_}), assuming that it has isolated singularities. The proof of this result is conceptual but quite difficult, it takes 3 sections (Sections \ref{_twisto_tra_Section_}--\ref{_modu_hyperho_Section_}), and uses arguments of quaternionic-K\"ahler geometry (\cite{_Swann_}, \cite{_Nitta:Y-M_}) and twistor transform (\cite{_NHYM_}). In our study of deformations of hyperholomorphic sheaves, we usually assume that a deformation of a hyperholomorphic sheaf over $(M, I)$ is again hyperholomorphic, i. e. that an induced complex structure $I$ is $C$-restricted, for $C$ sufficiently big (\ref{_C_restri_Definition_}). Since $C$-restrictness is a tricky condition, it is preferable to get rid of it. For this purpose, we use the theory of twistor paths, developed in \cite{_coho_announce_}, to show that the moduli spaces of hyperholomorphic sheaves are real analytic equivalent for different complex structures $I$ on $M$ (\ref{_iso_Bun_exists_gene_pola_Theorem_}). This is done as follows. A hyperk\"ahler structure on $M$ admits a 2-dimensional sphere of induced complex structures. This gives a rational curve in the moduli space $Comp$ of complex structures on $M$, so-called {\bf twistor curve}. A sequence of such rational curves connect any two points of $Comp$ (\ref{_twistor_connect_Theorem_}). A sequence of connected twistor curves is called {\bf a twistor path}. If the intersection points of these curves are generic, the twistor path is called {\bf admissible} (\ref{_admi_twi_path_Definition_}). It is known (\ref{_admi_twi_impli_Theorem_}) that an admissible twistor path induces a real analytic isomorphism of the moduli spaces of hyperholomorphic bundles. There exist admissible twistor paths connecting any two complex structures (\ref{_admi_pa_exist_for_gene_pol_Claim_}). Thus, if we prove that the moduli of deformations of hyperholomorphic bundles are compact for one generic hyperk\"ahler structure, we prove a similar result for all generic hyperk\"ahler structures (\ref{_iso_Bun_exists_gene_pola_Theorem_}). Applying this argument to the moduli of deformations of a tangent bundle, we obtain the following theorem. \hfill \theorem \label{_defo_tange_Hilb_compact_intro_Theorem_} Let $M$ be a Hilbert scheme of points on a K3 surface, $\dim_{\Bbb H}(M)>1$ and $\c H$ a generic hyperk\"ahler structure on $M$. Assume that for all induced complex structures $I$, except at most a finite many of, all semistable bundle deformations of the tangent bundle $T(M, I)$ are stable. Then, for all complex structures $J$ on $M$ and all generic polarizations $\omega$ on $(M, J)$, the deformation space $\c M_{J, \omega}(T(M, J))$ is singular hyperk\"ahler and compact, and admits a smooth compact hyperk\"ahler desingularization. {\bf Proof:} This is \ref{_defo_tange_compact_Theorem_}. \blacksquare \hfill In the course of this paper, we develop the theory of $C$-restricted complex structures (Sections \ref{_cohomo_hype_Section_} and \ref{_C_restri_Section_}) and another theory, which we called {\bf the Swann's formalism for vector bundles} (Sections \ref{_twisto_tra_Section_} and \ref{_C_equiv_twi_spa_Section_}). These themes are of independent interest. We give a separate introduction to $C$-restricted complex structures (Subsection \ref{_C_restri_intro_Subsection_}) and Swann's formalism (Subsection \ref{_Swann's_intro_Subsection_}). \subsection{$C$-restricted complex structures: an introduction} \label{_C_restri_intro_Subsection_} This part of the Introduction is highly non-precise. Our purpose is to clarify the intuitive meaning of $C$-restricted complex structure. Consider a compact hyperk\"ahler manifold $M$, which is {\bf simple} (\ref{_simple_hyperkahler_mfolds_Definition_}), that is, satisfies $H^1(M) =0$, $H^{2,0}(M) = {\Bbb C}$. A {\bf reflexive hyperholomorphic sheaf} is by definition a semistable sheaf which has a filtration of stable sheaves with $SU(2)$-invariant $c_1$ and $c_2$. A {\bf hyperholomorphic sheaf} is a torsion-free sheaf which has hyperholomorphic reflexization and has $SU(2)$-invariant $c_2$ (\ref{_hyperho_shea_Definition_}). If the complex structure $I$ is of general type, all coherent sheaves are hyperholomorphic (\ref{_generic_manifolds_Definition_}, \ref{_generic_are_dense_Proposition_}), because all integer $(p,p)$-classes are $SU(2)$-invariant. However, for generic complex structures $I$, the corresponding complex manifold $(M, I)$ is never algebraic. If we wish to compactify the moduli of holomorphic bundles, we need to consider algebraic complex structures, and if we want to stay in hyperholomorphic category, the complex structures must be generic. This paradox is reconciled by considering the $C$-restricted complex structures (\ref{_C_restri_Definition_}). Given a generic hyperk\"ahler structure $\c H$, consider an algebraic complex structure $I$ with $Pic(M, I) = {\Bbb Z}$. The group of rational $(p,p)$-cycles has form \begin{equation} \label{_Pic_1_decomposi_Equation_} \begin{split} H^{p,p}_I(M, {\Bbb Q}) = &H^{2p}(M ,{\Bbb Q})^{SU(2)} \oplus a \cdot H^{2p}(M ,{\Bbb Q})^{SU(2)} \\ & a^2 \cdot\oplus H^{2p}(M ,{\Bbb Q})^{SU(2)} \oplus... \end{split} \end{equation} where $a$ is a generator of $Pic(M, I)\subset H^{p,p}_I(M, {\Bbb Z})$ and $H^{2p}(M ,{\Bbb Q})^{SU(2)}$ is the group of rational $SU(2)$-invariant cycles. This decomposition follows from an explicit description of the algebra of cohomology given by \ref{_S^*H^2_is_H^*M_intro-Theorem_}. Let \[ \Pi:\; H^{p,p}_I(M, {\Bbb Q}){\:\longrightarrow\:} a \cdot H^{2p}(M ,{\Bbb Q})^{SU(2)}\oplus a^2 \cdot H^{2p}(M ,{\Bbb Q})^{SU(2)} \oplus ... \] be the projection onto non-$SU(2)$-invariant part. Using Wirtinger's equality, we prove that a fundamental class $[X]$ of a complex subvariety $X\subset (M, I)$ is $SU(2)$-invariant unless $\deg \Pi([X])\neq 0$ (\ref{_Wirti_hyperka_Proposition_}). A similar result holds for the second Chern class of a stable bundle (\ref{_stable_shea_degree_Corollary_},). A $C$-restricted complex structure is, heuristically, a structure for which the decomposition \eqref{_Pic_1_decomposi_Equation_} folds, and $\deg a>C$. For a $C$-restricted complex structure $I$, and a fundamental class $[X]$ of a complex subvariety $X\subset (M, I)$ of complex codimension 2, we have $\deg [X]>C$ or $X$ is trianalytic. A version of Wirtinger's inequality for vector bundles (\ref{_stable_shea_degree_Corollary_}) implies that a stable vector bundle $B$ over $(M, I)$ is hyperholomorphic, unless $|\deg c_2(B)| >C$. Therefore, over a $C$-restricted $(M, I)$, all torsion-free semistable coherent sheaves with bounded degree of the second Chern class are hyperholomorphic (\ref{_sheaf_on_C_restr_hyperho_Theorem_}). The utility of $C$-restricted induced complex structures is that they are algebraic, but behave like generic induced complex structures with respect to the sheaves $F$ with low $|\deg c_2(F)|$ and $|\deg c_1(F)|$. We prove that a generic hyperk\"ahler structure admits $C$-restricted induced complex structures for all $C$, and the set of $C$-restricted induced complex structures is dense in the set of all induced complex structures (\ref{_C_restri_dense_Theorem_}). We prove this by studying the algebro-geometric properties of the moduli of hyperk\"ahler structures on a given hyperk\"ahler manifold (Subsection \ref{_modu_and_C-restri_Subsection_}). \subsection{Quaternionic-K\"ahler manifolds and Swann's formalism} \label{_Swann's_intro_Subsection_} Quaternionic-K\"ahler manifolds (Subsection \ref{_B_2_bundles_Subsection_}) are a beautiful subject of Riemannian geometry. We are interested in these manifolds because they are intimately connected with singularities of hyperholomorphic sheaves. A stable hyperholomorphic sheaf is equipped with a natural connection, which is called {\bf hyperholomorphic connection}. By definition, a {hyperholomorphic connection} on a torsion-free coherent sheaf is a connection $\nabla$ defined outside of singularities of $F$, with square-integrable curvature $\nabla^2$ which is an $SU(2)$- invariant 2-form (\ref{_hyperholo_co_Definition_}). We have shown that a stable hyperholomorphic sheaf admits a hyperholomorphic connection, and conversely, a reflexive sheaf admitting a hyperholomorphic connection is a direct sum of stable hyperholomorphic sheaves (\ref{_hyperho_conne_exi_Theorem_}). Consider a reflexive sheaf $F$ over $(M, I)$ with an isolated singularity in $x\in M$. Let $\nabla$ be a hyperholomorphic connection on $F$. We prove that $F$ can be desingularized by a blow-up of its singular set. In other words, for $\pi:\; \tilde M {\:\longrightarrow\:} (M, I)$ a blow-up of $x\in M$, the pull-back $\pi^* F$ is a bundle over $\tilde M$. Consider the restriction $\pi^* F\restrict C$ of $\pi^* F$ to the blow-up divisor \[ C = {\Bbb P} T_x M \cong{\Bbb C} P^{2n-1}. \] To be able to deal with the singularities of $F$ effectively, we need to prove that the bundle $\pi^* F\restrict C$ is semistable and satisfies $c_1\left(\pi^* F\restrict C\right)=0$. The following intuitive picture motivated our work with bundles over quaternionic-K\"ahler manifolds. The manifold $C = {\Bbb P} T_x M$ is has a quaternionic structure, which comes from the $SU(2)$-action on $T_xM$. We know that bundles which are compatible with a hyperk\"ahler structure (hyperholomorphic bundles) are (semi-)stable. If we were able to prove that the bundle $\pi^* F\restrict C$ is in some way compatible with quaternionic structure on $C$, we could hope to prove that it is (semi-)stable. To give a precise formulation of these heuristic arguments, we need to work with the theory of quaternionic-K\"ahler manifolds, developed by Berard-Bergery and Salamon (\cite{_Salamon_}). A quaternionic-K\"ahler manifold (\ref{_q-K_Definition_}) is a Riemannian manifold $Q$ equipped with a bundle $W$ of algebras acting on its tangent bundle, and satisfying the following conditions. The fibers of $W$ are (non-canonically) isomorphic to the quaternion algebra, the map $W\hookrightarrow \operatorname{End}(TQ)$ is compatible with the Levi-Civita connection, and the unit quaternions $h\in W$ act as orthogonal automorphisms on $TQ$. For each quaternion-K\"ahler manifold $Q$, one has a twistor space $\operatorname{Tw}(Q)$ (\ref{_twi_q-K_Definition_}), which is a total space of a spherical fibration consisting of all $h\in W$ satisfying $h^2=-1$. The twistor space is a complex manifold (\cite{_Salamon_}), and it is K\"ahler unless $W$ is flat, in which case $Q$ is hyperk\"ahler. Further on, we shall use the term ``quaternionic-K\"ahler'' for manifolds with non-trivial $W$. Consider the twistor space $\operatorname{Tw}(M)$ of a hyperk\"ahler manifold $M$, \\ equipped with a natural map \[ \sigma:\;\operatorname{Tw}(M) {\:\longrightarrow\:} M.\] Let $(B, \nabla)$ be a bundle over $M$ equipped with a hyperholomorphic connection. A pullback $(\sigma^* B, \sigma^*\nabla)$ is a holomorphic bundle on $\operatorname{Tw}(M)$ (\ref{_autodua_(1,1)-on-twi_Lemma_}), that is, the operator $\sigma^*\nabla^{0,1}$ is a holomorphic structure operator on $\sigma^* B$. This correspondence is called {\bf the direct twistor transform}. It is invertible: from a holomorphic bundle $(\sigma^* B, \sigma^*\nabla^{0,1})$ on $\operatorname{Tw}(M)$ it is possible to reconstruct $(B, \nabla)$, which is unique (\cite{_NHYM_}; see also \ref{_dire_inve_twisto_Theorem_}). A similar construction exists on quaternionic-K\"ahler manifolds, due to T. Nitta (\cite{_Nitta:bundles_}, \cite{_Nitta:Y-M_}). A bundle $(B, \nabla)$ on a quaternionic-K\"ahler manifold $Q$ is called {\bf a $B_2$-bundle} if its curvature $\nabla^2$ is invariant with respect to the adjoint action of ${\Bbb H}^*$ on $\Lambda^2(M, \operatorname{End}(B))$ (\ref{_B_2_bu_Definition_}). An analogue of direct and inverse transform exists for $B_2$-bundles (\ref{_dire_inve_q-K_Theorem_}). Most importantly, T. Nitta proved that on a quaternionic-K\"ahler manifold of positive scalar curvature a twistor transform of a $B_2$-bundle is a Yang-Mills bundle on $\operatorname{Tw}(Q)$ (\ref{_twi_tra_YM_q-K_Theorem_}). This implies that a twistor transform of a Hermitian $B_2$-bundle is a direct sum of stable bundles with $\deg c_1 =0$. In the situation described in the beginning of this Subsection, we have a manifold $C= {\Bbb P} T_x M \cong {\Bbb C} P ^{2n-1}$ which is a twistor space of a quaternionic projective space \[ {\Bbb P}_{\Bbb H}T_x M = \bigg(T_x M\backslash 0\bigg)/{\Bbb H}^* \cong {\Bbb H} P^n. \] To prove that $\pi^* F\restrict C$ is stable, we need to show that $\pi^* F\restrict C$ is obtained as twistor transform of some Hermitian $B_2$-bundle on ${\Bbb P}_{\Bbb H}T_x M$. This is done using an equivalence between the category of qua\-ter\-ni\-onic-\-K\"ah\-ler manifolds of positive scalar curvature and the category of hyperk\"ahler manifolds equipped with a special type of ${\Bbb H}^*$-action, constructed by A. Swann (\cite{_Swann_}). Given a quaternionic-K\"ahler manifold $Q$, we consider a principal bundle $\c U(Q)$ consisting of all quaternion frames on $Q$ (\ref{_specia_and_q-K-Subsection_}). Then $\c U(Q)$ is fibered over $Q$ with a fiber ${\Bbb H}/\{\pm 1\}$. It is easy to show that $\c U(Q)$ is equipped with an action of quaternion algebra in its tangent bundle. A. Swann proved that if $Q$ has with positive scalar curvature, then this action of quaternion algebra comes from a hyperk\"ahler structure on $\c U(M)$ (\ref{_U(Q)_hyperk_Theorem_}). The correspondence $Q{\:\longrightarrow\:} \c U(Q)$ induces an equivalence of appropriately defined categories (\ref{_U(Q)_equiva_cate_Theorem_}). We call this construction {\bf Swann's formalism}. The twistor space $\operatorname{Tw}(\c U(Q))$ of the hyperk\"ahler manifold $\c U(Q)$ is equ\-ipped with a holomorphic action of ${\Bbb C}^*$. Every $B_2$-bundle corresponds to a ${\Bbb C}^*$-invariant holomorphic bundle on $\operatorname{Tw}(\c U(Q))$ and this correspondence induces an equivalence of appropriately defined categories, called {\bf Swann's formalism for budnles} (\ref{_B_2_to_C^*_equiva_Theorem_}). Applying this equivalence to the ${\Bbb C}^*$-equivariant sheaf obtained as an associate graded sheaf of a hyperholomorphic sheaf, we obtain a $B_2$ bundle on ${\Bbb P}_{\Bbb H}T_x M$, and $\pi^* F\restrict C$ is obtained from this $B_2$-bundle by a twistor transform. The correspondence between $B_2$-bundles on $Q$ and ${\Bbb C}^*$-invariant holomorphic bundles on $\operatorname{Tw}(\c U(Q))$ is an interesting geometric phenomenon which is of independent interest. We construct it by reduction to $\dim Q=0$, where it follows from an explicit calculation involving 2-forms over a flat manifold of real dimension 4. \subsection{Contents} The paper is organized as follows. \begin{itemize} \item Section \ref{_intro_Section_} is an introduction. It is independent from the rest of this apper. \item Section \ref{_basics_hyperka_Section_} is an introduction to the theory of hyperk\"ahler manifolds. We give a compenduum of results from hyperkaehler geometry which are due to F. Bogomolov (\cite{_Bogomolov_}) and A. Beauville (\cite{_Beauville_}), and give an introduction to the results of \cite{_Verbitsky:Hyperholo_bundles_}, \cite{_Verbitsky:hypercomple_}, \cite{_Verbitsky:Symplectic_II_}. \item Section \ref{_hyperho_shea_Section_} contains a definition and basic properties of hyperholomorphic sheaves. We prove that a stable hyperholomorphic sheaf admits a hyperholomorphic connection, and conversely, a reflexive sheaf admitting a hyperholomorphic connection is stable hyperholomorphic (\ref{_hyperho_conne_exi_Theorem_}). This equivalence is constructed using Bando-Siu theory of Yang--Mills connections on coherent sheaves. We prove an analogue of Wirtinger's inequality for stable sheaves (\ref{_stable_shea_degree_Corollary_}), which states that for any induced complex structure $J\neq \pm I$, and any stable reflexive sheaf $F$ on $(M,I)$, we have \[ \deg_I\left(2c_2(F) - \frac{r-1}{r} c_1(F)^2\right) \geq \left|\deg_J\left(2c_2(F) - \frac{r-1}{r} c_1(F)^2\right)\right|, \] and the equality holds if and only if $F$ is hyperholomorphic. \item Section \ref{_cohomo_hype_Section_} contains the preliminary material used for the study of $C$-restricted complex structures in Section \ref{_C_restri_Section_}. We give an exposition of various algebraic structures on the cohomology of a hyperk\"ahler manifold, which were discovered in \cite{_so(5)_} and \cite{_Verbitsky:cohomo_}. In the last Subsection, we apply the Wirtinger's inequality to prove that the fundamental classes of complex subvarieties and $c_2$ of stable reflexive sheaves satisfy a certain set of axioms. Cohomology classes satisfying these axioms are called CA-classes. This definition simplifies the work on $C$-restricted complex structures in Section \ref{_C_restri_Section_}. \item In Section \ref{_C_restri_Section_} we define $C$-restricted complex structures and prove the following. Consider a compact hyperk\"ahler manifold and an $SU(2)$-invariant class $a\in H^4(M)$. Then for all $C$-restricted complex structures $I$, with $C> \deg a$, and all semistable sheaves $I$ on $(M, I)$ with $c_2(F) =a$, the sheaf $F$ is hyperholomorphic (\ref{_sheaf_on_C_restr_hyperho_Theorem_}). This is used to show that a deformation of a hyperholomorphic sheaf is again hyperholomorphic, over $(M,I)$ with $I$ a $C$-restricted complex structure, $c> \deg c_2(F)$. We define the moduli space of hyperk\"ahler structures, and show that for a dense set $\c C$ of hyperk\"ahler structures, all $\c H \in \c C$ admit a dense set of $C$-induced complex structures, for all $C\in {\Bbb R}$ (\ref{_C_restri_dense_Theorem_}). \item In Section \ref{_desingu_she_Section_} we give a proof of Desingularization Theorem for stable reflexive hyperholomorphic sheaves with isolated singularities (\ref{_desingu_hyperho_Theorem_}). We study the natural ${\Bbb C}^*$-action on a local ring of a hyperk\"ahler manifold defined in \cite{_Verbitsky:DesinguII_}. We show that a sheaf $F$ admitting a hyperholomorphic connection is equivariant with respect to this ${\Bbb C}^*$-action. Then $F$ can be desingularized by a blow-up, because any ${\Bbb C}^*$-equivariant sheaf with an isolated singularily can be desingularized by a blow-up (\ref{_desingu_C^*_equi_Proposition_}). \item Section \ref{_twisto_tra_Section_} is a primer on twistor transform and quaternionic-K\"ahler geometry. We give an exposition of the works of A. Swann (\cite{_Swann_}), T. Nitta (\cite{_Nitta:bundles_}, \cite{_Nitta:Y-M_}) on quaternionic-K\"ahler manifolds and explain the direct and inverse twistor transform over hyperk\"ahler and qua\-ter\-ni\-onic-\--K\"ah\-ler manifolds. \item Section \ref{_C_equiv_twi_spa_Section_} gives a correspondence between $B_2$-bundles on a qua\-ter\-ni\-onic-\--K\"ah\-ler manifold, and ${\Bbb C}^*$-equivariant holomorphic bundles on the twistor space of the corresponding hyperk\"ahler manifold constructed by A. Swann. This is called ``Swann's formalism for vector bundles''. We use this correspondence to prove that an associate graded sheaf of a hyperholomorphic sheaf is equipped with a natural connection which is compatible with quaternions. This implies polystability of the bundle $\pi^* F\restrict C$ (see Subsection \ref{_Swann's_intro_Subsection_}). \item In Section \ref{_modu_hyperho_Section_}, we use the polystability of the bundle $\pi^* F\restrict C$ to show that a hyperholomorphic deformation of a hyperholomorphic bundle is again a bundle. Together with results on $C$-restricted complex structures and Maruyama's compactification (\cite{_Maruyama:Si_}), this implies that the moduli of semistable bundles are compact, under conditions of $C$-restrictness and non-existence of trianalytic subvarieties (\ref{_space_semista_bu_compa_Theorem_}). \item In Section \ref{_new_exa_Section_}, we apply these results to the hyperk\"ahler geometry. Using the desingularization theorem for singular hyperk\"ahler manifolds (\ref{_desingu_Theorem_}), we prove that the moduli of stable deformations of a hyperholomorphic bundle has a compact hyperk\"ahler desingularization (\ref{_space_sta_bu_compa_hyperka_Theorem_}). We give an exposition of the theory of twistor paths, which allows one to identify the categories of stable bundles for different K\"ahler structures on the same hyperk\"ahler manifold (\ref{_admi_twi_impli_Theorem_}). These results allow one to weaken the conditions necessary for compactness of the moduli spaces of vector bundles. Finally, we give a conjectural exposition of how these results can be used to obtain new examples of compact hyperk\"ahler manifolds. \end{itemize} \section{Hyperk\"ahler manifolds} \label{_basics_hyperka_Section_} \subsection{Hyperk\"ahler manifolds} \label{_hyperka_Subsection_} This subsection contains a compression of the basic and best known results and definitions from hyperk\"ahler geometry, found, for instance, in \cite{_Besse:Einst_Manifo_} or in \cite{_Beauville_}. \hfill \definition \label{_hyperkahler_manifold_Definition_} (\cite{_Besse:Einst_Manifo_}) A {\bf hyperk\"ahler manifold} is a Riemannian manifold $M$ endowed with three complex structures $I$, $J$ and $K$, such that the following holds. \begin{description} \item[(i)] the metric on $M$ is K\"ahler with respect to these complex structures and \item[(ii)] $I$, $J$ and $K$, considered as endomorphisms of a real tangent bundle, satisfy the relation $I\circ J=-J\circ I = K$. \end{description} \hfill The notion of a hyperk\"ahler manifold was introduced by E. Calabi (\cite{_Calabi_}). \hfill Clearly, a hyperk\"ahler manifold has a natural action of the quaternion algebra ${\Bbb H}$ in its real tangent bundle $TM$. Therefore its complex dimension is even. For each quaternion $L\in \Bbb H$, $L^2=-1$, the corresponding automorphism of $TM$ is an almost complex structure. It is easy to check that this almost complex structure is integrable (\cite{_Besse:Einst_Manifo_}). \hfill \definition \label{_indu_comple_str_Definition_} Let $M$ be a hyperk\"ahler manifold, and $L$ a quaternion satisfying $L^2=-1$. The corresponding complex structure on $M$ is called {\bf an induced complex structure}. The $M$, considered as a K\"ahler manifold, is denoted by $(M, L)$. In this case, the hyperk\"ahler structure is called {\bf combatible with the complex structure $L$}. Let $M$ be a compact complex variety. We say that $M$ is {\bf of hyperk\"ahler type} if $M$ admits a hyperk\"ahler structure compatible with the complex structure. \hfill \hfill \definition \label{_holomorphi_symple_Definition_} Let $M$ be a complex manifold and $\Theta$ a closed holomorphic 2-form over $M$ such that $\Theta^n=\Theta\wedge\Theta\wedge...$, is a nowhere degenerate section of a canonical class of $M$ ($2n=dim_{\Bbb C}(M)$). Then $M$ is called {\bf holomorphically symplectic}. \hfill Let $M$ be a hyperk\"ahler manifold; denote the Riemannian form on $M$ by $<\cdot,\cdot>$. Let the form $\omega_I := <I(\cdot),\cdot>$ be the usual K\"ahler form which is closed and parallel (with respect to the Levi-Civitta connection). Analogously defined forms $\omega_J$ and $\omega_K$ are also closed and parallel. A simple linear algebraic consideration (\cite{_Besse:Einst_Manifo_}) shows that the form $\Theta:=\omega_J+\sqrt{-1}\omega_K$ is of type $(2,0)$ and, being closed, this form is also holomorphic. Also, the form $\Theta$ is nowhere degenerate, as another linear algebraic argument shows. It is called {\bf the canonical holomorphic symplectic form of a manifold M}. Thus, for each hyperk\"ahler manifold $M$, and an induced complex structure $L$, the underlying complex manifold $(M,L)$ is holomorphically symplectic. The converse assertion is also true: \hfill \theorem \label{_symplectic_=>_hyperkahler_Proposition_} (\cite{_Beauville_}, \cite{_Besse:Einst_Manifo_}) Let $M$ be a compact holomorphically symplectic K\"ahler manifold with the holomorphic symplectic form $\Theta$, a K\"ahler class $[\omega]\in H^{1,1}(M)$ and a complex structure $I$. Let $n=\dim_{\Bbb C} M$. Assume that $\int_M \omega^n = \int_M (Re \Theta)^n$. Then there is a unique hyperk\"ahler structure $(I,J,K,(\cdot,\cdot))$ over $M$ such that the cohomology class of the symplectic form $\omega_I=(\cdot,I\cdot)$ is equal to $[\omega]$ and the canonical symplectic form $\omega_J+\sqrt{-1}\:\omega_K$ is equal to $\Theta$. \hfill \ref{_symplectic_=>_hyperkahler_Proposition_} follows from the conjecture of Calabi, pro\-ven by Yau (\cite{_Yau:Calabi-Yau_}). \blacksquare \hfill Let $M$ be a hyperk\"ahler manifold. We identify the group $SU(2)$ with the group of unitary quaternions. This gives a canonical action of $SU(2)$ on the tangent bundle, and all its tensor powers. In particular, we obtain a natural action of $SU(2)$ on the bundle of differential forms. \hfill \lemma \label{_SU(2)_commu_Laplace_Lemma_} The action of $SU(2)$ on differential forms commutes with the Laplacian. {\bf Proof:} This is Proposition 1.1 of \cite{_Verbitsky:Symplectic_II_}. \blacksquare Thus, for compact $M$, we may speak of the natural action of $SU(2)$ in cohomology. \hfill Further in this article, we use the following statement. \hfill \lemma \label{_SU(2)_inva_type_p,p_Lemma_} Let $\omega$ be a differential form over a hyperk\"ahler manifold $M$. The form $\omega$ is $SU(2)$-invariant if and only if it is of Hodge type $(p,p)$ with respect to all induced complex structures on $M$. {\bf Proof:} This is \cite{_Verbitsky:Hyperholo_bundles_}, Proposition 1.2. \blacksquare \subsection{Simple hyperk\"ahler manifolds} \definition \label{_simple_hyperkahler_mfolds_Definition_} (\cite{_Beauville_}) A connected simply connected compact hy\-per\-k\"ah\-ler manifold $M$ is called {\bf simple} if $M$ cannot be represented as a product of two hyperk\"ahler manifolds: \[ M\neq M_1\times M_2,\ \text{where} \ dim\; M_1>0 \ \ \text{and} \ dim\; M_2>0 \] Bogomolov proved that every compact hyperk\"ahler manifold has a finite covering which is a product of a compact torus and several simple hyperk\"ahler manifolds. Bogomolov's theorem implies the following result (\cite{_Beauville_}): \hfill \theorem\label{_simple_mani_crite_Theorem_} Let $M$ be a compact hyperk\"ahler manifold. Then the following conditions are equivalent. \begin{description} \item[(i)] $M$ is simple \item[(ii)] $M$ satisfies $H^1(M, {\Bbb R}) =0$, $H^{2,0}(M) ={\Bbb C}$, where $H^{2,0}(M)$ is the space of $(2,0)$-classes taken with respect to any of induced complex structures. \end{description} \blacksquare \subsection{Trianalytic subvarieties in hyperk\"ahler manifolds.} In this subsection, we give a definition and basic properties of trianalytic subvarieties of hyperk\"ahler manifolds. We follow \cite{_Verbitsky:Symplectic_II_}, \cite{_Verbitsky:DesinguII_}. \hfill Let $M$ be a compact hyperk\"ahler manifold, $\dim_{\Bbb R} M =2m$. \hfill \definition\label{_trianalytic_Definition_} Let $N\subset M$ be a closed subset of $M$. Then $N$ is called {\bf trianalytic} if $N$ is a complex analytic subset of $(M,L)$ for any induced complex structure $L$. \hfill \hfill Let $I$ be an induced complex structure on $M$, and $N\subset(M,I)$ be a closed analytic subvariety of $(M,I)$, $dim_{\Bbb C} N= n$. Consider the homology class represented by $N$. Let $[N]\in H^{2m-2n}(M)$ denote the Poincare dual cohomology class, so called {\bf fundamental class} of $N$. Recall that the hyperk\"ahler structure induces the action of the group $SU(2)$ on the space $H^{2m-2n}(M)$. \hfill \theorem\label{_G_M_invariant_implies_trianalytic_Theorem_} Assume that $[N]\in H^{2m-2n}(M)$ is invariant with respect to the action of $SU(2)$ on $H^{2m-2n}(M)$. Then $N$ is trianalytic. {\bf Proof:} This is Theorem 4.1 of \cite{_Verbitsky:Symplectic_II_}. \blacksquare \hfill The following assertion is the key to the proof of \ref{_G_M_invariant_implies_trianalytic_Theorem_} (see \cite{_Verbitsky:Symplectic_II_} for details). \hfill \proposition \label{_Wirti_hyperka_Proposition_} (Wirtinger's inequality) Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure and $X\subset (M, I)$ a closed complex subvariety for complex dimension $k$. Let $J$ be an induced complex structure, $J \neq \pm I$, and $\omega_I$, $\omega_J$ the associated K\"ahler forms. Consider the numbers \[ \deg_I X:= \int_X \omega_I^k, \ \ \deg_J X:= \int_X \omega_J^k \] Then $\deg_I X\geq |\deg_J X|$, and the inequality is strict unless $X$ is trianalytic. \blacksquare \hfill \remark \label{_triana_dim_div_4_Remark_} Trianalytic subvarieties have an action of quaternion algebra in the tangent bundle. In particular, the real dimension of such subvarieties is divisible by 4. \hfill \definition \label{_generic_manifolds_Definition_} Let $M$ be a complex manifold admitting a hyperk\"ahler structure $\c H$. We say that $M$ is {\bf of general type} or {\bf generic} with respect to $\c H$ if all elements of the group \[ \bigoplus\limits_p H^{p,p}(M)\cap H^{2p}(M,{\Bbb Z})\subset H^*(M)\] are $SU(2)$-invariant. We say that $M$ is {\bf Mumford--Tate generic} if for all $n\in {\Bbb Z}^{>0}$, all the cohomology classes \[ \alpha \in \bigoplus\limits_p H^{p,p}(M^n)\cap H^{2p}(M^n,{\Bbb Z})\subset H^*(M^n) \] are $SU(2)$-invariant. In other words, $M$ is Mumford--Tate generic if for all $n\in {\Bbb Z}^{>0}$, the $n$-th power $M^n$ is generic. Clearly, Mumford--Tate generic implies generic. \hfill \proposition \label{_generic_are_dense_Proposition_} Let $M$ be a compact manifold, $\c H$ a hyperk\"ahler structure on $M$ and $S$ be the set of induced complex structures over $M$. Denote by $S_0\subset S$ the set of $L\in S$ such that $(M,L)$ is Mumford-Tate generic with respect to $\c H$. Then $S_0$ is dense in $S$. Moreover, the complement $S\backslash S_0$ is countable. {\bf Proof:} This is Proposition 2.2 from \cite{_Verbitsky:Symplectic_II_} \blacksquare \hfill \ref{_G_M_invariant_implies_trianalytic_Theorem_} has the following immediate corollary: \corollary \label{_hyperkae_embeddings_Corollary_} Let $M$ be a compact holomorphically symplectic manifold. Assume that $M$ is of general type with respect to a hyperk\"ahler structure $\c H$. Let $S\subset M$ be closed complex analytic subvariety. Then $S$ is trianalytic with respect to $\c H$. \blacksquare \hfill In \cite{_Verbitsky:hypercomple_}, \cite{_Verbitsky:Desingu_}, \cite{_Verbitsky:DesinguII_}, we gave a number of equivalent definitions of a singular hyperk\"ahler and hypercomplex variety. We refer the reader to \cite{_Verbitsky:DesinguII_} for the precise definition; for our present purposes it suffices to say that all trianalytic subvarieties are hyperk\"ahler varieties. The following Desingularization Theorem is very useful in the study of trianalytic subvarieties. \hfill \theorem \label{_desingu_Theorem_} (\cite{_Verbitsky:DesinguII_}) Let $M$ be a hyperk\"ahler or a hypercomplex variety, $I$ an induced complex structure. Let \[ \widetilde{(M, I)}\stackrel n{\:\longrightarrow\:} (M,I)\] be the normalization of $(M,I)$. Then $\widetilde{(M, I)}$ is smooth and has a natural hyperk\"ahler structure $\c H$, such that the associated map $n:\; \widetilde{(M, I)} {\:\longrightarrow\:} (M,I)$ agrees with $\c H$. Moreover, the hyperk\"ahler manifold $\tilde M:= \widetilde{(M, I)}$ is independent from the choice of induced complex structure $I$. \blacksquare \hfill Let $M$ be a K3 surface, and $M^{[n]}$ be a Hilbert scheme of points on $M$. Then $M^{[n]}$ admits a hyperk\"ahler structure (\cite{_Beauville_}). In \cite{_Verbitsky:Hilbert_}, we proved the following theorem. \hfill \theorem\label{_no_triana_subva_of_Hilb_Theorem_} Let $M$ be a complex K3 surface without automorphisms. Assume that $M$ is Mumford-Tate generic with respect to some hyperka\"hler structure. Consider the Hilbert scheme $M^{[n]}$ of points on $M$. Pick a hyperk\"ahler structure on $M^{[n]}$ which is compatible with the complex structure. Then $M^{[n]}$ has no proper trianalytic subvarieties. \blacksquare \hfill \subsection{Hyperholomorphic bundles} \label{_hyperholo_Subsection_} This subsection contains several versions of a definition of hyperholomorphic connection in a complex vector bundle over a hyperk\"ahler manifold. We follow \cite{_Verbitsky:Hyperholo_bundles_}. \hfill Let $B$ be a holomorphic vector bundle over a complex manifold $M$, $\nabla$ a connection in $B$ and $\Theta\in\Lambda^2\otimes End(B)$ be its curvature. This connection is called {\bf compatible with a holomorphic structure} if $\nabla_X(\zeta)=0$ for any holomorphic section $\zeta$ and any antiholomorphic tangent vector field $X\in T^{0,1}(M)$. If there exists a holomorphic structure compatible with the given Hermitian connection then this connection is called {\bf integrable}. \hfill One can define a {\bf Hodge decomposition} in the space of differential forms with coefficients in any complex bundle, in particular, $End(B)$. \hfill \theorem \label{_Newle_Nie_for_bu_Theorem_} Let $\nabla$ be a Hermitian connection in a complex vector bundle $B$ over a complex manifold. Then $\nabla$ is integrable if and only if $\Theta\in\Lambda^{1,1}(M, \operatorname{End}(B))$, where $\Lambda^{1,1}(M, \operatorname{End}(B))$ denotes the forms of Hodge type (1,1). Also, the holomorphic structure compatible with $\nabla$ is unique. {\bf Proof:} This is Proposition 4.17 of \cite{_Kobayashi_}, Chapter I. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill This result has the following more general version: \hfill \proposition \label{_Newle_Nie_for_NH_bu_Proposition_} Let $\nabla$ be an arbitrary (not necessarily Hermitian) connection in a complex vector bundle $B$. Then $\nabla$ is integrable if and only its $(0,1)$-part has square zero. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill This proposition is a version of Newlander-Nirenberg theorem. For vector bundles, it was proven by Atiyah and Bott. \hfill \definition \label{_hyperho_conne_Definition_} Let $B$ be a Hermitian vector bundle with a connection $\nabla$ over a hyperk\"ahler manifold $M$. Then $\nabla$ is called {\bf hyperholomorphic} if $\nabla$ is integrable with respect to each of the complex structures induced by the hyperk\"ahler structure. As follows from \ref{_Newle_Nie_for_bu_Theorem_}, $\nabla$ is hyperholomorphic if and only if its curvature $\Theta$ is of Hodge type (1,1) with respect to any of complex structures induced by a hyperk\"ahler structure. As follows from \ref{_SU(2)_inva_type_p,p_Lemma_}, $\nabla$ is hyperholomorphic if and only if $\Theta$ is a $SU(2)$-invariant differential form. \hfill \example \label{_tangent_hyperholo_Example_} (Examples of hyperholomorphic bundles) \begin{description} \item[(i)] Let $M$ be a hyperk\"ahler manifold, and $TM$ be its tangent bundle equi\-p\-ped with the Levi--Civita connection $\nabla$. Consider a complex structure on $TM$ induced from the quaternion action. Then $\nabla$ is a Hermitian connection which is integrable with respect to each induced complex structure, and hence, is Yang--Mills. \item[(ii)] For $B$ a hyperholomorphic bundle, all its tensor powers are also hyperholomorphic. \item[(iii)] Thus, the bundles of differential forms on a hyperk\"ahler manifold are also hyperholomorphic. \end{description} \subsection{Stable bundles and Yang--Mills connections.} \label{_sta_bu_and_YM_Subsection_} This subsection is a compendium of the most basic results and definitions from the Yang--Mills theory over K\"ahler manifolds, concluding in the fundamental theorem of Uhlenbeck--Yau \cite{_Uhle_Yau_}. \hfill \definition\label{_degree,slope_destabilising_Definition_} Let $F$ be a coherent sheaf over an $n$-dimensional compact K\"ahler manifold $M$. We define $\deg(F)$ as \[ \deg(F)=\int_M\frac{ c_1(F)\wedge\omega^{n-1}}{vol(M)} \] and $\text{slope}(F)$ as \[ \text{slope}(F)=\frac{1}{\text{rank}(F)}\cdot \deg(F). \] The number $\text{slope}(F)$ depends only on a cohomology class of $c_1(F)$. Let $F$ be a coherent sheaf on $M$ and $F'\subset F$ its proper subsheaf. Then $F'$ is called {\bf destabilizing subsheaf} if $\text{slope}(F') \geq \text{slope}(F)$ A coherent sheaf $F$ is called {\bf stable} \footnote{In the sense of Mumford-Takemoto} if it has no destabilizing subsheaves. A coherent sheaf $F$ is called {\bf semistable} if for all destabilizing subsheaves $F'\subset F$, we have $\text{slope}(F') = \text{slope}(F)$. \hfill Later on, we usually consider the bundles $B$ with $deg(B)=0$. \hfill Let $M$ be a K\"ahler manifold with a K\"ahler form $\omega$. For differential forms with coefficients in any vector bundle there is a Hodge operator $L: \eta{\:\longrightarrow\:}\omega\wedge\eta$. There is also a fiberwise-adjoint Hodge operator $\Lambda$ (see \cite{_Griffi_Harri_}). \hfill \definition \label{Yang-Mills_Definition_} Let $B$ be a holomorphic bundle over a K\"ahler manifold $M$ with a holomorphic Hermitian connection $\nabla$ and a curvature $\Theta\in\Lambda^{1,1}\otimes End(B)$. The Hermitian metric on $B$ and the connection $\nabla$ defined by this metric are called {\bf Yang-Mills} if \[ \Lambda(\Theta)=constant\cdot \operatorname{Id}\restrict{B}, \] where $\Lambda$ is a Hodge operator and $\operatorname{Id}\restrict{B}$ is the identity endomorphism which is a section of $End(B)$. Further on, we consider only these Yang--Mills connections for which this constant is zero. \hfill A holomorphic bundle is called {\bf indecomposable} if it cannot be decomposed onto a direct sum of two or more holomorphic bundles. \hfill The following fundamental theorem provides examples of Yang-\--Mills \linebreak bundles. \theorem \label{_UY_Theorem_} (Uhlenbeck-Yau) Let B be an indecomposable holomorphic bundle over a compact K\"ahler manifold. Then $B$ admits a Hermitian Yang-Mills connection if and only if it is stable, and this connection is unique. {\bf Proof:} \cite{_Uhle_Yau_}. \blacksquare \hfill \proposition \label{_hyperholo_Yang--Mills_Proposition_} Let $M$ be a hyperk\"ahler manifold, $L$ an induced complex structure and $B$ be a complex vector bundle over $(M,L)$. Then every hyperholomorphic connection $\nabla$ in $B$ is Yang-Mills and satisfies $\Lambda(\Theta)=0$ where $\Theta$ is a curvature of $\nabla$. \hfill {\bf Proof:} We use the definition of a hyperholomorphic connection as one with $SU(2)$-invariant curvature. Then \ref{_hyperholo_Yang--Mills_Proposition_} follows from the \hfill \lemma \label{_Lambda_of_inva_forms_zero_Lemma_} Let $\Theta\in \Lambda^2(M)$ be a $SU(2)$-invariant differential 2-form on $M$. Then $\Lambda_L(\Theta)=0$ for each induced complex structure $L$.\footnote{By $\Lambda_L$ we understand the Hodge operator $\Lambda$ associated with the K\"ahler complex structure $L$.} {\bf Proof:} This is Lemma 2.1 of \cite{_Verbitsky:Hyperholo_bundles_}. \blacksquare \hfill Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure. For any stable holomorphic bundle on $(M, I)$ there exists a unique Hermitian Yang-Mills connection which, for some bundles, turns out to be hyperholomorphic. It is possible to tell when this happens. \hfill \theorem \label{_inva_then_hyperho_Theorem_} Let $B$ be a stable holomorphic bundle over $(M,I)$, where $M$ is a hyperk\"ahler manifold and $I$ is an induced complex structure over $M$. Then $B$ admits a compatible hyperholomorphic connection if and only if the first two Chern classes $c_1(B)$ and $c_2(B)$ are $SU(2)$-invariant.\footnote{We use \ref{_SU(2)_commu_Laplace_Lemma_} to speak of action of $SU(2)$ in cohomology of $M$.} {\bf Proof:} This is Theorem 2.5 of \cite{_Verbitsky:Hyperholo_bundles_}. \blacksquare \subsection{Twistor spaces} Let $M$ be a hyperk\"ahler manifold. Consider the product manifold $X = M \times S^2$. Embed the sphere $S^2 \subset {\Bbb H}$ into the quaternion algebra ${\Bbb H}$ as the subset of all quaternions $J$ with $J^2 = -1$. For every point $x = m \times J \in X = M \times S^2$ the tangent space $T_xX$ is canonically decomposed $T_xX = T_mM \oplus T_JS^2$. Identify $S^2 = {\Bbb C} P^1$ and let $I_J:T_JS^2 \to T_JS^2$ be the complex structure operator. Let $I_m:T_mM \to T_mM$ be the complex structure on $M$ induced by $J \in S^2 \subset {\Bbb H}$. The operator $I_x = I_m \oplus I_J:T_xX \to T_xX$ satisfies $I_x \circ I_x = -1$. It depends smoothly on the point $x$, hence defines an almost complex structure on $X$. This almost complex structure is known to be integrable (see \cite{_Salamon_}). \hfill \definition\label{_twistor_Definition_} The complex manifold $\langle X, I_x \rangle$ is called {\it the twistor space} for the hyperk\"ahler manifold $M$, denoted by $\operatorname{Tw}(M)$. This manifold is equipped with a real analytic projection $\sigma:\; \operatorname{Tw}(M){\:\longrightarrow\:} M$ and a complex analytic projection $\pi:\; \operatorname{Tw}(M) {\:\longrightarrow\:} {\Bbb C} P^1$. \hfill The twistor space $\operatorname{Tw}(M)$ is not, generally speaking, a K\"ahler manifold. For $M$ compact, it is easy to show that $\operatorname{Tw}(M)$ does not admit a K\"ahler metric. We consider $\operatorname{Tw}(M)$ as a Hermitian manifold with the product metric. \hfill \definition \label{_Li_Yau_condi_Definition_} Let $X$ be an $n$-dimensional Hermitian manifold and let $\sqrt{-1}\omega$ be the imaginary part of the metric on $X$. Thus $\omega$ is a real $(1,1)$-form. Assume that the form $\omega$ satisfies the following condition of Li and Yau (\cite{_Li_Yau_}). \begin{equation}\label{_Li_Yau_condi_Equation_} \omega^{n-2} \wedge d\omega = 0. \end{equation} Such Hermitian metrics are called {\bf metrics satisfying the condition of Li--Yau}. For a closed real $2$-form $\eta$ let $$ \deg\eta = \int_X \omega^{n-1} \wedge \eta. $$ The condition \eqref{_Li_Yau_condi_Equation_} ensures that $\deg\eta$ depends only on the cohomology class of $\eta$. Thus it defines a degree functional $\deg:H^2(X,{\Bbb R}) \to {\Bbb R}$. This functional allows one to repeat verbatim the Mumford-Takemoto definitions of stable and semistable bundles in this more general situation. Moreover, the Hermitian Yang-Mills equations also carry over word-by-word. Li and Yau proved a version of Uhlenbeck--Yau theorem in this situation (\cite{_Li_Yau_}; see also \ref{_UY_for_shea_Theorem_}). \hfill \proposition Let $M$ be a hyperk\"ahler manifold and $\operatorname{Tw}(M)$ its twistor space, considered as a Hermitian manifold. Then $\operatorname{Tw}(M)$ satisfies the conditions of Li--Yau. {\bf Proof:} \cite{_NHYM_}, Proposition 4.5. \blacksquare \section{Hyperholomorphic sheaves} \label{_hyperho_shea_Section_} \subsection{Stable sheaves and Yang-Mills connections} In \cite{_Bando_Siu_}, S. Bando and Y.-T. Siu developed the machinery allowing one to apply the methods of Yang-Mills theory to torsion-free coherent sheaves. In the course of this paper, we apply their work to generalise the results of \cite{_Verbitsky:Hyperholo_bundles_}. In this Subsection, we give a short exposition of their results. \hfill \definition\label{_refle_Definition_} Let $X$ be a complex manifold, and $F$ a coherent sheaf on $X$. Consider the sheaf $F^*:= \c Hom_{{\cal O}_X}(F, {\cal O}_X)$. There is a natural functorial map $\rho_F:\; F {\:\longrightarrow\:} F^{**}$. The sheaf $F^{**}$ is called {\bf a reflexive hull}, or {\bf reflexization} of $F$. The sheaf $F$ is called {\bf reflexive} if the map $\rho_F:\; F {\:\longrightarrow\:} F^{**}$ is an isomorphism. \hfill \remark For all coherent sheaves $F$, the map $\rho_{F^*}:\; F^* {\:\longrightarrow\:} F^{***}$ is an isomorphism (\cite{_OSS_}, Ch. II, the proof of Lemma 1.1.12). Therefore, a reflexive hull of a sheaf is always reflexive. \hfill \claim Let $X$ be a K\"ahler manifold, and $F$ a torsion-free coherent sheaf over $X$. Then $F$ (semi)stable if and only if $F^{**}$ is (semi)stable. {\bf Proof:} This is \cite{_OSS_}, Ch. II, Lemma 1.2.4. \blacksquare \hfill \definition Let $X$ be a K\"ahler manifold, and $F$ a coherent sheaf over $X$. The sheaf $F$ is called {\bf polystable} if $F$ is a direct sum of stable sheaves. \hfill The admissible Hermitian metrics, introduced by Bando and Siu in \cite{_Bando_Siu_}, play the role of the ordinary Hermitian metrics for vector bundles. \hfill Let $X$ be a K\"ahler manifold. In Hodge theory, one considers the operator $\Lambda:\; \Lambda^{p, q}(X) {\:\longrightarrow\:}\Lambda^{p-1, q-1}(X)$ acting on differential forms on $X$, which is adjoint to the multiplication by the K\"ahler form. This operator is defined on differential forms with coefficient in every bundle. Considering a curvature $\Theta$ of a bundle $B$ as a 2-form with coefficients in $\operatorname{End}(B)$, we define the expression $\Lambda\Theta$ which is a section of $\operatorname{End}(B)$. \hfill \definition \label{_admi_metri_Definition_} Let $X$ be a K\"ahler manifold, and $F$ a reflexive coherent sheaf over $X$. Let $U\subset X$ be the set of all points at which $F$ is locally trivial. By definition, the restriction $F\restrict U$ of $F$ to $U$ is a bundle. An {\bf admissible metric} on $F$ is a Hermitian metric $h$ on the bundle $F\restrict U$ which satisfies the following assumptions \begin{description} \item[(i)] the curvature $\Theta$ of $(F, h)$ is square integrable, and \item[(ii)] the corresponding section $\Lambda \Theta\in \operatorname{End}(F\restrict U)$ is uniformly bounded. \end{description} \hfill \definition \label{_Yang-Mills_sheaves_Definition_} Let $X$ be a K\"ahler manifold, $F$ a reflexive coherent sheaf over $X$, and $h$ an admissible metric on $F$. Consider the corresponding Hermitian connection $\nabla$ on $F\restrict U$. The metric $h$ and the connection $\nabla$ are called {\bf Yang-Mills} if its curvature satisfies \[ \Lambda \Theta\in \operatorname{End}(F\restrict U) = c\cdot \operatorname{\text{\sf id}} \] where $c$ is a constant and $\operatorname{\text{\sf id}}$ the unit section $\operatorname{\text{\sf id}} \in \operatorname{End}(F\restrict U)$. \hfill Further in this paper, we shall only consider Yang-Mills connections with $\Lambda \Theta=0$. \hfill \remark By Gauss-Bonnet formule, the constant $c$ is equal to $\deg(F)$, where $\deg(F)$ is the degree of $F$ (\ref{_degree,slope_destabilising_Definition_}). \hfill One of the main results of \cite{_Bando_Siu_} is the following analogue of Uhlenbeck--Yau theorem (\ref{_UY_Theorem_}). \hfill \theorem\label{_UY_for_shea_Theorem_} Let $M$ be a compact K\"ahler manifold, or a compact Hermitian manifold satisfying conditions of Li-Yau (\ref{_Li_Yau_condi_Definition_}), and $F$ a coherent sheaf without torsion. Then $F$ admits an admissible Yang--Mills metric is and only if $F$ is polystable. Moreover, if $F$ is stable, then this metric is unique, up to a constant multiplier. {\bf Proof:} In \cite{_Bando_Siu_}, \ref{_UY_for_shea_Theorem_} is proved for K\"ahler $M$ (\cite{_Bando_Siu_}, Theorem 3). It is easy to adapt this proof for Hermitian manifolds satisfying conditions of Li--Yau. \blacksquare \hfill \remark Clearly, the connection, corresponding to a metric on $F$, does not change when the metric is multiplied by a scalar. The Yang--Mills metric on a polystable sheaf is unique up to a componentwise multiplication by scalar multipliers. Thus, the Yang--Mills connection of \ref{_UY_for_shea_Theorem_} is unique. \hfill Another important theorem of \cite{_Bando_Siu_} is the following. \hfill \theorem\label{_YM_can_be_exte_Theorem_} Let $(F, h)$ be a holomorphic vector bundle with a Hermitian metric $h$ defined on a K\"ahler manifold $X$ (not necessary compact nor complete) outside a closed subset $S$ with locally finite Hausdorff measure of real co-dimension $4$. Assume that the curvature tensor of $F$ is locally square integrable on $X$. Then $F$ extends to the whole space $X$ as a reflexive sheaf $\c F$. Moreover, if the metric $h$ is Yang-Mills, then $h$ can be smoothly extended as a Yang-Mills metric over the place where ${\cal F}$ is locally free. {\bf Proof:} This is \cite{_Bando_Siu_}, Theorem 2. \blacksquare \subsection{Stable and semistable sheaves over hyperk\"ahler manifolds} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure, $F$ a torsion-free coherent sheaf over $(M,I)$ and $F^{**}$ its reflexization. Recall that the cohomology of $M$ are equipped with a natural $SU(2)$-action (\ref{_SU(2)_commu_Laplace_Lemma_}). The motivation for the following definition is \ref{_inva_then_hyperho_Theorem_} and \ref{_UY_for_shea_Theorem_}. \hfill \definition \label{_hyperho_shea_Definition_} Assume that the first two Chern classes of the sheaves $F$, $F^{**}$ are $SU(2)$-invariant. Then $F$ is called {\bf stable hyperholomorphic} if $F$ is stable, and {\bf semistable hyperholomorphic} if $F$ can be obtained as a successive extension of stable hyperholomorphic sheaves. \hfill \remark \label{_slope_hyperho_Remark_} The slope of a hyperholomorphic sheaf is zero, because a degree of an $SU(2)$-invariant 2-form is zero (\ref{_Lambda_of_inva_forms_zero_Lemma_}). \hfill \claim \label{_hyperho_suppo_of_F^**/F_Claim_} Let $F$ be a semistable coherent sheaf over $(M,I)$. Then the following conditions are equivalent. \begin{description} \item[(i)] $F$ is stable hyperholomorphic \item[(ii)] Consider the support $S$ of the sheaf $F^{**}/F$ as a complex subvariety of $(M,I)$. Let $X_1$, ... , $X_n$ be the set of irreducible components of $S$ of codimension 2. Then $X_i$ is trianalytic for all $i$, and the sheaf $F^{**}$ is stable hyperholomorphic. \end{description} {\bf Proof:} Consider an exact sequence \[ 0 {\:\longrightarrow\:} F {\:\longrightarrow\:} F^{**} {\:\longrightarrow\:} F^{**}/ F{\:\longrightarrow\:} 0. \] Let $[F / F^{**}]\in H^4(M)$ be the fundamental class of the union of all co\-di\-men\-sion-2 components of support of the sheaf $F / F^{**}$, taken with appropriate multiplicities. Then, $c_2(F^{**}/ F) =- [F / F^{**}]$. From the product formula for Chern classes, it follows that \begin{equation} \label{_c_2(F)_and_F^**_Equation_} c_2(F)= c_2(F^{**}_i) + c_2(F^{**}/ F) = c_2(F^{**}_i) - [F / F^{**}]. \end{equation} Clearly, if all $X_i$ are trianalytic then the class $[F / F^{**}]$ is $SU(2)$-invariant. Thus, if the sheaf $F^{**}$ is hyperholomorphic and all $X_i$ are trianalytic, then the second Chern class of $F$ is $SU(2)$-invariant, and $F$ is hyperholomorphic. Conversely, assume that $F$ is hyperholomorphic. We need to show that all $X_i$ are trianalytic. By definition, \[ [F / F^{**}] = \sum_i \lambda_i [X_i] \] where $[X_i]$ denotes the fundamental class of $X_i$, and $\lambda_i$ is the multiplicity of $F / F^{**}$ at $X_i$. By \eqref{_c_2(F)_and_F^**_Equation_}, ($F$ hyperholomorphic) implies that the class $[F / F^{**}]$ is $SU(2)$-invariant. Since $[F / F^{**}]$ is $SU(2)$-invariant, we have \[ \sum_i \lambda_i\deg_J(X_i) = \sum_i \lambda_i\deg_I(X_i). \] By Wirtinger's inequality (\ref{_Wirti_hyperka_Proposition_}), \[ \deg_J(X_i) \leq\deg_I(X_i), \] and the equality is reached only if $X_i$ is trianalytic. By definition, all the numbers $\lambda_i$ are positive. Therefore, \[ \sum_i \lambda_i\deg_J(X_i) \leq \sum_i \lambda_i\deg_I(X_i). \] and the equality is reached only if all the subvarieties $X_i$ are trianalytic. This finishes the proof of \ref{_hyperho_suppo_of_F^**/F_Claim_}. \blacksquare \hfill \claim Let $M$ be a compact hyperk\"ahler manifold, and $I$ an induced complex structure of general type. Then all torsion-free coherent sheaves over $(M, I)$ are semistable hyperholomorphic. {\bf Proof:} Let $F$ be a torsion-free coherent sheaf over $(M, I)$. Clearly from the definition of induced complex structure of general type, the sheaves $F$ and $F^{**}$ have $SU(2)$-invariant Chern classes. Now, all $SU(2)$-invariant 2-forms have degree zero (\ref{_Lambda_of_inva_forms_zero_Lemma_}), and thus, $F$ is semistable. \blacksquare \subsection{Hyperholomorphic connection in torsion-free sheaves} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure, and $F$ a torsion-free sheaf over $(M,I)$. Consider the natural $SU(2)$-action in the bundle $\Lambda^i (M,B)$ of the differential $i$-forms with coefficients in a vector bundle $B$. Let $\Lambda^i_{inv}(M, B)\subset \Lambda^i (M, B)$ be the bundle of $SU(2)$-invariant $i$-forms. \hfill \definition \label{_hyperholo_co_Definition_} Let $X\subset (M, I)$ be a complex subvariety of codimension at least 2, such that $F\restrict{M\backslash X}$ is a bundle, $h$ be an admissible metric on $F\restrict{M\backslash X}$ and $\nabla$ the associated connection. Then $\nabla$ is called {\bf hyperholomorphic} if its curvature \[ \Theta_\nabla = \nabla^2 \in \Lambda^2\left(M, \operatorname{End}\left(F\restrict{M\backslash X}\right)\right) \] is $SU(2)$-invariant, i. e. belongs to $\Lambda^2_{inv}\left(M, \operatorname{End}\left(F\restrict{M\backslash X}\right)\right)$. \hfill \claim\label{_singu_triana_Claim_} The singularities of a hyperholomorphic connection form a trianalytic subvariety in $M$. {\bf Proof:} Let $J$ be an induced complex structure on $M$, and $U$ the set of all points of $(M,I)$ where $F$ is non-singular. Clearly, $(F, \nabla)$ is a bundle with admissible connection on $(U,J)$. Therefore, the holomorphic structure on $F\restrict{(U,J)}$ can be extended to $(M,J)$. Thus, the singular set of $F$ is holomorphic with respect to $J$. This proves \ref{_singu_triana_Claim_}. \blacksquare \hfill \proposition \label{_conne_=>_hyperho_Proposition_} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf admitting a hyperholomorphic connection. Then $F$ is a polystable hyperholomorphic sheaf. \hfill {\bf Proof:} By \ref{_hyperho_co_YM_Remark_} and \ref{_UY_for_shea_Theorem_}, $F$ is polystable. We need only to show that the Chern classes $c_1(F)$ and $c_2(F)$ are $SU(2)$-invariant. Let $U\subset M$ be the maximal open subset of $M$ such that $F\restrict U$ is locally trivial. By \ref{_YM_can_be_exte_Theorem_}, the metric $h$ and the connection $\nabla$ can be extended to $U$. Let $\operatorname{Tw} U\subset \operatorname{Tw} M$ be the corresponding twistor space, and $\sigma:\; \operatorname{Tw} U {\:\longrightarrow\:} U$ the standard map. Consider the bundle $\sigma^* F\restrict U$, equipped with a connection $\sigma^* \nabla$. It is well known \footnote{See for instance the section ``Direct and inverse twistor transform'' in \cite{_NHYM_}.} that $\sigma^* F\restrict U$ is a bundle with an admissible Yang-Mills metric (we use Yang-Mills in the sense of Li-Yau, see \ref{_Li_Yau_condi_Definition_}). By \ref{_YM_can_be_exte_Theorem_}, $\sigma^* F\restrict U$ can be extended to a reflexive sheaf $\c F$ on $\operatorname{Tw} M$. Clearly, this extension coincides with the push-forward of $\sigma^* F\restrict U$. The singular set $\tilde S$ of $\c F$ is a pull-back of the singular set $S$ of $F$. Thus, $S$ is trianalytic. By desingularization theorem (\ref{_desingu_Theorem_}), $S$ can be desingularized to a hyperk\"ahler manifold in such a way that its twistors form a desingularization of $\c S$. {}From the exact description of the singularities of $\c S$, provided by the desingularization theorem, we obtain that the standard projection $\pi:\; \c S {\:\longrightarrow\:} {\Bbb C} P^1$ is flat. By the following lemma, the restriction of $\c F$ to the fiber $(M,I) = \pi^{-1}(\{I\})$ of $\pi$ coincides with $F$. \hfill \lemma \label{_exte_flat_Lemma_} Let $\pi:\; X {\:\longrightarrow\:} Y$ be a map of complex varieties, and $S\hookrightarrow X$ a subvariety of $X$ of codimension at least 2, which is flat over $Y$. Denote by $U\stackrel j \hookrightarrow X$ the complement $U = (X\backslash S)$. Let $F$ be a vector bundle over $U$, and $j_* F$ its push-forward. Then the restriction of $j_* F$ to the fibers of $\pi$ is reflexive. \hfill {\bf Proof:} Let $Z= \pi^{-1}(\{y\})$ be a fiber of $\pi$. Since $S$ is flat over $Y$ and of codimension at least 2, we have $j_*({\cal O}_{Z \cap U}) = {\cal O}_{Z}$. Clearly, for an open embedding $\gamma:\; T_1 {\:\longrightarrow\:} T_2$ and coherent sheaves $A, B$ on $T_1$, we have $\gamma_*(A\otimes B) = \gamma_* A \otimes \gamma_* B$. Thus, for all coherent sheaves $A$ on $U$, we have \begin{equation} \label{_j_*_commu_tenso_Equation_} j_* A \otimes {\cal O}_{Z} = j_*(A \otimes {\cal O}_{Z\cap U}). \end{equation} This implies that $j_*(F \restrict Z) = j_* F\restrict Z$. It is well known (\cite{_OSS_}, Ch. II, 1.1.12; see also \ref{_normal_refle_Lemma_}) that a push-forward of a reflexive sheaf under an open embedding $\gamma$ is reflexive, provided that the complement of the image of $\gamma$ has codimension at least 2. Therefore, $j_* F\restrict Z$ is a reflexive sheaf over $Z$. This proves \ref{_exte_flat_Lemma_}. \blacksquare \hfill Return to the proof of \ref{_conne_=>_hyperho_Proposition_}. Consider the sheaf $\c F$ on the twistor space constructed above. Since $\c F$ is reflexive, its singularities have codimension at least 3 (\cite{_OSS_}, Ch. II, 1.1.10). Therefore, $\c F$ is flat in codimension 2, and the first two Chern classes of $F= \c F\restrict{\pi^{-1}(I)}$ can be obtained by restricting the first two Chern classes of $\c F$ to the subvariety $(M,I) = \pi^{-1}(I) \subset \operatorname{Tw}(M)$. It remains to show that such restriction is $SU(2)$-invariant. Clearly, $H^2((M, I)) = H^2((M, I)\backslash S)$, and $H^4((M, I)) = H^4((M, I)\backslash S)$. Therefore, \[ c_1\left(\c F \restrict {(M,I)}\right) = c_1\left(\c F \restrict {(M,I)\backslash S}\right) \] and \[ c_2\left(\c F \restrict {(M,I)}\right) = c_2\left(\c F \restrict {(M,I)\backslash S}\right). \] On the other hand, the restriction $\c F \restrict {\operatorname{Tw}(M)\backslash S}$ is a bundle. Therefore, the classes \[ c_1\left(\c F \restrict {(M,I)\backslash S}\right), \ \ c_2\left(\c F \restrict {(M,I)\backslash S}\right) \] are independent from $I\in {\Bbb C} P^1$. On the other hand, these classes are of type $(p,p)$ with respect to all induced complex structures $I\in {\Bbb C} P^1$. By \ref{_SU(2)_inva_type_p,p_Lemma_}, this implies that the classes $c_1(\c F \restrict {(M,I)})$, $c_1(\c F \restrict {(M,I)})$ are $SU(2)$-invariant. As we have shown above, these two classes are equal to the first Chern classes of $F$. \ref{_conne_=>_hyperho_Proposition_} is proven. \blacksquare \hfill \subsection{Existence of hyperholomorphic connections} The following theorem is the main result of this section. \hfill \theorem\label{_hyperho_conne_exi_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf on $(M,I)$. Then $F$ admits a hyperholomorphic connection if and only if $F$ is polystable hyperholomorphic in the sense of \ref{_hyperho_shea_Definition_}. \hfill \remark \label{_hyperho_co_YM_Remark_} {}From \ref{_Lambda_of_inva_forms_zero_Lemma_}, it is clear that a hyperholomorphic connection is always Yang-Mills. Therefore, such a connection is unique (\ref{_UY_for_shea_Theorem_}). \hfill The ``only if'' part of \ref{_hyperho_conne_exi_Theorem_} is \ref{_conne_=>_hyperho_Proposition_}. The proof of ``if'' part of \ref{_hyperho_conne_exi_Theorem_} takes the rest of this subsection. \hfill Let $I$ be an induced complex structure. We denote the corresponding Hodge decomposition on differential forms by $\Lambda^*(M)= \oplus \Lambda^{p,q}_I(M)$, and the standard Hodge operator by $\Lambda_I:\; \Lambda^{p,q}_I(M) {\:\longrightarrow\:} \Lambda^{p-1,q-1}_I(M)$. All these structures are defined on the differential forms with coefficients in a bundle. Let $\deg_I \eta:= \int_M Tr (\Lambda_I)^k(\eta)$, for $\eta\in \Lambda^k(M, \operatorname{End} B)$. The following claim follows from an elementary linear-algebraic computation. \hfill \claim\label{_degree_2-forms-in-End(B)_line-alge_Claim_} Let $M$ be a hyperk\"ahler manifold, $B$ a Hermitian vector bundle over $M$, and $\Theta$ a 2-form on $M$ with coefficients in $\frak{su}(B)$. Assume that \[ \Lambda_I \Theta =0, \ \ \ \Theta\in \Lambda^{1,1}_I(M, \operatorname{End} B) \] for some induced complex structure $I$. Assume, moreover, that $\Theta$ is square-integrable. Let $J$ be another induced complex structure, $J\neq \pm I$. Then \[ \deg_I \Theta^2 \geq |\deg_J \Theta^2|, \] and the equality is reached only if $\Theta$ is $SU(2)$-invariant. {\bf Proof:} The following general argument is used. \hfill \sublemma\label{_deg_coeff_End(B)_Sublemma_} Let $M$ be a K\"ahler manifold, $B$ a Hermitian vector bundle over $M$, and $\Xi$ a square-integrable 2-form on $M$ with coefficients in $\frak{su}(B)$. Then: \begin{description} \item[(i)] For \[ \Lambda_I \Xi =0, \ \ \ \Xi\in \Lambda^{1,1}_I(M, \operatorname{End} B) \] we have \[ \deg_I \Xi^2 = C \int_M|\Xi|^2 \operatorname{Vol} M , \] where $C= (4\pi^2 n (n-1))^{-1} $ $M$. \item[(ii)] For \[ \Xi\in \Lambda^{2,0}_I(M, \operatorname{End} B)\oplus \Lambda^{0,2}_I(M, \operatorname{End} B) \] we have \[ \deg_I \Xi^2 = -C \int_M |\Xi|^2\operatorname{Vol} M, \] where $C$ is the same constant as appeared in (i). \end{description} {\bf Proof:} The proof is based on a linear-algebraic computation (so-called L\"ubcke-type argument). The same computation is used to prove Hodge-Riemann bilinear relations. \blacksquare \hfill Return to the proof of \ref{_degree_2-forms-in-End(B)_line-alge_Claim_}. Let $\Theta= \Theta^{1,1}_J + \Theta^{2,0}_J + \Theta^{0,2}_J$ be the Hodge decomposition associated with $J$. The following Claim shows that $\Theta^{1,1}_J$ satisfies conditions of \ref{_deg_coeff_End(B)_Sublemma_} (i). \hfill \claim Let $M$ be a hyperk\"ahler manifold, $I$, $L$ induced complex structures and $\Theta$ a 2-form on $M$ satisfying \[ \Lambda_I \Theta =0, \ \ \ \Theta\in \Lambda^{1,1}_I(M). \] Let $\Theta^{1,1}_L$ be the $(1,1)$-component of $\Theta$ taken with respect to $L$. Then $\Lambda_L \Theta^{1,1}_L =0$. \hfill {\bf Proof:} Clearly, $\Lambda_L \Theta^{1,1}_L= \Lambda_L\Theta$. Consider the natural Hermitian structure on the space of 2-forms. Since $\Theta$ is of type $(1,1)$ with respect to $I$, $\Theta$ is fiberwise orthogonal to the holomorphic symplectic form $\Omega= \omega_J +\sqrt{-1}\: \Omega_K\in \Lambda^{2,0}_I(M)$. By the same reason, $\Theta$ is orthogonal to $\bar\Omega$. Therefore, $\Theta$ is orthogonal to $\omega_J$ and $\omega_K$. Since $\Lambda_I \Theta =0$, $\Theta$ is also orthogonal to $\omega_I$. The map $\Lambda_L$ is a projection to the form $\omega_L$ which is a linear combination of $\omega_I$, $\omega_J$ and $\omega_K$. Since $\Theta$ is fiberwise orthogonal to $\omega_L$, we have $\Lambda_L \Theta =0$. \blacksquare \hfill By \ref{_deg_coeff_End(B)_Sublemma_}, we have \[ \deg_J \left(\Theta^{1,1}_J\right)^2 = C \int_M|\Theta^{1,1}_J|^2 \] and \[ \deg_J \left(\Theta^{2,0}_J + \Theta^{0,2}_J\right)^2 = - C \int_M|\Theta^{2,0}_J + \Theta^{0,2}_J|^2. \] Thus, \[ \deg_J \Theta^2 = C \int_M\left|\Theta^{1,1}_J\right|~2- C \int_M\left|\Theta^{2,0}_J + \Theta^{0,2}_J\right|~2. \] On the other hand, \[ \deg_I \Theta^2 = C \int_M\left|\Theta\right|~2 = C \int_M\left|\Theta^{1,1}_J\right|~2 + C \int_M\left|\Theta^{2,0}_J + \Theta^{0,2}_J\right|~2. \] Thus, $\deg_I \Theta^2> |\deg_J \Theta^2|$ unless $\Theta^{2,0}_J + \Theta^{0,2}_J=0$. On the other hand, $\Theta^{2,0}_J + \Theta^{0,2}_J=0$ means that $\Theta$ is of type $(1,1)$ with respect to $J$. Consider the standard $U(1)$-action on differential forms associated with the complex structures $I$ and $J$. These two $U(1)$-actions generate the whole Lie group $SU(2)$ acting on $\Lambda^2(M)$ (here we use that $I\neq \pm J$). Since $\Theta$ is of type $(1,1)$ with respect to $I$ and $J$, this form is $SU(2)$-invariant. This proves \ref{_degree_2-forms-in-End(B)_line-alge_Claim_}.\blacksquare \hfill Return to the proof of \ref{_hyperho_conne_exi_Theorem_}. Let $\nabla$ be the admissible Yang-Mills connection in $F$, and $\Theta$ its curvature. Recall that the form $Tr \Theta^2$ represents the cohomology class $2c_2(F) - \frac{r-1}{r} c_1(F)^2$, where $c_i$ are Chern classes of $F$. Since the form $Tr \Theta^2$ is square-integrable, the integral \[ \deg_J \Theta^2= \int_M Tr \Theta^2\omega_J^{n-2} \] makes sense. In \cite{_Bando_Siu_}, it was shown how to approximate the connection $\nabla$ by smooth connections, via the heat equation. This argument, in particular, was used to show that the value of integrals like $\int_M Tr \Theta^2\omega_J^{n-2}$ can be computed through cohomology classes and the Gauss--Bonnet formula \[ Tr \Theta^2 = 2c_2(F) - \frac{r-1}{r} c_1(F)^2.\] Since the classes $c_2(F)$, $c_1(F)$ are $SU(2)$-invariant, we have \[ \deg_I\Theta^2= \deg_J \Theta^2 \] for all induced complex structures $I$, $J$. By \ref{_degree_2-forms-in-End(B)_line-alge_Claim_}, this implies that $\Theta$ is $SU(2)$-invariant. \ref{_hyperho_conne_exi_Theorem_} is proven. \blacksquare \hfill The same argument implies the following corollary. \hfill \corollary\label{_stable_shea_degree_Corollary_} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure, $F$ a stable reflexive sheaf on $(M,I)$, and $J$ be an induced complex structure, $J\neq \pm I$. Then \[ \deg_I\left(2c_2(F) - \frac{r-1}{r} c_1(F)^2\right) \geq \left|\deg_J\left(2c_2(F) - \frac{r-1}{r} c_1(F)^2\right)\right|, \] and the equality holds if and only if $F$ is hyperholomorphic. \blacksquare \subsection{Tensor category of hyperholomorphic sheaves} This subsection is extraneous. Further on, we do not use the tensor structure on the category of hyperholomorphic sheaves. However, we need the canonical idenitification of the categories of hyperholomorphic sheaves associated with different induced complex structures. {}From Bando-Siu (\ref{_UY_for_shea_Theorem_}) it follows that on a compact K\"ahler manifold a tensor product of stable reflexive sheaves is polystable. Similarly, \ref{_hyperho_conne_exi_Theorem_} implies that a tensor product of polystable hyperholomorphic sheaves is polystable hyperholomorphic. We define the following category. \hfill \definition Let $M$ be a compace hyperk\"ahler manifold and $I$ an induced complex structure. Let $\c F_{st}(M,I)$ be a category with objects reflexive polystable hyperholomorphic sheaves and morphisms as in category of coherent sheaves. This category is obviously additive. The tensor product on $\c F_{st}(M,I)$ is induced from the tensor product of coherent sheaves. \hfill \claim The category $\c F_{st}(M,I)$ is abelian. Moreover, it is a Tannakian tensor category. {\bf Proof:} Let $\phi:\; F_1 {\:\longrightarrow\:} F_2$ be a morphism of hyperholomorphi sheaves. In \ref{_degree,slope_destabilising_Definition_} , we introduced {\bf a slope} of a coherent sheaf. Clearly, $sl(F_1)\leq sl(\operatorname{im} \phi)\leq sl(F_2)$. All hyperholomorphic sheaves have slope 0 by \ref{_slope_hyperho_Remark_}. Thus, $sl(\operatorname{im} \phi)=0$ and the subsheaf $\operatorname{im}\phi\subset F_2$ is destabilizing. Since $F_2$ is polystable, this sheaf is decomposed: \[ F_2 = \operatorname{im} \phi \oplus \operatorname{coker} \phi.\] A similar argument proves that $F_1 = \ker\phi \oplus \operatorname{coim} \phi$, with all summands hyperholomorphic. This proves that $\c F_{st}(M,I)$ is abelian. The Tannakian properties are clear. \blacksquare \hfill The category $\c F_{st}(M,I)$ does not depend from the choice of induced complex structure $I$: \hfill \theorem\label{_equi_cate_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $I_1$, $I_2$ induced complex structures and $\c F_{st}(M,I_1)$, $\c F_{st}(M,I_2)$ the associated categories of polystable reflexive hyperholomorphic sheaves. Then, there exists a natural equivalence of tensor categories \[ \Phi_{I_1, I_2}:\; \c F_{st}(M,I_1){\:\longrightarrow\:} \c F_{st}(M,I_2). \] {\bf Proof:} Let $F\in \c F_{st}(M,I_1)$ be a reflexive polystable hyperholomorphic sheaf and $\nabla$ the canonical admissible Yang-Mills connection. Consider the sheaf $\c F$ on the twistor space $\operatorname{Tw}(M)$ constructed as in the proof of \ref{_conne_=>_hyperho_Proposition_}. Restricting $\c F$ to $\pi^{-1}(I_2)\subset \operatorname{Tw}(M)$, we obtain a coherent sheaf $F'$ on $(M,I_2)$. As we have shown in the proof of \ref{_conne_=>_hyperho_Proposition_}, the sheaf $(F')^{**}$ is polystable hyperholomorphic. Let $\Phi_{I_1, I_2}(F):= (F')^{**}$. It is easy to check that thus constructed map of objects gives a functor \[ \Phi_{I_1, I_2}:\; \c F_{st}(M,I_1){\:\longrightarrow\:} \c F_{st}(M,I_2), \] and moreover, $\Phi_{I_1, I_2}\circ \Phi_{I_2, I_1} = Id$. This shows that $\Phi_{I_1, I_2}$ is an equivalence. \ref{_equi_cate_Theorem_} is proven. \blacksquare \hfill \definition \label{_hyperho_shea_on_M_Definition_} By \ref{_equi_cate_Theorem_}, the category $\c F_{st}(M,I_1)$ is independent from the choice of induced complex structure. We call this category {\bf the category of polystable hyperholomorphic reflexive sheaves on $M$} and denote it by $\c F(M)$. The objects of $\c F(M)$ are called {\bf hyperholomorphic sheaves on $M$}. For a hyperholomorphic sheaf on $M$, we denote by $F_I$ the corresponding sheaf from $\c F_{st}(M,I_1)$. \hfill \remark Using the same argument as proves \ref{_admi_twi_impli_Theorem_} (ii), it is easy to check that the category $\c F(M)$ is a deformational invariant of $M$. That is, for two hyperk\"ahler manifolds $M_1$, $M_2$ which are deformationally equivalent, the categories $\c F(M_i)$ are also equivalent, assuming that $Pic(M_1)= Pic(M_2)=0$. The proof of this result is essentially contained in \cite{_coho_announce_}. \hfill \remark As Deligne proved (\cite{_Deli:Tanna_}), for a each Tannakian category $\c C$ equipped with a fiber functor, there exists a natural pro-algebraic group $G$ such that $\c C$ is a group of representations of $G$. For $\c F(M)$, there are several natural fiber functors. The simplest one is defined for each induced complex structure $I$ such that $(M,I)$ is algebraic (such complex structures always exist, as proven in \cite{_Fujiki_}; see also \cite{_Verb:alge_} and Subsection \ref{_alge_indu_Subsection_}). Let $\c K(M,I)$ is the space of rational functions on $(M,I)$. For $F\in \c F_{st}(M, I)$, consider the functor $F{\:\longrightarrow\:} \eta_I(F)$, where $\eta_I(F)$ is the space of global sections of $F\otimes \c K(M,I)$. This is clearly a fiber functor, which associates to $\c F(M)$ the group $G_I$. The corresponding pro-algebraic group $G_I$ is a deformational, that is, topological, invariant of the hyperk\"ahler manifold. \section{Cohomology of hyperk\"ahler manifolds} \label{_cohomo_hype_Section_} This section contains a serie of preliminary results which are used further on to define and study the $C$-restricted complex structures. \subsection{Algebraic induced complex structures} \label{_alge_indu_Subsection_} This subsection contains a recapitulation of results of \cite{_Verb:alge_}. \hfill A more general version of the following theorem was proven by A. Fujiki (\cite{_Fujiki_}, Theorem 4.8 (2)). \hfill \theorem \label{_alge_dense_Theorem_} Let $M$ be a compact simple hyperk\"ahler manifold and $\c R$ be the set of induced complex structures $\c R \cong {\Bbb C} P^1$. Let $\c R_{alg}\subset \c R$ be the set of all algebraic induced complex structures. Then $\c R_{alg}$ is countable and dense in $\c R$. {\bf Proof:} This is \cite{_Verb:alge_}, Theorem 2.2. \blacksquare \hfill In the proof of \ref{_alge_dense_Theorem_}, the following important lemma was used. \hfill \lemma \label{_K_proje_on_R_Lemma_} \begin{description} \item [(i)] Let ${\cal O}\subset H^2(M, {\Bbb R})$ be the set of all cohomology classes which are K\"ahler with respect to some induced comples structure. Then ${\cal O}$ is open in $H^2(M, {\Bbb R})$. Moreover, for all $\omega\in {\cal O}$, the class $\omega$ is {\it not} $SU(2)$-invariant. \item[(ii)] Let $\eta\in H^2(M, {\Bbb R})$ be a cohomology class which is not $SU(2)$-invariant. Then there exists a unique up to a sign induced complex structure $I\in \c R/\{\pm 1\}$ such that $\eta$ belongs to $H^{1,1}_I(M)$. \end{description} {\bf Proof:} This statement is a form of \cite{_Verb:alge_}, Lemma 2.3. \blacksquare \subsection{The action of $\frak{so}(5)$ on the cohomology of a hyperk\"ahler manifold} \label{_so(5)_Subsection_} This subsection is a recollection of data from \cite{_so(5)_} and \cite{_Verbitsky:Symplectic_II_}. \hfill Let $M$ be a hyperk\"ahler manifold. For an induced complex structure $R$ over $M$, consider the K\"ahler form $\omega_R=(\cdot,R\cdot)$, where $(\cdot,\cdot)$ is the Riemannian form. As usually, $L_R$ denotes the operator of exterior multiplication by $\omega_R$, which is acting on the differential forms $A^*(M,{\Bbb C})$ over $M$. Consider the adjoint operator to $L_R$, denoted by $\Lambda_R$. One may ask oneself, what algebra is generated by $L_R$ and $\Lambda_R$ for all induced complex structures $R$? The answer was given in \cite{_so(5)_}, where the following theorem was proven. \hfill \theorem (\cite{_so(5)_}) Let $M, \c H$ be a hyperk\"ahler manifold, and $\frak{a}_{\c H}$ be a Lie algebra generated by $L_R$ and $\Lambda_R$ for all induced complex structures $R$ over $M$. Then the Lie algebra $\frak{a}_{\c H}$ is isomorphic to $\frak{so}(4,1)$. \blacksquare \hfill The following facts about a structure of $\frak{a}_{\c H}$ were also proven in \cite{_so(5)_}. Let $I$, $J$ and $K$ be three induced complex structures on $M$, such that $I\circ J=-J\circ I=K$. For an induced complex structure $R$, consider an operator $ad R$ on cohomology, acting on $(p,q)$-forms as a multiplication by $(p-q)\sqrt{-1}\:$. The operators $ad R$ generate a 3-dimensional Lie algebra ${\frak g}_{\c H}$, which is isomorphic to $\frak{su}(2)$. This algebra coincides with the Lie algebra associated to the standard $SU(2)$-action on $H^*(M)$. The algebra $\frak{a}_{\c H}$ contains ${\frak g}_{\c H}$ as a subalgebra, as follows: \begin{equation}\label{_ad_I_as_commu_Equation_} [\Lambda_J,L_K]=[L_J,\Lambda_K]= ad\: I\;\; \text{(etc)}. \end{equation} The algebra $\frak{a}_{\c H}$ is 10-dimensional. It has the following basis: $L_R,\Lambda_R$, $ad\: R$ $(R=I,J,K)$ and the element $H=[L_R,\Lambda_R]$. The operator $H$ is a standard Hodge operator; it acts on $r$-forms over $M$ as multiplication by a scalar $n-r$, where $n=dim_{\Bbb C} M$. \hfill \definition \label{_isoty_Definition_} Let ${\frak g}$ be a semisimple Lie algebra, $V$ its representation and $V= \oplus V_{\alpha}$ a ${\frak g}$-invariant decompostion of $V$, such that for all $\alpha$, $V_\alpha$ is a direct sum of isomorphic finite-dimensional representations $W_\alpha$ of $V$, and all $W_\alpha$ are distinct. Then the decomposition $V= \oplus V_{\alpha}$ is called {\bf the isotypic decomposition of $V$}. \hfill It is clear that for all finite-dimensional representations, isotypic decomposition always exists and is unique. \hfill Let $M$ be a compact hyperk\"ahler manifold. Consider the cohomology space $H^*(M)$ equipped with the natural action of $\frak{a}_{\c H}= \frak{so}(5)$. Let $H_o^*\subset H^*(M)$ be the isotypic component containing $H^0(M)\subset H^*(M)$. Using the root system written explicitly for $\frak{a}_{\c H}$ in \cite{_so(5)_}, \cite{_Verbitsky:cohomo_}, it is easy to check that $H^*_o(M)$ is an irreducible representation of $\frak{so}(5)$. Let $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ be the unique $\frak{so}(5)$-invariant projection, and $i:\; H^*_o(M) \hookrightarrow H^*(M)$ the natural embedding. \hfill Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure, and $\omega_I$ the corresponding K\"ahler form. Consider the {\bf degree map} $\deg_I:\; H^{2p}(M) {\:\longrightarrow\:} {\Bbb C}$, $\eta{\:\longrightarrow\:} \int_M \eta \wedge \omega_I^{n-p}$, where $n= \dim_C M$. \hfill \proposition\label{_degree_isotypic_Proposition_} The space \[ H^*_o(M)\subset H^*(M)\] is a subalgebra of $H^*(M)$, which is invariant under the $SU(2)$-action. Moreover, for all induced complex structures $I$, the degree map \[\deg_I:\; H^*(M) {\:\longrightarrow\:} {\Bbb C}\] satisfies \[ \deg_I(\eta) =\deg_I(i(p(\eta)), \] where $i:\; H^*_o(M) \hookrightarrow H^*(M)$, $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ are the $\frak{so}(5)$-invariant maps defined above. And finally, the projection $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ is $SU(2)$-invariant. \hfill {\bf Proof:} The space $H^*_o(M)$ is generated from $\pmb 1\in H^0(M)$ by operators $L_R$, $\Lambda_R$. To prove that $H^*_o(M)$ is closed under multiplication, we have to show that $H^*_o(M)$ is generated (as a linear space) by expressions of type $L_{r_1} \circ L_{R_2} \circ ... \circ \pmb 1$. By \eqref{_ad_I_as_commu_Equation_}, the commutators of $L_R$, $\Lambda_R$ map such expressions to linear combinations of such expressions. On the other hand, the operators $\Lambda_R$ map $\pmb 1$ to zero. Thus, the operators $\Lambda_R$ map expressions of type $L_{r_1} \circ L_{R_2} \circ ... \circ \pmb 1$ to linear combinations of such expressions. This proves that $H^*_o(M)$ is closed under multiplication. The second statement of \ref{_degree_isotypic_Proposition_} is clear (see, e. g. \cite{_Verbitsky:Symplectic_II_}, proof of Proposition 4.5). It remains to show that $H^*_o(M)\subset H^*(M)$ is an $SU(2)$-invariant subspace and that $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ is compatible with the $SU(2)$-action. From \eqref{_ad_I_as_commu_Equation_}, we obtain that the Lie group $G_A$ associated with $\frak a_{\c H}\cong \frak{so}(1,4)$ contains $SU(2)$ acting in a standard way on $H^*(M)$. Since the map $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ commutes with $G_A$-action, $p$ also commutes with $SU(2)$-action. We proved \ref{_degree_isotypic_Proposition_}. \blacksquare \subsection{Structure of the cohomology ring} \label{_cohomo_stru_Subsection_} In \cite{_Verbitsky:cohomo_} (see also \cite{_coho_announce_}), we have computed explicitly the subalgebra of cohomology of $M$ generated by $H^2(M)$. This computation can be summed up as follows. \hfill \theorem \label{_S^*H^2_is_H^*M_intro-Theorem_} (\cite{_Verbitsky:cohomo_}, Theorem 15.2) Let $M$ be a compact hyperk\"ahler manifold, $H^1(M)=0$, $\dim_{\Bbb C} M=2n$, and $H^*_r(M)$ the subalgebra of cohomology of $M$ generated by $H^2(M)$. Then \[\bigg\{\begin{array}{lr} H^{2i}_r(M)\cong S^i H^2(M)& \mbox{\ \ for $i\leq n$, and}\\ H^{2i}_r(M)\cong S^{2n-i} H^2(M) & \mbox{\ \ for $i\geq n$ } \end{array} \] \blacksquare \hfill \theorem\label{_gene_all_SU(2)_Theorem_} Let $M$ be a simple hyperk\"ahler manifold. Consider the group $G$ generated by a union of all $SU(2)$ for all hyperk\"ahler structures on $M$. Then the Lie algebra of $G$ is isomorphic to $\frak{so}(H^2(M))$, for a certain natural integer bilinear symmetric form on $H^2(M)$, called Bogomolov-Beauville form. {\bf Proof:} \cite{_Verbitsky:cohomo_} (see also \cite{_coho_announce_}). \blacksquare \hfill The key element in the proof of \ref{_S^*H^2_is_H^*M_intro-Theorem_} and \ref{_gene_all_SU(2)_Theorem_} is the following algebraic computation. \hfill \theorem \label{_r_proper_Theorem_} Let $M$ be a simple hyperk\"ahler manifold, and $\c H$ a hyperk\"ahler structure on $M$. Consider the Lie subalgebra \[ {\frak a}_{\c H}\subset \operatorname{End}(H^*(M)), \ \ {\frak a}_{\c H} \cong \frak{so}(1,4), \] associated with the hyperk\"ahler structure (Sub\-sec\-tion \ref{_so(5)_Subsection_}). Let \[ {\frak g}\subset \operatorname{End}(H^*(M)) \] be the Lie algebra generated by subalgebras ${\frak a}_{\c H}\subset \operatorname{End}(H^*(M))$, for all hyperk\"aher structures $\c H$ on $M$. Then \begin{description} \item[(i)] The algebra ${\frak g}$ is naturally isomorphic to the Lie algebra $\frak{so}(V\oplus \frak H)$, where $V$ is the linear space $H^2(M, {\Bbb R})$ equipped with the Bogomolov--Beauville pairing, and $\frak H$ is a 2-dimensional vector space with a quad\-ra\-tic form of signature $(1, -1)$. \item[(ii)] The space $H_r^*(M)$ is invariant under the action of ${\frak g}$, Moreover, \[ H_r^*(M)\subset H^*(M)\] is an isotypic \footnote{See \ref{_isoty_Definition_} for the definition of isotypic decomposition.} component of the space $H^*(M)$ considered as a representation of ${\frak g}$. \end{description} {\bf Proof:} \cite{_Verbitsky:cohomo_} (see also \cite{_coho_announce_}). \blacksquare \hfill As one of the consequences of \ref{_S^*H^2_is_H^*M_intro-Theorem_}, we obtain the following lemma, which will be used further on in this paper. \hfill \lemma\label{_p_multi_Lemma_} Let $M$ be a simple hyperk\"ahler manifold, $\dim_{\Bbb H} M =n$, and $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ the map defined in Subsection \ref{_so(5)_Subsection_}. Then, for all $x, y \in H^*_r(M)$, we have \[ p(x) p(y) = p(xy), \text{\ \ whenever\ \ } xy \in \bigoplus\limits_{i\leq 2n} H^i(M). \] {\bf Proof:} Let $\omega_I$, $\omega_J$, $\omega_K$, $x_1$, ..., $x_n$ be an orthonormal basis in $H^2(M)$. Clearly, the vectors $x_1$, ..., $x_k$ are $SU(2)$-invariant. Therefore, these vectors are highest vectors of the corresponding ${\frak a}_{\c H}$-representations, with respect to the root system and Cartan subalgebra for ${\frak a}_{\c H}$ which is written in \cite{_so(5)_} or \cite{_Verbitsky:cohomo_}. We obtain that the monomials \[ P_{k_1, k_2, k_3, \{n_i\} } = \omega_I^{k_1} \omega_J^{k_2} \omega_K^{k_3} \prod x_i^{n_i}, \sum n_i = N, \ \ P_{k_1, k_2, k_3, \{n_i\} } \in \bigoplus\limits_{i\leq 2n} H^i(M) \] belong to the different isotypical components for different $N$'s. By \ref{_S^*H^2_is_H^*M_intro-Theorem_}, a product of two such monomials $P_{k_1, k_2, k_3, \{n_i\} }$ and $P_{k_1', k_2', k_3', \{n_i'\} }$ is equal to $P_{k_1+k_1', k_2+k_2', k_3+k_3', \{n_i+n_i'\} }$, assuming that \[ P_{k_1, k_2, k_3, \{n_i\} } P_{k_1', k_2', k_3', \{n_i'\} }\in \bigoplus\limits_{i\leq 2n} H^i(M). \] Thus, the isotypical decomposition associated with the $\frak{a}_{\c H}$-action is compatible with multiplicative structure on $H^*(M)$, for low-dimensional cycles. This implies \ref{_p_multi_Lemma_}. \blacksquare \hfill We shall use the following corollary of \ref{_p_multi_Lemma_}. \hfill \corollary \label{_p_prods_H^2_Corollary_} Let $M$ be a simple hyperk\"ahler manifold, $\dim_{\Bbb H} M>1$, and $\omega_1, \omega_2 \in H^2(M)$ cohomology classes which are $SU(2)$-invariant. Then, for all induced complex structures $I$, we have $\deg_I(\omega_1 \omega_2) =0$. {\bf Proof:} By definition, the classes $\omega_1, \omega_2$ satisfy $\omega_i\in \ker p$. By \ref{_p_multi_Lemma_}, we have $\omega_1 \omega_2\in \ker p$. By \ref{_degree_isotypic_Proposition_}, $\deg_I\omega_1 \omega_2 =0$. \blacksquare \hfill Let $\omega$ be a rational K\"ahler form. The corresponding $\frak{sl}(2)$-action on $H^*(M)$ is clearly compatible with the rational structure on $H^*(M)$. It is easy to see (using, for instance, \ref{_K_proje_on_R_Lemma_}) that ${\frak g}$ is generated by $\frak{sl}(2)$-triples associated with rational K\"ahler forms $\omega$. Therefore, the action of ${\frak g}$ on $H^*(M)$ is compatible with the rational structure on $H^*(M)$. Using the isotypic decomposition, we define a natural ${\frak g}$-invariant map $r:\; H^*(M) {\:\longrightarrow\:} H^*_r(M)$. Further on, we shall use the following properties of this map. \hfill \claim \label{_r_H^*_to_H_r_Claim_} \begin{description} \item[(i)] The map $r:\; H^*(M) {\:\longrightarrow\:} H^*_r(M)$ is compatible with the rational structure on $H^*(M)$. \item[(ii)] For every $x\in \ker r$, and every hyperk\"ahler structure $\c H$, the corresponding map $p:\; H^*(M) {\:\longrightarrow\:} H^*_o(M)$ satisfies $p(x)=0$. \item[(iii)] For every $x\in \ker r$, every hyperk\"ahler structure $\c H$, and every induced complex structure $I$ on $M$, we have $\deg_I x =0$. \end{description} {\bf Proof:} \ref{_r_H^*_to_H_r_Claim_} (i) is clear, because the action of ${\frak g}$ on $H^*(M)$ is compatible with the rational structure on $H^*(M)$. To prove \ref{_r_H^*_to_H_r_Claim_} (ii), we notice that the space $H^*_r(M)$ is generated from $H^0(M)$ by the action of ${\frak g}$, and $H^*_o(M)$ is generated from $H^0(M)$ by the action of ${\frak a}_{\c H}$. Since ${\frak a}_{\c H}$ is by definition a subalgebra in ${\frak g}$, we have $H^*_o(M) \subset H^*_r(M)$. The isotypic projection $r:\; H^*(M) {\:\longrightarrow\:} H^*_r(M)$ is by definition compatible with the ${\frak g}$-action. Since ${\frak a}_{\c H}\subset {\frak g}$, the map $r$ is also compatible with the ${\frak a}_{\c H}$-action. Therefore, $\ker r\subset \ker p$. \ref{_r_H^*_to_H_r_Claim_} (iii) is implied by \ref{_r_H^*_to_H_r_Claim_} (ii) and \ref{_degree_isotypic_Proposition_}. \blacksquare \hfill Let $x_i$ be an basis in $H^2(M, {\Bbb Q})$ which is rational and orthonormal with respect to Bogomolov-Beauville pairing, $(x_i, x_i)_{\c B} = \epsilon_i = \pm 1$. Consider the cohomology class $\theta':= \epsilon_i x_i^2\in H^4(M, {\Bbb Q})$. Let $\theta \in H^4(M, {\Bbb Z})$ be a non-zero integer cohomology class which is proportional to $\theta'$. {}From results of \cite{_Verbitsky:cohomo_} (see also \cite{_coho_announce_}), the following proposition can be easily deduced. \hfill \proposition\label{_theta_SU(2)_inva_Proposition_} The cohomology class $\theta\in H^4(M, {\Bbb Z})$ is $SU(2)$-invariant for all hyperk\"ahler structures on $M$. Moreover, for a generic hyperk\"ahler structure, the group of $SU(2)$-invariant integer classes $\alpha\in H^4_r(M)$ has rank one, where $H^*_r(M)$ is the subalgebra of cohomology generated by $H^2(M)$. \hfill {\bf Proof:} Clearly, if an integer class $\alpha$ is $SU(2)$-invariant for a generic hyperk\"ahler structure, then $\alpha$ is $G$-invariant, where $G$ is the group defined in \ref{_gene_all_SU(2)_Theorem_}. On the other hand, $H^4_r(M)\cong S^2(H^2(M))$, as follows from \ref{_S^*H^2_is_H^*M_intro-Theorem_}. Clearly, the vector $\theta\in H^4_r(M)\cong S^2(H^2(M))$ is $\frak{so}(H^2(M))$-invariant. Moreover, the space of $\frak{so}(H^2(M))$-invariant vectors in $S^2(H^2(M))$ is one-dimensional. Finally, from an explicit computation of $G$ it follows that $G$ acts on $H^4(M)$ as $SO(H^2(M))$, and thus, the Lie algebra invariants coincide with invariants of $G$. We found that the space of $G$-invariants in $H^4_r(M)$ is one-dimensional and generated by $\theta$. This proves \ref{_theta_SU(2)_inva_Proposition_}. \blacksquare \hfill \remark It is clear how to generalize \ref{_theta_SU(2)_inva_Proposition_} from dimension 4 to all dimensions. The space $H^{2d}_r(M)^G$ of $G$-invariants in $H^{2d}_r(M)$ is 1-dimensional for $d$ even and zero-dimensional for $d$ odd. \subsection{Cohomology classes of CA-type} Let $M$ be a compact hyperk\"ahler manifold, and $I$ an induced complex structure. All cohomology classes which appear as fundamental classes of complex subvarieties of $(M, I)$ satisfy certain properties. Classes satisfying these properties are called classes of CA-type, from Complex Analytic. Here is the definition of CA-type. \hfill \definition Let $\eta\in H^{2,2}_I(M) \cap H^4(M, {\Bbb Z})$ be an integer (2,2)-class. Assume that for all induced complex structures $J$, satisfying $I\circ J = - J\circ I$, we have $\deg_I(\eta)\geq \deg_J(\eta)$, and the equality is reached only if $\eta$ is $SU(2)$-invariant. Assume, moreover, that $\deg_I(\eta)\geq |\deg_J(\eta)|$. Then $\eta$ is called {\bf a class of CA-type}. \hfill \theorem \label{_funda_and_Chern_CA_Theorem_} Let $M$ be a simple hyperk\"ahler manifold, of dimension $\dim_{\Bbb H} M>1$, $I$ an induced complex structure, and $\eta \in H^{2,2}_I(M) \cap H^4(M, {\Bbb Z})$ an integer (2,2)-class. Assume that one of the following conditions holds. \begin{description} \item[(i)] There exists a complex subvariety $X\subset (M, I)$ such that $\eta$ is the fundamental class of $X$ \item[(ii)] There exists a stable coherent torsion-free sheaf $F$ over $(M,I)$, such that the first Chern class of $F$ is zero, and $\eta=c_2(F)$. \end{description} Then $\eta$ is of CA-type. \hfill {\bf Proof:} \ref{_funda_and_Chern_CA_Theorem_} (i) is a direct consequence of Wirtinger's inequality (\ref{_Wirti_hyperka_Proposition_}). It remains to prove \ref{_funda_and_Chern_CA_Theorem_} (ii). We assume, temporarily, that $F$ is reflexive. By \ref{_stable_shea_degree_Corollary_}, we have \begin{equation} \label{_c_2,1_ineq_Equation_} \deg_I(2c_2(F) - \frac{r-1}{r} c_1(F)^2) \geq \left|\deg_J(2c_2(F) - \frac{r-1}{r} c_1(F)^2)\right|, \end{equation} and the equality happens only if $F$ is hyperholomorphic. Since $c_1(F)$ is $SU(2)$-invariant, we have $\deg_I(c_1(F)^2) = \deg_J (c_1(F)^2) =0$ (\ref{_p_prods_H^2_Corollary_}). Thus, \eqref{_c_2,1_ineq_Equation_} implies that \[ \deg_I 2c_2(F) \geq |\deg_J 2c_2(F)| \] and the inequality is strict unless $F$ is hyperholomorphic, in which case, the class $c_2(F)$ is $SU(2)$-invariant by definition. We have proven \ref{_funda_and_Chern_CA_Theorem_} (ii) for the case of reflexive $F$. \hfill For $F$ not necessary reflexive sheaf, we have shown in the proof of \ref{_hyperho_suppo_of_F^**/F_Claim_} that \[ c_2(F) = c_2(F^{**}) + \sum n_i [X_i], \] where $n_i$ are positive integers, and $[X_i]$ are the fundamental classes of irreducible components of support of the sheaf $F^{**}/F$. Therefore, the class $c_2(F)$ is a sum of classes of CA-type. Clearly, a sum of cohomology classes of CA-type is again a class of CA-type. This proves \ref{_funda_and_Chern_CA_Theorem_}. \blacksquare \section{$C$-restricted complex structures on hy\-per\-k\"ah\-ler manifolds} \label{_C_restri_Section_} \subsection{Existence of $C$-restricted complex structures} We assume from now till the end of this section that the hyperk\"ahler manifold $M$ is simple (\ref{_simple_hyperkahler_mfolds_Definition_}). This assumption can be avoided, but it simplifies notation. We assume from now till the end of this section that the hyperk\"ahler manifold $M$ is compact of real dimension $\dim_{\Bbb R} M \geq 8$, i. e. $\dim_{\Bbb H}M \geq 2$. This assumption is absolutely necessary. The case of hyperk\"ahler surfaces with $\dim_{\Bbb H}M =1$ (torus and K3 surface) is trivial and for our purposes not interesting. It is not difficult to extend our definitions and results to the case of a compact hyperk\"ahler manifold which is a product of simple hyperk\"ahler manifolds with $\dim_{\Bbb H}M \geq 2$. \hfill \definition\label{_C_restri_Definition_} Let $M$ be a compact hyperk\"ahler manifold, and $I$ an induced complex structure. As usually, we denote by $\deg_I:\; H^{2p}(M) {\:\longrightarrow\:} {\Bbb C}$ the associated degree map, and by $H^*(M)= \oplus H^{p,q}_I(M)$ the Hodge decomposition. Assume that $I$ is algebraic. Let $C$ be a positive real number. We say that the induced complex structure $I$ is {\bf $C$-restricted} if the following conditions hold. \begin{description} \item[(i)] For all non-$SU(2)$-invariant cohomology classes classes $\eta\in H^{1,1}_I(M) \cap H^2(M, {\Bbb Z})$, we have $|\deg_I(\eta)| > C$. \item[(ii)] Let $\eta \in H^{2,2}_I(M)$ be a cohomology class of CA-type which is not $SU(2)$-invariant. Then $|\deg_I(\eta)| > C$. \end{description} \hfill The heuristic (completely informal) meaning of this definition is the following. The degree map plays the role of the metric on the cohomology. Cohomology classes with small degrees are ``small'', the rest is ``big''. Under reasonably strong assumptions, there are only finitely many ``small'' integer classes, and the rest is ``big''. For each non-$SU(2)$-invariant cohomology class $\eta$ there exists at most two induced complex structures for which $\eta$ is of type $(p,p)$. Thus, for most induced complex structures, all non-$SU(2)$-invariant integer $(p,p)$ classes are ``big''. Intuitively, the $C$-restriction means that all non-$SU(2)$-invariant integer (1,1) and (2,2)-cohomology classes are ``big''. This definition is needed for the study of first and second Chern classes of sheaves. The following property of $C$-restricted complex structures is used (see \ref{_funda_and_Chern_CA_Theorem_}): for every subvariety $X\subset (M,I)$ of complex codimension 2, either $X$ is trianalytic or $\deg_I(X)>C$. \hfill \definition \label{_admitti_C_restri_Definition_} Let $M$ be a compact manifold, and $\c H$ a hyperk\"ahler structure on $M$. We say that $\c H$ {\bf admits $C$-restricted complex structures} if for all $C>0$, the set of all $C$-restricted algebraic complex structures is dense in the set $R_{\c H}= {\Bbb C} P^1$ of all induced complex structures. \hfill \proposition \label{_restri_for_H^11_1-dim-Proposition_} Let $M$ be a compact simple hyperk\"ahler manifold, $\dim_{\Bbb H}(M) >1$, and $r:\; H^4(M) {\:\longrightarrow\:} H_r^4(M)$ be the map defined in \ref{_r_H^*_to_H_r_Claim_}. Assume that for all algebraic induced complex structures $I$, the group $H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$ has rank one, and the group \[ H^{2,2}_I(M)\cap H^4(M, {\Bbb Z})/(\ker r)\] has rank 2. Then $M$ admits $C$-restricted complex structures. \hfill {\bf Proof:} The proof of \ref{_restri_for_H^11_1-dim-Proposition_} takes the rest of this section. Denote by $\c R$ the set $\c R\cong {\Bbb C} P^1$ of all induced complex structures on $M$. Consider the set $\c R/\{\pm 1\}$ of induced complex structures up to a sign (\ref{_K_proje_on_R_Lemma_}). Let $\alpha \in H^2(M)$ be a cohomology class which is not $SU(2)$-invariant. According to \ref{_K_proje_on_R_Lemma_}, there exists a unique element $c(\alpha)\in \c R/\{\pm 1\}$ such that $\alpha \in H^{1,1}_{c(\alpha)}(M)$. This defines a map \[ c:\; \left( H^2(M, {\Bbb R}) \backslash H^2_{inv}(M)\right) {\:\longrightarrow\:} \c R/\{\pm 1\}, \] where $H^2_{inv}(M)\subset H^2(M)$ is the set of all $SU(2)$-invariant cohomology classes. For induced complex structures $I$ and $-I$, and $\eta\in H^{2p}(M)$, the degree maps satisfy \begin{equation} \label{_deg_-I_Equation_} \deg_I(\eta) = (-1)^{p}\deg_{-I}(\eta). \end{equation} Thus, the number $|\deg_I(\eta)|$ is independent from the sign of $I$. Let $\eta\in H^*(M, {\Bbb Z})$ be a cohomology class. The {\bf largest divisor} of $\eta$ is the biggest positive integer number $k$ such that the cohomology class $\frak \eta k$ is also integer. Let $\alpha\in H^2(M, {\Bbb Z})$ be an integer cohomology class, which is not $SU(2)$-invariant, $k$ its largest divisor and $\tilde \alpha:= \frak \alpha k$ the corresponding integer class. Denote by $\widetilde\deg(\alpha)$ the number \[ \widetilde\deg(\alpha) := \left| \deg_{c(\alpha)}(\tilde \alpha)\right|. \] The induced complex structure $c(\alpha)$ is defined up to a sign, but from \eqref{_deg_-I_Equation_} it is clear that $\widetilde\deg(\alpha)$ is independent from the choice of a sign. \hfill \lemma \label{_C_restri_from_A,d_Lemma_} Let $M$ be a compact hyperk\"ahler manifold, and $I$ be an algebraic induced complex structure, such that the group $H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$ has rank one, and the group $H^{2,2}_I(M)\cap H^4(M, {\Bbb Z})/(\ker r)$ has rank 2. Denote by $\alpha$ the generator of $H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$. Since the class $\alpha$ is proportional to a K\"ahler form, $\alpha$ is not $SU(2)$-invariant (\ref{_K_proje_on_R_Lemma_}, (i)). Let $d:= \widetilde \deg\alpha$. Then, there exists a positive real constant $A$ depending on volume of $M$, its topology and its dimension, such that $I$ is $d\cdot A$-restricted. \hfill {\bf Proof:} This lemma is a trivial calculation based on results of \cite{_Verbitsky:cohomo_} (see also \cite{_coho_announce_} and Subsection \ref{_cohomo_stru_Subsection_}). Since $H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$ has rank one, for all $\eta \in H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$, $\eta\neq 0$, we have $| \deg_I \eta| \geq d$. This proves the first condition of \ref{_C_restri_Definition_}. Let $\theta$ be the $SU(2)$-invariant integer cycle $\theta\in H^4(M)$ defined in \ref{_theta_SU(2)_inva_Proposition_}. By \ref{_SU(2)_inva_type_p,p_Lemma_}, $\theta\in H^{2,2}_I(M)$. Consider $\alpha^2\in H^{2,2}_I(M)$, where $\alpha$ is the generator of $H^{1,1}_I(M)\cap H^2(M, {\Bbb Z})$. \hfill \sublemma\label{_degrees_theta_alpha^2_Sublemma_} Let $J$ be an induced complex structure, $J\circ I=-J\circ I$, and $\deg_I$, $\deg_J$ the degree maps associated with $I$, $J$. Then \[ \deg_I \alpha^2 >0, \deg_J\alpha^2 =0, \deg_I\theta=\deg_J\theta >0. \] {\bf Proof:} Since $\alpha$ is a K\"ahler class with respect to $I$, we have $\deg_I \alpha^2 >0$. Since the cohomology class $\theta$ is $SU(2)$-invariant, and $SU(2)$ acts transitively on the set of induced complex structures, we have $\deg_I\theta=\deg_J\theta$. It remains to show that $\deg_J\alpha^2 =0$ and $\deg_J\theta >0$. The manifold $M$ is by our assumptions simple; thus, $\dim H^{2,0}(M) =1$ (\cite{_Besse:Einst_Manifo_}). Therefore, in the natural $SU(2)$-invariant decomposition \begin{equation}\label{_H^2_isoty_Equation_} H^2(M) = H_{inv}^2(M) \oplus H^2_+(M), \end{equation} we have $\dim H^2_+(M) = 3$. In particular, the intersection $H^2_+(M)\cap H^{1,1}_I(M)$ is 1-dimensional. Consider the decomposition of $\alpha$, associated with \eqref{_H^2_isoty_Equation_}: $\alpha = \alpha_+ + \alpha_{inv}$. Since $\alpha$ is of type $(1,1)$ with respect to $I$, the class $\alpha_+$ is proportional to the K\"ahler class $\omega_I$, with positive coefficient. A similar argument leads to the following decomposition for $\theta$: \[ \theta = \omega_I^2 + \omega_J^2 +\omega_K^2 + \sum x_i^2, \] where $K= I\circ J$ is an induced complex structure, and the classes $x_i$ belong to $H_{inv}^2(M)$. {}From \ref{_p_prods_H^2_Corollary_}, we obtain that the classes $x_i^2$ satisfy $\deg_I(x_i^2)=0$ (here we use $\dim_{\Bbb H}(M)>1$). Thus, \[ \deg_I(\theta) = \deg_I(\omega_I^2 + \omega_J^2 +\omega_K^2) = \deg_I(\omega_I^2) >0. \] Similarly one checks that \[ \deg_J(\alpha^2) = \deg_J((\alpha_+ + \alpha_{inv})^2) = \deg_J(\alpha_+^2) = \deg_J (c^2 \omega_I) =0. \] This proves \ref{_degrees_theta_alpha^2_Sublemma_}. \blacksquare \hfill Return to the proof of \ref{_C_restri_from_A,d_Lemma_}. Since $\deg_I\alpha^2 \neq \deg_J \alpha^2$, the class $\alpha^2$ is {\it not} $SU(2)$-invariant. Since $\theta$ {\it is} $SU(2)$-invariant, $\theta$ is not collinear with $\alpha^2$. The degrees $\deg_I$ of $\theta$ and $\alpha^2$ are non-zero; we have $\deg_I(\theta)=\deg_J(\theta)$, $\deg_I(\alpha^2) \neq \deg_J(\alpha^2)$ for $J$ an induced complex structure, $J\neq \pm I$. By \ref{_degree_isotypic_Proposition_}, no non-trivial linear combination of $\theta$, $\alpha^2$ belongs to $\ker p$. By \ref{_r_H^*_to_H_r_Claim_} (ii), the classes $\theta$, $\alpha^2$ generate a 2-dimensional subspace in $H^4(M, {\Bbb Q})/\ker r$. By assumptions of \ref{_C_restri_from_A,d_Lemma_}, the group $H^{2,2}_I(M)\cap H^4(M, {\Bbb Z})/(\ker r)$ has rank 2. Therefore $\omega$ and $\alpha^2$ generate the space \[ H^{2,2}_I(M)\cap H^4(M, {\Bbb Q})/(\ker r). \] To prove \ref{_C_restri_from_A,d_Lemma_} it suffices to show that for all integer classes \[ \beta = a\alpha^2 + b \theta, \ \ a \in {\Bbb Q}\backslash 0, \ \ \deg_I \beta \geq \deg_J \beta, \] we have $|\deg_I \beta| >A \cdot d$, for a constant $A$ depending only on volume, topology and dimension of $M$. Since $\deg_I \beta \geq |\deg_J \beta|$, and $\deg_J \alpha^2 =0$ (\ref{_degrees_theta_alpha^2_Sublemma_}), we have \[ \deg_I (a \alpha^2 + b \theta) \geq |\deg_J b \theta|. \] Therefore, either $a$ and $b$ have the same sign, or $\deg_I (a \alpha^2) > 2\deg_I(b\theta)$. In both cases, \begin{equation}\label{_deg_beta_geq_deg_alpha_Equation_} |\deg_I \beta|\geq \frac{1}{2}\deg_I (a \alpha^2). \end{equation} Let $x\in {\Bbb Q}^{>0}$ be the smallest positive rational value of $a$ for which there exists an integer class $\beta = a\alpha^2 + b \theta$. We have an integer lattice $L_1$ in $H^4_r(M)$ provided by the products of integer classes; the integer lattice $L_2\supset L_1$ provided by integer cycles might be different from that one. Clearly, $x$ is greater than determinant $\det (L_1 /L_2)$ of $L_1$ over $L_2$, and this determinant is determined by the topology of $M$. Form the definition of $x$ and \eqref{_deg_beta_geq_deg_alpha_Equation_}, we have $|\deg_I \beta| > x^2 \deg_I (\alpha^2)$. On the other hand, $\deg_I (\alpha^2)> C \deg_I(\alpha)$, where $C$ is a constant depending on volume and dimension of $M$. Setting $A:= x^2 \cdot C$, we obtain $|\deg_I \beta| > x^2 \cdot C \cdot d$. This proves \ref{_C_restri_from_A,d_Lemma_}. \blacksquare \hfill Consider the maps \[ \widetilde\deg:\; H^2(M, {\Bbb Z})\backslash H^2_{inv}(M) {\:\longrightarrow\:} {\Bbb R}, \] \[ c:\; H^2(M)\backslash H^2_{inv}(M) {\:\longrightarrow\:} \c R/\{\pm 1\} \] introduced in the beginning of the proof of \ref{_restri_for_H^11_1-dim-Proposition_}. \hfill \lemma \label{_dense_big_tilde_deg_Lemma_} In assumptions of \ref{_restri_for_H^11_1-dim-Proposition_}, let \[ {\cal O}\subset H^2(M, {\Bbb R})\backslash H^2_{inv}(M) \] be an open subset of $H^2(M, {\Bbb R})$, such that for all $x\in {\cal O}$, $k\in {\Bbb R}^{>0}$, we have $k\cdot x \in {\cal O}$. Assume that ${\cal O}$ contains the K\"ahler class $\omega_I$ for all induced complex structures $I\in \c R$. For a positive number $C\in {\Bbb R}^{>0}$, consider the set $X_C\subset {\cal O}$ \[ X_C := \left\{ \vphantom\prod\alpha \in {\cal O} \cap H^2(M, {\Bbb Z})\;\; | \;\; \widetilde \deg(\alpha) \geq C \right\}. \] Then $c(X_C)$ is dense in $\c R/\{\pm 1\}$ for all $C\in {\Bbb R}^{>0}$. \hfill {\bf Proof:} The map $\widetilde \deg$ can be expressed in the following wey. We call an integer cohomology class $\alpha\in H^2(M, {\Bbb Z})$ {\bf indivisible} if its largest divisor is 1, that is, there are no integer classes $\alpha'$, and numbers $k\in {\Bbb Z}$, $k>1$, such that $\alpha = k \alpha'$. \hfill \sublemma Let $\alpha \in H^2(M)$ be an non-$SU(2)$-invariant cohomology class and $\alpha= \alpha_{inv} + \alpha_+$ be a decomposition associated with \eqref{_H^2_isoty_Equation_}. Assume that $\alpha$ is indivisible. Then \begin{equation}\label{_tilde_deg_Equation_} \widetilde \deg (\alpha) = C\sqrt{((\alpha_+,\alpha_+)_{\c B})}, \end{equation} where $(\cdot,\cdot)_{\c B}$ is the Bogomolov-Beauville pairing on $H^2(M)$ (\cite{_coho_announce_}; see also \ref{_gene_all_SU(2)_Theorem_}), and $C$ a constant depending on $\dim M$, $\operatorname{Vol} M$. {\bf Proof:} By \ref{_degree_isotypic_Proposition_}, \[ \deg_I(\alpha) = \deg_I(\alpha_+) \] (clearly, $p(\alpha) = \alpha_+$). By definition of $(\cdot,\cdot)_{\c B}$, we have \[ \deg_I(\alpha_+)=(\alpha_+,\omega_{c(\alpha)})_{\c B} \] On the other hand, $\alpha_+$ is collinear with $\omega_{c(\alpha)}$ by definition of the map $c$. Now \eqref{_tilde_deg_Equation_} follows trivially from routine properties of bilinear forms. \blacksquare \hfill Let $I$ be an induced complex structure such that the cohomology class $\omega_I$ is irrational: $\omega_I \notin H^2(M, {\Bbb Q})$. \begin{equation}\label{_sequence_x_i_Equation_} \begin{minipage}[m]{0.8\linewidth} To prove \ref{_dense_big_tilde_deg_Lemma_}, we have to produce a sequence $x_i \in {\cal O} \cap H^2(M, {\Bbb Z})$ such that \begin{description} \item[(i)] $c(x_i)$ converges to $I$, \item[(ii)] and $\lim\tilde \deg (x_i) = \infty$. \end{description} \end{minipage} \end{equation} We introduce a metric $(\cdot,\cdot)_{\c H}$ on $H^2(M, {\Bbb R})$, \[ (\alpha, \beta)_{\c H}:= (\alpha_+,\beta_+)_{\c B} - (\alpha_{inv},\beta_{inv})_{\c B}. \] It is easy to check that $(\cdot,\cdot)_{\c H}$ is positive definite (\cite{_Verbitsky:cohomo_}). For every $\epsilon$, there exists a rational class $\omega_\epsilon \in H^2(M, {\Bbb Q})$ which approximates $\omega_I$ with precision \[ (\omega_\epsilon - \omega_I, \omega_\epsilon - \omega_I)_{\c H} <\epsilon. \] Since ${\cal O}$ is open and contains $\omega_I$, we may assume that $\omega_\epsilon$ belongs to ${\cal O}$. Take a sequence $\epsilon_i$ converging to $0$, and let $\tilde x_i:= \omega_{\epsilon_i}$ be the corresponding sequence of rational cohomology cycles. Let $x_i:= \lambda_i\tilde x_i$ be the minimal positive integer such that $x_i \in H^2(M, {\Bbb Z})$. We are going to show that the sequence $x_i$ satisfies the conditions of \eqref{_sequence_x_i_Equation_}. First of all, $\tilde x_i$ converges to $\omega_I$, and the map \[ c:\; H^2(M)\backslash H^2_{inv}(M) {\:\longrightarrow\:} \c R/\{\pm 1\} \] is continuous. Therefore, $\lim c(\tilde x_i) = c(\omega_I) = I$. By construction of $c$, $c$ satisfies $c(x) = c(\lambda x)$, and thus, $c(x_i) = c(\tilde x_i)$. This proves the condition (i) of \eqref{_sequence_x_i_Equation_}. On the other hand, since $\omega_I$ is irrational, the sequence $\lambda_i$ goes to infinity. Therefore, \[ \lim (x_i, x_i)_{\c H} = \infty. \] It remains to compare $(x_i, x_i)_{\c H}$ with $\tilde \deg x_i$. By \eqref{_tilde_deg_Equation_}, \[ \tilde \deg x_i = \sqrt{((x_i)_+, (x_i)_+)_{\c B}}. \] On the other hand, since $(x_i)_+\in H^2_+(M)$, we have \[ ((x_i)_+, (x_i)_+)_{\c B} = ((x_i)_+, (x_i)_+)_{\c H}.\] To prove \eqref{_sequence_x_i_Equation_} (ii), it remains to show that \[ \lim ((x_i)_+, (x_i)_+)_{\c H} = \lim (x_i, x_i)_{\c H}. \] Since the cohomology class $\tilde x_i\in H^2(M, {\Bbb Q})$ $\epsilon$-approximates $\omega_I$, and $\omega_I$ belongs to $H^2_+(M)$, we have \[ (\tilde x_i - (\tilde x_i)_+, \tilde x_i - (\tilde x_i)_+)_{\c H} <\epsilon_i. \] Therefore, \begin{equation}\label{_x_i_close_x_i_+_Equation_} ( x_i - ( x_i)_+, x_i - ( x_i)_+)_{\c H} <\lambda_i \epsilon_i. \end{equation} On the other hand, for $i$ sufficiently big, the cohomology class $\tilde x_i$ approaches $\omega_I$, and \begin{equation}\label{_x_i_bigger_Equation_} (x_i, x_i)_{\c H} > \frac 1 2 \lambda_i (\omega_I, \omega_I)_{\c H} \end{equation} Comparing \eqref{_x_i_close_x_i_+_Equation_} and \eqref{_x_i_bigger_Equation_} and using the distance property for the distance given by $\sqrt{(\cdot,\cdot)_{\c H}}$, we find that \begin{equation}\label{_x_i_+_bigger_Equation_} \sqrt{(x_i)_+, (x_i)_+} > \sqrt{\frac 1 2 \lambda_i (\omega_I, \omega_I)_{\c H}} - \sqrt{\lambda_i \epsilon_i} = \sqrt{\lambda_i}\cdot \left(\sqrt{\frac 1 2 (\omega_I, \omega_I)_{\c H}} - \sqrt{\epsilon_i} \right). \end{equation} Since $\epsilon_i$ converges to 0 and $\lambda_i$ converges to infinity, the right hand side of \eqref{_x_i_+_bigger_Equation_} converges to infinity. On the other hand, by \eqref{_tilde_deg_Equation_} the left hand side of \eqref{_x_i_+_bigger_Equation_} is equal constant times $\tilde \deg x_i$, so $\lim \tilde \deg x_i = \infty$. This proves the second condition of \eqref{_sequence_x_i_Equation_}. \ref{_dense_big_tilde_deg_Lemma_} is proven. \blacksquare \hfill We use \ref{_C_restri_from_A,d_Lemma_} and \ref{_dense_big_tilde_deg_Lemma_} in order to finish the proof of \ref{_restri_for_H^11_1-dim-Proposition_}. Let $M$ be a compact hyperk\"ahler manifold, and ${\cal O}\subset H^2(M, {\Bbb R})$ be the set of all K\"ahler classes for the K\"ahler metrics compatible with one of induced complex structures. By \ref{_K_proje_on_R_Lemma_}, ${\cal O}$ is open in $H^2(M, {\Bbb R})$. Applying \ref{_dense_big_tilde_deg_Lemma_} to ${\cal O}$, we obtain the following. In assumptions of \ref{_restri_for_H^11_1-dim-Proposition_}, let $Y_C\subset \c R$ be the set of all algebraic induced complex structures $I$ with $\tilde \deg \alpha >C$, where $\alpha$ is a rational K\"ahler class, $\alpha \in H^{1,1}(M) \cap H^2(M, {\Bbb Z})$. Then $Y_C$ is dense in $\c R$. Now, \ref{_C_restri_from_A,d_Lemma_}, implies that for all $I \in Y_C$, the induced complex structure $I$ is $A \cdot C$-restricted, where $A$ is the universal constant of \ref{_C_restri_from_A,d_Lemma_}. Thus, for all $C$ the set of $C$-restricted induced complex structures is dense in $\c R$. This proves that $M$ admits $C$-restricted complex structures. We finished the proof of \ref{_restri_for_H^11_1-dim-Proposition_}. \blacksquare \subsection{Hyperk\"ahler structures admitting $C$-restricted complex structures} \label{_modu_and_C-restri_Subsection_} Let $M$ be a compact complex manifold admitting a hyperk\"ahler structure $\c H$. Assume that $(M, \c H)$ is a simple hyperk\"ahler manifold of dimension $\dim_{\Bbb H} M >1$. The following definition of (coarse, marked) moduli space for complex and hyperk\"ahler structures on $M$ is standard. \hfill \definition\label{_moduli_hyperka_Definition_} Let $M_{C^\infty}$ be the $M$ considered as a differential manifold, $\widetilde{Comp}$ be the set of all integrable complex structures, and $\widetilde{\mbox{\it Hyp}}$ be the set of all hyperk\"ahler structures on $M_{C^\infty}$. The set $\widetilde{\mbox{\it Hyp}}$ is equipped with a natural topology. Let $\widetilde{\mbox{\it Hyp}}^0$ be a connected component of $\widetilde{\mbox{\it Hyp}}$ containing $\c H$ and $\widetilde{Comp}^0$ be a set of all complex structures $I\in \widetilde{Comp}$ which are compatible with some hyperk\"ahler structure $\c H_1\in \widetilde{\mbox{\it Hyp}}^0$. Let $\mbox{\it Diff}$ be the group of diffeomorphisms of $M$ which act trivially on the cohomology. The coarse, marked moduli $\mbox{\it Hyp}$ of hyperk\"ahler structures on $M$ is the quotient $\mbox{\it Hyp}:= \widetilde{\mbox{\it Hyp}}^0/\mbox{\it Diff}$ equipped with a natural topology. The coarse, marked moduli $Comp$ of complex structures on $M$ is defined as $Comp:= \widetilde{Comp}^0/\mbox{\it Diff}$. For a detailed discussion of various aspects of this definition, see \cite{_Verbitsky:cohomo_}. \hfill Consider the variety \[ X \subset {\Bbb P} H^2(M, {\Bbb C}),\] consisting of all lines $l\in {\Bbb P} H^2(M, {\Bbb C})$ which are isotropic with respect to the Bogomolov-Beauville's pairing: \[ X:= \{ l \in H^2(M, {\Bbb C}) \; \; | \; (l, l)_{\c B} =0 \}. \] Since $M$ is simple, $\dim H^{2,0}(M, I) =1$ for all induced complex structures. Let $P_c:\; Comp {\:\longrightarrow\:} {\Bbb P} H^2(M, {\Bbb C})$ map $I$ to the line $H^{2,0}_I(M) \subset H^2(M, {\Bbb C})$. The map $P_c$ is called {\bf the period map}. It is well known that $Comp$ is equipped with a natural complex structure. From general properties of the period map it follows that $P_c$ is compatible with this complex structure. Clearly from the definition of Bogomolov-Beauville's form, $P_c(I)\in X$ for all induced complex structures $I\in Comp$ (see \cite{_Beauville_} for details). \hfill \theorem \label{_Bogomo_etale_Theorem_} \cite{_Besse:Einst_Manifo_} (Bogomolov) The complex analytic map \[ P_c:\; Comp {\:\longrightarrow\:} X\] is locally an etale covering. \footnote{ The space $Comp$ is smooth, as follows from \ref{_Bogomo_etale_Theorem_}. This space is, however, in most cases not separable (\cite{_Huybrechts_}). The space $\mbox{\it Hyp}$ has no natural complex structures, and can be odd-dimensional. } \blacksquare \hfill It is possible to formulate a similar statement about hyperk\"ahler structures. For a hyperk\"ahler structure $\c H$, consider the set $\c R_{\c H}\subset Comp$ of all induced complex structures associated with this hyperk\"ahler structure. The subset $\c R_{\c H}\subset Comp$ is a complex analytic subvariety, which is isomorphic to ${\Bbb C} P^1$. Let $S:= P_c(\c R_{\c H})$ be the corresponding projective line in $X$, and $\bar L(X)$ be the space of smooth deformations of $S$ in $X$. The points of $L(X)$ correspond to smooth rational curves of degree 2 in ${\Bbb P} H^2(M, {\Bbb C})$. For every such curve $s$, there exists a unique 3-dimensional plane $L(s) \subset H^2(M, {\Bbb C})$, such that $s$ is contained in ${\Bbb P} L$. Let $Gr$ be the Grassmanian manifold of all 3-dimensional planes in $H^2(M, {\Bbb C})$ and $Gr_0\subset Gr$ the set of all planes $L\in Gr$ such that the restriction of the Bogomolov-Beauville form to $L$ is non-degenerate. Let $L(X)\subset \bar L(X)$ be the space of all rational curves $s\in \bar L(X)$ such that the restriction of the Bogomolov-Beauville form to $L(s)$ is non-degenerate: $L(s)\in Gr_0$. The correspondence $s{\:\longrightarrow\:} L(s)$ gives a map $\kappa:\; L(X) {\:\longrightarrow\:} Gr_0$. \hfill \lemma The map $\kappa:\; L(X) {\:\longrightarrow\:} Gr_0$ is an isomorphism of complex varieties. {\bf Proof:} For every plane $L\in Gr_0$, consider the set $s(L)$ of all isotropic lines $l\in L$, that is, lines satisfying $(l, l)_{\c B}=0$. Since $(\cdot,\cdot)_{\c B}\restrict L$ is non-degenerate, the set $s(L)$ is a rational curve in ${\Bbb P}L$. Clearly, this curve has degree 2. Therefore, $s(L)$ belongs to $X(L)$. The map $L{\:\longrightarrow\:} s(L)$ is inverse to $\kappa$. \blacksquare \hfill Consider the standard anticomplex involution \[ \iota:\; H^2(M, {\Bbb C}) {\:\longrightarrow\:} H^2(M, {\Bbb C}), \ \ \ \eta{\:\longrightarrow\:} \bar \eta. \] Clearly, $\iota$ is compatible with the Bogomolov-Beauville form. Therefore, $\iota$ acts on $L(X)$ as an anticomplex involution. Let $L(X)_\iota\subset L(X)$ be the set of all $S\in L(X)$ fixed by $\iota$. \hfill Every hyperk\"ahler structure \[ \c H\in \mbox{\it Hyp} \] gives a rational curve $\c R_{\c H}\subset Comp$ with points corresponding all induced complex structures. Let $P_h(\c H)\subset X$ be the line $P_c(\c R_{\c H})$. Clearly from the definition, $P_h(\c H)$ belongs to $L(X)_\iota$. We have constructed a map $P_h:\; \mbox{\it Hyp} {\:\longrightarrow\:} L(X)_\iota$. Let $L(Comp)$ be the space of deformations if $\c R_{\c H}$ in $Comp$. Denote by \[ \gamma:\; \mbox{\it Hyp} {\:\longrightarrow\:} L(Comp) \] the map $\c H {\:\longrightarrow\:} \c R_{\c H}$. The following result gives a hyperk\"ahler analogue of Bogomolov's theorem (\ref{_Bogomo_etale_Theorem_}). \hfill \theorem \label{_hyperka_etale_Theorem_} The map $\gamma:\; \mbox{\it Hyp} {\:\longrightarrow\:} L(Comp)$ is an embedding. The map $P_h:\; \mbox{\it Hyp} {\:\longrightarrow\:} L(X)_\iota$ is locally a covering. {\bf Proof:} The first claim is an immediate consequence of Calabi-Yau Theorem (\ref{_symplectic_=>_hyperkahler_Proposition_}). Now, \ref{_hyperka_etale_Theorem_} follows from the Bogomolov's theorem (\ref{_Bogomo_etale_Theorem_}) and dimension count. \blacksquare \hfill Let $I\in Comp$ be a complex structure on $M$. Consider the groups \[ H_h^2(M, I):= H^{1,1}(M, I) \cap H^2(M, {\Bbb Z}) \] and \[ H_h^2(M, I):= H^{2,2}_r(M, I) \cap H^4(M, {\Bbb Z}). \] For a general $I$, $H^2_h(M, I)=0$ and $H^4_h(M, I)={\Bbb Z}$ as follows from \ref{_theta_SU(2)_inva_Proposition_}. Therefore, the set of all $I$ with $\operatorname{rk} H^2_h(M, I) =1$, $\operatorname{rk} H^4_h(M, I)=2$ is a union of countably many subvarieties of codimension 1 in $Comp$. Similarly, the set $V\subset Comp$ of all $I$ with $\operatorname{rk} H^2_h(M, I) >1$, $\operatorname{rk} H^4_h(M, I)>2$ is a union of countably many subvarieties of codimension more than 1. Together with \ref{_hyperka_etale_Theorem_}, this implies the following. \hfill \claim\label{_hype_1_2_dense_Claim_} Let $U\subset \mbox{\it Hyp}$ be the set of all $\c H\in Hyp$ such that $\c R_{\c H}$ does not intersect $V$. Then $U$ is dense in $\mbox{\it Hyp}$. \hfill {\bf Proof:} Consider a natural involution $i$ of $Comp$ which is compatible with the involution $\iota:\; X {\:\longrightarrow\:} X$ inder the period map $P_c:\; Comp {\:\longrightarrow\:} X$. This involution maps the complex structure $I$ to $-I$. \begin{equation}\label{_Hyp_identi_Equation_} \begin{minipage}[m]{0.8\linewidth} By \ref{_hyperka_etale_Theorem_}, $\mbox{\it Hyp}$ is identified with an open subset in the set $L(X)_\iota$ of real points of $L(Comp)$. \end{minipage} \end{equation} Let $L_U\subset L(Comp)$ be the set of all lines which do not intersect $V$. Since $V$ is a union of subvarieties of codimension at least 2, a general rational line $l \in L(Comp)$ does not intersect $V$. Therefore, $L_U$ is dense in $L(Comp)$. Thus, the set of real points of $L_U$ is dense $L(X)_\iota$. Using the idenitification \eqref{_Hyp_identi_Equation_}, we obtain the statement of \ref{_hype_1_2_dense_Claim_}. \blacksquare \hfill \ref{_hype_1_2_dense_Claim_} together with \ref{_restri_for_H^11_1-dim-Proposition_} imply the following theorem. \hfill \theorem\label{_C_restri_dense_Theorem_} Let $M$ be a compact simple hyperk\"ahler manifold, $\dim_{\Bbb H}M >1$, and $\mbox{\it Hyp}$ its coarse marked moduli of hyperk\"ahler structures. Let $U\subset \mbox{\it Hyp}$ be the set of all hyperk\"ahler structures which admit $C$-restricted complex structures (\ref{_admitti_C_restri_Definition_}). Then $U$ is dense in $\mbox{\it Hyp}$. \blacksquare \subsection{Deformations of coherent sheaves over manifolds with $C$-res\-t\-ric\-ted complex structures} The following theorem shows that a semistable deformation of a hyperholomorphic sheaf on $(M,I)$ is again hyperholomorphic, provided that $I$ is a $C$-restricted complex structure and $C$ is sufficiently big. \hfill \theorem \label{_sheaf_on_C_restr_hyperho_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, and $\c F \in \c F(M)$ a polystable hyperholomorphic sheaf on $M$ (\ref{_hyperho_shea_on_M_Definition_}). Let $I$ be a $C$-restricted induced complex structure, for $C= \deg_I c_2(\c F)$,\footnote{Clearly, since $\c F$ is hyperholomorphic, the class $c_2(\c F)$ is $SU(2)$-invariant, and the number $\deg_I c_2(\c F)$ independent from $I$.} and $F'$ be a semistable torsion-free coherent sheaf on $(M,I)$ with the same rank and Chern classes as $\c F$. Then the sheaf $F'$ is hyperholomorphic. \hfill {\bf Proof:} Let $F_1$, ..., $F_n$ be the Jordan-H\"older series for the sheaf $F'$. Since $\c F$ is hyperholomorphic, we have $\text{slope}(\c F)=0$ (\ref{_slope_hyperho_Remark_}). Therefore, $\operatorname{slope}(F_i)=0$, and $\deg_I(c_1(F_i))=0$. By \ref{_C_restri_Definition_} (i), then, the class $c_1(F_i)$ is $SU(2)$ invariant for all $i$. To prove that $F'$ is hyperholomorphic it remains to show that the classes $c_2(F_i)$, $c_2(F_i^{**})$ are $SU(2)$-invariant for all $i$. \hfill Consider an exact sequence \[ 0 {\:\longrightarrow\:} F_i {\:\longrightarrow\:} F^{**}_i {\:\longrightarrow\:} F_i / F_i^{**} {\:\longrightarrow\:} 0. \] Let $[F_i / F_i^{**}]\in H^4(M)$ be the fundamental class of the union of all components of $Sup(F_i / F_i^{**})$ of complex codimension 2, taken with appropriate multiplicities. Clearly, $c_2(F_i)= c_2(F^{**}_i) + [F_i / F_i^{**}]$. Since $[F_i / F_i^{**}]$ is an effective cycle, $\deg_I([F_i / F_i^{**}])\geq0$. By the Bogomolov-Miyaoka-Yau inequality (see \ref{_stable_shea_degree_Corollary_}), we have $\deg_I (c_2(F^{**}_i)\geq 0$. Therefore, \begin{equation}\label{_c_2_greater_Equation_} \deg_I c_2(F_i)\geq \deg_I c_2(F^{**}_i)\geq 0. \end{equation} \hfill Using the product formula for Chern classes, we obtain \begin{equation}\label{_pro_Chern_2_Equation_} c_2(F) = \sum_i c_2(F_i) + \sum_{i, j} c_2(F_i)\wedge c_2(F_j). \end{equation} By \ref{_p_prods_H^2_Corollary_}, $\deg_I(\sum_{i, j} c_2(F_i)\wedge c_2(F_j)) =0$. Since the numbers $\deg_I c_2(F_i)$ are non-negative, we have $\deg_I c_2(F_i)\leq \deg_I c_2(F) =C$. By \ref{_funda_and_Chern_CA_Theorem_}, the classes $c_2(F_i)$, $c_2(F_i^{**})$ are of CA-type. By \ref{_C_restri_Definition_} (ii), then, the inequality $\deg_I c_2(F_i)\leq C$ implies that the class $c_2(F_i)$ is $SU(2)$-invariant. By \eqref{_c_2_greater_Equation_}, $\deg_I c_2(F^{**}_i)\leq\deg_I c_2(F_i)$, so the class $c_2(F_i^{**})$ is also $SU(2)$-invariant. \ref{_sheaf_on_C_restr_hyperho_Theorem_} is proven. \blacksquare \section{Desingularization of hyperholomorphic sheaves} \label{_desingu_she_Section_} The aim of this section is the following theorem. \hfill \theorem\label{_desingu_hyperho_Theorem_} Let $M$ be a hyperk\"ahler manifold, not necessarily compact, $I$ an induced complex structure, and $F$ a reflexive coherent sheaf over $(M, I)$ equipped with a hyperholomorphic connection (\ref{_hyperholo_co_Definition_}). Assume that $F$ has isolated singularities. Let $\tilde M\stackrel \sigma{\:\longrightarrow\:} M$ be a blow-up of $(M,I)$ in the singular set of $F$, and $\sigma^* F$ the pullback of $F$. Then $\sigma^* F$ is a locally trivial sheaf, that is, a holomorphic vector bundle. \hfill We prove \ref{_desingu_hyperho_Theorem_} in Subsection \ref{_desingu_she_Subsection_}. \hfill The idea of the proof is the following. We apply to $F$ the methods used in the proof of Desingularization Theorem (\ref{_desingu_Theorem_}). The main ingredient in the proof of Desingularization Theorem is the existence of a natural ${\Bbb C}^*$-action on the completion $\hat {\cal O}_x(M, I)$ of the local ring ${\cal O}_x(M, I)$, for all $x\in M$. This ${\Bbb C}^*$-action identifies $\hat {\cal O}_x(M, I)$ with a completion of a graded ring. Here we show that a sheaf $F$ is ${\Bbb C}^*$-equivariant. Therefore, a germ of $F$ at $x$ has a grading, which is compatible with the natural ${\Bbb C}^*$-action on $\hat {\cal O}_x(M, I)$. Singularities of such reflexive sheaves can be resolved by a single blow-up. \subsection{Twistor lines and complexification} \label{_twi_lines_C^*_Subsection_} Further on, we need the following definition. \hfill \definition Let $X$ be a real analytic variety, which is embedded to a complex variety $X_{\Bbb C}$. Assume that the sheaf of complex-valued real analytic functions on $X$ coincides with the restriction of ${\cal O}_{X_{\Bbb C}}$ to $X\subset X_{\Bbb C}$. Then $X_{\Bbb C}$ is called {\bf a complexification of $X$}. \hfill For more details on complexification, the reader is referred to \cite{_GMT_}. There are the most important properties. \hfill \claim \label{_complexi_Claim_} In a neighbourhood of $X$, the manifold $X_{\Bbb C}$ has an anti-complex involution. The variety $X$ is identified with the set of fixed points of this involution, considered as a real analytic variety. Let $Y$ be a complex variety, and $X$ the underlying real analytic variety. Then the product of $Y$ and its complex conjugate is a complexification of $X$, with embedding $X\hookrightarrow Y \times \bar Y$ given by the diagonal. The complexification is unique in the following weak sense. For $X_{\Bbb C}$, $X'_{\Bbb C}$ complexifications of $C$, the complex manifolds $X_{\Bbb C}$, $X'_{\Bbb C}$ are naturally identified in a neighbourhood of $X$. \blacksquare \hfill Let $M$ be a hyperk\"ahler manifold, $\operatorname{Tw}(M)$ its twistor space, and $\pi:\; \operatorname{Tw}(M) {\:\longrightarrow\:} {\Bbb C} P^1$ the twistor projection. Let $l\subset \operatorname{Tw}(M)$ be a rational curve, such that the restriction of $\pi$ to $l$ is an identity. Such a curve gives a section of $\pi$, and vice versa, every section of $\pi$ corresponds to such a curve. The set of sections of the projection $\pi$ is called {\bf the space of twistor lines}, denoted by $\operatorname{Lin}$, or $\operatorname{Lin}(M)$. This space is equipped with complex structure, by Douady (\cite{_Douady_}). Let $m\in M$ be a point. Consider a twistor line ${s_m}:\; I {\:\longrightarrow\:} (I \times m)\in {\Bbb C} P^1 \times M = \operatorname{Tw}$. Then $s_m$ is called {\bf a horisontal twistor line}. The space of horisontal twistor lines is a real analytic subvariety in $\operatorname{Lin}$, denoted by $\operatorname{Hor}$, or $\operatorname{Hor}(M)$. Clearly, the set $\operatorname{Hor}$ is naturally identified with $M$. \hfill \proposition\label{_Lin_is_MxM_Proposition_} (Hitchin, Karlhede, Lindstr\"{o}m, Ro\v{c}ek) Let $M$ be a hyperk\"ahler manifold, $\operatorname{Tw}(M)$ its twistor space, $I$, $J\in {\Bbb C} P^1$ induced complex structures, and $\operatorname{Lin}$ the space of twistor lines. The complex manifolds $(M,I)$ and $(M, J)$ are naturally embedded to $\operatorname{Tw}(M)$: \[ (M,I) = \pi^{-1}(I),\ \ (M,J) = \pi^{-1}(J). \] Consider a point $s\in \operatorname{Lin}$, $s:\; {\Bbb C} P^1{\:\longrightarrow\:} \operatorname{Tw}(M)$. Let \[ ev_{I,J}:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, I)\times (M,J) \] be the map defined by $ev_{I,J}(s) = (s(I), s(J))$. Assume that $I\neq J$. Then there exists a neighbourhood $U$ of $\operatorname{Hor}\subset \operatorname{Lin}$, such that the restriction of $ev_{I,J}$ to $U$ is an open embedding. {\bf Proof:} \cite{_HKLR_}, \cite{_Verbitsky:hypercomple_}. \blacksquare \hfill Consider the anticomplex involution $i$ of ${\Bbb C} P^1 \cong S^2$ which corresponds to the central symmetry of $S^2$. Let $\iota:\; \operatorname{Tw} {\:\longrightarrow\:} \operatorname{Tw}$ be the corresponding involution of the twistor space $\operatorname{Tw}(M) = {\Bbb C} P^1\times M$, $(x, m) {\:\longrightarrow\:} (i(x), m)$. It is clear that $\iota$ maps holomorphic subvarieties of $\operatorname{Tw}(M)$ to holomorphic subvarieties. Therefore, $\iota$ acts on $\operatorname{Lin}$ as an anticomplex involution. For $J= -I$, we obtain a local identification of $\operatorname{Lin}$ in a neighbourhood of $\operatorname{Hor}$ with $(M, I) \times (M, -I)$, that is, with $(M, I)$ times its complex conjugate. Therefore, the space of twistor lines is a complexification of $(M,I)$. The natural anticomplex involution of \ref{_complexi_Claim_} coincides with $\iota$. This gives an identification of $\operatorname{Hor}$ and the real analytic manifold underlying $(M,I)$. \hfill We shall explain how to construct the natural ${\Bbb C}^*$-action on a local ring of a hyperk\"ahler manifold, using the machinery of twistor lines. \hfill Fix a point $x_0\in M$ and induced complex structures $I$, $J$, such that $I\neq \pm J$. Let $V_1$, $V_2$ be neighbourhoods of $s_{x_0}\in \operatorname{Lin}$, and $U_1$, $U_2$ be neighbourhoods of $(x_0,x_0)$ in $(M, I)\times (M,-I)$, $(M, J)\times (M,-J)$, such that the evaluation maps $ev_{I, -I}$, $ev_{J, -J}$ induce isomorphisms \[ ev_{I, -I}:\; V_1 \oldtilde{\:\longrightarrow\:} U_1, \ \ \ \ ev_{J, -J}:\; V_2\oldtilde{\:\longrightarrow\:} U_2. \] Let $B$ be an open neighbourhood of $x_0\in M$, such that $(B, I)\times (B,-I)\subset U_1$ and $(B, I)\times (B,-I)\subset U_2$. Denote by $V_I\subset V_1$ be the preimage of $(B, I)\times (B,-I)$ under $ev_{I, -I}$, and by $V_J\subset V_2$ be the preimage of $(B, J)\times (B,-J)$ under $ev_{J, -J}$. Let $p_I:\; V_I {\:\longrightarrow\:} (B, I)$ be the evaluation, $s{\:\longrightarrow\:} s(I)$, and $e_I:\; (B, I){\:\longrightarrow\:} V_1$ the map associating to $x\in B$ the unique twistor line passing through $(x, x_0) \subset (B, I)\times (B,-I)$. In the same fashion, we define $e_J$ and $p_J$. We are interested in the composition \[ \Psi_{I, J}:= e_I\circ p_J \circ e_J \circ p_I:\; (B_0, I) {\:\longrightarrow\:} (B,I) \] which is defined in a smaller neighbourhood $B_0\subset B$ of $x_0\in M$. \hfill The following proposition is the focal point of this Subsection: we explain the map $\Psi_{I, J}$ of \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_} is geometric terms (in \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_} this map was defined algebraically). \hfill \proposition \label{_Psi_acts_on_TM_Proposition_} Consider the map $\Psi_{I, J}:\; (B_0, I) {\:\longrightarrow\:} (B,I)$ defined above. By definition, $\Psi_{I, J}$ preserves the point $x_0\in B_0\subset B$. Let $d\Psi_{I, J}$ be the differential of $\Psi_{I, J}$ acting on the tangent space $T_{x_0}B_0$. Assume that $I\neq \pm J$. Then $d\Psi_{I, J}$ is a multiplication by a scalar $\lambda\in {\Bbb C}$, $0<|\lambda| <1$. \hfill {\bf Proof:} The map $\Psi_{I, J}$ was defined in \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_} using the identifications between the real analytic varieties underlying $(M,I)$ and $(M, J)$. We proved that $\Psi_{I, J}$ defined this way acts on $T_{x_0}B_0$ as a multiplication by the scalar $\lambda\in {\Bbb C}$, $0<|\lambda| <1$. It remains to show that the map $\Psi_{I, J}$ defined in \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_} coincides with $\Psi_{I, J}$ defined above. Consider the natural identification \[ (B, I)\times (B,-I)\sim (B, J)\times (B,-J), \] which is defined in a neighbourhood $B_{\Bbb C}$ of $(x_0, x_0)$. There is a natural projection $a_I:\; B_{\Bbb C} {\:\longrightarrow\:} (M,I)$. Consider the embedding $b_I:\; (B, I) {\:\longrightarrow\:} B_{\Bbb C}$, $x{\:\longrightarrow\:} (x, x_0)$, defined in a neighbourhood of $x_0\in (B,I)$. In a similar way we define $a_J$, $b_J$. In \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_} we defined $\Psi_{I, J}$ as a composition $b_I\circ a_J \circ b_J \circ a_I$. Earlier in this Subsection, we described a local identification of $(B, I)\times (B,-I)$ and $\operatorname{Lin}(B)$. Clearly, under this identification, the maps $a_I$, $b_I$ correspond to $p_I$, $e_I$. Therefore, the definition of $\Psi_{I,J}$ given in this paper is equivalent to the definition given in \cite{_Verbitsky:DesinguII_}, \cite{_Verbitsky:hypercomple_}. \blacksquare \subsection[The automorphism $\Psi_{I,J}$ acting on hyperholomorphic sheaves]{The automorphism $\Psi_{I,J}$ acting on hyperholomorphic \\ sheaves} \label{_Psi_on_shea_Subsection_} In this section, we prove that hyperholomorphic sheaves are equivariant with respect to the map $\Psi_{I, J}$, considered as an automorphism of the local ring ${\cal O}_{x_0}(M,I)$. \hfill \theorem \label{_Psi_equiv_hyperho_Theorem_} Let $M$ be a hyperk\"ahler manifold, not necessarily compact, $x_0\in M$ a point, $I$ an induced complex structure and $F$ a reflexive sheaf over $(M,I)$ equipped with a hyperholomorphic connection. Let $J\neq \pm I$ be another induced complex structure, and $B_0$, $B$ the neighbourhoods of $x_0\in M$ for which the map $\Psi_{I,J}:\; B_0 {\:\longrightarrow\:} B$ was defined in \ref{_Psi_acts_on_TM_Proposition_}. Assume that $\Psi_{I,J}:\; B_0 {\:\longrightarrow\:} B$ is an isomorphism. Then there exists a canonical functorial isomorphism of coherent sheaves \[ \Psi_{I,J}^F:\; F\restrict {B_0} {\:\longrightarrow\:} \Psi_{I,J}^*(F\restrict B). \] {\bf Proof:} Return to the notation introduced in Subsection \ref{_twi_lines_C^*_Subsection_}. Let $W:= V_I \cap V_J$. By definition of $V_I$, $V_J$, the evaluation maps produce open embeddings \[ ev_{I, -I}:\; \operatorname{Lin}(W) \hookrightarrow (W,I)\times (W, -I),\] and \[ ev_{J, -J}:\; \operatorname{Lin}(W) \hookrightarrow (W,J)\times (W, -J),\] Let $S\subset W$ be the singular set of $F\restrict W$, $\operatorname{Tw}(S) \subset \operatorname{Tw}(W)$ the corresponding embedding, and $L_0\subset \operatorname{Lin}(W)$ be the set of all lines $l\in \operatorname{Lin}(W)$ which do not intersect $\operatorname{Tw}(S)$. Consider the maps \[ p_I:\; L_0 \hookrightarrow (W,I)\backslash S \] and \[ p_J:\; L_0 \hookrightarrow (W,J)\backslash S \] obtained by restricting the evaluation map $p_I:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, I)$ to $L_0\subset \operatorname{Lin}(M)$. Since $F$ is equipped with a hyperholomorphic connection, the vector bundle $F\restrict{(M, J)\backslash S}$ has a natural holomorphic structure. Let $\underline F_1:= p_I^*\left(F\restrict{(M, I)\backslash S}\right)$ and $\underline F_2:= p_J^*\left(F\restrict{(M, J)\backslash S}\right)$ be the corresponding pullback sheaves over $L_0$, and $F_1$, $F_2$ the sheaves on $\operatorname{Lin}(W)$ obtained as direct images of $\underline F_1$, $\underline F_2$ under the open embedding $L_0 \hookrightarrow \operatorname{Lin}(W)$. \hfill \lemma \label{_F_1_=F_2_Lemma_} Under these assumptions, the sheaves $F_1$, $F_2$ are coherent reflexive sheaves. Moreover, there exists a natural isomorphism of coherent sheaves $\Psi_{1,2}:\;F_1{\:\longrightarrow\:} F_2$. \hfill {\bf Proof:} The complex codimension of the singular set $S$ in $(M,I)$ is at least 3, because $F$ is reflexive (\cite{_OSS_}, Ch. II, 1.1.10). Since $S$ is trianalytic (\ref{_singu_triana_Claim_}), this codimension is even. Thus, $\operatorname{codim}_{\Bbb C} (S, (M,I)) \geq 4$. Therefore, \[ \operatorname{codim}_{\Bbb C} (\operatorname{Tw}(S), \operatorname{Tw}(M)) \geq 4. \] Consider the set $L_S$ of all twistor lines $l\in \operatorname{Lin}(W)$ passing through $\operatorname{Tw}(S)$. For generic points $x,y\in \operatorname{Tw}(W)$, there exists a unique line $l\in \operatorname{Lin}(W)$ passing through $x, y$. Therefore, \[ \operatorname{codim}_{\Bbb C} (L_S, \operatorname{Lin}(W)) = \operatorname{codim}_{\Bbb C} (\operatorname{Tw}(S), \operatorname{Tw}(M))-1\geq 3. \] By definition, $L_0:= \operatorname{Lin}(W)\backslash L_S$. Since $F_1$, $F_2$ are direct images of bundles $\underline F_1$, $\underline F_2$ over a subvariety $L_S$ of codimension 3, these sheaves are coherent and reflexive (\cite{_OSS_}, Ch. II, 1.1.12; see also \ref{_normal_refle_Lemma_}). To show that they are naturally isomorphic it remains to construct an isomorphism between $\underline F_1$ and $\underline F_2$. Let $\c F$ be a coherent sheaf on $\operatorname{Tw}(W)$ obtained from $F\restrict W$ as in the proof of \ref{_conne_=>_hyperho_Proposition_}. The singular set of $\c F$ is $\operatorname{Tw}(S)\subset \operatorname{Tw}(W)$. Therefore, the restriction $\c F\restrict{\operatorname{Tw}(W)\backslash \operatorname{Tw}(S)}$ is a holomorphic vector bundle. For all horisontal twistor lines $l_x\subset \operatorname{Tw}(W)\backslash \operatorname{Tw}(S)$, the restriction $\c F\restrict {l_x}$ is clearly a trivial vector bundle over $l_x\cong {\Bbb C} P^1$. A small deformation of a trivial vector bundle is again trivial. Shrinking $W$ if necessary, we may assume that for all lines $l\in L_0$, the restriction of $\c F$ to $l\cong {\Bbb C} P^1$ is a trivial vector bundle. The isomorphism $\underline \Psi_{1,2}:\;\underline F_1{\:\longrightarrow\:}\underline F_2$ is constructed as follows. Let $l\in L_0$ be a twistor line. The restriction $\c F\restrict {l}$ is trivial. Consider $l$ as a map $l:\; {\Bbb C} P^1 {\:\longrightarrow\:} \operatorname{Tw}(M)$. We identify ${\Bbb C} P^1$ with the set of induced complex structures on $M$. By definition, the fiber of $F_1$ in $l$ is naturally identified with the space $\c F\restrict {l(I)}$, and the fiber of $F_2$ in $l$ is identifies with $\c F\restrict {l(J)}$. Since $\c F\restrict {l}$ is trivial, the fibers of the bundle $\c F\restrict {l}$ are naturally identified. This provides a vector bundle isomorphism $\underline \Psi_{1,2}:\;\underline F_1{\:\longrightarrow\:}\underline F_2$ mapping $\underline F_1 \restrict l = \c F\restrict {l(I)}$ to $\underline F_2 \restrict l = \c F\restrict {l(J)}$. It remains to show that this isomorphism is compatible with the holomorphic structure. Since the bundle $\c F\restrict l$ is trivial, we have an identification \[ \c F\restrict {l(I)}\cong\c F\restrict{l(J)}= \Gamma(\c F\restrict l), \] where $\Gamma(\c F\restrict l)$ is the space of global sections of $\c F\restrict l$. Thus, $F_i \restrict l = \Gamma(\c F\restrict l)$, and this identification is clearly holomorphic. This proves \ref{_F_1_=F_2_Lemma_}. \blacksquare \hfill We return to the proof of \ref{_Psi_equiv_hyperho_Theorem_}. Denote by $F_J$ the restriction of $\c F$ to $(M, J)= \pi^{-1}(J) \subset \operatorname{Tw}(M)$. The map $\Psi_{I, J}$ was defined as a composition $e_I\circ p_J \circ e_J \circ p_I$. The sheaf $p_I^* F$ is by definition isomorphic to $F_1$, and $p_J^* F_J$ to $F_2$. On the other hand, clearly, $e_J^* F_2= F_J$. Therefore, $(p_J \circ e_J)^* F_2 \cong F_2$. Using the isomorphism $F_1\cong F_2$, we obtain $(p_J \circ e_J)^* F_1 \cong F_1$. To sum it up, we have the following isomorphisms: \begin{equation*} \begin{split} p_I^* F &\cong F_1,\\ (p_J \circ e_J)^* F_1 &\cong F_1,\\ e_I^* F_1 &\cong F. \end{split} \end{equation*} A composition of these isomorphisms gives an isomorphism \[ \Psi_{I,J}^F:\; F\restrict {B_0} {\:\longrightarrow\:} \Psi_{I,J}^*(F\restrict B). \] This proves \ref{_Psi_equiv_hyperho_Theorem_}. \blacksquare \subsection{A ${\Bbb C}^*$-action on a local ring of a hyperk\"ahler manifold} \label{_C^*_action_on_loca_ring_Subsection_} Let $M$ be a hyperk\"ahler manifold, non necessarily compact, $x\in M$ a point and $I$, $J$ induced complex structures, $I\neq J$. Consider the complete local ring ${\cal O}_{x, I}:= \hat {\cal O}_x(M, I)$. Throughout this section we consider the map $\Psi_{I,J}$ (\ref{_Psi_acts_on_TM_Proposition_}) as an automorphism of the ring ${\cal O}_{x, I}$. Let $\frak m$ be the maximal ideal of ${\cal O}_{x, I}$, and $\frak m/\frak m^2$ the Zariski cotangent space of $(M,I)$ in $x$. \begin{equation} \label{_Psi_acts_on_cota_Equation_} \begin{minipage}[m]{0.8\linewidth} By \ref{_Psi_acts_on_TM_Proposition_}, $\Psi_{I,J}$ acts on $\frak m/\frak m^2$ as a multiplication by a number $\lambda\in{\Bbb C}$, $0<|\lambda|<1$. \end{minipage} \end{equation} Let $V_{\lambda^n}$ be the eigenspace corresponding to the eigenvalue $\lambda^n$, \[ V_{\lambda^n}:= \{ v\in {\cal O}_{x, I}\ \ |\ \ \Psi_{I,J}(v) = \lambda^n v\}. \] Clearly, $\oplus V_{\lambda^i}$ is a graded subring in ${\cal O}_{x, I}$. In \cite{_Verbitsky:DesinguII_}, (see also \cite{_Verbitsky:hypercomple_}) we proved that the ring $\oplus V_{\lambda^i}$ is dense in ${\cal O}_{x, I}$ with respect to the adic topology. Therefore, the ring ${\cal O}_{x, I}$ is identified with the adic completion of $\oplus V_{\lambda^i}$. \hfill Consider at action of ${\Bbb C}^*$ on $\oplus V_{\lambda^i}$, with $z\in {\Bbb C}^*$ acting on $V_{\lambda^i}$ as a multiplication by $z^i$. This ${\Bbb C}^*$-action is clearly continuous, with respect to the adic topology. Therefore, it can be extended to \[{\cal O}_{x, I} = \widehat {\oplus V_{\lambda^i}}.\] \hfill \definition \label{_Psi(z)_Definition_} Let $M$, $I$, $J$, $x$, ${\cal O}_{x, I}$ be as in the beginning of this Subsection. Consider the ${\Bbb C}^*$-action \[ \Psi_{I,J}(z):\; {\cal O}_{x, I} {\:\longrightarrow\:} {\cal O}_{x, I} \] constructed as above. Then $\Psi_{I,J}(z)$ is called {\bf the canonical ${\Bbb C}^*$-action associated with $M$, $I$, $J$, $x$.} \hfill In the above notation, consider a reflexive sheaf $F$ on $(M,I)$ equipped with a hyperholomorphic connection. Denote the germ of $F$ at $x$ by $F_x$, $F_x:= F\otimes_{{\cal O}_{(M,I)}} {\cal O}_{x, I}$. From \ref{_Psi_equiv_hyperho_Theorem_}, we obtain an isomorphism $F_x \cong \Psi^*_{I,J} F_x$. This isomorphism can be interpreted as an automorphism \[ \Psi^F_{I,J}:\; F_x{\:\longrightarrow\:} F_x\] satisfying \begin{equation}\label{_Psi^F_and_multi_Equation_} \Psi^F_{I,J}(\alpha v) = \Psi_{I,J}(\alpha) v, \end{equation} for all $\alpha \in {\cal O}_{x, I}$, $v\in F_x$. By \eqref{_Psi^F_and_multi_Equation_}, the automorphism $\Psi^F_{I,J}$ respects the filtration \[ F_x \supset \frak m F_x \supset \frak m^2 F_x \supset ... \] Thus, it makes sense to speak of $\Psi^F_{I,J}$-action on $\frak m^i F_x /\frak m^{i+1} F_x$. \hfill \lemma\label{_Psi^F_on_m^iF/m^i+1F_Lemma_} The automorphism $\Psi^F_{I,J}$ acts on $\frak m^i F_x /\frak m^{i+1} F_x$ as a multiplication by $\lambda^i$, where $\lambda\in {\Bbb C}$ is the number considered in \eqref{_Psi_acts_on_cota_Equation_}. \hfill {\bf Proof:} By \eqref{_Psi^F_and_multi_Equation_}, it suffices to prove \ref{_Psi^F_on_m^iF/m^i+1F_Lemma_} for $i=0$. In other words, we have to show that $\Psi^F_{I,J}$ acts as identity on $F_x /\frak m F_x$. We reduced \ref{_Psi^F_on_m^iF/m^i+1F_Lemma_} to the following claim. \hfill \claim \label{_Psi^F_identi_on_F/mF_Claim_} In the above assumptions, the automorphism $\Psi^F_{I,J}$ acts as identity on $F_x /\frak m F_x$. \hfill {\bf Proof:} In the course of defining the map $\Psi^F_{I,J}$, we identified the space $\operatorname{Lin}(M)$ with a complexification of $(M,I)$, and defined the maps \[ p_I:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, I),\ \ p_J:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, I) \] (these maps are smooth, in a neighbourhood of $\operatorname{Hor}\subset \operatorname{Lin}(M)$, by \ref{_Lin_is_MxM_Proposition_}), and \[ e_I:\; (B, I) {\:\longrightarrow\:} \operatorname{Lin}(M),\ \ e_J:\; (B, J) {\:\longrightarrow\:} \operatorname{Lin}(M) \] (these maps are locally closed embeddings). Consider $(M,J)$ as a subvariety of $\operatorname{Tw}(M)$, $(M, J) = \pi^{-1}(J)$. Let $\c F$ be the lift of $F$ to $\operatorname{Tw}(M)$ (see the proof of \ref{_conne_=>_hyperho_Proposition_} for details). Denote the completion of ${\cal O}_x(M, J)$ by ${\cal O}_{x, J}$. Let $F_J$ denote the ${\cal O}_{x, J}$-module $\left(\c F\restrict{(M,J)} \right)\otimes_{{\cal O}_{(M,J)}} \hat {\cal O}_{x, J}$. Consider the horisontal twistor line $l_x\in \operatorname{Lin}(M)$. Let $\operatorname{Lin}_x(M)$ be the spectre of the completion ${\cal O}_{x,\operatorname{Lin}}$ of the local ring of holomorphic functions on $\operatorname{Lin}(M)$ in $l_x$. The maps $p_I$, $p_J$, $e_I$, $e_J$ can be considered as maps of corresponding formal manifolds: \begin{equation*} \begin{split} p_I:\; \operatorname{Lin}_x(M) & {\:\longrightarrow\:} \operatorname{Spec}({\cal O}_{x, I}),\\ p_J:\; \operatorname{Lin}_x(M) & {\:\longrightarrow\:} \operatorname{Spec}({\cal O}_{x, J}),\\ e_I:\; \operatorname{Spec}({\cal O}_{x, I})& {\:\longrightarrow\:}\operatorname{Lin}_x(M),\\ e_J:\; \operatorname{Spec}({\cal O}_{x, J})& {\:\longrightarrow\:}\operatorname{Lin}_x(M), \end{split} \end{equation*} As in Subsection \ref{_Psi_on_shea_Subsection_}, we consider the ${\cal O}_{x,\operatorname{Lin}}$-modules $F_1:= p_I^* F_x$ and $F_2:= p_J^* F_J$. By \ref{_F_1_=F_2_Lemma_}, there exists a natural isomorphism $\Psi_{1,2}:\; F_1{\:\longrightarrow\:} F_2$. Let $\frak m_{l_x}$ be the maximal ideal of ${\cal O}_{x,\operatorname{Lin}}$. Since the morphism $p_I$ is smooth, the space $F_1/ \frak m_{l_x} F_1$ is naturally isomorphic to $F_x/\frak m F_x$. Similarly, the space $F_2/ \frak m_{l_x} F_2$ is isomorphic to $F_J / \frak m_J F_J$, where $\frak m_J$ is the maximal ideal of ${\cal O}_{x, J}$. We have a chain of isomorphisms \begin{equation} \label{_chain_iso_Psi_F/mF_Equation_} \begin{split} F_x/\frak m F_x &\stackrel{p_I^*}{\:\longrightarrow\:} F_1/ \frak m_{l_x} F_1 \stackrel {\Psi_{1,2}} {\:\longrightarrow\:} F_2/ \frak m_{l_x} F_2\\ &\stackrel{e_J^*}{\:\longrightarrow\:} F_J/ \frak m_{J} F_J \stackrel{p_J^*}{\:\longrightarrow\:} F_2/ \frak m_{l_x} F_2\\ &\stackrel {\Psi_{1,2}^{-1}} {\:\longrightarrow\:} F_1/ \frak m_{l_x} F_1 \stackrel{e_I^*}{\:\longrightarrow\:} F_x/ \frak m F_x. \end{split} \end{equation} By definition, for any $f\in F_x/\frak m F_x$, the value of $\Psi^F_{I,J}(f)$ is given by the composide map of \eqref{_chain_iso_Psi_F/mF_Equation_} applied to $f$. The composition \begin{equation} \label{_restri_F/mF_F_2_and_back_Equation_} F_2/ \frak m_{l_x} F_2 \stackrel{e_J^*}{\:\longrightarrow\:} F_J/ \frak m_{J} F_J \stackrel{p_J^*}{\:\longrightarrow\:} F_2/ \frak m_{l_x} F_2 \end{equation} is identity, because the spaces $F_2/ \frak m_{l_x} F_2$ and $F_J / \frak m_J F_J$ are canonically identified, and this identification can be performed via $e_J^*$ or $p_J^*$. Thus, the map \eqref{_chain_iso_Psi_F/mF_Equation_} is a composition \[ F_x/\frak m F_x \stackrel{p_I^*}{\:\longrightarrow\:} F_1/ \frak m_{l_x} F_1 \stackrel {\Psi_{1,2}} {\:\longrightarrow\:} F_2/ \frak m_{l_x} F_2 \stackrel {\Psi_{1,2}^{-1}} {\:\longrightarrow\:} F_1/ \frak m_{l_x} F_1 \stackrel{e_I^*}{\:\longrightarrow\:} F_x/ \frak m F_x. \] This map is clearly equivalent to a composition \[ F_x/\frak m F_x \stackrel{p_I^*}{\:\longrightarrow\:} F_1/ \frak m_{l_x} F_1 \stackrel{e_I^*}{\:\longrightarrow\:} F_x/ \frak m F_x, \] which is identity according to the same reasoning which proved that \eqref{_restri_F/mF_F_2_and_back_Equation_} is identity. We proved \ref{_Psi^F_identi_on_F/mF_Claim_} and \ref{_Psi^F_on_m^iF/m^i+1F_Lemma_}. \blacksquare \hfill Consider the $\lambda^n$-eigenspaces $F_{\lambda^n}$ of $F_x$. Consider the $\oplus V_{\lambda^n}$-submodule $\oplus F_{\lambda^n}\subset F_x$, where $\oplus V_{\lambda^n}\subset {\cal O}_{x, I}$ is the ring defined in Subsection \ref{_C^*_action_on_loca_ring_Subsection_}. From \ref{_Psi^F_identi_on_F/mF_Claim_} and \eqref{_Psi_acts_on_cota_Equation_} it follows that $\oplus F_{\lambda^n}$ is dense in $F_x$, with respect to the adic topology on $F_x$. For $z\in {\Bbb C}^*$, let $\Psi_{I,J}^F(z):\; \oplus F_{\lambda^n}{\:\longrightarrow\:} \oplus F_{\lambda^n}$ act on $F_{\lambda^n}$ as a multiplication by $z^n$. As in \ref{_Psi(z)_Definition_}, we extend $\Psi_{I,J}^F(z)$ to $F_x = \widehat{\oplus F_{\lambda^n}}$. This automorphism makes $F_x$ into a ${\Bbb C}^*$-equivariant module over ${\cal O}_{x, I}$ \hfill \definition\label{_C^*_stru_on_sge_Definition_} The constructed above ${\Bbb C}^*$-equivariant structure on $F_x$ is called {\bf the canonical ${\Bbb C}^*$-equivariant structure on $F_x$ associated with $J$}. \subsection{Desingularization of ${\Bbb C}^*$-equivariant sheaves} \label{_desingu_she_Subsection_} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf with isolated singularities over $(M,I)$, equipped with a hyperholomorphic connection. We have shown that the sheaf $F$ admits a ${\Bbb C}^*$-equivariant structure compatible with the canonical ${\Bbb C}^*$-action on the local ring of $(M,I)$. Therefore, \ref{_desingu_hyperho_Theorem_} is implied by the following proposition. \hfill \proposition \label{_desingu_C^*_equi_Proposition_} Let $B$ be a complex manifold, $x\in B$ a point. Assume that there is an action $\Psi(z)$ of ${\Bbb C}^*$ on $B$ which fixes $x$ and acts on $T_x B$ be dilatations. Let $F$ be a reflexive coherent sheaf on $B$, which is locally trivial outside of $x$. Assume that the germ $F_x$ of $F$ in $x$ is equipped with a ${\Bbb C}^*$-equivariant structure, compatible with $\Psi(z)$. Let $\tilde B$ be a blow-up of $B$ in $x$, and $\pi:\; \tilde B{\:\longrightarrow\:} B$ the standard projection. Then the pullback sheaf $\tilde F:= \pi^* F$ is locally trivial on $\tilde B$. \hfill {\bf Proof:} Let $C:= \pi^{-1}(x)$ be the singular locus of $\pi$. The sheaf $F$ is locally trivial outside of $x$. Let $d$ be the rank of $F\restrict{B\backslash x}$. To prove that $\tilde F$ is locally trivial, we need to show that for all points $y\in \tilde B$, the fiber $\pi^* F\restrict y$ is $d$-dimensional. Therefore, to prove \ref{_desingu_C^*_equi_Proposition_} it suffices to show that $\pi^* F\restrict C$ is a vector bundle of dimension $d$. The variety $C$ is naturally identified with the projectivization ${\Bbb P} T_x B$ of the tangent space $T_xB$. The total space of $T_x B$ is equipped with a natural action of ${\Bbb C}^*$, acting by dilatations. Clearly, coherent sheaves on ${\Bbb P} T_x B$ are in one-to-one correspondence with ${\Bbb C}^*$-equivariant coherent sheaves on $T_x B$. Consider a local isomorphism $\phi:\; T_x B {\:\longrightarrow\:} B$ which is compatible with ${\Bbb C}^*$-action, maps $0\in T_x B$ to $x$ and acts as identity on the tangent space $T_0 (T_x B)=T_x B$. The sheaf $\phi^* F$ is ${\Bbb C}^*$-equivariant. Clearly, the corresponding sheaf on ${\Bbb P} T_x B$ is canonically isomorphic with $\pi^*F \restrict C$. Let $l\in T_x B$ be a line passing through $0$, and $l\backslash 0$ its complement to $0$. Denote the corresponding point of ${\Bbb P} T_x B$ by $y$. The restriction $\phi^* F\restrict {l\backslash 0}$ is a ${\Bbb C}^*$-equivariant vector bundle. The ${\Bbb C}^*$-equivariant structure identifies all the fibers of the bundle $\phi^* F\restrict {l\backslash 0}$. Let $F_l$ be one of these fibers. Clearly, the fiber of $\pi^*F \restrict C$ in $y$ is canonically isomorphic to $F_l$. Therefore, the fiber of $\pi^*F \restrict C$ in $y$ is $d$-dimensional. We proved that $\pi^*F$ is a bundle. \blacksquare \section{Twistor transform and quaternionic-K\"ahler geometry} \label{_twisto_tra_Section_} This Section is a compilation of results known from the literature. Subsection \ref{_dire_inve_twi_Subsection_} is based on \cite{_NHYM_} and the results of Subsection \ref{_twi_tra_Hermi_Subsection_} are implicit in \cite{_NHYM_}. Subsection \ref{_B_2_bundles_Subsection_} is based on \cite{_Salamon_}, \cite{_Nitta:bundles_} and \cite{_Nitta:Y-M_}, and Subsection \ref{_specia_and_q-K-Subsection_} is a recapitulation of the results of A. Swann (\cite{_Swann_}). \subsection{Direct and inverse twistor transform} \label{_dire_inve_twi_Subsection_} In this Subsection, we recall the definition and the main properties of the direct and inverse twistor transform for bundles over hyperk\"ahler manifolds (\cite{_NHYM_}). \hfill The following definition is a non-Hermitian analogue of the notion of a hyperholomorphic connection. \hfill \definition Let $M$ be a hyperk\"ahler manifold, not necessarily compact, and $(B, \nabla)$ be a vector bundle with a connection over $M$, not necessarily Hermitian. Assume that the curvature of $\nabla$ is contained in the space $\Lambda^2_{inv}(M, \operatorname{End}(B))$ of $SU(2)$-invariant 2-forms with coefficients in $\operatorname{End}(B)$. Then $(B, \nabla)$ is called {\bf an autodual bundle}, and $\nabla$ {\bf an autodual connection}. \hfill Let $\operatorname{Tw}(M)$ be the twistor space of $M$, equipped with the standard maps $\pi:\; \operatorname{Tw}(M) {\:\longrightarrow\:} {\Bbb C} P^1$, $\sigma:\; \operatorname{Tw}(M) {\:\longrightarrow\:} M$. \hfill We introduce the direct and inverse twistor transforms which relate autodual bundles on the hyperk\"ahler manifold $M$ and holomorphic bundles on its twistor space $\operatorname{Tw}(M)$. \hfill Let $B$ be a complex vector bundle on $M$ equipped with a connection $\nabla$. The pullback $\sigma^*B$ of $B$ to $\operatorname{Tw}(M)$ is equipped with a pullback connection $\sigma^*\nabla$. \hfill \lemma \label{_autodua_(1,1)-on-twi_Lemma_} (\cite{_NHYM_}, Lemma 5.1) The connection $\nabla$ is autodual if and only if the connection $\sigma^*\nabla$ has curvature of Hodge type $(1,1)$. {\bf Proof:} Follows from \ref{_SU(2)_inva_type_p,p_Lemma_}. \blacksquare \hfill In assumptions of \ref{_autodua_(1,1)-on-twi_Lemma_}, consider the $(0,1)$-part $(\sigma^*\nabla)^{0,1}$ of the connection $\sigma^*\nabla$. Since $\sigma^*\nabla$ has curvature of Hodge type $(1,1)$, we have \[ \left((\sigma^*\nabla)^{0,1}\right)^2=0, \] and by \ref{_Newle_Nie_for_NH_bu_Proposition_}, this connection is integrable. Consider $(\sigma^*\nabla)^{0,1}$ as a holomorphic structure operator on $\sigma^* B$. \hfill Let $\c A$ be the category of autodual bundles on $M$, and $\c C$ the category of holomorphic vector bundles on $\operatorname{Tw}(M)$. We have constructed a functor \[ (\sigma^* \bullet)^{0,1}:\; \c A {\:\longrightarrow\:} \c C, \] $\nabla {\:\longrightarrow\:} (\sigma^*\nabla)^{0,1}$. Let $s\in \operatorname{Hor}\subset \operatorname{Tw}(M)$ be a horisontal twistor line (Subsection \ref{_twi_lines_C^*_Subsection_}). For any $(B, \nabla)\in \c A$, consider corresponding holomorphic vector bundle $(\sigma^* B, (\sigma^*\nabla)^{0,1})$. The restriction of $(\sigma^* B, (\sigma^*\nabla)^{0,1})$ to $s\cong {\Bbb C} P^1$ is a trivial vector bundle. A converse statement is also true. Denote by $\c C_0$ the category of holomorphic vector bundles $C$ on $\operatorname{Tw}(M)$, such that the restriction of $C$ to any horisontal twistor line is trivial. \hfill \theorem \label{_dire_inve_twisto_Theorem_} Consider the functor \[ (\sigma^* \bullet)^{0,1}:\; \c A {\:\longrightarrow\:} \c C_0\] constructed above. Then it is an equivalence of categories. {\bf Proof:} \cite{_NHYM_}, Theorem 5.12. \blacksquare \hfill \definition Let $M$ be a hyperk\"ahler manifold, $\operatorname{Tw}(M)$ its twistor space and $\c F$ a holomorphic vector bundle. We say that $\c F$ is {\bf compatible with twistor transform} if the restriction of $C$ to any horisontal twistor line $s\in \operatorname{Tw}(M)$ is a trivial bundle on $s\cong {\Bbb C} P^1$. \hfill Recall that a connection $\nabla$ in a vector bundle over a complex manifold is called $(1,1)$-connection if its curvature is of Hodge type $(1,1)$. \hfill \remark \label{_cano_conne_Remark_} Let $\c F$ be a holomorphic bundle over $\operatorname{Tw}(M)$ which is compatible with twistor transform. Then $\c F$ is equipped with a natural $(1,1)$-connection $\nabla_{\c F}= \sigma^* \nabla$, where $(B, \nabla)$ is the corresponding autodual bundle over $M$. The connection $\nabla_{\c F}$ is not, generally speaking, Hermitian, or compatible with a Hermitian structure. \subsection{Twistor transform and Hermitian structures on vector bundles} \label{_twi_tra_Hermi_Subsection_} Results of this Subsection were implicit in \cite{_NHYM_}, but in this presentation, they are new. Let $M$ be a hyperk\"ahler manifold, not necessarily compact, and $\operatorname{Tw}(M)$ its twistor space. In Subsection \ref{_dire_inve_twi_Subsection_}, we have shown that certain holomorphic vector bundles over $\operatorname{Tw}(M)$ admit a canonical (1,1)-connection $\nabla{_\c F}$ (\ref{_cano_conne_Remark_}). This connection can be non-Hermitian. Here we study the Hermitian structures on $(\c F, \nabla{_\c F})$ in terms of holomorphic properties of $\c F$. \hfill \definition Let $F$ be a real analytic complex vector bundle over a real analytic manifold $X_{\Bbb R}$, and $h:\; F\times F {\:\longrightarrow\:} {\Bbb C}$ a ${\cal O}_{X_{\Bbb R}}$-linear pairing on $F$. Then $h$ is called {\bf semilinear} if for all $\alpha \in {\cal O}_{X_{\Bbb R}}\otimes_{\Bbb R} {\Bbb C}$, we have \[ h(\alpha x, y) = \alpha\cdot h(x, y), \text{\ \ and\ \ } h(x,\alpha y) = \bar \alpha\cdot h(x, y). \] For $X$ a complex manifold and $F$ a holomorphic vector bundle, by a semilinear pairing on $F$ we understand a semilinear pairing on the underlying real analytic bundle. Clearly, a real analytic Hermitian metric is always semilinear. \hfill Let $X$ be a complex manifold, $I:\; TX {\:\longrightarrow\:} TX$ the complex structure operator, and $i:\; X {\:\longrightarrow\:} X$ a real analytic map. We say that $X$ is anticomplex if the induced morphism of tangent spaces satisfies $i\circ I = - I \circ I$. For a complex vector bundle $F$ on $X$, consider the complex adjoint vector bundle $\bar F$, which coincides with $F$ as a real vector bundle, with ${\Bbb C}$-action which is conjugate to that defined on $F$. Clearly, for every holomorphic vector bundle $F$, and any anticomplex map $i:\; X {\:\longrightarrow\:} X$, the bundle $i^* \bar F$ is equipped with a natural holomorphic structure. \hfill Let $M$ be a hyperk\"ahler manifold, and $\operatorname{Tw}(M)$ its twistor space. Recall that $\operatorname{Tw}(M)={\Bbb C} P^1\times M$ is equipped with a canonical anticomplex involution $\iota$, which acts as identity on $M$ and as central symmetry $I{\:\longrightarrow\:} -I$ on ${\Bbb C} P^1= S^2$. For any holomorphic bundle $\c F$ on $\operatorname{Tw}(M)$, consider the corresponding holomorphic bundle $\iota^* \bar{\c F}$. \hfill Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure, $F$ a vector bundle over $M$, equipped with an autodual connection $\nabla$, and $\c F$ the corresponding holomorphic vector bundle over $\operatorname{Tw}(M)$, equipped with a canonical connection $\nabla_{\c F}$. As usually, we identify $(M, I)$ and the fiber $\pi^{-1}(I)$ of the twistor projection $\pi:\; \operatorname{Tw}(M) {\:\longrightarrow\:} {\Bbb C} P^1$. Let $\nabla_{\c F} = \nabla^{1,0}_I + \nabla^{0,1}_I$ be the Hodge decomposition of $\nabla$ with respect to $I$. \begin{equation} \label{_nabla^1,0_as_holo_Equation_} \begin{minipage}[m]{0.8\linewidth} Clearly, the operator $\nabla^{1,0}_I$ can be considered as a holomorphic structure operator on $F$, considered as a complex vector bundle over $(M, -I)$. \end{minipage} \end{equation} Then the holomorphic structure operator on $\c F\restrict{(M,I)}$ is equal to $\nabla^{0,1}_I$, and the holomorphic structure operator on $\c F\restrict{(M,-I)}$ is equal to $\nabla^{1,0}_I$. Assume that the bundle $(\c F, \nabla_{\c F})$ is equipped with a non-degenerate semilinear pairing $h$ which is compatible with the connection. Consider the natural connection $\nabla_{\c F^*}$ on the dual bundle to $\c F$, and its Hodge decomposition (with respect to $I$) \[ \nabla_{\c F^*} = \nabla^{1,0}_{\c F^*} + \nabla^{0,1}_{\c F^*}. \] Clearly, the pairing $h$ gives a $C^\infty$-isomorphism of $\c F$ and the complex conjugate of its dual bundle, denoted as $\bar{\c F}^*$. Since $h$ is semilinear and compatible with the connection, it maps the holomorphic structure operator $\nabla^{0,1}_I$ to the complex conjugate of $\nabla^{1,0}_{\c F^*}$. On the other hand, the operator $\nabla^{1,0}_{\c F^*}$ is a holomorphic structure operator in $\c F^*\restrict{(M,-I)}$, as \eqref{_nabla^1,0_as_holo_Equation_} claims. We obtain that the map $h$ can be considered as an isomorphism of holomorphic vector bundles \[ h:\; \c F {\:\longrightarrow\:} (\iota^* \bar{\c F})^*.\] This correspondence should be thought of as a (direct) twistor transform for bundles with a semilinear pairing. \hfill \proposition\label{_twi_tra_for_semili_Proposition_} (direct and inverse twistor transform for bundles with semilinear pairing) Let $M$ be a hyperk\"ahler manifold, and $\c C_{sl}$ the category of autodual bundles over $M$ equipped with a non-degenerate semilinear pairing. Consider the category $\c C_{hol,sl}$ of holomorphic vector bundles $\c F$ on $\operatorname{Tw}(M)$, compatible with twistor transform and equipped with an isomorphism \[ {\frak h}:\; \c F {\:\longrightarrow\:} (\iota^* \bar{\c F})^*.\] Let $\c T:\; \c C_{sl}{\:\longrightarrow\:} \c C_{hol,sl}$ be the functor constructed above. Then $\c T$ is an isomorphism of categories. \hfill {\bf Proof:} Given a pair $\c F, {\frak h}:\; \c F {\:\longrightarrow\:} (\iota^* \bar{\c F})^*$, we need to construct a non-degenerate semilinear pairing $h$ on $\c F\restrict{(M,I)}$, compatible with a connection. Since $\c F$ is compatible with twistor transform, it is a pullback of a bundle $(F, \nabla)$ on $M$. This identifies the real analytic bundles $\c F\restrict{(M,I')}$, for all induced complex structures $I'$. Taking $I' = \pm I$, we obtain an identification of the $C^{\infty}$-bundles $\c F\restrict{(M,I)}$, $\c F\restrict{(M,-I)}$. Thus, $\frak h$ can be considered as an isomorphism of $F= \c F\restrict{(M,I)}$ and $(\bar {\c F})^*\restrict{(M,I)}$. This allows one to consider $\frak h$ as a semilinear form $h$ on $F$. We need only to show that $h$ is compatible with the connection $\nabla$. Since $\nabla_{\c F}$ is an invariant of holomorphic structure, the map ${\frak h}:\; \c F {\:\longrightarrow\:} (\iota^* \bar{\c F})^*$ is compatible with the connection $\nabla_{\c F}$. Thus, the obtained above form $h$ is compatible with the connection $\nabla_{\c F}\restrict{(M,I)}=\nabla$. This proves \ref{_twi_tra_for_semili_Proposition_}. \blacksquare \subsection{ $B_2$-bundles on quaternionic-K\"ahler manifolds} \label{_B_2_bundles_Subsection_} \definition \label{_q-K_Definition_} (\cite{_Salamon_}, \cite{_Besse:Einst_Manifo_}) Let $M$ be a Riemannian manifold. Consider a bundle of algebras $\operatorname{End}(TM)$, where $TM$ is the tangent bundle to $M$. Assume that $\operatorname{End}(TM)$ contains a 4-dimensional bundle of subalgebras $W\subset \operatorname{End}(TM)$, with fibers isomorphic to a quaternion algebra ${\Bbb H}$. Assume, moreover, that $W$ is closed under the transposition map $\bot:\; \operatorname{End}(TM){\:\longrightarrow\:} \operatorname{End}(TM)$ and is preserved by the Levi-Civita connection. Then $M$ is called {\bf quaternionic-K\"ahler}. \hfill \example Consider the quaternionic projective space \[ {\Bbb H} P^n= ({\Bbb H}^n \backslash 0) / {\Bbb H}^*. \] It is easy to see that ${\Bbb H} P^n$ is a quaternionic-K\"ahler manifold. For more examples of quaternionic-K\"ahler manifolds, see \cite{_Besse:Einst_Manifo_}. \hfill A quaternionic-K\"ahler manifold is Einstein (\cite{_Besse:Einst_Manifo_}), i. e. its Ricci tensor is proportional to the metric: $Ric(M) = c \cdot g$, with $c\in {\Bbb R}$. When $c=0$, the manifold $M$ is hyperk\"ahler, and its restricted holonomy group is $Sp(n)$; otherwise, the restricted holonomy is $Sp(n)\cdot Sp(1)$. The number $c$ is called {\bf the scalar curvature} of $M$. Further on, we shall use the term {\it quaternionic-K\"ahler manifold} for manifolds with non-zero scalar curvature. The quaternionic projective space ${\Bbb H} P^n$ has positive scalar curvature. \hfill The quaternionic projective space is the only example of quaternionic-K\"ahler manifold which we need, in the course of this paper. However, the formalism of quaternionic-K\"ahler manifolds is very beautiful and significantly simplifies the arguments, so we state the definitions and results for a general quaternionic-K\"ahler manifold whenever possible. \hfill Let $M$ be a quaternionic-K\"ahler manifold, and $W\subset \operatorname{End}(TM)$ the corresponding 4-dimensional bundle. For $x\in M$, consider the set $\c R_x\subset W\restrict x$, consisting of all $I\in W\restrict x$ satisfying $I^2=-1$. Consider $\c R_x$ as a Riemannian submanifold of the total space of $W\restrict x$. Clearly, $\c R_x$ is isomorphic to a 2-dimensional sphere. Let $\c R= \cup_x \c R_x$ be the corresponding spherical fibration over $M$, and $\operatorname{Tw}(M)$ its total space. The manifold $\operatorname{Tw}(M)$ is equipped with an almost complex structure, which is defined in the same way as the almost complex structure for the twistor space of a hyperk\"ahler manifold. This almost complex structure is known to be integrable (see \cite{_Salamon_}). \hfill \definition\label{_twi_q-K_Definition_} (\cite{_Salamon_}, \cite{_Besse:Einst_Manifo_}) Let $M$ be a quaternionic-K\"ahler manifold. Consider the complex manifold $\operatorname{Tw}(M)$ constructed above. Then $\operatorname{Tw}(M)$ is called {\bf the twistor space of $M$.} \hfill Note that (unlike in the hyperk\"ahler case) the space $\operatorname{Tw}(M)$ is K\"ahler. For quaternionic-K\"ahler manifolds with positive scalar curvature, the anticanonical bundle of $\operatorname{Tw}(M)$ is ample, so $\operatorname{Tw}(M)$ is a Fano manifold. \hfill Quaternionic-K\"ahler analogue of a twistor transform was studied by T. Nitta in a serie of papers (\cite{_Nitta:bundles_}, \cite{_Nitta:Y-M_} etc.) It turns out that the picture given in \cite{_NHYM_} for K\"ahler manifolds is very similar to that observed by T. Nitta. \hfill A role of $SU(2)$-invariant 2-forms is played by so-called $B_2$-forms. \hfill \definition Let $SO(TM)\subset \operatorname{End}(TM)$ be a group bundle of all orthogonal automorphisms of $TM$, and $G_M:= W\cap SO(TM)$. Clearly, the fibers of $G_M$ are isomorphic to $SU(2)$. Consider the action of $G_M$ on the bundle of 2-forms $\Lambda^2(M)$. Let $\Lambda^2_{inv}(M)\subset \Lambda^2(M)$ be the bundle of $G_M$-invariants. The bundle $\Lambda^2_{inv}(M)$ is called {\bf the bundle of $B_2$-forms}. In a similar fashion we define $B_2$-forms with coefficients in a bundle. \hfill \definition\label{_B_2_bu_Definition_} In the above assumptions, let $(B, \nabla)$ be a bundle with connection over $M$. The bundle $B$ is called {\bf a $B_2$-bundle}, and $\nabla$ is called {\bf a $B_2$-connection}, if its curvature is a $B_2$-form. \hfill Consider the natural projection $\sigma:\; \operatorname{Tw}(M){\:\longrightarrow\:} M$. The proof of the following claim is completely analogous to the proof of \ref{_SU(2)_inva_type_p,p_Lemma_} and \ref{_autodua_(1,1)-on-twi_Lemma_}. \hfill \claim\label{_B_2_=_holo_on_Tw_Claim_} \begin{description} \item[(i)] Let $\omega$ be a $2$-form on $M$. The pullback $\sigma^* \omega$ is of type $(1,1)$ on $\operatorname{Tw}(M)$ if and only if $\omega$ is a $B_2$-form on $M$. \item[(ii)] Let $B$ be a complex vector bundle on $M$ equipped with a connection $\nabla$, not necessarily Hermitian. The pullback $\sigma^*B$ of $B$ to $\operatorname{Tw}(M)$ is equipped with a pullback connection $\sigma^*\nabla$. Then, $\nabla$ is a $B_2$-connection if and only if $\sigma^*\nabla$ has curvature of Hodge type $(1,1)$. \end{description} \blacksquare There exists an analogue of direct and inverse twistor transform as well. \hfill \theorem \label{_dire_inve_q-K_Theorem_} For any $B_2$-connection $(B, \nabla)$, consider the corresponding holomorphic vector bundle \[ (\sigma^* B, (\sigma^*\nabla)^{0,1}). \] The restriction of $(\sigma^* B, (\sigma^*\nabla)^{0,1})$ to a line $\sigma^{-1}(m)\cong {\Bbb C} P^1$ is a trivial vector bundle, for any point $m\in M$. Denote by $\c C_0$ the category of holomorphic vector bundles $C$ on $\operatorname{Tw}(M)$, such that the restriction of $C$ to $\sigma^{-1}(m)$ is trivial, for all $m\in M$, and by $\c A$ the category of $B_2$-bundles (not necessarily Hermitian). Consider the functor \[ (\sigma^* \bullet)^{0,1}:\; \c A {\:\longrightarrow\:} \c C_0\] constructed above. Then it is an equivalence of categories. {\bf Proof:} It is easy to modify the proof of the direct and inverse twistor transform theorem from \cite{_NHYM_} to work in quaternionic-K\"ahler situation. \blacksquare \hfill We will not use \ref{_dire_inve_q-K_Theorem_}, except for its consequence, which was proven in \cite{_Nitta:bundles_}. \hfill \corollary \label{_twi_tra_q-K_Hermi_Corollary_} Consider the functor \[ (\sigma^* \bullet)^{0,1}:\; \c A {\:\longrightarrow\:} \c C_0\] constructed in \ref{_dire_inve_q-K_Theorem_}. Then $(\sigma^* \bullet)^{0,1}$ gives an injection $\kappa$ from the set of equivalence classes of Hermitian $B_2$-connections over $M$ to the set of equivalence classes of holomorphic connections over $\operatorname{Tw}(M)$. \blacksquare \hfill Let $M$ be a quaternionic-K\"ahler manifold. The space $\operatorname{Tw}(M)$ has a natural K\"ahler metric $g$, such that the standard map $\sigma:\; \operatorname{Tw}(M){\:\longrightarrow\:} M$ is a Riemannian submersion, and the restriction of $g$ to the fibers $\sigma^{-1}(m)$ of $\sigma$ is a metric of constant curvature on $\sigma^{-1}(m)= {\Bbb C} P^1$ (\cite{_Salamon_}, \cite{_Besse:Einst_Manifo_}). \hfill \example\label{_Tw_HP^n_Fu-St_Example_} In the case $M= {\Bbb H}P^n$, we have $\operatorname{Tw}(M) = {\Bbb C} P^{2n+1}$, and the K\"ahler metric $g$ is proportional to the Fubini-Study metric on ${\Bbb C} P^{2n+1}$. \hfill \theorem \label{_twi_tra_YM_q-K_Theorem_} (T. Nitta) Let $M$ be a quaternionic-K\"ahler manifold of positive scalar curvature, $\operatorname{Tw}(M)$ its twistor space, equipped with a natural K\"ahler structure, and $B$ a Hermitian $B_2$-bundle on $M$. Consider the pullback $\sigma^* B$, equipped with a Hermitian connection. Then $\sigma^* B$ is a Yang-Mills bundle on $\operatorname{Tw}(M)$, and $\deg c_1(\sigma^* B) =0$. {\bf Proof:} \cite{_Nitta:Y-M_}. \blacksquare \hfill Let $\kappa$ be the map considered in \ref{_twi_tra_q-K_Hermi_Corollary_}. Assume that $M$ is a compact manifold. In \cite{_Nitta:Y-M_}, T. Nitta defined the moduli space of Hermitian $B_2$-bundles. By Uhlenbeck-Yau theorem, Yang-Mills bundles are polystable. Then the map $\kappa$ provides an embedding from the moduli of non-decomposable Hermitian $B_2$-bundles to the moduli $\c M$ of stable bundles on $\operatorname{Tw}(M)$. The image of $\kappa$ is a totally real subvariety in $\c M$ (\cite{_Nitta:Y-M_}). \subsection{Hyperk\"ahler manifolds with special ${\Bbb H}^*$-action and qua\-ter\-ni\-o\-nic-\-K\"ah\-ler manifolds of positive scalar curvature} \label{_specia_and_q-K-Subsection_} Further on, we shall need the following definition. \hfill \definition {\bf An almost hypercomplex manifold} is a smooth manifold $M$ with an action of quaternion algebra in its tangent bundle For each $L\in \Bbb H$, $L^2 = -1$, $L$ gives an almost complex structure on $M$. The manifold $M$ is caled {\bf hypercomplex} if the almost complex structure $L$ is integrable, for all possible choices $L\in \Bbb H$. \hfill The twistor space for a hypercomplex manifold is defined in the same way as for hyperk\"ahler manifolds. It is also a complex manifold (\cite{_Kaledin_}). The formalism of direct and inverse twistor transform can be repeated for hypercomplex manifolds verbatim. \hfill Let ${\Bbb H}^*$ be the group of non-zero quaternions. Consider an embedding $SU(2)\hookrightarrow {\Bbb H^*}$. Clearly, every quaternion $h\in {\Bbb H}^*$ can be uniquely represented as $h= |h| \cdot g_h$, where $g_h\in SU(2)\subset {\Bbb H}^*$. This gives a natural decomposition ${\Bbb H}^*= SU(2)\times {\Bbb R}^{>0}$. Recall that $SU(2)$ acts naturally on the set of induced complex structures on a hyperk\"ahler manifold. \hfill \definition\label{_H^*_specia_Definition_} Let $M$ be a hyperk\"ahler manifold equipped with a free smooth action $\rho$ of the group ${\Bbb H}^*= SU(2)\times {\Bbb R}^{>0}$. The action $\rho$ is called {\bf special} if the following conditions hold. \begin{description} \item[(i)] The subgroup $SU(2)\subset {\Bbb H}^*$ acts on $M$ by isometries. \item[(ii)] For $\lambda\in {\Bbb R}^{>0}$, the corresponding action $\rho(\lambda):\; M{\:\longrightarrow\:} M$ is compatible with the hyperholomorphic structure (which is a fancy way of saying that $\rho(\lambda)$ is holomorphic with respect to any of induced complex structures). \item[(iii)] Consider the smooth ${\Bbb H^*}$-action $\rho_e:\; {\Bbb H^*}\times \operatorname{End}(TM) {\:\longrightarrow\:} \operatorname{End}(TM)$ induced on $\operatorname{End}(TM)$ by $\rho$. For any $x\in M$ and any induced complex structure $I$, the restriction $I\restrict x$ can be considered as a point in the total space of $\operatorname{End}(TM)$. Then, for all induced complex structures $I$, all $g\in SU(2)\subset {\Bbb H^*}$, and all $x\in M$, the map $\rho_e(g)$ maps $I\restrict{x}$ to $g(I)\restrict{\rho_e(g)(x)}$. Speaking informally, this can be stated as ``${\Bbb H}^*$-action interchanges the induced complex structures''. \item[(iv)] Consider the automorphism of $S^2 T^*M$ induced by $\rho(\lambda)$, where $\lambda\in {\Bbb R}^{>0}$. Then $\rho(\lambda)$ maps the Riemannian metric tensor $s\in S^2 T^*M$ to $\lambda^2 s$. \end{description} \hfill \example Consider the flat hyperk\"ahler manifold $M_{\rm fl}= {\Bbb H}^n \backslash 0$. There is a natural action of ${\Bbb H^*}$ on ${\Bbb H}^n\backslash 0$. This gives a special action of ${\Bbb H^*}$ on $M_{\rm fl}$. \hfill The case of a flat manifold $M_{\rm fl}= {\Bbb H}^n \backslash 0$ is the only case where we apply the results of this section. However, the general statements are just as difficult to prove, and much easier to comprehend. \hfill \definition\label{_speci_equi_Definition_} Let $M$ be a hyperk\"ahler manifold with a special action $\rho$ of ${\Bbb H^*}$. Assume that $\rho(-1)$ acts non-trivially on $M$. Then $M/\rho(\pm 1)$ is also a hyperk\"ahler manifold with a special action of ${\Bbb H^*}$. We say that the manifolds $(M, \rho)$ and $(M/\rho(\pm 1), \rho)$ are {\bf hyperk\"ahler manifolds with special action of ${\Bbb H^*}$ which are special equivalent}. Denote by $H_{sp}$ the category of hyperk\"ahler manifolds with a special action of ${\Bbb H^*}$ defined up to special equivalence. \hfill A. Swann (\cite{_Swann_}) developed an equivalence between the category of qua\-ter\-ni\-o\-nic-\-K\"ah\-ler manifolds of positive scalar curvature and the category $H_{sp}$. The purpose of this Subsection is to give an exposition of Swann's formalism. \hfill Let $Q$ be a quaternionic-K\"ahler manifold. The restricted holonomy group of $Q$ is $Sp(n)\cdot Sp(1)$, that is, $(Sp(n)\times Sp(1))/\{\pm 1\}$. Consider the principal bundle $\c G$ with the fiber $Sp(1)/\{\pm 1\}$, corresponding to the subgroup \[ Sp(1)/\{\pm 1\}\subset (Sp(n)\times Sp(1))/\{\pm 1\}.\] of the holonomy. There is a natural $Sp(1)/\{\pm 1\}$-action on the space \\ ${\Bbb H^*}/\{\pm 1\}$. Let \[ \c U(Q):= \c G\times_{Sp(1)/\{\pm 1\}}{\Bbb H^*}/\{\pm 1\}.\] Clearly, $\c U(Q)$ is fibered over $Q$, with fibers which are isomorphic to \\ ${\Bbb H^*}/\{\pm 1\}$. We are going to show that the manifold $\c U(Q)$ is equipped with a natural hypercomplex structure. \hfill There is a natural smooth decomposition $\c U(Q)\cong \c G \times {\Bbb R}^{>0}$ which comes from the isomorphism ${\Bbb H^*}\cong Sp(1)\times {\Bbb R}^{>0}$. \hfill Consider the standard 4-dimensional bundle $W$ on $Q$. Let $x\in Q$ be a point. The fiber $W\restrict q$ is isomorphic to $\Bbb H$, in a non-canonical way. The choices of isomorphism $W\restrict q\cong \Bbb H$ are called {\bf quaternion frames in $q$}. The set of quaternion frames gives a fibration over $Q$, with a fiber $\operatorname{Aut}({\Bbb H})\cong Sp(1)/\{\pm 1\}$. Clearly, this fibration coincides with the principal bundle $\c G$ constructed above. Since $\c U(Q)\cong \c G \times {\Bbb R}^{>0}$, a choice of $u\in \c U(Q)\restrict q$ determines an isomorphism $W\restrict q\cong \Bbb H$. \hfill Let $(q, u)$ be the point of $\c U(Q)$, with $q\in Q$, $u\in \c U(Q)\restrict q$. The natural connection in $\c U(Q)$ gives a decomposition \[ T_{(q, u)}U(Q) = T_u \bigg(\c U(Q)\restrict q\bigg) \oplus T_q Q. \] The space $\c U(Q)\restrict q \cong {\Bbb H^*}/\{\pm 1\}$ is equipped with a natural hypercomplex structure. This gives a quaternion action on $T_u \bigg(\c U(Q)\restrict q\bigg)$ The choice of $u\in \c U(Q)\restrict q$ determines a quaternion action on $T_q Q$, as we have seen above. We obtain that the total space of $\c U(Q)$ is an almost hypercomplex manifold. \hfill \proposition\label{_U(Q)_hypercomple_Proposition_} (A. Swann) Let $Q$ be a quaternionic-K\"ahler manifold. Consider the manifold $\c U(Q)$ constructed as above, and equipped with a quaternion algebra action in its tangent space. Then $\c U(Q)$ is a hypercomplex manifold. \hfill {\bf Proof:} Clearly, the manifold $\c U(Q)$ is equipped with a ${\Bbb H}^*$-action, which is related with the almost hypercomplex structure as prescribed by \ref{_H^*_specia_Definition_} (ii)-(iii). Pick an induced complex structure $I\in {\Bbb H}$. This gives an algebra embedding ${\Bbb C} {\:\longrightarrow\:} {\Bbb H}$. Consider the corresponding ${\Bbb C}^*$-action $\rho_I$ on an almost complex manifold $(\c U(Q), I)$. This ${\Bbb C}^*$-action is compatible with the almost complex structure. The quotient $\c U(Q)/\rho(I)$ is an almost complex manifold, which is naturally isomorphic to the twistor space $\operatorname{Tw}(Q)$. Let $L^*$ be a complex vector bundle of all $(1,0)$-vectors $v\in T (\operatorname{Tw}(Q))$ tangent to the fibers of the standard projection $\sigma:\; \operatorname{Tw}(Q){\:\longrightarrow\:} Q$, and $L$ be the dual vector bundle. Denote by $Tot_{\neq 0}(L)$ the complement $\operatorname{Tot}(L)\backslash N$, where $N=\operatorname{Tw}(Q)\subset \operatorname{Tot}(L)$ is the zero section of $L$. Using the natural connection in $L$, we obtain an almost complex structure on $\operatorname{Tot}(L)$. Consider the natural projection $\phi:\; Tot_{\neq 0}(L){\:\longrightarrow\:} Q$. The fibers $\phi^{-1}(q)$ of $\phi$ are identified with the space of non-zero vectors in the total space of the cotangent bundle $T^* \sigma^{-1}(q)\cong T^*({\Bbb C} P^1)$. This space is naturally isomorphic to \[ \c G\restrict q \times {\Bbb R}^{>0}= \c U(Q)\restrict q\cong {\Bbb H}^*/\{\pm 1\}. \] This gives a canonical isomorphism of almost complex manifolds \[ (\c U(Q), I){\:\longrightarrow\:} Tot_{\neq 0}(L).\] Therefore, to prove that $(\c U(Q), I)$ is a complex manifold, it suffices to show that the natural almost complex structure on $Tot_{\neq 0}(L)\subset \operatorname{Tot}(L)$ is integrable. Consider the natural connection $\nabla_L$ on $L$. To prove that $\operatorname{Tot}(L)$ is a complex manifold, it suffices to show that $\nabla_L$ is a holomorphic connection. The bunlde $L$ is known under the name of {\bf holomorphic contact bundle}, and it is known to be holomorphic (\cite{_Salamon_}, \cite{_Besse:Einst_Manifo_}). \blacksquare \hfill \remark The result of \ref{_U(Q)_hypercomple_Proposition_} is well known. We have given its proof because we shall need the natural identification $Tot_{\neq 0}(L)\cong \c U(Q)$ further on in this paper. \hfill \theorem\label{_U(Q)_hyperk_Theorem_} Let $Q$ be a quaternionic-K\"ahler manifold of positive scalar curvature, and $\c U(Q)$ the hypercomplex manifold constructed above. Then $\c U(Q)$ admits a unique (up to a scaling) hyperk\"ahler metric compatible with the hypercomplex structure. {\bf Proof:} \cite{_Swann_}. \blacksquare \hfill Consider the action of ${\Bbb H}^*$ on $\c U(M)$ defined in the proof of \ref{_U(Q)_hypercomple_Proposition_}. This action satisfies the conditions (ii) and (iii) of \ref{_H^*_specia_Definition_}. The conditions (i) and (iv) of \ref{_H^*_specia_Definition_} are easy to check (see \cite{_Swann_} for details). This gives a functor from the category $\c C$ of quaternionic-K\"ahler manifolds of positive scalar curvature to the category $H_{sp}$ of \ref{_speci_equi_Definition_}. \hfill \theorem \label{_U(Q)_equiva_cate_Theorem_} The functor $Q{\:\longrightarrow\:} \c U(Q)$ from $\c C$ to $H_{sp}$ is an equivalence of categories. {\bf Proof:} \cite{_Swann_}. \blacksquare \hfill The inverse functor from $H_{sp}$ to $C$ is constructed by taking a quotient of $M$ by the action of ${\Bbb H}^*$. Using the technique of quaternionic-K\"ahler reduction anf hyperk\"ahler potentials (\cite{_Swann_}), one can equip the quotient $M/{\Bbb H}^*$ with a natural quaternionic-K\"ahler structure. \section{${\Bbb C}^*$-equivariant twistor spaces} \label{_C_equiv_twi_spa_Section_} \hfill In Section \ref{_twisto_tra_Section_}, we gave an exposition of the twistor transform, $B_2$-bundles and Swann's formalism. In the present Section, we give a synthesis of these theories, obtaining a construction with should be thought of as Swann's formalism for vector bundles. Consider the equivalence of categories $Q{\:\longrightarrow\:} \c U(Q)$ constructed in \ref{_U(Q)_equiva_cate_Theorem_} (we call this equivalence ``Swann's formalism''). We show that $B_2$-bundles on $Q$ are in functorial bijective correspondence with ${\Bbb C}^*$-equivariant holomorphic bundles on $\operatorname{Tw}(\c U(Q))$ (\ref{_B_2_to_C^*_equiva_Theorem_}). In Subsection \ref{_hyperho_shea_C^*_equiv_Y-M_on_blow-up_Subsection_}, this equivalence is applied to the vector bundle $\pi^*(F)$ of \ref{_desingu_hyperho_Theorem_}. We use it to construct a canonical Yang-Mills connection on $\pi^*(F)\restrict C$, where $C$ is a special fiber of $\pi:\; \tilde M{\:\longrightarrow\:} (M, I)$ (see \ref{_desingu_hyperho_Theorem_} for details and notation). This implies that the holomorphic bundle $\pi^*(F)\restrict C$ is polystable (\ref{_hyperho_blow-up_stable_Theorem_}). \subsection[$B_2$-bundles on quaternionic-K\"ahler manifolds and ${\Bbb C}^*$-equi\-va\-ri\-ant holomorphic bundles over twistor spaces]{$B_2$-bundles on quaternionic-K\"ahler manifolds and \\${\Bbb C}^*$-equi\-va\-ri\-ant holomorphic bundles over twistor spaces} \label{_B_2_to_C^*-invaholo_ove_twi_Subsection_} For the duration of this Subsection, we fix a hyperk\"ahler manifold $M$, equipped with a special ${\Bbb H}^*$-action $\rho$, and the corresponding quaternionic-K\"ahler manifold $Q= M/{\Bbb H}^*$. Denote the natural quotient map by $\phi:\; M{\:\longrightarrow\:} Q$. \hfill \lemma \label{_phi^*_B_2-forms_1,1_Lemma_} Let $\omega$ be a 2-form over $Q$, and $\phi^* \omega$ its pullback to $M$. Then the following conditions are equivalent \begin{description} \item[(i)] $\omega$ is a $B_2$-form \item[(ii)] $\phi^* \omega$ is of Hodge type $(1,1)$ with respect to some induced complex structure $I$ on $M$ \item[(iii)] $\phi^* \omega$ is $SU(2)$-invariant. \end{description} {\bf Proof:} Let $I$ be an induced complex structure on $M$. As we have shown in the proof of \ref{_U(Q)_hypercomple_Proposition_}, the complex manifold $(M, I)$ is idenified with an open subset of the total space $\operatorname{Tot}(L)$ of a holomorphic line bundle $L$ over $\operatorname{Tw}(Q)$. The map $\phi$ is represented as a composition of the projections $h:\; \operatorname{Tot}(L){\:\longrightarrow\:} \operatorname{Tw}(Q)$ and $\sigma_Q:\; \operatorname{Tw}(Q) {\:\longrightarrow\:} Q$. Since the map $h$ is smooth and holomorphic, the form $\phi^* \omega$ is of Hodge type $(1,1)$ if and only if $\sigma_Q^*\omega$ is of type $(1,1)$. By \ref{_B_2_=_holo_on_Tw_Claim_} (i), this happens if and only if $\omega$ is a $B_2$-form. This proves an equivalence (i) $\Leftrightarrow$ (ii). Since the choice of $I$ is arbitrary, the pullback $\phi^* \omega$ of a $B_2$-form is of Hodge type $(1,1)$ with respect to all induced complex structures. By \ref{_SU(2)_inva_type_p,p_Lemma_}, this proves the implication (i) $\Rightarrow$ (iii). The implication (iii) $\Rightarrow$ (ii) is clear. \blacksquare \hfill \proposition Let $(B, \nabla)$ be a complex vector bundle with connection over $Q$, and $(\phi^*B, \phi^*\nabla)$ its pullback to $M$. Then the following conditions are equivalent \begin{description} \item[(i)] $(B, \nabla)$ is a $B_2$-form \item[(ii)] The curvature of $(\phi^*B, \phi^*\nabla)$ is of Hodge type (1,1) with respect to some induced complex structure $I$ on $M$ \item[(iii)] The bundle $(\phi^*B, \phi^*\nabla)$ is autodual \end{description} {\bf Proof:} Follows from \ref{_phi^*_B_2-forms_1,1_Lemma_} applied to $\omega = \nabla^2$. \blacksquare \hfill For any point $I\in {\Bbb C} P^1$, consider the corresponding algebra embedding ${\Bbb C} \stackrel {c_I} \hookrightarrow {\Bbb H}$. Let $\rho_I$ be the action of ${\Bbb C}^*$ on $(M, I)$ obtained as a restriction of $\rho$ to $c_I({\Bbb C}^*)\subset {\Bbb H}^*$. Clearly from \ref{_H^*_specia_Definition_} (ii), $\rho_I$ acts on $(M, I)$ by holomorphic automorphisms. Consider $\operatorname{Tw}(M)$ as a union \[ \operatorname{Tw}(M) = \bigcup_{I\in {\Bbb C} P^1} \pi^{-1}(I), \ \ \pi^{-1}(I)= (M, I) \] Gluing $\rho(I)$ together, we obtain a smooth ${\Bbb C}^*$-action $\rho_{\Bbb C}$ on $\operatorname{Tw}(M)$. \hfill \claim \label{_C^*_acti_on_Tw_holo_Claim_} Consider the action $\rho_{\Bbb C}:\; {\Bbb C}^*\times \operatorname{Tw}(M) {\:\longrightarrow\:} \operatorname{Tw}(M)$ constructed above. Then $\rho_{\Bbb C}$ is holomorphic. {\bf Proof:} It is obvious from construction that $\rho_{\Bbb C}$ is compatible with the complex structure on $\operatorname{Tw}(M)$. \blacksquare \hfill \example Let $M = {\Bbb H}^n \backslash 0$. Since $\operatorname{Tw}({\Bbb H}^n)$ is canonically isomorphic to a total space of the bundle ${\cal O}(1)^n$ over ${\Bbb C} P^1$, the twistor space $\operatorname{Tw}(M)$ is $\operatorname{Tot}({\cal O}(1)^n)$ without zero section. The group ${\Bbb C}^*$ acts on $\operatorname{Tot}({\cal O}(1)^n)$ by dilatation, and the restriction of this action to $\operatorname{Tw}(M)$ coincides with $\rho_{\Bbb C}$. \hfill Consider the map $\sigma:\; \operatorname{Tw}(M) {\:\longrightarrow\:} M$. Let $(B, \nabla)$ be a $B_2$-bundle over $Q$. Since the bundle $(\phi^*B, \phi^*\nabla)$ is autodual, the curvature of $\sigma^*\phi^*\nabla$ has type $(1,1)$. Let $(\sigma^* \phi^*B, (\sigma^*\phi^*\nabla)^{0,1})$ be the holomorphic bundle obtained from $(\phi^*B, \phi^*\nabla)$ by twistor transform. Clearly, this bundle is ${\Bbb C}^*$-equivariant, with respect to the natural ${\Bbb C}^*$-action on $\operatorname{Tw}(M)$. It turns out that any ${\Bbb C}^*$-equivariant bundle $\c F$ on $\operatorname{Tw}(M)$ can be obtained this way, assuming that $\c F$ is compatible with twistor transform. \hfill \theorem\label{_B_2_to_C^*_equiva_Theorem_} In the above assumptions, let $\c C_{B_2}$ be the category of of $B_2$-bundles on $Q$, and $\c C_{\operatorname{Tw}, {\Bbb C}^*}$ the category of ${\Bbb C}^*$-equivariant holomorphic bundles on $\operatorname{Tw}(M)$ which are compatible with the twistor transform. Consider the functor \[ (\sigma^* \phi^*)^{0,1}: \c C_{B_2}{\:\longrightarrow\:} \c C_{\operatorname{Tw}, {\Bbb C}^*}, \] $(B, \nabla){\:\longrightarrow\:} (\sigma^* \phi^*B, (\sigma^*\phi^*\nabla)^{0,1})$, constructed above. Then $(\sigma^* \phi^*)^{0,1}$ establishes an equivalence of categories. \hfill We prove \ref{_B_2_to_C^*_equiva_Theorem_} in Subsection \ref{_twi_tra_H^*_Subsection_}. \hfill \remark Let $Q$ be an arbitrary quaternionic-K\"ahler manifold, and $M= \c U(Q)$ the corresponding fibration. Then $M$ is hypercomplex, and its twistor space is equipped with a natural holomorphic action of ${\Bbb C}^*$. This gives necessary ingredients needed to state \ref{_B_2_to_C^*_equiva_Theorem_} for $Q$ with negative scalar curvature. The proof which we give for $Q$ with positive scalar curvature will in fact work for all quaternionic-K\"ahler manifolds. \hfill \question What happens with this construction when $Q$ is a hyperk\"ahler manifold? \hfill In this paper, we need \ref{_B_2_to_C^*_equiva_Theorem_} only in the case $Q={\Bbb H}P^n$, $M = {\Bbb H}^n\backslash 0$, but the general proof is just as difficult. \hfill \subsection{${\Bbb C}^*$-equivariant bundles and twistor transform} \label{_C^*_equiva_and_twistor_Subsection_} Let $M$ be a hyperk\"ahler manifold, and $\operatorname{Tw}(M)$ its twistor space. Recall that $\operatorname{Tw}(M)={\Bbb C} P^1\times M$ is equipped with a canonical anticomplex involution $\iota$, which acts as identity on $M$ and as central symmetry $I{\:\longrightarrow\:} -I$ on ${\Bbb C} P^1= S^2$. \hfill \proposition \label{_conne_flat_along_leave_C^*_Proposition_} Let $M$ be a hyperk\"ahler manifold, and $\operatorname{Tw}(M)$ its twistor space. Assume that $\operatorname{Tw}(M)$ is equipped with a free holomorphic action $\rho(z):\; \operatorname{Tw}(M){\:\longrightarrow\:} \operatorname{Tw}(M)$ of ${\Bbb C}^*$, acting along the fibers of $\pi:\; \operatorname{Tw}(M){\:\longrightarrow\:} {\Bbb C} P^1$. Assume, moreover, that $\iota\circ \rho(z) = \rho(\bar z) \circ \iota$, where $\iota$ is the natural anticomplex involution of $\operatorname{Tw}(M)$.% \footnote{These assumptions are automatically satisfied when $M$ is equipped with a special ${\Bbb H}^*$-action, and $\rho(z)$ is the corresponding ${\Bbb C}^*$-action on $\operatorname{Tw}(M)$.} Let $\c F$ be a ${\Bbb C}^*$-equivariant holomorphic vector bundle on $\operatorname{Tw}(M)$. Assume that $\c F$ is compatible with the twistor transform. Let $\nabla_{\c F}$ be the natural connection on $\c F$ (\ref{_cano_conne_Remark_}). Then $\nabla_{\c F}$ is flat along the leaves of $\rho$. \hfill {\bf Proof:} First of all, let us recall the construction of the natural connection $\nabla_{\c F}$. Let $\c F$ be an arbitrary bundle compatible with the twistor transform. We construct $\nabla_{\c F}$ in terms of the isomorphism $\Psi_{1,2}$ defined in \ref{_F_1_=F_2_Lemma_}. \hfill Consider an induced complex structure $I$. Let $F_I$ be the restriction of $\c F$ to $(M, I)= \pi^{-1}(I)\subset \operatorname{Tw}(M)$. Consider the evaluation map \[ p_I:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, I)\] (Subsection \ref{_twi_lines_C^*_Subsection_}). In a similar way we define the holomorphic vector bundle $F_{-I}$ on $(M, -I)$ and the map $p_{-I}:\; \operatorname{Lin}(M) {\:\longrightarrow\:} (M, -I)$. Denote by $F_1$, $F_{-1}$ the sheaves $p_I^*(F_I)$, $p_{-I}^*(F_{-I})$. In \ref{_F_1_=F_2_Lemma_}, we constructed an isomorphism $\Psi_{1,-1}:\; F_1 {\:\longrightarrow\:} F_{-1}$. Let us identify $\operatorname{Lin}(M)$ with $(M, I)\times (M, I)$ (this idenitification is naturally defined in a neighbourhood of $\operatorname{Hor}\subset \operatorname{Lin}(M)$ -- see \ref{_Lin_is_MxM_Proposition_}). Then the maps $p_I$, $p_{-I}$ became projections to the relevant components. Let \begin{equation*} \begin{split} \bar\partial:\; F_1 &{\:\longrightarrow\:} F_1 \otimes p_I^* \Omega^1(M, -I), \\ \partial:\; F_{-1} &{\:\longrightarrow\:} F_{-1} \otimes p_{-I}^* \Omega^1(M, I), \end{split} \end{equation*} be the sheaf maps obtained as pullbacks of de Rham differentials (the tensor product is taken in the category of coherent sheaves over $\operatorname{Lin}(M)$). Twisting $\partial$ by an isomorphism $\Psi_{1,-1}:\; F_1 {\:\longrightarrow\:} F_{-1}$, we obtain a map \[ \partial^\Psi:\; F_1 {\:\longrightarrow\:} F_1 \otimes p_I^* \Omega^1(M, I). \] Adding $\bar \partial$ and $\partial^\Psi$, we obtain \[ \nabla:\; F_1 {\:\longrightarrow\:} F_1 \otimes \bigg (p_I^* \Omega^1(M, I) \oplus p_I^* \Omega^1(M, -I)\bigg). \] Clearly, $\nabla$ satisfies the Leibniz rule. Moreover, the sheaf $p_I^* \Omega^1(M, I) \oplus p_I^* \Omega^1(M, -I)$ is naturally isomorphic to the sheaf of differentials over \[ \operatorname{Lin}(M) = (M,I)\times (M, -I).\] Therefore, $\nabla$ can be considered as a connection in $F_1$, or as a real analytic connection in a real analytic complex vector bundle underlying $F_I$. From the definition of $\nabla_{\c F}$ (\cite{_NHYM_}), it is clear that $\nabla_{\c F}\restrict{(M, I)}$ equals $\nabla$. \hfill Return to the proof of \ref{_conne_flat_along_leave_C^*_Proposition_}. Consider a ${\Bbb C}^*$-action $\rho_I(z)$ on $(M, I)$, $(M, -I)$ induced from the natural embeddings $(M,I)\hookrightarrow \operatorname{Tw}(M)$, $(M,-I)\hookrightarrow \operatorname{Tw}(M)$. Then $F_I$ is a ${\Bbb C}^*$-equivariant bundle. Since $\iota\circ \rho(z) = \rho(\bar z) \circ \iota$, the identification $\operatorname{Lin}(M)=(M, I)\times (M, I)$ is compatible with ${\Bbb C}^*$-action. Let ${\mathbf r}= \frac{d}{dr}$ be the holomorphic vector field on $(M, I)$ corresponding to the ${\Bbb C}^*$-action. To prove \ref{_conne_flat_along_leave_C^*_Proposition_}, we have to show that the operator \[ [\nabla_{\mathbf r},\nabla_{\bar {\mathbf r}}]:\; F_I {\:\longrightarrow\:} F_I\otimes \Lambda^{1,1}(M, I) \] vanishes. Consider the equivariant structure operator \[ \rho(z)^F: \rho_I(z)^* F_I {\:\longrightarrow\:} F_I. \] Let $U$ be a ${\Bbb C}^*$-invariant Stein subset of $(M, I)$. Consider $\rho(z)^F$ an an endomorphism of the space of global holomorphic sections $\Gamma_U(F_I)$. Let \[ D_r(f):= \lim\limits_{\epsilon\rightarrow 0} \frac{\rho_I(1+\epsilon)}{\epsilon}, \] for $f\in \Gamma_U(F_I)$. Clearly, $D_r$ is a well defined sheaf endomorphism of $F_I$, satisfying \[ D_r(\alpha \cdot f) = \frac{d}{dr}\alpha \cdot f +\alpha \cdot D_r(f), \] for all $\alpha\in {\cal O}_{(M,I)}$. We say that a holomorphic section $f$ of $F_I$ is {\bf ${\Bbb C}^*$-invariant} if $D_r(f)=0$. Clearly, the ${\cal O}_{(M,I)}$-sheaf $F_I$ is generated by ${\Bbb C}^*$-invariant sections. Therefore, it suffices to check the equality \[ [\nabla_{\mathbf r},\nabla_{\bar {\mathbf r}}] (f)=0 \] for holomorphic ${\Bbb C}^*$-invariant $f\in F_I$. Since $f$ is holomorphic, we have $\nabla_{\bar {\mathbf r}} f =0$. Thus, \[ [\nabla_{\mathbf r},\nabla_{\bar {\mathbf r}}] (f) = \nabla_{\bar {\mathbf r}} \nabla_{\mathbf r}(f). \] We obtain that \ref{_conne_flat_along_leave_C^*_Proposition_} is implied by the following lemma. \hfill \lemma \label{_conne_on_C^*_inva_Lemma_} In the above assumptions, let $f$ be a ${\Bbb C}^*$-invariant section of $F_I$. Then $\nabla_{\mathbf r}(f)=0$. \hfill {\bf Proof:} Return to the notation we used in the beginning of the proof of \ref{_conne_flat_along_leave_C^*_Proposition_}. Then, $\nabla(f) = \bar \partial(f) +\partial^\Psi(f)$. Since $f$ is holomorphic, $\bar \partial(f)=0$, so we need to show that $\partial^\Psi(f)({\mathbf r})=0$. By definition of $\partial^\Psi$, this is equivalent to proving that \[ \partial\Psi_{1,-1}(f)({\mathbf r})=0.\] Consider the ${\Bbb C}^*$-action on $\operatorname{Lin}(M)$ which is induced by the ${\Bbb C}^*$-action on $\operatorname{Tw}(M)$. Since the maps $p_I$, $p_{-I}$ are compatible with the ${\Bbb C}^*$-action, the sheaves $F_1$, $F_{-1}$ are ${\Bbb C}^*$-equivariant. We can repeat the construction of the operator $D_r$ for the sheaf $F_{-I}$. This allows one to speak of holomorphic ${\Bbb C}^*$-invariant sections of $F_{-I}$. Pick a ${\Bbb C}^*$-invariant Stein subset $U\subset (M, -I)$. Since the statement of \ref{_conne_on_C^*_inva_Lemma_} is local, we may assume that $M=U$. Let $g_1, ... , g_n$ be a set of ${\Bbb C}^*$-invariant sections of $F_I$ which generated $F_I$. Then, the sections $p_{-I}^*(g_1), ..., p_{-I}^*(g_n)$ generate $F_{-1}$. Consider the section $\Psi_{1,-1}(f)$ of $F_{-1}$. Clearly, $\Psi_{1,-1}$ commutes with the natural ${\Bbb C}^*$-action. Therefore, the section $\Psi_{1,-1}(f)$ is ${\Bbb C}^*$-invariant, and can be written as \[ \Psi_{1,-1}(f)= \sum \alpha_i p_{-I}^* (g_i), \] where the functions $\alpha_i$ are ${\Bbb C}^*$-invariant. By definition of $\partial$ we have \[ \partial\left(\sum \alpha_i p_{-I}^* (g_i)\right) = \sum \partial(\alpha_i p_{-I}^* (g_i))+\sum \alpha_i \partial (p_{-I}^* (g_i)). \] On the other hand, $g_i$ is a holomorphic section of $F_{-I}$, so $\partial p_{-I}^* (g_i)=0$. We obtain \[ \partial\left(\sum \alpha_i p_{-I}^* \cdot (g_i)\right) = \sum \partial\alpha_i p_{-I}^* (g_i). \] Thus, \[ \partial\Psi_{1,-1}(f)({\mathbf r}) = \sum \frac{\partial\alpha_i}{\partial r} p_{-I}^* (g_i), \] but since the functions $\alpha_i$ are ${\Bbb C}^*$-invariant, their derivatives along ${\mathbf r}$ vanish. We obtain $\partial\Psi_{1,-1}(f)({\mathbf r})=0$. This proves \ref{_conne_on_C^*_inva_Lemma_}. \ref{_conne_flat_along_leave_C^*_Proposition_} is proven. \blacksquare \subsection{Twistor transform and the ${\Bbb H}^*$-action} \label{_twi_tra_H^*_Subsection_} For the duration of this Subsection, we fix a hyperk\"ahler manifold $M$, equipped with a special ${\Bbb H}^*$-action $\rho$, and the corresponding quaternionic-K\"ahler manifold $Q= M/{\Bbb H}^*$. Denote the natural quotient map by $\phi:\; M{\:\longrightarrow\:} Q$. Clearly, \ref{_B_2_to_C^*_equiva_Theorem_} is an immediate consequence of the following theorem. \hfill \theorem \label{_C^*_equi_cano_conne_Theorem_} Let $\c F$ be a ${\Bbb C}^*$-equivariant holomorphic bundle over $\operatorname{Tw}(M)$, which is compatible with the twistor transform. Consider the natural connection $\nabla_{\c F}$ on $\c F$. Then $\nabla_{\c F}$ is flat along the leaves of ${\Bbb H}^*$-action. \hfill {\bf Proof:} The leaves of ${\Bbb H}^*$-action are parametrized by the points of $q\in Q$. Consider such a leaf $M_q:=\phi^{-1}(q)\subset M$. Clearly, $M_q$ is a hyperk\"ahler submanifold in $M$, equipped with a special action of ${\Bbb H}^*$. Moreover, the restriction of $\c F$ to $\operatorname{Tw}(M_q)\subset \operatorname{Tw}(M)$ satisfies assumptions of \ref{_C^*_equi_cano_conne_Theorem_}. To prove that $\nabla_{\c F}$ is flat along the leaves of ${\Bbb H}^*$-action, we have to show that ${\c F}\restrict{\operatorname{Tw}(M_q)}$ is flat, for all $q$. Therefore, it suffices to prove \ref{_C^*_equi_cano_conne_Theorem_} for $\dim_{\Bbb H} M=1$. \hfill \lemma\label{_cano_conne_flat_on_4-dim_Lemma_} We work in notation and assumptions of \ref{_C^*_equi_cano_conne_Theorem_}. Assume that $\dim_{\Bbb H} M=1$. Then the connection $\nabla_{\c F}$ is flat. \hfill {\bf Proof:} Let $I$ be an induced complex structure, and $F_I:= F\restrict{(M,I)}$ the corresponding holomorphic bundle on $(M, I)$. Denote by $z_I$ the vector field corresponding to the ${\Bbb C}^*$-action $\rho_I$ on $(M, I)$. By definition, the connection $\nabla\restrict{F_I}$ has $SU(2)$-invariant curvature $\Theta_I$. On the other hand, $\Theta_I(z_I, \bar z_I)=0$ by \ref{_conne_flat_along_leave_C^*_Proposition_}. Since $\nabla_{\c F}=\sigma^*\nabla$ is a pullback of an autodual connection $\nabla$ on $M$, its curvature is a pullback of $\Theta_I$. In particular, $\Theta= \Theta_I$ is independent from the choice of induced complex structure $I$. We obtain that $\Theta(z_I, \bar z_I)=0$ for all induced complex structures $I$ on $M$. Now \ref{_cano_conne_flat_on_4-dim_Lemma_} is implied by the following linear-algebraic claim. \hfill \claim \label{_SU_2_inva_2-form_z_bar_z=0_is_zero_Claim_} Let $M$ be a hyperk\"ahler manifold equipped with a special ${\Bbb H}^*$-action, $\dim_{\Bbb H}M=1$. Consider the vectors $z_I$, $\bar z_I$ defined above. Let $\Theta$ be a smooth $SU(2)$-invariant 2-form, such that for all induced complex structures, $I$, we have $\Theta(z_I, \bar z_I)=0$. Then $\Theta=0$. \hfill {\bf Proof:} The proof of \ref{_SU_2_inva_2-form_z_bar_z=0_is_zero_Claim_} is an elementary calculation. Fix a point $m_0\in M$. Consider the flat hyperk\"ahler manifold ${\Bbb H}\backslash 0$, equipped with a natural special action of ${\Bbb H}^*$. From the definition of a special action, it is clear that the map $\rho$ defines a covering ${\Bbb H}\backslash 0{\:\longrightarrow\:} M$, $h{\:\longrightarrow\:} \rho(h)m_0$ of hyperk\"ahler manifolds, and this covering is compatible with the special action. Therefore, the hyperk\"ahler manifold $M$ is flat, and the ${\Bbb H}^*$-action is linear in the flat coordinates. Let \[ \Lambda^2(M) = \Lambda^+(M) \oplus\Lambda^-(M) \] be the standard decomposition of $\Lambda^2(M)$ according to the eigenvalues of the Hodge $*$ operator. Consider the natural Hermitian metric on $\Lambda^2(M)$. Then $\Lambda^-(M)$ is the bundle of $SU(2)$-invariant 2-forms (see, e. g., \cite{_Verbitsky:Hyperholo_bundles_}), and $\Lambda^+(M)$ is its orthogonal complement. Consider the corresponding orthogonal projection $\Pi:\; \Lambda^2(M) {\:\longrightarrow\:} \Lambda^-(M)$. Denote by $dz_I\wedge d \bar z_I$ the differential form which is dual to the bivector $z_I\wedge \bar z_I$. Let $R\subset \Lambda^-(M)$ be the $C^\infty(M)$-subsheaf of $\Lambda^-(M)$ generated by $\Pi(dz_I\wedge d \bar z_I)$, for all induced complex structures $I$ on $M$. Clearly, $\Theta\in \Lambda^-(M)$ and $\Theta$ is orthogonal to $R\subset \Lambda^-(M)$. Therefore, to prove that $\Theta=0$ it suffices to show that $R= \Lambda^-(M)$. Since $M$ is covered by ${\Bbb H}\backslash 0$, we may prove $R= \Lambda^-(M)$ in assumption $M={\Bbb H}\backslash 0$. \hfill Let $\gamma$ be the real vector field corresponding to dilatations of $M={\Bbb H}\backslash 0$, and $d\gamma$ the dual 1-form. Clearly, \[ dz_I\wedge d \bar z_I = 2\sqrt{-1}\: d\gamma\wedge I(d\gamma). \] Averaging $d\gamma\wedge I(d\gamma)$ by $SU(2)$, we obtain \[ \Pi(dz_I\wedge d \bar z_I) = \sqrt{-1}\:\bigg( d\gamma\wedge I(d\gamma) - J(d\gamma) \wedge K(d\gamma)\bigg) \] where $I$, $J$, $K$ is the standard triple of generators for quaternion algebra. Similarly, \[ \Pi(dz_J\wedge d \bar z_J) = \sqrt{-1}\:\bigg( d\gamma\wedge J(d\gamma) + K(d\gamma) \wedge I(d\gamma)\bigg) \] and \[ \Pi(dz_K\wedge d \bar z_K) = \sqrt{-1}\:\bigg( d\gamma\wedge K(d\gamma) + I(d\gamma) \wedge J(d\gamma)\bigg) \] Thus, $\Pi(R)$ is a 3-dimensional sub-bundle of $\Lambda^-(M)$. Since $\dim \Lambda^-(M) =3$, we have $\Pi(R)= \Lambda^-(M)$. This proves \ref{_SU_2_inva_2-form_z_bar_z=0_is_zero_Claim_}. \ref{_cano_conne_flat_on_4-dim_Lemma_} and \ref{_C^*_equi_cano_conne_Theorem_} is proven. \blacksquare \subsection{Hyperholomorphic sheaves and ${\Bbb C}^*$-equivariant bundles over $M_{\rm fl}$} \label{_hyperho_shea_C^*_equiv_Y-M_on_blow-up_Subsection_} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf over $(M, I)$, equipped with a hyperholomorphic connection. Assume that $F$ has an isolated singularity in $x\in M$. Consider the sheaf $\c F$ on $\operatorname{Tw}(M)$ corresponding to $\c F$ as in the proof of \ref{_conne_=>_hyperho_Proposition_}. Let $s_x\subset \operatorname{Tw}(M)$ be the horizontal twistor line corresponding to $x$, and $\frak m$ its ideal. Consider the associated graded sheaf of $\frak m$. Denote by $\operatorname{Tw}^{gr}$ the spectre of this associated graded sheaf. Clearly, $\operatorname{Tw}^{gr}$ is naturally isomorphic to $\operatorname{Tw}(T_x M)$, where $T_xM$ is the flat hyperk\"ahler manifold corresponding to the space $T_xM$ with induced quaternion action. Consider the natural ${\Bbb H}^*$-action on $T_xM$. This provides the hyperk\"ahler manifold $T_xM\backslash 0$ with a special ${\Bbb H}^*$-action. Let $s_0\subset \operatorname{Tw}^{gr}$ be the horisontal twistor line corresponding to $s_x$. The space $\operatorname{Tw}^{gr}\backslash s_0$ is equipped with a holomorphic ${\Bbb C}^*$-action (\ref{_C^*_acti_on_Tw_holo_Claim_}). Denote by $\c F^{gr}$ the sheaf on $\operatorname{Tw}^{gr}$ associated with $\c F$. Clearly, $\c F^{gr}$ is ${\Bbb C}^*$-equivariant. In order to be able to apply \ref{_B_2_to_C^*_equiva_Theorem_} and \ref{_C^*_equi_cano_conne_Theorem_} to $\c F^{gr}\restrict{\operatorname{Tw}^{gr}\backslash s_0}$, we need only to show that $\c F^{gr}$ is compatible with twistor transform. \hfill \proposition \label{_F^gr_compa_twi_tra_Proposition_} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf over $(M, I)$, equipped with a hyperholomorphic connection. Assume that $\c F$ has an isolated singularity in $x\in M$. Let $\c F^{gr}$ be the ${\Bbb C}^*$-equivariant bundle on $\operatorname{Tw}^{gr}\backslash s_0$ constructed above. Then \begin{description} \item[(i)] the bundle $\c F^{gr}$ is compatible with twistor transform. \item[(ii)] Moreover, the natural connection $\nabla_{\c F^{gr}}$ (\ref{_cano_conne_Remark_}) is Hermitian. \end{description} {\bf Proof:} The argument is clear, but cumbersome, and essentially hinges on taking associate graded quotients everywhere and checking that all equations remain true. We give a simplified version of the proof, which omits some details and notation. Consider the bundle $\c F\restrict {M\backslash s_x}$. This bundle is compatible with twistor transform, and therefore, is equipped with a natural connection $\nabla_{\c F}$. This connection is constructed using the isomorphism $\Psi_{1,-1}:\; F_1{\:\longrightarrow\:} F_{-1}$ (see the proof of \ref{_conne_flat_along_leave_C^*_Proposition_}). We apply the same consideration to $\c F^{gr}\restrict{(T_x M, I)}$, and show that the resulting connection $\nabla_{\c F^{gr}}$ is hyperholomorphic. This implies that $\c F^{gr}$ admits a $(1,1)$-connection which is a pullback of some connection on $\c F^{gr}\restrict{(T_x M, I)}$. This argument is used to prove that $\c F^{gr}$ is compatible with the twistor transform. We use the notation introduced in the proof of \ref{_conne_flat_along_leave_C^*_Proposition_}. Let $\operatorname{Lin}^{gr}$ be the space of twistor maps in $\operatorname{Tw}^{gr}$. Consider the maps $p_{\pm I}^{gr}:\; \operatorname{Lin}^{gr}{\:\longrightarrow\:} (T_x M, \pm I)$ and the sheaves $F^{gr}_{\pm1}:=(p_{\pm I}^{gr})^*\c F^{gr}_{\pm I}$ obtained in the same way as the maps $p_{\pm I}$ and the sheaves $F_{\pm1}$ from the corresponding associated graded objects. Taking the associated graded of $\Psi_{1,-1}$ gives an isomorphism $\Psi^{gr}_{1,-1}:\; F^{gr}_1{\:\longrightarrow\:} F^{gr}_{-1}$. Using the same construction as in the proof of \ref{_conne_flat_along_leave_C^*_Proposition_}, we obtain a connection operator \[ \bar \partial^{gr}+\partial^{\Psi^{gr}}=\nabla^{gr}_I:\; F_1^{gr} {\:\longrightarrow\:} F_1^{gr} \otimes \bigg ((p^{gr}_{-I})^* \Omega^1(T_x M, I) \oplus (p^{gr}_I)^* \Omega^1(T_x M, -I)\bigg). \] Since $(\bar \partial^{gr})^2 = (\partial^{\Psi^{gr}})^2=0$, the curvature of $\nabla^{gr}_I$ has Hodge type $(1,1)$ with respect to $I$. To prove that $\nabla^{gr}_I$ is hyperholomorphic, we need to show that the curvature of $\nabla^{gr}_I$ has type $(1,1)$ with respect to every induced complex structure. Starting from another induced complex structure $J$, we obtain a connection $\nabla^{gr}_J$, with the curvature of type $(1,1)$ with respect to $J$. To prove that $\nabla^{gr}_J$ is hyperholomorphic it remains to show that $\nabla^{gr}_J=\nabla^{gr}_I$. Let $\nabla_I$, $\nabla_J$ be the corresponding operators on $F_1$. From the construction, it is clear that $\nabla^{gr}_I$, $\nabla^{gr}_J$ are obtained from $\nabla_I$, $\nabla_J$ by taking the associated graded quotients. On the other hand, $\nabla_I = \nabla_J$. Therefore, the connections $\nabla^{gr}_I$ and $\nabla^{gr}_J$ are equal. We proved that the bundle $\c F^{gr}\restrict{\operatorname{Tw}^{gr}\backslash s_0}$ is compatible with the twistor structure. To prove \ref{_F^gr_compa_twi_tra_Proposition_}, it remains to show that the natural connection on $\c F^{gr}$ is Hermitian. The bundle ${\c F} \restrict{\operatorname{Tw}(M\backslash{x_0})}$ is by definition Hermitian. Consider the corresponding isomorphism $\c F {\:\longrightarrow\:} (\iota^*\bar{\c F})^*$ (\ref{_twi_tra_for_semili_Proposition_}). Taking an associate graded map, we obtain an isomorphism \[ \c F^{gr}\oldtilde\rightarrow (\iota^*\bar{\c F}^{gr})^*.\] This gives a non-degenerate semilinear form $h^{gr}$ on $\c F^{gr}$. It remains only to show that $h^{gr}$ is pseudo-Hermitian (i. e. satisfies $h(x, y) = \overline{h(y,x)}$) and positive definite. \hfill Let $M^{gr}_{\Bbb C}$ be a complexification of $M^{gr}=T_x M$, $M^{gr}_{\Bbb C} = \operatorname{Lin}(M^{gr})$. Consider the corresponding complex vector bundle $\c F^{gr}_{\Bbb C}$ over $M^{gr}_{\Bbb C}$ underlying $\c F^{gr}$. The metric $h^{gr}$ can be considered as a semilinear form $\c F^{gr}_{\Bbb C}\times \c F^{gr}_{\Bbb C}{\:\longrightarrow\:} {\cal O}_{M^{gr}_{\Bbb C}}$. This semilinear form is obtained from the corresponding form $h$ on $\c F$ by taking the associate graded quotients. Since $h$ is Hermitian, the form $h^{gr}$ is pseudo-Hermitian. To prove that $h^{gr}$ is positive semidefinite, we need to show that for all $f\in \c F^{gr}_{\Bbb C}$, the function $h^{gr}(f, \bar f)$ belongs to ${\cal O}_{M^{gr}_{\Bbb C}}\cdot\bar{\cal O}_{M^{gr}_{\Bbb C}}$, where ${\cal O}_{M^{gr}_{\Bbb C}}\cdot\bar{\cal O}_{M^{gr}_{\Bbb C}}$ denotes the ${\Bbb R}^{>0}$-semigroup of ${\cal O}_{M^{gr}_{\Bbb C}}$ generated by $x\cdot\bar x$, for all $x\in {\cal O}_{M^{gr}_{\Bbb C}}$. A similar property for $h$ holds, because $h$ is positive definite. Clearly, taking associated graded quotient of the semigroup ${\cal O}_{M_{\Bbb C}}\cdot\bar{\cal O}_{M_{\Bbb C}}$, we obtain ${\cal O}_{M^{gr}_{\Bbb C}}\cdot\bar{\cal O}_{M^{gr}_{\Bbb C}}$. Thus, \[ h^{gr}(f, \bar f)\in \left({\cal O}_{M_{\Bbb C}}\cdot\bar{\cal O}_{M_{\Bbb C}}\right)^{gr} = {\cal O}_{M^{gr}_{\Bbb C}}\cdot\bar{\cal O}_{M^{gr}_{\Bbb C}} \] This proves that $h^{gr}$ is positive semidefinite. Since $h^{gr}$ is non-degenerate, this form in positive definite. \ref{_F^gr_compa_twi_tra_Proposition_} is proven. \blacksquare \hfill \remark\label{_exte_conne_conje_Remark_} Return to the notations of \ref{_desingu_hyperho_Theorem_}. Consider the bundle $\pi^* F\restrict C$, where $C= {\Bbb P}T_xM$ is the blow-up divisor. Clearly, this bundle corresponds to the graded sheaf $F_I^{gr}= \c F^{gr}\restrict{(M,I)}$ on $(T_xM, I)$. By \ref{_F^gr_compa_twi_tra_Proposition_} (see also \ref{_C^*_equi_cano_conne_Theorem_}), the bundle $\pi^* F\restrict C$ is equipped with a natural ${\Bbb H}^*$-invariant connection and Hermitian structure.% \footnote{As usually, coherent sheaves over projective variety $X$ correspond to finitely generated graded modules over the graded ring $\oplus \Gamma({\cal O}_X(i))$.} The sheaf $\pi^*F \restrict{\tilde M \backslash C}$ is a hyperholomorphic bundle over $\tilde M \backslash C\cong M\backslash x_0$. Therefore, $\pi^*F \restrict{\tilde M \backslash C}$ is equipped with a natural metric and a hyperholomorphic connection. It is expected that the natural connection and metric on $\pi^* F\restrict{\tilde M\backslash C}$ can be extended to $\pi^* F$, and the rectriction of the resulting connection and metric to $\pi^* F\restrict C$ coincides with that given by \ref{_F^gr_compa_twi_tra_Proposition_} and \ref{_C^*_equi_cano_conne_Theorem_}. This will give an alternative proof of \ref{_F^gr_compa_twi_tra_Proposition_} (ii), because a continuous extension of a positive definite Hermitian metric is a positive semidefinite Hermitian metric. \subsection{Hyperholomorphic sheaves and stable bundles on ${\Bbb C} P^{2n+1}$} The purpose of the current Section was to prove the following result, which is a consequence of \ref{_F^gr_compa_twi_tra_Proposition_} and \ref{_C^*_equi_cano_conne_Theorem_}. \hfill \theorem \label{_hyperho_blow-up_stable_Theorem_} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure and $F$ a reflexive sheaf on $(M,I)$ admitting a hyperholomorphic connection. Assume that $F$ has an isolated singularity in $x\in M$, and is locally trivial outside of $x$. Let $\pi:\; \tilde M{\:\longrightarrow\:} (M,I)$ be the blow-up of $(M,I)$ in $x$. Consider the holomorphic vector bundle $\pi^* F$ on $\tilde M$ (\ref{_desingu_hyperho_Theorem_}). Let $C\subset (M,I)$ be the blow-up divisor, $C= {\Bbb P}T_xM$. Then the holomorphic bundle $\pi^* F\restrict{ C}$ admits a natural Hermitian connection $\nabla$ which is flat along the leaves of the natural ${\Bbb H}^*$-action on ${\Bbb P}T_xM$. Moreover, the connection $\nabla$ is Yang-Mills, with respect to the Fubini-Study metric on $C= {\Bbb P}T_xM$, the degree $\deg c_1\left(\pi^* F\restrict{C}\right)$ vanishes, and the holomorphic vector bundle $\pi^* F\restrict{C}$ is polystable. \hfill {\bf Proof:} By definition, coherent sheaves on $C= {\Bbb P}T_xM$ correspond bijectively to ${\Bbb C}^*$-equivariant sheaves on $T_xM\backslash 0$. Let $F^{gr}$ be the associated graded sheaf of $F$ (Subsection \ref{_hyperho_shea_C^*_equiv_Y-M_on_blow-up_Subsection_}). Consider $F^{gr}$ as a bundle on $T_xM\backslash 0$. In the notation of \ref{_F^gr_compa_twi_tra_Proposition_}, $F^{gr}= \c F^{gr}\restrict{(M, I)}$. By \ref{_F^gr_compa_twi_tra_Proposition_}, the sheaf $\c F^{gr}$ is ${\Bbb C}^*$-equivariant and compatible with the twistor structure. According to the Swann's formalism for bundles (\ref{_C^*_equi_cano_conne_Theorem_}), the bundle $F^{gr}\restrict{T_xM\backslash 0}$ is equipped with a natural Hermitian connection $\nabla_{F^{gr}}$ which is flat along the leaves of ${\Bbb H}^*$-action. Let $(B, \nabla_{\Bbb H})$ be the corresponding $B_2$-bundle on \[ {\Bbb P}_{\Bbb H} T_x M:= \bigg(T_xM\backslash 0\bigg)/{\Bbb H}^*\cong {\Bbb H}P^n. \] Then $\pi^*F\restrict C$ is a holomorphic bundle over the corresponding twistor space $C= \operatorname{Tw}({\Bbb P}_{\Bbb H} T_x M)$, obtained as a pullback of $(B, \nabla_{\Bbb H})$ as in \ref{_B_2_=_holo_on_Tw_Claim_}. The natural K\"ahler metric on the twistor space $C= \operatorname{Tw}({\Bbb P}_{\Bbb H} T_x M)$ is the Fubini-Study metric (\ref{_Tw_HP^n_Fu-St_Example_}). By \ref{_twi_tra_YM_q-K_Theorem_}, the bundle $\pi^*F\restrict C$ is Yang-Mills and has $\deg c_1\left(\pi^* F\restrict{C}\right) =0$. Finally, by Uhlenbeck-Yau theorem (\ref{_UY_Theorem_}), the bundle $\pi^*F\restrict C$ is polystable. \blacksquare \section{Moduli spaces of hyperholomorphic sheaves and bundles} \label{_modu_hyperho_Section_} \subsection{Deformation of hyperholomorphic sheaves with isolated singularities} The following theorem is an elementary consequence of \ref{_hyperho_blow-up_stable_Theorem_}. The proof uses well known results on stability and reflexization (see, for instance, \cite{_OSS_}). The main idea of the proof is the following. Given a family of hyperholomorphic sheaves with an isolated singularity, we blow-up this singularity and restrict the obtained family to a blow-up divisor. We obtain a family of coherent sheaves $\frak V_s$, $s\in S$ over ${\Bbb C} P^{2n+1}$, with fibers semistable of slope zero. Assume that for all $s\in S$, $s\neq s_0$, the sheaf $\frak V_s$ is trivial. Then the family $\frak V$ is also trivial, up to a reflexization. \hfill We use the following property of reflexive sheaves. \hfill \definition Let $X$ be a complex manifold, and $F$ a torsion-free coherent sheaf. We say that $F$ is {\bf normal} if for all open subvarieties $U\subset X$, and all closed subvarieties $Y\subset U$ of codimension 2, the restriction \[ \Gamma_U(F) {\:\longrightarrow\:} \Gamma_{U\backslash Y}(F)\] is an isomorphism. \hfill \lemma\label{_normal_refle_Lemma_} Let $X$ be a complex manifold, and $F$ a torsion-free coherent sheaf. Then $F$ is reflexive if and only if $F$ is normal. {\bf Proof:} \cite{_OSS_}, Lemma 1.1.12. \blacksquare \hfill \theorem \label{_reflexi_defo_loca_trivi_Theorem_} Let $M$ be a hyperk\"ahler manifol, $I$ an induced complex structure, $S$ a complex variety and $\frak F$ a family of coherent sheaves over $(M, I)\times S$. Consider the sheaf $F_{s_0}:= {\frak F}\restrict{(M, I)\times \{s_0\}}$. Assume that the sheaf $F_{s_0}$ is equipped with a filtration $\xi$. Let $F_i$, $i= 1, ..., m$ denote the associated graded components of $\xi$, and $F_i^{**}$ denote their reflexizations. Assume that $\frak F$ is locally trivial outside of $(x_0, s_0)\in (M, I)\times S$. Assume, moreover, that all sheaves $F_i^{**}$, $i= 1, ..., m$ admit a hyperholomorphic connection. Then the reflexization ${\frak F}^{**}$ is locally trivial. \hfill {\bf Proof:} Clearly, it suffices to prove \ref{_reflexi_defo_loca_trivi_Theorem_} for ${\frak F}$ reflexive. Let $\tilde X$ be the blow-up of $(M, I)\times S$ in $\{x_0\} \times S$, and $\tilde {\frak F}$ the pullback of $\frak F$ to $\tilde X$. Clearly, $\tilde X= \tilde M\times S$, where $\tilde M$ is a blow-up of $(M,I)$ in $x_0$. Denote by $C\subset \tilde M$ the blow-up divisor of $\tilde M$. Taking $S$ sufficiently small, we may assume that the bundle $\frak F\restrict{\{x_0\}\times(S\backslash\{s_0\})}$ is trivial. Thus, the bundle $\tilde{\frak F}\restrict {(C\times S)\backslash (C\times \{s_0\})}$, which is a pullback of $\frak F\restrict{\{x_0\}\times(S\backslash\{s_0\})}$ under the natural projection $(C\times S)\backslash (C\times \{s_0\}) {\:\longrightarrow\:}\{x_0\}\times(S\backslash\{s_0\}$ is trivial. To prove that $\frak F$ is locally trivial, we have to show that $\tilde {\frak F}$ is locally trivial, and that the restriction of $\tilde {\frak F}$ to $C\times S$ is trivial along the fibers of the natural projection $C \times S{\:\longrightarrow\:} S$. Clearly, to show that $\tilde {\frak F}$ is locally trivial we need only to prove that the fiber $\tilde {\frak F}\restrict z$ has constant dimension for all $z\in C \times S$. Thus, $\tilde {\frak F}$ is locally trivial if and only if $\tilde {\frak F}\restrict {C\times S}$ is locally trivial. This sheaf is reflexive, since it corresponds to an associate graded sheaf of a reflexive sheaf, in the sense of Footnote to \ref{_exte_conne_conje_Remark_}. It is non-singular in codimension 2, because all reflexive sheaves are non-singular in codimension 2 (\cite{_OSS_}, Ch. II, Lemma 1.1.10). By \ref{_hyperho_blow-up_stable_Theorem_}, the sheaf $\tilde {\frak F}\restrict {C\times\{s\}}$ is semistable of slope zero. \ref{_reflexi_defo_loca_trivi_Theorem_} is implied by the following lemma, applied to the sheaf $\tilde {\frak F}\restrict {C\times S}$. \hfill \lemma \label{_F_to_blow-up_stable=>loc_triv_Lemma_} Let $C$ be a complex projective space, $S$ a complex variety and $\frak F$ a torsion-free sheaf over $C\times S$. Consider an open set $U\stackrel j\hookrightarrow C\times S$, which is a complement of $C\times\{s_0\}\subset C\times S$. Assume that the sheaf ${\frak F}\restrict U$ is trivial: ${\frak F}\restrict U\cong {\cal O}_U^n$. Assume, moreover, that $\frak F$ is non-singular in codimension 2, the sheaf $\left({\frak F}\restrict{C \times\{s_0\}}\right)^{**}$ is semistable of slope zero and \[ \operatorname{rk} {\frak F}=\operatorname{rk} {\frak F}\restrict {C\times \{s_0\}}. \] Then the reflexization ${\frak F}^{**}$ of $\frak F$ is a trivial bundle. \hfill {\bf Proof:} Using induction, it suffices to prove \ref{_F_to_blow-up_stable=>loc_triv_Lemma_} assuming that it is proven for all $\frak F'$ with $\operatorname{rk} {\frak F}'<\operatorname{rk} {\frak F}$. We may also assume that $S$ is Stein, smooth and 1-dimensional. \hfill {\bf Step 1:} {\it We construct an exact sequence \[ 0{\:\longrightarrow\:}\frak F_2{\:\longrightarrow\:}\frak F{\:\longrightarrow\:} \operatorname{im} p_{O_1} {\:\longrightarrow\:} 0 \] of sheaves of positive rank, which, as we prove in Step 3, satisfy assumptions of \ref{_F_to_blow-up_stable=>loc_triv_Lemma_}.} \hfill Consider the pushforward sheaf $j_* {\cal O}_U^n$. From the definition of $j_*$, we obtain a canonical map \begin{equation}\label{_embe_tilde_F_to_j_*_Lemma_} {\frak F}{\:\longrightarrow\:} j_* {\cal O}_U^n, \end{equation} and the kernel of this map is a torsion subsheaf in ${\frak F}$. Let $f$ be a coordinate function on $S$, which vanishes in $s_0\in S$. Clearly, \[ j_* {\cal O}_U^n\cong {\cal O}_{C\times S}^n\left[\frac{1}{f}\right]. \] On the other hand, the sheaf ${\cal O}_{C\times S}\left[\frac{1}{f}\right]$ is a direct limit of the following diagram: \[ {\cal O}_{C\times S}^n\stackrel{\cdot f}{\:\longrightarrow\:} {\cal O}_{C\times S}^n\stackrel{\cdot f}{\:\longrightarrow\:} {\cal O}_{C\times S}^n\stackrel{\cdot f}{\:\longrightarrow\:} ..., \] where $\cdot f$ is the injection given by the multiplication by $f$. Thus, the map \eqref{_embe_tilde_F_to_j_*_Lemma_} gives an embedding \[ {\frak F}\stackrel p \hookrightarrow {\cal O}_{C\times S}^n,\] which is idenity outside of $(x_0, s_0)$. Multiplying $p$ by $\frac{1}{f}$ if necessary, we may assume that the restriction $p\restrict {C\times \{s_0\}}$ is non-trivial. Thus, $p$ gives a map \begin{equation} \label{_goth_F_to_calo_Equation_} {\frak F}\restrict {C\times \{s_0\}}{\:\longrightarrow\:} {\cal O}^n_{C\times \{s_0\}}. \end{equation} with image of positive rank. Since both sides of \eqref{_goth_F_to_calo_Equation_} are semistable of slope zero, and ${\cal O}^n_{C\times \{s_0\}}$ is po\-ly\-stable, the map \eqref{_goth_F_to_calo_Equation_} satisfies the following conditions. (see \cite{_OSS_}, Ch. II, Lemma 1.2.8 for details). \hfill \begin{minipage}[m]{0.8\linewidth} Let $F_1:= \operatorname{im} p\restrict {C\times \{s_0\}}$, and $F_2:= \ker p\restrict {C\times \{s_0\}}$. Then the reflexization of $F_1$ is a trivial bundle ${\cal O}^k_{C\times \{s_0\}}$, and $p$ maps $F_1$ to the direct summand $O_1'= {\cal O}^k_{C\times \{s_0\}}\subset {\cal O}^n_{C\times \{s_0\}}$. \end{minipage} \hfill Let $O_1 = {\cal O}^k_{C\times S}\subset {\cal O}^n_{C\times S}$ be the corresponding free subsheaf of ${\cal O}^n_{C\times S}$. Consider the natural projection $\pi_{O_1}$ of ${\cal O}^n_{C\times S}$ to $O_1$. Let $p_{O_1}$ be the composition of $p$ and $\pi_{O_1}$, $\frak F_1$ the image of $p_{O_1}$, and $\frak F_2$ the kernel of $p_{O_1}$. \hfill {\bf Step 2:} {\it We show that the sheaves $\frak F_2$ and $\frak F_1$ and non-singular in codimension 2.} \hfill Consider the exact sequence \[ Tor^1({\cal O}_{C\times \{s_0\}}, {\frak F}_1){\:\longrightarrow\:} \frak F_2\restrict {C\times \{s_0\}}{\:\longrightarrow\:} \frak F\restrict {C\times \{s_0\}} {\:\longrightarrow\:} {\frak F}_1\restrict {C\times \{s_0\}} {\:\longrightarrow\:} 0 \] obtained by tensoring the sequence \[ 0{\:\longrightarrow\:}\frak F_2{\:\longrightarrow\:}\frak F{\:\longrightarrow\:} {\frak F}_1 {\:\longrightarrow\:} 0\] with ${\cal O}_{C\times \{s_0\}}$. From this sequence, we obtain an isomorphism ${\frak F}_1\restrict {C\times \{s_0\}}\cong F_1$. A torsion-free coherent sheaf over a smooth manifold is non-singular in codimension 1 (\cite{_OSS_}, Ch. II, Corollary 1.1.8). Since $\frak F$ is non-singular in codimension 2, the restriction $\frak F\restrict {C\times \{s_0\}}$ is non-singular in codimension 1. Therefore, the torsion of $\frak F\restrict {C\times \{s_0\}}$ has support of codimension at least 2 in $C\times \{s_0\}$. Since the sheaf $F_2$ is a subsheaf of $\frak F\restrict {C\times \{s_0\}}$, its torsion has support of codimension at least 2. Therefore, the singular set of $F_2$ has codimension at least 2 in $C\times \{s_0\}$. The rank of $F_2$ is by definition equal to $n-k$. Since $F_1$ has rank $k$, the singular set of $\frak F_1$ coincides with the singular set of $F_1$. Since the restriction ${\frak F}_1\restrict {C\times \{s_0\}}=F_1$, is a subsheaf of a trivial bundle of dimension $k$ on $C\times \{s_0\}$, it is torsion-free. Therefore, the singularities of ${\frak F}_1$ have codimension at least 2 in $C\times \{s_0\}$. We obtain that the support of $Tor^1({\cal O}_{C\times \{s_0\}}, {\frak F}_1)$ has codimension at least 2 in $C\times \{s_0\}$. Since the quotient sheaf \begin{equation}\label{_F_2_quoti_Equation_} \frak F_2\restrict {C\times \{s_0\}}\bigg / Tor^1({\cal O}_{C\times \{s_0\}}, {\frak F}_1)\cong F_2 \end{equation} is isomorphic to the sheaf $F_2$, this quotient is non-singular in codimension 1. Since we proved that $F_2$ is non-singular in codimension 1, the sheaf $\frak F_2 \restrict {C\times \{s_0\}}$ is also non-singular in codimension 1, and its rank is equal to the rank of $F_2$. Let $R$ be the union of singular sets of the sheaves $\frak F_2$, $\frak F$, ${\frak F}_1$. Clearly, $R$ is contained in $C\times \{s_0\}$, and $R$ coincides with the set of all $x\in C\times \{s_0\}$ where the dimension of the fiber of the sheaves $\frak F_2$, $\frak F$, ${\frak F}_1$ is not equal to $n-k$, $n$, $k$. We have seen that the restrictions of $\frak F_2$, ${\frak F}_1$ to ${C\times \{s_0\}}$ have ranks $n-k$, $k$. Therefore, the singular sets of $\frak F_2$, ${\frak F}_1$ coincide with the singular sets of $\frak F_2\restrict{C\times \{s_0\}}$, ${\frak F}_1\restrict{C\times \{s_0\}}$. We have shown that these singular sets have codimension at least 2 in $C\times \{s_0\}$. On the other hand, $\frak F$ is non-singular in codimension 2, by the conditions of \ref{_F_to_blow-up_stable=>loc_triv_Lemma_}. Therefore, $R$ has codimension at least 3 in $C\times S$. \hfill {\bf Step 3:} {\it We check the assumptions of \ref{_F_to_blow-up_stable=>loc_triv_Lemma_} applied to the sheaves $\frak F_2$, ${\frak F}_1$.} \hfill Since the singular set of ${\frak F}_1$ has codimension 2 in $C\times \{s_0\}$, the ${\cal O}_{C\times \{s_0\}}$-sheaf $Tor^1({\cal O}_{C\times \{s_0\}},{\frak F}_1)$ is a torsion sheaf with support of codimension 2 in $C\times \{s_0\}$. By \eqref{_F_2_quoti_Equation_}, the reflexization of $\frak F_2\restrict {C\times \{s_0\}}$ coincides with the reflexization of $F_2$. Thus, the sheaf $\left(\frak F_2\restrict {C\times \{s_0\}}\right)^{**}$ is semistable. On the other hand, outside of ${C\times \{s_0\}}$, the sheaf $\frak F_2$ is a trivial bundle. Thus, $\frak F_2$ satisfies assumptions of \ref{_F_to_blow-up_stable=>loc_triv_Lemma_}. Similarly, the sheaf $\frak F_1$ is non-singular in codimension 2, its restriction to $C\times \{s_0\}$ has trivial reflexization, and it is free outside of $C\times \{s_0\}$. \hfill {\bf Step 4:} {\it We apply induction and prove \ref{_F_to_blow-up_stable=>loc_triv_Lemma_}.} \hfill By induction assumption, the reflexization of $\frak F_2$ is isomorphic to a trivial bundle ${\cal O}^{n-k}_{C\times S}$. and reflexization of $\frak F_1$ is ${\cal O}^k_{C\times S}$. We obtain an exact sequence \begin{equation}\label{_F_2_to_F_exa_Equation_} 0 {\:\longrightarrow\:} \frak F_2{\:\longrightarrow\:} \frak F {\:\longrightarrow\:} {\frak F}_1 {\:\longrightarrow\:} 0, \end{equation} where the sheaves $\frak F_2$ and ${\frak F}_1$ have trivial reflexizations. Let $V:= C\times S\backslash R$. Restricting the exact sequence \eqref{_F_2_to_F_exa_Equation_} to $V$, we obtain an exact sequence \begin{equation}\label{_F_2_to_F_restri_V_exa_Equation_} 0 {\:\longrightarrow\:} {\cal O}^{n-k}_{V} \stackrel a {\:\longrightarrow\:} {\frak F}\restrict V \stackrel b{\:\longrightarrow\:} {\cal O}^k_{V} {\:\longrightarrow\:} 0. \end{equation} Since $V$ is a complement of a codimension-3 complex subvariety in a smooth Stein domain, the first cohomology of a trivial sheaf on $V$ vanish. Therefore, the sequence \eqref{_F_2_to_F_restri_V_exa_Equation_} splits, and the sheaf ${\frak F}\restrict V$ is a trivial bundle. Consider the pushforward $\zeta_*{\frak F}\restrict V$, where $\zeta:\; V{\:\longrightarrow\:} C\times S$ is the standard map. Then $\zeta_*{\frak F}\restrict V$ is a reflexization of ${\frak F}$ (a pushforward of a reflexive sheaf over a subvariety of codimension 2 or more is reflexive -- see \ref{_normal_refle_Lemma_}). On the other hand, since the sheaf ${\frak F}\restrict V$ is a trivial bundle, its push-forward over a subvariety of codimension at least 2 is also a trivial bundle over $C\times S$. We proved that the sheaf ${\frak F}^{**}=\zeta_*{\frak F}\restrict V$ is a trivial bundle over $C\times S$. The push-forward $\zeta_*{\frak F}\restrict V$ coincides with reflexization of $\frak F$, by \ref{_normal_refle_Lemma_}. This proves \ref{_F_to_blow-up_stable=>loc_triv_Lemma_} and \ref{_reflexi_defo_loca_trivi_Theorem_}. \blacksquare \subsection{The Maruyama moduli space of coherent sheaves} This Subsection is a compilation of results of Gieseker and Maruyama on the moduli of coherent sheaves over projective manifolds. We follow \cite{_OSS_}, \cite{_Maruyama:Si_}. To study the moduli spaces of holomorphic bundles and coherent sheaves, we consider the following definition of stability. \hfill \definition (Gieseker--Maruyama stability) (\cite{_Gieseker_}, \cite{_OSS_}) Let $X$ be a projective variety, ${\cal O}(1)$ the standard line bundle and $F$ a torsion-free coherent sheaf. The sheaf $F$ is called {\bf Gieseker--Maruyama stable} (resp. Gieseker--Maruyama semistable) if for all coherent subsheaves $E\subset F$ with $0<\operatorname{rk} E<\operatorname{rk} F$, we have \[ p_F(k) < p_E(k) \ \ (\text{resp.}, \ \ p_F(k) \leq p_E(k)) \] for all sufficiently large numbers $k\in {\Bbb Z}$. Here \[ p_F(k) = \frac{\dim \Gamma_X(F\otimes {\cal O}(k))}{\operatorname{rk} F}. \] \hfill Clearly, Gieseker--Maruyama stability is weaker than the Mum\-ford-\-Ta\-ke\-mo\-to stability. Every Gieseker--Maruyama semistable sheaf $F$ has a so-called Jordan-H\"older filtration $F_0\subset F_1\subset ...\subset F$ with Gieseker--Maruyama stable successive quotients $F_i/F_{i-1}$. The corresponding associated graded sheaf \[ \oplus F_i/F_{i-1} \] is independent from a choice of a filtration. It is called {\bf the associate graded quotient of the Jordan-H\"older filtration on $F$}. \hfill \definition Let $F$, $G$ be Gieseker--Maruyama semistable sheaves on $X$. Then $F$, $G$ are called {\bf $S$-equivalent} if the corresponding associate graded quotients $\oplus F_i/F_{i-1}$, $\oplus G_i/G_{i-1}$ are isomorphic. \hfill \definition Let $X$ be a complex manifold, $F$ a torsion-free sheaf on $X$, and $Y$ a complex variety. Consider a sheaf $\c F$ on $X\times Y$ which is flat over $Y$. Assume that for some point $s_0\in Y$, the sheaf $\c F\restrict {X\times \{s_0\}}$ is isomorphic to $F$. Then $\c F$ is called {\bf a deformation of $F$ parametrized by $Y$}. We say that a sheaf $F'$ on $X$ is {\bf deformationally equivalent} to $F$ if for some $s\in Y$, the restriction $\c F\restrict {X\times \{s\}}$ is isomorphic to $F'$. Slightly less formally, such sheaves are called {\bf deformations of $F$}. If $F'$ is a (semi-)stable bundle, it is called {\bf a (semi-)stable bundle deformation of $F$.} \hfill \remark\label{_Chern_deforma_equal_} Clearly, the Chern classes of deformationally equivalent sheaves are equal. \hfill \definition \label{_coarse_modu_Definition_} Let $X$ be a complex manifold, and $F$ a torsion-free sheaf on $X$, and $\c M_{mar}$ a complex variety. We say that $\c M_{mar}$ is a {\bf coarse moduli space of deformations of $F$} if the following conditions hold. \begin{description} \item[(i)] The points of $s\in \c M_{mar}$ are in bijective correspondence with $S$-equi\-va\-lence classes of coherent sheaves $F_s$ which are deformationally equivalent to $F$. \item[(ii)] For any flat deformation $\c F$ of $F$ parametrized by $Y$, there exists a unique morphism $\phi:\; Y{\:\longrightarrow\:} \c M_{mar}$ such that for all $s\in Y$, the restriction $\c F\restrict {X\times \{s\}}$ is $S$-equivalent to the sheaf $F_{\phi(s)}$ corresponding to $\phi(s)\in \c M_{mar}$. \end{description} Clearly, the coarse moduli space is unique. By \ref{_Chern_deforma_equal_}, the Chern classes of $F_s$ are equal for all $s\in \c M_{mar}$. \hfill It is clear how to define other kinds of moduli spaces. For instance, replacing the word {\it sheaf} by the word {\it bundle} throughout \ref{_coarse_modu_Definition_}, we obtain a definition of {\bf the coarse moduli space of semistable bundle deformations of $F$}. Further on, we shall usually omit the word ``coarse'' and say ``moduli space'' instead. \hfill \theorem\label{_Maruya_exists_} (Maruyama) Let $X$ be a projective manifold and $F$ a coherent sheaf over $X$. Then the Maruyama moduli space $\c M_{mar}$ of deformations of $F$ exists and is compact. {\bf Proof:} See, e. g., \cite{_Maruyama:Si_}. \blacksquare \hfill \subsection{Moduli of hyperholomorphic sheaves and $C$-restricted comples structures} Usually, the moduli space of semistable bundle deformations of a bundle $F$ is not compact. To compactify this moduli space, Maruyama adds points corresponding to the deformations of $F$ which are singular (these deformations can be non-reflexive and can have singular reflexizations). Using the desingularization theorems for hyperholomorphic sheaves, we were able to obtain \ref{_reflexi_defo_loca_trivi_Theorem_}, which states (roughly speaking) that a deformation of a semistable hyperholomorphic bundle is again a semistable bunlde, assuming that all its singularities are isolated. In Section \ref{_C_restri_Section_}, we showed that under certain conditions, a deformation of a hyperholomorphic sheaf is again hyperholomorphic (\ref{_sheaf_on_C_restr_hyperho_Theorem_}). This makes it possible to prove that a deformation of a semistable hyperholomorphic bundle is locally trivial. \hfill In \cite{_Verbitsky:Hilbert_}, we have shown that a Hilbert scheme of a K3 surface has no non-trivial trianalytic subvarieties, for a general hyperk\"ahler structure. \hfill \theorem\label{_space_semista_bu_compa_Theorem_} Let $M$ be a compact hyperk\"ahler manifold without non-trivial trianalytic subvarieties, $\dim_{\Bbb H}\geq 2$, and $I$ an induced complex structure. Consider a hyperholomorphic bundle $F$ on $(M, I)$ (\ref{_hyperho_shea_Definition_}). Assume that $I$ is a $C$-restricted complex structure, $C= \deg_I c_2(F)$. Let $\c M$ be the moduli space of semistable bundle deformations of $F$ over $(M, I)$. Then $\c M$ is compact. \hfill {\bf Proof:} The complex structure $I$ is by definition algebraic, with unique polarization. This makes it possible to speak of Gieseker--Maruyama stability on $(M, I)$. Denote by $\c M_{mar}$ the Maruyama moduli of deformations of $F$. Then $\c M$ is naturally an open subset of $\c M_{mar}$. Let $s\in \c M_{mar}$ be an arbitrary point and $F_s$ the corresponding coherent sheaf on $(M, I)$, defined up to $S$-equivalence. According to \ref{_Chern_deforma_equal_}, the Chern classes of $F$ and $F_s$ are equal. Thus, by \ref{_sheaf_on_C_restr_hyperho_Theorem_}, the sheaf $F_s$ is hyperholomorphic. Therefore, $F_s$ admits a filtration with hyperholomorphic stable quotient sheaves $F_i$, $i= 1, ..., m$. By \ref{_singu_triana_Claim_}, the singular set $S$ of $F_s$ is trianalytic. Since $M$ has no proper trianalytic subvarieties, $S$ is a collection of points. We obtain that $F_s$ has isolated singularities. Let $\frak F$ be a family of deformations of $F$, parametrized by $Y$. The points $y\in Y$ correspond to deformations $F_y$ of $F_s$. Assume that for all $y\in Y$, $y\neq s$, the sheaf $F_y$ is a bundle. Since $\c M$ is open in $\c M_{mar}$, such a deformation always exists. The sheaf $F_s$ has isolated singularities and admits a filtration with hyperholomorphic stable quotient sheaves. This implies that the family $\frak F$ satisfies the conditions of \ref{_reflexi_defo_loca_trivi_Theorem_}. By \ref{_reflexi_defo_loca_trivi_Theorem_}, the reflexization $\frak F^{**}$ is locally trivial. To prove that $\c M= \c M_{mar}$, we have to show that for all $s\in \c M_{mar}$, the corresponding coherent sheaf $F_s$ is locally trivial. Therefore, to finish the proof of \ref{_space_semista_bu_compa_Theorem_}, it remains to prove the following algebro-geometric claim. \hfill \claim \label{_defo_shea-w-holes-has-holes_Claim_} Let $X$ be a compact complex manifold, $\dim_{\Bbb C} X>2$, and $\frak F$ a torsion-free coherent sheaf over $X\times Y$ which is flat over $Y$. Assume that the reflexization of $\frak F$ is locally trivial, $\frak F$ has isolated singularities, and for some point $s\in Y$, the restriction of $\frak F$ to the complement $(X\times Y)\backslash (X\times \{s\})$ is locally trivial. Then the reflexization $\left(\frak F\restrict X\times\{s\}\right)^{**}$ is locally trivial. \hfill \remark We say that a kernel of a map from a bundle to an Artinian sheaf is {\bf a bundle with holes}. In slightly more intuitive terms, \ref{_defo_shea-w-holes-has-holes_Claim_} states that a flat deformation of a bundle with holes is again a bundle with holes, and cannot be smooth, assuming that $\dim_{\Bbb C} X>2$. \hfill {\bf Proof of \ref{_defo_shea-w-holes-has-holes_Claim_}:} \ref{_defo_shea-w-holes-has-holes_Claim_} is well known. Here we give a sketch of a proof. Consider a coherent sheaf $F_s= \frak F \restrict{X\times \{s\}}$, and an exact sequence \[ 0{\:\longrightarrow\:} F_s{\:\longrightarrow\:} F_s^{**} {\:\longrightarrow\:} k{\:\longrightarrow\:} 0, \] where $k$ is an Artinian sheaf. By definition, the sheaf $F_s^{**}$ is locally trivial. The flat deformations of $F_s$ are infinitesimally classified by $Ext^1(F_s, F_s)$. Replacing $F_s$ by a quasi-isomorphic complex of sheaves $F_s^{**} {\:\longrightarrow\:} k$, we obtain a spectral sequence converging to $Ext^\bullet(F_s, F_s)$. In the $E_2$-term of this sequence, we observe the group \[ Ext^1(F_s^{**}, F_s^{**})\oplus Ext^1(k, k) \oplus Ext^2(k, F_s^{**})\oplus Ext^0(F_s^{**}, k). \] which is responsible for $Ext^1(F_s, F_s)$. The term $Ext^1(F_s^{**}, F_s^{**})$ is responsible for deformations of the bundle $F_s^{**}$, the term $Ext^0(F_s^{**}, k)$ for the deformations of the map $F_s^{**} {\:\longrightarrow\:} k$, and the term $Ext^1(k, k)$ for the deformations of the Artinian sheaf $k$. Thus, the term $Ext^2(k, F_s^{**})$ is responsible for the deformations of $F_s$ which change the dimension of the cokernel of the embedding $F_s{\:\longrightarrow\:} F_s^{**}$. We obtain that whenever $Ext^2(k, F_s^{**})=0$, all deformations of $F_s$ are singular. On the other hand, $Ext^2(k, F_s^{**})=0$, because the $i$-th $Ext$ from the skyscraper to a free sheaf on a manifold of dimension more than $i$ vanishes (this is a basic result of Grothendieck's duality, \cite{_Hartshorne:Grothendieck's_}). \blacksquare \section{New examples of hyperk\"ahler manifolds} \label{_new_exa_Section_} \subsection{Twistor paths} \label{_twi_paths_Subsection_} This Subsection contains an exposition and further elaboration of the results of \cite{_coho_announce_} concerning the twistor curves in the moduli space of complex structures on a complex manifold of hyperk\"ahler type. Let $M$ be a compact manifold admitting a hyperk\"ahler structure. In \ref{_moduli_hyperka_Definition_}, we defined the coarse, marked moduli space of complex structures on $M$, denoted by $Comp$. For the duration of this section, we fix a compact simple hyperk\"ahler manifold $M$, and its moduli $Comp$. \hfill Further on, we shall need the following fact. \hfill \claim \label{_simple=hyperho_Claim_} Let $M$ be a hyperk\"ahler manifold, $I$ an induced complex structure of general type, and $B$ a holomorphic vector bundle over $(M, I)$. Then $B$ is stable if an only if $B$ is simple.\footnote{Simple sheaves are coherent sheaves which have no proper subsheaves} {\bf Proof:} By \ref{_Lambda_of_inva_forms_zero_Lemma_}, for all $\omega \in Pic(M, I)$, we have $\deg_I(\omega)=0$. Therefore, every subsheaf of $B$ is destabilising. \blacksquare \hfill \remark In assumptions of \ref{_simple=hyperho_Claim_}, all stable bundles are hyperholomorphic (\ref{_inva_then_hyperho_Theorem_}). Therefore, \ref{_simple=hyperho_Claim_} implies that $B$ is hyperholomorphic if it is simple. \hfill In Subsection \ref{_modu_and_C-restri_Subsection_}, we have shown that every hyperk\"ahler structure $\c H$ corresponds to a holomorphic embedding \[ \kappa(\c H):\; {\Bbb C} P^1 {\:\longrightarrow\:} Comp, \ \ L {\:\longrightarrow\:} (M, L).\] \definition A projective line $C \subset Comp$ is called {\bf a twistor curve} if $C= \kappa(\c H)$ for some hyperk\"ahler structure $\c H$ on $M$. \hfill The following theorem was proven in \cite{_coho_announce_}. \hfill \theorem \label{_twistor_connect_Theorem_} (\cite{_coho_announce_}, Theorem 3.1) Let $I_1, I_2\in Comp$. Then there exist a sequence of intersecting twistor curves which connect $I_1$ with $I_2$. \blacksquare \hfill \definition Let $P_0$, ..., $P_n\subset Comp$ be a sequence of twistor curves, supplied with an intersection point $x_{i+1}\in P_i\cap P_{i+1}$ for each $i$. We say that $\gamma= P_0, ..., P_n, x_1, ..., x_n$ is a {\bf twistor path}. Let $I$, $I'\in Comp$. We say that $\gamma$ is {\bf a twistor path connecting $I$ to $I'$} if $I\in P_0$ and $I'\in P_n$. The lines $P_i$ are called {\bf the edges}, and the points $x_i$ {\bf the vertices} of a twistor path. \hfill Recall that in \ref{_generic_manifolds_Definition_}, we defined induced complex structures which are generic with respect to a hyperk\"ahler structure. \hfill Given a twistor curve $P$, the corresponding hyperk\"ahler structure $\c H$ is unique (\ref{_hyperka_etale_Theorem_}). We say that a point $x\in P$ is {\bf of general type}, or {\bf generic with respect to $P$} if the corresponding complex structure is generic with respect to $\c H$. \hfill \definition \label{_admi_twi_path_Definition_} Let $I$, $J\in Comp$ and $\gamma= P_0, ..., P_n$ be a twistor path from $I$ to $J$, which corresponds to the hyperk\"ahler structures $\c H_0$, ..., $\c H_n$. We say that $\gamma$ is {\bf admissible} if all vertices of $\gamma$ are of general type with respect to the corresponding edges. \hfill \remark In \cite{_coho_announce_}, admissible twistor paths were defined slightly differently. In addition to the conditions above, we required that $I$, $J$ are of general type with respect to $\c H_0$, $\c H_n$. \hfill \ref{_twistor_connect_Theorem_} proves that every two points $I$, $I'$ in $Comp$ are connected with a twistor path. Clearly, each twistor path induces a diffeomorphism $\mu_\gamma:\; (M,I){\:\longrightarrow\:} (M,I')$. In \cite{_coho_announce_}, Subsection 5.2, we studied algebro-geometrical properties of this diffeomorphism. \hfill \theorem \label{_admi_twi_impli_Theorem_} Let $I$, $J\in Comp$, and $\gamma$ be an admissible twistor path from $I$ to $J$. Then \begin{description} \item[(i)] There exists a natural isomorphism of tensor cetegories \[ \Phi_{\gamma}:\; Bun_I(\c H_0){\:\longrightarrow\:} Bun_J(\c H_n),\] where $Bun_I(\c H_0)$, $Bun_J(\c H_n)$ are the categories of polystable hyperholomorphic vector bundles on $(M, I)$, $(M, J)$, taken with respect to $\c H_0$, $\c H_n$ respectively. \item[(ii)] Let $B\in Bun_I(\c H_0)$ be a stable hyperholomorphic bundle, and \[ \c M_{I, \c H_0}(B)\] the moduli of stable deformations of $B$, where stability is taken with respect to the K\"ahler metric induced by $\c H_0$. Then $\Phi_{\gamma}$ maps stable bundles which are deformationally equivalent to $B$ to the stable bundles which are deformationally equivalent to $\Phi_\gamma(B)$. Moreover, obtained this way bijection \[ \Phi_\gamma:\; \c M_{I, \c H_0}(B){\:\longrightarrow\:} \c M_{J, \c H_n}(\Phi_\gamma(B))\] induces a real analytic isomorphism of deformation spaces. \end{description} {\bf Proof:} \ref{_admi_twi_impli_Theorem_} (i) is a consequence of \cite{_coho_announce_}, Corollary 5.1. Here we give a sketch of its proof. Let $I$ be an induced complex structure of general type. By \ref{_simple=hyperho_Claim_}, a bundle $B$ over $(M, I)$ is stable if and only if it is simple. Thus, the category $Bun_I(\c H)$ is independent from the choice of $\c H$ (\ref{_simple=hyperho_Claim_}). In \ref{_equi_cate_Theorem_}, we constructed the equivalence of categories $\Phi_{I, J}$, which gives the functor $\Phi_\gamma$ for twistor path which consists of a single twistor curve. This proves \ref{_admi_twi_impli_Theorem_} (i) for $n=1$. A composition of isomorphisms $\Phi_{I, J}\circ \Phi_{J, J'}$ is well defined, because the category $Bun_I(\c H)$ is independent from the choice of $\c H$. Taking successive compositions of the maps $\Phi_{I, J}$, we obtain an isomorphism $\Phi_\gamma$. This proves \ref{_admi_twi_impli_Theorem_} (i). The variety $\c M_{I, \c H}(B)$ is singular hyperk\"ahler (\cite{_Verbitsky:Hyperholo_bundles_}), and the variety $\c M_{J, \c H}(B)$ is the same singular hyperk\"ahler variety, taken with another induced complex structure. By definition of singular hyperk\"ahler varieties, this implies that $\c M_{I, \c H}(B)$, $\c M_{J, \c H}(B)$ are real analytic equivalent, with equivalence provided by $\Phi_{I, J}$. This proves \ref{_admi_twi_impli_Theorem_} (ii). \blacksquare \hfill For $I\in Comp$, denote by $Pic(M, I)$ the group $H^{1,1}(M, I)\cap H^2(M,{\Bbb Z})$, and by $Pic(I, {\Bbb Q})$ the space $H^{1,1}(M, I)\cap H^2(M, {\Bbb Q})\subset H^2(M)$. Let $Q\subset H^2(M, {\Bbb Q})$ be a subspace of $H^2(M, {\Bbb Q})$, and \[ Comp_Q:= \{ I\in Comp \;\; | \;\; Pic(I, {\Bbb Q}) =Q\}. \] \theorem\label{_admi_exi_Theorem_} Let $\c H$, $\c H'$ be hyperk\"ahler structures, and $I$, $I'$ be complex structures of general type to and induced by $\c H$, $\c H'$. Assume that $Pic(I, {\Bbb Q}) = Pic(I', {\Bbb Q}) =Q$, and $I$, $I'$ lie in the same connected component of $Comp_Q$. Then $I$, $I'$ can be connected by an admissible path. {\bf Proof:} This is \cite{_coho_announce_}, Theorem 5.2. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill For a general $Q$, we have no control over the number of connected components of $Comp_Q$ (unless global Torelli theorem is proven), and therefore we cannot directly apply \ref{_admi_exi_Theorem_} to obtain results from algebraic geometry.\footnote{Exception is a K3 surface, where Torelli holds. For K3, $Comp_Q$ is connected for all $Q\subset H^2(M, {\Bbb Q})$.} However, when $Q=0$, $Comp_Q$ is clearly connected and dense in $Comp$. This is used to prove the following corollary. \hfill \corollary \label{_I_conne_w_admi_Corollary_} Let $I$, $I'\in Comp_0$. Then $I$ can be connected to $I'$ by an admissible twistor path. {\bf Proof} This is \cite{_coho_announce_}, Corollary 5.2. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition\label{_gene_pola_Definition_} Let $I\in Comp$ be a complex structure, $\omega$ be a K\"ahler form on $(M, I)$, and $\c H$ the corresponding hyperk\"ahler metric, which exists by Calabi-Yau theorem. Then $\omega$ is called {\bf a generic polarization} if any of the following conditions hold \begin{description} \item[(i)] For all $a\in Pic(M, I)$, the degree $\deg_\omega(a)\neq 0$, unless $a=0$. \item[(ii)] For all $SU(2)$-invariant integer classes $a\in H^2(M, {\Bbb Z})$, we have $a=0$. \end{description} The conditions (i) and (ii) are equivalent by \ref{_Lambda_of_inva_forms_zero_Lemma_}. \hfill \claim\label{_omega_gene_otho_Claim_} Let $I\in Comp$ be a complex structure, $\omega$ be a K\"ahler form on $(M, I)$, and $\c H$ the corresponding hyperk\"ahler structure, which exists by Calabi-Yau theorem. Then $\omega$ is generic if and only if for all integer classes $a\in H^{1,1}(M, I)$, the class $a$ is not orthogonal to $\omega$ with respect to the Bogomolov-Beauville pairing. {\bf Proof:} Clearly, the map $\deg_\omega:\; H^2(M) {\:\longrightarrow\:} {\Bbb R}$ is equal (up to a scalar multiplier) to the orthogonal projection onto the line ${\Bbb R}\cdot\omega$. Then, \ref{_omega_gene_otho_Claim_} is equivalent to \ref{_gene_pola_Definition_}, (i). \blacksquare \hfill From \ref{_omega_gene_otho_Claim_} it is clear that the set of generic polarizations is a complement to a countable union of hyperplanes. Thus, generic polarizations are dense in the K\"ahler cone of $(M, I)$, for all $I$. \hfill \claim \label{_admi_pa_exist_for_gene_pol_Claim_} Let $I, J\in Comp$, and $a$, $b$ be generic polarizations on $(M, I)$. Consider the corresponding hyperk\"ahler structures $\c H_0$ and $\c H_n$ inducing $I$ and $J$. Then there exists an admissible twistor path starting from $I, \c H_0$ and ending with $\c H_n, J$. \hfill {\bf Proof:} Consider the twistor curves $P_0$, $P_n$ corresponding to $\c H_0$, $\c H_n$. Since $a$, $b$ are generic, the curves $P_0$, $P_n$ intersect with $Comp_0$. Applying \ref{_I_conne_w_admi_Corollary_}, we connect the curves $P_0$ and $P_n$ by an admissible path. \blacksquare \hfill Putting together \ref{_admi_pa_exist_for_gene_pol_Claim_} and \ref{_admi_twi_impli_Theorem_}, we obtain the following result. \hfill \theorem\label{_iso_Bun_exists_gene_pola_Theorem_} Let $I$, $J\in Comp$ be complex structures, and $a, b$ be generic polarizations on $(M, I)$, $(M, J)$. Then \begin{description} \item[(i)] There exist an isomorphism of tensor cetegories \[ \Phi_{\gamma}:\; Bun_I(a){\:\longrightarrow\:} Bun_J(a),\] where $Bun_I(a)$, $Bun_J(b)$ are the categories of polystable hyperholomorphic vector bundles on $(M, I)$, $(M, J)$, taken with respect to the hyperk\"ahler structures defined by the K\"ahler classes $a$, $b$ as in \ref{_symplectic_=>_hyperkahler_Proposition_}. \item[(ii)] Let $B\in Bun_I(a)$ be a stable hyperholomorphic bundle, and \[ \c M_{I, a}(B)\] the moduli of stable deformations of $B$, where stability is taken with respect to the polarization $a$. Then $\Phi_{\gamma}$ maps stable bundles which are deformationally equivalent to $B$ to the stable bundles which are deformationally equivalent to $\Phi_\gamma(B)$. Moreover, obtained this way bijection \[ \Phi_\gamma:\; \c M_{I, a}(B){\:\longrightarrow\:} \c M_{J, b}(\Phi_\gamma(B))\] induces a real analytic isomorphism of deformation spaces. \end{description} \blacksquare \hfill \lemma \label{_Phi_of_tange_Lemma_} In assumptions of \ref{_admi_twi_impli_Theorem_}, let $B$ be a holomorphic tangent bundle of $(M, I)$. Then $\Phi_\gamma(B)$ is a holomorphic tangent bundle of $(M, J)$. {\bf Proof:} Clear. \blacksquare \hfill \corollary \label{_mod_of_tange_compa_Corollary_} Let $I, J\in Comp$ be complex structures, and $a$, $b$ generic polarizations on $(M, I)$, $(M, J)$. Assume that the moduli of stable deformations $\c M_{I, a}(T(M, I))$ of the holomorphic tangent bundle $T^{1,0}(M, I)$ is compact. Then the space $\c M_{J, b}(T(M, J))$ is also compact. {\bf Proof:} Let $\gamma$ be the twistor path of \ref{_admi_pa_exist_for_gene_pol_Claim_}. By \ref{_Phi_of_tange_Lemma_}, $\Phi_\gamma(T(M, I)) = T(M, J)$. Applying \ref{_admi_twi_impli_Theorem_}, we obtain a real analytic equivalence from $\c M_{I, a}(T(M, I))$ to $\c M_{J, b}(T(M, J))$. \blacksquare \subsection{New examples of hyperk\"ahler manifolds} \theorem\label{_space_sta_bu_compa_hyperka_Theorem_} Let $M$ be a compact hyperk\"ahler manifold without non-trivial trianalytic subvarieties, $\dim_{\Bbb H}M\geq 2$, and $I$ an induced complex structure. Consider a hyperholomorphic bundle $F$ on $M$ (\ref{_hyperho_shea_on_M_Definition_}). Let $F_I$ be the corresponding holomorphic bundle over $(M, I)$. Assume that $I$ is a $C$-restricted complex structure, $C= \deg_I c_2(F)$. Assume, moreover, that all semistable bundle deformations of $F_I$ are stable.\footnote{This may happen, for instance, when $\operatorname{rk} F= \dim_{\Bbb C} M=n$, and the number $c_n(F)$ is prime.} Denote by $\c M_F^I$ the moduli of stable bundle deformations of $F_I$ over $(M, I)$. Then \begin{description} \item[(i)] the normalization $\tilde{\c M}_F^I$ is a compact and smooth complex manifold equipped with a natural hyperk\"ahler structure. \item[(ii)] Moreover, for all induced complex structures $J$ on $M$, the the variety $\c M_F^J$ is compact, and has a smooth normalization $\tilde{\c M}_F^J$, which is also equipped with a natural hyperk\"ahler structure. \item[(iii)] Finally, the hyperk\"ahler manifolds $\tilde{\c M}_F^J$, $\tilde{\c M}_F^I$ are naturally isomorphic. \end{description} {\bf Proof:} The variety $\c M_F^I$ is compact by \ref{_space_semista_bu_compa_Theorem_}. In \cite{_Verbitsky:Hyperholo_bundles_}, it was proven that the space $\c M_F^I$ of stable deformations of $F$ is a singular hyperk\"ahler variety (see also \cite{_NHYM_} for an explicit construction of the twistor space of $\c M_F^I$). Then \ref{_space_sta_bu_compa_hyperka_Theorem_} is a consequence of the Desingularization Theorem for singular hyperk\"ahler varietiess (\ref{_desingu_Theorem_}). \blacksquare \hfill The assumptions of \ref{_space_sta_bu_compa_hyperka_Theorem_} are quite restrictive. Using the technique of twistor paths, developed in Subsection \ref{_twi_paths_Subsection_}, it is possible to prove a more accessible form of \ref{_space_sta_bu_compa_hyperka_Theorem_}. \hfill Let $M$ be a hyperk\"ahler manifold, and $I$, $J$ induced complex structures. Given an admissible twistor path from $I$ to $J$, we obtain an equivalence $\Phi_\gamma$ between the category of hyperholomorphic bundles on $(M, I)$ and $(M, J)$. \hfill \theorem \label{_twi_pa_space_sta_compa_Theorem_} Let $M$ be a compact simple hyperk\"ahler manifold, $\dim_{\Bbb H}M >1$, and $I$ a complex structure on $M$. Consider a generic polarization $a$ on $(M, I)$. Let $\c H$ be the corresponding hyperk\"ahler structure, and $F$ a hyperholomorphic bundle on $(M, I)$. Fix a hyperk\"ahler structure $\c H'$ on $M$ admitting $C$-restricted complex structures, such that $M$ has no trianalytic subvarieties with respect to $\c H'$. Assume that for some $C$-restricted complex structure $J$ induced by $\c H'$, $C=\deg_I c_2(F)$, all admissible twistor paths $\gamma$ from $I$ to $J$, and all semistable bundles $F'$ which are deformationally equivalent to $\Phi_\gamma(F)$, the bundle $F'$ is stable. Then the space of stable deformations of $F$ is compact. \hfill \remark The space of stable deformations of $F$ is singular hyperk\"ahler (\cite{_Verbitsky:Hyperholo_bundles_}) and its normalization is smooth and hyperk\"ahler (\ref{_desingu_Theorem_}). \hfill {\bf Proof of \ref{_twi_pa_space_sta_compa_Theorem_}:} Clearly, $F'$ satisfies assumptions of \ref{_space_sta_bu_compa_hyperka_Theorem_}, and the moduli space of its stable deformations is compact. Since $\Phi_\gamma$ induces a homeomorphism of moduli spaces (\ref{_admi_twi_impli_Theorem_}), the space of stable deformations of $F$ is also compact. \blacksquare \hfill Applying \ref{_twi_pa_space_sta_compa_Theorem_} to the holomorphic tangent bundle $T(M, I)$, we obtain the following corollary. \hfill \theorem \label{_defo_tange_compact_Theorem_} Let $M$ be a compact simple hyperk\"ahler manifold, $\dim_{\Bbb H}(M)>1$. Assume that for a generic hyperk\"ahler structure $\c H$ on $M$, this manifold admits no trianalytic subvarieties.\footnote{This assumption holds for a Hilbert scheme of points on a K3 surface.} Assume, moreover, that for some $C$-restricted induced complex structure $I$, all semistable bundle deformations of $T(M, I)$ are stable, for $C> \deg_I c_2(M)$. Then, for all complex structures $J$ on $M$ and all generic polarizations $\omega$ on $(M, J)$, the deformation space $\c M_{J, \omega}(T(M, J))$ is compact. {\bf Proof:} Follows from \ref{_twi_pa_space_sta_compa_Theorem_} and \ref{_mod_of_tange_compa_Corollary_}. \blacksquare \subsection{How to check that we obtained new examples of hyperk\"ahler manifolds?} \label{_new_exa_F-M_checking_Subsection_} A. Beauville \cite{_Beauville_} described two families of compact hyperk\"ahler manifolds, one obtained as the Hilbert scheme of points on a K3-surface, another obtained as the Hilbert scheme of a 2-dimensional torus factorized by the free torus action. \hfill \conjecture\label{_anti_Beauville_Conjecture} There exist compact simple hyperk\"ahler manifolds which are not isomorphic to deformations of these two fundamental examples. \hfill Here we explain our strategy of a proof of \ref{_anti_Beauville_Conjecture} using results on compactness of the moduli space of hyperholomorphic bundles. \hfill The results of this subsection are still in writing, so all statements below this line should be considered as conjectures. We give an idea of a proof for each result and label it as ``proof'', but these ``proofs'' are merely sketches. \hfill First of all, it is possible to prove the following theorem. \hfill \theorem \label{_no_rk-2-bu_on_Hilb_Theorem_} Let $M$ be a complex K3 surface without automorphisms. Assume that $M$ is Mumford-Tate generic with respect to some hyperk\"ahler structure. Consider the Hilbert scheme $M^{[n]}$ of points on $M$, $n>1$. Pick a hyperk\"ahler structure $\c H$ on $M^{[n]}$ which is compatible with the complex structure. Let $B$ be a hyperholomorphic bundle on $(M^{[n]}, \c H)$, $\operatorname{rk} B=2$. Then $B$ is a trivial bundle. {\bf Proof:} The proof of \ref{_no_rk-2-bu_on_Hilb_Theorem_} is based on the same ideas as the proof of \ref{_no_triana_subva_of_Hilb_Theorem_}. \blacksquare \hfill For a compact complex manifold $X$ of hyperk\"ahler type, denote its coarse, marked moduli space (\ref{_moduli_hyperka_Definition_}) by $Comp(X)$. \hfill \corollary\label{_no_rk-2-bu_on_def_Theorem_} Let $M$ be a K3 surface, $I\in Comp(X)$ an arbitrary complex structure on $X = M^{[n]}$, $n>1$, and $a$ a generic polarization on $(X, J)$. Consider the hyperk\"ahler structure $\c H$ which corresponds to ($I$, $a$) as in \ref{_symplectic_=>_hyperkahler_Proposition_}. Let $B$, $\operatorname{rk} B=2$ be a hyperholomorphic bundle over $(X, \c H)$. Then $B$ is trivial. {\bf Proof:} Follows from \ref{_no_rk-2-bu_on_Hilb_Theorem_} and \ref{_iso_Bun_exists_gene_pola_Theorem_}. \blacksquare \hfill \corollary\label{_defo_4-dim_bu_Corollary_} Let $M$ be a K3 surface, $I\in Comp(X)$ an arbitrary complex structure on $X = M^{[n]}$, $n>1$, and $a$ a generic polarization on $(X, I)$. Consider the hyperk\"ahler structure $\c H$ which corresponds to ($J$, $a$) (\ref{_symplectic_=>_hyperkahler_Proposition_}). Let $B$, $\operatorname{rk} B\leq 6$ be a stable hyperholomorphic bundle on $(X, \c H)$. Assume that the Chern class $c_{\operatorname{rk} B}(B)$ is non-zero. Assume, moreover, that $I$ is $C$-restricted, $C = \deg_I(c_2(B))$. Let $B'$ be a semistable deformation of $B$ over $(X, I)$. Then $B'$ is stable. \hfill {\bf Proof:} Consider the Jordan--H\"older serie for $B'$. Let $Q_1 \oplus Q_2 \oplus ...$ be the associated graded sheaf. By \ref{_sheaf_on_C_restr_hyperho_Theorem_}, the stable bundles $Q_i$ are hyperholomorphic. Since $c_{\operatorname{rk} B}(B)\neq 0$, we have $c_{\operatorname{rk} Q_i}(Q_i)\neq 0$. Therefore, the bundles $Q_i$ are non-trivial. By \ref{_no_rk-2-bu_on_def_Theorem_}, $\operatorname{rk} Q_i >2$. Since all the Chern classes of the bundles $Q_i$ are $SU(2)$-invariant, the odd Chern classes of $Q_i$ vanish (\ref{_SU(2)_inva_type_p,p_Lemma_}). Therefore, $\operatorname{rk} Q_i\geq 4$ for all $i$. Since $\operatorname{rk} B\leq 6$, we have $i=1$ and the bundle $B'$ is stable. \blacksquare \hfill Let $M$ be a K3 surface, $X= M^{[i]}$, $i= 2$, $3$ be its second or third Hilbert scheme of points, $I\in Comp(X)$ arbitrary complex structure on $X$, and $a$ a generic polarization on $(X, I)$. Consider the hyperk\"ahler structure $\c H$ which corresponds to $J$ and $a$ by Calabi-Yau theorem (\ref{_symplectic_=>_hyperkahler_Proposition_}). Denote by $TX$ the tangent bundle of $X$, considered as a hyperholomorphic bundle. Let $\operatorname{Def}(TX)$ denote the hyperk\"ahler desingularization of the moduli of stable deformations of $TX$. By \ref{_iso_Bun_exists_gene_pola_Theorem_}, the real analytic subvariety underlying $\operatorname{Def}(TX)$ is independent from the choice of $I$. Therefore, its dimension is also independent from the choice of $I$. The dimension of the deformation space $\operatorname{Def}(TX)$ can be estimated by a direct computation, for $X$ a Hilbert scheme. We obtain that $\dim \operatorname{Def}(TX)> 40$. \hfill \claim \label{_Def_TX_compa_} In these assumptions, the space $\operatorname{Def}(TX)$ is a compact hyperk\"ahler manifold. {\bf Proof:} By \ref{_defo_4-dim_bu_Corollary_}, all semistable bundle deformations of $TX$ are stable. Then \ref{_Def_TX_compa_} is implied by \ref{_defo_tange_compact_Theorem_}. \blacksquare \hfill Clearly, deforming the complex structure on $X$, we obtain a deformation of complex structures on $\operatorname{Def}(TX)$. This gives a map \begin{equation} \label{_Comp_X_to_Comp_Def_Equation_} Comp(X) {\:\longrightarrow\:} Comp(\operatorname{Def}(TX)). \end{equation} It is easy to check that the map \eqref{_Comp_X_to_Comp_Def_Equation_} is complex analytic, and maps twistor curves to twistor curves. \hfill \claim \label{_maps_pre_tw_curves_Claim_} Let $X$, $Y$ be hyperk\"ahler manifolds, and \[ \phi:\; Comp(X) {\:\longrightarrow\:} Comp(Y)\] be a holomorphic map of corresponding moduli spaces which maps twistor curves to twistor curves. Then $\phi$ is locally an embedding. {\bf Proof:} An elementary argument using the period maps, in the spirit of Subsection \ref{_modu_and_C-restri_Subsection_}. \blacksquare \hfill The following result, along with \ref{_no_rk-2-bu_on_Hilb_Theorem_}, is the major stumbling block on the way to proving \ref{_anti_Beauville_Conjecture}. The other results of this Subsection are elementary or routinely proven, but the complete proof of \ref{_no_rk-2-bu_on_Hilb_Theorem_} and \ref{_defo_simple_Theorem_} seems to be difficult. \hfill \theorem \label{_defo_simple_Theorem_} Let $X$ be a simple hyperk\"ahler manifold without proper trianalytic subvarieties, $B$ a hyperholomorphic bundle over $X$, and $I$ an induced complex structure. Denote the corresponding holomorphic bundle over $(X, I)$ by $B_I$. Assume that the space $\c M$ of stable bundle deformations of $B$ is compact. Let $\operatorname{Def}(B)$ be the hyperk\"ahler desingularization of $\c M$. Then $\operatorname{Def}(M)$ is a simple hyperk\"ahler manifold. \hfill {\bf Proof:} Given a decomposition $\operatorname{Def}(M) = M_1\times M_2$, we obtain a parallel 2-form on $\Omega_1$ on $\operatorname{Def}(B)$, which is a pullback of the holomorphic symplectic form on $M_1$. Consider the space $\c A$ of connections on $B$, which is an infinitely-dimensional complex analytic Banach manifold. Then $\Omega_1$ corresponds to a holomorphic 2-form $\tilde \Omega_1$ on $\c A$. Since $\Omega_1$ is parallel with respect to the natural connection on $\operatorname{Def}(B)$, the form $\tilde \Omega_1$ is also a parallel 2-form on the tangent space to $\c A$, which is identified with $\Omega^1(X, \operatorname{End}(B))$. It is possible to prove that this 2-form is obtained as \[ A, B {\:\longrightarrow\:} \int_{Y} \Theta\left(A\restrict Y, B\restrict Y\right)\operatorname{Vol}(Y) \] where \[ \Theta:\;\Omega^1(Y, \operatorname{End}(B)) \times \Omega^1(Y, \operatorname{End}(B)){\:\longrightarrow\:} {\cal O}_Y \] is a certain holomorphic pairing on the bundle $\Omega^1(Y, \operatorname{End}(B))$, and $Y$ is a trianalytic subvariety of $X$. Since $X$ has no trianalytic subvarieties, $\tilde \Omega_1$ is obtained from a ${\cal O}_X$-linear pairing \[ \Omega^1(X, \operatorname{End}(B))\times \Omega^1(X, \operatorname{End}(B)) {\:\longrightarrow\:} {\cal O}_X. \] Using stability of $B$, it is possible to show that such a pairing is unique, and thus, $\Omega_1$ coincides with the holomorphic symplectic form on $\operatorname{Def}(B)$. Therefore, $\operatorname{Def}(B) = M_1$, and this manifold is simple. \blacksquare \hfill Return to the deformations of tangent bundles on $X= M^{[i]}$, $i=2,3$. Recall that the second Betti number of a Hilbert scheme of points on a K3 surface is equal to $23$, and that of the generalized Kummer variety is 7 (\cite{_Beauville_}). Consider the map \eqref{_Comp_X_to_Comp_Def_Equation_}. By \ref{_defo_simple_Theorem_}, the manifold $\operatorname{Def}(TX)$ is simple. By Bogomolov's theorem (\ref{_Bogomo_etale_Theorem_}), we have \[ \dim Comp(\operatorname{Def}(TX)) = \dim H^2(\operatorname{Def}(TX)) -2.\] Therefore, either $\dim H^2(\operatorname{Def}(TX))> \dim H^2(X)=23$, or the map \eqref{_Comp_X_to_Comp_Def_Equation_} is etale. In the first case, the second Betti number of $\operatorname{Def}(TX)$ is bigger than that of known simple hyperk\"ahler manifolds, and thus, $\operatorname{Def}(TX)$ is a new example of a simple hyperk\"ahler manifold; this proves \ref{_anti_Beauville_Conjecture}. Therefore, to prove \ref{_anti_Beauville_Conjecture}, we may assume that $\dim H^2(\operatorname{Def}(TX))=23$, the map \eqref{_Comp_X_to_Comp_Def_Equation_} is etale, and $\operatorname{Def}(TX)$ is a deformation of a Hilbert scheme of points on a K3 surface. \hfill Consider the universal bundle $\tilde B$ over $X\times \operatorname{Def}(TX)$. Restricting $\tilde B$ to $\{x\} \times \operatorname{Def}(TX)$, we obtain a bundle $B$ on $\operatorname{Def}(TX)$. Let $\operatorname{Def}(B)$ be the hyperk\"ahler desingularization of the moduli space of stable deformations of $B$. Clearly, the manifold $\operatorname{Def}(B)$ is independent from the choice of $x\in X$. Taking the generic hyperk\"ahler structure on $X$, we may assume that the hyperk\"ahler structure $\c H$ on $\operatorname{Def}(TX)$ is also generic. Thus, $(\operatorname{Def}(TX),\c H)$ admits $C$-restricted complex structures and has no trianalytic subvarieties. In this situation, \ref{_defo_4-dim_bu_Corollary_} implies that the hyperk\"ahler manifold $\operatorname{Def}(B)$ is compact. Applying \ref{_maps_pre_tw_curves_Claim_} again, we obtain a sequence of maps \[ Comp(X) {\:\longrightarrow\:} Comp(\operatorname{Def}(TX)){\:\longrightarrow\:} Comp(\operatorname{Def}(B)) \] which are locally closed embeddings. By the same argument as above, we may assume that the composition $Comp(X) {\:\longrightarrow\:} Comp(\operatorname{Def}(B))$ is etale, and the manifold $\operatorname{Def}(B)$ is a deformation of a Hilbert scheme of points on K3. Using Mukai's version of Fourier transform (\cite{_Orlov:K3_}, \cite{_BBH-R_}), we obtain an embedding of the corresponding derived categories of coherent sheaves, \[ D(X) {\:\longrightarrow\:} D(\operatorname{Def}(TX)){\:\longrightarrow\:} D(\operatorname{Def}(B)). \] Using this approach, it is easy to prove that \[ \dim X\leq \dim \operatorname{Def}(TX)\leq \dim \operatorname{Def}(B). \] Let $x\in X$ be an arbitrary point. Consider the complex $C_x \in D(\operatorname{Def}(B))$ of coherent sheaves on $\operatorname{Def}(B)$, obtained as a composition of the Fourier-Mukai transform maps. It is easy to check that the lowest non-trivial cohomology sheaf of $C_x$ is a skyscraper sheaf in a point $F(x)\in \operatorname{Def}(B)$. This gives an embedding \[ F:\; X{\:\longrightarrow\:} \operatorname{Def}(B). \] The map $F$ is complex analytic for all induced complex structure. We obtained the following result. \hfill \lemma \label{_double_Fou_embedding_Lemma_} In the above assumptions, the embedding \[ F:\; X{\:\longrightarrow\:} \operatorname{Def}(B) \] is compatible with the hyperk\"ahler structure. \blacksquare \hfill By \ref{_double_Fou_embedding_Lemma_}, the manifold $\operatorname{Def}(B)$ has a trianalytic subvariety $F(X)$, of dimension $0<\dim F(X)< 40< \dim \operatorname{Def}(B)$. On the other hand, for a hyperk\"ahler structure on $X$ generic, the corresponding hyperk\"ahler structure on $\operatorname{Def}(B)$ is also generic, so this manifold has no trianalytic subvarieties. We obtained a contradiction. Therefore, either $\operatorname{Def}(TX)$ or $\operatorname{Def}(B)$ is a new example of a simple hyperk\"ahler manifold. This proves \ref{_anti_Beauville_Conjecture}. \hfill \hfill {\bf Acknowledegments:} I am grateful to V. Batyrev, A. Beilinson, P. Deligne, D. Gaitsgory, D. Kaledin, D. Kazhdan, M. Koncevich and T. Pantev for valuable discussions. My gratitude to D. Kaledin, who explained to me the results of \cite{_Swann_}. This paper uses many ideas of our joint work on direct and inverse twistor transform (\cite{_NHYM_}). {\small
1999-03-08T22:37:57
9712
alg-geom/9712011
en
https://arxiv.org/abs/alg-geom/9712011
[ "alg-geom", "dg-ga", "hep-th", "math.AG", "math.DG" ]
alg-geom/9712011
Bong Lian
B. Lian, K. Liu, and S.T. Yau
Mirror Principle I
Typos corrected, Plain Tex 50 pages with t.o.c option
null
null
null
null
We propose and study the following Mirror Principle: certain sequences of multiplicative equivariant characteristic classes on Kontsevich's stable map moduli spaces can be computed in terms of certain hypergeometric type classes. As applications, we compute the equivariant Euler classes of obstruction bundles induced by any concavex bundles -- including any direct sum of line bundles -- on $\P^n$. This includes proving the formula of Candelas-de la Ossa-Green-Parkes hence completing the program of Candelas et al, Kontesevich, Manin, and Givental, to compute rigorously the instanton prepotential function for the quintic in $\P^4$. We derive, among many other examples, the multiple cover formula for Gromov-Witten invariants of $\P^1$, computed earlier by Morrison-Aspinwall and by Manin in different approaches. We also prove a formula for enumerating Euler classes which arise in the so-called local mirror symmetry for some noncompact Calabi-Yau manifolds. At the end we interprete an infinite dimensional transformation group, called the mirror group, acting on Euler data, as a certain duality group of the linear sigma model.
[ { "version": "v1", "created": "Thu, 11 Dec 1997 15:15:07 GMT" } ]
2009-09-25T00:00:00
[ [ "Lian", "B.", "" ], [ "Liu", "K.", "" ], [ "Yau", "S. T.", "" ] ]
alg-geom
1997-12-17T13:14:09
9712
alg-geom/9712019
en
https://arxiv.org/abs/alg-geom/9712019
[ "alg-geom", "math.AG" ]
alg-geom/9712019
Thomas Bauer
Thomas Bauer
On the cone of curves of an abelian variety
null
null
null
null
null
Let $X$ be a smooth projective variety over the complex numbers. One knows by the Cone Theorem that the closed cone of curves of $X$ is rational polyhedral whenever $c_1(X)$ is ample. For varieties $X$ such that $c_1(X)$ is not ample, however, it is in general difficult to determine the structure of $\bar NE(X)$. The purpose of this paper is to study the cone of curves of abelian varieties. Specifically, the abelian varieties $X$ are determined such that the closed cone $\bar NE(X)$ is rational polyhedral. The result can also be formulated in terms of the nef cone of $X$ or in terms of the semi-group of effective classes in the N\'eron-Severi group of $X$.
[ { "version": "v1", "created": "Wed, 17 Dec 1997 12:14:09 GMT" } ]
2007-05-23T00:00:00
[ [ "Bauer", "Thomas", "" ] ]
alg-geom
\section{Introduction} Let $X$ be a smooth projective variety over the complex numbers and let $N_1(X)$ be the real vector space $$ N_1(X)=_{\operatorname{def}}\{\mbox{1-cycles on $X$ modulo numerical equivalence}\}\otimes{\Bbb R} \ . $$ As usual denote by $N\hskip-0.25em E(X)$ the {\em cone of curves} on $X$, i.e.\ the convex cone in $N_1(X)$ generated by the effective 1-cycles. The {\em closed cone of curves} ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$ is the closure of $N\hskip-0.25em E(X)$ in $N_1(X)$. One knows by the Cone Theorem \cite{Mor82} that it is rational polyhedral whenever $c_1(X)$ is ample. For varieties $X$ such that $c_1(X)$ is not ample, however, it is in general difficult to determine the structure of ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$, since it may depend in a subtle way on the geometry of $X$ (cf.\ \cite[\S4]{CKM}). This becomes already apparent in the surface case, as work of Kov\'acs on K3 surfaces shows (see \cite{Kov94}). The purpose of this paper is to study the cone of curves of abelian varieties. Specifically, we focus on the problem of determining the abelian varieties $X$ such that the closed cone ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$ is rational polyhedral. Attacking the question from the dual point of view, one is lead to consider the nef cone of $\operatorname{Nef}(X)$ or the semi-group $\NP(X)$ of homology classes of effective line bundles, i.e.\ the subset $$ \NP(X)=_{\operatorname{def}}\{\lambda\inN\hskip-0.2em S(X)\mid\lambda=c_1(L)\mbox{ for some }L\in\operatorname{Pic}(X)\mbox{ with }h^0(X,L)> 0\} $$ of the N\'eron-Severi group of $X$. In fact, Rosoff has studied this semi-group in \cite{Ros81}, where he gives examples of abelian varieties for which $\NP(X)$ is finitely generated, as well as examples where finite generation fails. He shows: \begin{items} \item[(1)] If $X$ is a singular abelian variety, i.e.\ if $\rkN\hskip-0.2em S(X)=(\dim X)^2$, and if $\dim X\ge 2$, then $\NP(X)$ is not finitely generated. \item[(2)] For elliptic curves $E_1$ and $E_2$, $\NS^+(E_1\times E_2)$ is finitely generated if and only if $\rkN\hskip-0.2em S(E_1\times E_2)=2$. \end{items} Considering these examples it is natural to ask if the abelian varieties $X$ for which $\NP(X)$ is finitely generated or, equivalently, for which ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$ is rational polyhedral, can be characterized in a simple way. Our main result shows that this is in fact the case: \pagebreak \begin{varthm*}{Theorem} Let $X$ be an abelian variety over the field of complex numbers. Then the following conditions are equivalent: \begin{items} \item[(ia)] The closed cone of curves ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$ is rational polyhedral. \item[(ib)] The nef cone $\operatorname{Nef}(X)$ is rational polyhedral. \item[(ic)] The semi-group $\NP(X)$ is finitely generated. \item[(ii)] $X$ is isogenous to a product $$ X_1\times\dots\times X_r $$ of mutually non-isogenous abelian varieties $X_i$ with $N\hskip-0.2em S(X_i)\cong{\Bbb Z}$ for $1\le i\le r$. \end{items} \end{varthm*} Note that, since on abelian varieties the nef cone coincides with the effective cone, the equivalence of (ia), (ib) and (ic) follows from elementary properties of cones and is stated here merely for the sake of completeness (see Sect.\ \ref{sect proof}). Observe that the theorem of course contains statement (2) above, while statement (1) follows from the theorem plus the fact that by \cite{ShiMit74} a singular abelian variety is isogenous to a product $E^n$ for some elliptic curve $E$. \begin{varthm*}{\it Notation and Conventions} \rm We work throughout over the field ${\Bbb C}$ of complex numbers. We will always use additive notation for the tensor product of line bundles, since this is more convenient for our purposes (for example when working with ${\Bbb Q}$- or ${\Bbb R}$-line bundles). Numerical equivalence of divisors or line bundles, which for abelian varieties coincides with algebraic equivalence, will be denoted by $\equiv$. If $X$ is an abelian variety and $L$ is a line bundle on $X$, then $\phi_L$ denotes the homomorphism $X\longrightarrow\widehat X$, $x\mapstochar\longrightarrow t_x^*L - L$, where $t_x$ is the translation map $y\mapstochar\longrightarrow x+y$ and $\widehat X=\operatorname{Pic}^0(X)$ is the dual abelian variety. Recall that $\phi_L$ depends only on the algebraic equivalence class of $L$. \end{varthm*} \section{Effective classes on simple abelian varieties} In this section we consider the semi-group of effective divisor classes on simple abelian varieties. We start by stating alternative characterizations of $\NP(X)$ which we will use in the sequel. While these are at least implicitly well-known, we include a proof for the convenience of the reader. \begin{lemma}\label{nef and effective} Let $X$ be an abelian variety of dimension $n$ and let $A$ be an ample line bundle on $X$. Then the following conditions on a line bundle $L\in\operatorname{Pic}(X)$ are equivalent: \begin{items} \item[(i)] $L$ is algebraically equivalent to some effective line bundle. \item[(ii)] $L$ is nef. \item[(iii)] $L^iA^{n-i}\ge 0$ for $1\le i\le n$. \end{items} \end{lemma} \startproof{\it Proof. } Condition (i) certainly implies (ii), since if $L\equiv L'$ for some effective line bundle $L'$, then a suitable translate of an effective divisor in $|L'|$ will intersect any given curve properly. The implication (ii) $\Rightarrow$ (iii) is clear, since an intersection product of nef line bundles is non-negative. For (iii) $\Rightarrow$ (ii) it is enough to show that the line bundle $A+mL$ is ample for all $m\ge 0$. But this follows from $$ (A+mL)^i A^{n-i}=A^n+\sum_{k=0}^{i-1} {i \choose k} A^{n+k-i}L^{i-k} m^{i-k} > 0 \ , $$ and the version of the Nakai-Moishezon Criterion given in \cite[Corollary 4.3.3]{LB}. Finally, for the implication (ii) $\Rightarrow$ (i), suppose that $L$ is nef. Then $A+mL$ is ample for all $m\ge 0$, so that the first Chern class of $L$, viewed as a hermitian form on $T_0X$, cannot have negative eigenvalues. But this implies that there is a line bundle $P\in\operatorname{Pic}^0(X)$ such that $L+P$ descends to an ample line bundle on a quotient of $X$ and is therefore effective (cf.\ \cite[Sect.\ 3.3]{LB} and \cite[p.\ 95]{Mum70}). \nopagebreak\hspace*{\fill}\qedsymbol\par\addvspace{\bigskipamount} We show next that on a simple abelian variety the existence of two algebraically independent line bundles already prevents $\NP(X)$ from being finitely generated: \begin{proposition}\label{simple case} Let $X$ be a simple abelian variety such that $\NP(X)$ is finitely generated. Then $N\hskip-0.2em S(X)\cong{\Bbb Z}$. \end{proposition} \startproof{\it Proof. } Assume to the contrary that $\rkN\hskip-0.2em S(X)>1$ and choose ample line bundles $L_1$ and $L_2$ whose classes are not proportional in $\NS_{\bbQ}(X)$. Consider then the positive real number $$ s =_{\operatorname{def}}\inf\Big\{t\in{\Bbb R}\ \Big\vert\ tL_1-L_2\mbox{ is nef }\Big\} \ . $$ Here $tL_1-L_2$ is considered as an ${\Bbb R}$-line bundle and nefness means that $tL_1C\ge L_2C$ for every irreducible curve $C$ in $X$. We assert that \begin{equation}\label{s rational} s \not\in{\Bbb Q} \ . \end{equation} Suppose to the contrary that $s$ is rational and consider the line bundle $$ L=sL_1-L_2 \in\operatorname{Pic}_{{\Bbb Q}}(X) \ . $$ We choose an integer $n$ such that $ns\in{\Bbb Z}$. The line bundle $nL$ is then algebraically equivalent to an effective (and integral) line bundle. But $L$, and hence $nL$, is certainly not ample, so that the kernel $K(nL)$ of $\phi_{nL}$ is of positive dimension. On the other hand, since $L_1$ and $L_2$ are not proportional, $nL$ is not algebraically equivalent to $0$, and hence $K(nL)$ cannot be the whole of $X$. So we find that the neutral component of $K(nL)$ is a non-trivial abelian subvariety of $X$, contradicting the simplicity assumption on $X$. This establishes the assertion \eqnref{s rational} One checks next that, since $\NP(X)$ is finitely generated, its intersection with ${\Bbb Z}[L_1]\oplus{\Bbb Z}[L_2]$ is finitely generated as well (cf.\ for example \cite[Sect.\ 1.3]{Zie95}). Choose generators $\liste N1k$ for this intersection. Let now $0<\varepsilon\ll 1$ and fix large integers $p_1,p_2$ such that \begin{equation}\label{s inequality} s < \frac{p_1}{p_2} < s+\varepsilon \ . \end{equation} The line bundle $$ A=p_1L_1-p_2L_2=p_2L+p_2\(\frac{p_1}{p_2}-s\)L_1 $$ is then ample and therefore effective, so that we have $A\equiv\sum_{i=1}^k \ell_i N_i$ with integers $\ell_i\ge 0$. Thus, writing $N_i\equiv a_iL_1-b_iL_2$ with $a_i,b_i\ge 0$, we get $$ \frac{p_1}{p_2} = \frac{\sum_{i=1}^k \ell_i a_i}{\sum_{i=1}^k \ell_i b_i} \ , $$ which, upon letting $$ q =_{\operatorname{def}}\min\left\{\frac{a_i}{b_i}\ \Big\vert\ 1\le i \le k\right\} \ , $$ yields the lower bound $$ \frac{p_1}{p_2} \ge q \ . $$ But, due to the fact that $s$ is irrational, which implies $q>s$, and since $\varepsilon$ can be taken arbitrarily small, this is incompatible with \eqnref{s inequality} \nopagebreak\hspace*{\fill}\qedsymbol\par\addvspace{\bigskipamount} \section{Classes on products} We study in this section effective divisor classes on the self-product $X\times X$ of an abelian variety $X$. Suppose for a moment that $X$ is an elliptic curve. Then, since $N\hskip-0.2em S(X\times X)$ is of rank $\ge 3$, statement (2) of the introduction says that $\NS^+(X\times X)$ is not finitely generated. The argument given in \cite{Ros81} revolves around the alternating matrices associated with effective line bundles. Our aim here is to prove by different methods that the analogous statement holds in any dimension. To simplify the proof, we only consider abelian varieties of Picard number $1$ for now, as the general case will follow with no effort from the proof of the theorem in Sect.\ \ref{sect proof}. \begin{proposition}\label{product case} Let $X$ be an abelian variety with $N\hskip-0.2em S(X)\cong{\Bbb Z}$. Then $\NS^+(X\times X)$ is not finitely generated. \end{proposition} \startproof{\it Proof. } We denote by $\iota_1,\iota_2,\iota_3$ the closed embeddings of $X$ in $X\times X$ given by $$ \iota_1:x\mapstochar\longrightarrow(x,0),\ \iota_2:x\mapstochar\longrightarrow(0,x),\ \iota_3:x\mapstochar\longrightarrow(x,x) \ . $$ Further, fix an ample line bundle $M$ whose algebraic equivalence class generates $N\hskip-0.2em S(X)$ and let $n$ denote the dimension of $X$. Supposing to the contrary that $\NS^+(X\times X)$ is finitely generated, our first claim is then the following boundedness statement: \begin{items} \item[(*)] There is an integer $c>0$ such that for all effective line bundles $B$ on $X\times X$ with $\iota_1^* B\equiv M$ the inequality $$ \(\iota_2^* B-\iota_3^* B\)^n \le c $$ holds. \end{items} To prove (*), choose a finite set of generators $\liste N1k$ of $\NS^+(X\times X)$ and write $$ B\equiv\sum_{i=1}^k b_iN_i $$ with integers $b_i\ge 0$. Because of $N\hskip-0.2em S(X)={\Bbb Z}\cdot[M]$ we have $\iota_1^* N_i\equiv n_i M$ with integers $n_i\ge 0$ for $1\le i\le k$. The equivalences $$ M\equiv\iota_1^* B\equiv\sum_{i=1}^k b_i \iota_1^* N_i \equiv\(\sum_{i=1}^k b_in_i\)M $$ show that there is a subscript $i_0$ with the property $$ b_i n_i = \left\{ \begin{array}{ll} 1 & \mbox{, if $i=i_0$} \\ 0 & \mbox{, if $i\ne i_0$} \ . \end{array} \right. $$ If now $N$ is any effective line bundle on $X\times X$ with $\iota_1^* N\equiv 0$, then it follows (for instance using the Seesaw Principle) that $N$ is a multiple of $pr_2^* M$, where $pr_2:X\times X\longrightarrow X$ is the second projection, so that $\iota_2^* N\equiv\iota_3^* N$. In particular we therefore obtain $$ \iota_2^* B-\iota_3^* B\equiv \sum_{i=1}^k b_i\( \iota_2^* N_i-\iota_3^* N_i\) \equiv \iota_2^* N_{i_0} - \iota_3^* N_{i_0} \ , $$ so that (*) will hold, if we take the integer constant $c$ to be $$ c =_{\operatorname{def}}\max\Big\{ \(\iota_2^* N_i-\iota_3^* N_i\)^n\ \Big\vert\ 1\le i\le k\Big\} \ . $$ Having established (*), the idea is now to construct a contradiction by exhibiting a sequence of nef line bundles $B_m$, $m\ge 1$, satisfying \begin{equation}\label{B conditions} \iota_1^* B_m\equiv M \quad\mbox{ and }\quad \lim_{m\rightarrow\infty}\(\iota_2^* B_m-\iota_3^* B_m \)^n =\infty \ . \end{equation} To this end we set $$ L_1=pr_1^*M\ ,\ L_2=pr_2^*M\ ,\ L_3=\mu^*M\ , $$ where $pr_1,pr_2$ are the projections and $\mu$ is the addition map $X\times X\longrightarrow X$. We then consider the line bundles $$ B_m=_{\operatorname{def}} (1-m)L_1 + (m^2-m)L_2 + mL_3 \ . $$ One checks that with this choice of the bundles $B_m$ the conditions \eqnref{B conditions} are satisfied. So we will be done as soon as have shown that $B_m$ is nef for $m\ge 1$. Now recall that an ample line bundle $A$ on $X\times X$ defines an injective homomorphism of vector spaces $$ \NS_{\bbQ}(X\times X)\longrightarrow\End_{\bbQ}(X\times X)\ , \ L\mapstochar\longrightarrow\phi_A^{-1}\phi_L \ , $$ whose image consists of the elements of $\End_{\bbQ}(X\times X)$ which are symmetric with respect to the Rosati involution $f\mapstochar\longrightarrow f'=\phi_A^{-1}\widehat f\phi_A$. In particular, for an endomorphism $f$ of $X\times X$, the pullback $f^*A$ corresponds to the symmetric endomorphism $f'f$. Let now $A=L_1+L_2$. One checks that, thanks to the fact that $A$ is a product polarization, an endomorphism $$ f=\(\begin{array}{cc} f_1 & f_2 \\ f_3 & f_4 \end{array}\): X\times X\longrightarrow X\times X $$ is symmetric if and only if both $f_1$ and $f_4$ are symmetric and $f_2'=f_3$. Therefore the endomorphisms $\alpha_1,\alpha_2,\alpha_3$, which are defined by \begin{eqnarray*} \alpha_1:(x,y)&\mapstochar\longrightarrow&(x,0) \\ \alpha_2:(x,y)&\mapstochar\longrightarrow&(0,x) \\ \alpha_3:(x,y)&\mapstochar\longrightarrow&(x+y,x+y) \end{eqnarray*} are symmetric and, upon using $\alpha_1^2=\alpha_1$, $\alpha_2^2=\alpha_2$ and $\alpha_3^2=2\alpha_3$, one finds that they correspond to the line bundles $L_1,L_2,L_3$. This in turn shows that the line bundle $B_m$ corresponds to the endomorphism $$ \beta_m=(1-m)\alpha_1 + (m^2-m)\alpha_2 + m\alpha_3 \ . $$ The point is now that $\beta_m^2=(m^2+1)\beta_m$, so that $\beta_m|\operatorname{im}\beta_m$ is just multiplication by $m^2+1$. Therefore, if we denote by $Y_m$ the complementary abelian subvariety of $\operatorname{im}\beta_m$, then the differential of $\beta_m$ at the point $0$ is the map $$ d_0\beta_m: T_0\operatorname{im}\beta_m\oplus T_0 Y_m \longrightarrow T_0\operatorname{im}\beta_m\oplus T_0 Y_m, \ (u,v)\mapstochar\longrightarrow((m^2+1)u,0) \ , $$ so that the analytic characteristic polynomial of $\beta_m$ is $$ P_m(t)=t^n\(t-(m^2+1)\)^n \ , $$ But the alternating coefficients of $P_m(t)$ are positive multiples of the intersection numbers $A^iB_m^{2n-i}$, so that $B_m$ is nef, as claimed. This completes the proof of the proposition. \nopagebreak\hspace*{\fill}\qedsymbol\par\addvspace{\bigskipamount} \section{The cone of curves and the nef cone of an abelian variety}\label{sect proof} Finally, we give in this section the proof of the theorem stated in the introduction. So let $X$ be an abelian variety and denote by $N_1(X)$ the vector space of numerical equivalence classes of real-valued 1-cycles on $X$, and by $N\hskip-0.25em E(X)$ the convex cone in $N_1(X)$ generated by irreducible curves. Through the intersection product the vector space $N_1(X)$ is dual to the N\'eron-Severi vector space $\NS_{\bbR}(X)=N\hskip-0.2em S(X)\otimes{\Bbb R}$. The dual cone of $N\hskip-0.25em E(X)$ is the nef cone $$ \operatorname{Nef}(X) =\{\lambda\in\NS_{\bbR}(X)\mid\lambda\xi\ge 0\mbox{ for all }\xi\inN\hskip-0.25em E(X)\} \ , $$ which in the case of abelian varieties coincides with the effective cone (cf.\ Lemma \ref{nef and effective}). The dual of $\operatorname{Nef}(X)$ in turn is the closed cone ${}\hskip0.25em\bar{\hskip-0.25em N\hskip-0.25em E}(X)$, so that one of these two cones is rational polyhedral if and only if the other is. By Gordon's Lemma this is equivalent to the semi-group $\NP(X)$ being finitely generated. (See e.g.\ \cite[Theorem 14.1 and \S\S19,20]{Roc70} for the elementary properties of cones used here). The idea is now, given an abelian variety, to first apply Poincar\'e`s Complete Reducibility Theorem, i.e.\ to decompose it up to isogenies into a product of powers of non-isogenous simple abelian varieties, and to apply Proposition \ref{simple case} and Proposition \ref{product case} subsequently. One needs then that finite generation of $\NP(X)$ is a property which is invariant under isogenies: \begin{lemma}[\cite{Ros81}]\label{isog} Let $X$ and $Y$ be isogenous abelian varieties. Then $\NS^+(X)$ is finitely generated if and only if $\NS^+(Y)$ is. \end{lemma} Since this observation is crucial for our approach, let us briefly indicate a proof, before we proceed to the proof of the theorem. So suppose that $\NP(X)$ is finitely generated and that there is an isogeny $f:X\longrightarrow Y$. Thanks to the fact that $f^*$ embeds $\NS^+(Y)$ into $\NP(X)$ and to the symmetry of the situation, it is enough to show that $f^*\NS^+(Y)$ is finitely generated. Let then $\liste N1k$ be generators for $\NP(X)$ and put for $1\le i\le k$ $$ n_i =_{\operatorname{def}}\min\{ n\in{\Bbb Z}\mid nN_i\in f^*\NS^+(Y) \} \ . $$ (The set on the right-hand side is non-empty, since $f$ is an isogeny.) Then $f^*\NS^+(Y)$ is generated by the elements $n_1N_1,\dots,n_rN_r$ together with those elements $\sum_{i=1}^k m_iN_i$, $0\le m_i<n_i$, which belong to $f^*\NS^+(Y)$. \nopagebreak\hspace*{\fill}\qedsymbol\par\addvspace{\bigskipamount} Consider now for an abelian variety $X$ its decomposition $$ X_1^{n_1}\times\dots\times X_r^{n_r} $$ up to isogenies, where $\liste X1r$ are mutually non-isogenous simple abelian varieties and $\liste n1r$ are positive integers. In view of the remarks made at the beginning of this section, the theorem stated in the introduction will follow from \begin{theorem} The semi-group $\NP(X)$ is finitely generated if and only if $N\hskip-0.2em S(X_i)\cong{\Bbb Z}$ and $n_i=1$ for $1\le i\le r$. \end{theorem} \startproof{\it Proof. } Suppose first that the conditions on the factors $X_i$ and the exponents $n_i$ are satisfied for $X$. By Lemma \ref{isog} we may assume that $X$ is the product $X_1\times\dots\times X_r$. Fix for $1\le i\le r$ an ample generator $N_i$ of $\operatorname{Pic}(X_i)$ and let $A$ be the product polarization $A=\sum_{i=1}^r pr_i^*N_i$. Due to the fact that the $X_i$ are non-isogenous, we have $$ \NS_{\bbQ}\(\prod_{i=1}^r X_i\) \cong\End_{\bbQ}^s\(\prod_{i=1}^r X_i\) \cong\bigoplus_{i=1}^r\End_{\bbQ}^s(X_i) \cong\bigoplus_{i=1}^r\NS_{\bbQ}(X_i) \ , $$ where $\End_{\bbQ}^s\(\prod_{i=1}^r X_i\)$ and $\End_{\bbQ}^s(X_i)$ denote the subgroups of symmetric endomorphisms with respect to the Rosati involutions associated with $A$ and $N_i$ respectively. Therefore $$ \NP(X)=\bigoplus_{i=1}^r {\Bbb Z}^+\cdot[N_i] $$ is finitely generated. Now suppose conversely that $\NP(X)$ is finitely generated. By Lemma \ref{isog} again, we may assume that $X$ is the product $X_1^{n_1}\times\dots\times X_r^{n_r}$. Note that if $V_1$ and $V_2$ are varieties such that $\NS^+(V_1\times V_2)$ is finitely generated, then $\NS^+(V_1)$ and $\NS^+(V_2)$ are finitely generated as well. So in particular Proposition \ref{simple case} applies to the factors $X_i$ and shows that we have $N\hskip-0.2em S(X_i)\cong{\Bbb Z}$ for all $i$. Further, if we had $n_i>1$ for some $i$, i.e.\ if a multiple factor $X_i$ appeared in the product decomposition of $X$, then $\NS^+(X_i\times X_i)$ would be finitely generated, which however is impossible according to Proposition \ref{product case}. This completes the proof of the theorem. \nopagebreak\hspace*{\fill}\qedsymbol\par\addvspace{\bigskipamount} \begin{varthm*}{\it Acknowledgements} \rm This research was supported by DFG contract Ba 1559/2-1. I would like to thank R.\ Lazarsfeld for helpful discussions and the University of California, Los Angeles, for its hospitality. \end{varthm*}
1998-01-06T12:30:51
9712
alg-geom/9712020
en
https://arxiv.org/abs/alg-geom/9712020
[ "alg-geom", "math.AG" ]
alg-geom/9712020
Carlos Simpson
Carlos Simpson
Secondary Kodaira-Spencer classes and nonabelian Dolbeault cohomology
75 pages. Correction-existence and functoriality of decomposition of an infinite loop stack into product of Eilenberg-MacLane stacks don't hold in general. However, what we need for the calculation is still true
null
null
null
null
If $X$ is a smooth projective variety moving in a family, we define a secondary Kodaira-Spencer class for nonabelian Dolbeault cohomology $Hom(X_{Dol}, T)$ of $X$ with coefficients in the complexified 2-sphere $T=S^2\otimes \cc$ (which is a 3-stack on $Sch /\cc$). Let $Z$ be a simply connected projective surface with $h^{2,0}\neq 0$, and let $X$ be the blow-up of $Z$ at a point $P$. As $P$ moves in $Z$, the blow-up $X$ moves in a family and we show that the secondary Kodaira-Spencer class is nontrivial. This contrasts with the fact that the variations of mixed Hodge structures on the homotopy groups of $X$ are constant. We discuss various surrounding notions, including two appendices where we give some details about the Breen calculations in characteristic zero and representability of simply connected complex shapes.
[ { "version": "v1", "created": "Thu, 18 Dec 1997 21:25:20 GMT" }, { "version": "v2", "created": "Tue, 6 Jan 1998 11:30:50 GMT" } ]
2007-05-23T00:00:00
[ [ "Simpson", "Carlos", "" ] ]
alg-geom
\section*{Secondary Kodaira-Spencer classes and nonabelian Dolbeault cohomology} Carlos Simpson\newline CNRS, Laboratoire Emile Picard\newline Universit\'e Toulouse III\newline 31062 Toulouse CEDEX, France\newline [email protected] \bigskip One of the nicest things about variations of Hodge structure is the ``infinitesimal variation of Hodge structure'' point of view \cite{IVHS}. A variation of Hodge structure $(V=\bigoplus V^{p,q}, \nabla )$ over a base $S$ gives rise at any point $s\in S$ to the {\em Kodaira-Spencer map} $$ \kappa _s: T(S)_s \rightarrow Hom (V^{p,q}_s, V^{p-1, q+1}_s). $$ In the geometric situation of a family $X\rightarrow S$ we have $$ V^{p,q}_s = H^q(X_s, \Omega ^p_{X_s}), $$ and the Kodaira-Spencer map is given by cup-product with the Kodaira-Spencer deformation class $$ T(S)_s \rightarrow H^1(X_s, T(X_s)). $$ The Kodaira-Spencer map is a component of the connection $\nabla$. In particular, this implies that if $\kappa _s\neq 0$ then the connection $\nabla$ is nontrivial with respect to the Hodge decomposition. Various Hodge-theory facts imply that the global monodromy must be nontrivial in this case. We can be a bit more precise: if $u\in V^{p,q}$ is a vector such that $\kappa _s(v)(u)\neq 0$ for some tangent vector $v\in T(S)_s$, then $u$ cannot be preserved by the global monodromy. Thus a local calculation (which actually only depends on the first-order deformation of $X_s$) implies a global fact. In particular this global fact would hold for any family of varieties $X'$ over any base $S'$, such that the new family osculates to order $1$ with the original one (say as a map from $S'$ into the moduli stack of the fibers). A particularly nice aspect of this situation is that the Kodaira-Spencer map is defined on the Dolbeault cohomology $H^q(X_s, \Omega ^p)$ and in particular it is obtained involving only algebraic-geometric calculations (just a cup-product with the deformation class)---no analytic considerations are needed. The goal of this paper is to calculate an example showing a similar type of behavior with a secondary Kodaira-Spencer class coming from nonabelian cohomology with coefficients in the complexified $2$-sphere $T=S^2 \otimes {\bf C}$. For such a $T$ (or any other coefficient stack similar in nature) we define the {\em nonabelian Dolbeault cohomology of $X$ with coefficients in $T$}, denoted $Hom (X_{Dol}, T)$. When $X$ varies in a family parametrized by a base $S$ then we show how to define a secondary class which is a map $$ \bigwedge ^2T(S)_s\rightarrow \pi _1(Hom (X_{Dol}, T)) = H^2_{Dol}(X)/(\eta ) $$ where $\eta$ is the class in $H^2_{Dol}(X)$ pulled back from the tautological class on $T$ by the map $X_{Dol}\rightarrow T$ which we take as basepoint. The secondary class is defined when the primary Kodaira-Spencer classes vanish. It seems likely, although we don't show that here, that our secondary class is just a quadruple Massey product $$ \alpha \wedge \beta \mapsto \{ \eta , \eta , \alpha , \beta \} $$ for $\alpha , \beta \in H^1(X, TX)$ and $\eta \in H^2_{Dol}(X)$ with $\eta \cup \eta = 0$. Instead of interpreting the secondary class this way, we calculate it directly using Mayer-Vietoris arguments for nonabelian cohomology and reducing to calculations in abelian cohomology. Our main purpose in the present paper is to calculate the secondary class in a specific, somewhat instructive, example. We look at the family of varieties $X$ obtained by blowing up a point $P\in Z$ on a smooth surface $Z$. This family is parametrized in an obvious way by $Z$. If $Z$ is simply connected, then any standard Hodge-theoretic information related to $X$ must be independent of the basepoint because it would vary in a variation of (mixed) Hodge structure parametrized by $Z$, and such a variation is forcibly constant. However, we will show that the secondary Kodaira-Spencer class for nonabelian Dolbeault cohomology is nonzero, if $Z$ has a nonzero holomorphic $2$-form. Before getting to the example, we will discuss some general aspects of nonabelian cohomology with emphasis on the case of Dolbeault cohomology. We start with a quick review of $n$-stacks in \S 1. Then we define the notion of ``connected very presentable shape''in \S 2 by looking at maps to $n$-stacks $T$ which are $0$-connected, with $\pi _1(T)$ an affine algebraic group scheme, and $\pi _i(T)$ vector spaces, for $i\geq 2$. The next step in \S 3 is to set $X_{Dol}$ equal to the $1$-stack over $X$ whose fiber is $K(\widehat{TX}, 1)$ where $\widehat{TX}$ is the completion of the tangent bundle of $X$ along the zero-section. We define the {\em nonabelian Dolbeault cohomology of $X$ with coefficients in an $n$-stack $T$} to be $Hom (X_{Dol},T)$. We look at this particularly for connected very presentable $T$. This gives rise to the {\em Dolbeault shape} which is the $(n+1)$-functor $$ T\mapsto Hom (X_{Dol}, T) $$ on connected very presentable $T$. Sections 4 and 5 treat group actions and secondary classes in the $n$-stack situation. In \S 5 we define the secondary Kodaira-Spencer class in general. Then in \S 6 we look at a particular case of $T$, namely the {\em complexified $2$-sphere}, defined by the conditions that $\pi _2=\pi _3 = {\cal O} $ and having nontrivial Whitehead product. With these preliminary steps done, we get to our example in \S\S 7-8. The main result is Theorem \ref{calculation}. It says that if $X$ is the blowup of a surface $Z$ at a point $P$ then the secondary Kodaira-Spencer class for nonabelian Dolbeault cohomology $Hom (X_{Dol}, T)$ with coefficients in the complexified $2$-sphere $T$, is Serre dual to the evaluation map of holomorphic $2$-forms at $P$. In particular if $H^0(Z,\Omega ^2_Z)\neq 0$ then this class is nonzero. At the end in two appendices, we will discuss some topics from \cite{kobe} but in greater detail. First, the ``Breen calculations'' giving the cohomology of $K({\cal O} , n)$ or more generally $K(V/S,n)$ for a vector sheaf $V$ over a base scheme $S$. Then we discuss representability of certain shapes. These topics come up in a few places in the body of the paper, which is the reason for the appendices; for an introduction we refer the reader to the appendices. Without going into great bibliographic detail, we point out here some recent papers which seem to be somewhat related. Karpishpan \cite{Karpishpan1} \cite{Karpishpan2} treats higher-order Kodaira-Spencer mappings, i.e. the higher order derivatives of the period map. The same is treated by Ran \cite{Ran} and Esnault-Viehweg \cite{EsnaultViehweg}. Biswas defines secondary invariants for families of Higgs bundles \cite{Biswas}. Also Bloch and Esnault treat algebraic Cherns-Simons classes \cite{BlochEsnault}, which are types of secondary classes. I would like to thank Mark Green for an inspiring question---albeit one which the present paper doesn't answer. He asked whether there are examples of families of varieties where the variation of Hodge structure on the cohomology is constant, but where the variation of mixed Hodge structure on the homotopy groups is nontrivial. In the absence of an answer to that question (which is very interesting), the typical mathematician's reply is to change the question---in this case, to look for an example where even the mixed Hodge structures on the homotopy groups remain constant, but where the Hodge filtration on the full homotopy type is nonconstant. \footnote{ After the first version of this paper, Richard Hain pointed out to me the following example answering Mark Green's original question. The example comes from a paper of Carlson, Clemens, and Morgan (\cite{CarlsonEtAl}, p. 330). Let $C\subset {\bf P} ^3$ be an embedded curve of positive genus. For points $p,q\in C$, let $X_{p,q}$ be the $3$-fold obtained by first blowing up $p$ and $q$ and then blowing up the strict transform of $C$. The family of $X_{p,q}$ parametrized by (an open subset of) $C\times C$ has constant variation of Hodge structures on the cohomology but, according to \cite{CarlsonEtAl} the variation of MHS on the homotopy is nonconstant.} \subnumero{Notation} We always work in characteristic $0$. In order to simplify notation we use ${\bf C}$ as the ground field (i.e. $Spec ({\bf C} )$ as base scheme), but everything we say would work equally well over any ground field of characteristic $0$. Let $Sch /{\bf C}$ denote the site of schemes of finite type over $Spec ({\bf C} )$ with the etale topology. The structure sheaf ${\cal O}$ on $Sch /{\bf C}$ is the sheaf defined by $$ {\cal O} (Y):= \Gamma (Y, {\cal O} _Y). $$ It is represented by the affine line ${\bf A}^1$, in other words it is represented by the $1$-dimensional vector space ${\bf C}$. A finite dimensional vector space represents a sheaf of the form ${\cal O} ^a$. \numero{Basic remarks about $n$-stacks} We make some brief remarks about $n$-stacks as we shall use them in this paper. For all details the reader is referred to the following references: \newline ---for the history and basic notions of simplicial presheaves: Brown \cite{Brown}, Illusie \cite{Illusie}, Jardine \cite{Jardine}; \newline ---for the history and basic notions of $1$-stacks: Artin \cite{ArtinInventiones}, Deligne-Mumford \cite{Deligne-Mumford}, and Laumon-Moret-Bailly \cite{LMB}; \newline ---for cohomological theory using simplicial presheaves: Thomason \cite{Thomason}; \newline ---for a ``homotopy coherent'' approach: Cordier-Porter \cite{CordierPorter} and also \cite{flexible}; \newline ---for $n$-categories and $n$-stacks: Grothendieck \cite{PursuingStacks}, Breen \cite{BreenAsterisque}, Gordon-Power-Street \cite{GordonPowerStreet}, Tamsamani \cite{Tamsamani}, Baez-Dolan \cite{Baez-Dolan}, and several papers of the author. The first main remark is that we shall almost always be concerned with $n$-stacks of $n$-groupoids, and following the intuition put forth in \cite{PursuingStacks}, an $n$-groupoid is the same thing (up to homotopy) as a topological space whose homotopy groups vanish in degrees $i>n$---we call such a space {\em $n$-truncated}. Thus it is safe to replace $n$-groupoids everywhere by $n$-truncated topological spaces or, again equivalently, $n$-truncated simplicial sets. In this point of view, an $n$-stack (on the site $Sch /{\bf C} $ which is fixed throughout) is just a presheaf of simplicial sets otherwise known as a {\em simplicial presheaf}, which is object-by-object $n$-truncated. Once one has made the passage to simplicial presheaves, the truncation condition is no longer crucial (although it often facilitates arguments and many things in the literature are only stated in this case or under a complementary hypothesis about vanishing cohomological dimension). Thus, when we speak of ``$n$-stacks'', one way to read this is in terms of the homotopy theory of simplicial presheaves, see \cite{Brown} \cite{Illusie} \cite{Jardine} \cite{Thomason}. Another reading would plunge directly into the theory of $n$-categories and $n$-stacks, imposing the groupoid condition along the way. For $n\leq 3$ (which in the end is the case we treat in the present paper) this can be had in a relatively formulaic way in \cite{BreenAsterisque} and \cite{GordonPowerStreet}. For arbitrary $n$, see \cite{Tamsamani} \cite{nCAT}, but unfortunately the $n$-stack part of this theory still needs to be worked out a bit more. The only place where we make reference to $n$-categories which are not $n$-groupoids is when we look at the $n+1$-category $nSTACK$ of $n$-stacks (of groupoids). This $n+1$-category has the property of being {\em $1$-groupic}, i.e. the morphism $n$-categories are $n$-groupoids. One can safely replace the morphism $n$-groupoids by spaces or simplicial sets, and one obtains the notion of {\em Segal category} \cite{effective}, motivated by Segal's delooping machine \cite{Segal}. This notion came into higher category theory in Tamsamani's definition of $n$-category \cite{Tamsamani}. Thus $nSTACK$ may be considered as a Segal category. This fits in relatively nicely with the simplicial presheaf point of view; in fact this Segal category comes from the simplicial category of fibrant and cofibrant objects in Jardine's closed model category of simplicial presheaves. This comes up in looking at the functoriality in $T$ of the construction $Hom (X_{Dol}, T)$. In one other place we refer to the $n+1$-stack $n\underline{STACK}$ of $n$-stacks, which is discussed somewhat in \cite{nCAT} and \cite{limits}; we don't get any further into the general theory here. Currently, the ``simplicial presheaves'' alternative is the most accessible (it is also historically the first, dating from \cite{Brown}). In terms of simplicial presheaves, an {\em $n$-stack} is a simplicial presheaf on $Sch/{\bf C} $ which is object-by-object $n$-truncated. If $X,Y$ are simplicial presheaves, then we obtain a simplicial presheaf $Hom(X,Y)$ by first replacing $Y$ by a fibrant object \cite{Jardine}, then looking at the internal $Hom$ of simplicial presheaves. In particular this gives a simplicial set $Hom(X,Y)(Spec ({\bf C} ))$ and this family of simplicial sets makes the simplicial presheaves into a simplicial category (or Segal category) which we denote $nSTACK$. When we speak of morphisms between $n$-stacks, the above procedure is always understood, i.e. we always replace the target by a fibrant object. Given a presheaf of $n$-groupoids or presheaf of spaces, we often want to take the ``associated $n$-stack''. What this means depends somewhat on the point of view which is taken. If one works with objects in a closed model category such as that of simplicial presheaves, then this just means to consider the object as an element of the closed model category. One might also want to say that it means to replace the object by a weakly equivalent fibrant object. Finally there is an intermediate notion based on enforcing the global descent condition but not the local fibrant condition. It doesn't really matter which point of view we adopt, since when looking at morphisms to a given object, we always replace it by an equivalent fibrant object anyway. The ``yoga'' of the situation is that one can do topology with $n$-stacks instead of spaces (or more precisely $n$-truncated spaces). In particular, all standard constructions and results in algebraic topology carry over to $n$-stacks. Most of these are contained somewhere in the literature referred to above; but if not, we don't give proofs here as that would get beyond the scope of the present paper. There is basically only one slight ``twist'' which is not present in the topological case: this is that the $0$-truncated objects can be topologically nontrivial, i.e. can have cohomology. In the usual topological case, the $0$-truncated objects are just the disjoint unions of contractible components and these make no significant contribution to homotopy. In the case of $n$-stacks over a site, one can have $0$-stacks, i.e. sheaves of sets, with nontrivial cohomology (this is the case of a smooth projective variety $X$, for example); and similarly there are sheaves of groups over $\ast$ (the site itself) which can have nontrivial cohomology. The upshot of all this is that when it comes to choosing basepoints for an $n$-stack $T$, one must choose first an object $Y\in Sch /{\bf C}$ and then choose a basepoint $t\in T(Y)$. (In other words, if we look only at basepoints in $T(Spec {\bf C} )$ we might be missing some topology.) The first place where the previous paragraph has an impact is in the notion of {\em homotopy groups}. If $T$ is an $n$-stack then for any $Y\in Sch /{\bf C}$ and $t\in T(Y)$ we obtain a presheaf, denoted in utmost precision by $$ \pi _i^{\rm pre}(T|_{Sch /Y}, t) $$ but which we often shorten to $\pi _i^{\rm pre}(T,t)$. This is a presheaf of groups over the site $Sch /Y$, abelian if $i\geq 2$. On the other hand the presheaf $\pi _0^{\rm pre}(T)$ is defined absolutely as a presheaf of sets over $Sch /{\bf C} $. The definition which is fundamental to the theory is that we define $$ \pi _i(T|_{Sch /Y}, t) $$ to be the sheaf associated to the presheaf $\pi _i^{\rm pre}(T|_{Sch /Y}, t)$. Similarly $\pi _0(T)$ is the sheaf of sets associated to the presheaf $\pi _0^{\rm pre}(T)$. These, and not the presheaf versions, are the only thing we care about. This is formalized by saying that a morphism $f:T\rightarrow T' $ is called a {\em weak equivalence} \cite{Illusie} if for all $Y\in Sch /{\bf C}$ and $t\in T(Y)$, the resulting morphisms $$ \pi _i(T|_{Sch /Y}, t)\rightarrow \pi _i(T'|_{Sch /Y}, f(t)) $$ are isomorphisms of sheaves on $Sch /Y$ (resp. $\pi _0(T)\rightarrow \pi _0(T')$ is an isomorphism of sheaves of sets on $Sch /{\bf C} $). The theory is localized by this notion of equivalence, in other words $T$ and $T'$ are thought of as equivalent if there is a weak equivalence between them. Jardine constructs a closed model category which takes this into account \cite{Jardine}. This leads, in particular, to the right notion of morphism, namely we only look at morphisms whose target is a fibrant object; if necessary, a target object is replaced by a weakly equivalent fibrant object. Without further mentionning this, we make the convention that whenever we speak of morphisms between $n$-stacks, the target object is made fibrant. For a general site, one can have a connected stack $T$ (i.e. $\pi _0(T)=\ast$) but where the global section space of $T$ is empty, or nonconnected. However, in the present case we are working in the etale topology over an algebraically closed field ${\bf C}$. In this case we have the implication $$ \pi _0(T)= \ast \; \Rightarrow \; \pi _0(T(Spec ({\bf C} ))= \ast . $$ Indeed, the etale coverings of $Spec ({\bf C} )$ are trivial, so there is no change over the object $Spec ({\bf C} )$ when one passes from the presheaf $\pi _0^{\rm pre}(T)$ to the associated sheaf. If $T$ is connected, then, we can choose a basepoint $t\in T(Spec ({\bf C} ))$ which is unique up to homotopy, so the sheaf of groups $\pi _1(T,t)$ is uniquely defined up to global conjugacy. If $Y$ is any scheme and $t'\in T(Y)$ then locally on $Y$, $t'$ is equivalent to $t|_Y$ so $\pi _1(T, t')$ is locally over $Y$ equivalent to the restriction of $\pi _1(T,t)$ to $Y$. Thus in this case, the fundamental group sheaf $\pi _1(T,t)$ over $Sch /{\bf C} $ gives a relatively accurate picture of the $1$-type of $T$. In particular, it makes sense to require that $T$ be {\em $1$-connected}, that is that $\pi _0(T)=\ast $ and $\pi _1(T,t)=\{ 1\} $ for the basepoint $t\in T(Spec ({\bf C} ))$. If $T$ is $1$-connected then for any scheme $Y$ and $t'\in T(Y)$, $\pi _1(T,t')=\{ 1\}$. Furthermore, in this case the fundamental group of $T(Spec ({\bf C} ))$ is trivial (the cohomological contributions from the higher homotopy vanish because etale cohomology of $Spec ({\bf C} )$ is trivial). Therefore the basepoint $t\in T(Spec ({\bf C} ))$ is well-defined up to unique homotopy. We now describe the standard topological constructions which we shall use for $n$-stacks. The first is the notion of {\em homotopy fiber product}. If $A\rightarrow B \leftarrow C$ are morphisms of $n$-stacks then we obtain the {\em homotopy fiber product} $A\times _BC$ with a diagram $$ \begin{array}{ccc} A\times _BC&\rightarrow & A\\ \downarrow && \downarrow \\ C & \rightarrow & B \end{array} $$ together with a homotopy of commutativity of the diagram. These data are essentially well-defined (in the sense that they are well-defined up to homotopy which is itself well-defined up to homotopy \ldots ). In the simplicial presheaf theory, the homotopy fiber product is obtained by replacing one of the two morphisms by a fibrant morphism and then taking the usual fiber product. In the $n$-category theory, see \cite{limits}. Suppose $f:A\rightarrow B$ is a morphism and suppose $b\in B(Spec ({\bf C} ))$. We can think of $b$ as a morphism $b: \ast \rightarrow B$ where $\ast$ denotes the constant presheaf with values the $1$-point topological space. Define the {\em fiber of $f$ over $b$} to be the homotopy fiber product $$ Fib(f,b):= \ast \times _B A. $$ If the base $B$ is $0$-connected, then as mentionned above, the choice of basepoint $b$ exists and is unique up to a global homotopy (i.e. a path in $B(Spec ({\bf C} ))$. Thus we can denote by $Fib(f)$ the fiber over this $b$, bearing in mind that it is defined up to the conjugation action of $\pi _1(B (Spec ({\bf C} )))$. If $B$ is $1$-connected the choice of basepoint $b$ is unique up to unique homotopy, so $Fib(f)$ is well-defined up to homotopy. We call this the {\em homotopy fiber of $f$}. We say that $$ A\rightarrow B \rightarrow C $$ is a {\em fiber sequence} if $C$ is $1$-connected or if we are otherwise given a basepoint $c\in C(Spec ({\bf C} ))$, if we are given a homotopy between the composition $A\rightarrow C$ and the constant map at the basepoint $c$, and if the map $A\rightarrow B$ together with this homotopy induce an equivalence between $A$ and $Fib(B\rightarrow C, c)$. A morphism $T\rightarrow R$ is said to be a {\em locally constant fibration with fiber $F$} if for every scheme $Y\rightarrow R$, locally on $Y$ (in the etale topology) we have $Y\times _RT \cong Y \times F$. In the usual topological case with connected base, this is vacuous. In the case of $n$-stacks over a site, we still have that if $R$ is $0$-connected then any morphism $T\rightarrow R$ is a locally constant fibration. However, we are often interested in cases where $R$ is not connected (i.e. $\pi _0(R)$ is some nontrivial sheaf of sets). In these cases, being locally constant condition is a nontrivial additional condition. The next general type of operation we discuss is {\em truncation}. If $T$ is an $n$-stack and if $m\leq n$ then we obtain an $m$-stack $\tau_{\leq m}T$ (i.e. a simplicial presheaf which is $m$-truncated, in other words has homotopy group sheaves vanishing in degrees $>m$) together with a morphism of $n$-stacks $$ T\rightarrow \tau_{\leq m}T $$ which induces an isomorphism on homotopy group sheaves in degrees $i\leq m$. The $m$-stack $\tau_{\leq m}T$ together with this morphism are essentially well-defined. We can construct $\tau _{\leq m}T$ as the $m$-stack associated to the presheaf of spaces $$ (\tau _{\leq m}^{\rm pre}T)(Y):= \tau _{\leq m}(T(Y)) $$ where the truncation on the left is just truncation of topological spaces (also known as the coskeleton operation). A first example of truncation is the sheaf of sets $\pi _0(T)= \tau _{\leq 0}T$. If $T\rightarrow R$ is a morphism of $n$-stacks then there is a relative (or ``fiberwise'') version of the truncation denoted $\tau _{\leq m/R}(T)\rightarrow R$. This is defined by the property that for any scheme $Y$ and morphism $Y\rightarrow R$, $$ \tau _{\leq m/R}(T)\times _R Y = \tau _{\leq m}(T\times _RY). $$ Using the operations of truncation and homotopy fiber products, we obtain the {\em Postnikov tower}. If $T$ is an $n$-stack then we have morphisms $$ T\rightarrow \ldots \rightarrow \tau _{\leq m}T $$ $$ \rightarrow \tau _{\leq m-1}T \rightarrow \ldots \rightarrow \pi _0(T). $$ In order to describe the stages in this tower of maps, we need a few more notions. Suppose $Y$ is a scheme and $L$ is a sheaf of groups over $Y$. Fix $m\leq n$ and suppose $L$ is abelian if $m\geq 1$. Then we can construct the simplicial presheaf $K^{\rm pre}(L,m)$ on $Sch/Y$ by the standard construction applied to $L$; let $K(L,m)$ be the associated stack. Note that $K(L,m)$ has a chosen basepoint section (over $Y$) which we denote by $0$, and $\pi _i(K(L,m),0)=0$ for $i\neq m$, and it is $=L$ for $i=m$. Furthermore these properties characterize $K(L,m)$ essentially uniquely. The $K(L,m)$ on the site $Sch /Y$ corresponds to an $n$-stack on $Sch /{\bf C}$ with morphism to $Y$ (where $Y$ is considered as a $0$-stack or sheaf of sets), which we denote by $K(L/Y,m)\rightarrow Y$. We can do the same construction relative to any $n$-stack but for this we need to have a notion corresponding to sheaf of (abelian) groups. If $A$ is an $n$-stack then a {\em local system of (abelian) groups)} on $A$ is a morphism $L\rightarrow A$ with relative group structure, which is relatively $0$-truncated (i.e. for any scheme $Y$ and map $Y\rightarrow A$, the homotopy fiber product $Y\times _AL$ is $0$-truncated). This is equivalent to the data for every $Y$ of a local system of (abelian) groups $L_Y$ over $A(Y)$, together with restriction morphisms $L_Y|_{A_{Y'}}\rightarrow L_{Y'}$ for $Y'\rightarrow Y$, satisfying the obvious associativity condition. {\em Caution:} if $X$ is a sheaf of sets represented by a scheme, then a local system over $X$ (according to the above terminology) is the same thing as a sheaf of (abelian) groups over $X$. It doesn't have anything to do with the notion of ``flat vector bundle'' over $X$. If $L\rightarrow A$ is a local system of abelian groups then we obtain a morphism $$ K(L/A,n)\rightarrow A, $$ whose homotopy fiber over any $a\in A(Y)$ is the Eilenberg-MacLane $n$-stack $K(L|_Y,n)$ over $Y$. For $n=1$ we can make do with any local system of groups not necessarily abelian. There is a standard fibration sequence relative to $A$ $$ K(L/A,m)\rightarrow A \rightarrow K(L/A, m+1), $$ in other words $$ K(L/A,m)= A \times _{K(L/A, m+1)}A. $$ Using this we obtain the usual description of the stages in the Postnikov tower: if $m\geq 2$ then, setting $A:= \tau _{\leq m-1}T$ there is a local system $L$ of abelian groups over $A$ and a section $ob:A\rightarrow K(L/A, m+1)$ such that the morphism in the Postnikov tower $$ \tau _{\leq m}T \rightarrow \tau _{\leq m-1}T =A $$ is equivalent to $$ A\times _{K(L/A,m+1)} A\rightarrow A $$ where the first morphism in the fiber product is $ob$ and the second is $0$. The description of the first stage $\tau _{\leq 1}T\rightarrow \pi _0(T)$ is much more complicated and is basically the subject of Giraud's book \cite{Giraud}. Using the notion of local system we can define a relative version of the homotopy group sheaves. If $T\rightarrow R$ is a morphism of $n$-stacks and $s: R\rightarrow T$ is a section then we obtain local systems of groups $\pi _i(T/R, s)$ over $R$. Suppose $X\rightarrow Z$ and $Y\rightarrow Z$ are morphisms of $n$-stacks. Then we obtain a {\em relative internal $Hom$} which is an $n$-stack with morphism to $S$, $$ Hom (X/Z,Y/Z)\rightarrow S. $$ It is defined by the universal property that maps $A\rightarrow Hom (X/Z,Y/Z)$ are the same (in an essentially well-defined way) as maps $X\times _ZA\rightarrow Y$ over $Z$. For existence, if the proof isn't contained somewhere in the literature then one might have to apply the techniques of \cite{limits}. If $Z=\ast$ then we get back to the usual internal $Hom(X,Y)$. Similarly if $Y\rightarrow X \rightarrow Z$ then we obtain the {\em relative section stack} $$ \Gamma (X/Z, Y) $$ which is defined to be the fiber product $$ Hom (X/Z,Y/Z)\times _{Hom (Y/Z,Y/Z)}Z $$ where the second map in the fiber product is that corresponding to the identity of $Y$, and the first map is induced by $Y\rightarrow X$. Again if $Z=\ast$ we denote this simply by $\Gamma (X,Y)$. We now come to one of the main types of observations, namely the relationship between the above objects and cohomology. See for example Thomason \cite{Thomason} for much of this. If $A$ is an $n$-stack and $L$ a local system of abelian groups over $A$ then we define $$ H^i(A, L):= \pi _0\Gamma (A, K(L/A,i)). $$ It is a sheaf of abelian groups on the site $Sch /{\bf C} $. Similarly if $Z$ is an $n$-stack, $p:A\rightarrow Z$ a morphism and $L$ a local system of abelian groups over $A$ then we define $$ H^i(A/Z,L):= \tau _{\leq 0 /Z}\Gamma (A/Z, K(L/A,i)) $$ where $\tau _{\leq 0 /Z}$ is the relative version of the truncation operation for $n$-stacks over $Z$. Note that $H^i(A/Z,L)$ is a local system of abelian groups on $Z$. We can also denote it by $R^ip_{\ast} (L)$. One has the result that the cohomology defined above coincides with sheaf cohomology over simplicial objects (representing $A$ by a simplicial object in the topos of $Sch /{\bf C}$). See \cite{Thomason} or \cite{flexible}. In particular the notation $R^ip_{\ast} (L)$ coincides with the usual meaning (particularly when we are looking at $A$ and $Z$ which are represented by schemes, for example). We have the formulae $$ \pi _i(\Gamma (A, K(L/A,m)), 0) =H^{m-i}(A,L) $$ and (using the relative version of homotopy groups) $$ \pi _i(\Gamma (A/Z, K(L/A,m))/Z, 0) =H^{m-i}(A/Z,L). $$ The usual results concerning cohomology of topological spaces hold for cohomology as defined above. In particular, we have cup-products, corresponding to the following operations on Eilenberg-MacLane spaces. If $L$, $L'$ and $L''$ are local systems of abelian groups over $A$ and if $$ L\times _AL'\rightarrow L'' $$ is a bilinear morphism (of relative abelian group objects) then we obtain morphisms $$ K(L/A, i)\times K(L'/A, j)\rightarrow K(L'' /A, i+j). $$ These give cup-products in cohomology which are bilinear morphisms $$ H^i(A, L)\times H^j(A,L')\rightarrow H^{i+j}(A, L''). $$ We also have a K\"unneth formula. The case which we use in the present paper is as follows. Suppose $X$ and $Y$ are $n$-stacks. Then $$ H^m(X\times Y, {\cal O} )= \bigoplus _{i+j= m} H^i(X, H^j(Y,{\cal O} )). $$ If $H^j(Y,{\cal O} )$ are represented by finite dimensional vector spaces then we can write the more usual formula $$ H^m(X\times Y, {\cal O} )= \bigoplus _{i+j= m} H^i(X, {\cal O} )\otimes _{{\cal O}} H^j(Y,{\cal O} ). $$ This extends to the relative case of morphisms $X\rightarrow S$ and $Y\rightarrow S$ if these families are locally trivial over the etale topology of $S$. Finally, we have a Leray-Serre spectral sequence. See \cite{Thomason} for one way to set this up. If $f: X\rightarrow Y$ is a morphism of $n$-stacks and if $L$ is a local system of abelian groups on $X$ then we obtain a ``complex'' $R^{\cdot} f_{\ast}(L)$ on $Y$ and the cohomology of $X$ is the ``hypercohomology'' of this complex. These terms are put in quotations because one should actually interpret the notion of complex as being a fibration in spectra over $Y$ (the raw notion of complex of local systems is not adapted to the higher homotopy involved if $Y$ is not $0$-truncated and locally cohomologically trivial). In any case we get the cohomology objects of the direct image, which are the relative homotopy group sheaves of the fibration of spectra, denoted $R^if_{\ast}(L)$. These are local systems over $Y$. We have the Leray-Serre spectral sequence (cf Thomason \cite{Thomason}) $$ E^{i,j}_2 = H^i(Y, R^jf_{\ast}(L))\Rightarrow H^{i+j}(X, L). $$ One way of approaching all of the above details is to replace any $n$-stack i.e. simplicial presheaf (particularly those coming in as the base in questions about cohomology) by a simplicial object whose stages are formal, possibly infinite, disjoint unions of schemes (this technique was pointed out to me by C. Teleman). Similarly, we can replace morphisms of $n$-stacks by morphisms of such objects. Then questions about cohomology become just questions about cohomology of simplicial schemes with obvious modifications made to allow for the infinite number of components. The $n$-stack $K({\cal O} , n)$ has an infinite loop-space structure or $E_{\infty}$-structure, in other words it has an infinite delooping (the $m$-fold delooping is just $K({\cal O} , m+n)$). This structure is the homotopical analogue of an abelian group structure: it contains a homotopy class of maps $$ K({\cal O} , n)\times K({\cal O} , n)\rightarrow K({\cal O} ,n) $$ but also higher homotopies of associativity, commutativity etc. We often think of $K({\cal O} , n)$ as a homotopical group object and speak of things such as ``principal bundles'' for it. The infinite loop-space structure is inherited by $Hom ({\cal F} , K({\cal O} , n))$ for any sheaf ${\cal F}$. In the first version of the paper, it was stated that the infinite loop-space structure provides a functorial decomposition into products of Eilenberg-MacLane stacks. This is not true in general. For one thing, such a decomposition may not exist, and for another it is never completely functorial. For existence, the obstruction is in the $Ext$ between the various homotopy group sheaves. If such a decomposition exists, then it can be functorial up to one homotopy, but this homotopy itself will not be uniquely determined up to a second homotopy. The statement which we actually need for the calculation in \S 8 below, does work, and we state it as a proposition. \begin{proposition} \label{decomp} (A)\, Suppose $(Y,y)$ is a basepointed $n$-stack with infinite loop structure. Suppose that $\pi _i(Y,y)$ are represented by finite-dimensional vector spaces. Then there exists an equivalence $$ \varepsilon : Y\cong Y_0 \times \ldots \times Y_n $$ where $$ Y_i = K( {\cal O} ^{a_i}, i) = K(\pi _i(Y,y), i). $$ (B)\, Suppose $f:Y\rightarrow Y'$ is a map of basepointed $n$-stacks both as in (A), and suppose $f$ is compatible with the infinite loop structure. Let $$ \varepsilon : Y\cong Y_0\times \ldots \times Y_n,\;\;\; \varepsilon ': Y'\cong Y'_0\times \ldots \times Y'_n $$ denote the maps given by (A) (chosen independantly of $f$). Then there is a homotopy making the square $$ \begin{array}{ccc} Y&\cong &Y_0\times \ldots \times Y_n\\ \downarrow &&\downarrow \\ Y'&\cong &Y'_0\times \ldots \times Y'_n \end{array} $$ commute, where the vertical arrow on the right is a product of the morphisms of Eilenberg-MacLane stacks $Y_i\rightarrow Y'_i$ induced by $f_{\ast}:\pi _i(Y,y)\rightarrow \pi _i(Y', y')$. \end{proposition} {\em Proof:} The short way of saying this is that the $Ext^j(V,W)$ vanish for $j>0$ for finite dimensional vector spaces, so the obstruction to splitting vanishes. We give the following more concrete argument (which is in a certain sense just repeating the argument which will be given in \ref{ext} below for the vanishing of the $Ext^j$). For $N>n$ we are given an $N-1$-connected pointed $n+N$-stack $(Z,z)$ with $(Y,y)= \Omega ^N(Z,z)$ (this is the {\em ad hoc} definition of ``infinite loop structure'' which we use). In part (B) the map $f$ comes from a map $g:Z\rightarrow Z'$. Thus for part (A) it suffices to obtain a decomposition $$ Z\cong Z_0\times \ldots \times Z_n $$ with $$ Z_i = K(V_i, N+i),\;\;\; V_i := \pi _i(Y,y) = \pi _{N+i}(Z,z), $$ and for part (B) it suffices to obtain the homotopy of functoriality on the level of the morphism $g$. Calculate the cohomology of $Z$ with coefficients in ${\cal O}$. Using Leray-Serre spectral sequences for the stages in the Postnikov tower, and using the Breen calculations \ref{bc} in view of the hypothesis that the $\pi _j$ are finite dimensional vector spaces, we find that for $i\leq N+n$, $$ H^{j}(Z, {\cal O} ) = Hom (\pi _j(Z,z), {\cal O} ) =V_{N-j}^{\ast}. $$ Thus we have tautological classes $$ \varepsilon _i \in H^{N+i}(Z, \pi _{N+i}(Z,z))= H^{N+i}(Z, V_i) $$ which together provide us with a map $$ \varepsilon = (\varepsilon _0,\ldots , \varepsilon _n) : Z \rightarrow K(V_0,N)\times \ldots \times K(V_n, N+n). $$ This map induces an isomorphism on homotopy groups (in degrees up to $N+n$). Thus it provides the required splitting for (A). For part (B) suppose we have a map $$ g:Z= Z_0\times \ldots \times Z_n \rightarrow Z'_0\times \ldots \times Z'_n $$ with $Z_i= K(V_i, N+i)$ and $Z'_i= K(W_i, N+i)$. Such a map corresponds, up to homotopy, to a collection of classes in $H^{N+i}(Z, W_i)$. From the K\"unneth formula (which can be seen by a collection of Leray-Serre spectral sequences for the projections onto the factors) we have $$ H^{N+i}(Z, W_i) = H^{N+i}(Z_i, W_i). $$ The other factors vanish again using the Breen calculations \ref{bc} from the fact that $V_j$ and $W_i$ are represented by finite dimensional vector spaces. Our map $g$ is therefore homotopic to a map given by the classes in $H^{N+i}(Z_i, W_i)$, i.e. a map compatible with the product decomposition. \hfill $\Box$\vspace{.1in} The homotopy in part (B) is not unique: it can be changed by a map $Y\rightarrow \Omega Y'$ in other words by a collection of morphisms $Y_i\rightarrow Y'_{i+1}$. \medskip Suppose that ${\cal F}$ is an $n$-stack such that the $H^i({\cal F} , {\cal O} )$ are represented by finite dimensional vector spaces. Then we can apply the above proposition to $$ Y:= Hom ({\cal F} , K({\cal O} ,n)). $$ The decomposition of $Hom ({\cal F} , K({\cal O} ,n))$ into a product of Eilenberg-MacLane spaces $$ Hom ({\cal F} , K({\cal O} ,n)) = \prod _i K(H^{n-i}({\cal F} , {\cal O} ), i), $$ is related to the K\"unneth formula. A morphism $Z\rightarrow Hom ({\cal F} , K({\cal O} , n))$ corresponds (by the definition of internal $Hom$) to a morphism $Z\times {\cal F} \rightarrow K({\cal O} ,n)$, in other words to a class $f\in H^n(Z\times {\cal F} , {\cal O} )$. By the above product structure this class decomposes into a collection of classes $f_i\in H^i(Z, H^{n-i}({\cal F} , {\cal O} ))$. The $f_i$ are the K\"unneth components of $f$. \begin{center} $\ast$ \hspace*{2cm}$\ast$ \hspace*{2cm}$\ast$ \end{center} For the remainder of the paper, we look at $n$-stacks of groupoids on $Sch /{\bf C}$ and unless specified otherwise, the reader may fix any $n\geq 3$ (for the calculation it suffices to take $n=3$.) \numero{Connected very presentable shape} We isolate some special $n$-stacks $T$ and then use them to measure the ``shape'' of an arbitrary $n$-stack ${\cal F}$. In other words, take a sub-$n+1$-category ${\cal P} \subset nSTACK$ of the $n+1$-category of $n$-stacks (of groupoids, say), and look at the $n+1$-functor $$ {\cal P} \rightarrow nSTACK $$ $$ T\mapsto Hom ({\cal F} , T). $$ We call this $n+1$-functor the {\em shape of ${\cal F}$ as measured by ${\cal P}$}. There are many possible ways to choose ${\cal P}$. Some reasonable parameters are to require that ${\cal P} \subset nSTACK$ be a full sub-$n+1$-category (in other words that we make no limitation on the morphisms of ${\cal P}$); and that the condition $T\in {\cal P}$ should be measured only by looking at the homotopy group sheaves $\pi _i(T,t)$. For our present purposes we start by requiring that $T$ be connected, i.e. $\pi _0(T)=\ast$. Among other things, this insures that the isomorphism classes of the higher homotopy group sheaves $\pi _i(T,t)$ be well defined. Before describing our choice of conditions for the $\pi _i(T,t)$ we take note of the following: any $T\in {\cal P}$ will decompose in a Postnikov tower whose stages are $K(\pi _i , i)$. Morphisms ${\cal F} \rightarrow K(\pi _i,i)$ are classified by $H^i({\cal F} ,\pi _i)$ and more generally, one has obstruction theory for classifying the morphisms ${\cal F} \rightarrow T$ going up in the Postnikov tower; the obstruction classes are in $H^{i+1}({\cal F} ,\pi _i)$. Thus, one should choose the class of possible $\pi _i$ to be a class of sheaves such that, for the ${\cal F}$ we are interested in, the cohomology $H^j({\cal F} ,\pi _i)$ has reasonable properties. For the topic of Dolbeault cohomology, we already know how to take nonabelian $H^1$ with coefficients in an affine algebraic group \cite{Moduli} \cite{hbls}, and we know how to take higher Dolbeault cohomology with coefficients in ${\bf C}$ or more generally in a finite-dimensional complex vector space. This suggests that our condtions should be that $\pi _1$ be an affine algebraic group, and $\pi _i$ be represented by finite-dimensional vector spaces for $i\geq 2$. Recall from \cite{RelativeLie} and \cite{GeometricN} that a {\em connected very presentable $n$-stack $T$} is an $n$-stack of groupoids $T$ on $Sch /{\bf C}$ subject to the following conditions: \newline {\bf (connectedness):} $\pi _0(T) = \ast$ as a sheaf of sets on $Sch /{\bf C}$; \newline {\bf (very presentability):} if $t\in T(Spec ({\bf C} ))$ is a basepoint (which we assume exists) then $\pi _i(T,t)$ are representable by group schemes of finite type over $Spec ({\bf C} )$, which are required to be affine for $i=1$ and vector spaces (i.e. affine unipotent abelian) for $i\geq 2$. (This is the ``very presentability'' condition of \cite{RelativeLie} under the additional hypothesis of connectedness; in the non-connected case the definition is more complicated and that is basically the subject of the paper \cite{RelativeLie}.) We now choose ${\cal P}$ for our shape theory to be the $n+1$-category of connected very presentable $n$-stacks of groupoids. Suppose ${\cal F}$ is an $n$-stack on $Sch/{\bf C}$. The{\em shape of ${\cal F}$} is defined as the $n+1$-functor from the $n+1$-category of connected very presentable $n$-stacks $T$, to the $n+1$-category $nSTACK$ of all $n$-stacks, given by the formula $$ Shape ({\cal F} )(T):= Hom ({\cal F} , T). $$ This contains all information about ${\cal F}$ which one can extract by looking at $G$-torsors over ${\cal F}$ and cohomology of associated vector bundles. In many cases, $Hom({\cal F} , T)$ will be a geometric or locally geometric $n$-stack \cite{GeometricN}. For example in the case ${\cal F} = X_{Dol}$ we look at below, $Hom(X_{Dol} , T)$ will be locally geometric. In these cases the shape of ${\cal F}$ may be considered as an $n+1$-functor from connected very presentable $n$-stacks to the $n+1$-category of (locally) geometric $n$-stacks, sitting inside $nSTACK$. \subnumero{Examples} We explain how to understand the structure of connected very presentable $T$. First is the simply connected case. Here $T$ is given by a Postnikov tower where the stages are of the form $K({\cal O} ^a, m)$. The only question is how they are put together. The fibration $$ K({\cal O} ^a, m)\rightarrow \tau _{\leq m}T \rightarrow \tau _{\leq m-1}T $$ is classified by a map $$ \tau _{\leq m-1}T\rightarrow K({\cal O} ^a, m+1), $$ in other words $\tau _{\leq m}T$ is the pullback by this map of the standard fibration $$ K({\cal O} ^a, m)\rightarrow \ast \rightarrow K({\cal O} ^a, m+1). $$ We can write $$ \tau _{\leq m}T= \tau _{\leq m-1}T\times _{K({\cal O} ^a, m+1)}\;\; \ast . $$ The classifying map is a class in $H^{m+1}(\tau _{\leq m-1}T, {\cal O} ^a)$. In turn, this cohomology can be ``calculated'' by the Leray-Serre spectral sequence applied to the previous part of the Postnikov tower for $\tau _{\leq m-1}T$. The basic pieces that we need to know are the cohomology of the Eilenberg-MacLane spaces. These are given by the {\em Breen calculations} \cite{Breen1} \cite{Breen2}, which we recall in Appendix I (giving a relative version). For the present discussion the answer is that $H^{\ast} (K({\cal O} ^a, m), {\cal O} )$ is a graded-symmetric algebra on ${\cal O}^a$ in degree $m$. Note that this answer is the same as the classical answer for rational cohomology of rational Eilenberg-MacLane spaces $H^{\ast} (K({\bf Q} ^a, m),{\bf Q} )$. As we shall explain below and also in Appendix II, if $Y$ is a finite simply-connected $CW$-complex then we obtain a $1$-connected very presentable $T= Y\otimes {\bf C}$ whose homotopy group sheaves are $\pi _i(Y,y)\otimes _{{\bf Q}} {\cal O}$. We now look at connected but not simply connected very presentable $T$. Note that since $T$ is connected, we can choose a basepoint $t\in T(Spec ({\bf C} ))$. Let $G:= \pi _1(T,t)$; by hypothesis it is an affine algebraic group scheme over ${\bf C}$. We have a fiber sequence $$ T'\rightarrow T \rightarrow K(G,1) $$ where $T'$ is simply connected very presentable. The homotopy group sheaves $\pi _i(T', t)= \pi _i(T,t)$ are vector spaces ${\cal O} ^a$ for $i\geq 2$; but notice also that $G$ acts on these vector spaces. The action is an action of sheaves on the site $Sch /{\bf C} $ so it is automatically algebraic; we can write $\pi _i(T', t)= V^i$ with $V^i$ a linear representation of $G$. The same Postnikov tower discussion as above, works here. The only difference is that in calculating the cohomology of the $\tau _{\leq m}T$ we may have coefficients which are linear representations of $G$, and at the end we get down to a step where we have to calculate $H^i(K(G,1), V)$. If $G$ is reductive, this ``algebraic cohomology'' vanishes for $i\geq 1$, whereas if $G$ is unipotent then it is equal to the Lie algebra cohomology. A simple example of non-simply connected $T$ may be obtained as follows. Start with a linear algebraic group $G$ with a linear representation $V$, corresponding to a local system $\underline{V}\rightarrow K(G,1)$. Use the notational shorthand $$ K(V/G; n):= K(\underline{V}/K(G,1), n). $$ We have a fibration sequence $$ K(V, n)\rightarrow K(V/G,n) \rightarrow K(G,1). $$ Maps ${\cal F} \rightarrow T$ correspond to pairs $(E, \eta )$ where $E$ is a $G$-torsor over ${\cal F}$ and \newline $\eta \in H^n({\cal F}, E\times ^GV)$. An advantage of the nonabelian cohomological formulation of the secondary Kodaira-Spencer classes we define is that the definition works for cohomology with coefficients in a connected very presentable $T$, even non-simply connected. However, for the calculation we will do, we look at a particular simply-connected $T$ (the ``complexified $2$-sphere'' $S^2\otimes {\bf C}$). \subnumero{Representability} Under certain circumstances, basically when the shape of ${\cal F}$ is simply connected and has reasonable cohomology sheaves, then $Shape ({\cal F} )$ is {\em representable}. By this we mean that there is a morphism ${\cal F} \rightarrow \Sigma$ from ${\cal F}$ to a very presentable $n$-stack $\Sigma$ such that for any other very presentable $n$-stack $T$ we have $$ Hom (\Sigma , T)\stackrel{\cong}{\rightarrow} Hom ({\cal F} , T). $$ For example we have the following precise statement. \begin{theorem} \mylabel{representable0} Suppose ${\cal F}$ is an $n$-stack on $Sch /{\bf C} $ such that for any affine algebraic group $G$, $$ K(G,1)\stackrel{\cong}{\rightarrow}Hom ({\cal F} , K(G,1)). $$ Suppose that the $H^i({\cal F} , {\cal O} )$ are representable by finite dimensional vector spaces. Then there is a morphism ${\cal F} \rightarrow \Sigma$ to a $1$-connected very presentable $n$-stack $\Sigma$, such that for any connected very presentable $n$-stack $T$ we have $$ Hom (\Sigma , T)\stackrel{\cong}{\rightarrow} Hom ({\cal F} , T). $$ \end{theorem} The proof will be given in Appendix II. On the other hand, for $n$-stacks ${\cal F}$ whose shape is not $1$-connected, the shape will not in general be representable. For example, suppose $W$ is a finite $CW$ complex (considered as a constant $n$-stack) such that $\pi _1(W)=\Gamma := {\bf Z} $ and such that some $\pi _i(W)\otimes {\bf Q}$ is a ${\bf Q} [\Gamma ]$-module which is not completely torsion (hence infinite-dimensional over ${\bf Q}$). Then $Shape (W)$ is not representable. Indeed, the infinite dimensionality of $\pi _i(W)\otimes {\bf Q}$ is seen by the shape, since all irreducible representations of $\Gamma$ are finite ($1$-) dimensional. \numero{Nonabelian Dolbeault cohomology} Suppose $X$ is a smooth quasiprojective variety. Recall that one defines the {\em Dolbeault cohomology} of $X$ as the hypercohomology of the trivial complex $\Omega ^{\cdot}_X$ with differential equal to $0$: $$ H^i_{Dol}(X):= {\bf H}^i(X, \Omega ^0_X \stackrel{0}{\rightarrow} \ldots ) = \bigoplus _{p+q=i} H^q(X, \Omega ^q_X) $$ (generally speaking this is only motivated by topology when $X$ is projective; but we make the notation for quasiprojective $X$ too, for use in Mayer-Vietoris arguments). A nonabelian version for $H^1$ may be defined by setting $H^1_{Dol}(X, G)$ equal to the moduli stack of Higgs principal $G$-bundles (\cite{Hitchin0} \cite{Hitchin} \cite{hbls} \cite{Moduli}) $(P,\theta )$. One can impose semistability conditions and vanishing of Chern classes to get a version more closely related to topology, but we don't need that for the present algebraic discussion. If $(P,\theta )$ is a principal Higgs bundle and if $V$ is a representation of $G$ then we obtain an associated Higgs bundle $(E,\theta )$. Recall that we define the {\em Dolbeault cohomology of $(E,\theta )$} as the hypercohomology of the {\em Dolbeault complex} (cf \cite{hbls}) $$ \ldots \stackrel{\theta}{\rightarrow} E\otimes _{{\cal O} _X}\Omega ^i_X \stackrel{\theta}{\rightarrow} \ldots , $$ $$ H^i_{Dol}(X, (E,\theta )):= {\bf H}^i(X, (E\otimes _{{\cal O}_X}\Omega ^{\cdot}_X,\theta )). $$ We present a way of unifying these definitions into a notion of {\em nonabelian Dolbeault cohomology}. For nonabelian $H^1$ the present interpretation was explained in \cite{SantaCruz}. Let $\widehat{TX}$ denote the formal completion of $TX$ along the zero section. Considered as a presheaf on $Sch /{\bf C} $ it associates to any ${\bf C}$-scheme $Y$, the set of maps $Y\rightarrow TX$ which map the underlying reduced subscheme $Y^{\rm red}$ to the zero-section $X\subset TX$. Define the $1$-stack $X_{Dol}$ to be the relative $K(\widehat{TX}/X, 1)$, i.e. the relative classifying stack for the group scheme $\widehat{TX} \rightarrow X$. A variant which is technically easier to work with is $$ X_{UDol} := K(TX/X, 1). $$ The $U$ in the notation stands for ``unipotent'': as we shall see below (Proposition \ref{calcDol}), a morphism $X_{UDol} \rightarrow K(G,1)$ corresponds to a principal Higgs bundle with structure group $G$, over $X$, such that the Higgs field is a section of unipotent elements of the Lie algebra of $G$. However, for morphisms to simply connected $T$, we can safely replace $X_{Dol}$ by $X_{UDol}$ and for the purposes of the present paper this is what we shall do. If $T$ is an $n$-stack then we define the {\em nonabelian Dolbeault cohomology of $X$ with coefficients in $T$} to be the $n$-stack (of $n$-groupoids) $$ Hom (X_{Dol}, T). $$ Of course this includes the classical abelian case, when we take $T= K({\cal O} , n)$ where ${\cal O}$ denotes the structural sheaf, equal to ${\bf G}_a$, represented by the affine line. It also includes the somewhat classical case of nonabelian Dolbeault $H^1$ with coefficients in a group scheme $G$. More generally we will be most interested in the case where $T$ is a connected very presentable $n$-stack. The calculation of the present paper involves a $1$-connected $T$, so it doesn't refer to the case of Higgs principal $G$-bundles or Dolbeault cohomology with coefficients in Higgs bundles. These aspects are only presented to show the unified character of the definition. \begin{proposition} \mylabel{calcDol} We have that $$ \pi _0 (Hom (X_{Dol}, K({\cal O} , n))) = H^n_{Dol}(X, {\bf C} ) $$ is the usual Dolbeault cohomology of $X$. The same holds for $X_{UDol}$. If $G$ is an affine algebraic group scheme then $$ Hom (X_{Dol} , K(G,1))= {\cal M}_{Dol}(X, G) $$ is the moduli $1$-stack of principal Higgs bundles with structure group $G$. On the other hand, $$ Hom (X_{UDol} , K(G,1)) $$ is the moduli $1$-stack of principal Higgs bundles $(P, \theta )$ with structure group $G$, such that for every $x\in X$ the element $\theta _x \in ad(P)_x\cong {\bf g} $ is a unipotent element of the Lie algebra ${\bf g}$ of $G$ (this condition is of course independant of the isomorphism $ad(P)_x\cong {\bf g} $ chosen). \end{proposition} {\em Proof:} Let $X^{\rm fc}_{Dol}$ denote the formal category defined in \cite{SantaCruz} which gives the $1$-stack $X_{Dol}$. We have $$ Ob\, X^{\rm fc}_{Dol} = X, $$ and $$ Mor (X^{\rm fc}_{Dol}) = \widehat{TX} \rightarrow X \hookrightarrow X\times X $$ (the morphism object lies over the diagonal in $X\times X$). The composition of morphisms is just addition in $\widehat{TX}$ (which is a formal group scheme over $X$). This formal groupoid gives in an obvious way a presheaf of groupoids on $Sch /{\bf C} $, whose associated stack is $X_{Dol}$. A morphism $X_{Dol}\rightarrow K(G,1)$ is the same thing as a $G$-torsor over $X_{Dol}^{\rm fc}$, which in turn is the same thing as a principal $G$-bundle $P$ over $X$ together with action of the formal group scheme $\widehat{TX}$. The action may be interpreted as a morphism of sheaves of groups over $X$, $$ \widehat{TX} \rightarrow Ad (P). $$ Since the domain is formal, this is the same thing as a morphism of Lie algebras over $X$, $$ \theta : TX \rightarrow ad(P). $$ This proves the second statement. For the third statement, the same proof works but with $\widehat{TX}$ replaced by $TX$. Note that $TX$ is a unipotent group scheme over $X$, so $$ TX\rightarrow Ad(P) $$ corresponds to a morphism of Lie algebras $$ \theta : TX \rightarrow ad(P) $$ with image in the unipotent elements of $ad(P)$. For the first statement, one way to proceed is to notice that the cohomology of $X_{Dol}$ with coefficients in ${\cal O}$ is the same as the cohomology of the formal category $X^{\rm fc}_{Dol}$ as considered by Berthelot \cite{Berthelot} and Illusie \cite{Illusie}. From those references, one gets a generalized de Rham complex calculating the cohomology, which in our case is seen to be exactly the Dolbeault complex. We also need to prove that the morphism $X_{Dol}\rightarrow X_{UDol}$ induces an isomorphism on cohomology. For this it suffices to look locally over $X$, so we can assume that $TX$ is trivial. Thus it suffices to prove that if $V$ is a vector space of dimension $n$ then the morphism $$ K(\widehat{V},1)\rightarrow K(V,1) $$ induces an isomorphism of cohomology. This morphism has homotopy fiber the sheaf $V_{DR}$ defined by $V_{DR}(Y)=V(Y^{\rm red})$ (cf \cite{kobe}). The cohomology of $V_{DR}$ with coefficients in ${\cal O}$ is the algebraic de Rham cohomology of $V$ (\cite{kobe} Theorem 6.2) which is trivial because $V$ is an affine space. Since the statement of the first part for $X_{UDol}$ is what we actually use, we indicate a somewhat more elementary proof. Let $V$ be a vector space of dimension $n=dim(X)$. We have a natural (split) extension $$ 1\rightarrow V \rightarrow G \rightarrow GL(V) \rightarrow 1 $$ which gives a fiber sequence of $1$-stacks $$ K(V,1)\rightarrow K(G,1)\stackrel{p}{\rightarrow} K(GL(V), 1). $$ Let $C^{\cdot} = R^{\cdot} p_{\ast}({\cal O} )$. Using the Breen calculations \cite{Breen2} \cite{kobe}, which we recall in Theorem \ref{bc} in Appendix I below, it is easy to see that $$ H^i(C^{\cdot}) = \bigwedge ^i(V^{\ast}). $$ However, the fact that $GL(V)$ is a reductive group implies that all of the algebraic group cohomology, in other words the cohomology of $K(GL(V), 1)$ with coefficients in any local system associated to a representation of $GL(V)$, vanishes except in degree $0$. Therefore the invariants of the complex $C^{\cdot}$ vanish so $$ C^{\cdot} \sim \bigoplus _{i} \bigwedge ^i(V^{\ast}). $$ Now notice that the vector bundle $TX\rightarrow X$ corresponds to a map $X\rightarrow K(GL(V), 1)$ and $$ X_{UDol} = K(G,1)\times _{K(GL(V),1)} X. $$ Let $q: X_{UDol}\rightarrow X$ denote the projection; then $R^{\cdot} q_{\ast}{\cal O} $ is the pullback of $C^{\cdot}$, hence it splits as $$ R^{\cdot} q_{\ast}{\cal O} = \bigoplus _i\Omega ^{i}_X. $$ The cohomology of $X_{UDol}$ with coefficients in ${\cal O}$ is equal to the hypercohomology of $X$ with coefficients in this complex, which is the Dolbeault cohomology. \hfill $\Box$\vspace{.1in} {\bf Remark:} We have the following which are sometimes useful: $$ \pi _i (Hom (X_{Dol}, K({\cal O} , n))) = H^{n-i}_{Dol}(X, {\bf C} ). $$ \begin{proposition} \mylabel{calcDol2} Suppose $(P,\theta )$ is a Higgs principal $G$-bundle on $X$ corresponding to a map $X_{Dol}\rightarrow K(G,1)$. Suppose $V$ is a representation of $G$ and let $(E,\theta )$ be the associated Higgs bundle. Then the cohomology of $X_{Dol}$ with coefficients in the local system $(P,\theta )^{\ast}(V)$ is naturally isomorphic to the Dolbeault cohomology $H^n_{Dol}(X, (E,\theta ))$. The same works for a family of principal bundles parametrized by a base scheme $S$. \end{proposition} {\em Proof:} As before, we interpret $(P,\theta )$ as a $G$-torsor over the formal category $X_{Dol}^{\rm fc}$ which defines the stack $X_{Dol}$. Associated to the representation $V$ we get a local system over the formal category, and its cohomology (which is that mentionned in the statement of the proposition) is calculated by a de Rham complex (cf \cite{Berthelot} \cite{Illusie}). This de Rham complex is exactly the Dolbeault complex for $(E,\theta )$. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{UdolDol} Suppose $T$ is a connected very presentable $n$-stack such that $\pi _1(T)$ is a unipotent group scheme over $Spec ({\bf C} )$. Then the morphism $$ Hom (X_{UDol} , T)\rightarrow Hom (X_{Dol} , T) $$ is an equivalence. \end{corollary} {\em Proof:} This follows immediately from the previous propositions using a Postnikov tower for $T$, and noting that if $G$ is a unipotent group scheme then all principal $G$-Higgs bundles satisfy the condition that the Higgs field be a section of unipotent elements. \hfill $\Box$\vspace{.1in} \subnumero{Cohomology classes of subschemes} A particular detail of (usual) Dolbeault cohomology which enters into our calculations is the cohomology class of a subscheme. Suppose $X$ is a smooth projective variety and suppose that $Z\subset X$ is a subscheme of codimension $d$. Then we obtain the class $$ [Z] \in H^d(X, \Omega ^d_X)\subset H^{2d}_{Dol}(X). $$ In fact this class has a canonical trivialization over $U:=X-Z$, in other words we are actually given a lifting to $$ [Z] \in H^d((X,U), \Omega ^d_X)\subset H^{2d}((X_{Dol}, U_{Dol}),{\cal O} ). $$ Suppose for example $d=1$. Take an open covering $X=\bigcup U_i$ with $U_0=U=X-Z$, and the remaining $U_i$ affine with defining equations $z_i\in {\cal O} (U_i)$ having zeros of order one along $Z$ and nonvanishing elsewhere. Put $$ g_{ij}:= \frac{dz_i}{z_i} - \frac{dz_j}{z_j}. $$ This gives a $1$-cocycle which determines the cohomology class $[Z]$. Since $U_0=X-Z$ is a part of the covering, it gives a class in the cohomology of the pair $(X,X-Z)$. The case of higher codimension is treated similarly. The only thing we need to know is that if $\lambda , \lambda ' \in {\cal O} (X)$ are regular functions such that $\lambda |_Z= \lambda '|_Z$, then $\lambda [Z] = \lambda '[Z]$. In particular if $P$ is a point then $\lambda [P]=\lambda (P)[P]$. \numero{Group actions} Suppose $W$ is a sheaf of groups on $Sch /{\bf C} $, and suppose $R$ is an $n$-stack on $Sch /{\bf C} $. Then an {\em action of $W$ on $R$} is a morphism $$ \rho : A \rightarrow K(W, 1) $$ together with an identification of the fiber $\rho ^{-1}(0)$ (by which we mean the homotopy fiber product $\{ 0\} \times _{K(W, 1)} A$) with $R$, $$ R\cong \rho ^{-1}(0). $$ To put this more briefly, an action of $W$ on $R$ is a fiber sequence $$ R\rightarrow A \rightarrow K(W,1). $$ If $T$ is another $n$-stack, and if $\rho$ is an action of $W$ on $R$ then we obtain an associated action $Hom (\rho , T)$ of $W$ on $Hom (R,T)$. It is given by the relative $Hom$, $$ Hom (A/K(W,1), T)\rightarrow K(W,1), $$ whose fiber is naturally identified with $Hom(R,T)$. {\bf Example 1:} If $W$ acts on a sheaf of sets $R$ in the usual sense, then this can be interpreted as an action in the above sense. The stack $A$ (which in this case is a $1$-stack) is the stack-theoretical quotient $R/W$ with its canonical principal $W$-bundle $R\rightarrow R/W$ which corresponds to a morphism $R/W \rightarrow K(W,1)$. {\bf Example 2:} The group ${\bf G}_m$ acts on $\widehat{TX}$ (resp. $TX$) over $X$ by scalar multiplication. Therefore ${\bf G}_m$ acts on $X_{Dol}$ (resp. $X_{UDol}$) and on $Hom (X_{Dol}, T)$ (resp. $Hom(X_{UDol}, T)$). In the case $T=K(G,1)$ this is the usual action of ${\bf G}_m$ on the moduli stack of principal Higgs bundles \cite{Hitchin} \cite{hbls} \cite{Moduli}. In the case $T=K({\cal O} , n)$ this action gives the decomposition of $H^i_{Dol}(X)$ into pieces $H^q(X, \Omega ^p_X)$. {\bf Example 3:} A different example, more closely related to what we are interested in, is the following. Suppose $$ 1\rightarrow V \rightarrow E \rightarrow W \rightarrow 1 $$ is an exact sequence of sheaves of groups on $Sch /{\bf C}$. This gives a fibration sequence $$ K(V,1)\rightarrow K(E,1)\rightarrow K(W,1), $$ hence by definition it is an action of $W$ on $K(V,1)$. More generally if $X$ is a sheaf of sets and if $V\rightarrow X$ is a sheaf of groups over $X$, then we can look at the relative Eilenberg-MacLane stack $K(V/X, 1)\rightarrow X$. Suppose $$ 1\rightarrow V \rightarrow E \rightarrow p^{\ast}(W)\rightarrow 1 $$ is an exact sequence of sheaves of groups on $X$ (where $p:X\rightarrow \ast$ denotes the projection). Then we obtain a fibration sequence $$ K(V/X, 1)\rightarrow K(E/X, 1)\rightarrow K(W,1), $$ the latter map being the composition of the map induced by the second map in the exact sequence, with the projection $$ K(p^{\ast}W /X,1)= K(W\times X/X, 1)= K(W, 1)\times X \rightarrow K(W,1). $$ Thus, again by definition, our exact sequence corresponds to an action of $W$ on $K(V/X,1)$ (lying over the trivial action of $W$ on $X$). \subnumero{Secondary classes} We first discuss classifying spaces. If $R$ is an $n$-stack then $Aut (R)$ is an $n$-stack with ``group'' structure, more precisely with the structure of a loop space. To make this statement precise, we construct a pointed $n+1$-stack $B\, Aut(R)$ with basepoint denoted $0$, with an equivalence $$ \Omega (B\, Aut(R), 0) \cong Aut (R). $$ Construct the presheaf of spaces $B^{\rm pre}\, Aut(R)$ as the realization of a simplicial $n$-stack $B_{\cdot}Aut(R)$ defined as follows: let $\overline{I}^{(i)}$ be the $1$-category with $i+1$ isomorphic objects $0,\ldots , i$ (which we shall call ``vertices''); then set $B_iAut(R)$ equal to the $n$-stack of morphisms $\overline{I}^{(i)}\rightarrow n\underline{STACK}$ sending the vertices to the object $R\in nSTACK$. Note that the component $B_iAut(R)$ is homotopic to $Aut(R)\times \ldots \times Aut(R)$. This simplicial $n$-stack may be interpreted as a presheaf of $n+1$-categories. Note that $Aut(R)$ is defined to be the $n$-stack of morphisms $\overline{I}\rightarrow n\underline{STACK}$ sending $0,1$ to $R$, and it is a stack of $n$-groupoids. It may safely be confused with a presheaf of $n$-truncated spaces, and $B_{\cdot}\, Aut(R)$ becomes a presheaf of simplicial spaces. The component simplicial spaces satisfy Segal's condition (cf \cite{Segal} \cite{Tamsamani} \cite{nCAT} \cite{effective}) so if we set $B^{\rm pre}\, Aut(R)$ equal to the realization into a presheaf of spaces then Segal's Proposition 1.5 \cite{Segal} implies that the natural map $$ Aut(R) = B_1Aut(R) \rightarrow \Omega (B^{\rm pre}\, Aut (R), 0) $$ is an equivalence (object-by-object). Finally, let $B\, Aut(R)$ be the $n+1$-stack associated to $B^{\rm pre}Aut(R)$. If $R$ is an $n$-stack of groupoids, the locally constant fibrations with fiber $R$ $$ R \rightarrow E \rightarrow S $$ are classified exactly by maps $S\rightarrow B\, Aut(R)$. In other words, given $S$ and $R$, the $n+1$-category of such fiber sequences is equivalent to the $n+1$-category $Hom (S, B\, Aut(R))$. In particular, an action of a sheaf of groups $W$ on an $n$-stack of groupoids $R$ is the same thing as a morphism $$ f:K(W,1)\rightarrow B\, Aut (R). $$ This leads to the notion of {\em characteristic classes} for the action: if ${\cal G}$ is a sheaf of groups and if $c\in H^i(B\, Aut (R), {\cal G} )$ then for any action we can pull back $c$ to obtain a class in $H^i(K(W,1), {\cal G} )$. One can also obtain {\em secondary invariants}, which only become defined when some primary invariants vanish. For example, suppose $c\in H^i(B\, Aut (R), {\cal G} )$ is a cohomology class corresponding to a map $$ c: B\, Aut(R) \rightarrow K({\cal G} , i). $$ Suppose that $f: K(W, 1)\rightarrow B\, Aut(R)$ is an action of $W$ on $R$, such that $c\circ f : K(W, 1)\rightarrow K({\cal G}, i)$ is trivial, i.e. homotopic to the constant map at the basepoint. Let $Fib(c)$ denote the homotopy fiber of the morphism $c$ over $0$. A choice of trivialization $\psi$ of $c\circ f$ gives rise to a morphism $$ f_{\psi} : K(W, 1)\rightarrow Fib(c). $$ The choices of trivialization $\psi$ are, up to homotopy, classified by $H^{i-1}(K(W,1), {\cal G} )$. If $c'$ is a class in $H^j(Fib(c), {\cal G} ')$ then the composition $c'\circ f_{\psi}$ gives a class in $H^j(K(W, 1), {\cal G} ')$. This is (a typical example of) a secondary characteristic class for the action. We don't get any further into the general theory of secondary classes, because insofar as the formalism is concerned, there is no difference between the present case of $n$-stacks and the classical topological version of the theory. Instead, we give a more explicit description of the secondary classes that we are interested in. These concern only classes of degrees $1$ and $2$ on $B\, Aut (R)$, so a more concrete discussion is possible. The first step is to replace $R$ by its truncation down to a $1$-stack, denoted $\tau _{\leq 1}R$. We fix a point $r\in R (Spec ({\bf C} ))$ (and denote also by $r$ its image in $\tau _{\leq 1}(R)$). Note that the ``group'' $Aut(R)$ acts on $\tau _{\leq 1}R$. The first of the primary classes comes from a morphism $$ Aut(R)\rightarrow Aut(\pi _0R). $$ On the left is an actual sheaf of groups ${\cal G}$, and this map corresponds to a class in $H^1(B\, Aut(R), {\cal G} )$. Suppose that we have an action by $W$ such that this class vanishes. This implies that $W$ fixes the point $r\in \pi _0(R)(Spec ({\bf C} ))$. (Vanishing of this class actually implies that $W$ fixes every point but for what follows we only need that it fixes $r$). Let $R^r$ denote the component of $R$ containing $r$, more precisely it is the fiber product $$ R^r:= \ast \times _{\pi _0(R)} R. $$ Under our preliminary vanishing hypothesis, we obtain an action of $W$ on $R^r$. Now $\tau _{\leq 1}(R^r)= K(G, 1)$ where $G= \pi _1(R,r)$ (the point $r$ is a basepoint defined over $Spec ({\bf C} )$ giving this trivialization of the $1$-gerb $\tau _{\leq 1}(R^r)$---the basepoint $r$ corresponds to $0\in K(G,1)$). For any group $G$ we have a fibration sequence $$ K(Center(G), 2)\rightarrow B\, Aut (K(G,1)) \rightarrow K(Out(G), 1). $$ In our case above, $W$ acts on $R^r$, hence on $\tau _{\leq 1}R^r=K(G,1)$. The map $$ K(W,1)\rightarrow B\, Aut(\tau _{\leq 1}R^r)= B\, Aut (K(G,1)) $$ thus projects first of all to a morphism of groups $$ W\rightarrow Out \left( \pi _1(R, r )\right) . $$ This is in a certain sense again a primary invariant. Suppose that this invariant vanishes. Then, given that $$ K(W,1)\rightarrow K(Out(G), 1) $$ is a pointed map, there is a canonical homotopy of this map to the constant map at the basepoint, so we canonically can identify our map $$ K(W,1)\rightarrow B\, Aut(\tau _{\leq 1}R^r) $$ as a map $$ K(W,1)\rightarrow K(Center \, \pi _1(R, r) , 2), $$ in other words as a class in $H^2(K(W,1), Center)$ where $Center$ is the center of $\pi _1(R, r)$. This is the secondary class we will be interested in calculating below. We can describe the secondary class a bit more concretely in the following way. Suppose $\alpha , \beta$ are elements of $W$ thought of as paths in $K(W,1)$. Choose trivializations of the fibration $$ R\rightarrow A\rightarrow K(W,1) $$ above the paths $\alpha$ and $\beta$. These trivializations lead in particular to liftings of our paths starting with the basepoint $r$, and ending at points we denote $\alpha ^{\ast}(r)$ and $\beta ^{\ast}(r)$. By the hypothesis of vanishing of primary classes, $\alpha ^{\ast}(r)$ and $\beta ^{\ast}(r)$ are homotopic to $r$. Thus we can choose homotopies which we denote $h_{\alpha}$ and $h_{\beta}$ (these are paths in $R$ joining $\alpha ^{\ast}(r)$ resp. $\beta ^{\ast}(r)$ to $r$). Applying the trivializations chosen above to these paths we obtain paths in $R$ $$ \alpha ^{\ast}(h_{\beta}): \alpha^{\ast} \beta^{\ast} (r) \rightarrow \alpha ^{\ast} (r), $$ $$ \beta ^{\ast}(h_{\alpha}): \beta^{\ast} \alpha^{\ast} (r) \rightarrow \beta ^{\ast} (r). $$ Finally, the commutativity of $W$ means that there is a torus obtained by attaching a $2$-cell along the commutator of $\alpha$ and $\beta$. Lifting this $2$-cell (more precisely, trivializing the family above this $2$-cell) provides a path between $\alpha^{\ast} \beta^{\ast} (r)$ and $\alpha^{\ast} \beta^{\ast} (r)$. Combining these all together we get a $5$-sided loop based at $r$; this is the secondary class evaluated on $\alpha \wedge \beta$. The loop is in the center of $\pi _1(R,r)$ because of the fact that we chose trivializations of the family over our paths (or $2$-cell) rather than just liftings starting at the basepoint. A slightly different and more geometric way of looking at this is to look at the torus in $K(W,1)$ given by the commutator $2$-cell for $\alpha$ and $\beta$. Attaching two $2$-cells to the torus, one along $\alpha$ and the other along $\beta$, gives a $2$-sphere. The vanishing of the primary classes means that the fibration $A\rightarrow K(W,1)$ can be extended across these new $2$-cells, so we get a family over $S^2$. Trivializing over the northern and southern hemispheres, the family is determined by a morphism from the equator to $Aut(R)$: this element of $\pi _1(Aut(R))=Center$ is the image of $\alpha \wedge \beta$ under the secondary class. \numero{The secondary Kodaira-Spencer map} We now come to the situation which gives a ``secondary Kodaira-Spencer map''. Suppose $f:X_S\rightarrow S$ is a smooth projective morphism. Fix a basepoint $s\in S$ and denote by $X$ the fiber of $X_S$ over $s$. Define $$ W:= T(S)_s $$ which is a vector space, thus unipotent abelian group scheme, considered as a sheaf of groups on $Sch /{\bf C} $. We have an exact sequence (of unipotent abelian group schemes i.e. vector bundles over $X$) $$ 0\rightarrow TX \rightarrow T(X_S)|_X \rightarrow f^{\ast}(T(S)_s)\rightarrow 0. $$ This is an action of $W=T(S)_s$ on the stack $K(TX/X, 1)= X_{UDol}$. In particular, for any $n$-stack $T$ we obtain an action of $W$ on $Hom (X_{UDol}, T)$. The primary invariant in this situation is an action of $W$ on the sheaf of sets \linebreak $\pi _0Hom (X_{UDol}, T)$. For example, if $T=K({\cal O} , n)$ then by Proposition \ref{calcDol} $$ \pi _0Hom (X_{UDol}, T)= H^n_{Dol}(X)= \bigoplus _{p+q=n}H^q(X, \Omega ^p_X) $$ and we obtain an action of $W=T(S)_s$ on $H^n_{Dol}(X)$. This action is of course just the usual Kodaira-Spencer map which decomposes into components $$ T(S)_s \rightarrow Hom \left( H^q (X, \Omega ^p_X), H^{q+1}(X, \Omega ^{p-1}_X) \right) . $$ To obtain secondary invariants, we proceed as described above, using vanishing of the primary invariants if we want to (but bearing in mind that the secondary invariants will then only be defined when the primary invariants vanish). For example, suppose that $T$ is an $n$-stack and suppose $\eta \in Hom (X_{UDol}, T)$ such that the point $[\eta ] \in \pi _0Hom (X_{UDol}, T)$ is fixed by the action of $W$. Then $W$ acts on the connected $n$-stack $Hom ^{\eta}(X_{UDol}, T)$ which is the connected component containing $\eta$. We obtain a morphism $$ K(W,1)\rightarrow B\, Aut (Hom ^{\eta}(X_{UDol}, T)), $$ and as remarked above, cohomology classes on the right can be pulled back to give classes on $K(W,1)$. At this point we refer to the Breen calculations \cite{Breen2}. In Appendix II we prove a relative version in characteristic zero which was stated in \cite{kobe}; the reader may refer there for the general statement. In our case, as $W$ is represented by a finite-dimensional vector space (in particular, $W\cong {\cal O} ^a$) we have $$ H^i(K(W,1), {\cal O} ) = \bigwedge ^iW^{\ast} $$ where $W^{\ast} = Hom (W, {\cal O} )$ is the sheaf represented by the dual vector space. To get down to the concrete example we would like to consider, we boil things down a bit farther, following the discussion at the end of the previous section. Namely instead of looking at the full $Hom ^{\eta}(X_{UDol}, T)$ we truncate it down to a $1$-stack (which is connected, too) by looking at $$ \tau _{\leq 1}Hom ^{\eta}(X_{UDol}, T). $$ In the very presentable case, this automatically has a base point over $Spec ({\bf C} )$ (but also we have chosen a basepoint $\eta$) so it is equivalent to something of the form $K(G,1)$ with $$ G= \pi _1(Hom (X_{UDol}, T), \eta ). $$ Again, in the very presentable case (i.e. if $T$ is very presentable which implies the same for $Hom ^{\eta}(X_{UDol}, T)$) then $G$ is an affine algebraic group. In our example below $G$ will itself be a vector space. The first invariant is $$ W\rightarrow Out \left( \pi _1(Hom (X_{UDol}, T), \eta )\right) . $$ Suppose that this invariant vanishes. Then we get a map $$ K(W,1)\rightarrow K(Center \, \pi _1(Hom (X_{UDol}, T), \eta ) , 2), $$ which may be interpreted as a class in $H^2(K(W,1), Center)$ where $Center$ is the center of $\pi _1(Hom (X_{UDol}, T), \eta )$. If $T$ is very presentable then this center will be an affine abelian group scheme. This class in $H^2$ is the class we are interested in calculating below. The secondary class can be described concretely by choosing homotopies trivializing the primary classes and combining them together using the commutativity homotopy for the action of $W$, as at the end of the previous section. \numero{The complexified $2$-sphere} We discuss in more detail the example of a $3$-stack $T$ for which we make our calculation. Recall from ``standard topology'' how to describe the rational homotopy type of $S^2$. The only nontrivial stages in the Postnikov tower are $K({\bf Q} , 2)$ and $K({\bf Q} , 3)$. Thus the rational homotopy type is described by the fibration sequence $$ K({\bf Q} , 3) \rightarrow S^2 \otimes {\bf Q} \rightarrow K({\bf Q} , 2). $$ In turn, the classifying space for fibrations with fiber $K({\bf Q} , 3)$ is the base for the universal fibration $$ K({\bf Q} , 3)\rightarrow \ast \rightarrow K({\bf Q} , 4). $$ Thus the above fibration is determined by a morphism $$ K({\bf Q} , 2)\rightarrow K({\bf Q} , 4), $$ in other words a class $\sigma \in H^4(K({\bf Q} , 2), {\bf Q} )$. The classical calculations give $$ H^{2m}(K({\bf Q} , 2), {\bf Q} ) = Sym ^m({\bf Q} ). $$ In particular $$ H^4(K({\bf Q} , 2), {\bf Q} )\cong Sym ^2({\bf Q} ) = {\bf Q} . $$ Up to change of basis element for $\pi _3$, there are only two possibilities: either $\sigma = 0$ or $\sigma \neq 0$. The case $\sigma = 0$ corresponds to the direct product $K({\bf Q} , 2)\times K({\bf Q} , 3)$ but in this case $H^3$ would be nonzero, whereas $H^3(S^2\otimes {\bf Q} , {\bf Q} )=0$. Therefore we must have $\sigma \neq 0$, in other words $\sigma$ is the cup product $\eta \cup \eta$ of the canonical class $\eta \in H^2(K({\bf Q} , 2), {\bf Q} )$ with itself. In view of this discussion, we can do the same with very presentable stacks. We will construct $T= S^2\otimes {\bf C}$ with two stages $K({\cal O} , 2)$ and $K({\cal O} , 3)$ in the Postnikov tower. We construct it as the pullback of the universal fibration $$ K({\cal O} , 3)\rightarrow \ast \rightarrow K({\cal O} , 4) $$ by a morphism $$ \sigma : K({\cal O} , 2) \rightarrow K({\cal O} , 4). $$ The Breen calculations say that the cohomology of $K({\cal O} , n)$ with coefficients in ${\cal O}$ has the same answer as the cohomology of $K({\bf Q} , n)$ with coefficients in ${\bf Q}$. In other words, $$ H^{2m} (K({\cal O} , 2), {\cal O} ) = Sym ^m_{{\cal O}} ({\cal O} ) \cong {\cal O} $$ and $$ H^4(K({\cal O} , 2), {\cal O} )= {\cal O} . $$ Let $\sigma$ be the generator of ${\cal O} (Spec \, {\bf C} ) = {\bf C}$. This gives a map $$ \sigma : K({\cal O} , 2)\rightarrow K({\cal O} , 4). $$ In terms of cohomology operations, a morphism ${\cal F} \rightarrow K({\cal O} , 2)$ corresponds to a cohomology class $\eta \in H^2({\cal F} , {\cal O} )$ and the composition of such a map with $\sigma$ corresponds to the cup-product square $\eta \cup \eta \in H^4({\cal F} , {\cal O} )$. Now set $$ T:= S^2\otimes {\bf C} := K({\cal O} , 2) \times _{K({\cal O} , 4)} \ast , $$ (this choice of $T$ shall be in vigor for the rest of the paper unless explicitly mentionned otherwise). For any $n$-stack ${\cal F}$, a morphism $$ {\cal F} \rightarrow T $$ corresponds to a pair $(\eta , \varphi )$ where $\eta : {\cal F} \rightarrow K({\cal O} , 2)$ and $\varphi$ is a homotopy between $$ \sigma \circ \eta = \eta \cup \eta : {\cal F} \rightarrow K({\cal O} , 4) $$ and the constant map at the basepoint $$ \underline{o}: {\cal F} \rightarrow K({\cal O} , 4). $$ Alternatively, $\varphi$ may be thought of as a section of the fibration $$ K({\cal O} , 3) \rightarrow {\cal F} \times _{K({\cal O} , 4)} \ast \rightarrow {\cal F} . $$ A first remark is that for a given class $\eta \in H^2({\cal F} , {\cal O} )$ there exists a lifting to a map ${\cal F} \rightarrow T$ if and only if $\eta \cup \eta = 0$ in $H^4({\cal F} , {\cal O} )$. If this cup-product is zero so that there exists one lifting, then the fiber of the map $$ Hom ({\cal F} , T) \rightarrow Hom ({\cal F} , K({\cal O} , 2)) $$ over the point $\eta$ is equivalent to $Hom ({\cal F} , K({\cal O} , 3))$ This is because if there exists one lifting then we can choose a lifting to trivialize the fibration $$ {\cal F} \times _{K({\cal O} , 4)} \ast \rightarrow {\cal F} $$ i.e. to make this fibration equivalent to $$ {\cal F} \times K({\cal O} , 3) \rightarrow {\cal F} $$ and then the stack of other liftings is just $Hom ({\cal F} , K({\cal O} , 3))$. {\bf Example:} We give an example of a topological space $Y$ which admits no nonconstant maps to the actual $2$-sphere $S^2$ but which admits maps to the complexified $2$-sphere $S^2\otimes {\bf C}$. Construct $Y$ by taking a wedge of two $2$-spheres and adding on a $4$-cell via an attaching map $$ f: S^3\rightarrow S^2 \vee S^2. $$ Note that $$ \pi _3(S^2 \vee S^2)= Sym ^2 \pi _2(S^2\vee S^2) = {\bf Z} ^3, $$ and the class of the attaching map $f$ determines the cup product. We can think of the class of $f$ as a symmetric $2\times 2$ matrix, which can be chosen arbitrarily. Choose the matrix to be diagonal with $(r,s)$ on the diagonal. The two obvious classes $e,f\in H^2(S^2\vee S^2,{\bf Q} )$ persist as classes in $H^2(Y,{\bf Q} )$. Note that $H^4(Y,{\bf Q} ) \cong {\bf Q} $ and we have the formulae $$ e\cup e = r, \;\; f\cup f = s,\;\; e\cup f = 0. $$ Now a map $Y\rightarrow S^2$ (or even to $S^2\otimes {\bf Q}$) corresponds to a class $$ \eta = ae +bf \in H^2(Y,{\bf Q} ) $$ with $a,b\in {\bf Q} $, such that $\eta \cup \eta =0$. The lifting to a map into $S^2$ is unique because in our case $H^3(Y,{\bf Q} )=0$. However, if $r$ and $s$ are chosen to be relatively prime and having no square prime factors, then the equation $$ \eta \cup \eta = a^2r + b^2s = 0 $$ doesn't have any nonzero solutions with $a,b\in {\bf Q}$ (the same holds if $r$ and $s$ have the same sign). Thus there are no nontrivial maps $Y\rightarrow S^2$. On the other hand, the above equation always has nonzero complex solutions, so there is always a nontrivial map $Y\rightarrow S^2\otimes {\bf C} $. \numero{Our example} We now come down to the example which we would like to calculate. Fix $T= S^2\otimes {\bf C}$ as defined above. Let $Z$ be a smooth projective surface with $H^1(Z, {\cal O} ) = 0$ (hence $H^1(Z, {\bf C} )= H^3(Z, {\bf C} )=0$). Let $P\in Z$ and let $X$ be the blow-up of $Z$ at $P$. Let $W\subset H^1(X, TX)$ be the rank two subspace of deformations of $X$ corresponding to moving the point $P$ which is blown up (thus canonically $W\cong T(Z)_P$). Then $W$ acts on $X_{UDol}$, via the exact sequence $$ 0\rightarrow TX\rightarrow T(Tot)|_X \rightarrow W\otimes _{{\bf C}}{\cal O} _X \rightarrow 0 $$ where $Tot$ refers to the total space of the family. Let $$ R:= Hom (X_{UDol}, T). $$ Then $W$ acts on $R$ and we will look at secondary classes for this action. Because of our hypothesis $H^3_{Dol}(X)=0$, the map $T\rightarrow K({\cal O} , 2)$ induces an injection $$ \pi _0(R) \hookrightarrow H^2_{Dol}(X) $$ with the image the sheaf represented by the subscheme defined by the equation $\eta \cup \eta = 0$. The variation of Hodge structure of $H^2(X, {\bf C} )$ parametrized by $P\in Z$ as we move the point which is blown up, is trivial. Thus the Kodaira-Spencer class is trivial, in other words $W$ acts trivially on $H^2_{Dol}(X)$ and hence it acts trivially on $\pi _0(R)$. If we fix $\rho = (\eta , \varphi )\in R(Spec \, {\bf C} )$ then we obtain a secondary class $$ \kappa \in H^2(K(W, 1), \pi _1(R, \rho )). $$ The first task is to calculate $\pi _1(R, \rho )$ (which will be a group scheme over $Spec ({\bf C} )$ since $\rho$ is defined over $Spec ({\bf C} )$). Recall that we have the fibration $$ R\rightarrow Hom (X_{UDol}, K({\cal O} , 2)) $$ whose fiber over a point in the target is either empty or else equivalent to \newline $Hom (X_{UDol}, K({\cal O} , 3))$. Thus the long exact sequence for this fibration gives $$ {\bf C} = H^0_{Dol}(X) \rightarrow H^2_{Dol}(X)\rightarrow \pi _1(R, \rho ) \rightarrow H^1_{Dol}(X) = 0. $$ the first term being $\pi _2Hom (X_{UDol}, K({\cal O} , 2))$, the second term being \newline $\pi _1Hom (X_{UDol}, K({\cal O} , 3))$, and so on. {\bf Claim:} The connecting morphism $H^0_{Dol}(X) \rightarrow H^2_{Dol}(X)$ in the above exact sequence is multiplication by a nonzero multiple of $\eta$. {\em Proof:} For any $n$-stack $Y$ we can define $Hom (Y, T)$ and look at the long exact sequence for the fibration $$ Hom(Y,T)\rightarrow Hom(Y,K({\cal O} , 2)). $$ The connecting morphism will be functorial in $Y$. Apply this to $Y=T$ itself; then for any other $Y$ (such as $X_{UDol}$ considered in the claim) the connecting morphism for the long exact sequence at a basepoint $Y\rightarrow T$ is obtained by pulling back the connecting morphism for $Hom (T,T)$ at the identity map. In other words, looking at $Y=T$ gives a universal version of the connecting morphism. On the other hand, note that $$ \pi _1 Hom (T, K({\cal O} , 3))=H^2(T, {\cal O} )={\cal O} $$ and $$ \pi _2Hom (T,K({\cal O} , 2)) = H^0(T,{\cal O} )={\cal O} . $$ Thus the universal connecting morphism is a scalar constant $C$, and for any $\rho : Y\rightarrow T$ the connecting morphism for $Hom (Y,T)$ based at $\rho$ fits into the diagram $$ \begin{array}{ccc} H^0(T,{\cal O} ) & \rightarrow & H^2(T, {\cal O} )\\ \downarrow && \downarrow \\ H^0(Y,{\cal O} ) & \rightarrow & H^2(Y, {\cal O} ). \end{array} $$ The vertical maps are those induced by $\rho$. It follows that the connecting map for $Hom (Y,T)$ based at $\rho$ is the same constant $C$ multiplied by the class $\eta$ which is the image of $\rho$ in $H^2(Y,{\cal O} )$. For the claim, apply this to $Y=X_{UDol}$. To finish proving the claim we just have to show that $C$ is nonzero. But if $C$ were zero then the generator for $\pi _2(Hom (T,K({\cal O} , 2)), 1_T)$ would lift to an element of $\pi _2(Hom(T,T), 1_T)$. This would give a map $$ S^2 \times T \rightarrow T $$ (where $S^2$ denotes the constant presheaf with values $S^2$), which is nontrivial on $S^2\times \{ 0\} $ and $\{ 0\} \times T$. This map would correspond to an element $\mu \in H^2(S^2\times T, {\cal O} )$ with $\mu \cup \mu = 0$. But we have $$ H^2(S^2\times T, {\cal O} )= H^2(S^2, {\cal O} )\oplus H^2(T, {\cal O} ) = {\cal O} \oplus {\cal O} $$ and the cup product of the two components is nontrivial (by K\"unneth). Therefore it is impossible to have a class $\mu$ which is nontrivial in both components but with $\mu \cup \mu = 0$. This contradiction implies that $C\neq 0$, giving the claim. \hfill $\Box$\vspace{.1in} With the claim we obtain $$ \pi _1(R,\rho ) \cong H^2_{Dol}(X)/(\eta ). $$ Thus our characteristic class $\kappa $ becomes a map $$ \kappa : \bigwedge ^2(W) \rightarrow H^2_{Dol}(X)/(\eta ). $$ Next, we choose $\eta$ (which fixes the choice of $\rho$ up to homotopy). Let $E$ be the exceptional divisor on $X$ and let $H$ denote the pullback of an ample divisor on $Z$ not meeting the point $P$. Let $[E]$ and $[H]$ denote their Chern classes in $H^1(X, \Omega ^1_X)\subset H^2_{Dol}(X)$. We set $$ \eta = m[E] + n[H] \in H^2_{Dol}(X). $$ We have to choose $n$ and $m$ so that $\eta \cup \eta = 0$. Note that $H^4_{Dol}(X)\cong {\bf C} $ with natural morphism given by the residue map, normalized so that the cohomology class of a point is equal to $1$. Via this isomorphism, $[E] \cup [E] = E.E = -1$ and $[H] \cup [H] = H.H\in {\bf Z} $. Note that $[E]\cup [H]=0$ since the two divisors don't intersect. Thus $$ \eta \cup \eta = n^2H.H - m^2. $$ We choose $m,n$ so that this is equal to $0$. {\bf Remark:} If $H.H$ is not the square of an integer, then the pair $(m,n)$ can not be chosen in ${\bf Q} ^2$, and in particular our map will not exist as a topological map $X^{\rm top} \rightarrow S^2$. However, our map will exist as a map from the constant presheaf with values $X^{\rm top}$, to $S^2\otimes {\bf C} $. (This is a heuristic remark since in the present paper we don't treat the question of the relationship between Betti cohomology and Dolbeault cohomology). Here is the result of the calculation which will be done below. \begin{theorem} \mylabel{calculation} With the above choices of $T$, $X$, $R:= Hom (X_{UDol}, T)$, $\eta = m[E] + n[H]$ ($\eta \neq 0$), and hence $\rho \in R(Spec \, {\bf C} )$, the secondary Kodaira-Spencer class $$ \kappa : \bigwedge ^2 W \rightarrow \pi _1(R, \rho )= H^2_{Dol}(X)/(\eta ) $$ lands in $H^2(X, {\cal O} ) \subset H^2_{Dol}(X)/(\eta )$ and the map $$ \bigwedge ^2 W \rightarrow H^2(X,{\cal O} ) $$ is dual (using Serre duality, the isomorphism $H^0(X, \Omega ^2_X)\cong H^0(Z, \Omega ^2_Z)$, and the isomorphism $W^{\ast} \cong T^{\ast}(Z)_P$) to $m^2ev_P$ where $$ ev_P:H^0(Z, \Omega ^2_Z) \rightarrow \bigwedge ^2T^{\ast}(Z)_P $$ is the evaluation morphism. In particular, $\kappa \neq 0$ if $h^{2,0}(X)=h^{2,0}(Z) > 0$ and $m\neq 0$. \end{theorem} Before getting on with the proof in \S 8 below, we make a few general remarks about this result. A similar thing can be stated for the ``Dolbeault homotopy type of $X$''. One way of defining this (which wouldn't be the historical way, though) is as the $1$-connected very presentable $n$-stack $\Sigma$ representing the very presentable shape of $X_{Dol}$ (cf Theorem \ref{representable1} in Appendix II below). In this point of view, we get an action of $W$ on the Dolbeault homotopy type. The theorem says that this action of $W$ is nontrivial. Note however that the action of $W$ on the homotopy group sheaves (which are the homotopy groups of $X$ tensored with ${\bf C}$) will be trivial. It is certainly possible to define the action of $W$, and to make the same calculation as below to show that the action is nontrivial, using the algebra of forms $$ A^{\cdot} _{Dol}(X)= (\bigoplus _{p,q} A^{p,q}(X), \overline{\partial} ). $$ In fact, my first heuristic version of the calculation was done using forms. However, the technical details relating a differential-forms version of nonabelian cohomology, with the version presented here, seem for the moment somewhat difficult, so we restrict in the present paper to an algebraic version of the calculation. The secondary class $\kappa$ is a natural map, so it doesn't really have any choice other than to be a multiple of the dual of the evaluation map $ev_P$. The only question is whether this multiple is nonzero or not. Here is a heuristic global argument to see why, in principle, the constant should be nonzero. Let ${\cal X} \rightarrow Z$ be the total space of the family of blow-ups of points moving in $Z$. It is obtained by blowing up the diagonal in $Z\times Z$. The secondary Kodaira-Spencer class we calculate here is (or should be, at least) the $(2,0)\times (0,2)$ Hodge component (i.e. the component of type $(2,0)$ on the base and $(0,2)$ on the fiber) of the following globally defined invariant. Fix $\eta \in H^2(X,{\bf C} )$, which is invariant under the monodromy since the monodromy is trivial (the base $Z$ being simply connected); the degeneration of the Leray spectral sequence says that $\eta$ comes from restriction of a global class $\tilde{\eta} \in H^2({\cal X} , {\bf C} )$; the cup product $\tilde{\eta} \cup \tilde{\eta}$ restricts to zero on the fibers, so it lies in the next step for the Leray filtration, which in our case is $H^2(Z, H^2(X, {\bf C} ))$. This cup product is therefore a globally defined class. It is the obstruction to extending $\rho : X^{\rm top}\rightarrow T$ to a map ${\cal X} ^{\rm top}\rightarrow T$. The $(2,0)\times (0,2)$ component of this class, which is a holomorphic $2$-form on $Z$ with coefficients in $H^2(X, {\cal O} )$, {\em should} give $\kappa$ when evaluated at the point $P\in Z$. I don't currently have a proof of this, though. In our example, it is relatively easy to see by looking at the cohomology class of the diagonal that the $(2,0)\times (0,2)$ component of the global class is nonzero, and in fact it is the identity matrix (via Serre duality). Thus if one could prove the above statement that the global class gives $c$ under evaluation at $P\in Z$, then this would prove that $\kappa\neq 0$ when $H^2(X, {\cal O} )\neq 0$. With the previous paragraphs as heuristic argument, the result that $\kappa\neq 0$ doesn't look all that surprising. Still, it means that the ``variation of nonabelian Hodge structure'' \footnote{ This terminology is put in quotes because the current discussion of Dolbeault cohomology is only a first step towards defining what a ``variation of nonabelian Hodge structure'' is.} on the family of $Hom (X^{\rm top}, T)$, when $X$ is a variable fiber in the family ${\cal X} \rightarrow Z$, is nontrivial, and even nontrivial for infinitesimal reasons. The base of this ``variation'' is $Z$ which is simply connected. In particular, the variations of mixed Hodge structure on the homotopy groups (or anything else you could think of) are trivial. From a topological point of view, it is never a surprise to find a family where the homotopy groups are constant but the family nontrivial. On the other hand, this goes against the commonly held intuition for projective algebraic varieties that ``formality means that everything is determined by the cohomology ring'': in the example ${\cal X} \rightarrow Z$, the locally constant family---parametrized by $Z^{\rm top}$---of cohomology rings of the fibers $X$ is trivial. What remains true of course is that the topology of the family is determined by the cohomology ring of the total space ${\cal X}$. Our secondary Kodaira-Spencer invariant is a local invariant which contributes to nontriviality of the global cohomology ring of the total space of the family. Our class detects the motion of a point $P\in Z$ exactly when $H^0(Z,\Omega ^2_Z)\neq 0$. This seems to fit in with the standard intuition that $H^0(Z,\Omega ^2_Z)\neq 0$ causes the class group of zero cycles to be big (Mumford's and Clemens' results, Bloch conjecture etc. cf Voisin \cite{Voisin}). I don't see a precise connection, though. \numero{The calculation} We keep the above notations $T$, $P\in Z$, $X$, $E$, $H$, $\eta$, $\rho$. We establish some more: let $N$ be an affine neighborhood of $P$ in $Z$ such that $TZ$ is trivialized over $N$. Assume that $N$ doesn't meet the divisor image of $H$ in $Z$. Let $\alpha$ and $\beta$ denote basis sections in $TZ(N)$. Let $B$ be the inverse image of $N$ in $X$, and let $C=X-E \cong Z-\{ P\}$. Then put $A:= B\cap C$. Note that $\{ B,C\}$ is an open covering of $X$. We can write $$ X= B\cup ^AC $$ (this is true as a pushout of sheaves of sets on $Sch/{\bf C} $). Similarly we have $$ X_{UDol}= B_{UDol} \cup ^{A_{UDol}}C_{UDol}. $$ The basis vectors $\alpha , \beta$ give cocycles for elements in $H^1(X, TX)$ (actually in the \v{C}ech cohomology relative to our covering) and it is easy to see that these cocycles project to basis elements of our $2$-dimensional space $W$. We denote the basis vectors of $W$ also by $\alpha$ and $\beta$. We can describe the action of $W$ on $X_{UDol}$ concretely in the following way. Set $K:= K(W, 1)$ with basepoint denoted $0\in K$. Then the trivialization of $TZ|_N$ gives an equivalence $$ A_{UDol} \cong A\times K. $$ There is a group structure on $K$, that is a morphism $K\times K\rightarrow K$ corresponding to the addition on $W$, and this gives $$ A\times K\times K\rightarrow A\times K, $$ which we can rewrite as $$ \mu : A_{UDol}\times K \rightarrow A_{UDol}. $$ Putting the identity $1_K$ in the second variable we get a map $$ \Phi: A_{UDol} \times K \rightarrow A_{UDol} \times K $$ which is an equivalence. Note also that $\Phi |_{A_{UDol} \times 0}$ is the identity of $A_{UDol}$. Let $i,j$ be the inclusions from $A_{UDol}$ to $B_{UDol}$ and $C_{UDol}$ respectively. Then use the inclusions $$ (i\times 1_K)\circ \Phi : A_{UDol}\times K \rightarrow B_{UDol}\times K $$ and $$ j\times 1_K: A_{UDol}\times K \rightarrow C_{UDol}\times K $$ to construct the pushout $$ P:= B_{UDol}\times K\cup ^{A_{UDol}\times K}C_{UDol}\times K. $$ This comes equipped with a morphism $$ P\rightarrow K $$ and the fiber over the basepoint $0\in K$ is just $$ B_{UDol}\cup ^{A_{UDol}}C_{UDol} = X_{UDol}. $$ Therefore according to our definition, $P\rightarrow K$ is an action of $W$ on $X_{UDol}$. The corresponding action of $K$ on $R:= Hom (X_{UDol}, T)$ is by definition $$ Hom (P/K, T)\rightarrow K. $$ Using the Mayer-Vietoris principle we get $$ Hom (P/K, T) = $$ $$ Hom (B_{UDol}\times K/K,T) \times _{Hom (A_{UDol}\times K/K,T)} Hom (C_{UDol}\times K/K,T). $$ However, note that $$ Hom (B_{UDol}\times K/K,T) = Hom (B_{UDol}, T) \times K $$ and similarly for the other factors. The morphism $$ Hom (C_{UDol}, T) \times K\rightarrow Hom (A_{UDol}, T) \times K $$ induced by $j\times 1_K$ is just the product of the morphism induced by $j$, with $1_K$. (On the other hand, the same is not true of the first morphism in the fiber product, as it is induced by $(i\times 1_K)\circ \Phi$.) We can now write $$ Hom (P/K, T) = $$ $$ (Hom (B_{UDol},T)\times K) \times _{Hom (A_{UDol},T)} Hom (C_{UDol},T). $$ The first morphism in the fiber product is the composition of the product-compatible morphism $$ (j^{\ast} \times 1_K): Hom (B_{UDol},T)\times K \rightarrow Hom (A_{UDol},T)\times K, $$ with the morphism $$ \Psi : Hom (A_{UDol},T)\times K\rightarrow Hom (A_{UDol}, T). $$ This map is equivalent (by the definition of internal $Hom$) to $$ Hom (A_{UDol},T)\rightarrow Hom (K, Hom (A_{UDol},T)), $$ which in turn is equivalent to a map $$ Hom (A_{UDol},T)\rightarrow Hom (A_{UDol}\times K, T), $$ this latter being induced by our action $\mu : A_{UDol}\times K \rightarrow A_{UDol}$. The first step is to notice that the map $\rho |_{A_{UDol}}$ from $A_{UDol}$ to $T$ factors through a map $$ h: A_{UDol} \rightarrow K({\cal O} , 3)\rightarrow T. $$ This factorization is given by the fact that $\eta$ is a class in $H^2((X_{Dol}, A_{Dol}), {\cal O} )$, in other words we are given a trivialization of $\eta$ over $A_{Dol}$. Recall that we write $\rho = (\eta , \varphi )$ where $\eta : X_{UDol}\rightarrow K({\cal O} , 2)$ and $\varphi$ is a section of the pullback bundle $$ L_{\eta}:= X_{UDol} \times _{K({\cal O} , 4)} \ast $$ which is a bundle with fiber $K({\cal O} , 3)$ over $X_{UDol}$. The section $\varphi$ determines a trivialization $$ L_{\eta} \cong X_{UDol}\times K({\cal O} , 3) $$ such that $\varphi$ corresponds to the $0$-section. This trivialization is uniquely determined by the condition that it be compatible with the structure of ``principal bundle'' under the ``group'' $K({\cal O} , 3)$. We adopt the following strategy for calculating the secondary class. We will look at new $n$-stacks $P^i\rightarrow K$ with $K$-maps $P^i \rightarrow P$, and points $\rho ^i\in P^i_0$ (where $P^i_0$ means the fiber of $P^i$ over $0\in K$), such that $\rho ^i$ maps to $\rho$. We have to arrange so that the primary class is trivial, in other words that the class of $\rho ^i$ in $\pi _0(P^i_0)$ should be invariant under the action of $W=\pi _1(K)$. We also have to insure that the other primary class, the action of $W$ on $\pi _1(P^i_0, \rho ^i)$ by outer automorphisms, should be trivial. In this case, we obtain a secondary class for $\rho ^i$, which is an element of $$ H^2(K, \pi _1(P^i_0, \rho ^i)) $$ and this secondary class maps to our class for $P$. First of all, let $\ast \rightarrow Hom (B_{UDol}, T)$ be the morphism corresponding to the point $\rho |_{B_{UDol}}$. We get $$ K= \ast \times K \rightarrow Hom (B_{UDol}, T)\times K. $$ Thus we obtain a morphism $$ P^1:= K\times _{Hom (A_{UDol}, T)}Hom (C_{UDol}, T) \rightarrow $$ $$ (Hom (B_{UDol}, T)\times K) \times _{Hom (A_{UDol}, T)}Hom (C_{UDol}, T) =P. $$ This morphism is compatible with the projections to $K$. The first morphism $$ u: K\rightarrow Hom (A_{UDol}, T) $$ in the fiber product is obtained by the composition $$ K \rightarrow Hom (B_{UDol}, T)\times K\stackrel{(i\times 1_K)\circ \Phi}{\rightarrow} $$ $$ Hom (A_{UDol}, T)\times K \stackrel{p_1}{\rightarrow} Hom (A_{UDol}, T). $$ Thus $u$ corresponds to the map $K\times A_{UDol} \rightarrow T$ obtained by composing $$ K\times A_{UDol} \stackrel{\mu }{\rightarrow} A_{UDol} \stackrel{\rho}{\rightarrow} T. $$ The map $u$ factors through a morphism $$ K\rightarrow \Gamma (A_{UDol}, L_{\eta})\rightarrow Hom (A_{UDol}, T) $$ (technically speaking what should enter into the above notation is $L_{\eta}|_{A_{UDol}}$ but for brevity we omit the restriction since it is implicitly determined by the notation $\Gamma (A_{UDol}, -)$.) The factorization comes about because the composition $$ A_{UDol} \stackrel{\rho}{\rightarrow} T \rightarrow K({\cal O} , 2), $$ is given as the constant map at the basepoint. Thus pulling back by $$ A_{UDol}\times K \rightarrow A_{UDol} \rightarrow K({\cal O} , 2) $$ is again constant at the basepoint so the map $$ K\rightarrow Hom( A_{UDol}, T) $$ factors through a map $$ \tilde{u}:K\rightarrow Hom( A_{UDol}, K({\cal O} , 3)). $$ Following above, $\tilde{u}$ corresponds to the composition $$ K\times A_{UDol}\stackrel{\mu}{\rightarrow} A_{UDol} \stackrel{h}{\rightarrow } K({\cal O} , 3), $$ which we can write as $\tilde{u} = \mu ^{\ast}(h)$. Now we obtain a morphism (over $K$) $$ P^2:= K\times _{\Gamma (A_{UDol}, L_{\eta})}\Gamma (C_{UDol}, L_{\eta}) $$ $$ \rightarrow K\times _{Hom (A_{UDol}, T)}Hom (C_{UDol}, T) = P^1. $$ We have used the section $\varphi$ of $L_{\eta}$ to obtain a trivialization $$ L_{\eta} \cong X_{UDol} \times K({\cal O} , 3). $$ Via this equivalence the section corresponding to $\rho$ (that is, the section $\varphi$) corresponds to the zero-section. Using the trivialization given by $\varphi$ we can write $$ \Gamma (A_{UDol}, L_{\eta})\cong Hom (A_{UDol}, K({\cal O} , 3)) $$ ({\em Caution:} this is not the same trivialization as given by saying that $A_{UDol}$ maps to the basepoint of $K({\cal O} , 2)$ so the pullback fibration $L_{\eta}$ is trivial over $A_{UDol}$; these two trivializations differ by translation by $h$, a point which will come up below); and $$ \Gamma (C_{UDol}, L_{\eta})\cong Hom (C_{UDol}, K({\cal O} , 3)). $$ These are compatible with the restriction $j^{\ast}$ so we can write $$ P^2 = K\times _{Hom (A_{UDol}, K({\cal O} , 3))} Hom (C_{UDol}, K({\cal O} , 3)), $$ with the second morphism in the fiber product being the restriction $j^{\ast}$ acting on maps to $K({\cal O} , 3)$. We have to re-calculate the first morphism in the fiber product $$ a:K\rightarrow Hom (A_{UDol}, K({\cal O} , 3)). $$ It is no longer equal to $\tilde{u}=\mu ^{\ast}(h)$, because when we set the section $\varphi$ of $L_{\eta}$ equal to the zero-section to get an equivalence between $L_{\eta}$ and $K({\cal O} , 3)$, this made a translation on $L_{\eta} |_{A_{UDol}}$---which was already trivial due to the fact that $\eta |_{A_{UDol}}=0$---this translation has the effect of setting $h$ equal to the $0$-section. This translation gives us the formula $$ a = \mu ^{\ast}(h)-p_2^{\ast}(h) $$ where $p_2^{\ast}(h)$ is the map $K\rightarrow Hom (A_{UDol}, K({\cal O} , 3))$ corresponding to the composition $$ K\times A_{UDol} \stackrel{p_2}{\rightarrow} A_{UDol} \stackrel{h}{\rightarrow} K({\cal O} , 3). $$ Note that $p_2$ denotes the second projection. The minus sign in the equation for $a$ is subtraction using the ``abelian group'' (i.e. $E_{\infty}$) structure of $Hom(A_{UDol}, K({\cal O} , 3))$ induced by the ``abelian group'' structure of $K({\cal O} , 3)$. We now turn back to the second morphism in the fiber product, $$ j^{\ast} : Hom( C_{UDol}, K({\cal O} , 3))\rightarrow Hom( A_{UDol}, K({\cal O} , 3)). $$ The infinite loop space structure of $K({\cal O} , 3)$ which is inherited by $Hom ({\cal F} , K({\cal O} , 3))$ for ${\cal F} = C_{UDol}$ and ${\cal F} = A_{UDol}$. By Proposition \ref{decomp}, this delooping structure gives a decomposition of $Hom ({\cal F} , K({\cal O} , 3))$ into a product of Eilenberg-MacLane stacks. The restriction morphism $j^{\ast}$ above is compatible with the delooping structures, so by Proposition \ref{decomp} (B), it is homotopic to a map compatible with the decomposition into a product of Eilenberg-MacLane stacks. Recall that $$ \pi _i(Hom (C_{UDol}, K({\cal O} , 3)), 0) = H^{3-i}(C_{UDol}, {\cal O} )= H^{3-i}_{Dol}(C) $$ and $$ \pi _i(Hom (A_{UDol}, K({\cal O} , 3)), 0) = H^{3-i}(A_{UDol}, {\cal O} )= H^{3-i}_{Dol}(A). $$ From the splitting given by the delooping structures via Proposition \ref{decomp} with homotopy of functoriality of part (B) of that proposition (choose one), we obtain a homotopy-commutative diagram $$ \begin{array}{ccc} K(H^2_{Dol}(C), 1)& \rightarrow & K(H^2_{Dol}(A), 1) \\ \downarrow && \downarrow \\ Hom( C_{UDol}, K({\cal O} , 3))&\rightarrow &Hom( A_{UDol}, K({\cal O} , 3)). \end{array} $$ All maps are infinite loop maps. Using this diagram, we get the map $$ P^3:= K\times _{Hom( A_{UDol}, K({\cal O} , 3))}K(H^2_{Dol}(C), 1) \rightarrow $$ $$ K\times _{Hom (A_{UDol}, K({\cal O} , 3))} Hom (C_{UDol}, K({\cal O} , 3))=P^2. $$ Set $\rho ^3:= (0,0)$ in $P^3_0$. It maps to $\rho ^2$ (a point which we didn't specify because it was always obviously given by $\rho$), because we have normalized so that $\varphi$ becomes the zero-section. The map $P^3\rightarrow P^2$ (obviously a map over $K$) takes $\rho ^3$ to $\rho ^2$. We have $$ H^2_{Dol}(C)= H^2(C, {\cal O} _C)\oplus H^1(C, \Omega ^1_C)\oplus H^0(C, \Omega ^2_C), $$ and similarly $$ H^2_{Dol}(A)= H^2(A, {\cal O} _A)\oplus H^1(A, \Omega ^1_A)\oplus H^0(A, \Omega ^2_A). $$ Recall that $C$ and $A$ are isomorphic to open subsets of $Z$: together $N$ and $C$ form a covering of $Z$ and $N\cap C= A$. On the other hand, cohomology of coherent sheaves on $N$ vanishes because $N$ is affine. Therefore Mayer-Vietoris gives an exact sequence $$ H^1(C,\Omega ^1_C) \rightarrow H^1(A, \Omega ^1_A) \rightarrow H^2(Z, \Omega ^1_Z). $$ We are supposing that $H^1_{Dol}(Z)=0$ so by duality $H^3_{Dol}(Z)=0$. Thus the term on the right is zero, and the morphism of vector spaces $$ H^1(C,\Omega ^1_C) \rightarrow H^1(A, \Omega ^1_A) $$ is surjective. Choose a splitting. This gives a morphism $$ K(H^1(A, \Omega ^1_A), 1)\rightarrow K(H^2_{Dol}(C), 1) $$ such that the projection into $K(H^2_{Dol}(A), 1)$ is equal to the inclusion from the Dolbeault direct sum decomposition for $A$. Using our choice of splitting we obtain a map $$ P^4:= K\times _{Hom( A_{UDol}, K({\cal O} , 3))}K(H^1(A, \Omega ^1_A), 1) \rightarrow $$ $$ K\times _{Hom( A_{UDol}, K({\cal O} , 3))}K(H^2_{Dol}(C), 1) = P^3. $$ Again set $\rho ^4 = (0,0)\in P^4_0$, which maps to $\rho ^3$. Now $P^4$ is defined entirely in terms of $A$. We decompose things a bit more. Using the infinite loop-space structure of $Hom( A_{UDol}, K({\cal O} , 3))$ and again Proposition \ref{decomp} we get the decomposition into a product of Eilenberg-MacLane spaces $$ Hom( A_{UDol}, K({\cal O} , 3))\cong J^0\times J^1 \times J^2 \times J^3 $$ where $$ J^i = K(H^{3-i}_{Dol}(A), i). $$ The second map in the fiber product defining $P^4$ is homotopic (choose a homotopy) to one which is compatible with this decomposition, coming from the morphism $$ K(H^1(A, \Omega ^1_A), 1)\rightarrow J^1= K(H^2_{Dol}(A), 1) $$ (the other components are the maps sending $\ast$ to the basepoints $0\in J^i$, $i\neq 2$). Thus we can write $$ P^4 = Q^0\times Q^1 \times Q^2 \times Q^3 $$ where $$ Q^i= K\times _{J^i}\ast = K\times _{K(H^{3-i}_{Dol}(A), i)}\ast $$ for $i\neq 1$ and where $$ Q^1 = K\times _{K(H^{2}_{Dol}(A), 1)}K(H^1(A, \Omega ^1_A), 1). $$ This means that the action of $W$ on $P^4_0$ decomposes into a product of actions on the $Q^i_0$. Note that the $Q^i_0$ are themselves Eilenberg-MacLane spaces, $$ Q^i_0 = \Omega J^i = K( H^{3-i}_{Dol}(A), i-1) $$ for $i\geq 2$, and $$ Q^1_0 = K(H^{2}_{Dol}(A)/H^1(A, \Omega ^1_A), 0). $$ The classifying maps for the actions of $W$ on the components $Q^i_0$ factor as $$ K\rightarrow J^i \rightarrow B\, Aut (Q^i_0) $$ for $i\geq 2$, and $$ K\rightarrow K(H^{2}_{Dol}(A)/H^1(A, \Omega ^1_A), 1)\rightarrow B\, Aut (Q^1_0) $$ for $i=1$. Thus the classifying map for the action $K\rightarrow B\, Aut (P^4_0)$ factors through our above map $$ a: K\rightarrow K(H^{2}_{Dol}(A)/H^1(A, \Omega ^1_A), 1)\times J^2\times J^3. $$ We use this to calculate the characteristic classes for the action: the primary invariant is the first component, corresponding to a map $$ W\rightarrow H^{2}_{Dol}(A)/H^1(A, \Omega ^1_A). $$ The other primary invariant giving the action of $W$ on $\pi _1$ by outer automorphisms, is trivial: it is the action on $\Omega J^2$ induced by the classifying map $K\rightarrow J^2$, and $J^2$ is simply connected. The secondary invariant which we are interested in is the map $K\rightarrow J^2$ which corresponds to a class in $H^2(K, H^1_{Dol}(A))$. A preliminary remark is that the formula $a= \mu ^{\ast}(h)-p_2^{\ast}(h)$ means that the component of $a$ in $J^0$ is equal to $0$. In fact, $p_2^{\ast}(h)$ is exactly the $J^0$-component of $\mu ^{\ast}(h)$. On the other hand, the remaining components of $a$ are the same as those of $\mu ^{\ast}(h)$; these are the K\"unneth components of $$ h\circ \mu : K\times A_{UDol} \rightarrow K({\cal O} , 3). $$ The first thing to check is that the primary invariant is trivial for $(P^4, \rho ^4)$. As we have said above, it is the map $$ K(W, 1)\rightarrow K(H^{2}_{Dol}(A)/H^1(A, \Omega ^1_A), 1). $$ To check that it is trivial, it suffices to prove the \newline {\bf Claim} ${\bf (\ast )}:$ the map $$ W\rightarrow \pi _1(Hom( A_{UDol}, K({\cal O} , 3)), 0)= H^2_{Dol}(A) $$ takes $W$ into the component $H^1(A, \Omega ^1_A)$. For this we must again get back to the description of the map $$ \mu ^{\ast}(h): K\rightarrow Hom (A_{UDol}, K({\cal O} , 3)) $$ (note as remarked above that all of the components except the $J^0$ component, are the same for $\mu ^{\ast}(h)$ or for $a$). This map, which is equivalent to a map $$ A_{UDol}\times K \rightarrow K({\cal O} , 3), $$ is the pullback of $$ h: A_{UDol}\rightarrow K({\cal O} , 3) $$ by the map $$ \mu : A_{UDol}\times K \rightarrow A_{UDol}. $$ Now we have $$ A_{UDol} = A\times K $$ so $$ H^3(A_{UDol}, {\cal O} ) = H^3(A, {\cal O} )\oplus H^2(A, H^1(K, {\cal O} )) \oplus H^1(A, H^2(K, {\cal O} ) ) \oplus H^3(K , {\cal O} ). $$ The last term $H^3(K , {\cal O} )$ vanishes because it would be $\bigwedge ^3W^{\ast}$ but $W$ is $2$-dimensional. Similarly, one can arrange (by an appropriate choice of $N$) so that $A$ has an open covering by two affine open sets. Thus $H^3(A, {\cal O} ) = H^2(A, {\cal O} )= 0$. The only remaining term is $$ H^3(A_{UDol}, {\cal O} ) = H^1(A, H^2(K, {\cal O} ))= H^1(A, \bigwedge ^2W^{\ast}) = H^1(A, \Omega ^2_A). $$ Our map $$ A_{UDol}\times K \rightarrow K({\cal O} , 3) $$ is obtained by pulling back the above, using the map $K\times K\rightarrow K$. Pullback for this map is $$ \bigwedge ^2W^{\ast} \rightarrow $$ $$ H^2(K,{\cal O})\otimes _{{\cal O}}{\cal O} \oplus H^1(K,{\cal O} )\otimes _{{\cal O}}H^1(K, {\cal O} ) \oplus {\cal O} \otimes _{{\cal O}}H^2(K,{\cal O} ). $$ Each of the factors is nontrivial, with the middle being (up to a multiple which depends on normalizations for notation in exterior products) the standard map $$ \bigwedge ^2 W^{\ast} \rightarrow W^{\ast}\otimes _{{\cal O}}W^{\ast}. $$ The morphism induced by pulling back $h$ to $A_{UDol}\times K$ thus decomposes into K\"unneth components $$ H^1(A, \Omega ^2_A)\otimes _{{\cal O}}{\cal O} \; \; \oplus \; \; H^1(A, \Omega ^1_A)\otimes _{{\cal O}}W^{\ast} \; \; \oplus \; \; H^1(A, {\cal O} _A)\otimes _{{\cal O}}\bigwedge ^2W^{\ast}. $$ Each component is induced by $h\in H^1(A,\Omega ^2_A)$. The map $$ \pi _1(K)=W\rightarrow \pi _1(Hom (A_{UDol}, K({\cal O} , 3)) = H^2_{Dol}(A) $$ corresponds to the middle component, which is in fact a map $$ W\rightarrow H^1(A, \Omega ^1_A). $$ This was exactly the claim ${\bf (\ast )}$ we needed to prove to show that the primary invariant for $(P^4, \rho ^4)$ was trivial. The other primary invariant, the action of $W$ on $\pi _1(P^4_0, \rho ^4)$, is trivial as remarked above. To restate the argument, the map $a'$ induces an injection on $\pi _1$, therefore the $\pi _1$ of the fiber is the image of $\pi _2$ of the base; but since the base is an infinite loop space, the action of $\pi _1$ of the base on $\pi _2$ of the base is trivial; thus the action of $\pi _1$ of the base and in particular of $W$ on $\pi _1(P^4_0, \rho ^4)$ is trivial. We can now look at the secondary invariant for $(P^4, \rho ^4)$. It is the map $$ K\rightarrow J^2 = K(H^1_{Dol}(A), 2), $$ which is the next K\"unneth component of the pullback of $h$ to $A_{UDol}\times K$, $$ \bigwedge ^2W \rightarrow H^1(A, {\cal O} _A)\subset H^1_{Dol} (A) = \pi _2(Hom (A_{UDol}, K({\cal O} , 3))). $$ This claim tells us that the secondary class for $P^4$ is just $h$ considered as an element of $$ H^1(A, \Omega ^2_A)= H^1(A, \bigwedge ^2W^{\ast})= H^1(A,{\cal O} _A)\otimes _{{\cal O}} \bigwedge ^2W^{\ast}. $$ Recalling that $$ P^4_0 = \ast \times _{Hom (A_{UDol}, K({\cal O} , 3))}K(H^1(A, {\cal O} _A), 1) $$ and $\rho ^4= (0,0)$, we have that $P^4_0$ is just the homotopy fiber of the second morphism. The long exact sequence for the fibration gives $$ 0\rightarrow \pi _2(Hom (A_{UDol}, K({\cal O} , 3)))\rightarrow \pi _1(P^4_0, \rho ^4) \rightarrow 0 $$ (the morphism from $\pi _1$ of the total space to $\pi _1$ of the base being injective, and $\pi _2$ of the total space being zero). This long exact sequence persists under the maps $$ P^4_0\rightarrow P^3_0\rightarrow P^2_0\rightarrow P^1_0 $$ and furthermore, even into $P_0$ where the long exact sequence for the fiber of a morphism is replaced by a long exact sequence for the homotopy fiber product. It follows that the secondary class for $(P, \rho )$ is the image of our above class (basically $h$) $$ \bigwedge ^2W \rightarrow H^1(A, {\cal O} _A) $$ under composition with the map $$ H^1(A, {\cal O} _A)=\pi _2(Hom (A_{UDol}, K({\cal O} , 3)), 0) \rightarrow \pi _2(Hom (A_{UDol}, T), \rho |_{A_{UDol}}) $$ $$ \rightarrow \pi _1(Hom (X_{UDol}, T), \rho )=H^2_{Dol}(X)/(\eta ). $$ One slight twist to notice is that the morphism $K({\cal O} , 3)\rightarrow T$ in question (over $A_{UDol}$) is shifted by $h$. This shift is recovered in the last equality, where we undo a shift by $\varphi$. This morphism from the long exact sequence for the fiber product (of $Hom$'s) is equal to the connecting morphism $$ H^1(A, {\cal O} _A)\rightarrow H^2(X, {\cal O} _X) \subset H^2_{Dol}(X)/(\eta ). $$ Finally, we have concluded that our secondary class is the composition $$ \bigwedge ^2W^{\ast} \rightarrow H^1(A, {\cal O} _A) \rightarrow H^2(X, {\cal O} _X) $$ where the first map is $h$ (which depends on $\eta$ and which we investigate below) and the second map is the connecting morphism. The remaining problem is to calculate $h\in H^1(A, \Omega ^2_A)$ or more precisely its image by the connecting morphism. We have the exact sequence of the cohomology of the pair $(X, A)$ with coefficients in $\Omega ^2_X$: $$ 0= H^1(X, \Omega ^2_X) \rightarrow H^1(A, \Omega ^2_A) \rightarrow $$ $$ H^2((X, A), \Omega ^2_X) \rightarrow H^2(X, \Omega ^2_X). $$ The class $\eta$ may be considered as lying in $H^1((X, A), \Omega ^1_X)$, which is the statement that our map to $T$ factors, over $A$, through a map to $K({\cal O} , 3)$. Therefore the cup product $\eta \cup \eta$ can be considered as lying in $H^2((X, A), \Omega ^2_X)$ and mapping to zero in $H^2(X, \Omega ^2_X)$. The class $h\in H^1(A, \Omega ^2_A)$ is the preimage of $\eta \cup \eta$ in the exact sequence for the pair $(X,A)$. Now recall that $X = B\cup C$. This means that the pair $(X,A)$ decomposes as a ``disjoint union'' of the pairs $(X,B)$ and $(X,C)$ (after applying excision). Thus we can write $$ H^2((X, A), \Omega ^2_X) = H^2((X, B), \Omega ^2_X) \oplus H^2((X, C), \Omega ^2_X) . $$ Our class $h\in H^1(A, \Omega ^2_A)$ corresponds to $b+c$ with $$ b= -n^2 [H]^2\in H^2((X, B), \Omega ^2_X) $$ and $$ c= m^2 [E]^2\in H^2((X, C), \Omega ^2_X). $$ We now do the same thing for $Z= N\cup C$. The class $h$ corresponds here to an element of $$ H^2((Z, N), \Omega ^2_Z) \oplus H^2((Z, C), \Omega ^2_Z) . $$ This is again of the form $b'+c'$ but now with $$ b'= -n^2 [H]^2\in H^2((Z, N), \Omega ^2_Z) $$ and $$ c'= m^2 [P] \in H^2((Z, C), \Omega ^2_Z). $$ From the long exact sequence of the pair $(Z, N)$ and the fact that $N$ is affine we find that $$ H^2((Z, N), \Omega ^2_Z) \cong H^2(Z, \Omega ^2_Z)={\bf C} . $$ To calculate our secondary class we have to contract $h$ with $\alpha \wedge \beta$ to obtain a class in $H^1(A, {\cal O} _A)$ and then take its image by the connecting map in $H^2(X,{\cal O} _X)$ or equivalently in $H^2(Z, {\cal O} _Z)$. To measure this image, use Serre duality: we will choose a form $\omega \in H^0(Z, \Omega ^2_Z)$ and take the cup-product to end up with a class in $H^2(Z, \Omega ^2_Z)$ (of which we then take the residue to end up in ${\bf C}$). The result of this procedure is the same as if we first contract $\omega$ with $\alpha \wedge \beta$ and then multiply this section of $H^0(A, {\cal O} _A)$ by $h$ getting a class in $H^1(A, \Omega ^2_A)$. Then take the image of this class by the connecting map and take its residue. Note that the contraction of $\omega$ with $\alpha \wedge \beta$ is defined over all of the neighborhood $N$. Call this section $\lambda \in {\cal O} (N)$. We are now reduced to the problem of calculating the image in $H^2(Z, \Omega ^2_Z)$ under the connecting map for the covering $Z=N\cup C$, of $\lambda h \in H^1(A, \Omega ^2_A)$. We have written that the image of $h$ in $H^2((Z, A), \Omega ^2_Z)$ (which we shall denote $[h]$) is equal to $b'+c'$ where $b'\in H^2((Z, N), \Omega ^2_Z)$ and $c'\in H^2((Z, C),\Omega ^2_Z)$. The components $b'$ and $c'$ are obtained by residue maps for the class $h$, along respectively $H$ and $P$ (noting that $H= Z-N$ and $P= Z-C$). The form of the residue map is not important for us, just the fact that the classes have poles of order $1$; it follows that if $\lambda$ is a regular function on $N$ (a neighborhood of $P$) then the residue of $\lambda h$ at $P$, is equal to $\lambda $ times the residue of $h$ at $P$. Therefore we can write $$ [\lambda h] = b'' + \lambda c', $$ with $$ b''\in H^2((Z,N), \Omega ^2_Z) $$ and $$ \lambda c'\in H^2((Z,C), \Omega ^2_Z) $$ both being obtained by excision. The value of $b''\in H^2((Z,N), \Omega ^2_Z)\cong {\bf C}$ is determined by the condition that the image of $[\lambda h]$ in $H^2(Z, \Omega ^2_Z)$ be zero (from the long exact sequence for the pair $(Z,A)$). We now look at the image of $\lambda h$ by the connecting map in the long exact sequence of the covering $Z= N\cup C$, $$ H^1(N, \Omega ^2_N)\oplus H^1(C, \Omega ^2_C) \rightarrow H^1(A, \Omega ^2_A) $$ $$ \rightarrow H^2(Z, \Omega ^2_Z)\rightarrow \ldots . $$ We can decompose this connecting map as a composition $$ H^1(A, \Omega ^2_A) \rightarrow H^2((Z, A), \Omega ^2_Z) \rightarrow H^2((Z, C), \Omega ^2_Z) \rightarrow H^2(Z, \Omega ^2_Z), $$ where the second arrow is the projection onto the first factor in the excision decomposition $$ H^2((Z, A), \Omega ^2_Z) = H^2((Z, C), \Omega ^2_Z) \oplus H^2((Z, N), \Omega ^2_Z) $$ One could equally well use the second factor, with a sign change; our calculations are not accurate insofar as signs are concerned. Thus the image of $\lambda h$ by the connecting map for the covering $Z= N\cup C$, is equal to the class of either $\lambda c'$ or of $-b''$. We don't know how to calculate $b''$ so we use the representation as $\lambda c'$. This image is then equal to $$ \lambda \cdot (m^2 [P]) $$ which is just $m^2\lambda (P)$ (because as noted above, $[P]$ is represented by cocycles with poles of order $1$). We have established the formula that the image of $\alpha \wedge \beta $ under the map $$ \bigwedge ^2W \rightarrow H^2(Z, {\cal O} _Z) $$ is a class which, when paired with a form $\omega \in H^0(Z, \Omega ^2_Z)$, gives $$ m^2\omega (\alpha \wedge \beta )(P). $$ This completes the proof of Theorem \ref{calculation}. \hfill $\Box$\vspace{.1in} \numero{APPENDIX I: Relative Breen calculations in characteristic $0$} Crucial to the reasonable working of a theory of nonabelian cohomology is the possibility of calculating the invariants in the Postnikov tower of the spaces which measure the ``shape''. In our setup, this means that we would like to calculate $H^i(K({\cal O} , m), {\cal O} )$. This calculation is the algebraic analogue of the classical Eilenberg-MacLane calculations. The algebraic version is the subject of Breen's work \cite{Breen1}, \cite{Breen2}. His motivation came mostly from arithmetic geometry, so he concentrated on the case of characteristic $p$ in \cite{Breen2}. The characteristic $0$ version, while not explicitly stated in \cite{Breen2}, is implicit there because it is strictly easier than the characteristic $p$ case: there are no new classes coming from Frobenius. These calculations for the case of base scheme $Spec ({\bf C} )$ are sufficient for the purposes of the present paper, but eventually a relative version will also be useful. In the context of calculation of $Ext$ sheaves (i.e. the stable part of the calculation) this relative version was already evoked in \cite{Breen1}, where Breen states that the $Ext ^i(G,\cdot )$ for representable group schemes $G$ can always be calculated. This part of the topic was not really taken up afterward, probably for lack of a reasonable category of sheaves over a base scheme $S$. Such a category of sheaves is provided by Hirschowitz's notion of {\em $U$-coherent sheaf} \cite{Hirschowitz}, see also Jaffe's recent paper \cite{Jaffe}. \footnote{ Hirschowitz's notion is similar to, but not quite the same as Auslander's theory of ``coherent functors'' developed in the 1960's. Jaffe's paper \cite{Jaffe} views $U$-coherent sheaves as a modification or generalization of Auslander's theory---one looks at functors of algebras rather than functors of modules. Jaffe, who seems to have been unaware of Hirschowitz's paper, cites an unpublished letter from Artin to Grothendieck, dating from the 1960's, as a reference for the generalized version of Auslander's theory. In order to straighten out the history of this notion, one would have to compare Artin's letter with \cite{Hirschowitz}.} We change Hirschowitz's notation and call these objects {\em vector sheaves}. The category of vector sheaves over a base $S$ is defined in \cite{Hirschowitz} as the smallest abelian subcategory of sheaves of ${\cal O}$-modules on the big site $Sch /S$, containing ${\cal O}$ and stable under localization of the base. Thus, locally over $S$ vector sheaves are obtained starting with ${\cal O}$ by repeated applications of taking direct sums, kernels and cokernels. The abelian category of vector sheaves has several nice properties \cite{Hirschowitz}. The coherent sheaves on $S$, which are defined as cokernels $$ {\cal O} ^a \rightarrow {\cal O} ^b \rightarrow {\cal F} \rightarrow 0, $$ are vector sheaves. Coherent sheaves are injective objects. Their duals, which we call {\em vector schemes}, are the group-schemes with vector space structure (but not necessarily flat) over $S$. These admit dual presentations as kernels of maps ${\cal O} ^b\rightarrow {\cal O} ^a$. They are projective objects (at least if the base $S$ is affine). If $S$ is affine, then any vector sheaf $U$ admits resolutions $$ 0\rightarrow V\rightarrow V' \rightarrow V'' \rightarrow U \rightarrow 0 $$ with $V$, $V'$ and $V''$ vector schemes; and $$ 0\rightarrow U \rightarrow {\cal F} \rightarrow {\cal F} '\rightarrow {\cal F} '' \rightarrow 0 $$ with ${\cal F}$, ${\cal F} '$, ${\cal F} ''$ coherent sheaves. {\bf Example:} The motivating example for the definition of vector sheaf in \cite{Hirschowitz} was the following example. If $E^{\cdot}$ is a complex of vector bundles over $S$ then the cohomology sheaves defined on the big site $Sch /S$ are vector sheaves. Indeed they are vector sheaves of a special type which Hirschowitz calls ``cohomologies'': quotients of vector schemes by coherent sheaves. This example is important because it arises from the cohomology of flat families of coherent sheaves parametrized by $S$: if $f:X\rightarrow S$ is a projective morphism and ${\cal F}$ is a coherent sheaf on $X$ flat over $S$ then a classical result says that the higher direct image complex $R^{\cdot} f_{\ast}({\cal F} )$ (calculated on the big site) is quasiisomorphic to a complex of vector bundles. The notion of vector sheaf (``$U$-coherent sheaf'' in \cite{Hirschowitz}) thus keeps track of the jumping of cohomology of flat families of cohoerent sheaves. \footnote{ This type of example comes up in relation with Dolbeault cohomology: for example let $$ M:= {\cal M}_{Dol}(X,G) = Hom (X_{Dol}, K(G,1)) $$ be the moduli stack of principal Higgs $G$-bundles. If $V$ is a representation of $G$ then we obtain $$ T:= K(V/G, n) \rightarrow K(G,1) $$ with fiber $K(V,n)$. There is a universal local system $E$ on $X_{Dol}\times M$, and $$ \pi _i(Hom (X_{Dol}, T)/M, 0) = H^i(X_{Dol}\times M/M, E). $$ This is calculated by a Dolbeault complex for $E$ on $X/M$, and the general discussion of cohomology in flat families applies. Therefore the $H^i(X_{Dol}\times M/M, E)$ are vector sheaves over $M$ (here $M$ is an algebraic stack; the condition of being a vector sheaf means that the pullback to any scheme $Y\rightarrow M$ is a vector sheaf on $Y$). } The most surprising property from \cite{Hirschowitz} is that the duality functor $U^{\ast} := Hom (U,{\cal O} )$ is exact, and is an involution. This is due to the fact that we take the big site $Sch /S$ rather than the small Zariski or etale sites. Another interesting point is that there are two different types of tensor products of vector sheaves: the {\em tensor product} $$ U\otimes _{{\cal O}} V := Hom (U, V^{\ast} )^{\ast}, $$ and the {\em cotensor product} $$ U\otimes ^{{\cal O}} V := Hom (U^{\ast}, V). $$ These are not the same (although they coincide for coherent sheaves cf Lemma \ref{cohtensor} below) and in particular they don't have the same exactness properties. Neither of them is equal to the tensor product of sheaves of ${\cal O}$-modules. See \cite{kobe} and \cite{RelativeLie} for further discussion. The above facts work in any characteristic and depend on the ${\cal O}$-module structure. However, in characteristic zero vector sheaves have the additional property that the morphisms $U\rightarrow V$ of sheaves of abelian groups over $Sch /{\bf C} $ are automatically morphisms of ${\cal O}$-modules, see \cite{kobe} \cite{RelativeLie}. Similarly, extensions of sheaves of abelian groups, between two vector sheaves, are again vector sheaves. These properties persist for the higher $Ext^i$, see Corollary \ref{ext} below. These properties do not remain true in characteristic $p$, as shown precisely by Breen's calculations of \cite{Breen2}. The basic problem is that Frobenius provides a morphism ${\cal O} \rightarrow {\cal O}$ of sheaves of abelian groups, which is not a morphism of sheaves of ${\cal O}$-modules. This difficulty in characteristic $p$ seems to be the main obstacle to realizing a reasonable analogue of rational homotopy theory, for homotopy in characteristic $p$. So we stick to characteristic $0$! We don't give a detailed introduction to vector sheaves, rather we refer the reader to \cite{Hirschowitz}, \cite{kobe} and \cite{RelativeLie}. However, we do take this opportunity to correct an omission from \cite{kobe} and \cite{RelativeLie}. Without the following lemma, the discussion in those references often seems contradictory, as tensor products and cotensor products are interchanged when the coefficients are coherent sheaeves. For example, in the statement of Corollary 3.9 of \cite{kobe} (which we restate as Theorem \ref{bc} and prove in more detail below), all terms occuring are coherent sheaves. Thus the tensor product which appears in the notation is also equal to the cotensor product. It is the cotensor product which appears most naturally in that situation. Indeed, Lemma \ref{cohtensor} below explains (i.e. justifies) the seemingly erroneous statement ${\cal F} \otimes _{{\cal O}} {\cal G} =\underline{Hom}({\cal F} ^{\ast}, {\cal G} )$ in the proof of Corollary 3.9 of \cite{kobe}. \begin{lemma} \mylabel{cohtensor} Suppose ${\cal F}$ and ${\cal G}$ are coherent sheaves. Then the tensor product ${\cal F} \otimes _{{\cal O}} {\cal G}$ and the cotensor product ${\cal F} \otimes ^{{\cal O}}{\cal G}$ coincide. \end{lemma} {\em Proof:} Choose a presentation $$ {\cal O} ^a \rightarrow {\cal O} ^b \rightarrow {\cal G} \rightarrow 0. $$ Now ${\cal F} ^{\ast} := Hom ({\cal F} , {\cal O} )$ is a vector scheme, in particular it is a scheme affine over $S$. Therefore the functor $$ U\mapsto Hom ({\cal F} ^{\ast} , U) $$ is exact in $U$. Applying this functor to the above presentation we obtain $$ {\cal F} ^a \rightarrow {\cal F} ^b \rightarrow Hom ({\cal F} ^{\ast} , {\cal G} )\rightarrow 0. $$ The term $Hom ({\cal F} ^{\ast} , {\cal G} )$ is by definition the cotensor product ${\cal F} \otimes ^{{\cal O}} {\cal G}$. Taking the dual of the above presentation we obtain $$ 0\rightarrow {\cal G} ^{\ast} \rightarrow {\cal O} ^b \rightarrow {\cal O} ^a. $$ Applying the functor $$ U\mapsto Hom ({\cal F} , U) $$ which is exact on the left, we obtain $$ 0\rightarrow Hom ({\cal F} , {\cal G} ^{\ast})\rightarrow ({\cal F} ^{\ast})^a \rightarrow ({\cal F} ^{\ast})^b. $$ Taking the dual we get $$ {\cal F} ^a \rightarrow {\cal F} ^b \rightarrow Hom ({\cal F} , {\cal G} ^{\ast})^{\ast}\rightarrow 0. $$ This time the term $Hom ({\cal F} , {\cal G} ^{\ast})^{\ast}$ is by definition the tensor product ${\cal F} \otimes _{{\cal O}} {\cal G}$. In general there is a natural morphism $$ U\otimes _{{\cal O}} V = Hom (U,V^{\ast})^{\ast} \rightarrow Hom (U^{\ast}, V)= U\otimes ^{{\cal O}} V. $$ To define this map we define a trilinear morphism $$ Hom (U,V^{\ast})^{\ast}\times U^{\ast} \times V^{\ast} \rightarrow {\cal O} , $$ by $(\lambda , \mu , \nu )\mapsto \lambda (\nu \mu )$, the product $\nu \mu$ being the composed morphism $$ U\stackrel{\mu}{\rightarrow} {\cal O} \stackrel{\nu}{\rightarrow} V^{\ast}. $$ This trilinear map gives a morphism $$ Hom (U,V^{\ast})^{\ast}\rightarrow Hom (U^{\ast}, (V^{\ast})^{\ast}) $$ then note that $(V^{\ast})^{\ast}=V$. The above presentations for ${\cal F} \otimes ^{{\cal O}} {\cal G}$ and ${\cal F} \otimes _{{\cal O}} {\cal G}$ (which are the same) are compatible with this natural morphism, so the natural morphism is an isomorphism $$ {\cal F} \otimes ^{{\cal O}} {\cal G}\cong {\cal F} \otimes ^{{\cal O}} {\cal G} . $$ \hfill $\Box$\vspace{.1in} We now come to the statement of the ``relative Breen calculations in characteristic $0$''. The case $S=Spec (k)$ for $k$ a field is due to \cite{Breen2}, and the relative case for a group scheme was suggested in \cite{Breen1} (where the case of cohomology with coefficients in the multiplicative group scheme was treated). \begin{theorem} \mylabel{bc} Suppose $S$ is a scheme over $Spec ({\bf Q} )$. Suppose $V$ is a vector scheme over $X$ and suppose ${\cal F}$ is a coherent sheaf over $S$. Then for $n$ odd we have $$ H^i(K(V/S, n)/S, {\cal F} ) = {\cal F} \otimes _{{\cal O}} \bigwedge _{{\cal O}}^{i/n} (V^{\ast}). $$ For $n$ even we have $$ H^i(K(V/S, n)/S, {\cal F} ) = {\cal F} \otimes _{{\cal O}} Sym _{{\cal O}}^{i/n} (V^{\ast}). $$ In both cases the answer is $0$ of $i/n$ is not an integer. The multiplicative structures on the left sides, in the case ${\cal F} = {\cal O}$, coincide with the obvious ones on the right sides. In the case of arbitrary ${\cal F}$, the natural structures of modules over the cohomology with coefficients in ${\cal O}$, on both sides, coincide. \end{theorem} We first recall the following result. \begin{proposition} \mylabel{somecomplexes} Suppose $S$ is a scheme and $V$ is a vector sheaf on $S$. Then the complexes $$ \ldots {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^j \otimes _{{\cal O}} Sym _{{\cal O}}^k V \rightarrow {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^{j-1} \otimes _{{\cal O}} Sym _{{\cal O}}^{k+1} V \ldots $$ and $$ \ldots {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^j \otimes _{{\cal O}} Sym _{{\cal O}}^k V \rightarrow {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^{j+1} \otimes _{{\cal O}} Sym _{{\cal O}}^{k-1} V \ldots $$ are exact as sequences of vector sheaves (i.e. as sequences of sheaves on the site $Sch /{\bf C} $). \end{proposition} {\em Proof:} For ${\cal F} = {\cal O}$ this is Proposition 3.8 of \cite{kobe}. The proof is easy, obtained by taking the graded symmetric powers of the cohomologically trivial complex $V\rightarrow V$ (placing this complex starting in odd or even degrees, leads to the two cases of the statement). For a general ${\cal F}$ note that the cotensor product with a coherent sheaf is exact---indeed, for any vector sheaf $U$, $Hom ({\cal F} ^{\ast}, U)$ is exact in $U$ because ${\cal F} ^{\ast}$ is represented by a scheme, i.e. an element of the site $Sch /{\bf C} $. The tensor product is equal to the cotensor product because both sides are coherent sheaves (here is where we use the hypothesis that $V$ is a vector scheme, i.e. $V^{\ast}$ is a coherent sheaf). \hfill $\Box$\vspace{.1in} \subnumero{Proof of Theorem \ref{bc}} Now we start the proof of Theorem \ref{bc}. Suppose that it is true for $n\leq m-1$, and we prove it for $n=m$. (The case $m=1$ to start the induction will be treated separately at the end.) Look at the fiber sequence $$ K(V/S, m-1)\rightarrow \ast \stackrel{p}{\rightarrow} K(V/S, m). $$ We will look at the Leray spectral sequence for the morphism $p$, for cohomology with coefficients in ${\cal F}$. We may assume that $S$ is affine Let $A_m$ denote the algebra $H^{\ast}(K(V/S, m)/S, {\cal O} )$ (this is a sheaf of algebras over $S$) and let $A_m({\cal F} )$ denote the $A_m$-module $H^{\ast}(K(V/S, m)/S, {\cal F} )$. Note that if ${\cal F}$ itself is an algebra-object then $A_m({\cal F} )$ is an $A_m$-algebra (graded-commutative). By the inductive hypothesis, the $A_k$ and $A_k({\cal F} )$ are direct sums of coherent sheaves (coherent in each degree) for $k\leq m-1$. The $E_2$ term of our spectral sequence is $$ H^i(K(V/S, m)/S, H^j(K(V/S, m-1)/S, {\cal F} ))\Rightarrow H^{i+j}(\ast , {\cal F} ). $$ In the case ${\cal F} = {\cal O}$ the $E_2$ term has a structure of algebra, and for arbitrary ${\cal F}$, a structure of module over that algebra. These are respectively $$ A_m(A_{m-1}) $$ and $$ A_m(A_{m-1}({\cal F} )). $$ By induction we know that $$ A_{m-1}({\cal F} ) = A_{m-1}\otimes _{{\cal O}} {\cal F} . $$ The first possible nonzero differential in the spectral sequence is $$ H^i(K(V/S, m)/S, H^j(K(V/S, m-1)/S, {\cal F} ))\rightarrow $$ $$ H^{i+m}(K(V/S, m)/S, H^{j+1-m}(K(V/S, m-1)/S, {\cal F} )). $$ We prove by a second induction on $k$, that for all ${\cal F}$ the answer is as given in the theorem, for $H^i(K(V/S, m)/S,{\cal F} )$ for all $i\leq k$. Suppose this is true for $i\leq k-1$. Then the elements of the diagonal complex for the above differential, ending at $(i,j)= (k, 0)$, are all in the region $i\leq k-1$, except for the term $(k,0)$. By our second inductive hypothesis (applied to the cohomology of $K(V/S, m)$ with coefficients in the coherent sheaves $H^j(K(V/S, m-1), {\cal F} )$, this complex coincides with one of the two complexes appearing in Proposition \ref{somecomplexes}, except maybe for the last term. However, the complex must be exact at the last stage because otherwise, what is left over would persist into $E_{\infty}$ contradicting the answer of the spectral sequence (which must be ${\cal F}$ in degree $0$ and $0$ otherwise). Proposition \ref{somecomplexes} gives exactness of the complexes appearing there. Therefore the $E^{k,0}_2$-term of our spectral sequence must also coincide with the last term of the complex from Propositon \ref{somecomplexes}. For example in the case where $m$ is odd, the end of the spectral sequence is $$ \ldots \rightarrow {\cal F} \otimes _{{\cal O}}Sym ^2_{{\cal O}}(V^{\ast})\otimes_{{\cal O}}\bigwedge _{{\cal O}} ^{(k-2m)/m} (V^{\ast}) $$ $$ \rightarrow {\cal F} \otimes _{{\cal O}}V^{\ast} \otimes_{{\cal O}}\bigwedge _{{\cal O}} ^{(k-m)/m} (V^{\ast}) $$ $$ \stackrel{d}{\rightarrow} E^{k,0}_2({\cal F} ) \rightarrow 0 $$ (where we denote by $d$ the last differential), whereas the end of the complex of Proposition \ref{somecomplexes} is $$ \ldots \rightarrow {\cal F} \otimes _{{\cal O}}Sym ^2_{{\cal O}}(V^{\ast})\otimes_{{\cal O}}\bigwedge _{{\cal O}} ^{(k-2m)/m} (V^{\ast}) $$ $$ \rightarrow {\cal F} \otimes _{{\cal O}}V^{\ast} \otimes_{{\cal O}}\bigwedge _{{\cal O}} ^{(k-m)/m} (V^{\ast}) $$ $$ \rightarrow {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^{k/m} (V^{\ast}) \rightarrow 0. $$ Therefore if $m$ is odd, $$ E^{k,0}_2 = {\cal F} \otimes _{{\cal O}} \bigwedge _{{\cal O}}^{k/n} (V^{\ast}). $$ The same holds with symmetric power instead of exterior power if $m$ is even. Cup product gives a bilinear morphism $$ \mu : E^{k-m,0}_2\times H^m(K(V/S, m), {\cal O} ) = E^{k-m,0}_2\times V^{\ast} \rightarrow E^{k,0}_2 . $$ Let $$ d': {\cal F} \otimes _{{\cal O}}V^{\ast} \otimes_{{\cal O}}\bigwedge _{{\cal O}} ^{(k-2m)/m} (V^{\ast}) \rightarrow E^{k-m,0}_2({\cal F} ) $$ denote the previous differential. We know by induction that $d'$ establishes an isomorphism between $E^{k-m,0}_2({\cal F} ) $ and $$ {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^{k/m} (V^{\ast}) $$ where this latter is considered as a quotient of the range of $d'$ via Proposition \ref{somecomplexes}. Refering to the cup-product morphism $\mu$ considered above and its precursor $$ \mu ': E^{k-2m,m-1}_2\times V^{\ast} \rightarrow E^{k-m,m-1}_2 , $$ we have the Leibniz formula $$ d\mu '(a,v) = \mu (d'(a), v) $$ noting that the term $V^{\ast}$ appearing in the formulas is $E^{m,0}_2$ so the differential acts trivially on the variable $v$. We have exactly the same formula when the terms $E^{i,j}_2$ are replaced by their counterparts from the sequences of Proposition \ref{somecomplexes}. Denote the multiplication in these counterparts by $\nu$ and $\nu '$ and the differentials by $\delta$ and $\delta '$. Call the isomorphism established by $d$, $$ \psi ^{k,0}: {\cal F} \otimes _{{\cal O}}\bigwedge _{{\cal O}} ^{k/m} (V^{\ast}) \cong E^{k,0} _2 $$ and similarly we have isomorphisms (by the inductive hypothesis) $\psi ^{k-m, 0}$, $\psi ^{k-m, m-1}$ and $\psi ^{k-2m, m-1}$. The definition of $\psi^{k,0}$ is given by the equation $$ \psi ^{k,0}(\delta (b))= d \psi ^{k-m, m-1} (b). $$ Similarly we have $$ \psi ^{k-m,0}(\delta '(b))= d' \psi ^{k-2m, m-1} (b). $$ We get $$ \psi ^{k,0}(\nu (\delta '(a), v))= \psi ^{k,0}(\delta \nu '(a,v)) = d\psi ^{k-m, m-1} (\nu '(a,v)). $$ On the other hand, in this region we know by induction that the $\psi$ are compatible with products. Therefore we get $$ d\psi ^{k-m, m-1} (\nu '(a,v)) = d \mu '(\psi ^{k-2m,m-1}(a), v) $$ $$ = \mu (d' \psi ^{k-2m,m-1}(a), v) $$ $$ =\mu (\psi ^{k-m,0}(\delta '(a)), v). $$ In all we have established the formula $$ \psi ^{k,0}(\nu (b, v)) =\mu (\psi ^{k-m,0}(b), v) $$ for any $b= \delta '(a)$. But $\delta '(a)$ is surjective. This establishes the compatibility of our isomorphism $\psi ^{k,0}$ with products (given already the compatibility of $\psi ^{k-m, 0}$). We note in the above proof that elements of the tensor products are always (locally on $S$) finite sums of tensors. This can be seen for example from the proof of Lemma \ref{cohtensor}. Thus to check compatibility with products, for example, it suffices to check it on elementary tensors as we have done above. This completes the proof of the theorem, modulo the case $m=1$ which we now treat. We have the fiber sequence $$ V\rightarrow S \stackrel{p}{\rightarrow} K(V,1). $$ The higher direct images vanish for coefficients in a coherent sheaf ${\cal F}$ so the Leray spectral sequence implies that $$ R^ip_{\ast}({\cal F} ) = 0, \;\; i> 0 $$ and $R^0p_{\ast}({\cal F} ) $ is a local system on $K(V,1)$ which when restricted to $S$ gives $$ {\cal F} [V] \cong {\cal F} \otimes _{{\cal O}} Sym ^{\cdot}_{{\cal O}} (V^{\ast}). $$ Now we use the complex given in Proposition \ref{somecomplexes}, which is basically a de Rham complex in our situation: $$ 0\rightarrow {\cal O} \rightarrow Sym ^{\cdot}_{{\cal O}} (V^{\ast}) \ldots \rightarrow Sym ^{\cdot}_{{\cal O}} (V^{\ast})\otimes _{{\cal O}} \bigwedge ^i_{{\cal O}} (V^{\ast})\ldots $$ which can then be tensored by ${\cal F}$ and remains exact. We can define the translation action of $V$ on all of the terms, and the exact sequence remains an exact sequence of sheaves with action of $V$. All of the terms except for the first one are acyclic by the previous result. Therefore the cohomology of $K(V,1)$ with coefficients in ${\cal F}$ is equal to the cohomology of the complex $$ {\cal F} \rightarrow \ldots \rightarrow {\cal F} \otimes _{{\cal O}} \bigwedge ^i_{{\cal O}} (V^{\ast})\ldots . $$ One can check that the differentials are actually zero, so the cohomology is as desired. One should check that the cup-product is equal to the obvious product structure on the exterior-algebra side of the answer. Instead of doing this (which as such would seem to be a difficult task) we proceed as follows. The above construction is functorial (contravariantly) for morphisms $V\rightarrow V'$. It is easy to see that if one considers an injection $V\hookrightarrow {\cal O} ^a$ (which exists locally on $S$ by the definition of vector scheme) then the morphism of functoriality induces a surjection on cohomology, coming from the surjection ${\cal O} ^a \rightarrow V^{\ast}$. Thus to establish a formula for cup-products in cohomology, it suffices to establish the formula for the case $V={\cal O} ^a$. In that case we can apply the K\"unneth formula, or more precisely remark that the same K\"unneth formula holds for the cohomology as for the exterior algebra, and that these formulas are compatible via the above isomorphism. The K\"unneth formulae are both compatible with cup-products. Thus we can reduce to the case $V={\cal O}$, but here the cohomology is concentrated in degrees $0$ and $1$ so there are no cup-products to verify (excepting the product with a degree $0$ class but this case is easy). \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{cohVPisVS} Suppose $S$ is a scheme and $T\rightarrow S$ is a relatively $1$-connected very presentable $n$-stack. Then for any vector sheaf $V$ on $S$, $$ H^i(T/S, V) $$ is a vector sheaf. \end{corollary} {\em Proof:} We use systematically (without further mention) the fact that the category of vector sheaves is closed under kernels, cokernels and extensions cf \cite{kobe} \cite{RelativeLie}. The case of $T= K(U/S,n)$ for $U$ a vector scheme and $V$ a coherent sheaf is given by Theorem \ref{bc}. The case of coefficients in any vector sheaf $V$ is obtained by taking a resolution of $V$ by coherent sheaves and using the long exact sequence of cohomology. The case of $T=K(U/S,n)$ for any vector sheaf $U$ is obtained by taking a resolution of $U$ by vector schemes (divided into two short exact sequences which give rise to two fibration sequences) and then applying the Leray spectral sequence. Finally, any relatively $1$-connected very presentable $T$ has a Postnikov tower (relative to $S$) whose stages are $K(U/S,n)$. Repeated application of the Leray spectral sequence gives the result. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{homVP} If $S$ is a scheme and $T\rightarrow S$ and $T'\rightarrow S$ are relatively $1$-connected very presentable $n$-stacks then $Hom (T/S, T'/S)$ is very presentable. \end{corollary} For this one has to use the fact that $Aut(V)$ is a very presentable group sheaf when $V$ is a vector sheaf, see \cite{RelativeLie}. \hfill $\Box$\vspace{.1in} The following application was the original motivation for Breen's calculations of the cohomology of the Eilenberg-MacLane sheaves \cite{Breen1} \cite{Breen2}. From our version Theorem \ref{bc}, we obtain the corresponding result in the relative case in characteristic zero. Similar corollaries were stated for example for cohomology with coefficients in the multiplicative group ${\bf G}_m$, in \cite{Breen1}. \begin{corollary} \mylabel{ext} {\rm (\cite{kobe} Corollary 3.11)} Suppose $U,V$ are vector sheaves over a scheme $S$. Let $Ext^i_{\rm gp} (U,V)$ denote the $Ext$ sheaves between $U$ and $V$ considered as sheaves of abelian groups on $Sch /S$, let $Ext^i_{\rm vs}(U,V)$ denote the $Ext$ sheaves between $U$ and $V$ considered as vector sheaves on $S$. Then the natural morphisms are isomorphisms $$ Ext^i_{\rm vs} (U,V) \stackrel{\cong}{\rightarrow}Ext^i_{\rm gp} (U,V). $$ The $Ext^i$ vanish for $i>2$. \end{corollary} {\em Proof:} Let $K_{\cdot}(U,n)$ denote the simplicial presheaf $$ Y\mapsto K_{\cdot}(U(Y), n ) $$ given by the standard simplicial Eilenberg-MacLane construction (i.e. Dold-Puppe applied to the complex with $U$ placed in degree $n$). We don't take the associated stack (as doing this or not doesn't affect the morphisms to an $m$-stack). Let ${\bf Z} K_{\cdot}(U,n)$ denote the associated presheaf of simplicial free abelian groups. Finally let $N{\bf Z} K_{\cdot}(U,n)$ be the presheaf of normalized complexes (in the homology direction i.e. with differential of degree $-1$) of this simplicial abelian group. For each $Y$, $$ N{\bf Z} K_{\cdot}(U,n)(Y) $$ is a complex with homology group $U(Y)$ in degree $n$, and with all other cohomology groups equal to $0$ in degrees $<2n$. Thus if ${\cal F}$ is an injective sheaf of groups then the (cohomological) complex of sheaves $$ Hom (N{\bf Z} K_{\cdot}(U,n), {\cal F} ) $$ has homology sheaf $Hom (U, {\cal F} )$ in degree $n$ and zero homology in all other degrees $< 2n$. It follows that if ${\cal F}$ is any sheaf of groups then for $i<n$, $$ H^{n+i}Hom (N{\bf Z} K_{\cdot}(U,n), {\cal F} ) = Ext ^i_{\rm gp}(U, {\cal F} ). $$ On the other hand, this complex of sheaves also calculates $H^{\cdot}(K(U, n), {\cal F} )$. Thus we find that $$ H^{n+i}(K(U,n), {\cal F} ) = Ext ^i_{\rm gp}(U,{\cal F} ), \;\; i<n. $$ This holds true for any sheaves of groups $U$ and ${\cal F}$. This is one of the motivating observations of Breen's paper \cite{Breen2}---we have repeated the proof here for the reader's convenience. Now suppose that $U$ and ${\cal F}$ are vector sheaves. If $U$ is a vector scheme and ${\cal F}$ is a coherent sheaf then, via the above observation, the relative Breen calculations (Theorem \ref{bc}) show that $Ext ^i_{\rm gp}(U,{\cal F} )=0$ for $i>0$. A coherent sheaf ${\cal F}$ is an injective object in the category of vector sheaves, and the functor $Hom (\cdot , {\cal F} )$ is exact (cf the discussion of vector sheaves in \cite{RelativeLie} for example). Thus if $U$ is any vector sheaf, we can (locally on $S$) resolve it by vector schemes and apply the previous paragraph. The functor $Ext ^0(\cdot , {\cal F} )= Hom (\cdot , {\cal F} )$ is exact (recall from \cite{kobe} Lemma 3.2 or \cite{RelativeLie} Lemma 4.5 that morphisms of sheaves of abelian groups are the same as morphisms of vector sheaves so the $Hom$ is the same in the two categories). Using this exactness we get that $Ext ^i_{\rm gp}(U,{\cal F} )=0$ for $i> 0$. Finally, if $V$ is any vector sheaf then we can resolve it by coherent sheaves, which is an injective resolution in the category of vector sheaves. This is also an acyclic resolution for $Ext$ in the category of sheaves of abelian groups, so we obtain the isomorphism $$ Ext^i_{\rm vs} (U,V) \stackrel{\cong}{\rightarrow}Ext^i_{\rm gp} (U,V). $$ The vanishing of the $Ext^i$ for $i>2$ comes from the fact that any vector sheaf $V$ has a resolution of length $2$ (i.e. with terms in degrees $0,1,2$) by coherent sheaves (cf \cite{kobe} \cite{RelativeLie}. \hfill $\Box$\vspace{.1in} We can apply \ref{ext} to the example discussed at the start of the appendix. Suppose $S$ is a scheme and suppose $E^{\cdot}$ is a complex of vector bundles on $S$. The cohomology sheaves $V^i= {\cal H}^i(E^{\cdot})$ are vector sheaves. In general, a complex with given cohomology objects is determined by higher extension classes in $Ext^i$ for all values of $i\geq 2$. However, by virtue of the above theorem the $Ext ^i(V^j, V^k)$ vanish for $i\geq 3$. Thus the complex $E^{\cdot}$ is determined completely by the successive extension classes $$ \delta _{j,j+1} \in Ext ^2(V^{j+1}, V^j). $$ The same is true for any complex of vector sheaves with $V^j$ as cohomology objects. {\bf Problem:} describe the conditions which must be satisfied by the classes $\delta _{j,j+1}$ for the complex determined by these classes to be (quasiisomorphic to) a complex of vector bundles. \numero{APPENDIX II: Representability of very presentable shape} The following result was stated without proof in (\cite{kobe}, the discussion above Theorem 5.7). Since we refer anew to this result in our discussion after Theorem \ref{calculation} of the present paper, I felt it to be an opportune time to give a proof. \begin{theorem} \mylabel{representable1} Suppose ${\cal F}$ is a connected $n$-stack on $Sch /{\bf C} $ such that the cohomology sheaves $H^i({\cal F} , {\cal O} )$ are represented by finite dimensional vector spaces. Suppose furthermore that $H^0({\cal F} , {\cal O} ) = {\cal O} $ and $H^1({\cal F} , {\cal O} )= 0$. Then the $n+1$-functor $$ T\mapsto Hom ({\cal F} , T) $$ from $1$-connected very presentable $n$-stacks of groupoids $T$ to the same, is representable by a morphism ${\cal F} \rightarrow \Sigma$, with $\Sigma$ being a $1$-connected very presentable $n$-stack. \end{theorem} {\em Proof:} It suffices to have a morphism ${\cal F} \rightarrow \Sigma$ which induces an isomorphism $$ H^i(\Sigma ,{\cal O} )\stackrel{\cong}{\rightarrow} H^i({\cal F} ,{\cal O} ) $$ for any $i$. We say that a morphism ${\cal F} \rightarrow \Sigma _m$ is {\em $m$-arranged} if the induced morphisms on cohomology with coefficients in ${\cal O}$ are isomorphisms for $k\leq i < m$ and injective for $k\leq i = m$. Note that the morphism ${\cal F} \rightarrow \ast$ is $1$-arranged because of the hypothesis that $H^1({\cal F} , {\cal O} )=0$. The strategy of the proof (taken from E. Brown \cite{EBrown}) will be to suppose that we have constructed ${\cal F} \rightarrow \Sigma _m$ which is $m$-arranged. Then we will construct a factorization $$ {\cal F} \rightarrow \Sigma _{m+1} \rightarrow \Sigma _m $$ where the first morphism is $m+1$-arranged. By induction this suffices to prove the theorem (we can stop as soon as we get to $m>n$). So start with the situation of ${\cal F} \rightarrow \Sigma _m$, $m$-arranged, $m\geq 1$. Let $$ {\cal C} := Cone ({\cal F} \rightarrow \Sigma _m) $$ so we have a map $\Sigma _m \rightarrow {\cal C}$ which restricted to ${\cal F}$ gives a map homotopic to the basepoint $\ast \rightarrow {\cal C}$ (this basepoint is included in the definition of $Cone$---it is the vertex of the cone over ${\cal F}$). It is easy to see using the $m$-arrangedness of our map, that the cohomology of ${\cal C}$ with coefficients in ${\cal O}$ vanishes in degrees $\leq m$. Furthermore, the $m+1$-st cohomology fits into a long exact sequence with those of ${\cal F} $ and $\Sigma _m$: $$ 0\rightarrow H^m(\Sigma _m,{\cal O} )\rightarrow H^m({\cal F} , {\cal O} ) $$ $$ \rightarrow H^{m+1}({\cal C} ,{\cal O} ) \rightarrow H^{m+1}(\Sigma _m, {\cal O} ) $$ $$ \rightarrow H^{m+1}({\cal F} , {\cal O} ) \rightarrow H^{m+2}({\cal C} , {\cal O} ) $$ $$ \rightarrow H^{m+2}(\Sigma _m, {\cal O} )\rightarrow \ldots . $$ The cohomology of $\Sigma _m$ with coefficients in ${\cal O}$ is a finite dimensional vector space, by Theorem \ref{bc} (this case is contained in the original characteristic $0$ version obtainable from \cite{Breen2}). The property of being represented by a finite dimensional vector space is closed under extensions, kernels and cokernels (\cite{kobe} Theorem 3.3 and \cite{RelativeLie} Corollary 4.10 and Theorem 4.11). Therefore the cohomology of ${\cal C}$ with coefficients in ${\cal O}$ is again a (sheaf represented by a) finite dimensional vector space. Now let $$ W^{\ast} := H^{m+1} ({\cal C} , {\cal O} ) $$ define the finite dimensional vector space $W$. We get a morphism ${\cal C} \rightarrow K(W , m+1)$ which is universal for morphisms to $K(U, m+1)$ with $U$ a finite dimensional vector space. In particular it induces an isomorphism $$ H^{m+1}(K(W, m+1), {\cal O} )\stackrel{\cong}{\rightarrow} H^{m+1}({\cal C} , {\cal O} ). $$ We will compare the previous long exact sequence with the long exact sequence that occurs at the start of the Leray-Serre spectral sequence for the morphism $p:\Sigma _m\rightarrow K(W, m+1)$. Set $$ Fib:=Fib(\Sigma _m \rightarrow K(W , m+1) ). $$ Note that we have a morphism ${\cal F} \rightarrow Fib$. The higher direct images occuring in the Leray-Serre spectral sequence for $p$ are constant local systems over the base, because $K(W,m+1)$ is $1$-connected. In other words, $$ R^ip_{\ast} {\cal O} = H^i(Fib ,{\cal O} ) \times K(W,m+1) \rightarrow K(W,m+1) . $$ Do a standard type of spectral sequence argument. First of all, for $k<m$ we prove by induction on $k$ that for all $i\leq k$, the $H^i(Fib,{\cal O} ) $ are finite dimensional vector spaces. Suppose we know this for $k-1$. Then in view of the vanishing of the cohomology of $K(W, m+1)$ with vector space coefficients (the sheaves represented by vector spaces are ${\cal O} ^a$) the terms $E^{i,j}_2$ with $1\leq i\leq m$ and $j<k$ vanish; whereas for $j=0$ we have $$ E^{i,0}_2 = H^i(Fib, {\cal O} ) $$ because of the fact that $K(W,m+1)$ is $1$-connected. Therefore the terms $E^{k,0}_2$ persist to $E_{\infty}$ and we have $$ H^k(Fib, {\cal O} ) = H^k(\Sigma _m , {\cal O} )= H^i({\cal F} , {\cal O} ). $$ In view of the hypothesis of the theorem, this proves the induction step of the first part of the argument. Incidentally we get that the morphism ${\cal F} \rightarrow Fib$ induces an isomorphism on cohomology with coefficients in ${\cal O}$ in degrees $k< m$. Now look at the term $E^{m,0}_2= H^m(Fib, {\cal O} )$. The only differential concerning it is $$ d_{m+1}: H^m(Fib , {\cal O} )\rightarrow H^{m+1}(K(W,m+1) , {\cal O} ) $$ (noting that we already have $H^0(Fib , {\cal O} ) = {\cal O} $). From our hypothesis which implies that $H^1(Fib, {\cal O} )=0$ we get, similarly, that the only differential concerning the term $E^{m+1,0}_2$ is $$ d_{m+2}: H^{m+1}(Fib , {\cal O} )\rightarrow H^{m+2}(K(W,m+1) , {\cal O} ). $$ From these and the fact that the spectral sequence abuts to the cohomology of $\Sigma _m$, we get the long exact sequence $$ 0\rightarrow H^m(\Sigma _m,{\cal O} )\rightarrow H^m(Fib, {\cal O} ) $$ $$ \rightarrow H^{m+1}(K(W,m+1) ,{\cal O} ) \rightarrow H^{m+1}(\Sigma _m, {\cal O} ) \rightarrow $$ $$ H^{m+1}(Fib , {\cal O} ) \rightarrow H^{m+2}(K(W,m+1) , {\cal O} )\rightarrow H^{m+2}(\Sigma _m, {\cal O} ). $$ Remark that $H^{m+2}(K(W, m+1),{\cal O} )=0$---this comes from Theorem \ref{bc} and it is here where we use $m\geq 1$. In particular, the morphism $$ H^{m+2}(K(W, m+1), {\cal O} )\rightarrow H^{m+2}({\cal C} ,{\cal O} ) $$ is injective for the trivial reason that the left side is $0$. Recall that the same induced morphism in degree $m+1$ was an isomorphism (by the construction of $W$). Therefore, comparing with the previous long exact sequence and using the $5$-lemma, we get that the morphism $$ {\cal F} \rightarrow Fib $$ induces isomorphisms on cohomology in degrees $\leq m$ and an injection in degree $m+1$. In other words this morphism is $m+1$-arranged. Thus we can set $$ \Sigma _{m+1}:= Fib $$ and we have completed our inductive construction to prove the theorem. \hfill $\Box$\vspace{.1in} {\em Definition:} If ${\cal F}$ satisfies the condition of Theorem \ref{representable1} then we obtain the representing $1$-connected very presentable $\Sigma ({\cal F} )$ with universal morphism $$ {\cal F} \rightarrow \Sigma ({\cal F} ). $$ We define (for any basepoint $f:Y\rightarrow {\cal F}$) $$ \pi ^{\rm vp}_i({\cal F} \times Y/Y, f):= \pi _i(\Sigma \times Y/Y, f). $$ In the latter case we usually just take a basepoint $f\in {\cal F} (Spec \, {\bf C} )$ and then denote this by $\pi ^{\rm vp}_i({\cal F} , f)$. \begin{theorem} \mylabel{fibration} Suppose ${\cal F}$ and ${\cal G}$ are $n$-stacks with basepoint $g\in {\cal G} (Spec ({\bf C} )$, which satisfy the criterion of Theorem \ref{representable1} so their shapes are representable. Suppose $$ f:{\cal F} \rightarrow {\cal G} $$ is a morphism of $n$-stacks with the following property (we denote by ${\cal H}$ the fiber over $g$): the local systems $R^if_{\ast} ({\cal O} )$ are isomorphic to ${\cal O} ^{a_i}$ on ${\cal G}$, and that the morphisms $$ R^if_{\ast}({\cal O} ) |_g \rightarrow H^i({\cal H} , {\cal O} ) $$ are isomorphisms. Suppose that $$ H^1({\cal F}, {\cal O} ) =H^1({\cal G} ,{\cal O} ) = H^1({\cal H} ,{\cal O} )={\cal O} $$ and $$ H^1({\cal F}, {\cal O} ) =H^1({\cal G} ,{\cal O} ) = H^1({\cal H} ,{\cal O} )=0. $$ Then we have a fiber sequence for the representing objects $$ \Sigma ({\cal H} )\rightarrow \Sigma ({\cal F} )\rightarrow \Sigma ({\cal G} ). $$ \end{theorem} {\em Proof:} The Leray spectral sequence for $f$ is $$ H^i({\cal G} , R^jf_{\ast}({\cal O} ))\Rightarrow H^{i+j}({\cal F} , {\cal O} ). $$ In view of the hypothesis, this becomes $$ H^i({\cal G} , {\cal O} )\otimes _{ {\cal O} } H^j({\cal H} , {\cal O} )\Rightarrow H^{i+j}({\cal F} , {\cal O} ). $$ On the other hand, we obtain a morphism of representing shapes $$ \Sigma ({\cal F} )\rightarrow \Sigma ({\cal G} ) $$ (by the universal property of $\Sigma ({\cal F} )$). Let $Fib$ denote the fiber of $\Sigma ({\cal F} )$ (over the image of the point $g$). Note that the $R^i\Sigma (f)_{\ast} ({\cal O} )$ are constant on $\Sigma ({\cal G} )$ because $\Sigma ({\cal G} )$ is $1$-connected. In particular $$ R^if_{\ast}({\cal O} ) |_g \stackrel{\cong}{\rightarrow} H^i(Fib , {\cal O} ). $$ We obtain the spectral sequence $$ H^i(\Sigma ({\cal G} ) , {\cal O} )\otimes _{{\cal O}} H^j(Fib , {\cal O} ) \Rightarrow H^{i+j} (\Sigma ({\cal F} ), {\cal O} ). $$ Note that the $\pi _i(Fib )$ are finite dimensional vector spaces (using the long exact sequence of homotopy groups of a fibration, and \cite{RelativeLie} Theorem 4.11 applied to the case of trivial base $S=\ast $). Thus $H^j(Fib , {\cal O} )$ are finite dimensional vector spaces. The composition $$ {\cal H} \rightarrow \Sigma ({\cal F} )\rightarrow \Sigma ({\cal G} ) $$ is homotopic to the constant map at the basepoint, so we get a map $$ {\cal H} \rightarrow Fib. $$ This gives maps $H^i(Fib, {\cal O} )\rightarrow H^i({\cal H} , {\cal O} )$. We claim that these are isomorphisms, which would imply that ${\cal H} \rightarrow Fib$ is a map representing the very presentable shape of ${\cal H}$, in other words $\Sigma ({\cal H} )\cong Fib$, thus giving the desired result. To prove the claim, note that the maps in question are compatible via the previous identifications, with the maps $$ R^i\Sigma ({\cal F} )_{\ast} {\cal O} \rightarrow R^if_{\ast} ({\cal O} ). $$ These in turn fit into a morphism of Leray spectral sequences. We show using a spectral sequence argument, by induction on $k$, that for all $i\leq k$ we have $$ H^i(Fib, {\cal O} )\stackrel{\cong}{\rightarrow} H^i({\cal H} , {\cal O} ) $$ or equivalently $$ R^i\Sigma ({\cal F} )_{\ast} {\cal O} \stackrel{\cong}{\rightarrow} R^if_{\ast} ({\cal O} ). $$ Suppose this is true for $k-1$. Then look at the term $E^{0,k}_2 = H^k({\cal H} , {\cal O} )$. When we look at the $r$th differential $$ d^{0,k}_r:E^{0,k}_r\rightarrow E^{r,k+1-r}_r $$ the term $E^{r,k+1-r}_r$ has not yet been touched by any term $E^{i,j}$ with $j\geq k$, and after this differential, the term $E^{r,k+1-r}$ is no longer touched by any further differentials. We have a morphism of spectral sequences (the above remarks apply to both) which induces an isomorphism on the abuttments. It follows that the morphism between spectral sequences induces an isomorphism on images of $d^{0,k}_r$. Furthermore, the morphism induces an isomorphism on the intersection (for all $r$) of the kernels of the $d^{0,k}_r$. This implies that the morphism induces an isomorphism on $E^{0,k}_2$ and we obtain the inductive step for $k$. This proves the claim and hence the theorem. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{complexify} Suppose $Y$ is a simply connected finite CW complex. Let ${\cal F} $ be the constant $n$-stack associated to the constant prestack with values $Y$ (or more precisely, with values the $n$-type $\tau _{\leq n}Y$). Fix a basepoint $y\in Y$ which also gives a basepoint section of ${\cal F}$. Let ${\cal F} \rightarrow \Sigma$ be the morphism representing the shape of ${\cal F}$. The morphisms induced by $Y\rightarrow \Sigma (Spec {\bf C} )$, $$ \pi _i(Y,y)\rightarrow \pi _i(\Sigma , y) $$ induce isomorphisms $$ \pi _i(Y,y)\otimes _{{\bf Z}} {\cal O} \cong \pi _i(\Sigma , y). $$ \end{corollary} {\em Proof:} Using the previous theorem, we can reduce by the Postnikov tower to the case $Y= K(A, n)$ for a finitely generated abelian group $A$. Then the Breen calculations imply that the morphism $$ K(A, n)\rightarrow K(A\otimes _{{\bf Z}} {\cal O} , n) $$ induces an isomorphism on cohomology with coefficients in ${\cal O}$. This implies that $$ \Sigma (K(A,n))= K(A\otimes _{{\bf Z}}{\cal O} , n), $$ which gives the statement of the corollary. \hfill $\Box$\vspace{.1in} {\bf Definition:} Fix $n$. If $Y$ is a $1$-connected finite CW complex, then we define the {\em complexification of $Y$} denoted $Y\otimes {\bf C}$ to be the $n$-stack $\Sigma ({\cal F} )$ representing the very presentable shape of the constant $n$-stack ${\cal F}$ with values $\tau _{\leq n}Y$. Note that this notion depends on $n$ because we have chosen not to treat the questions arising if we try to take $n=\infty$. {\bf Example:} If we apply this to $Y=S^2$ then we obtain $\Sigma = S^2\otimes {\bf C} $ as defined in \S 6 above. This is easy to see because, using the previous theorem, the homotopy sheaves of $\Sigma$ are ${\cal O}$ in degrees $2$ and $3$; then there are only two possibilities for $\Sigma$ and they are distinguished by the vanishing or nonvanishing of the Whitehead product. As the Whitehead product is nonzero for $S^2$ and the isomorphisms of the previous theorem are compatible with the Whitehead product (exercise), this implies that the Whitehead product for $\Sigma$ is nontrivial, therefore $\Sigma$ is equal to the $T$ defined in \S 6. We propose the above results as a way of interpreting what it means to look at the ``complexified homotopy type of a space $Y$''. We could do the same thing over the ground field ${\bf Q}$, and then we propose that this is what it means to look at the ``rational homotopy type'' of $Y$. This notion is preserved by base extension of the ground field. Of course this should all be related to the usual definitions of Quillen, Sullivan, Morgan, Hain et.al. which refer (excepting Quillen) to algebras of differential forms. In those theories, base extension is obtained by tensoring the algebra of forms with the field extension. It has always been somewhat unclear what geometric interpretation to put on this base-extension process, and we propose the above theory as a way of obtaining a reasonable interpretation. We don't, however, get into details of the relationship between the above theory and the differential-forms theories. One advantage of the present formulation is that it explains what is going on in the non-simply connected case: the shape of the constant sheaf ${\cal F}=Const(Y)$ is no longer representable by a very presentable object (except in fairly restricted cases such as finite fundamental group). Thus, the object which carries the ``rational homotopy'' information of $Y$ is the shape itself, rather than the representing object which may not exist. The shape, i.e. the functor $$ T\mapsto Hom ({\cal F} , T) $$ exists even when $Y$ is not simply connected. \subnumero{Proof of Theorem \ref{representable0}} The statement of Theorem \ref{representable0} from \S 2 is very slightly different from the statement \ref{representable1} given above. We indicate here how to get \ref{representable0}. Suppose that ${\cal F}$ is an $n$-stack on $Sch /{\bf C}$ such that for any affine algebraic group $G$, $$ K(G,1)\stackrel{\cong}{\rightarrow} Hom ({\cal F} , K(G,1)). $$ In particular this implies that $H^0({\cal F} , {\cal O} )={\cal O} $ and $H^1({\cal F} , {\cal O} )=0$. With the hypothesis that $H^i({\cal F} , {\cal O} )$ are represented by finite dimensional vector spaces, we can apply Theorem \ref{representable1} to get a morphism $$ {\cal F} \rightarrow \Sigma ({\cal F} ) $$ universal for morphisms to $1$-connected very prepresentable $T$. Note that $\Sigma ({\cal F} )$ is $1$-connected. We have to show that it is also universal for morphisms to $0$-connected very presentable $T$; suppose that $T$ is one such. We may choose a basepoint $t$, and let $G= \pi _1(T,t)$ (which is an affine algebraic group). We have a fiber sequence $$ T'\rightarrow T \rightarrow K(G,1). $$ This gives a diagram $$ \begin{array}{ccccc} Hom (\Sigma ({\cal F} ), T')& \rightarrow &Hom (\Sigma ({\cal F} ), T)&\rightarrow &Hom (\Sigma ({\cal F} ), K(G,1))\\ \downarrow & & \downarrow && \downarrow \\ Hom ({\cal F} , T')&\rightarrow & Hom ({\cal F} ,T) & \rightarrow & Hom ({\cal F} , K(G,1)), \end{array} $$ where the horizontal sequences are fiber sequences. Since $\Sigma ({\cal F} )$ is $1$-connected we have $$ K(G, 1) \stackrel{\cong}{\rightarrow } Hom (\Sigma ({\cal F} ), K(G,1)), $$ and the same holds for ${\cal F}$ by hypothesis. Therefore the vertical map on the right is an equivalence between $K(G,1)$ and our diagram becomes $$ \begin{array}{ccccc} Hom (\Sigma ({\cal F} ), T')& \rightarrow &Hom (\Sigma ({\cal F} ), T)&\rightarrow &K(G,1) \\ \downarrow & & \downarrow && \downarrow =\\ Hom ({\cal F} , T')&\rightarrow & Hom ({\cal F} ,T) & \rightarrow & K(G,1). \end{array} $$ Now note that $T'$ is a $1$-connected very presentable $n$-stack, so the vertical arrow on the left is an equivalence. Since the base $K(G,1)$ is $0$-connected, we can use the long exact sequences of homotopy for these fibrations to conclude that the vertical morphism in the middle is an equivalence. This is what we needed to know to establish the universal property of ${\cal F} \rightarrow \Sigma ({\cal F} )$ for Theorem \ref{representable0}. \hfill $\Box$\vspace{.1in} \subnumero{A relative version} While we are on the subject, we give a relative version of Theorem \ref{representable1}. Recall \cite{kobe} \cite{RelativeLie} that if $Y$ is a scheme then a $1$-connected $n$-stack ${\cal F} \rightarrow Y$ (which can also be thought of as a $1$-connected $n$-stack on the site $Sch /Y$ of schemes over $Y$) is said to be {\em very presentable} if for any basepoint section $f: Y'\rightarrow {\cal F}$ for a scheme $Y'\rightarrow Y'$, the homotopy group sheaves $\pi _i({\cal F} |_{Y'}, f)$ are vector sheaves on $Y'$. Since (for the present discussion) we have assumed ${\cal F}$ to be relatively $1$-connected, the homotopy group sheaves don't depend on the choice of basepoint (indeed, the choice of basepoint is locally unique up to homotopy which itself is unique up to---nonunique---homotopy). Therefore they descend to sheaves of abelian groups $\pi _i({\cal F} /Y)$ on $Y$. For ${\cal F}$ to be very presentable, it is equivalent to require that these be vector sheaves. We introduce the following terminology. We say that a covariant endofunctor $F$ from the category of vector sheaves on $Y$ to itself, is {\em anchored} if the natural map $$ F(U)\rightarrow Hom (Hom (F({\cal O} ),{\cal O} ), U) $$ is an isomorphism for any coherent sheaf $U$ (recall that the coherent sheaves are the injective objects in the category of vector sheaves). The above natural map comes from the trilinear map $$ F(U)\times Hom (F({\cal O} ), {\cal O} ) \times Hom (U,{\cal O} ) \rightarrow {\cal O} $$ defined by $(f,g,h)\mapsto g( F(h)(f))$. \begin{lemma} \mylabel{anchored1} (A) If $$ 0\rightarrow F' \rightarrow F \rightarrow F'' \rightarrow 0 $$ is a short exact sequence of natural transformations between covariant endofuncturs on the category of vector sheaves over $Y$, then if any two of the three endofunctors is anchored, so is the third. \newline (B) If $F$ is an anchored endofunctor which is also left exact, then $F$ is representable $F(V)= Hom (W, V)$ for a vector sheaf $W=Hom (F({\cal O} ), {\cal O} )$. \end{lemma} {\em Proof:} (A) follows from the $5$-lemma. For (B) suppose $F$ is a left exact anchored endofunctor. Set $W:= Hom (F({\cal O} ), {\cal O} )$. The natural map $F (U)\rightarrow Hom (W, U)$ is an isomorphism for coherent sheaves $U$. On the other hand, both sides are left exact in $U$. Suppose $$ 0\rightarrow U \rightarrow U' \rightarrow U'' $$ is an exact sequence with $U'$ and $U''$ being coherent sheaves. Then we obtain exact sequences $$ 0\rightarrow F(U) \rightarrow F(U' )\rightarrow F(U'') $$ and $$ 0\rightarrow Hom(W,U) \rightarrow Hom(W,U' )\rightarrow Hom(W,U''), $$ and our natural map is a morphism between these exact sequences inducing isomorphisms on the last two terms. Thus $F(U)\rightarrow Hom (W,U)$ is an isomorphism. This completes the proof in view of the fact that (locally on $S$) any vector sheaf $U$ fits into such a short exact sequence. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{anchored2} Suppose $V$ is a vector sheaf. Then the endofunctor on the category of vector sheaves defined by $$ U\mapsto H^i(K(V,m), U) $$ is anchored. \end{lemma} {\em Proof:} This follows immediately from Theorem \ref{bc} if $V$ is a vector scheme. Now suppose that we have an exact sequence $$ 0\rightarrow V' \rightarrow V'' \rightarrow V \rightarrow 0 $$ where $V''$ is a vector scheme, and where we know the lemma for $V'$. This gives a fibration sequence $$ K(V'' , m)\rightarrow K(V,m)\rightarrow K(V', m+1), $$ and taking the cohomology with coefficients in a coherent sheaf $U$ leads to a Leray spectral sequence $$ H^i(K(V', m+1), H^j(K(V'', m),U))\Rightarrow H^{i+j}(K(V,m), U). $$ The cohomology of the fiber are again coherent sheaves by Theorem \ref{bc}, so by the lemma for $V'$ the natural map occuring in the definition of ``anchored'' induces an isomorphism on the $E_2$ terms of the spectral sequence. Since the property of being anchored is preserved by kernels, cokernels and extensions, we get that the cohomology of $K(V,m)$ is anchored. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{anchored3} Suppose $T$ is a relatively $1$-connected very presentable $n$-stack over a scheme $S$. Then the endofunctor $$ U\mapsto H^i(T/S, U) $$ is anchored. \end{corollary} {\em Proof:} Decompose $T$ into a Postnikov tower where the pieces are of the form $K(V,m)$ for vector sheaves $V$ \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{representable2} Suppose $S$ is a scheme and ${\cal F} \rightarrow S$ is a morphism of $n$-stacks on $Sch /{\bf C}$. Suppose that for any vector sheaf $V$ over $S$, the cohomology $H^i({\cal F} /S, V)$ is again a vector sheaf over $S$. Suppose furthermore that $H^0({\cal F} /S, V) = V$ and $H^1({\cal F} /S, V)= 0$ for any vector sheaf $V$. Finally suppose that the functors $V\mapsto H^i({\cal F} /S, V)$ are anchored. Then the functor $$ T\mapsto Hom ({\cal F} /S , T/S ) $$ from relatively $1$-connected very presentable $n$-stacks of groupoids $T\rightarrow S$ to the same, is represented by a morphism ${\cal F} \rightarrow \Sigma$ over $S$, with $\Sigma \rightarrow S$ being a relatively $1$-connected and very presentable $n$-stack over $S$. \end{theorem} {\em Proof:} Follow the same outline as for the proof of Theorem \ref{representable1}. We try to find a relatively $1$-connected very presentable $\Sigma \rightarrow Y$ with a morphism $$ {\cal F} \rightarrow \Sigma $$ inducing an isomorphism on cohomology with coefficients in any coherent sheaf $U$ on $S$ (the isomorphism for coefficients in any vector sheaf $U$ then follows by resolving $U$ by coherent sheaves). We say that a morphism ${\cal F} \rightarrow \Sigma _m$ is {\em $m$-arranged} if the induced morphisms on cohomology with coefficients in any coherent sheaf $U$ on $S$ are isomorphisms for $k\leq i < m$ and injective for $k\leq i = m$. Note that the morphism ${\cal F} \rightarrow S$ is $1$-arranged because of the hypothesis that $H^1({\cal F} /S , U )=0$. Thus we may take $\Sigma _1 := S$. The strategy of the proof will be to suppose for some $m\geq 1$ that we have constructed ${\cal F} \rightarrow \Sigma _m$ which is $m$-arranged. Then we will construct a factorization $$ {\cal F} \rightarrow \Sigma _{m+1} \rightarrow \Sigma _m $$ where the first morphism is $m+1$-arranged. By induction this suffices to prove the theorem. Make the same constructions, using the same notations (which we won't repeat here) as in the proof of Theorem \ref{representable1}. Along the way, replace the cohomology with coefficients in ${\cal O}$ (and the higher direct images of ${\cal O}$ etc.) by cohomology with coefficients in any coherent sheaf $U$ on $S$. We obtain the first long exact sequence (actually valid for any vector sheaf $U$ as coefficients) $$ 0\rightarrow H^m(\Sigma _m/S,U )\rightarrow H^m({\cal F} /S, U ) $$ $$ \rightarrow H^{m+1}({\cal C} /S,U) \rightarrow H^{m+1}(\Sigma _m/S, U) $$ $$ \rightarrow H^{m+1}({\cal F} /S, U ) \rightarrow H^{m+2}({\cal C} /S, U ) $$ $$ \rightarrow H^{m+2}(\Sigma _m/S, U )\rightarrow \ldots . $$ The cohomology of $\Sigma _m$ with coefficients in a vector sheaf is again a vector sheaf, by Corollary \ref{cohVPisVS} above. The property of being represented by a finite dimensional vector space is closed under extensions, kernels and cokernels (\cite{kobe} Theorem 3.3 and \cite{RelativeLie} Corollary 4.10 and Theorem 4.11). Therefore the cohomology of ${\cal C}$ with coefficients in a vector sheaf $U$ is again a vector sheaf. When we come to the construction of $W$ we need to say something more---this is the reason for introducing the notion of ``anchored'' above. The functor $$ U\mapsto H^{m+1}({\cal C} , U) $$ is anchored. This comes from the facts that the cohomology of $\Sigma _m$ is anchored by Corollary \ref{anchored3}, that the cohomology of ${\cal F}$ is anchored by hypothesis, and the fact that being anchored is preserved by kernels, cokernels and extensions (Lemma \ref{anchored1}). On the other hand, the fact that the cohomology of ${\cal C}$ vanishes in degrees $0<i\leq m$ (note that $m\geq 1$) implies that the above functor is left-exact. Therefore by Lemma \ref{anchored1} (B) it is representable by a vector sheaf $W$: we have $$ H^{m+1}({\cal C} , U) = Hom (W, U). $$ In particular there is a tautological class in $H^{m+1}({\cal C} , W)$ corresponding to a morphism ${\cal C} \rightarrow K(W, m+1)$, and this morphism is universal for morphisms from ${\cal C}$ to things of the form $K(U,m+1)$. In particular it induces an isomorphism $$ H^{m+1}(K(W, m+1), U )\stackrel{\cong}{\rightarrow} H^{m+1}({\cal C} , U ). $$ Again set $$ Fib:=Fib(\Sigma _m \rightarrow K(W , m+1) ). $$ Note that we have a morphism ${\cal F} \rightarrow Fib$. Compare the first long exact sequence with the long exact sequence that occurs at the start of the Leray-Serre spectral sequence for the morphism $p:\Sigma _m\rightarrow K(W, m+1)$, using the same argument as in the proof of Theorem \ref{representable1}. We need to know that the morphism $$ H^{m+2}(K(W, m+1), U )\rightarrow H^{m+2}({\cal C} ,U ) $$ is injective for cohomology with coefficients in a coherent sheaf $U$ (recall that only coherent sheaves occur as coefficients for the cohomology in the definition of arrangedness---one goes back to the general case after the induction on $m$ is finished). To prove this we again show that $H^{m+2}(K(W, m+1), U )=0$ (note that this wouldn't be true if $U$ were not a coherent sheaf and that is the reason why we restrict to coherent sheaves in the definition of arrangedness). In fact, using that $m\geq 1$ and following the argument of \ref{representable1} we get that $$ H^{m+2}(K(W, m+1), U )=Ext ^1(W, U). $$ However, a coherent sheaf $U$ is an injective object in the category of vector sheaves \cite{RelativeLie} Lemma 4.17, so $Ext ^1(W, U)=0$. This gives a proof of the desired statement. Alternatively one can obtain a proof using a spectral sequence argument with a resolution $$ 0\rightarrow V \rightarrow V' \rightarrow V'' \rightarrow W \rightarrow 0 $$ of $W$ by vector schemes (decompose this into two short exact sequences and use a Leray spectral sequence argument for each of the corresponding fibration sequences). After that the rest of the argument works exactly the same way as in Theorem \ref{representable1} (calling upon Theorem \ref{bc} in the relative case as necessary). We don't repeat this. \hfill $\Box$\vspace{.1in}
1997-11-30T02:08:36
9712
alg-geom/9712001
en
https://arxiv.org/abs/alg-geom/9712001
[ "alg-geom", "math.AG" ]
alg-geom/9712001
Richard Mayer
Richard Mayer
Coupled Contact Systems and Rigidity of Maximal Dimensional Variations of Hodge Structure
24 pages, latex2e uses amsart.cls
null
null
null
null
In this article we prove that locally Griffiths' horizontal distribution on the period domain is given by a generalized version of the familiar contact differential system. As a consequence of this description we obtain strong local rigidity properties of maximal dimensional variations of Hodge structure. For example, we prove that if the weight is odd then there is a unique germ of maximal dimensional variation of Hodge stucture though every point of the period domain. Similar results hold if the weight is even with the exception of one case.
[ { "version": "v1", "created": "Sun, 30 Nov 1997 01:08:37 GMT" } ]
2007-05-23T00:00:00
[ [ "Mayer", "Richard", "" ] ]
alg-geom
\section{Introduction} Variations of Hodge structure are integral manifolds of Griffiths' horizontal distribution on the period domain \(D\) of a fixed integral lattice which are stable under the action of a discrete subgroup of the isometry group of \(D\). This distribution is defined in terms of the filtrations \(\{F^p\}\) by \[ \frac{\partial F^p}{\partial z_i}\subset F^{p-1}. \] The horizontal distribution is not completely integrable in general if the weight is greater than one, so integral manifolds must have lower dimension than the distribution. There are two fundamental questions for not completely integrable differential systems: \begin{itemize} \item What is the maximal dimension an integral manifold can have? \\ \item What is the structure of the integral manifolds that attain the dimension bound? \end{itemize} We shall examine these two questions for the horizontal system in detail. The tool that makes this possible is the fact that the horizontal differential system is locally given by coupled matrix valued contact systems. (See Section 4.) We derive two consequences of this description. The first one is a new proof of the main result of \cite{C-K-T} which says that there are explicit quadratic functions \(q^{even}_1, q^{even}_2, q^{even}_3\) (weight even) and \(q^{odd}_1, q^{odd}_2\) (weight odd) of the Hodge numbers \(h^{i,j}\) which give sharp upper bounds for the dimension of a variation of Hodge structure. (Theorem~\ref{dimboundthm}). In fact, we cover a case which is missing from \cite{C-K-T}, this will be explained in Section 4. The second consequence is the main result of this work which generalizes Carlson's rigidity theorem for weight two variations (cf. \cite{C1}) to arbitrary weights. The precise statement is as follows: \begin{thm}\label{mainthm} Let \(D\) denote the period domain and let \(w\) be the weight. \begin{enumerate} \item Assume that one of the following holds: \begin{enumerate} \item \(w=2k+1\), \(h^{k,k+1}>2\) and all the other Hodge numbers are greater than one \item \(w=2k\), all the Hodge numbers are greater than one and the maximum dimension is \(q^{even}_1\) \end{enumerate} then there is a unique maximal dimensional germ of variations of Hodge structure through each point of \(D\), i.e., there is a unique maximal dimensional foliation the leaves of which are the maximal dimensional variations of Hodge structure. \item Let \(S_1, S_2\subset D\) be two maximal dimensional variations of Hodge structure. Assume that one of the following holds: \begin{enumerate} \item \(w=2k\), \(dim(S_1)=dim(S_2)=q^{even}_2\), \(h^{k+1}>2\), the other Hodge numbers are greater than one and \(h^k\geq 4\) is even \item \(w=2k\), \(dim(S_1)=dim(S_2)=q^{even}_3\) and \(h^{k+1}>2\) \end{enumerate} then there is an element \(g\in Aut(D)\) such that \(g\cdot S_1=S_2\) holds in a neighborhood. \item Let \(w=2k\) and let \(E\) denote a maximal dimensional integral element. If \(dim(E)=q^{even}_2\) and \(h^k\) is odd then there is an infinite dimensional family of germs of maximal dimensional integral manifolds tangent to \(E\), i.e. flexibility holds. \end{enumerate} \end{thm} In the remaining special cases when the Hodge numbers are small (i.e., they do not satisfy the requirements of the theorem) rigidity does not hold. In fact, in these cases we have the behavior of the classical contact system explained in Section~\ref{contactsection}. The outline of the paper is as follows. In Section 2 we give the necessary definitions of exterior differential systems and treat the case of the classical contact system. In Section 3 we introduce local coordinate systems on the period domain. Using these coordinates we show in Section 4 that the horizontal system is locally equivalent to a system given by analogs of the classical contact system. Analyzing this new system leads to the proof of the main theorem in Section 5. The author would like to thank his advisor James A. Carlson for suggesting this problem to him and for the support throughout his graduate studies. \section{Exterior Differential Systems} The main reference for this section is \cite{B}. Let \(X\) be a manifold, let \(T^*\) denote the cotangent sheaf of \(X\) and \(\bigwedge ^* T^*\) the associated deRham algebra. We will be concerned with complex manifolds and the holomorphic cotangent sheaf, but most of the results remain true even if we consider \(C^\infty \) real manifolds. \subsection{Integral Elements and Manifolds} \begin{defi} A differential system is defined by an ideal \(\mathcal{I} \subset \bigwedge ^* T^*\) which is closed under exterior differentiation, i.e., \(d\mathcal{I}\subset\mathcal{I}\). An integral manifold of \(\mathcal{I}\) is a holomorphic mapping \(i\):~ \(S \longrightarrow X\) such that \(i^*\omega = 0\) for each germ \(\omega\) of \(\mathcal{I}\). \end{defi} Often we will take \(i\) to be the inclusion and we will talk about the {\it germ of an integral manifolds} by which we will mean an equivalence class of integral manifolds where two integral manifolds are equivalent if their images agree in a neighborhood of a point. Let \(i\):~ \(S \longrightarrow X\) be an integral manifold of \(\mathcal{I}\). If \(s \in S \) and \(E = T_s S \subset T_s M\) is the tangent space to \(S\) at \(s\) then for each \(\omega \in \mathcal{I}\) \(\omega_E = 0\) where \(\omega_E\) denotes the restriction of \(\omega\) to \(E\). It is clear that the vanishing of \(i^*\omega\) at each point depends only on the tangent space of \(S\). This leads to the following definition. \begin{defi} A linear subspace \(E \subset M\) is an integral element of \(\mathcal{I}\) if \(\omega_E = 0\) for each \(\omega \in \mathcal{I}\). \end{defi} Note that in general an integral element is not necessarily tangent to an integral manifold, see \cite{B} for examples. Suppose now that \(\mathcal{I}\) is generated algebraically by the 1-forms \(\{\omega^1, \ldots, \omega^{n-k}\}\) where \(n\) is the dimension of \(M\) and let \({\mathcal{D}}_{v} \subset T_{v} M\) denote the linear space of tangent vectors that are annihilated by \(\mathcal{I}\). If \({\mathcal{D}}_{v}\) has the same dimension for each \(v\in M\) then these subspaces define a {\it distribution} in the tangent bundle. The dimension of the distribution \(\mathcal{D}\) is the dimension of \({\mathcal{D}}_{v}\). From now on we will consider differential systems that are defined by 1-forms of this type and we will use the words differential system and distribution interchangeably. The condition that \(\mathcal{I}\) is closed means that \(d\omega^i\) is given as an algebraic combination of \(\{\omega^1, \ldots, \omega^{n-k}\}\) for each \(i\). This condition \((F)\) is called the {\it Frobenius condition}. It is not hard to see (cf. \cite{W} Proposition 2.30) that the Frobenius condition for differential forms is equivalent to the following for a distribution \(\mathcal{D}\): for any two holomorphic vector fields \(X\) and \(Y\) lying in \(\mathcal{D}\), the Lie bracket [\(X,Y\)] also lies in \(\mathcal{D}\). The fundamental integrability result for such systems is the following theorem of Frobenius. \begin{thm}[Frobenius]\label{frobenius} Let \(\mathcal{I}\) be a differential ideal generated algebraically by the linearly independent 1-forms \(\{\omega^1, \ldots, \omega^{n-k}\}\). There is a local coordinate system \((y^1,\ldots ,y^n)\) at each point of \(M\) such that \(\mathcal{I}\) is generated by \(\{dy^{k+1}, \ldots, dy^{n}\}\) if and only if the condition \((F)\) holds. \end{thm} If such coordinate system exists then clearly there is a \(k\)-dimensional integral manifold through each point of \(M\). For this reason such differential systems are called {\it completely integrable}. In this work we will be interested in differential systems that are not completely integrable. If this is the case then the maximal possible dimension of an integral manifold must be strictly less than the dimension of the distribution. It is an interesting question to determine this dimension. \subsection{Contact Differential Systems}\label{contactsection} Let us start with a differential system defined by a single 1-form. \begin{defi} Let \(M=\C^{2n+1}\) with coordinates \((x_1,\ldots,x_n,y_1,\ldots,y_n,z)\). The differential system whose differential ideal is generated by the 1-form \[ \omega = dz - \sum_{i=1}^{n}x_idy_i \] is called the contact system. \end{defi} \begin{rem} Sometimes we will refer to this system as the classical contact system in order to distinguish it from matrix valued analogs that will be discussed later. \end{rem} Let us observe that the distribution defined by this system has dimension \(2n\). However, the system is not completely integrable (as we will see in a moment) so any integral manifold must have dimension less than \(2n\). An easy computation shows that \[\omega \wedge (dw)^{\wedge n} = \pm dz\wedge dx_1\wedge \ldots \wedge dx_n \wedge dy_1 \wedge \ldots \wedge dy_n \] so \(d\omega \wedge \omega\) is clearly nonzero which implies that the Frobenius condition can not hold for the contact system. In light of this, our next task is going to be to determine the maximal dimension an integral manifold can have. In what follows the notation \(f(x_I,y_J)\) means that the function \(f\) depends on the variables \(x_i, i \in I\) and \(y_j, j \in J\). \begin{prop}\label{contactmax} Let \(S \subset {\C}^{2n+1}\) be a \(k\)-dimensional integral manifold of the contact system going through the point \(s\). Then there exist two disjoint sets of indices \(I,J\subset \{1,\ldots,n\}\) of total length \(k\) such that in a neighborhood of \(s\), \(S\) is parameterized by a set of holomorphic functions \((g_1(x_I,y_J),\ldots,g_{2n+1}(x_I,y_J))\) where the variables \(x_{i}\) and \(y_{j}\) \(({i\in I},{j \in J})\) are independent on \(S\). \end{prop} \begin{rem} By definition the variables \(x_{i}\), \(y_{j}\) \(({i\in I},{j \in J})\) are independent on \(S\) if the differential 1-forms \(dx_{i}\) and \(dy_{j}\) \(({i\in I},{j \in J})\) are linearly independent at each point of a neighborhood of \(S\). This also means that independent variables form part of a coordinate system in a neighborhood. \end{rem} \begin{proof} In a neighborhood \(U\) of \(s\) the manifold \(S\) is given by a set of holomorphic functions \((g_1,\ldots,g_{2n+1})\). Let \(L\subset \{x_i, y_j,z\}\) denote the set of {\em independent} coordinates that the functions \(g_i\) depend on, (\(|L|=k\)). Let \(I\subset L\), \(J\subset L\) be the sets consisting of variables \(x_{i}\), \(y_{j}\) respectively. Since \(S\) is an integral manifold, \(\omega\) and \(d\omega\) must vanish on \(S\). This implies that the coordinate \(z\) can be expressed as a function of the other variables so we can assume that the functions \(g_i\) do not depend on \(z\), i.e., \(z\notin L\). Suppose that there is a pair of coordinates \((x_i,y_i)\), \(x_i,y_i\in L\) such that \(dx_i \wedge dy_i \) does not vanish on \(U\). Since \[ 0 = d\omega = \sum_{l=1}^{n}dx_l \wedge dy_l \;\;\;\;\;\;\mathrm{on}\;\;\;S \] there must be another index \(j\neq i\) such that \(dx_j \wedge dy_j \) does not vanish on \(U\) and the variables \(x_j,y_j\) are not in \(L\) since they depend on \(x_i,y_i\). Now \[0 \neq dx_j\wedge dy_j = (\frac{\partial x_j}{\partial x_i} \frac{\partial y_j}{\partial y_i} - \frac{\partial y_j}{\partial x_i} \frac{\partial x_j}{\partial y_i})dx_i\wedge dy_i + \eta\] where \(\eta\) is a 2-form that does not contain \(dx_i\wedge dy_i\). This means that one of the coefficient functions, say \(\frac{\partial x_j}{\partial x_i}\), must be nonvanishing so we can introduce the coordinate change \(x_i \longleftrightarrow x_j\) in a neighborhood of \(s\), by which we mean that we replace the coordinate \(x_i\) by \(x_j\) and leave the other coordinates intact. This means that we replaced the independent pair of variables \((x_i,y_i)\) by another independent pair \((x_j,y_i)\) thereby decreasing \(|I\cap J|\). Continuing in the same manner, after at most \(k\) steps we arrive at two {\em disjoint} sets of indices \(I,J\subset \{1,\ldots,n\}\) that satisfy the requirements of the proposition. \end{proof} \begin{rem}\label{fixj} It is clear from the proof that if we choose the set \(J\) to be maximal in the sense that no other independent set of variables that \(S\) depends on contains more \(y_j\)'s then we can find the set \(I\) with \(|I\cap J|=0\) without changing the set \(J\). This simple remark will be used quite frequently later, sometimes without explicit reference to it. \end{rem} \begin{cor} Let \(S \subset {\C}^{2n+1}\) be an integral manifold of the contact system. Then \(dim(S) \leq n\). \end{cor} \begin{proof} By Proposition~\ref{contactmax}, \(S\) is locally defined by functions \(g_i\) depending on at most \(n\) independent variables. This clearly implies the claim. \end{proof} The following result of \cite{A} (Appendix 4) gives a complete characterization of maximal dimensional integral manifolds of the contact system. \begin{thm}\label{arnold} For any partition \(I+J\) of the set of indices \(\{1,\ldots,n\}\) into two disjoint subsets and for any function \(f(x_I,y_J)\) of \(n\) variables \(x_i,y_j\) \((i\in I, j\in J)\), the formulas \[y_i = \frac{\partial f}{\partial x_i} \; (i\in I),\; \; \; x_j = -\frac{\partial f}{\partial y_j}\; (j\in J),\; \; \; z = f - \sum_{i\in I}x_i\frac{\partial f}{\partial x_i} \] define an \(n\)-dimensional integral manifold of the contact system. Conversely, every \(n\)-dimensional integral manifold is defined in a neighborhood of every point by these formulas for a choice of the subset \(I\) and for some generating function \(f\). \end{thm} \begin{proof} A simple calculation shows that the manifold defined by the formulas in the first part of the Theorem is indeed an integral manifold of the contact system. Assume now that \(S\) is an \(n\)-dimensional integral manifold. By Proposition~\ref{contactmax} we can find \(n\) {\em independent} variables \(x_I,y_J\) such that \(S\) is defined by functions of these variables. Let \[f = z + \sum_{i\in I} x_iy_i\] and notice that \[df = \sum_{i\in I}y_idx_i - \sum_{j\in J}x_jdy_j \] since \(0 = \omega = dz + \sum x_ldy_l\) on \(S\). Also, the differential forms \(dx_i,\, dy_j\) are independent so it follows from the defining equation of \(f\) that \[y_i = \frac{\partial f}{\partial x_i} \; (i\in I),\; \; \; x_j = -\frac{\partial f}{\partial y_j}\; (j\in J)\] which finishes the proof. \end{proof} \begin{rem}\label{contactflexible} Let us mention here that if we fix an \(n\)-dimensional integral element at the point \(s\), then there exists an integral manifold that is tangent to the given integral element (see Theorem~\ref{exptrick}) but this integral manifold is by no means unique. This follows from the fact that we can choose \(f\) in the previous theorem to be an arbitrary holomorphic function and fixing an integral element specifies only the linear part of \(f\). This means that to any integral element of dimension \(n\) there is an infinite dimensional family of germs of integral manifolds that are tangent to the integral element at the point \(s\). \end{rem} \section{Variations of Hodge Structure} In this section we will consider variations of Hodge structure as integral manifolds of the horizontal distribution. First let us recall the necessary definitions. \subsection{Hodge Structures and Classifying Spaces} For a more detailed discussion, proofs of the basic results, as well as geometric motivation for some of the following definitions, see \cite{G1} and \cite{G2}. Let us fix a finite dimensional real vector space \(H_{\mathbb R}\) and a lattice \(H_{\Z}\subset H_{\mathbb R}\) together with an integer \(w\), (\(w\) will be referred to as the weight). Suppose that we are given a non-degenerate bilinear form \(Q\) on \(H_{\mathbb R}\) which satisfies \(Q(x,y) = (-1)^w Q(y,x)\) for all \(x,y\in H_{\mathbb R}\) and takes rational values on \(H_{\Z}\). Let \(H_{\C} = H_{\mathbb R} \otimes_{\mathbb R} \C\) denote the complexification of \(H_{\mathbb R}\). \begin{defi} A Hodge structure of weight \(w\) on \(H_{\mathbb R}\) is a decomposition \[H_{\C} = \bigoplus_{p+q = w}H^{p,q}\] such that \(H^{p,q}\) and \(H^{q,p}\) are complex conjugate to each other with respect to \(H_{\mathbb R}\). The integers \(h^{p,q} = dim\,H^{p,q}\) are the Hodge numbers. \end{defi} Let \(h(x,y) = (-1)^{w(w-1)/2}Q(x,\overline{y})\) and let \(C\) denote the Weil operator that acts on \(H^{p,q}\) by multiplication by \(i^{p-q}\). \begin{defi} The Hodge structure is weakly polarized if \[Q(H^{p,q},H^{r,s})=0 \; \; \; \;\rm{for} \;\;\;\; (r,s) \neq (q,p).\] It is strongly polarized if, in addition, the form \(h_C(x,y) = h(Cx,y)\) is positive hermitian. \end{defi} To each Hodge structure of weight \(w\) we can assign the {\it Hodge filtration} \[H_{\C} \supset \ldots \supset F^{p-1} \supset F^{p} \supset F^{p+1}\supset \ldots \supset 0\] where \[F^{p} = \bigoplus_{i\geq p}H^{i,w-i}.\] This filtration has the property \[H_{\C} = F^{p} \oplus \bar{F}^{w-p+1} \; \; \; \rm{for}\; \;\rm{each} \; \;p\] where the bar denotes complex conjugation. If the Hodge structure is strongly polarized then the associated Hodge filtration satisfies \(Q(F^p,F^{w-p+1})=0\) for all \(p\), and the form \(h_C(x,y)\) is positive hermitian. It is not hard too see that a Hodge filtration with the listed properties determines a strongly polarized Hodge structure so these two notions are in fact equivalent. Let \(\check{D}\) denote the set of weakly polarized Hodge structures (or Hodge filtrations) with fixed Hodge numbers \(h^{p,q}\), \((p+q=w)\) and let \(D \subset \check{D}\) be the subset of strongly polarized Hodge structures. There is a natural complex structure on the set \(\check{D}\) which can be described as follows. Since a Hodge structure is determined by its associated filtration, we can view \(\check{D}\) as a subset of a product of Grassmannian manifolds. The conditions \(F^{p} \supset F^{p+1}\) and \(Q(F^p,F^{w-p+1})=0\) are algebraic so \(\check{D}\) is a subvariety of a product of Grassmannians. As such it has the structure of a complex projective variety. It can be checked that the special orthogonal group \(G_{\C} = SO(Q,\C)\) acts transitively on \(\check{D}\), in particular \(\check{D}\) is a complex manifold. The subset \(D\subset \check{D}\) is open in the complex topology and it is homogeneous for the corresponding real group \(G_{\mathbb R}\). The classifying space \(D\) is sometimes referred to as the {\it period domain} of Hodge structures. \subsection{Canonical Coordinates on the Period Domain} In this section we will describe two sets of local coordinate systems on the period domain. These coordinate systems will be used to investigate the local properties of differential systems on the period domain. The two coordinate systems are equivalent to each other and the reason they will both be used is that certain results can be described more conveniently in one of these systems. \subsubsection{Lie algebra coordinates} Let \(\mathfrak{g}_{\C}\) denote the Lie algebra of \(G_{\C}\) and let us fix a reference Hodge structure \(H\in D\). \(H\) defines a Hodge structure of weight 0 on \(\mathfrak{g}_{\C}\) (cf. \cite{S}) where \[\mathfrak{g}^{p,-p} = \{\phi \in \mathfrak{g}_{\C}\,|\, \phi(H^{r,s})\subset H^{r+p,s-p}\; \rm{for \;all} \;(r,s)\}.\] \(\mathfrak{g}_{\C}\) can be decomposed as \[\mathfrak{g}_{\C} = \mathfrak{g}^{-}\oplus \mathfrak{g}^{0}\oplus \mathfrak{g}^{+}\] where the subalgebras \(\mathfrak{g}^{-}\), \(\mathfrak{g}^{0}\) and \(\mathfrak{g}^{+}\) are defined as \[\mathfrak{g}^{-} = \bigoplus_{p<0}\mathfrak{g}^{p,-p},\;\;\;\; \mathfrak{g}^{0} = \mathfrak{g}^{0,0},\;\;\;\; \mathfrak{g}^{+} = \bigoplus_{p>0}\mathfrak{g}^{p,-p}.\] Recall that \(\check{D} \cong G_{\C}/B\) as a homogeneous space where \[B = \{g\in G_{\C} \,|\, g(F^{p}) \subset F^{p}\}\] is the isotropy subgroup of the reference Hodge structure. The Lie algebra of \(B\) is \[\mathfrak{b} = \mathfrak{g}^{0}\oplus \mathfrak{g}^{+}\] so the complement \(\mathfrak{g}^{-}\) can be identified with the holomorphic tangent space of \(\check{D}\). Note that the Lie bracket is compatible with the Hodge decomposition, i.e., \begin{equation}\label{liecompat} [\mathfrak{g}^{p,q},\mathfrak{g}^{r,s}]\subset \mathfrak{g}^{p+r,q+s}. \end{equation} \begin{defi}\label{algebracoord} The local Lie algebra coordinates in a neighborhood of the reference Hodge structure \(H\in D\) are given by the map \[\Phi:\mathfrak{g}^{-}\longrightarrow D,\;\;\;\; N\longmapsto (exp(N))\cdot H\] \end{defi} Note that the image will lie in \(D\) if the norm of \(N\) is small since \(D\subset \check{D}\) is an open subset. Also, the differential of the map is the identity so \(\Phi\) defines local coordinates in a neighborhood. The actual local coordinates will be chosen by specifying a convenient basis in \(\mathfrak{g}^{-}\). This is our next task (cf. \cite{C-K-T}). \subsubsection{Hodge Frames and Block Decomposition} \begin{defi} A Hodge frame for \(H\) is a set of bases \(B^{p,q} = \{B^{p,q}_j \,|\, j=1,\ldots,h^{p,q}\}\) such that \begin{enumerate} \item[(i)] \(B^{p,q}\) is an \(h_C\)-unitary basis of \(H^{p,q}\) \item[(ii)] \(B^{p,q} = \overline{B^{q,p}}.\) \end{enumerate} \end{defi} It is clear that the matrix of the bilinear form \(Q\) relative to a Hodge frame has a block decomposition such that the only nonzero blocks are on the antidiagonal and these blocks are identity matrices up to a sign. Let us denote by \(M[i,j]\) the matrix whose only nonzero block is the matrix \(M\) of size \(h^{w-i,i}\times h^{w-j,j}\) in the \((i,j)\) position. Then the matrix of the polarization \(Q\) can be written as \[Q=\sum_{k=0}^{w} (-1)^kI[k,w-k]\] where \(I\) stands for the identity matrix. Similarly, the matrix of an endomorphism \(X\) of \(H_{\C}\) relative to a Hodge frame also has a block decomposition \(\sum X_{i,j}[i,j]\) where the matrix \(X_{i,j}\) represents an endomorphism from \(H^{w-j,j}\) to \(H^{w-i,i}\). Now let us consider the block decomposition of a general element \(X\in \mathfrak{g}^{-}\). Clearly the matrix of \(X\) is strictly lower triangular and because it is and element of the orthogonal Lie algebra it satisfies the relation \(X^tQ+QX=0\). In terms of the \(X_{i,j}\) we have \begin{equation}\label{ortcond} (-1)^{w-i}(X_{i,j})^t+(-1)^{w-j}X_{w-j,w-i}=0\;\;\;\;i,j\in \{0,\ldots ,w\}. \end{equation} \begin{ex} To illustrate the definitions above let us consider the weight two case. The matrix of \(Q\) will be \[Q= \begin{pmatrix} 0&0&I_{h^{2,0}}\\ 0&-I_{h^{1,1}}&0\\ I_{h^{2,0}}&0&0 \end{pmatrix} \] and a general element \(X\in \mathfrak{g}^{-}\) can be written as \[X= \begin{pmatrix} 0&0&0\\ X_{1,0}&0&0\\ X_{2,0}&X_{2,1}&0 \end{pmatrix} \] where \((X_{1,0})^t=X_{2,1}\) and \((X_{2,0})^t=-X_{2,0}\) because of Equation~\ref{ortcond}. In this case \(dim(D) = h^{2,0}h^{1,1} + \frac{1}{2}h^{2,0}(h^{2,0}-1)\) since this is the number of coordinates in the Lie algebra coordinate system given by the entries of the matrices \(X_{i,j}\). \end{ex} \subsubsection{Lie Group Coordinates}\label{groupcord} Now it is easy to describe the other coordinate system. Let \(G^-\) denote the unipotent Lie group which corresponds to the Lie algebra \(\mathfrak{g}^{-}\). \begin{defi}\label{groupcoord} The local Lie group coordinates in a neighborhood of the reference Hodge structure \(H\in D\) are given by the map \[\Psi:G^{-}\longrightarrow D,\;\;\;\; g\longmapsto g\cdot H\] \end{defi} It follows from the discussion of the Lie algebra coordinates that this is a coordinate system in a neighborhood of \(H\). If we fix a Hodge frame of \(H\) then an element \(Y\in G^-\) has a block decomposition \(\sum Y_{i,j}[i,j]\) which is induced by the corresponding Lie algebra element \(log(Y)\). \begin{rem} Let us note that the blocks \(Y_{i,j}\) will satisfy the condition corresponding to Equation~\ref{ortcond} and we will analyze this later. \end{rem} \begin{rem} To visualize the matrices in the Lie group coordinate system see Example~\ref{visex}. \end{rem} \subsection{Variations of Hodge Structure} \subsubsection{Horizontal Distribution} In order to define a distribution we have to specify a fixed dimensional linear subspace of the holomorphic tangent space at each point of \(\check{D}\). The holomorphic tangent space was identified by \(\mathfrak{g}^{-}\) so we can proceed as follows. \begin{defi} The horizontal distribution on \(\check{D}\) is given by the vector space \(\mathfrak{g}^{-1,1}\subset \mathfrak{g}^{-}\) at each point of \(\check{D}\). \end{defi} This distribution is holomorphic and homogeneous as can be seen by elementary Lie group theory (cf. \cite{S}). Integral manifolds of the horizontal distribution are sometimes referred to as {\it horizontal maps}. We have the following \begin{prop}\label{nonintegrable} The horizontal differential system is not completely integrable provided that \(\mathfrak{g}^{-2,2}\neq 0\). \end{prop} \begin{proof} Since \([\mathfrak{g}^{-1,1},\mathfrak{g}^{-1,1}]=\mathfrak{g}^{-2,2}\), the Proposition follows from Theorem~\ref{frobenius}. \end{proof} \begin{defi} Let \(\Gamma\) denote a properly discontinuous group of automorphisms of \(D\) and let \(M\) be a complex manifold. A holomorphic map \(\Phi : M \longrightarrow \Gamma~\backslash~ D\) is a variation of Hodge structure if the map \(\Phi\) is locally liftable to \(D\) and the local liftings are horizontal (i.e., integral manifolds of the horizontal differential system). \end{defi} In this work we will be interested in the local properties of variations of Hodge structure. \begin{defi} A germ of a variation of Hodge structure is an equivalence class of integral manifolds \(\Phi :U \longrightarrow D\) of the horizontal system, where two integral manifolds are equivalent if they have the same image in a neighborhood of a point of \(D\). \end{defi} \subsubsection{Local Description of the Horizontal Differential System} Let us consider the description of the horizontal differential system in terms of the Lie algebra and Lie group coordinate systems. At the reference Hodge structure the horizontal system is defined by the differential 1-form entries of the matrices that are at least two steps below the diagonal \begin{equation}\label{systemorigin} dX_{i,j}\;\;\;\;\rm{for}\;\;\;\; i\geq j+2. \end{equation} This is clear since at the reference point the horizontal distribution is defined by the subspace \(\mathfrak{g}^{-1,1}\) and the differential forms in Equation~\ref{systemorigin} give the dual of this subspace. Let us now determine the differential forms that define the horizontal differential system in a neighborhood of the reference structure. Since the horizontal system is homogeneous it is given by the left invariant extensions of the differential forms in Equation~\ref{systemorigin}. Let \(\Omega\) denote the Maurer-Cartan matrix, which can be written as \(\Omega = exp(-X)d\,exp(X)\) or \(\Omega = Y^{-1}d\,Y\) in the Lie algebra and Lie group coordinate systems, respectively. Then, in light of the above discussion, we have the following. \begin{prop}\label{maurer} The horizontal differential system in a neighborhood of the reference Hodge structure is given by the differential form entries of the matrices \begin{equation}\label{maurerequation} \Omega_{i,j}\;\;\;\;\rm{for}\;\;\;\; i\geq j+2. \end{equation} \end{prop} \subsubsection{Integral Elements of the Horizontal Differential System} Now that we have a local description of the horizontal system we can examine the question: for which integral elements can we find an integral manifold tangent to it? First let us give a characterization of the integral elements. \begin{prop}\label{commutative} Let \(E\subset \mathfrak{g}^{-1,1}\) be a linear subspace. Then \(E\) is an integral element of the horizontal differential system if and only if it is an abelian subalgebra. \end{prop} \begin{proof} See \cite{C-K-T} Proposition 2.2 or Section~\ref{proofcom} for another proof. \end{proof} Using this result we find the following theorem. \begin{thm}\label{exptrick} Let \(E\subset \mathfrak{g}^{-1,1}\) be an integral element at \(v\in D\). Then there is a germ of an integral manifold through \(v\) which is tangent to \(E\). \end{thm} \begin{proof} Let \((e_1,\ldots,e_k)\) be a basis of \(E\). According to Proposition~\ref{commutative} the Lie brackets \begin{equation}\label{bracketcomm} [e_i,e_j]=0\;\;\;\; \rm{for} \;\;\;\; i,j\in (1,\ldots,k) \end{equation} and since \((e_1,\ldots,e_k)\subset \mathfrak{g}^{-1,1}\) the basis elements \(e_i\) are represented by matrices with all blocks zero except the ones one step below the diagonal. Our aim is to show that the following map \(\varphi\) is an integral manifold in a neighborhood \(U\) of the origin of \(\C^{k}\). \begin{equation} \varphi : \C^{k} \longrightarrow D,\;\;\;\; (x_1,\ldots,x_k)\longmapsto \Phi (x_1\cdot e_1,\ldots,x_k\cdot e_k) \end{equation} where the map \(\Phi\) is the Lie algebra coordinate system given in Definition~\ref{algebracoord}. Note that the map \(\varphi\) is well defined and injective in a neighborhood of the origin and its tangent space at the origin is \(E\). What remains to be shown is that \(\varphi\) is horizontal. By Proposition~\ref{maurer} it needs to be checked that the entries \(\Omega_{i,j}\;\;\rm{for}\;\; i\geq j+2\) of the Maurer-Cartan matrix vanish along the image of \(\varphi\). Now \[\Omega |\varphi (x_1,\ldots,x_n) = exp(-\sum x_i\cdot e_i)\;d\,exp(\sum x_i\cdot e_i)=\] \[=\prod exp(-x_i\cdot e_i)\;d\,\prod exp(x_i\cdot e_i)=e_1dx_1+\ldots +e_kdx_k\] where the second equality holds because of Equation~\ref{bracketcomm}. Since the blocks \((e_l)_{i,j} = 0\) for \(i\geq j+2\) the proposition follows. \end{proof} \section{Coupled Contact Systems and Dimension Bounds} In the previous section we gave a local description of the horizontal differential system and we saw that it is given by the 1-form entries in those blocks of the Maurer-Cartan matrix that are at least two steps below the diagonal (Equation~\ref{maurerequation}). Now we will see that this system can be replaced locally by another equivalent system which is easier to deal with. A closer examination of this system will lead to an upper bound on the dimension of integral manifolds. \subsection{Coupled Contact Differential Systems} It will be more convenient to work in the coordinate system provided by the Lie group coordinates \(Y\in G^-\) of Section~\ref{groupcord}. In this coordinate system the identity matrix corresponds to the reference Hodge structure \(H\). Recall that the matrix \(Y\) has a block decomposition \(\sum Y_{i,j}[i,j]\) where the block \(Y_{i,j}\) at position \((i,j)\) has size \(h^{w-i,i}\times h^{w-j,j}\). Note that the matrices \(Y_{i,j}\) must satisfy the Lie group coordinate version of Equation~\ref{ortcond}, namely \begin{equation}\label{grouportcond} (log(Y))^tQ+Qlog(Y)=0 \end{equation} As a first step let us write the generators of the differential ideal of the horizontal differential system in terms of the entries of the blocks \(Y_{i,j}\). \begin{prop}\label{firstreduction} The differential ideal of the horizontal system is generated algebraically by the 1-form entries of the matrices \begin{equation}\label{almostcontact} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;i\geq j+2 \end{equation} \end{prop} \begin{proof} By Proposition~\ref{maurer} we have to show that the entries of the matrices \(\Omega_{i,j}\) for \(i\geq j+2\) can be generated by the 1-forms in Equation~\ref{almostcontact}. We will proceed by induction on \(i-j\). If \(i-j=2\) then an easy computation shows that \begin{equation}\label{contactfirst} \Omega_{i,j}=(Y^{-1}dY)_{i,j}=dY_{i,j}-Y_{i,j+1}dY_{j+1,j} \end{equation} so the result holds in this case. Let us now assume that the entries of \(\Omega_{i,j}\) are generated by the forms in Equation~\ref{almostcontact} for \(i-j<k\) and consider \(\Omega_{a,b}\) with \(a-b=k\). Let us compute the \((a,b)\) blocks of both sides of the the defining equation of the Maurer-Cartan matrix \[dY=Y\cdot\Omega .\] We get \[dY_{a,b}=\sum_{l=0}^{a}Y_{a,l}\Omega_{l,b}=Y_{a,b+1}dY_{b+1,b}+\sum_{l=b+2}^{a-1}Y_{a,l}\Omega_{l,b}+\Omega_{a,b}.\] So \[\Omega_{a,b}=dY_{a,b}-Y_{a,b+1}dY_{b+1,b}-\sum_{l=b+2}^{a-1}Y_{a,l}\Omega_{l,b}\] and since in the sum we only have \(\Omega_{l,b}\) with \(l-b\leq a-1-b=k-1\) the right hand side of the above equation is generated by forms in Equation~\ref{almostcontact} by the induction hypothesis and this implies the result. \end{proof} As the next step let us define a new differential system on a submanifold of the period domain in a neighborhood of the reference Hodge structure. Let \(W\subset D\) denote the submanifold defined by the equations \begin{equation} Y_{i,j}=0 \;\;\;\;\rm{for}\;\;\;\;i>j+2. \end{equation} \begin{defi}\label{contactdef} The coupled contact system on the submanifold \(W\) is given by the differential 1-form entries of the matrices \begin{equation}\label{contactsystem} (Y^{-1}d\,Y)_{i,j}=dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;i=j+2. \end{equation} \end{defi} \begin{rem} Let us emphasize that the blocks \(Y_{i,j}\) in the previous definition are of course subject to the condition specified by Equation~\ref{grouportcond}. \end{rem} \begin{rem}\label{hhell} Note that the matrices in the definition are exactly the blocks \(\Omega_{i,j}\) for \(i=j+2\) of the Maurer-Cartan matrix written in terms of the Lie group coordinates, i.e., the blocks which are two steps below the diagonal. In other words we defined a differential systems which is the projection of the horizontal system to the submanifold \(W\) (cf. Equation~\ref{almostcontact}). What is remarkable is that this system is locally equivalent to the horizontal system as we will see shortly. \end{rem} \begin{rem} Each of the equations in Definition~\ref{contactdef} is a matrix valued contact system. In fact, in the weight two case when \(h^{2,0}=2\) this system is the classical contact system. Note furthermore that each matrix \(Y_{i,j}\) appears in two consecutive matrix valued contact systems and so these systems are coupled through these matrices, hence the name in the definition. \end{rem} \begin{thm}\label{contactequiv} There is a one--to--one dimension preserving correspondence between germs of integral manifolds of the horizontal differential system on \(D\) and germs of integral manifolds of the coupled contact system on \(W\). The correspondence also identifies integral elements of the two systems. \end{thm} \begin{rem} The theorem implies that instead of studying the uniqueness and dimension properties of germs of variations of Hodge structure we can study the corresponding properties of integral manifolds of the coupled contact system. The results we arrive at will remain true for variations of Hodge structure. \end{rem} \begin{proof} By Proposition~\ref{firstreduction} the horizontal system is given by the differential form entries of the matrices \begin{equation}\label{variation} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\; i\geq j+2 \end{equation} and by Definition~\ref{contactdef} the coupled contact system is given by \begin{equation}\label{contact} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\; i=j+2. \end{equation} Let \(S\subset D\) be a germ of an integral manifold of the horizontal system. Then because of the above equations the projection \(pr_{W}(S)\) to the submanifold \(W\) defines an integral manifold of the contact system. To prove the theorem we have to show that we can go the other way, namely given an integral manifold \(T\subset W\) of the contact system we need to find a {\it locally unique} extension of \(T\) to an integral manifold \(\bar{T}\subset D\) of the horizontal system. This amounts to showing that the differential equations \begin{equation}\label{needtosolve} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}=0\;\;\;\;\rm{for}\;\;\;\; i>j+2 \end{equation} have unique local solutions, provided that we have a solution of \begin{equation}\label{given} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}=0\;\;\;\;\rm{for}\;\;\;\; i=j+2. \end{equation} Since \(T\) is an integral manifold we also have the equations \begin{equation}\label{givenwedge} dY_{i,j+1}\wedge dY_{j+1,j}=0\;\;\;\;\rm{for}\;\;\;\; i=j+2. \end{equation} Let us proceed by induction on \(i-j\). Consider the equation \begin{equation}\label{solve1} dY_{a,b}=Y_{a,b+1}dY_{b+1,b}\;\;\;\;\rm{where}\;\;\;\; a-b=3. \end{equation} Note that the right hand side of Equation~\ref{solve1} consists of matrices which are already determined at this point since they appear in Equation~\ref{given}. This implies that if we can solve \ref{solve1} for \(Y_{a,b}\) then the solution will be unique. By the Poincar\'e Lemma there is a local solution of \ref{solve1} if and only if the 1-form entries of the right hand side are {\it closed} 1-forms, i.e., it needs to be checked that \begin{equation} dY_{a,b+1}\wedge dY_{b+1,b}=0\;\;\;\;\rm{for}\;\;\;\; a-b=3 \end{equation} holds. Now, since \(a-(b+1)=2\) we have \(dY_{a,b+1}=Y_{a,b+2}dY_{b+2,b+1}\) by Equation~\ref{given} so \[ dY_{a,b+1}\wedge dY_{b+1,b}=(Y_{a,b+2}dY_{b+2,b+1})\wedge dY_{b+1,b}=Y_{a,b+2}(dY_{b+2,b+1}\wedge dY_{b+1,b})=0 \] where the last equality holds by Equation~\ref{givenwedge}. This completes the first step of the induction. Assume now that the equations \begin{equation}\label{givengeneral} dY_{i,j}=Y_{i,j+1}dY_{j+1,j} \end{equation} are solvable for \(i-j<k\). This also implies that \begin{equation}\label{wedgegeneralgiven} dY_{i,j+1}\wedge dY_{j+1,j}=0\;\;\;\;\rm{for}\;\;\;\;i-j<k. \end{equation} By the same argument as above we see that to solve \begin{equation}\label{solve2} dY_{a,b}=Y_{a,b+1}dY_{b+1,b}\;\;\;\;\rm{where}\;\;\;\; a-b=k \end{equation} we need that \begin{equation}\label{wedgegeneral} dY_{a,b+1}\wedge dY_{b+1,b}=0\;\;\;\;\rm{for}\;\;\;\; a-b=k. \end{equation} Note again that the matrices on the right hand side of \ref{solve2} have already appeared in Equation~\ref{givengeneral} since \(a-(b+1)=k-1<k\) so uniqueness holds. To check \ref{wedgegeneral}, note that \(dY_{a,b+1}=Y_{a,b+2}dY_{b+2,b+1}\) by \ref{givengeneral} since \(a-(b+1)<k\) so \[ dY_{a,b+1}\wedge dY_{b+1,b}=Y_{a,b+2}(dY_{b+2,b+1}\wedge dY_{b+1,b})=0 \] by Equation~\ref{wedgegeneralgiven}. \end{proof} \subsection{Integral Elements and Abelian Lie Algebras}\label{proofcom} In this section we will prove Proposition~\ref{commutative}. Let us recall the statement. \begin{prop} Let \(E\subset \mathfrak{g}^{-1,1}\) be a linear subspace. Then \(E\) is an integral element of the horizontal differential system if and only if it is an abelian subalgebra. \end{prop} \begin{proof} By definition \(E\) is an integral element if and only if \(\omega_E = 0\) for each \(\omega\) in the differential ideal. By Theorem~\ref{contactequiv} this is equivalent to the vanishing of the two forms \(\varphi_{i,j}=dY_{i,j+1}\wedge dY_{j+1,j}\) on \(E\) for \(i=j+2\). Now let \(X^1,X^2\in E\) be two general tangent vectors. Then \begin{equation} 0=\varphi_{i,j}(X^1,X^2)=dY_{i,j+1}\wedge dY_{j+1,j}(X^1,X^2)=X^1_{i,j+1}X^2_{j+1,j}-X^2_{i,j+1}X^1_{j+1,j} \end{equation} which is exactly the condition for the commutator \([X^1,X^2]\) to vanish since the matrices \(X^1,X^2\) have nonzero blocks only at the \(i=j+1\) positions. \end{proof} \begin{rem} In the language of Hodge theory integral elements correspond to infinitesimal variations of Hodge structure and the above commutativity condition is part of the definition of an infinitesimal variation of Hodge structure (see \cite{C1}). In this context Proposition~\ref{exptrick} says that to every infinitesimal variation of Hodge structure there is a germ of a variation of Hodge structure tangent to it. Apparently this simple result was not well known. \end{rem} \subsection{Dimension Bound for Variations of Hodge Structure}\label{dimbound} In this section we will use the contact differential system to give sharp upper bounds for the dimension of variations of Hodge structure. This is the main result of \cite{C-K-T}. The proof is different, however, and less elementary than the one in this section. Also a case is not covered in \cite{C-K-T} so it seems worthwhile to reprove this result here. Up to this point we did not need to use the conditions that Equation~\ref{grouportcond} specifies in terms of the blocks \(Y_{i,j}\) but to give the upper bound we will have to compute these at least for the contact system. By doing this it is possible to further reduce the contact system to an equivalent differential system. \begin{prop}\label{realreduced} a) If the weight \(w=2k+1\) is odd then there is a one-to-one correspondence between germs of integral manifolds of the coupled contact differential system and the system given by the 1-form entries of the matrices \begin{equation}\label{reducedcontactodd} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;i=j+2\;\;\;\;\rm{with}\;\;\;\;j<k. \end{equation} where the matrix \(Y_{k+1,k}\) is symmetric. b) If \(w=2k\) is even then the above correspondence is between the coupled contact system and the system given by the entries of \begin{equation}\label{reducedcontacteven} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;i=j+2\;\;\;\;\rm{with}\;\;\;\;j<k-1 \;\;\;\;\rm{and} \end{equation} \begin{equation}\label{weight2add} dY_{k+1,k-1}-Y_{k,k-1}^tdY_{k,k-1} \end{equation} and we also have \begin{equation}\label{weight2cond} Y_{k+1,k-1}+Y_{k+1,k-1}^t=Y_{k,k-1}^tY_{k,k-1}. \end{equation} \end{prop} \begin{rem} The content of this proposition is that we need to consider only blocks of the decomposition that are above the main antidiagonal and we have an explicit description of the dependencies between the entries of the matrices (e.g., the condition that \(Y_{k+1,k}\) is symmetric when the weight is odd). To make this easier to visualize let us consider the \(w=3\) and \(w=4\) cases. \end{rem} \begin{ex}\label{visex}(\(w=3\)) In this case a general element \(Y\in G^-\) has the form \[Y= \begin{pmatrix} I & 0 & 0 & 0\\ Y_{1,0} & I & 0 & 0\\ Y_{2,0}& Y_{2,1} & I &0\\ Y_{3,0}& Y_{3,1} & Y_{3,2} &I \end{pmatrix} \] where the blocks satisfy Equation~\ref{grouportcond}. Now the proposition says that the contact differential system is determined by the 1-forms in \[ dY_{2,0}-Y_{2,1}dY_{1,0}\] where \(Y_{2,1}\) is symmetric and this is the only dependence between the entries of the matrices giving local coordinates. So to exhibit local variations of Hodge structure in the weight three case it is enough to solve the differential equations in \(dY_{2,0}=Y_{2,1}dY_{1,0}\). \end{ex} \begin{ex}(\(w=4\)) Now a general element \(Y\in G^-\) has the form \[Y= \begin{pmatrix} I & 0 & 0 & 0 & 0\\ Y_{1,0} & I & 0 & 0 & 0\\ Y_{2,0}& Y_{2,1} & I &0 & 0\\ Y_{3,0}& Y_{3,1} & Y_{3,2} &I & 0\\ Y_{4,0}& Y_{4,1} & Y_{4,2} & Y_{4,3}& I \end{pmatrix} \] where the blocks satisfy Equation~\ref{grouportcond}. By the proposition this means that the contact system is determined by \[ dY_{2,0}-Y_{2,1}dY_{1,0}\;\;\;\;\rm{and}\;\;\;\;dY_{3,1}-Y_{2,1}^tdY_{2,1}\] where there is one more restriction of the form \[ Y_{3,1}^t + Y_{3,1}=Y_{2,1}^tY_{2,1}\] \end{ex} \begin{proof} Since the contact system involves blocks that are one or two steps below the main diagonal, to compute these blocks of \(log(Y)\) only the first two terms of the power series expansion will have to be considered: \((Y-I)-(Y-I)^2/2\). If we consider the blocks that are one step below the diagonal it is immediate from Equation~\ref{ortcond} that \begin{equation}\label{symmetry} Y_{i,i-1}^t=Y_{w-i+1,w-i}\;\;\;\;i\in \{0,\ldots ,w\}. \end{equation} If \(w\) is odd this implies that \(Y_{k+1,k}\) is symmetric. Now we have to show that an equation consisting of blocks below the antidiagonal is automatically solved if we have solutions for the equations above the antidiagonal. This is a simple consequence of Equation~\ref{ortcond}. In detail, let us consider the equation \begin{equation} dY_{j+2,j}=Y_{j+2,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;j>k\;\;\;\;w=2k+1\;\;\;\;\rm{or}\;\;\;\;j>k-1\;\;\;\;w=2k. \end{equation} Computing \(log(Y)_{j+2,j}\) and substituting it to Equation~\ref{ortcond} we get \begin{equation}\label{reducedcontact} Y_{j+2,j} = Y_{j+2,j+1}Y_{j+1,j}-Y_{w-j,w-j-2}^t \end{equation} from which \begin{equation} dY_{j+2,j}=Y_{j+2,j+1}dY_{j+1,j}+dY_{j+2,j+1}Y_{j+1,j}-Y_{w-j,w-j-2}^t=Y_{j+2,j+1}dY_{j+1,j} \end{equation} by Equation~\ref{symmetry} and Equation~\ref{reducedcontactodd} (or Equation~\ref{reducedcontacteven}). \end{proof} Using this reduced form of the contact system we will proceed to give upper bounds for the dimension of variations of Hodge structure. Let us define the following quadratic functions that depend on the Hodge numbers \(h^{i,j}\). Let \(h^l\) stand for \(h^{l,w-l}\). If the weight is odd \((w=2k+1>1)\) let \begin{eqnarray*} q^{odd}_1& = &\sum_{i=0}^{\infty}h^{k+2+2i}h^{k+3+2i}\cr q^{odd}_2& = &\frac{1}{2}h^{k+1}(h^{k+1}+1)+\sum_{i=0}^{\infty}h^{k+3+2i}h^{k+4+2i}. \end{eqnarray*} If the weight is even \((w=2k)\) let \begin{eqnarray*} q^{even}_1& =&\sum_{i=0}^{\infty}h^{k+1+2i}h^{k+2+2i}\cr q^{even}_2& = &\overline{q_2}+\sum_{i=0}^{\infty}h^{k+2+2i}h^{k+3+2i}\cr q^{even}_3& = &h^k+h^{k+2}(h^{k+1}-1)+\sum_{i=0}^{\infty}h^{k+3+2i}h^{k+4+2i}\;\;\;\;\rm{if}\;\;\;\;w\geq 4. \end{eqnarray*} where \begin{equation} \overline{q_2}= \left\{\begin{array}{ll} h^k & \mbox{if \(h^{k+1}=1\),}\cr \frac{1}{2}h^{k+1}h^k& \mbox{if \(h^k\) is even and \(h^{k+1}>1\),}\cr \frac{1}{2}h^{k+1}(h^k-1)+1& \mbox{if \(h^k\) is odd and \(h^{k+1}>1\).} \end{array} \right. \end{equation} \begin{rem} The sums are of course finite since \(h^i=0\) if \(i>w\). \(q^{even}_3\) is defined only if the weight is at least four. \end{rem} \begin{thm}\label{dimboundthm} Let \(S\subset D\) be a variation of Hodge structure. \begin{enumerate} \item If \(w=2k+1\) then \(dim(S)\leq max\{q^{odd}_1,q^{odd}_2\}\). \item If \(w=2k\) then \(dim(S)\leq max\{q^{even}_1,q^{even}_2,q^{even}_3\}\). \end{enumerate} \end{thm} \begin{proof}1) According to Proposition~\ref{realreduced} the contact system is equivalent to the system defined by \begin{equation}\label{dimproofcontact} dY_{i,j}-Y_{i,j+1}dY_{j+1,j}\;\;\;\;\rm{for}\;\;\;\;i=j+2\;\;\;\;\rm{with}\;\;\;\;j<k. \end{equation} with \(Y_{k+1,k}\) symmetric. We will analyze this system to derive the dimension bound. In a neighborhood of the reference Hodge structure \(S\) is defined by holomorphic functions depending on a set \(L\) of independent coordinate variables. Clearly \(dim(S)\leq |L|\) where \(|L|\) denotes the cardinality of the set \(L\). We will give a bound for \(|L|\). At the origin the system is defined by the vanishing of the differential form entries of \(dY_{j+2,j}\) \((j<k)\) so the coordinate entries of the matrices \(Y_{j+2,j}\) are not elements of \(L\), i.e., all elements of \(L\) appear in the matrices \(dY_{j+1,j}\). Let \(y_{1,0}\) denote the maximal number of the coordinates in the set \(L\) that appear in a column of the matrix \(Y_{1,0}\). Then, clearly, \(Y_{1,0}\) contains at most \(y_{1,0}\cdot h^{w}\) elements of \(L\) since \(Y_{1,0}\) has \(h^{w}\) columns. Notice that a single entry of the matrix valued differential form \(dY_{2,0}-Y_{2,1}dY_{1,0}\) defines a classical contact system so we can use Proposition~\ref{contactmax} and Remark~\ref{fixj} to partition the variables of this classical contact system into two disjoint subsets. More precisely, suppose that the \(l^{th}\) column in the matrix \(Y_{1,0}\) contains \(y_{1,0}\) elements of \(L\), i.e., it is one of the columns containing the maximal possible number of elements of \(L\). Consider the classical contact system defined by the \((i,l)\) entry of the system \(dY_{2,0}-Y_{2,1}dY_{1,0}\) for every \(i\in \{0,\ldots,h^{w-2}\}\): \begin{equation}\label{classcont} (dY_{2,0})^{i,l}-\sum_{t}(Y_{2,1})^{i,t}(dY_{1,0})^{t,l} \end{equation} where superscripts denote matrix entries. By Proposition~\ref{contactmax} the \(i^{th}\) row of the matrix \(Y_{2,1}\) can contain at most \(h^{w-1}-y_{1,0}\) elements of \(L\). This is because the classical contact system (\ref{classcont}) already contains \(y_{1,0}\) independent variables coming from the \(l^{th}\) column of \(Y_{1,0}\), (the set \(J\) in Proposition~\ref{contactmax}), so since \(I\) and \(J\) are disjoint we get \(|I|\leq h^{w-1}-y_{1,0}\). This implies that there are at most \(h^{w-1}-y_{1,0}\) columns of the matrix \(Y_{2,1}\) that contain elements of \(L\) so if \(y_{2,1}\) denotes the maximal number of the coordinates in the set \(L\) that appear in a column of the matrix \(Y_{2,1}\) then this matrix can contain at most \(y_{2,1}\cdot (h^{w-1}-y_{1,0}) \) elements of \(L\). Applying the same argument to all of the matrix valued contact systems in Equation~\ref{dimproofcontact} and taking into consideration that \(Y_{k+1,k}\) is symmetric, we arrive at the following upper bound for \(|L|\): \begin{eqnarray*}\label{boundd} |L|& \leq & y_{1,0}\cdot h^{w}+y_{2,1}\cdot (h^{w-1}-y_{1,0})+\ldots+\\ & + & y_{k,k-1}\cdot (h^{w-(k-1)}-y_{k-1,k-2})+\frac{1}{2}(h^{w-k}-y_{k,k-1})(h^{w-k}-y_{k,k-1}+1) \end{eqnarray*} The next step is to determine the maximum value of the right hand side of the above equation for \(y_{j+1,j}\in [1,h^{w-(j+1)}]\) (\(j<k)\). This is a quadratic programming problem in a given rectangle that can be solved as follows. Let \(f(y_{i,j})\) denote the right hand side of this equation. Then the function \(f(y_{i,j})\) can not have interior maximum because its Hessian matrix is not negative definite and it is nowhere negative semidefinite. If we restrict \(f\) to a face than we get a function which is linear or has a non-negative Hessian. Applying this repeatedly we can see that the maximum must occur at a vertex and examining the function \(f(vertex)\) we conclude the theorem in the odd weight case. 2) First let us consider the \(w=2\) case. By Proposition~\ref{realreduced} the contact system is defined by the equation \begin{equation}\label{w2} dY_{2,0}=Y_{1,0}^tdY_{1,0} \end{equation} which is subject to the condition \begin{equation}\label{weight2cond2} Y_{2,0}+Y_{2,0}^t=Y_{1,0}^tY_{1,0}. \end{equation} \(Y_{1,0}\) determines the symmetric part of \(Y_{2,0}\) by Equation~\ref{weight2cond2} so it is enough to consider the 1-form entries \((i,j)\) of Equation~\ref{w2} for which \(i<j\). The rest of the equations are automatically satisfied as it can be seen by taking the exterior derivative of Equation~\ref{weight2cond2}. Let the set \(L\) be as in the first part of the proof and let \(y_{1,0}\) denote the maximal number of the coordinates in the set \(L\) that appear in a column of the matrix \(Y_{1,0}\). (Just like above we can assume that the entries of the matrix \(Y_{2,0}\) do not appear in \(L\)). Then we have the following bounds for \(L\): \begin{equation} |L|\leq y_{1,0}\cdot h^{2,0} \end{equation} since the matrix \(Y_{1,0}\) has \(h^{2,0}\) columns. We also have \begin{equation} |L|\leq y_{1,0}+(h^{1,1}-y_{1,0})\cdot (h^{2,0}-1) \end{equation} which follows by fixing a column having \(y_{1,0}\) coordinate entries in \(L\) and applying the argument of Proposition~\ref{contactmax} to each of the classical contact systems determined by this fixed column and the other columns of \(Y_{1,0}\). Then in the fixed column we have \(y_{1,0}\) independent variables and each of the other columns can have at most \(h^{1,1}-y_{1,0}\) of them. The above two bounds for \(L\) imply that the maximum occurs if \(y_{1,0}=\frac{1}{2}h^{1,1}\). If \(h^{1,1}\) is even then this implies that \(|L|\leq \frac{1}{2}h^{1,1}\cdot h^{2,0}\). If \(h^{1,1}\) is odd then an easy computation shows that \(\frac{1}{2}h^{2,0}(h^{1,1}-1)+1\) is the largest value satisfying both of the bounds. This proves the theorem in the weight two case. Assume now that \(w\geq 4\). The argument will be the same as in the odd weight case except that when considering the last of the matrix valued differential forms in Equation~\ref{reducedcontacteven} it will be necessary to use similar arguments as we did in the weight two case. Let us consider the last two equations: \begin{equation}\label{4one} dY_{k,k-2}=Y_{k,k-1}dY_{k-1,k-2} \end{equation} and \begin{equation}\label{4two} dY_{k+1,k-1}=Y_{k,k-1}^tdY_{k,k-1}. \end{equation} As in the weight two case we need to consider only the antisymmetric part of Equation~\ref{4two}. Using the same arguments as above we have the following bounds for the number of independent variables in the matrix \( Y_{k,k-1}\), (\(dim(Y_{k,k-1})\) will denote this number): \begin{equation} dim(Y_{k,k-1})\leq y_{k,k-1}\cdot (h^{k+1}-y_{k-1,k-2}). \end{equation} which follows from the fact that by Equation~\ref{4one} \( Y_{k,k-1}\) can have at most \(h^{k+1}-y_{k-1,k-2}\) columns that contain elements of \(L\) and by definition one column can contain at most \(y_{k,k-1}\) independent variables. We also have \begin{equation} dim(Y_{k,k-1})\leq y_{k,k-1} + (h^{k+1}-y_{k-1,k-2}-1)(h^{k}-y_{k,k-1}) \end{equation} from Equation~\ref{4two} by applying the argument we had in the weight two case. Now we have to distinguish two cases.\newline a) If \(h^{k+1}-y_{k-1,k-2}-1\neq 0\) then from the above two bounds we get that the maximum occurs at \(y_{k,k-1}=\frac{1}{2}h^{k}\) which means that \begin{equation} dim(Y_{k,k-1})\leq \frac{1}{2}h^{k}(h^{k+1}-y_{k-1,k-2})\;\;\;\;\rm{if}\;\;\;\;h^{k}\;\;\;\;\rm{even} \end{equation} \begin{equation} dim(Y_{k,k-1})\leq \frac{1}{2}(h^{k}-1)(h^{k+1}-y_{k-1,k-2})+1\;\;\;\;\rm{if}\;\;\;\;h^{k}\;\;\;\;\rm{odd}. \end{equation} Applying arguments as in the odd weight case to the remaining matrix valued contact systems we arrive at the following upper bound for \(L\): \begin{eqnarray*} |L| & \leq & y_{1,0}\cdot h^{w}+y_{2,1}\cdot (h^{w-1}-y_{1,0})+\ldots+\\ & + & y_{k-1,k-2}\cdot (h^{w-(k-2)}-y_{k-2,k-3})+dim(Y_{k,k-1}) \end{eqnarray*} Now the usual quadratic programming argument applies to this function to give the result in this case.\newline b) If \(h^{k+1}-y_{k-1,k-2}-1= 0\) then \begin{equation} dim(Y_{k,k-1})\leq h^{k} \end{equation} since there can only be one column in \(Y_{k,k-1}\) that contains elements of \(L\). In this case we get the following upper bound: \begin{eqnarray*} |L| & \leq & y_{1,0}\cdot h^{w}+y_{2,1}\cdot (h^{w-1}-y_{1,0})+\ldots+\\ & + & (h^{k+1}-1)\cdot (h^{w-(k-2)}-y_{k,k-1})+h^{k} \end{eqnarray*} and applying quadratic programming again we get the claim of the theorem. \end{proof} \begin{rem} Let \(\Phi : M \longrightarrow \Gamma~\backslash~ D\) be a global variation of Hodge structure. Then \(rank(\Phi)\) also satisfies the dimension bound given in the theorem. \end{rem} \begin{rem} For sharpness of this result see \cite{C-K-T}. Note that the case when the maximal dimension is \(q^{even}_3\) is missing from the main theorem in \cite{C-K-T}. It is easy to see that there are Hodge numbers for which \(q^{even}_3\) is indeed the maximal dimension so this function is needed in the proper form of the dimension bound. \end{rem} \section{Rigidity of Maximal Dimensional Variations of Hodge Structure} In this section we will examine maximal dimensional variations. We will proceed in two steps. First, we fix a maximal dimensional integral element of the horizontal system and investigate the nature of germs of integral manifolds tangent to this integral element. It turns out that in most cases there is a unique germ of an integral manifold tangent to the given integral element. This is sometimes referred to as local rigidity. In the second step we will fix a point and investigate the nature of integral elements through the given point hence obtaining infinitesimal rigidity. These two steps together lead to the main rigidity results. \subsection{Germs of Integral Manifolds} Let \(E\) be a maximal dimensional integral element of the horizontal differential system going through the point \(p\). By Theorem~\ref{exptrick} there is a maximal dimensional integral manifold tangent to \(E\) so the only question we have to investigate is uniqueness of such integral manifolds. The following result answers this question. \begin{thm}\label{localrigid} Let \(E\) be a maximal dimensional integral element of the horizontal differential system. \begin{enumerate} \item Let \(w=2k+1\). If \(h^{k,k+1}>2\) and all the other Hodge numbers are greater than one then there is a unique germ of an integral manifold whose tangent space at \(x\) is \(E\). \item Let \(w=2k\). Assume that one of the following conditions holds: \begin{enumerate} \item \(dim(E)=q^{even}_1\) and all the Hodge numbers are greater than one \item \(dim(E)=q^{even}_2\), \(h^{k+1}>2\), the other Hodge numbers are greater than one and \(h^k\geq 4\) is even, \item \(dim(E)=q^{even}_3\) and \(h^{k+1}>2\) \end{enumerate} then there is a unique germ of an integral manifold whose tangent space at \(x\) is \(E\). \item If \(dim(E)=q^{even}_2\) and \(h^k\) is odd then there is an infinite dimensional family of germs of maximal dimensional integral manifolds tangent to \(E\). \end{enumerate} \end{thm} \begin{proof} 1) Assume first that the maximum is \(q^{odd}_1\) i.e., \(dim(E)=q^{odd}_1\). Consider the first matrix valued differential form of Equation~\ref{reducedcontactodd} \begin{equation}\label{egy} dY_{k+1,k-1}=Y_{k+1,k}dY_{k,k-1} \end{equation} and its exterior derivative \begin{equation}\label{ketto} 0=dY_{k+1,k}\wedge dY_{k,k-1} \end{equation} From the proof of Theorem~\ref{dimboundthm} we see that in this case the entries of the matrix \(Y_{k,k-1}\) are independent variables (elements of the set \(L\)). We would like to see that the entries of the matrix \(Y_{k+1,k}\) are at most linear functions of the variables in \(L\) since this implies that they are uniquely determined by fixing the tangent space \(E\). To this end, consider an entry of Equation~\ref{ketto} (e.g., the \((1,1)\) entry): \begin{equation}\label{harom} 0=\sum_{i}(dY_{k+1,k})^{1,i}\wedge (dY_{k,k-1})^{i,1}. \end{equation} This equation involves the first row of \(dY_{k+1,k}\) and the first column of \(dY_{k,k-1}\). Since all the entries of \(dY_{k,k-1}\) are independent, Cartan's lemma (\cite{W} Exercise 2.16) implies that \begin{equation} (dY_{k+1,k})^{1,i}=\sum_jA_{i,j}(dY_{k,k-1})^{j,1}\;\;\;\;\rm{for}\;\;\;\;\rm{each}\;\;\;\;i \end{equation} for some functions \(A_{i,j}\). Since the matrix \(Y_{k,k-1}\) has at least two columns because of the condition on the Hodge numbers in the theorem, we can do the same for the (1,2) entry of Equation~\ref{ketto}. The same argument shows that \begin{equation} (dY_{k+1,k})^{1,i}=\sum_jB_{i,j}(dY_{k,k-1})^{j,2}\;\;\;\;\rm{for}\;\;\;\;\rm{each}\;\;\;\;i \end{equation} for some functions \(B_{i,j}\). Since the entries in the first and second column of the matrix \(Y_{k,k-1}\) are independent from each other this is possible only if all the functions \(A_{i,j}\) and \(B_{i,j}\) are zero. This implies that entries in the first row of the matrix \(Y_{k+1,k}\) must be constant functions. The same argument applies to all the rows of this matrix so we conclude that all the entries of the matrix \(Y_{k+1,k}\) are constant functions. Consider now the next matrix valued differential form of Equation~\ref{reducedcontactodd} \begin{equation}\label{egyy} dY_{k,k-2}=Y_{k,k-1}dY_{k-1,k-2}. \end{equation} Notice that this equation contains the matrix \(Y_{k,k-1}\) which has independent entries so applying the above argument we conclude that the entries of the matrix \(dY_{k-1,k-2}\) are constant functions. Continuing in the same manner we see that the matrices involved in Equation~\ref{reducedcontactodd} are either constant matrices or contain entries that are independent coordinate functions on the integral manifold. This implies the claim of the theorem showing that there is a unique germ of an integral manifold tangent to the fixed integral element \(E\). If the maximum is \(q^{odd}_2\) (i.e., \(dim(E)=q^{odd}_2\)) the above argument applies but we have to be careful about the equation containing the symmetric matrix \(Y_{k+1,k}\). Consider this equation \begin{equation}\label{hello} dY_{k+1,k-1}=Y_{k+1,k}dY_{k,k-1} \end{equation} and its exterior derivative \begin{equation} 0=dY_{k+1,k}\wedge dY_{k,k-1}. \end{equation} From Theorem~\ref{dimboundthm} we see that the entries of the symmetric matrix \(Y_{k+1,k}\) are independent variables and we would like to conclude that this implies that \(Y_{k,k-1}\) has constant entries. Consider the first column of \(Y_{k,k-1}\). As before from the \((1,1)\) entry of Equation~\ref{hello} we have \begin{equation} (dY_{k,k-1})^{i,1}=\sum_jA_{i,j}(dY_{k+1,k})^{1,j}\;\;\;\;\rm{for}\;\;\;\;\rm{each}\;\;\;\;i \end{equation} and from the \((2,1)\) entry we have \begin{equation} (dY_{k,k-1})^{i,1}=\sum_jB_{i,j}(dY_{k+1,k})^{2,j}\;\;\;\;\rm{for}\;\;\;\;\rm{each}\;\;\;\;i. \end{equation} Since \(Y_{k+1,k}\) is symmetric this implies only that the entries in the first column of \(Y_{k,k-1}\) are functions of the \((1,2)\) entry of \(Y_{k+1,k}\). On the other hand considering the \((1,1)\) and \((1,3)\) entries of Equation~\ref{hello} (\(h^{k,k+1}>2\)) we conclude that these same entries are functions of the \((1,3)\) entry of \(Y_{k+1,k}\). These two facts together imply that the entries must be constant functions. The same argument applies to all the remaining entries and to all of the remaining matrix valued contact equations and this implies the claim of the theorem in this case. 2) a) The difference between the odd weight and even weight cases is the appearance of the equation \begin{equation}\label{extraeq} dY_{k+1,k-1}=Y_{k,k-1}^tdY_{k,k-1}. \end{equation} However, if the maximum dimension occurs at \(dim(E)=q^{even}_1\) then there are no independent variables among the entries of matrix \(Y_{k,k-1}\); consequently the same argument applies as in the odd weight, \(q^{odd}_1\) case. b) Assume now that \(dim(E)=q^{even}_2\). According to Theorem~\ref{dimboundthm} this implies that each column of the matrix \(Y_{k,k-1}\) has exactly \(\frac{1}{2}h^k\) independent entries. Without loss of generality we can assume that the first \(\frac{1}{2}h^k\) entries in the first column (\((Y_{k,k-1})^{l,1}\) with \(l\leq \frac{1}{2}h^k\)) are independent. Consider the \((1,2)\) entry in Equation~\ref{extraeq}: \begin{equation} (dY_{k+1,k-1})^{1,2}=\sum_l(Y_{k,k-1})^{l,1}(dY_{k,k-1})^{l,2}. \end{equation} Applying Proposition~\ref{contactmax} to this classical contact system it follows that the independent entries in the second column are \begin{equation} (Y_{k,k-1})^{l,2}\;\;\;\;\rm{with}\;\;\;\;l>\frac{1}{2}h^k. \end{equation} Furthermore, by Cartan's lemma, the entries \((dY_{k,k-1})^{l,1}\) with \(l>\frac{1}{2}h^k\) can be expressed in terms of the independent 1-forms \((dY_{k,k-1})^{l,1}\) with \(l\leq \frac{1}{2}h^k\) and \((dY_{k,k-1})^{l,2}\) with \(l>\frac{1}{2}h^k\). Considering the \((1,3)\) entry of Equation~\ref{extraeq} we can similarly conclude that the same entries can be expressed in terms of \((dY_{k,k-1})^{l,1}\) with \(l\leq \frac{1}{2}h^k\) and \((dY_{k,k-1})^{l,3}\) with \(l>\frac{1}{2}h^k\). This implies that these entries depend only on \((dY_{k,k-1})^{l,1}\) with \(l\leq \frac{1}{2}h^k\). Similarly, the entries \((dY_{k,k-1})^{l,2}\) with \(l\leq \frac{1}{2}h^k\) depend only on \((dY_{k,k-1})^{l,2}\) with \(l>\frac{1}{2}h^k\). It follows from this and the \((1,2)\) entry of the exterior derivative of Equation~\ref{extraeq} \begin{equation} 0=\sum_l(dY_{k,k-1})^{l,1}\wedge (dY_{k,k-1})^{l,2} \end{equation} that the entries \((Y_{k,k-1})^{l,1}\) with \(l>\frac{1}{2}h^k\) can be at most linear functions which is what we wanted to prove. Consider now the exterior derivative of the next matrix valued contact system in Equation~\ref{reducedcontacteven} \begin{equation} 0=dY_{k,k-1}\wedge dY_{k-1,k-2} \end{equation} By applying changes of coordinates we can assume that the independent variable entries in matrix \(Y_{k,k-1}\) are the first \(\frac{1}{2}h^k\) entries in each column. This implies that at least the first two rows of matrix \(Y_{k,k-1}\) consist of independent variables, which in turn implies that the entries of the matrix \(Y_{k-1,k-2}\) must be constant functions by applying the usual Cartan's lemma type argument. All the remaining matrix valued contact systems in Equation~\ref{reducedcontacteven} can be treated exactly the same way as in the odd weight case so we conclude that the entries in these matrices which are not independent variables must be constants. Together with the above results this concludes the proof for this case. c) In this case \(w\geq 4\) and \(dim(E)=q^{even}_3\). By Theorem~\ref{dimboundthm} the matrix \(Y_{k,k-1}\) has exactly one column that consists entirely of independent variables and the other columns do not contain independent variables at all. We can assume that the entries of the first column are independent. Considering the entries of Equation~\ref{weight2add} it follows that the remaining entries of \(Y_{k,k-1}\) can be expressed as functions of the variables in the first column. In fact, from the \((1,2)\) entry of the exterior derivative of Equation~\ref{weight2add} \begin{equation} 0=\sum_l(dY_{k,k-1})^{l,1}\wedge (dY_{k,k-1})^{l,2} \end{equation} it follows that the entries \((Y_{k,k-1})^{l,2}\) depend only on the independent variables \((Y_{k,k-1})^{l,1}\). Similarly, the same holds for the other columns. In this case the matrix \(Y_{k-1,k-2}\) consists of independent variables except for its first row (cf. proof of Theorem~\ref{dimboundthm}). From the \((1,1)\) entry of the exterior derivative of Equation~\ref{4one} \begin{equation} 0=\sum_l(dY_{k,k-1})^{1,l}\wedge (dY_{k-1,k-2})^{l,1} \end{equation} we can conclude that the entries \((Y_{k,k-1})^{1,l}\) for \(l\geq 2\) depend only on \((Y_{k,k-1})^{1,1}\) and in fact, they must be linear functions in this variable. Similarly, the entry \((Y_{k-1,k-2})^{1,1}\) must be a linear function of the variables \((Y_{k-1,k-2})^{1,l}\) \((l\geq 2)\). These claims easily follow from the fact that the differential 1-forms \((dY_{k,k-1})^{1,1}\) and \((dY_{k-1,k-2})^{1,l}\), \((l\geq 2)\) are linearly independent. Considering the other entries we conclude that the entries that are not independent variables must be linear functions. In fact, let us remark here that the coefficients of the linear functions in a given row of the matrix \(Y_{k,k-1}\) must be the same for each entry since for example in the second column each coefficient is equal to \(\frac{\partial((Y_{k-1,k-2})^{1,1})}{\partial((Y_{k-1,k-2})^{2,1})}\). It also follows that the coefficients of the linear functions \((Y_{k-1,k-2})^{1,i}\) are the same for each \(i\). The remaining matrix valued contact systems in Equation~\ref{reducedcontacteven} can now be treated the same way as in the odd weight case which concludes the proof of the theorem in this case. 3) What remains to be considered to complete the proof of the theorem is the case when \(h^k\) is odd. We will exhibit an infinite dimensional family of germs of maximal dimensional integral manifolds tangent to an integral element \(E\). To this end let us specify the entries of the matrices in Equation~\ref{reducedcontacteven}. To give a maximal dimensional integral manifold we must specify \(\frac{1}{2}h^{k+1}(h^k-1)+1\) independent variables among the entries of \(Y^{k-1,k}\), according to Theorem~\ref{dimboundthm}. Let us choose the first \(\frac{1}{2}(h^k-1)\) variables in each column of \(Y^{k-1,k}\) to be independent and let the next \(\frac{1}{2}(h^k-1)\) entries in each column to be \(\sqrt{-1}\) times the first \(\frac{1}{2}(h^k-1)\) independent entries. This leaves the last row to be considered. Let the first element of the last row be an independent variable and let the remaining entries of the last row be arbitrary holomorphic functions of of the first entry. For example, if \(h^k=5\) and \(h^{k+1}=4 \) then \[ Y_{k-1,k}= \begin{pmatrix} y_{1,1}& y_{1,2} & y_{1,3} & y_{1,4}\\ y_{2,1}& y_{2,2} & y_{2,3} & y_{2,4}\\ i\cdot y_{1,1}& i\cdot y_{1,2} & i\cdot y_{1,3} & i\cdot y_{1,4}\\ i\cdot y_{2,1}& i\cdot y_{2,2} & i\cdot y_{2,3} & i\cdot y_{2,4}\\ y_{5,1}& f_{5,2} & f_{5,3} & f_{5,4} \end{pmatrix} \] where the functions \(f_{a,b}\) are arbitrary holomorphic functions of the single variable \(y_{5,1}\) and \(i=\sqrt{-1}\). Furthermore, let the matrices \(Y_{k-3-2j,k-2-2j}\) (for each \(j\)) consist entirely of independent variables and let the remaining matrices \(Y_{k-2-2j,k-1-2j}\) (for each \(j\)) be zero. This means that we have \(q^{even}_2\) independent variables. It is an easy computation to verify that all the equations in (\ref{reducedcontacteven}) are satisfied so we have defined a maximal dimensional integral manifold. Fixing an integral element to which this integral manifold has to be tangent to can determine only the linear parts of the functions \(f_{a,b}\) which implies that by varying these functions we can exhibit an infinite dimensional family of germs of integral manifolds tangent to a fixed integral element. \end{proof} \subsection{Maximal Dimensional Integral Elements} In this section we examine how maximal dimensional integral elements are related to each other. First let us recall a theorem of Carlson (\cite{C1}) that will be used in this section. \begin{thm}[Carlson] Let \(D\) be the period domain of weight two Hodge structures with \(h^{2,0}>2\) and \(h^{1,1}\) even. Let \(E_1\) and \(E_2\) be maximal dimensional integral elements of the horizontal distribution. Then there is an element \(g\in Aut(D)\) such that \(g\cdot E_1=E_2\). \end{thm} We will consider the same question for higher weight, namely what can be said about the relationship between maximal dimensional integral elements of the horizontal differential system. \begin{thm}\label{elementrigid} Let \(D\) denote the period domain. \begin{enumerate} \item Assume that one of the following holds: \begin{enumerate} \item \(w=2k+1\), \(h^{k,k+1}>2\) and all the other Hodge numbers are greater than one \item \(w=2k\), all the Hodge numbers are greater than one and the maximum dimension is \(q^{even}_1\) \end{enumerate} then there is a unique maximal dimensional integral element of the horizontal system through each point of \(D\). \item Let \(E_1\), \(E_2\) be two maximal dimensional integral elements. Assume that one of the following holds: \begin{enumerate} \item \(w=2k\), \(dim(E_1)=dim(E_2)=q^{even}_2\), \(h^{k+1}>2\), the other Hodge numbers are greater than one and \(h^k\geq 4\) is even \item \(w=2k\), \(dim(E_1)=dim(E_2)=q^{even}_3\) and \(h^{k+1}>2\) \end{enumerate} then there is an element \(g\in Aut(D)\) such that \(g\cdot E_1=E_2\). \end{enumerate} \end{thm} \begin{proof} 1) a) and b) The horizontal system is homogeneous so we can always assume that the integral elements are at the reference Hodge structure. In Theorem~\ref{localrigid} we proved that in cases 1.a and 1.b the matrix entries are either independent variables or constants, and these constants are determined be the point of the period domain. If we are at the reference Hodge structure then all these constants are zero. To compute the tangent space to the integral manifold we have to take partial derivatives of the matrix entries by the independent variables. From this it is clear that there can be only one maximal dimensional integral element through each point of \(D\). 2) a) In this case it follows from the proof of Theorem~\ref{localrigid} that the matrices besides \(Y_{k,k-1}\) consist of independent variables or constant functions, so the subspace of the tangent space coming from these matrices is unique. Now, it is clear that the equations \(Y_{k+1,k}\) has to satisfy are exactly the weight two equations so Carlson's theorem applies to conclude the theorem in this case. 2) b) Again, from the proof of Theorem~\ref{localrigid} we see that the subspace of the tangent space coming from the matrices besides \(Y_{k,k-1}\) and \(Y_{k-1,k-2}\) is unique, so we need to consider these two matrices. About these matrices we proved that they contain linear entries such that the coefficients of the columns are the same. This implies that we can conjugate the possible tangent spaces into each other by multiplying by elements of \(Aut(D)\). \end{proof} \begin{rem} It is true that result 2) of the above theorem holds even if \(h^{k}\) is odd. However, the author does not know a simple proof of this along the lines of these other results. Since there is an infinite dimensional family of integral manifolds to a given integral element in this case, we could not use this result to conclude further rigidity properties; hence it is not proved here. \end{rem} \begin{rem} Theorem~\ref{localrigid} and Theorem~\ref{elementrigid} immediately imply our main result Theorem~\ref{mainthm}. Note that if the Hodge numbers are smaller than what is required for these results, then we are reduced to the classical contact case and so flexibility holds by Theorem~\ref{arnold}. \end{rem} \begin{rem} It is not known whether the maximal dimensional variations are geometric if the weight is bigger than two. In the weight two, \(h^{1,1}\) even case this is true by \cite{C-S}. It is an interesting question in the author's opinion whether the reduction of the horizontal system to the much smaller coupled contact system is merely a local possibility or there is some global geometric reason that would explain this result. \end{rem}
1997-12-03T00:45:43
9712
alg-geom/9712003
en
https://arxiv.org/abs/alg-geom/9712003
[ "alg-geom", "math.AG" ]
alg-geom/9712003
Janos Kollar
J\'anos Koll\'ar
Real Algebraic Surfaces
LATEX2e, 27 pages
null
null
null
null
These are the notes for my lectures at the Trento summer school held September 1997. The aim of the lectures is to provide an introduction to real algebraic surfaces using the minimal model program. This leads to a fairly complete understanding of real rational surfaces and to a complete topological classification of real Del Pezzo surfaces. Almost all the results are contained in the works of Comessatti and Silhol.
[ { "version": "v1", "created": "Tue, 2 Dec 1997 23:45:42 GMT" } ]
2007-05-23T00:00:00
[ [ "Kollár", "János", "" ] ]
alg-geom
\section{Minimal models of real algebraic surfaces} \begin{defn} Let $X$ be a variety over a field $k$. A {\it 1--cycle} on $X$ is a formal linear combination $C=\sum c_iC_i$, where the $C_i\subset X$ are irreducible, reduced and proper curves. A 1--cycle is called {\it effective} if $c_i\geq 0$ for every $i$. Two 1--cycles $C,C'$ are {\it numerically equivalent} if $(C\cdot D)=(C'\cdot D)$ for every Cartier divisor $D$ on $X$. 1--cycles with real coefficients modulo numerical equivalence form a vectorspace, denoted by $N_1(X)$. $N_1(X)$ is finite dimensional by the Theorem of the base of N\'eron--Severi (cf. \cite[p.447]{Hartshorne77}). Its dimension, denoted by $\rho(X)$, is called the {\it Picard number} of $X$. Effective 1--cycles generate a cone $NE(X)\subset N_1(X)$. Its closure in the Euclidean topology $\overline{NE}(X)\subset N_1(X)$ is called the {\it cone of curves} of $X$. If $K_X$ is Cartier (or at least some multiple of $K_X$ is Cartier) then set $$ \overline{NE}(X)_{K\geq 0}:=\{z\in \overline{NE}(X)\vert (z\cdot K_X)\geq 0\}. $$ Let $V\subset \r^n$ be a closed convex cone. For $v\in V$, a ray $\r^{\geq 0}v\subset V$ is called {\it extremal} if $u,u'\in V$, $u+u'\in \r^{\geq 0}v$ implies that $u,u'\in\r^{\geq 0}v$. Intuitively: $\r^{\geq 0}v$ is an edge of $V$. An extremal ray $\r^{\geq 0}z\subset \overline{NE}(X)$ is called {\it $K_X$-negative} if $(z\cdot K_X)<0$. This does not depend on the choice of $z$ in the ray. Let $R\subset \overline{NE}(X)$ be a ray. A {\it contraction} of $R$ is a morphism $f_R:X\to X'$ such that $(f_R)_*\o_X=\o_{X'}$ and a curve $C\subset X$ is mapped to a point iff $[C]\in R$. \end{defn} \begin{exrc}\label{er.noncontr.ex} Show that the contraction of a ray is unique (if it exists). Also, if $X'$ is projective then $R$ is an extremal ray. Find examples of extremal rays which can not be contracted. \end{exrc} \begin{exrc}\label{int.cone.exrc} Let $F$ be a smooth projective surface and $z\in \overline{NE}(F)$ a 1--cycle such that $(z^2)>0$. Then $z$ or $-z$ is in the interior of $\overline{NE}(F)$. Thus if $\r^{\geq 0}z$ is extremal and $(z^2)>0$ then $\overline{NE}(F)$ is 1--dimensional. \end{exrc} We use the following description of the cone of curves of smooth surfaces over $\c$. The result is essentially equivalent to the theory of minimal models of surfaces developed around the turn of the 20th century. This formulation (and its higher dimensional generalization) is due to \cite{Mori82}. See also \cite{koll96, km98} for proofs. \begin{thm}[Cone Theorem]\label{cone.thm.c} Let $F$ be a smooth projective surface over an algebraically closed field. Then there are curves $C_i\subset F$ such that $$ \overline{NE}(F)=\overline{NE}(F)_{K\geq 0}+\sum \r^{\geq 0}[C_i], $$ and the $\r^{\geq 0}[C_i]$ are $K_F$-negative extremal rays of $\overline{NE}(F)$. Moreover, we can assume that each $C_i\subset F$ is a smooth rational curve and $(C_i^2)\in\{-1,0,1\}$. If $(C_i^2)=1$ (resp. $(C_i^2)=0$) for some $i$ then $F\cong \p^2$ (resp. $F$ is a minimal ruled surface over a curve). \qed \end{thm} Let now $F$ be a smooth projective surface over $\r$. If $C$ is a 1--cycle on $F_{\c}$ then $C+\bar C$ is a 1--cycle on $F$, and every 1--cycle on $F$ arises this way, at least if we use rational or real coefficients. Thus (\ref{cone.thm.c}) immediately gives: \begin{thm}[Cone Theorem over $\r$]\label{cone.thm.r} Let $F$ be a smooth projective surface over $\r$. Then there are smooth rational curves $C_i\subset F_{\c}$ with $(C_i^2)\in\{-1,0,1\}$ such that $$ \overline{NE}(F)=\overline{NE}(F)_{K\geq 0}+\sum \r^{\geq 0}[C_i+\bar C_i], $$ and the $\r^{\geq 0}[C_i+\bar C_i]$ are $K_F$-negative extremal rays of $\overline{NE}(F)$. \end{thm} Proof. There is one point that we need to be careful about. Namely, it happens frequently that $C_i$ gives an extremal ray but $C_i+\bar C_i$ does not. So we have to throw away some of the $C_i$ appearing in (\ref{cone.thm.c}). \qed \medskip \begin{say}[Geometric irreducibility] Let $X\subset \p^n$ be a variety over $\c$ and $\bar X$ the variety defined by conjugate equations. The disjoint union of $X$ and $\bar X$ is invariant under conjugation, and so there is a real variety $Y_{\r}$ such that $Y_{\c}\cong X\cup \bar X$. Such real varieties are not particularly interesting since the theory of $Y_{\r}$ over $\r$ is equivalent to the theory of $X$ over $\c$. Thus it is reasonable to restrict our attention to real varieties $Y$ such that $Y_{\c}$ is irreducible, that is, $Y$ is geometrically irreducible. Of course, during a proof we may run into a subvariety of $Y_{\r}$ which is geometrically reducible, and these have to be dealt with appropriately. Thus we can not ignore such varieties completely. \end{say} \begin{defn} Let $S$ be a smooth projective surface over a field $k$. $S$ is called a {\it Del Pezzo surface} if $S$ is geometrically irreducible and $-K_S$ is ample. It is called {\it minimal} (over $k$) if $\rho(S)=1$. $S$, together with a morphism to a smooth curve $f:S\to B$ is called a {\it conic bundle} if every fiber is isomorphic to a plane conic. A conic bundle is called {\it minimal} if $\rho(S)=2$. \end{defn} The geometric description and meaning of the extremal rays occurring in (\ref{cone.thm.r}) is given in the next result: \begin{thm}\label{mmp.over.R} Let $F$ be a smooth projective geometrically irreducible surface over $\r$ and $R\subset \nec{F}$ a $K_F$-negative extremal ray. Then $R$ can be contracted $f:F\to F'$, and we obtain one of the following cases: \begin{enumerate} \item[(B)] (Birational) $F'$ is a smooth projective surface over $\r$ and $\rho(F')=\rho(F)-1$. $F$ is the blow up of $F'$ at a closed point $P$. We have two cases: \begin{enumerate} \item $P\in F'(\r)$, or \item $P$ is a pair of conjugate points. \end{enumerate} \item[(C)] (Conic bundle) $B:=F'$ is a smooth curve, $\rho(F)=2$ and $F\to B$ is a conic bundle. The fibers $f^{-1}(P): P\in B(\c)$ are smooth, except for an even number of ponts $P_1,\dots,P_{2m}\in B(\r)$. $(K_F^2)=8(1-g(B))-2m$. \item[(D)] (Del Pezzo surface) $F'$ is a point, $\rho(F)=1$, $-K_F$ is ample and we have one of the following cases \begin{enumerate} \item $(K_F^2)=9$ and $F\cong \p^2$. \item $(K_F^2)=8$ and $F\cong (x_0^2+x_1^2+x_2^2-x_3^2=0)\subset \p^3$. \item $(K_F^2)=2$. \item $(K_F^2)=1$. \end{enumerate} \end{enumerate} \end{thm} Proof. By (\ref{mmp.over.R}), there is a curve $C\cong \p^1$ over $\c$ such that $C+\bar C$ generates $R$ and $(C^2)\in \{-1,0,1\}$. We consider various possibilities. Assume first that $(C^2)=-1$. If $C=\bar C$ then the contraction of $C$ in $F_{\c}$ is defined over $\r$, thus $F$ is the blow up of a surface at a real point. (This is Castelnuovo's contraction theorem, cf. \cite[V.5.7]{Hartshorne77}.) If $C$ and $\bar C$ are disjoint, then we can contract them simultaneously over $\r$ to obtain $f:F\to F'$ which is an isomorphism near $F(\r)$. If $(C^2)=-1$ and $(C\cdot \bar C)=1$ or $(C^2)= (C\cdot \bar C)=0$, then $C+\bar C$ has selfintersection 0. From Riemann--Roch we obtain that $$ h^0(F, \o_F(m(C+\bar C)) \geq \chi (F, \o_F(m(C+\bar C))= m+\chi(\o_F), $$ thus $m(C+\bar C)$ moves in a linear system for $m\gg 1$. It's moving part is base point free by (\ref{bpf.exer}). Let $f:F\to B$ be the Stein factorization of the resulting morphism. Let $A\subset F$ be an irreducible fiber. If $A$ is a multiple fiber, write it as $A=mA_1$. Since $[A_1]\in R$, $$ 2g(A_1)-2=(A_1^2)+(A_1\cdot K_F)=(A_1\cdot K_F)<0. $$ Thus $A_1$ is isomorphic to a smooth conic over $\r$ (\ref{conics.r}) and $(A_1\cdot K_F)=-2$. The generic fiber $A_g$ is not multiple, so $(mA_1\cdot K_F)=(A_g\cdot K_F)=-2$ which shows that there are no multiple fibers. Let $A_1+ A_2=f^{-1}(b)$ be a reducible fiber over $\c$, where $A_1$ is an irreducible and reduced curve. In particular, $(A_1^2)<0$. $\bar A_1+ \bar A_2=f^{-1}(\bar b)$ is also a fiber. If $b\neq \bar b$ then $A_1$ is disjoint from $\bar A_1$, thus $((A_1+\bar A_1)^2)=(A_1^2)+(\bar A_1^2)<0$. $[A_1+\bar A_1]+[A_2+\bar A_2]\in R$ and $[A_1+\bar A_1]\not\in R$, a contradiction. Thus all singular fibers lie over real points $P_1,\dots, P_r\in B(\r)$. Therefore, $A_1+\bar A_1\subset f^{-1}(b)$. Every fiber of $f$ is irreducible over $\r$, thus $A_1+\bar A_1= f^{-1}(b)$. We get that $(A_1\cdot K_X)=-1$ and so $A_1$ is a $-1$-curve and $A_1+\bar A_1$ is isomorphic to a pair of conjugate lines in $\p^2$. $F$ is a conic bundle over $B$ by (\ref{conic.bund}). Over $\c$ we can contract one of the components of every singular fiber to obtain a minimal ruled surface. The selfintersection number of the canonical class of a minimal ruled surface is $8(1-g(B))$, and each singular fiber drops this number by 1. We see in (\ref{mmp.over.R.top}) that the number of singular fibers is even. Assume next that $(C^2)=-1$ and $r:=(C\cdot \bar C)\geq 2$. Then $$ ((C+\bar C)^2)=-2+2(C\cdot \bar C)=2r-2>0. $$ By (\ref{int.cone.exrc}) this implies that $\nec{F}$ is 1-dimensional, hence $-K_F\equiv a(C+\bar C)$ for some $a>0$. $(-K_F\cdot (C+ \bar C))=2$, thus $$ (C+\bar C)\equiv (1-r)K_F\qtq{and} 2r-2=(1-r)^2(K_F^2). $$ This gives the possibilities $r=2, (K_F^2)=2$ or $r=3, (K_F^2)=1$. If $(C^2)=0$ and $r:=(C\cdot \bar C)\geq 1$ then a computation as above gives that $8=r(K_F^2)$, which allows too many cases. It is better to consider this geometrically. By (\ref{cone.thm.c}), $C$ is a fiber of a $\p^1$-bundle $g:F_{\c}\to D$ over $\c$ and $\bar C$ is a (possibly multiple) section of $g$. Thus $D$ is rational. By the classification of minimal ruled surfaces, either $F_{\c}\cong \p^1\times \p^1$ and we are done by (\ref{quadric.lem}), or $g$ has a unique section $E$ with negative selfintersection. $E$ is then defined over $\r$, thus $\rho(F)=2$, a contradiction. We are left with the case when $F_{\c}\cong \c\p^2$ and $C$ is a line in $\c\p^2$. Then $\bar C$ is another line and $C$ and $\bar C$ intersect in a unique point, which is therefore real. We can get another real point, and so also a real line. Thus $\o_F(1)$ is defined over $\r$ and $F\cong \r\p^2$. \qed \medskip As a consequence we obtain the minimal model program (MMP for short) for real algebraic surfaces: \begin{thm}[MMP for surfaces]\label{mmp.surf.thm} Let $F$ be a smooth projective geometrically irreducible surface over $\r$. Then there is a sequence of morphisms $$ F=F_0\stackrel{f_0}{\to} F_1 \to \cdots F_{m-1}\stackrel{f_{m-1}}{\to} F_m=F^* $$ such that each $f_i:F_i{\to} F_{i+1}$ is a birational contraction as in (\ref{mmp.over.R}.B) and $F^*$ satisfies precisely one of the following properties: \begin{enumerate} \item[(M)] (Minimal model) $K_{F^*}$ is nef. (That is, it has nonnegative intersection number with every curve in $F^*$.) \item[(C)] (Conic bundle) $F^*$ is a conic bundle over a curve $f:F^*\to B$. In particular, $\rho(F^*)=2$ \item[(D)] (Del Pezzo surface) $\rho(F^*)=1$, $-K_{F^*}$ is ample and $F^*$ is among those listed in (\ref{mmp.over.R}.D). \end{enumerate} \end{thm} Proof. We do the steps of (\ref{mmp.over.R}.B) as long as we can. $\rho(F_{i+1})=\rho(F_i)-1$, so eventually we reach $F^*=F_m$ where we can not perform a contraction as in (\ref{mmp.over.R}.B). If $K_{F^*}$ is nef then we have a minimal model. If $K_{F^*}$ is not nef, then by (\ref{mmp.over.R}) we can perform a contraction as in (\ref{mmp.over.R}.C--D). This gives our last two cases.\qed \begin{exrc}\label{conics.r} Let $k$ be a field and $C$ a smooth projective curve over $k$ such that $C_{\bar k}\cong \p^1$. Show that $C$ is isomorphic to a smooth conic over $k$. Also, $C\cong \p^1_k$ iff $C(k)\neq\emptyset$. \end{exrc} \begin{exrc}\label{conic.bund} Let $S$ be a smooth projective surface over $\c$ and $f:S\to B$ a morphism to a smooth curve. Assume that $f^{-1}(b)\cong \p^1$ for some $b\in B$ and every fiber has at most 2 irreducible components. Show that $-K_F$ is very ample on the fibers, $f_*\o_F(-K_F)$ is a rank 3 vector bundle over $B$ and we have an injection $$ F\DOTSB\lhook\joinrel\rightarrow \proj_B f_*\o_F(-K_F). $$ Under this injection the fibers of $f$ become conics. Such a surface is called a {\it conic bundle} over $B$. \end{exrc} \begin{exrc}\label{projspace.ex} Let $X$ be a variety over $\r$ such that $X_{\c}\cong \p^n$. Show that $X\cong \p^n$ if $n$ is even, but not necessarily if $n$ is odd. \end{exrc} \begin{exrc}[Cohomology commutes with base change]\label{coh.b.c} Let $X$ be a variety over $\r$ and $F$ a coherent sheaf on $X$. Show that $$ H^i(X,F)\otimes_{\r}\c\cong H^i(X_{\c}, F_{\c}). $$ \end{exrc} \begin{exrc}\label{bpf.exer} Let $F$ be a smooth projective surface and $D\subset F$ an irreducible curve such that $(D^2)=0$. Then the moving part of $|mD|$ is either empty or base point free. \end{exrc} \begin{notation} $Q^{r,s}$ denotes the quadric hypersurface $(x_1^2+\cdots +x_r^2-x_{r+1}^2-\cdots-x_{r+s}^2=0)$. $Q^{2,1}$ is isomorphic to $\p^1$ by (\ref{conics.r}). \end{notation} \begin{lem}\label{quadric.lem} Let $F$ be a smooth projective surface over $\r$ such that $F_{\c}\cong \p^1\times \p^1$. Then one of the following holds: \begin{enumerate} \item $F\cong Q^{2,2}\cong Q^{2,1}\times Q^{2,1}$, $\rho(F)=2$ and $F(\r)\sim S^1\times S^1$, \item $F\cong Q^{3,1}$, $\rho(F)=1$ and $F(\r)\sim S^2$, \item $F\cong Q^{4,0}\cong Q^{3,0}\times Q^{3,0}$, $\rho(F)=2$ and $F(\r)=\emptyset$, \item $F\cong Q^{3,0}\times \p^1$, $\rho(F)=2$ and $F(\r)= \emptyset$. \end{enumerate} \end{lem} Proof. Let $C\subset F_{\c}$ be one of the rulings. Then $\bar C$ is another ruling, thus either $(C\cdot \bar C)=0$ or $(C\cdot \bar C)=1$. If $(C\cdot \bar C)=0$ then the linear system $|C+\bar C|$ is defined over $\r$ and maps $F$ onto a conic. Similarly for the other rulings, thus $F$ is the product of two conics. All 3 possibilities are listed. If $(C\cdot \bar C)=1$ then $\o_F(C+\bar C)$ is a line bundle on $F$ which is of type $\o_{F_{\c}}(1,1)$ over $\c$. Thus its global sections embed $F$ as a quadric. $Q^{3,1}$ is the only quadric not yet accounted for.\qed \medskip In the above proofs we had to establish several times that certain line bundles on $F_{\c}$ are defined over $\r$. This is frequently a quite subtle point. Some aspects of it are treated in the next exercise. \begin{exrc}\label{pic=gal.inv} Let $X$ be a scheme over $\r$ and $L$ a line bundle on $X$. Then $L_{\c}$ is a line bundle on $X_{\c}$ and $L_{\c}\cong \bar L_{\c}$. Thus if $M$ is a line bundle on $X_{\c}$ and $M\not\cong \bar M$, then $M$ is not the complexification of a real line bundle. Find a curve $C$ over $\r$ and a line bundle $M$ on $C_{\c}$ such that $M\cong \bar M$ but $M$ is not the complexification of a real line bundle. Let $X$ be a scheme over $\r$ and $M$ a line bundle on $X_{\c}$ such that $M\cong \bar M$. Show that $M^{\otimes 2}$ is the complexification of a real line bundle. If $X$ is connected, reduced and $X(\r)\neq \emptyset$ then $M$ itself is the complexification of a real line bundle. More generally, let $X_K$ be an integral scheme defined over a field $K$ and $L\supset K$ a Galois extension with Galois group $G$. Show that if $X_K$ has a $K$-point then $\pic (X_K)=\pic (X_L)^G$. \end{exrc} \section{The topology of $F(\r)$} In this section we study the MMP from the topological point of view. The main results of this section are already in \cite{Comessatti14}. \begin{notation} $M\uplus N$ denotes the disjoint union of $M$ and $N$. $\uplus rN$ denotes the disjoint union of $r$ copies of $N$. $M\# N$ denotes the connected sum of two manifolds $M$ and $N$ (which are assumed to have the same dimension). $\# rN$ denotes the connected sum of $r$ copies of $N$. (By definition, $\#0M=S^{\dim M}$.) $M\sim N$ denotes that $M$ and $N$ are homeomorphic. \end{notation} One can give a complete topological description of the various contractions in (\ref{mmp.over.R}): \begin{thm}\label{mmp.over.R.top} Let $F$ be a smooth projective geometrically irreducible surface over $\r$ and $R\subset \nec{F}$ a $K_F$-negative extremal ray. The following is the topological description of the corresponding contraction: \begin{enumerate} \item[(B)] (Birational) $F$ is the blow up of $F'$ at a closed point $P$. We have two cases: \begin{enumerate} \item If $P\in F'(\r)$ then $F(\r)\sim F'(\r)\#\r\p^2$. \item If $P$ is a pair of conjugate points then $F(\r)\sim F'(\r)$. \end{enumerate} \item[(C)] (Conic bundle) $f:F\to B $ is a conic bundle with singular fibers $f^{-1}(P_1),\dots,f^{-1}(P_{2m})$. Then $$ F(\r)\sim \uplus mS^2\uplus N_1\uplus\cdots \uplus N_b, $$ where $b$ is the number of connected components of $B(\r)$ which do not contain any of the points $P_i$ and each $N_i$ is either a torus or a Klein bottle. \item[(D)] (Del Pezzo surface) There are 4 cases: \begin{enumerate} \item If $(K_F^2)=9$ then $F(\r)\sim \r\p^2$. \item If $(K_F^2)=8$ then $F(\r)\sim S^2$. \item If $(K_F^2)=2$ then $F(\r)\sim \uplus 4S^2$. \item If $(K_F^2)=1$ then $F(\r)\sim \r\p^2\uplus 4S^2$. \end{enumerate} \end{enumerate} \end{thm} Proof. Blowing up replaces a point with all tangent directions through that point. So we remove a disc and put in an interval bundle over $S^1$ whose boundary is connected. This is a M\"obius strip and so $F(\r)\sim F'(\r)\#\r\p^2$. In the conic bundle case, let $M\sim S^1$ be a connected component of $B(\r)$. If none of the $P_i$ lie on $M$ then $F(\r)\to B(\r)$ is a smooth $S^1$-bundle over $M$, this gives either a torus or a Klein bottle. If $k$ of the points $P_1,\dots,P_k\in M\sim S^1$ correspond to singular fibers then, after reindexing, they divide $M$ into $k$ intervals $[P_i,P_{i+1}]$ (subscript $\mod k$). $F(\r)$ is alternatingly empty or a copy of $S^2$ over the intervals. Thus $k$ is even. In the Del Pezzo case we are done if $(K_F^2)=9,8$. The cases $(K_F^2)=2,1$ are considerably harder. They follow from (\ref{deg2.equiv.thm}) and (\ref{deg1.equiv.thm}).\qed \medskip Using (\ref{mmp.over.R.top}) it is easy to determine which 2-manifolds occur as $F(\r)$ for geometrically rational surfaces $F$. The conclusion is that orientable surfaces of genus $>1$ do not occur. This is the main result of \cite{Comessatti14}. \begin{thm}\label{c-rat.top} Let $F$ be a smooth, projective surface over $\r$ such that $F_{\c}$ is rational. Then one of the following holds: \begin{enumerate}\setcounter{enumi}{-1} \item $F(\r)=\emptyset$. \item $F(\r)\sim S^1\times S^1$. \item $F(\r)\sim \#r_1\r\p^2\uplus \cdots \uplus \#r_m\r\p^2$ for some $r_1,\dots,r_m\geq 0$. \end{enumerate} All these cases do occur. \end{thm} Proof. Apply the MMP over $\r$ to get $F=F_1\to F_2\to\cdots$. We prove the theorem by induction on the number of blow ups in the sequence. If $F_i\to F_{i+1}$ is the inverse of the blowing up of a real point, then $F_i(\r)\sim F_{i+1}(\r)\#\r\p^2$. If $F_i\to F_{i+1}$ is the inverse of the blowing up of a pair of conjugate points, then $F_i(\r)\sim F_{i+1}(\r)$. The induction works since $(S^1\times S^1)\#\r\p^2\sim \#3\r\p^2$. Thus we are reduced to one of the following two cases: \begin{enumerate} \item $F$ has a conic bundle structure $F\to B$, or \item $F$ is Del Pezzo and $\rho(F)=1$. \end{enumerate} In the first case, $B_{\c}\cong \c\p^1$ since $F_{\c}$ is rational. Thus either $B(\r)=\emptyset$ and so $F(\r)=\emptyset$, or $B\cong \r\p^1$. Thus $F(\r)$ is the torus or the Klein bottle if there are no singular fibers and $F(\r)\sim \uplus mS^2$ if there are $2m>0$ singular fibers by (\ref{mmp.over.R.top}.C). Note that $S^2=\#0\r\p^2$ by convention. In the second case we use (\ref{mmp.over.R.top}.D). \qed \begin{exrc}\label{comp.bir.inv} Let $X$ and $Y$ be smooth projective varieties over $\r$. Assume that $X$ and $Y$ are birational to each other (over $\r$). Show that $X(\r)$ and $Y(\r)$ have the same number of connected components. \end{exrc} \begin{say}[Vector bundles over real varieties] Let $X$ be a veriety over $\r$ and $p:V\to X$ a vector bundle of rank $n$. Locally $V$ is like $U\times {\mathbb A}_{\r}^n\to U$ where $U\subset X$ is Zariski open and ${\mathbb A}_{\r}^n=\spec_{\r}\r[t_1,\dots,t_n]$ is affine $n$-space over $\r$. (Which should {\em not} be identified with $\r^n$!) As usual, to $V$ one can associate a vector bundle $p_{\c}:V_{\c}\to X_{\c}$ and also a real vector bundle $p(\r):V(\r)\to X(\r)$ which is obtained by taking the $\r$-valued points of ${\mathbb A}_{\r}^n$ which is exactly $\r^n$. (To complete the picture, any real vector bundle on a manifold can be complexified, and $V(\r)\otimes_{\r}\c\cong V_{\c}|_{X(\r)}$.) \end{say} \begin{say}[Degrees of line bundles over $\r$ and $\c$] \label{dolb.exrc}{\ } Let $B$ be a smooth projective curve over $\c$ and $L$ a line bundle on $B$. Let $s$ be a nonzero meromorphic section of $L$. The number of zeros minus the number of poles of $s$ on $B$ (counted with multiplicity) is called the {\it degree of $L$}. Let $Y$ be a smooth projective variety over $\c$ and $L$ a line bundle on $Y$. For any curve $B\subset Y$ the degree of $L|_B$ is defined. It is also called the {\it intersection number} of $B$ and $L$ and denoted by $(B\cdot L)$. Let $A\sim S^1$ be a compact 1--dimensional manifold and $L$ a real line bundle on $M$. Let $s$ be a nonzero section of $L$. The number of zeros of $s$ on $A$ (counted with multiplicity) makes sense only mod 2. If $M$ is a compact manifold and $L$ a real line bundle on $M$ then for any 1-cycle $A\subset M$ we obtain the $\z_2$-valued {\it intersection number} of $A$ and $L$. It is denoted by $(A\cap L)$. (To be precise, I should write $(A\cap w_1(L))$ where $w_1(L)$ stands for the first Stiefel--Whitney class of $L$. This is a class in $H^1(X(\r),\z_2)$ analogous to the first Chern class of a complex line bundle, cf. \cite[Sec. 4]{milnor-s74}.) Let now $X$ be a smooth projective variety over $\r$, $C\subset X$ a curve and $L$ a line bundle on $X$. We obtain two numbers: $$ (L(\r)\cap C(\r)) \qtq{and} (C_{\c}\cdot L_{\c}). $$ What is the relationship between them? To answer this, take a real meromorphic section $s$ of $L$ which has only finitely many zeros and poles on $C$. When we count the real zeros and poles of $s$ on $C(\r)$, we miss the complex zeros and poles of $s$ on $C_{\c}$. Since $s$ is real, the complex zeros and poles come in conjugate pairs. Thus we conclude that $$ (C(\r)\cap L(\r))\equiv (C_{\c}\cdot L_{\c}) \mod 2, $$ which is best possible since the left hand side is defined only mod 2 anyhow. \end{say} \begin{say}[Orientability of $X(\r)$ and the canonical class]{\ } Let $M$ be a differentiable manifold, $0\in M$ a point and $x_1,\dots,x_n$ local coordinates. A {\it local orientation} of $M$ at $0$ is a choice of an $n$-form $f(x)dx_1\wedge \dots\wedge dx_n$ with $f(0)\neq 0$ up to multiplication by a positive function. An {\it orientation} of $M$ is a nowhere zero global $n$-form on $M$, up to multiplication by a positive function. $n$-forms are sections of the real line bundle $\det T^*_M$. If $S^1\sim A\subset M$ is a loop then one can choose a consistent oreintation of $M$ along $A$ $\Leftrightarrow$ $\det T^*_M$ has a nowhere zero section along $A$ $\Leftrightarrow$ $(\det T^*_M\cap A)=0$. If $X$ is a smooth variety over $\r$ then $n$-forms appear as sections of the canonical line bundle. This proves that $$ \det T^*_{X(\r)} \cong K_X(\r). $$ In many cases this gives a way to decide if $X(\r)$ is orientable or not. \end{say} \begin{cor} Let $X$ be a smooth projective variety over $\r$. Assume that there is a curve $C\subset X$ such that $(C\cdot K_X)$ is odd. Then $X(\r)$ is not orientable. \end{cor} Proof. We have proved above that $$ (\det T^*_{X(\r)}\cap C(\r))\equiv (C\cdot K_X) \equiv 1\mod 2. $$ $C(\r)$ may have several components, but along one of them $\det T^*_{X(\r)}$ has odd degree, so we can not choose a consistent orientation along that component.\qed \begin{exrc} Show that $\r\p^n$ is orientable iff $n$ is odd. Let $X\subset \p^n$ be a smooth hypersurface of degree $d$. Show that $X(\r)$ is orientable if $n-d$ is odd. Show that $X(\r)$ is not orientable if $n$ and $d$ are both odd. If $n$ and $d$ are both even, then $X(\r)$ may or may not be orientable. \end{exrc} If $(C\cdot K_X)$ is even, then it can happen that $X(\r)$ is not orientable along an even number of components of $C(\r)$. In some cases we are still able to conclude orientability of $X(\r)$ using stronger assumptions: \begin{exrc}\label{gen.orient.ex} Let $X$ be a smooth projective variety over $\r$. Assume that $K_X\cong L^{\otimes 2}$ for a real line bundle $L$. Show that $X(\r)$ is orientable. More generally, assume that $K_X\cong \o_X(2D+D')$ where $D,D'$ are divisors over $\r$ and $D'(\r)$ has codimension at least 2 in $X(\r)$. (This is equivalent to assuming that every irreducible component of $D'$ is geometrically reducible.) Show that $X(\r)$ is orientable. \end{exrc} \section{Birational classification} \begin{defn} Let $F$ be a smooth real algebraic surface. A surface obtained from $F$ by blowing up $a$ real points and $b$ pairs of conjugate complex points (possibly infinitely near) is denoted by $(F,a,2b)$. Given $F$ and $a,b$, the surfaces of the form $(F,a,2b)$ consitute a connected family if $F(\r)$ and $F_{\c}$ are both connected. \end{defn} \begin{lem}\label{elem.bir.lem} We have the following elementary birational equivalences between the minimal models in (\ref{mmp.over.R}). \begin{enumerate} \item $(\p^2,2,0)\cong (Q^{2,2},1,0)$. \item $(\p^2,0,2)\cong (Q^{3,1},1,0)$. \item $(Q^{4,0},0,2)$ is isomorphic to the blow up of $Q^{3,0}\times \p^1$ at a pair of conjugate points on the same section $Q^{3,0}\times P$, $P\in \r\p^1$. \item Any minimal conic bundle over a rational curve with $2$ singular fibers is isomorphic to $(Q^{3,1},0,2)$. \end{enumerate} \end{lem} Proof. In the first two cases we blow up the 2 points in $\p^2$ and then contract the line through them to get a quadric. $Q^{4,0}\cong Q^{3,0}\times Q^{3,0}$, let $\pi_1$ be the first projection. The pencil of planes through the 2 points gives a map $p:Q^{4,0}\map \p^1$. $$ (\pi_1,p):Q^{4,0}\map Q^{3,0}\times \p^1 \qtq{is birational} $$ and becomes a morphism after blowing up the 2 points. Finally assume that $F\to B$ is a minimal conic bundle over a rational curve with $2$ singular fibers. By (\ref{mmp.over.R}.C), $B(\r)\neq\emptyset$, thus $B\cong \p^1$. $F_{\c}$ is the blow up of a minimal ruled surface $F''$ at 2 points. We can even assume that $F''$ has a section $E$ with negative selfintersection $(E^2)=-k$ and the two points are not on $E$. If $k\geq 2$ then all other sections of $F''$ have selfintersection at least 2, so $E\subset F$ is the unique section with negative selfintersection. Thus $E$ is defined over $\r$ and $F\to B$ is not minimal. Thus $k=1$ and there is a unique section $E'\subset F''$ such that ${(E'}^2)=1$ and $E'$ passes through the two blown up points. Let $\bar E\subset F$ be the birational transform. Then $E$ and $\bar E$ have to be conjugate. Contracting them gives the quadric $Q^{3,1}$.\qed \begin{lem}\label{minruled.lem} Let $F$ be a smooth projective surface over $\r$ such that $F_{\c}$ is a minimal ruled surface over $\c\p^1$. Then one of the following holds: \begin{enumerate} \item $F_{\c}\cong \p^1\times \p^1$ (these cases were enumerated in (\ref{quadric.lem})), or \item $F\cong \proj_B(\o_B+\o_B(-r))$ is a minimal ruled surface over a smooth real conic $B$ for some $r>0$. \end{enumerate} \end{lem} Proof. By the classification of minimal ruled surfaces, either $F_{\c}\cong \p^1\times \p^1$, or $F_{\c}$ has a unique irreducible curve $E$ with negative selfintersection $-r$. $E$ and the ruling $g:F\to B$ are then defined over $\r$ and $g_*\o_F(E)\cong \o_B+\o_B(-r)$. \qed \medskip As a corollary, we obtain the following birational classification of real surfaces such that $F_{\c}$ is rational: \begin{cor}\label{birclass.over.R} Let $F$ be a smooth real projective surface such that $F_{\c}$ is rational. Then $F$ is birationally equivalent over $\r$ to a surface in exactly one of the following classes: \begin{enumerate} \item[1.] $Q^{3,0}\times \p^1$. In this case $F(\r)=\emptyset$. \item[2.] $\p^2$. In this case $F(\r)$ is connected. \item[3$_m$.] Minimal conic bundle with $2m$ ($m\geq 2$) singular fibers. In this case $F(\r)$ has $m$ connected components. \item[4.] Degree 2 minimal Del Pezzo surface. \item[5.] Degree 1 minimal Del Pezzo surface. \end{enumerate} \end{cor} Remark. In (\ref{deg2.equiv.thm}) and (\ref{deg1.equiv.thm}) we prove that $F(\r)$ has 4 (resp. 5) connected components if $F$ is a minimal Del Pezzo surface of degree 2 (resp. 1). \medskip Proof. Let $F\to F^*$ be the minimal model of $F$. By (\ref{mmp.over.R}) $F^*$ is either one of those listed above, or $F^*$ is a conic bundle with $0$ or $2$ singular fibers. The former are treated in (\ref{minruled.lem}). The latter are birational to $Q^{3,0}\times \p^1$ by (\ref{elem.bir.lem}). The number of connected components of the real part is a birational invariant (\ref{comp.bir.inv}), hence the cases (1--3$_m$) are all different birationally. The cases (4--5) differ birationally from the other ones by (\ref{seg-man.deg123}).\qed We use, without proof, the following result about the birational classification of low degree Del Pezzo surfaces over any field. Lectures 2--3 of \cite{KS97} serve as a good introduction. \begin{thm}\cite{Segre51, Manin66}\label{seg-man.deg123} Let $k$ be a field (of characteristic zero) and $F$ a minimal Del Pezzo surface of degree 1,2 or 3 over $k$. Then \begin{enumerate} \item $F$ is not rational (over $k$), \item $F$ is not birational (over $k$) to any conic fibration, \item $F$ is birational to another minimal Del Pezzo surface $F'$ of degree 1,2 or 3 over $k$ iff $F$ is isomorphic to $F'$. \qed \end{enumerate} \end{thm} This theorem, (\ref{birclass.over.R}) and (\ref{comp.bir.inv}) imply the following: \begin{cor}\label{rat.char.over.R}\cite[VI.6.5]{Silhol89} Let $F$ be a smooth projective surface over $\r$. The following are equivalent: \begin{enumerate} \item $F$ is birational to $\p^2$ over $\r$. \item $F_{\c}$ is birational to $\c\p^2$ and $F(\r)$ is connected. \qed \end{enumerate} \end{cor} \section{Birational Classification of Conic Fibrations} \begin{defn} Let $F$ be a smooth projective surface over a field $k$. A morphism $f:F\to B$ to a smooth curve is called a {\it conic fibration} if the generic fiber is isomorphic to a plane conic (over $k(B)$). By (\ref{conics.r}) this is equivalent to assuming that $f^{-1}(b)\cong \p^1_{\bar k}$ for a general $b\in B(\bar k)$. \end{defn} In this section we discuss the birational classification of those surfaces over $\r$ which admit a conic fibration. This covers all surfaces where the MMP ends with the case (\ref{mmp.surf.thm}.C). This is done in two steps. First we consider those birational maps which preserve the conic fibration. To be precise: \begin{defn} Two conic fibrations $f:F\to B$ and $f':F'\to B'$ are called {\it birational} if there is a birational map $\phi:F\map F'$ and an isomorphism $\tau: B \cong B'$ (both over $k$) which give a commutative diagram $$ \begin{array}{ccc} F & \stackrel{\phi}{\map} & F'\\ f\downarrow{\ } & & {\ }\downarrow f'\\ B & \stackrel{\tau}{\map} & B' \end{array} $$ \end{defn} The second step is to understand the birational maps between $F$ and $F'$ which do not preserve the conic fibration. Fortunately, in many cases there are no such maps. (For a proof see \cite[V--VI]{Silhol89} or the original paper of \cite{Iskovskikh67}.) \begin{thm}\label{cb.unique.thm} Let $k$ be a field and $f:F\to B$ a relatively minimal conic bundle over $B$. Let $f':F'\to B'$ be any conic fibration and $\phi:F\map F'$ any birational map (over $k$). Then \begin{enumerate} \item If $(K_F^2)\leq 0$ then $\phi$ is a birational map of the conic fibrations. \item If $(K_F^2)=2$ then $F$ and $F'$ are birational conic fibrations (though $\phi$ itself need not respect the fibration structure).\qed \end{enumerate} \end{thm} \begin{defn} Let $f:F\to B$ be a conic fibration over $\r$. The image of the set of real points $f(F(\r))\subset B(\r)$ is a union of finitely many closed intervals. Let us denote it by $I(F)$. \end{defn} The main theorem of the section shows that $I(f)$ characterizes $f$: \begin{thm} Two conic fibrations $f:F\to \p^1$ and $f':F'\to \p^1$ over $\r$ are birational iff there is an isomorphism $\tau:\p^1 \cong \p^1$ such that $\tau(I(f))=I(f')$. \end{thm} Proof. Let $\phi:F\map F'$ and $\tau:\p^1 \cong \p^1$ be a birational map of the two conic fibrations. Then $F'(\r)$ and $\phi(F(\r))$ agree outside finitely many fibers, thus $I(f')$ and $\tau(I(f))$ differ only at finitely many points. Unions of closed intervals can not differ at finitely many points only, thus in fact $\tau(I(f))=I(f')$. The converse is established by bringing each conic fibration to a normal form. (The roots $a_i$ in (\ref{cb.standard.form.thm}) are the boundary points of $I(f)$. This leaves two choices for $I(f)$ itself, corresponding to the two choices of the sign on the right hand side.) \qed \begin{thm}\label{cb.standard.form.thm} Let $f:F\to \p^1$ be a conic fibration over $\r$. Then $f$ is birational to a conic fibration $f':F'\to \p^1$ with affine equation $$ x^2+y^2=\pm \prod_{i=1}^{2m}(z-a_i)\subset {\mathbb A}^3, $$ where the $a_i$ are distinct real numbers. \end{thm} The proof rests on the following simple lemma about quadratic forms: \begin{lem}\label{cb.standard.form.lem} Let $k$ be a field (of characteristic different from 2) and $Q(x_0,\dots,x_n)$ a quadratic form over $k$ which is anisotropic (that is $Q=0$ has no nontrivial solution over $k$). For any $a\in k$ the following are equivalent \begin{enumerate} \item $Q=0$ has a nontrivial solution over $k(\sqrt{a})$. \item After a suitable coordinate change, $Q$ can be written as \noindent $b(y_0^2-ay_1^2)+Q'(y_2,\dots,y_n)$. \end{enumerate} \end{lem} Proof. (2) $\Rightarrow$ (1) is shown by the substitution ${\bf y}:=(\sqrt{a}:1:0\cdots:0)$. Conversely, assume that ${\bf v}\in k(\sqrt{a})^{n+1}$ satisfies $Q({\bf v})=0$. Let $\bar {\bf v}$ denote the conjugate of ${\bf v}$. Then $Q(\bar{\bf v})=0$. ${\bf v}$ and $\bar{\bf v}$ span a 2-dimensional linear subspace of $k(\sqrt{a})^{n+1}$ which is defined over $k$. That is, there is a linear subspace $V\subset k^{n+1}$ such that $V\otimes_kk(\sqrt{a})=\langle {\bf v}, \bar{\bf v}\rangle$. $Q$ is nondegenerate on $V$ (since $Q=0$ has no solutions in $k$), thus $k^{n+1}=V+V^{\perp}$ where $V^{\perp}$ is the orthogonal complement of $V$ with respect to $Q$. Let $y_2,\dots,y_n$ be coordinates on $V^{\perp}$ and choose coordinates $y_0,y_1$ on $V$ such that $$ {\bf v}=(\sqrt{a},1)\qtq{and} \bar{\bf v}=(-\sqrt{a},1).\qed $$ \begin{say}[Proof of (\ref{cb.standard.form.thm})]{\ } Let $k:=\r(t)$ be the quotient field of $\p^1_{\r}$. The generic fiber of $F\to \p^1$ is birational to a plane conic $C_k$ over $k$ (\ref{conics.r}). If $C_k$ has a $k$-point (equivalently, if $F\to \p^1$ has a section) then $F$ is birational to $\p^1\times \p^1$ by (\ref{conics.r}). $C_k$ has a point over $k(\sqrt{-1})=\c(t)$ (equivalently, $F\to \p^1$ has a section over $\c$). Thus by (\ref{cb.standard.form.lem}), in suitable coordinates the equation of $C_k$ becomes $x_0^2+x_1^2=g(t)x_2^2$ for some $g(t)\in \r(t)$. We can multiply through with the square of the denominator of $g$, thus we may assume that $g(t)\in \r[t]$. Write $g=f(t)g_1(t)^2\prod_a (z-a)(z-\bar a)$ where $f$ has only simple real roots and the $a$ are nonreal complex numbers. We can divide by $g_1(t)^2$. If $a=u+iv$ then $(z-a)(z-\bar a)=(z-u)^2+v^2$. Note that \begin{eqnarray*} g(t)(h_0^2+h_1^2)x_2^2&=&x_0^2+x_1^2\qtq{is equivalent to}\\ g(t)x_2^2&=& \left(\frac{x_0h_0-x_1h_1}{h_0^2+h_1^2}\right)^2+ \left(\frac{x_0h_1+x_1h_0}{h_0^2+h_1^2}\right)^2. \end{eqnarray*} Using this, we can get rid of the complex factors $\prod_a (z-a)(z-\bar a)$ one at a time. At the end we obtain the required normal form, except that we may have an odd number of factors on the right hand side: $$ x^2+y^2=\pm \prod_{i=1}^{2m-1}(z-a_i). $$ In this case we first apply a translation to ensure that $0$ is not among the $a_i$ and then make a substitution $(x,y,z)\mapsto (xz^{-n},yz^{-n}, z^{-1})$ to get the equation $$ x^2+y^2=\pm \prod_{i=1}^{2m}(z-a'_i), $$ where $a'_i=a_i^{-1}$ for $i<2m$ and $a'_{2m}=0$.\qed \end{say} Putting things together, we obtain the following criterion for birational equivalence of conic fibrations. The result corrects a slight inaccuracy in \cite[VI.3.15]{Silhol89}. \begin{exrc}\label{cb.isom.ex} Two conic bundles \begin{eqnarray*} F & = & (x^2+y^2= c\prod_{i=1}^{2m}(z-a_i))\to \p^1\qtq{and}\\ F' & = & (x^2+y^2= c'\prod_{i=1}^{2m}(z-a'_i))\to \p^1 \end{eqnarray*} are birational to each other iff there is a permutation $\sigma\in S_{2m}$ and a matrix $$ \left(\begin{array}{cc} \alpha & \beta \\ \gamma & \delta \end{array} \right)\in GL(2,\r) \qtq{such that} a'_{\sigma(i)}=\frac{\alpha a_i+\beta }{\gamma a_i+\delta}, \qtq{and} $$ $c'$ and $c\prod_i(\gamma a_i+\delta)$ have the same sign. \end{exrc} \begin{exrc}\label{geom.cb.ex} Using elementary transformations of conic bundles, give a geometric proof of the results in this section. \end{exrc} \begin{say}[Moduli of conic fibrations]{\ } Let $F$ be a smooth projective surface over $\r$ which admits a conic fibration $f:F\to \p^1$. We proved that if $I(f)\subset \r\p^1$ has at least 3 components then $I(f)$ (modulo the action of $GL(2,\r)$) determines $F$ up to birational equivalence (over $\r$). The space of $m$ disjoint closed intervals in $\r\p^1$ is a connected manifold of real dimension $2m$. The quotient by the $GL(2,\r)$ action gives a $2m-3$ dimensional topological space (it has some quotient singularities). With some more care, we could even realize this space as the set of real points of a $(2m-3)$-dimensional algebraic variety. For $m=0$ there are 2 conic fibrations up to birational equivalence: $\p^1\times \p^1$ gives $I(f)=\r\p^1$ and $Q^{3,0}\times \p^1$ gives $I(f)=\emptyset$. For $m=1$ we have only one birational equivalence class by (\ref{elem.bir.lem}.4). For $m=2$ we see in (\ref{deg4.cb.ex}) that all such surfaces are birational to each other (though they are not birational as conic fibrations). \end{say} \section{Del Pezzo Surfaces of Degree $\geq 3$} In this section we describe all Del Pezzo surfaces of degree $d\geq 3$ over $\r$. \begin{exrc}\label{amp.image.ex} Let $g:S\to S'$ be a birational morphism of smooth surfaces. Show that if $H$ is ample on $S$ then $f(H)$ is ample on $S'$. Thus if $S$ is Del Pezzo then $S'$ is also Del Pezzo. \end{exrc} \begin{prop}\label{bireg.empty} Let $F$ be a smooth real Del Pezzo surface which is birational to $Q^{3,0}\times \p^1$. Set $d:=(K_F^2)$. Then $d\in\{8,6,4,2\}$. If $d=8$ then $F$ is isomorphic to either $Q^{3,0}\times Q^{3,0}$ or to $Q^{3,0}\times \p^1$. If $d<8$ then $F$ is isomorphic to to $Q^{3,0}\times \p^1$ blown up in $\frac12(8-d)$ pairs of conjugate points. Therefore, for $d\in\{6,4,2\}$ such surfaces form a connected family. \end{prop} Proof. Apply the MMP over $\r$ to obtain $F\to \cdots\to F^*$. If $F$ is Del Pezzo then so is $F^*$ by (\ref{amp.image.ex}), and $(K_{F^*}^2)\geq (K_F^2)$. Hence in our case $F^*$ is either $Q^{3,0}\times Q^{3,0}$ or $Q^{3,0}\times \p^1$. By (\ref{elem.bir.lem}.3) any blow up of $Q^{3,0}\times Q^{3,0}$ at a pair of conjugate points is also a blow up of $Q^{3,0}\times \p^1$. \qed \begin{prop}\label{bireg.rtl} Let $F$ be a smooth real Del Pezzo surface which is birational to $\p^2$. Set $d:=(K_F^2)$. Then $9\geq d\geq 1$ and we have one of the following cases: If $d=9$ then $F$ is isomorphic to $\p^2$. If $d=8$ then $F$ is isomorphic to either $Q^{3,1}$ or to $Q^{2,2}$ or to $\p^2$ blown up at a real point. If $d<8$ then $F$ is isomorphic to one of the following: \begin{enumerate} \item $\p^2$ blown up at $a\geq 0$ real points and $b\geq 0$ pairs of conjugate points for some $a+2b=9-d$. Thus $F(\r)\sim \#(a+1)\r\p^2$. \item $Q^{3,1}$ blown up at $b=\frac12(8-d)$ pairs of conjugate points (so $d$ is even). Thus $F(\r)\sim S^2$. \item $Q^{2,2}$ blown up at $b=\frac12(8-d)$ pairs of conjugate points (so $d$ is even). Thus $F(\r)\sim S^1\times S^1$. Therefore, for any $d<8$, such surfaces with a given topological type $F(\r)$ form a connected family. \end{enumerate} \end{prop} Proof. The minimal model of such a surface is either $\p^2$, $Q^{3,1}$ or $Q^{2,2}$. By (\ref{elem.bir.lem}.1--2) any blow up of $Q^{3,1}$ or to $Q^{2,2}$ at a real point is also a blow up of $\p^2$. \qed \medskip The two propositions above account for all Del Pezzo surfaces of degrees $d\geq 5$. The results are summarized in the next statement: \begin{cor}\label{d>4.dp.top} The following table lists all topological types of the real points of Del Pezzo surfaces of degrees $9\geq d\geq 5$. All surfaces of a fixed degree and topological type form a connected family, except for $d=8$ and $F(\r)=\emptyset$ when there are 2 such surfaces. $$ \begin{tabular}{|c|l|} \hline degree & \qquad\qquad topological types\\ \hline 9 & $\r\p^2$\\ \hline 8 & $S^2$ or $S^1\times S^1$ or $\r\p^2\#\r\p^2$ or $\emptyset$ \\ \hline 7 & $\r\p^2$ or $\#3\r\p^2$ \\ \hline 6 & $S^2$ or $S^1\times S^1$ or $\r\p^2\#\r\p^2$ or $\#4\r\p^2$ or $\emptyset$\\ \hline 5 & $\r\p^2$ or $\#3\r\p^2$ or $\#5\r\p^2$ \\ \hline \end{tabular} $$ \end{cor} The following result shows that odd degree Del Pezzo surfaces over $\r$ are relatively easy to understand: \begin{lem}\label{d1=d2.blowup} Every degree $2d-1$ Del Pezzo surface $F$ over $\r$ with $\rho(F)\geq 2$ is the blow up of a degree $2d$ Del Pezzo surface at a real point. \end{lem} Proof. Since $(K_F^2)$ is odd, $F$ is not a minimal conic bundle. Thus $F$ is either the blow up of a degree $2d$ Del Pezzo surface at a real point or the blow up of a degree $2d+1$ Del Pezzo $F'$ surface at a conjugate pair of complex points $P+\bar P$. If $F'\cong \p^2$ then let $L\subset \p^2$ be the line through the two points. Its birational transform on $F$ is a line. Otherwise, $F'$ is again not minimal by (\ref{mmp.over.R}), hence $F'$ contains a line $L$ over $\r$ by induction. $P,\bar P\not\in L$ since otherwise the birational transform of $L$ on $F$ would have a nonnegative intersection number with $K_F$. Thus $L$ gives a real line on $F$. Contracting a real line on $F$ we get a degree $2d$ Del Pezzo surface.\qed \medskip This shows that the study of degree 3 Del Pezzo surfaces is reduced to the study of degree 4 cases. The classification of these two classes is summarized next. These results were obtained by \cite{Schlafli1863}, who actually worked directly with cubic surfaces. \begin{cor}\label{d=4,3.dp.top} The following table lists all topological types of the real points of Del Pezzo surfaces of degrees 4 and 3. All surfaces of a fixed degree and topological type form a connected family. $$ \begin{tabular}{|c|l|} \hline degree & \qquad\qquad topological types\\ \hline 4 & $S^2$ or $S^1\times S^1$ or $\r\p^2\#\r\p^2$ or $\#4\r\p^2$ or $\emptyset$ or $S^2\uplus S^2$\\ \hline 3 & $\r\p^2$ or $\#3\r\p^2$ or $\#5\r\p^2$ or $\#7\r\p^2$ or $S^2\uplus \r\p^2$ \\ \hline \end{tabular} $$ Moreover, in the $S^2\uplus S^2$ case the monodromy interchanges the two components. \end{cor} Proof. As we noted above, it is sufficient to describe all degree 4 Del Pezzo surfaces. If a degree 4 Del Pezzo surface $F$ is obtained from a higher degree surface by blowing up then we are reduced to (\ref{d>4.dp.top}). Otherwise $F$ is a conic bundle over $\p^1$ with 4 singular fibers. These 4 singular fibers give 8 lines on $F$. By looking at the set of all lines over $\c$, we see that the remaining 8 lines again form 4 pairs and determine another morphism to $\p^1$. Thus $F$ is a double cover of $\p^1\times \p^1$ ramified along a curve $D\subset \p^1\times \p^1$ of type $(2,2)$. $D$ has 4 horizontal and 4 vertical tangents and the 16 lines are sitting over these tangents. The rest is a special case of (\ref{(2,2)-curves}). \qed \begin{exrc}\label{(2,2)-curves} Show that the space of all smooth real curves of type $(2,2)$ on $\p^1\times \p^1$ has 7 connected components. They are determined by the homotopy classes of the components of $D(\r)$: $\emptyset$ or $(0,0)$ or $(0,0)\uplus (0,0)$ or $(1,1)\uplus (1,1)$ or $(1,-1)\uplus (1,-1)$ or $(1,0)\uplus (1,0)$ or $(0,1)\uplus (0,1)$. \end{exrc} \begin{exrc}\label{deg4.cb.ex}\cite[VI.3.5]{Silhol89} Using the correspondence between $(2,2)$-curves on $\p^1\times\p^1$ and degree 4 Del Pezzo surfaces show that any two minimal conic bundles $F,F'$ with 4 singular fibers are birational, by producing examples $S\to \p^1\times \p^1$ such that one conic bundle structure of $S$ is birational to $F$ and the other to $F'$. (This should be easier after the next section.) \end{exrc} \begin{exrc}\cite{Schlafli1863}\label{cubic.lines.ex} Show that a smooth cubic over $\r$ has 27,15,7 or 3 real lines. \end{exrc} \section{Del Pezzo Surfaces of Degree 2 and 1} \begin{notation} Let $D\subset \p^2$ be a degree 4 smooth real curve. $D(\r)$ divides $\r\p^2$ into connected open sets and precisely one of these is nonorientable (denoted by $U_D$). We choose an equation $f(x,y,z)\in \r[x,y,z]$ of $D$ such that $f$ is negative on $U_D$. We can associate two different degree 2 Del Pezzo surfaces to $D$. One is $F^+_D:=(u^2=f(x,y,z)\subset \p^3(1,1,1,2)$ and the other $F^-_D:=(u^2=-f(x,y,z)\subset \p^3(1,1,1,2)$. The correspondence $F^+_D\leftrightarrow F^-_D$ is a natural involution on the space of degree 2 real Del Pezzo surfaces. $D$ has 28 bitangents over $\c$ and over each bitangent of $D$ we get a pair of lines on $F^{\pm}_D$. This gives a total of 56 lines. \end{notation} The topological classification of degree 4 plane curves over $\r$ is very old, it is already contained in \cite{Plucker1839}. (See \cite{Viro90} for a recent survey of the study of low degree real plane curves.) This implies the topological classification of degree 2 real Del Pezzo surfaces. The following proposition summarizes these results. \begin{prop}\label{deg2.dp.top} There are 6 topological types of degree 4 smooth real plane curves. Correspondingly there are 12 topological types of degree 2 real Del Pezzo surfaces. The following table gives the complete list. The types in the same row correspond to each other under $D \leftrightarrow F^+_D\leftrightarrow F^-_D$. $$ \begin{tabular}{|c|c|c|} \hline $D(\r)$ & $F^+_D(\r)$ & $F^-_D(\r)$\\ \hline $\bigcirc\bigcirc\bigcirc\bigcirc{}$ & $\uplus 4S^2$ & $\#8\r\p^2$\\ \hline $\bigcirc\bigcirc\bigcirc$ & $\uplus 3S^2$ & $\#6\r\p^2$\\ \hline $\bigcirc\bigcirc{}$ & $S^2\uplus S^2$ & $\#4\r\p^2$\\ \hline $\bigcirc$ & $S^2$ & $ \#2\r\p^2$\\ \hline $\emptyset$ & $\emptyset$ & $\r\p^2\uplus \r\p^2$\\ \hline $\bigcirc \!\!\!\!\!\circ$ & $S^1\times S^1$ & $S^2\uplus \#2\r\p^2$\\ \hline \end{tabular} $$ \end{prop} \cite{Zeuthen1874} studied the bitangents of degree 4 plane curves. He proved the equivalence of (\ref{deg2.equiv.thm}.1) and (\ref{deg2.equiv.thm}.5). He understood the relationship between degree 4 plane curves and cubic surfaces. (Projecting a cubic surface from one of its points, the branch curve is a plane quartic. Equivalently, blowing up the cubic at a point we get a degree 2 Del Pezzo surface.) This is, however, not the natural thing to do from the modern viewpont. Most of (6.3) is proved in \cite{Comessatti13}. \begin{thm} \label{deg2.equiv.thm} Let $D\subset \p^2$ be a degree 4 smooth real curve. The following are equivalent: \begin{enumerate} \item All 28 bitangents of $D$ are real. \item All 56 lines of $F^-_D$ are real. \item $F^-_D$ is isomorphic to $\p^2$ blown up in 7 real points. \item $F^-_D(\r)\sim \#8\r\p^2$. \item $D(\r)\sim \uplus 4S^1$. \item $F^+_D(\r)\sim \uplus 4S^2$. \item $F^+_D$ has Picard number 1 over $\r$. \end{enumerate} \end{thm} Proof. (1) $\Rightarrow$ (2): A neighborhood of a line in $\r\p^2$ is not orientable, thus any bitangent is contained in $U_D$ (except for the points of tangency). $f$ is negative on any bitangent and so $u^2=-f$ has real solutions, giving 56 real lines on $F^-_D$. (2) $\Rightarrow$ (3): Over $\c$, $F^-_D$ is the blow up of $\p^2$ at 7 points, hence it has 7 disjoint lines. If all lines are real, we have 7 disjont real lines. Contracting these we get a Del Pezzo surface of degree 9 over $\r$. By (\ref{mmp.over.R}.D) it is $\p^2_{\r}$. (3) $\Rightarrow$ (4): Topologically, each blowing up is connected sum with $\r\p^2$. (4) $\Rightarrow$ (5): This follows from (\ref{deg2.dp.top}). (5) $\Rightarrow$ (6): This also follows from (\ref{deg2.dp.top}). (6) $\Rightarrow$ (7): Assume to the contrary that $F^+_D$ has Picard number $\geq 2$ over $\r$. By (\ref{mmp.over.R}) we have one of 2 cases: \begin{enumerate} \item $F^+_D$ is a minimal conic bundle with 6 singular fibers. In this case $F^+_D(\r)\sim \uplus 3S^2$, a contradiction. \item $F^+_D$ is the blow up of a Del Pezzo surface of degree 3 or 4 over $\r$. By (\ref{d=4,3.dp.top}) $F^+_D(\r)$ has at most 2 connected components, a contradiction. \end{enumerate} (7) $\Rightarrow$ (1): Assume that $D$ has a complex bitangent $L$. Its conjugate $\bar L$ is again a bitangent. Let $C\subset F^+_D$ be a complex line over $L$. Its conjugate $\bar C$ lies over $\bar L$. Then $(C\cdot \bar C)\leq 1$ (the only possible intersection point lies over $L\cap \bar L$). Thus $F^+_D$ has either a disjoint pair of conjugate lines or a conic bundle structure, a contradiction.\qed \begin{prop}\label{deg2.dp.top.conn}\cite{Klein1876} The space of degree 4 smooth real plane curves has 6 connected components corresponding to the 6 topological types in (\ref{deg2.dp.top}). The space of degree 2 real Del Pezzo surfaces has 12 connected components corresponding to the 12 topological types in (\ref{deg2.dp.top}). \end{prop} Proof. The two parts are equivalent and it is sufficient to treat the $F^-_D$ cases. $D$ always has a real bitangent (\ref{deg.4.bitang}), thus $F^-_D$ contains a real line. So $F^-_D$ is obtained by blowing up a degree 3 Del Pezzo surface at a real point. (\ref{deg2.dp.top.conn}) now follows from (\ref{d=4,3.dp.top}).\qed \begin{exrc}\cite{Zeuthen1874}\label{deg.4.bitang} Let $C$ be a smooth real plane curve of degree 4. Let $d$ be the number of ovals of $D(\r)$ which are not contained in another oval. Show that $C$ has $4+2d(d-1)$ real bitangents. \end{exrc} \begin{say}[Degree 1 Del Pezzo surfaces]\label{deg1.dp.say}{\ } Let $F$ be a degree 1 Del Pezzo surface over any field $k$. $|-K_F|$ is a pencil with exactly one base point. So this is a $k$-point and $F(k)\neq \emptyset$. $|-2K_F|$ is base point free and exhibits $F$ as a double cover of a quadric cone $Q\subset \p^3$, ramified along a curve $D\subset Q$ which is a complete intersection of $Q$ with a cubic surface with equation $(f=0)$. $D$ does not pass through the vertex of the cone. $F_{\bar k}$ contains 240 lines (that is $-1$-curves); cf. \cite[IV.4.3]{Manin72}. We obtain these as follows. Take a plane $H\subset \p^3$ which is tangent to $D$ at 3 points. The preimage of $H\cap Q$ in $F$ has 2 irreducible components, each is a line. Thus we conclude that there are 120 planes which are tangent to $D$ at 3 points. Assume now that $k=\r$. Since $Q(\r)\neq \emptyset$, we can write $Q=(x^2+y^2=1)$ in suitable affine coodinates $(x,y,z)$ on ${\Bbb A}^3$. That is, $Q(\r)$ is a cylinder with a singular point at infinity. As in the degree 2 case, for each (nonhomogeneous) cubic $f(x,y,z)$ we obtain two degree 1 Del Pezzo surfaces, given by affine equations $$ F^{\pm}_f:=(x^2+y^2-1=u^2\mp f(x,y,z)=0)\subset {\Bbb A}^4. $$ There are 2 types of simple closed loops on a cylinder: null homotopic ones (I call them ovals) and those homotopic to a plane section (I call them big circles). Since $D(\r)$ is the intersection of a cylinder with a cubic, it has 3 or 1 intersection points with any ruling line of the cylinder. Thus $D(\r)$ contains either 3 big circles (and no ovals) or 1 big circle. $D$ has genus 4, hence by Harnack's theorem (cf. \cite[VII.4]{Shafarevich72}), $D(\r)$ has at most 5 connected components. An oval can not be inside another oval since this would give 4 points on a ruling. Furthermore, we can not have an oval on one side the big circle and at least two ovals on the other side. Indeed, choosing points $P_1,P_2,P_3$ inside the 3 ovals, there is a plane $H$ through them. Then $H$ intersects each oval in at least 2 points, and also the big circle. So $(H\cdot D)\geq 8$, but $D$ has degree 6, a contradiction. If all the ovals are on the same side of the big circle, we can normalize $f$ so that it is positive on the other side. The other cases are symmetrical and it makes little sense to normalize $f$. We can summarize these results in the following table: \end{say} \begin{prop}\label{deg1.dp.top} There are 7 topological types of degree 6 smooth real complete intersection curves on the cylinder $(x^2+y^2=1)$, not passing through the point at infinity. Correspondingly there are 11 topological types of degree 1 real Del Pezzo surfaces. The following table gives the complete list. The types in the same row correspond to each other under $D=(f=0)\cap Q \leftrightarrow F^+_f\leftrightarrow F^-_f$. $$ \begin{tabular}{|c|c|c|} \hline $D(\r)$ & $F^+_f(\r)$ & $F^-_f(\r)$\\ \hline 1 big circle + 4 ovals & $\r\p^2\uplus 4S^2$ & $ \#9\r\p^2$\\ \hline 1 big circle + 3 ovals & $\r\p^2\uplus 3S^2$ & $ \#7\r\p^2$\\ \hline 1 big circle + 2 ovals & $\r\p^2\uplus 2S^2$ & $ \#5\r\p^2$\\ \hline 1 big circle + 1 oval & $\r\p^2\uplus S^2$ & $ \#3\r\p^2$\\ \hline 1 big circle + 0 oval & $\r\p^2$ & $\r\p^2$\\ \hline 1 big circle + 1+1 ovals & $ \#3\r\p^2\uplus S^2$ & $ \#3\r\p^2\uplus S^2$\\ \hline 3 big circles & $\r\p^2\uplus \#2\r\p^2$ & $\r\p^2\uplus \#2\r\p^2$\\ \hline \end{tabular} $$ \end{prop} The following theorem, due to \cite{Comessatti13}, is the degree 1 version of (\ref{deg2.equiv.thm}). I thank F. Russo for checking the arguments of Comessatti. \begin{thm} \label{deg1.equiv.thm} Let $D=(f=0)\subset Q$ be a degree 6 smooth real complete intersection curve on the cylinder $Q=(x^2+y^2=1)$. The following are equivalent: \begin{enumerate} \item All 120 triple tangents of $D$ are real and $f$ is negative on all of them. \item All 240 lines of $F^-_f$ are real. \item $F^-_f$ is isomorphic to $\p^2_{\r}$ blown up in 8 real points. \item $F^-_f(\r)\sim \#9\r\p^2$. \item $D(\r)\sim \uplus 5S^1$. \item $F^+_f(\r)\sim \r\p^2\uplus 4S^2$. \item $F^+_f$ has Picard number 1 over $\r$. \end{enumerate} \end{thm} Proof. (1) $\Rightarrow$ (2): If $f$ is negative on a triple tangent then $u^2=-f$ has real solutions, giving a pair of real lines on $F^-_f$. (2) $\Rightarrow$ (3): Over $\c$, $F^-_f$ is the blow up of $\p^2$ at 8 points. Thus it has 8 disjoint lines. If all lines are real, we have 8 disjont real lines. Contracting these we get a Del Pezzo surface of degree 9 over $\r$. By (\ref{mmp.over.R}.D) it is $\p^2_{\r}$. (3) $\Rightarrow$ (4): Topologically, each blowing up is connected sum with $\r\p^2$. (4) $\Rightarrow$ (5): This follows from (\ref{deg1.dp.top}). (5) $\Rightarrow$ (6): This also follows from (\ref{deg1.dp.top}). (6) $\Rightarrow$ (7): Assume to the contrary that $F^+_f$ has Picard number $\geq 2$ over $\r$. $F^+_f$ can not be a minimal conic bundle since $(K^2)$ is odd. Thus $F^+_f$ is the blow up of Del Pezzo surface of degree 2 or 3 over $\r$. By (\ref{deg2.dp.top}, \ref{d=4,3.dp.top}) $F^+_f(\r)$ has at most 4 connected components, a contradiction. (7) $\Rightarrow$ (1): If $D$ has a complex triple tangent, we can argue as in (\ref{deg2.equiv.thm}.(7) $\Rightarrow$ (1)) that $F^+_f$ contains a conjugate pair of lines $C,\bar C$ such that $(C\cdot \bar C)\leq 2$. $(C+\bar C)\equiv rK$ for some $r\in \z$, thus $2(C\cdot \bar C)-2=r^2$. This is impossible. If there is a real triple tangent such that $f$ is positive on it then as in (1) $\Rightarrow$ (2) we get real lines on $F^+_f$. \qed \begin{prop}\label{deg1.dp.top.conn} The space of degree 1 real Del Pezzo surfaces has 11 connected components corresponding to the 11 topological types in (\ref{deg1.dp.top}). \end{prop} Proof. For the nonminimal ones, this follows from (\ref{deg1.dp.top.conn}) and (\ref{d1=d2.blowup}). The minimal ones are in one--to-one correspondence with the blow ups of $\p^2$ at 8 real points, this is again a connected space.\qed \begin{ack} I thank my audience at the Trento sumer school, especially L. Bonavero, S. Cynk and S. Endrass for numerous comments and improvements. A. Marin directed me to several 19th century references. F. Russo checked the arguments of Comessatti about degree 1 and 2 Del Pezzo surfaces and pointed out some misunderstandings on my part. Partial financial support was provided by the NSF under grant number DMS-9622394. \end{ack} \section{Hints to selected exercises} (\ref{er.noncontr.ex}) Blow up $\geq 10$ general points (over $\c$) on a smooth plane cubic. The birational transform of the cubic generates an extremal ray which is not contractible. To see this show that $\pic$ of the blown up surface injects into $\pic$ of the cubic. \medskip (\ref{int.cone.exrc}) Write down Riemann--Roch for $mz$ and $-mz$ to get the first part. Then use this for $z+\epsilon z'$ for any $z'\in N_1(F)$. (cf. \cite[V.1.8]{Hartshorne77}.) \medskip (\ref{conics.r}) $|-K|$ embeds $C$ as a conic. \medskip (\ref{projspace.ex}) Let $H\in X_{\c}$ be a hyperplane. Show that $H\cap \bar H$ is real and use induction. Even degree symmetric powers of the empty conic give examples. \medskip (\ref{coh.b.c}) This is a special case of \cite[III.9.3]{Hartshorne77}. \medskip (\ref{pic=gal.inv}) The empty conic gives a good example. For the rest, the key point is to understand that we know more than the existence of an isomorphism $\tau:L_{\c}\cong \bar L_{\c}$. Namely, by conjugation this induces $\bar\tau:\bar L_{\c}\cong \bar{\bar L_{\c}}\cong L_{\c}$, and the composite of these two gives the identity of $L_{\c}$ (and not just an isomorphism of $L_{\c}$ to itself). Thus we have to choose a specific isomorphism $\tau:M\cong \bar M$. If $X$ has a real point $P$, then on the fiber over $P$ we can choose $\tau$ to be conjugation (and not just some constant times conjugation). Once things are set up right, the real sections of $M$ are those sections $s$ of $M$ such that $\tau(s)=\bar s$. \medskip (\ref{comp.bir.inv}) Use the fact that a birational map between projective varieties is defined outside a codimension 2 subset. \medskip (\ref{gen.orient.ex}) Let $S^1\sim A\subset X(\r)$ be any loop. Perturb it to achieve that $A$ intersects $D(\r)$ transversally at smooth points and is disjoit from $D'(\r)$. \medskip (\ref{cb.isom.ex}) $z$ is transformed by the inverse of the matrix. Then do the explicit computation. \medskip (\ref{geom.cb.ex}) Let $F\to \p^1$ be a minimal conic bundle. There are 2 types of elementary transformations: blow up a real point in a fiber and then contract the fiber, or blow up conjugate points in conjugate fibers and then contract the fibers. Pick any section $C$ over $\c$. Using elementary transformations get to the situation when $C$ and $\bar C$ are disjoint. Show that $(C^2)=(\bar C^2)=-m$ if there are $2m$ singular fibers. The normal form is an affine piece of representing $F$ as a conic bundle in $\proj f_*\o_F(C+\bar C)$. \medskip (\ref{(2,2)-curves}) The case when $D(\r)=\emptyset$ is easy. There are many ways to study the remaining cases. For instance, pick a point $P\in D(\r)$ and blow it up. By contracting the birational transforms of the two sections through $P$, we obtain a correspondance between pairs $(P\in D(\r))$ and triplets $(Q_1,Q_2\in E(\r))$ where $E\subset \p^2$ is an elliptic curve. One has to be a little careful since $\r\p^2$ is not orientable. If we fix the orientations of the two copies of $\r\p^1$ in $\r\p^1\times \r\p^1$ then they give local orientations of $\r\p^2$ at the points $Q_1,Q_2$. There is an ambiguity of changing both orientations (since this does not change the orientation of $\r\p^1\times \r\p^1$). We have to study various cases according to the location of $Q_1,Q_2$ on $E(\r)$. Moreover, we have to see how the local orientations match up if we move from $Q_1$ to $Q_2$ along $E(\r)$. If $Q_1,Q_2$ are both on a pseudo line, then the two local orientations are consistent if we move in one direction and inconsistent in the other direction. However, if $Q_1,Q_2$ are both on an oval, then the two local orientations are either consistent in both directions or inconsistent if both directions. This gives 2 cases. \medskip (\ref{deg4.cb.ex}) Given a ramification curve $D\subset \p^1\times \p^1$ we get 2 different degree 4 Del Pezzo surfaces. We want one surface $S^+$ where none of the lines are real. Then, in the other surface $S^-$, all lines are real. One can construct $S^-$ as follows. Start with $\p^1\times \p^1$, points $P_1,\dots,P_4$ in the first factor and $P'_1,\dots,P'_4$ in the second factor. Blow up the 4 points $(P_i,P'_i)$. Show that we get a degree 4 Del Pezzo surface iff there is no isomorphism $\tau:\p^1\times \p^1$ such that $\tau(P_i)=P'_i$. Projecting to the first factor gives a conic fibration with singular fibers over $P_i$. Projection to the second factor is not the right thing to do. Instead, the second conic fibration is given by the linear system of curves of type $(2,2)$ passing through the 4 points $(P_i,P'_i)$. Also keep in mind that we have to take care not only of the 4 points but also the set $I(f)$. \medskip (\ref{cubic.lines.ex}) and (\ref{deg.4.bitang}) can both be seen from the classification. One has to prove that we can not blow up a point on a line. \cite{Schlafli1863} proved first that a cubic can be written as $C_1-C_2=0$ where $C_i$ are products of linear factors and then studied the various cases when the linear factors are all real or some are conjugates. \cite{Zeuthen1874} notes that 2 ovals have 4 tangents, thus we need to show that there are 4 more which are either tangents to the same oval or at complex points. He does this by a continuity argument. This is a bit tricky since these 4 tangents are not invariant under deformations of the curve, just their number is. \medskip
1997-12-15T00:53:17
9712
alg-geom/9712013
en
https://arxiv.org/abs/alg-geom/9712013
[ "alg-geom", "math.AG" ]
alg-geom/9712013
Christopher Thomas Woodward
Sharad Agnihotri (Amsterdam) and Chris Woodward (Harvard)
Eigenvalues of products of unitary matrices and quantum Schubert calculus
18 pages, uses amssymb
null
null
null
null
We describe the inequalities on the possible eigenvalues of products of unitary matrices in terms of quantum Schubert calculus. Related problems are the existence of flat connections on the punctured two-sphere with prescribed holonomies, and the decomposition of fusion product of representations of SU(n), in the large level limit. In the second part of the paper we investigate how various aspects of the problem (symmetry, factorization) relate to properties of the Gromov-Witten invariants.
[ { "version": "v1", "created": "Sun, 14 Dec 1997 23:53:17 GMT" } ]
2016-08-30T00:00:00
[ [ "Agnihotri", "Sharad", "", "Amsterdam" ], [ "Woodward", "Chris", "", "Harvard" ] ]
alg-geom
\section{Introduction} Beginning with Weyl \cite{we:ei}, many mathematicians have been interested in the following question: given the eigenvalues of two Hermitian matrices, what are the possible eigenvalues of their sum? In a recent preprint \cite{kl:sb}, Klyachko observes that a complete solution to this problem is given by an application of Mumford's criterion in geometric invariant theory. The eigenvalue inequalities are derived from products in Schubert calculus. In particular, Weyl's inequalities correspond to Schubert calculus in projective space. The necessity of these conditions is due to Helmke and Rosenthal \cite{hr:ei}. One of the fascinating points about the above problem are several equivalent formulations noted by Klyachko. For instance, the problem is related to the following question in representation theory: Given a collection of irreducible representations of $SU(n)$, which irreducibles appear in the tensor product? A second equivalent problem involves toric vector bundles over the complex projective plane. In this paper we investigate the corresponding problem for {\em products} of {\em unitary} matrices. This question also has a relationship with a representation-theoretic problem, that of the decomposition of the fusion product of representations. The solution to the multiplicative problem is also derived from geometric invariant theory, namely from the Mehta-Seshadri theory of parabolic bundles over the projective line. The main result of this paper, Theorem \ref{final}, shows that the eigenvalue inequalities are derived from products in {\em quantum} Schubert calculus. This improves a result of I. Biswas \cite{bi:ex}, who gave the first description of these inequalities. A similar result has been obtained independently by P. Belkale \cite{bl:ip}. The proof is an application of the Mehta-Seshadri theorem. A set of unitary matrices $A_1,\ldots ,A_l$ such that each $A_i$ lies in a conjugacy class $\mathcal{C}_i$ and such that their product is the identity is equivalent to a unitary representation of the fundamental group of the $l$ times punctured sphere, with each generator $\gamma_i$ being mapped to the conjugacy class $\mathcal{C}_i.$ By the Mehta-Seshadri theorem such a representation exists if and only if there exists a semi-stable parabolic bundle on $\P^1$ with $l$ parabolic points whose parabolic weights come from the choice of conjugacy classes $\mathcal{C}_i.$ This last interpretation of the original eigenvalue problem can be related to the Gromov-Witten invariants of the Grassmannian and this is done in Section 5 below. In Sections 6 and 7 we investigate how factorization and hidden symmetries of these Gromov-Witten invariants relate to the multiplicative eigenvalue problem. \section{Additive inequalities (after Klyachko and Helmke-Rosenthal)} Let $\lie{su}(n)$ denote the Lie algebra of $SU(n)$, and $$\t = \{ (\lambda_1,\ldots,\lambda_n) \in \mathbb{R}^n \ | \ \sum \lambda_i = 0 \} $$ its Cartan subalgebra. Let $$ \t_+ = \{ (\lambda_1,\ldots,\lambda_n) \in \t \ | \ \lam_i \ge \lam_{i+1}, \ \ i=1,\ldots,n-1 \} $$ be a choice of closed positive Weyl chamber. For any matrix $A \in \lie{su}(n)$ let $$ \lam(A) = (\lam_1(A),\lam_2(A),\cdots,\lam_n(A)) \in \t_+ $$ be the eigenvalues of the Hermitian matrix $-i A$ in non-increasing order. Let $\Delta(l) \subset (\t_+)^l$ denote the set $$ \Delta(l) = \{ (\lam(A_1),\lam(A_2),\ldots,\lam(A_l)) \ \vert \ A_1,\ldots,A_l \in \lie{su}(n), \ A_1 + A_2 + \ldots + A_l = 0 \} .$$ Define an involution $$ *: \ \t_+ \cong \t_+, \ \ (\lambda_1,\ldots,\lambda_n) \mapsto (-\lambda_n, \ldots, -\lambda_1) .$$ For any $A \in \lie{su}(n)$ the matrix $-A$ has eigenvalues $ \lam(-A) = *\lam(A)$. The set $\Delta(l)$ is invariant under the map $$ *^l:\ (\t_+)^l \to (\t_+)^l, \ \ (\xi_1,\ldots,\xi_l) \mapsto (*\xi_1,\ldots,*\xi_l) $$ and also under the action of the symmetric group $S_l$ on $(\t_+)^l$. The set $\Delta(l)$ has interesting interpretations in symplectic geometry and representation theory. Consider the cotangent bundle $ T^*SU(n)^{l-1} $ with the action of $SU(n)^l$ given by $SU(n)$ acting diagonally on the left and $SU(n)^{l-1}$ on the right. The moment polytope of this action may be identified with $\Delta(l)$ (see Section 5.) From convexity theorems in symplectic geometry (see e.g. \cite{sj:co} and \cite{le:co}) it follows that $\Delta(l)$ is a finitely-generated convex polyhedral cone. In particular there are a finite number of inequalities defining $\Delta(l) $ as a subset of the polyhedral cone $(\t_+)^l$. The set $\Delta(l) $ may also be described in terms of the tensor product of representations. Let $$(\ ,\ ) : \ \lie{su}(n) \times \lie{su}(n) \to \mathbb{R}, \ \ \ (A,B) \mapsto - \Tr(AB) $$ denote the basic inner product on $\lie{su}(n)$, which induces an identification $\lie{su}(n) \cong \lie{su}(n)^*$. Let $ \Lambda = \mathbb{Z}^n \cap \t$ denote the integral lattice and $ \Lambda^* \subset \t$ its dual, the weight lattice. For each $\lam \in \Lambda^*\cap \t_+$, let $V_\lam$ denote the corresponding irreducible representation of $SU(n)$. We will see in equation \eqref{inv_eqn} that $\Delta(l) \cap \mathbb{Q}^l$ is the set of $(\lam^1,\ldots,\lam^l)$ such that for some $N$ such that $N\lam^1,\ldots, N \lam^l \in \Lambda^*$, we have $$V_{N \lam^1} \otimes \ldots \otimes V_{N\lam^{l-1}} \supset V^*_{N \lam^l},$$ that is, $V_{N \lam^1} \otimes \ldots \otimes V_{N \lam^l}$ contains a non-zero invariant vector. The work of Klyachko and Helmke-Rosenthal gives a complete set of inequalities describing $\Delta(l)$ in terms of Schubert calculus. Let $$ \mathbb{C}^n = F_n \supset F_{n-1} \supset \ldots \supset F_0 = \{ 0 \} $$ be a complete flag in $\mathbb{C}^n$, $G(r,n)$ the Grassmanian of $r$-planes in $\mathbb{C}^n$, and for any subset $I = \{i_1,\ldots,i_r \} \subset \{1, \ldots n \}$ let $$ \sig_I = \{ W \in G(r,n) \ \vert \ \dim(W \cap F_{i_j}) \ge j, \ j = 1,\ldots,r \} $$ denote the corresponding Schubert variety. The Schubert cell $C_I \subset \sigma_I$ is defined as the complement of all lower-dimensional Schubert varieties contained in $\sig_I$: $$ C_I = \sigma_I \backslash \bigcup_{\sigma_J \subset \sigma_I} \sigma_J. $$ We say that $W$ is in {\em position $I$} with respect to the flag $F_*$ if $W \in C_I$. The homology classes $[ \sig_I ]$ form a basis of $H_*(G(r,n),\mathbb{Z})$. Given two Schubert cycles $\sig_I,\sig_J$, we can expand the intersection product $[\sig_I] \cap [\sig_J]$ in terms of this basis. We say $[\sig_I] \cap [\sig_J]$ contains $[\sig_K]$ if $[\sig_K]$ appears in this expansion with non-zero (and therefore positive) coefficient. Equivalently, let $$*K = \{n + 1 - i_r,n+ 1 - i_{r-1},\ldots, n+1-i_1 \},$$ so that $[\sig_{*K}]$ is the Poincare dual of $[\sig_K]$. Then $[\sig_I] \cap [\sig_J]$ contains $[\sig_K]$ if and only if the intersection of general translates of the Schubert cycles $\sig_I,\sig_J,\sig_{*K}$ is non-empty and finite. \begin{theorem}[Klyachko, resp. Helmke-Rosenthal] \labell{klyachko} A complete (resp. necessary) set of inequalities describing $\Delta(l)$ as a subset of $(\t_+)^l$ are \begin{equation} \labell{add_ineq} \sum_{i \in I_1} \lam_i(A_1) + \sum_{i \in I_2} \lam_i(A_2) + \ldots + \sum_{i \in I_l} \lam_i(A_l) \leq 0, \end{equation} where $I_1,\ldots,I_l$ are subsets of $\{ 1 ,\ldots, n \}$ of the same cardinality $r$ such that $[\sig_{I_1}] \cap \ldots \cap [\sig_{I_{l-1}}] \supset [\sig_{*I_l}]$, and $r$ ranges over all values between $1$ and $n-1$. \end{theorem} Note that the cases $l =1,2$ are trivial: $\Delta(1) = \{ 0 \},$ and $ \Delta(1) = \{ (\mu,*\mu) \ |\ \mu \in \t_+ \}.$ Klyachko also claims that these inequalities are independent. From Theorem \ref{klyachko} follows a complete set of inequalities for the possible eigenvalues of a sum of skew-Hermitian matrices. For instance, for $l=3$ one obtains the inequalities \begin{equation} \labell{add_ineq2} \sum_{i \in I} \lam_i(A) + \sum_{j \in J} \lam_j(B) \le \sum_{k \in K} \lam_k(A + B), \end{equation} where $I,J,K \subset \{ 1 ,\ldots, n \}$ range over subsets such that $[\sig_I] \cap [\sig_J]$ contains $[\sig_K]$. \begin{example} Let $r=1$ so that $G(r,n) \cong \P^{n-1}$ and $I = \{ n - i + 1 \}, \ J = \{ n - j + 1 \}$. Then $\sigma_I \cong \P^{n-i},\ \sigma_J \cong \P^{n-j}$ so that $[\sigma_I] \cap [\sigma_J] \cong \P^{n-i-j+1} = \sigma_K$ where $K = \{n - i -j + 2\}$. One obtains \begin{equation} \labell{dual_Weyl} \lam_{n-i+1}(A) + \lam_{n-j+1}(B) \le \lam_{n-i-j+2}(A + B) . \end{equation} \end{example} \subsection{Duality} Let $A_1,\ldots, A_l \in \lie{su}(n)$. From \eqref{add_ineq2} applied to $-A_1,\ldots ,- A_l$ one obtains \begin{equation} \labell{dual_ineq} - \sum_{i \in *I_{1}} \lam_i(A_1) -\ldots - \sum_{i \in *I_{l}} \lam_i(A_l)\leq 0 \end{equation} or equivalently $ \sum_{i \in *I_{1}} \lam_i(A_1) +\ldots + \sum_{i \in *I_{l}} \lam_i(A_l)\geq 0. $ By the trace condition, \eqref{dual_ineq} is equivalent to $$ \sum_{i \notin *I_{1}} \lam_i(A_1) +\ldots + \sum_{i \notin *I_{l}} \lam_i(A_l)\leq 0 .$$ Let $I_i^c = \{ 1,\ldots,n \} \backslash * I_i$. Then $[\sigma_{I_i^c}]$ is the image of $[\sigma_{I_{i}}]$ under the isomorphism of homology induced by $G(r,n) \cong G(n-r,n)$ (see page 197 onwards of Griffiths and Harris \cite{gr:pr}). Thus the appearance of \eqref{dual_ineq} in \eqref{add_ineq} corresponds to a product in the Schubert calculus of $G(n-r,n)$. \begin{example} \labell{Weyl_example} The dual equation to \eqref{dual_Weyl} is Weyl's 1912 \cite{we:ei} inequality \begin{equation} \labell{Weyl} \lam_i(A) + \lam_j(B) \ge \lam_{i + j -1}(A + B). \end{equation} \end{example} \section{Multiplicative Inequalities} Let $\lie{A} \subset \t_+$ be the fundamental alcove of $SU(n)$: $$\lie{A} = \{ \lam \in \t_+ \ | \ \lam_1 - \lam_n \leq 1 \}.$$ Let $A \in SU(n)$ be a unitary matrix with determinant $1$. Its eigenvalues may be written $$ e^{2\pi i \lam_1(A)}, e^{2\pi i \lam_2(A)}, \ldots ,e^{2\pi i \lam_n(A)} $$ where $\lam(A) = (\lam_1(A),\ldots,\lam_n(A)) \in \lie{A}$. The map $A \mapsto \lam(A)$ induces a homeomorphism $$ \lie{A} \cong SU(n) / \Ad(SU(n)) .$$ Let $\Delta_q(l) \subset \lie{A}^l$ ($q$ for quantum) denote the set $$ \Delta_q(l) = \{ (\lam(A_1),\ldots,\lam(A_l)) \ \vert \ A_1,\ldots,A_l \in SU(n), \ A_1A_2\ldots A_l = I \}. $$ As before, $\Delta_q(l)$ is invariant under the involution, $*^l: \lie{A}^l \to \lie{A}^l$, and the action of the symmetric group $S_l$ on $\lie{A}^l$. The set $\Delta_q(l)$ has an interpretation as a moment polytope. Let $\M$ be the space of flat $SU(n)$-connections on the trivial $SU(n)$ bundle over the $l$-holed two-sphere, modulo gauge transformations which are the identity on the boundary (see \cite{me:lo}). The gauge group of the boundary acts on $\M$ in Hamiltonian fashion and the set $\Delta_q(l)$ is the moment polytope for this action. By \cite[Theorem 3.19]{me:lo}, $\Delta_q(l)$ is a convex polytope. In fact, an analogous statement holds for arbitrary compact, simply-connected Lie groups. In particular, a finite number of inequalities describe $\Delta_q(l)$. In the case $n=2$, these inequalities were given explicitly for $l=3$ in Jeffrey-Weitsman \cite{jw:bs} and for arbitrary numbers of marked points in Biswas \cite{bi:r2}. A description of the inequalities in the arbitrary rank case was given in \cite{bi:ex} but the description given there does not seem to be computable. There is also an interpretation of $\Delta_q(l)$ in terms of fusion product. Let $\fus_N$ denote the fusion product on the Verlinde algebra $R(SU(n)_N)$ of $SU(n)$ at level $N$. Then $\Delta_q(l) \cap \mathbb{Q}^l$ is the set of $(\lam^1,\ldots,\lam^l) \in \lie{A} \cap \mathbb{Q}^l$ such that for some $N$ such that $N\lam^1,\ldots, N \lam^l \in \Lambda^*$, we have \begin{equation} \labell{fus} V_{N \lam^1} \fus_N \ldots \fus_N V_{N\lam^{l-1}} \supset V_{N * \lam^l}.\end{equation} See Section \ref{Verlinde}. \subsection{Quantum Schubert calculus} Quantum cohomology is a deformation of the ordinary cohomology ring that was introduced by the physicists Vafa and Witten. Quantum cohomology of the Grassmannian (quantum Schubert calculus) was put on a rigorous footing by Bertram \cite{be:qs}. Recall that the degree of a holomorphic map $\varphi: \ \P^1 \rightarrow G(r,n)$ is the homology class $[\varphi] \in H^2(G(r,n),\mathbb{Z}) \cong \mathbb{Z}$. Let $p_1,\ldots,p_l$ be distinct marked points in $\P^1$. The quantum intersection product $\star$ on $H_*(G(r,n),\mathbb{C}) \otimes \mathbb{C}[q]$ is defined by $$ [\sigma_{I_1} ] \star \ldots \star [\sigma_{I_l}] = \sum_{J} \langle[\sig_{I_1}],\ldots, [\sigma_{I_l}], [\sig_J] \rangle_d \ [\sigma_{*J}] q^d,$$ where the Gromov-Witten invariant $\langle[\sig_{I_1}],\ldots, [\sigma_{I_l}], [\sig_J] \rangle_d$ is equal to the number of holomorphic maps $\P^1 \mapsto G(r,n)$ sending $p_1,\ldots,p_l,p$ to general translates of $\sigma_{I_1},\ldots,\sigma_{I_l},\sigma_J$ if this number is finite, and is otherwise zero. Our main result is the following description of $\Delta_q(l)$: \begin{theorem}\labell{final} A complete set of inequalities for $\Delta_q(l)$ are given by \begin{equation} \sum_{i \in I_1} \lam_i(A_1) + \sum_{i \in I_2} \lam_i(A_2) + \ldots + \sum_{i \in I_l} \lam_i(A_l) \leq d \end{equation} for $(I_1,\ldots,I_l,d)$ such that $\l [\sig_{I_1}] \star \ldots \star [ \sig_{I_l} ] \r_d \neq 0,$ that is, $\ [\sig_{I_1}] \star \ldots \star [ \sig_{I_{l-1}}] \supset q^d [\sig_{*I_l}] .$ \end{theorem} In the last few years several techniques have been developed for computing the coefficients of quantum Schubert calculus. See for instance Bertram, Ciocan-Fontanine, Fulton \cite{be:qm}. Therefore the above theorem makes the question of which inequalities occur computable in practice. One recovers the inequalities for $\Delta(l)$ from the degree $0$ Gromov-Witten invariants. This shows that $\Delta(l) $ is the cone on $\Delta_q(l)$ at the $0$-vertex, i.e. $$ \Delta(l) = \mathbb{R}_+ \cdot \Delta_q(l) $$ This may be verified by several alternative methods, e.g. Remark \ref{small}. The simplest example of a positive degree inequality is given by the following: \begin{example} \labell{quantum_Weyl_example} Let $r = 1$ so that $G(r,n) = \P^{n-1}$, and $U,V,W \subset \mathbb{C}^n$ be subspaces in general position of dimensions $i,j,n+1-i-j$. There is a unique degree $1$ map $\P^1 \rightarrow \P^{n-1}$ mapping $p_1,p_2,p_3$ to $\P(U),\P(V),\P(W)$ respectively. Together with the degree $0$ inequality mentioned before, this gives \begin{equation} \labell{quant_Weyl} \lam_{i+j-1}(AB) \le \lam_i(A) + \lam_j(B) \le \lam_{i+j}(AB) + 1. \end{equation} We will see in Section \ref{symmetry} that these inequalities are related by a symmetry of $\Delta_q(l)$. \end{example} \vskip .1in {\noindent \em Question:} Are the inequalities in Theorem \ref{final} independent? \vskip .1in \section{Moduli of flags and Mumford's criterion} As a warm-up we review some of the ideas involved in Klyachko's proof. For any $\xi \in \t_+$ let $$\O_\xi = SU(n) \cdot \xi = \{ A \in \lie{su}(n) \ | \ \lam(A) = \xi \} $$ denote the corresponding adjoint orbit. Via the identification $\lie{su}(n) \cong \lie{su}(n)^*$, $\O_\xi$ inherits a canonical symplectic structure, called the Kostant-Kirillov-Souriau two-form, and the action of $SU(n)$ on $\O_\xi$ is Hamiltonian with moment map given by inclusion into $\lie{su}(n)$. The diagonal action of $SU(n)$ on $\O_{\xi_1} \times \ldots \times \O_{\xi_l}$ has moment map given by $$ (A_1,\ldots,A_l) \mapsto A_1 + A_2 + \ldots + A_l.$$ The symplectic quotient $\mathcal{N}({\xi_1,\ldots,\xi_l}) = \O_{\xi_1} \times \ldots \times \O_{\xi_l} \qu SU(n)$ is given by $$ \mathcal{N}({\xi_1,\ldots,\xi_l}) = \{(A_1,\ldots,A_l) \in \O_{\xi_1} \ldots \O_{\xi_l} \ | \ A_1 + A_2 + \ldots + A_l = 0 \} / SU(n) .$$ For generic $(\xi_1,\ldots,\xi_l),$ that is, values where the moment map has maximal rank, the quotient $\mathcal{N}({\xi_1,\ldots,\xi_l})$ is a symplectic manifold. The $l$-tuple $(\xi_1,\ldots,\xi_l)$ lies in $\Delta(l) $ if and only if $\mathcal{N}({\xi_1,\ldots,\xi_l})$ is non-empty. The quotients $\mathcal{N}({\xi_1,\ldots,\xi_l})$ may be viewed as symplectic quotients of the cotangent bundle $ T^*SU(n)^{l-1} $. Indeed, the symplectic quotient $$ (T^*SU(n) \times \O_\xi) \qu SU(n) \cong \O_\xi .$$ Therefore, the quotient of $T^*SU(n)^{l-1}$ by the right action of $SU(n)^{l-1}$ and the diagonal left action of $SU(n)$ is $$ (T^*SU(n)^{l-1} \times \O_{\xi_1} \times \ldots \times \O_{\xi_{l}}) \qu SU(n)^{l} \cong \mathcal{N}({\xi_1,\ldots,\xi_l}).$$ It follows that $\Delta(l)$ is the moment polytope of the action of $SU(n)^l$ on $ T^*SU(n)^{l-1}$. One can determine whether $\mathcal{N}({\xi_1,\ldots,\xi_l})$ is empty by computing its symplectic volume. This is given by a formula derived from the Duistermaat-Heckman theorem due to Guillemin-Prato (see \cite{gu:he} or \cite[(4)]{me:wi}). Unfortunately the formula involves cancelations and it is not apparent what the support of the volume function is, or even that the support is a convex polytope. The manifolds $\O_{\xi_i}$ have canonical complex structures (induced by the choice of positive Weyl chamber) and are isomorphic to (possibly partial) flag varieties. Suppose that $\xi_1,\ldots,\xi_l$ lie in the weight lattice $\Lambda^*$, so that there exist pre-quantum line bundles $L_{\xi_i} \to \O_{\xi_i}$; i.e., equivariant line bundles with curvature equal to $2 \pi i$ times the symplectic form. The sections of $L_{\xi_1}\boxtimes \ldots \boxtimes L_{\xi_l}$ define a K\"ahler embedding $$\O_{\xi_1} \times \ldots \times \O_{\xi_l} \to \P(V_{\xi_1}^*) \times \ldots \times \P(V_{\xi_l}^*),$$ where $V_{\xi_1},\ldots,V_{\xi_l}$ are the irreducible representations with highest weights $\xi_1,\ldots,\xi_l$. By an application of a theorem of Kirwan and Kempf-Ness (which holds for arbitrary smooth projective varieties, see \cite[page 109]{ki:coh}) the symplectic quotient is homeomorphic to the geometric invariant theory quotient $$\mathcal{N}({\xi_1,\ldots,\xi_l}) \cong \O_{\xi_1} \times \ldots \times \O_{\xi_l} \qu SL(n,\mathbb{C}).$$ By definition, $\O_{\xi_1} \times \ldots \times \O_{\xi_l} \qu SL(n,\mathbb{C}) $ is the quotient of the set of semi-stable points in $\O_{\xi_1} \times \ldots \times \O_{\xi_l}$ by the action of $SL(n,\mathbb{C})$, where $(F_{\xi_1},\ldots,F_{\xi_l}) \in \O_{\xi_1} \times \ldots \times \O_{\xi_l}$ is called semi-stable if and only if for some $N$ there is an invariant section of $(L_{\xi_1} \boxtimes \ldots \boxtimes L_{\xi_l})^{\otimes N} $ which is non-vanishing at $(F_{\xi_1},\ldots,F_{\xi_l})$. The quotient $\mathcal{N}({\xi_1,\ldots,\xi_l})$ is therefore non-empty if and only if there exists a non-zero $SU(n)$-invariant vector in \begin{equation} \labell{inv_eqn} H^0((L_{\xi_1} \boxtimes \ldots \boxtimes L_{\xi_l})^{\otimes N}) = V_{N \xi_1} \otimes \ldots \otimes V_{N \xi_l}.\end{equation} This explains the representation-theoretic interpretation of $\Delta(l) $ alluded to in the introduction. In order to obtain the inequalities in Theorem \ref{klyachko}, one applies the criterion of Mumford, which says that {\em a point is semi-stable if and only if it is semi-stable for all one-parameter subgroups} \cite[Chapter 2]{mu:ge}, see also \cite[Lemma 8.8]{ki:coh}. Let us assume that the $\xi_i$ are generic. An application of the criterion gives that an $l$-tuple of complete flags $(F_{1},\ldots,F_{l}) \in \O_{\xi_1} \times \ldots \times \O_{\xi_l}$ is semi-stable if and only if for all subspaces $W \subset \mathbb{C}^n$, one has $$ \sum_{i \in I_1} \xi_{1,i} + \ldots + \sum_{i \in I_l} \xi_{l,i} \le 0, $$ where $I_j$ is the position of $W$ with respect to the flag $F_{j}$. The proof similar to that for Grassmannians given in Section 4.4 of \cite{mu:ge}. The set of semi-stable points is dense if non-empty. It follows that $\mathcal{N}({\xi_1,\ldots,\xi_l})$ is non-empty if and only if the above inequality holds for every intersection $\sigma_{I_1} \cap \ldots \cap \sigma_{I_l}$ of Schubert cycles in general position. Any inequality corresponding to a positive dimensional intersection must be redundant. Indeed, since the intersection is a projective variety, it cannot be contained in any of the Schubert cells. The boundary of $\sigma_{I_l}$ consists of Schubert varieties $\sigma_J$ with $J$ such that $j_k \le i_k$ for $k = 1,\ldots,r$, where $i_1,\ldots,i_r$ and $j_1,\ldots, j_r$ are the elements of $I_l$ and $J$ in increasing order. The inequality obtained from an intersection $\sigma_{I_1} \cap \ldots \cap \sigma_{I_{l-1}} \cap \sigma_J \neq \emptyset$ therefore implies the inequality obtained from $\sigma_{I_1} \cap \ldots \cap \sigma_{I_l} \neq \emptyset$. \section{Application of the Mehta-Seshadri theorem} For any $\xi \in \lie{A}$, let $$\mathcal{C}_{\xi} = \{ A \in SU(n) \ | \ \lam(A) = \xi \} $$ denote the corresponding conjugacy class. The mapping $A \mapsto \lam(A)$ induces a homeomorphism $SU(n)/\Ad(SU(n)) \cong \lie{A} $. Let $p_1,\ldots,p_l \in \P^1$ be distinct marked points and $\M(\xi_1,\ldots,\xi_l)$ the moduli space of flat $SU(n)$-connections on $\P^1 \backslash \{p_1,\ldots, p_l \} $ with holonomy around $p_i$ lying in $\mathcal{C}_{\xi_i}$. Since the fundamental group of $\P^1 \backslash \{p_1,\ldots, p_l\}$ has generators the loops $\gamma_1, \ldots, \gamma_l$ around the punctures, with the single relation $\gamma_1 \cdot \ldots \cdot \gamma_l=1$, $$ \M({\xi_1,\ldots,\xi_l}) \cong \{ (A_1,\ldots,A_l) \in \mathcal{C}_{\xi_1} \times \ldots \times \mathcal{C}_{\xi_l} \ | \ A_1A_2 \cdots A_l = I \} / SU(n) .$$ In particular $\M({\xi_1,\ldots,\xi_l})$ is non-empty if and only if $(\xi_1,\ldots,\xi_l) \in \Delta_q(l)$. In theory one can determine if $\M({\xi_1,\ldots,\xi_l})$ is non-empty by computing its symplectic volume by the formulae stated in Witten \cite[(4.11)]{wi:tw}, Szenes \cite{sz:vo}, and \cite[Theorem 5.2]{me:co}. For rational $\xi_1,\ldots,\xi_l$ the space $\M({\xi_1,\ldots,\xi_l})$ has an algebro-geometric description due to Mehta-Seshadri \cite{ms:pb}. Let $C$ be a Riemann surface with marked points $p_1,\ldots,p_l \in C$ and let $\cE \rightarrow C$ be a holomorphic bundle. A parabolic structure without multiplicity on $\cE$ consists of the following data at each marked point $p_i$: a complete ascending flag $$ 0 = \cE_{p_i,0} \subset \cE_{p_i,1} \subset \cE_{p_i,2} \ldots \subset \cE_{p_i,n} = \cE_{p_i} $$ in the fiber $\cE_{p_i}$ and a set of {\em parabolic weights} $$ \lam_{i,1} > \lam_{i,2} > \ldots > \lam_{i,n} $$ satisfying $\lam_{i,1} - \lam_{i,n} \leq 1$. In \cite{ms:pb} the weights are required to lie in the interval $[0,1)$, but the definitions work without this assumption. A parabolic bundle is a holomorphic bundle with a parabolic structure. Recall that the degree $\deg(\cE)$ of $\cE$ is the first Chern class $c_1(\cE) \in H^2(C,\mathbb{Z}) \cong \mathbb{Z}$. The parabolic degree $\pardeg(\cE)$ is defined by $$ \pardeg(\cE) = \deg(\cE) + \sum_{i=1,j=1}^{l,n} \lam_{i,j} .$$ The parabolic slope $\mu(\cE)$ is $$ \mu(\cE) = \frac{\pardeg(\cE)}{\rk(\cE)} .$$ Given a holomorphic sub-bundle $\mathcal{F} \subset \cE$ of rank $r$ one obtains a parabolic structure on $\mathcal{F}$ as follows. An ascending flag in the fiber $\mathcal{F}_{p_i}$ at each marked point $p_i$ is obtained by removing from $$ \mathcal{F}_{p_i} \cap \cE_{p_i,1} \subseteq \mathcal{F}_{p_i} \cap \cE_{p_i,2} \subseteq \ldots \subseteq \mathcal{F}_{p_i} \cap \cE_{p_i,n} = \mathcal{F}_{p_i} $$ those terms for which the inclusion is not strict. The parabolic weights for $\mathcal{F}$ are $$ \mu_{i,j} = \lam_{i,k_j}, $$ where $k_j$ is the minimal index such that $\mathcal{F}_{p_i,j} \subseteq \cE_{p_i,k_j}$. Let $K_i = \{ k_1, \ldots, k_r \}$. The fiber $\mathcal{F}_{p_i}$ may be viewed as a element of the Grassmannian of $r$-planes in $\cE_{p_i}$, and $K$ is the position of $\mathcal{F}_{p_i}$ with respect to the flag $\cE_{p_i,*}$. The parabolic degree of $\mathcal{F}$ is $$ \pardeg(\mathcal{F}) = \deg(\mathcal{F}) + \sum_{i,\ k \in K_i} \lam_{i,k} .$$ A parabolic sub-bundle of $\cE$ is a holomorphic sub-bundle $\mathcal{F} \subset E$ whose parabolic structure is the one induced from the inclusion. A parabolic bundle $\cE \to C$ is called parabolic semi-stable if $\mu(\mathcal{F}) \le \mu(\cE)$ for all parabolic sub-bundles $\mathcal{F} \subset \cE$. There is a natural equivalence relation on parabolic bundles: Two bundles are said to be grade equivalent if the associated graded bundles are isomorphic as parabolic bundles. See \cite{ms:pb} for more details. \begin{theorem}[Mehta-Seshadri] Suppose the parabolic weights $\lam_{i,j}$ are rational and lie in the interval $[0,1)$. Then the moduli space $\M(\lam_1, \ldots,\lam_l)$ of grade equivalence classes of semi-stable parabolic bundles with parabolic weights $\lam_{i,j}$ and parabolic degree $0$ is a normal, projective variety, homeomorphic to the moduli space of flat unitary connections over $C \backslash \{ p_1,\ldots,p_r \}$ such that the holonomy of a small loop around $p_i$ lies in $\mathcal{C}_{\lam_{i}}$. \end{theorem} In fact, the Mehta-Seshadri theorem also holds without the assumption that the parabolic weights lie in $[0,1)$. One can see this either through the theory of elementary transformations, or through the extension of the Mehta-Seshadri theorem to non-zero parabolic degree given in Boden \cite{bo:re}. The explanation using elementary transformations goes as follows. Let $\cQ$ denote the skyscraper sheaf with fiber $\cE_{p_i}/\cE_{p_i,n-1}$ at $p_i$. One has an exact sequence of sheaves $$ 0 \to \cE' \to \cE \to \cQ \to 0 .$$ The kernel $\cE'$ is a sub-sheaf of a locally free sheaf and therefore locally free. Since degree is additive in short exact sequences $\deg(\cE') = \deg(\cE)-1$. One calls the $\cE'$ an elementary transformation of $\cE$ at $p_i$. There is a canonical line $\cE'_{p_i,1}$ in the fiber $\cE'_{p_i}$ which is the kernel of the fiber map $\pi: \cE'_{p_i} \to \cE_{p_i}$. One extends the canonical line to a complete flag by taking $\cE'_{p_i,j} = \pi^{-1}(\cE'_{p_i,j-1})$ for $j>1$. Finally one takes as parabolic weights at $p_i$ the set $ \lam_{i,n} + 1,\lambda_{i,1}, \ldots,\lam_{i,n-1}$. With this parabolic structure the bundle $\cE'$ is parabolic semi-stable of the same parabolic degree as $\cE$. Details, in a slightly different form, can be found in Boden and Yokogawa \cite{bo:ra}. The following is the key lemma in the derivation of Theorem \ref{result} from Mehta-Seshadri. Let $d = \deg(\cE) = -\sum \lam_{i,j}$ denote the degree of any element $\cE \in \M(\lam_1, \ldots,\lam_l)$. \begin{lemma} \labell{key} Suppose that there is some ordinary semi-stable bundle on $C$ of degree $d$. Then the set of equivalence classes of parabolic semi-stable bundles of parabolic degree $0$ whose underlying holomorphic bundle is ordinary semi-stable is Zariski dense in $\M(\lam_1,\ldots,\lam_l)$. \end{lemma} \begin{proof} Recall from the construction of $\M(\lam_1,\ldots,\lam_l)$ in \cite{ms:pb} that for some integer $N$ there exists an $SL(N)$-equivariant bundle $ \tilde{R} \stackrel{\pi}{\to} R $ whose fibers are products of $l$ complete flag varieties, such that the geometric invariant theory quotients of $\tilde{R},R$ are $$ \tilde{R} \qu SL(N) = \M(\lam_1,\ldots,\lam_l) , \ \ \ \ R \qu SL(N) = \M,$$ where $\M$ denotes the moduli space of ordinary semi-stable bundles on $C$ of degree $d$. By definition, $$ \tilde{R} \qu SL(N) = \tilde{R}^{\ss} / SL(N), \ \ \ \ R \qu SL(N) = R^{\ss} / SL(N) $$ where $\tilde{R}^{\ss},R^{\ss}$ denote the Zariski dense set of semi-stable points in $\tilde{R},R$ respectively. The inverse image $\pi^{-1}(R^{\ss}) \cap \tilde{R}^{\ss} / SL(N)$ is therefore dense in $\tilde{R}^{\ss} / SL(N) = \M(\lam_1,\ldots,\lam_l)$. \end{proof} Now we specialize to the case $C = \P^1$ with $l$ marked points $p_1,p_2,\ldots,p_l$. Let $\xi_1,\ldots,\xi_l \in \lie{A}^l \cap \mathbb{Q}^l$. By Lemma \ref{key}, $\M({\xi_1,\ldots,\xi_l})$ is non-empty if and only there exists a parabolic semi-stable $\cE$ with parabolic degree $0$ and weights $\xi_1,\ldots,\xi_l$ whose underlying holomorphic bundle is semi-stable. Since the sum of the parabolic weights is zero, the degree of $\cE$ is also zero. By Grothendieck's theorem, $\cE$ is holomorphically trivial. A sub-bundle $\cF \subset \cE$ of rank $r$ is given by a holomorphic map $$ \varphi_{\cF}: \ \P^1 \to G(r,n).$$ Since $\varphi_{\cF}$ is the classifying map of the quotient $\cE/\cF$, the degree of $\cF$ is minus the degree of $\varphi_{\cF}$. The parabolic slope of $\cF$ is given by $$ \mu(\cF) = -\deg(\varphi_{\cF}) + \sum_{i \in I_1(\varphi)} \xi_{1,i} + \ldots + \sum_{i \in I_1(\varphi)} \xi_{l,i}, $$ where $I_i(\varphi)$ is the position of the subspace $\varphi(p_i) \subset \cE_{p_i}$ with respect to the flag $\cE_{p_i,* }$ above. The parabolic bundle $\cE$ is called parabolic semi-stable if and only if for all such $F$, $ \mu(F) \leq 0 $, that is, $$ \sum_{i \in I_1(\varphi)} \xi_{1,i} + \ldots + \sum_{i \in I_1(\varphi)} \xi_{l,i} \leq \deg(\varphi) $$ for all maps $\varphi:\P^1 \to G(r,n)$. The following result was obtained independently by P. Belkale \cite{bl:ip}. \begin{theorem} \labell{result} A complete set of inequalities for $\Delta_q(l)$ as a subset of $\lie{A}^l$ is given by \begin{equation} \labell{mult_ineq} \sum_{i \in I_1} \lam_i(A_1) + \sum_{i \in I_2} \lam_i(A_2) + \ldots + \sum_{i \in I_l} \lam_i(A_l) \leq d \end{equation} for subsets $I_1,\ldots,I_l \subset \{ 1 ,\ldots, n \}$ of the same cardinality $r$ and non-negative integers $d$ such that there exists a rational map $\P^1 \to G(r,n)$ of degree $d$ mapping $p_1,\ldots,p_l$ to the Schubert cells $C_{I_1},\ldots,C_{I_l}$ in general position. \end{theorem} \begin{proof} If $\M(\xi_1,\ldots,\xi_l)$ is non-empty, then a trivial bundle with a general choice of flags will be parabolic semi-stable. Indeed, by the above discussion the fiber $\operatorname{Flag}^l$ of $\pi :\tilde{R} \to R$ over a trivial bundle intersects $\tilde{R}^{ss}$, so $\tilde{R}^{ss}\cap \operatorname{Flag}^{l}$ is open in $\operatorname{Flag}^{l}$. Therefore, $\M(\xi_1,\ldots,\xi_l)$ is non-empty if and only if $$ \sum_{i \in I_1} \xi_{1,i} + \ldots + \sum_{i \in I_l} \xi_{l,i} \leq d $$ for all subsets $I_1,\ldots,I_l$ and integers $d$ such that there exists a degree $d$ map sending $p_1,\ldots,p_l$ to general translates of the Schubert cells $C_{I_1},\ldots,C_{I_l}$. \end{proof} \begin{remark} \labell{small} For sufficiently small parabolic weights $\lam_{i,j}$ any parabolic semi-stable bundle on $\P^1$ is necessarily ordinary semi-stable of degree $0$, and therefore trivial. It follows that the moduli spaces $\M(\lam_1,\ldots,\lam_l)$ and $\mathcal{N}(\lam_1,\ldots,\lam_l)$ are isomorphic. This shows that Klyachko's result is implied by Theorem \ref{result}. \end{remark} We now show that the existence of the maps described in Theorem \ref{result} may be detected by Gromov-Witten invariants. Let $\sig_{I_1},\ldots,\sig_{I_l}$ be some collection of Schubert varieties, and consider the expansion $$ [\sig_{I_1}] \star [\sig_{I_2}] \ldots \star [\sig_{I_l}] = \sum_i q^i \alpha_i $$ where $\alpha_i \in H_*(G(r,n))$. (Question: is this product always non-zero?) We say that $q^d $ divides $ [\sig_{I_1}] \star [\sig_{I_2}] \ldots \star [\sig_{I_l}]$ if $\alpha_i = 0$ for all $i < d$. The following lemma is stated in Ravi \cite{ra:in}. \begin{lemma} \labell{compute} Let $d$ be the lowest degree of a map $\P^1 \to G(r,n)$ sending $p_1,\ldots,p_l$ to general translates of $\sig_{I_1},\ldots,\sig_{I_l}$ respectively. Then $q^d$ is the maximal power of $q$ dividing $[\sig_{I_1}] \star \ldots \star [\sig_{I_l}]$. \end{lemma} \begin{proof} Let $\M_d$ denote the space of maps $\P^1 \to G(r,n)$ of degree $d$, $\operatorname{ev}^l: \, \M_d \to G(r,n)^l$ the evaluation map, and $\sig_{I_*}(p_*) = (\operatorname{ev}^l)^{-1}( \sigma_{I_*}) $ the subset of maps sending $p_j$ to $\sigma_{I_j}$ for $j=1,\ldots,l$. By \cite[Moving Lemma 2.2A]{be:qs}, $ \sig_{I_*}(p_*) $ is a quasi-projective variety, of the expected codimension in $\M_d$. By choosing enough additional marked points $p_1',\ldots,p_m'$, we can insure that the corresponding evaluation map $\operatorname{ev}^m: \ \M_d \to G(r,n)^m$ is injective when restricted to $\sig_{I_*}(p_*).$ Let $Y \subset G(r,n)^{l}\times G(r,n)^{m}$ be the closure of $(\operatorname{ev}^l \times \operatorname{ev}^m)(\M_{d})$, and let $\phi: \ G(r,n)^{l}\times G(r,n)^{m} \to G(r,n)^m $ be the projection. Since the homology class $[\phi(Y \cap \sigma_{I_*})]$ is non-trivial \cite[page 64]{gr:pr}, $\phi(Y \cap \sigma_{I_*})$ must intersect some Schubert variety $$ \sigma_{J_*} = \sigma_{J_1} \times \sigma_{J_2} \times \ldots \times \sigma_{J_m} \subset G(r,n)^m $$ of complementary dimension. By Kleiman's lemma, \cite[Theorem 10.8 page 273]{ha:al}, the singular locus of $\phi(Y \cap \sigma_{I_*})$ does not intersect a general translate of $\sigma_{J_*}$, and similarly the singular locus of $\sigma_{J_*}$ does not intersect $\phi(Y \cap \sigma_{I_*})$. Therefore the intersection occurs in the smooth loci of $\phi(Y \cap \sigma_{I_*})$ and $\sigma_{J_*}$, and another application of the lemma implies that the intersection is finite. For generic translates of $\sigma_{J_*}$, the intersection is contained in $\operatorname{ev}^m(\sigma_{I_*}(p_*))$. Indeed, let $ {\overline{\sig}}_{I_*}(p_*) $ be the compactification of $\sig_{I_*}(p_*)$ given in \cite{be:qs}, and $\Gamma \subset \overline{ \sig}_{I_*}(p_*) \times G(r,n)^m$ the closure of the graph of $\operatorname{ev}^m$. Let $Z \subset \Gamma$ be the complement of the graph of $\operatorname{ev}^m$. The projection $\pi(Z)$ of $Z$ in $ G(r,n)^m$ is a closed sub-variety of $\phi(Y \cap \sigma_{I_*})$. By Kleiman's lemma, for generic translates of $\sig_{J_*}$ the intersection of $\pi(Z)$ and $\sig_{J_*}$ is empty, so the intersection is contained in $\operatorname{ev}^m(\sigma_{I_*}(p_*))$. Because $\operatorname{ev}^m \ | \ \sigma_{I_*}(p_*)$ is injective, the intersection $\sig_{I_*}(p_*) \cap \sig_{J_*}(p_*') $ is finite and non-empty. Since the homology class $[\phi(Y \cap \sigma_{I_*})]$ is independent of the choice of general translate of $\sigma_{I_*}$, the above intersection is finite and non-empty for general translates of the $\sig_{I_i}$ and $\sig_{J_j}$. This implies that Gromov-Witten invariant $$ \l [\sigma_{I_1}],\ldots,[\sig_{I_l}],[\sig_{J_1}],\ldots,[\sig_{J_m}] \r_d \neq 0 .$$ In terms of the quantum product $$ [\sigma_{I_1}] \star \ldots \star [\sig_{I_l}] \star [\sig_{J_1}] \star \ldots \star [\sig_{J_{m-1}}] \supset q^d \sig_{*J_m} $$ which implies that $ [\sigma_{I_1}] \star \ldots \star [\sig_{I_l}]$ contains a term with coefficient $q^i$ with $i \leq d$. That is, for some Schubert variety $\sigma$, $$ \l [\sig_{I_1}], [\sig_{I_2}], \ldots , [\sig_{I_l}],[\sigma] \r_i \neq 0 .$$ To prove the lemma it suffices to show that $i=d$. By \cite[Moving Lemma 2.2]{be:qs}, for general translates of the Schubert varieties the degree $i$ moduli space $\sig_{I_1}(p_1) \cap \ldots \cap \sig_{I_l}(p_l) \cap \sigma(p)$ is finite and consists of maps sending $p_1,\ldots,p_l,p$ to the corresponding Schubert cells. Since $d$ is minimal, $i = d$. \end{proof} \section{Factorization} In this section we show that a relationship between the polytopes for different numbers of marked points is related to factorization of Gromov-Witten invariants (i.e. associativity of quantum multiplication). A similar, easier, discussion holds for the additive polytopes $\Delta(l)$. A consideration of a ``trivial'' factorization completes the proof of Theorem \ref{final}. Suppose that $l$ can be written $l = j + k -2 $ for positive integers $j,k \ge 2$. It is easy to see that $\Delta_q(l)$ are projections of a section of $\Delta_q(j) \times \Delta_q(k)$\footnote{In fact, the volume functions satisfy the factorization properties $$ \on{Vol}(\mathcal{N}(\mu_1,\ldots,\mu_{j-1},\nu_1,\ldots,\nu_{k-1})) = \int_{\t_+} \on{Vol}(\mathcal{N}(\mu_1,\ldots,\mu_{j-1},*\lam)) \on{Vol}(\mathcal{N}(* \lam, \nu_1,\ldots,\nu_{k-1}) )\d \lam $$ $$ \on{Vol}(\M(\mu_1,\ldots,\mu_{j-1},\nu_1,\ldots,\nu_{k-1})) = \int_{\lie{A}} \on{Vol}(\M(\mu_1,\ldots,\mu_{j-1},*\lam) ) \on{Vol}(\M(* \lam, \nu_1,\ldots,\nu_{k-1})) \d \lam .$$ The second formula is implicit in Witten \cite[p.51]{wi:tw}, proved in \cite{jw:va}, and generalized in \cite{me:lo}.}. $$ \Delta_q(l) = \{ (\mu_1,\ldots,\mu_{j-1},\nu_1,\ldots,\nu_{k-1}) \ | \ (\mu,\nu) \in \Delta_q(j) \times \Delta_q(k), \ \ \mu_j = *\nu_k \} $$ To show the forward inclusion, note that if $A_1 A_2 \ldots A_l = I$ then letting $ B = A_j A_{j+1} \ldots A_l$ we have $$(\lam(A_1),\ldots,\lam(A_{j-1}),\lam(B)) \in \Delta_q(j), $$ $$ (\lam(B^{-1}),\lam(A_j), \ldots, \lam(A_l) ) \in \Delta(k).$$ In particular this means that any face of $\Delta_q(l)$ is a projection of a face (usually not of codimension $1$) of $\Delta_q(j) \times \Delta_q(k)$. Any face is the intersection of codimension $1$ faces. This shows that any defining inequality of $\Delta_q(l)$ is implied by a finite set of defining inequalities for $\Delta_q(j)$ and $\Delta_q(k)$. Using associativity of quantum cohomology one can be more specific about which inequalities for $\Delta_q(j),\Delta_q(k)$ are needed to imply an inequality for $\Delta_q(l)$. Suppose that a Gromov-Witten invariant $ \l \sigma_{I_1},\ldots,\sigma_{I_l}, \sigma_J \r_d \neq 0 $ so that one has an inequality for $\Delta(l)$ given by \begin{equation} \labell{desired} \sum_{i \in I_1} \lam_{1,i} + \ldots + \sum_{i \in I_{l}} \lam_{l,i} \leq d .\end{equation} Associativity of quantum multiplication says that $$ \l \sigma_{I_1},\ldots,\sigma_{I_l}, \sigma_J \r_d = \sum_{d_1 + d_2 = d, \ \ |K|=r } \l \sigma_{I_1},\ldots,\sigma_{I_{j-1}}, \sigma_K \r_{d_1} \l \sigma_{*K}, \sigma_{I_j},\ldots,\sigma_{I_l}, \sigma_J \r_{d_2} .$$ In particular there exist some $d_1,d_2$ with $d_1 + d_2 = d$ and some Schubert variety $\sigma_K$ such that $$ \l \sigma_{I_1},\ldots,\sigma_{I_{j-1}}, \sigma_K \r_{d_1} \neq 0, \ \ \ \l \sigma_{*K} , \sigma_{I_j},\ldots,\sigma_{I_l}, \sigma_J \r_{d_2} \neq 0 .$$ From the non-vanishing of these Gromov-Witten invariants one deduces the inequalities for $\Delta_q(j),\Delta_q(k):$ \begin{equation} \labell{add1} \sum_{i \in I_1} \mu_{1,i} + \ldots + \sum_{i \in I_{j-1}} \mu_{j-1,i} + \sum_{k \in K} \mu_{j,k} \leq d_1 ;\end{equation} \begin{equation} \labell{add2} \sum_{k \in *K} \nu_{1,k} + \sum_{i \in I_{j}} \nu_{2,i} + \ldots + \sum_{i \in I_{l}} \nu_{k,i} \leq d_2 .\end{equation} Restricting to the section $\mu_j = *\nu_1$ one has that $$ \sum_{k \in *K} \nu_{1,k} = - \sum_{k \in K} (* \nu_1)_k = - \sum_{k \in K} \mu_{j,k}, $$ so by adding the two inequalities one obtains \eqref{desired}. Using the trivial factorization $l = (l + 2) - 2$ we complete the proof of Theorem \ref{final}. \begin{lemma} \labell{factor} Any inequality for $\Delta_q(l)$ corresponding to a Gromov-Witten invariant $$ \l [\sigma_{I_1}],...,[\sigma_{I_l}],[\sigma_K]\r_d \neq 0 $$ is a consequence of an inequality corresponding to a Gromov-Witten invariant of the form $$ \l [\sigma_{I_1}],...,[\sigma_{I_{l-1}}],[\sigma_J]\r_{d_{1}} \neq 0 $$ for some $J \subset \{ 1,\ldots, n\}$ and $d_1\leq d.$ \end{lemma} \begin{proof} Suppose that $$ \l [\sigma_{I_1}],...,[\sigma_{I_l}],[\sigma_K]\r_d \neq 0 .$$ Taking $k=2$ we obtain that for some $J$ and $d_1\leq d$ $$ \l [\sigma_{I_1}],...,[\sigma_{I_{l-1}}],[\sigma_J]\r_{d_{1}} \l [\sigma_{*J}],[\sigma_{I_l}],[\sigma_K]\r_{d_2} \neq 0 .$$ Thus the inequality \begin{equation} \labell{want} \sum_{i \in I_1} \lam_{1,i} + \ldots + \sum_{i \in I_{l}} \lam_{l,i} \leq d .\end{equation} follows from the inequalities \begin{equation} \labell{first} \sum_{i \in I_1} \lam_{1,i} + \ldots + \sum_{i \in I_{l-1}} \lam_{l-1,i} + \sum _{j\in J} \lambda_{l,j} \leq d_1 \end{equation} for $\lam \in \Delta_q(l)$ and \begin{equation} \labell{last} \sum _{j\in *J} (*\lam_l)_j + \sum _{i\in I_l} (\lam_l)_i \leq d_2. \end{equation} The last equation is a tautology for $\lam_l \in \lie{A}$ by the $l=2$ case of Theorem \ref{result}. In other words, \eqref{last} is implied by the equations $\lam_{l,i} \ge \lam_{l,i+1}, \ \lam_{l,1} - \lam_{l,n} \le 1$. Thus \eqref{want} follows from \eqref{first} and the inequalities defining $\lie{A}^l$. \end{proof} \section{Hidden symmetry} \labell{symmetry} An interesting aspect of the multiplicative problem is that it possesses a symmetry not present in the additive case, related to the symmetry of the fundamental alcove $\lie{A}$ of $SU(n)$. Let $Z \cong \mathbb{Z}/n\mathbb{Z}$ denote the center of $SU(n)$, with generator $c \in SU(n)$ the unique element of $SU(n)$ with $$ \lambda(c) = (1/n,1/n, \ldots, 1/n, (1-n)/n) .$$ The action of $Z$ on $SU(n)$ induces an action on $\lie{A} \cong SU(n) / \Ad(SU(n))$, given by $$ c \cdot (\lambda_1,\ldots,\lambda_n) = (\lambda_2 + 1/n, \lambda_3 + 1/n,\ldots, \lambda_n + 1/n, \lambda_1 - (n-1)/n) .$$ Let $C(l) \subset SU(n)^l$ denote the subgroup $$ C(l) = \{ (z_1,\ldots,z_l) \subset Z^l \ | \ z_1z_2 \ldots z_l = 1 \} \cong Z^{l-1}.$$ The action of $C(l)$ on $\lie{A}^l$ leaves the polytope $\Delta_q(l)$ invariant. This symmetry of the polytope $\Delta_q(l)$ implies a symmetry on the facets of $\Delta_q(l)$. Let $c$ act on subsets of $\{ 1,2,\ldots,n \}$ via the action of $(12\ldots n)^{-1} \in S_n$: $$ c^m \{ i_1, \ldots, i_r \} = \{i_{s+1} - m, \ldots, i_r - m, i_1 - m + n, \ldots, i_s - m + n\} $$ where $s$ is the largest index for which $i_s - m \ge 1$ Suppose an $l+1$-tuple $(I_1,\ldots,I_l,d)$ defines a facet of $\Delta_q(l)$ via the inequality \eqref{mult_ineq}. Under the action of $ (c^{m_1},\ldots,c^{m_l}) \in C(l)$, \eqref{mult_ineq} becomes the inequality corresponding to $(c^{m_1}I_1,\ldots,c^{m_l}I_l,d')$ where $d'$ is defined by \begin{equation} \labell{4ac} \sum_{i=1}^l | c^{m_i}I_i | + nd' = \sum_{i=1}^l | I_i | + nd. \end{equation} \begin{example} From the degree $0$ inequality $\lambda_n(A) + \lambda_n(B) \le \lambda_n(AB)$ we obtain by the action of $(c^{-i},c^{-j},c^{i+j}), \ i + j \le n$ the degree $1$ inequality \eqref{quant_Weyl}. \end{example} Equation \eqref{4ac} defines a $C(l)$ action on the set of $l+1$-tuples $(I_1,\ldots,I_l,d)$ defining facets of $\Delta_q(l)$. It is an interesting fact that the Gromov-Witten invariants $\l \sig_{I_1},\ldots,\sig_{I_l}\r_d$ are invariant under this action: \begin{proposition} Let $(c^{m_1},\ldots,c^{m_l}) \in C(l)$. Then $\l \sig_{I_1},\ldots,\sig_{I_l} \r_d = \l \sigma_{c^{m_1} I_1}, \ldots, \sigma_{c^{m_l} I_l} \r_{d'}.$ \end{proposition} \begin{proof} Let $\sigma_c = \sigma_{r,r+1,\ldots,n-1}$ denote the Schubert variety isomorphic to the Grassmannian $G(r,n-1)$ of $r$-planes contained in $n-1$-space. We claim that quantum multiplication by $\sigma_c$ is given by the following formula: \begin{equation} \labell{Cox_mult} [ \sigma_c ] \star [ \sigma_I ] = q^{(|cI| + r - |I|)/n} [ \sigma_{cI} ] . \end{equation} The exponent $(|cI| + r - |I|)/n$ equals $1$ if $1 \in I$, and equals $0$ otherwise. In particular $[\sig_c]^{\star n}=q^r.$ The lemma then follows by associativity of the quantum product. Without loss of generality it suffices to show that the Gromov-Witten invariants are invariant under an element of the form $(c,c^{-1},1,\ldots,1) \in C(l)$. Given that $$ [\sig _{I_1}]\star \ldots \star [\sig _{I_{l-1}}]\supset \l \sig_{I_1},\ldots,\sig_{I_l} \r_d [\sigma_{*I_{l}}]q^d$$ multiplying by $[\sig_c]$ on both sides yields $$[\sig _{cI_1}]\star \ldots \star [\sig _{I_{l-1}}]\supset \l \sig_{I_1},\ldots,\sig_{I_l} \r_d [\sigma_{c(*I_{l})}]q^{d'}=\l \sig_{I_1},\ldots,\sig_{I_l} \r_d [\sig_{*c^{-1}I_l}]q^{d'}. $$ The formula \eqref{Cox_mult} may be proved using either the canonical isomorphism of quantum Schubert calculus with the Verlinde algebra of $U(r)$, $$ QH^*(G(r,n))/(q=1) \cong R(U(r)_{n-r,n}).$$ given a mathematical proof in Agnihotri \cite{ag:th}, or using the combinatorial formula of Bertram, Ciocan-Fontanine and Fulton \cite{be:qm}. $R(U(r)_{n-r,n})$ denotes the Verlinde algebra of $U(r)$ at $SU(r)$ level $n-r$ and $U(1)$ level $n$, and is the quotient of the tensor algebra $R(U(r))$ by the relations $$ V_{\lam} \sim (-1)^l(w) V_{w(\lam+ \rho)-\rho}, w \in \Waff $$ and if $\lam_1 - \lam_r \le n -r $ then $$ V_{(\lam_1,\ldots,\lam_r)} \sim V_{(\lam_2-1,\lam_3-1, \ldots,\lam_r-1,\lam_1 - (n-r+1)} .$$ Here $\Waff$ acts on $\Lambda^*$ at level $n$, and $\rho$ is the half-sum of positive roots. The Verlinde algebra $R(U(r)_{n-r,n})$ has as a basis the (equivalence classes of the) representations $V_{\lam}$, where $\lam = (\lambda_1,\ldots,\lambda_r) \in \mathbb{Z}^r, \ 0 \leq \lambda_i \leq n- r $ are dominant weights of $U(r)$ at level $n-r$. The canonical isomorphism is given by $\sigma_I \mapsto V_{\lambda}$, where $\lambda$ is defined by $$ \lambda_j = n-r + j - i_j .$$ The key point is that the sub-algebra $R(U(1)) \subset R(U(r))$ descends to a sub-algebra $ R(U(1)_n) \subset R(U(r)_{n-r,r})$ generated by the representation $V_c := V_{(1,1,\ldots,1)}$, which maps under the isomorphism to the Schubert variety $\sigma_c$. From the description of the algebra given above one sees that $ V_c \fus V_{\lam} = V_{\lambda'} $ where $$ \lambda' = \begin{array}{cl} (\lambda_1 + 1,\lambda_2 +1,\ldots, \lambda_r + 1) & \hbox{\ if\ }\lambda_1 < n-r \\ (\lambda_2,\ldots,\lambda_r,\lambda_1 - n +r) & \hbox{\ if\ }\lambda_1 = n-r \end{array} .$$ Since $V_{\lambda'}$ maps to $\sigma_{cI}$ under the canonical isomorphism, this proves \eqref{Cox_mult}. Alternatively, \eqref{Cox_mult} can be derived from the combinatorial rim-hook formula of \cite[p. 8]{be:qm}. Let $\lambda^t$ denote the transpose of $\lambda$, so that $\sig_{\lambda^t}$ is the image of $\sig_\lambda$ under the isomorphism $G(r,n) \cong G(n-r,n)$. The ordinary (resp. quantum) Littlewood-Richardson numbers are invariant under transpose $$ N_{\lam^t \mu^t}^{\rho^t} = N_{\lam \mu}^{\rho}, \ \ \ N_{\lam^t \mu^t}^{\rho^t}(n-r,r) = N_{\lam \mu}^{\rho}(r,n-r). $$ It follows from \cite[Corollary]{be:qm} that $$ N_{\lam \mu}^{\rho}(r,n-r) =\sum \eps(\rho^t/\nu^t) N_{\lam \mu}^{\rho} $$ where $\rho$ ranges over all diagrams of height $\leq r$ that can be obtained by adding $m$ rim-hooks. If $\mu = (1,1,\ldots,1)$ then $$ V_\mu \otimes V_{\lam} = V_\rho, \ \ \rho = (\lam_1 + 1, \ldots, \lam_r + 1) .$$ If $\lam_1 < n-r$, then since the height of $\rho$ is $\leq r$, there are no rim $n$-hooks in $\rho$. On the other hand, if $\lam_1 = n-r$, then it is easy to see that there is a unique rim $n$-hook in $\rho$, whose complement is $\lam'$ above. We have learned from A. Postnikov that formula similar to \eqref{Cox_mult} holds for the full flag variety \cite{po:hs}. A deeper reason for the appearance of symmetry is given by Seidel \cite{se:pi}. \end{proof} This symmetry simplifies the computation of many Gromov-Witten invariants. For sufficiently small $n$ and $l$ all Gromov-Witten invariants are equivalent to degree $0$ ones. An example of a Gromov-Witten invariant not equivalent via symmetry to a degree $0$ invariant is the degree $1$ invariant for $G(5,10)$ $$ \l \sigma_{\{ 2,4,6,8,10 \}}, \sigma_{\{ 2,4,6,8,10 \}}, \sigma_{\{ 1,3,5,7,9 \}} \r_1 $$ which may be computed using the formula of \cite{be:qm}. That is, for sufficiently large $r,n$, not all of the inequalities are related to ``classical'' inequalities via symmetry. \section{Verlinde algebras} \labell{Verlinde} Finally we want to explain the representation-theoretic interpretation of $\Delta_q(l)$ in terms of the Verlinde algebra of $SU(n)$. Denote by $\Lambda^*_N$ the set of dominant weights of $SU(n)$ at level $N$: $$ \Lambda^*_N = \{ (\lambda_1,\ldots,\lambda_n) \in (\mathbb{Z}/n)^n \ | \ \lam_i - \lam_{i+1} \in \mathbb{Z}_{\ge 0},\ \ \lam_1 - \lam_n \le N \} .$$ The Verlinde algebra $R(SU(n)_N)$ is the free group $\mathbb{Z}[\Lambda_N^*]$ on the generators $V_{\xi}, \xi \in \Lambda_N^*$. The algebra structure is given by ``fusion product'' $$ V_{\xi_1} \fus_N \ldots \fus_N \mathcal{V}_{\xi_l} = \sum_{\nu \in \Lambda^*_N} m^N({\xi_1,\ldots,\xi_l,\nu}) \ V_{* \nu} $$ where the coefficients $m^N(\xi_1,\ldots,\xi_l)$ are defined as follows. There is a positive line bundle $L^N({\xi_1,\ldots,\xi_l}) \to\M(\xi_1/N ,\ldots ,\xi_l/N)$ which descends from the polarizing line bundle on $\tilde{R}$ (see Pauly \cite[Section 3]{pa:em}). The coefficient $m^N({\xi_1,\ldots,\xi_l})$ is defined by $$ m^N({\xi_1,\ldots,\xi_l}) = \dim(H^0(L^N(\xi_1,\ldots,\xi_l)) .$$ Since $L^N$ is positive, $\M(\xi_1/N,\ldots,\xi_l/N)$ is non-empty if and only if for some $k$ $$\dim(H^0(L^k(\xi_1,\ldots,\xi_l)^{\otimes N})=\dim(H^0(L^{kN}(k\xi_1,\ldots,k\xi_l))\neq 0,$$ that is, $m^{kN}({k\xi_1,\ldots,k\xi_l}) \neq 0.$
1999-06-25T17:31:55
9712
alg-geom/9712026
en
https://arxiv.org/abs/alg-geom/9712026
[ "alg-geom", "math.AG" ]
alg-geom/9712026
Dr G. K. Sankaran
K. Hulek, I. Nieto, G. K. Sankaran
Degenerations of (1,3) abelian surfaces and Kummer surfaces
LaTeX, 16 pages, 2 figures. Final version, with minor corrections: to appear in Algebraic Geometry - Hirzebruch 70 (AMS Contemporary Mathematics)
null
null
null
null
We continue our study of the geometry of Nieto's quintic threefold, looking at degenerate surfaces that correspond to certain loci and showing how they arise from a toroidal compactification of a suitable moduli space.
[ { "version": "v1", "created": "Fri, 19 Dec 1997 22:02:57 GMT" }, { "version": "v2", "created": "Fri, 25 Jun 1999 15:31:55 GMT" } ]
2007-05-23T00:00:00
[ [ "Hulek", "K.", "" ], [ "Nieto", "I.", "" ], [ "Sankaran", "G. K.", "" ] ]
alg-geom
\section{Theta functions} In this section we will give an explicit description of a basis of the space $H^0({\mathcal L}^{\otimes 2})^-$ which defines the map from $A$ to ${\mathbb{P}}^3$, factoring through the Kummer surface, in terms of theta functions. Our standard reference for theta functions is Igusa's book \cite {I}. We shall denote points of the Siegel upper half plane ${\mathbb{H}}_2$ by $\tau=\left( \begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array} \right),$ and $z=(z_1, z_2)$ will denote the coordinates on ${\mathbb{C}}^2$. For every pair $(m', m'')\in{\mathbb{R}}^2\times{\mathbb{R}}^2$ we define the theta function $$ \Theta_{m' m''}(\tau, z)=\sum\limits_{q\in{\mathbb{Z}}^2} e^{2\pi i[\frac 12 (q+m')\tau^t(q+m')+(q+m')^t(z+m'')]}. $$ Given a point $\tau=\left( \begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array} \right)\in {\mathbb{H}}_2$ we associate to it a period matrix $$ \Omega_{\tau}=\left( \begin{array}{cccc} 2\tau_1 & 2\tau_2 & 2 & 0\\ 2\tau_2 & 2\tau_3 & 0 & 6 \end{array} \right) $$ and the lattice $$ L_{\tau}={\mathbb{Z}}^4\Omega_{\tau}={\mathbb{Z}} e_1+{\mathbb{Z}} e_2 + {\mathbb{Z}} e_3 +{\mathbb{Z}} e_4 $$ generated by the columns $e_i$ of the period matrix $\Omega_{\tau}$. The abelian surface $$ A_{\tau}={\mathbb{C}}^2/L_{\tau} $$ has a $(1,3)$--(and hence also a $(2,6)$--)polarization. Normally $0\in A_{\tau}$ is chosen as the origin and the involution given by taking the inverse is $\iota:x\mapsto -x$. For reasons which will become apparent later we shall want to define the origin of $A_{\tau}$ as the image of the point $$ \omega=\frac 12(1,1)\left( \begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array} \right)=\frac 12 (\tau_1+\tau_2, \tau_2+\tau_3). $$ Note that with respect to $0$ this is a $4$-torsion point. Then the involution with respect to $\omega$ is given by $$ \iota_{\omega}(z)=-z+2\omega. $$ Finally we set $$ \tau'=\left( \begin{array}{cc} \tau_1/2 & \tau_2/6\\ \tau_2/6 & \tau_3/18 \end{array} \right),\quad z'=(z_1/2,z_2/6). $$ The main objects of this section are the functions $$ {\widehat{\Theta}}_{\alpha \beta}(\tau,z): = \Theta_{00\frac{\alpha}{2}\frac{\beta}{6}}(\tau',z'-{\omega}'), \quad \alpha=0,1;\ \beta=0,\ldots,5. $$ Note that ${\widehat{\Theta}}_{\alpha+2, \beta}={\widehat{\Theta}}_{\alpha \beta}$ and $\widehat{\Theta}_{\alpha, \beta+6}=\widehat{\Theta}_{\alpha \beta}$ so that we can read the indices cyclically. \begin{lemma}\label{lem11} {\rm(i)} The functions ${\widehat{\Theta}}_{\alpha\beta}$ are all sections of the same line bundle ${\mathcal L}_{\tau}$ on $A_{\tau}$.\\ {\rm(ii)} ${\mathcal L}_{\tau}$ represents a polarization of type $(2,6)$. \end{lemma} \begin{proof} (i)\ We must prove that the automorphy factor of the functions ${\widehat{\Theta}}_{\alpha \beta}$ with respect to $z\mapsto z+e_i$ does not depend on $(\alpha, \beta)$. This follows immediately from the formulae $(\Theta 1)-(\Theta 5)$ of \cite[pp. 49, 50]{I}.\\ \noindent (ii)\ Since the type of a polarization is constant in families it is enough to prove the statement for $\tau_2=0$ where $$ A_{\tau}=E({\tau_1}) \times E({\tau_3}) $$ with $$ E(\tau_1)={\mathbb{C}}/({\mathbb{Z}} 2\tau_1+{\mathbb{Z}} 2),\ E(\tau_3)={\mathbb{C}}/({\mathbb{Z}} 2\tau_3+{\mathbb{Z}} 6). $$ In this case $$ {\widehat{\Theta}}_{\alpha \beta}(\tau,z)=\vartheta_{0\frac{\alpha}{2}}(\tau_1/2, z_1/2-{\omega_1}/2)\ \vartheta_{0\frac{\beta}{6}}(\tau_3/18, z_2/6-{\omega_2}/6) $$ where we use $\vartheta$ to denote theta functions in one variable. We claim that the degree on $E(\tau_1)$ is $2$ and that the degree on $E(\tau_3)$ is $6$. Indeed the first claim follows since $$ \begin{array}{rcl} \vartheta_{0\frac{\alpha}{2}}(\tau_1/2, z_1/2-{\omega_1}/2)=0 & \Leftrightarrow & z_1/2 \in ({\mathbb{Z}} \tau_1/2 +{\mathbb{Z}})-\alpha/2+{\omega_1}/2\\ &\Leftrightarrow & z_1\in ({\mathbb{Z}} \tau_1+{\mathbb{Z}} 2)-\alpha+{\omega_1}. \end{array} $$ This means that $\vartheta_{0\frac{\alpha}{2}}(\tau_1/2, z_1/2-{\omega_1}/2)$ has two zeroes on $E(\tau_1)$. The other claim follows in exactly the same way. \hfill \end{proof} We shall denote the sections of ${\mathcal L}_{\tau}$ defined by ${\widehat{\Theta}}_{\alpha\beta}(\tau, z)$ by ${\widehat{s}}_{\alpha \beta}$. By general theory the twelve sections ${\widehat{s}}_{\alpha\beta};\quad \alpha=0,1,\ \beta=0,\ldots,5$ form a {\em basis} of $H^0({\mathcal L}_{\tau})$. (Cf. \cite[p.75]{I} for an analogous statement.) We now want to describe the symmetry properties of the line bundle ${\mathcal L}_{\tau}$ and the sections ${\widehat{s}}_{\alpha\beta}$. The kernel of the map $$ \begin{array}{rcl} \lambda:\ A_{\tau} & \rightarrow & \mbox{Pic}^0A_{\tau}\\ x & \mapsto & t_x^*{\mathcal L}_{\tau} \otimes{\mathcal L}_{\tau}^{-1} \end{array} $$ where $t_x$ is translation by $x$ is equal to $$ \operatorname{ker } \lambda =({\mathbb{Z}} \frac{e_1}{2} +{\mathbb{Z}}\frac{e_2}{6}+{\mathbb{Z}}\frac{e_3}{2}+{\mathbb{Z}}\frac{e_4}{6}) L_{\tau}\cong({\mathbb{Z}}/2)^2 \times({\mathbb{Z}}/6)^2. $$ \noindent We set $\rho_6:=e^{2\pi i/6}$. \begin{proposition}\label{prop12} {\rm{(i)}} The group $\operatorname{ker} \lambda$ acts on the sections ${\widehat{s}}_{\alpha \beta}$ as follows $$ \begin{array}{rlcllclcl} e_1 /2 & : & {\widehat{s}}_{\alpha\beta} & \mapsto & (-1)^{\alpha}\ {\widehat{s}}_{\alpha\beta}\ , & e_2 /6 & : {\widehat{s}}_{\alpha\beta} & \mapsto & \rho_6^{-\beta}\ {\widehat{s}}_{\alpha\beta}\\ e_3 /2 & : & {\widehat{s}}_{\alpha\beta} & \mapsto & {\widehat{s}}_{\alpha+1, \beta}\ , & e_4 /6 & : {\widehat{s}}_{\alpha\beta} & \mapsto & {\widehat{s}}_{\alpha, \beta+1}. \end{array} $$ {\rm{(ii)}} The involution $\iota_{\omega}$ acts on the sections ${\widehat{s}}_{\alpha\beta}$ by $$ \iota_{\omega}:{\widehat{s}}_{\alpha\beta}\mapsto {\widehat{s}}_{-\alpha, -\beta}. $$ \end{proposition} \begin{proof} (i)\ We shall prove this for $e_1/2$, the other cases being similar. Again using \cite[pp. 49, 50]{I} we find $$ \begin{array}{rcl} {\widehat{\Theta}}_{\alpha\beta}(\tau, z+\frac{e_1}{2}) & = & \Theta_{00\frac{\alpha}{2}\frac{\beta}{6}}(\tau', z'-{\omega}'+(\tau_1/2, \tau_2/6))\\[2mm] & {\begin{array}{c} (\Theta 3)\\ {\sim} \end{array}} & e^{2\pi i (-(1,0)({\begin{array}{c}\alpha/2\\ \beta/6 \end{array})})} \Theta_{10\frac{\alpha}{2}\frac{\beta}{6}}(\tau', z'-{\omega}')\\[2mm] & {\begin{array}{c} (\Theta 1)\\ = \end{array}} & (-1)^{\alpha} \Theta_{00\frac{\alpha}{2}\frac{\beta}{6}}(\tau', z'-{\omega}')\\ [2mm] & = & (-1)^{\alpha}{\widehat{\Theta}}_{\alpha\beta}(\tau, z). \end{array} $$ \noindent Here $\sim$ denotes equality up to a nowhere vanishing function which is independent of $\alpha$ and $\beta$.\\ \noindent (ii) Here we have that $$ \begin{array}{rcl} {\widehat{\Theta}}_{\alpha\beta}(\tau, -z+2\omega) & = & \Theta_{00\frac{\alpha}{2}\frac{\beta}{6}}(\tau', -z'+{\omega}')\\ & {\begin{array}{c} (\Theta 1)\\ = \end{array}} & \Theta_{00-\frac{\alpha}{2}-\frac{\beta}{6}}(\tau', z'-{\omega}')\\ & = & {\widehat{\Theta}}_{-\alpha -\beta}(\tau, z) \end{array} $$ where indices are to be read cyclically.\hfill \end{proof} \begin{remark}\label{rem13} {\rm{(i)}} Part (i) of the above proposition gives an explicit description of the lifting of the group $({\mathbb{Z}} /2)^2 \times ({\mathbb{Z}} /6)^2$ to the Heisenberg group $H_{26}$.\\ {\rm{(ii)}} Note that part (ii) of the above proposition is true for any choice of the point $\omega$ and hence in particular also for the involution $\iota$ itself. \end{remark} We can now describe a basis of the eigenspaces $H^0({\mathcal L}_\tau)^+$ and $H^0({\mathcal L}_\tau)^-$ as follows: $$ \begin{array}{ll} {\widehat{u}}_{\alpha\beta}={\widehat{s}}_{\alpha\beta}+{\widehat{s}} _{-\alpha,-\beta}\in H^0({\mathcal L}_\tau)^+; & \alpha \in \{0,1\}, \beta \in \{0,1,2,3\}\\ {\widehat{t}}_{\alpha\beta}={\widehat{s}}_{\alpha\beta}-{\widehat{s}} _{-\alpha,-\beta}\in H^0({\mathcal L}_\tau)^-; & (\alpha,\beta)= (0,1),(0,2),(1,1),(1,2). \end{array} $$ \noindent For our purposes it is, however, better to work with a different basis of $H^0({\mathcal L}_\tau)^-$. $$ \begin{array}{ccrclclcl} {\widehat{g}}_0 & := & {\widehat{t}}_{01} & + &{\widehat{t}}_{11}& - & {\widehat{t}}_{02} & - & {\widehat{t}}_{12}\\ {\widehat{g}}_1 & := & -{\widehat{t}}_{01} & - &{\widehat{t}}_{11}& - & {\widehat{t}}_{02} & - & {\widehat{t}}_{12}\\ {\widehat{g}}_2 & := & {\widehat{t}}_{01} & - &{\widehat{t}}_{11}& - & {\widehat{t}}_{02} & + & {\widehat{t}}_{12}\\ {\widehat{g}}_3 & := & -{\widehat{t}}_{01} & + &{\widehat{t}}_{11}& - & {\widehat{t}}_{02} & + & {\widehat{t}}_{12}. \end{array} $$ Recall the Heisenberg group $H_{22}$ from \cite {BN}. The group $H_{22}$ has order~$32$ and $$ H_{22}/\mbox{ centre } \cong ({\mathbb{Z}}/2)^4 $$ is the group generated by the elements $$ \begin{array}{cclcl} \sigma_1 & : & (z_0: z_1: z_2: z_3) & \mapsto & (z_2: z_3: z_0: z_1)\\ \sigma_2 & : & (z_0: z_1: z_2: z_3) & \mapsto & (z_1: z_0: z_3: z_2)\\ \tau_1 & : & (z_0: z_1: z_2: z_3) & \mapsto & (z_0: z_1: -z_2:-z_3)\\ \tau_2 & : & (z_0: z_1: z_2: z_3) & \mapsto & (z_0:-z_1: z_2: -z_3). \end{array} $$ The group ${}_2A_{\tau}$ of $2$-torsion points of $A_{\tau}$ is contained in $\operatorname{ker} \lambda$. Here we identify ${}_2A_{\tau}$ with translations of $A_{\tau}$ of order~$2$. Using the translations $x\mapsto x+e_i/2$ as generators we obtain an identification of ${}_2A_{\tau}$ with $({\mathbb{Z}}/2)^4$. A straightforward calculation using Proposition \ref{prop12} and the definition of the basis ${\widehat{g}}_0,\ldots,{\widehat{g}}_3$ shows that $$ \begin{array}{clclll} e_1/2 : &(\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3) & \mapsto & (\widehat{g}_2: \widehat{g}_3: \widehat{g}_0: \widehat{g}_1) & = & \sigma_1 (\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3)\\ e_2/2 : &(\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3) & \mapsto & (\widehat{g}_1: \widehat{g}_0: \widehat{g}_3: \widehat{g}_2) & = & \sigma_2 (\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3)\\ e_3/2 : &(\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3) & \mapsto & (\widehat{g}_0: \widehat{g}_1: -\widehat{g}_2: -\widehat{g}_3) & = & \tau_1 (\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3)\\ e_4/2 : &(\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3) & \mapsto & (\widehat{g}_0: -\widehat{g}_1: \widehat{g}_2: -\widehat{g}_3) & = & \tau_2 (\widehat{g}_0: \widehat{g}_1: \widehat{g}_2: \widehat{g}_3) \end{array} $$ We can, therefore, summarize our results as follows: \begin{theorem}\label{theo13} {\rm(i)} The basis $\widehat{g}_0,\ldots,\widehat{g}_3 \in H^0({\mathcal L}_{\tau})^-$ defines a rational map from $A_{\tau}$ to ${\mathbb{P}}^3$ which factors through $\widetilde{\operatorname{Km}}(A_{\tau})$. This map is equivariant with respect to the action of ${}_2A_{\tau}\cong ({\mathbb{Z}}/2)^4$ on $A_{\tau}$ and of $H_{22}/\mbox{centre }\cong ({\mathbb{Z}}/2)^4$ on ${\mathbb{P}}^3$. In particular the image is $H_{22}$-invariant.\\ {\rm(ii)} The Kummer surface $\widetilde{\operatorname{Km}}(A_{\tau})$ is embedded as a smooth quartic surface if and only if $A_{\tau}$ is neither a product nor a bielliptic abelian surface. If $A_{\tau}$ is bielliptic then $\widetilde{\operatorname{Km}}(A_{\tau})$ is mapped to a quartic with four nodes; if $A_{\tau}$ is a product, then $\widetilde{\operatorname{Km}}(A_{\tau})$ is mapped $2:1$ onto a quadric. \end{theorem} \begin{proof} {\rm(i)} Follows immediately from our above calculations.\\ {\rm(ii)} This was shown in \cite{HNS}. \hfill \end{proof} \begin{remark} ${\mathcal L}_{\tau}$ is the unique totally symmetric line bundle with respect to the involution ${\iota}_{\omega}$. \end{remark} \section{Degenerations} In this section we construct degenerations of $(1,3)$--polarized abelian surfaces which correspond to points on the S-planes. The construction of degenerating families of abelian varieties is in general technically complicated (see e.g. \cite{FC}, \cite{AN}). Although we cannot avoid these technicalities entirely, we have tried to present our construction in a way which uses only a minimum of technical steps. These, however, cannot be avoided. We consider the group $$ P=\left\{ \left( \begin{array}{c|c} \left.{\bf 1}\right. & \begin{array}{rr} 2 {\mathbb{Z}} & 6{\mathbb{Z}}\\ 6 {\mathbb{Z}} & 18 {\mathbb{Z}} \end{array}\\ \hline 0 & {\bf 1} \end{array} \right) \right\} \subset \mbox{Sp}(4,{\mathbb{Z}}). $$ Note that this is the lattice contained in the parabolic subgroup of $\Gamma_{1,3}(2)\cap \Gamma_{1,3}^{\mbox{\scriptsize lev}} $ which fixes the isotropic plane $h=(0, 0, 1, 0)\wedge (0, 0, 0, 1)$. Here $\Gamma_{1,3}(2)$ is the group which defines the moduli space of abelian surfaces with a $(1,3)$--polarization and a level-$2$ structure, whereas $\Gamma_{1,3}^{\mbox{\scriptsize lev}}$ belongs to the moduli space of $(1,3)$--polarized abelian surfaces with a canonical level structure (cf. \cite[I.1] {HKW}). There are two reasons for considering this group. One is that we can then make use of the constructions in \cite{HKW} which from our point of view is the most economical way to construct the degenerations which we are interested in; the second reason is that, at least with the known constructions of degenerations of abelian surfaces, the presence of a canonical level structure is necessary. We could also use the method of Alexeev and Nakamura \cite{AN} which likewise goes back to Mumford's construction \cite{M}, and \cite{Nak}, \cite{Nam}. For the surfaces which we are interested in it makes, however, little difference which of these methods we choose. The group $P$ acts on ${\mathbb{H}}_2$ by $$ \left( \begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array} \right)\mapsto\left( \begin{array}{cc} \tau_1+2{\mathbb{Z}} & \tau_2+6{\mathbb{Z}}\\ \tau_2+6{\mathbb{Z}} & \tau_3+18{\mathbb{Z}} \end{array}\right). $$ The partial quotient of ${\mathbb{H}}_2$ by $P$ is given by $$ \begin{array}{ccl} {\mathbb{H}}_2 & \rightarrow & {\mathbb{H}}_2/P\subset ({\mathbb{C}}^*)^3\\ \left( \begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array} \right) & \mapsto & (e^{2\pi i \tau_1/2}, e^{2\pi i \tau_2/6},e^{2\pi i \tau_3/18})=(t_1,t_2,t_3). \end{array} $$ Recall that $$ A_{\tau}={\mathbb{C}}^2/L_{\tau} $$ where $$ L_{\tau}={\mathbb{Z}} \left( \begin{array}{l} 2\tau_1\\ 2\tau_2 \end{array} \right)+ {\mathbb{Z}} \left( \begin{array}{l} 2\tau_2\\ 2\tau_3 \end{array} \right)+ {\mathbb{Z}} \left( \begin{array}{c} 2\\ 0 \end{array} \right)+ {\mathbb{Z}} \left( \begin{array}{c} 0\\ 6 \end{array} \right)=L'_{\tau}+L''. $$ Here $L_{\tau}'$ is spanned by the first two columns of $\Omega_{\tau}$ and $L''$ by the last two. Obviously $L''$ does not depend on $\tau$ and $$ {\mathbb{C}}^2/L''=({\mathbb{C}}^*)^2. $$ We shall use the coordinates $$ w_1=z_1/2,\quad w_2=z_2/6 $$ on $({\mathbb{C}}^*)^2$. The lattice $L'_{\tau}$ acts on the trivial torus bundle ${\mathbb{H}}_2/P\times({\mathbb{C}}^*)^2$ by $$ (m, n): (t_1, t_2, t_3; w_1, w_2)\mapsto (t_1, t_2, t_3;\ t_1^{2m}\ t_2^{6n} w_1, t_2^{2m}\ t_3^{6n} w_2). $$ We have to extend the trivial bundle ${\mathbb{H}}_2/P\times({\mathbb{C}}^*)^2$ to the boundary in such a way that the action of $L'_{\tau}$ also extends. The general theory of toroidal compactifications of moduli spaces of abelian surfaces (the material which is relevant in our situation can be found in \cite{HKW}) leads us to consider first the map $$ \begin{array}{ccl} ({\mathbb{C}}^*)^3 & \rightarrow & {\mathbb{C}}^3\\ (t_1, t_2, t_3) & \mapsto & (t_1 t_2, t_2 t_3, t_2^{-1})=(T_1, T_2, T_3). \end{array} $$ Let $$ B:=\overset{\circ}{(\overline{{\mathbb{H}}_2/P})} $$ be the interior of the closure of ${\mathbb{H}}_2/P$ in ${\mathbb{C}}^3$ in the ${\mathbb{C}}$--topology. (What we have considered here is an open part of the partial compactification in the direction of the cusp corresponding to $h=(0,0,1,0) \wedge (0,0,0,1)$. The surfaces $B\cap\{T_i=0\}$ are mapped to boundary surfaces in the Igusa compactification of the moduli space ${\mathcal A}_{1,3}^{\mbox{\scriptsize lev}}(2)$ of $(1,3)$--polarized abelian surfaces with both a level-$2$ and a canonical level structure.) In terms of the coordinates $T_i$ the action of $L'_{\tau}$ is now given by $$ (m, n): (T_1, T_2, T_3; w_1, w_2)\mapsto (T_1, T_2, T_3;\ T_1^{2m} T_3^{2m-6n} w_1, T_2^{6n} T_3^{6n-2m} w_2). $$ Here we are particularly interested in degenerations which are given by $\tau_1\rightarrow i\infty$. This is equivalent to $t_1=0$ and hence corresponds to points on the surface $T_1=0$. We now consider the space $$ {\tilde {\mathcal P}}=\mbox{Proj } R_{\Phi,\Sigma}\rightarrow \mbox{Spec } {\mathbb{C}}[T_1, T_2, T_3]\cong {\mathbb{C}}^3 $$ which was defined in \cite[p.210]{HKW}. Let $$ {\mathcal P}:={\tilde {\mathcal P}}|_{B}. $$ Then ${\mathcal P}$ is a partial compactification of the trivial torus bundle ${\mathbb{H}}_2/P\times({\mathbb{C}}^*)^2$ over ${\mathbb{H}}_2/P\subset B$. Moreover the action of $L_{\tau}$ on the trivial torus bundle extends to an action on ${\mathcal P}$. The construction of ${\tilde {\mathcal P}}$ is originally due to Mumford \cite[final example]{M}. Let $$ {\bar A}:={\mathcal P}/L_{\tau}. $$ Then we have a diagram $$ \begin{array}{ccc} A & \subset & {\bar A}\\ \pi\downarrow & & \downarrow\pi\\ {\mathbb{H}}_2/P & \subset & B \end{array} $$ where $A=({\mathbb{H}}_2/P\times({\mathbb{C}}^*)^2)/L_{\tau}$ is the universal family. In particular ${\bar A}$ extends the universal family $A$ to the boundary. The fibres $$ {\bar A}_{u}=\pi^{-1}(u) $$ over ``boundary points'' $u\in B\backslash({\mathbb{H}}_2/P)$ are degenerate abelian surfaces. We are interested in the fibres ${\bar A}_u$ over points $u=(0, T_2, T_3)$ with $T_2 T_3\neq 0$. These are the corank~$1$ degenerations associated to the boundary component given by $\tau_1\rightarrow i\infty$. Note that if $T_2T_3\neq 0$ then this gives $t_2=T_3^{-1}$ and $t_3=T_2T_3$. In particular the point $u$ determines a point $(\tau_2, \tau_3)\in{\mathbb{C}}\times {\mathbb{H}}_1$ where $\tau_2$ and $\tau_3$ are uniquely defined up to $6{\mathbb{Z}}$ and $18{\mathbb{Z}}$ respectively. We can now formulate the main result of this section. \begin{theorem}\label{theo21} Let $u=(0, T_2, T_3)\in B$. Then ${\bar A}_u$ is a degenerate abelian surface with the following properties:\\ {\rm(i)} ${\bar A}_u$ is a corank $1$ degeneration. More precisely ${\bar A}_u$ is a chain of two elliptic ruled surfaces $A_{u,1}, A_{u,2}$ i.e. there exists an elliptic curve $E$ and a line bundle ${\mathcal M}_u\in \operatorname{Pic}^{0}(E)$ such that $A_{u,i}={\mathbb{P}}({\mathcal O}_E\oplus {\mathcal M}_u), \ i=1,2$. The surfaces $A_{u,i}$ are glued with a glueing parameter $e$ as shown below in Figure~$1$.\\ {\rm(ii)} The base curve $E\cong E(\tau_3)={\mathbb{C}}/({\mathbb{Z}} 2\tau_3 + {\mathbb{Z}} 6)$.\\ {\rm(iii)} The line bundle ${\mathcal M}_u={\mathcal O}_{E(\tau_3)}(6[\tau_2]-6[0])$ where $[\tau_2],[0]$ are the points of $E(\tau_3)$ given by $\tau_2$ and~$0$.\\ {\rm(iv)} The glueing parameter $e=[2\tau_2] \in E(\tau_3).$ \unitlength1cm \begin{figure}[htb] \begin{picture}(13.5,8.5) \put(-0.7,0){\includegraphics{bild13.eps}} \end{picture} \caption{Glueing of the surface $\bar A_u$} \end{figure} \end{theorem} \begin{proof} We can derive this from \cite[part II]{HKW}. There the quotient ${\hat A}={\mathcal P}/{\hat L}$ was considered where ${\hat L}\cong {\mathbb{Z}}^2$ acts on the trivial torus bundle ${\mathbb{H}}_2/P\times({\mathbb{C}}^*)^2$ by $$ (m, n): (T_1, T_2, T_3; w_1, w_2) \mapsto (T_1, T_2, T_3; T_1^{m} T_3^{m-n} w_1, T_2^{n} T_3^{n-m} w_2). $$ Hence $L'_{\tau}$ is a subgroup of ${\hat L}$ with ${\hat L}/L_{\tau}'\cong({\mathbb{Z}}/2)\times({\mathbb{Z}}/6)$. This means that we can use the description of ${\hat A}$ given in \cite[part II]{HKW} to give a description of ${\bar A}$. In the terminology of \cite{HKW} the group $L_{\tau}'=<s^{-2}, r^{-6}>$. The statements (i) and (ii) now follow exactly as in the proof of \cite [Theorem II.3.10]{HKW}. In particular, the fact that $s^2\in L_{\tau}'$, but $s \not\in L_{\tau}'$ implies that ${\bar A}_u$ has two irreducible components. The statement about the base curve $E$ follows from diagram \cite[II.3.13]{HKW}. Statements (iii) and (iv) are an immediate consequence of the proof of \cite[Proposition (II.3.20)]{HKW}. \end{proof} Our next task is to study the involutions $\iota$ and $\iota_{\omega}$ on $A$ and their extensions to ${\bar A}$. If we choose $0\in A_{\tau}={\mathbb{C}}^2/L_{\tau}$ as the origin, then this defines a section of $A$ which extends to a section of ${\bar A}$. Moreover, the involution $\iota:z\mapsto -z$ defines an involution of $A$ which extends to ${\bar A}$ (this is the involution given by \cite [Lemma (II.2.9)(ii)]{HKW}. But we said in section 1 that we wanted to choose $\omega=[(\tau_1+\tau_2)/2, (\tau_2 + \tau_3)/2]$ as the origin. This point is a $4$-torsion point of $A_{\tau}$ if we choose~$0$ as the origin. This choice of origin will be necessary for what follows, but at this point it has the disadvantage that it only defines a multisection of $A$, not a section. Nevertheless this multisection extends to ${\bar A}$. We also claim that the involution $\iota_{\omega}(z)=-z+2\omega$ extends to ${\bar A}$. Since $z\mapsto -z$ is defined on ${\bar A}$ it is enough to show that the translation $z\mapsto z+2\omega$ is defined on $A$ and extends to ${\bar A}$. This is easy to see, since $z\mapsto z+2\omega$ in terms of the coordinates $w_1, w_2$ is given by $$ (w_1, w_2)\mapsto(t_1 t_2^3 w_1, t_2 t_3^3 w_2)=(T_1 T_3^{-2} w_1, T_2^3 T_3^2 w_2). $$ This is the element $s^{-1} r^{-3} \in {\hat L}$ and hence acts on ${\mathcal P}$ and on ${\bar A}={\mathcal P}/L'_{\tau}$. Since $s^{-2} r^{-6}\in L'_{\tau}$, this is an involution. In particular $\iota_{\omega}$ defines an involution on the fibres ${\bar A}_u$ of ${\bar A}$. Recall that for $u=(0, T_2, T_3)$ with $T_2 T_3\neq 0$ the surface ${\bar A}_u$ has two irreducible components $A_{u,i},\ i=1,2$ and that the singular locus of ${\bar A}_u$ consists of two disjoint elliptic curves $E_1$ and $E_2$ with $E_i \cong E(\tau_3)={\mathbb{C}}/({\mathbb{Z}} 2\tau_3+{\mathbb{Z}} 6)$. \begin{proposition}\label{prop22} The involution $\iota_{\omega}$ interchanges the two components $A_{u,1}$ and $A_{u,2}$ of ${\bar A}_u$ and induces an involution on each of the two curves $E_1$ and $E_2$ with four fixed points on each of these curves. \end{proposition} \begin{proof} The involution $\iota$ fixes each of the surfaces $A_{u,1}$ and $A_{u,2}$ and interchanges $E_1$ and $E_2$. Addition by the $2$-torsion point $2\omega$ also interchanges $A_{u,1}$ and $A_{u,2}$ as well as $E_1$ and $E_2$. Hence $\iota_{\omega}$ interchanges $A_{u,1}$ and $A_{u,2}$ but induces non-trivial involutions on $E_1$ and $E_2$. In order to determine the fixed points of $\iota_{\omega}$ it is sufficient to compute the limit of the fixed points of $\iota_{\omega}$ in $A_{\tau}$ as $\tau_1\rightarrow i\infty$. The sixteen fixed points of $\iota_{\omega}$ on $A_{\tau}$ are given by $$ \left[(\tau_1+\tau_2, \tau_2+\tau_3)/2+\varepsilon_1(\tau_1,\tau_2)+\varepsilon_2(\tau_2,\tau_3)+ \varepsilon_3 (1,0)+ \varepsilon_4(0,3)\right]\in A_{\tau} $$ where $\varepsilon_i=0$ or $1$. As $\tau_1\rightarrow i\infty$ these 16 points come together in pairs; more precisely any two points which only differ by $\varepsilon_3$ have the same limit. This gives us eight points of which four lie on each of the curves $E_i$ (depending on whether $\varepsilon_1=0$ or $1$). These points are given by $$ \begin{array}{ll} \left[(\tau_2+\tau_3)/2+\varepsilon_2\tau_3+\varepsilon_4 3\right] \in E(\tau_3) & (\varepsilon_1=0)\\[2mm] \left[(\tau_2+\tau_3)/2+\tau_2+\varepsilon_2\tau_3+\varepsilon_4 3\right]\in E(\tau_3) & (\varepsilon_1=1). \end{array} $$ \hfill \end{proof} Figure~$2$ indicates the action of $\iota_{\omega}$ and the position of the eight fixed points on $\bar A_u$. \unitlength1cm \begin{figure}[htb] \begin{center} {\includegraphics[width=0.5\columnwidth]{bild16.eps}} \end{center} \caption{The involution $\iota_{\omega}$} \end{figure} Note that two fixed points lie on one ruling if and only if $[\tau_2]$ is a $2$-torsion point on $E(\tau_3)$, i.e. if and only if the glueing parameter $e=[2\tau_2]=0$. The next step is to extend the polarization to the degenerate abelian surfaces. Ideally we would like to glue the line bundle ${\mathcal L}_{\tau}$ on $A_{\tau}$ to a line bundle ${\mathcal L}$ on $A$ and to extend this line bundle to ${\bar A}$ in such a way that the sections ${\widehat s}_{\alpha \beta}$ as well as the action of the symmetry group (see Proposition \ref {prop12}) extend. At this point, however, we encounter a fundamental difficulty. We have seen that it is possible to extend $A$ to ${\bar A}$ in such a way that the symmetries, and here in particular the involution $\iota_{\omega}$, extend. It is also possible to define a suitable line bundle ${\mathcal L}$ and extend it to a line bundle ${\bar {\mathcal L}}$ on ${\bar A}$ (see \cite[II.5]{HKW}). But it is not possible to do this in such a way that the action of the symmetry group also extends to ${\bar{\mathcal L}}$. (This leads in particular to a numerical contradiction on the fibre over the origin $0\in B$.) For this reason we shall now restrict ourselves to taking the partial quotient with respect to the group $$ P'=\left\{ \left( \begin{array}{c|c} {\bf 1} & \begin{array}{cc} 2{\mathbb{Z}}\ & 0\\ 0 & 0 \end{array}\\ \hline 0 & {\bf 1} \end{array} \right) \right\} \subset \mbox {Sp}(4,{\mathbb{Z}}). $$ This is the lattice contained in the stabilizer of the line generated by $l_0=(0, 0, 1, 0)$ in the group $\Gamma_{1,3}(2)\cap \Gamma_{1,3}^{\mbox{\scriptsize lev}} $. The partial quotient defined by this group is given by the map $$ \begin{array}{rcl} {\mathbb{H}}_2 & \rightarrow & {\mathbb{C}}^*\times{\mathbb{C}}\times{\mathbb{H}}_1\subset{\mathbb{C}}\times{\mathbb{C}}\times{\mathbb{H}}_1\\[2mm] \left(\begin{array}{cc} \tau_1 & \tau_2\\ \tau_2 & \tau_3 \end{array}\right) & \mapsto & (t_1=e^{2\pi i\tau_1/2}, \tau_2, \tau_3). \end{array} $$ Partial compactification of ${\mathbb{H}}_2/P'$ in ${\mathbb{C}}\times{\mathbb{C}}\times{\mathbb{H}}_1$ is given by $$ B':=\overset{\circ}{(\overline{{\mathbb{H}}_2/P'})} $$ The two partial quotients with respect to $P$ and $P'$ are related by the glueing map $$ \begin{array}{rcl} \varphi:\quad B' & \rightarrow & B\\ (t_1, \tau_2, \tau_3) & \mapsto & (t_1 t_2, t_2 t_3, t_2^{-1}) \end{array} $$ where $t_2=e^{2\pi i\tau_2/6}$ and $t_3=e^{2\pi i \tau_3/18}$. The image of $\varphi$ is $B\backslash (B\cap\{T_2T_3=0\})$ and the map $\varphi$ is unramified onto its image. We can pull the family ${\bar A}$ over $B$ back to $B'$ via $\varphi$ and we shall denote the resulting family by ${\bar A'}$. This family extends the universal family $A'$ over ${\mathbb{H}}_2/P'$. We shall denote the projection from ${\bar A'}$ to $B'$ by $\pi'$. \begin{proposition}\label{prop23} {\rm(i)} The line bundles ${\mathcal L}_{\tau}$ on $A_{\tau}$ glue to a line bundle ${\mathcal L}'$ on~$A'$.\\ {\rm(ii)} The line bundle ${\mathcal L}'$ can be extended to a line bundle ${\bar{\mathcal L'}}$ on ${\bar A'}$ in such a way that the sections ${\widehat s}_{\alpha \beta}$ as well as the action of the Heisenberg group $H_{2 6}$ and the involution $\iota_{\omega}$ extend. \end{proposition} \begin{proof} A straightforward computation shows that with respect to the coordinates $t_1=e^{2\pi i\tau_1/2}$ and $w_1=e^{2\pi i z_1/2}, w_2=e^{2\pi i z_2/6}$ the functions ${\widehat\Theta}_{\alpha \beta}(\tau, z)$ are given by: $$ \begin{array}{rcl} {\widehat\Theta}_{\alpha \beta}(\tau, z) & = & {\Theta}_{00}\frac{\alpha}{2}\frac{\beta}{6}(\tau', z' -{\omega}')\\[2mm] & = & \sum\limits_{q\in{\mathbb{Z}}^2} t_1^{\frac 12 q_1(q_1-1)} \exp\{9\pi i q_2(q_2-3)\tau_3\}\\[3mm] & & \qquad\cdot\exp\{6\pi i(2q_1 q_2 -3 q_1 -q_2)\tau_2\}w_1^{q_1} w_2^{q_2} (-1)^{\alpha q_1}\rho_6^{\beta q_2}. \end{array} $$ In particular this shows that we can consider these functions as functions on $({\mathbb{H}}_2/P')\times ({\mathbb{C}}^*)^2$. Similarly we find that with respect to the lattice $L_{\tau}$ the functions ${\widehat\Theta}_{\alpha \beta}(\tau, z)$ have the following transformation behaviour. For $(k, l)\in {\mathbb{Z}}^2$: $$ {\widehat\Theta}_{\alpha\beta} (\tau, z+(2k,6l))={\widehat\Theta}_{\alpha\beta}(\tau,z). $$ For $(m,n)\in {\mathbb{Z}}^2$: $$ \begin{array}{l} {\widehat\Theta}_{\alpha\beta} \left(\tau, z+(m,n)\left( \begin{array}{cc} 2\tau_1 & 2\tau_2\\ 2\tau_2 & 2\tau_3 \end{array} \right)\right) \\ \qquad= \Theta_{00\frac{\alpha}{2}\frac{\beta}{6}}\left(\tau', z'-{\omega}'+(2m,6n) \left( \begin{array}{cc} \tau'_1 & \tau'_2\\ \tau'_2 & \tau'_3 \end{array}\right)\right)\\[2mm] \qquad=\exp\left\{2\pi i\left[-\frac 12 (2m, 6n)\left( \begin{array}{cc} \tau'_1 & \tau'_2\\ \tau'_2 & \tau'_3 \end{array}\right)\left(\begin{array}{c} 2m\\ 6n\end{array}\right)-(2m, 6n)(z'-{\omega}')\right]\right\}\\[2mm] \qquad\qquad\cdot\Theta_{2m, 6n,\frac {\alpha}{2},\frac{\beta}{6}} (\tau', z'-{\omega}')\\[2mm] \qquad= t_1^{-2m^2+m} w_1^{-2m} w_2^{-6n} e^{2\pi i[(-2mn+\frac{m^2}{2})\tau_2+(-n^2+\frac n2)\tau_3]} {\widehat\Theta}_{\alpha \beta}(\tau, z). \end{array} $$ \noindent These calculations show claim (i). To prove (ii) we have to consider the limit as $t_1\rightarrow 0$. Here we find $$ \begin{array}{rcl} \lim\limits_{t_1\rightarrow 0}{\widehat\Theta}_{\alpha \beta}(\tau, z) & = &\sum\limits_{q_2\in {\mathbb{Z}}} e^{9\pi i q_2(q_2-3)} e^{6\pi i (-q_2\tau_2)} w_2^{q_2} \rho_6^{\beta q_2}\\ &&\quad+(-1)^{\alpha} w_1\sum\limits_{q_2\in{\mathbb{Z}}} e^{9\pi i q_2(q_2-3)} e^{2\pi i (q_2-3)\tau_2} w_2^{q_2} \rho_6^{\beta q_2}\\ & = &\vartheta_{0\frac k6} \left(\tau_3 / 6,(z_2-\tau_3/2-\tau_2/2)/6\right ) \\ &&\quad+(-1)^{\alpha} w_1 e^{2\pi i\left(-\frac{\tau_2}{4}\right)}\vartheta_{0\frac k6}\left(\tau_3 /6,(z_2-\tau_3/2+\tau_2/2)/6 \right). \end{array} $$ To prove that ${\mathcal L}'$ can be extended to a line bundle ${\bar{\mathcal L}'}$ on $\bar A'$ we can argue exactly as in the proof of \cite[Proposition (II.5.13)]{HKW}, the only difference being that we took the partial quotient with respect to a smaller group. (Note that if we take the quotient with respect to $P$ we no longer obtain integer exponents of $t_2$.) The extension of the action of the symmetry group follows as in the proof of \cite[Proposition (II.5.41)]{HKW}. \hfill \end{proof} \section {The map to ${\mathbb{P}}^3$} We consider boundary points $u'=(0,\tau_2, \tau_3) \in B'$ and $u=(0,t_2t_3, t_2^{-1})\in B$ (so $u=\varphi(u')$) and the associated degenerate abelian surfaces $$ {\bar A'}_{u'}={\bar A}_u=A_{u,1}\cup A_{u,2} $$ where $A_{u,1}=A_{u,2}$ is an elliptic ruled surface. We gave a precise description of the surfaces $A_{u,i}$ and the way they are glued in Theorem \ref{theo21}. Recall that $A_{u,i}$ is an elliptic ruled surface with two disjoint sections $E_i, i=1,2$ of self-intersection number $E_i^2=0$. We shall denote the fibre over a point $P$ of the base curve by $f_P$. Recall also the line bundle ${\bar{\mathcal L}}'$ on ${\bar A}'$. We set $$ {\bar{\mathcal L}}_u:={\bar{\mathcal L}}'|_{{\bar A}'_u}={\bar{\mathcal L}}'|_{{\bar A}_u}, \quad {\mathcal L}_{u,i}:={\bar{\mathcal L}}'|_{A_{u,i}}. $$ \noindent In the proof of Proposition \ref{prop23} we computed that $$ \begin{array}{rl} \lim\limits_{t_1\rightarrow 0} {\widehat\Theta}_{\alpha\beta}(\tau,z)=& \vartheta_{0\frac{k}{6}}(\tau_3/6,(z_2-\tau_3/2-\tau_2/2)/6)\\[2mm] &+w^{-1}\vartheta_{0\frac{k}{6}}(\tau_3/6,(z_2-\tau_3/2+\tau_2/2)/6). \end{array} $$ The theta function $\vartheta_{0\frac k6}(\tau_3/6, (z_2-\tau_3/2+\tau_2/2)/6)$ has six zeroes on the elliptic curve $E(\tau_3)={\mathbb{C}}/({\mathbb{Z}} 2\tau_3+\ZZ6)$. Hence $$ \operatorname{deg} {\mathcal L}_{u,i}|_{E_i}=6. $$ Since the exponent of $w$ is $-1$ it follows that $$ \operatorname{deg} {\mathcal L}_{u,i}|_{f_P}=1. $$ (See \cite[Proposition(II.5.35)]{HKW} for similar considerations in the $(1,p)$ case.) Hence $$ {\mathcal L}_{u,i}={\mathcal O}_{A_{u,i}}(E_1+6 f_{P_i}) $$ for a suitable point $P_i\in E(\tau_3)$. (The point $P_i$ can be computed from $\lim\limits_{t_{1\rightarrow 0}}{\widehat\Theta}_{\alpha\beta}(t,z)$ and the normal bundle of $E_1$ in $A_{u,i}$, but we shall not need this later.) Standard arguments using Riemann-Roch show that $$ h^0(A_{u,i}, {\mathcal L}_{u,i})=h^0(A_{u,i}, {\mathcal O}_{A_{u,i}}(E_1+6f_{P_i}))=12. $$ \begin{proposition}\label{prop31} The restriction ${\bar{\mathcal L}}_u={\bar{\mathcal L}}'|_{{\bar A}'_{u'}} ={\bar{\mathcal L}}'|_{{\bar A}_{u}}$ of ${\bar{\mathcal L}}'$ to the degenerate abelian surface ${\bar A}_u$ has the following properties:\\ {\rm (i)} $h^0({\bar A}_u,{\bar{\mathcal L}}_u)=12,$\\ {\rm (ii)} The restriction map $\operatorname{rest}: H^0({\bar A}',{\bar{\mathcal L}}')\rightarrow H^0({\bar A}_u, {\bar{\mathcal L}}_u)$ is surjective.\\ {\rm (iii)} The restriction map $\operatorname{rest}: H^0({\bar A}_u,{\bar{\mathcal L}}_u)\rightarrow H^0 (A_{u,i} {\mathcal L}_{u,i})$ is an isomorphism. \end{proposition} \begin{proof} We consider the space $V\subset H^0({\bar A}',{\bar{\mathcal L}}')$ which is spanned by the twelve sections ${\widehat s}_{\alpha \beta}$. Since these sections are a basis of $H^0(A_{\tau},{\mathcal L}_{\tau})$ for every $\tau$ the space $V$ has dimension~$12$. We claim that the restriction map $$ \mbox{rest}: V\rightarrow H^0({\bar A}_u, {\bar{\mathcal L}}_u) $$ is injective. By our computation of $\lim\limits_{t\rightarrow 0} {\widehat\Theta}_{\alpha \beta}(\tau, z)$ it follows that this map is not identically zero. It is also $H_{26}$-equivariant and hence our claim follows if we can show that $V$ is irreducible as an $H_{26}$--module. But this is easy to see: As an $H_6$-module $V=V_0\oplus V_1$ where $V_i=\mbox{span } (\widehat s_{i\beta}, \beta=0,\ldots, 5)$. The $H_6$--modules $V_0$ and $V_1$ are irreducible, Moreover addition by $e_3/2$ interchanges $V_0$ and $V_1$. Hence $h^0({\bar A}_u, {\bar{\mathcal L}}_u)\ge 12$. We have already remarked that $h^0(A_{u,i},{\mathcal O}_{A_{u,i}}(E_1+6f_{P_i}))=12$. Our next claim is that the map $$ \mbox{rest }: H^0(A_{u,i}, {\mathcal L}_{u,i})\rightarrow H^0(E_1,{\mathcal L}_{u,i}|_{E_1})\oplus H^0(E_2,{\mathcal L}_{u,i}|_{E_2}) $$ is an isomorphism. Since the vector spaces on both side have the same dimension, namely~$12$, it is enough to prove injectivity. This follows from $h^0(A_{u,i},{\mathcal O}_{A_{u,i}}(-E_2+6f_{P_i}))=0$. But now this implies that glueing sections on $A_{u,1}$ and $A_{u,2}$ along $E_1$ and $E_2$ gives at least $2\times 6=12$ conditions. Hence $h^0({\bar A}_u, {\bar{\mathcal L}}_u)\le 12$. With our previous argument this shows that $h^0({\bar A}_u, {\bar{\mathcal L}}_u)=12$ and hence both (i) and (ii) are proved. This also shows that $$ \mbox{rest}: H^0({\bar A}_u,{\bar{\mathcal L}}_u)\rightarrow H^0(A_{u,i}, {\mathcal L}_{u,i}) $$ is an isomorphism and hence we have shown (iii).\hfill \end{proof} Since we are interested in the map to ${\mathbb{P}}^3$ given by $H^0({\mathcal L})^-$ we consider the subspace $$ V^-=\langle {\widehat g}_0, {\widehat g}_1, {\widehat g}_2, {\widehat g}_3\rangle\subset V \subset H^0({\bar A}', {\bar{\mathcal L}}') $$ and $$ \begin{array}{lcl} V^-_{u}&=&\mbox{ rest } (V^-\rightarrow H^0 ({\bar A}_u, {\bar {\mathcal L}}_u)),\\[2mm] V^-_{u,i}&=&\mbox{ rest } (V^-\rightarrow H^0 (A_{u,i}, {\mathcal L}_{u,i})). \end{array} $$ The spaces $V^-_u$ and $V^-_{u,i}$ are $4$--dimensional. We want to study the map $$ \varphi_{V^-_u}:{\bar A}_u--\rightarrow {\mathbb{P}}^3. $$ The sections ${\widehat g}_i$ vanish at the eight ``$2$-torsion'' points $P_1,\ldots ,P_8$ on ${\bar A}_u$ and hence $$ V^-_u\subset H^0(A_{u,i}, {\mathcal O}_{A_{u,i}}(E_1+6f_{P_i}-\sum\limits^8_{j=1} P_j)). $$ Again using restriction to $E_1$ and $E_2$ it follows that the vector space on the right hand side has dimension~$4$ and hence $$ V^-_u=H^0(A_{u,i},{\mathcal O}_{A_{u,i}}( E_1+6f_{P_i}-\sum\limits^8_{j=1} P_j)). $$ We shall first consider the {\em product case}, i.e. $e=[\tau_2]=0$. In this case $A_{u,i}=E(\tau_3)\times {\mathbb{P}}^1$ and there are four rulings which contain two of the points $P_j$ each. These four rulings are, therefore, in the base locus of the linear system $|V^-_u|$. Removing this base locus we obtain the complete linear system of a line bundle on $A_{u,i}$ which has degree~$2$ on the sections $E_i$ and degree $1$ on the fibres. This maps $A_{u,i}$ $2:1$ onto a quadric. Since the map ${\bar A}_u--\rightarrow {\mathbb{P}}^3$ factors through $A_{u,i}$ this shows that the ``Kummer surface'' ${\bar A}_u /\iota_{\omega}$ is mapped $2:1$ onto a quadric. It should be noted that double quadrics arise not only from degenerations of abelian surfaces, but also from special abelian surfaces, namely products (cf. Theorem \ref{theo13}). In fact what happens is that the map from the moduli space ${\mathcal A}_{1,3}(2)$ (or its extension to a toroidal compactification) contracts each Humbert surface parametrizing product surfaces to a double point in $N$ corresponding to a quadric. {}From now on we shall assume $e\neq 0$. Then no fibre of the ruling contains two of the points $P_i$. We have to recall the notion of an {\em elementary transformation} of a ruled surface $S$ at a point $P$. This consists of first blowing up $S$ in $P$ and then blowing down the strict transform of the fibre through $P$. The result is again a ruled surface $\mbox{elm}_{P}S$. Let $$ {\hat A}_u:=\mbox{elm}_{P_1,\ldots,P_8}(A_{u,i}). $$ Then ${\hat A}_u$ has again two disjoint sections $E_1$ and $E_2$ with $E_i^2=0$. In particular $$ {\hat A}_u={\mathbb{P}}({\mathcal O}_{E(\tau_3)}\oplus{\widehat{\mathcal M}}_u) \mbox { for some } {\widehat{\mathcal M}}_u\in \mbox{Pic}^{0}(E(\tau_3)). $$ Note that ${\widehat{\mathcal M}}_u$ and ${{\widehat{\mathcal M}}_u}^{-1}$ define the same ${\mathbb{P}}_1$-bundle. It is straightforward to compute the normal bundle of the sections $E_1$ and $E_2$ in $A_{u,i}$ (cf. \cite[p. 229] {HKW}. Since we can control the self-intersection of a section under blowing up and blowing down and since we know the points $P_j$ it is straightforward to show that ${{\widehat{\mathcal M}}_u}^{\pm 1}={\mathcal O}_{E(\tau_3)}(2[\tau_2]-2[0])$, and hence $$ {\hat A}_u={\mathbb{P}}({\mathcal O}_{E(\tau_3)}\oplus {\mathcal O}_{E(\tau_3)}(2[\tau_2]-2[0])). $$ Consider the diagram $$ \unitlength1pt \begin{picture}(100,60)(0,0) \put(45,55){$\tilde A_u$} \put(16,35){$\pi_1$} \put(75,35){$\pi_2$} \put(43,51){\vector(-1,-1){35}} \put(53,51){\vector(1,-1){35}} \put(0,5){$A_u$} \put(90,5){$\hat A_u$} \end{picture} $$ where $A_u=A_{u,i}$ and ${\tilde A}_u$ is $A_u$ blown up in $P_1,\ldots, P_8$. We denote the exceptional divisors over $P_1,\ldots, P_8$ by $E^1,\ldots, E^8$. The line bundle $\pi^*_1{\mathcal L}_{u,i}\otimes{\mathcal O}_{{\tilde A}_u}(-E^1-\ldots-E^8)$ has degree $0$ on the strict transforms of the fibres through the points $P_j$. Hence $$ {\hat{\mathcal L}}_u:=\pi_{2*}(\pi^*_1{\mathcal L}_{u,i}\otimes {\mathcal O}_{{\tilde A}_u}(-E^1-\ldots-E^8))\in \mbox{Pic } {\hat A}_u $$ is a line bundle on ${\hat A}_u$. The degree of ${\hat{\mathcal L}}_u$ is $1$ on a ruling and $2$ on the sections $E_i$. Hence $$ {\hat{\mathcal L}}_u={\mathcal O}_{{\hat A}_u}(E_1+2f_Q) $$ for a suitable point $Q$ on the base curve. (Clearly $Q$ can be computed explicitly, but this is immaterial for our purposes.) By the usual arguments $$ h^{0}({\hat A}_u, {\hat{\mathcal L}}_u)=4 $$ and the rational map from $A_u$ to ${\hat A}_u$ defines an isormorphism $$ \pi_{2*} \pi^*_1:V^-\cong H^{0}({\hat A}_u, {\hat{\mathcal L}}_u). $$ \begin{proposition}\label{prop32} Let $e\neq 0$. The linear system $|V^-|$ on ${\hat A}_u$ has the following properties:\\ {\rm(i)} $|V^-|$ is base point free.\\ {\rm(ii)} $|V^-|$ maps the two sections $E_i$ each $2:1$ onto two skew lines.\\ {\rm(iii)} $|V^-|$ is very ample outside the sections $E_i$. More precisely, if a cluster $\zeta$ of length~$2$ (i.e. two points or a point and a tangent direction) is not embedded, then $\zeta$ is contained in $E_1$ or $E_2$. \end{proposition} \begin{proof} This follows easily from Reider's theorem. We write $$ |V^-|=|E_1+2f_Q|=|K+L| $$ where the canonical divisor $K=-E_1-E_2$ and $L=2E_1+E_2+2f_Q$. Then $L^2=12$. If $|V^-|$ is not base point free, then there exists a curve $D$ with $L.D=0$ or $1$. Clearly such a curve cannot exist. If $|V ^-|$ fails to embed a cluster $\zeta$ then there exists a curve $D\supset\zeta$ with $L.D=2$ and $D^2=0$. Then $D$ must be a section with $D.E_i=0$. Since ${\widehat{\mathcal M}}_u\neq{\mathcal O}_{E(\tau_3)}$ (here we use $e=2[\tau_2]\neq 0)$ it follows that $D=E_1$ or $D=E_2$. Finally note that the restriction of $|V^-|$ to the elliptic curves $E_i$ gives a complete linear system of degree~$2$. Hence these curves are mapped $2:1$ to lines. Since pairs $(x,y)$ with $x\in E_1$ and $y\in E_2$ are separated, there lines are skew. \hfill \end{proof} We can now summarize our results as follows. \begin{theorem}\label{theo33} Let ${\bar A}_u$ be a corank~$1$ degenerate abelian surface over a point $u=(0,T_2,T_3)\in B$ with $T_2T_3\neq0$ and consider the Kummer map given by the linear system $|V^-|$: $$ \phi_{|V^-|}:{\bar A}_u--\rightarrow{\mathbb{P}}^3. $$ {\rm(i)} If $e=0$ then ${\bar A}_u$ is mapped $4:1$ onto a smooth quadric.\\ {\rm(ii)} If $e\neq 0$ then there is a commutative diagram. $$ \unitlength1pt \begin{picture}(100,60)(0,0) \put(0,46){$\bar A_u$} \put(13,50){\line(1,0){15}} \put(43,50){\line(1,0){15}} \put(73,50){\vector(1,0){20}} \put(100,46){${\mathbb{P}}^3$} \put(13,45){\line(1,-1){10}} \put(27,31){\line(1,-1){10}} \put(12,20){$2:1$} \put(41,17){\vector(1,-1){12}} \put(54,-3){$\hat A_u$} \put(67,8){\vector(1,1){30}} \end{picture} $$ The image of ${\bar A}_u$ is an elliptic ruled surface which has double points along two skew lines but no other singularities. \end{theorem} \begin{proof} By construction the Kummer map ${\bar A}_u--\rightarrow {\mathbb{P}}^3$ factors through ${\bar A}_u/{\iota_{\omega}}=A_{u,i}$. All other statements follow from our above discussion of the map $\phi_{|V^-|}: {\hat A}_u \to {\mathbb{P}}^3$ given by the linear system $|V^-|$. \hfill \end{proof} This theorem explains the irreducible singular quartic surfaces which are parametrized by the S-planes, appearing already in \cite[Section 5-2]{Ni}. We want to conclude this paper with some remarks. We have already seen that the degenerate surfaces with $e=0$ correspond to products. The limits of {\em bielliptic} abelian surfaces are characterised by $2e=0, e\neq 0$. Geometrically this means that the surfaces ${\bar A}_u$ contain degenerate elliptic curves which are $4$-gons, i.e. cycles consisting of $4$ rulings. The degenerate abelian surfaces ${\bar A}_u$ where $u$ is on another boundary component or where more than one of the $T_i$ vanishes can also be described. If $T_3=0$ and $T_1 T_2\neq 0$ then ${\bar A}_u$ is a chain of~$6$ elliptic ruled surfaces. If two of the $T_i$ are zero, then ${\bar A}_u$ consists of $12$ quadrics, whereas ${\bar A}_0$ has $36$ components, of which $24$ are ${\mathbb{P}}^2$ and $12$ are ${\mathbb{P}}^2$ blown up in three points. Limits of polarizations of type $(2,6)$ exist on these surfaces, but as we pointed out before, there is no possibility of defining the Kummer map globally over $B$. On the other hand it is easy to construct degenerations of the ruled surfaces $A_{u,i}$ which lead to a union of two quadrics intersecting along a quadrangle or to a tetrahedron. Finally we want to comment on the boundary components of the Igusa compactification of the moduli space ${\mathcal A}_{1,3}(2)$. These are enumerated by the {\em Tits building} of the group $\Gamma_{1,3}(2)$, i.e. by the equivalence classes modulo $\Gamma_{1,3}(2)$ of the lines and isotropic planes in ${\mathbb{Q}}^4$. The Tits building was calculated by Friedland in \cite{F}: for details, and for some other cases, see \cite{FS}. There are $30$ equivalence classes of lines. These correspond to the $15$ equivalence classes of short, respectively long vectors. Each set of $15$ lines is naturally parametrized by $({\mathbb{Z}}/2)^4 \setminus \{0\}={\mathbb{P}}^3({\mathbb{F}}_2)$. The $15$ planes are parametrized by the $15$ isotropic planes in $\operatorname{Gr}(1,{\mathbb{P}}^3({\mathbb{F}}_2))$. The isotropic planes are a hyperplane section of $\operatorname{Gr}(1,{\mathbb{P}}^3({\mathbb{F}}_2))$ embedded as a quadric via the Pl\"ucker embedding. The $15$ short and the $15$ long vectors as well as the $15$ planes are identified under the group $\Gamma_{1,3}/\Gamma_{1,3}(2)\cong \operatorname{Sp}(4,{\mathbb{F}}_2)\cong S_6$. That is, there are two equivalence classes of lines modulo $\Gamma_{1,3}$ and one plane (see also \cite [Theorem(I.3.40)]{HKW}). Finally the involution $V_3$ of the maximal arithmetic subgroup $\Gamma^*_{1,3}$ identifies short and long vectors (see \cite[Folgerung 3.7]{G} and \cite[Section 2]{HNS}). In our computations above the boundary component given by $T_1=0$ corresponds to a short vector, whereas the boundary component given by $T_3=0$ corresponds to a long vector. We described the degenerate abelian surfaces associated to points on a boundary component correspronding to a short vector. The matrix $V_3$ (and similarly any involution $gV_3$ where $g$ is an element of $\Gamma_{1,3}$ -- cf. \cite[Theorem 2.4]{HNS}) interchanges boundary components associated to short vectors with boundary components associated to long vectors. It should, however, be pointed out that the induced action of $V_3$ on the Igusa compactification ${\mathcal A}_{1,3}^*$ is only a rational map, not a morphism. This follows since the boundary components associated to long, and short vectors are not isomorphic: although their open parts (i.e. away from the corank-$2$ boundary components) are isomorphic (namely to the open Kummer modular surface $K^{0}(1)$), they contain different configurations of rational curves in the corank-$2$ boundary components. This follows from \cite[Satz III.5.19]{B} and \cite[Theorem 4.13]{W}. The degenerate abelian surfaces belonging to points on $T_3=0$ are different from those associated to points on $T_1=0$: they are a cycle of six elliptic ruled surfaces rather than two. At first this looks like a contradiction to \cite[Theorem 2.4]{HNS}, but this is not the case. The polarization on each of the six components of the surface $A_P$ where $P\in \{T_3=0\}$ is of the form ${\mathcal O}(E_1+2f_P)$. On four of the six components the linear system $|V^-|$ has a base locus consisting of a section. These four components are contracted. The other two components are identified under $|V^-|$ and are mapped to a quartic which is an elliptic ruled surface singular along two skew lines. In this way we find the same images in ${\mathbb{P}}^3$ as in the case $T_1=0$. \bibliographystyle{amsalpha}
1997-12-03T00:51:33
9712
alg-geom/9712005
en
https://arxiv.org/abs/alg-geom/9712005
[ "alg-geom", "math.AG" ]
alg-geom/9712005
Janos Kollar
J\'anos Koll\'ar
Real Algebraic Threefolds II: Minimal Model Program
LATEX2e, 61 pages
null
null
null
null
This is the second of a series of papers studying real algebraic threefolds using the minimal model program. The main result is the following. Let $X$ be a smooth projective real algebraic 3-fold. Assume that the set of real points is an orientable 3-manifold (this assumption can be weakened considerably). Then there is a fairly simple description on how the topology of real points changes under the minimal model program. The first application is to study the topology of real projective varieties which are birational to projective 3-space (Nash conjecture). The second application is a factorization theorem for birational maps.
[ { "version": "v1", "created": "Tue, 2 Dec 1997 23:51:33 GMT" } ]
2007-05-23T00:00:00
[ [ "Kollár", "János", "" ] ]
alg-geom
\section{Introduction} In real algebraic geometry, one of the main directions of investigation is the topological study of the set of real solutions of algebraic equations. The first general result was proved in \cite{Nash52}, and later developed by many others (see \cite{AK92} for some recent directions). One of these theorems says that every compact differentiable manifold can be realized as the set of real points of an algebraic variety. \cite{Nash52} posed the problem of obtaining similar results using a restricted class of varieties, for instance rational varieties. For real algebraic surfaces this question was settled in \cite{Comessatti14}. The aim of this series of papers is to utilize the theory of minimal models to investigate this question for real algebraic threefolds. This approach is very similar in spirit to the one employed by \cite{Comessatti14}. (See \cite{Silhol89, ras} for introductions to real algebraic surfaces from the point of view of the minimal model program.) For algebraic threefolds over $\c$, the minimal model program (MMP for short) provides a very powerful tool. The method of the program is the following. (See \cite{koll87, CKM88} or \cite{KM98} for introductions.) Starting with a smooth projective 3-fold $X$, we perform a series of ``elementary" birational transformations $$ X=X_0\map X_1\map \cdots \map X_n=:X^* $$ until we reach a variety $X^*$ whose global structure is ``simple". (Neither the intermediate steps $X_i$ nor the final $X^*$ are uniquely determined by $X$.) In essence the minimal model program allows us to investigate many questions in two steps: first study the effect of the ``elementary" transformations and then consider the ``simple" global situation. In practice both of these steps are frequently rather difficult. For instance, we still do not have a complete list of all possible ``elementary" steps, despite repeated attempts to obtain it. A somewhat unpleasant feature of the theory is that the varieties $X_i$ are not smooth, but have so called terminal singularities. This means that $X_i(\r)$ is not necessarily a manifold. In developing the theory of minimal models for real algebraic threefolds, we again have to understand the occurring terminal singularities. This was done in the first paper of this series \cite{rat1}. If $X$ is defined over a field $K$, then there is a variant of the MMP where the intermediate varieties $X_i$ are also defined over $K$. I refer to this as the MMP over $K$. This suggests the following two step approach to understand the topology of $X(\r)$: \begin{enumerate} \item Study the topological effect of the ``elementary" transformations. \item Investigate the topology of $X^*(\r)$. \end{enumerate} \noindent The aim of this paper is to complete the first of these two steps. I am unable to say much about this question in general. There are serious problems coming from algebraic geometry and also from 3-manifold topology. Some of these are discussed in section 4. My aim is therefore more limited: find reasonable conditions which ensure that the steps of the MMP can be described topologically. The simplest case to study is contractions $f:X\to Y$ where $X$ is smooth. Over $\c$ the complete list of such contractions is known \cite{Mori82}, and it is not hard to obtain a complete list over $\r$. From this list one can see that in all such examples where $X(\r)\to Y(\r)$ is complicated, $X(\r)$ contains a special surface of nonnegative Euler characteristic. This turns out to be a general pattern, though the proof presented here relies on a laborious case analysis. The precise technical theorem is stated in (\ref{int.nonorient.thm}). None of the complicated examples occur if $X(\r)$ is orientable, and this yields the following: \begin{thm}\label{int.orient.thm} Let $X$ be a smooth, projective, real algebraic $3$-fold and $X^*$ the result of the MMP over $\r$. Assume that $X(\r)$ is orientable. Then the topological normalization $\overline{X^*(\r)}$ of $X^*(\r)$ is a PL-manifold, and $X(\r)$ can be obtained from $\overline{X^*(\r)}$ by repeated application of the following operations: \begin{enumerate} \setcounter{enumi}{-1} \item throwing away all isolated points of $\overline{X^*(\r)}$, \item taking connected sums of connected components, \item taking connected sum with $S^1\times S^2$, \item taking connected sum with $\r\p^3$. \end{enumerate} \end{thm} \begin{rem} $X^*$ uniquely determines (\ref{int.orient.thm}.0) and also (\ref{int.orient.thm}.1). The latter can be seen by analyzing real analytic morphisms $h:[0,1]\to X^*(\r)$ where the endpoints map to different connected components of $\overline{X^*(\r)}$. In practice this may be quite hard, and it could be easier to work through the MMP backwards. $X^*$ contains some information about the steps (\ref{int.orient.thm}.2--3), but these are by no means unique. Even if $X^*$ is smooth, both of these steps are possible, as shown by the next example. \end{rem} \begin{exmp}\label{connsum.exmp} It is well known how to create connected sum with $\r\p^3$ algebraically. Let $X$ be a smooth 3-fold over $\r$ and $0\in X(\r)$ a real point. Set $Y=B_0X$. Then $Y(\r)\sim X(\r)\ \#\ \r\p^3$. (The connected sum of two nonoriented manifolds is, in general, not unique. It is, however, unique if one of the summands has an automorphism with an isolated fixed point which reverses local orientation there.) Connected sum with $S^1\times S^2$ is somewhat harder. Let $X$ be a smooth 3-fold over $\r$ and $D\subset X$ a real curve which has a unique real point $\{0\}= D(\r)$. Assume furthermore that near $0$ the curve is given by equations $(z=x^2+y^2=0)$. Set $Y_1=B_DX$. $Y_1$ has a unique singular point $P$; set $Y=B_PY_1$. It is not hard to see that $Y$ is smooth and $Y(\r)\sim X(\r)\ \#\ (S^1\times S^2)$. \end{exmp} \begin{rem}\label{1.basic.top.facts} As (\ref{int.orient.thm}) already shows, we have to move between topological, PL and differentiable manifolds. In dimension 3 every compact topological 3--manifold carries a unique PL--manifold structure (cf.\ \cite[Sec.\ 36]{Moise77}) and also a unique differentiable structure (cf.\ \cite[p.3]{Hempel76}). I mostly use the PL--structure since most algebraic constructions are natural in the PL--category. For instance, $\r^1\to \r^2$ given by $t\mapsto (t^2,t^3)$ is a PL--embedding but not a differentiable embedding in the natural differentiable structures. In dimension 3 the PL--structure behaves very much like a differentiable structure. For instance, let $M^3$ be a PL 3--manifold, $N$ a compact PL--manifold of dimension 1 or 2 and $g:N\DOTSB\lhook\joinrel\rightarrow M$ a PL--embedding. Then a suitable open neighborhood of $g(N)$ is PL--homeomorphic to a real vector bundle over $N$ (cf.\ \cite[Secs.\ 24 and 26]{Moise77}). (Note that a similar result fails for topological 3--manifolds (cf.\ \cite[Sec.\ 18]{Moise77}), and it also fails for PL 4--manifolds: take any nontrivial knot in $S^3$ and suspend it in $S^4$.) \end{rem} \begin{say}[Surfaces in 3--manifolds]\label{surf.in.3-man} Let $M$ be a PL 3--manifold without boundary, $N$ a compact PL 2--manifold without boundary and $g:N\DOTSB\lhook\joinrel\rightarrow M$ a PL--embedding. As we noted above, a neighborhood of $N$ is an $\r$-bundle over $N$. $\r$-bundles over $N$ are classified by group homomorphisms $\rho:\pi_1(N)\to \{\pm 1\}$. If $\rho $ is trivial then $N$ is 2--sided in $M$, otherwise it is 1--sided. We also allow self homeomorphisms of $N$, thus we get the following possibilities when $N$ has nonnegative Euler characteristic: \begin{description} \item[$S^2$] Always 2--sided, many such surfaces in every $M^3$. \item[$\r\p^2$] $M^3$ is not orientable in the 2--sided case. Such manifolds are called $\p^2$-reducible (cf.\ \cite[p.88]{Hempel76}). In the 1--sided case the boundary of a regular neighborhood is $S^2$, thus $M\sim M'\ \#\ \r\p^3$ for some 3--manifold $M'$. Most 3--manifolds do not contain any $\r\p^2$. \item[{\rm Torus}] The 2--sided case occurs in any 3--manifold as the boundary of a regular neighborhood of any $S^1$ along which $M$ is orientable. There is a unique 1--sided case. For these $M$ is not orientable. Most nonorientable 3--manifolds do not contain 1--sided tori, see section 12. \item[{\rm Klein bottle}] $M$ is nonorientable in the 2--sided case. The boundary of a regular neighborhood of any $S^1$ along which $M$ is nonorientable is such. There are two different 1--sided cases, depending on whether $M$ is orientable near $N$ or not. These are again rare, see section 12. \end{description} \noindent This shows that there are many 3--manifolds which do not contain $\r\p^2$, 1--sided tori or Klein bottles. These correspond to 6 different cases on the above list. It turns out that we need to exclude only 3 of these for our main theorem. \end{say} \begin{condition}\label{int.no.cond} Let $M$ be a PL 3-manifold without boundary. Consider the following properties: \begin{enumerate} \item $M$ does not contain a 2-sided $\r\p^2$, \item $M$ does not contain a 1-sided torus, \item $M$ does not contain a 1-sided Klein bottle with nonorientable neighborhood. \end{enumerate} Failure of any of these properties implies that $M$ is not orientable, but there are many nonorientable 3-manifolds which do satisfy all 3 of the above conditions. For instance, this holds if $M$ is hyperbolic (\ref{hyp.doesnotcont.thm}). \end{condition} \begin{thm}\label{int.nonorient.thm} Let $X$ be a smooth, projective, real algebraic $3$-fold and $X^*$ the result of the MMP over $\r$. Assume that $X(\r)$ satisfies the 3 conditions (\ref{int.no.cond}.1--3). Then the conclusions of (\ref{int.orient.thm}) hold. \end{thm} \begin{rem} It would seem that we also need to allow connected sum with $S^1\tilde{\times} S^2$ (cf.\ (\ref{5.notation})), corresponding to attaching a nonorientable 1--handle. This, however, would give a 1--sided torus which we excluded. All 3 conditions (\ref{int.no.cond}.1--3) are necessary for the theorem to hold. My feeling is that essentially nothing can be said without (\ref{int.no.cond}.1) or (\ref{int.no.cond}.3). (\ref{int.no.cond}.2) has a twofold role in the proof. First, it ensures that $X$ is not obtained as a blow up of a smooth 3-fold $Y$ along a curve. This in itself would not be a problem, but it may happen that $Y(\r)$ contains a 2-sided $\r\p^2$ but $X(\r)$ does not. It seems to me that this leads to rather complicated topological questions. Still, a suitable reformulation of the theorem may get around this problem. Second, (\ref{int.no.cond}.2) is also used to exclude a few singularities on the $X_i$. These cases are of index 1 and they can be described very explicitly. It should be possible to work with them. \end{rem} The technical heart of the proof is a listing of the possible singularities that occur in the course of the MMP and a fairly detailed description of the steps of the MMP. The final result is relatively easy to state but the proof is a case-by-case examination. \begin{thm}\label{int.mmp.sings} Let $X$ be a smooth, projective, real algebraic $3$-fold and assume that $X(\r)$ satisfies the 3 conditions (\ref{int.no.cond}.1--3). Let $X_i$ be any of the intermediate steps of the MMP over $\r$ starting with $X$ and $0\in X_i(\r)$ a real point. Then a neighborhood of $0\in X_i$ is real analytically equivalent to one of the following standard forms: \begin{enumerate} \item ($cA_0$) Smooth point. \item ($cA_{>0}^+$) $(x^2+y^2+g_{\geq 2}(z,t)=0)$, where $g$ is not everywhere negative in a punctured neighborhood of $0$. \item ($cE_6$) $(x^2+y^3+(z^2+t^2)^2+ yg_{\geq 4}(z,t)+g_{\geq 6}(z,t)=0)$. \end{enumerate} \end{thm} \begin{rem} The symbol $g_{\geq m}$ denotes a power series of multiplicity at least $m$. The name of the cases is explained in \cite{rat1}. The above points of type $cE_6$ form a codimension 7 family in the space of all $cE_6$ singularities. They all occur, even if $X(\r)$ is orientable. Points of type $cA_{>0}^+$ occur for many choices of $g$. Section 10 gives an algorithm to decide which cases of $g$ do occur, but I was unable to write the condition in closed form. For the applications this does not seem to matter. \end{rem} Using \cite[4.3, 4.4, 4.9]{rat1}, this immediately implies: \begin{cor}\label{int.mmp.sings.top} Notation and assumptions as in (\ref{int.mmp.sings}). Then $\overline{X_i(\r)}\setminus\{\mbox{isolated points}\}$ is a compact PL 3-manifold without boundary.\qed \end{cor} The next step is to understand the ``elementary" steps of the MMP over $\r$. (\ref{int.nonorient.thm}) turns out to be a consequence of (\ref{int.mmp.steps}). (See (\ref{8.wbup}) for the definition of weighted blow-ups.) \begin{thm}\label{int.mmp.steps} Let $X$ be a smooth, projective, real algebraic $3$-fold such that $X(\r)$ satisfies the conditions (\ref{int.no.cond}.1--3). Let $f_i:X_i\map X_{i+1}$ be any of the intermediate steps of the MMP over $\r$ starting with $X$. Then the induced map $f_i:X_i(\r)\to X_{i+1}(\r)$ is everywhere defined and the following is a complete list of possibilities for $f_i$: \begin{enumerate} \item ($\r$-trivial) $f_i$ is an isomorphism in a (Zariski) neighborhood of the set of real points. \item ($\r$-small) $f_i:X_i(\r)\to X_{i+1}(\r)$ collapses a 1-complex to points and there are small perturbations $\tilde f_i$ of $f_i$ such $\tilde f_i: \overline{X_i(\r)}\to \overline{X_{i+1}(\r)}$ is a PL-homeomorphism. \item (smooth point blow up) $f_i$ is the inverse of the blow up of a smooth point $P\in X_{i+1}(\r)$. \item (singular point blow up) $f_i$ is the inverse of a (weighted) blow up of a singular point $P\in X_{i+1}(\r)$. There are two cases: \begin{enumerate} \item ($cA_{>0}^+$, $\mult_0g$ even) Up to real analytic equivalence near $P$, $X_{i+1}\cong (x^2+y^2+g_{\geq 2m}(z,t)=0)$ where $g_{2m}(z,t)\neq 0$, $m\geq 1$ and $X_i$ is the weighted blow up $B_{(m,m,1,1)}X_{i+1}$. \item ($cA_{>0}^+$, $\mult_0g$ odd) Up to real analytic equivalence near $P$, $X_{i+1}\cong (x^2+y^2+g_{\geq 2m+1}(z,t)=0)$ where $m\geq 1$, $z^{2m+1}\in g$ and $z^it^j\not\in g$ for $2i+j< 4m+2$. $X_i$ is the weighted blow up $B_{(2m+1,2m+1,2,1)}X_{i+1}$. \end{enumerate} \end{enumerate} \end{thm} \begin{rem} The more precise results in sections 9--11 give a description of the various cases when $f_i$ is $\r$-small (though so far I have not excluded some cases). The $\r$-trivial steps do not change anything in a neighborhood of the real points, but it is in these steps that the full complexity of the MMP appears. All the difficulties involving higher index terminal singularities and flips are present, but they always appear in conjugate pairs. For the topological questions these have no effect, but in other applications of (\ref{int.mmp.steps}) this should be taken into account. \end{rem} \begin{rem} The lists in (\ref{int.mmp.sings}) and (\ref{int.mmp.steps}) are fairly short, but I do not see a simple conceptual way of stating the results, let alone proving them by general arguments. The appearence of the singularities of type $cE_6$ in (\ref{int.mmp.sings}) was rather unexpected for me. The formulations also hide the cicumstance that there does not seem to be a single method of excluding all other a priori possible cases. The algebraic method of the proof of (\ref{int.mmp.sings}) ends with a much longer list (\ref{ge.gwextr.thm}). The topological method excludes many of these right away, but in a few cases several steps of the MMP need to be analyzed. \end{rem} \begin{say}[Method of the proof of (\ref{int.mmp.steps})]{\ } The proof relies on rather extensive computations. The first step is a classification of all 3--dimensional terminal singularities over $\r$ and the study of their topological properties. This was carried out in \cite{rat1}. The next step is to gain a good understanding of the resolutions of these singularities. More precisely, we need to understand the ``simplest" exceptional divisors in these resolutions. (Simplicity is measured by the discrepancy, cf.\ (\ref{mmp.discr.def}).) Over $\c$ the first step in this direction is \cite{Markushevich96}. A much more detailed study of such exceptional divisors was completed by \cite{Hayakawa97}. Our main emphasis is over $\r$, and it turns out that there is very little overlap between the computations of \cite{Hayakawa97} and those in sections 9--11. Nonetheless, the basic underlying principles are exactly the same. \end{say} \begin{ack} I thank M. Bestvina, S. Gersten, M. Kapovich and G. Mikhalkin for answering my numerous questions about 3-manifold topology and real algebraic geometry. The existence of $cE_6$ type points in (\ref{int.mmp.sings}) was established with the help of V. Alexeev. I have received helpful comments and questions from A. Bertram, M. Fried, L. Katzarkov and B. Mazur. Partial financial support was provided by the NSF under grant number DMS-9622394. \end{ack} \section{Applications and Speculations} \subsection{Factorization of Birational Morphisms} Let $f:Y\to X$ be a birational morphism between smooth and projective varieties. It is a very old problem to factor $f$ as a composite of ``elementary" birational morphisms. In dimension 2 this is easy to do: $f$ is the composite of blow ups of points. In dimension 3 and over $\c$, the MMP factors $f$ as a composition of divisorial contractions and flips, but these intermediate steps are rather complicated and not too well understood. If $f:Y\to X$ is a birational morphism between smooth and projective threefolds over $\r$, then one would like to get a factorization where the intermediate steps are also defined over $\r$. It turns out that if $Y(\r)$ is orientable, the answer is very simple. As with minimal models in general, the intermediate steps involve singular varieties though in this case the real singularites are very mild. \begin{defn} A real 3--fold $X$ is said to have a $cA_1$ singularity at $0\in X(\r)$ if in suitable real analytic cordinates $X$ can be given by an equation $(\pm x^2\pm y^2\pm z^2\pm t^m=0)$ for a suitable choice of signs and $m\geq 1$. \end{defn} \begin{thm}\label{bir.morph.factor} Let $f:Y\to X$ be a birational morphism between smooth and projective threefolds over $\r$. Assume that $Y(\r)$ satisfies the conditions (\ref{int.no.cond}). Then $f$ can be factored as $$ f: Y=X_n\stackrel{f_n}{\to} X_{n-1}\to \cdots\to X_1 \stackrel{f_1}{\to} X_0=X, $$ where each $X_i$ has only $cA_1$ singularities at real points and the following is a complete list of possibilities for the $f_i$: \begin{enumerate} \item (smooth point blow up) $f_i$ is the blow up of a smooth point $P\in X_{i-1}(\r)$. \item (singular point blow up) $f_i$ is the blow up of a singular point $P\in X_{i-1}(\r)$. \item (curve blow up) $f_i$ is the blow up of a real curve $C\subset X_{i-1}$. $C$ has only finitely many real points, $X_{i-1}$ is smooth at each of these and in suitable real analytic coordinates $C$ can be written as $(z=x^2+y^{2m}=0)$. \item ($\r$-trivial) $f_i$ is an isomorphism in a (Zariski) neighborhood of the set of real points. \end{enumerate} \end{thm} \begin{rem} As in (\ref{int.mmp.steps}), it is in the $\r$-trivial steps that the full complexity of the MMP appears. In particular, the $\r$-trivial steps may be flips where the flipping curve has no real points. \end{rem} Proof. For purposes of induction we consider the more general case when $X$ is allowed to have $cA_1$-type singularities at real points and terminal singularities at complex points. We assume that $X$ is $\q$-factorial (that is, a suitable multiple of every Weil divisor is Cartier). Run the real MMP for $Y$ over $X$ to obtain $$ f: Y=X_n\stackrel{f_n}{\map} X_{n-1}\map \cdots\map X_1 \stackrel{f_1}{\to} X_0=X. $$ The proof is by induction on the number of steps it takes the MMP to reach $X$. The last step, $f_1:X_1\to X_0=X$, is a contraction since we work over $X$. The possibilities for $f_1$ are described in (\ref{int.mmp.steps}). We are done by induction if $f_1$ is $\r$-trivial or a smooth point blow up. Assume that $f_1$ is a singular point blow up. Since $X_0$ has only $cA_1$ points, we are in case (\ref{int.mmp.steps}.4a) with $m=1$. $f_1$ is the ordinary blow up and by explicit computation we see that $X_1$ still has only $cA_1$ singularities. The case when $f_1$ is $\r$-small (\ref{int.mmp.steps}.2) needs to be studied in greater detail. $f_1$ can not be a g--extraction (\ref{gw.g-e.def}) since $cA_1$ type points do not have g--extractions other than the one listed above by (\ref{ge.gwextr.thm}). Thus $f_1$ is the blow up of a curve $C\subset X_0$. Moreover, $X_0$ is smooth along $C(\r)$ and $C$ is locally planar along $C(\r)$ by (\ref{sm.thm}). $C(\r)$ is finite since $f_1$ is $\r$-small. Pick any point $P\in C(\r)$ and assume that $C$ is given by real analytic equations $(z=g(x,y)=0)$. By explicit computation, $B_CX_0$ has a unique singular point with equation $(st-g(x,y)=0)$ which is equivalent to $(u^2-v^2-g(x,y)=0)$. $X_1$ is an intermediate step of an MMP starting with $Y$, hence its singularities are among those listed (\ref{int.mmp.sings}). Thus $g$ has multiplicity 2 and so it can be written as $\pm x^2\pm y^r$. Since $(g=0)$ has only the origin as its real solution, $g=\pm(x^2+y^{2m})$. \qed \subsection{Application to the Nash Conjecture} The main conclusion of (\ref{int.orient.thm}) and (\ref{int.nonorient.thm}) is that if we want to understand the topology of $X(\r)$ (say when it is orientable), it is sufficient to study the topology of $X^*(\r)$ instead. $X^*$ has various useful properties, depending on the conditions imposed on $X$. Consider, for instance, the original Nash question: what happens if $X$ is rational. Since the fifties it has been understood that being rational is a very subtle condition and it is very hard to work with. \cite{KoMiMo92} introduced the much more general notion of being {\it rationally connected}. A $X$ is rationally connected if two general points of $X(\c)$ can be connected by an irreducible rational curve. The lines show that $\p^n$ is rationally connected. The structure theory of \cite{KoMiMo92} implies that a 3-fold $X$ is rationally connected iff $X^*$ falls in one of 3 classes: \begin{enumerate} \item (Conic fibrations) There is a morphism (over $\r$) $g:X^*\to S$ onto a surface such that the general fiber is a conic. Correspondingly there is a morphism $X^*(\r)\to S(\r)$ whose general fiber is $S^1$ or empty. These cases will be studied in a subsequent paper. \item (Del Pezzo fibrations) There is a morphism (over $\r$) $g:X^*\to C$ onto a curve such that the general fiber is a Del Pezzo surface. If $X(\r)$ is orientable, then this induces a morphism $X^*(\r)\to C(\r)$ whose general fiber is a torus or a union of some copies of $S^2$. These cases will be studied later. \item (Fano varieties) The anticanonical bundle of $X^*$ is ample. There is a complete list of such varieties if $X^*$ is also smooth \cite{Iskovskikh80}. Even if $X^*$ is known rather explicitly, a topological description of $X^*(\r)$ may not be easy. It would be interesting to work out at least some of the cases, for instance hypersurfaces of degree 3 or 4 in $\p^4$. (Mikhalkin pointed out that the degree 3 cases can be understood using the classification of degree 4 real surfaces in $\r\p^3$ \cite{Kharlamov76}.) In general it is known that there are only finitely many families of singular Fano varieties in dimension 3 \cite{Kawamata92}. Thus we can get only finitely many different topological types for $X^*(\r)$ in this case. \end{enumerate} \subsection{Homology Spheres} It is interesting to consider if we can get further simplifications of the real MMP if we pose further restrictions on $X(\r)$. We may assume, for instance, that $X(\r)$ is a homology sphere. This was in fact the assumption I considered first. One can ask if under this assumption $X(\r)\to X^*(\r)$ is a homeomorphism. Unfortunately this is not the case. Consider for instance the singular real threefold $X^*$ given by affine equation $$ x^2+y^2+z^2+(t-a_0)(t-a_m)\prod_{i=1}^{m-1}(t-a_i)^{2r}=0, $$ where $a_0<a_1<\cdots<a_m$ are reals. This has $m-1$ singular points of the form $x^2+y^2+z^2-u^{2r}=0$, which can be resolved by $r$ successive blow ups. Resolving all singular points we obtain the 3--fold $X$. One can easily see that $X(\r)\sim S^3$, but $\overline{X^*(\r)}$ is the disjoint union of $m$ copies of $S^3$. One may also study the types of singularities that occur if we pose stronger restrictions on $X(\r)$. It seems to me that the best one can get is the following: \begin{conj}\label{no.E6.conj} Let $X$ be a smooth projective 3--fold over $\r$. Assume that $X(\r)$ satisfies the conditions (\ref{int.no.cond}) and $X(\r)$ can not be written as a connected sum with $S^1\times S^2$. Let $X_i$ be any of the intermediate steps of the MMP over $\r$ starting with $X$ and $0\in X_i(\r)$ a real point. Then a neighborhood of $0\in X_i$ is real analytically equivalent to one of the following standard forms: \begin{enumerate} \item ($cA_0$) Smooth point. \item ($cA_{>0}^+$) $(x^2+y^2+g_{\geq 2}(z,t)=0)$, where $\mult_0g$ is even and $g$ is not everywhere negative in a punctured neighborhood of $0$. \end{enumerate} \end{conj} In fact, most $cA_{>0}^+$-type singularities should not occur. It is possible that one can write down a complete list. Also, one can be more precise about how the singular points separate $X_i(\r)$. The results in sections 8--11 come close to proving (\ref{no.E6.conj}), but two points remain unresolved. In order to exclude $cE_6$ type points, one needs to show that the only possible g--extraction is the one described in (\ref{cE6.g--extr.exist}). This should be a feasible computation. The main problem is that in (\ref{cA+.multg-odd}) I could not exclude certain $\r$-small contractions. I do not see how to deal with this case. \subsection{Beyond the Nash Conjecture} One can refine the 3--dimensional Nash conjecture in two ways. \medskip First, one can study the topology of $X(\r)$ for other classes of real algebraic varieties. The simplest cases may be those whose minimal models admit a natural fibration. This should be very helpful in their topological study. One such class is elliptic threefolds, where we have a morphism $X^*\to S$ whose general fiber is an elliptic curve. A study of the singular fibers occurring in codimension 1 was completed by \cite{Silhol84}. Another, probably more difficult class are Calabi--Yau 3--folds. It would be very interesting to find some connection between the topology of $X(\r)$ and mirror symmetry. The following question is consistent with the examples that I know: \begin{question} Let $X$ be a smooth projective real 3--fold. Asume that $X(\r)$ is hyperbolic. Does this imply that $X$ is of general type? \end{question} \medskip One can also start with a 3--manifold $M$ and look for a ``simple" real projective 3--fold $X$ such that $X(\r)\sim M$. Ideally one would like to find a solution where certain topological structures on $M$ are reflected by the algebraic properties of $X$. There are hyperbolic 3--manifolds which embedd into $\r^4$. Ths implies that they can be realized by real algebraic hypersurfaces in $\r^4$. It would be interesting to find such examples. \medskip The methods of this paper require a very detailed study of the steps of the MMP, which is currently feasible only in dimension 3. It would be, however, interesting to develop some examples in higher dimensions. Example (\ref{connsum.exmp}) describing connected sum with $S^1\times S^2$ should have interesting higher dimensional versions. There may be other, more complicated examples as well. The first steps of the 4--dimensional MMP over $\c$ have been recently classified by \cite{andwis}. It should be possible to obtain the complete list over $\r$ and to study their topology. \section{The Minimal Model Program over $\r$} This section is intended to provide a summary of the MMP over $\r$. More generally, I discuss the MMP over an arbitrary field $K$ of characteristic zero, since there is no difference in the general features. Conjecturally the whole program works in all dimensions but at the moment it is only established in dimensions $\leq 3$. \cite{koll87, koll90} provide general introductions. The minimal model program for real algebraic surfaces is explained in detail in \cite{ras}. For more comprehensive treatments (mostly over $\c$) see \cite{CKM88, kolletal92, KoMo98}. One of the special features of the 3-dimensional MMP is that we have to work with certain singular varieties in the course of the program. \begin{defn}\label{mmp.qf.def} Let $X$ be a normal variety defined over a field $K$. A {\it (Weil) divisor} over $K$ is a formal linear combination $D:=\sum a_iD_i$ ($a_i\in \z$) of codimension 1 subvarieties, each defined and irreducible over $K$. A {\it $\q$-divisor} is defined similarly, except we allow $a_i\in \q$. A divisor $D$ is called {\it Cartier} if it is locally definable by one equation and {\it $\q$-Cartier} if $mD$ is Cartier for some $m\in\n$. The smallest such $m>0$ is called the {\it index} of $D$. We say that $X$ is {\it factorial} (resp. {\it $\q$-factorial}) if every Weil divisor is Cartier (resp. $\q$-Cartier). A divisor $D$ defined over $K$ is Cartier (resp. $\q$-Cartier) iff it is Cartier (resp. $\q$-Cartier) after some field extension. However, a variety may be $\q$-factorial over $K$ and not $\q$-factorial over $\bar K$. For instance, the cone $x^2+y^2+z^2-t^2$ is factorial over $\r$ but not over $\c$. (For instance, $(x-\sqrt{-1}y=z-t=0)$ is a not $\q$-Cartier.) \end{defn} \begin{defn}\label{mmp.KX.def} For a normal variety $X$, let $K_X$ denote its {\it canonical class}. $K_X$ is a linear equivalence class of Weil divisors. The corresponding reflexive sheaf $\o_X(K_X)$ is isomorphic to the {\it dualizing sheaf} $\omega_X$ of $X$. The index of $K_X$ is called the {\it index } of $X$. \end{defn} \begin{defn}\label{mmp.discr.def} Let $X,Y$ be normal varieties and $f:Y\to X$ a birational morphism with exceptional set $\ex(f)$. Let $E_i\subset \ex(f)$ be the exceptional divisors. If $mK_X$ is Cartier, then $f^*\o_X(mK_X)$ is defined and there is a natural isomorphism $$ f^*\o_X(mK_X)|(Y\setminus \ex(f))\cong \o_Y(mK_Y)|(Y\setminus \ex(f)). $$ Hence there are integers $b_i$ such that $$ \o_Y(mK_Y)\cong f^*\o_X(mK_X)(\sum b_iE_i). $$ Formally divide by $m$ and write this as $$ K_Y\equiv f^*(K_X)+\sum a(E_i, X)E_i,\qtq{where $a(E_i,X)\in \q$.} $$ The rational number $a(E_i,X)$ is called the {\it discrepancy} of $E_i$ with respect to $X$. The closure of $f(E_i)\subset X$ is called the {\it center} of $E_i$ on $X$. It is denoted by $\cent_XE_i$. If $f':Y'\to X$ is another birational morphism and $E'_i:=(f'\circ f^{-1})(E_i)\subset Y'$ is a divisor then $a(E'_i,X)=a(E_i,X)$ and $\cent_XE_i=\cent_XE'_i$. Thus the discrepancy and the center depend only on the divisor up to birational equaivalence, but not on the particular variety where the divisor appears. \end{defn} \begin{defn} Let $X$ be a normal variety such that $K_X$ is $\q$-Cartier. We say that $X$ is {\it terminal} (or that it has {\it terminal singularities}) if for every $f:Y\to X$, the discrepancy of every exceptional divisor is positive. \end{defn} The following result makes it feasible to decide if $X$ is terminal or not. \begin{lem}\label{mmp.term.lem} For a normal variety $X$ the following are equivalent: \begin{enumerate} \item $X$ is terminal, \item $a(E,X)>0$ for every resolution of singularities $f:Y\to X$ and for every exceptional divisor $E\subset \ex(f)$. \item There is a resolution of singularities $f:Y\to X$ such that $a(E,X)>0$ for every exceptional divisor $E\subset \ex(f)$.\qed \end{enumerate} \end{lem} \begin{exmp}\label{mmp.comp.discr.exmp} It is frequently not too hard to compute discrepancies. Assume for instance that $X$ is a hypersurface defined by $(F(x_1,\dots,x_n)=0)$. A local generator of $\o_X(K_X)$ is given by any of the forms $$ \eta_i:=\frac1{\partial F/\partial x_i}dx_1\wedge\cdots\wedge dx_{i-1}\wedge dx_{i+1}\wedge\cdots\wedge dx_n. $$ Let $f:Y\to X$ be a resolution of singularities and $P\in Y$ a point with local coordinates $y_1,\dots,y_{n-1}$. $f$ is given by coordinate functions $x_i=f_i(y_1,\dots,y_{n-1})$ and so we can write \begin{eqnarray*} f^*\eta_n&=&f^*\left(\frac1{\partial F/\partial x_n}\right) \operatorname{Jac} dy_1\wedge\cdots\wedge dy_{n-1}, \qtq{where}\\ \operatorname{Jac}&=& \operatorname{Jac} \left(\frac{f_1,\dots,f_{n-1}}{x_1,\dots,x_{n-1}}\right) \end{eqnarray*} denotes the determinant of the Jacobian matrix. Hence the discrepancies can be computed as the order of vanishing of the Jacobian minus the order of vanishing of $f^*(\partial F/\partial x_n)$. If $X$ is smooth then we conclude that $a(E,X)\geq 1$ for every exceptional divisor. Thus smooth varieties are terminal. \end{exmp} Next we define various birational maps which have special role in the MMP. \begin{defn}\label{mmp.extremal.def} Let $X$ be a variety over $K$ and assume that $K_X$ is $\q$-Cartier. A proper morphism $g:X\to Y$ is called an {\it extremal contraction} if the following conditions hold: \begin{enumerate} \item $g_*\o_X=\o_Y$, \item $X$ is $\q$-factorial, \item Let $C\subset X$ be any irreducible curve such that $g(C)=\mbox{point}$. Then a $\q$-divisor $D$ on $X$ is the pull back of a $\q$-Cartier $\q$-divisor $D'$ on $Y$ iff $(D\cdot C)=0$. (Necessarily, $D'=g_*(D)$.) \end{enumerate} \end{defn} \begin{defn}\label{mmp.excontypes.def} Let $g:X\to Y$ be an extremal contraction. We say that $g$ is of {\it fiber type} if $\dim Y<\dim X$. We say that $g$ is a {\it divisorial} contraction if the exceptional set $\ex(g)$ is the support of $\q$-Cartier divisor. In this case $\ex(g)$ is irreducible over $K$. We say that $g$ is a {\it small} contraction if $\dim \ex(g)\leq \dim X-2$. One can see that every extremal contraction is in one of these 3 groups. \end{defn} \begin{defn}\label{mmp.KXneg.def} A proper morphism $f:X\to Y$ is called {\it $K_X$-negative} if $-K_X$ is $f$-ample. \end{defn} \begin{defn}\label{mmp.flip.def} Let $f: X \to Y$ be a small $K_X$-negative extremal contraction. A variety $X^+$ together with a proper birational morphism $f^+: X^+ \to Y$ is called a {\it flip} of $f$ if \begin{enumerate} \item $K_{X^+}$ is $\q$-Cartier, \item $K_{X^+}$ is $f^+$-ample, and \item the exceptional set $\ex(f^+)$ has codimension at least two in $X^+$. \end{enumerate} By a slight abuse of terminology, the rational map $\phi: X \map X^+$ is also called a flip. A flip gives the following diagram: $$ \begin{array}{rcl} X &\stackrel{\phi}{\map} &X^+\\ \mbox{$-K_{X}$ is $f$-ample} &\searrow \quad\swarrow & \mbox{$K_{X^+}$ is $f^+$-ample}\\ &Y& \end{array} $$ It is not hard to see that a flip is unique and the main question is its existence. \end{defn} We are ready to state the 3-dimensional MMP over an arbitrary field: \begin{thm}[MMP over $K$]\label{mmp.mmp.thm} Let $X$ be a smooth projective 3-fold defined over a field $K$ (of characteristic zero). Then there is a sequence $$ X=X_0\stackrel{f_0}{\map} X_1\map \cdots \map X_i \stackrel{f_i}{\map} X_{i+1}\map \cdots \stackrel{f_{n-1}}{\map} X_n=:X^* $$ with the following properties \begin{enumerate} \item Each $X_i$ is a terminal projective 3-fold over $K$ which is $\q$-factorial over $K$. \item Each $f_i$ is either a $K_X$-negative divisorial extremal contraction or the flip of a $K_X$-negative small extremal contraction. \item One of the following holds for $X^*$: \begin{enumerate} \item $K_{X^*}$ is nef (that is $(C\cdot K_{X^*})\geq 0$ for any curve $C\subset X^*$), or \item there is a fiber type extremal contraction $X^*\to Z$. \end{enumerate} \end{enumerate} \end{thm} \begin{rem} For the purposes of this paper one can handle the MMP as a black box. It is sufficient to know that it works, but I will use very few of its finer properties. In particular, there is no need to know anything about flips beyond believing their existence. The rest of the section is devoted to explicitly stating all further results from minimal model theory that I use later. The most significant among these is the classification of terminal 3-fold singularities over nonclosed fields, established in \cite{rat1}. \end{rem} \begin{notation}\label{mmp.ps.not} For a field $K$ let $K[[x_1,\dots,x_n]]$ denote the ring of formal power series in $n$ variables over $K$. For $K=\r$ or $K=\c$, let $K\{x_1,\dots,x_n\}$ denote the ring of those formal power series which converge in some neighborhood of the origin. For a power series $F$, $F_d$ denotes the degree $d$ homogeneous part. The multiplicity, denoted by $\mult_0F$, is the smallest $d$ such that $F_d\neq 0$. If we write a power series as $F_{\geq d}$ then it is assumed that its multiplicity is at least $d$. For $F\in \r\{x_1,\dots,x_n\}$ let $(F=0)$ denote the germ of its zero set in $\c^n$ with its natural real structure. I always think of it as a complex analytic germ with a real structure and not just as a real analytic germ in $\r^n$. $(F=0)/\frac1{n}(a,b,c,d)$ means the following. Define a $\z_n$-grading of $\c\{x,y,z,t\}$ by $x\mapsto a, y\mapsto b, z\mapsto c,t\mapsto d$. If $F$ is graded homogeneous, then $(F=0)/\frac1{n}(a,b,c,d)$ denotes the germ whose ring of holomorphic functions is the ring of grade zero elements of $\c\{x,y,z,t\}/(F)$. If $(F=0)$ is terminal then $n$ coincides with the index (\ref{mmp.KX.def}) of the singularity. \end{notation} \begin{exmp} In case $X=(x^2+y^2+z^2+ t^2=0)/\frac12(1,1,1,0)$ the ring is $$ \o_X=\c\{x^2,y^2,z^2,t,xy,yz,zx\}/(x^2+y^2+z^2+ t^2), $$ with the natural real structure. $X$ can also be realized as the image of the hypersurface $(x^2+y^2+z^2+ t^2=0)$ under the map $$ \phi: \c^4\to \c^7:\quad (x,y,z,t)\mapsto (x^2,y^2,z^2,t,xy,yz,zx), $$ which has degree 2 over its image. Although $(x^2+y^2+z^2+ t^2=0)$ has only the origin as its real solution, $X$ has plenty of real points. Indeed, any real solution of $x^2+y^2+z^2- t^2=0$ gives a {\it real} point $P=\phi(\sqrt{-1}x,\sqrt{-1}y, \sqrt{-1}z, t)\in X(\r)$. $\phi^{-1}(P)$ is a pair of conjugate points on the hypersurface $(x^2+y^2+z^2+ t^2=0)$. All the real elements of $\o_X$ take up real values at $P$. This way we see that $X(\r)$ is a cone over 2 copies of $\r\p^2$. \end{exmp} The following is a summary of the classification of terminal singularities obtained in \cite{rat1}. As it turns out, the classification closely follows the earlier results over algebraically closed fields. The choice of the subdivison into cases is dictated by the needs of the proof in sections 9--11, rather than the internal logic of the classification. \begin{thm}\label{mmp.ts.thm} Let $X$ be a real algebraic or analytic 3-fold and $0\in X(\r)$ a real point. Then $X$ has a terminal singularity at $0$ iff a neighborhood of $0\in X$ is real analytically equivalent to one of the following: $$ \begin{tabular}{ll} name & \qquad equation \\ $cA_0$ &$(t=0)$\\ $cA_1$ &$(x^2+y^2\pm z^2\pm t^m=0)$\\ $cA_{>1}^+$ &$(x^2+y^2+g_{\geq 3}(z,t)=0)$ \\ $cA_{>1}^-$ &$(x^2-y^2+g_{\geq 3}(z,t)=0)$ \\ $cD_4$ &$(x^2+f_{\geq 3}(y,z,t)=0)$, where $f_3\neq l_1^2l_2$ for linear forms $l_i$\\ $cD_{>4}$ &$(x^2+y^2z+f_{\geq 4}(y,z,t)=0)$,\\ $cE_6$ &$(x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)=0)$, where $h_4\neq 0$\\ $cE_7$ &$(x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)=0)$, where $g_3\neq 0$\\ $cE_8$ &$(x^2+y^3+yg_{\geq 4}(z,t)+h_{\geq 5}(z,t)=0)$, where $h_5\neq 0$\\ $cA_0/n$ & $(t=0)/\frac1{n}(r,-r,1,0)$ where $n\geq 2$ and $(n,r)=1$\\ $cA_1/2$ & $(x^2+y^2\pm z^n\pm t^m=0)/\frac12(1,1,1,0)$ where $\min\{n,m\}=2$\\ $cA_{>1}^+/2$ & $(x^2+y^2+f_{\geq 3} (z,t)=0)/\frac12(1,1,1,0)$\\ $cA_{>1}^-/2$ & $(x^2-y^2+f_{\geq 3}(z,t)=0)/ \frac12(1,1,1,0)$\\ $cA/n$ & $(xy+f(z,t)=0)/\frac1{n}(r,-r,1,0)$ where $n\geq 3$ and $(n,r)=1$\\ $cAx/2$ & $(x^2\pm y^2 +f_{\geq 4} (z,t)=0)/\frac12(0,1,1,1)$\\ $cAx/4$ & $(x^2\pm y^2 +f_{\geq 2}(z,t)=0)/\frac14(1,3,1,2)$ where $f_2(0,1)=0$\\ $cD/2$ &$(x^2+f_{\geq 3}(y,z,t)=0)/\frac12(1,1,0,1)$\\ $cD/3$ &$(x^2+f_{\geq 3}(y,z,t)=0)/\frac13(0,1,1,2)$ where $f_3(0,0,1)\neq 0$\\ $cE/2$ &$(x^2+y^3+f_{\geq 4}(y,z,t)=0)/\frac12(1,0,1,1)$\\ \end{tabular} $$ \end{thm} \section{The Topology of Real Points and the MMP} Starting with a projective variety $X$ over $\r$, let us run the MMP over $\r$. We obtain a sequence of birational maps $$ X=X_0\map X_1\map \cdots\map X_i\stackrel{f_i}{\map} X_{i+1}\map\cdots\map X^*. $$ These in turn induce (not necessarily everywhere defined) maps between the sets of real points $$ X(\r)=X_0(\r)\map \cdots\map X_i(\r)\stackrel{f_i}{\map} X_{i+1}(\r)\map\cdots\map X^*(\r). $$ Our aim is to see if there is a way of describing $X(\r)$ in terms of $X^*(\r)$ and a local description of the maps $X_i(\r)\map X_{i+1}(\r)$ in a neighborhood of their exceptional sets. \begin{prop}\label{mmpt.cases.prop} Every step $f_i$ of the MMP over $\r$ is among the following five: \begin{enumerate} \item (divisor--to--point) $f_i$ contracts a geometrically irreducible divisor $E_i\subset X_i$ to a point $P_{i+1}\in X_{i+1}(\r)$. \item (divisor--to--curve) $f_i$ contracts a geometrically irreducible divisor $E_i\subset X_i$ to a real curve $C_{i+1}\subset X_{i+1}$. \item ($\r$-small) $f_i:X_i(\r)\to X_{i+1}(\r)$ collapses a 1-complex to points and is a homeomorphism elsewhere. \item (flip) $f_i$ is the flip of a curve $C_i\subset X_i$. \item ($\r$-trivial) $f_i$ is an isomorphism in a (Zariski) neighborhood of the set of real points. \end{enumerate} \end{prop} Proof. If $f_i$ is a flip then we have case (4). Thus we may assume that $f_i$ is the contraction of a divisor $E_i\subset X_i$ and $E_i$ is irreducible over $\r$. If $E_i$ is irreducible over $\c$ then we have one of the cases (1--2). If $E_i$ is reducible over $\c$ then $E_i(\r)$ is a 1-complex by (\ref{mmpt.red.small.lem}) and so we are in case (3). Any of the above cases can also be of type (5).\qed \begin{lem}\label{mmpt.red.small.lem} Let $X$ be an n-dimensional scheme over $\r$ (that is, an algebraic variety possibly with several irreducible components and with singularities). Assume that if $X_i\subset X$ is any $\r$-irreducible component then $X_i$ is reducible over $\c$. Then $X(\r)= (\sing X)(\r)$, that is, every real point is singular. In particular, $\dim X(\r)\leq n-1$. \end{lem} Proof. Assume that $P\in X(\r)$ is a smooth real point. Then $P$ lies on a unique irreducible component $Y\subset X_{\c}$, thus $Y$ is invariant under complex conjugation. So $Y$ is an irreducible real component which stays irreducible over $\c$, a contradiction.\qed \medskip Each of the 5 steps $X_i(\r)\map X_{i+1}(\r)$ have different topological behaviour. The following informal discussion intends to emphasize their main features. \begin{say}[Divisor--to--point] Let $M=X_i(\r)$ be a 3-complex (with only finitely many singular points) and $F=E_i(\r)\subset M$ a 2-complex. We collapse $F$ to a point: $$ \begin{array}{ccc} F& \subset & M\\ \downarrow && \ \ \downarrow f\\ P&\in &N. \end{array} $$ In practice we are frequently able to describe a regular neighborhood $F\subset U\subset M$ (this is a local datum) and by assumption we know a regular neighborhood $P\in V\subset N$. Thus we see that M is obtained from $U$ and $N\setminus \inter V$ by gluing them together along the boundaries $\partial U$ and $\partial V$. The gluing is determined by a PL-homeomorphism $\phi: \partial U\to \partial V$. Thus, besides knowing $U$ and $N$, we also need to know $\phi$ up to PL-isotopy. If one of the connected components of $\partial U$ has genus at least 2, this is a very hard problem. In fact, as the example of Heegard splittings shows (cf. \cite[Ch.2]{Hempel76}), the choice of $\phi$ is usually the most significant information. Unfortunately, $\phi$ can be described only in terms of global data. If $\partial U$ is a union of $m$ copies of $S^2$, then $\phi$ is classified by an element of the symmetric group on $m$ elements (which $S^2$ maps where) and a sign for each $S^2$ (describing whether the map is orientation preserving or reversing on that $S^2$). Hence, knowing $U$ and $N$, we can determine $M$ up to finite ambiguity. In many cases $U$ is so simple that different choices of $\phi$ give the same $M$, giving even fewer possibilities for $M$. If $P\in N$ is an isolated singular point, then $\partial V$ is a union of spheres iff $\bar N$ (the topological normalization of $N$) is a manifold. The situation is similarly simple if $\partial U$ is a union of copies of $\r\p^2$ and of $S^2$, and still manageable if $\partial U$ also contains tori and Klein bottles. For us these more general cases do not come up. \end{say} \begin{say}[Divisor--to--curve] This time we construct it bottom up. Assume for simplicity that $N=X_{i+1}(\r)$ is a 3-manifold and $L=C_{i+1}(\r)\subset N$ a link. The projectivized normal bundle is an $S^1$-bundle $S\to L$. The blow up of $L$ in $N$ replaces $L$ by $S$ to obtain: $$ \begin{array}{ccc} S& \subset & M\\ \downarrow && \ \ \downarrow f\\ L&\in &N. \end{array} $$ (In general $N$ may have finitely many singular points and $L$ is only a 1-complex, but I believe that a similar description is possible in all cases.) Here $M$ is uniquely determined, once we know $N$ and $L$. By assumption we know $N$ but $L$ is a free choice. The Jaco--Johannson--Shalen decomposition (cf. \cite[p.483]{Scott83}) shows that in most cases $B_LN$ determines $M\setminus L$. Thus the description of all possible $B_LN$ is essentially equivalent to the description of all links. For us $L$ has to come from an algebraic curve, thus we are led to the question: Which links in a real algebraic 3-fold can be realized by algebraic curves? In some cases every link is realized (cf. \cite{AK81}), thus we again run into a hard topological problem. So $M$ can be described in terms of $N$, though the answer depends on the choice of a link, which is a very complicated object. \end{say} \begin{say}[$\r$-small contraction] $N=X_{i+1}(\r)$ is obtained from $M=X_i(\r)$ by collapsing a 1-complex $C=(\sing E_i)(\r)=E_i(\r)$ to a point: $$ \begin{array}{ccc} C& \subset & M\\ \downarrow && \ \ \downarrow f\\ P&\in &N. \end{array} $$ If the normalizations $\bar M$ and $\bar N$ are manifolds, then we see in (\ref{top.alg.homeo.lem}) that a suitable small perturbation of $f$ is a homeomorphism between $\bar M$ and $\bar N$. Thus this step (which is actually more complicated from the point of view of algebraic geometry than the previous two cases) is easy to analyze topologically. \end{say} \begin{say}[Flip] Assume for simplicity that $M=X_i(\r)$ is an orientable 3-manifold and $C(\r)\sim S^1$. $N=X_{i+1}(\r)$ is obtained from $M$ by a surgery along $S^1$. The boundary of a regular neighborhood of $S^1$ is $S^1\times S^1$, and the surgery is determined by a diffeomorphism of $S^1\times S^1$ up to isotopy. These are classified by $SL(2,\z)$. (In general $M$ may have finitely many singular points and $C(\r)$ is a 1-complex, but I believe that a similar description is possible in all cases.) A complete classification of flips is known \cite{KoMo92}, thus it should be possible to compute the resulting diffeomorphism of $S^1\times S^1$. Here again we run into a global problem. $S^1\subset M$ may be very knotted, and the result of the surgery depends mostly on the knot $S^1\subset M$. The usual descriptions of flips characterize a complex analytic neighborhood of $C$, thus they say nothing about how its real part is knotted. From the point of view of algebraic geometry, this is a global invariant. We have the additional problem that flipping curves are rigid objects, thus we can not hope to get a flipping curve by approximating a real curve algebraically. Furthermore, it is very hard to determine which curves are obtained by a flip. (Even if $Z$ is a smooth complex 3-fold and $C\subset Z$ a smooth $\c\p^1$, I know of no practical way of determining if $C\subset Z$ is obtained as a result of a flip.) \end{say} \begin{say}[Conclusion]\label{mmpt.concl} Start with a projective 3-fold $X$ over $\r$ and run the MMP over $\r$: $$ X=X_0\map X_1\map \cdots\map X_i\stackrel{f_i}{\map} X_{i+1}\map\cdots\map X^*. $$ If we would like to understand the topology of $X(\r)$ in terms of $X^*(\r)$, then we have to ensure that the MMP has the following properties: \begin{enumerate} \item $\overline{X_i(\r)}$ is a manifold for every $i$. \item Each $f_i$ is either $\r$-trivial or $\r$-small or a divisor--to--point contraction. \end{enumerate} \end{say} (\ref{int.mmp.steps}) asserts that both of these conditions can be satisfied by imposing certain mild conditions on the topology of $X(\r)$. \section{The Topology of Divisorial Contractions} The aim of this section is to describe some examples where the change of the topology of a real algebraic variety under a divisorial contraction can be readily understood by topological methods. \begin{notation}\label{5.notation} The disjoint union of two topological spaces is denoted by $M\uplus N$. Direct product is denoted by $M\times N$. The unique nontrivial $S^2$-bundle over $S^1$ is denoted by $S^1\tilde{\times} S^2$. This is obtained from $[0,1]\times S^2$ by indentifying the 2 ends via an orientation reversing homeomorphism. Homeomorphism of two topological spaces is denoted by $M\sim N$. \end{notation} We start with the study of $\r$-small contractions: \begin{lem}\label{top.homeo.lem} Let $f:M\to N$ be a proper PL-map between PL-manifolds of dimension $n\geq 3$. Assume that there is a 1-complex $C\subset M$ and a finite set of points $P\subset N$ such that $f:M\setminus C\to N\setminus P$ is a PL-homeomorphism. Then $M$ and $N$ are PL-homeomorphic (by a small perturbation of $f$). \end{lem} Proof. If $C$ is collapsible to points, then a regular neighborhood of $C$ in $M$ is a union of disjoint $n$-cells \cite[1.8]{Hempel76} and we are done. In order to see that $C$ is collapsible to points, we may assume that $P$ is a point and $N=S^n$. Thus $M$ is also a compact PL-manifold. $M$ is orientable outside the codimenison $\geq 2$ subset $C$, hence it is orientable. Consider the exact homology sequences $$ \begin{array}{ccccccc} H_i(C)&\to &H_i(M)&\to & H_i(M,C)&\to & H_{i-1}(C)\\ \downarrow &&\downarrow &&\downarrow &&\downarrow \\ H_i(P)&\to &H_i(S^n)&\to & H_i(S^n,P)&\to & H_{i-1}(P) \end{array} $$ We compute $H_i(M,C)=H_i(S^n,P)$ from the second sequence and substitute into the first to obtain that $$ H_1(C)\cong H_1(M), \qtq{and} 0=H_{n-1}(C)\cong H_{n-1}(M). $$ By Poincar\'e duality we conclude that $H_1(C)=0$, thus $C$ is contractible. \qed \begin{cor}\label{top.alg.homeo.lem} Let $f:X\to Y$ be a morphism a $n$-dimensional real algebraic varieties, $n\geq 3$. Assume that \begin{enumerate} \item $\overline{X(\r)}=M\uplus R$ and $\overline{Y(\r)}=N\uplus R'$ where $M,N$ are PL-manifolds and $\dim R,\dim R'<n$. \item $f$ induces an isomorphism $R\cong R'$. \item $\ex(f)(\r)$ is a 1-complex. \end{enumerate} \noindent Then $\overline{X(\r)}$ is PL-homeomorphic to $\overline{Y(\r)}$. \end{cor} Proof. Set $C=\ex(f)(\r)$ and $\bar C\subset M\subset \overline{X(\r)}$ its preimage. Since $\bar C$ has dimension 1, there is a one--to--one correspondence between the connected components of $\bar C$ and the connected components of the boundary of a regular neighborhood of $\bar C$. Hence $f$ lifts to a morphism $\bar f: \overline{X(\r)}\to \overline{Y(\r)}$. Thus (\ref{top.homeo.lem}) implies (\ref{top.alg.homeo.lem}), and the homeomorphism is given by a small perturbation of $\bar f$.\qed \begin{rem} A real algebraic curve is a union of copies of $S^1$. The proof of (\ref{top.homeo.lem}) shows that the preimage of $\ex(f)(\r)$ in $\overline{X(\r)}$ is contractible. Thus if $f$ itself is not a homeomorphism, then $X(\r)$ is not a manifold. \end{rem} Next we look at divisor--to--point contractions. \begin{prop}\label{top.normsurf.nbd.boundary} Let $M$ be a 3-dimensional PL-manifold and $F\subset M$ a connected 2-complex with only finitely many singular points. Let $F\subset \inter U\subset M$ be a regular neighborhood of $F$. Then $$ \dim H_1(\bar F,\q)\leq \dim H_1(\partial U,\q), $$ and strict inequality holds unless every connected component of $\bar F$ is one of the following: \begin{enumerate} \item $S^2$ or $\r\p^2$, \item a one-sided $S^1\times S^1$, \item a one-sided Klein bottle whose neighborhood is not orientable. \end{enumerate} \end{prop} Proof. Let $F$ be a compact 2-complex with only finitely many singular points. Its normalization $\bar F$ can be written as $F^{(2)}\uplus F^{(1)}$ where $F^{(2)}$ is a compact 2-manifold and a $ F^{(1)}$ is a 1-complex. Pick a point $P\in F$ whose link in $F$ consists of at least 2 circles. Locally $F$ looks like the cone over parallel plane sections $(z=a_i)\cap (x^2+y^2+z^2=1)$ of the unit sphere in $\r^3$ (plus a few 1-cells). By a homotopy we can replace this by the parallel plane sections of the unit ball $(z=a_i)\cap (x^2+y^2+z^2\leq 1)$ and add the interval $[\min_i\{a_i\},\max_i\{a_i\}]$ on the $z$-axis. This does not change the boundary of the regular neighborhood. Thus we may assume that $F^{(2)}\to M$ is an embedding. Let us take a point or a 1-cell $e$ in $F^{(1)}$. If $e$ does not intersect the rest of $F$, then a regular neighborhood of $e$ is a 3-cell. $e$ can be deleted from $F$ without changing the inequality. If $e$ intersects the rest of $F$ in one endpoint only, then we can delete $e$ from $F$ without changing the regular neighborhood. If $e$ intersects the rest of $F$ at both endpoints, then removing $e$ creates a new 2-complex $F'$, and $F^{(2)}={F'}^{(2)}$. Let $F'\subset\inter U'$ be its regular neighborhood. $\partial U$ is obtained from $\partial U'$ by attaching a handle $[0,1]\times S^1$. Thus $H_1(\partial U)\geq H_1(\partial U')$, and it is sufficient to verify our inequality for $F'$. At the end we are reduced to the situation when $F$ is the disjoint union of embedded 2-manifolds, and it is sufficient to check the ineqality for each connected component of $F$ separately. $\partial U \to F$ is a 2 sheeted cover, thus $H_1(\partial U)\geq H_1(F)$ with equality only if $F\sim S^2, F\sim \r\p^2, F\sim S^1\times S^1\sim \partial U$ or $F$ and $\partial U$ are both Klein bottles.\qed \begin{prop}\label{top.normsurf.collapse} Let $M$ be a 3-dimensional PL-manifold and $F\subset M$ a compact 2-complex with only finitely many singular points. Let $0\in N$ be obtained from $M$ by collapsing $F$ to a point. Assume that $\bar N$ is a 3-manifold. Then $M$ can be obtained from $\bar N$ by repeated application of the following operations: \begin{enumerate} \item taking connected sums of connected components, \item taking connected sum with $S^1\times S^2$, \item taking connected sum with $S^1\tilde{\times} S^2$, or \item taking connected sum with $\r\p^3$. \end{enumerate} \end{prop} Proof. We use the notation of (\ref{top.normsurf.nbd.boundary}) and of its proof. Let $F\subset \inter U\subset M$ and $0\in \inter V\subset N$ be regular neighborhoods such that $U=f^{-1}(V)$. Then $\partial U=\partial V$. Since $\bar N$ is a manifold, this implies that $\partial U$ is a union of 2-spheres. We also see that $\bar N$ is obtained from $M\setminus\inter U$ by attaching a 3-ball to each $S^2$ in $\partial U$. As in the proof of (\ref{top.normsurf.nbd.boundary}) we may assume that $F^{(2)}\to M$ is an embedding. If $e$ is a point or a 1-cell in $F^{(1)}$ which intersects the rest of $F$ in zero or one point only, then we can delete $e$ from $F$. If $e$ intersects the rest of $F$ at both endpoints, then removing $e$ creates a new 2-complex $F'$ such that $\bar F=\bar F'$. Let $F'\subset\inter U'$ be its regular neighborhood. $\partial U$ is obtained from $\partial U'$ by attaching a handle $[0,1]\times S^1$. The two ends $\{0\}\times S^1$ and $\{1\}\times S^1$ can not attach to the same connected component of $\partial U'$ since that would create a torus or a Klein bottle in $\partial U$. Thus $\partial U'$ has one more copies of $S^2$ than $\partial U$. $\bar N$ is obtained from $\bar N'$ by collapsing the image of $e$ to a point, hence $\bar N$ and $\bar N'$ are homeomorphic by (\ref{top.homeo.lem}). At the end we are reduced to the situation when $F$ is the disjoint union of embedded copies of $S^2$ and $\r\p^2$. An $S^2$ is necessarily 2-sided. Removing it from $F$ corresponds to taking connected sums of connected components (if $S^2$ separates $M$) or to taking connected sum with $S^1\times S^2$ or $S^1\tilde{\times} S^2$ (if $S^2$ does not separate $M$) (cf. \cite[Chap.\ 3]{Hempel76}). If $\r\p^2$ is 2-sided, then the boundary of its regular neighborhood consists of two copies of $\r\p^2$, so this can not happen. A 1-sided $\r\p^2$ corresponds to taking connected sum with $\r\p^3$. \qed \begin{cor}\label{top.alg.normsurfup.lem} Let $f:X\to Y$ be a morphism a real algebraic $3$-folds. Assume that \begin{enumerate} \item $\overline{X(\r)}$ and $\overline{Y(\r)}$ are PL-manifolds, and \item $\ex(f)$ is a geometrically irreducible normal surface which is contracted to a point. \end{enumerate} \noindent Then $\overline{X(\r)}$ can be obtained from $\overline{Y(\r)}$ by repeated application of the following operations: \begin{enumerate} \setcounter{enumi}{2} \item removing an isolated point from $\overline{Y(\r)}$, \item taking connected sums of connected components, \item taking connected sum with $S^1\times S^2$, \item taking connected sum with $S^1\tilde{\times} S^2$, or \item taking connected sum with $\r\p^3$. \end{enumerate} \end{cor} Proof. If $\ex(f)(\r)=\emptyset$ then the image of $\ex(f)$ is an isolated real point of $Y(\r)$ which has to be thrown away to obtain $X(\r)$. If $\ex(f)(\r)\neq\emptyset$, then isolated points of $X(\r)$ correspond to isolated points of $Y(\r)$, hence they can be ignored. Let $M$ be the topological normalization of $X(\r)\setminus(\mbox{isolated points})$, $N$ the topological normalization of $Y(\r)\setminus(\mbox{isolated points})$ and $F$ the preimage of $\ex(f)(\r)$ in $M$. $F$ is a 2-complex with isolated singularities since $\ex(f)$ is normal. Thus (\ref{top.alg.normsurfup.lem}) follows from (\ref{top.normsurf.collapse}).\qed \begin{complement} It is worthwhile to note that condition (\ref{top.alg.normsurfup.lem}.2) can be weakened to: \begin{enumerate} \item[2'] $\ex(f)$ contains a unique geometrically irreducible surface $S$. $S$ has only isolated singularities and $S$ is contracted to a point by $f$. \end{enumerate} \end{complement} It would be very useful to have a version of (\ref{top.alg.normsurfup.lem}) which works if $\ex(f)$ is an irreducible but nonnormal surface. In the topological version (\ref{top.normsurf.collapse}) esentially nothing can be said if $F$ is allowed to become an arbitrary compact 2-complex. For instance, let $M$ be an arbitrary compact 3-manifold and $F$ the 2-skeleton of a triangulation of $M$. Then $\bar N$ is the union of some copies of $S^3$ (one for each 3-simplex). This example usually can not arise as the real points of an algebraic surface, but it is not hard to modify this example by approximating each simplex with a sphere to get the following. (This is not used in the sequel and so no proof is given here.) \begin{prop}\label{top.exmp} Let $M$ be a compact differentiable manifold of dimension $n$. Then there is a smooth real algebraic variety $X$ and a morphism $f:X\to Y$ with the following properties: \begin{enumerate} \item $X(\r)\sim M$, \item $\overline{Y(\r)}$ is a disjoint union of copies of $S^n$, \item $\ex(f)$ is a geometrically irreducible divisor and $\ex(f)(\r)$ is a union of copies of $S^{n-1}$ intersecting transversally. \end{enumerate} \end{prop} \section{The Gateway Method} At the beginnings of the MMP, divisorial contractions were considered to be the easily understandable part of the program and flips the hard part. Lately, however, more and more questions require a detailed understanding of all the steps of the MMP. A fairly complete description of all flips is known \cite{KoMo92}, but it seems very difficult to obtain a list of all divisorial contractions. One can try to study the MMP in two basic ways: \begin{say}[Analysis of the MMP] Starting with a projective variety $X$, let us run the MMP. We obtain a sequence of birational maps $$ X=X_0\map X_1\map \cdots\map X_i\map X_{i+1}\map\cdots\map X^*. $$ Assume that $X$ has some nice property that we would like to preserve. We need some way of proving that $X^*$ also has this property, at least under some additional assumptions. One way is to prove this directly, by analyzing each step of the MMP. This would sometimes require knowing each step of the MMP, and even in dimension 3 the list is not yet available. Still there are many results that can be established this way, for instance the existence of the MMP itself. In this approach one starts with a variety $X$ and tries to understand every possible way an MMP can {\it start} with $X$. This is oftentimes manageable if $X$ has only mild singularities. Another way is to look at each step of the MMP backwards. In dimension 3 we have a pretty good description of the possible singularities that arise in the course of an MMP. Thus we can start with a variety $Y$ and try to understand every possible way an MMP can {\it end} with $Y$. This also seems rather hard. Even the case when $Y$ is smooth is not at all understood, but in some other cases this approach has been carried through \cite{Kawamata9?}. It seems that this method is easier to apply when $X$ is fairly singular. The {\it gateway method} attempts to solve the original problem in an intermediate way. In the above chain of maps there is a smallest index $i$ such that $X_i$ is still ``nice" but $X_{i+1}$ is not. Hence $X_i\map X_{i+1}$ is a ``gateway" through which the process leaves the set of ``nice" varieties. Analysing these ``gateways" should be easier since the direct approach tends to work for the nice variety $X_i$ and the backwards method tends to work for more complicated singularities of $X_{i+1}$. Once such a list of ``gateways" is obtained, it is a matter of checking the list to see if some additional properties ensure that this step does not happen. \end{say} One of the simplest examples where these ideas yield a nontrivial result is the following. \begin{exmp} Assume that we want to stay within the class of varieties of index 1. In this case there is only one gateway: {\bf Proposition.} {\it Let $f:X\map X'$ be a step of the 3-dimensional MMP where $X$ has index 1 but $X'$ has higher index. Then $f$ is the contraction of a divisor $E\subset X$ to a point. Furthermore, $E\cong \p^2$, $X$ is smooth along $E$ and $E$ has normal bundle $\o_E(-2)$.} This result is a special case of \cite{Mori88} and \cite{Cutkosky88}, though they did not approach this from the point of view of gateways. A proof along the lines suggested by the gateway method is not hard to construct, but this is not any shorter then the direct proofs. As a consequence we obtain: {\bf Corollary} {\it Let $X$ be a projective 3-fold with index 1 terminal singularities. Assume that $X$ does not contain any surface $S\subset X$ which admits a birational morphism onto $\p^2$. Then each step of the MMP starting with $X$ is a projective 3-fold with index 1 terminal singularities.} Unfortunately the above condition needs to be checked for every surface $S$, even for very singular ones. Thus in practice this does not seem to be a useful observation. \end{exmp} \begin{say} Our aim is to develop a similar theory for real algebraic threefolds. Thus we have to decide which varieties are ``nice" and then describe all possible gateways through which the MMP can leave the class of ``nice" varieties. (\ref{mmpt.concl}) naturally suggest a topological choice: $X$ is ``nice" if $X(\r)$ or $\overline{X(\r)}$ is a 3-manifold, maybe with some additional properties. This was my first attempt, but I was unable to make it work. The main problem seems to be that, as the computations of \cite{rat1} show, there is basically no relationship between the algebraic complexity of a terminal singularity $0\in X$ and the topological complexity of its real points $X(\r)$. Eventually I settled at a completely algebraic choice: $X$ is nice if it has index 1 along $X(\r)$. There are two main reason for adopting this definition: \begin{enumerate} \item Most complications of 3-dimensional birational geometry come from the appearance of points of index $>1$. Hence this is likely to be the right choice algebraically. \item One of the first things I realized was that under this condition there would be no flips. Indeed, flips need higher index singular points to exist. If we have only index 1 points along $X(\r)$, then all higher index points appear in conjugate pairs. A look at the list of flips \cite{KoMo92} shows that the singularities appearing along a flipping curve are {\it always} asymmetrical. \end{enumerate} Thus our task is to get a list of all steps $f:Y\to X$ of the MMP over $\r$ such that $Y$ has index 1 along $Y(\r)$. The case of divisor--to--curve contraction is relatively easy. Most of the work is devoted to studying the divisor--to--point contractions. Let $0\in X(\r)$ be the point in question. The existence of $f$ is local in the Euclidean topology. I will go through the classification (up to real analytic equivalence) of 3-dimensional terminal singularities over $\r$ and for each describe all possible $f:Y\to X$. There is one subtle point here: the condition of $\q$-factoriality is not preserved under analytic equivalence. Thus first we need to develop a notion of ``extremal contraction without $\q$-factoriality". \end{say} \begin{defn}\label{elem.extr.defn} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. A proper birational morphism $f:Y\to X$ is called an {\it elementary extraction} of $X$ if \begin{enumerate} \item $Y$ is normal and $K_Y$ is $\q$-Cartier. \item The exceptional set $\ex(f)$ contains a unique $K$-irreducible divisor $E$. \item $-K_Y$ is $f$-ample. \end{enumerate} If we start with $Y$ and construct $f:Y\to X$ then $f$ is usually called an {\it elementary contraction} of $Y$. We can write $K_Y\equiv f^*K_X+a(E,X)E$ where $a(E,X)$ is the discrepancy of $E$. Thus $-a(E,X)E$ is $f$-ample. An exceptional divisor can never be relatively ample (or nef) (cf. \cite[3.35]{KoMo98}), thus $a(E,X)>0$ and so $-E$ is $f$-ample. This implies that $\ex(f)=\supp E$. $f(E)$ is also called the {\it center} of $f$ on $X$. \end{defn} A crucial property of elementary extractions is that they are determined by their exceptional divisors: \begin{prop}\label{gw.mats-mumf.lem} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. Let $f_i:Y_i\to X$ be elementary extractions with exceptional divisors $E_i\subset Y_i$ for $i=1,2$. Assume that $E_1$ and $E_2$ correspond to each other under the birational map $f_2^{-1}\circ f_1:Y_1\map Y_2$. Then $Y_1$ and $Y_2$ are isomorphic (over $X$). \end{prop} Proof. Let $\phi:f_2^{-1}\circ f_1:Y_1\map Y_2$ be the composition. $\phi$ is birational, and $\ex(\phi)$, $\ex(\phi^{-1})$ have codimension at least 2. Furthermore, $K_{Y_1}$ and $K_{Y_2}=\phi_*(K_{Y_1})$ are relatively ample. Thus $\phi$ is an isomorphism by an argument of \cite[p.671]{Matsusaka-Mumford64}. \qed \medskip In some sense this gives a way of enumerating all elementary extractions of $X$. We try to list all exceptional divisors over $X$ and for each construct the corresponding unique elementary extraction. Usually there are infinitely many elementary extractions for a given $X$ and there does not seem to be an easy way to predict for which divisors does the corresponding elementary extraction exist. The next definition singles out a special class of elementary extractions, by restricting the singularities allowed on $Y$. The aim is to formalize a special case of the gateway method: we assume that $Y$ is ``nice". \begin{defn}\label{gw.g-e.def} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. A proper birational morphism $f:Y\to X$ is called a {\it gateway--extraction} or {\it g--extraction} if \begin{enumerate} \item $f$ is an elementary extraction with exceptional divisor $E\subset Y$. \item $Y$ has terminal singularities. \item $K_X$ and $E$ are Cartier at the generic point of every geometrically irreducible $K$-subvariety of $\ex(f)$. \end{enumerate} \noindent In dimension three $Y$ has only isolated singularities, hence (3) is equivalent to the apparently weaker condition: \begin{enumerate} \item[3'] $K_X$ and $E$ are Cartier at every $K$-point of $\ex(f)$. \end{enumerate} If we start with $Y$ and construct $f:Y\to X$ then $f$ is usually called a {\it g--contraction} of $Y$. \end{defn} The main technical aim of this article is to obtain a list of g--extractions for threefolds with terminal singularities. The project turns out to be feasible since the discrepancy $a(E,X)$ is always quite small. I have no a priori proof of this, but in every case the study of low discrepancy divisors leads to a description of all g--extractions. The relationship between low discrepancy divisors and g--extractions rests on the following easy observation: \begin{prop}\label{gw.d-ineq.prop} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. Let $f:Y\to X$ be a g--extraction with exceptional divisor $F\subset Y$. Let $E$ be a geometrically irreducible $K$-divisor over $X$ such that $\cent_XE\subset \cent_XF$. Then $$ a(E,X)\geq a(E,Y)+a(F,X). $$ \end{prop} Proof. Let $g:Z\to X$ be a proper birational morphism such that $\cent_ZE$ is a divisor on $Z$. We may assume that the induced rational map $h:Z\map Y$ is a morphism. $h(E)$ is a geometrically irreducible $K$-subvariety of $Y$ which is contained in $\ex(f)$. Write $$ \begin{array}{rcl} K_Z&\equiv& g^*K_X+a(E,X)E+(\mbox{other exceptional divisors}),\\ K_Z&\equiv& h^*K_Y+a(E,Y)E+(\mbox{other exceptional divisors}),\\ K_Y&\equiv& h^*K_X+a(F,Y)F, \qtq{and}\\ h^*F&\equiv& cE+(\mbox{other exceptional divisors}), \end{array} $$ where $c>0$ since $h(E)\subset \ex(f)=\supp F$ and $c$ is an integer by (\ref{gw.g-e.def}.3). Making the substitutions we obtain that $a(E,X)= a(E,Y)+c\cdot a(F,X)\geq a(E,Y)+a(F,X)$. \qed \medskip The same method also proves the following result: \begin{prop}\label{gw.d-ineq.prop.cor} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. Let $f:Y\to X$ be a morphism with exceptional divisor $F=\cup F_i\subset Y$. Assume that $Y$ has terminal singularities and $K_X$ and $F$ are Cartier at the generic point of every geometrically irreducible $K$-subvariety of $\ex(f)$. Let $E$ be a geometrically irreducible $K$-divisor over $X$ such that $\cent_XE\subset \cup_i\cent_XF_i$. Then $$ a(E,X)\geq a(E,Y)+\min_i\{a(F_i,X)\}.\qed $$ \end{prop} \begin{cor}\label{gw.discr1.cor} Let $X$ be a normal variety over a field $K$ such that $K_X$ is $\q$-Cartier. Let $f:Y\to X$ be an elementary extraction with exceptional divisor $E\subset Y$. Assume that $E$ is geometrically irreducible and $a(E,X)\leq 1$. Then either $f:Y\to X$ is a g--extraction, or $X$ has no g--extractions whose center contains $f(E)$. \end{cor} Proof. Let $g:Z\to X$ be a g--extraction of $X$ whose center contains $f(E)$. Let $F\subset Z$ be the exceptional divisor. Then $a(E,X)\geq a(E,Z)+a(F,X)$. If $a(E,Z)=0$ then $\cent_ZE$ is a divisor which is contained in $F$. Since $F$ is an irreducible divisor, $\cent_ZE=F$, hence $Y=Z$ by (\ref{gw.mats-mumf.lem}). Otherwise $a(E,Y)\geq 1$ which would force $a(F,X)\leq 0$. This contradicts (\ref{gw.g-e.def}.2). \qed \begin{rem} This corollary gives a very efficient way of finding all g--extractions of a given $X$ in some cases. We have to find {\it one} geometrically irreducible divisor $E$ such that $a(E,X)\leq 1$ and construct the corresponding elementary extraction $f:Y\to X$. Then it is usually easy to determine the singularities of $Y$. \cite{Markushevich96} proved that if $0\in X$ is a terminal threefold singularity which is not smooth, then there is a divisor $E$ over $\bar K$ with $\cent_XE=\{0\}$ and $a(E,X)\leq 1$. Thus there is always such an irreducible $K$-divisior, but it may not be geometrically irreducible. Still, in many cases we are able to apply (\ref{gw.discr1.cor}) directly. In the remaining cases we show that there is always a geometrically irreducible divisor $E$ with $\cent_XE=\{0\}$ and $a(E,X)\leq 3$. This is still very useful, thanks to the following: \end{rem} \begin{cor}\label{gw.discr2.cor} Let $X$ be a normal variety over a field $K$ such that $K_X$ is Cartier. Let $f:Y\to X$ be an elementary extraction with exceptional divisor $E\subset Y$. Assume that $E$ is geometrically irreducible. Let $g:Z\to X$ be any g--extraction with exceptional divisor $F$ whose center contains $f(E)$. Then either $g=f$ or $a(F,X)\leq a(E,X)-1$. \end{cor} Proof. By (\ref{gw.d-ineq.prop}), $a(E,X)\geq a(E,Z)+a(F,X)$. If $a(E,Z)=0$ then $E$ and $F$ correspond to each other, hence $Y=Z$ by (\ref{gw.mats-mumf.lem}). Otherwise $a(E,Y)\geq 1$, thus $a(F,X)\leq a(E,X)-1$. \qed \section{Small and Divisor--to--Curve Contractions} In this section we look at those steps $f:X\to Y$ of the MMP over $\r$ which are either small contractions or contract a divisor to a curve. The two cases can be treated together in the following setting: \begin{notation}\label{sm.not} Let $K$ be a field of characteristic 0. Let $X$ be a 3-fold over $K$ with terminal singularities and $f:X\to Y$ a proper birational morphism over $K$ such that $-K_X$ is $f$-ample and $f_*\o_X=\o_Y$. Let $0\in Y(K)$ be a closed point such that $\dim f^{-1}(0)=1$. We will need that under these assumptions $R^1f_*\o_X=R^1f_*\o_X(K_X)=0$ by the generalized Grauert--Riemenschneider vanishing theorem (see, for instance, \cite[8.8]{CKM88} or \cite[2.65]{KoMo98}). \end{notation} In keeping with the principles of the gateway method, we are interested in the case when $X$ has index 1 at all points of $X(K)$. The following theorem gives a complete description of such contractions: \begin{thm}\label{sm.thm} Notation and assumptions as in (\ref{sm.not}). Assume in addition that $X$ has index 1 at all points of $X(K)$. Then $Y$ is smooth at $0$ and one can choose local (analytic or formal) coordinates $(x,y,z)$ at $0\in Y$ such that $X$ is the blow up of the curve $(z=g(x,y)=0)\subset Y$ for some $g\in K[[x,y]]$. In particular, $f$ can not be small. \end{thm} This theorem has some very useful consequences for the MMP over $\r$: \begin{cor}\label{sm.mmp.cor1} Starting with a projective variety $X$ over $\r$, let $$ X=X_0\map X_1\map \cdots\map X_{i}\stackrel{f_{i}}{\map} X_{i+1} $$ be the beginning of an MMP over $\r$. Assume that $X_j$ has index 1 at all points of $X_j(\r)$ for $j\leq i$. Then the induced maps between the sets of real points $$ X(\r)=X_0(\r)\to \cdots\to X_{i}(\r)\stackrel{f_{i}}{\to} X_{i+1}(\r) $$ are everywhere defined. \end{cor} Proof. The only steps of the MMP over $\r$ which are not everywhere defined are the flips of small contractions (\ref{mmpt.cases.prop}). By (\ref{sm.thm}) there are no flips in the sequence.\qed \medskip The topological behavior of divisor--to--curve contractions can also be determined using (\ref{sm.thm}): \begin{thm}\label{sm.top.thm} Let $X$ be a proper 3-fold over $\r$ with terminal singularities such that $X$ has index 1 at all points of $X(\r)$ and $\overline{X(\r)}$ is a 3-manifold. Let $f:X\to Y$ be a proper birational morphism over $\r$ such that $-K_X$ is $f$-ample and $f_*\o_X=\o_Y$. Assume that $\dim f^{-1}(y)\leq 1$ for every $y\in Y$. Then either \begin{enumerate} \item $f$ is $\r$-small, or \item $\overline{X(\r)}$ contains a 1-sided torus or Klein bottle with nonorientable neighborhood. \end{enumerate} \end{thm} Proof. By (\ref{sm.thm}), there is a real curve $D\subset Y$ such that $Y$ is smooth along $D$ and $X=B_DY$ (at least in a neighborhood of $Y(\r)$). Pick $0\in D(\r)$ and let $(z=g(x,y)=0)$ be a local equation of $D$. By (\ref{sm.ci.blowup}), either $D$ is smooth at $0$ or $X$ has a unique singular point over $0$ with local equation $st=g(x,y)$, which is equivalent to $s^2-t^2-g(x,y)=0$. These are of type $cA_{>1}^-$ or $cA_1$ in the classification of \cite{rat1}. If $g$ does not change sign on the $(x,y)$-plane then $X(\r)\setminus f^{-1}(0)\to Y\setminus\{0\}$ is one--to--one near $0$, hence $f$ is $\r$-small near $0$. If $g$ does change sign on the $(x,y)$-plane, then from \cite[sec. 4]{rat1} we see that (after a coordinate change) $g=\pm(x^2+y^{2r+1})$ and $X(\r)$ is a manifold near $f^{-1}(0)$. In particular, $D(\r)$ is the disjoint union of some isolated points and some copies of $S^1$. If $D(\r)$ is finite then $f$ is $\r$-small. Otherwise $D(\r)$ has a connected component $M\sim S^1$. Let $E\subset X$ be the exceptional divisor of $f$. By explicit computation we see that $E(\r)\to D(\r)$ is an $S^1$-bundle. Hence there is a unique connected component $N\subset E(\r)$ such that $N$ is an $S^1$-bundle over $M$. Thus $N$ is either a torus or a Klein bottle. $N$ is 1-sided with nonorientable neighborhood, since these hold locally for the blow up of a smooth curve in a smooth 3-fold.\qed \begin{exmp}\label{sm.ci.blowup} Set $Y={\Bbb A}^3$ with coordinates $(x,y,z)$. Let $X$ be the blow up of the curve $(z=g(x,y)=0)\subset Y$. Then $X$ has a unique singular point which is given by an equation $st-g(x,y)=0$. \end{exmp} \begin{cor}\label{sm.mmp.cor2} Starting with a projective variety $X$ over $\r$, let $$ X=X_0\map X_1\map \cdots\map X_{i}\stackrel{f_{i}}{\map} X_{i+1} $$ be the beginning of an MMP over $\r$. Assume that \begin{enumerate} \item $X_j$ has index 1 at all points of $X_j(\r)$ for $j\leq i$, \item $\overline{X_j(\r)}$ is a PL-manifold for $j\leq i$, \item $\overline{X(\r)}$ satisfies the conditions (\ref{int.no.cond}). \end{enumerate} \noindent Then: \begin{enumerate} \setcounter{enumi}{3} \item The induced maps between the sets of real points $f_j: X_{j}(\r)\to X_{j+1}(\r)$ are everywhere defined for $j\leq i$, \item For every $j\leq i+1$, there is a finite set $S_j\subset X_{j}(\r)$ such that $\overline{X_{j}(\r)}\setminus S_j$ is homeomorphic to an open subset of $\overline{X(\r)}$, \item The smooth part of $\overline{X_{i+1}(\r)}$ also satisfies the conditions (\ref{int.no.cond}). \end{enumerate} \end{cor} Proof. The steps of an MMP are everywhere defined by (\ref{sm.mmp.cor1}). If $g:U\to V$ is any divisorial contraction over $\r$ then $\ex(g^{-1})(\r)$ is finite unless $g$ is a divisor--to--curve contraction which is not $\r$-small. Let $f_j$ be the first divisor--to--curve contraction in the sequence which is not $\r$-small. By the above remark, (5) holds for $j$. By (\ref{sm.top.thm}), $X_j(\r)$ contains a surface $F$ which is either a 1-sided torus or Klein bottle with nonorientable neighborhood. We can move $F$ away from any finitely many points, thus by (5) $X(\r)$ also contains a 1-sided torus or Klein bottle with nonorientable neighborhood. This is a contradiction. Hence among the steps there is no divisor--to--curve contraction which is not $\r$-small. This gives (5) and (6).\qed \medskip The proof of (\ref{sm.thm}) relies on two results: \begin{prop}\cite[Thm. 4]{Cutkosky88} \label{sm.cut} (\ref{sm.thm}) holds if $K$ is algebraically closed and $C$ is irreducible.\qed \end{prop} \begin{lem}\label{sm.tree}(cf. \cite[1.14]{Mori88}) Let $f:X\to Y$ be a proper morphism and $0\in Y$ a closed point such that $\dim f^{-1}(0)=1$. Set $\red f^{-1}(0)=C=\cup C_i$. \begin{enumerate} \item If $R^1f_*\o_X=0$ then $C$ is a tree of smooth rational curves. \item Let $D$ be a $\q$-Cartier Weil divisor on $X$ such that $D$ is Cartier at all but finitely many points of $f^{-1}(0)$. Assume that $(D\cdot C_i)<0$ for every $i$ and $R^1f_*\o_X(D)=0$. Then $-1\leq (D\cdot C_i)<0$ for every $i$ and $D$ is not Cartier at the singular points of $C$. \end{enumerate} \end{lem} Proof. By replacing $Y$ with a neighborhood of $0$, we may assume that every fiber of $f$ has dimension at most 1. Let $G$ be a sheaf on $X$ such that $R^1f_*G=0$ and $Q=G/F$ a quotient of $G$ whose support is in $f^{-1}(0)$. We get an exact sequence $$ R^1f_*G\to R^1f_*Q\to R^2f_*F. $$ The left hand side is zero by assumption and the right hand side is zero since every fiber of $f$ has dimension at most 1. Thus $R^1f_*Q=0$. Applying this with $G=\o_X$ and $Q=\o_C$ we conclude that $H^1(C,\o_C)=R^1f_*\o_C=0$, hence $C$ is a tree of smooth rational curves. This proves (1). In order to see the second part, we may assume that the residue field of $0$ is algebraically closed. Then a point $P\in C$ is singular iff there are at least 2 irreducible components through $P$. $\o_X(D)\otimes \o_{C_i}$ is a rank one locally free sheaf except possibly at the ponts where $D$ is not Cartier. Let $L_i$ denote its quotient by the torsion subsheaf. Then $L_i$ is an invertible sheaf and we have a surjection $\o_X(D)\to L_i$. Applying $R^1f_*$ we obtain as above that $H^1(C_i,L_i)=0$. Thus $\deg L_i\geq -1$. On the other hand, for every $m>0$ we have an injection $$ L_i^m\cong (\o_X(D)^{\otimes m}\otimes \o_{C_i})/(\mbox{torsion}) \DOTSB\lhook\joinrel\rightarrow (\o_X(mD)\otimes \o_{C_i})/(\mbox{torsion}). $$ If $mD$ is Cartier then the right hand side has negative degree, thus $L_i^m$ has negative degree. Therefore $\deg L_i=-1$ for every $i$. Furthermore, $m(D\cdot C_i)\geq m\deg L_i=-m$, so $(D\cdot C_i)\geq -1$. Set $M:=(\o_X(D)\otimes \o_{C})/)\mbox{torsion})$. $H^1(C,M)=0$ as above. We have an exact sequence $$ 0\to M\to \sum L_i\to Q\to 0, $$ where $Q$ is supported at the singular points of $C$. Taking cohomologies, we conclude that $H^0(C,Q)=0$ thus $Q=0$. If $D$ is Cartier at a singular point $P$ of $C$ then $M$ is locally free at $P$ and $M\to \sum L_i$ can not be surjective at $P$ (it is not even surjective when tensored with the residue field at $P$).\qed \begin{cor}\label{sm.tree2} Notation and assumptions as in (\ref{sm.not}). Then $C$ is a tree of smooth rational curves and $K_X$ is not Cartier at the singular points of $C$. \end{cor} Proof. Apply (\ref{sm.tree}) with $D=K_X$. $R^1f_*\o_X=R^1f_*\o_X(K_X)=0$ by (\ref{sm.not}). \qed \begin{say}[Proof of (\ref{sm.thm})] The assumptions and conclusions are local near $0$, thus we may replace $Y$ by a suitable analytic or formal neighborhood of $0$. By (\ref{sm.tree2}), $C$ is a connected tree of smooth rational curves. $\gal(\bar K/K)$ acts on $C$, thus $C$ either has a singular $K$-point or a geometrically irreducible component defined over $K$. If $P\in C$ is a singular point then $K_X$ is not Cartier at $P$ by (\ref{sm.tree2}), but if $P\in X(K)$ then $K_X$ is Cartier at $P$ by assumption. Thus $C$ can not have a singular $K$-point. Let $C_0\subset C$ be a geometrically irreducible component defined over $K$. Let $H\subset X$ be a divisor defined over $K$ which intersects all irreducible components of $C\setminus C_0$ transversally but is disjoint from $C_0$. A large multiple of $H$ defines a morphism $X\to Y'\to Y$ such that $C_0$ is contracted to a point in $Y'$. If (\ref{sm.thm}) holds for $X\to Y'$, then $X$ has index one along $C_0$. By (\ref{sm.tree2}) this implies that $C_0$ is a connected component of $C$. On the other hand, $C$ is connected since $f_*\o_X=\o_Y$. Thus $C=C_0$ and $Y'=Y$. Therefore it is sufficient to prove (\ref{sm.thm}) under the additional assumption that $C$ is geometrically irreducible. First we show that (\ref{sm.thm}) holds if $X$ has only index 1 points along $C$. By (\ref{sm.cut}), $Y$ is smooth at $0$ and $X=B_DY$ where $D\subset Y$ is a curve of embedding dimension 2. $D$ is the image of the exceptional divisor of $f$, hence $D$ is defined over $K$. Since $D$ has embedding dimension 2, its ideal is of the form $(z,g(x,y))$. Finally we show that $X$ has only index 1 points along $C$. We start with the case when $K=\r$. Let $P_1,\bar P_1, \dots, P_k,\bar P_k$ be all the conjugate pairs of points of index $>1$. At each $P_i$ pick a local member $D_i\in |K_X|$ such that $C\cap D_i=P_i$. (In order to do this, we may need to replace $X_{\bar K}$ with a smaller analytic neighborhood of $C$.) Let $\bar D_i$ be the conjugates. Set $D=\sum D_i$. Let $m>1$ be the smallest natural number such that $mD$ is Cartier. $D-\bar D$ is a Weil divisor and $\o_X(m(D-\bar D))\cong \o_X$ since the Picard group of a neighborhood of $C$ is isomorphic to $H^2(C(\c),\z)$ (cf. \cite[4.13]{KoMo98}). Corresponding to $1\in H^0(X,\o_X)$ we obtain an $m$-sheeted cyclic cover $\pi: \tilde X\to X$ which is unramified outside the points of index $>1$. Thus $K_{\tilde X}=\pi^*K_X$ and $\tilde X$ has index 1 terminal singularities. Let $\tilde f:\tilde X\to \tilde Y$ be the Stein factorization of $\tilde X\to Y$. By the already discussed index 1 case, $\tilde Y$ is smooth and one can choose local analytic coordinates $(x,y,z)$ at $0\in \tilde Y$ such that $\tilde X$ is the blow up of the curve $(z=g(x,y)=0)\subset \tilde Y$. The group of $m^{th}$ roots of unity (denoted by $\z_m$) acts on $\tilde f:\tilde X\to \tilde Y$ and the quotient is $ f: X\to Y$. If $\mult_0g\geq 2$ then $\tilde X$ has a unique singular point (\ref{sm.ci.blowup}), which is necessarily fixed by the $\z_m$-action. Thus $X$ would have a unique point (of index $m$) which is the quotient of a singular point. On the other hand, the index $>1$ singularities of $X$ come in conjugate pairs. Therefore $\tilde X$ is smooth and $\tilde f:\tilde X\to \tilde Y$ is the blow up of a smooth curve $(z=y=0)\subset \tilde Y$. We can choose local coordinates $(x,y,z)$ on $\tilde Y$ such that the action is $$ (x,y,z)\mapsto (\epsilon^a x,\epsilon^by, \epsilon^cz) $$ where $\epsilon$ is a primitive $m^{th}$ root of unity. The corresponding action on $\tilde X$ has two fixed points (or a fixed curve) and the corresponding quotients are $$ \c^3/{\textstyle \frac1{m}}(a,b-c,c)\qtq{and} \c^3/{\textstyle \frac1{m}}(a,b,c-b). $$ These are both of type $cA_0/n$ on the list (\ref{mmp.ts.thm}). A simple checking shows that both of these can not be simultaneously terminal. If $K$ is arbitrary, we can still proceed as above if we can find local divisors $D_i\in |K_X|$ at the index $>1$ points such that $(C\cdot \sum D_i)=0$. Finding the $D_i$ needs a little case by case analysis, and sometimes it can be done only after first taking an auxiliary cover. It is probably easier to observe that there can be at most 2 points of index $>1$ along $C$ (see, for instance, \cite[14.5.5]{CKM88}), thus in fact the only case we need to handle is when there is precisely one pair of conjugate points of index $>1$. \qed \end{say} \section{Proof of the Main Theorems} The determination of all divisor--to--point g--extractions is rather technical and lengthy. In this section I state a summary of the list of all g--extractions, and then use it to prove the main theorems stated in the introduction. The proofs of (\ref{ge.gwsing.thm}) and of (\ref{ge.gwextr.thm}) are given in sections 9--11. \begin{notation}\label{7.notation} Let $g(x_1,\dots,x_m)$ be a polynomial or power series and let $M$ be a monomial in the $x_i$. $M\in g$ means that $M$ appears in $g$ with nonzero coefficient. \end{notation} \begin{thm}\label{ge.gwsing.thm} Let $0\in X$ be a three dimensional terminal singularity over $\r$. If $X$ has a g--extraction then $0\in X$ is one of the following (up to real analytic equivalence near $0$). \begin{enumerate} \item ($cA_0$) Smooth point. \item ($cA_0/2$) Quotient of a smooth point by the $\z_2$-action $(x,y,z)\mapsto (-x,-y,-z)$. \item ($cA_{>0}^+$) Given as $(x^2+y^2+g_{\geq m}(z,t)=0)$, where $g_m(z,t)\neq 0$ and $m\geq 2$. \item ($cA_{>0}^+/2$) Given as $(x^2+y^2+g_{\geq m}(z,t)=0)/\z_2$, where $g_m(z,t)\neq 0$, $m\geq 2$ and the $\z_2$-action is $(x,y,z,t)\mapsto (-x,-y,-z,t)$. Furthermore, one of the following two conditions has to be satisfied: \begin{enumerate} \item $m$ is divisible by $4$ and $z^m,t^m\in g$, or \item\label{7.bad.index.2} $m$ is odd. \end{enumerate} \item ($cE_6$) Given as $(x^2+y^3+(z^2+t^2)^2+ yg_{\geq 4}(z,t)+h_{\geq 6}(z,t)=0)$. \end{enumerate} \end{thm} \begin{complement} All the singularities on the above list have g--extractions, with the possible exception of types (\ref{7.bad.index.2}). These singularities have not been analyzed completely. \end{complement} \begin{thm}\label{ge.gwextr.thm} Let $0\in X$ be a three dimensional terminal singularity over $\r$ and $f:Y\to X$ a g--extraction with exceptional divisor $E=\red f^{-1}(0)$. If $E$ is geometrically irreducible then $f:Y\to X$ is on the following list (up to real analytic equivalence near $0$). \begin{enumerate} \item ($cA_0$, point blow up) $B_0{\Bbb A}^3\to {\Bbb A}^3$, $E\cong \p^2$. \item ($cA_0$, curve blow up) $B_C{\Bbb A}^3\to {\Bbb A}^3$ where $C\subset {\Bbb A}^3$ is a geometrically irreducible, real and locally planar curve. \item ($cA_0/2$) $B_0{\Bbb A}^3/\z_2\to {\Bbb A}^3/\z_2$, where the $\z_2$-action on ${\Bbb A}^3$ is $(x,y,z)\mapsto (-x,-y,-z)$. $E\cong \p^2$. \noindent Furthermore, in this case there are no other g--extractions whose center contains the origin. \item ($cA_{>0}^+$, $\mult_0g$ even) $X=(x^2+y^2+g_{\geq 2m}(z,t)=0)$ where $g_{2m}(z,t)\neq 0$ and $m\geq 1$. $Y=B_{(m,m,1,1)}X$ and $E=(x^2+y^2+g_{2m}(z,t)=0)\subset \p^3(m,m,1,1)$. \noindent Furthermore, in this case there are no other g--extractions whose center contains the origin. \item ($cA_{>0}^+$, $\mult_0g$ odd) $X=(x^2+y^2+g_{\geq 2m+1}(z,t)=0)$ where $g_{2m+1}(z,t)\neq 0$ and $m\geq 1$. This case occurs only if there is a linear change of the $(z,t)$-coordinates such that $t^{2m+1}\in g$ and $z^it^j\not\in g$ for $i+2j< 4m+2$. In this coordinate system, $Y=B_{(2m+1,2m+1,1,2)}X$ and $E=(x^2+y^2+g_{2m+1}(z,t)=0)\subset \p^3(2m+1,2m+1,1,2)$. \item ($cA_{>0}^+/2$, $\mult_0g$ even) $X=(x^2+y^2+g_{\geq 2m}(z,t)=0)/\z_2$, where $g_{2m}(z,t)\neq 0$, $m\geq 1$ and the $\z_2$-action is $(x,y,z,t)\mapsto (-x,-y,-z,t)$. This case occurs only if $m$ is even and $z^{2m},t^{2m}\in g$. Then $Y=B_{(m,m,1,1)}\tilde X/\z_2$ and $E=\tilde E/\z_2$, where \noindent $\tilde X=(x^2+y^2+g_{\geq 2m}(z,t)=0)$ and \noindent $\tilde E=(x^2+y^2+g_{2m}(z,t)=0)\subset \p^3(m,m,1,1)$. Furthermore, in this case there are no other g--extractions whose center contains the origin. \item ($cA_{>0}^+/2$, $\mult_0g$ odd) $X=(x^2+y^2+g_{\geq 2m+1}(z,t)=0)/\z_2$ where $g_{2m+1}(z,t)\neq 0$, $m\geq 1$ and the $\z_2$-action is $(x,y,z,t)\mapsto (-x,-y,-z,t)$. In this case I do not have a complete list. \end{enumerate} \end{thm} \begin{cor}\label{ge.Enorm.cor} Let $0\in X$ be a three dimensional terminal singularity over $\r$ and $f:Y\to X$ a g--extraction with exceptional divisor $E=\red f^{-1}(0)$. Assume that we are not in case (\ref{ge.gwextr.thm}.7). If $E$ is geometrically irreducible then $E$ is normal. \end{cor} Proof. Equations for $E$ are given in (\ref{ge.gwextr.thm}). $E\cong \p^2$ in the first two cases. In the remaining cases $E$ is (or is the quotient of) a surface of the form $$ F:=(x^2+y^2+p(z,t)=0)\subset \p^3(r,r,1,s). $$ All the singularities of $F$ are contained in the $(x=y=0)$ line. Thus we get only finitely many singularities if $p$ is not identically zero, which is always the case in (\ref{ge.gwextr.thm}).\qed \begin{say}[Proof of (\ref{ge.gwsing.thm}) and (\ref{ge.gwextr.thm}) $\Rightarrow$ (\ref{int.mmp.sings}) and (\ref{int.mmp.steps})]{\ } Under the additional assumption that $X(\r)$ satisfies the conditions (\ref{int.no.cond}), we need to exclude the cases (\ref{ge.gwsing.thm}.2), (\ref{ge.gwsing.thm}.4), (\ref{ge.gwextr.thm}.2) and in (\ref{ge.gwsing.thm}.3) we need to show that $g$ is not everywhere negative in a punctured neighborhood of $0$. Starting with $X$, let us run the MMP over $\r$. We get a sequence $$ X=X_0\map X_1\map \cdots\map X_{i}\stackrel{f_{i}}{\map} X_{i+1}. $$ Assume by induction that (\ref{int.mmp.sings}) holds for $X_j$ for $j\leq i$ and (\ref{int.mmp.steps}) holds for $f_j:X_j\map X_{j+1}$ for $j\leq i-1$. We need to show that (\ref{int.mmp.sings}) holds for $X_{i+1}$ and (\ref{int.mmp.steps}) holds for $f_i:X_i\map X_{i+1}$. By (\ref{sm.mmp.cor2}) $f_j:X_j(\r)\to X_{j+1}(\r)$ are everywhere defined for $j\leq i-1$ and $\overline{X_{i}(\r)}$ does not contain a 1-sided torus or Klein bottle with nonorientable neighborhood. Furthermore, by (\ref{sm.top.thm}), $f_i$ is either $\r$-small or a divisor--to--point contraction. By induction $X_i$ has index 1 along $X_i(\r)$, thus $f_i$ is a g--extraction, hence it is one of the cases listed in (\ref{ge.gwsing.thm}). We excluded several cases one at a time. \medskip {\it Excluding (\ref{ge.gwsing.thm}.2).} By (\ref{ge.gwextr.thm}.3) $X_i\to X_{i+1}$ is the blow up of the singular point $\a^3/\z_2$. This gives a 1--sided $\r\p^2$ in $X_i(\r)$. This is a contradiction by (\ref{sm.mmp.cor2}). \medskip {\it Excluding (\ref{ge.gwextr.thm}.2).} In this case $X_i$ contains a 1-sided torus or Klein bottle with nonorientable neighborhood by (\ref{sm.top.thm}). This is again a contradiction by (\ref{sm.mmp.cor2}). \medskip {\it Excluding (\ref{ge.gwsing.thm}.3) with $g<0$.} That is, we consider the case when $0\in X_{i+1}$ is of the form $(x^2+y^2+g_{\geq m}(z,t)=0)$ and $g$ is everywhere negative in a punctured neighborhood of $0$. (These are called $cA^+_{>0}(0,-)$ in \cite{rat1}.) By \cite[4.4]{rat1} the link of $0\in X_{i+1}(\r)$ is a torus. This gives only a 2-sided torus in $X(\r)$ which is allowed. I proceed to prove, however, that we still get a 1-sided torus in $X(\r)$ coming from the exceptional divisor of $X(\r)\to X_{i+1}(\r)$. This contradicts (\ref{int.no.cond}.2). $m=\mult_0g(z,t)$ is necessarily even, say $m=2r$. By (\ref{ge.gwextr.thm}.4) the only g--extraction is the $(r,r,1,1)$-blow up. Thus $X_i\to X_{i+1}$ is this blow up. We distinguish two cases: General case: $g_{2r}(z,t)$ is negative on $\r^2\setminus\{0\}$. The exceptional divisor $E$ of the above g--extraction is the weighted hypersurface $$ E=(x^2+ y^2+g_{2r}(z,t)=0)\subset \p(r,r,1,1). $$ Its canonical divisor is $K_E=\o_E(-2)$, thus $E$ is orientable. The projection $(x:y:z:t)\mapsto (z:t)$ exhibits $E$ as an $S^1$-bundle over $\r\p^1$, thus $E\sim S^1\times S^1$. $L(0\in X_{i+1}(\r))$ is connected, thus $E(\r)\subset X_i(\r)$ is a 1-sided torus. By (\ref{sm.mmp.cor2}), $X(\r)$ also contains a 1-sided torus, a contradiction. Special case: $g_{2r}(z,t)$ is not negative on $\r^2\setminus\{0\}$. $g_{2r}(z,t)$ is the leading term of $g_{\geq m}(z,t)$, which is negative on $\r^2\setminus\{0\}$. Thus $g_{2r}(z,t)$ is nonpositive on $\r^2\setminus\{0\}$. The $t$-chart on $X_i\cong B_{(r,r,1,1)}X_{i+1}$ is $x_1^2+ y_1^2+ t_1^{-2m}g(z_1t_1,t_1)$. Set $g'(z_1,t_1):=t_1^{-2m}g(z_1t_1,t_1)$. Then $g'(z_1,t_1)$ is strictly negative outside the $z_1$-axis, and is not identically zero on the $z_1$-axis. Thus $g'(z_1,t_1)$ is everywhere nonpositive with only finitely many zeros. Thus at each zero of $g'(z_1,t_1)$, $X_i$ has a singular point of type $cA^+_{>0}(0,-)$. This contradicts the inductive assumption. Thus $X_{i+1}$ does not contain any $cA^+_{>0}(0,-)$ type points. \medskip {\it Excluding (\ref{ge.gwsing.thm}.4).} \cite[5.9]{rat1} shows that the link of a singularity of type $cA^+/2$ contains a connected component homeomorphic to $\r\p^2$, except when we can write the singularity as $(x^2+y^2+g_{\geq m}(z,t)=0)/\z_2$, where the $\z_2$-action is $(x,y,z,t)\mapsto (-x,-y,-z,t)$ and $g$ is everywhere negative in a punctured neighborhood of $0$. If the link of a singularity of $X_{i+1}(\r)$ contains a connected component homeomorphic to $\r\p^2$, then $X_i(\r)$ contains a 2-sided $\r\p^2$. Hence by (\ref{sm.mmp.cor2}), $X(\r)$ also contains a 2-sided $\r\p^2$ which is excluded by (\ref{int.no.cond}.1). Thus we are reduced to the case $(x^2+y^2+g_{\geq m}(z,t)=0)/\z_2$, where the $\z_2$-action is $(x,y,z,t)\mapsto (-x,-y,-z,t)$ and $g$ is everywhere negative in a punctured neighborhood of $0$. As in the previous case, $m=\mult_0g(z,t)$ is necessarily even, say $m=2r$ and by (\ref{ge.gwextr.thm}.6) the only g--extraction is the $(r,r,1,1)$-blow up. We proceed to prove that the exceptional divisor of $X_i(\r)\to X_{i+1}(\r)$ contains a 1-sided Klein bottle with nonorientable neighborhood, contradicting (\ref{int.no.cond}.3). We again distinguish two cases: General case: $g_{2r}(z,t)$ is negative on $\r^2\setminus\{0\}$. The exceptional divisor $E$ is the $\z_2$ quotient of the weighted hypersurface $$ \tilde E=(x^2+ y^2+g_{2r}(z,t)=0)\subset \p(r,r,1,1). $$ We already determined that $\tilde E(\r)$ is a torus and a choice of orientation is given by $(dy\wedge dz)/x$. The $\z_2$-action sends this to $(d(-y)\wedge d(-z))/(-x)=-(dy\wedge dz)/x$, hence $\tilde E(\r)/\z_2$ is not orientable. We conclude that one of the connected components of $E(\r)$ is a Klein bottle. (There may be other connected components.) The Klein bottle is 1-sided since $\tilde E(\r)$ is 1-sided. The regular neighborhood is nonorientable since its boundary, the link of $0\in X(\r)$, is again a Klein bottle. Special case: $g_{2r}(z,t)$ is not negative on $\r^2\setminus\{0\}$. The same computation as above shows that this leads to a $cA^+/2$ point of the same type that we started with on $X_i$, which contradicts the inductive assumption. Thus we conclude that $X_{i+1}$ does not contain any $cA^+/2$ type points.\qed \end{say} \begin{say}[Proof of (\ref{int.orient.thm}) and (\ref{int.nonorient.thm})]{\ } We follow the steps of an MMP over $\r$, using (\ref{int.mmp.steps}). $f_i:X_i(\r)\to X_{i+1}(\r)$ is a homeomorphism in cases (\ref{int.mmp.steps}.1--2) while (\ref{int.mmp.steps}.3) gives a connected sum with $\r\p^3$. In the cases (\ref{int.mmp.steps}.4) the exceptional divisor is normal by (\ref{ge.Enorm.cor}), hence we get various cases of (\ref{int.orient.thm}) by (\ref{top.alg.normsurfup.lem}). \qed \end{say} \begin{exmp}\label{ge.2extr.exmp} Consider the singularity $X:=(x^2+y^2+z^{2m+1}+t^{4m+2}=0)$. The $(2m+1,2m+1,2,1)$ blow up $X_1\to X$ is a g--extraction which is smooth along the $\r$-points. The $(m,m,1,1)$-blow up is another g--extraction whith one singular point $(x_1^2+y_1^2+z_1^{2m+1}t_1+t_1^{2m+2}=0)$ on the $t$-chart. After the $(m+1,m+1,1,1)$-blow up we obtain a variety $X_2\to X$ which is smooth along its $\r$-points. These two resolutions are indeed quite different. Using the methods of section 5, we see that $X_1(\r)\sim X(\r)\ \#\ \r\p^3$ and $X_2(\r)\sim X(\r)\ \#\ S^1\times S^2$. \end{exmp} \section{$cAx$ and $cD$-type Points} In this section we begin to classify g--extractions (\ref{gw.g-e.def}) of terminal singularities over any field. The classification of 3--fold terminal singularities over nonclosed fields is done in \cite{rat1}. The results are summarized in (\ref{mmp.ts.thm}). We work through the list of the singularities. In most cases it is easy to see that there are no g--extractions. This is done by exhibiting an elementary extraction (\ref{elem.extr.defn}) which is not a g--extraction. If the discrepancy of the exceptional divisor is $\leq 1$ then there are no g--extractions by (\ref{gw.discr1.cor}). In this section we deal with the cases $cAx/2, cAx/4, cD, cD/2, cD/3$. Among terminal singularities these are somewhat esoteric but the proofs work well for them: in each case (\ref{gw.discr1.cor}) applies. The remaining terminal singularities are considered in the next 2 sections. In some cases much more complicated arguments are needed to classify all g--extractions. \begin{defn}\label{8.wbup}[Weighted blow-ups]{\ } Let $x_1,\dots,x_n$ be coordinates on $\a^n$. The usual blow up of the origin is patched together from affine charts with morphisms of the form $$ x_j=x'_jx'_i\qtq{if} j\neq i\qtq{and} x_i=x'_i. $$ I refer to this as the {\it $x_i$-chart}. Let $a_1,\dots,a_n$ be a sequence of positive integers. For every $1\leq i\leq n$ we can define a morphism $\Pi_i:\a^n\to \a^n$ by $$ x_j=x'_j(x'_i)^{a_j}\qtq{if} j\neq i\qtq{and} x_i=(x'_i)^{a_i}. $$ This morphism is birational iff $a_i=1$ and has degree $a_i$ in general. One can easily notice that $\Pi_i$ is invariant under the action $$ \a^n(x'_1,\dots,x'_n)/\textstyle{\frac1{a_i}} (-a_1,\dots,-a_{i-1},1,-a_{i+1},\dots,-a_n) $$ and it descend to a birational morphism $\pi_i$ $$ \Pi_i:\a^n(x'_1,\dots,x'_n)\to \a^n(x'_1,\dots,x'_n)/\z_{a_i} \stackrel{\pi_i}{\longrightarrow} \a^n(x_1,\dots,x_n). $$ Furthermore, these charts patch together to give a projective morphism $$ \pi:B_{(a_1,\dots,a_n)}\a^n\to \a^n. $$ This is called the {\it weighted blow up} of $\a^n$ with weights $a_1,\dots,a_n$. \end{defn} \begin{notation}\label{8.notation} In the proofs in sections 9--11 I use the following conventions. Firts I state the name of the singularity $X$ from (\ref{mmp.ts.thm}) and possibly some other restrictions. Then I write down the normal form of the equation $X=F(x,y,z,t)/\frac1{r}(b_x,b_y,b_z,b_t)$. Any restrictions on $F$ are explained in detail here. Then I specify the weights $(a_x,a_y,a_z,a_t)$ for a weighted blow up and write down the equation of the birational transform of $X$ on one of the charts on the weighted blow up. Before taking quotients, this has the form $t_1^{-m}F(x_1t_1^{a_x}, y_1t_1^{a_y}, z_1t_1^{a_z}, t_1^{a_t})$ if I use the $t$-chart. This is denoted by $B\tilde X$. I need to take quotient by 2 actions. First is the $\frac1{a_t}(-a_x,-a_y,-a_z,1)$-action coming from the weighted blow up. Second, the $\frac1{r}(b_x,b_y,b_z,b_t)$-action needs to be lifted to the $(x_1,y_1,z_1,t_1)$-space. In some cases this lifts as a $\z_r$-action but in other cases the actions combine into a $\z_{(ra_t)}$-action. The quotient of $B\tilde X$ by these 2 actions is a chart on the weighted blow up of $X$; it is denoted by $BX$. All these can be done in 4 different charts. I chose the chart where the singularities are most visible or the discrepancy computation is the clearest. Finally I compute the exceptional divisor of the blow up, the singularities of $BX$ and the discrepancy of the exceptional divisor. \end{notation} \begin{say}[$cAx/2$] {\ } \newline Normal form: $ax^2+by^2+g_{\geq 4}(z,t)/{\textstyle \frac12(1,0,1,1)}$, where $ab\neq 0$. Weights for blow-up: (1,1,1,1) $t$-chart: $ax_1^2+by_1^2+t_1^{-2}g_{\geq 4}(z_1t_1,t_1)/{\textstyle \frac12(0,1,0,1)}$ Exceptional divisor: $(t_1=ax_1^2+by_1^2=0)$. Over $\bar K$ this is reducible and the two irreducible components are $(t_1=\sqrt{a}x_1\pm \sqrt{-b}y_1=0)$. The $\z_2$-action interchanges these two, so on the quotient we get a geometrically irreducible exceptional divisor. Singularity: The $\z_2$-action has a fixed curve on $B\tilde X$: the intersection with the $(x_1=z_1=0)$-plane. Thus we get a curve of nonterminal singularities on $BX$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. \end{say} \begin{say}[$cAx/4$] {\ } \newline Normal form: $ax^2+by^2+g_{\geq 2}(z,t)/{\textstyle \frac14(1,3,1,2)}$, where $ab\neq 0$ and $g_2(0,1)=0$ for weight reasons. Weights for blow-up: (1,1,1,1) $t$-chart: $ax_1^2+by_1^2+t_1^{-2}g_{\geq 2}(z_1t_1,t_1)/{\textstyle \frac14(3,1,3,2)}$ Exceptional divisor: $\tilde E:=(t_1=ax_1^2+by_1^2+g_2(1,0)z_1^2=0)$. $\tilde E$ is geometrically irreducible if $g_2(1,0)\neq 0$. If $g_2(1,0)= 0$, then $\tilde E$ is reducible over $\bar K$, and the two irreducible components are $(t_1=\sqrt{a}x_1\pm \sqrt{-b}y_1=0)$. The $\z_4$-action interchanges these two, so on the quotient we get a geometrically irreducible exceptional divisor $E$. Singularity: The origin is on $B\tilde X$ since $g_2(0,1)=0$ and it is a fixed point. We get an index 4 point on $BX$ Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. \end{say} \begin{say}[$cD_4$ main series] {\ } \newline Normal form: $x^2+f_{\geq 3}(y,z,t)$, where we assume that $f_3(y,z,t)$ is irreducible over $\bar K$. Weights for blow-up: (2,1,1,1) $x$-chart: $x_1+x_1^{-3}f_{\geq 3}(y_1x_1, z_1x_1,t_1x_1)/{\textstyle \frac12(1,1,1,1)}$ Exceptional divisor: $\tilde E:=(x_1=f_3(y_1,z_1,t_1)=0)$. $\tilde E$ is geometrically irreducible by our assumption. Singularity: The origin is a fixed point on $B\tilde X$, hence we get an index 2 point on $BX$ Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= x_1\cdot dy_1\wedge dz_1\wedge dt_1$, so $a(E,X)=1$. \end{say} \begin{say}[$cD_4/2$ main series] {\ } \newline Normal form: $x^2+f_{\geq 3}(y,z,t)/{\textstyle \frac12(1,1,0,1)}$, where we assume that $f_3(y,z,t)$ is irreducible over $\bar K$. However, for weight reasons $z|f_3(y,z,t)$, so this can not happen. \end{say} \begin{say}[$cD/3$] {\ } \newline Normal form: $x^2+f_{\geq 3}(y,z,t)/{\textstyle \frac13(0,1,1,2)}$, where $f_3(0,0,t)\neq 0$. Since this is not a $cE$ point and for weight reasons, also $f_3(y,z,0)\neq 0$. We can write $f_3=t^3+f_3(y,z,0)$. Weights for blow-up: (2,1,1,1) $x$-chart: $x_1+x_1^{-3}f_{\geq 3}(y_1x_1, z_1x_1,t_1x_1)/{\textstyle \frac12(1,1,1,1)}$, and then take the $\z_3$-action. Lifting of the $\z_3$-action: It lifts to $\frac16(3,5,5,1)$. Exceptional divisor: $\tilde E:=(x_1=f_3(y_1,z_1,t_1)=0)$. $\tilde E$ is geometrically irreducible if $f_3(y,z,0)$ is not a cube. If $f_3(y,z,0)=-L(y,z)^3$ over $\bar K$, then $\tilde E$ has three geometrically irreducible components $(x_1=t_1-\eta L(y_1,z_1)=0)$ where $\eta^3=1$. The $\z_6$-action permutes these, so on $BX$ we get a geometrically irreducible exceptional divisor. Singularity: The origin is a $\z_6$-fixed point which has multiplicity 1 on $B\tilde X$. $BX$ has a terminal quotient singularity of index 6. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= x_1\cdot dy_1\wedge dz_1\wedge dt_1$, so $a(E,X)=1$. \end{say} \begin{say}[$cD_{>4}$ and special $cD_4$] {\ } \newline Normal form: $x^2+Q_2(y,z,t)z+g_{\geq 4}(y,z,t)$, where $Q_2(y,0,t)\neq 0$. In the $cD_{>4}$ we always have this form (with $Q_2(y,z,t)=y^2$). In the $cD_4$-case we can achieve this form iff $f_3(y,z,t)$ has a simple linear factor over $K$. Weights for blow-up: (2,1,2,1) $z$-chart: $x_1^2+Q_2(y_1,z_1,t_1)+z_1^{-4}g_{\geq 4}(y_1z_1, z_1^2,t_1z_1)/{\textstyle \frac12(0,1,1,1)}$ Exceptional divisor: $\tilde E:=(z_1=x_1^2+Q_2(y_1,0,t_1)+g_4(y_1,0,t_1)=0)$. $\tilde E$ is geometrically irreducible iff $Q_2(y_1,0,t_1)$ is not a square over $\bar K$ or $g_4(y_1,0,t_1)\neq 0$. If $Q_2(y_1,0,t_1)=-L_1(y_1,t_1)^2$ (over $\bar K$) and $g_4(y_1,0,t_1)=0$ then $\tilde E$ is reducible over $\bar K$, and the two irreducible components are $(z_1=x_1\pm L_1(y_1,t_1)=0)$. The $\z_2$-action interchanges these two, so $E\subset BX$ is geometrically irreducible. Singularity: The origin is a fixed point on $B\tilde X$, hence we get an index 2 point on $BX$ Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2z_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. \end{say} \begin{say}[$cD_{>4}/2$ and special $cD_4/2$] {\ } \newline Normal form: $x^2+Q_2(y,z,t)z+g_{\geq 4}(y,z,t)/{\textstyle \frac12(1,1,0,1)}$. Weights for blow-up: (2,1,2,1) $z$-chart: $x_1^2+Q_2(y_1,z_1,t_1)+z_1^{-4}g_{\geq 4}(y_1z_1, z_1^2,t_1z_1)/{\textstyle \frac12(0,1,1,1)}$ and then take the $\z_2$-action. Lifting of the $\z_2$-action: We get a pair of commuting $\z_2$-action on $B\tilde X$, given by $\frac12(0,1,1,1)$ and $\frac12(1,0,1,0)$. Singularity: The second action has a fixed curve on $B\tilde X$, so $BX$ is singular along a curve. Exceptional divisor and discrepancy: as in the $cD_{>4}$-case. \end{say} \begin{say}[$cD$-cases, conclusion] {\ } \newline We have settled all the $cD_{>4}$, $cD/2$ and $cD/3$ cases, they have no g--extractions. In the $cD_4$ cases there are no g--extractions if $f_3$ is irreducible or if it has a simple linear factor over $K$. The only remaining case is when $f_3$ is the product of 3 linear factors which are conjugate over $K$. This can not happen when $K=\r$, so over $\r$ points of type $cD, cD/2$ and $cD/3$ do not have g--extractions. The situation is more complicated over fields which do have cubic extensions, as the following example shows. I have not classified all cases. \end{say} \begin{exmp} Consider $x^2+y^3+az^3+t^6$, where $a\in K$ is not a cube. The exceptional divisor of the $(3,2,2,1)$-blow up is irreducible and has discrepancy 1. It has three points of index 2 which are conjugate over $K$, and no other singularities. Hence this is a g--extraction. \end{exmp} \begin{exmp} We obtain an interesting example from the equation $x^2+(y^2+z^2)z+t^5$. The $(2,1,2,1)$ blow up has terminal singularities (one with index 2). The exceptional divisor $E$ is singular along a curve. \end{exmp} \section{$cA$-type Points} In this section we study g--extractions of $cA$ type terminal singularities. The conventions of (\ref{7.notation}), (\ref{8.wbup}) and of (\ref{8.notation}) are used throughout. \begin{say}[$cA_0$] (That is, smooth points.) {\ } \newline Normal form: ${\Bbb A}^3$. The blow up of the origin is smooth with exceptional divisor $E\cong \p^2$. $a(E,X)=2$, and by (\ref{gw.d-ineq.prop}) $a(F,X)\geq 2$ for every exceptional divisor $F$ with $\cent_XF=\{0\}$. Therefore by (\ref{gw.discr2.cor}), the blow up of the origin is the only g--extraction. The exceptional divisor is $E\cong \p^2$ with normal bundle $\o_{\p^2}(-1)$. \end{say} \begin{say}[$cA_0/n,\ n\geq 2$] {\ } \newline Normal form: ${\Bbb A}^3/{\textstyle \frac1{n}(r,-r,1)}$, where $(r,n)=1$ and $1\leq r\leq n-1$. Weights for blow-up: (r,n-r,1). $x$-chart: ${\Bbb A}^3(x_1,y_1,z_1)/{\textstyle \frac1{r}(1,-n,-1)}$. Exceptional divisor: $\tilde E:=(x_1=0)$. Geometrically irreducible and invariant under the $\z_r$-action. Lifting of the $\z_n$-action: The $\z_n$-action lifts to $\frac1{n}(1,0,0)$. Its invariants are $x_2:=x_1^n$ and $y_1,z_1$. The $\z_r$-action descends to the quotient of the $\z_n$-action as ${\Bbb A}^3(x_2,y_1,z_1)/{\textstyle \frac1{r}(n,-n,-1)}$. Singularity: We obtain an index $r$ point on the $x$-chart, and similarly an index $n-r$ point on the $y$-chart. Discrepancy: $\pi^*dx\wedge dy\wedge dz= rx_1^ndx_1\wedge dy_1\wedge dz_1= \frac{r}{n}x_1dx_2\wedge dy_1\wedge dz_1$. Since $x_1=x_2^{1/n}$, we obtain that $a(E,X)=1/n$. Conclusion: The above blow up is the only possible g--extraction. If $n\geq 3$ then either $r\geq 2$ or $n-r\geq 2$, and we obtain a singular point of index $\geq 2$ on $BX$. If $r=2$ then $BX$ is smooth, the exceptional divisor is $E\cong \p^2$ with normal bundle $\o_{\p^2}(-2)$. $BX\to X$ is the unique g--extraction. \end{say} \begin{say}[$cA_1$] {\ } \newline Normal form: $ax^2+by^2+cz^2 +dt^m$, where $abcd\neq 0$. Weights for blow-up: (1,1,1,1) $t$-chart: $ax_1^2+by_1^2+cz_1^2 +dt_1^{m-2}$. Exceptional divisor: $E:=(t_1=ax_1^2+by_1^2+cz_1^2=0)$ for $m\geq 3$ and $(t_1=ax_1^2+by_1^2+cz_1^2+d=0)$ for $m=2$. $E$ is geometrically irreducible. Singularity: $BX$ has exactly one singular point for $m\geq 4$, it lies on the $t$-chart. $BX$ is smooth for $m=2,3$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: The only g--extraction is this blow up. The singularities can be resolved by repeatedly blowing up the unique singular point. \end{say} \begin{say}[$cA_1/2$] {\ } \newline Normal form: $ax^2+by^2+cz^n +dt^m/{\textstyle \frac12(1,1,1,0)}$, where $abcd\neq 0$ and $\min\{n,m\}=2$. Weights for blow-up: (1,1,1,1) $z$-chart: $ax_1^2+by_1^2+cz_1^{n-2} +dt_1^mz_1^{m-2}$. Exceptional divisor: $\tilde E:=(z_1=ax_1^2+by_1^2+c=0)$ for $m\geq 3$, $(z_1=ax_1^2+by_1^2+dt_1^2=0)$ for $n\geq 3$ and $(z_1=ax_1^2+by_1^2+c+dt_1^2=0)$ for $n=m=2$. $E$ is geometrically irreducible. Singularity: The $\frac12(1,1,1,0)$ action lifts to a $\frac12(0,0,1,1)$ action. Thus we get a fixed curve, where the blow up intersects the plane $(z_1=t_1=0)$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= z_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: The only possible g--extraction is this blow up. It has nonterminal singularities, so this does not occur. \end{say} \begin{say}[$cA_1/n,\ n\geq 3$] {\ } \newline Normal form: $xy+cz^{pm}+dt^2/{\textstyle \frac1{n}(r,-r,1,0)}$, where $(r,n)=1$ and $cd\neq 0$. Weights for blow-up: (1,1,1,1) $z$-chart: $x_1y_1+cz_1^{pm-2}+dt_1^2/{\textstyle \frac1{n}(r-1,1-r,1,-1)}$ Exceptional divisor: $\tilde E:=(z_1=x_1y_1+dt_1^2=0)$, it is geometrically irreducible. Singularity: The $\z_n$-action has an isolated fixed point at the origin on $B\tilde X$. Thus $BX$ has an index $n$ point. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= z_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: The only possible g--extraction is this blow up. It has a higher index point, so this does not occur. \end{say} \begin{say}[$cA_{>1}^-$] {\ } \newline Normal form: $xy+g_{\geq 3}(z,t)$. Weights for blow-up: (1,1,1,1) $t$-chart: $x_1y_1+t_1^{-2}g_{\geq 3}(z_1t_1,t_1)$. Exceptional divisor: $E:=(t_1=x_1y_1=0)$. It has two geometrically irreducible components. Singularity: Not important Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: At least 2 geometrically irreducible divisors with discrepancy $\leq 1$. \end{say} \begin{say}[$cA_{>1}/n,\ n\geq 3$ and $cA_{>1}^-/2$] {\ } \newline Normal form: $xy+g_{\geq 3}(z,t)/{\textstyle \frac1{n}(r,-r,1,0)}$, where $(r,n)=1$. Weights for blow-up: (1,1,1,1) $t$-chart: $x_1y_1+t_1^{-2}g_{\geq 3}(z_1t_1,t_1)/{\textstyle \frac1{n}(r,-r,1,0)}$ Exceptional divisor: $\tilde E:=(t_1=x_1y_1=0)$. It is reducible and both irreducible components are geometrically irreducible and invariant under the $\z_n$-action. Singularity: Not important Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= z_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: At least 2 geometrically irreducible divisors with discrepancy $\leq 1$. \end{say} \begin{say}[$cA_{>1}^+$, $\mult_0g$ even] {\ } \newline Normal form: $ax^2+by^2+g_{\geq 2m}(z,t)$, where $m\geq 2$, $-ab$ is not a square and $g_{2m}\neq 0$. Weights for blow-up: (m,m,1,1) $t$-chart: $ax_1^2+by_1^2+t_1^{-2m}g_{\geq 2m}(z_1t_1,t_1)$. Exceptional divisor: $E:=(t_1=ax_1^2+by_1^2+g_{2m}(z_1,1)=0)$, it is geometrically irreducible. Singularity: The $t$-chart on $BX$ is singular only at points $P$ corresponding to the multiple roots of $g_{2m}(z,1)$. The singularity at $P$ again has type $cA_{>1}^+$, but the multiplicity of the corresponding $g^P(z_1,t_1)$ is not necessarily even. The $z$ chart is similar. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. $x$-chart: $a+by_1^2+x_1^{-2m}g_{\geq 2m}(z_1x_1,t_1x_1)/ {\textstyle \frac1{m}(1,0,-1,-1)}$. Singularity: The fixed points of the $\z_m$-action are along the $y_1$-axis, this itersects $B\tilde X$ in two points $(0,\sqrt{-a/b},0,0)$ which are conjugate over $K$. Thus $BX$ has 2 index $m$ terminal singularities which are conjugate over $K$. No other new singular points. The $y$-chart is similar. Conclusion: The only g--extraction is the above weighted blow up. The exceptional divisor is geometrically irreducible with a pair of conjugate index $m$-points. The other singular $K$-points of $BX$ are again of type $cA_{>1}^+$ or $cA_1$. \end{say} \begin{say}[$cA^+_{>1}/2$, $\mult_0g$ even] {\ } \newline Normal form: $ax^2+by^2+ g_{\geq 2m}(z,t)/{\textstyle \frac12(1,1,1,0)}$, where $-ab$ is not a square and $g_{2m}\neq 0$. Weights for blow up: $(m,m,1,1)$. $z$-chart: $ax_1^2+by_1^2+z_1^{-2m}g_{\geq 2m}(z_1,t_1z_1) /{\textstyle \frac12(1-m,1-m,1,1)}$. Exceptional divisor: $\tilde E:=(z_1=ax_1^2+by_1^2+g_{2m}(1,t_1)=0)$ is geometrically irreducible. Singularities: If $m$ is odd then $(z_1=t_1=0)$ intersects $B\tilde X$ in a fixed curve of the $\z_2$-action, thus we get a singular curve on $BX$. If $m$ is even, then on the $z$-chart the only $\z_2$-fixed point is the origin. This is not on $B\tilde X$ iff $z^{2m}\in g$. $t$-chart: $ax_1^2+by_1^2+t_1^{-2m}g_{\geq 2m}(z_1t_1, t_1) /{\textstyle \frac12(1,1,1,0)}$. Singularities: On the $t$-chart the fixed point set is the $t_1$-axis. This intersects the exceptional divisor at the origin. This is not on the blow up iff $t^{2m}\in g$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. $x$-chart: $a+by_1^2+x_1^{-2m}g_{\geq 2m}(z_1x_1, t_1x_1) /{\textstyle \frac1{m}(1,0,-1,-1)}$ and we also need to take the quotient by the $\z_2$-action. Lifting the $\z_2$-action: $a+by_1^2+x_1^{-2m}g_{\geq 2m}(z_1x_1, t_1x_1) /{\textstyle \frac1{2m}(1,0,m-1,-1)}$ Singularities: On the $x$-chart the fixed point set is the $y_1$-axis. This intersects $B\tilde X$ at two points $(0,\pm\sqrt{-a/b},0,0)$. We get a conjugate pair of terminal singularities of index $2m$ on $BX$. $y$-chart: Similar to the $x$-chart. Conclusion: $ax^2+by^2 +g_{\geq 2m}(z,t)/{\textstyle \frac12(1,1,1,0)}$ where $-ab$ is not a square has a g--extraction iff $m$ is even and $z^{2m},t^{2m}\in g_{2m}(z,t)$. Under these assumptions, the unique g--extraction is the $(m,m,1,1)$-blow up. \end{say} \begin{say}[$cA_{>1}^+$, $\mult_0g$ odd]\label{cA+.multg-odd} {\ } \newline Normal form: $ax^2+by^2+g_{\geq 2m+1}(z,t)$, where $m\geq 1$, $-ab$ is not a square and $g_{2m+1}\neq 0$. Weights for blow-up: $(s,s,1,1)$ for $1\leq s\leq m$, giving $B_sX\to X$. $t$-chart: $ax_1^2+by_1^2+t_1^{-2s}g_{\geq 2m+1}(z_1t_1,t_1)$. Exceptional divisor: $E:=(t_1=ax_1^2+by_1^2=0)$, it is irreducible over $K$ but geometrically reducible. Singularity: The exceptional divisor itself has only smooth or normal crossing points, thus $B_sX$ has only $cA$ type points. The $(x_1=y_1=0)$ line is singular if $s<m$ and generically smooth for $s=m$. $B_mX$ is terminal. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Divisors with discrepancy 1: Take the $(1,1,1,1)$-blow up. $BX$ is singular along a line with an $A_{2m-2}$ transversal section. We can blow up the line $(m-1)$-times. At each time the exceptional divisor is a pair of transversally intersecting planes, thus we have only $cA$ type singularities. After $(m-1)$ blow ups we obtain $g:Y\to X$ and $Y$ has only isolated $cA$ points, hence terminal. By (\ref{gw.d-ineq.prop.cor}), all the exceptional divisors over $0\in X$ with discrepancy $1$ are birational to divisors on $Y$. They all come in conjugate pairs and have been enumerated by the above $(s,s,1,1)$ blow ups. Conclusion: There is a unique g--extraction whose exceptional divisor has discrepancy 1. It is the $(m,m,1,1)$-blow up $B_mX\to X$. Its exceptional divisor is geometrically reducible, so we need to look further. \medskip Divisors with discrepancy 2: Let $F$ be a geometrically irreducible exceptional divisor over $0\in X$ with discrepancy $2$. Then $\cent_{B_mX}F$ is real. The center can not be the whole $(x_1=y_1=0)$ line or a smooth point on it since both would give $a(F,X)\geq 3$. Thus it is one of the singular points, corresponding to a linear factor of $g_{2m+1}$. By a linear change of the $z,t$-coordinates we may assume that this linear factor is $z$. Thus $\cent_{B_mX}F$ is the origin of the $t$ chart, where $B_mX$ has equation $ax_1^2+by_1^2+t_1^{-2m}g_{\geq 2m+1}(z_1t_1,t_1)$. This is again a $cA_{>1}^+$ type point, ($\mult_0g$ can be even or odd) and $a(F,B_mX)=1$. We have already enumerated all these cases, and we know that $F$ is obtained by an $(r,r,1,1)$-blow up. Putting the two steps together, we see that $F$ is obtained from $X$ by an $(m+r,m+r,2,1)$-blow up. Next we compute these. Normal form: $ax^2+by^2+g_{\geq 2m+1}(z,t)$, where $m\geq 1$, $-ab$ is not a square, $g_{2m+1}\neq 0$ and $\mult_0g(Z^2,T)\geq 2(m+r)$. Weights for blow-up: $(m+r,m+r,2,1)$, giving $B^rX\to X$. $z$-chart: $ax_1^2+by_1^2+z_1^{-2(m+r)}g_{\geq 2m+1}(z_1^2,t_1z_1)/{\textstyle \frac12(m+r,m+r,1,1)}$. Singularity: If $m+r$ is even, then the action has a fixed curve on $B^r\tilde X$, so $B^rX$ is not terminal. If $m+r$ is odd and the origin is in $B^r\tilde X$, then we get an index 2 point. $z_1^{-2(m+r)}g_{\geq 2m+1}(z_1^2,t_1z_1)$ does not vanish at the origin iff $z^{m+r}\in g_{\geq 2m+1}(z,t)$. This implies that $r\geq m+1$. But $g_{2m+1}(z_1^2,t_1z_1)$ itself is not divisible by $z_1^{4m+3}$, hence $r=m+1$. Conclusion: Assume that there is a linear change of the $(z,t)$-coordinates such that $$ g_{\geq 2m+1}(z,t)=\sum_{2i+j\geq 2m+2r} \gamma_{ij}z^it^j,\qtq{and $\gamma_{ij}\neq 0$ for some $2i+j=2m+2r$.} $$ In this coordinate system, the $(m+r,m+r,2,1)$ blow-up gives an elementary extraction whose exceptional divisor is geometrically irreducible and has discrepancy 2. Thus the only possible g--extractions are this weighted blow up and the $(m,m,1,1)$ blow up found earlier. The $(m+r,m+r,2,1)$ blow up is a g--extraction only in the $r=m+1$ case: $$ g_{\geq 2m+1}(z,t)=\sum_{2i+j\geq 4m+2} \gamma_{ij}z^it^j,\qtq{and} \gamma_{2m+1,0}\neq 0. $$ In some cases (cf. (\ref{cA.d=3.ex})), we do not have any geometrically irreducible exceptional divisor over $0\in X$ with discrepancy $2$. Then we have to compute further with discrepancy 3. Fortunately, we can stop there. \medskip Divisors with discrepancy 3: Normal form: $ax^2+by^2+g_{\geq 2m+1}(z,t)$, where $m\geq 1$, $-ab$ is not a square and $g_{2m+1}\neq 0$. Weights for blow-up: $(2m+1,2m+1,2,2)$, giving $Y\to X$. $z$-chart: $ax_1^2+by_1^2+z_1^{-4m-2}g_{\geq 2m+1}(z_1^2,t_1z_1^2)/{\textstyle \frac12(1,1,1,0)}$. Exceptional divisor: $E:=(t_1=ax_1^2+by_1+g_{2m+1}(1,t_1)=0)$, it is geometrically irreducible. Singularity: We get an index 2 point corresponding to the linear factors of $g_{2m+1}(z,t)$. Thus over $\r$ there is always an index 2 point. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2z_1^3\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=3$. \medskip Final conclusion: These singularities always have g--extractions. One is the $(m,m,1,1)$-blow up. Its exceptional divisor is geometrically reducible. This is the only g--extraction with discrepancy 1. In some cases after a suitable coordinate change we can also perform the $(2m+1,2m+1,2,1)$ blow up. This is the only g--extraction whose exceptional divisor is geometrically irreducible and has discrepancy 2. If $g_{2m+1}(z,t)$ has no linear factors over $K$, then the $(2m+1,2m+1,2,2 )$ blow up is a g--extraction whose exceptional divisor is geometrically irreducible and has discrepancy 3. This is the only one such. This case never happens over $\r$. There may be other g--extractions whose exceptional divisor is geometrically reducible and has discrepancy 2. I have no such examples. \end{say} \begin{exmp}\label{cA.d=3.ex} Consider the singularity $X:=(x^2+y^2+z^m+t^n=0)$ for $m,n$ odd and $m+2\leq n\leq 2m-1$. The above computations show that there is no geometrically irreducible exceptional divisor over $0\in X$ with discrepancy $\leq 2$. \end{exmp} \begin{say}[$cA^+_{>1}/2$, $\mult_0g$ odd] {\ } \newline Normal form: $ax^2+by^2+ g_{\geq 2m+1}(z,t)/{\textstyle \frac12(1,1,0,1)}$, where $-ab$ is not a square and $g_{2m+1}\neq 0$. Weights for blow up: For weight reasons, only even powers of $t$ appear in $g$. Thus we can define an integer $r$ by $2m+2r=\mult_0g(Z^2,T)$. $r\leq m+1$ since $g_{2m+1}\neq 0$. We consider the $(m+r,m+r,2,1)$ blow up. $z$-chart: $ax_1^2+by_1^2+z_1^{-2m-2r}g_{\geq 2m+1}(z_1^2,t_1z_1) /{\textstyle \frac12(m+r,m+r,1,1)}$ and then we have to take the quotient by the $\frac12(1,1,0,1)$-action. This lifts to a $\frac12(1,1,0,1)$-action on $B\tilde X$. We get a pair of commuting $\z_2$-actions. Exceptional divisor: $\tilde E:=(z_1=ax_1^2+by_1^2+\sum_{2i+j=2m+2r} \gamma_{ij}t_1^j=0)$ is geometrically irreducible. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2z_1^2\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(\tilde E,X)=2$. Singularities: If $m+r$ is even then the $\z_2\times \z_2$-action is free in codimension one. One of the elements acts by $(0,0,1,1)$, thus we get a singular curve in $BX$. If $m+r$ is odd then one of the elements acts by $(0,0,1,0)$. Coordinates on the quotient are given by $x_1,t_1, z_2=z_1^2,t_1$ and we get the equation $$ ax_1^2+by_1^2+z_2^{-m-r}h_{\geq m+r}(z_2,t_1^2z_2) /{\textstyle \frac12(1,1,0,1)} $$ where $h(Z,T^2)=g(Z,T)$. At the origin we get a $\z_2$-fixed point unless $z^{m+r}\in g$. Thus $r\geq m+1$. On the other hand $r\leq m+1$, thus $r=m+1$. Computing the $t$-chart shows that $BX$ has an index 2 point unless $t^{4m+2}\in g$. Discrepancy: From this we see that $a(E,X)=1/2$ if $m+r$ is odd and $a(E,X)=2$ if $m+r$ is even. Conclusion: If $m+r$ is odd then a g--extraction exists iff $$ g_{\geq 2m+1}(z,t)=\sum_{2i+j\geq 4m+2} \gamma_{ij}z^it^j,\qtq{and} \gamma_{2m+1,0}\neq 0 \neq \gamma_{0,4m+2}. $$ If this holds then the $(2m+1,2m+1,2,1)$ blow-up is the unique g--extractions. It has a geometrically irreducible exceptional divisor with discrepancy $1/2$. If $m+r$ is even then there may exist g--extractions with discrepancy 1/2 or 1. These can be determined by classifying all $\z_2$-invariant divisors of disrepancy 1 and pointwise $\z_2$-fixed divisors of disrepancy 2 or 3 over $\tilde X$. The first task is easy, and we never get any g--extractions this way. The second task is harder and it seems to require separate consideration of about a dozen cases; I have not done all of them. Fortunately, these singularities can be easily excluded in the main theorems using topological considerations. \end{say} \section{$cE$-type Points} In this section we study g--extractions of $cE$ type terminal singularities. The conventions of (\ref{7.notation}), (\ref{8.wbup}) and of (\ref{8.notation}) are used throughout. \begin{say}[$cE_6$ main series] {\ } \newline Normal form: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$. Weights for blow-up: (2,2,1,1) $y$-chart: $x_1^2+y_1^2+h_{4}(z_1,t_1)+y_1\Phi(y_1,z_1,t_1)/{\textstyle \frac12(0,1,1,1)}$. Exceptional divisor: $\tilde E:=(y_1=x_1^2+h_{4}(z_1,t_1)=0)$. $\tilde E$ is geometrically irreducible iff $h_4$ is not a square over $\bar K$. If $-h_4$ is a square over $K$, then $\tilde E$ has 2 geometrically irreducible components. In the other cases $\tilde E$ is irreducible over $K$ but reducible over $\bar K$. Both of the components are fixed by the $\z_2$-action, so the same 3 cases happen for $E$. Singularity: The origin is a fixed point of the $\z_2$-action which is on $B\tilde X$. So we get an index 2 point on $BX$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2y_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. Conclusion: If $-h_4$ is a square over $K$ then there are 2 geometrically irreducible divisors with discrepancy 1, so no g--extractions. If $h_4$ is not a square over $\bar K$ then we get an index 2 point, so again there are no g--extractions. \end{say} \begin{say}[$cE/2$] {\ } \newline Normal form: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)/{\textstyle \frac12(1,0,1,1)}$. By weight considerations $g_3=0$ and $h_5=0$. $h_4\neq 0$ since otherwise we would not have a terminal point. This is a $cE_6/2$ point. Weights for blow-up: (2,2,1,1) $y$-chart: $x_1^2+y_1^2+h_{4}(z_1,t_1)+y_1\Phi(y_1,z_1,t_1)/{\textstyle \frac12(0,1,1,1)}$. Lifting of the $\z_2$-action. The $\z_2$-action lifts to $\frac12(1,1,0,0)$. Thus on $B\tilde X$ we have two commuting $\z_2$-actions. Exceptional divisor: $\tilde E:=(y_1=x_1^2+h_{4}(z_1,t_1)=0)$. It is geometrically irreducible iff $h_4$ is not a square over $\bar K$. If $h_4=-Q_2(z_1,t_1)^2$ then $\tilde E$ has 2 geometrically irreducible components $(y_1=x_1\pm Q_2(z_1,t_1)=0)$. The $\frac12(1,1,0,0)$ action interchanges the 2 components, thus $E\subset BX$ is geometrically irreducible. Singularity: The $\frac12(1,1,0,0)$ action has a fixed curve, thus we get a nonterminal singular curve on $BX$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2y_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. \end{say} \begin{say}[$cE_7$ main series]\label{cE7.main.ser} {\ } \newline Normal form: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)$. Weights for blow-up: (3,2,1,1) $x$-chart: $x_1+y_1^3x_1+y_1g_{3}(z_1,t_1)+h_{5}(z_1,t_1)+x_1\Phi(y_1,z_1,t_1)/{\textstyle \frac13(1,1,2,2)}$. Exceptional divisor: $\tilde E:=(x_1=y_1g_{3}(z_1,t_1)+h_{5}(z_1,t_1)=0)$. It is geometrically irreducible iff $g_3$ and $h_5$ have no common factors. Singularity: The origin is a fixed point of the $\z_3$-action which is on $B\tilde X$. So we get an index terminal 3 point on $BX$. In fact, it is the index 3 terminal point ${\Bbb A}^3/\frac13(1,1,2)$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2x_1\cdot dy_1\wedge dz_1\wedge dt_1$, so $a(E,X)=1$. Conclusion: If $g_3$ and $h_5$ have no common factors then $E$ is irreducible and there are no g--extractions. \end{say} \begin{say}[$cE$ with common linear factors]\label{cE.comm.l.f} {\ } \newline Normal form: $$ x^2+y^3+yzG_2(z,t)+z^2Q_2(z,t)+zH_4(z,t)+yg_{\geq 4}(z,t)+h_{\geq 6}(z,t). $$ The following cases are of this form: $cE_8$: $x^2+y^3+yg_{\geq 4}(z,t)+h_{\geq 5}(z,t)$, if $h_5$ has a linear factor over $K$, which we can call $z$. $cE_7$: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)$, if $g_3$ and $h_5$ have a common linear factor over $K$, which we can call $z$. $cE_6$: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$ if there is a linear factor over $K$, which we can call $z$, such that $z^2|h_4,\ z|g_3$ and $z|h_5$. Weights for blow-up: (3,2,2,1) $z$-chart: $$ \begin{array}{rl} x_1^2+y_1^3&+y_1G_2(0,t_1)+Q_2(0,t_1)+H_4(0,t_1)\\ &+y_1g_4(0,t_1)+h_6(0,t_1)+z_1\Phi(y_1,z_1,t_1)/{\textstyle \frac12(1,0,1,1)}. \end{array} $$ Exceptional divisor $\tilde E$ is geometrically irreducible: $$ (z_1= x_1^2+y_1^3+y_1G_2(0,t_1)+Q_2(0,t_1)+H_4(0,t_1)+y_1g_4(0,t_1)+h_6(0,t_1)=0). $$ Singularity: The origin is a fixed point of the $\z_2$-action which is on $B\tilde X$. So we get an index 2 point on $BX$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 2z_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1$. \end{say} \begin{say}[$cE_6$ with $h_4$ a square] {\ } \newline Normal form: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 4}(z,t)$. Weights for blow-up: (1,1,1,1) $t$-chart: $$ x_1^2+y_1^3t_1+y_1t_1^2g_3(z_1,1)+y_1t_1^3g_4(z_1,1)+ t_1^2h_{4}(z_1,1)+t_1^3h_5(z_1,1)+t_1^4\Phi(y_1,z_1,t_1). $$ The $z$-chart is similar. Exceptional divisor: $E:=(t_1=x_1=0)$, and the scheme theoretic exceptional divisor is $2E$. Singularity: On the $t$-chart the singular set is the line $L:=(x_1=y_1=t_1=0)$. We determine the singularities along this line. For a fixed value $z_1=b\in \bar K$ we get a $cA$-point if $h_4(b,1)\neq 0$. If $h_4(z,1)$ has a simple root at $b$ then we get a $cD$-point. If $h_4(z,1)$ has a multiple root at $b$ then we still get a $cD$ point if $g_3(b,1)\neq 0$ and a $cE$-point if $h_5(b,1)\neq 0$. Hence, if $b\in K$ and we do not have a cDV point, then $h_4$ has a multiple linear factor which also divides $g_3$ and $h_5$. This case was settled in (\ref{cE.comm.l.f}). Assuming that this is not the case, we obtain that $BX$ has $cDV$ points along $L$. The $z$-chart is similar and easy computations show that the $x$ and $y$-charts are smooth along $E$. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= t_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=2$. First conclusion: $BX$ is not a g--extraction since it has a singular curve. $E$ is geometrically irreducible and $a(E,X)=2$, thus if $g:Z\to X$ is a g--extraction with exceptional divisor $F$ then $a(F,X)=1$ by (\ref{gw.discr2.cor}). Computations: Here we determine all divisors $F$ over $0\in X$ with $a(F,X)=1$. If $\cent_{BX} F$ is not on $L$ then $a(F,X)\geq 3$, and if $\cent_{BX} F$ is a point on $L$ then $a(F,X)\geq 2$. Thus if $a(F,X)=1$ then $\cent_{BX}F=L$ and $a(F,BX)=0$. Along $L$ the threefold $BX$ has transversal type $A_5$ whose singularity is resolved by blowing up the line 3-times. By explicit computation we see that only the first of these produces an exceptional divisor $F$ with $a(F,X)=1$. This is the same divisor that we encountered in the $(2,2,1,1)$-blow up and so it was already accounted for. Final conclusion: There is no g--extraction except possibly when there is a $b\in \bar K\setminus K$ such that $(z-bt)^2|h_4, (z-bt)|g_3,(z-bt)|h_5$. In these cases the same divisibilities hold if we replace $b$ by its conjugates over $K$. Thus $b$ is quadratic over $K$, a root of $Q_2(z,1)$. If $F$ is any divisor over $0\in X$ with $a(F,X)=1$ then its center in $BX$ is $(z_1-b=0)\in L$ or its conjugate. Thus $F$ is geometrically reducible. \end{say} \begin{say}[$cE_6$ last case]\label{cE6.g--extr.exist} {\ } \newline Normal form: $$ x^2+y^3+cQ_2(z,t)^2+yL_1(z,t)Q_2(z,t)+C_3(z,t)Q_2(z,t)+ yg_{\geq 4}(z,t)+h_{\geq 6}(z,t), $$ where $Q_2$ is a quadratic form which is irreducible over $K$ and $-c$ is not a square in $K$. By a coordinate change as in \cite[I.12.6]{AGV85} we can bring this to the simpler form $$ x^2+y^3+cQ_2(z,t)^2+ yg_{\geq 4}(z,t)+h_{\geq 6}(z,t), $$ though this is not important. Normal form and topology over $\r$: We can choose $Q_2$ to be positive definite and diagonalize it. $-c\in \r$ is not a square, so we can choose $c=1$. Thus we get the normal form $$ x^2+y^3+(z^2+t^2)^2+ yg_{\geq 4}(z,t)+h_{\geq 6}(z,t). $$ By (\cite[4.9]{rat1}) we obtain that $X(\r)$ is homeomorphic to $\r^3$. g--extractions: As we discussed above, all the g--extractions of $X$ have geometrically reducible exceptional divisors. Construction of g--extractions: It turns out that in these cases there is a g--extraction. By above remark we do not need to know this for certain to understand the topology over $\r$, thus I only outline the construction. Basic constructions of toric geometry are used without reference; see \cite{Fulton93} for an introduction. Over $\bar K$ we can bring the equation to the form $$ x^2+y^3+z^2t^2+ yg_{\geq 4}(z,t)+h_{\geq 6}(z,t). $$ Let $e_x,e_y,e_z,e_t$ be a basis of $\r^4$. Consider the vectors $w_z=\frac18(3,2,2,1)$ and $w_t=\frac18(3,2,1,2)$. These vectors give a triangulation of the simplex with vertices $e_x,e_y,e_z,e_t$ where the edges are $$ (e_x,w_z), (e_x,w_t), (e_y,w_z), (e_y,w_t), (e_z,w_z), (e_t,w_t). $$ Let us take the corresponding toric blow up. One can check by a rountine but tedious computation that all singularities of $BX$ are terminal and we get two index 3 points on the chart corresponding to the simplex $(e_x,e_y,w_z,w_t)$. Note that the above construction is symmetric in $z$ and $t$. Thus if we start with a quadratic form $Q_2=z^2+qt^2$ and introduce new coordinates $z'=z+\sqrt{q}t$ and $t'=z-\sqrt{q}t$ then $Q_2=z't'$ and any blow up which is symmetric in $z',t'$ can be transformed back to a blow up of $X$ defined over $K$. We need to check that the two index 3 points become conjugates over $K$, but this is easy to see from the explicit equations. \end{say} \begin{say}[$cE_7$ with common nonlinear factor] {\ } \newline Normal form: $x^2+y^3+yg_{\geq 3}(z,t)+h_{\geq 5}(z,t)$, where we assume that the greatest common divisor of $g_3$ and $h_5$ is $K$-irreducible (and nonconstant). We write $g_3=Q(z ,t)G(z,t)$ and $h_5=Q(z,t)H(z,t)$. ($Q$ is allowed to be linear, though this case is treated already.) Weights for blow-up: (3,2,1,1) $x$-chart: $x_1+y_1^3x_1+y_1g_{3}(z_1,t_1)+h_{5}(z_1,t_1)+x_1\Phi(y_1,z_1,t_1)/{\textstyle \frac13(1,1,2,2)}$. Exceptional divisor: It has two irreducible components: \begin{eqnarray*} \tilde E&:=&(x_1=y_1G(z_1,t_1)+H(z_1,t_1)=0),\qtq{and}\\ \tilde F&:=&(x_1=Q(z_1,t_1)=0). \end{eqnarray*} $\tilde E$ is geometrically irreducible, $\tilde F$ is irreducible but geometrically reducible if $Q$ is not linear. Discrepancy: $\pi^*\frac{dy\wedge dz\wedge dt}{x}= 3x_1\frac{dy_1\wedge dz_1\wedge dt_1}{x_1}$, so $a(E,X)=1=a(F,X)$. (The latter equality uses that $Q$ is not a multiple factor.) Further aim: We would like to construct a birational morphism $g:Z\to X$ whose exceptional divisor corresponds to $E$, and determine the singularities of $Z$. Thus in $BX$ we have to contract $F$. $F$ is not $\q$-Cartier in $BX$ and $F$ can not be contracted in $BX$. First we have to correct this problem. Singularities of $BX$: I claim that $BX$ has only canonical singularities. This can be done 2 ways. One can compute each chart explicitly, which is rather tedious. I found it easier to use a degeneration argument as follows. Let $F$ be the normal form of the equation as above. We may assume that $g_3(1,0)=1$. Consider the substitution $$ F(x,y,z,t)\mapsto \epsilon^{-24}F(\epsilon^{12}x,\epsilon^{8}y,\epsilon^{6}z, \epsilon^{7}t). $$ The exponents are chosen so that for $\epsilon\to 0$ the limit is $X_0:=(x^2+y^3+yz^3=0)$. The $(3,2,1,1)$-blow up $BX_0$ is easy to compute. We find an index 3 terminal point ${\Bbb A}^3/\frac13(1,1,2)$, a curve of $cA$-points and a curve of $cE_7$-points corresponding to the $t$-axis. Thus $BX$, as a small deformation of $BX_0$, has an index 3 point at the origin and some $cDV$ singularities. (These turn out to be isolated points but we do not need this.) As in (\ref{cE7.main.ser}) we see that the index 3 point is at the origin of the $x$-chart and it is ${\Bbb A}^3/\frac13(1,1,2)$. In particular it is $\q$-factorial. Small blow up: Let $p:Y\to BX$ be the blow up of $F$ in $BX$. Let $F'\subset Y$ denote the birational transform of $F$. Away from the index 3 point $BX$ is locally isomorphic to $B\tilde X$. $\tilde F$ is defined by 2 equations $(x_1=Q(z_1,t_1)=0)$, thus $p:Y\to BX$ is small and is an isomorphism at all points where $F$ is $\q$-Cartier. The index 3 point is $\q$-factorial, so $F$ is $\q$-Cartier there. Thus $p:F'\to F$ is an isomorphism. Contracting $F'$: $F$ is a cone over a $K$-irreducible curve, hence its cone of curves over $K$ is 1-dimensional. If $C\subset F'$ is a general curve then $(C\cdot K_Y)=(p(C)\cdot K_{BX})<0$ and $(C\cdot F')=(p(C)\cdot F)<0$. Thus the curves in $F'$ generate a $K_Y$-negative extremal ray of $Y/X$, which can be contracted. We obtain $f: Y\to Z$ and $g:Z\to X$. $P:=f(F')$ is a $K$-point since $F'$ is connected. Conclusion: $g:Z\to X$ has a geometrically irreducible exceptional divisor corresponding to $E$ and it has discrepancy 1. Furthermore, by (\ref{gw.discr2.cor}) the index of $P$ can not be one since $a(F,X)=1$. Hence there are no g--extractions. \end{say} \begin{say}[Conclusion] The $cE_8$ case is settled if $h_5(z,t)$ has a linear factor over $K$. This always holds if $K=\r$, hence at least in this case there are no g--extractions. I do not know what happens if $K\neq \r$. The $cE_7$ case is settled if $g_3(z,t)$ and $h_5(z,t)$ have no common factor, or if they have a common linear factor over $K$ or if they have a unique common factor over $K$. This accounts for all the possibilities, hence there are no g--extractions. The $cE_6$ case is settled if $h_4(z,t)$ is not a square over $\bar K$, if $-h_4(z,t)$ is a square over $K$ or if $h_4(z,t)$ is divisible by the square of a linear form over $K$. In these cases there are no g--extractions. The remaining case is treated in (\ref{cE6.g--extr.exist}) and the unique g--extraction is written down explicitly. For the applications in this paper the existence is not crucial. \end{say} \begin{exmp} Let $X$ be the $cE_7$ type singularity $x^2+y^3+yg_3(z,t)+h_5(z,t)$, where $g_3$ and $h_5$ do not have a common factor. It is not hard to see that $X$ is an isolated singular point and its $(3,2,1,1)$-blow up has only terminal singularities. As in (\ref{cE7.main.ser}), the $y$ chart on the blow up gives the exceptional divisor $$ E=(g_3(z,t)+h_5(z,t)=0)/{\textstyle \frac12(1,1,1,1)}. $$ This gives examples of extremal contractions whose exceptional divisor $E$ has a quite complicated singularity along the $(z=t=0)$-line. \begin{enumerate} \item $x^2+y^3+yz^3+t^5$. $E$ is singular along $(z=t=0)$, with a transversal singularity type $z^3+t^5$, that is $E_8$. \item $x^2+y^3+y(z-at)(z-bt)(z-ct)+t^5$. $E$ has triple selfintersection along $z=t=0$. \end{enumerate} \end{exmp} \section{Hyperbolic 3--manifolds} The aim of this section is to show that every hyperbolic 3--manifold satisfies the conditions (\ref{int.no.cond}). \begin{thm}\label{hyp.doesnotcont.thm} Let $M$ be a compact hyperbolic 3--manifold. Then $M$ does not contain any PL submanifold of the following types: \begin{enumerate} \item $\r\p^2$ \item 1--sided $S^1\times S^1$ \item 1--sided Klein bottle. \end{enumerate} \end{thm} We use two facts about hyperbolic 3--manifolds. First, that their universal cover is homeomorphic to $\r^3$. Second, that their fundamental group does not contain a subgroup isomorphic to $\z^2$ (see, for instance, \cite[4.6]{Scott83}). \medskip More generally, we see how these conditions fit in the framework of Thurston's geometrization conjecture. This version was pointed out to me by Kapovich. \begin{thm}\label{gen.doesnotcont.thm} Let $M$ be a compact 3--manifold. Assume that $M=M_1\ \#\ \cdots \ \#\ M_k$, where \begin{enumerate} \item[(i)] each $M_i$ is aspherical, and \item[(ii)] the Seifert fibered part of the Jaco--Shalen--Johannson decomposition of $M_i$ is orientable. \end{enumerate} Then $M$ does not contain any PL submanifold of the following types: \begin{enumerate} \item $\r\p^2$ \item 1--sided $S^1\times S^1$ \item 1--sided Klein bottle. \end{enumerate} \end{thm} We consider the 3 types of submanifolds separately. Condition (\ref{int.no.cond}.1) is closely related to the notion of $\p^2$-irreducibility (cf.\ \cite[p.88]{Hempel76}). \begin{lem} Let $M$ be a 3--manifold with universal cover $\tilde M$. \begin{enumerate} \item If $M\sim M_1\ \#\ M_2$, then $M$ contains a 2--sided $\r\p^2$ iff one of the summands does. \item Assume that $\tilde M$ is homeomorphic to $\r^3$. Then $M$ does not contain an $\r\p^2$ and $M$ can not be written as a nontrivial connected sum. \end{enumerate} \end{lem} Proof. Assume that $F\subset M$ is a 2--sided $\r\p^2$. We may assume that $F$ is transversal to the gluing $S^2$. Thus $C=F\cap S^2$ is an embedded curve in $F$. Assume first that $F$ has a connected component $C_1\subset C$ which is not null homotopic in $F$. Then $F$ is not orientable along $C_1$, and the same holds for $M$ along $C_1$. But $M$ is orientable along $S^2$, a contradiction. Take any connected component $C_i\subset C$ such that $C_i\subset S^2$ bounds a disc $D_i$ which is disjoint from $C$. $C_i$ also bounds a disc $D'_i$ in $F$ (since it is null homotopic in $F$). Thus we can change the embedding $\r\p^2\to M$ by replacing $D'_i$ with $D_i$ and then pushing it to one side. The new embedding is still 2--sided. Repeating if necessary, we eventally get an embedding which is disjoint from $S^2$, proving (1). $\r\p^2$ can not be embedded into $\r^3$ (cf. \cite[27.11]{GrHa81}), thus the preimage of $\r\p^2$ in $\r^3$ is a union of copies of $S^2$. Fix one of these and call it $N$. By the Schoenflies theorem (cf.\ \cite[Sec. 17]{Moise77}) $N$ bounds a 3--ball $B^3$. At least one element of $\pi_1(M)$ maps $N$ to itself. It can not map the inside of $N$ to its outside since these are not homeomorphic. If it maps $B^3$ to itself, then by the Borsuk--Ulam theorem (cf.\ \cite[23.20]{Fulton95}) we have a covering transformation with a fixed point, a contradiction. Assume that we have $S^2\sim N'\subset M$ and let $S^2\sim N\subset \tilde M$ be one of the preimages. Then $N$ bounds a 3--ball and so does $N'$. \qed \medskip In order to study the conditions (\ref{int.no.cond}.2--3) we have to distinguish two cases. \begin{say}[Incompressible case]\label{12.incompr.say} Let $M$ be a compact 3--manifold and $S\subset M$ a compact 1--sided torus or Klein bottle. Assume that $\pi_1(S)\DOTSB\lhook\joinrel\rightarrow \pi_1(M)$. Let $\partial U$ be the boundary of a regular neighborhood of $S$. Then $\partial U$ is a 2--sided torus or Klein bottle and $\pi_1(\partial U)\DOTSB\lhook\joinrel\rightarrow \pi_1(S)\DOTSB\lhook\joinrel\rightarrow \pi_1(M)$ is an injection. This implies that $\partial U$ is incompressible in $M$ (cf.\ \cite[pp.88-89]{Hempel76}). Thus $U$ is one of the pieces of the Jaco--Shalen--Johannson decomposition of $M$ (cf.\ \cite[p.483]{Scott83}). We have to be a little more careful since $U$ is Seifert fibered, thus it may sit inside one of the Seifert fibered components. The fundamental group of a hyperbolic 3--manifold does not contain a subgroup isomorphic to $\z^2$ (see, for instance, \cite[4.6]{Scott83}), hence the incompressible case does not happen for hyperbolic 3--manifolds. \end{say} \begin{say}[Compressible case]\label{12.compr.say} In this case we show that $M$ can be written as a connected sum with a very special summand. \begin{prop}\label{1-s.torus} Let $M$ be a compact 3--manifold. Then $M$ contains a 1--sided torus $T$ such that $\pi_1(T)\to \pi_1(M)$ is not an injection iff $M\sim N\ \#\ (S^1\tilde{\times}S^2)$ or $M\sim N\ \#\ (S^1 \times\r\p^2)$ \end{prop} Proof. Let $T\subset U\subset M$ be a regular neighborhood. Set $V=M\setminus U$. Then $\partial U=\partial V\sim S^1\times S^1$. We know that $\pi_1(\partial U)$ injects into $\pi_1(U)$. If $\pi_1(\partial U)\DOTSB\lhook\joinrel\rightarrow \pi_1(V)$, then $\pi_1(\partial U)\DOTSB\lhook\joinrel\rightarrow \pi_1(U)\DOTSB\lhook\joinrel\rightarrow \pi_1(M)$ by Schreier's theorem (cf. \cite[IV.2.6]{Lyndon-Schupp77}). $\pi_1(\partial U)$ is an index 2 subgroup of $\pi_1(T)$ an $\pi_1(T)$ is torsion free. Thus $\pi_1(T)\to \pi_1(M)$ is also an injection, a contradiction. Therefore, by the Loop theorem (cf.\ \cite[4.2]{Hempel76}), there is an embedding of the disc $j:(B,\partial B)\DOTSB\lhook\joinrel\rightarrow (V,\partial V)$ such that the image of $j(\partial B)$ is not contractible in $\partial V$. Let us cut $V$ along $j(B)$ to get $W$. The boundary of $W$ is $\partial V$ cut along $j(\partial B)$ (which is a cylinder) with two copies of $B$ pasted to the ends. That is, $\partial W\sim S^2$. Therefore $M$ is obtained by pasting $W$ to a 3--manifold (with boundary) $K$, which is obtained from $U$ by attaching a 2--handle. There are two cases corresponding to whether $j(\partial B)$ is a primitive element of $\pi_1(U)\cong \z^2$ (hence $\pi_1(K)\cong \z$) or is contained in $2\z^2$ (hence $\pi_1(K)\cong \z+\z_2$).\qed \begin{prop}\label{1-s.kb} Let $M$ be a compact 3--manifold which does not contain a 2--sided $\r\p^2$. Then $M$ contains a 1--sided Klein bottle $K$ such that $\pi_1(K)\to \pi_1(M)$ is not an injection iff $M\sim N\ \#\ (S^1\tilde{\times}S^2)$ or $ M\sim N\ \#\ (\r\p^3\ \#\ \r\p^3)$. \end{prop} Proof. Let $K\subset U\subset M$ be a regular neighborhood. As in the proof of (\ref{1-s.torus}) we obtain an embedding of the disc $j:(B,\partial B)\DOTSB\lhook\joinrel\rightarrow (V,\partial V)$ such that the image of $j(\partial B)$ is not contractible in $\partial V$. We again cut $V$ along $j(B)$ to get $W$. Let $\partial V^*$ denote $\partial V$ cut along $j(\partial B)$. There are 3 cases to consider corresponding to what $\partial V^*$ is: \begin{enumerate} \item ($\partial V^*$ is a cylinder). Then we obtain a connected sum decomposition as in (\ref{1-s.torus}). \item ($\partial V^*$ consists of two Moebius bands). Then $\partial W$ is two disjoint projective planes, hence $M$ contains a 2--sided projective plane. This can not happen by assumption. \item ($\partial V^*$ is a Moebius band). In this case $j(\partial B)$ is 1--sided in $\partial V$, thus $M$ is not orientable along $j(\partial B)$. Then $j(\partial B)$ can not be the boundary of an embedded disc.\qed \end{enumerate} \end{say} \begin{rem} So far we have excluded Seifert fiber spaces from considerations. Many Seifert fiber spaces do contain 1--sided tori or Klein bottles. If $p:M\to F$ is a Seifert fiber space and $C\subset F$ a 1--sided curve not passing through any critical value, then $p^{-1}(C)\subset M$ is a 1--sided torus or Klein bottle. Another example can be obtained as follows. Let $x,x'\in F$ be two points such that the fibers over them have multiplicity 2. Let $I\subset F$ be a simple path connecting $x$ and $x'$. Then $p^{-1}(I)$ is a 1--sided Klein bottle. It is not hard to see that if $T\subset M$ is a 1--sided torus or Klein bottle such that $p(T)$ is 1--dimensional (these are called vertical) then $T$ is obtained by one of the above constructions. Assume now in addition that $M$ has a geometry modelled on ${\mathbb H}^2\times \r$ (cf.\ \cite[p.459]{Scott83}). Then by \cite[5.6]{Johannson79}, every 1--sided torus or Klein bottle in $M$ is isotopic to a vertical one. This way we obtain many examples of nonorientable Seifert fiber spaces which satisfy the conditions (\ref{int.no.cond}). \end{rem}
1998-01-02T06:55:25
9712
alg-geom/9712008
en
https://arxiv.org/abs/alg-geom/9712008
[ "alg-geom", "math.AG" ]
alg-geom/9712008
Bumsig
Bumsig Kim (University of California - Davis)
Quantum Hyperplane Section Theorem For Homogeneous Spaces
24 pages, LaTeX 2e. The presentation is improved a lot
null
null
null
null
We formulated a mirror-free approach to the mirror conjecture, namely, quantum hyperplane section conjecture, and proved it in the case of nonnegative complete intersections in homogeneous manifolds. For the proof we followed the scheme of Givental's proof of a mirror theorem for toric complete intersections.
[ { "version": "v1", "created": "Fri, 5 Dec 1997 21:07:48 GMT" }, { "version": "v2", "created": "Tue, 16 Dec 1997 23:46:55 GMT" }, { "version": "v3", "created": "Fri, 2 Jan 1998 05:55:25 GMT" } ]
2007-05-23T00:00:00
[ [ "Kim", "Bumsig", "", "University of California - Davis" ] ]
alg-geom
\section{Introduction} Quantum cohomology of a symplectic manifold is a certain deformed ring of the ordinary cohomology ring with parameter space given by the second cohomology group. It encodes enumerative geometry of rational curves on the manifold. In general it is difficult to compute the quantum cohomology structure. On the other hand, mirror symmetry predicts an answer to traditional questions of counting the virtual numbers of rational curves of a given degree on a three-dimensional Calabi-Yau manifold, which amounts to knowing the structure of the quantum cohomology. In the large class of Calabi-Yau manifolds, the complete intersections in toric manifolds or homogeneous spaces, this mirror symmetry prediction \cite{Ca, BV, BCKV1, BCKV2} can be interpreted as a quantum cohomology counterpart of the weak Lefschetz hyperplane section theorem relating cohomology algebras of the ambient manifolds and their hyperplane sections. As it is mentioned in \cite{GS}, \lq\lq quantum hyperplane section conjecture" can be formulated in intrinsic terms of Gromov-Witten theory on the ambient manifold and does not require a reference to its mirror partner. In this paper we formulate and prove the conjecture for homogeneous spaces. It would be one of the highly nontrivial functorial properties enjoyed by quantum cohomology algebras. One can compute the virtual numbers of rational curves on a Calabi-Yau 3-fold complete intersection, provided one knows the quantum cohomology algebra of the ambient space. In fact, one needs to know the quantum differential equations of the space, which are certain linear differential equations arising from the flat connection in the quantum cohomology algebra. The mirror symmetry prediction is that the quantum differential equations of a Calabi-Yau manifold are equivalent (in a sense) to the Picard-Fuchs differential equations of another Calabi-Yau manifold. In contrast, the proposed conjecture is that there is a certain relation between quantum differential equations of a manifold and those of a certain complete intersection. When the ambient space is a symplectic toric manifold, the conjecture is a corollary of the Givental mirror theorem \cite{GT}. \bigskip Let $X$ be a compact homogeneous space of a semi-simple complex Lie group and let $V$ be a vector bundle over $X$. Suppose $V'_\beta :=\pi _*e_1^*V$ becomes a vector orbi-bundle over Kontsevich moduli space $\overline{M}_{0,0}(X,\beta)$ where $e_1$ is the evaluation map at the (first) marked point from $\overline{M}_{0,1}(X,\beta )$ to $X$ and $\pi$ is the map from $\overline{M}_{0,1}(X,\beta)$ to $\overline{M}_{0,0}(X,\beta)$ associated with \lq\lq forgetting the marked point" \cite{Ko}. Then one might want compute \[ \int _{\overline{M}_{0,0}(X,\beta)}Euler(V'_{\beta}). \] Introduce a formal parameter $\hb$. Then it turns out that the classes \[ G_{\beta}^V :=(e_1)_*\frac{Euler(V_{\beta})}{\hb (\hb -c)}\] would be better considered \cite{GE}, where $V_\beta =\pi ^*(V'_\beta )$ and $c$ (depending on $\beta$) are the first Chern classes of the universal cotangent line bundles. The classes are in $H^*(X) [\hb ^{-1}]$. They recover the original integrals which we want: \[ \int _X G^V_\beta = \frac{-2}{\hb ^3}\int _{\overline{M}_{0,0}(X,\beta)}Euler(V'_{\beta}) +o(\hb ^{-3}).\] Consider the classes \[ G_{\beta}^X:=(e_1)_* \frac{1}{\hb(\hb -c)}\] corresponding to $X$ itself (without $V$). When $V$ is a convex, decomposable, vector bundle $\oplus L_j$ of line bundles $L_j$, the main result of this paper proves some explicit relationship between $A:=\{ G_{\beta}^V |\ \beta\in H_2(X,\ZZ ) \} $ and $B:=\{ H^V_{\beta}\cup G_{\beta}^X | \ \beta\in H_2(X,\ZZ ) \}$, where \[ H^V_\beta = \prod _j\prod _{m=0}^{<c_1(L_j) ,\beta >} (c_1(L_j) + m\hb ) \] which is the key object introduced in this sequel. \bigskip We now formulate the precise result of this paper. Let $\{ p_i \}_{i=1}^k$ denote the $\ZZ _+$ basis of the closed integral K\"ahler (ample) cone of $X$. Let us introduce formal parameters $q_i$, $i=1,...,k$, and the ring $\QQ [[ q_1,...,q_k]]$ of formal power series of $q_i$. Denote by $q^\beta$ \[ \prod _{i=1}^k q_i^{<p_i,\beta >}.\] For simplicity, let $G^X_0 =1$ and $G^V_0=Euler(V)$. We want to compare generating functions $J^V$ and $I^V$ from $A$ and $B$, respectively: \begin{eqnarray*} S^V &:=& \sum _{\beta} q^\beta G^V_\beta \\ \Phi ^V &:=& \sum _{\beta }q^\beta H_\beta ^V \cup G^X_\beta .\end{eqnarray*} We prove that one can be transformed to another by a unique \lq\lq mirror" transformation. To describe the transformation, let \[ q_i = e^ {t_i}, \text{ for } i=1,...,k, \] and introduce another formal variable $t_0$. Define degree of $q_i$ by \[ c_1(TX ) -c_1(V)=\sum (\deg q_i) p_i .\] Let \begin{eqnarray*} J^V (t_0,...,t_k) &:=& e^{(t_0+\sum _i p_it_i)/\hb }S^V \end{eqnarray*} and \begin{eqnarray*} I^V (t_0,...,t_k)&:=& e^{(t_0+\sum _i p_it_i)/\hb }\Phi ^V ,\end{eqnarray*} which are formal power series of $t_1,...,t_k, e^{t_0},...,e^{t_k}$ over $H^*(X)[\hb ^{-1}]$. {\theorem\label{thmmain} Assume that $\deg q_i\ge 0$ for all $i$. Then $J^V$ and $I^V$ coincide up to a unique weighted homogeneous change of variables: $t_0\mapsto t_0+f_0\hb +f_{-1}$ and $t_i\mapsto t_i +f_i$, where $f_{-1},...,f_k$ are power series of $q_1,...,q_k$ over $\QQ$ without constant terms, $\deg f_i=0$, $i=0,...,k,$ and $\deg f_{-1}=1$.} \bigskip {\it Remarks:} 0. $J^V$ will be shown to be the cohomological expression of solutions to quantum differential equations associated to $(X,V)$, which is closely related to the quantum differential equations of the smooth zero locus of $V$. The theorem can be extended to the case of decomposable concavex vector bundles $V$. 1. The change of variables is uniquely determined by coefficients of $1=(\frac{1}{\hb})^0$ and $\frac{1}{\hb}$ in the expansions of $J^V$ and $I^V$ as power series of $\frac 1\hb$. 2. In the case of a symplectic toric manifold $X$ the similar statement is a corollary of a mirror theorem in \cite{GT}, where $\Phi ^X$ is explicitly known. 3. For the proof of \ref{thmmain} we follow the scheme of Givental's proof \cite{GE, GT} of the mirror theorem for nonnegative complete intersections in toric manifolds. 4. The theorem verifies the prediction \cite{BCKV1} of virtual numbers in Calabi-Yau 3-fold complete intersections in Grassmannians. 5. A mirror construction is established for complete intersections in partial flag manifolds \cite{BCKV1, BCKV2}. Because of the known quantum cohomology structure \cite{Ionut}, in principle there is no essential difficulty in finding $G^X_\beta$ for each partial flag manifold $X$, even though a general formula of it is unknown. 6. The quantum hyperplane section principle is applied to a nonconvex manifold in \cite{To}. \smallskip {\it Notation:} $X$ will always be a generalized flag manifold $G/P$, where $G$ is a complex semi-simple Lie group and $P$ is a parabolic subgroup. Let $T$ be a maximal torus of $G$ in $P$ and let $T$ act on $X$ on the left. Let a complex torus $T'$ act on $X$ trivially and let $V$ be a $T\ti T'$-equivariant convex vector bundle over $X$. Consider $E$ a multiplicative class and suppose that $E(V)\in H_{T\ti T'}(X)$ is invertible in $H_{(T\ti T')}(X):= H_{T\ti T'}(X)\ot H_{T\ti T'}$, where $H_{(T\ti T')}$ is the quotient field of $H_{T\ti T'}(pt)$. In section 2, we will not consider $T$-action on $X$. In section 6, additionally we will assume that $V$ is decomposable. The convexity of $V$ is by definition that $H^1(\PP ^1, f^*V)=0$ for any morphism $f:\PP ^1\ra X$. Let $T\ti T'$ equivariant line bundles $U_i$, $i=1,...,k$, form ample basis of ordinary Picard group. We denote $\int _X ABE(V)$ by $<A,B>^V_{0}$, for $A, B\in H^*_{(T\ti T')}(X)$ and also we use $\int _V A:=\int AE(V)$ (equivariant push forwards). The Mori cone of $X$ will be denote by $\Lambda$, which can be identified with $\ZZ _+^k$ with respect to coordinates $p_i:=c_1(U_i)$. On the additive group $\ZZ ^{k}$ we will give the standard partial ordering, so that $d:=(d_1,...,d_k)\ge 0$ means $d_i\ge 0$. Let $\phi _v$ denote the equivariant pushforward of $1$ under the embedding $i_v$ of the fixed point $v$ to $(X,V,E)$; this $(X,V,E)$ has the Frobenius structure by pairing $<,>^V_0$, so that $A_v:=<A,\phi _v>_0^V=i_v^*(A)$ for $A\in H^*_{(T\ti T')}(X)$. For a $G$-manifold $M$, let $M^G$ denote the set of $G$-fixed points of $M$. We will say simply degree and dimension for complex degree and complex dimension, respectively. Let $\sum _a T_a\ot T^a$ be the equivariant diagonal class of $(X,V,E)$ in $X\ti X$. That is, $<T_a, T^b>^V_0=\delta _{a,b}$. In the paper we will consider various rings $H^*_{T\ti T'}[[\hb ^{-1}]][[q]]$ formal power series ring of $\hb ^{-1}, q$ over $H^*_{T\ti T'}$, $H^*_{(T\ti T')}[[\hb ^{-1}]][[q]]$ formal power series ring of $\hb ^{-1}, q$ over $H^*_{(T\ti T')}$, and $H^*_{(T\ti T')}(\hb )[[q]]$ formal power series ring of $q$ over quotient field of $H^*_{T\ti T'}[\hb ]$. \smallskip {\it Structure of the paper:} In section 2, we recall a general theory of Gromov-Witten invariants and quantum cohomology. We introduce the Givental Correlators $S^V$. In section 3, we show that the equivariant correlators satisfy certain \lq\lq almost recursion relations." In section 4, we introduce the double construction and show that the correlators satisfy certain polynomiality in the double construction. In section 5, we introduce certain class $\mathcal{P}(X,V,E)$ of series of $q=(q_1,...,q_k), \hb ^{-1}$ over $H^*_{T\ti T'}(X)$, where a \lq mirror' group acts freely and transitively. In section 6, we introduce a modified correlator of $S^X$. It will also belong to the class $\mathcal P (X,V,E)$. The modification is given by the hypergeometric correcting Euler classes $H^V_\beta$ according to the decomposition type of $V$. In sections 7 and 8, we analyze the torus $T$ action on a generalized flag manifold and its one dimensional orbits, the representations of the section spaces of equivariant line bundles restricted to the orbits. The analysis would be useful to find the explicit expression of $\Phi ^X$. {\it Acknowledgments:} I am grateful to A. Givental and Y.-P. Lee for helping me to understand the paper \cite{GE}; and V. Batyrev, I. Ciocan-Fontaine, B. Fulton, B. Kreu\ss ler, E. Tj\o tta, K. Wirthm\"uller for useful discussions on the papers \cite{GE, GT}. Also, I would like to thank Institut Mittag-Leffler for the financial support during the year-long program, \lq\lq enumerative geometry and its interactions with theoretical physics" in 1996/1997. My special thank goes to D. van Straten for numerous comments and help to improve the clarity of the paper. \section{Mirror Symmetry}\label{setup} \subsection{The moduli space of stable maps} To fix notation we recall the definition of stable maps and some elementary properties of the moduli spaces of stable maps to $X$ \cite{Ko, FuP, BM}. The notion of stable maps is due to M. Kontsevich. We recommend the (survey) paper of W. Fulton and R. Pandharipande \cite{FuP}. A prestable rational curve $C$ is a connected arithmetic genus $0$ projective curve with possibly nodes. The curve is not necessary irreducible. A prestable map $(f,C; x_1,...,x_n )$ is a morphism $f$ from $C$ to $X$ with fixed ordered $n$-many marked distinct smooth points $x_i\in C$. We will identify $(f,C; \{ x_i\} )$ with $(f',C',\{ x_i'\})$ if there is an isomorphism $h$ from $C$ to $C'$ preserving the configuration of marked points such that $f=f'\circ h$. A stable map $(f,C; \{ x_i\} )$ is a prestable map with only finitely many automorphisms. Let $\overline{M}_{0,n}(X,\beta )$ be the (coarse moduli) space of all stable maps $(f, C; \{ x_i \} _{i=1}^n)$ with the fixed homology type $\beta = f_*([C])\in H_2(X,\ZZ )$. Whenever it is nonempty, the moduli space is a connected\footnote{For a proof of the connectedness see \cite{Th}.} compact complex {\it orbifold} with complex dimension $\dim X + <c_1(TX),\beta > + n -3$. \bigskip More precisely, locally near a stable map the moduli space has data of a quotient of a holomorphic domain by the (finite) group action of all automorphisms of the stable map. In the paper \cite{FuP} are constructed smooth open complex domains $V$ with finite groups $\Gamma$ which act on $V$ such that $V/\Gamma$ are naturally glued together in the moduli space of stable maps. Let $X\subset _i \PP ^N$, $\beta\ne 0$, and $(X,i_*(\beta ))\ne (\PP ^1, \kline )$. Here $\kline$ denotes the line class of $H_2(\PP ^N)$. Given a stable map $(f,C)$ (without marked points for simplicity), choose hyperplanes $H_j$ in $\PP ^N$ satisfying that $\{ H_j\}$ gives rise to a basis of $H^0(\PP ^N ,{\mathcal O}(1))$, $f$ is transversal to the hyperplanes, and their inverse images $\{ x_{i,j} \} _i=f^{-1}(H_j)$ contain no nodes of $C$. Then the data $(C;\{ x_{i,j}\})$ determines a point in the moduli space of marked stable curves. Conversely, a point in a suitable closed subvariety of an open smooth domain of the moduli space of marked stable curves naturally determines a stable map $f$ with the extra choices of elements in $(\CC ^\ti )^N$. If $G$ is the product of the symmetric group of the elements of the each group $\{ x_{i,j}\}_i$, then this $G$ has an action sending the data $(f,C; \{ x_{i,j}\})$ to another by permutations of the new marked points. A $(\CC ^\ti )^N$ - bundle of the smooth closed subvariety is an algebraic local chart of the moduli space of stable maps at $f$ with the induced $G$ action. \smallskip {\it Example: } Let $X=\PP ^2$ and $f$ be a stable map without marked points such that $f$ is transversal to the hyperplanes $x=0$, $y=0$, and $z=0$. Assume no singular points of $C$ are mapped into the hyperplanes, and $f_*[C]=2[\text{line}]$. Consider their inverse images (Cartier divisors), $a_1, a_2$, $b_1, b_2$, $c_1, c_2$ in $C$. This information $(C; a_1,...,c_2)$ as a stable curve will determine $f$ uniquely with $(\CC ^\ti )^2$ ambiguity. This $(\CC ^\ti )^2$-bundle over some open subset of the smooth space $\overline{M}_{0,6}$ is the local smooth chart. Notice that for instance, $(C; a_2, a_1, b_1, b_2, c_2, c_1)$ gives rise to the same $f$ up to isomorphism. Thus we have to take account of the quotient by the finite group permuting the elements of sets $\{ a_1, a_2 \}$, $\{ b_1, b_2 \}$ and $\{ c_1, c_2\}$. {\it Claim:} The stabilizer subgroup $G _{(C;\{ x_{i,j}\} )}$ of $G$ is exactly the automorphism group $Aut(f,C)$ of $(f,C)$. {\it Proof:} We shall construct a correspondence between $G _{(C;\{ x_{i,j}\})}$ and $Aut(f,C)$. Let $g\in G _{(C;\{ x_{i,j}\} )}$ which is given by one of the suitable permutations of $x_{i,j}$. So, $g(C;\{ x_{i,j}\} ) = (C; \{ g(x_{i,j})\} )$. Since the permutation does not change the stable curve $(C;\{ x_{i,j}\} )$, there is an isomorphism $h$ from $(C;\{ x_{i,j}\})$ to $(C;\{ g(x_{i,j})\} )$. The isomorphism $h$ is unique since there is no nontrivial automorphism in the stable curve of genus $0$. Of course this $h$ gives rise to an automorphism of $(f,C)$. Conversely, if $h$ is an automorphism of $(f,C)$, then it induces an isomorphism from $(C;\{ x_{i,j}\} )$ to $(C;\{ g(x_{i,j})\})$ for a unique permutation $g$ which we allow. Thus we established 1-1 correspondence, which can be easily seen to be a group homomorphism. {\it Remark:} The action of $Aut(f,C)$ may not be effective in general. For instance, see $\overline{M}_{0,0}(\PP ^1,2\kline )$. \subsection{Gromov-Witten Invariants and $QH_{(T')}^*(V)$} There are natural morphisms on the moduli spaces, namely, evaluation maps $e_i$ at the $i$-th marked points and forgetting-marked-point maps $\pi$: \[\begin{CD} \overline{M}_{0,n+1}(X,\beta ) @> e_{n+1} >> X \\ @V\pi VV \\ \overline{M}_{0,n}(X,\beta ). \end{CD} \] If $s_i$ are the universal sections for the marked points, then $e_i=e_{n+1}\circ s_i$ (here we assume that $\pi$ is the forgetful map of the last marked point). In the orbifold charts, $\pi$ gives the universal family of stable maps as a fine moduli space. Consider, for a second homology class $\beta \ne 0$ and an integer $n\ge 0$, the vector orbi-bundle $V_{\beta}=\pi _*(e_{n+1}^*(V ))$. Here $\pi$ is a flat morphism in the level of orbifold charts. Thus indeed, $V_{\beta}$ is vector orbi-bundle with the fiber $H^0(C, f^*(V ))$ at $(f, C; \{ x_i\})$. Notice that $V_{\beta} =\pi ^*(V_{\beta})$ (it has nothing to do with marked points). \bigskip {\bf Notation:} for $A_i\in H_{(T')}^*(X)$, \begin{eqnarray*} V_0&:=& V \\ \overline{M}_{0,i}(X,0)&:=& X \ \ \text{ for } i=0,1,2 \\ <A_1,...,A_N>^V_{\beta} &:=& \int _{\overline{M}_{0,N}(X,\beta )}e_1^*(A_1) \cup ....\cup e_N^*(A_N) \cup E(V_\beta ) \end{eqnarray*} Then one can show that for all $\beta$ $$\sum _{\beta _1+\beta _2=\beta}\sum _a <A_1,A_2,T_a>^V<T^a,A_3,A_4>^V $$ are totally symmetric in $A_i$. This property will be equivalent to the associativity of the quantum cohomology of $QH_{(T')}^*(V)$ which we define in the below. \bigskip Let us choose a basis $\{ p_i \}_{i=1}^k$ of $H^2(X)$ by classes in the closed K\"ahler cone. \bigskip {\bf Notation:} \begin{eqnarray*} q^\beta &:=& \prod _i q_i^{<p_i,\beta>} \\ <A_1,...,A_N>^V &:=& \sum q^\beta <A_1,....,A_N>^V_\beta \end{eqnarray*} \bigskip The quantum multiplication $\circ$ is defined by the following simple requirement; for $A, B ,C\in H^*_{(T')}(X)$ $$<A\circ B,C >_0^V = <A,B,C>^V$$ which is a formal power series of parameters $q_i$. Thus our quantum cohomology $QH^*_{(T')}(V)$ is defined as $H^*_{(T')}(X)\ot _{\QQ }\QQ [[q_1,...,q_k]]$ with a product structure. \subsection{Givental's Correlators} We review the topic after \cite{Du, GE}. \subsubsection{The flat connections and the fundamental solutions}\label{fund} Now let $q_i=e^{t_i}$ with the formal parameters $t_i$. We have a one-parameter family of the formal $\mathcal D$-module structures on $QH^*_{(T')}(V)$ by giving a flat connection $\nabla _i= \hb\frac{\partial}{\partial t_i}-p_i\circ$ for any nonzero $\hb$, $i=1,...,k$. For the fundamental solutions we introduce $c_i\in H^*_{T'}( \overline{M}_{0,N}(X,\beta ))$, so-called Gravitational descendents. These $c_i$ are the first Chern classes of the universal cotangent line bundles at the $i$-th marked points. The line bundles are, by definition, the dual of the normal bundle of $s_i(\overline{M}_{0,N}(X,\beta ))$ in $\overline{M}_{0,N+1}(X,\beta )$. \bigskip {\bf Notation:} Let $f_i(x)\in H^*_{(T')} [y][[x]]$ for indeterminant $x,y$. Through out the paper, \begin{eqnarray*} <A_1f_1(c),...,A_Nf_N(c);B>_{\beta }^X &:= &\int _{\overline{M}_{0,N}(X,\beta )} e_1^*(A_1)f_1(c_1) ... e_N^*(A_N)f_N(c_N) B \\ <A_1f_1(c),...,A_Nf_N(c);B>_{\beta }^V &:=& <A_1f_1(c),...,A_Nf_N(c);BE(V_\beta )>_{\beta}^X \\ <A_1f_1(c),...,A_Nf_N(c);B>^X &=& \sum _{\beta} q^{\beta} <A_1f_1(c),...,A_Nf_N(c);B>_{\beta }^X \\ <A_1f_1(c),...,A_Nf_N(c);B>^V &:=& \sum _{\beta} q^{\beta}<A_1f_1(c),...,A_Nf_N(c);B>_{\beta }^V \end{eqnarray*} where $A_i\in H^*_{(T')}(X),$ $B\in H^*_{(T')}(\overline{M}_{0,N}(X,\beta ))$. \bigskip The system of the first order equations $\nabla _i \vec{s}=0$, $i=1,...,k$, has the following complete set of ($\dim H^*(X)$)-many solutions \cite{GE}, \begin{eqnarray*} \vec{s}_a &:=& \sum _b<\frac{e^{pt/\hb}T_a}{\hb -c}, T_b>^VT^b , \end{eqnarray*} where $pt$ denotes $\sum _{i=1}^k p_it_i$ and $\hb$ is a formal variable (but when $\beta =0$, set $\hb =1$). The following two formulas show that $\vec{s}_a$ are indeed solutions to the quantum differential system $\nabla _i\vec{s}=0$. Using that $c_i-\pi ^*(c_i )$ is the fundamental class $\Delta _i$ represented by the section $s_i: \overline{M}_{0,n}(X,\beta )\ra \overline{M}_{0,n+1}(X,\beta )$ and $c_i\cup\Delta _i =0$ (the image of $s_i$ is isomorphic to $\overline{M}_{0,3}(X,0)\ti _X\overline{M}_{0,n}(X,\beta )$), it is easy to derive so-called the fundamental class axiom and the divisor axiom \cite{W, GE}. Let $f_i(x)$ be polynomial with coefficients in $\pi ^*(H^*_{(T')}(\overline{M}_{0,n}(X,d)))$. Let $D$ be a divisor class in $H^*_{(T')}(X)$. Then (for $n>0$) \begin{eqnarray*} <f_1(c),...,f_n(c),1>^V_{\beta} &=&\sum _i<f_1(c),...,\frac{f_i(c)-f_i(0)}{c},...,f_n(c)>^V_{\beta} , \end{eqnarray*} (where we abuse the notation \lq$f_i(\pi ^*(c))=f_i(c)$',) and \begin{eqnarray*} <f_1(c), ..., f_n(c), D>^V_{\beta} &=& <D,\beta > <f_1(c), ...,f_n(c)>^V_{\beta} \\ + \sum _i <f_1(c),...,f_{i-1}(c), & D\frac{f_i(c)-f_i(0)}{c} &,f_{i+1}(c),..., f_n(c)>^V_{\beta}. \end{eqnarray*} \bigskip Consider \[e^{pt/\hb }S^V := \sum _a<\vec{s}_{a},1>_0^VT^a= e^{pt/\hb }\sum _a <\frac{T_a}{\hb(\hb -c)}>^VT^a=e^{pt/\hb }(1+o(1/\hb )). \] which is the main object in this paper. This $S^V$ will be called the Givental's correlator for $(X,V,E)$. It is an element in $H^*_{(T')}(X)[\hb ^{-1}][[q_1,...,q_k]]$. Notice that $S^V$ for $(X,V,Euler)$ is homogeneous of degree $0$ if we let $\sum (\deg q _i)p_i= c_1(TX)-c_1(V)$, $\deg \hb =1$, and $\deg A=b$ if $A\in H^{2b}_{T'}(X)$. \bigskip The quantum $\mathcal D$-module of $QH^*_{(T')}(V)$ is defined by the $\mathcal D$-module generated by $<\vec{s},1>^V_{0}$ for all flat sections $\vec{s}$. When there is no $V$ considered, we denote by $QH^*(X)$ the quantum cohomology. That is, using $<...>^X$, we define $QH^*(X)$. \bigskip {\it Remark:} Suppose a differential operator $P(\hbar\frac{\partial}{\partial t_i} ,e^{t_i} ,\hb )$ with coefficients in $H^*_{(T')}$ annihilates $<\vec{s},1>^V_0$ for all flat sections $\vec{s}$, then $P(p_1,...,p_k,q_1,...q_k,0)$ holds in $QH^*_{(T')}(V)$ \cite{GE}. \subsubsection{Examples} The projective space $\PP ^{n}$: It is well known that in the quantum cohomology ring $QH^*(\PP ^n)$, $(p\circ )^{n+1}=q$, where $p=c_1({\mathcal O}(1))$ and $q$ is given with respect to the line class dual to $p$. The corresponding operator is $(\hb\frac d{dt})^{n+1} - e^t$ . The solutions are explicitly known in \cite{GH}. $S^X$ is \[ 1+ \sum _{d>0}e^{dt} \frac 1{((p+\hb )(p+2\hb )...(p+d\hb ))^{n+1}}.\] The complete flag manifolds $F(n):$ Let $F(n)$ be the set of all complete flags $(\CC ^1\subset ...\subset \CC ^n)$ in $\CC ^n$. The usual cohomology ring is $\QQ [x_1, x_2,..., x_n]/(I_1, ..., I_n)$ where $x_i$ are the Chern classes of $(S_i/S_{i-1})^*$, $S_i$ are the universal subbundles with fibers $\CC ^i$ and $I_i$ are the $i$-th elementary symmetric polynomials of $x_1,...,x_n$. Let us use as a basis of $H_2(F(n),\ZZ )$ duals of the first Chern classes of $(S_i)^*$, $i=1,...,n-1$. They are in the edges of the closed K\"ahler cone. Let $A(x_i)$ be a matrix \[\left( \begin{array}{cccccc} x_1 & q_1 & 0 & 0 & ... & 0 \\ -1 & x_2 & q_2 & 0 & ... & \\ & ... & & & ... & \\ 0 & ... & & -1 & x_{n-1} & q_{n-1} \\ 0 & ... & & 0 & -1 & x_n \end{array}\right).\] Then the quantum relations are generated by the coefficients of the characteristic polynomial of the matrix $A(x_i)$. The corresponding differential operators turn out to be obtained by the same method using $A(x_i)$ with arguments $\hb\frac{\partial}{ \partial t_1}$ instead of $x_1$, $\hb\frac{\partial}{ \partial t_{i}}-\hb\frac{\partial}{ \partial t_{i-1}}$ instead of $x_i$, and $-\hb\frac{\partial}{ \partial t_{n-1}}$ instead of $x_n$ \cite{GS, KT}. These differential operators are the integrals of the quantized Toda lattices. The quadratic differential operator of them can be easily derived. In fact, given a quantum relation of $F(n)$ between the divisors $x_i$, there is a unique operator satisfying that its symbol becomes the relation and it annihilates $<\vec{s},1>_0$ for all flat sections $\vec{s}$. In general, the explicit cohomological expression $S^X$ of solutions to the quantum differential operators are not known. \subsubsection{The general quintic hypersurface in $\PP ^4$} Let $Y$ be a smooth degree $5$ hypersurface in $\PP ^4$. $Y$ is not a homogeneous space. However, using virtual fundamental class $[\overline{M}_{0,n}(Y,\beta )]$ \cite{Be, BF, LT}, one can define also the quantum cohomology $QH^*(Y)$ of $Y$. It is expected that \[ <A_1f_1(c),...,A_Nf_N(c)>^Y = <A_1f_1(c),...,A_Nf_N(c)>^{{\mathcal O}(5)}. \] Let $p$ be the induced class of the hyperplane divisor in $\PP ^4$. The quantum relation is $(p\circ ) ^4=0$. The corresponding operator is, however, {\it not} $(\hb\frac d{dt})^4$, but $$(\hb\frac d{dt})^2 \left( \frac{(\hb\frac d{dt})^2}{5+f(q)}\right),$$ where $<p\circ p,p>^Y_{0} = 5+f(q)$. Notice that in this $Calabi-Yau$ 3-fold case, we lose the whole information of quantum cohomology when one concerns only the quantum relation, $(p\circ )^4=0$. The unknown $f(q)$ was conjectured by physicists \cite{Ca}. The general idea of the prediction is the following. Roughly speaking, in theoretical physics, there are quantum field theories associated to Calabi-Yau three-folds by $A$-model and $B$-model. What we have constructed so far are $A$-model objects for Calabi-Yau three-folds. On the other hand using a family of the so-called mirror manifolds which are also Calabi-Yau three-folds, conjecturally one may construct the equivalent quantum field theory by $B$-model. The corresponding mirror partner of a quantum differential equation / quantum ${\mathcal D}$ module is the Picard-Fuchs differential equation / Gauss-Manin connection of the mirror family. It was predicted that they are equivalent by a certain transformation. In \cite{Ca} are obtained the conjectural mirror family of quintics, the Picard-Fuchs differential equation and the transformation. That is how the prediction is made. The prediction is now proven to be correct by Givental \cite{GE}. \subsection{The idea of the proof of theorem \ref{thmmain}} To describe the idea, let us notice that Givental's proof \cite{GT} of the mirror conjecture for the nonnegative toric complete intersections can be divided into three parts. (He shows in the paper that the mirror phenomenon occurs also in non-Calabi-Yau manifolds.) Let $X$ be a Fano toric manifold with a big torus $T$, and $V$ be a $T\ti T'$-equivariant decomposable convex vector bundle over $X$, where $T'$ acts on $X$ trivially. \begin{enumerate} \item In $A$-part, it is proven that \begin{enumerate} \item the $T\ti T'$-equivariant solution vector $S^V\in H^*_{(T\ti T')}(X)[[q,\hb ^{-1}]]$ has an \lq\lq almost recursion relation," \item it satisfies the polynomiality in the so-called \lq\lq double construction," and \item it is uniquely determined by the above two properties with the aymptotical behavior $S^V=1+o(\frac 1{\hb})$. \end{enumerate} \item In $B$-part, another ($T\ti T'$-equivariant hypergeometric) vector $\Phi ^V$, presumably given by the mirror symmetry conjecture, is constructed. It is verified that it also satisfies (a) and (b) using a toric (naive) compactification of holomorphic maps from $\PP ^1$ to $X$. \item When $c_1(X)-c_1(V)$ is nonnegative and $E=Euler$, there is a suitable equivalence transformation between $\Phi ^V$ and $S^V$. \end{enumerate} In this paper, for a $T\ti T'$ equivariant decomposable convex vector bundle $V$ over any compact homogeneous $X$ of a semi-simple complex Lie group $G$, we will show that $S^V$ satisfies property 1 above. In this case, $T$ is a maximal torus of $G$. We define $\Phi ^V$ which corresponds $\Phi ^V$ of the toric case in property 2: Let $\Phi ^X=S^X=\sum _d\Phi ^X_{d}q^d$. For $\Phi ^V$, we will find a modification $H'_d\in H^*_{T\ti T'}(X)[\hb ]$ (depend on $V$ and $d$) such that if $\Phi ^V :=\sum _d \Phi ^X_{d}H'_dq^d$, then (A) $\Phi ^V$ (after the restriction to the fixed points) has the almost recursion relation exactly like $S^V$ and (B) $\Phi ^V$ has the polynomial property in the double construction. In fact, we design $H'_d$ to satisfy (A) and (B). Finally, when $E=Euler$ and $c_1(TX)-c_1(V)$ is nonnegative, we will prove that a certain operation will transform $S^V$ to $\Phi ^V$, since they satisfy the same almost recursion relation and the polynomiality of the double construction. \section{The almost recursion relations}\label{sectionre} As in section \ref{setup} let $X$ be a homogeneous manifold $G/P$ where $G$ is a complex semi-simple Lie group and $P$ is a parabolic subgroup. Let $T$ be a maximal torus. The $T$ action has only isolated fixed points $\{ v,w,... \}$. The one dimensional invariant orbit of $T$ is analyzed in detail in section \ref{grass} and \ref{flag}. For a moment we need the fact that the closures of orbits are {\em finite} $\PP ^1$'s connecting a fixed point $v$ to another fixed point $w$. For a given equivariant vector bundle $W$ over a $T\ti T'$-space $M$, we use $[W]$ which denote the element in the $K$-group $K^0_{T\ti T'}(M)$ corresponding to the $T\ti T'$ vector bundle $W$. \bigskip The torus action on $X$ induces the natural action on the moduli space of stable maps by the functorial property. Since the evaluation maps are $T\ti T'$-equivariant, the pullbacks of $T\ti T'$-bundles have natural actions in the orbifold sense. In turn, $V_\beta$ has the induced $T\ti T'$-action. {\it All ingredients in section 2 are from now on the equivariant ones.} We would like to evaluate $S^V$ as a specialization of the equivariant one corresponding to $S^V$. We use the same notation $S^V\in H^*_{(T\ti T')}(X)[[\hb ^{-1}]][[q_1,...,q_k]]$ for the equivariant one. Notice that $S^V$ might have power series of $\hb ^{-1}$ in each coefficient of $q^d$, since $c$ are not anymore nilpotent. Using the localization theorem, we shall find an \lq\lq almost recursion relation" on the equivariant Givental correlator. To begin with, we summarize the fixed points of the induced action on the moduli space of stable maps. If a stable map represents a fixed point in the moduli space, the image of the map should lie in the closure of the 1-dimensional orbits. The special points are mapped to isolated fixed points. Let us denote by $\ka _{v,w}$ the character of the tangent space of a 1-dimensional orbit connecting an isolated fixed point $v$ to another $w$. Then, $-\ka _{v, w}$ is the character of the tangent line of the 1-dimensional orbit $o(v,w)$ at $w$. We use $\beta _{v,w}$ to stand for the second homology class represented by the ray. Denote by $o(v)$ the set of all fixed points $w\ne v$ which can be connected by a one-dimensional orbit with $v$. \bigskip {\lemma\label{lemmare} {\bf Recursion Lemma} {\em (\cite {GE})} Denote by $\phi _{v}$ the equivariant classes $i_*(1)$ at $v$, here $i_v$ denotes the $T\ti T'$-equivariant inclusion of the point $v$ into $(X,V)$. Then $S^V\in H^*_{T\ti T'}(X)[[\hb ^{-1}]][[q_1,...,q_k]]$ has an \lq\lq almost" recursion relation, namely, for any $v\in X^{T}$, 0) $S^V_v(q,\hb ):=<S^V,\phi _v>^V_0 \in H^*_{(T\ti T')}(\hb )[[q]]$ and the substitution $S_w(q,-\ka _{v,w}/m)$ of $\hb$ with $-\ka _{v,w}$ in $S_w(q,\hb )$ is well-defined, 1) The difference $R_v$ of $S^V_v(q,\hb )$ and the \lq\lq recursion part" is a power series of $q$ over the {\em polynomial} ring of $1/\hb$, that is, \[ R_v:= S^V_v(q,\hb ) - \sum _{w\in o(v),\ m>0} q^{m\beta _{v,w}} \frac {(-\ka _{v,w})/m}{\hb (\ka _{v,w}+m\hb)} \frac{E(V_{v, w,m})i_v^*(\phi _v)} {Euler(N_{v,w,m})}S_w(q,-\ka _{v,w}/m), \] is in $H^*_{(T\ti T')}[\hb ^{-1}][[q]]$, where $V_{v,w,m}$ is $T\ti T'$ representation space $H^0(\PP ^1,f^*V)$, here $f$ is the totally ramified $m$-fold map onto $o(v,w)$ over $v$ and $w$; and $N_{v,w,m}$ is the $T\ti T'$-representation space $[H^0(\PP ^1,f^*TX)]-[0]$; and 2) furthermore, for $S^X$ itself, the first term $R_v$ is $1$.} \bigskip We will say that the statement 1) reveals the almost recursion relation of $S^V$. The statement 2) shows that $S_v^X$ have recursion relations in the ordinary sense. \bigskip {\bf Proof.} First of all, using the short exact sequence \[ 0 \ra Ker \ra V_d \ra e_1^*(V) \ra 0 \] over $\overline{M}_{0,1}(X,d)$, we see that $S^V$ is indeed in $H^*_{T\ti T'}[[\hb ^{-1}]][[q_1,...,q_k]]$. (The last map in the sequence is given by the evaluation of global sections at the marked point.) A connected component of the $T$-fixed loci of the moduli space $X_d:=\overline{M}_{0,1}(X,d)$ is isomorphic to a product of Deline-Mumford spaces with marked points from the special points of the inverse image $f^{-1}(v)$ of the generic $f$ in the component for all $v\in X^{T}$. Now fix a $v$ and consider $S^V_v$. It is enough to count the fixed locus $F^{d,v}$ where the marked point $x$ should be mapped to the fixed point $v$ since $\phi _v$ can be supported only near the point. For a stable map $(f,C;x)$ denote by $C_1$ the irreducible component of $C$ containing the marked point $x$. Then $F^{d,v}$ is the disjoint union of \[ F^{d,v}_1:=\{ (f,C;x)\in F^{d,v} \ | \ f(C_1)=v \}\] and \[ F^{d,v}_2:= \bigcup _{w\in o(v), m=1,...,m\beta _{v,w}\le d} F^{d,v,w,m}, \] where $F^{d,v,w,m}$ is \[ \{ (f,C;x)\in F^{d,v} \ | \ w\in f(C_1),\ \deg f|_{C_1}=m \}. \] $S^V_v$ is an integral over $F^{d,v}$'s by a localization theorem for orbifolds. We claim that the integral of \[ \frac{E(V_d)e_1^*(\phi _v)}{\hb (\hb -c )} \] over $F^{d,v}_1$ is in $H^*_{(T\ti T')}[\hb ^{-1}]$. The reason is that the universal cotangent line bundle over $F^{d,v}_1$ in the moduli space has the trivial action. It implies that the equivariant class $c$ restricted to $F^{d,v}_1$ is nilpotent. \smallskip Now we shall obtain the \lq almost recursion relation' from the contribution of the fixed loci $F^{d,v}_2$. Denote $d-m\beta {v,w}$ by $d'$. Since $C_1$ is always one end of $C$ for any $(f,C;x)\in F^{d,v}_2$, we can have a natural isomorphism from $F^{d',w}$ to $F^{d,v,w,m}$, where $F^{d',w}$ are fixed loci in $X_{d'}:= \overline{M}_{0,1}(X,d')$, consisting of the stable maps sending the marked points to $w$. We obtain the morphism, joining the $m$-covering of $o(v,w)$ to stable maps in $F^{d',w}$. By the $m$-covering of $o(v,w)$, we mean a totally $m$-ramified map from $\PP ^1\cong C_1$ to $o(v,w)$ over $v$ and $w$. Let $x'=f^{-1}(w)\cap C_1$. \bigskip We claim that the normal bundles as in $K^0(F^{d, v,w,m}\cong F^{d',w})$ satisfy the equality \begin{eqnarray} [N_{X_d/F^{d, v,w,m}}]-[N_{X_{d'}/F^{d',w}}] &=& [N_{v,w,m}] -[T_wX] +[T_{x'}C_1\ot L|_{F^{d',w}}] \label{normal} \end{eqnarray} where $L$ is the universal tangent line bundle over $X_{d'}$. The reason of the claim is as follows: Recall that each fixed component is isomorphic to the product of moduli space of stable curves (see section 3 in \cite{Ko} for detail). Hence, we conclude that $[N_{X_d/F^{d, v,w,m}}]-[N_{X_{d'}/F^{d',w}}] -[L|_{F^{d',w}}]$ (over each fixed components) is equal to a trivial bundle with nontrivial actions. The twister by action can be computed by study of action on normal spaces at $(f,C_1\cup C_2 ; x)\in F^{d,v,w,m}$. Let $N_1$ be the normal space of $F^{d,v,w,m}$ at $(f,C_1\cup C_2 ;x )$ and $N_2$ be the normal space of $F^{d',w}$ at $(f|_{C_2}, C_2; x':=C_1\cap C_2)$. Then as representation spaces \begin{eqnarray*} [N_1] &= &[N_2]+([H^0(C_1, f|_{C_1}^*TX)]-[H^0(C_1,TC_1)]) -[T_{w}X] \\ &+& [T_{x'}C_1\ot T_{x'}C_2] +[T_{x'}C_1]+[T_xC_1].\end{eqnarray*} Hence we conclude the claim (\ref{normal}) after canceling of $[H^0(C_1,TC_1)] =[0]+[T_{x'}C_1]+[T_xC_1]$. \bigskip On the other hand, the direct sum of the fiber of $V_d$ at $(f,C_1\cup C_2 ; x)\in F^{d,v,w,m}$ and $V|_w$ is equal to the direct sum of the fiber of $V_{d'}$ at $(f|_{C_2}, C_2;x')$ and $H^0(C_1, (f|_{C_1})^*V)$. Thus, applying the localization theorem we obtain \begin{eqnarray*} &&\int _{X_d}\frac{E(V_d)e_1^*(\phi _v)}{\hb(\hb -c)}=I+ \sum _{w\in o(m), 0<m; m\beta _{v,w}\le d} \frac{E(V_{v,w,m})i^*_v(\phi _v)(-\ka _{v,w}/m)}{m\hb (\ka _{v,w}/m +\hb ) Euler(N_{v,w,m})} \\ && \ti \int _{X_{d-m\beta _{v,w}}} \frac{E(V_{d-m\beta _{v,w}})e_1^*(\phi _w)}{(-\ka _{v,w}/m)(-\ka _{v,w}/m - c)}, \end{eqnarray*} where $I$ is the integral over $F^{d,v}_1$. The factor $m$ in $m\hb (\ka _{v,w}/m +\hb ) $ comes from the nature of orbifold localization theorem. (There are $m$ automorphisms of $f|_{C_1}$.) \bigskip Using induction on $|d|=\sum d_i$, we may assume that the integral factors in the second term are well-defined and belong to $H^*_{(T\ti T')}$. (Localization theorem itself also explains them.) So, statements 0) and 1) in the lemma are proven. \smallskip Now let us prove statement 2). Since $<c_1(TX),\beta >\ge 2$ for all $\beta$, by degree counting we see that there are no contributions from the integral over $F^{d,v}_1$. The reason is that $\dim \overline{M}_{0,\sum d_i+1} =(\sum d_i)-2$ is less than $2(\sum d_i )-2$ if $(d_1,..,d_k)\ne 0$ and $\dim \overline{M}_{0,1}(X,d)\ge 2\sum d_i+\dim X -2$. So, in the case of $S^X$, $R_v =1$. \section{The double construction} {\lemma\label{lemmado}{\bf Double Construction Lemma} The double construction \[ W(S^V):=\int _{V} S^V (qe^{\hb z},\hb )e^{\sum p_iz_i}S^V (q,-\hb )\] is a power series of $q_1,...,q_k$ and $z_1,...,z_k$ with coefficients in $H_{T\ti T'}^*[\hb ]$.} \bigskip A priori $W(S^V)$ has coefficients in Laurent power series ring of $\hb ^{-1}$ over $H_{T\ti T'}^*$. For the proof we will make use of graph spaces and universal classes defined in the below. \subsection{The main lemma} Let $L_d$ be the projective space of the collection of all $(f_0,...,f_N)$ such that $f_i(z_0,z_1)$ are homogeneous polynomials of degree $d$. $L_d$ is isomorphic to $\PP ^{(d+1)(N+1)-1}$. Given a stable map of degree $(d,1)$ from a prestable curve $C$ to $\PP ^N\ti \PP ^1$, there is a special irreducible component $C_0$ of $C$ such that $C_0$ has degree $(d_0,1)$ under the stable map. This special component $C_0$ is parameterized by $\PP ^1$ in the target space. Thus we can identify $C_0$ with $\PP ^1$ and keep track where the other components intersect. Suppose the other connected components $C_1,...,C_l$ of $C-C_0$ intersect with $C_0=\PP ^1$ at $[x_1: y_1],...,[x_l: y_l]$. If the degrees of $C_i$ are $d_i$ under the stable map, we now associate the stable map to $$\prod _{i=1}^l(y_iz_0 -x_iz_1)^{d_i}(f^0_0,...,f_N^0),$$ where $(f^0_0,...,f_N^0)$ are the polynomials coming from the data of the restriction of $f$ to $C_0$. \bigskip {\bf Main lemma: } (Givental \cite{GE}) {\em The above \lq\lq polynomial" mapping from $G_d(\PP ^N):=\overline{M}_{0,0}( \PP ^N\ti \PP ^1, (d,1))$ to $L_d$ is a $(\CC ^\ti)^N\ti \CC ^\ti$-equivariant morphism, where $\PP ^N$ has the diagonal $(\CC ^\ti)^N$ action and $\PP ^1$ has the $\CC ^\ti$ action by $[z_0:z_1]\mapsto [tz_0:z_1]$ for $t\in \CC ^\ti$.} \bigskip Notice that the $\CC^\ti$ action on $L_d$ is given by \[ [f_0(z_0,z_1):...:f_N(z_0,z_1)] \mapsto [f_0(t^{-1}z_0,z_1):...:f_N(t^{-1}z_0,z_1)]\] for $t\in \CC ^\ti$. \subsection{The universal class} The $T\ti T'$-equivariant spanned line bundle $U_i$ over $X$ gives rise to the $T\ti T'$-equivariant morphism $\mu ^i_0: X\ra\PP ^N$, and so we obtain: \[ \begin{CD} (\mu _d^i) ^*({\mathcal O}(1)) @. @. {\mathcal O}(1) \\ @VVV @. @VVV \\ G_d(X)@>>> G_{d_i}(\PP ^N)@>>> L_{d_i} \\ \end{CD}, \] where $G_d(X)$ is the graph space $\overline{M}_{0,0}(X\ti \PP ^1, (d,1))$, and $\mu _d^i$ is the $T\ti T'\ti \CC ^\ti$-equivariant map from $G_d(X)$ to $L_{d_i}$. On ${\mathcal O}(1)$ we choose the lifted $\CC ^{\ti}$-action coming from the action on the vector space of $N+1$ $d_i$-homogeneous polynomials by \[ [f_0(z_0,z_1):...:f_N(z_0,z_1)] \mapsto [f_0(z_0,tz_1):...:f_N(z_0,tz_1)]\] for $t\in \CC ^\ti$. Denote by $P_i=c_1((\mu ^i_d) ^*{\mathcal O}(1))$, the $T\ti T'\ti \CC ^\ti $-equivariant Chern class. It is said to be a universal class in the paper \cite{GT}. Denote by $W_d$ the vector orbi-bundle over $G_d(X)$ with the fiber $H^0(C,\psi ^*\pi _1^*V)$ at $(C,\psi )$: Consider \[\begin{CD} G_{d,1}(X)@>>{e_1}> X\ti \PP ^1 \\ @V\pi VV @AAA \\ G_d(X) @. \pi _1^*V, \end{CD}\] where $G_{d,1}(X)$ denotes the graphs space with one marked point and $\pi _1$ is the projection of $X\ti \PP ^1$ to the first factor $X$. Then $W_d:=\pi _* e_1^*\pi _1^*V$. \subsection{Proof of Lemma \ref{lemmado}}\label{pfdo} It is enough to show the equality \[\sum _dq^d\int _{G_d(X)}e^{Pz}E(W_d) =\int _{V} S^V (q,\hb )e^{pz}S^V (qe^{-\hb z},-\hb ).\] The left integral is a $T\ti T'\ti \CC ^{\ti}$-equivariant push forward with $\hb$ as $c_1({\mathcal O}(1))$ over $\PP ^{\infty}$ and the right one is a $T\ti T'$-equivariant push forward with a formal variable $\hb$. \bigskip We will apply localization theorem. Let us analyze the $\CC ^\ti$-action fixed loci $G_d(X)^{\CC ^\ti}$ of $G_d(X)$. $G_d(X)^{\CC ^\ti}$ is isomorphic to $\sum _{d^{(1)}+d^{(2)}=d} \overline{M}_{0,1}(X,d^{(1)})\ti _X\overline{M}_{0,1}(X,d^{(2)})$. \bigskip Suppose $|d^{(1)}|+|d^{(2)}|\ne 0$. The normal bundle is as follows: When $|d^{(1)}||d^{(2)}|=0$: The codimension is 2 (one from the nodal condition and the other from the condition of the image of the nodal point). Then the Euler class of the normal bundle is $\hb (\hb -c_0)$, or $-\hb (-\hb -c_\infty)$, where $c_0$ and $c_\infty$ are the Chern classes of universal cotangent line bundles of the first marked point over $\overline{M}_{0,1}(X, d^{(1)})$ and $\overline{M}_{0,1}(X, d^{(2)})$, respectively. Here we assume the following convention: $0=[0:1], \infty =[1:0]$, the associated equivariant line bundle to the character 1 of the group $\CC ^\ti$ has $\hb$ as its equivariant Chern class. When $|d^{(1)}||d^{(2)}|\ne 0$: The codimension is $4$ and the Euler class is $\hb (\hb -c_0)(-\hb )(-\hb -c_\infty )$. Here, for instance, $c_0\in H^2(\overline{M}_{0,1}(X,d^{(1)})\ti _X\overline{M}_{0,1}(X,d^{(2)}))$ is the pull-back of the Chern class of the universal cotangent line bundle of the first factor of $\overline{M}_{0,1}(X,d^{(1)})\ti _X\overline{M}_{0,1}(X,d^{(2)})$. \bigskip Let us analyze $P_i$ restricted to $G_d(X)^{\CC ^\ti}$. Consider the commutative diagram, \[\begin{CD} G_{d^{(1)},d^{(2)}}(X):=\overline{M}_{0,1}(X,d^{(1)})\ti _X \overline{M}_{0,1}(X,d^{(2)}) @>>\mu _d^i > L_{d_i} @. \ni z_0^{d^{(1)}_i}z_1^{d^{2)}_i}[x_0:...:x_N]\\ @V{\pi _2} VV @AAA @AAA\\ \overline{M}_{0,1}(X,d^{(2)}) @>>{\mu ^i_0\circ e_1}> \PP ^N @. \ni [x_0:...:x_N], \end{CD}\] where the first vertical map $\pi _2$ is the projection and under the second vertical map $\PP ^N$ is embedded into $L_{d_i}$ as the $\CC ^\ti$ -action fixed locus of the part $\{ z_0^{d^{(1)}}z_1^{d^{(2)}}[x_0:...:x_N] | [x_0:...:x_N]\in \PP ^N\} $. One concludes that $e_1^*\circ (\mu _0^i)^* (c_1({\mathcal O}(1)|_{\PP ^N}))= e_1^*(p_i)-d_i^{(2)}\hb $ and so \[ \sum P_iz_i|_{G_{d^{(1)},d^{(2)}}} =\sum (\pi _2^*e_1^*(p_i) -d^{(2)}_i\hb )z_i. \] Since \begin{eqnarray*} && S^V(q,\hb )e^{pz} S^V(qe^{-\hb z},-\hb) \\ &=& \sum _{a ,b,d^{(1)},d^{(2)} } <\frac{T_{a}}{\hb (\hb -c)}>^V_{d^{(1)}} T^{a}q^{d^{(1)}} e^{pz}<\frac{T^{b}}{-\hb (-\hb -c)}>^V_{d^{(2)}}T_{b}q^{d^{(2)}}e^{-d^{(2)}\hb z}, \end{eqnarray*} we see that \begin{eqnarray*} & & \int _{V} S^V(q,\hb)e^{pz}S^V(qe^{-\hb z},-\hb) \\ &=& \sum _{a ,d^{(1)},d^{(2)}}<\frac{T_{a}}{\hb (\hb -c)}>^V_{d^{(1)}} <\frac{T^{a}e^{pz-d^{(2)}\hb z}}{-\hb (-\hb -c)}>^V_{d^{(2)}}q^{d^{(1)}+d^{(2)}} \\ &=& \sum _d q^d\int _{G_{d^{(1)},d^{(2)}}(X)} \frac{e^{(\pi _2^*e_1^*p-d^{(2)}\hb)z}E(W_d)} {[N_{G_d(X)/G_{d^{(1)},d^{(2)}}(X)}]} \\ &=& \sum _d q^d\int _{G_d(X)}e^{Pz}E(W_d), \end{eqnarray*} after applying the localization theorem only for $\CC ^\ti$ action on $G_d(X)$. \section{The class ${\mathcal P}({\mathcal C})$ and mirror transformations}\label{sectiontr} \subsection{The class $\mathcal P({\mathcal C})$} Let ${\mathcal C}$ be the collection of given data of $C_{v,w,m}\in H^*_{(T\ti T')}$, $\ka _{v,w}\in H^*_{T\ti T'}$, and $\beta _{v,w}\in \Lambda -0$, for all $(v,w,m)\in X^{T}\ti X^{T}\ti \NN$ with $v\in o(w)$. Here $\NN$ is the set of positive integers. Assume that $(p_i)_w-(p_i)_v=-<p_i,\beta _{v,w}>\ka _{v,w}$ for all $i=1,...,k$. Define degree of $\hb$ as 1. Let $q_1,...,q_k$ be formal parameters with some given nonnegative degrees. Define the degree of a homogeneous class of $H_{T\ti T'}^b(X)$ as $b/2$. Let $\mathcal P({\mathcal C})$ be the class of all $Z(q,\hb )\in H^*_{T\ti T'}(X)[[\hb ^{-1}, q]]$ of homogeneous degree 0 such that \begin{description} \item[a)] $Z(0,\hb )=1$, $Z_v(q,\hb ):=<Z,\phi _v>^V_0$ (this is not depend on $V$) is in $H^*_{(T\ti T')}(\hb )[[q]]$ for any fixed point $v$, and $Z_w(q,-\ka _{v,w}/m )$ are well-defined for all $v\in o(w),m>0$ ($m$ are positive integers), \item[b)] the almost recursion relation for each fixed point $v$ holds, that is by definition, \begin{eqnarray*} R_v:= Z_v(q, \hb ) - \sum _{m>0,w\in o(v)} q^{m\beta _{v,w}}\frac{C_{v,w,m}}{\hb (\ka _{v,w} +m\hb )}Z_w(q,-\ka _{v,w}/m), \end{eqnarray*} is in $H^*_{(T\ti T')}[\hb ^{-1}][[q]]$, where $$ q^{m\beta _{v,w}}:=\prod _i q_i^{m<p_i,\beta_{v,w}>}$$; and \item[c)] in the double construction \[ W(Z)(q,z):=\int _{V}Z(qe^{\hb z}, \hb )e^{\sum p_iz_i}Z(q,-\hb ), \] is in $H^*_{T\ti T'}[\hb][[q,z]]$. (We use the multi-index notation for $z=(z_1,...,z_k)$ and $q=(q_1,...,q_k)$.) \end{description} \bigskip Whenever the data $\mathcal C$ comes from $(X,V,E)$ as in lemma \ref{lemmare}, we denote the class by ${\mathcal P}(X,V,E)$. So, in the case \[ C_{v,w,m} :=C_{v,w,m}^V:= \frac{(-\ka _{v,w})/m\ E(V_{v, w,m})i_v^*(\phi _v)} {Euler(N_{v,w,m})}, \] $\ka _{v,w}$ is the character of $T_vo(v,w)$, $\beta _{v,w}=[o(v,w)]\in H_2(X,\ZZ )$, and $$c_1(TX)-c_1(V)=\sum _{i=1,...,k}(\deg q_i) p_i.$$ So far, we proved that $S^V$ for $E=Euler$ is in class ${\mathcal P}(X,V,Euler)$. \bigskip In the below we introduce on ${\mathcal P}({\mathcal C})$ a transformation group generated by the following three types of operations. \begin{description} \item[1) Multiplication by $f(q)$] Let $f(q)=\sum _{d\ge 0}f_dq^d$, where $f_d\in \QQ$, $f(q)$ is homogeneous of degree $0$, and $f(0)=1$. Then $f(q)Z\in {\mathcal P}({\mathcal C})$. \item[2) Multiplication by $\mathrm{exp} (f(q)/\hb )$] Let $f(q)=\sum _{d>0}f_d q^d$, where $f_d$ are in $H^*_{T\ti T'}$. Suppose $\deg (f(q))=1$. Then $Z^{new}:=\exp (f(q)/\hb )Z$ is still in $\mathcal P({\mathcal C})$. \item[3) Coordinate changes] Consider a transformation: $$Z\ra Z^{new}:= \exp (\sum _if_i(q)p_i/\hb )Z(q\exp (f(q)), \hb )$$ where $f_i(q)=\sum _{d>0} f^{(d)}_iq^d$ of homogeneous degree $0$, $f_i^{(d)}\in \QQ$, and $q\exp (f(q))=(q_1\exp (f_1(q)),...,q_k\exp (f_k(q)))$. Then $Z^{new}$ is still in $\mathcal P({\mathcal C})$. \end{description} Let us call the transformation group the mirror group. {\theorem\label{thmtr} {(\em \cite{GT})} Suppose $\deg q $ are nonnegative and there is at least one element of form $1+o(\hb ^{-1} )$ in the class ${\mathcal P}({\mathcal C})$. Then the mirror group action on ${\mathcal P}({\mathcal C})$ is free and transitive.} \bigskip First, we will check 1), 2) and 3); and prove the so-called uniqueness lemma and then theorem above. {\it Proof of 1).} First, $Z^{new}:=fZ$ is homogeneous of degree $0$, $f(0)Z(0,\hb )=1$, and $fZ_v$ are in $H_{(T\ti T')}^*(\hb )[[q]]$, and of course $Z^{new}_w(q,-\ka _{v,w}/m)$ are well-defined. Second, \begin{eqnarray*} Z_v^{new} &=& f(q)R_v + \sum q^{m\beta _{v,w}}\frac{C_{v,w,m}}{\hb (\ka _{v,w} +m\hb )} Z^{new}_w(q,-\ka _{v,w}/m). \end{eqnarray*} Thus $fZ$ has the almost recursion relation. Finally, \begin{eqnarray*} W^{new} &:=&\int _{V} Z^{new}(qe^{\hb z},\hb )e^{pz}Z^{new}(q,-\hb ) \\ &=& f(qe^{\hb z})f(q)W, \end{eqnarray*} which still has the polynomial coefficients in $H^*_{T\ti T'}[\hb ]$. {\it Proof of 2).} The new $Z^{new}$ is homogeneous of degree $0$, $Z^{new}(0,\hb)=1$, $Z^{new}_v$ are in $H^*_{(T\ti T')}(\hb )[[q]]$, and $Z^{new}_w(q,-\ka _{v,w}/m)$ are well-defined. Since $\exp (\frac{f(q)}{\hb }+\frac{mf(q )}{\ka _{v,w}}) =1+(\ka _{v,w}+m\hb ) g_{\ka _{v,w},m}$ and $g_{\ka _{v,w},m}$ is a $q$-series with polynomial coefficients in $H^*_{(T\ti T')}[\hb ^{-1}]$, $Z^{new}$ has the almost recursion relation. Once again, \begin{eqnarray*} W^{new} &=& \exp(\frac {1}{\hb} (f(qe^{\hb z})-f(q)))W .\end{eqnarray*} But $f(qe^{\hb z} )-f(q ) =\sum_{d>0} f_d((e^{\hb z})^d -1) q^d$ is a $(z,q)$-series with polynomial coefficients in $\hb H^*_{T\ti T'}[\hb ]$. {\it Proof of 3).} The $Z^{new}$ is homogeneous of degree $0$, $Z^{new}(0,\hb )=1$, $Z^{new}_v$ are in $H^*_{(T\ti T')}( \hb )[[q]]$, and $Z_w^{new}(q,-\ka _{v,w}/m)$ make sense. Since $(p_i)_w-(p_i)_v=-<p_i,\beta _{v,w}>\ka _{v,w}$, \begin{eqnarray*} && \sum _if_i(q)(p_i)_v/\hb = \sum _i f_i(q)(p_i)_w/(-\ka _{v,w}/m) \\ && - m\sum _i <p_i,\beta _{v,w}>f_i(q) + \sum _i\frac{f_i(q)(p_i)_v }{\ka _{v,w}\hb }(m\hb +\ka _{v,w}) . \end{eqnarray*} The exponential of the last term on the right can be denoted by $1+(\ka _{v,w}+m\hb )g_{\ka _{v,w},m}$ where $g_{\ka _{v,w},m}$ is a $q$-series with coefficients which are in $H^*_{(T\ti T')}[\hb ^{-1}]$. $Z^{new}$ satisfies the almost recursion relation. Consider the double construction \begin{eqnarray*} W^{new}(q,z) &=& \int _{V}e^{f(qe^{\hb z})p/\hb } Z(qe^{\hb z}e^{f(qe^{\hb z})},\hb ) e^{pz}e^{-f(q)p/\hb }Z(qe^{f(q)},-\hb ) \\ &=& W(qe^{f(q)}, z+\frac{f(qe^{\hb z})-f(q)}{\hb }) . \end{eqnarray*} But since $f(qe^{\hb z})-f(q)$ is divisible by $\hb $, $W^{new}$ is a polynomial $(q,z)$-series. \bigskip {\lemma\label{lemmaun} {\bf Uniqueness Lemma} Let $Z=\sum _{d\ge 0}Z_dq^d$ and $Z'=\sum _{d\ge 0} Z'_dq^d$ be series in ${\mathcal P}({\mathcal C})$. Suppose $Z\equiv Z'$ modulo $(\frac 1\hb )^2$. Then $Z'=Z$.} \bigskip {\bf Proof.} We may suppose that $Z'_{d}=Z_{d}$ for all $0\le d <d_0$ for some $d_0\ge 1$. Let \begin{eqnarray*} D(\hb ):=Z'_{d_0}-Z_{d_0} &=& A\hb ^{-2r-1}+B\hb ^{-2r}+....\\ &=&\hb ^{-2r}(A/\hb +B + O(\hb )) ,\end{eqnarray*} where $A, \ B \in H^*_{T\ti T'}(X)$. ($A$ might be $0$.) This is possible since $<D,\phi _v>_0$ for all $v$ are polynomials of $1/\hb$ over $H^*_{(T\ti t')}$ and so $D$ is a polynomial of $1/\hb$ over $H^*_{T\ti T'}(X)$. Consider the coefficient of $q^{d_0}$ in $W(Z')-W(Z)$, which can be set $\delta (D) =\int _{V}e^{(p+d_0\hb )z}D(\hb )+e^{pz}D(-\hb ) $. If $r=0$, then $D=0$ since $D\equiv 0$ modulo $(1/\hb )^2$. Assume $r\ge 1$. We shall show that $A=0=B$, which implies by induction that $D=0$. Notice that, since $\kd (D)$ is a polynomial of $\hb $, \begin{eqnarray*} O(\hb ^2)=\hb ^{2r}\kd (D) &=&\int _{V}e^{(p+d_0\hb )z} (A/\hb + B+ O(\hb )) +e^{pz}(-A/\hb +B +O(\hb )) \\ &=&\int _{V}e^{pz}Ad_oz +2Be^{pz} +O(\hb ). \end{eqnarray*} So, \begin{eqnarray*} 0 &=& d_0z\int _{V}e^{pz}A + 2 \int _{V}Be^{pz} \\ &=& \sum _{v\in X^T}(d_0ze^{p_vz}A_v+ 2 e^{p_vz}B_v)\frac {1}{i_v^*(\phi _v)}, \end{eqnarray*} where $A_v$, $B_v$, and $Euler(V)_v$ are the restrictions of $A$, $B$, and $Euler(V)$ to the fixed point $v$, respectively. Since $p_vz$ are different as $v$ are different (this can be seen in section \ref{grass} and \ref{flag}), $e^{p_v z}$ and $ze^{p_vz}$ are independent over $H^*_{(T\ti T')}$. So we conclude that $A_v=0=B_v$ for all $v$, and hence $A=0=B$. \subsection{Proof of theorem \ref{thmtr}} It suffices to show the transitivity of the action. Let $Z_1$ and $Z_2$ be in class ${\mathcal P}({\mathcal C})$ and let $Z_1=1+o(1/\hb )$. Since $\deg q\ge 0$, we may let $$ Z_2=Z_2^{(0)} +Z_2^{(1)}\frac 1\hb + o(\frac 1\hb ),$$ where $Z_2^{(0)}\in H^*_{T\ti T'}(X)[[q]]$ is of homogeneous of degree $0$ and $Z_2^{(1)}$ is homogeneous of degree $1$. Furthermore, $Z_2^{(0)}(q)\in H^*_{T\ti T'}(X)[[q]]$ is a $q$-series with coefficients in $\QQ$, $Z_2^{(0)}(0)=1$, and $Z_2^{(1)}(q)$ is a $q$-series with coefficients in $H^*_{T\ti T'} [p]$ by degree counting. We may let $$\frac{Z_2^{(1)}(q )}{Z _2^{(0)}(q)}= \sum _i (f_i(q)\cdot p_i) + g(q),$$ where $f_i(q)$ are pure $q$-series over $\QQ$ of degree 0 and $g(q )$ are degree 1 in $H^*_{T\ti T'}[[q]]$. In addition, $f_i(0)=0=g(0)$. Now, consider operations on $Z_1$: first, coordinate changes, \[ Z_1'=\exp (f(q)p/\hb )Z_1(q\exp (f(q)), \hb ) = 1+ f(q)p/\hb +o(1/\hb ), \] second, multiplication by $\exp (g(q)/\hb )$, \[Z_1''=\exp (g(q) /\hb ) Z_1' = 1 + \frac 1\hb (f(q)p + g(q)) + o(1/\hb ), \] finally, multiplication by $Z_2^{(0)}(q)$, \[ Z_1''' = Z_2 ^{(0)}(q)Z_1'' =Z_2 ^{(0)} +\frac 1\hb Z_2 ^{(1)} + o (1/\hb ).\] According to the uniqueness lemma, the last one $Z_1'''$ must be equal to $Z_2$ since $Z_1'''\cong Z_2$ modulo $(1/\hb )^2$. \subsection{Transformation from $J^V$ to $I^V$} We explain the transformation introduced in the introduction. Let $\tilde{Z}$ be the nonequivariant specialization of $Z$. Let $Z_1$ and $Z_2$ be in class ${\mathcal P}({\mathcal C})$ and let $Z_1=1+o(1/\hb )$. Now let us specialize the equivariant setting to nonequivariant one. Let $J^V=e^{(t_0+pt)/\hb }\tilde{Z_1}(q)$ and $I^V=e^{(t_0+pt)/\hb }\tilde{Z_2}(q)$. Then, they are equivalent up to the unique coordinate change $t_0\mapsto t_0 + f_0(q)\hb + f_{-1}(q)$ and $t\mapsto t_i + f_i(q)$, $i=1,...,k$, where $f_j\in \QQ [[q]]$ for all $j$, $f_0$ and $f_i$ ($i=1,...,k$) have degree $0$, $f_{-1}$ has degree $1$; and $f_j(0)=0$ for all $j$. \section{The modified B series} Let $X$ be a homogeneous manifold with the torus $T\ti T'$ action. From now on let $V=L_1\oplus ...\oplus L_l$ be an equivariant decomposable convex vector bundle over $X$, where $L_i$ are line bundles. \subsection{The correcting Euler classes} Let $x=(x_1,...,x_l)$ be indeterminant. Define a polynomial of $x$ over $\ZZ [\hb ]$ for $\beta\in \Lambda$: $$ H_{\beta}(x,\hb ) :=\prod _{i=1}^{l}\prod _{m=0}^{<c_1(L_i),\beta >}(x_i+m\hb ).$$ Set $$H'_{\beta}(x,\hb ):=\frac{H_{\beta}(x,\hb )}{\prod x_i}. $$ We treat each linear factor $(x_i+m\hb)$ of $H_\beta$ as a Chern character. Define \[ \Phi ^V (q,\hb ):=\sum _{d\in\Lambda }\sum _a q^d<\frac{T_a}{\hb (\hb -c)}>^X_dT^a E(H'_d(x,\hb ))(c_1(L),\hb ),\] where $c_1(L)=(c_1(L_1),...,c_1(L_k))$. \bigskip {\it Claim} 1. $(p_i)_w = (p_i)_v-<p_i, \beta _{v,w}>\ka _{v,w}$, 2. $c_1(L)_w = c_1(L)_v - < c_1(L), \beta _{v,w}>\ka _{v,w}$ 3. $E(V_{v,w,m})=E(H_{m\beta _{v,w}})(c_1(L)_v,-\frac{\ka _{v,w}}{m}).$ {\it Proof:} Let $U$ be any equivariant convex line bundle. On the ray $o(v,w)$ ($\cong \PP ^1$), we have a homogeneous coordinate $[z_0:z_1]$ such that the induced action on the ray is linear (because of the equivariant embedding theorem). We have also global sections $z_0^n,z_0^{n-1}z_1,...,z_1^n$ of the restriction $U|_{\PP ^1}$ of $U$ to the ray, where $[1:0]=w$, $[0:1]=v$ and $n=<c_1(U), \beta _{v,w}>$. We know that $z_0^n$, $z_1^n$, $z_0/z_1$ have the characters $c_1(U)_w$, $c_1(U)_v$ and $-\ka _{v,w}$, respectively. This concludes the proof. \bigskip The first one in the claim shows that we have the well-defined classes ${\mathcal P}(X,V,E)$. (Otherwise, the mirror group transformation may not preserve the class ${\mathcal P}(X,V,E)$.) \bigskip {\theorem\label{thmmo} Suppose $c_1(TX)-c_1(V)$ is in ample cone. Then $\Phi ^V$ is in the class ${\mathcal P}(X,V,Euler)$. } \bigskip Notice that for $\beta ' \le \beta$ \begin{eqnarray} H_\beta (x-<c_1(L), \beta ' >\hb , \hb ) &= &H_{\beta '}(x,-h)H'_{\beta -\beta '}(x,\hb ) \label{Hdo} \\ H'_\beta (x, \hb ) &=&H'_{\beta '}(x,h)H'_{\beta -\beta '}(x+<c_1(L),\beta '>\hb ,\hb ) , \label{Hre} \end{eqnarray} which will show the polynomiality of double construction and the almost recursion relation for $\Phi ^V$, respectively. \subsection{The proof of theorem} The homogeneous of $\Phi ^V$ is clear when $E=Euler$ and the rest properties will be proven for general $E$. First of all, it is easy check to see \[ \Phi ^V \in H^*_{T\ti T'}[[\hb ^{-1}]][[q]] .\] For the polynomiality, consider \begin{eqnarray} && \int _V \Phi ^V(q,\hb ) e^{pz}\Phi ^V(qe^{-\hb z},-\hb ) \label{Hdoo} \\ &=& \sum _d \sum _{d^{(1)}+d^{(2)}=d, a} q^{d^{(1)}} <\frac{T_a E(H_{d^{(1)}})(c_1(L),\hb )}{E(V)\hb (\hb -c)}>^X_{d^{(1)}} \nonumber \\ && q^{d^{(2)}}<\frac{T^ae^{(p-d^{(2)})z} E(H_{d^{(2)}})(c_1(L),-\hb )}{ -\hb (-\hb -c)}>^X_{d^{(2)}}, \nonumber \end{eqnarray} where $<T_a, T^b>_0^X=\delta _{a,b}$. Let us use the notation and facts in \ref{pfdo}. Since \begin{eqnarray*} && E(H_{d^{(1)}})(c_1(L),\hb )E(H_{d^{(2)}})(c_1(L),-\hb ) \\ &=& E(H_d(x-<c_1(L),d^{(2)}>\hb,\hb ))(c_1(L),\hb )E(V) \end{eqnarray*} from (\ref{Hdo}), the universal class $U(c_1(L))$ corresponding to $c_1(L)$ restricted to $G_{d_1,d_2}(X)$ is \[ c_1(L)-<c_1(L),d^{(2)}>\hb ,\] and $e_1\circ \pi _1 = e_1\circ \pi _2$, (\ref{Hdoo}) is equal to \begin{eqnarray*} && \sum _d q^d\int _{G_{d^{(1)},d^{(2)}}(X)} \frac{ e^{(\pi _2e^*_1p-d^{(2)}\hb)z}E(H_d)(U(c_1(L)),\hb )} {[N_{G_d(X)/G_{d^{(1)},d^{(2)}}(X)}]} \\ &=& \sum _d q^d \int _{G_d(X)} e^{Pz}E(H_d)(U(c_1(L)),\hb ), \end{eqnarray*} which shows the polynomiality. Now let us check the almost recursion relation. Let $$S _v^X(q,\hb ):= <S ^X,\phi ^X_v>^X_0=\sum _d S _{v,d}^X(\hb )q^d.$$ Since (if $d\ne 0$) $$S _{v,d}^X(\hb )=\sum _{w\in o(v), 0<m; m\beta _{v,w}\le d} \frac {C^X_{v,w,m}}{\hb (\ka _{v,w} + m\hb )} S ^X_{w,d-m\beta _{v,w}}( -\frac{\ka _{v,w}}{m}) $$ and \[ E(H'_{\beta})(c_1(L)_v,-\frac{\ka _{v,w}}{m}) =\frac{E(V_{v, w,m})}{E(V)_v} E(H'_{\beta -m\beta _{v,w}})(c_1(L)_w, -\frac{\ka _{v,w}}{m}) \] from (\ref{Hre}) and the Claim, we obtain that \begin{eqnarray*} &&\Phi ^{V}_{v,d}(\hb ):=<\Phi ^V_d(\hb ),\phi _v>^V_0 = R_{v,d} \\ &+&\sum _{w\in o(v), 0<m; m\beta _{v,w}\le d} \frac{C^X_{v,w,m}E(V_{v,w,m})} {E(V)_v\hb (\ka _{v,w} +m \hb )} \\ && \ti \Phi ^X_{w,d-m\beta _{v,w}} (-\frac{\ka _{v,w}}{m}) E(H'_{d-m\beta _{v,w}})(c_1(L)_w,-\frac{\ka _{v,w}}{m}), \end{eqnarray*} where $R_{v,d}$ is indeed a polynomial of $1/\hb $ over $H^*_{(T\ti T')}$. However, since $$ C^V_{v,w,m} = \frac{C^X_{v,w,m}E(V_{v,w,m})}{E(V)_v},$$ $\Phi ^V_v(q,\hb )$ has the same almost recursion coefficients $C^V_{v,w,m}$ with $S^V$. \subsection{Proof of main theorem \ref{thmmain}} Recursion lemma \ref{lemmare} and double construction lemma \ref{lemmado} show that $S^V$ is in class ${\mathcal P}(X,V,Euler)$. Certainly $S^V$ is form of $1+o(\hb ^{-1})$. According to theorem \ref{thmmo}, $\Phi ^V$ also belongs to ${\mathcal P}(X,V,Euler)$. Then theorem \ref{thmtr} concludes the proof. (We use the condition that $E=Euler$, in order to make sure that $S^V$ and $\Phi ^V$ are homogeneous of degree 0.) \section{Grassmannians}\label{grass} \subsection{Notation} Let $e_1,...,e_n$ form the standard basis of $\CC ^n$, $T=(\CC ^\ti )^n$ the complex torus, and $X:=Gr(k,n)$ the Grassmannian, the set of all $k$-subspaces in $\CC ^n$. As usual, let $T$ act on $Gr(k,n)$ by the diagonal action. The fixed points $v=(i_1,...,i_k)$ are then the $k$-planes generated by vectors $e_{i_1},...,e_{i_k}$. Denote by $\CC ^n\ti X$ the trivial vector bundle with the standard action. Then we may consider $L$, the determinant of the bundle dual to the $T$-equivariant universal $k$-subbundle of $\CC ^n\ti X$. Define $V=L^{\ot l}$, $l>0$. Denote by $p$ the equivariant class $c_1(L)$. We may identify $H^*(BT)$ with $\QQ [\ke _1,...,\ke _n]$ by the correspondence that $\ke _i$ is also denoted the equivariant Chern class of the line bundle over a point equipped with $T$ action as the representation of the character $\ke _i$. With respect to the Chern class of $L$, we shall write $d\in \ZZ=H_2(X, \ZZ )$. \subsection{A series} \subsubsection{Fixed points} Let $v$ be, say, $(1,2,..,k)$. Then around the point, a local chart can be described by \[\left( \begin{array}{cccc} 1 & 0 & ... & 0\\ 0 & 1 & & 0\\ 0 & 0 & ...& 1 \\ x_{1,1} & x_{1,2} & ...& x_{n-k,k} \\ & ... & \\ x_{n-k,1} & x_{n-k,2} & ... & x_{n-k,k} \end{array}\right).\] For each complex value $(x_{i,j})$ the column vectors in the matrix span a $k$-plane which stands for a point in $Gr(k,n)$. Then in the chart the action by $(t_1,...,t_n)\in T$ is described as follows: \[\left(\begin{array}{cccc} x_{1,1} & x_{1,2} & ...& x_{n-k,k} \\ & ... & \\ x_{n-k,1} & x_{n-k,2} & ... & x_{n-k,k} \end{array}\right) \mapsto \left(\begin{array}{cccc} t_1^{-1}t_{k+1} x_{1,1} & t_2 ^{-1}t_{k+1} x_{1,2} & ...& t_k^{-1}t_{k+1} x_{1,k}\\ & & ...& \\ t_1^{-1}t_n x_{n-k,1} & t_2 ^{-1}t_{n} x_{n-k,2} & ... & t_k^{-1}t_{n}x_{n-k,k} \end{array}\right).\] In each isolated fixed point of the Grassmannian there is $\dim Gr(k,n)$-many 1-dimensional orbit (ray) passing through the point. For instance, if $v=(1,2,...,k)$, then there is only one ray $(v,w)$ from $v$ to $w=(...,\hat{i},..., j)$ for any $i\le k <j$. These rays have degree $1\in H_2(X, \ZZ )$. \subsubsection{The Euler classes} Notice that the tangent space at $v=(1,2,...,k)$ of the ray connecting $v$ to $w=(...,\hat {i},...,j)$ has the character $\ka (v,w)=\ke _j - \ke _i$, where $j>k\ge i$. Similarly one can find out the characters for the other cases. Let $f:\PP ^1\ra Gr(k,n)$ be a $m$-fold morphism totally ramifying the ray over $v$ and $w$. The $T$ representation space $H^0(f^*L^{\ot l})$ has the orbi-characters \begin{eqnarray*} \frac {ap_v + b p_w}{m} = lp_v -\frac{\ke _j-\ke _i}{m}b, \text{ for } a+b=lm, \ a\ge 0,\ b \ge 0, \end{eqnarray*} where $p_v=-(\ke _1+...+\ke _k)$ and $p_w=-(\ke _1+...+\hat{\ke _i} +....+\ke _k + \ke _j)$ are $p=c_1(L)$ restricted to the fixed points $v$ and $w$, respectively. \section{The flag manifolds}\label{flag} We analyze fixed points of the maximal torus actions and the invariant curves connecting two fixed points. This explicit description would be useful also to find $S^X$ explicitly. \subsection{The complete flag manifolds} Let $X$ be the set of all Borel subgroups of a simply connected semi-simple Lie group $G$. It is a homogeneous space with the $G$-action by conjugation. Then the maximal torus $T$-action (---fix one---) has isolated fixed points. They are exactly Borel subgroups containing $T$. The fixed points are naturally one-to-one corresponding to the set of Weyl chambers. Each Borel subgroup containing $T$ gives rise to a negative roots (---our convention---) of $B$ and so a chamber associated to the positive roots. Let $C$ be the set of chambers. The tangent line subspace associated to the positive root $\ka$ has the character $\ka$. There is, if one fix a fixed point $v$, a natural correspondence between the $H^2(X,\ZZ )$ and the characters of $T$. Then the K\"ahler cone is exactly the positive Weyl chamber $v$. Notice that the fundamental roots span the K\"ahler cone. Consider co-roots $\ka ^\vee$. They span the Mori cone. We can identify the Mori cone $\Lambda$ with the non-negative integer span of co-roots. \subsection{The generalized flag manifolds} Let $X$ be the set of all parabolic subgroups with a given conjugate type. Let $T$ be a maximal torus of $G$. Then the fixed loci are isolated fixed points consisting of parabolic subgroups containing $T$. \subsubsection{Rays}\label{ray} Let us choose a fixed point $P\supset T$. Then the rays at the fixed points are described by the following way. (The rays are by definition the 1-dimensional orbits of $T$ passing through $P$.) Fix $B$ a Borel subgroup in $P$ containing $T$. First consider the $T$-equivariant fibration, $G/B \ra G/P$ and the rational map to $G/B$ by $\exp (zX_\ka )\in G$, $z\in \CC$, where $X_\ka$ is an eigenvector of the positive root $\ka$. Since $\exp H\exp (zX_\ka )\exp -H =\exp (z\exp\ad H (X_\ka ))=\exp (z\exp (\ka (H)) X_\ka )$ for $H\in \Lie T$, we conclude that it is a $T$ - invariant stable map. By the composition of the fibration, we obtain all the rays. They are effectively labeled by the positive roots which are not roots of $P$. So there are exactly $\dim X$-many rays at each fixed point. The tangent line at the ray has the character $\ka$. \subsubsection{The K\"ahler cone} Here we need the Levi-decomposition of $P$ and then consider simple roots $\{ \ka _i\} _{i\in P(\Delta )}$ which are not roots of the semi-simple part of $P$. Then the fundamental roots with respect to $P$ is, by definition, $\{ \kl _i \} _{i\in P(\Delta )}$, where $\kl _i$ are dual to $\ka _i^\vee $. Choose a fixed point $P$. We may identify $H^2(X,\ZZ )$ with the set of integral weights according to Borel-Weil theorem. Then the K\"ahler cone is the set of all dominant integral weights {\em with respect to} $P$. \subsubsection{Homogeneous line bundles} One can produces all very ample line bundles by homogeneous line bundles associated to irreducible representations of $P$ with highest weights $\kl$. The weights corresponding to the very ample line bundles are exactly the positive integral combination of the fundamental weights with respect to $P$. We shall denote by ${\mathcal O}(\kl )$ the homogeneous line bundle associated to the (1-dimensional) highest weight $\kl =\sum _{i\in P(\Delta )} a_i\kl _i$ representation of $P$. It is a very ample bundle if and only if $a_i >0$ for all $i$. This also shows that the ray $\PP ^1$ associated to $\ka$ has the homology class \lq\lq $\ka ^\vee$," in the sense that $<\PP ^1,c_1({\mathcal O}(\kl )>=(\ka ^\vee ,\kl )$. {\em We shall use $\ka ^\vee$ to denote the homology class.} \subsubsection{$\sum (p_i)_vz_i$ are different for different $v$.} Consider a line bundle $L$ associated to $\kl =\sum _{i\in P(\Delta )}a_i\kl _i$. (Here in advance, we have to fix $P\supset T$.) Let $S_{\ka}$ denote the Weyl group element of the reflection associated to the root $\ka$. Then the line bundle is $L={\mathcal O}(S_{\ka}(\kl ))$ if one look at it with respect to another \lq\lq origin " $P'=\exp (\frac{\pi}{2}(X_\ka -Y_\ka)) P \exp (-\frac{\pi}{2}(X_\ka -Y_\ka))$ where $[X_\ka ,Y_\ka]=H_\ka$, $[H_\ka ,X_\ka ]=2X_\ka$ and $[H_\ka ,Y_\ka]=-2Y_\ka$. This $P'$ is the other $T$-fixed point lies in the ray associated to $\ka$ which passes through $P$. (Because of the $SL(2,\CC )$-equivariant map from $\PP ^1$ to the ray, it is enough to check it when $G=SL(2,\CC )$, which is obvious.) \subsubsection{$V_{v,w,m}$ and $N_{v,w,m}$} Let $V={\mathcal O}(\kl )$. Let $\psi :\PP ^1\ra X$ be a stable map totally ramifying one of rays, passing through $P\supset T$. Suppose the ray is associated to a positive root $\ka $ with respect to $P$ and $f$ is a $m$-multiple branched cover representing an isolated $T$-fixed point of $\overline{M}_{0,0}(X,m\ka ^\vee)$. Then the $T$-representation space $H^0(\PP ^1, f^*({\mathcal O}(\kl ))$ has the characters, \[ \kl - a\frac{\ka}{m}, \ a=0,...,m(\kl ,\ka ^\vee). \] To see it, use the coordinate $z\in \CC$ around the fixed point and $\exp H\exp (zE_\ka )\exp -H =\exp (z\exp\ad H (E_\ka ))=\exp (z\exp (\ka (H)) E_\ka )$. Similarly, $N_{v,w,m}$ has characters \begin{eqnarray*} \delta -a\frac{\ka }{m} && \text{ for } \ka \ne \delta >0, a=0,...,m(\delta ,\ka ^\vee ), \\ \ka - a\frac{\ka }{m} && \text{ for } a=0,...,\hat{m},...,2m, \end{eqnarray*} where $\delta >0$ means $\delta$ is a positive root with respect to $P$.
1997-12-29T21:48:21
9712
alg-geom/9712030
en
https://arxiv.org/abs/alg-geom/9712030
[ "alg-geom", "math.AG" ]
alg-geom/9712030
Terrence Napier
T. Napier, M. Ramachandran
The L^2 dbar method, weak Lefschetz theorems, and the topology of Kahler manifolds
30 pages
null
null
null
null
A new approach to Nori's weak Lefschetz theorem is described. The new approach, which involves the dbar-method, avoids moving arguments and gives much stronger results. In particular, it is proved that if X and Y are connected smooth projective varieties of positive dimension and f is a holomorphic immersion of Y into X with ample normal bundle, then the image of the fundamental group of Y in that of X is of finite index. This result is obtained as a consequence of a direct generalization of Nori's theorem. The second part concerns a new approach to the theorem of Burns which states that a quotient of the unit ball in complex Euclidean space (of dimension at least 3) by a discrete group of automorphisms which has a strongly pseudoconvex boundary component has only finitely many ends. The following generalization is obtained. If a complete Hermitian manifold X of dimension at least 3 has a strongly pseudoconvex end E and the Ricci curvature of X is bounded above by a negative constant, then, away from E, X has finite volume.
[ { "version": "v1", "created": "Mon, 29 Dec 1997 20:48:21 GMT" } ]
2007-05-23T00:00:00
[ [ "Napier", "T.", "" ], [ "Ramachandran", "M.", "" ] ]
alg-geom
\section{Introduction} \label{intro} In~[No], Nori studied the fundamental group of complements of nodal curves with ample normal bundle in smooth projective surfaces. The main tool was the following weak Lefschetz theorem: \begin{norithm} Suppose $\Phi : U @>>> X$ is a local biholomorphism from a connected complex manifold~$U$ into a connected smooth projective variety~$X$ of dimension at least~$2$ and $U$ contains a connected effective divisor~$Y$ with compact support and ample normal bundle. Then, for every Zariski open subset~$Z$ of~$X$, the image of $\pi _1(\Phi ^{-1} (Z))$ in $\pi _1(Z)$ is of finite index. \end{norithm} For $X$ a surface, he obtained sharp bounds for the index using the Hodge index theorem. A striking corollary of this result is the following: \begin{noricor} If $X$ and $Y$ are connected smooth projective varieties with $$ \dim X=\dim Y +1>1 $$ and $f : Y @>>> X$ is a holomorphic immersion with ample normal bundle, then the image of $\pi_1(Y)$ in $\pi _1(X)$ is of finite index. \end{noricor} Nori's proof of these results depends heavily on deformations. The first step is to show that a large multiple of the divisor~$Y$ in the theorem moves in a family in which the general member is irreducible and meets~$Y$ and the union of these members contains an open subset of~$U$. Unfortunately, moving arguments do not seem to apply in the higher codimensional case, because Fulton and Lazarsfeld~[FL2] have observed that for a certain smooth projective $4$-fold and a smooth surface~$Y$ in~$X$ with ample normal bundle constructed by Gieseker, no multiple of~$Y$ in~$X$ moves. Given the existence of a sufficiently large number of deformations, the rest of the proof of Nori's weak Lefschetz theorem has been streamlined by Campana~[C1] and Koll\'ar~[K]. In~[NR], another proof of Nori's theorem was given when $Z=X$ using harmonic functions , but it was the same in spirit as the earlier arguments. A survey on Lefschetz type theorems can be found in Fulton~[F]. In this paper we introduce a new approach which avoids moving arguments and which gives much stronger results. In particular, the new approach allows one to address the case of higher codimension. Before giving precise statements, we recall some terminology. Let $Y$ be a complex analytic subspace of complex space~$U$. We denote the structure sheaf of~$U$ by ${\cal O} _U$ and the ideal sheaf of~$Y$ in~$U$ by~$\cal I_Y$. The {\it formal completion}~$\widehat U $~{\it of}~$U$ {\it with respect to }~$Y$ is the ringed space $$ (\widehat U , {\cal O} _{\widehat U })=(Y, \lim _{ @<<< }{\cal O} _U/\cal I_Y^n). $$ If $\cal F$ is an analytic sheaf on~$U$ we denote by $\widehat {\cal F}$ the associated analytic sheaf on~$\widehat U $ given by $$ \widehat {\cal F} = \lim _{ @<<< }(\cal F \otimes {\cal O} _U/\cal I_Y^n). $$ If $\cal F$ is coherent, then $\widehat {\cal F}$ is also coherent over~${\cal O} _{\widehat U }$. The main result is the following generalization of Nori's weak Lefschetz theorem: \begin{thm} \it Suppose $\Phi : U @>>> X$ is a holomorphic map from a connected complex manifold~$U$ into a connected smooth projective variety~$X$ of dimension at least~$2$ which is a submersion at some point. Let~$Y\subset U$ be a connected compact analytic subspace such that $\dim H^0(\widehat U , \widehat {\cal L}) <\infty $ for every locally free analytic sheaf~$\cal L$ on~$U$. Then, for every Zariski open subset~$Z$ of~$X$, the image of $\pi _1(\Phi ^{-1} (Z))$ in $\pi _1(Z)$ is of finite index. \end{thm} \begin{rems} 1. For example, by a theorem of Hartshorne~[H] (and Grothendieck~[Gr]), $H^0(\widehat U , \widehat {\cal L})$ is finite dimensional when $Y$ is a connected compact analytic subspace which is locally a complete intersection and which has ample normal bundle (or even $k$-ample normal bundle in the sense of Sommese~[So] where $k=\dim Y-1$). \noindent 2. Theorem~0.1 also holds for $U$ irreducible and reduced and $X$ normal and projective. Moreover, as will be shown in Sect.~3 (Corollary~3.4), in the smooth case one only needs finite dimensionality for $\cal L$ the analytic pullback of an invertible sheaf on~$X$. \noindent 3. As a consequence of Theorem~0.1, one can remove the dimension restriction on the subspace~$Y$ in the corollary to Nori's theorem. More precisely, we get the following: \end{rems} \begin{cor} If $X$ and $Y$ are connected smooth projective varieties of positive dimension and $f : Y @>>> X$ is a holomorphic immersion with ample normal bundle, then the image of $\pi_1(Y)$ in $\pi _1(X)$ is of finite index. \end{cor} Hironaka and Matsumura~[HM] proved the analogous result for algebraic fundamental groups when $f$ is an inclusion with ample normal bundle. However, the result for topological fundamental groups (as stated in the above corollary) is new (provided $\dim X > \dim Y +1$) even for $f$ an inclusion. Moreover, simple examples show that, if $\dim X > \dim Y +1$, then, even if $f$ is an inclusion (with ample normal bundle), the map $\pi_1(Y) @>>> \pi _1(X)$ is not necessarily surjective. The idea of the proof of Theorem~0.1 is to form a covering space~$\widetilde Z @>>> Z$ with fundamental group equal to the image~$G$ of $\pi _1(\Phi ^{-1} (Z))$ and then construct $L^2$ holomorphic sections of a suitable line bundle which separate the sheets of the covering. This construction is a standard application of the $L^2$~$\bar \partial $-method (Andreotti-Vesentini~[AV], H\"ormander~[Ho], Skoda~[Sk], Demailly~[D1]). Pulling these sections back to $\Phi ^{-1} (Z)$ by a lifting of~$\Phi $, the finite dimensionality of the space of holomorphic sections on the formal completion gives a bound on the dimension of the space of sections on~$\widetilde Z $ and hence a bound on the degree of the covering space (i.e.~on the index of~$G$). \begin{rem} Campana~[C2] has independently applied $L^2$-methods to study exceptional curves on coverings of surfaces. \end{rem} The second main result of this paper generalizes a theorem of Burns~[B] which states that a quotient of the unit ball in~$\C ^n$ ($n\geq 3$) by a discrete group of automorphisms which has a strongly pseudoconvex boundary component has only finitely many ends. The main tools are a theorem of Lempert on the compactification of a pseudoconvex boundary from the pseudoconcave side~[L], a finiteness theorem of Andreotti for pseudoconcave manifolds~[A], and the $L^2$ Riemann-Roch inequality of Nadel and Tsuji~[NT]. The precise statement is as follows: \begin{thm} If a complete Hermitian manifold~$X$ of (complex) dimension at least~$3$ has a strongly pseudoconvex end and $\text {Ricci}\, (X) \leq -C$ for some positive constant~$C$, then, away from the strongly pseudoconvex end, the manifold has finite volume. \end{thm} As in the proof of Theorem~0.1, the idea is to apply finite dimensionality of the space of holomorphic sections of a line bundle. By Lempert's theorem, one can cap off the strongly pseudoconvex end by a domain in a smooth projective variety. Andreotti's finiteness theorem applied to the resulting pseudoconcave manifold gives finite dimensionality of the space of holomorphic sections of a suitable line bundle. Finally, the $L^2$ Riemann-Roch inequality of Nadel and Tsuji gives a (finite) upper bound for the volume in terms of the dimension of this space of sections. \begin{rem} One natural question which arises is might there be an improved version of the $L^2$ Riemann-Roch inequality which would give improved bounds for the volume in Theorem~0.3 as well as the index in~Theorem~0.1? Also, for $X$ a surface in the corollary to Nori's weak Lefschetz theorem, Nori~[No] found bounds for the index in terms of certain intersection numbers. It is therefore natural to look for analogous bounds in more general cases. \end{rem} Sect.~1 begins with a proof of Theorem~0.1 in the case where~$\Phi $ is a local biholomorphism. The main idea of the new approach is easy to see in this context, and, although a few technicalities arise in the general case, the proof is essentially the same. The proof of Theorem~0.1 is then given. The required result from the $L^2$ \ $\bar \partial $-method is discussed in Sect.~2. Further generalizations of the weak Lefschetz theorem for $X$ not necessarily projective are considered in Sect.~3. Theorem~0.3 is proved in Sect.~4, which may be read independently of Sects.~1--3. \noindent {\it Acknowledgements}. Madhav Nori suggested we reformulate Theorem~0.1 in terms of formal completions, which considerably widened its scope. Charles Epstein told us about Lempert's result. For this and other useful advice, we would like to thank them both. We would also like to thank Alan Nadel for bringing the $L^2$ Riemann-Roch inequality to our attention, Dan Burns for useful discussions on his theorem, and Raghavan Narasimhan for his interest in this work. Finally, we would like to thank the referee for helpful suggestions. \section{Weak Lefschetz theorems for a projective variety} This section contains the proof Theorem~0.1. We first prove the theorem for the case of a local biholomorphism. This is a direct generalization to immersed complex spaces of {\it arbitrary} codimension (Nori proved the theorem stated below for $Y$ an ample divisor in~$U$). More general versions will be stated later. Aside from a few minor technical problems, however, the proofs of all of the generalizations are the same in spirit as the proof of this special case. \begin{thm} Let $U$ be a connected complex manifold, let~$X$ be a connected smooth projective variety of dimension~$n>1$, let $\Phi : U @>>> X$ be a holomorphic map, let~$Y$ be a connected compact analytic subspace (not necessarily reduced) of~$U$, and let $\widehat U $ be the formal completion of~$U$ with respect to~$Y$. Assume that \begin{enumerate} \item[(i)] $\Phi $ is locally biholomorphic , and \item[(ii)] $\dim H^0(\widehat U , \widehat {{\cal O} (\Phi ^*L)})<\infty $ for every holomorphic line bundle~$L$ on~$X$. \end{enumerate} Then there is a positive constant~$b$ depending only on the mapping $\Phi :U @>>> X$ and the subspace~$Y\subset U$ such that , if $R\subset X$ is a nowhere dense analytic subset of~$X$ and~$V$ is a connected neighborhood of~$Y$ in~$U$, then the image~$G$ of $\pi _1(V\setminus \Phi ^{-1} (R)) @>>> \pi _1(X\setminus R)$ is of index at most~$b$ in~$\pi _1(X\setminus R)$. Moreover, if $\Phi (Y)\cap R=\emptyset $, then the image of $\pi _1(Y) @>>> \pi _1(X\setminus R)$ is also of index at most~$b$. \end{thm} \begin{pf} Given $R$,$V$, and~$G$ as in the statement of the theorem, let $S=\Phi ^{-1} (R)$, let $M=X\setminus R$, let $W=V\setminus S$, and let $\pi : \widetilde M @>>> M$ be a connected covering space with $\pi _* (\pi _1(\widetilde M ))=G$. Thus $\pi : \widetilde M @>>> M$ has degree $d=[\pi _1(M):G]$ and we have the following commutative diagram of holomorphic mappings: \begin{center}\begin{picture}(250,80) \put(5,10){$W=V\setminus S\subset V $} \put(90,14){\vector(1,0){50}} \put(145,10){$X \supset X\setminus R=M$} \put(110,3){$\Phi $} \put(125,55){\vector(3,-1){90}} \put(110,60){$\widetilde M $} \put(17,25){\vector(3,1){90}} \put(57,45){$\tilde \Phi $} \put(170,45){$\pi $} \end{picture} \end{center} Since $X$ is projective, there exists a Hermitian holomorphic line bundle~$(L,h)$ with positive curvature and a K\"ahler metric~$g$ on~$X$. As will be shown in Sect.~2 (see Corollary~2.3), the $L^2$~$\bar \partial $-method, in the form given by Skoda~[Sk] and Demailly~[D1], enables one to prove that there is a positive integer~$\nu $ independent of~$R$~and~$V$ such that $$ d\leq \dim H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (L^{\nu }\otimes K_M))); $$ where $K_M$ is the canonical bundle on~$M$ and $H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (L^{\nu }\otimes K_M)))$ is the space of holomorphic sections of $\pi ^* (L^{\nu }\otimes K_M)$ which are in~$L^2$ with respect to the Hermitian metrics $\pi ^*(h\otimes g^*)$ on $\pi ^* (L^{\nu }\otimes K_M)$ and~$\pi ^* g$ on~$\widetilde M $. If $s\in H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (L^{\nu }\otimes K_M)))$, then $\tilde \Phi ^* s$ is a holomorphic section of $\Phi ^* (L^{\nu }\otimes K_X)$ on $W$. Given a point $x_0\in S\cap V$, $\Phi $ maps a neighborhood~$Q$ of~$x_0$ in~$V$ biholomorphically onto $\Phi (Q) \subset X$. Hence $\tilde \Phi $ maps $Q\setminus S$ biholomorphically onto its image in~$\widetilde M $ and, therefore, $\tilde \Phi ^*s$ is in~$L^2$ on $Q\setminus S$ with respect to the Hermitian metrics $\Phi ^*(h\otimes g^*)$ in $\Phi ^* (L^{\nu }\otimes K_X)$ and $\Phi ^*g$ on~$U$. Since these metrics are defined over the entire set~$U$ and a square integrable function which is holomorphic outside a nowhere dense analytic set in a manifold extends holomorphically past the analytic set, $\tilde \Phi ^*s$ extends to a holomorphic section of $\Phi ^* (L^{\nu }\otimes K_X)$ on~$V$. Therefore $$ d\leq \dim H^0(V, {\cal O} (\Phi ^* (L^{\nu }\otimes K_X))). $$ On the other hand, by a general fact about formal completions, if $\cal F$ is a coherent analytic sheaf on~$V$, then the kernel of the mapping $$ H^0(V,\cal F) @>>> H^0(\widehat V ,\widehat {\cal F})=H^0(\widehat U ,\widehat {\cal F}) $$ consists of all of the sections of~$\cal F$ on~$V$ which vanish on a neighborhood of~$Y$ in~$V$ (see [BS, Proposition VI.2.7]). In particular, if~$\cal F$ is locally free, then this mapping is injective. Therefore, taking $\cal F={\cal O} (\Phi ^*(L^{\nu }\otimes K_X))$, we get $$ d\leq \dim H^0(V,\cal F) \leq \dim H^0(\widehat U ,\widehat {\cal F})<\infty . $$ Thus $b=\dim H^0(\widehat U ,\widehat {\cal F})$ is a uniform bound for~$d$ independent of~$R$ and~$V$. Finally, if $\Phi (Y)\cap R=\emptyset $, then we may choose the neighborhood~$V$ so that $V\subset U\setminus S$ and the map $\pi _1(Y) @>>> \pi _1(V)$ is a surjective isomorphism. Hence the image of $\pi _1(Y)$ in $\pi _1(M)$ is equal to the image of $\pi _1(V)=\pi _1(V\setminus S)$ and therefore is of index at most~$b$. \end{pf} We now consider generalizations. If in the above theorem one assumes only that~$\Phi $ is a generic submersion (or a generic local biholomorphism), then a slight technical problem arises. While (as one may easily check) the section $\tilde \Phi ^*s$ of $\Phi ^*(L^\nu \otimes K_X)$ extends holomorphically past~$S$ near points at which~$\Phi $ is submersive, $\tilde \Phi ^*s$ need not extend near points where $\text {rank} \, \Phi _* < n$. However, as we will see, $\tilde \Phi ^*s$ does extend as a holomorphic $n$-form with values in~$\Phi ^*L^\nu $. A simple illustration is given by $$ U=\Delta \ni z \overset {\Phi } \mapsto \zeta =z^2 \in \Delta , \quad \widetilde M =\Delta ^* \ni z \overset {\pi } \mapsto \zeta =z^2 \in \Delta ^*=M, \quad \text {and } s=z ^{-1} \pi ^* d\zeta . $$ Here, $\tilde \Phi ^*s$ does not extend as a section of the pullback of the canonical bundle, but the corresponding holomorphic $1$-form~$2dz$ does extend. In fact, by passing to desingularizations, one also gets this extension property for $U$ and $X$ singular. Given an irreducible reduced complex space~$A$ and a positive integer~$n$, we denote by~$\Omega ^n_A$ the coherent analytic sheaf on~$A$ obtained by forming a desingularization $\check A @>>> A$ of~$A$ and taking the direct image of $\Omega ^n_{\check A}$. By the following lemma, this sheaf is independent of the choice of the desingularization. \begin{lem}[Grauert and Riemenschneider [GR, Sect.~2.1{]}] Let $A$ be an irreducible reduced complex space of dimension~$m$ and let~$n$ be a positive integer. Suppose that, for $i=1,2$, $B_i$ is a connected complex manifold of dimension~$m$ and $\Psi _i : B_i @>>> A$ is a proper modification. Then $(\Psi _1)_*\Omega ^n_{B_1}=(\Psi _2)_*\Omega ^n_{B_2}$. \end{lem} The proof is similar to the proof for $\dim A=n$ given in~[GR]. The main point is that if $A$ is smooth, then $\Psi _1$ is biholomorphic outside an analytic set of codimension at least~$2$ in~$A$. For the general case, one passes to a common proper modification of $B_1$~and~$B_2$. We may now state the extension property as follows: \begin{lem} \it Let $\Phi :U @>>> X$ be a holomorphic mapping of irreducible reduced complex spaces $U$~and~$X$ of dimensions $m$~and~$n$, respectively, such that $\Phi (U)$ has nonempty interior. Suppose \begin{center}\begin{picture}(250,80) \put(5,10){$W=U\setminus S\subset U$} \put(90,14){\vector(1,0){50}} \put(145,10){$X \supset X\setminus R=M$} \put(110,3){$\Phi $} \put(125,55){\vector(3,-1){90}} \put(110,60){$\widetilde M $} \put(17,25){\vector(3,1){90}} \put(57,45){$\tilde \Phi $} \put(170,45){$\pi $} \end{picture} \end{center} is a commutative diagram of holomorphic mappings where $R\subset X$ is a nowhere dense analytic subset which contains~$\sing X$, $S=\Phi ^{-1} (R)$, and $\pi : \widetilde M @>>> M$ is a connected holomorphic covering space. Let $L$ be a holomorphic line bundle on~$X$ and let $\theta $ be a holomorphic $n$-form with values in $\pi^*L$ on $\widetilde M $ which is in $L^2$ with respect to the liftings of a Hermitian metric~$h$ in~$L$ on~$X$ and a Hermitian metric~$g$ on~$M$. Then the pullback $(\tilde \Phi | _{\reg W})^*\theta$ of~$\theta$ to a holomorphic $n$-form with values in $\Phi ^*L$ on $\reg W$ extends to a (unique) section in $H^0(U, {\cal O} (\Phi ^*L) \otimes \Omega ^n_U)$. \end{lem} The proof uses standard methods but will be postponed until the end of this section (see also Sakai~[S]). We may now apply the argument given in the proof of Theorem~1.1 to get Theorem~0.1 of the introduction. In fact, we get the following: \begin{thm} Let $U$ be an irreducible reduced complex space, let $X$ be a connected normal projective variety of dimension~$n>1$, let $\Phi :U @>>> X$ be a holomorphic map, let~$Y$ be a connected compact analytic subspace (not necessarily reduced) of~$U$, and let $\widehat U $ be the formal completion of~$U$ with respect to~$Y$. Assume that \begin{enumerate} \item[(i)] $\Phi (U)$ has nonempty interior, and \item[(ii)] $\dim H^0(\widehat U , \widehat {{\cal O} (\Phi ^*L)}\otimes \widehat {\Omega ^n_U}) <\infty $ for every holomorphic line bundle~$L$ on~$X$. \end{enumerate} Then there exists a positive constant~$b$ depending only on the mapping $\Phi : U @>>> X$ and the subspace~$Y$ such that , if~$R\subset X$ is a nowhere dense analytic subset and $V$ is a connected neighborhood of~$Y$ in~$U$, then the image of $\pi _1(V\setminus \Phi ^{-1} (R)) @>>> \pi _1 (X\setminus R)$ is of index at most~$b$. Moreover, if $\Phi (Y)\cap R=\emptyset $, then the image of~$\pi _1(Y)$ in $\pi _1(X\setminus R)$ is also of index at most~$b$. \end{thm} \begin{pf} Given $R$ and $V$ as in the statement of the theorem, we get a commutative diagram \begin{center}\begin{picture}(250,80) \put(5,10){$W=V\setminus S\subset V$} \put(90,14){\vector(1,0){50}} \put(145,10){$X \supset X\setminus R=M$} \put(110,3){$\Phi $} \put(125,55){\vector(3,-1){90}} \put(110,60){$\widetilde M $} \put(17,25){\vector(3,1){90}} \put(57,45){$\tilde \Phi $} \put(170,45){$\pi $} \end{picture} \end{center} as in the proof of Theorem~1.1. Since $X$ is normal, the map $\pi _1(M\setminus \sing X) @>>> \pi _1(M)$ is surjective. Therefore, by replacing~$R$ by $R\cup\sing X$, we may assume that $\sing X\subset R$; i.e.~that $M$ is a complete K\"ahler manifold. If $s\in H^0_{L^2}(\widetilde M , {\cal O} ((\pi ^*L^\nu )\otimes K_{\widetilde M }))$ for some~$\nu $ (with respect to metrics lifted from the base), then, by Lemma~1.3, the pullback to $\reg W$ {\it as a holomorphic $n$-form with values in~$\Phi ^*L^\nu $} extends to a unique section in $H^0(V, {\cal O} (\Phi ^*L^\nu ) \otimes \Omega ^n_U)$. By applying Corollary~2.3 as in the proof of Theorem~1.1, one now gets the required bound on the index. \end{pf} A finiteness theorem of Hartshorne~[H, Theorem~III.4.1] and Grothendieck~[Gr] and the above theorem together imply immediately that, in Nori's weak Lefschetz theorem, one may take the mapping to be a generic submersion and the subvariety to be of arbitrary codimension. More precisely, we have the following: \begin{cor} Let $U$ be a connected complex manifold, let~$X$ be a connected normal projective variety of dimension~$n>1$, let $\Phi : U @>>> X$ be a holomorphic map, and let~$Y$ be a positive dimensional connected compact analytic subspace (not necessarily reduced) of~$U$. Assume that \begin{enumerate} \item[(i)] $\Phi (U)$ has nonempty interior, \item[(ii)] $Y$ is locally a complete intersection in~$U$, and \item[(iii)] The normal bundle $N_{Y/U}$ is ample. \end{enumerate} Then there is a positive constant~$b$ depending only on the mapping $\Phi :U @>>> X$ and the subspace~$Y\subset U$ such that , if~$Z$ is a nonempty Zariski open subset of~$X$ and $V$ is a connected neighborhood of~$Y$ in~$U$, then the image of $\pi _1(V\cap \Phi ^{-1} (Z)) @>>> \pi _1 (Z)$ is of index at most~$b$ in~$\pi _1(Z)$. Moreover, if $\Phi (Y)\subset Z$, then the image of~$\pi _1(Y)$ in $\pi _1(Z)$ is also of index at most~$b$. \end{cor} \begin{rems} 1. The approach of considering sections of vector bundles on formal completions fits well with Grothendieck's approach to the Lefschetz theorems~[Gr] (see also~[H]). In a sense, the results of this paper extend to the topological fundamental group Grothendieck's Lefschetz theorems concerning the algebraic fundamental group. \noindent 2. Further generalizations in which $X$ is not necessarily projective will be stated and proved in Sect.~3. A slightly more precise bound for the index in terms of the dimension of a space of sections will also be obtained. \end{rems} We conclude this section with the proof of the extension property. \begin{pf*}{Proof of Lemma~1.3} We first observe that we may assume that $U$~and~$X$ are smooth and that $R$ is a divisor with normal crossings by passing to desingularizations. More precisely, we may form a commutative diagram \begin{center}\begin{picture}(250,90) \put(60,10){$U$} \put(90,14){\vector(1,0){50}} \put(150,10){$X$} \put(115,0){$\Phi $} \put(65,55){\vector(0,-1){30}} \put(50,60){$U\times _XX'$} \put(45,35){$\text{pr}_U$} \put(100,64){\vector(1,0){40}} \put(150,60){$X'$} \put(115,75){$\text {pr}_{X'}$} \put(160,35){$\beta $} \put(155,55){\vector(0,-1){30}} \end{picture} \end{center} where $X'$ is a connected complex manifold, $R'=\beta ^{-1} (R)$ of $R$ is a divisor with normal crossings, and $\beta : X' @>>> X$ is a proper modification which maps $M'=X'\setminus R'$ biholomorphically onto $M=X\setminus R$. Since $\text {pr}_U ^{-1} (W)=W\times _MM'$ is just the graph of the restriction of~$\Phi $ to a mapping $W @>>> M'=M$ and $U$ is irreducible , $\text {pr}_U ^{-1} (W)$ is an open irreducible subset of $U\times _XX'$ which is mapped isomorphically onto~$W$. In particular, $\text {pr}_U ^{-1} (W)$ lies in a unique irreducible component~$C$ of $U\times _XX'$; and, since $\text {pr}_U$ is a proper mapping, we must have $\text {pr}_U(C)=U$. Passing to a desingularization of~$C$, we get a commutative diagram of holomorphic mappings \begin{center}\begin{picture}(250,90) \put(70,10){$U$} \put(95,14){\vector(1,0){50}} \put(155,10){$X$} \put(115,0){$\Phi $} \put(75,55){\vector(0,-1){30}} \put(70,60){$U'$} \put(60,35){$\alpha $} \put(95,64){\vector(1,0){50}} \put(155,60){$X'$} \put(115,72){$\Phi '$} \put(165,35){$\beta $} \put(160,55){\vector(0,-1){30}} \end{picture} \end{center} where $U'$ is a connected complex manifold of dimension~$m$, $\alpha : U' @>>> U$ is a proper modification, $S'\equiv \alpha ^{-1} (S)=(\Phi ') ^{-1} (R')$, and, if $W'=U'\setminus S'=\alpha ^{-1} (W)$, then $\alpha $ maps the set $W'\setminus \alpha ^{-1} (\sing U)$ biholomorphically onto~$\reg W$. We also get a connected covering space $\pi '=(\beta |_{M'}) ^{-1} \circ \pi : \widetilde M @>>> M'$ and a lifting $\tilde \Phi ' = \tilde \Phi \circ (\alpha | _{W'}) : W' @>>> \widetilde M $ of $\Phi '| _{W'}$. Therefore, if $L'=\beta ^*L$, then $\theta $ is a holomorphic $n$-form with values in $\pi^*L=(\pi ')^*L'$ which is in~$L^2$ with respect to the metrics $\pi ^*h=(\pi ')^*\beta ^*h$ in $(\pi ')^*L'$ and $\pi ^*g=(\pi ')^*\beta ^*g$ on~$\widetilde M $. Suppose the pullback of $\theta $ to~$W'$ extends to a section $$ \eta \in H^0(U', {\cal O} ((\Phi')^*L')\otimes \Omega ^n_{U'}). $$ Since $(\Phi')^*L'=\alpha ^*\Phi ^*L$ and $\alpha :U' @>>> U$ is a proper modification, we have (by the definition of $\Omega ^n_U$ and Lemma~1.2) $$ \alpha _*\bigl( {\cal O} ((\Phi')^*L')\otimes \Omega ^n_{U'}\bigr) ={\cal O} (\Phi^*L)\otimes \Omega ^n_{U}. $$ Hence $\eta $ determines an extension of $(\tilde \Phi | _{\reg W})^*\theta $ to a section in $H^0(U, {\cal O} (\Phi^*L)\otimes \Omega ^n_{U})$. Thus we may assume that $U$~and~$X$ are smooth and that $R$ is a divisor with normal crossings in~$X$. In particular, $S=\Phi ^{-1} (R)$ is a divisor in~$U$. Since the lemma is entirely local, it suffices to extend the section near each point $x_0\in S$ and we may assume that $U=\Delta ^m$ is the unit polydisk centered at~$x_0=0$ in~$\C ^m$, that $X=\Delta ^n$ is the unit polydisk centered at~$\Phi (x_0)=0$ in~$\C ^n$, that $L$ is the trivial line bundle with the trivial metric on~$X$ (since all metrics are comparable on relatively compact subsets), and that $g$ is the restriction of the Euclidean metric $g_{\C ^n}$ to~$M$ (since the $L^2$ condition on forms of type $(n,0)$ is independent of the choice of the metric on an $n$-dimensional manifold). We denote the coordinates in~$\C ^m$ by $z=(z_1,\dots , z_m)$, the coordinates in~$\C ^n$ by $\zeta =(\zeta _1,\dots , \zeta _n)$, and the coordinate functions of the mapping by $\Phi =(\Phi _1, \dots , \Phi _n)$. Thus $\theta =fd(\zeta _1\circ \pi ) \wedge \dots \wedge d(\zeta _n\circ \pi )$ for some holomorphic function~$f$ which is square integrable on~$\widetilde M $ with respect to $\pi ^*g_{\C ^n}$ and $\tilde \Phi ^*\theta =(f\circ \tilde \Phi )d\Phi _1\wedge \dots \wedge d\Phi _n$ on $W$. Since holomorphic sections extend past analytic sets of codimension at least~$2$, we may assume that~$x_0\in \reg S$ and hence that~$S$ is the zero set of~$z_1$. Since $R$ is a divisor with normal crossings, we may also assume that $R$ is the zero set of $\zeta _1\cdots \zeta _k$. Finally, if $\setof {x\in S}{\Phi _j(x)=0}$ is nowhere dense in~$S$ for some~$j$, then, again, it suffices to consider a point~$x_0$ which avoids this zero set. Thus we may assume that $$ S=\Phi ^{-1} (R)=\{ \Phi _j=0 \} \text { for } j=1, \cdots , k. $$ We now show that $(\Phi _1\cdots \Phi _k) \cdot (f\circ \tilde \Phi )$ extends to a holomorphic function which vanishes along $S$. If~$x=(x_1,\dots ,x_m)$ is a point in~$W=U\setminus S=\Delta ^*\times \Delta ^{m-1}$ near~$x_0=0$ and $y=\Phi (x)=(y_1, \dots ,y_n)$, then we have $r_j=|y_j|< 1/2$ for $j=1,\dots , n$ and $r_j>0$ for $j=1, \dots ,k$. Thus the polydisk $$ P =\Delta (y_1;r_1) \times \dots \times \Delta (y_k;r_k) \times \Delta (y_{k+1};1/2) \times \Delta (y_n;1/2) $$ centered at~$y$ is contained in $M=(\Delta ^*)^k\times \Delta ^{n-k}$ and is therefore evenly covered by $\pi : \widetilde M @>>> M$. Hence~$\pi $ maps the connected component~$\widetilde P $ of~$\pi ^{-1} (P)$ containing $\tilde y =\tilde \Phi (x)$ isomorphically onto~$P$. The $L^{\infty }/L^2$-estimate now gives $$ |f(\tilde y )|^2\leq (\text {\rm vol} \, (\widetilde P )) ^{-1} \int _{\widetilde P } |f|^2 \, dV_{\pi ^* g_{\Cn }}. $$ As $x$ approaches a point~$x_1$ in~$S$ near~$x_0$, $\text {\rm vol} \, (\widetilde P )=\text {\rm vol} \, (P)$ will approach~$0$. Therefore, after multiplying both sides of the above inequality by $(r_1\cdots r_k )^2$ we get, since $|f|^2$ is integrable on~$\widetilde M $, $$ |\Phi _1(x)\cdots \Phi _k(x) f(\tilde \Phi (x))|^2=(r_1\cdots r_k)^2|f(\tilde y )|^2 \leq \pi ^{-n}4^{(n-k)}\int _{\widetilde P } |f|^2 \, dV_{\pi ^* g_{\Cn }} @>>> 0 $$ and the claim follows. For each $j=1,\dots ,k$, we have $\Phi _j=z_1^{\mu _j}h_j$ where $\mu _j =\text {ord}_S\Phi _j$ and $h_j$ is a unit. Therefore, setting $\mu =\mu _1+\dots + \mu _k $ and $\psi =d\Phi _{k+1}\wedge \dots \wedge d\Phi _n$, we get \begin{align*} d\Phi _1\wedge \dots \wedge d\Phi _n &=z_1^\mu dh_1\wedge \dots \wedge dh_k \wedge \psi \\ &\quad+z_1^{\mu -1} \sum _{j=1}^k \mu _j h_j (-1)^{j-1}dz_1 \wedge dh_1\wedge \dots \wedge \widehat {dh_j}\wedge \dots \wedge dh_k \wedge \psi . \end{align*} Since $z_1^{\mu }(f\circ \tilde \Phi )$ extends to a holomorphic function which vanishes along~$S$, it follows that the $n$-form $\tilde \Phi ^*\theta =(f\circ \tilde \Phi )d\Phi _1\wedge \dots \wedge d\Phi _n$ also extends holomorphically as claimed. \end{pf*} \section{Results from the $L^2$~$\bar \partial $-method} As described in Sect.~1, the proofs of the weak Lefschetz theorems rely on a consequence of the $L^2$~$\bar \partial $-method (Andreotti-Vesentini~[AV], H\"ormander~[Ho], Skoda~[Sk], Demailly~[D1]) which will be described in this section. Given a real-valued function~$\varphi $ of class~$C^2$ on a complex manifold~$M$ of dimension~$n$, the {\it Levi form} $\lev \varphi $ of~$\varphi $ is the Hermitian tensor defined by $$ \lev \varphi = \sum _{i,j=1}^n \frac {\partial ^2\varphi }{\partial z_i\partial \bar z_j} dz_id\bar z_j $$ in local holomorphic coordinates $(z_1,\dots ,z_n)$. The function $\varphi $ is said to be plurisubharmonic if $\lev \varphi \geq 0$ and strictly plurisubharmonic if $\lev \varphi > 0$. If $(L,h)$ is a Hermitian holomorphic line bundle on a complex manifold $M$, then the {\it curvature tensor}~$\cal C (L,h)$ of~$(L,h)$ is given by $$ \cal C(L,h)= \lev {-\log |s|^2} $$ for any nonvanishing local holomorphic section~$s$ of~$L$. We will need the following special case of a theorem of Demailly~[D1, Theorem~5.1] concerning the $\bar \partial $-method for singular metrics with semi-positive curvature. \begin{thm}[Demailly] \it Let $(E,h)$ be a Hermitian holomorphic line bundle with semi-positive curvature (i.e.~$\cal C(L,h)\geq 0$) on a complete K\"ahler manifold~$(M,g)$ of dimension~$n$. Suppose $\varphi : M @>>> [-\infty ,0]$ is a function which is of class~$C^{\infty } $ outside a discrete subset~$S$ of~$M$ and, near each point $p\in S$, $\varphi (z)=A_p\log |z|^2$ where $A_p$ is a positive constant and $z=(z_1,\dots ,z_n)$ are local holomorphic coordinates centered at~$p$. Assume that $\cal C (E,he^{-\varphi })=\cal C (E,h)+\lev {\varphi }\geq 0$ on~$M\setminus S$ (and hence on~$M$ as the curvature of a singular metric) and let $\lambda : M @>>> [0,1]$ be a continuous function such that $\cal C (E,h)+\lev {\varphi }\geq \lambda g$ on~$M\setminus S$. Then, for every $C^{\infty } $~form~$\theta $ of type~$(n,1)$ with values in~$E$ on~$M$ which satisfies $$ \bar \partial \theta =0 \quad \text {and} \quad \int _M\lambda ^{-1} |\theta |^2_{h\otimes g^*}e^{-\varphi } \, dV_g <\infty, $$ there exists a $C^{\infty } $ form~$\eta $ of type~$(n,0)$ with values in~$L$ on~$M$ such that $$ \bar \partial \eta =\theta \quad \text {and} \quad \int _M |\eta |^2_{h\otimes g^*}e^{-\varphi }\, dV_g \leq \int _M\lambda ^{-1} |\theta |^2_{h\otimes g^*}e^{-\varphi } \, dV_g . $$ \end{thm} \begin{rem} Demailly's theorem is much stronger than the above special case. This special case also follows from Theorem~4.1 of~[D1], since one can approximate~$\varphi (z)$ by functions which locally have the form~$A_p\log (|z|^2+\epsilon )$ near the nonsmooth points; or one can complete the metric on~$M\setminus S$. \end{rem} A well-known technique for producing sections with prescribed values on a discrete set gives the following: \begin{thm} Suppose $(L,h)$ is a Hermitian holomorphic line bundle on an irreducible reduced complex space~$X$ of dimension~$n$ and the curvature of~$h$ is semipositive on~$X$ and positive at some point in~$X$. Then there exist a positive integer~$\nu _0$ and a positive constant~$c_0$ which depend only on~$X$ and~$\cal C(L,h)$ and which have the following property. If~$\nu $ is an integer with $\nu \geq \nu _0$, $R$ is a nowhere dense analytic subset of~$X$ whose complement~$M=X\setminus R$ is smooth and admits a complete K\"ahler metric, $(F,k)$ is a Hermitian holomorphic line bundle on~$X$ with semi-positive curvature, $E_{\nu }=L^{\nu }\otimes F$, and $\pi : \widetilde M @>>> M$ is a connected covering space of degree~$d$ ($1\leq d\leq \infty $), then $$ c_0\nu ^nd \leq \dim H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (E_{\nu } \otimes K_M))). $$ The $L^2$ condition is taken with respect to the Hermitian metric $\pi ^*(h^{\nu }\otimes k)$ in $\pi ^*E_{\nu }$ and, for any choice of a Hermitian metric~$g$ on~$\widetilde M $, with respect to the Hermitian metric $g^*$~in~$K_{\widetilde M }=\pi ^*K_M$ and~$g$ on~$\widetilde M $ (the $L^2$-norm of an $(n,0)$-form does not depend on the choice of the metric on the manifold). \end{thm} \begin{rems} 1. The curvature condition on~$L$ means that if $s$ is a nonvanishing holomorphic section of~$L$ on an open set~$W$, then $-\log |s|^2_h$ is plurisubharmonic and, for some choice of~$W\neq \emptyset $, $-\log |s|^2_h$ is strictly plurisubharmonic . \noindent 2. The proof will also show that \begin{align*} c_0\nu ^nd &\leq c_0\nu ^n(d-1)+\dim H^0(X,{\cal O} (E_{\nu })\otimes \Omega ^n_X) \\ &\leq \dim \biggl( H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (E_{\nu } \otimes K_M))) +\pi ^*H^0(X,{\cal O} (E_{\nu }) \otimes \Omega ^n_X) \biggl); \end{align*} where the sum in the last expression takes place in $H^0(\widetilde M ,{\cal O} (\pi ^*(E_{\nu }\otimes K_M)))$. \noindent 3. By a theorem of Demailly~[D1], $M=X\setminus R$ admits a complete K\"ahler metric if, for example, $X$ is a complete K\"ahler manifold and~$R$ is a compact analytic subset. In particular, any smooth quasiprojective variety admits a complete K\"ahler metric. Thus we get as a special case the following: \end{rems} \begin{cor} Suppose $(L,h)$ is a positive Hermitian holomorphic line bundle on an irreducible reduced projective variety~$X$ of dimension~$n$. Then there exist a positive integer~$\nu _0$ which depends on~$X$ and~$\cal C(L,h)$ and which has the following property. If~$\nu $ is an integer with $\nu \geq \nu _0$, $R$ is a nowhere dense analytic subset of~$X$ with smooth complement~$M=X\setminus R$, and $\pi : \widetilde M @>>> M$ is a connected covering space of degree~$d$, then $$ d\leq \dim H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (L^{\nu } \otimes K_M))). $$ The $L^2$ condition is taken with respect to the Hermitian metric $\pi ^*h^{\nu }$ in $\pi ^*L^{\nu }$ and, for any choice of a Hermitian metric~$g$ on~$\widetilde M $, with respect to the Hermitian metric $g^*$~in~$K_{\widetilde M }=\pi ^*K_M$ and~$g$ on~$\widetilde M $. \end{cor} \begin{pf*}{Proof of Theorem~2.2} By hypothesis, $\cal C(L,h) \geq 0$ on~$X$ and $\cal C(L,h)>0$ on some relatively compact open subset~$W$ of~$X$. We may assume that $W\subset \reg X$ and that there exist holomorphic coordinates $z=(z_1,\dots ,z_n)$ with $|z|<1/2$ on~$W$. Fix a nonempty relatively compact open subset~$V$ of~$W$ and a~$C^{\infty } $ function~$\rho $ with compact support in~$W$ such that $\rho \equiv 1$ on a neighborhood of~$\overline V$, and, for each point~$p\in V$, let~$\varphi _p$ be the~$C^{\infty } $ function on~$X$ defined by $$ \varphi _p(x) = \left\{ \begin{alignedat}{2} &\rho (x)\log (|z(x)-z(p)|^2)& \quad \text { if } x\in W \\ &0&\quad \text { if } x\in X\setminus W \end{alignedat} \right. $$ Then $\text {supp}\, \varphi _p =\text {supp}\, \rho \subset W$, $\varphi _p =\log (|z-z(p)|^2)$ (a plurisubharmonic function ) on~$V$, and there is a positive constant~$a_0$ which does not depend on~$p$ such that $a_0\cal C(L,h)+\lev {\varphi _p}$ is semipositive on~$X\setminus \{ p \} $ and positive on~$W\setminus \{ p \} $. Fix an integer~$\nu _0 > na_0$ (later, we will also choose $c_0$ to depend only on~$a_0$ and~$n$). Let $\sing X\subset R\subset X$, $\pi : \widetilde M @>>> M=X\setminus R$, and d be as in the statement of the theorem and fix a point~$p$ in the nonempty open set~$V\setminus R$. Given a multi-index $\alpha =(\alpha _1,\dots ,\alpha _n)\in \Bbb Z _{\geq 0}^n$, a nonnegative integer~$\nu $, and a $C^{\infty } $ section~$s$ of $E_{\nu }\otimes K_X=L^{\nu }\otimes F\otimes K_X$ on a neighborhood of~$p$, we denote by $\partial ^{|\alpha |}s/\partial z^{\alpha }$ the corresponding multiple derivative of~$s$ with respect to some fixed trivialization in~$L$~and~$F$ on a neighborhood of~$p$ and the trivialization in~$K_X$ induced by the holomorphic coordinates $z=(z_1,\dots ,z_n)$ on~$W$. Similarly, if~$s$ is a~$C^{\infty } $ section of $\pi ^* (E_{\nu }\otimes K_M)=\pi ^*E_{\nu } \otimes K_{\widetilde M }$ on a neighborhood of a point $q\in \pi ^{-1} (p)$, then we denote by $\partial ^{|\alpha |}s/\partial z^{\alpha }$ the corresponding multiple derivative of~$s$ with respect to the trivialization and local coordinates lifted from~$X$. We will now apply the $\bar \partial $-method to show that if $\nu \geq \nu _0$, $\alpha =(\alpha _1,\dots ,\alpha _n)$ is a multi-index with $|\alpha |=\sum \alpha _j \leq (\nu /a_0)-n$, and $q\in \pi ^{-1} (p)$, then there exists a section $$ s\in H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* (E_{\nu } \otimes K_M))) =H^0_{L^2}(\widetilde M , {\cal O} (\pi ^* E_{\nu }\otimes K_{\widetilde M })) $$ such that , for every multi-index~$\beta $ with $|\beta | \leq (\nu /a_0)-n$ and for every point $r\in \pi ^{-1} (p)$, we have $$ \frac {\partial ^{|\beta |}s}{\partial z^{\alpha }}(r)= \left\{ \begin{alignedat}{2} &1& \quad \text { if } \beta =\alpha \text { and } r=q \\ &0&\quad \text { otherwise} \end{alignedat} \right. $$ By hypothesis, there exists a complete K\"ahler metric~$g$ on $M=X\setminus R$. Let $\widetilde E _{\nu }=\pi ^*E_{\nu }$ for each~$\nu $, let $\tilde h =\pi ^*h$, let $\tilde k =\pi ^*k$, let $\tilde g =\pi ^*g$, and let $\tilde \varphi _p=\pi ^*\varphi _p$. We may choose a relatively compact neighborhood~$U$ of~$q$ in $\pi ^{-1} (V) \setminus (\pi ^{-1} (p)\setminus \{ q \} )$ and a~$C^{\infty } $ section~$u$ of $\widetilde E _\nu \otimes K_{\widetilde M }$ with compact support in~$U$ such that $u$ is holomorphic on a neighborhood of~$q$ in~$\widetilde M $ and, for every multi-index~$\beta $, $$ \frac {\partial ^{|\beta |}u}{\partial z^{\alpha }}(q)= \left\{ \begin{alignedat}{2} &1& \quad \text { if } \beta =\alpha \\ &0&\quad \text { if } \beta \neq \alpha \end{alignedat} \right. $$ Hence the form $\theta =\bar \partial u$ is a~$C^{\infty } $~$\bar \partial $-closed $(n,1)$-form with values in~$\widetilde E _\nu $ and the support of~$\theta $ is a compact subset of $U\setminus \pi ^{-1} (p)$ (since $\bar \partial u=0$ near~$q$). By construction, there is also a continuous function $\lambda : \widetilde M @>>> [0,1]$ such that $\lambda >0$ on~$\pi ^{-1} (V)$ and $$ \cal C(\widetilde E _\nu ,{\exp ({-\frac {\nu }{a_0}\tilde \varphi _p}) \tilde h ^\nu \otimes \tilde k }) = \nu \cal C(\pi ^*L,\tilde h ) + \cal C(\pi ^*F,\tilde k ) +\frac {\nu }{a_0}\lev {\tilde \varphi _p} \geq \lambda \tilde g $$ on~$\widetilde M \setminus \pi ^{-1} (p) $. Moreover, $$ \int _{\widetilde M } \lambda ^{-1} |\theta |^2_{\tilde h ^\nu \otimes \tilde k \otimes \tilde g ^*} e^{-\frac {\nu }{a_0}\tilde \varphi _p} \, dV_{\tilde g } < \infty , $$ because $\theta $ has compact support in $U\setminus \pi ^{-1} (p)$, $\lambda >0$ on~$U$, and $\tilde \varphi _p$ is smooth on $U\setminus \{ q \}= U\setminus \pi ^{-1} (p)$. Applying Demailly's theorem (Theorem~2.1), one gets a $C^{\infty } $ form~$\eta $ of type~$(n,0)$ with values in $\widetilde E _\nu $ on~$\widetilde M $ such that $$ \bar \partial \eta =\theta \quad \text {and} \quad \int _{\widetilde M } |\eta |^2_{\tilde h ^\nu \otimes \tilde k \otimes \tilde g ^*} e^{-\frac {\nu }{a_0}\tilde \varphi _p} \, dV_{\tilde g } < \infty . $$ In particular, the $(n,0)$-form $s=u-\eta $ is in $H^0_{L^2}(\widetilde M , {\cal O} (\widetilde E_{\nu }\otimes K_{\widetilde M }))$ because $\bar \partial s=0$ and $\tilde \varphi _p\leq 0$. Since $u$ is holomorphic near each point $r\in \pi ^{-1} (p)$, so is~$\eta $. Moreover, in suitable local holomorphic coordinates $w=(w_1,\dots ,w_n)$ centered at~$r$ in a neighborhood ~$Q$, we have $\tilde \varphi _p(w)=\log |w|^2$ and hence $$ \int _Q|\eta |^2|w|^{-2\nu/a_0} \, dV<\infty ; $$ where the notation for the metrics has been suppressed. Therefore $\eta $ vanishes at~$r$ to an order greater than $(\nu /a_0)-n$. Thus, if $\beta $ is a multi-index with $|\beta |\leq (\nu /a_0)-n$, then $\partial ^{|\beta |}\eta /z^\beta $ vanishes at each point $r\in \pi ^{-1} (p)$ and the claim follows. The claim implies that if, for each $c\geq 0$, $b_c=\binom {[c]+n}{n}$ denotes the number of multi-indices~$\alpha $ satisfying $|\alpha | \leq c $, then we have $$ \dim H^0_{L^2}(\widetilde M , {\cal O} (\widetilde E _{\nu }\otimes K_{\widetilde M })) \geq b_{(\nu /a_0)-n}\cdot d $$ for each integer $\nu \geq \nu _0$. It is easy to see that $b_{(\nu /a_0)-n}\geq c_0\nu ^n$ for some positive constant~$c_0$ depending only on $a_0$~and~$n$ and the theorem now follows. \end{pf*} \begin{rem} To obtain the inequalities given in the remark~2, we fix a point $r\in \pi ^{-1} (p)$ and, for each point $q\in \pi ^{-1} (p)\setminus \{ r \} $ and each multi-index~$\alpha $ with $|\alpha |\leq (\nu /a_0)-n$, we form a section in $H^0_{L^2}(\widetilde M , {\cal O} (\widetilde E _{\nu }\otimes K_{\widetilde M }))$ as in the above proof. We then get a collection of $b_{(\nu /a_0)-n}\cdot (d-1)$ linearly independent sections and the span of this collection meets $\pi ^*H^0(X,{\cal O} (E_\nu ) \otimes \Omega ^n_X)$ only in the zero section. Therefore \begin{align*} c_0\nu ^n(d-1)&+\dim H^0(X,{\cal O} (E_{\nu }) \otimes \Omega ^n_X) \\ &\leq \dim \biggl( H^0_{L^2}(\widetilde M , {\cal O} (\widetilde E _{\nu } \otimes K_{\widetilde M })) + \pi ^*H^0(X,{\cal O} (E_{\nu }) \otimes \Omega ^n_X)\biggl) . \end{align*} Finally, observe that if we take $\widetilde M =M$ and $s\in H^0_{L^2}(M , {\cal O} (E _{\nu }\otimes K_M))$, then, since the $L^2$~condition in the canonical bundle is independent of the choice of the metric on the base manifold (provided one also takes the associated metric in the canonical bundle), the pullback~$s'$ of $s$ to a desingularization~$X'$ of~$X$ is locally in~$L^2$ with respect to a metric on~$X'$. Therefore $s'$ extends to a section in~$H^0(X', \Omega ^n_{X'})$ and hence $s$ extends to a section in~$H^0(X, \Omega ^n_X)$. It follows that $$ c_0\nu ^n \leq \dim H^0(X,{\cal O} (E_{\nu }) \otimes \Omega ^n_X). $$ Thus we get all of the desired inequalities. \end{rem} \section{Further generalizations of the weak Lefschetz theorem} Theorem~2.2 and the arguments given in the proofs of Theorem~1.1 and Theorem~1.4 now give the following generalization: \begin{thm} Let $U$ be an irreducible reduced complex space, let $X$ be connected normal complex space of dimension $n>1$, let $\Phi :U @>>> X$ be a holomorphic mapping, and let $(L,h)$ be a Hermitian holomorphic line bundle on~$X$. Assume that \begin{enumerate} \item[(i)] $\Phi (U)$ has nonempty interior, and \item[(ii)] The curvature of~$(L,h)$ is semipositive everywhere on~$X$ and positive at some point in~$X$. \end{enumerate} Then there exist a positive integer~$\nu _0$ and a positive constant~$c_0$ which depend only on~$X$ and (the curvature of)~$(L,h)$ such that , if~$R$ is a nowhere dense analytic subset of~$X$ whose complement~$\reg{X}\setminus R$ in $\reg{X}$ admits a complete K\"ahler metric, $V$~is a (nonempty) domain in~$U$, and~$\nu $~is an integer with $\nu \geq \nu _0$, then \begin{align*} c_0\nu ^n d &\leq \dim H^0(V, {\cal O} (\Phi ^*L^\nu )\otimes \Omega ^n_U); \end{align*} where $d$ is the index of the image~$G$ of $\pi_1(V\setminus \Phi ^{-1} (R)) @>>> \pi _1(X\setminus R)$. In particular, if $H^0 (V, {\cal O} (\Phi ^*(L^\nu )) \otimes \Omega ^n_U)$ is finite dimensional for some choice of a sufficiently large~$\nu $, then $G$ is of finite index. \end{thm} Next, we show that for $U$ and $X$ smooth, there exists a bound on the index in terms of the dimension of the space of sections of an {\it invertible} sheaf. We first prove an elementary fact which relates sections of the pullback of the canonical bundle to holomorphic $n$-forms (see also Sakai~[S]). \begin{lem} Suppose $\Phi =(\Phi _1,\dots ,\Phi _n) : \Delta ^m @>>> \C ^n$ is a holomorphic mapping and $\Phi _*$ has rank~$n$ at each point in $\Delta ^*\times\Delta ^{m-1}$. We denote the coordinates in~$\C ^m$ and the coordinates in~$\Cn$ by $z=(z_1,\dots ,z_m)$ and $\zeta =(\zeta _1,\dots ,\zeta _n)$, respectively. Let~$l\geq 0$ be the order of vanishing of the holomorphic $n$-form $d\Phi _1\wedge \dots \wedge d\Phi _n$ along $\{ 0 \} \times \Delta ^{m-1}$. Then the mapping ${\cal O} (\Phi ^*K_{\C ^n}) @>>> \Omega ^n_{ \Delta ^m }$ given by $$ s =f\Phi ^* (d\zeta _1\wedge \dots \wedge d\zeta _n) \mapsto fz_1^{-l}d\Phi _1\wedge \dots \wedge d\Phi _n $$ maps ${\cal O} (\Phi ^*K_{\C ^n}) $ isomorphically onto the sheaf of holomorphic $n$-forms~$\theta $ such that $\theta _z\in \C (d\Phi _1\wedge \dots \wedge d\Phi _n)_z$ for each point $z\in \Delta ^* \times \Delta ^{m-1}$ at which~$\theta $ is defined. \end{lem} \begin{rem} Here $\Phi ^* (d\zeta _1\wedge \dots \wedge d\zeta _n)$ denotes the pullback of $d\zeta _1\wedge \dots \wedge d\zeta _n$ as a section of $\Phi ^*K_{\Cn }$ while $d\Phi _1\wedge \dots \wedge d\Phi _n$ is the pullback as a form of type~$(n,0)$. \end{rem} \begin{pf} Clearly, $z_1^{-l}d\Phi _1\wedge \dots \wedge d\Phi _n$ is a holomorphic $n$-form on~$\Delta ^m$, so we get an injective mapping as described above. Conversely, suppose~$\theta $ is a holomorphic $n$-form on an open set $V\subset \Delta ^m$ and there exists a holomorphic function~$h$ on $V\cap (\Delta ^* \times \Delta ^{m-1})$ with $\theta =hd\Phi _1\wedge \dots \wedge d\Phi _n$ on $V\cap (\Delta ^* \times \Delta ^{m-1})$. We have $$ \theta =\Sigma ' \theta _Idz_I \text { on } V \quad \text {and} \quad d\Phi _1\wedge \dots \wedge d\Phi _n =\Sigma ' \beta _Idz_I \text { on } \Delta ^m; $$ where $\sum '$ denotes the sum over increasing multi-indices. In particular, $$ l=\min _I (\text {ord} _{ \{ 0 \} \times \Delta ^{m-1} }\beta _I ) =\text {ord} _{ \{ 0 \} \times \Delta ^{m-1} }\beta _{I_0} $$ for some multi-index~$I_0$, and, for each nonzero coefficient~$\beta _I$, we have $h=\theta _I/ \beta _I$ on $V\cap (\Delta ^* \times \Delta ^{m-1})$. Therefore $h$ is a meromorphic function on~$V$ with pole set contained in $\{ 0 \} \times \Delta ^{m-1}$ and $z_1^lh= \theta _{I_0}/ (\beta _{I_0}/z_1^l)$. Since the intersection of the zero set of $\beta _{I_0}/z_1^l$ and $\{ 0 \} \times \Delta ^{m-1}$ has codimension at least~$2$ in~$\Delta ^m$ and since the the pole set of $z_1^lh$ lies in this intersection, the pole set must be empty. Therefore $z_1^lh$ is holomorphic on~$V$ and the holomorphic section $s=z_1^lh\Phi ^*(d\zeta _1\wedge \dots \wedge d\zeta _n)$ of $\Phi ^*K_{\C ^n}$ maps to $\theta $ (on $V\cap (\Delta ^* \times \Delta ^{m-1})$ and hence on~$V$). \end{pf} \begin{thm} Let $U$~and~$X$ be connected complex manifolds of dimensions~$m$ and $n>1$, respectively, let $\Phi :U @>>> X$ be a holomorphic mapping, and let~$(L,h)$ be a Hermitian holomorphic line bundle on~$X$. Assume that \begin{enumerate} \item[(i)] $\Phi $ has rank~$n$ at some point (i.e.~$\Phi (U)$ has nonempty interior), and \item[(ii)] The curvature of $(L,h)$ is semipositive everywhere on~$X$ and positive at some point in~$X$. \end{enumerate} Then there exist a positive integer~$\nu _0$ and a positive constant~$c_0$ which depend only on~$X$ and (the curvature of)~$(L,h)$ and there exists an effective divisor~$D_0$ in~$U$ which depends only on the mapping $\Phi : U @>>> X$ such that , if~$R$ is a nowhere dense analytic subset of~$X$ whose complement~$X\setminus R$ admits a complete K\"ahler metric, $(F,k)$ is a Hermitian holomorphic line bundle on~$X$ with semipositive curvature, $V$~is a (nonempty) domain in~$U$, $\nu $~is an integer with $\nu \geq \nu _0$, $E_\nu =L^\nu \otimes F$, and $d$ is the index of the image~$G$ of $\pi_1(V\setminus \Phi ^{-1} (R)) @>>> \pi _1(X\setminus R)$, then we have the estimates \begin{align*} c_0\nu ^nd &\leq c_0\nu ^n (d-1) +\dim H^0 (X, {\cal O} (E_\nu \otimes K_X)) \\ &\leq \dim H^0 (V, {\cal O} (\Phi ^*(E_\nu \otimes K_X)\otimes [D_0])) \tag 1 \\ &\leq \dim H^0(V, {\cal O} (\Phi ^*E_\nu )\otimes \Omega ^n_U). \end{align*} \end{thm} \begin{rems} 1. If, for some positive integer~$k$, $L^k\otimes K ^{-1} _X$ is semipositive, then we may take $F=L^k\otimes K ^{-1} _X$ and we get estimates which do not involve the canonical bundle~$K_X$. \noindent 2. The divisor~$D_0$ which will be constructed is probably not the optimal choice. \end{rems} \begin{pf*}{Proof of Theorem~3.3} Guided by Lemma~3.2, we first describe~$D_0$. The set $$ B=\setof {x\in U}{\text {rank}\, (\Phi _*)_x < n } $$ is a nowhere dense analytic subset of~$U$. Let $\{ A_i \} $ be the collection of all of the irreducible components of~$B$ of dimension~$m-1$ whose image $\Phi (A_i)$ lies in some nowhere dense analytic subset of~$X$, let $A=\cup _i A_i$, and, for each~$i$, let~$l_i$ be the minimal order of vanishing along~$A_i$ of the $(n\times n)$-minor determinants of~$\Phi _*$. In other words, if $\Phi =(\Phi _1,\dots , \Phi _n)$ and $d\Phi _1\wedge \dots \wedge d\Phi _n=\sum 'a_Jdz_J$ with respect to local coordinates $(z_1,\dots ,z_m)$ near $x_0\in A_i$ in~$U$ and $(\zeta _1,\dots , \zeta _n)$ near~$\Phi (x_0)$ in~$X$, then $$ l_i=\min _J (\text {ord}_{A_i} a_J). $$ We define $$ D_0=\sum l_iA_i. $$ Given a holomorphic line bundle~$E$ on~$X$, Lemma~3.2 implies that we have an injective linear mapping $$ H^0(U, {\cal O} ((\Phi ^* (E\otimes K_X))\otimes [D_0])) @>>> H^0(U, {\cal O} (\Phi ^* E)\otimes \Omega ^n_U) $$ given as follows. Let~$t$ be a global defining section for~$[D_0]$ on~$U$. To each section $s\in H^0(U, {\cal O} ((\Phi ^* (E\otimes K_X))\otimes [D_0]))$, we may associate a holomorphic $n$-form~$\theta $ with values in $(\Phi ^* E)\otimes [D_0]$ and $\theta /t$ is a holomorphic $n$-form on~$U\setminus A$ with values in $\Phi ^* E$. But the lemma implies that, near points of $A\setminus \sing B$, $\theta /t$ extends holomorphically past~$A$. Thus $\theta /t$ extends to a holomorphic $n$-form on~$U\setminus (A\cap \sing B)$ with values in~$\Phi ^* E$. Since co$\dim \sing B \geq 2$, $\theta /t$ extends holomorphically to the entire manifold~$U$. Thus we get a mapping $s \mapsto \theta /t$ (similarly, this mapping surjects onto the space of holomorphic $n$-forms with values in~$\Phi ^* E$ whose restriction to $U\setminus A$ comes from a section of $\Phi ^*(E\otimes K_X)$). In particular, the third of the inequalities in~(1) holds. Let \begin{center}\begin{picture}(250,80) \put(5,10){$W=V\setminus S\subset V $} \put(90,14){\vector(1,0){50}} \put(145,10){$X \supset X\setminus R=M$} \put(110,3){$\Phi $} \put(125,55){\vector(3,-1){90}} \put(110,60){$\widetilde M $} \put(17,25){\vector(3,1){90}} \put(57,45){$\tilde \Phi $} \put(170,45){$\pi $} \end{picture} \end{center} be a commutative diagram as in the proof of Theorem~1.1. We will show that if $s\in H^0_{L^2}(\widetilde M , {\cal O} ((\pi ^*E_\nu )\otimes K_{\widetilde M }))$ for some~$\nu $ (with respect to metrics lifted from the base), then $(\tilde \Phi ^*s)\otimes t$ extends to a unique holomorphic section of $(\Phi ^* (E_\nu \otimes K_X))\otimes [D_0]$ on~$V$. By Lemma~1.3, the pullback of $s$ {\it as an $E_\nu $-valued holomorphic $n$-form} extends to a $\Phi ^*E_\nu $-valued holomorphic $n$-form on~$V$. In particular, $\tilde \Phi ^*s$ extends holomorphically as a section of $\Phi ^*(E_\nu \otimes K_M)$ near each point at which $\Phi _*$ is of maximal rank (by Lemma~3.2 with $l=0$). Moreover, $\tilde \Phi ^*s$ extends holomorphically past analytic sets of codimension at least~$2$. Therefore, it suffices to show that $(\tilde \Phi ^*s)\otimes t$ extends holomorphically near each point $x_0\in \reg S\cap \reg B$ at which $S\cap B$ is of dimension~$m-1$. An irreducible component of~$B$ containing such a point~$x_0$ must also be an irreducible component of~$S=\Phi ^{-1} (R)$ and must therefore be one of the irreducible components~$A_i$ of the support~$A$ of~$D_0$. Since the pullback of $s$ as an $E_\nu $-valued holomorphic $n$-form extends to~$V$, Lemma~3.2 and the definition of $D_0$ and $t$ now imply the claim. Clearly, if $s$ is a holomorphic section of $E_\nu \otimes K_X$ on $X$, then $(\Phi ^*s)\otimes t$ is a holomorphic section whose restriction to~$V$ is an extension of $(\tilde \Phi ^*\pi ^* s)\otimes t$. Thus we get an injective linear mapping of the subspace $$ \cal S=H^0_{L^2}(\widetilde M , {\cal O} ((\pi ^*E_\nu )\otimes K_{\widetilde M }))+ \pi ^* H^0(X, {\cal O} (E_\nu \otimes K_X)) $$ of $H^0(\widetilde M , {\cal O} ((\pi ^*E_\nu )\otimes K_{\widetilde M }))$ into $H^0(V,{\cal O} ((\Phi ^*(E_\nu \otimes K_X))\otimes [D_0]))$. We have, therefore \begin{align*} \dim \cal S &\leq \dim H^0 (V, {\cal O} (\Phi ^*(E_\nu \otimes K_X)\otimes [D_0])) \\ &\leq \dim H^0(V, {\cal O} (\Phi ^*E_\nu )\otimes \Omega ^n_U). \end{align*} The second remark following Theorem~2.2 now gives the inequalities~(1) for $\nu $ sufficiently large and for some constant~$c_0$ (both depending only on $(L,h)$ and $X$). \end{pf*} \begin{rems} 1. The proofs of Lemma~1.3 and Theorem~3.3 show that one can form a divisor $D_R$ which depends on~$R$, but which satisfies $D_R\leq D_0$ and gives a sharper estimate for the index. For example, it suffices to include only those irreducible components $A_i$ which are contained in $S=\Phi ^{-1} (R)$, so one may choose $D_R$ to have support contained in~$S$. Moreover, the proof of Lemma~1.3 shows that if $\Phi (A_i)$ contains a point~$p\in R$ at which $R$ is a divisor with normal crossings and $u$ is a defining function for $R$ near~$p$, then one may take the coefficient of~$A_i$ to be $-1+ \text {ord}_{A_i}(u\circ \Phi )$. The proof also shows that, by choosing~$p$ so that this coefficient is minimal, we get $D_R\leq D_0$. \noindent 2. Similarly, for $U$ a normal neighborhood of a compact complex space~$Y$ and $X$ a smooth projective variety, one can find a uniform bound on the index in terms of the dimension of a space of sections of a line bundle pulled back from $X$ as in Theorem~1.1. More precisely, we have the following: \end{rems} \begin{cor} Let $\Phi :U @>>> X$ be a holomorphic mapping of a connected normal complex space~$U$ into a connected smooth projective variety~$X$ of dimension~$n>1$, let~$Y$ be a connected compact analytic subspace (not necessarily reduced) of~$U$, and let $\widehat U $ be the formal completion of~$U$ with respect to~$Y$. Assume that \begin{enumerate} \item[(i)] $\Phi (U)$ has nonempty interior, and \item[(ii)] $\dim H^0(\widehat U , \widehat {{\cal O} (\Phi ^*L)}) <\infty $ for every holomorphic line bundle~$L$ on~$X$. \end{enumerate} Then there is a positive constant~$b$ depending only on the mapping $\Phi :U @>>> X$ and the subspace~$Y\subset U$ such that , if $R\subset X$ is a nowhere dense analytic subset of~$X$ and~$V$ is a connected neighborhood of~$Y$ in~$U$, then the image~$G$ of $\pi _1(V\setminus \Phi ^{-1} (R)) @>>> \pi _1(X\setminus R)$ is of index at most~$b$ in~$\pi _1(X\setminus R)$. \end{cor} \begin{pf*}{Sketch of the proof} First suppose $U$ is smooth and let $D_0=\sum l_iA_i$ be the associated divisor in~$U$ as in the proof of Theorem~3.3. By construction, each of the sets $\Phi (A_i)$ is contained in some nowhere dense analytic subset of~$X$. By replacing $U$ by a relatively compact neighborhood of~$Y$, we may assume that there is a nowhere dense analytic subset~$C$ in~$X$ which contains all of these sets and that the collection of coefficients $\seq li$ is bounded. Hence we may choose a positive holomorphic line bundle~$L$ on $X$ and a holomorphic section $t$ of~$L$ such that the divisor $D_1$ of the section $\Phi ^* t$ satisfies $D_1\geq D_0$. Now let $R$, $V$, and $G$ be as in the statement of the corollary and let \begin{center}\begin{picture}(250,80) \put(5,10){$W=V\setminus S\subset V $} \put(90,14){\vector(1,0){50}} \put(145,10){$X \supset X\setminus R=M$} \put(110,3){$\Phi $} \put(125,55){\vector(3,-1){90}} \put(110,60){$\widetilde M $} \put(17,25){\vector(3,1){90}} \put(57,45){$\tilde \Phi $} \put(170,45){$\pi $} \end{picture} \end{center} be a commutative diagram as in the proof of Theorem~1.1. By the proof of Theorem~3.3 and the above remarks, if $s\in H^0_{L^2}(\widetilde M ,{\cal O} (\pi ^* L\otimes K_{\widetilde M }))$, then $(\tilde \Phi ^*s)\otimes (\Phi ^* t)$ extends to a holomorphic section of $\Phi ^*(L^2\otimes K_X)$ on~$V$. If $U$ is connected and normal (but not necessarily smooth), then we may form a desingularization $\alpha : U' @>>> U$ of~$U$ and a commutative diagram \begin{center}\begin{picture}(250,80) \put(70,10){$U$} \put(85,14){\vector(1,0){65}} \put(155,10){$X$} \put(115,0){$\Phi $} \put(75,55){\vector(0,-1){30}} \put(70,60){$U'$} \put(60,35){$\alpha $} \put(85,60){\vector(2,-1){70}} \put(115,50){$\Phi '$} \end{picture} \end{center} We may associate to $\Phi ' : U' @>>> X$ (after shrinking~$U$) a line bundle~$L$ and a section~$t$ as above, and we get the extension property for pullbacks of $L^2$ sections as described. On the other hand, $U$ is normal, so $\alpha _*{\cal O} ((\Phi ')^*(L^2\otimes K_X))={\cal O} (\Phi ^*(L^2\otimes K_X))$. Therefore the extension property also holds in~$U$, and the usual argument now applies. \end{pf*} We close this section by observing that Theorem~3.1 has immediate consequences for pseudoconcave spaces. An open subset $\Omega $ of a complex space~$X$ is said to have {\it pseudoconcave boundary in the sense of Andreotti}~[A] if each point~$x_0\in \partial \Omega $ admits a fundamental system of neighborhoods~$W$ in~$X$ such that $x_0$ is an interior point of $$ \widehat {(W\cap \Omega )}_X =\setof {x\in X}{|f(x)|\leq \sup _{W\cap \Omega }|f|\quad \forall \, f\in {\cal O} (X) }. $$ For example, by Proposition~10 of~[A], if each irreducible component of~$X$ has dimension at least~$k>1$ and, for each point~$x_0\in \partial \Omega $, there is a $C^{\infty } $~$(k-1)$-convex function~$\varphi $ on a neighborhood~$W$ of~$x_0$ in~$X$ such that $$ \Omega \cap W =\setof {x\in W}{\varphi (x) >0}, $$ then $\Omega $ has pseudoconcave boundary in the sense of Andreotti. A connected complex space~$X$ is said to be {\it pseudoconcave in the sense of Andreotti}~[A] if there exists a nonempty relatively compact open subset~$\Omega $ which has pseudoconcave boundary in the sense of Andreotti and which meets each irreducible component of~$X$. By a finiteness theorem of Andreotti~[A, Theorem~1], if $\cal F$ is a torsion-free coherent analytic sheaf on a locally irreducible connected complex space~$X$ and $X$ is pseudoconcave in the sense of Andreotti, then $\dim H^0(X,\cal F) < \infty $ (the case in which $X$ admits a $C^{\infty } $~$(k-1)$-convex exhaustion function, where the dimension of~$X$ is at least $k>1$ at each point, is due to Andreotti and Grauert~[AG]). Theorem 3.1 and Andreotti's finiteness theorem together give the following: \begin{cor} Let $U$ be an irreducible reduced complex space, let $X$ be a connected normal projective variety of dimension~$n>1$, and let $\Phi :U @>>> X$ be a holomorphic mapping. Assume that $\Phi (U)$ has nonempty interior and that $U$ is pseudoconcave in the sense of Andreotti. Then there is a positive constant~$b$ depending only on the mapping $\Phi : U @>>> X$ such that , if~$Z$ is a nonempty Zariski open subset of~$X$, then the image of $\pi _1(\Phi ^{-1} (Z)) @>>> \pi _1 (Z)$ is of index at most~$b$ in~$\pi _1(Z)$. \end{cor} \begin{rems} 1. Clearly, Theorem~3.1 also gives a version of the above theorem in which $X$ is not necessarily projective. 2. There are many results concerning when a compact analytic subset~$Y$ of an $m$-dimensional complex space~$U$ admits a strongly $(m-1)$-concave neighborhood , and hence when one may apply Andreotti's [A] (or Andreotti and Grauert's~[AG]) finiteness theorem as above. For example, Okonek~[O] proved that $Y$ admits a fundamental system of such neighborhoods if $N_{Y/U}$ is Finsler-$q$-positive, where $q=\dim Y$. \end{rems} \section{Burns' theorem} The goal of this section is the following theorem: \begin{thm} Let $(X,g)$ be a connected complete Hermitian manifold and let $M\subset X$ be a domain with nonempty smooth compact boundary~$\partial M$ in~$X$. Assume that \begin{enumerate} \item[(i)] $M$ is strongly pseudoconvex at each point of~$\partial M$; \item[(ii)] There exists a Hermitian metric~$a$ in~$K_M$ and a constant $c>0$ such that $\cal C(K_M,a)\geq cg$ on~$M$; and \item[(iii)] $X$ has dimension~$n\geq 3$. \end{enumerate} Then $\text {\rm vol} \, _g(M)<\infty $. \end{thm} \begin{rems} 1. Since $M$ admits a complete K\"ahler metric, $\partial M$ is necessarily connected (see, for example, Proposition~4.4 below). \noindent 2. If, for example, the Ricci curvature of~$g$ is bounded above by $-c$ on~$M$, then the associated metric $a=g^*$ in~$K_M$ satisfies the condition~(ii) since $$ \cal C(K_M,g^*)=-\text {Ric} \, (g) \geq cg. $$ \noindent 3. Clearly, it is not necessary to assume that $M$ is a domain in some larger manifold~$X$. The conclusion also holds if $M=X$ and $M$ admits a $C^{\infty } $ function which, along some end, is strictly plurisubharmonic and exhaustive; since one can then replace $M$ by a suitable sublevel set of the function. It will, however, be more convenient to have Theorem~4.1 stated for a domain as above. \noindent 4. As in the proofs of the weak Lefschetz theorems, the idea is to apply finite dimensionality of a space of holomorphic sections of a line bundle to obtain a result about the manifold. \end{rems} Theorem~4.1 and an analysis of the thick-thin decomposition as in~[BGS] together give as a conclusion the following theorem: \begin{thm}[Burns~[B{]}] Let $\Gamma $ be a torsion-free discrete group of automorphisms of the unit ball~$B$ in~$\C ^n$ with $n \geq 3$ and let $M =\Gamma \setminus B$. Assume that the limit set~$\Lambda $ is a proper subset of~$\partial B$ and that the quotient $\Gamma \setminus ((\partial B)\setminus \Lambda )$ has a compact component~$A$. Then $M$ has only finitely many ends; all of which, except for the (unique) end corresponding to~$A$, are cusps. In fact, $M$ is diffeomorphic to a compact manifold with boundary. \end{thm} \begin{rem} By applying a theorem of Lempert~[L] as in the proof of Theorem~4.1 below and the argument given by Siu and Yau~[SY], one can close up the cusps projectively. In other words, $M\cong \Omega \setminus D$, where~$\Omega $ is a strongly pseudoconvex domain in a smooth projective variety and~$D$ is a (compact) divisor contained in~$\Omega $. The boundary component~$A$ corresponds to~$\partial \Omega $. \end{rem} The main tool in the proof of Theorem~4.1 is Nadel and Tsuji's~[NT] $L^2$ version of Demailly's~[D2] asymptotic Riemann-Roch inequality. \begin{thm}[Nadel-Tsuji~[NT{]}] Suppose $(X,g)$ is a connected complete K\"ahler manifold of dimension~$n$ and $(L,h)$ is a Hermitian holomorphic line bundle on~$X$ such that $$ \cal C(L,h)\geq cg $$ for some constant $c>0$. Then $$ \liminf _{\nu @>>> \infty } \nu ^{-n} \dim H^0_{L^2}(X, {\cal O} (K_X\otimes L^\nu )) \geq \frac {1}{n!}\int _X \bigl( c_1(L,h)\bigr) ^n. $$ \end{thm} \begin{rems} 1. The Chern form $c_1(L,h)$ is the real form of type~$(1,1)$ (associated to the Hermitian tensor $\cal C(L,h)$) given by $$ c_1(L,h)=-\frac {\sqrt {-1}}{2\pi }\partial \bar \partial \log |s|^2_h $$ for any local nonvanishing holomorphic section~$s$ of~$L$. \noindent 2. As Nadel and Tsuji observed (see~[NT, Lemma~2.5]), if, in particular, $X$ is pseudoconcave in the sense of Andreotti~[A] (see Sect.~3), then it follows that~$X$ has finite volume. \noindent 3. The theorem is only stated in~[NT] for~$L$ the canonical bundle, but the proof of the general case is the same. The first point is that, for a smooth relatively compact domain~$\Omega $ in~$X$ and for $\lambda >0$, one has Demailly's~[D2] generalization of Weyl's asymptotic formula for the number of eigenvalues~$N_\Omega (\lambda )$ less than or equal to~$\nu \lambda $ for the Dirichlet problem for the Laplacian in $K_X\otimes L^\nu $: $$ \liminf _{\nu @>>> \infty } \nu ^{-n} N_\Omega (\lambda )\geq \frac {1}{n!}\int _\Omega \bigl( c_1(L,h)\bigr) ^n. $$ The second point is that for a $C^{\infty } $ compactly supported form~$\alpha $ of type~$(n,1)$ with values in~$L^\nu $, the Bochner-Kodaira formula implies that $$ \| \bar \partial \alpha \| _{L^2}^2+\| \bar \partial ^*\alpha \| _{L^2}^2 \geq c\nu \| \alpha \| _{L^2}^2. $$ With these slight changes in mind, the proof given in~[NT] goes through. \end{rems} We will also apply the following Hartogs type extension property: \begin{prop} Let $(X,g)$ be a connected complete Hermitian manifold of dimension~$n>1$ and let $M\subset X$ be a domain with nonempty smooth compact strongly pseudoconvex boundary. Assume that the restriction $g| _M$ of~$g$ to~$M$ is K\"ahler. Suppose~$f$ is a holomorphic function on $U\cap M$ for some neighborhood~$U$ of~$\partial M$ in~$X$. Then there exists a holomorphic function~$h$ on~$M$ such that $h=f$ near~$\partial M$. In particular, $\partial M$ is connected . \end{prop} \begin{pf} We may assume that $M=\setof {x\in X}{\varphi (x) <0}$ for some $C^{\infty } $ function~$\varphi $ on~$X$ which is strictly plurisubharmonic on a neighborhood of~$X\setminus M$ in~$X$. Since $g| _M$ is K\"ahler and $g$~is complete on~$X$, a theorem of Nakano~[N] and of Demailly~[D1] implies that~$M$ admits a complete K\"ahler metric~$g'$. Moreover, the existence of~$\varphi $ implies that $(M,g')$ admits a positive Green's function~$G$ which vanishes along~$\partial M$. We normalize~$G$ so that, for each point $x_0\in M$, $$ \Delta _{\text {distr.}}G(\cdot , x_0) =-(2n-2)\sigma _{2n-1}\delta _{x_0}; $$ where $n=\dim X$, $\sigma _{2n-1}=\text {\rm vol} \, (S^{2n-1})$, and $\delta _{x_0}$ is the Dirac function at~$x_0$. Fix a $C^{\infty } $ function~$\lambda $ with compact support in~$U$ such that $\lambda \equiv 1$ on a neighborhood of~$\partial M$ and let~$\alpha $ be the $\bar \partial $-closed compactly supported form of type~$(0,1)$ on~$M$ given by $\alpha =\bar \partial (\lambda f)$ (extended by~$0$ to~$M$). Then the function~$\beta $ defined by $$ \beta (x) =-\frac {1}{(2n-2)\sigma _{2n-1}} \int _M G(x,y) \bar \partial ^* \alpha (y)\, dV_{g'}(y) $$ is a $C^{\infty } $ bounded function with finite energy (i.e.~$\int _M |\nabla \beta | ^2 \, dV_{g'} <\infty $), $\Delta \beta = \bar \partial ^* \alpha $, and $\beta $ vanishes on~$\partial M$. Hence $\gamma \equiv\alpha -\bar \partial \beta $ is an $L^2$ harmonic form of type~$(0,1)$ and the Gaffney theorem~[G] implies that~$\gamma $ is closed (and coclosed). In particular, $\bar \gamma $ is a holomorphic $1$-form on~$M$ and $\beta $~is pluriharmonic on $W\cap M$ for some neighborhood~$W$ of~$X\setminus M$ in~$X$. We will show that~$\beta $ vanishes near~$\partial M$. Fix $a<0$ so close to~$0$ that~$\varphi $ is strictly plurisubharmonic on $V=\setof {x\in M}{\varphi (x) >a}$ and $V\subset\subset W$. If $\rho $ is the real part or the imaginary part of~$\beta $ and $\rho $ does not vanish identically near~$\partial M$, then we may choose a nonzero regular value~$b$ of~$\rho $ contained in~$\rho (V)$. Since $b\neq 0$ and $\rho $ vanishes on~$\partial M$, $\rho ^{-1} (b)$ avoids~$\partial M$. Thus the restriction of~$\varphi $ to~$\rho ^{-1} (b)$ assumes its maximum at some point~$x_0\in V\subset W\cap M$ (with $\varphi (x_0)>a$). But the leaf~$L$ through~$x_0$ of the foliation determined by the holomorphic $1$-form~$\partial \rho $ on~$V\cap M$ is contained in~$\rho ^{-1} (b)$, so $\varphi | _L$ also assumes its maximum at~$x_0$. Since~$\varphi $ is strictly plurisubharmonic on~$V$, we have arrived at a contradiction. Therefore $\beta $ vanishes near~$\partial M$. Hence $\gamma =\alpha -\bar \partial \beta $ vanishes near~$\partial M$ and, therefore, on all of~$M$, since~$\bar \gamma $ is a holomorphic $1$-form. Thus the function~$h\equiv \lambda f-\beta $ is holomorphic on~$M$ (since $\bar \partial h=\gamma =0$) and equal to~$f$ near~$\partial M$. In particular, since one can take $f$ to be a locally constant function which separates distinct components of~$\partial M$, $\partial M$ is connected . \end{pf} \begin{pf*}{Proof of Theorem~4.1} Since $n\geq 3$, one can apply a theorem of Rossi~[R] to ``fill in the holes'' and obtain a connected Stein space~$Y$ with isolated singularities, a relatively compact pseudoconvex domain~$N$ in~$Y$ containing~$\sing Y$, and a biholomorphic mapping $\Phi : U @>>> V$ of a neighborhood~$U$ of~$\partial M$ in~$X$ onto a neighborhood~$V$ of~$\partial N$ in~$Y$ such that $\Phi (U\cap M)=V\cap N$. Since $N$ may be embedded into a Euclidean space, Proposition~4.4 implies that~$\Phi $ extends to a holomorphic mapping $M\cup U @>>> Y$, which we also denote by~$\Phi $, and $\Phi (M)\subset N$. Next, by a theorem of Lempert~[L], one can form a ``cap'' on~$N$. That is, we may assume that~$Y$ is an affine algebraic variety. By forming the closure~$\overline {Y}$ of~$Y$ in a projective space and desingularizing~$\overline {Y}$ at infinity, we get a projective variety~$Z$ with isolated singularities such that $\sing Z\subset N\cup V \subset Z$. Finally, by replacing~$X$ by $$ (M\cup U) \cup (V\cup (Z\setminus \overline N))\bigg/ x\in U \sim \Phi (x) \in V $$ and by replacing the metric $g$ by any extension of $g| _M$ to the new manifold, we may assume that we have a holomorphic mapping $\Phi : X @>>> Z$ such that $\Phi (M) \subset N$ and $\Phi $ maps $(X\setminus M)\cup U$ biholomorphically onto $(Z\setminus N)\cup V$. In particular, since $X\setminus \overline M \subset\subset X$, it follows that $X$ is pseudoconcave in the sense of Andreotti. Now let~$H$ be a positive Hermitian holomorphic line bundle on~$Z$. Then $\Phi ^*H$ is semipositive on~$X$ and positive on $(X\setminus M)\cup U$. On the other hand, by shrinking~$M$ slightly and extending the Hermitian metric~$a$, we may assume that $K_X$ admits a Hermitian metric whose curvature is greater than or equal to $cg$ at each point of~$M$. It follows that if~$m$ is a sufficiently large positive integer and $L=K_X\otimes \Phi ^*H^m$, then~$L$ admits a Hermitian metric~$h$ such that $\cal C(L,h)\geq cg$ on~$X$. In particular, $g'=\cal C(L,h)$ is a complete K\"ahler metric on~$X$. Therefore, by the $L^2$ Riemann-Roch inequality of Nadel and Tsuji (Theorem~4.3), we have, for every sufficiently large positive integer~$\nu $, $$ 1+ \, \nu ^{-n} \dim H^0_{L^2}(X, {\cal O} (K_X\otimes L^\nu )) \geq \frac {1}{n!}\int _X \bigl( c_1(L,h)\bigr) ^n \geq c^n\pi ^{-n} \int _X \, dV_{g}; $$ (where the Hermitian metric in $K_X\otimes L^\nu $ is $(g')^*\otimes h^\nu $). Since, by Andreotti's finiteness theorem~[A] (or by~[AG]), the left-hand side is finite, we get $$ \text {\rm vol} \, _g(M)\leq \text {\rm vol} \, _g(X) <\infty . $$ \end{pf*} \begin{rem} By a version of the $L^2$~Riemann-Roch inequality due to Takayama~[T], it is only necessary to assume in the hypothesis~(ii) that $\cal C(K_M,a)\geq cg$ outside a relatively compact neighborhood of $\partial M$ in $X$. \end{rem} \bibliographystyle{amsplain}
1997-12-17T19:30:53
9712
alg-geom/9712002
en
https://arxiv.org/abs/alg-geom/9712002
[ "alg-geom", "math.AG" ]
alg-geom/9712002
Yuri Tschinkel
Victor V. Batyrev and Yu. Tschinkel
Tamagawa numbers of polarized algebraic varieties
54 pages, minor corrections
null
null
null
null
Let ${\cal L} = (L, \| \cdot \|_v)$ be an ample metrized invertible sheaf on a smooth quasi-projective algebraic variety $V$ defined over a number field. Denote by $N(V,{\cal L},B)$ the number of rational points in $V$ having ${\cal L}$-height $\leq B$. We consider the problem of a geometric and arithmetic interpretation of the asymptotic for $N(V,{\cal L},B)$ as $B \to \infty$ in connection with recent conjectures of Fujita concerning the Minimal Model Program for polarized algebraic varieties. We introduce the notions of ${\cal L}$-primitive varieties and ${\cal L}$-primitive fibrations. For ${\cal L}$-primitive varieties $V$ over $F$ we propose a method to define an adelic Tamagawa number $\tau_{\cal L}(V)$ which is a generalization of the Tamagawa number $\tau(V)$ introduced by Peyre for smooth Fano varieties. Our method allows us to construct Tamagawa numbers for $Q$-Fano varieties with at worst canonical singularities. In a series of examples of smooth polarized varieties and singular Fano varieties we show that our Tamagawa numbers express the dependence of the asymptotic of $N(V,{\cal L},B)$ on the choice of $v$-adic metrics on ${\cal L}$.
[ { "version": "v1", "created": "Mon, 1 Dec 1997 17:38:41 GMT" }, { "version": "v2", "created": "Wed, 17 Dec 1997 18:30:53 GMT" } ]
2007-05-23T00:00:00
[ [ "Batyrev", "Victor V.", "" ], [ "Tschinkel", "Yu.", "" ] ]
alg-geom
\section{Introduction} \bigskip Let $F$ be a number field (a finite extension of ${ \bf Q }$), ${\rm Val}(F)$ the set of all valuations of $F$, $F_v$ the $v$-adic completion of $F$ with respect to $v \in {\rm Val}(F)$, and $|\cdot |_v\, : \, F_v \rightarrow {\bf R }$ the $v$-adic norm on $F_v$ normalized by the conditions $| x|_v = |N_{F_v/{\bf Q}_p}(x)|_p$ for $p$-adic valuations $v \in {\rm Val}(F)$. Consider a projective space ${ \bf P }^m$ with standard homogeneous coordinates $(z_0,...,z_m)$ and a locally closed quasi-projective subvariety $V \subset { \bf P }^m$ defined over $F$ (we want to stress that $V$ is not assumed to be projective). Let $V(F)$ be the set of points in $V$ with coordinates in $F$. A {\bf standard height function} $H\,:\, { \bf P }^m(F) \rightarrow {\bf R }_{>0}$ is defined as follows $$ H(x):=\prod_{v \in {\rm Val}(F)} \max_{j=0, \ldots, m} \{ |z_j(x)|_v \}. $$ A basic fact about the standard height function $H$ claims that the set \[ \{x\in { \bf P }^m(F)\,:\, H(x)\le B\} \] is finite for any real number $B$ \cite{lang}. We set $$ N(V,B)=\#\{x\in V(F)\; : \; H(x)\le B\}. $$ It is an experimental fact that whenever one succeeds in proving an asymptotic formula for the function $N(V,B)$ as $B\rightarrow \infty$, one obtains the asymptotic \begin{equation} N(V,B)= c(V) B^{a(V)}(\log B)^{b(V)-1}(1+o(1)) \label{formula} \end{equation} with some constants $a(V) \in {\bf Q},$ $ b(V) \in \frac{1}{2}{\bf Z}$, and $c(V) \in {\bf R}_{>0}$. We want to use this observation as our starting point. It seems natural to ask the following: \medskip {\bf Question A.} {\em For which quasi-projective subvarieties $V \subset {\bf P}^m$ defined over $F$ do there exist constants $a(V) \in {\bf Q},$ $ b(V) \in \frac{1}{2}{\bf Z}$ and $c(V) \in {\bf R}_{>0}$ such that the asymptotic formula $(1)$ holds? } \medskip {\bf Question B.} {\em Does there exist a quasi-projective variety $V$ over $F$ with an asymptotic which is different from {\rm (\ref{formula})}?} \medskip In this paper we will be interested not in Questions $A$ and $B$ themselves but in a related to them another natural question: \bigskip \noindent {\bf Question C.} {\em Assume that $V$ is an irreducible quasi-projective variety over a number field $F$ such that the asymptotic formula {\rm (\ref{formula})} holds. How to compute the constants $a(V),b(V)$ and $c(V)$ in this formula via some arithmetical properties of $V$ over $F$ and geometrical properties of $V$ over ${\bf C}$? } \bigskip To simplify our terminology, it will be convenient for us to postulate: \bigskip \noindent {\bf Assumption.} For all quasi-projective $V', V$ with $V'\subset V\subset { \bf P }^m$ and $|V(F)|=\infty $ there exists the limit $$ \lim_{B\rightarrow \infty} \frac{N(V',B)}{N(V,B)}. $$ \bigskip The following definitions have been useful to us: \bigskip \noindent {\bf Definition ${\bf S_1}$.} A smooth irreducible quasi-projective subvariety $V \subset {\bf P}^m$ over a number field $F$ is called {\bf weakly saturated}, if $|V(F)| = \infty$ and if for any locally closed subvariety $W \subset V$ with ${\rm dim}\, W < {\rm dim}\, V$ one has \[ {\lim}_{B \rightarrow \infty} \frac{N(W,B)}{N(V,B)} < 1. \] \bigskip It is important to remark that Question C really makes sense {\em only for weakly saturated} varieties. Indeed, if there were a locally closed subvariety $W \subset V$ with ${\rm dim}\, W < {\rm dim}\, V$ and \[ {\lim}_{B \rightarrow \infty} \frac{N(W,B)}{N(V,B)} = 1, \] then it would be enough to answer Question C for each irreducible component of $W$ and for all possible intersections of these components (i.e., one could forget about the existence of $V$ and reduce the situation to a lower-dimensional case). In general, it is not easy to decide whether or not a given locally closed subvariety $V \subset {\bf P}^m$ is weakly saturated. We expect (and our assumption implies this) that the orbits of connected subgroups $G \subset PGL(m+1)$ are examples of weakly saturated varieties $V \subset {\bf P}^m$ (see \ref{equiv-sat}). \bigskip \noindent {\bf Definition} ${\bf S_2.}$ A smooth irreducible quasi-projective subvariety $V \subset {\bf P}^m$ with $|N(V,B)| = \infty$ is called {\bf strongly saturated}, if for all dense Zariski open subsets $U \subset V$, one has \[ {\lim}_{B \rightarrow \infty} \frac{N(U,B)}{N(V,B)} = 1. \] \bigskip First of all, if $V \subset {\bf P}^m$ is a strongly saturated subvariety, then for any locally closed subvariety $W \subset V$ with ${\rm dim}\, W < {\rm dim}\, V$, one has \[ {\lim}_{B \rightarrow \infty} \frac{N(W,B)}{N(V,B)} =0, \] i.e., $V$ is weakly saturated. On the other hand, if $V \subset {\bf P}^m$ is weakly saturated, but not strongly saturated, then there must be an infinite sequence $W_1, W_2, \ldots $ of pairwise different locally closed irreducible subvarieties $W_i \subset V$ with ${\rm dim}\, W_i < {\rm dim}\, V$ and $|W_i(F)| = \infty$ such that for an arbitrary positive integer $k$ one has \[ 0 < {\lim}_{B \rightarrow \infty} \frac{N(W_1 \cup \cdots \cup W_k,B)}{N(V,B)} < 1. \] Moreover, in this situation one can always choose the varieties $W_i$ to be strongly saturated (otherwise one could find $W_i' \subset W_i$ with ${\rm dim}\, W_i' < {\rm dim}\, W_i$ with the same properties as $W_i$ etc.). The strong saturatedness of each $W_i$ implies that \[ {\lim}_{B \rightarrow \infty} \frac{N(W_{i_1} \cap \cdots \cap W_{i_l},B)}{N(V,B)} = 0 \] for all pairwise different $i_1, \ldots, i_l$ and $l \geq 2$. In particular, one has \[ \sum_{i=1}^k {\lim}_{B \rightarrow \infty} \frac{N(W_i,B)}{N(V,B)} = {\lim}_{B \rightarrow \infty} \frac{N(W_1 \cup \cdots \cup W_k,B)}{N(V,B)} < 1\;\; \forall k >0. \] \medskip \noindent {\bf Definition} ${\bf F.}$ Let $V$ be a weakly saturated quasi-projective variety in ${\bf P}^m$ and $W_1, W_2, \ldots$ an infinite sequence of strongly saturated irreducible subvarieties $W_i$ having the property \[ 0 < \theta_i := {\lim}_{B \rightarrow \infty} \frac{N(W_i,B)}{N(V,B)} < 1\;\; \forall i > 0. \] We say that the set $\{W_1, W_2, \ldots \}$ forms an {\bf asymptotic arithmetic fibration} on $V$, if the following equality holds \[ \sum_{i=1}^{\infty} \theta_i = 1. \] \medskip The main purpose of this paper is to explain some geometric and arithmetic ideas concerning weakly saturated varieties and their asymptotic arithmetic fibrations by strongly saturated subvarieties. It seems that the cubic bundles considered in \cite{BaTschi4} are examples of such a fibration. We want to remark that most of the above terminology grew out of our attempts to restore a conjectural picture of the interplay between the geometry of algebraic varieties and the arithmetic of the distribution of rational points on them after we have found in \cite{BaTschi4} an example which contradicted general expectations formulated in \cite{BaMa}. \medskip In section 2 we consider smooth quasi-projective varieties $V$ over ${\bf C}$ together with a polarization ${\cal L} = (L, \| \cdot \|_h)$ consisting of an ample line bundle $L$ on $V$ equipped with a positive hermitian metric $\| \cdot \|_h$. Our main interest in this section is a discussion of geometric properties of $V$ in connection with the Minimal Model Program \cite{KMM} and its version for polarized algebraic varieties suggested by Fujita \cite{fujita0,fujita01,fujita1}. We introduce our main geometric invariants $\alpha_{\cal L}(V)$, ${\beta}_{\cal L}(V)$, and ${\delta}_{\cal L}(V)$ for an arbitrary ${\cal L}$-polarized variety $V$. It is important to remark that we will be only interested in the case $\alpha_{\cal L}(V) > 0$. The number $\alpha_{\cal L}(V)$ was first introduced in \cite{BaMa,Ba}, it equals to the opposite of the so called {\em Kodaira energy} (investigated by Fujita in \cite{fujita0,fujita01,fujita1}). Our basic geometric notion in the study of ${\cal L}$-polarized varieties $V$ with $\alpha_{\cal L}(V) >0$ is the notion of an ${\cal L}$-{\em primitive} variety. In Fujita's program for polarized varieties with negative Kodaira energy ${\cal L}$-{primitive} varieties play the same role as ${\bf Q}$-Fano varieties in Mori's program for algebraic varieties with negative Kodaira dimension. In particular, one expects the existence of so called ${\cal L}$-{\em primitive fibrations}, which are analogous to ${\bf Q}$-Fano fibrations in Mori's program. We show that on ${\cal L}$-primitive varieties there exists a canonical volume measure. Moreover, this measure allows us to construct a descent of hermitian metrics to the base of ${\cal L}$-primitive fibrations. Many geometric ideas of this section are inspired by \cite{BaMa,Ba}. In section 3 we introduce our main arithmetic notions of {\em weakly} and {\em strongly ${\cal L}$-saturated} varieties. Our first main diophantine conjecture claims that if an adelic ${\cal L}$-polarized quasi-projective algebraic variety $V$ over a number field $F$ is strongly ${\cal L}$-saturated, then the corresponding ${\cal L}$-polarized complex algebraic variety $V({\bf C})$ is ${\cal L}$-primitive. Moreover, we conjecture that if an adelic ${\cal L}$-polarized quasi-projective algebraic variety $V$ over a number field $F$ is weakly ${\cal L}$-saturated, then the corresponding ${\cal L}$-polarized complex algebraic variety $V({\bf C})$ admits an ${\cal L}$-primitive fibration having infinitely many fibers $W$ defined over $F$ which form an asymptotic arithmetic fibration. These conjectures allow us to establish a connection between the geometry of $V({\bf C})$ and the arithmetic of $V$. Following this idea, we explain a construction of an adelic measure on an arbitrary ${\cal L}$-primitive variety $V$ with ${\alpha}_{\cal L}(V) > 0$ and of the corresponding Tamagawa number $\tau_{\cal L}(V)$ as a regularized adelic integral of this measure. Our construction generalizes the definition of Tamagawa measures associated with a metrization of the canonical line bundle due to Peyre \cite{peyre}. We expect that for strongly ${\cal L}$-saturated varieties $V$ the number $\tau_{\cal L}(V)$ reflects the dependence of the constant $c(V)$ in the asymptotic formula (\ref{formula}) on the adelic metrization of the ample line bundle ${L}$. We discuss the natural question about the behavior of the adelic constant $\tau_{\cal L}(W)$ for fibers $W$ in ${\cal L}$-primitive fibrations on weakly ${\cal L}$-saturated varieties. In section 4 we show that our diophantine conjectures agree with already known examples of asymptotic formulas established for polarized algebraic varieties through the study of analytic properties of height zeta functions. In Section 5 we illustrate our expectations for the constants $a(V), b(V)$ and $c(V)$ in the asymptotics of $N(V,B)$ on some examples of smooth Zariski dense subsets $V$ in Fano varieties with singularities. We would like to thank J.-L. Colliot-Th\'el\`ene for his patience and encouragement. We are very grateful to B. Mazur, Yu. I. Manin, L. Ein and A. Chambert-Loir for their comments and suggestions. We thank the referee for several useful remarks. \section{Geometry of ${\cal L}$-polarized varieties} \subsection{${\cal L}$-closure } Let $V$ be a smooth irreducible quasi-projective algebraic variety over ${\bf C}$, $V({\bf C})$ the set of closed points of $V$, $L$ an ample invertible sheaf on $V$, i.e., $L^{\otimes k} = i^* {\cal O}_{{\bf P}^m}(1)$ for some $k >0$ and some embedding $i \,: \, V \hookrightarrow {\bf P}^m$. Since we don't assume $V$ to be compact, the invertible sheaf $L$ on $V$ itself contains too little information about the embedding $i \,: \, V \hookrightarrow {\bf P}^m$. For instance, let $V$ be an affine variety of positive dimension. Then the space of global sections of $L$ is infinite dimensional and we don't know anything about the projective closure of $V$ in ${\bf P}^m$ even though we know that the invertible sheaf $L^{\otimes k}$ is isomorphic to $i^*{\cal O}_{{\bf P}^m}(1)$. This situation changes if one considers $L$ together with a positive hermitian metric, i.e., an ample metrized invertible sheaf ${\cal L}$ associated with $L$. Let us choose a positive hermitian metric $h$ on ${\cal O}_{{\bf P}^m}(1)$ (e.g. Fubini-Study metric) and denote by $\|\cdot \|_h$ the induced metric on $L^{\otimes k}$. Thus we obtain a metric $\| \cdot \|$ on $L$ by putting $\|s(x)\|: = \| s^k(x) \|_h^{1/k}$ for any $x \in V({\bf C})$ and any section $s \in { \rm H }^0(U, L)$ over an open subset $U \subset V$. \begin{dfn} {\rm We call a pair ${\cal L} = ( L, \|\cdot \|)$ {\bf an ample metrized invertible sheaf} associated with $L$. We denote by ${\cal L}^{\otimes \nu}$ the pair $( L^{\otimes \nu}, \|\cdot \|^{\nu })$.} \end{dfn} Our next goal is to show that an ample metrized invertible sheaf contains almost complete information about the projective closure of $V$ in ${\bf P}^m$. \begin{dfn} {\rm Let ${\cal L} =( L, \|\cdot \|) $ be an ample metrized invertible sheaf on a complex irreducible quasi-projective variety $V$. We denote by $$ { \rm H }^0_{\rm bd}(V, {\cal L}) $$ the subspace of ${ \rm H }^0(V, L)$ consisting of those global sections $s$ of $L$ over $V$ such that the corresponding continuous function $x \mapsto \|s(x) \|$ $(x \in V({\bf C}))$ is globally bounded on $V({\bf C})$ from above by a positive constant $C(s)$ depending only on $s$. We call ${ \rm H }^0_{\rm bd}(V, {\cal L})$ the {\bf space of globally bounded sections} of ${\cal L}$. } \end{dfn} \begin{prop} Let $\overline{V}$ be the normalization of the projective closure of $V$ with respect to the embedding $i\; : \; V \hookrightarrow {\bf P}^m$ with ${L}^{\otimes k} = i^*{\cal O}_{{\bf P}^m}(1)$. Denote by ${e}\, : \,\overline{V} \rightarrow {\bf P}^m$ the corresponding finite projective morphism. Then one has a natural isomorphism $$ { \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes k}) \cong { \rm H }^0(\overline{V}, {e}^*{\cal O}_{{\bf P}^m}(1)). $$ \label{l-bd} \end{prop} \noindent {\em Proof.} Since $\overline{V}({\bf C})$ is compact, the continuous function $x \mapsto \|s(x)\|$ is globally bounded on $\overline{V}({\bf C})$ for any $s \in { \rm H }^0(\overline{V}, {e}^*{\cal O}_{{\bf P}^m}(1))$. Therefore, we obtain that ${ \rm H }^0(\overline{V}, {e}^*{\cal O}_{{\bf P}^m}(1))$ is a subspace of ${ \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes k})$. Now let $f \in { \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes k})$ be a globally bounded on $V({\bf C})$ section of $i^*{\cal O}_{{\bf P}^m}(1)$. Since ${ \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes k})$ is a subspace of ${ \rm H }^0(V, L^{\otimes k})$, the section $f$ uniquely extends to a global meromorphic section $\overline{f} \in { \rm H }^0(\overline{V}, {e}^*{\cal O}_{{\bf P}^m}(1))$. Since a bounded meromorphic function is holomorphic, $\overline{f}$ is a global regular section of ${e}^*{\cal O}_{{\bf P}^m}(1)$ (we apply the theorem of Riemann to some resolution of singularities $\rho\, : \, X \rightarrow \overline{V}$ and use the fact that $\rho_* {\cal O}_X = {\cal O}_{\overline{V}}$). Thus we have $$ { \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes k}) \subset { \rm H }^0(\overline{V}, {e}^*{\cal O}(1)). $$ \hfill $\Box$ \begin{dfn} {\rm We define a the graded ${\bf C}$-algebra $$ {\rm A}(V, {\cal L}) = \bigoplus_{ \nu \geq 0} { \rm H }^0_{\rm bd}(V, {\cal L}^{\otimes \nu } ). \] } \end{dfn} Using \ref{l-bd}, one immediately obtains: \begin{coro} The graded algebra $ {\rm A}(V, {\cal L})$ is finitely generated. \end{coro} \begin{dfn} {\rm We call the normal projective variety \[ \overline{V}^{\cal L} = {\rm Proj}\, {\rm A}(V, {\cal L}). \] the ${\cal L}$-{\bf closure} of $V$ with respect to an ample metrized invertible sheaf ${\cal L}$. } \end{dfn} \begin{rem} {\rm By \ref{l-bd}, $\overline{V}^{\cal L}$ is isomorphic to $\overline{V}$. Therefore, we have obtained a way to define the normalization of the projective closure of $V$ with respect to an $L^{\otimes k}$-embedding via a notion of an ample metrized invertible sheaf ${\cal L}$ on $V$.} \end{rem} \subsection{Kodaira energy and $\alpha_{\cal L}(V)$} Let $X$ be a normal irreducible algebraic variety of dimension $n$. We denote by ${\rm Div}(X)$ (resp. by ${\rm Z}_{n-1}(X)$) the group of Cartier divisors (resp. Weil divisors) on $X$. An element of ${\rm Div}(X) \otimes {\bf Q}$ (resp.${\rm Z}_{n-1} (X) \otimes {\bf Q}$) is called a ${\bf Q}$-Cartier divisor (resp. a ${\bf Q}$-divisor). By $K_X$ we denote a divisor of a meromorphic differential $n$-form on $X$, where $K_X$ is considered as an element of ${\rm Z}_{n-1}(X)$. \begin{dfn} {\rm Let $X$ be a projective variety and $L$ be an invertible sheaf on $X$. The {\bf Iitaka-dimension} $\kappa(L)$ is defined as \[ \kappa(L) = \left\{ \begin{array}{ll} - \infty & \mbox{\rm if ${ \rm H }^0(X, L^{\otimes \nu}) = 0$ for all $\nu > 0$} \\ {\rm Max}\, \dim \phi_{L^{\otimes \nu}}(X) : & { \rm H }^0(X, L^{\otimes \nu}) \neq 0 \end{array} \right. \] where $\phi_{L^{\otimes \nu}}(X)$ is the closure of the image of $X$ under the rational map \[ \phi_{L^{\otimes \nu}} \, : \, X \rightarrow {\bf P}({ \rm H }^0(X, L^{\otimes \nu})). \] A Cartier divisor $L$ is called {\bf semi-ample} (resp. {\bf effective}), if $L^{\otimes \nu}$ is generated by global sections for some $\nu > 0$ (resp. $\kappa(L) \geq 0)$. } \end{dfn} \begin{rem} {\rm The notions of Iitaka-dimension, ampleness and semi-ample\-ness obviously extend to ${\bf Q}$-Cartier divisors. Let $L$ be a Cartier divisor. Then for all $ \kappa_1, k_2 \in {\bf N}$ we set \[ \kappa(L^{\otimes k_1/k_2} ): = \kappa(L),\] \[ \mbox{\rm $L^{\otimes k_1/k_2}$ is ample} \, \Leftrightarrow \, \mbox{\rm $L$ is ample}, \] and \[ \mbox{\rm $L^{\otimes k_1/k_2}$ is semi-ample} \, \Leftrightarrow \, \mbox{\rm $L$ is semi-ample}. \] } \end{rem} \begin{dfn} {\rm Let $X$ be a smooth projective variety. We denote by ${ \rm NS}(X)$ the group of divisors on $X$ modulo numerical equivalence and set ${ \rm NS}(X)_{\bf R} = { \rm NS}(X) \otimes {\bf R}$. By $[L]$ we denote the class of a divisor $L$ in ${ \rm NS}(X)$. The {\bf cone of effective divisors} $\Lambda_{\rm eff}(X) \subset { \rm NS}(X)_{\bf R}$ is defined as the closure of the subset \[ \bigcup_{ \kappa(L) \geq 0} {\bf R}_{\geq 0} [L] \subset { \rm NS}(X)_{\bf R}. \] } \end{dfn} \begin{dfn} {\rm Let $V$ be a smooth quasi-projective algebraic variety with an ample metrized invertible sheaf ${\cal L}$, $\overline{V}^{\cal L}$ the ${\cal L}$-closure of $V$ and $\rho$ some resolution of singularities \[ \rho \; : \; X \rightarrow \overline{V}^{\cal L}. \] We define the number \[ \alpha_{\cal L}(V) = \inf \{ t \in {\bf Q}\; : \; t [ \rho^*L] + [ K_X] \in \Lambda_{\rm eff}(X) \}. \] and call it the ${\cal L}$-{\bf index} of $V$. } \end{dfn} \begin{rem} {\rm It is easy to see that the ${\cal L}$-{index} does not depend on the choice of $\rho$.} \end{rem} \begin{rem} {\rm The ${\cal L}$-index $\alpha_{\cal L}(V)$ for smooth projective varieties $V$ was first introduced in \cite{BaMa} and \cite{Ba}. We remark that the opposite number $- \alpha_{\cal L}(V)$ coincides with the notion of {\bf Kodaira energy} introduced and investigated by Fujita in \cite{fujita0,fujita01,fujita1}: \[ \kappa\epsilon (V, L) = - \alpha_{\cal L}(V) = - \inf \{ t \in {\bf Q}\; : \; \kappa ( (L)^{\otimes t} \otimes K_V) \geq 0 \}. \] From the viewpoint of our diophantine applications it is much more natural to consider $\alpha_{\cal L}(V)$ instead of its opposite $-\alpha_{\cal L}(V)$. The only reason that we could see for introducing the number $-\alpha_{\cal L}(V)$ instead of $\alpha_{\cal L}(V)$ is some kind of compatibility between the notions of {\em Kodaira energy} and {\em Kodaira dimension}, e.g. Kodaira energy must be positive (resp. negative) iff the Kodaira dimension is positive (resp. negative). } \end{rem} \noindent The following statement was conjectured in \cite{BaMa} (see also \cite{Ba,fujita00,fujita000}): \begin{conj} {\sc (Rationality)} Assume that $\alpha_{\cal L}(V)> 0$. Then $\alpha_{\cal L}(V)$ is rational. \end{conj} \begin{rem} {\rm It was shown in \cite{Ba} that this conjecture follows from the Minimal Model Program. In particular, it holds for ${\rm dim}\,V \leq 3$. If ${\rm dim}\, V =1$, then the only possible values of $\alpha_{\cal L}(V)$ are numbers $2/k$ with $(k \in {\bf N})$. If ${\rm dim}\, V =2$, then $\alpha_{\cal L}(V) \in \{ 2/k, 3/l \}$ with $( k, l \in {\bf N})$. } \label{a-values} \end{rem} \begin{dfn} {\rm A normal irreducible algebraic variety $W$ is said to have at worst {\bf canonical} (resp. {\bf terminal}) singularities if $K_W$ is a ${\bf Q}$-Cartier divisor and if for some (or every) resolution of singularities \[ \rho \; : \; X \rightarrow W \] one has \[ K_X = \rho^*(K_W)\otimes {\cal O}(D) \] where $D$ is an effective ${\bf Q}$-Cartier divisor (resp. the support of the effective divisor $D$ coincides with the exceptional locus of $\rho$). Irreducible components of the exceptional locus of $\rho$ which are not contained in the support of $D$ are called {\bf crepant divisors} of the resolution $\rho$.} \end{dfn} \begin{dfn} {\rm A normal irreducible algebraic variety $W$ is called a {\bf ca\-no\-nical} ${\bf Q}$-{\bf Fano} variety, if $W$ has at worst { canonical} singularities and $K_W^{-1}$ is an ample ${\bf Q}$-Cartier divisor. A maximal positive rational number $r(W)$ such that $K_W^{-1} = L^{\otimes r(W)}$ for some Cartier divisor $L$ is called the {\bf index} of a canonical ${\bf Q}$-Fano variety $W$ (obviously, one has $r(W) = \alpha_{\cal L}(W)$ for some positive metric on $L$). } \label{fano-c} \end{dfn} The following conjecture is due to Fujita \cite{fujita0}: \begin{conj} {\sc (Spectrum Conjecture)} Let $S(n)$ be the set all possible values of $\alpha_{\cal L}(V)$ for smooth quasi-projective algebraic varieties $V$ of dimension $\leq n$ with an ample metrized invertible sheaf ${\cal L}$. Then for any $\varepsilon > 0$ the set \[ \{ \alpha_{\cal L}(V) \in S(n) \; : \; \alpha_{\cal L}(V) > \varepsilon \} \] is finite. \end{conj} This conjecture follows from the Minimal Model Program \cite{KMM} and from the following conjecture on the boundedness of index for Fano varieties with canonical singularities: \begin{conj} {\sc (Boundedness of Index)} The set of possible values of index $r(W)$ for canonical ${\bf Q}$-Fano varieties $W$ of dimension $n$ is finite. \end{conj} In particular, both conjectures are true for ${\bf Q}$-Fano varieties of dimension $n \leq 3$ \cite{A,fujita0,fujita01,fujita1,Ka,shin}. \subsection{${\cal L}$-primitive varieties} \begin{dfn} {\rm Let $X$ be a projective algebraic variety. We call an effective ${\bf Q}$-divisor $D$ {\bf rigid}, if $\kappa(D) = 0$. } \end{dfn} \begin{prop} {\rm An effective ${\bf Q}$-divisor $D$ on $X$ is rigid if and only if there exist finitely many irreducible subvarieties $D_1, \ldots, D_l \subset X$ $(l \geq 0)$ of codimension $1$ such that $D =r_1 D_1 + \ldots + r_l D_l$ with $r_1, \ldots, r_l \in {\bf Q}_{>0}$ and \[ \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_1 D_1 + \ldots + n_l D_l)) =1 \;\; \forall\; (n_1, \ldots, n_l ) \in {\bf Z}^l_{\geq 0}. \] } \label{rigid2} \end{prop} \noindent {\em Proof.} Let $D$ be rigid. Take a positive integer $m_0$ such that $m_0D$ is a Cartier divisor and ${\rm dim}\, { \rm H }^0(X, {\cal O}(m_0D) ) =1$. Denote by $D_1, \ldots, D_l$ the irreducible components of the divisor $(s)$ of a non-zero section $s \in { \rm H }^0(X, {\cal O}(mD) )$. One has \[ (s) = m_1 D_1 + \cdots + m_l D_l\;\;\; m_1, \ldots, m_l \in {\bf N}. \] Since ${\cal O}(D_i)$ admits at least one global non-zero section we obtain that \[ \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_1' D_1 + \ldots + n_l' D_l)) \geq \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_1 D_1 + \ldots + n_l D_l)), \] whenever $n_1 \geq n_1', \ldots, n_l \geq n_l'$ for $(n_1', \ldots, n_l' ), \; (n_1, \ldots, n_l ) \in {\bf Z}^l_{\geq 0}$. This implies that \[ \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_1 D_1 + \ldots + n_l D_l)) \geq 1\;\; \forall\; (n_1, \ldots, n_l ) \in {\bf Z}^l_{\geq 0}. \] On the other hand, for any $(n_1, \ldots, n_l ) \in {\bf Z}^l_{\geq 0}$ there exists a positive integer $n_0$ such that $n_0m_1 \geq n_1, \ldots, n_0m_l \geq n_l$. Therefore, \[ \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_1 D_1 + \ldots + n_l D_l)) \leq \mbox {\rm dim}\, { \rm H }^0 (X, {\cal O}(n_0m_0D) ) =1, \] since $\kappa(n_0m_0D) =\kappa(D) =0$. \hfill $\Box$ \begin{coro} Let $D_1, \ldots, D_l \subset X$ be all irreducible components of the support of a rigid ${\bf Q}$-Cartier divisor $D$. Then a linear combination \[ n_1 D_1 + \ldots + n_l D_l, \; \; n_1, \ldots, n_l \in {\bf Z} \] is a principal divisor, iff $n_1 = \cdots = n_l = 0$. \label{rigid-l} \end{coro} \noindent {\em Proof.} Assume that $n_1 D_1 + \ldots + n_l D_l$ is linearly equivalent to $0$. Then the effective Cartier divisor $D_0 = \sum_{n_i \geq 0} n_i D_i$ is linearly equivalent to the effective Cartier divisor $D_0' = \sum_{n_j <0} (-n_j) D_j$. Since $D_0$ and $D_0'$ have different supports we have ${\rm dim}\, { \rm H }^0(X, {\cal O}(D_0)) \geq 2$. Contradiction to \ref{rigid2}. \hfill $\Box$ \begin{dfn} {\rm Let $V$ be a smooth quasi-projective algebraic variety with an ample metrized invertible sheaf ${\cal L}$ and $\overline{V}^{\cal L}$ the projective ${\cal L}$-closure of $V$. The variety $V$ is called ${\cal L}$-{\bf primitive}, if the number $\alpha_{\cal L}(V)$ is rational and if for some resolution of singularities \[ \rho \; : \; X \rightarrow \overline{V}^{\cal L} \] one has $\rho^*(L)^{\otimes \alpha_{\cal L}(V)} \otimes K_X = {\cal O}(D)$, where $D$ is a rigid effective ${\bf Q}$-Cartier divisor on $X$. } \end{dfn} \begin{rem} {\rm It is easy to see that the notion of an ${\cal L}$-primitive variety doesn't depend on the choice of a resolution of singularities $\rho$. Since $V$ is smooth, we can always assume that the natural mapping $$\rho\; : \; \rho^{-1}(V) \rightarrow V$$ is an isomorphism. } \end{rem} \begin{exam} {\rm Let $V_1$ and $V_2$ be two smooth quasi-projective varieties with ample metrized invertible sheaves ${\cal L}_1$ and ${\cal L}_2$ (resp. on $V_1$ and $V_2$). Assume that $V_1$ (resp. $V_2$ ) is ${\cal L}_1$-primitive (resp. ${\cal L}_1$-primitive). Then the product $V = V_1 \times V_2$ is ${\cal L}$-primitive, where ${\cal L} = \pi_1^*{\cal L}_1 \otimes \pi_2^*{\cal L}_2$. } \end{exam} \noindent Our main list of examples of ${\cal L}$-primitive varieties is obtained from canonical ${\bf Q}$-Fano varieties: \begin{exam} {\rm Let $V$ be the set of nonsingular points of a canonical ${\bf Q}$-Fano variety $W$ with an ample metrized invertible sheaf ${\cal L} = (L,\|\cdot\|)$ such that $K_W^{-1} = L^{\otimes r(W)}$ ( $W = \overline{V}^{\cal L} $). Then $V$ is an ${\cal L}$-primitive variety with ${\cal L}$-index $r(W)$. Indeed, let $\rho\, : \, X \rightarrow W$ be a resolution of singularities. By \ref{fano-c}, we have $\rho^*(L)^{\otimes r(W)} \otimes K_X = {\cal O}(D)$, where $D$ is an effective ${\bf Q}$-Cartier divisor. Since the support of $D$ consists of exceptional divisors with respect to $\rho$, $D$ is rigid (see \ref{rigid2}). } \end{exam} We expect that the above examples cover all ${\cal L}$-primitive varieties: \begin{conj} {\sc (Canonical ${\bf Q}$-Fano contraction)} Let $V$ be an ${\cal L}$-primitive variety with $\alpha_{\cal L}(V) >0$. Then there exists a resolution of singularities $\rho\, : \, X \rightarrow \overline{V}^{\cal L}$ and a birational projective morphism $\pi\, : \, X \rightarrow W$ to a canonical ${\bf Q}$-Fano variety $W$ such that $\pi^*K_W^{-1} \cong \rho^*(L)^{\alpha_{\cal L}(V)}$ (i.e., $\alpha_{\cal L}(V) = r(W)$) and the support of $D$ ($\rho^*(L)^{\otimes r(W)} \otimes K_X = {\cal O}(D)$) is contained in the exceptional locus of $\pi$. \label{conj-cont} \end{conj} The above conjecture is expected to follow from the Minimal Model Program using the existence and termination of flips (in particular, it holds for toric varieties). The following statement will be important in our construction of Tamagawa numbers for ${\cal L}$-primitive varieties defined over a number field: \begin{conj} {\sc (Vanishing)} {\rm For $V$ an ${\cal L}$-primitive variety such that $\alpha_{\cal L}(V) >0$ we have \[ {\rm h}^i(X, {\cal O}_X) = 0 \;\; \forall \; i > 0 \] for any resolution of singularities $$ \rho \; : \; X \rightarrow \overline{V}^{\cal L} $$ such that the support of the ${\bf Q}$-Cartier divisor $\rho^*(L)^{\otimes \alpha_{\cal L}(V) } \otimes K_X$ is a ${\bf Q}$-Cartier divisor with normal crossings. In particular, $ {\rm Pic}(X)$ is a finitely generated abelian group and one has a canonical isomorphism \[ {\rm Pic}(X) \otimes {\bf Q} \cong { \rm NS}(X) \otimes {\bf Q}. \] \label{vanish} } \end{conj} \begin{rem} {\rm Theorem 1-2-5 in \cite{KMM} implies the vanishing for the structure sheaf for ${\bf Q}$-Fano varieties with canonical singularities (even with log-terminal singularities). All canonical (and log-terminal singularities) are rational and it follows that the higher cohomology of the structure sheaf on any desingularization of a canonical or a log-terminal Fano variety must also vanish (by Leray spectral sequence). Therefore, we would obtain the vanishing \ref{vanish} for all ${\cal L}$-primitive varieties which are birationally equivalent to a Fano variety with at worst log-terminal singularities. The existence of a canonical ${\bf Q}$-contraction \ref{conj-cont} would insure this. \label{vanish-rem} } \end{rem} \begin{dfn} {\rm Let $V$ be an ${\cal L}$-primitive variety with $\alpha_{\cal L}(V) >0$, $\rho \, : \, X \rightarrow \overline{V}^{\cal L}$ any resolution of singularities, $D_1, \ldots, D_l$ irreducible components of the support of the rigid effective ${\bf Q}$-Cartier divisor $D$ with ${\cal O}(D) = \rho^*(L)^{\otimes \alpha_L(V)} \otimes K_X$. We shall call $$ {\rm Pic}(V, {\cal L}) : = {\rm Pic}(X \setminus \bigcup_{i =1}^l D_i) $$ the ${\cal L}$-{\bf Picard group} of $V$. The number \[ \beta_{\cal L}(V): = {\rm rk}\, {\rm Pic}(V, {\cal L}) \] will be called the ${\cal L}$-{\bf rank} of $V$. We define the ${\cal L}$-{\bf cone of effective divisors} $\Lambda_{\rm eff}(V, {\cal L}) \subset {\rm Pic}(V,{\cal L})\otimes {\bf R}$ as the image of $\Lambda_{\rm eff}(X) \subset { \rm NS}(X)_{\bf R} = {\rm Pic}(X) \otimes {\bf R}$ under the natural surjective ${\bf R}$-linear mapping \[ \tilde{\rho}\; : \;{\rm Pic}(X) \otimes {\bf R} \rightarrow {\rm Pic}(V,{\cal L})\otimes {\bf R}. \] } \end{dfn} \begin{rem} {\rm By \ref{rigid-l}, one obtains the exact sequence \begin{equation} 0 \rightarrow {\bf Z}[D_1] \oplus \cdots \oplus {\bf Z}[D_l] \rightarrow {\rm Pic}(X) \stackrel{\tilde{\rho}}{\rightarrow} {\rm Pic}(V, {\cal L}) {\rightarrow} 0 \end{equation} and therefore \[ \beta_{\cal L}(V)= {\rm rk}\, {\rm Pic} (X) - l. \] Using these facts, it is easy to show that the group $ {\rm Pic}(V, {\cal L})$ and the cone $\Lambda_{\rm eff}(V, {\cal L})$ do not depend on the choice of a resolution of singularities $\rho \, : \,X \rightarrow \overline{V}^{\cal L}$. } \end{rem} \noindent The above conjecture holds in dimension $n \leq 3$ as a consequence of the Minimal Model Program. More precisely, it is a consequence of Conjecture \ref{conj-cont} and the following weaker statement: \begin{conj} {\sc (Polyhedrality)} Let $V$ be an ${\cal L}$-primitive variety with $\alpha_{\cal L}(V) >0$. Then $\Lambda_{\rm eff}(V, {\cal L})$ is a rational finitely generated polyhedral cone. \label{polyhed} \end{conj} \medskip \begin{dfn}{\rm Let $ (A,A_{{\bf R }}, \Lambda )$ be a triple consisting of a finitely generated abelian group $A$ of rank $k$, a $k$-dimensional real vector space $A_{{\bf R }}=A\otimes {\bf R }$ and a convex $k$-dimensional finitely generated polyhedral cone $ \Lambda \in A_{{\bf R }}$ such that $ \Lambda \cap - \Lambda =0\in A_{{\bf R }}$. For ${\rm Re}({\bf s})$ contained in the interior of the cone $ \Lambda $ we define the ${\cal X}$-function of $ \Lambda $ by the integral $$ {\cal X}_{ \Lambda }({\bf s}):= \int_{ \Lambda ^*}e^{-<{\bf s},{\bf y}>}{\bf d}{\bf y} $$ where $ \Lambda ^*\in A^*_{{\bf R }}$ is the dual cone to $ \Lambda $ and ${\bf d}{\bf y}$ is the Lebesgue measure on $A^*_{{\bf R }}$ normalized by the dual lattice $A^*\subset A^*_{{\bf R }}$ where $A^* := {\rm Hom}(A, {\bf Z})$. } \end{dfn} \begin{rem}{\rm If $ \Lambda $ is a finitely generated rational polyhedral cone the function ${\cal X}_{ \Lambda }({\bf s})$ is a rational function in ${\bf s}$. However, the explicit determination of this function might pose serious computational problems. } \end{rem} \begin{dfn} {\rm Let $V$ be an ${\cal L}$-primitive smooth quasi-projective algebraic variety with a metrized invertible sheaf ${\cal L}$ and $ \alpha _{\cal L}(V)>0$. Let $X$ be any resolution of singularities $\rho\;:\; X\rightarrow \overline{V}^{\cal L}$. We consider the triple $$ ({\rm Pic}(V,{\cal L}),{\rm Pic}(V,{\cal L})_{{\bf R }}, \Lambda _{\rm eff}(V,{\cal L})) $$ and the corresponding ${\cal X}$-function. Assuming that ${\rm Pic}(V, {\cal L})$ is a finitely generated abelian group (cf. \ref{vanish}) and that $ \Lambda _{\rm eff}(V,{\cal L})$ is a polyhedral cone (cf. \ref{polyhed}), we define the constant $\gamma_{\cal L}(V)\in { \bf Q }$ by $$ \gamma_{\cal L}(V):= {\cal X}_{ \Lambda _{\rm eff}(V,{\cal L})}(\tilde{\rho}(-[K_X])). $$ \label{gamma-dfn} } \end{dfn} \subsection{${\cal L}$-primitive fibrations and descent of metrics} Let $V$ be an ${\cal L}$-primitive variety of dimension $n$. We show that there exists a canonical measure on $V({\bf C})$ which is uniquely defined up to a positive constant. In order to construct this measure we choose a resolution of singularities $\rho \; : \; X \rightarrow \overline{V}^{\cal L}$ and a positive integer $k_2$ such that $k_2D$ is a Cartier divisor, where ${\cal O}(D) \cong (\rho^*L)^{\otimes \alpha_{\cal L}(V) } \otimes K_X$. Then $k_1 = k_2 \alpha_{\cal L}(V)$ is a positive integer. Let $g \in { \rm H }^0(X, {\cal O}(k_2D))$ be a non-zero global section (by \ref{rigid2}, it is uniquely defined up to a non-zero constant). We define a measure ${\bf \omega}_{{\cal L}}(g)$ on $X({\bf C})$ as follows. Choose local complex analytic coordinates $z_{1}, \ldots, z_{n}$ in some open neighborhood $U_x \subset X({\bf C})$ of a point $x \in V({\bf C})$. We write the restriction of the global section $g$ to $U_x$ as \[ g = s^{k_2\alpha_{\cal L}(V)} ( dz_{1} \wedge \cdots \wedge dz_{n})^{\otimes k_2} = s^{k_1} ( dz_{1} \wedge \cdots \wedge dz_{n})^{\otimes k_2}, \] where $s$ is a local section of $L$ . Then we set \[ {\bf \omega}_{{\cal L}}(g) : = \left(\frac{\sqrt{-1}}{2} \right)^n \|s\|^{ \alpha_{\cal L}(V)} (dz_1 \wedge d\overline{z}_1) \wedge \cdots \wedge (dz_n \wedge d\overline{z}_n). \] By a standard argument, one obtains that ${\bf \omega}_{{\cal L}}(g)$ doesn't depend on the choice of local coordinates in $U_x$ and that it extends to the whole complex space $X({\bf C})$. It remains to notice that the restriction of the measure ${\bf \omega}_{{\cal L}}(g)$ to $V({\bf C}) \subset X({\bf C})$ does not depend on the choice of $\rho$. So we obtain a well-defined measure on $V({\bf C})$. \begin{rem} {\rm We note that the measure $\omega_{\cal L}(g)$ depends on the choice of $g \in { \rm H }^0(X, {\cal O}(k_2D))$. More precisely, it multiplies by $|c|^{1/k_2}$ if we multiply $g$ by some non-zero complex number $c$. Thus we obtain that the mapping \[ \| \cdot \|_{\cal L} \; : \; { \rm H }^0(X, {\cal O}(k_2D)) \rightarrow {\bf R}_{\geq 0} \] \[ g \mapsto \int_{X({\bf C})} \omega_{\cal L}(g) = \int_{V({\bf C})} \omega_{\cal L}(g) \] satisfies the property \[ \| c g \|_{\cal L} = |c|^{1/k_2} \| g \|_{\cal L}\;\; \forall c \in {\bf C}^*. \] } \end{rem} \begin{dfn} {\rm Let $V$ be a smooth quasi-projective variety with an ample metrized invertible sheaf ${\cal L}$, $\rho \, : \,X \rightarrow \overline{V}^{\cal L}$ a resolution of singularities. A regular projective morphism $\pi\, :\, X \rightarrow Y$ to a projective variety $Y$ $({\rm dim}\, Y < {\rm dim}\, X)$ is called an ${\cal L}$-{\bf primitive fibration on} $V$ if there exists a Zariski dense open subset $U \subset Y$ such that the following conditions are satisfied: {(i)} for any point $y \in U({\bf C})$ the fiber $V_y = \pi^{-1}({y}) \cap V$ is a smooth quasi-projective ${\cal L}$-primitive subvariety; (ii) $\alpha_{\cal L}(V) = \alpha_{\cal L}(V_y) > 0$ for all $y \in U({\bf C})$; (iii) for any $k \in {\bf N}$ such that $k\alpha_{\cal L}(V) \in {\bf Z}$, \[ L_k := {\rm R}^0\pi_* \left(\rho^*(L)^{\otimes \alpha_L(V)} \otimes K_X \right)^{\otimes k} \] is an ample invertible sheaf on $Y$. \label{prim-fb}} \end{dfn} \noindent We propose the following version of the Fibration Conjecture of Fujita (see \cite{fujita0}): \begin{conj} {\sc (Existence of Fibrations)} Let $V$ be an arbitrary smooth quasi-projective variety with an ample metrized invertible sheaf ${\cal L}$ and $\alpha_{\cal L}(V)> 0$. Then there exists a resolution of singularities $\rho \, : \,X \rightarrow \overline{V}^{\cal L}$ such that $X$ admits an ${\cal L}$-primitive fibration $\pi\, :\, X \rightarrow Y$ on some dense Zariski open subset $V' \subset V$. \label{conj-fb} \end{conj} \begin{rem} {\rm From the viewpoint of the Minimal Model Program, Conjecture \ref{conj-fb} is equivalent to the statement about the conjectured existence of ${\bf Q}$-Fano fibrations for algebraic varieties of negative Kodaira-dimension (cf. \cite{KMM}). The existence of an ${\cal L}$-primitive fibration is equivalent to the fact that the graded algebra \[ {\rm R}(V, {\cal L}) = \oplus_{\nu \geq 0} { \rm H }^0(X, M^{k \nu}), \;\; M:= \rho^*(L)^{\otimes \alpha_L(V)} \otimes K_X \] is finitely generated (cf. 2.4 in \cite{BaMa}). One can define $Y$ as ${\rm Proj}\,{\rm R}(V, {\cal L})$ and $X$ as a common resolution of singularities of $\overline{V}^{\cal L}$ and of the indeterminacy locus and generic fiber of the natural rational map $\overline{V}^{\cal L} \rightarrow Y$ (cf. \cite{H}). } \label{proj-fb} \end{rem} It is important to observe that a metric $\| \cdot \|$ on $L$ induces natural metrics on all ample invertible sheaves $L_k$ on $Y$: \begin{dfn} {\rm Let $L_k$ be an ample invertible sheaf on $Y$ as above. We define a metric $\| \cdot \|_{{\cal L}, k}$ on $L_k$ as follows. Let $y \in Y({\bf C})$ be a closed point, $U \subset Y$ be a Zariski open subset containing $y$, and $s \in { \rm H }^0(U, L_k)$ is a section with $s(y) \neq 0$. Then we set \[ \|s(y)\|_{{\cal L},k} := \left( \int_{V_y({\bf C})} \omega_{\cal L}(\pi^* s) \right)^k , \] where $V_y({\bf C})$ is the fiber over $y$ of the ${\cal L}$-primitive fibration $\pi^*$, $\pi^* s$ the $\pi$-pullback of $s$ restricted to ${\cal L}$-primitive variety $V_y({\bf C})$, and $\omega_{\cal L}(\pi^* s)$ the corresponding to $\pi^* s$ volume measure on $V_y({\bf C})$. We call $\| \cdot \|_{{\cal L}, k}$ a $k$-{\bf adjoint descent} to $Y$ of a metric $\| \cdot \|$ on $L$. } \end{dfn} \section{Heights and asymptotic formulas} \subsection{Basic terminology and notations} Let $F$ be a number field, ${\cal O}_F \subset F$ the ring of integers in $F$, ${\rm Val}(F)$ the set of all valuations of $F$, $F_v$ the completion of $F$ with respect to a valuation $v \in {\rm Val}(F)$, ${\rm Val}(F)_{\infty} = \{ v_1, \ldots, v_r \}$ the set of all archimedean valuations of $F$. For any algebraic variety $X$ over a field $F$ we denote by $X(F)$ the set of its $K$-rational points. \begin{dfn} {\rm Let $E$ be a vector space of dimension $m+1$ over $F$, ${\cal O}_E \subset E$ a projective ${\cal O}_F$-module of rank $m+1$ and $\| \cdot \|_{v_1}, \ldots, \| \cdot \|_{v_r}$ the set of Banach norms on the real or complex vector spaces $E_{v_i} = E \times_F F_{v_i}$ corresponding to elements of ${\rm Val}(F)_{\infty} = \{ v_1, \ldots, v_r \}$. It is well-known that the above data for $E$ define a family $\{ \| \cdot \|_v, \; v \in {\rm Val}(F) \}$ of $v$-adic metrics for a standard invertible sheaf ${\cal O}(1)$ on ${\bf P}(E)$. If $x \in X(F)$ is a point and $s \in { \rm H }^0(U, {\cal O}(1))$ is a section over an open subset $U \subset {\bf P}(E)$ containing $x$, then we denote by $\|s(x)\|_v$ the corresponding $v$-adic norm of $s$ at $x$. We set $\tilde{\cal O}(1) = ( {\cal O}(1), \|\cdot \|_v) $ to be the standard invertible sheaf ${\cal O}(1)$ on ${\bf P}(E)$ together with a family $v$-adic metrics $\|\cdot \|_v$ defined by the above data and we call $\tilde{\cal O}(1)$ the {\bf standard ample metrized invertible sheaf} on ${\bf P}(E)$. } \end{dfn} \begin{dfn} {\rm Let $X$ be an algebraic variety over $F$. For any point $x \in X(F)$ and any regular function $f \in { \rm H }^0(U, {\cal O}_X)$ on an open subset $U \subset {\bf P}(E)$ containing $x$, we define the $v$-adic norm $\|f(x)\|_v := |f(x)|_v$. We call this family of $v$-adic metrics on ${\cal O}_X$ the {\bf canonical metrization} of the structure sheaf ${\cal O}_X$. } \end{dfn} \begin{dfn} {\rm Let $X$ be a quasi-projective algebraic variety, $L$ a very ample invertible sheaf on $X$, $i \; :\; X \hookrightarrow {\bf P}(E)$ an embedding with $L = i^* {\cal O}(1)$. We denote by ${\cal L} = (L, \|\cdot \|_v )$ the sheaf $L$ together with $v$-adic metrics induced from a family of $v$-adic metrics on the standard ample metrized invertible sheaf $\tilde{\cal O}(1)$ on ${\bf P}(E)$. In this situation we call ${\cal L}$ {\bf a very ample metrized sheaf} on $X$ and write ${\cal L} = i^*\tilde{O}(1)$. } \end{dfn} \begin{dfn} {\rm Let ${\cal L} =( L, \|\cdot \|_v) $ be a very ample metrized invertible sheaf on $X$. Then for any point $x \in X(F)$, the ${\cal L}$-{\bf height} of $x$ is defined as \[ H_{\cal L}(x) = \prod_{v \in {\rm Val}(F)} \|s(x)\|_v^{-1}, \] where $s \in \Gamma(U, L)$ is a nonvanishing at $x$ section of $L$ over some open subset $U \subset X$. } \end{dfn} \begin{rem} {\rm Using a canonical metrization of the structure sheaf ${\cal O}_X$, the linear mapping \[ S^k(\Gamma(U, L)) \rightarrow \Gamma(U, L^{\otimes k}) \;\; (k > 0), \] and the $F$-bilinear mapping \[ \Gamma(U, L^{\otimes k}) \times \Gamma(U, L^{\otimes -k}) \rightarrow \Gamma(U, {\cal O}_X), \] one immediately sees that a family of $v$-adic metrics on an invertible sheaf $L$ allows to define a family of $v$-adic metrics on $L^{\otimes k}$ and on any invertible sheaf $M$ such that there exist integers $k_1, k_2$ $(k_2 \neq 0)$ with $L^{\otimes k_1} = M^{\otimes k_2}$. In this situation we write $$ {\cal M} = (M, \|\cdot \|_v) := (L^{\otimes{k_1/k_2}}, \|\cdot \|_v^{k_1/k_2}), $$ or simply $ {\cal M} ={\cal L}^{k_1/k_2}$. Obviously, one obtains \[ H_{\cal M}(x) = (H_{\cal L}(x))^{k_1/k_2}\; \; \;\forall \; x \in X(F). \]} \label{fr-norms} \end{rem} \begin{dfn} {\rm Let $L$ be an ample invertible sheaf on a quasi-projective variety $X$ and $k$ a positive integer such that $L^{\otimes k}$ is very ample. We define an {\bf ample metrized invertible sheaf} ${\cal L}=( L, \|\cdot \|_v)$ on $X$ associated with $L$ by considering ${\cal L}^{\otimes k}: = (L^{\otimes k}, \| \cdot \|_v^k)$ as a very ample metrized invertible sheaf on $X$. } \end{dfn} \subsection{Weakly and strongly ${\cal L}$-saturated varieties} Let $V$ be an arbitrary quasi-projective algebraic variety over $F$ with an ample metrized invertible sheaf ${\cal L}$. We always assume that $V(F)$ is infinite and set $$ N(V,{\cal L}, B):=\#\{x\in V(F)\; : \; H_{\cal L}(x)\le B\}. $$ Here and in 3.4 we will work under the following \begin{assume} {\rm For all quasi-projective $V',V$ with $V'\subset V$ and $|V(F)|=\infty$ there exists the limit $$ \lim_{B\rightarrow \infty} \frac{N(V',{\cal L}, B)}{N(V,{\cal L}, B)}. $$ } \label{assumption} \end{assume} \begin{dfn} {\rm We call an irreducible quasi-projective algebraic variety $V$ with an ample metri\-zed invertible sheaf ${\cal L}$ {\bf weakly ${\cal L}$-saturated} if for any Zariski locally closed subset $W \subset V$ with ${\rm dim}\, W < {\rm dim}\, V$, one has \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} < 1. \] } \end{dfn} \begin{dfn} {\rm We call an irreducible quasi-projective algebraic variety $V$ with an ample metri\-zed invertible sheaf ${\cal L}$ {\bf strongly ${\cal L}$-saturated} if for any dense Zariski open subset $U \subset V$, one has \[ {\lim}_{B \rightarrow \infty} \frac{N(U,{\cal L},B)}{N(V,{\cal L},B)} = 1. \] } \end{dfn} \begin{dfn} {\rm Let $V$ be a weakly ${\cal L}$-saturated variety, $W \subset V$ a locally closed strongly saturated subvariety of smaller dimension. Then we call $W$ an ${\cal L}$-{\bf target} of $V$, if $$ 0 < {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} < 1. $$ } \end{dfn} \begin{theo} Let $V$ be an arbitrary quasi-projective algebraic variety with an ample metrized invertible sheaf ${\cal L}$. Assume that $|V(F)|= \infty$ and that \ref{assumption} holds. Then we have: {\rm (i)} if $V$ is strongly ${\cal L}$-saturated then $V$ is weakly ${\cal L}$-saturated; {\rm (ii)} $V$ contains finitely many weakly ${\cal L}$-saturated subvarieties $W_1, \ldots, W_k$ with \[ {\lim}_{B \rightarrow \infty} \frac{N(W_1 \cup \cdots \cup W_k,{\cal L}, B)}{N(V,{\cal L},B)} = 1. \] {\rm (iii)} $V$ contains a strongly ${\cal L}$-saturated subvariety $W$ having the property \[ 0 < {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} < 1. \] {\rm (iv)} if $V$ is weakly saturated and if it doesn't contain a dense Zariski open subset $U \subset V$ which is strongly saturated then $V$ contains infinitely many ${\cal L}$-targets. \label{ws-sat} \end{theo} \noindent {\em Proof.} (i) Let $W \subset V$ be a Zariski closed subset with ${\rm dim}\, W < {\rm dim}\, V$ and $U = V \setminus W$. Then \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} + {\lim}_{B \rightarrow \infty} \frac{N(U,{\cal L},B)}{N(V,{\cal L},B)} = {\lim}_{B \rightarrow \infty} \frac{N(V,{\cal L},B)}{N(V,{\cal L},B)} =1. \] Since $V$ is strongly ${\cal L}$-saturated, we have \[ {\lim}_{B \rightarrow \infty} \frac{N(U,{\cal L},B)}{N(V,{\cal L},B)} =1 \] and therefore \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} =0 <1. \] (ii) Let $W \subset V$ be a minimal Zariski closed subset such that \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} = 1 \] and $W_1, \ldots, W_k$ irreducible components of $W$. It immediately follows from the minimality of $W$ that each $W_i$ is weakly saturated. (iii) Let $W \subset V$ be an irreducible Zariski closed subset of minimal dimension such that \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} =1. \] The minimality of $W$ implies that $W$ is weakly saturated. (iv) By (iii) the set of ${\cal L}$-targets is nonempty. Assume that the set of all ${\cal L}$-targets is finite: $\{W_1, \ldots, W_k \}$. The strong saturatedness of each $W_i$ implies that \[ {\lim}_{B \rightarrow \infty} \frac{N(W_{i_1} \cap \cdots \cap W_{i_l},{\cal L}, B)}{N(V,{\cal L}, B)} = 0 \] for all pairwise different $i_1, \ldots, i_l \in \{1,\ldots, k\}$ and $l \geq 2$. In particular, one has \[ \sum_{i=1}^k {\lim}_{B \rightarrow \infty} \frac{N(W_i,{\cal L},B)}{N(V,{\cal L}, B)} = {\lim}_{B \rightarrow \infty} \frac{N(W_1 \cup \cdots \cup W_k,{\cal L},B)}{N(V,{\cal L},B)} < 1. \] We set $U:= V \setminus (W_1 \cup \cdots \cup W_k)$. Then \[ {\lim}_{B \rightarrow \infty} \frac{N(U,{\cal L},B)}{N(V,{\cal L},B)} > 0. \] Since $U$ is not strongly saturated, there exists an irreducible Zariski closed subset $W_0 \subset V$ of minimal dimension $< {\rm dim}\, V$ such that \[ {\lim}_{B \rightarrow \infty} \frac{N(W_0\cap U,{\cal L},B)}{N(U,{\cal L},B)} > 0. \] It follows from the minimality of $W_0$ that $W_0$ is strongly saturated. On the other hand, one has \[ {\lim}_{B \rightarrow \infty} \frac{N(W_0,{\cal L},B)}{N(V,{\cal L},B)} \geq {\lim}_{B \rightarrow \infty} \frac{N(W_0\cap U,{\cal L},B)}{N(V,{\cal L},B)} > 0, \] i.e., $W_0$ is an ${\cal L}$-target and $W_0 \not\in \{ W_1, \ldots, W_k \}$. Contradiction. \hfill $\Box$ \begin{dfn} {\rm Let $V$ be a weakly ${\cal L}$-saturated variety and $W_1, W_2, \ldots$ an infinite sequence of strongly saturated irreducible subvarieties $W_i$ having the property \[ 0 < \theta_i := {\lim}_{B \rightarrow \infty} \frac{N(W_i,{\cal L},B)}{N(V,{\cal L},B)} < 1\;\; \forall i > 0. \] We say that the set $\{W_1, W_2, \ldots \}$ forms an {\bf asymptotic arithmetic ${\cal L}$-fibration} on $V$, if the following equality holds \[ \sum_{i=1}^{\infty} \theta_i = 1. \] } \end{dfn} \medskip We expect that the main source of examples of weakly and strongly saturated varieties should come from the following situation: \begin{prop} {\rm Assume \ref{assumption} and let $G \subset PGL(n+1)$ be a connected linear algebraic group acting on ${ \bf P }^n$ and $V :=Gx \subset { \bf P }^n$ a $G$-orbit of a point $x \in { \bf P }^n(F)$. Then $V$ is weakly $\tilde{O}(1)$-saturated. \label{sat1} } \end{prop} \noindent {\em Proof.} Let $W \subset V$ be an arbitrary locally closed subset with ${\rm dim}\, W < {\rm dim}\, V$, $\overline{W} \subset V$ its Zariski closure in $V$ and $ U := V \setminus \overline{W} \subset V$ the corresponding dense Zariski open subset of $V$. Then $V$ is covered by the open subsets $gU$, where $g$ runs over all elements in $G(F)$ (this follows from the fact that $G$ is unirational and that $G(F)$ is Zariski dense in $G$ \cite{borel}). Therefore, the orbit of $x\in V(F)$ under $G(F)$ is Zariski dense in $V$. Since the Zariski topology is noetherian we can choose a finite subcovering: $V = \bigcup_{i =1 }^k g_iU$ ($g_i$ in $G(F)$). Considering $g_i\in G(F)$ as matrices in $PGL(n+1)$ and using standard properties of heights \cite{lang}, one obtains positive constants $c_i$ such that $$ H_{\cal L}(g_i(x)) \le c_iH_{\cal L}(x) $$ for all $x\in { \bf P }^n(F)$. It is clear that for $c_0: =\sum_{i =1}^k c_i$ we have $$ N(U,{\cal L},B) \le N(V, {\cal L},B) \le c_0 N(U, {\cal L},B). $$ It follows that $$ \frac{1}{c_0} \leq {\lim}_{B \rightarrow \infty} \frac{N(U, {\cal L},B)}{N(V,{\cal L},B)} \leq 1. $$ Hence $$ {\lim}_{B \rightarrow \infty} \frac{N(W, {\cal L},B)}{N(V,{\cal L},B)} \leq {\lim}_{B \rightarrow \infty} \frac{N(\overline{W}, {\cal L},B)}{N(V,{\cal L},B)} \leq 1 -\frac{1}{c_0} <1. $$ \vskip 0,5cm \noindent We can reformulate the statement of \ref{sat1} as follows: \begin{prop} {\rm Let $G$ be a connected linear algebraic group, $H \subset G$ a closed subgroup and $V : = G/H$. If \ref{assumption} holds then $V$ is weakly saturated with respect to any $G$-equivariant projective embedding of $V$. \label{equiv-sat} } \end{prop} \noindent It is easy to see that $V = G/H$ is not necessarily weakly saturated with respect to projective embeddings which are not $G$-equivariant: \begin{exam} {\rm Let $S \subset { \bf P }^8$ be the anticanonically embedded Del Pezzo surface which is a blow up of a rational point in ${ \bf P }^2$. Denote by ${\cal L}$ the metrized anticanonical sheaf on $S$. The unique exceptional curve $C \subset S$ is contained in the union of two open subsets $U_0, U_1 \subset S$ where $U_0 \cong U_1 \cong {\bf A}^2$. Therefore, $S$ can be considered as a projective compactification of the algebraic group ${\bf G}_a^2$ (after an identification of ${\bf G}_a^2$ with $U_0$ or $U_1$). This compactification is not ${\bf G}_a^2$-equivariant. One has $$ a_{\cal L}(U_0) = a_{\cal L}(U_1) = a_{\cal L}(C) = 2, $$ but $$ a_{\cal L}(U_0 \setminus C) = a_{\cal L}(U_1 \setminus C) =1. $$ Hence, $U_0$ and $U_1$ are not weakly ${\cal L}$-saturated. } \end{exam} It is easy to show that an equivariant compactification of $G/H$ is not necessarily strongly ${\cal L}$-saturated: \begin{exam} {\rm Let $V = {\bf P}^1 \times {\bf P}^1$. Then $V$ is a $G$-homogeneous variety with $G = GL(2) \times GL(2)$. However, $V$ is not strongly ${\cal L}$-saturated for $L := \pi_1^*{\cal O}(k_1) \otimes \pi_2^*{\cal O}(k_2)$ ($k_1,k_2 \in {\bf N}$), if $k_1 \neq k_2$. } \end{exam} \subsection{Adelic ${\cal L}$-measure and $\tau_{\cal L}(V)$} Now we define an adelic measure ${\bf \omega}_{\cal L}$ corresponding to an ample metrized invertible sheaf ${\cal L}$ on an ${\cal L}$-primitive variety $V$ with $\alpha_{\cal L}(V) >0$ which satisfies the assumption \ref{vanish}. This is a generalization of a construction due to Peyre (\cite{peyre}) for $V$ being a smooth projective variety and ${\cal L}$ the metrized canonical line bundle, which in its turn is a generalization of the classical construction of Tamagawa measures on the adelic points of algebraic groups. Let $V$ be an ${\cal L}$-primitive variety of dimension $n$, $\rho \, : \, X \rightarrow \overline{V}^{\cal L}$ a resolution of singularities, $k_2$ a positive integer such that $k_1 = k_2 \alpha_{\cal L}(V) \in {\bf Z}$ and $$ ( \rho^*(L)^{\alpha_{\cal L}(V)} \otimes K_X)^{\otimes k_2} \cong {\cal O}(D), $$ where $D$ is a rigid effective Cartier divisor on $X$. \begin{dfn} {\rm Let $g$ be a non-zero element of the $1$-dimensional $F$-vector space ${ \rm H }^0(X, {\cal O} (D))$ ($g$ is defined uniquely up to an element of $F^*$). Let $v \in {\rm Val}(F)$. We define a measure ${\bf \omega}_{{\cal L},v}(g)$ on $V(F_v)$ as follows. Choose local $v$-analytic coordinates $x_{1,v}, \ldots, x_{n,v}$ in some open neighborhood $U_x \subset X(F_v)$ of a point $x \in X(F_v)$. We write the restriction of the global section $g$ to $U_x$ as \[ g = s^{k_1} ( dx_{1,v} \wedge \cdots \wedge dx_{n,v})^{k_2} \] where $s$ is a local section of $L$. Define a $v$-adic measure on $U_x$ as \[ {\bf \omega}_{{\cal L},v}(g) : = \|s\|_v^{k_1/k_2} dx_{1,v} \cdots dx_{n,v} = \|s\|_v^{ \alpha_{\cal L}(V)} dx_{1,v} \cdots dx_{n,v}, \] where $dx_{1,v}\cdots dx_{n,v}$ is the usual normalized Haar measure on $F_v^n$. By a standard argument, one obtains that ${\bf \omega}_{{\cal L},v}(g)$ doesn't depend on the choice of local coordinates in $U_x$ and that it extends to the whole $v$-adic space $X(F_v)$. The restriction of ${\bf \omega}_{{\cal L},v}(g)$ doesn't depend on the choice of $\rho$. So we obtain a well-defined $v$-adic measure on $V(F_v)$. } \end{dfn} \begin{rem} {\rm We remark that ${\bf \omega}_{{\cal L},v}(g)$ depends on the choice of a global section $g \in { \rm H }^0(X, {\cal O} (D))$: if $g' = cg$ $( c\in F^*)$ is another global section, then \[ {\bf \omega}_{{\cal L},v}(g') =|c|_v^{1/k} {\bf \omega}_{{\cal L},v}(g). \] } \end{rem} \noindent Our next goal is to obtain an explicit formula for the integral \[ d_v(V): = \int_{V(F_v)}{\bf \omega}_{{\cal L},v}(g) = \int_{X(F_v)}{\bf \omega}_{{\cal L},v}(g) \] for almost all $v \in {\rm Val}(F)$. We use the $p$-adic integral formula of Denef (\cite{denef} Th. 3.1) in the same way as the $p$-adic formula of A. Weil (\cite{weil} Th. 2.2.3) was used by Peyre in \cite{peyre}. \begin{rem} {\rm By \cite{AW,H}, we can choose $\rho$ in such a way that $\rho$ is defined over $F$ and all irreducible components $D_1, \ldots, D_l$ $(l \geq 0)$ of the support of $D = \sum_{i =1}^l m_i D_i$ are smooth divisors with normal crossings over the algebraic closure $\overline{F}$.} \end{rem} \begin{dfn} {\rm Let $G$ be the image of ${\rm Gal}(\overline{F}/F)$ in the symmetric group $S_l$ that acts by permutations on $D_1, \ldots, D_l$. We set $I: = \{ 1, \ldots, l\}$ and denote by $I/G$ the set of all $G$-orbits in $I$. For any $J \subset I$ we set \[ D_J : = \left\{ \begin{array}{ll} \bigcap_{j \in J} D_j, \; & \; \mbox{if $J \neq \emptyset$} \\ X \setminus \bigcup_{j \in I} D_j, \; & \; \mbox{if $J = \emptyset$} \end{array} \right. , \] \[ D_J^{\circ} = D_J \setminus \bigcup_{j \not\in J} D_j. \] ($D_J$ is defined over $F$ if and only if $J$ is a union of some of $G$-orbits in $I/G$.). } \end{dfn} We can extend $X$ to a projective scheme ${\cal X}$ of finite type over ${\cal O}_F$ and divisors $D_1, \ldots, D_l$ to codimension-$1$ subschemes ${\cal D}_1, \ldots, {\cal D}_l$ in ${\cal X}$ such that for almost all non-archimedean $v \in {\rm Val}(F)$ the reductions of $X$ and ${\cal D}_1, \ldots, {\cal D}_l$ modulo $\wp_v \subset {\cal O}_F$ are smooth projective varieties ${\cal X}_v$ and ${\cal D}_{v,1}, \ldots, {\cal D}_{v,l}$ over the algebraic closure $\overline{k_v}$ of the residue field $k_v$ with ${\cal D}_{v,i} \neq {\cal D}_{v,j}$ for $i \neq j$. Moreover, we can assume that $$ {\cal D}_{v,J} : = \left\{ \begin{array}{ll} \bigcap_{j \in J} {\cal D}_{v,j}, \; & \; \mbox{if $J \neq \emptyset$} \\ X \setminus \bigcup_{j \in I} {\cal D}_{v, j}, \; & \; \mbox{if $J = \emptyset$} \end{array} \right. $$ are also smooth over $\overline{k_v}$. \begin{dfn} {\rm A non-archimedean valuation $v \in {\rm Val}(F)$ which satisfies all the above assumptions will be called a {\bf good valuation} for the pair $({\cal X}, \{ {\cal D}_i \}_{i \in I})$. } \end{dfn} \begin{dfn} {\rm Let $G_v \subset G$ be a cyclic subgroup generated by a representative of the Frobenius element in ${\rm Gal}(\overline{k_v}/k_v)$. We denote by $I/G_v$ the set of all $G_v$-orbits in $I$. If $j \in I/G_v$, then we set $b_j$ to be the length of the corresponding $G_v$-orbit and put $r_j = m_j/k_2$, where $m_j$ is the multiplicity of irreducible components of $D$ corresponding to the $G_v$-orbit $j$. } \end{dfn} The following theorem is a slightly generalized version of Th. 3.1 in \cite{denef}: \begin{theo} Let $v \in {\rm Val}(F)$ be a good non-archimedean valuation for $({\cal X}, \{ {\cal D}_i \}_{i \in I})$. Then \[ d_v(V) = \int_{X(F_v)} {\bf \omega}_{{\cal L},v}(g) = \frac{c_{\emptyset}}{q_v^n} + \frac{1}{q_v^n} \sum_{\emptyset \neq J \subset I/{G_v} } c_J \prod_{j \in J} \left( \frac{q_v^{b_j} -1}{q_v^{b_j(r_j + 1)} -1} \right), \] where $q_v$ is the cardinality of $k_v$, and $c_J$ is the cardinality of the set of $k_v$-rational points in ${\cal D}^{\circ}_J$. \end{theo} Let us consider an exact sequence of Galois ${\rm Gal}(\overline{F}/F)$-modules: \begin{equation} 0 \rightarrow {\bf Z}[{\cal D}_1] \oplus \cdots \oplus {\bf Z}[{\cal D}_l] \rightarrow {\rm Pic}({\cal X}) \stackrel{\tilde{\rho}}{\rightarrow} {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i ) {\rightarrow} 0 \label{sh-3} \end{equation} \begin{theo} Assume that $X$ has the property ${\rm h}^1(X, {\cal O}_X) =0$. Then \[ d_v(V) = 1 + \frac{1}{q_v} {\rm Tr}(\Phi_v | {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )\otimes {\bf Q}_l) + O\left(\frac{1}{q_v^{1+ \varepsilon}} \right), \] where \[ \varepsilon = \min \{ 1/2, r_1, \ldots, r_l \} \] and $\Phi_v$ is the Frobenius morphism. \end{theo} \noindent {\em Proof.} By conjectures of Weil proved by Deligne \cite{deligne}, one has \[ \frac{c_J}{q_v^n} = O\left(\frac{1}{q_v} \right) \; J \neq \emptyset \] (since ${\rm dim}\, D_J \leq n-1$ for $J \neq \emptyset$) and \[ \frac{c_{\emptyset}}{q_v^n} = \sum_{k =0}^{2n} (-1)^k {\rm Tr}(\Phi_v | { \rm H }^k_c({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i, {\bf Q}_l)), \] where ${ \rm H }^k_c(\cdot , {\bf Q}_l)$ denotes the \'etale cohomology group with compact supports. Using long cohomology sequence of the pair $({\cal X}, \bigcup_i {\cal D}_i)$, one obtains isomorphisms \[ { \rm H }^{k}_c({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i, {\bf Q}_l) = { \rm H }^{k}_c({\cal X}, {\bf Q}_l)\;\; \mbox{\rm for $k = 2n, 2n-1$} \] and the short exact sequence \[ 0 \rightarrow { \rm H }^{2n-2}_c({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i, {\bf Q}_l) \rightarrow { \rm H }^{2n-2}_c({\cal X}, {\bf Q}_l) \rightarrow \bigoplus_{i =1}^l { \rm H }^{2n-2}_c({\cal D}_i, {\bf Q}_l) \rightarrow 0. \] Using isomorphisms \[ { \rm H }^{2n-2}_c({\cal D}_i, {\bf Q}_l(n-1)) \cong {\bf Q}_l, { \rm H }^{2n}_c({\cal X}, {\bf Q}_l(n)) \cong {\bf Q}_l, \] Poincar\'e duality \[ { \rm H }^{2n-2}_c({\cal X}, {\bf Q}_l(n-1)) \times {\rm Pic}({\cal X}) \stackrel{\sim}{\rightarrow} {\bf Q}_l \] and the vanishing property ${\rm h}^1(X, {\cal O}_X) =0$, we obtain \[ { \rm H }^{2n-1}_c({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i, {\bf Q}_l) = 0 \] and \[ \frac{c_{\emptyset}}{q_v^n} = 1 + \frac{1}{q_v} {\rm Tr}(\Phi_v | {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )\otimes {\bf Q}_l) \] \[ + \sum_{k =0}^{2n-3} \frac{ (-1)^k}{q_v^n} {\rm Tr}(\Phi_v | { \rm H }^k_c({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i, {\bf Q}_l)) \] \[ = 1 + \frac{1}{q_v} {\rm Tr}(\Phi_v | {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )\otimes {\bf Q}_l) + O\left(\frac{1}{q_v^{3/2}} \right). \] On the other hand, for $J \neq \emptyset$ one has \[ \prod_{j \in J} \left( \frac{q_v^{b_j} -1}{q_v^{b_j(r_j + 1)} -1} \right) = O\left(\frac{1}{q_v^{ 1 + \varepsilon_0}} \right), \] where $\varepsilon_0 = \min \{ r_1, \ldots, r_l \}$. \hfill $\Box$ \begin{dfn} {\rm We define the convergency factors \[ \lambda_v : = \left\{ \begin{array}{ll} L_v(1, {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )), \;& \; \mbox{\rm if $v$ is good} \\ 1 \; & \; \mbox{\rm otherwise} \end{array} \right. \] where $L_v$ is the local factor of the Artin $L$-function corresponding to the $G$-module ${\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i)$ and we set \[ \omega_{{\cal L},S} : = \sqrt{|{\rm disc}(F)|}^{-n} \prod_{v \in {\rm Val}(F)} \lambda_v^{-1} {\bf \omega}_{{\cal L},v}(g), \] where ${\rm disc}(F)$ is the absolute discriminant of ${\cal O}_F$ and $S$ is the set of bad valuations. } \end{dfn} By the product formula, $\omega_{{\cal L},S}$ doesn't depend on the choice of $g$. \begin{dfn} {\rm Denote by ${\bf A}_F$ the adele ring of $F$. Let $\overline{X(F)}$ be the closure of $X(F)$ in $X({\bf A}_F)$ (in direct product topology). Under the vanishing assumption ${\rm h}^1(X,{\cal O}_X) =0$, we define the constant \[ \tau_{\cal L}(V) = \lim_{s \rightarrow 1}(s-1)^{{\beta}_{\cal L}(V)} L_S(s, {\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )) \int_{\overline{X(F)}} \omega_{{\cal L},S}, \] where ${\beta}_{\cal L}(V)$ is the rank of the submodule of ${\rm Gal}(\overline{F}/F)$-invariants of the module ${\rm Pic}({\cal X} \setminus \bigcup_{i =1}^l {\cal D}_i )$. \label{tau-dfn} } \end{dfn} \subsection{Main strategy} Now we proceed to discuss our main strategy in understanding the asymptotic for the number $N(V, {\cal L}, B)$ as $B \to \infty$ for an arbitrary ${\cal L}$-polarized quasi-projective variety. Again, we shall make the assumption \ref{assumption}. Our approach consists in $4$ steps including $3$ subsequent simplifications of the situation: \medskip \noindent {\bf Step 1 (reduction to weakly ${\cal L}$-saturated varieties):} By \ref{ws-sat} (ii), every quasi-projective ${\cal L}$-polarized variety $V$ contains a finite number of weakly ${\cal L}$-saturated varieties $W_1, \ldots W_k$ such that \[ {\lim}_{B \rightarrow \infty} \frac{N(W,{\cal L},B)}{N(V,{\cal L},B)} = 1. \] Therefore, it would be enough to understand separately the asymptotics of $N(W_i, {\cal L}, B)$, $i \in \{1, \ldots, k\}$ modulo the asymptotics $N(W_{i_1} \cap \cdots \cap W_{i_l}, {\cal L}, B)$ for low-dimensional subvarieties $W_{i_1} \cap \cdots \cap W_{i_l}$, where $\{i_1, \ldots, i_l\} \subset \{1, \ldots, k\}$ are subsets of pairwise different elements with $l \geq 2$. \medskip For our next reduction step, we need: \begin{conj} Let $V$ be a weakly ${\cal L}$-saturated variety which doesn't contain an open Zariski dense and strongly ${\cal L}$-saturated subset $U \subset V$. Then the set of ${\cal L}$-targets of $V$ forms an asymptotic arithmetic ${\cal L}$-fibration. \label{aaf} \end{conj} \noindent {\bf Step 2 (reduction to strongly ${\cal L}$-saturated varieties):} Let $V$ be an arbitrary weakly ${\cal L}$-saturated variety. Then either $V$ contains a strongly ${\cal L}$-saturated Zariski open subset or, according to \ref{aaf}, we obtain an asymptotic arithmetic ${\cal L}$-fibration of $V$ by ${\cal L}$-targets. In the first situation, it is enough to understand the asymptotic of $N(U, {\cal L}, B)$ for the strongly ${\cal L}$-saturated variety $U$ (we note that the complement $V \setminus U$ consists of low-dimensional irreducible components). In the second situation, it is enough to understand the asymptotic of $N(W_i, {\cal L}, B)$ for each of the ${\cal L}$-targets $W_i \subset V$. \medskip For our next reduction step, we need: \begin{conj} Let $V$ be a smooth strongly ${\cal L}$-saturated quasi-projective variety. Then the complex analytic variety $V({\bf C})$ is ${\cal L}$-primitive. \end{conj} \noindent {\bf Step 3 (reduction to ${\cal L}$-primitive varieties):} Every quasi-projective algebraic variety $V$ is a disjoint union of finitely many locally closed smooth subvarieties $V_i$. Therefore, if one knows the asymptotic for each $N(V_i, {\cal L}, B)$ then one immediately obtains the asymptotic for $N(V, {\cal L}, B)$. \begin{dfn} {\rm Let $V$ be an ${\cal L}$-primitive algebraic variety over a number field $F$, $\rho\,:\, X \rightarrow \overline{V}^{\cal L}$ a desingularization over $F$ of the closure of $V$ with the exceptional locus consisting of smooth irreducible divisors $D_1,\ldots, D_l$. We consider ${\rm Pic}(X)$ and ${\rm Pic}(V,{\cal L})$ as ${\rm Gal}(\overline{F}/F)$-modules and we denote by $\beta_{\cal L}(V)$ the rank of ${\rm Gal}(\overline{F}/F)$-invariants in ${\rm Pic}(V,{\cal L})$ and by $\delta_{\cal L}(V)$ the cardinality of the cohomology group \[ { \rm H }^1({\rm Gal}(\overline{F}/F), {\rm Pic}(V, {\cal L})). \]} \label{kohom} \end{dfn} \begin{rem} {\rm Using the long exact Galois-cohomology sequence associated with (\ref{sh-3}), one immediately obtains that $\beta_{\cal L}(V)$ and $\delta_{\cal L}(V)$ do not depend on the choice of the resolution $\rho$. } \end{rem} \noindent {\bf Step 4 (expected asymptotic formula): } Let $V$ be a strongly ${\cal L}$-saturated (and ${\cal L}$-primitive) smooth quasi-projective variety. Assume that $a_{\cal L}(V) >0$. Then we expect that the following asymptotic formula holds: $$ N(V, {\cal L}, B) = c_{\cal L}(V)B^{\alpha_{\cal L}(V)} (\log B)^{ \beta_{\cal L}(V)-1} \left( 1 + o(1) \right), $$ where $$ c_{\cal L}(V) := \frac{\gamma_{\cal L}(V)}{ \alpha_{\cal L}(V)(\beta_{\cal L}(V)-1)!} \delta_{\cal L}(V)\tau_{\cal L}(V), $$ $\gamma_{\cal L}(V)$ is an invariant of the triple $({\rm Pic}(V,{\cal L}), {\rm Pic}(V,{\cal L})_{{\bf R }}, \Lambda_{\rm eff}(V,{\cal L}))$ (\ref{gamma-dfn}), $\delta_{\cal L}(V)$ is a cohomological invariant of the ${\rm Gal}(\overline{F}/F)$-module ${\rm Pic}(V, {\cal L})$ (\ref{kohom}) and $\tau_{\cal L}(V)$ is an adelic invariant of a family of $v$-adic metrics $\{ \|\cdot \|_v \}$ on $L$ (\ref{tau-dfn}). \medskip In sections 4 and 5 we discuss some examples which show how the constants $$ \alpha _{\cal L}(V), \beta _{\cal L}(V),\delta_{\cal L}(V), \gamma_{\cal L}(V),\tau_{\cal L}(V) $$ appear in asymptotic formulas for the number of rational points of bounded ${\cal L}$-height on algebraic varieties. Naturally, we expect that the exhibited behavior is typical. However, we also feel that one should collect more examples which could help to clarify the general situation. \subsection{${\cal L}$-primitive fibrations and $\tau_{\cal L}(V)$} We proceed to discuss our observations concerning the arithmetic conjecture \ref{aaf} at its relation to the geometric conjecture \ref{conj-fb}. Let $V$ be a weakly ${\cal L}$-saturated smooth quasi-projective variety with $a_{\cal L}(V) >0$ which is not strongly saturated and which doesn't contain Zariski open dense strongly saturated subvarieties. We distinguish the following two cases: \noindent {\bf Case 1. $V$ is not ${\cal L}$-primitive. } In this case we expect that some Zariski open dense subset $U \subset V$ admits an ${\cal L}$-primitive fibration which is defined by a projective regular morphism $\pi\, : \, X \rightarrow Y$ over $F$ to a low-dimensional normal irreducible projective variety $Y$ satisfying the conditions (i)-(iii) in \ref{prim-fb}, for an appropriate smooth projective compactification $X$ of $U$ (see \ref{conj-fb}). It seems natural to expect that all fibers satisfy the vanishing assumption \ref{vanish}. Thus we see that for any $y \in Y(F)$ such that $V_y = \pi^{-1}(y) \cap V$ is ${\cal L}$-primitive we can define the adelic number $\tau_{\cal L}(V_y)$. Furthermore, we expect that every ${\cal L}$-target $W$ is contained in an appropriate ${\cal L}$-primitive subvariety $V_y$ which is a fiber of the ${\cal L}$-primitive fibration $\pi \; :\; V \rightarrow U$ on $V$. In particular, Step 4 of our main strategy implies that if every ${\cal L}$-target $W$ {\em coincides} with a suitable ${\cal L}$-primitive fiber $V_y$ then one should expect the asymptotic $$ N(V, {\cal L}, B) = c_{\cal L}(V)B^{\alpha_{\cal L}(V)} (\log B)^{ \beta_{\cal L}(V)-1} \left( 1 + o(1) \right), $$ where the numbers $\alpha_{\cal L}(V)$ (resp. $\beta_{\cal L}(V)$) coincide with the numbers $\alpha_{\cal L}(V_y)$ (resp. $\beta_{\cal L}(V_y)$) for the corresponding ${\cal L}$-targets $V_y$ and the constant $c_{\cal L}(V)$ is equal to the sum $$ \sum_{y} c_{\cal L}(V_y) = \sum_{y} \frac{\gamma_{\cal L}(V_y)}{ \alpha_{\cal L}(V_y)(\beta_{\cal L}(V_y)-1)!} \delta_{\cal L}(V_y)\tau_{\cal L}(V_y), $$ where $y$ runs over all points in $Y(F)$ such that $V_y$ is an ${\cal L}$-target of $V$. It is natural to try to understand the dependence of $\tau_{\cal L}(V_y)$ on the choice of a point $y \in Y(F)$. We expect that the number $\tau_{\cal L}(V_y)$ can be interpreted as a ``height'' of $y$. More precisely, the examples we considered suggest the following: \begin{conj} There exist a family of $v$-adic metrics on $K_Y$ and two positive constants $c_2 > c_1 > 0$ such that \[ c_1 H_{\cal F}(y) \leq \tau_{\cal L}(V_y) \leq c_2 H_{\cal F}(y) \;\; \forall y \in Y(F) \cap U, \] where $U \subset Y$ is some dense Zariski open subset and ${\cal F}$ is a metrized ${\bf Q}$-invertible sheaf associated with the ${\bf Q}$-Cartier divisor $L_1^{-1} \otimes K_Y$ (recall that $L_1$ is the tautological ample ${\bf Q}$-Cartier divisor on $Y$ defined by the graded ring ${\rm R}(V, {\cal L})$ $($see \ref{proj-fb}$))$. \end{conj} \noindent {\bf Case 2. $V$ is ${\cal L}$-primitive (but not strongly saturated!).} We don't know examples of a precise asymptotic formula in this situation. \begin{exam} {\rm Let $V$ be a Fano diagonal cubic bundle over ${\bf P}^3$ with the homogeneous coordinates $(X_0:X_1:X_2:X_3)$ defined as a hypersurface in ${\bf P}^3 \times {\bf P}^3$ by the equation \[ X_0Y_0^3 + X_1Y_1^3 + X_2Y_2^3+ X_3Y_3^3= 0 \] in ${\bf P}^3 \times {\bf P}^3$ (see \cite{BaTschi4}). We expect that $V$ is not strongly saturated with respect to a metrized anticanonical sheaf $L:= {\cal O}(3,1)$ and that the corresponding ${\cal L}$-targets are the splitting diagonal cubics in fibers of the natural projection $\pi\,: \, V \to {\bf P}^3$ (this leads to the failure of the expected asymptotic formula in Step 4 for this example). } \end{exam} The next example was suggested to us by Colliot-Th\'el\`ene: \begin{exam} {\rm Let $V$ be an analogous diagonal quadric bundle over ${\bf P}^3$ defined as a hypersurface in ${\bf P}^3 \times {\bf P}^3$ by the equation \[ X_0Y_0^2 + X_1Y_1^2 + X_2Y_2^2 + X_3Y_3^2 = 0. \] For infinitely many fibers $V_x = \pi_1^{-1}(x)$ ($x \in {\bf P}^3(F)$) we have ${\rm rk}\,{\rm Pic}(V_x) = 2$. At the same time, we have also ${\rm rk}\,{\rm Pic}(V) = 2$. We consider the height function associated to some metrization of the line bundle $L:= {\cal O}(3,2)$. On the one hand, we think that the asymptotic on the whole variety is $c(V)B\log B(1+o(1))$ for $B\rightarrow \infty$ with some $c(V)>0$. On the other hand, if $X_0 X_1 X_2 X_3$ is a square in $F$ we get already about $B\log B$ solutions. Another important observation is the {\em expected} convergency of the series \[ \sum_{x \in {\bf P}^3(F) : V_x \cong {\bf P}^1 \times {\bf P}^1} c(V_x). \] The latter would be a consequence of the following two facts. First, the condition $V_x \cong {\bf P}^1 \times {\bf P}^1$ ($x = (X_0:X_1:X_2:X_3)$) is equivalent to the conditions that $V_x$ contains an $F$-rational point and that the product $X_0 X_1 X_2 X_3$ is a square in $F$. The number of $F$-rational points $x'= (X_0:X_1:X_2:X_3:Z)$ with $H_{{\cal O}(1)}(x') \leq B$ lying on the hypersurface with the equation $X_0X_1X_2X_3 = Z^2$ in the weighted projective space ${\bf P}^4(1,1,1,1,2)$ can be estimated from above by $B^2 (\log B)^3( 1 + o(1))$. \noindent Secondly, we expect $$ c(V_x) = H^{-1}_{{\cal O}(4)}(x) ( 1 + o(1) ), $$ {\em uniformly} over the base ${\bf P}^3(F)$. This would imply the claimed convergency. } \end{exam} \section{Height zeta-functions} \subsection{Tauberian theorem} One of the main techniques in the proofs of asymptotic formulas for the counting function $$ N(V,{\cal L},B):= \#\{x\in V(F)\,\,:\,\, H_{\cal L}(x)\le B\,\} $$ has been the use of {\em height zeta functions}. Let ${\cal L}$ be an ample metrized invertible sheaf on a smooth quasi-projective algebraic variety $X$. We define the height zeta function by the series $$ Z(X,{\cal L},s):=\sum_{x\in X(F)}H_{\cal L}(x)^{-s} $$ which converges absolutely for ${\rm Re}(s)\gg 0$. After establishing the analytic properties of $Z(X,{\cal L},s)$ one uses the following version of a Tauberian theorem: \begin{theo}(\cite{delange}) Suppose that there exist an $ \varepsilon >0$ and a real number $\Theta({\cal L})>0$ such that $$ Z(X,{\cal L},s)=\frac{\Theta({\cal L})}{(s-a)^{b}} + \frac{f(s)}{(s-a)^{b-1}} $$ for some $a>0$, $b\in { \bf N }$ and some function $f(s)$ which is holomorphic for ${\rm Re}(s)>a- \varepsilon $. Then we have the following asymptotic formula $$ N(X,{\cal L}, B)= \frac{\Theta({\cal L})}{a\cdot (b-1)!}B^a(\log B)^{b-1}(1+o(1)) $$ for $B\rightarrow \infty$. \end{theo} \subsection{Products} Let $X_1$ and $X_2$ be two smooth quasi-projective varieties with ample metrized invertible sheaves ${\cal L}_1$ and ${\cal L}_2$ (resp. on $X_1$ and $X_2$). Denote by $X=X_1\times X_2$ the product and by ${\cal L}$ the product of ${\cal L}_1$ and ${\cal L}_2$ (with the obvious metrization). Clearly, \[ Z(X,{\cal L}, s)= Z(X_1,{\cal L}_1, s)\cdot Z(X_2,{\cal L}_2, s). \] Assume that for $i=1,2$ we have $$ Z(X_i,{\cal L}_i,s)= \frac{\Theta({\cal L}_i)}{ (s-\alpha_{{\cal L}_i}(X_i))^{ \beta _{{\cal L}_i}(X_i)}} + \frac{f_i(s)}{(s-\alpha_{{\cal L}_i}(X_i))^{ \beta _{{\cal L}_i}(X_i)}} $$ with some functions $f_i(s)$ which are holomorphic in the domains ${\rm Re}(s_i)>\alpha_{{\cal L}_i}(X_i)- \varepsilon $ for some $ \varepsilon >0$. There are two possibilities: \smallskip Case 1: $\alpha_{{\cal L}_1}(X) = \alpha_{{\cal L}_2}(X)$. In this situation the constant $\Theta({\cal L})$ at the pole of highest order $ \beta _{{\cal L}_1}(X_1)+ \beta _{{\cal L}_2}(X_2)$ is given by $\Theta({\cal L})=\Theta({\cal L}_1)\Theta({\cal L}_2)$. \medskip Case 2: $\alpha_{{\cal L}_1}(X) < \alpha_{{\cal L}_2}(X)$. In this situation the constant is a sum $$ \Theta({\cal L}) = \sum_{x \in X_1(F)} H_{{\cal L}_1}^{-\alpha_{{\cal L}_2}}(x)\Theta({\cal L}_2). $$ Consider the projection $X\rightarrow X_1$ and denote by $V_x$ the fiber over $x\in X_1(F)$. We notice $$ \tau_{\cal L}(V_x) = H_{{\cal L}_1}^{-\alpha_{{\cal L}_2}}(x) \tau_{{\cal L}_2}(X_2). $$ We denote by $M$ the ${\bf Q}$-Cartier divisor $\pi_1^*{\cal L}_1^{\alpha_{{\cal L}_2}} \otimes K_{V_1}$. We obtain that $\pi_1^*{\cal L}_1^{\alpha_{{\cal L}_2}} = M \otimes K_{V_1}^{-1}$. So we have $\tau_{\cal L}(V_x) \sim H_{M \otimes K_{V_1}^{-1}}^{-1}$. We observe that Tamagawa numbers of fibers depend on the height of the points on the base. \subsection{Symmetric product of a curve} Let $C$ be a smooth irreducible curve of genus $g\ge 2$ over $F$. We denote by $X=C^{(m)}$ the $m$-th symmetric product of $C$ and by $Y:={\rm Jac}(C)$ the Jacobian of $C$. We fix an $m>2g-2$. We have a fibration $$ \pi\, :\; C^{(m)} \,\rightarrow \, Y, $$ with ${ \bf P }^{m-g}$ as fibers. We denote by $V_y$ a fiber over $y\in Y(F)$. Let $\tilde{C}\rightarrow {\rm Spec}({\cal O}_F)$ be a smooth model of $C$ over the integers and ${\cal L}$ an ample hermitian line bundle on $\tilde{C}$. It defines a height function $$ H_{\cal L}\,:\,C(F)\rightarrow {\bf R }_{>0} $$ which extends to $X(F)$. Observe that $ \alpha _{\cal L}(X)=(m+1-g)/d$, where $d:={\rm deg}_{{ \bf Q }}({\cal L})$. Consider the height zeta function $$ Z(X,{\cal L},s):=\sum_{x\in X(F)}H_{\cal L}(x)^{-s}. $$ This function was introduced by Arakelov in \cite{arakelov}. \begin{theo}\cite{faltings} Let ${\cal L}$ be an ample hermitian line bundle on $\tilde{C}$. There exist an $ \varepsilon ({\cal L})>0$ and a real number $\Theta({\cal L})\neq 0$ such that the height zeta function has the following representation $$ Z(X,{\cal L},s)=\frac{\Theta({\cal L})}{ (s- \alpha _{\cal L}(X))} +f(s) $$ with some function $f(s)$ which is holomorphic for ${\rm Re}(s)> \alpha _{\cal L}(X)- \varepsilon ({\cal L})$. \end{theo} This is Theorem 8 in (\cite{faltings}, p. 422). Arakelov gives an explicit expression for the constant $\Theta({\cal L})$ (\cite{arakelov}). We are very grateful to J.-B. Bost for pointing out to us that Arakelov's formula is not correct and for allowing us to use his notes on the Arakelov zeta function. \begin{theo} With the notations above we have $$ \Theta({\cal L})=\sum_{y\in Y(F)}\tau_{\cal L}(V_y). $$ \end{theo} \noindent {\em Proof.} We outline the proof for $F={ \bf Q }$. For ${\rm Re}(s)\gg 0$ one can rearrange the order of summation and one obtains $$ Z(X,{\cal L},s):=\sum_{y\in Y(F)} \sum_{x\in V_y(F)}H_{\cal L}(x)^{-s}. $$ It is proved in (\cite{faltings}, p. 420-422) that the sums $$ Z(V_y,{\cal L},s):=\sum_{x\in V_y(F)}H_{\cal L}(x)^{-s} $$ have simple poles at $s= \alpha _{\cal L}(X)$ with non-zero residues and that one can ``sum'' these expressions to obtain a function with a simple pole at $s= \alpha _{\cal L}(X)$ and meromorphic continuation to ${\rm Re}(s)> \alpha _{\cal L}(X)- \varepsilon ({\cal L})$ for some $ \varepsilon ({\cal L})>0$. Moreover, the residue at this pole is obtained as a sum over $y\in Y(F)$ of the residues of $Z(V_y,{\cal L},s)$. Choosing an element $z(y)$ in the class of $y\in {\rm Jac}(C)(F)$, and denoting by $E(y):=\Gamma (C,{\cal O}(z(y)))$ one can identify the fiber as $V_y= { \bf P }(E(y))$, where $f\in E(y)= \Gamma (C,{\cal O}(z(y)))$ is mapped to $z(y)+{\rm div}(f)$. The height is given by the formula $$ H_{\cal L}(z(y)+{\rm div}(f)):= H_{\cal L}(z(y)) \exp(\int_{C({\bf C })}\log |f|c_1({\cal L})). $$ This defines a metrization of the anticanonical line bundle on $V_y= { \bf P }(E(y))$. Assuming the Tamagawa number conjecture for ${ \bf P }^n$ (for suitable metrizations of the line bundle ${\cal O}(1)$ on ${ \bf P }^n$) we obtain $$ \lim_{s\rightarrow \alpha _{\cal L}(X)}(s- \alpha _{\cal L}(X))Z(V_y,{\cal L},s)= \tau_{\cal L}(V_y). $$ \hfill $\Box$ \medskip \noindent One can write down an explicit formula for $\tau_{\cal L}(V_y)$. For $f\in E(y)_{{\bf C }}\backslash \{0\}$ we define $$ \Phi(f):=\exp(\frac{1}{d}\int_{C({\bf C })}\log |f|c_1({\cal L})) $$ and we put $\Phi(0)=0$. It follows that $$ \tau_{\cal L}(V_y)= \frac{1}{2} \alpha _{\cal L}(X)H_{\cal L}(y)^{- \alpha _{\cal L}(X)} \cdot \frac{{\rm vol}(\{ f\in E(y)_{{\bf R }} \,|\, \Phi(f)\le 1\})}{ {\rm vol}(E(y)_{{\bf R }}/ E(y))}, $$ where the volumes are calculated with respect to some Lebesgue measure on $E(z(y))_{{\bf R }}$. Arakelov relates this last expression to the Neron-Tate height of $y\in {\rm Jac}(C)$. A detailed calculation due to Bost indicates that Arakelov's formula is correct only up-to $O(1)$. \subsection{Homogeneous spaces $G/P$} Let $G$ be a split semisimple linear algebraic group defined over a number field $F$. It contains a Borel subgroup $P_0$ defined over $F$ and a maximal torus which is split over $F$. Let $P$ be a standard parabolic. Denote by $Y_P=P\backslash G$ (resp. $X=P_0\backslash G $) the corresponding flag variety. A choice of a maximal compact subgroup ${\bf K}$ such that $G({\bf A}_F)=P_0({\bf A}_F){\bf K}$ defines a metrization on every line bundle $L$ on the flag varieties $Y_P$ (\cite{FMT}, p. 426). We will denote by $$ H_{\cal L}\,:\, P(F)\backslash G(F)\rightarrow {\bf R }_{>0} $$ the associated height. We consider the height zeta function $$ Z(X,{\cal L},s):= \sum_{x \in X(F)} H_{\cal L}(x)^{-s}. $$ \begin{theo} Let ${\cal L}$ be an ample metrized line bundle on $X$. There exist an $ \varepsilon ({\cal L})>0$ and a real number $\Theta({\cal L})\neq 0$ such that the height zeta function has the following representation $$ Z(X,{\cal L},s)=\frac{\Theta({\cal L})}{(s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)}} + \frac{f(s)}{(s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)-1}} $$ with some function $f(s)$ which is holomorphic for ${\rm Re}(s)> \alpha _{\cal L}(X)- \varepsilon ({\cal L})$. \end{theo} This theorem follows from the identification of the height zeta function with an Eisenstein series and from the work of Langlands. The formula (2.10) in (\cite{FMT}, p. 431) provides an expression for $\Theta({\cal L})$ which we will now analyze. There is a canonical way to identify the faces of the closed cone of effective divisors $ \Lambda _{\rm eff}(X)\subset {\rm Pic}(X)_{{\bf R }}$ with $ \Lambda _{\rm eff}(Y_P)$ as $P$ runs through the set of standard parabolics. A line bundle $L$ such that its class is contained in the interior of the cone $\in \Lambda _{\rm eff}(X)$ defines a line bundle $$ [L_Y]:= \alpha _{\cal L}(X)[L]+ [K_{X}] $$ which is contained in the interior of the face $ \Lambda _{\rm eff}(Y)\subset \Lambda _{\rm eff}(X)$ for some $Y=Y_P$. We have a fibration $\pi_{\cal L}\,: X\rightarrow Y$ with fibers isomorphic to the flag variety $V:=P_0\backslash P$. A fiber over $y\in Y(F)$ will be denoted by $V_y$. Denote by ${\cal K}_Y$ the canonical line bundle on $Y$ with the metrization defined above. \begin{theo} We have $$ \Theta({\cal L})= \sum_{y\in Y(F)}\gamma_{\cal L}(X)\tau_{\cal L}(V_y). $$ \end{theo} \noindent {\em Proof.} In the domains of absolute and uniform convergence we can rearrange the order of summation and we obtain $$ Z(X,{\cal L},s)=\sum_{y\in Y(F)} \sum_{x\in P_0(F)\backslash P(F)} H_{\cal L}(yx)^{-s}. $$ One can check that the sums $$ Z(V_y,{\cal L},s):=\sum_{x\in V_y(F)}H_{\cal L}(x)^{-s} =\sum_{x\in P_0(F)\backslash P(F)} H_{\cal L}(yx)^{-s} $$ have poles at $s= \alpha _{\cal L}(X)$ of order $ \beta _{\cal L}(X)$ with non-zero residues, and that they admit meromorphic continuation to $ \alpha _{\cal L}(X)- \varepsilon ({\cal L})$ for some $ \varepsilon ({\cal L})>0$. Moreover, the constant $\Theta({\cal L})$ is obtained as a sum over $y\in Y(F)$ of the residues of $Z(V_y,{\cal L},s)$. From the Tamagawa number conjecture for $P_{0}\backslash P$ (with varying metrizations of the anticanonical line bundle) we obtain $$ \lim_{s\rightarrow \alpha _{\cal L}(X)} (s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)}Z(V_y,{\cal L},s)=\gamma_{\cal L}(X) \tau_{\cal L}(V_y). $$ The sum $$ \sum_{y\in Y(F)}\gamma_{\cal L}(X) \tau_{\cal L}(V_y) $$ converges for $[L_Y]$ contained in the interior of $ \Lambda _{\rm eff}(Y)$. \hfill $\Box$ \medskip \noindent Let us recall the explicit formula for $\tau_{\cal L}(V_y)$ (see (2.9) in \cite{FMT}, p. 431): $$ \lim_{s\rightarrow \alpha _{\cal L}(X)}(s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)} \sum_{x\in P_0(F)\backslash P(F)} H_{\cal L}(xy)^{-s} = \gamma_1 c_P H_{{\cal L}_Y^{-1}\otimes {\cal K}_{Y}}(y) $$ where $ \gamma_1\in { \bf Q }$ is an explicit constant and the constant $c_P$ is defined in (\cite{FMT}, p. 430). It follows that $$ \Theta({\cal L})= \gamma_1 c_P\sum_{y\in Y(F)} H_{{\cal L}_Y^{-1}\otimes {\cal K}_Y}(y). $$ Next we observe that there is an explicit constant $ \gamma_2\in { \bf Q }$ such that we have $$ \gamma_2\tau_{\cal L}(V_{y})= c_P\cdot H_{{\cal L}^{-1}_Y\otimes {\cal K}_Y}(y) $$ for all $y\in Y(F)$. To see this, we first identify $c_P= \gamma_2\tau_{\cal L}(V)$, this is done by a computation of local factors of intertwining operators (\cite{peyre}, p.160-161). The next step involves the comparison of Tamagawa measures on $V_{y}$ for varying $y\in Y(F)$. Finally, we have $ \gamma_{\cal L}(X)= \gamma_1 \gamma_2$. \subsection{Toric varieties} There are many equivalent ways to describe a toric variety $X$ over a number field $F$ together with some projective embedding (see, for example, \cite{cox,fulton,danilov}). For us, it will be useful to view $X=X_{ \Sigma }$ as a collection of the following data (\cite{BaTschi1}): 1. A splitting field $E$ of the algebraic torus $T$ and the Galois group $G= {\rm Gal} (E/F)$. 2. The lattice of $E$-rational characters of $T$, which we denote by $M$ and its dual lattice $N$. 3. A $G$-invariant complete fan $ \Sigma $ in $N_{{\bf R }}=N\otimes {\bf R }$. There is an isomorphism between the group of $G$-invariant integral piecewise linear functions $\varphi\in PL( \Sigma )^G$ and classes of $T$-linearized line bundles on $X_{ \Sigma }$. For $\varphi\in PL( \Sigma )^G$ we denote the corresponding line bundle by $L(\varphi)$. We define metrizations of line bundles as follows. Let $G_v\subset G$ be the decomposition group at $v$. We put $N_v=N^{G_v}$ for the lattice of $G_v$-invariants of $N$ for non-archimedean valuations $v$ and $N_v=N^{G_v}_{{\bf R }}$ for archimedean $v$. We have the logarithmic map $$ T(F_v)/T({\cal O}_v)\rightarrow N_v $$ which is an embedding of finite index for all non-archimedean $v$, an isomorphism of lattices for almost all non-archimedean valuations and an isomorphism of real vector spaces for archimedean valuations. We denote by $\overline{t}_v$ the image of $t_v\in T(F_v)$ under this map. \begin{dfn}(\cite{BaTschi1}, p. 607) {\rm For every $\varphi\in PL( \Sigma )^G$ and $t_v\in T(F_v)$ we define the local height function $$ H_{ \Sigma ,v}(t_v,\varphi):=e^{\varphi(\overline{t}_v)\log q_v} $$ where $q_v$ is the cardinality of the residue field of $F_v$ for non-archimedean valuations and $\log q_v=1 $ for archimedean valuations. For $t\in T({\bf A}_F)$ we define the global height function as $$ H_{ \Sigma }(t,\varphi):=\prod_{v\in {\rm Val}(F)}H_{ \Sigma ,v}(t_v,\varphi). $$ } \end{dfn} \noindent We proved in (\cite{BaTschi1}, p. 608) that this pairing can be extended to a pairing $$ H_{ \Sigma }\,:\, T({\bf A}_F)\times PL( \Sigma )^G_{{\bf C }}\rightarrow {\bf C } $$ and that it defines a simultaneous metrization of $T$-linearized line bundles on $X$. We will denote such metrized line bundles by ${\cal L}={\cal L}(\varphi)$. We consider the height zeta function $$ Z(T,{\cal L},s)=\sum_{t\in T(F)}H_{\cal L}(t)^{-s}. $$ \begin{theo}(\cite{BaTschi3}) Let ${\cal L}$ be an invertible sheaf on $X$ (with the metrization introduced above) such that its class $[L]$ is contained in the interior of $ \Lambda _{\rm eff}(X)$. There exist an $ \varepsilon ({\cal L})>0$ and a $\Theta({\cal L})>0$ such that the height zeta function has the following representation $$ Z({\cal L},T,S)=\frac{\Theta({\cal L})}{ (s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)}}+ \frac{f(s)}{(s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)-1}} $$ where $f(s)$ is a function which is holomorphic for ${\rm Re}(s)> \alpha _{\cal L}(X)- \varepsilon ({\cal L})$. \end{theo} \begin{rem}{\rm The computation of the constants $ \alpha _{\cal L}(X)$ and $ \beta _{\cal L}(X)$ in specific examples is a problem in linear programming. For the anticanonical line bundle on a smooth toric variety $X$ we have $ \alpha _{{\cal K}^{-1}}(X) =1$ and $ \beta _{{\cal K}^{-1}}(X) = \dim PL( \Sigma )^G_{{\bf R }}-\dim M^G_{{\bf R }}$. } \end{rem} Our goal is to identify the constant $\Theta({\cal L})$. Let us recall some properties of toric varieties and introduce more notations (see \cite{BaTschi1}). The cone of effective divisors $ \Lambda _{\rm eff}(X)$ is generated by the classes of irreducible components of $X\backslash T$ which we denote by $[D_1],...,[D_r]$. These divisors correspond to Galois orbits $ \Sigma _1(1),..., \Sigma _r(1)$ on the set of $1$-dimensional cones in $ \Sigma $. The line bundle ${\cal L}$ defines a face $ \Lambda ({\cal L})$ of $ \Lambda _{\rm eff}(X)$. We denote by $J=J({\cal L})$ the maximal set of indices $J\subset [1,...,r]$ such that we have $$ \alpha _{\cal L}(X)[L]+[K_X]=\sum_{j\in J}r_j[D_j] $$ with $[D_j]\in \Lambda ({\cal L})$ and some $r_j\in { \bf Q }_{>0}$. We denote by $I=I({\cal L})$ the set of indices $i\not\in J({\cal L})$. We denote by $M_J$ the lattice given by $$ M_J:=\{ m\in M\,|\, <e,m>=0\hskip 0,3cm {\bf R }_{\ge 0}e\in \cup_{i\in I } \Sigma _i(1) \}, $$ We denote by $M_I:=M/M_J$ and by $N_*$ the corresponding dual lattices. We have an exact sequence of algebraic tori $$ 1\rightarrow T_I\rightarrow T\rightarrow T_J\rightarrow 1 $$ which induces a map $\pi_{\cal L}\,:\, T(F)\rightarrow T_J(F)$ with finite cokernel and an exact sequence of lattices $$ 0\rightarrow N_I\rightarrow N\rightarrow N_J\rightarrow 0. $$ The restriction of the fan $ \Sigma \subset N_{{\bf R }}$ to $N_{I,{\bf R }}$ will be denoted by $ \Sigma _I$. It is again a $G$-invariant fan and it will define an equivariant compactification $X_I$ of $T_I$. The class of the piecewise linear function $\varphi_I\in PL( \Sigma _I)^G$ with $\varphi_I(e)=1$ for $e\in \cup_{i\in I} \Sigma _i$ corresponds to the class of the anticanonical line bundle $[-K_I]\in {\rm Pic}(X_I)$. The line bundle ${\cal L}$ defines a fibration of varieties $\pi_{\cal L}\,:\, X_{ \Sigma }\rightarrow Y$ with fibers isomorphic to $X_I$, which, when restricted to $T$, gives rise to the exact sequence of tori above. We denote the fiber over $y\in T_J(F)$ by $X_{I,y}$. \begin{theo} We have $$ \Theta({\cal L})= \gamma_{\cal L}(X) \delta _{\cal L}(X)\sum_{y\in \pi_{\cal L}(T(F))}\tau_{\cal L}(X_{I,y}). $$ \end{theo} {\em Proof.} Let ${\cal L}={\cal L}(\varphi)$ be an invertible sheaf on $T$ with the metrization introduced above. In the domain of absolute and uniform convergence we can rearrange the order of summation and we obtain $$ Z(T,{\cal L},s)=\sum_{y\in \pi_{\cal L}(T(F))} \sum_{x\in T_{I,y}(F)} H_{ \Sigma }(yx,\varphi)^{-s}. $$ From the proof of our main theorem in \cite{BaTschi3} it follows that the sums $$ Z(T_{I,y},{\cal L},s)=\sum_{x\in T_{I}(F)} H_{ \Sigma }(yx,\varphi)^{-s} $$ have a pole at $ \alpha _{\cal L}(X)$ of order $ \beta _{\cal L}(X)$. Moreover, the constant $\Theta({\cal L})$ is obtained as a sum over $y\in \pi_{\cal L}(T(F))$ of residues of $Z(T_{I,y},{\cal L},s)$. Now we want to use the Tamagawa number conjecture for the anticanonical line bundle (with varying metrizations) on the toric variety $X_I$ to conclude the proof. In \cite{BaTschi2} we proved this conjecture for a specific metrization and under the assumption that the fan $ \Sigma $ is {\em regular}. We want to demonstrate that our proof goes through in the general case needed above. \smallskip Our main idea was to use the Poisson summation formula on the adelic group $T({\bf A}_F)$ and to obtain an integral representations for the height zeta function. We denote by ${\cal A}_I=(T_I({\bf A}_F)/{\bf K}T_I(F))^*$ the group of unitary characters of $T_I({\bf A}_F)$ which are trivial on $T_I(F)$ and on the maximal compact subgroup ${\bf K}\subset T_I({\bf A}_F)$. Using the the adelic definition of the height function we obtain $$ Z(T_{I,y},{\cal L},s)=\sum_{t\in T_I(F)}H_{\cal L}(yt)^{-s} =\int_{{\cal A}_I}d\chi \int_{T_I({\bf A}_F)} H_{\cal L}(yt)^{-s}\chi(t)d\mu, $$ where $d\mu$ is a Haar measure on $T({\bf A}_F)$ and $d\chi$ is the orthogonal Haar measure on ${\cal A}_I$. To apply our technical theorem in \cite{BaTschi2} about the analytic continuation and the residues of such integrals we need to know that $$ \int_{T_I({\bf A}_F)} H_{\cal L}(yt)^{-s}\chi(t)d\mu =\prod_{i\in I} L_i(\chi_i,s) \cdot \zeta_{ \Sigma _I}(\chi,s)\cdot \zeta_{\infty}(\chi,s) $$ where $$ \zeta_{ \Sigma _I}(\chi,s)=\prod_{v\in {\rm Val}(F)}\zeta_{ \Sigma _I,v}(\chi,s) $$ is an absolutely convergent Euler product for ${\rm Re}(s)> \alpha _{\cal L(X)}- \varepsilon ({\cal L})$, $L_i(\chi_i,s)$ are Artin $L$-functions (with some induced characters $\chi_i$) and $\zeta_{\infty}(\chi,s)$ satisfies certain growth conditions. First we observe that it is unnecessary to assume that the fan $ \Sigma _I$ is regular. Using our definition of the height function we see that the calculation of the Fourier transform (see \cite{BaTschi1}) reduces to summations of the function $q_v^{{\varphi}(\overline{t_v}) +im_v}$ ($m_v\in M_{I,{\bf R }}$) over the lattice $N_{I,v}$ (resp. to integrations in cones for archimedean valuation). A piecewise linear function $\varphi$ induces a piecewise linear function on any subdivision of the fan. Clearly, the result of such summations and integrations does not depend on any subdivisions. Next we see that for a fixed $y\in T_J(F)$ we have $H_{{\cal L},v}(yt)=H_{{\cal L},v}(t)$ for almost all $v$ and all $t\in T_I({\bf A}_F)$. Now we can refer to lemma 5.10 in \cite{BaTschi3} which proves the required statement. The local integrals for the remaining finitely many non-archimedean valuations will be absorbed into $\zeta_{ \Sigma _I}(\chi,s)$. And finally, we need to check that the estimates of the Fourier transform of $H_{{\cal L},v}(yt)$ at archimedean valuations are still satisfied for any $y\in T_J(F)$. This is straightforward. We can now apply the main technical theorem of \cite{BaTschi2} and obtain $$ \lim_{s\rightarrow \alpha _{\cal L}}(s- \alpha _{\cal L}(X))^{ \beta _{\cal L}(X)}Z(T_{I,y},{\cal L},s) = \gamma_{\cal L}(X) \delta_{\cal L}(X)\tau_{\cal L}(X_{I,y}). $$ \hfill $\Box$ \begin{rem} {\rm It is possible to compute $\tau_{\cal L}(X_{I,y})$ and to observe that it is related to the height of the point $y\in T_J(F)$. } \end{rem} \begin{rem} {\rm Similar statements hold for equivariant compactifications of homogeneous spaces $G/U$ where $G$ is a split reductive group and $U$ is its maximal unipotent subgroup \cite{StrTschi}. We hope that these results can be extended to equivariant compactifications of other homogeneous spaces, in particular, to equivariant compactifications of reductive and non-reductive groups. } \end{rem} \section{Singular Fano varieties} \subsection{Weighted projective spaces} Let $W:= {\bf P}(w) = {\bf P}(w_0, \ldots, w_n)$ be a weighted projective space of dimension $n$ with weights $w = (w_0, \ldots, w_n)$. We remark that $W$ is a rational variety over ${\bf Q}$ with ${\rm Pic}(W) \cong {\bf Z}$. Moreover, the anticanonical class $K_W^{-1}$ is an ample ${\bf Q}$-Cartier divisor. So $W$ is a (singular) Fano variety of index \[ r = \frac{ w_0 + \cdots + w_n} { l.c.m.\{ w_0, \ldots , w_n \}}. \] One could try to generalize the method of Schanuel \cite{schan} for counting ${\bf Q}$-rational points of bounded height on usual projective spaces to the case of weighted projective spaces. Let $z_0, z_1, \ldots, z_n$ be homogeneous coordinates on $Y$. Then a first approximation to counting points of bounded height would be a counting of all $(n+1)$-tuples $(x_0, x_1, \ldots, x_n) \in {\bf Z}^n \setminus \{0 \}$ satisfying the conditions \[ |x_i| \leq B^{\frac{w_i}{w_0 + w_1 + \cdots + w_n}}\;\; i =0, \ldots, n. \] Since the volume of the domain restricted by these inequalities is \[ B = \prod_{i =0}^n B^{\frac{w_i}{w_0 + w_1 + \cdots + w_n}}, \] one could expect that the asymptotic number of solutions of these inequalities agrees with ``expected'' linear growth for the anticanonical height. However, this ``intuition'' turns out to be wrong, in general, because the singularities of $Y$ could be even worse than canonical. A typical class of singularities that appear on $Y$ are so called log-terminal singularities introduced by Kawamata \cite{Ka}. We give below a simple example of a Del Pezzo surface with a log-terminal singularity and we show that for every dense Zariski open subsets $U \subset W$ the number $N(U,B)$ of $F$-rational points of anticanonical height $\leq B$ in $U$ has more than linear growth: \[ N(U, B) = c(U)B^{2 - \frac{4}{m+2}}(1 + o(B)). \] Moreover, there are no dense Zariski open subsets $U' \subset X$ such that the adelic term in the constant $c(U)$ in the asymptotic formula for $N(U,B)$ would be independent of $U$ for $U \subset U'$. \begin{exam} {\rm {\sc (Del-Pezzo surface with a log-terminal singularity)} Let $W = {\bf P}(1,1,m)$ be a singular weighted projective plane with weights $(1,1,m)$, $m \geq 2$. Then the anticanonical class of $W$ is an ample ${\bf Q}$-Cartier divisor (i.e., $W$ is Del Pezzo surface) and $p = (0:0:1)$ is the unique singularity of $W$. Let $X \rightarrow W$ be the minimal resolution of the singularity at $p \in W$. Then $X$ is isomorphic to a ruled surface ${\bf F}_m = {\bf P}({\cal O}_{{\bf P}^1} \oplus {\cal O}_{{\bf P}^1}(m))$ and the exceptional divisor $E = f^{-1}(p)$ is a smooth rational curve which is a section of the ${\bf P}^1$-bundle over ${\bf P}^1$ and $\langle E, E \rangle = -m$. Then we have \[ K_X = f^*K_W + \frac{2-m}{m} E. \] Therefore, $p$ is canonical $\Leftrightarrow$ $m = 2$ and $p$ is log-terminal $\Leftrightarrow$ $m \geq 3$. The group ${\rm Pic}(X)$ is isomorphic to ${\bf Z}^2$ where $\{ [E], [C] \}$ is a ${\bf Z}$-basis. Moreover, $ [E], [C]$ are generators for the cone $\Lambda_{\rm eff}(X)$ of effective divisors in ${\rm Pic}(X)_{\bf R}$. We have \[ K_X = -2E - (m+2 ) C, \] \[ L := f^*(-K_X) = \frac{m + 2}{m}E + (m + 2)C \] and \[ {\alpha}_L(W) = \inf \{ t \in {\bf Q} \; : \; t[L] +[K_X] \in \Lambda_{\rm eff}(X) \} = \frac{2m}{m+2}. \] } \end{exam} Since $X$ is a smooth toric variety, we can apply our main result in \cite{BaTschi3} and obtain the following: \begin{theo} Let $\pi: W \setminus p \rightarrow {\bf P}^1$ be the natural projection, $C_x$ the fiber of $\pi$ over $x \in {\bf P}^1({\bf Q})$. Then for any dense Zariski open subset in $W \setminus p$, one has \[ N(U,B) = c(U)B^{2 - \frac{4}{m+2}}(1 + o(B)) \] Moreover, \[ c(U) = \sum_{x \in {\bf P}^1({\bf Q})\cap \pi(U)} c(C_x). \] \end{theo} \subsection{Vaughan-Wooley cubic} \begin{exam} {\rm Let $Y \subset {\bf P}^5$ be a singular cubic defined by the equation $z_0z_1z_2 - z_3z_4z_5 = 0$, $X$ the intersection of $Y$ with the linear subspace in ${\bf P}^5$ with the equation: \[ z_0 + z_1 + z_2 - z_3 - z_4 - z_5 = 0. \] Vaughan and Wooley proved \cite{VW}: \begin{theo} Let $U \subset X$ be the complement in $X$ to the following $15$ divisors $D_{i_1 i_2 i_3}, D_{ij} \subset X$ $(\{ i_1, i_2, i_3 \} = \{ 0,1,2 \}, \; i \in \{ 0,1,2 \},\; j \in \{ 3,4,5\})$, where \[ D_{i_1 i_2 i_3} = \{ (z_0: \ldots : z_5) \in {\bf P}^5\; : \; z_{i_1} = z_3, \, z_{i_2} = z_4, \, z_{i_3} = z_5 \} \] \[ D_{ij} = \{ (z_0: \ldots : z_5) \in X\; : \; z_i = z_j = 0 \}.\] If $N(U,B)$ is the number of ${\bf Q}$-rational points in $U$ of the anticanonical height $\leq B$, then there exist some constants $c_1 > c_2 > 0$ such that \[ c_2 B^2 (\log B)^5 \leq N(U,B) \leq c_1 B^2 (\log B)^5. \] \end{theo} We want to show that this result is compatible with predictions in \cite{BaMa}. First of all we note that $Y$ is a $4$-dimensional toric Fano variety: an equivariant compactification of a $4$-dimensional algebraic torus $T$ with respect to a $4$-dimensional polyhedron $\Delta$ with $6$ lattice vertices \[ v_0 = (0,0,0,0), \, v_1 = (1,0,0,0), \, v_2 = (0,1,0,0), \] \[ v_3 = (0,0,1,0), \, v_4 = (0,0,0,1), \, v_5 = (1,1,-1,-1) \] ($\Delta$ is the support of global sections of a very ample divisor $Y$ corresponding to the embedding $Y \hookrightarrow {\bf P}^4$). The polyhedron $\Delta$ has $9$ faces $\Theta_{ij}$ of codimension $1$: \[ \Theta_{ij} = {\rm Conv}(\{ v_0,v_1,v_2,v_3, v_4,v_5 \} \setminus \{v_i, v_j \} ). \] Each face $\Theta_{ij}$ defines an torus invariant divisor $Y_{ij} \subset Y \setminus T$ such that $D_{ij} = Y_{ij} \cap X$. It is easy to check that all singularities of $Y$ are at worst terminal and that the hypersurface $X \subset Y$ intersects all strata $Y_{ij}$ transversally. From these facts we obtain that the only exceptional divisors with the discrepancy $0$ that appear in a resolution of singularities of $X$ come from singularities in $X \cap T$. We write down the affine equation of $X \cap T$ as \[ 1 + x + y - z -t - \frac{xy}{zt}, \] where $T = {\rm Spec}\, {\bf Q}[x^{\pm 1}, y^{\pm 1},z^{\pm 1}, t^{\pm 1}]$. From this equation one immediately sees that the only singularities in $X \cap T$ are the $A_1$-double points lying on the curve $C\;:\; x=y=z=t$. Therefore, we obtain ${\rm rk} \, {\rm Cl}(X) = {\rm rk}\, {\rm Cl}(Y) = 9 - {\rm dim}\,T = 5$. Moreover, there exists exactly one crepant divisor (over $C$) in a resolution of singularities of $X$. So the predicted power of $\log B$ in the asymptotic formula for $N(U, B)$ is $({\rm rk}\, {\rm Cl}(X) -1) + 1 = 5$. } \end{exam} \subsection{Cubic $xyz=u^3$} We consider the singular cubic surface $X\subset { \bf P }^3$ over ${ \bf Q }$ given by the homogeneous equation $xyz=u^3$. This is a toric variety, an equivariant compactification of the torus $$ T = {\rm Spec} {\bf Q}[x,y,z]/(xyz - 1) $$ given by the condition $u\neq 0$. We can fix an isomorphism $T \cong {\bf G}_m^2$ by choosing $\{ x, y\} $ as a basis of the group of algebraic characters of $T$. Consider the problem of the computation of the asymptotic of $$ N(T,B) = \mbox{\rm Card}\{ (x,y) \in ({ \bf Q }^*)^2\,\,:\,\, H(x,y) \le B\,\,\} $$ for $B\rightarrow \infty$, where \[ H(x,y) = \prod_{v \in {\rm Val}({\bf Q})} \max \{\|x\|_v,\|y\|_v,\|(xy)^{-1}\|_v,\|1\|_v \} \] This problem is addressed in \cite{fouvry}. We would like to use this problem as a down-to-earth illustration of our general theory of height zeta functions of toric varieties. First of all we note that the relation $\|x\|_v \|y\|_v \|(xy)^{-1}\|_v = \| 1 \|_v =1$ implies that $$ H_v(x,y): = \max \{\|x\|_v,\|y\|_v,\|(xy)^{-1}\|_v\}. $$ Since $l_1 =\log \|x\|_v$, $l_2 = \log \|y\|_v$, $l_3 = \log \|(xy)^{-1}\|_v$ are linear functions on the logarithmic space $N_{{\bf R},v} \cong {\bf R}^2$: $$ N_{{\bf R},v} = \left\{ \begin{array}{ll} T({\bf Q}_v) /T({\cal O}_v) \otimes_{\bf Z} {\bf R}, & \mbox{\rm if $v = p \in {\rm Spec}\,{\bf Z}$} \\ T({\bf Q}_v) /T({\cal O}_v), & \mbox{\rm if $v = \infty$,} \end{array} \right. $$ we can consider $\log h_v(x,y)$ as a piecewise linear function on $N_{{\bf R},v}$. Let $e_1 = (-2,1)$, $e_2 = (1,-2)$ and $e_3 = (1,1)$ be lattice vectors in ${\bf Z}^2 \subset {\bf R}^2$. We define the following $3$ convex cones in ${\bf R}^2$: \[ \sigma_1 = {\bf R}_{\geq 0} e_2 + {\bf R}_{\geq 0} e_3, \] \[ \sigma_2 = {\bf R}_{\geq 0} e_1 + {\bf R}_{\geq 0} e_3, \] \[ \sigma_3 = {\bf R}_{\geq 0} e_1 + {\bf R}_{\geq 0} e_2. \] Then $N_{{\bf R},v} = \bigcup_{i =1}^3 \sigma_i$ and the restriction of $\log H_v(x,y)$ to $\sigma_i$ coincides with the linear function $l_i$. Let \[ A := \bigoplus_{v \in {\rm Val}({\bf Q}} T({\bf Q}_v) /T({\cal O}_v) \] be the logarithmic adelic group. In order to compute the height zeta function \[ Z(s) = \sum_{(x,y) \in {({\bf Q}^*)^2 = T({\bf Q})}} H(x,y)^{-s} \] we use the natural homomorphism $Log$ of $T({\bf Q})$ to $A$. Denote by $B$ the subgroup $Log(T({\bf Q})) \subset A$. We remark that the kernel of $Log$ consists of $4$ elements of finite order in $({\bf Q}^*)^2$ and the quotient $A/B$ is isomorphic to ${\bf R}^2$. Moreover the functions $H_v(x,y)^{-s}$ on each $T({\bf Q}_v) /T({\cal O}_v)$ define a natural extension of $h(x,y)^{-s}$ to a function on $A$. So we obtain: \begin{equation} Z(s) = 4 \sum_{ b \in B} \prod_{v \in {\rm Val}({\bf Q})} H_v(b_v)^{-s} \label{poiss1} \end{equation} The main idea of our proof in \cite{BaTschi1} is to apply the Poisson summation formula on the group $A$ and to express the height zeta function $Z(s)$ as an integral $$ Z(s)=\frac{4}{(2\pi )^2} \int_{{{\bf R }}^2} \left( \prod_{p} Q_p(s,i{\bf m})\cdot Q_{\infty}(s,i{\bf m}) \right) {\bf d}{\bf m}, $$ where ${\bf m} =(m_1, m_2) \in {\bf R}^2$, ${\bf dm} = dm_1dm_2$, $$ Q_{p}(s,i{\bf m}) = \sum_{(b_{1,p},b_{2,p}) \in {\bf Z}^2} h_p(b_p)^{-s} p^{ i<b, {\bf m}>}, $$ and $$ Q_{\infty}(s,i{\bf m}) = \int_{{\bf R}^2} H_{\infty}(b)^{-s} \exp ( i<b, {\bf m}> ). $$ An exact computation of $Q_{p}(s,i{\bf m})$ and $Q_{\infty}(s,i{\bf m})$ can be obtained by a subdivision of each of the cones $\sigma_1, \sigma_2, \sigma_3$ into a union of $3$ subcones generated by a basis of the lattice ${\bf Z}^2 \subset {\bf R}^2$. (From the viewpoint of toric geometry this means that we reduce the counting problem for rational points in a torus with respect to a singular compactification to a counting problem for rational points in a torus with respect to the minimal resolution of singularities of this compactification). This calculation is done in \cite{BaTschi1}, Section 2 for arbitrary smooth toric varieties. What remains is the analytic continuation of the integral and the identification of the constant at the leading pole. For this it is necessary to work on the whole complexified space $PL( \Sigma )_{{\bf C }}$ and to invoke the technical theorems in Section 6 in \cite{BaTschi2}. Applying the main theorem of \cite{BaTschi2}, we obtain $$ N(T,B)= \frac{ \gamma_{{\cal K}^{-1}}(X) \delta_{{\cal K}^{-1}}(X) \tau_{{\cal K}^{-1}}(X)}{6!} B (\log B)^6(1+o(1)) $$ for $B\rightarrow \infty$. The constants are as follows: $ \gamma_{{\cal K}^{-1}} (X)=1/36$, $\delta(X)=1$ and $\tau_{{\cal K}^{-1}}(X)=\tau_{{\cal K}^{-1}}(X)_{\infty} \prod_p\tau_{{\cal K}^{-1}}(X)_p$ where $$ \tau_{{\cal K}^{-1}}(X)_p= \left(1 + \frac{7}{p} + \frac{1}{p^2} \right)\cdot (1-1/p)^{7} $$ for all primes $p$ and $\tau_{{\cal K}^{-1}}(X)_{\infty}=9\cdot 4$. Similar statements hold over any number field. One can compute the constant $\gamma_{{\cal K}^{-1}}(X)$ by observing that $\Lambda_{\rm eff}^*(X)$ (the dual cone to the cone of effective divisors) is a union of two simplicial cones.
1997-12-10T12:01:51
9712
alg-geom/9712010
fr
https://arxiv.org/abs/alg-geom/9712010
[ "alg-geom", "math.AG" ]
alg-geom/9712010
Francois Ducrot
Francois Ducrot
Structures du cube et fibres d'intersection
37 pages, latex2e with XYPic
null
null
UPRESA6093/46
null
We define the notion of a hypercube structure on a functor between two strictly commutative Picard categories which generalizes the notion of a cube structure on a $G_m$-torsor over an abelian scheme. We use this notion to define the intersection bundle of $n+1$ line bundles on a relative scheme $X/S$ of relative dimension $n$ and to construct an additive structure on the functor $I_{X/S}:PIC(X/S)^{n+1}\F PIC(S)$. Finally, we study a section of $I_{X/S}(L_1,...,L_{n+1})$ which generalizes the resultant of $n+1$ polynomials in $n$ variables and we interprete some classical formulas with this formalism.
[ { "version": "v1", "created": "Wed, 10 Dec 1997 11:01:51 GMT" } ]
2007-05-23T00:00:00
[ [ "Ducrot", "Francois", "" ] ]
alg-geom
\section{Introduction} \label{a} \subsection{} Soit $G$ un groupe alg{\'e}brique commutatif et $L$ un $G_m$-torseur sur $G$ , le classique th\'eor\`eme du cube (\cite{M3},p58) affirme que le $G_m$-torseur suivant sur $G^{3}$ est trivial \begin{equation} \label{cub1} \t (L) \; = \; m^{\ast }L \wedge (m_{12}^{\ast }L)^{-1} \wedge (m_{23}^{\ast }L)^{-1} \wedge (m_{31}^{\ast }L)^{-1} \wedge p_{1}^{\ast }L \wedge p_{2}^{\ast }L \wedge p_{3}^{\ast }L \end{equation} o\`u $m$ , $m_{ij}$ , $p_{i}$ sont les applications $G\times G \times G \longrightarrow G $ d\'efinies par $ m(x_{1},x_{2},x_{3})=x_{i}+x_{j}+x_{k} $, $ m_{ij}(x_{1},x_{2},x_{3})=x_{i}+x_{j} $ et $ p_{i}(x_{1},x_{2},x_{3})=x_{i} $. \subsection{} En r{\'e}alit{\'e} si $G$ est une vari{\'e}t{\'e} ab{\'e}lienne, une trivialisation du $G_m$-torseur $\t (L)$ devra v{\'e}rifier des conditions de compatibilit{\'e} qui s'expriment en termes de biextensions (notion introduite par \textsc{Mumford} pour l'{\'e}tude des groupes formels et {\'e}tudi{\'e}e par \textsc{Grothendieck} dans un cadre plus g{\'e}n{\'e}ral). Consid{\'e}rons en effet le $G_m$-torseur $\L =\L (L)$ sur $G \times G$ d{\'e}fini par: \[ \L (L) = m^{\ast }L \wedge ( p_1^{\ast }L)^{-1} \wedge ( p_2^{\ast }L)^{-1} \] Une trivialisation $t$ de $\t (l)$ induit deux lois (afin d'all{\'e}ger les notations on consid{\'e}rera les fibres de $\L$ au dessus d'un point g{\'e}n{\'e}rique de $G\times G$): \[ \begin{array}{lcl} *_1 : \L _{x,y} \wedge \L _{x,z}& \longrightarrow &\L _{x,y+z}\\ *_2 : \L _{x,z} \wedge \L _{y,z}& \longrightarrow &\L _{x+y,z} \end{array} \] Ces deux lois v{\'e}rifient alors des propri{\'e}t{\'e}s d'associativit{\'e}, de commutativit{\'e} (que le lecteur pourra {\'e}crire sans difficult{\'e}) et et de compatibilit{\'e} entre elles (qui traduit l'{\'e}galit{\'e} des deux fa\c{c}ons de d{\'e}velopper $ \L _{x+y,x^{\prime} + y^{\prime} }$ ). On dit alors que $\L$ est une biextension de $G\times G$ par $G_m$.\\ De plus $\L $ est muni d'un isomorphisme de sym{\'e}trie $s: \tau ^{\ast } \L \simeq \L$ ; on parle alors de biextension sym{\'e}trique. \subsection{} La notion de structure du cube, introduite par \textsc{Breen} dans \cite{B2}, explicite les notions pr{\'e}c{\'e}dentes. Une struture du cube sur un $G_m$-torseur $L$ sur $G$ est la donn{\'e}e d'une trivialisation $t$ du $G_m$-torseur $\t (L)$ sur $G^3$ telle que les lois partielles $\ast _1$ et $\ast _2$ induites par $t$ sur $\L (L)$ font de $\L (L)$ une biextension sym{\'e}trique. \subsection{} Si $\pi : X \longrightarrow S $ est une courbe relative lisse et $J$ d{\'e}signe la composante de degr{\'e} $0$ du sch{\'e}ma de Picard relatif $\text{PIC} (X/S)$, \textsc{Moret-Bailly} d{\'e}duit de l'existence d'une structure de biextension sur le faisceau de Poincar{\'e} $\mathcal{P}$ sur $J \times J^{\vee }$ et de l'existence d'une polarisation canonique sur $J$ l'existence d'une biextension $\mathcal{B}$ canonique de $J\times J$ par $G_m$ et il montre ensuite l'existence d'isomorphismes canoniques \begin{equation} \label{pairing} \mathcal{B} _{\text{cl} (L) ,\text{cl} (M)} \simeq \det \text{R}\pi \lst (L\otimes _{\ox} M)^{-1} \otimes _{\os} \det \text{R}\pi \lst (L) \otimes _{\os} \det \text{R}\pi \lst (M) \otimes _{\os} \det \text{R}\pi \lst (\mathcal{O}_{X} ) ^{-1} \end{equation} \subsection{}\textsc{Deligne} propose dans \cite{d3} un programme dont une {\'e}tape est la construction pour tout morphisme projectif plat de dimension $d$ : $\pi : X \longrightarrow S$, du fibr{\'e} d'intersection relativement {\`a} $S$ , de $d+1$ faisceaux inversibles $L_0 , \cdots , L_d$ sur $X$. Il s'agit de construire un $\mathcal{O}_{S}$-module inversible $ I_{X/S} (L_1 , \cdots , L_{n+1})$ , dont la construction est fonctorielle en les faisceaux $L_i$ (pour les isomorphismes de faisceaux) et compatible aux changements de base et qui est multiplicatif en les faisceaux $L_i$. De fa\c{c}on pr{\'e}cise, on veut construire un syst{\`e}me d'isomorphismes: \begin{multline} I_{X/S} (L_1 , \cdots , L_i \otimes _{\ox} L_i ^{\prime} , \cdots , L_{n+1}) \simeq \\ I_{X/S} (L_1 , \cdots , L_i , \cdots , L_{n+1}) \otimes _{\os} I_{X/S} (L_1 , \cdots , L_i ^{\prime} , \cdots , L_{n+1}) \end{multline} munis de donn{\'e}es de commutativit{\'e}, d'associativit{\'e} et de compatibilit{\'e} entre ces diff{\'e}rentes lois partielles. Un tel faisceau d'intersection est construit par \textsc{Deligne} dans \cite{De} dans le cas d'une courbe relative lisse et cette m{\'e}thode est {\'e}tendue par \textsc{Elkik} dans \cite{elkik1} au cas des morphismes plats de Cohen-Macaulay purement de dimension $n$.\\ \textsc{Moret-Bailly} propose une autre m{\'e}thode dans \cite{MB} dans le cas des courbes lisses, construisant le faisceau d'intersection par la formule (\ref{pairing}), bas{\'e}e sur le fibr{\'e} d{\'e}terminant, et montrant ensuite la multiplicativit{\'e} de cette construction en utilisant la propri{\'e}t{\'e} d'autodualit{\'e} de la jacobienne. Il explique ensuite comment {\'e}tendre cette construction au cas de courbes de Cohen-Macaulay.\\ Enfin une autre construction est propos{\'e}e par \textsc{Deligne} dans \cite{d3}, bas{\'e}e sur des symboles. Cette id{\'e}e est utilis{\'e}e par \textsc{Aitken} \cite{a} dans le cas d'une courbe singuli{\`e}re quelconque sur une base r{\'e}duite. \subsection{} Dans ce travail on introduit une notion de structure du $p$-cube sur un foncteur entre deux cat{\'e}gories de Picard strictement commutatives, qui {\'e}tend les d{\'e}finitions de \textsc{Breen} {\`a} un cadre l{\'e}g{\`e}rement plus g{\'e}n{\'e}ral (pour $p=3$, si $G$ est un groupe alg{\'e}brique commutatif sur un corps $k$ et $L$ est un $G_m$-torseur sur $G$, une structure du 3-cube sur le foncteur $\gd :G\longrightarrow\text{Vect}_k,g\mapsto L_g$ coincide avec la d{\'e}finition, donn{\'e}e par \textsc{Breen}, d'une structure du cube sur $L$). La donn{\'e}e d'une structure du $p$-cube sur un foncteur $\gd :\mathcal{C}\longrightarrow\mathcal{D}$ entre cat{\'e}gories de Picard strictement commutatives munit la "dif{\'e}rence sym{\'e}trique $(p-1)$-i{\`e}me de $\gd$", qui est un foncteur $\mathcal{C} ^{p-1}\longrightarrow \mathcal{D}$, de donn{\'e}es d'additivit{\'e} en chaque variables. On applique ces notions au cas $X/S$ est un sch{\'e}ma relatif projectif de dimension $n$ quelconque sur une base localement noeth{\'e}rienne et $\gd$ est le foncteur d{\'e}terminant de l'image directe d{\'e}riv{\'e}e $PIC (X/S)\longrightarrow PIC (S)$. Le th{\'e}or{\`e}me principal de ce travail montre alors alors l'existence d'une structure du $(n+2)$-cube canonique sur $\gd$. Sa d{\'e}monstration est bas{\'e}e sur une r{\'e}currence sur la dimension des sch{\'e}mas consid{\'e}r{\'e}s, qui construit une structure du $(n+2)$-cube sur le foncteur d{\'e}terminant associ{\'e} {\`a} un sch{\'e}ma relatif de dimension $n$ {\`a} partir de la donn{\'e}e de structures du $(n+1)$-cube sur le foncteur d{\'e}terminant de tout sous-sch{\'e}ma relatif de $X$ de dimension $n-1$. Une telle r{\'e}currence impose donc, m{\^e}me si on veut montrer le r{\'e}sultat uniquement pour un sch{\'e}ma relatif lisse, de savoir le montrer pour des sous-sch{\'e}mas qui n'ont aucune raison d'{\^e}tre lisses ou m{\^e}me simplement r{\'e}duits et donc de travailler avec des sch{\'e}mas $X/S$ assez g{\'e}n{\'e}raux. Ceci nous permet de montrer que le fibr{\'e} d'intersection $I_{X/S}:PIC (X/S)^{n+1}\longrightarrow PIC (S)$, d{\'e}fini, en suivant la m{\'e}thode de \textsc{Moret-Bailly}, comme la diff{\'e}rence sym{\'e}trique $(n+1)$-i{\`e}me du foncteur d{\'e}terminant est bien muni de donn{\'e}es d'additivit{\'e} en chaque variable. On {\'e}tudie alors une section canonique de $I_{X/S}(L_1,\cdots ,L_n)$ qui g{\'e}n{\'e}ralise la notion de r{\'e}sultant de $n$ polyn{\^o}mes. On a fait appel de fa\c{c}on syst{\'e}matique dans les raisonnements {\`a} une notion de $n$-cube dans une cat{\'e}gorie de Picard strictement commutative. Cette notion est de nature est de nature combinatoire et permet de repr{\'e}senter de mani{\`e}re commode des syst{\`e}mes d'isomorphismes de la forme $a\otimes b \stackrel{\sim }{\F } c\otimes d$. Son seul int{\'e}r{\^e}t est de permettre une repr{\'e}sentation graphique des raisonnements de r{\'e}currence sur la dimension. Le d{\'e}faut de cette approche est qu'il cache l'aspect g{\'e}om{\'e}trique li{\'e} {\`a} la notion de multiextension. \subsection{Plan de l'article} \begin{enumerate} \item Un exemple des m{\'e}thodes utilis{\'e}es: Les propri{\'e}t{\'e}s d'additivit{\'e} en chaque variable du "nombre d'intersection" de plusieurs diviseurs. \item Pr{\'e}liminaires techniques sur les diviseurs de Cartier relatifs et introduction d'une notion ad hoc de faisceau inversible suffisamment positif. \item Rappels sur les cat{\'e}gories de Picard et introduction de la notion de structure du cube. \item Pr{\'e}sentation des cat{\'e}gories et foncteurs utilis{\'e}s. \item Th{\'e}or{\`e}me pricipal: l'existence d'une structure du cube sur le fibr{\'e} d{\'e}terminant. \item Applications au fibr{\'e} d'intersection. Construction et {\'e}tude du r{\'e}sultant. \end{enumerate} \subsection{Remerciements} On reconna{\^\i}tra dans ce travail l'influence de Larry \textsc{Breen}, qui m'a introduit dans ce domaine et qui m'a expliqu{\'e} avec patience les subtilit{\'e}s des structures du cube. Je l'en remercie vivement. \section{Un exemple introductif} Soit $X$ un sch{\'e}ma projectif. Pour tout entier $p$ et tous faisceaux inversibles $L_1 , \cdots , L_p$ sur X, posons: \[<L_1 , \cdots , L_p>_X = (-1)^p \chi (\mathcal{O}_{X} ) + \sum_{k=1}^p (-1)^{p-k}\sum_{1 \leq i_1< \cdots < i_k\leq p} \chi (L_{i_1} \otimes \cdots \otimes L_{i_k}) \] Si $p$ est {\'e}gal {\`a} la dimension $n$ de $X$, on parlera alors de {\em nombre d'intersection} de $L_1 , \cdots , L_n$.\\ Notons d'abord que la d{\'e}finition du nombre d'intersection est sym{\'e}trique en les $L_i$ et que pour tout entier $k$ et tous faisceaux inversibles $L_1 , \cdots , L_k,L,M$ , on a: \begin{multline} <L_1 , \cdots , L_{n-1},L, M>_X =\\ <L_1 , \cdots , L_{n-1},L>_X +<L_1 , \cdots , L_{n-1}, M>_X -<L_1 , \cdots , L_{n-1},L \otimes M>_X \; . \end{multline} Le r{\'e}sultat suivant est bien connu (cf. par exemple \cite{beauville},Th 1.4 pour le cas de la dimension 2), mais sa d{\'e}monstration introduit dans un cadre simple les id{\'e}es utilis{\'e}es dans ce travail et il sera utilis{\'e}, dans sa reformulation (\ref{caracteristique2}) pour la d{\'e}monstration du th{\'e}or{\`e}me principal. \begin{lemme} \label{caracteristique} Soit $X$ un sch{\'e}ma projectif de dimension $n$, alors: \begin{enumerate} \item Le morphisme {\em nombre d'intersection}: $PIC(X)^n\longrightarrow \mathbb{Z}$ est $n$-lin{\'e}aire. \item Si $(\sigma _1,\cdots ,\sigma _n)$ est une suite r{\'e}guli{\`e}re de sections de $L_1 , \cdots , L_n$, alors $<L_1 , \cdots , L_n>_X$ est {\'e}gal {\`a} la longueur du sch{\'e}ma $Z$, de dimension 0, des z{\'e}ros de la section $\sigma =\sum_{i=1}^n \sigma _i:\mathcal{O}_{X}\longrightarrow\bigoplus_{i=1}^{n}L_i$. \end{enumerate} \end{lemme} \begin{proof}[Preuve] 1. Il suffit de montrer par r{\'e}currence sur $n$ que si $X$ est un sch{\'e}ma de dimension $n$ et si $L_1 , \cdots , L_{n+1}$ sont $n+1$ faisceaux inversibles sur X, on a: $<L_1 , \cdots , L_{n+1}>_X =0$.\\ Cette assertion est {\'e}vidente dans le cas d'un sch{\'e}ma de dimension 0 puisque dans ce cas, pour tout faisceau inversible $L$, on a $\chi (L)= \text{long}(X)$.\\ Supposons maintenant l'assertion v{\'e}rifi{\'e}e pour tout sch{\'e}ma de dimension inf{\'e}rieure ou {\'e}gale {\`a} $n$ et consid{\'e}rons un sch{\'e}ma projectif $X$ de dimension $n$. Pour des faisceaux inversibles $L_1 , \cdots , L_n$ et un diviseur effectif $D$, on a: \begin{multline} <L_1 , \cdots , L_n,\mathcal{O}_{X} (D)>_X =(-1)^n (\chi (\mathcal{O}_{X} (D)) - \chi (\mathcal{O}_{X} )\\ + \sum_{k=1}^n (-1)^{n-k}\sum_{1 \leq i_1< \cdots < i_k\leq n} (\chi (L_{i_1} \otimes \cdots \otimes L_{i_k}\otimes \mathcal{O}_{X} (D)) -\chi (L_{i_1} \otimes \cdots \otimes L_{i_k})) \end{multline} En appliquant la propri{\'e}t{\'e} d'additivit{\'e} de la caract{\'e}ristique d'Euler-Poincar{\'e} {\`a} des suites exactes de la forme $ 0 \longrightarrow L\longrightarrow L(D)\longrightarrow L(D)/L \longrightarrow 0 $ et en identifiant $\chi _X (L(D)/L)$ et $\chi_D (L(D)|_D)$, on en d{\'e}duit: \[ <L_1 , \cdots , L_n,\mathcal{O}_{X} (D)>_X = <L_1(D)|_D , \cdots , L_n(D)|_D>_D \] En appliquant l'hypoth{\`e}se de r{\'e}currence au diviseur effectif $D$, on obtient donc: \begin{equation} <L_1 , \cdots , L_n,\mathcal{O}_{X} (D)>_X =0 \end{equation} On en d{\'e}duit donc, pour tout diviseur effectif $D$, les {\'e}galit{\'e}s: \begin{equation} \label{add1} <L_1 , \cdots , L_{n-1},L_n(D)>_X= <L_1 , \cdots , L_{n-1},L_n>_X + <L_1 , \cdots , L_{n-1},\mathcal{O}_{X} (D)>_X \end{equation} et \begin{multline} \label{add2} <L_1 , \cdots ,L_i \otimes L_i^{\prime} ,\cdots , L_{n-1},\mathcal{O}_{X} (D)>_X\\ = <L_1 , \cdots , L_i^{\prime} ,\cdots , L_{n-1},\mathcal{O}_{X} (D)>_X + <L_1 , \cdots ,L_i ,\cdots , L_{n-1},\mathcal{O}_{X} (D)>_X \end{multline} Comme $X$ est projectif, si $L_n$ est un faisceau inversible sur $X$, il peut s'{\'e}crire $\mathcal{O}_{X} (D-E)$, o{\`u} $D$ et $E$ sont des diviseurs effectifs. On obtient alors en appliquant (\ref{add1}): \[ <L_1 , \cdots , L_{n-1},L_n>_X= <L_1 , \cdots , L_{n-1},\mathcal{O}_{X} (D)_X>- <L_1 , \cdots , L_{n-1},\mathcal{O}_{X} (E)>_X \] En appliquant (\ref{add2}) {\`a} chacun des termes de droite de l'{\'e}galit{\'e} pr{\'e}c{\'e}dente, on obtient l'additivit{\'e} de $<L_1 , \cdots ,L_{n-1},L_n>_X$ en chacune des variables $L_1 , \cdots , L_{n-1}$, et donc, par sym{\'e}trie, en toutes les variables.\\ 2. Consid{\'e}rons le complexe de Koszul: \[ K_{\bullet}:\; 0 \longrightarrow \L ^n (E) \longrightarrow \L ^{n-1}(E) \longrightarrow \cdots \longrightarrow E \longrightarrow \mathcal{O}_{X} \longrightarrow \mathcal{O}_{X} /I_Z \longrightarrow 0 \] associ{\'e} au morphisme $\sigma^{\vee } : E=\bigoplus_{i=1}^{n}L_i ^{-1} \longrightarrow \mathcal{O}_{X}$. Par hypoth{\`e}se, $K_{\bullet}$ est exact, donc $\chi (K_{\bullet})=0$ et le r{\'e}sultat s'en d{\'e}duit, compte tenu de l'isomorphisme \[ \L ^p (E) \simeq \bigoplus_{1 \leq i_1< \cdots < i_k\leq p} L_{i_1} \otimes \cdots \otimes L_{i_k}. \] \end{proof} Dans la suite de ce travail on {\'e}tudiera, non plus une application ensembliste de $Pic(X)^n$ dans $\mathbb{Z}$, mais un foncteur de la cat{\'e}gorie de Picard $\text{PIC}(X)$ dans une autre cat{\'e}gorie de Picard, et les difficult{\'e}s proviennent de la n{\'e}cessit{\'e} de faire des constructions fonctorielles. \section{pr{\'e}liminaires techniques} Dans cette partie nous d{\'e}taillons quelques propri{\'e}t{\'e}s des sections d'un faisceau inversible sur un sch{\'e}ma relatif et nous introduisons une notion technique de faisceau suffisamment positif, utile par la suite. $S$ est ici un sch{\'e}ma localement noeth{\'e}rien et $\pi : X \longrightarrow S$ est un morphisme projectif et plat. \subsection{Diviseur relatif d{\'e}fini par une section d'un fibr{\'e}} \label{div} Soit $L$ un faisceau inversible sur le sch{\'e}ma relatif $X/S$, une section $\sigma : \mathcal{O}_{X} \longrightarrow L$ de $L$ sera dite $\pi$-r{\'e}guli{\`e}re si elle d{\'e}finit elle d{\'e}finit un diviseur de Cartier relatif effectif de $X/S$. $\sigma$ est $\pi$-r{\'e}guli{\`e}re si et seulement si pour tout point $x\in X$, $\sigma (\mathcal{O} _{X,x}) \subset L_x$ est un $\mathcal{O} _{X,x}$-module plat et le quotient $(L \otimes _{\ox} \mathcal{O} _{X,x}) / \sigma (\mathcal{O} _{X,x})$ est un $\mathcal{O}_{S,\pi (x)}$-module plat.\\ L'ensemble $U^{\prime}$ des points $x$ de $X$ v{\'e}rifiant ces deux propri{\'e}t{\'e}s apparait comme un ensemble de platitude et est donc un ouvert de $X$ par (\cite{EGA4}, 11.1.1). Notons $Z^{\prime}$ le compl{\'e}mentaire de $U^{\prime}$. Comme $\pi$ est projectif, $Z= \pi (Z^{\prime} )$ est un ferm{\'e} de $S$ dont nous noterons $U$ le compl{\'e}mentaire.\\ Le th{\'e}or{\`e}me \cite{ma},22.5 entra{\^\i}ne que si $A\longrightarrow B$ est un morphisme d'anneaux locaux et $k$ d{\'e}signe le corps r{\'e}siduel de $A$, pour tout $u \in B$, on a l'{\'e}quivalence des assertions: \begin{enumerate} \item $u$ est non diviseur de z{\'e}ro dans $B$ et $B/uB$ est un $A$-module plat. \item $u \otimes _A 1 $ est non diviseur de z{\'e}ro dans $B \otimes _A k$. \end{enumerate} On d{\'e}duit de ceci qu'un point $s$ de $S$ est dans $U$ si et seulement si $\sigma \mid _{X_s}$ d{\'e}finit un diviseur de Cartier sur $X_s$, soit encore si $\sigma $ ne s'annulle sur aucun point associ{\'e} de $X_s$. \subsection{Suites r{\'e}guli{\`e}res} Soient $L_1,\cdots , L_p$ des faisceaux inversibles sur $X$, muni de sections $\sigma_i$. Notons $D_i$ le lieu des z{\'e}ros de $\sigma _i$. On dira que que la suite (ordonn{\'e}e) $(\sigma _1,\cdots ,\sigma _p)$ est une {\em suite $\pi$-r{\'e}guli{\`e}re} si les deux conditions suivantes sont v{\'e}rifi{\'e}es: \begin{enumerate} \item $D_1\cap\cdots\cap D_p \neq \emptyset$. \item Pour tout entier $i$ compris entre 1 et $p$, $\sigma _i$ d{\'e}finit une section $\pi$-r{\'e}guli{\`e}re sur le sch{\'e}ma relatif $D_1\cap\cdots\cap D_{i-1}/S$, si $i>1$, ou sur $X/S$, si $i=1$. \end{enumerate} On notera que \begin{enumerate} \item Si $(\sigma _1,\cdots ,\sigma _p)$ est une suite $\pi$-r{\'e}guli{\`e}re, alors pour tout $1\leq i \leq p$, $D_1\cap\cdots\cap D_i$ est plat sur $S$. \item Soit $(\sigma _1,\cdots ,\sigma _p)$ une suite de sections de $(L_i)$, le sous-ensemble de $S$ au dessus duquel $(\sigma _1,\cdots ,\sigma _p)$ est une suite r{\'e}guli{\`e}re est un ouvert de $S$. \end{enumerate} \subsection{Faisceaux inversibles suffisamment positifs} On dira qu'un faisceau inversible $L$ sur $X$ est {\em suffisamment positif} (on notera $L \gg 0$) si $L$ est tr{\`e}s ample relativement {\`a} $\pi$ et si pour tout $i>0$, on a $R^i \pi _{\ast} L =0$. Ces faisceaux v{\'e}rifient les propri{\'e}t{\'e}s suivantes: \subsubsection{} La propri{\'e}t{\'e} pour un faisceau inversible un faisceau inversible d'{\^e}tre suffisamment positif est conserv{\'e}e par tout changement de base $f:T\longrightarrow S$. Ceci r{\'e}sulte de l'invariance par changement de base de la notion de faisceau relativement tr{\`e}s ample (\cite{EGA2}, 4.4.10,iii ) et du th{\'e}or{\`e}me de changement de base dans la cohomologie (\cite{Ha},12.11). \subsubsection{} Si $L_1 , \cdots , \L_k$ sont des faisceaux inversibles sur $X$, il existe des faisceaux inversibles suffisamment positifs $M_1 , \cdots , M _k$ tels que les $L_i \otimes _{\ox} M_i$ sont tous isomorphes et suffisamment positifs. En effet si $M$ est un faisceau inversible tr{\`e}s ample relativement {\`a} $\pi$, pour tout faisceau inversible $L$, il existe un entier $N$ tel que pour tout $n \geq N$, on a $L \otimes _{\ox} M^{\otimes n} \gg 0$ (\cite{EGA2},4.4.10.{\em ii} et \cite{Ha},th III.8.8.c). En appliquant ceci aux faisceaux $\bigotimes_{j\neq i}L_j$ pour $1\leq i\leq k$ et $ L_1 \otimes \cdots \otimes L_k $ on trouve un entier $n$ tel que les faisceaux $\left( \bigotimes_{j\neq i}L_j \right) \otimes M^{\otimes n}$ et $\left( \bigotimes_{1\leq j \leq k}L_j \right) \otimes M^{\otimes n}$ soit suffisamment positifs. Il suffit alors de prendre $M_i =\left( \bigotimes_{j\neq i}L_j \right) \otimes M^{\otimes n}$. \subsubsection{} \label{cb} Si $L$ est suffisamment positif, $\pi _{\ast} L$ est un faisceau localement libre sur $S$ et pour tout $s\in S$, la fibre $(\pi _{\ast} L)_s$ est isomorphe {\`a} $H^0 (X_s ,L)$ (ceci r{\'e}sulte du th{\'e}or{\`e}me de changement de base \cite{Ha},12.11). Si de plus $X$ est {\`a} fibres de dimension au moins 1, $\pi _{\ast} L$ est de rang au moins 2. \subsection{Le diviseur universel d'un faisceau suffisamment positif} \label{div2} Si $L$ est suffisamment positif, $E=(\pi_{\ast } L)^{\vee }$ est un $\mathcal{O}_{X}$-module localement libre. Consid{\'e}rons alors le fibr{\'e} projectif $P_L = \mathbb{P}(E)$ et effectuons le changement de base: \[ \begin{CD} X_{P_L} @>g>> X \\ @V\pi VV @VV\pi V\\ P_L @>f>> S \end{CD} \] Par d{\'e}finition de $P_L$, on a un morphisme surjectif $\xymatrix{f^{\ast } E \ar@{->>}[r] &{\mathcal{O}}_{P_L}(1)}$, qui induit un morphisme $\xymatrix{{\mathcal{O}}_{P_L}(-1) \ar@{^{(}->}[r] & (f^{\ast } E)^{\vee }}$. Or on a $(f^{\ast } E)^{\vee } = f^{\ast }(E^{\vee })=f^{\ast }(\pi_{\ast } L)=\pi_{\ast }(g^{\ast } L)$, o{\`u} la derni{\`e}re {\'e}galit{\'e} provient de l'hypoth{\`e}se $L\gg 0$. Le morphisme $\xymatrix{{\mathcal{O}}_{P_L}(-1) \ar@{^{(}->}[r] & \pi_{\ast }(g^{\ast } L)}$ ainsi obtenu induit par adjonction un morphisme $\pi^{\ast }\mathcal{O}_{P_L}(-1) \longrightarrow g^{\ast } L$. On obtient donc finalement une section canonique $\sigma_L$ de $g^{\ast } L \otimes \pi^{\ast }\mathcal{O}_{P_L}(1)$. En appliquant au sch{\'e}ma relatif $X_{P_L}/P_L$ les constructions de la section (\ref{div}), on construit un ouvert $U_L$ de $P_L$, au dessus duquel $\sigma _L$ d{\'e}finit un diviseur de Cartier relatif $D_L$ et on note $Z_L$ son compl{\'e}mentaire. On obtient ainsi un isomorphisme canonique $(L\otimes\pi^{\ast }\mathcal{O}_{P_L}(1))| _{X _{U_L}} \simeq \mathcal{O} (D_L )$. La situation est d{\'e}crite par le diagramme suivant: \[ \xymatrix{ &(L\otimes\pi^{\ast }\mathcal{O}_{P_L}(1)\simeq \mathcal{O} (D_L))\ar@{.}[d] &(\mathcal{O}_{X} \stackrel{\sigma_L}{\longrightarrow} L\otimes\pi^{\ast }\mathcal{O}_{P_L}(1))\ar@{.}[d] &L\ar@{.}[d] \\ D_L \ar[dr] \ar@{^{(}->} [r] &X_{U_L} \ar[r] \ar[d] &X_{P_L} \ar[r] \ar[d] & X\ar[d]_{\pi} \\ &U_L \ar@{^{(}->}[r] &P_L \ar[r] &S } \] \begin{lemme} \label{genericite1} Pour tout point $s$ de $S$, la fibre de $Z_L$ au dessus de $s$ est une union finie de sous-espaces lin{\'e}aires propres de $P_{L,s}$. \end{lemme} \begin{proof}[Preuve] En effet, soit $s \in S$, on d{\'e}duit de (\ref{div}) l'{\'e}galit{\'e}: \[ Z_{L,s} = \bigcup_{x \in \text{Ass}(X_s)} \mathbb{P} \left( \left\{\sigma \in H^0 (X_s ,L) | \sigma (x)=0 \right\} \right) \subset \mathbb{P} \left( H^0 (X_s ,L) \right) = (P_L)_s \] L'ensemble des points associ{\'e}s de $X_s$ est fini puisque $X_s$ est projectif et comme $L\mid _{X_S}$ est tr{\`e}s ample, pour tout $x \in \text{Ass}(X_s)$, l'inclusion $ \left\{\sigma \in H^0 (X_s ,L) | \sigma (x)=0 \right\} \subset H^0 (X_s ,L) $ est stricte. \end{proof} \addtocounter{subsubsection}{1} \subsubsection{} Donnons nous de plus une suite $\pi$-r{\'e}guli{\`e}re de sections $(\sigma_i)_{i=1,\cdots ,p}$ de faisceaux inversibles $L_i$ sur $X$. Soit $V$ l'ouvert de $U_L$ au dessus duquel $(\sigma_1,\cdots,\sigma_p,\sigma_L)$ est une suite $\pi$-r{\'e}guli{\`e}re et notons $Y$ son compl{\'e}mentaire. \addtocounter{theo}{1} \begin{lemme}\label{genericite2} Pour tout point $s$ de $S$, la fibre de $Y$ au dessus de $s$ est contenue dans une union finie de sous-espaces lin{\'e}aires propres de $(P_L)_s$. \end{lemme} \begin{proof}[Preuve] Soit $D$ le lieu des z{\'e}ros de la section $\bigoplus\sigma_i$ de $\bigoplus L_i$. Un point $u\in U_L$ est {\'e}l{\'e}ment de $V$ si et seulement si $\sigma |_{D_u}$ d{\'e}finit un diviseur de Cartier effectif sur $D_u$. On raisonne donc comme pour le lemme pr{\'e}c{\'e}dent en {\'e}crivant: \[ Y_s = \bigcup_{x\in \text{Ass}(D_s)} \mathbb{P}\{ \sigma \in H^0 (X_s ,L) |\sigma (x)=0 \} \] \end{proof} \section{Structure du cube sur des cat{\'e}gories de Picard} Si $(\mathcal{C},\otimes )$ et $(\mathcal{D},\otimes ) $ sont des cat{\'e}gories de Picard strictement commutatives et $\gd $ est un foncteur de $\mathcal{C}$ vers $\mathcal{D}$, on d{\'e}finit ici la notion de structure du cube sur $\gd $. \subsection{Rappels sur les cat{\'e}gories de Picard commutatives} \subsubsection{D{\'e}finitions} Une {\em cat{\'e}gorie de Picard} est une cat{\'e}gorie $\mathcal{C}$ dont toutes les fl{\`e}ches sont des isomorphismes, et qui est munie d'un bifoncteur $\otimes : \mathcal{C} \times \mathcal{C} \longrightarrow \mathcal{C}$ tel que, pour tout $L \in \text{ob} (\mathcal{C} )$, les foncteurs $X \mapsto L\otimes X $ et $X \mapsto X\otimes L $ soient des {\'e}quivalences de cat{\'e}gories, et munie de plus de donn{\'e}es d'associativit{\'e} pour $\otimes$, c'est {\`a} dire d'un syst{\`e}me fonctoriel d'isomorphismes \[ (L \otimes M) \otimes N \simeq L \otimes (M \otimes N) \] v{\'e}rifiant des conditions de compatibilit{\'e} d{\'e}crites par l'axiome du pentagone.\\ Une cat{\'e}gorie de Picard $\mathcal{C}$ est dite {\em commutative} si elle est munie de plus de donn{\'e}es de commutativit{\'e}, c'est {\`a} dire d'un syst{\`e}me fonctoriel d'isomorphismes \[ L\otimes M \stackrel{\sim }{\F } M \otimes L \] compatible avec les donn{\'e}es d'associativit{\'e} (axiome de l'hexagone). On prendra garde que l'isomorphisme de commutativit{\'e} $L\otimes L \stackrel{\sim }{\F } L \otimes L$ n'est en g{\'e}n{\'e}ral pas le morphisme identit{\'e} (quand c'est le cas pour toul $L$, on dit que la cat{\'e}gorie est {\em strictement commutative}).\\ On d{\'e}duit des axiomes d'une cat{\'e}gorie de Picard commutative l'existence d'un objet unit{\'e} $\mathcal{O} $, munis de morphismes $L \otimes \mathcal{O} \longrightarrow L \longleftarrow \mathcal{O} \otimes L$ compatibles aux contraintes de commutativit{\'e}.\\ On obtient de m{\^e}me l'existence {\`a} isomorphismes uniques pr{\`e}s d'objets inverses $L^{\vee }$ munis de morphismes $L \otimes L^{\vee } \longrightarrow \mathcal{O} \longleftarrow L^{\vee } \otimes L$. \subsubsection{Produit d'une famille index{\'e}e par un ensemble fini} Soit $(L_i)_{i\in I}$ une famille d'objets d'une cat{\'e}gorie de Picard $\mathcal{C}$ index{\'e}s par un ensemble fini $I$. Si $I$ est muni d'un ordre total $<$, les donn{\'e}es d'associativit{\'e} de $\mathcal{C}$ permettent de d{\'e}finir de fa\c{c}on fonctorielle un objet $\bigotimes_{I,<} L_i$ de $\mathcal{C}$.\\ Supposons maintenant que $\mathcal{C}$ est une cat{\'e}gorie de Picard commutative. Si $<_1$ et $<_2$ sont deux ordres totaux sur $I$, les donn{\'e}es de commutativit{\'e} de $\mathcal{C}$ d{\'e}terminent un isomorphisme $\bigotimes_{I,<_1} L_i \stackrel{\sim }{\F } \bigotimes_{I,<_2} L_i$. On peut alors d{\'e}finir $\bigotimes_{i\in I}L_i$ comme la limite inductive (ou projective) des $\bigotimes_{I,<} L_i$ sur tous les ordres totaux $<$ sur $I$. Pour d{\'e}finir un morphisme $\bigotimes_{i\in I}L_i \longrightarrow M $ dans $\mathcal{C}$, il suffira donc de choisir un ordre $<$ sur $I$ et de d{\'e}finir un morphisme $\bigotimes_{I,<} L_i \longrightarrow M$. Par ailleurs pour deux ensembles d'indices disjoints $I$ et $J$, on un isomorphisme canonique $\bigotimes_{i\in I}L_i \otimes \bigotimes_{i\in J}L_i \simeq \bigotimes_{i\in I\cup J}L_i$, obtenu en consid{\'e}rant un ordre sur $I\cup J$ tel que $I<J$ et les ordres induits sur $I$ et $J$. \subsubsection{Foncteur additif.} Soit $F: \mathcal{C} \longrightarrow \mathcal{D}$ un foncteur additif entre deux cat{\'e}gories de Picard, une donn{\'e}e d'additivit{\'e} $\mu$ pour $F$ sera la donn{\'e}e pour tout couple d'objets $L,M \in \mathcal{C}$ d'un isomorphisme fonctoriel en $L$ et $M$: \[ \mu _{L,M}: F(L) \otimes F(M) \longrightarrow F(L\otimes M) \] La donn{\'e}e d'additivit{\'e} $\mu$ sera dite compatible aux donn{\'e}es d'associativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$ si le diagramme suivant, dont les fl{\`e}ches verticales sont donn{\'e}es par les morphismes d'associativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$ \[ \begin{CD} F(L) \otimes (F(M) \otimes F(N)) @>{\text{Id}\otimes\mu}>>F(L) \otimes F(M \otimes N) @>{\mu}>>F(L\otimes (M \otimes N)) \\ @VVV @. @VVV \\ (F(L) \otimes F(M)) \otimes F(N) @>{\mu\otimes\text{Id}}>> F(L \otimes M) \otimes F(N) @>{\mu}>> F((L\otimes M) \otimes N) \end{CD} \] est commutatif.\\ Si de plus $\mathcal{C}$ et $\mathcal{D}$ sont des cat{\'e}gories de Picard commutatives, $\mu$ sera dite compatible aux donn{\'e}es de commutativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$ si le diagramme \[ \begin{CD} F(L) \otimes F(M) @>>> F(L \otimes M)\\ @VVV @VVV \\ F(M) \otimes F(L) @>>> F(M \otimes L) \end{CD} \] est commutatif (les fl{\`e}ches verticales sont donn{\'e}es par les morphismes de commutativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$). \subsection{n-Cube dans une cat{\'e}gorie de Picard strictement commutative} On introduit ici une notion de nature combinatoire, qui traduit des calculs de fractions, du genre $ab ^{-1}= cd ^{-1}$, dans une cat{\'e}gorie de Picard strictement commutative $\mathcal{C}$, en indexant des objets de $\mathcal{C}$ par les sommets de diff{\'e}rents hypercubes de $\mathbb{R} ^n$, ce qui nous permettra dans la suite de repr{\'e}senter graphiquement certains raisonnements. \subsubsection{Cubes standard de $\mathbb{R}^n$} Pour tout entier positif $n$, on consid{\`e}re l'ensemble $C_n=\{0,1\}^n$ des sommets du n-cube standard de $\mathbb{R} ^n$.\\ On consid{\'e}rera pour tout entier $i\leq n$, les inclusions de $C_{n-1}$ dans $C_n$: $\phi _i^{\prime} : (s_1,\cdots ,s_{n-1}) \mapsto (s_1,\cdots ,s_{i-1} ,0,s_{i+1},\cdots ,s_n)$ et $\phi _i\sec : (s_1,\cdots ,s_{n-1}) \mapsto (s_1,\cdots ,s_{i-1} ,1,s_{i+1},\cdots ,s_n)$. L'image de $C_{n-1}$ par $\phi _i^{\prime}$ (resp. $\phi _i\sec$) sera appel{\'e}e la i-{\`e}me $(n-1)$-face arri{\`e}re (resp. avant) de $C_n$. De m{\^e}me, soient $\ge_1,\cdots ,\ge_k \in \{ 0,1 \} $ et des indices distincts $i_1,\cdots,i_k\leq n$, on peut consid{\'e}rer l'inclusion $\phi_{i_1=\ge_1,\cdots,i_k=\ge_k}$ de $C_{n-k}$ dans $C_n$.\\ Pour tout sommet $ s=(s_1 , \cdots , s_n)\in C_n$, on notera $\ge (s)= (-1)^{n- \sum s_i}$.\\ Enfin il sera commode d'introduire un ordre total sur les sommets de $C_n$ par: \[ (s \leq s^{\prime} ) \Leftrightarrow \begin{cases} \sum s_i < \sum s_i^{\prime} &\\ \text{ou}&\\ (\sum s_i = \sum s_i^{\prime} ) & \text{et}\; ( \exists i\leq n ,(s_i > s_i^{\prime} )\; \text{et} \; (\forall j<i\, , \, s_i=s_i^{\prime} )) \end{cases} \] On notera que l'ordre induit par $\leq $ sur une k-face du cube $C_n$ est encore l'ordre $\leq$ sur $C_k$. \subsubsection{Arrangements cubiques} On appellera alors {\em n-arrangement cubique} dans une cat{\'e}gorie de Picard commutative $\mathcal{C}$ la donn{\'e}e de $2^n$ objets $K_s$ index{\'e}s par les sommets de $C_n$.\\ Toute permutation $\sigma\in S_n$ agit sur $\mathbb{R}^n$ par permutation des coordonn{\'e}es, et induit donc $\sigma : C_n \longrightarrow C_n$. Pour tout $n$-arrangement cubique $K$ dans $\mathcal{C}$, on notera alors $\sigma^{\ast } K$ le compos{\'e} de $K$ avec la permutation de $C^n$ induite par $\sigma$.\\ Pour tout n-arrangement cubique $K$ et tout entier $1\leq i\leq n$, on consid{\'e}rera alors les $(n-1)$-arrangements cubiques $\phi^{\prime\ast}_i K$ et $\phi^{\prime \prime\ast}_i K$ (i-{\`e}me face arri{\`e}re et avant de $K$). Si $A$ et $B$ sont deux $(n-1)$-arrangements cubiques, on notera pour tout indice $1\leq i\leq n$, $(\xymatrix{A \ar@{-}[r]_-i&B})$ le n-arrangement cubique tel que $\phi^{\prime\ast}_i K=A$ et $\phi^{\prime \prime\ast}_i K=B$. De m{\^e}me, si $A_{00}$, $A_{01}$, $A_{10}$ et $A_{11}$ sont des $(n-2)$-arrangements cubiques, on utilisera, pour deux indices distincts $1\leq i,j \leq n$, la notation $K= \left( \begin{array}{c} \xymatrix{ A_{01}\ar@{-}[d]_-j \ar@{-}[r] &A_{11}\ar@{-}[d] \\ A_{00} \ar@{-}[r]_-i &A_{10} } \end{array} \right)$ pour d{\'e}signer le $n$-arrangement cubique $K$ tel que $\phi_{i=\ge ,j=\ge^{\prime}}=A_{\ge ,\ge^{\prime}}, \forall \ge,\ge^{\prime}\in \{0,1\}$.\\ Soient $A$ et $B$ deux $n$-arrangements cubiques tels que $\phi^{\prime \prime\ast}_iA=\phi^{\prime\ast}_iB$. Ecrivons les alors sous la forme: $A= (\xymatrix{U \ar@{-}[r]_-i&V})$ et $B=(\xymatrix{V \ar@{-}[r]_-i&W}) $ et posons: \[ (\xymatrix{U \ar@{-}[r]_-i&V}) \ast_i (\xymatrix{V\ar@{-}[r]_-i&W}) =(\xymatrix{U\ar@{-}[r]_-i&W}) \] On dira que $A\ast_iB$ est obtenu par recollement de $A$ et $B$ le long de leur i-{\`e}me face. \subsubsection{} Soit $\gd$ un foncteur de $\mathcal{C}$ dans une cat{\'e}gorie de Picard commutative $\mathcal{D}$, pour tout $n$-arrangement cubique $K$ dans $\mathcal{C}$, on pose \[ \gd (K) = \bigotimes_{s\in C_n} \gd (K_s)^{\ge (s)} \] Si $A$ et$B$ sont des $n$-arrangements cubiques recollables dans la i-{\`e}me direction, les isomorphismes de commutativit{\'e} et de contraction dans $\mathcal{D}$ induisent un isomorphisme canonique \[ \gd (A\ast_iB)\stackrel{\sim }{\F } \gd(A) \otimes \gd (B) \] Pour toute permutation $\sigma\in S_n$, les isomorphismes de commutativit{\'e} dans $\mathcal{D}$ induisent de m{\^e}me: \[ \gd (\sigma^{\ast } A) \stackrel{\sim }{\F } \sigma (A) \] Ces deux isomorphismes sont compatibles entre eux, ce qu'on exprime en disant que le diagramme suivant est commutatif: \[ \xymatrix{ \gd (\sigma^{\ast }( A\ast_iB))\ar[d]& \gd (A\ast_i B)\ar[l] \ar[r] & \gd (A)\otimes \gd (B)\ar[d]\\ \gd (\sigma^{\ast } A\ast_{\sigma^{-1}(i)} \sigma^{\ast } B)\ar[rr] && \gd (\sigma^{\ast } A)\otimes\gd (\sigma^{\ast } B) } \] \subsubsection{Cubes dans la cat{\'e}gorie de Picard strictement commutative $\mathcal{C}$} \label{def-cube} On appellera $1$-cube dans $\mathcal{C}$ un 1-arrangement cubique quelconque $(\xymatrix{L\ar@{-}[r]&M})$.\\ On appellera $2$-cube (carr{\'e}) dans $\mathcal{C}$ la donn{\'e}e d'un 2-arrangement cubique $K$ et d'un isomorphisme $m_K: \mathcal{O} \stackrel{\sim }{\F } \otimes_{s\in C_2} K_s^{\ge (s)}$.\\ On appellera $3$-cube dans $\mathcal{C}$ la donn{\'e}e d'un 3-arrangement cubique \[ K=\left( \begin{array}{c} \xymatrix{ &K_{001}\ar@{-}[dl]\ar@{-}'[d][dd]\ar@{-}[rr]& &K_{011}\ar@{-}[dl]\ar@{-}[dd] \\ K_{101}\ar@{-}[dd]\ar@{-}[rr]&&K_{111}\ar@{-}[dd]&\\ &K_{000}\ar@{-}[dl]\ar@{-}'[r][rr]&&K_{010}\ar@{-}[dl]\\ K_{100}\ar@{-}[rr]&&K_{110} } \end{array} \right) , \] et pour chacune des six 2-faces $F$ de $K$, d'un isomorphisme $m_F: \mathcal{O} \stackrel{\sim }{\F } \otimes_{s\in F} K_s^{\ge (s)}$ qui v{\'e}rifient la condition de compatibilit{\'e} suivante: Si $F$ d{\'e}signe l'une des faces de $K$ et $F^{\prime}$ d{\'e}signe la face oppos{\'e}e, on dispose alors d'un morphisme: \[ \begin{CD} \mathcal{O} @>>> \mathcal{O} \otimes \mathcal{O} @>m_F \otimes m_{F^{\prime} }>> \bigotimes _{s\in F} K_s^{\ge (s)} \otimes \bigotimes _{s\in {F^{\prime}}} K_s^{\ge (s)} @>>> \bigotimes _{s\in K} K_s^{\ge (s)} \end{CD} \] On impose que le morphisme ainsi obtenu soit ind{\'e}pendant du choix de la face $F$.\\ Pour $n\geq 3$, un {\em n-cube} dans $\mathcal{C}$ sera la donn{\'e}e d'un n-arrangement cubique $K$ et d'un isomorphisme $m_F$ associ{\'e} {\`a} chaque 2-face $F$ de $K$, tel que tout sous 3-arrangement cubique de $K$ est un cube. \subsubsection{Construction de cubes} (a) Pour tout $n$-cube $K$ dans $\mathcal{C}$, tout objet $L$ de $\mathcal{C}$ et tout indice $1\leq i\leq n$, on munit le $(n+1)$-arrangement cubique $(\xymatrix{K \ar@{-}[r]_-i&K\otimes L})=A$ d'un syst{\`e}me de morphismes $m_F$ associ{\'e}s {\`a} chaque 2-face de $A$, qui en fait un n-cube:\\ Si $F$ n'est pas parall{\`e}le {\`a} la direction $i$, elle appartient {\`a} l'un des deux $n$-cubes $K$ ou de $K\otimes L$ et est donc d{\'e}ja muni d'un morphisme $m_F$. Sinon elle s'{\'e}crit $F= \left( \begin{array}{c} \xymatrix{Y\ar@{-}[d]_-j\ar@{-}[r] &Y\otimes L\ar@{-}[d] \\ X\ar@{-}[r]_-i &X\otimes L } \end{array} \right) $ avec $X,Y\in\text{ob}(\mathcal{C} )$ et les morphismes de commutativit{\'e} et d'associativit{\'e} de $\mathcal{C}$ induisent un morphisme $ m_F:(X\otimes L)\otimes Y\stackrel{\sim }{\F } X\otimes (Y\otimes L) $. On v{\'e}rifie sans peine, en se ramenant au cas o{\`u} $K$ est un cube, que les relations de compatibilit{\'e} de (\ref{def-cube}) entre les $m_F$ sont v{\'e}rifi{\'e}es.\\ (b) Pour tout $n>0$, construisons un $n$-cube $K \stackrel{\sim }{\F } K_{L_0}(L_1, \cdots ,L_n)$ de la mani{\`e}re suivante: On pose $K_{L_0}(L_1)= (\xymatrix{L_0 \ar@{-}[r]_-1&L_0\otimes L_1})$ et par r{\'e}currence \[ K_{L_0}(L_1, \cdots ,L_{n+1})= (\xymatrix{K_{L_0}(L_1, \cdots ,L_n) \ar@{-}[r]_-{n+1}& K_{L_0}(L_1, \cdots ,L_n)\otimes L_{n+1}}). \] On a alors, pour tout $s \in C_n$ : \[ ( K_{L_0}(L_1, \cdots ,L_n))_s = L_0 \otimes \left( \bigotimes_{i=1}^n L_i^{s_i} \right) . \] (c) Notons enfin que si $K$ est un $n$-cube, si $K_0$ d{\'e}signe le $(n-1)$-cube $(\phi _i^{\prime} )^{\ast } K$, il existe un objet $L$ de $\mathcal{C}$ et un isomorphisme $K \stackrel{\sim }{\F } (\xymatrix{K_0\ar@{-}[r]_i&K_0\otimes L})$, uniques {\`a} isomorphisme unique pr{\`e}s. On en d{\'e}duit donc qu'il existe des objets $L_0, \cdots ,L_n$ et un isomorphisme $K \stackrel{\sim }{\F } K_{L_0}(L_1, \cdots ,L_n)$, uniques {\`a} isomorphisme unique pr{\`e}s. On dira que l'objet $L_i$ est la {\em i-{\`e}me ar{\^e}te} de K. \subsubsection{Recollement de cubes} Consid{\'e}rons deux $n$-cubes $K$ et $K^{\prime}$ dans $\mathcal{C}$ tels que les deux sous-cubes $\phi^{\prime \prime\ast}_i (K)$ et $\phi^{\prime\ast}_i (K^{\prime} )$ sont {\'e}gaux en tant que cubes. Munissons le n-arrangement cubique $K\ast _i K^{\prime}$ d'un syst{\`e}me de morphismes $m_F$, pour chaque 2-face $F$ de $K\ast _i K^{\prime}$, en faisant un un $n$-cube: Si $F$ n'est pas parall{\`e}le {\`a} la direction $i$, $F$ est une 2-face de l'un des deux $n$-cubes $K$ ou $K^{\prime}$ et on prend le $m_F$ correspondant. Si $F$ est parall{\`e}le {\`a} la direction $i$, elle est obtenue en recollant une 2-face de $K$ et une de $K^{\prime}$: \[ \xymatrix{ c\ar@{-}[d] \ar@{-}[r] &d\ar@{}[dr]|{\ast _i} &d\ar@{-}[d] \ar@{-}[r] &f\ar@{}[dr]|{=} &c\ar@{-}[d] \ar@{-}[r] \ar@{}[rd]|{F} &f\\ a&b\ar@{-}[u] \ar@{-}[l]_-i &b&e\ar@{-}[u] \ar@{-}[l]_-i &a&e\ar@{-}[u] \ar@{-}[l]_-i } \] et le morphisme $m_F$ est d{\'e}fini par: \[ \mathcal{O} \stackrel{\sim }{\F } \mathcal{O} \otimes \mathcal{O} \stackrel{\sim }{\F } (a \otimes b^{\vee } \otimes c^{\vee } \otimes d ) \otimes (b \otimes e^{\vee } \otimes d^{\vee } \otimes f) \stackrel{\sim }{\F } (a \otimes c^{\vee } \otimes c^{\vee } \otimes f) \] Pour montrer que ces $m_F$ v{\'e}rifient les relations de compatibilit{\'e}, il suffit de le faire dans le cas d'un recollement de deux 3-cubes $K$ et $K^{\prime}$ le long d'une 2-face commune $F$. Cela provient alors imm{\'e}diatement du fait que, par hypoth{\`e}se, $K$ et $K^{\prime}$ sont des cubes et les morphismes $m_F$ et $m_F^{\prime}$ associ{\'e}s {\`a} $F$, vu comme face de $K$ et $K^{\prime}$, sont les m{\^e}mes. \subsection{Structure de n-cube} Soient $\mathcal{C}$ et $\mathcal{D}$ des cat{\'e}gories de Picard strictement commutatives et $\gd :\mathcal{C}\longrightarrow\mathcal{D}$ un foncteur. \begin{Def} Une structure du n-cube sur le foncteur $\gd$ est la donn{\'e}e pour tout n-cube $K$ de $\mathcal{C}$ d'un morphisme $\psi _K:\mathcal{O} \stackrel{\sim }{\F } \gd (K)$ dans $\mathcal{C}$ v{\'e}rifiant les propri{\'e}t{\'e}s suivantes: \begin{enumerate} \item {\em Fonctorialit{\'e}.} Pour tout isomorphisme de n-cubes $f: K \stackrel{\sim }{\F } K^{\prime}$, le diagramme induit \[ \xymatrix{ & \gd (K)\ar[dd]^-{\gd (f)} \\ {\mathcal{O}} \ar[ur]^-{\psi _K} \ar[dr]_-{\psi _{K^{\prime}}} \\ & \gd (K^{\prime} ) } \] est commutatif. \item {\em Recollements de cubes.} Soient $K$ et $K^{\prime}$ deux n-cubes ayant leur i-{\`e}me (n-1)-face en commun, l'isomorphisme naturel $\gd (K) \otimes \gd (K^{\prime} ) \stackrel{\sim }{\F } \gd (K \ast _i K^{\prime} )$ induit un diagramme commutatif: \[ \xymatrix{ & \gd (K) \otimes \gd (K^{\prime} ) \ar[dd] \\ {\mathcal{O}} \ar[ur]^-{\psi _K \otimes\psi _{K^{\prime}} } \ar[dr]_-{\psi _{(K \ast _i K^{\prime} )}} & \\ &\gd (K \ast _i K^{\prime} ) } \] \item {\em Propri{\'e}t{\'e} de sym{\'e}trie.} Pour tout {\'e}l{\'e}ment $\sigma $ de $S_n$, le diagramme suivant, dont la fl{\`e}che verticale est donn{\'e}e par les isomorphismes de commutativit{\'e} de $\mathcal{D}$, est commutatif: \[ \xymatrix{ & \gd (K) \ar[dd] \\ {\mathcal{O}} \ar[ur]^-{\psi _K} \ar[dr]_-{\psi _{\sigma ^{\ast } K}} & \\ &\gd (\sigma ^{\ast } K ) } \] \end{enumerate} \end{Def} \begin{ex} \label{caracteristique2} Soit $X$ un sch{\'e}ma projectif de dimension $n$, consid{\'e}rons les cat{\'e}gories $\mathcal{C} = PIC (X)$ et $\mathcal{D} = \mathbb{Z}$ (cat{\'e}gorie discr{\`e}te) et le foncteur $ \gd : \mathcal{C} \longrightarrow \mathcal{D} , L \mapsto \chi (L)$. On a vu (Lemme \ref{caracteristique}) que $\gd$ est muni d'une structure du $(n+1)$-cube. \end{ex} \subsection{Le n-foncteur multilin{\'e}aire associ{\'e} {\`a} une structure du (n+1)-cube} \subsubsection{} Consid{\'e}rons un foncteur $\gd : \mathcal{C} \longrightarrow \mathcal{D}$ entre deux cat{\'e}gories de Picard strictement commutatives. Pour tout (n-1)-cube $K$ dans $\mathcal{C}$ et tout entier $1\leq i\leq n$, d{\'e}finissons un foncteur: \[ \L _{K,i}: \mathcal{C} \longrightarrow \mathcal{D} , L \mapsto \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes L}). \] De m{\^e}me, pour tout objet $L$ de $\mathcal{C}$, d{\'e}finissons un $n$-foncteur: \[ \L _L : \mathcal{C} ^n \longrightarrow \mathcal{D} , (L_1, \cdots , L_n) \mapsto \gd (K_L (L_1, \cdots , L_n)). \] \addtocounter{theo}{1} \begin{rem} $\L _L $ est canoniquement muni de donn{\'e}es de sym{\'e}trie $\L _L \stackrel{\sim }{\F } \sigma ^{\ast } \L _L$, pour tout $\sigma \in S^n$. \end{rem} \addtocounter{subsubsection}{1} \begin{rem}\label{identification} Pour tout entier $1\leq i\leq n$ on peut identifier canoniquement: \[ \L _L (L_1, \cdots , L_n) \stackrel{\sim }{\F } \L _{K_L(L_1,\cdots ,L_{i-1},L_{i+1},\cdots ,L_n),i} (L_i) \] \end{rem} \addtocounter{subsubsection}{1} \subsubsection{} \label{def-multifonct} \addtocounter{theo}{1} Donnons nous une structure de $(n+1)$-cube $S$ sur $\gd$. Remarquons d'abord que $S$ induit, pour tous $L,M \in \text{Ob} (\mathcal{C})$, un isomorphisme canonique de foncteurs $ \L_L \stackrel{\sim }{\F } \L_M$. En effet, pour tous $L, L_1, \cdots , L_n \in \text{Ob}(\mathcal{C} )$, on a $ K_L (L_1, \cdots , L_n) \simeq K_{\mathcal{O}} (L_1, \cdots , L_n)\otimes L $ et la structure du cube $S$, appliqu{\'e}e au $(n+1)$-cube $ (\xymatrix{ K_{\mathcal{O}} (L_1, \cdots , L_n) \ar@{-}[r]_-{n+1} & K_{\mathcal{O}} (L_1, \cdots , L_n)\otimes L}) $ {\'e}tablit donc un isomorphisme canonique: \[ \gd ( K_{\mathcal{O}} (L_1, \cdots , L_n)) \stackrel{\sim }{\F } \gd ( K_L (L_1, \cdots , L_n)) \] On peut donc d{\'e}sormais parler du foncteur $\L$ (en ommettant l'indice $L$). \subsubsection{} \addtocounter{theo}{1} La donn{\'e}e de la structure du cube sur $\gd$ permet de munir chaque foncteur $\L _{K,i}$ d'une donn{\'e}e d'additivit{\'e}: \[ \mu _{K,i}:\L _{K,i}(L) \otimes \L _{K,i}(M) \stackrel{\sim }{\F } \L _{K,i}(L\otimes M) \] d{\'e}finie par: \begin{multline} \label{linearite} \L _{K,i}(L) \otimes \L _{K,i}(M) = \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes L}) \otimes \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes M}) \stackrel{\sim }{\F } \\ \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes L}) \otimes \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes M}) \otimes \gd \left( \begin{array}{c} \xymatrix{K \otimes M \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes L } \end{array} \right)\\ \stackrel{\sim }{\F } \gd (\xymatrix{K\ar@{-}[r]_-i&K\otimes (L\otimes M)}) =\L _{K,i}(L\otimes M) \end{multline} En utilisant l'identification \ref{identification}, on voit que $S$ munit $\L$ de $n+1$ donn{\'e}es d'additivit{\'e} partielles: \begin{multline*} *_i :\L (L_1, \cdots ,L_{i-1} ,L_i ,L_{i+1}, \cdots ,L_n) \otimes \L (L_1, \cdots ,L_{i-1} ,L_i^{\prime} ,L_{i+1}, \cdots ,L_n)\\ \longrightarrow \L (L_1, \cdots ,L_{i-1} ,L_i \otimes L_i^{\prime} ,L_{i+1}, \cdots ,L_n). \end{multline*} \begin{rem} \label{strict.com} On notera que la cat{\'e}gorie de Picard $\mathcal{D}$ {\'e}tant {\em strictement} commutative, le diagramme: \[ \xymatrix{ (A^{\vee } \otimes A) \otimes A^{\vee } \ar[dr] \ar[dd]\\ & A^{\vee } \\ A^{\vee } \otimes (A \otimes A^{\vee } ) \ar[ur] } \] est commutatif, ce qui permet de ne pas pr{\'e}ciser comment on effectue les contractions dans (\ref{linearite}). \end{rem} \begin{lemme}\label{comm-assoc} Les donn{\'e}es d'additivit{\'e} $\mu _{K,i}$ pour $\L _{K,i}$ sont compatibles aux donn{\'e}es d'as\-so\-cia\-ti\-vi\-t{\'e} et de commutativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$. \end{lemme} \begin{proof}[Preuve] Ces deux assertions se traduisent en disant que si $L$, $M$ et $N$ sont des objets de $\mathcal{D}$, les isomorphismes canoniques \[ \gd \left( \begin{array}{c} \xymatrix{K \otimes M \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes L } \end{array} \right) \stackrel{\sim }{\F } \gd \left( \begin{array}{c} \xymatrix{K \otimes L \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (M \otimes L) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes M } \end{array} \right) \] et \begin{multline} \label{associativite1} \gd \left( \begin{array}{c} \xymatrix{K \otimes M \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes L } \end{array} \right) \otimes \gd \left( \begin{array}{c} \xymatrix{K \otimes N \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M\otimes N) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes (L\otimes M) } \end{array} \right)\\ \stackrel{\sim }{\F } \gd \left( \begin{array}{c} \xymatrix{K \otimes N \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (M \otimes N) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes M } \end{array} \right) \otimes \gd \left( \begin{array}{c} \xymatrix{K \otimes (M\otimes N) \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M\otimes N) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes L } \end{array} \right) \end{multline} identifient les trivialisations de chacun des termes, donn{\'e}es par la structure du (n+1)-cube.\\ Le premier point provient de l'hypoth{\`e}se que la structure du cube est sym{\'e}trique et de la remarque (\ref{strict.com}).\\ Pour l'associativit{\'e}, remarquons d'abord que chacun des membres de \ref{associativite1} est isomorphe {\`a} \begin{multline} \label{associativite3} \gd \left( \begin{array}{c} \xymatrix{K \otimes N \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (M \otimes N) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes M } \end{array} \right) \otimes \gd \left( \begin{array}{c} \xymatrix{K \otimes (M\otimes N) \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M\otimes N) \ar@{-}[d] \\ K\otimes M \ar@{-}[r]_-i &K\otimes (L\otimes M) } \end{array} \right)\\ \otimes \gd \left( \begin{array}{c} \xymatrix{K \otimes M \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (L \otimes M) \ar@{-}[d] \\ K\ar@{-}[r]_-i &K\otimes L } \end{array} \right). \end{multline} Consid{\'e}rons alors l'empilement de (n+1)-cubes suivant: \begin{equation} \label{associativite2} \xymatrix{ K \otimes N \ar@{-}[d]_-{i+1}\ar@{-}[r] &K\otimes (M \otimes N) \ar@{-}[d]_-{i+1} \ar@{-}[r] &K\otimes (L\otimes M \otimes N) \ar@{-}[d] \\ K \ar@{-}[r]_-i &K\otimes M \ar@{-}[d] \ar@{-}[r]_-i \ar@{-}[d]_-{i+1} &K\otimes (L\otimes M ) \ar@{-}[d] \\ &K \ar@{-}[r]_-i &K\otimes L } \end{equation} Il r{\'e}sulte alors de la propri{\'e}t{\'e} 2. d'une structure du cube, appliqu{\'e}e aux deux fa\c{c}ons d'effectuer des recollements dans (\ref{associativite2}) que les trivialisations des membres de (\ref{associativite1}) sont identifi{\'e}es aux trivialisations de (\ref{associativite3}). \end{proof} De ces diff{\'e}rents r{\'e}sultats, on d{\'e}duit la: \begin{prop} La donn{\'e}e d'une structure du (n+1)-cube $S$ sur $\gd: \mathcal{C} \longrightarrow \mathcal{D}$ munit le $n$-foncteur associ{\'e} $\L : \mathcal{C} ^n \longrightarrow \mathcal{D}$ de donn{\'e}es d'additivit{\'e} $\ast _i$ en chacune des $n$ variables. Ces donn{\'e}es sont compatibles aux donn{\'e}es d'associativit{\'e} et de commutativit{\'e} de $\mathcal{C}$ et $\mathcal{D}$ ainsi qu'aux donn{\'e}es de sym{\'e}trie de $\L$ et sont compatibles entre elles. \end{prop} \begin{proof}[Preuve] Les questions d'associativit{\'e} et de commutativit{\'e} proviennent des r{\'e}sultats analogues pour $\L _{K,i}$ (lemme \ref{comm-assoc}).\\ Pour simplifier les notations exprimons la compatibilit{\'e} des $\ast _i$ entre elles dans le cas $n=2$. Cela se traduit par la commutativit{\'e} du diagramme: \[ \begin{CD} \L (L,N) \otimes \L (L,P) \otimes \L (M,N) \otimes \L (M,P) @>{\ast _1 \otimes \ast _1}>> \L (LM,N) \otimes \L (LM,P)\\ @V{\ast _2 \otimes \ast _2}VV @V{\ast _2 }VV \\ \L (L,NP) \otimes \L (M,NP) @>{\ast _1}>> \L (LM,NP) \end{CD} \] \textsc{Breen} montre dans \cite{B2},2.5 que la commutativit{\'e} de ce diagramme est une cons{\'e}quence de l'associativit{\'e} de $\ast _1$ et $\ast _2$. En effet, cela se traduit en disant que l'isomorphisme canonique d{\'e}duit des morphismes de contraction: \begin{multline*} \ga : \gd (K(LM,N,P)) \otimes \gd (K(L,M,N)) \otimes \gd (K(L,M,P))\\ \stackrel{\sim }{\F } \gd (K(L,M,NP)) \otimes \gd (K(L,N,P)) \otimes \gd (K(M,N,P)) \end{multline*} identifie les trivialisations de chacun des deux termes qui sont d{\'e}duites de la structure du cube. Cette assertion provient alors du fait que $\ga$ se d{\'e}compose en \[ \begin{CD} \gd (K(LM,N,P)) \otimes \gd (K(L,M,N)) \otimes \gd (K(L,M,P))\\ @VVV\\ \gd (K(L,MN,P)) \otimes \gd (K(M,N,P)) \otimes \gd (K(L,M,N))\\ @AAA\\ \gd (K(L,M,NP)) \otimes \gd (K(L,N,P)) \otimes \gd (K(M,N,P)) \end{CD} \] provenant des morphismes d'associativit{\'e} de $\ast _1$ et $\ast _2$.\\ Pour traduire la compatibilit{\'e} des donn{\'e}es d'additivit{\'e} avec les donn{\'e}es de sym{\'e}trie, consid{\'e}rons des objets $L_1, \cdots , L_n,L,L^{\prime}$ de $\mathcal{C}$ et $\sigma \in S_n$ et notons $\underline{L}_i= (L_1,\cdots,L_{i-1},L,L_{i+1},\cdots ,L_n)$, $\underline{L^{\prime}}_i= (L_1,\cdots,L_{i-1},L^{\prime},L_{i+1},\cdots ,L_n)$ et $\underline{L\sec}_i= (L_1,\cdots,L_{i-1},L\otimes L^{\prime} ,L_{i+1},\cdots ,L_n)$ et $j=\sigma ^{-1} (i)$. On veut montrer que le diagramme suivant, dont les fl{\`e}ches verticales proviennent des donn{\'e}es de sym{\'e}trie de $\L$ \[ \begin{CD} \L (\sigma ^{\ast } \underline{L}_i ) \otimes \L (\sigma ^{\ast } \underline{L^{\prime}}_i ) @>{\ast _j}>> \L (\sigma ^{\ast } \underline{L\sec}_i ) \\ @VVV @VVV \\ \L ( \underline{L}_i ) \otimes \L ( \underline{L^{\prime}}_i ) @>{\ast _j}>> \L ( \underline{L\sec}_i ) \end{CD} \] est commutatif. Ceci provient de la propri{\'e}t{\'e} de sym{\'e}trie de la structure de $(n+1)$-cube, appliqu{\'e}e au $(n+1)$-cube $C= (K(L_1,\cdots,L_{i-1},L, L^{\prime},L_{i+1},\cdots ,L_n)$ et {\`a} la permutation $\tau \in S_{n+1} $ telle que $\tau ^{\ast } C = K(L_{\sigma (1)},\cdots,L_{\sigma (i-1)},L, L^{\prime},L_{\sigma (i+1)},\cdots ,L_{\sigma (n)})$. \end{proof} \section{Les cat{\'e}gories de Picard consid{\'e}r{\'e}es} Dans ce chapitre, on consid{\`e}re un morphisme projectif et plat $\pi : X \longrightarrow S$ sur un sch{\'e}ma localement noeth{\'e}rien et on introduit les cat{\'e}gories de Picard qui nous serviront dans la suite. \subsection{} $\mathcal{C}$ d{\'e}signera la cat{\'e}gorie $PIC (X)$ dont les objets sonts les faisceaux inversibles sur $X$, les fl{\`e}ches sont les isomorphismes de $\mathcal{O}_{X}$-modules, $\otimes$ d{\'e}signe le produit tensoriel usuel de deux $\mathcal{O}_{X}$-modules et les morphismes d'associativit{\'e} et de commutativit{\'e} sont ceux usuels. Les objets unit{\'e}s et inverses seront simplement le faisceau structural $\mathcal{O}_{X}$ et le faisceau dual $L^{-1}$. \subsection{} $\mathcal{D}$ d{\'e}signera la cat{\'e}gorie $PICgr (S)$ dont les objets sont les couples $(L,d)$ form{\'e}s d'un faisceau inversible $L$ sur $S$ et d'une application localement constante $d: X \longrightarrow \mathbb{Z}$, les fl{\`e}ches entre deux objets $(L,d)$ et $(M,e)$ n'existent que si $d=e$ et sont dans ce cas les isomorphismes entre les $\mathcal{O}_{S}$-modules $L$ et $M$. Le produit tensoriel sera alors d{\'e}fini par \[ (L,d) \otimes (M,e) = (L\otimes _{\ox} M , d+e). \] Les morphismes de commutativit{\'e} $\psi : (L,d) \otimes (M,e) \longrightarrow (M,e) \otimes (L,d)$ sont donn{\'e}s par: $\psi : L \otimes _{\os} M \longrightarrow M \otimes _{\os} L : l \otimes m \mapsto (-1)^{d.e} m\otimes l$. On notera que la cat{\'e}gorie $PICgr (S)$ n'est pas strictement commutative.\\ L'objet unit{\'e} de $\mathcal{D}$ sera alors $\mathcal{O} = (\mathcal{O}_{S} ,0)$ et l'inverse de $(L,d)$ sera $(L^{-1} ,-d)$. Le morphisme d'{\'e}valuation $(L,d) \otimes (L^{-1} ,-d)\longrightarrow \mathcal{O}$ est donn{\'e} par l'{\'e}valuation usuelle $L\otimes _{\os} L^{-1} \simeq \mathcal{O}_{S}$. On prendra garde que le morphisme $(L^{-1} ,-d) \otimes (L,d)\longrightarrow \mathcal{O}$ est donn{\'e} par le morphisme usuel d'{\'e}valuation multipli{\'e} par $(-1)^d$.\\ On consid{\`e}rera enfin le foncteur $\gd : \mathcal{C} \longrightarrow \mathcal{D}$ donn{\'e} par le d{\'e}terminant de l'image directe d{\'e}riv{\'e}e dont l'existence, annonc{\'e}e par \textsc{Grothendieck}, est montr{\'e}e par \textsc{Knudsen} et \textsc{Mumford} dans \cite{KM}. $\gd$ associe {\`a} tout $\mathcal{O}_{X}$-module inversible $L$, le $\mathcal{O}_{S}$-module $\det \text{R}\pi \lst L $, gradu{\'e} par la fonction localement constante $ s \mapsto \chi (L | _{X_s})$.\\ En r{\'e}alit{\'e}, le foncteur $\gd$ est d{\'e}fini sur la cat{\'e}gorie plus {\'e}tendue $COH (X/S)$ des $\mathcal{O}_{X}$-modules coh{\'e}rents et plats sur $S$. Il v{\'e}rifie les propri{\'e}t{\'e}s suivantes: \subsubsection{} Toute suite exacte dans $COH (X/S)$: \[ 0 \longrightarrow E \longrightarrow F \longrightarrow G \longrightarrow 0 \] induit un isomorphisme canonique de multiplicativit{\'e} $\gd (E) \otimes \gd (G) \stackrel{\sim }{\F } \gd (F)$ et pour tout diagramme de suites exactes courtes: \[ \begin{CD} @. 0 @. 0 @. 0 \\ @.@VVV @VVV @VVV\\ 0 @>>> E^{\prime} @>>>F^{\prime} @>>> G^{\prime} @>>>0\\ @. @VVV @VVV @VVV\\ 0 @>>> E @>>>F @>>> G @>>>0\\ @. @VVV @VVV @VVV\\ 0 @>>> E\sec @>>>F\sec @>>> G\sec @>>>0\\ @. @VVV @VVV @VVV\\ @. 0 @. 0 @.0 \end{CD} \] le diagramme \begin{equation} \label{diagramme-des-neuf} \begin{CD} \gd (E^{\prime} ) \otimes \gd (E\sec ) \otimes \gd (G^{\prime} )\otimes \gd (G\sec ) @>>> \gd (E) \otimes \gd (G)\\ @VVV @VVV \\ \gd (F^{\prime} ) \otimes \gd (F\sec ) @>>> \gd (F) \end{CD} \end{equation} qu'on en d{\'e}duit par application des morphismes de multiplicativit{\'e} et de commutativit{\'e} est commutatif (\cite{KM},prop.1). Notons que c'est la d{\'e}finition des isomorphismes de commutativit{\'e} dans $\mathcal{D}$ qui rend possible la commutativit{\'e} du diagramme pr{\'e}c{\'e}dent. \subsubsection{}Si $E$ est un $\mathcal{O}_{X}$-module coh{\'e}rent et plat sur $S$, {\`a} support dans un sous-sch{\'e}ma $Y$ de $X$, on a: $\det \text{R}\pi \lst (E) = \det \text{R} (\pi _{Y/S}) _{\ast } (E|_Y)$. \subsubsection{} La formation de $\det \text{R}\pi \lst $ commute aux changements de base. \subsection{} On consid{\`e}rera enfin la cat{\'e}gorie de Picard strictement commutative $\mathcal{D}^{\prime} =PIC (S)$ et on notera $\delta ^{\prime}$ le compos{\'e} du foncteur $\gd = \det \text{R}\pi \lst : \mathcal{C} \longrightarrow \mathcal{D}$ avec le foncteur oubli de la graduation $\mathcal{D} \longrightarrow \mathcal{D}^{\prime}$. On notera qu'en travaillant dans cette cat{\'e}gorie $\mathcal{D}^{\prime}$ on gagne le fait qu'elle est strictement commutative, mais on perd la possibilit{\'e} d'{\'e}crire certains diagrammes commutatifs du paragraphe pr{\'e}c{\'e}dent, comme par exemple (\ref{diagramme-des-neuf}), qui ne s'expriment naturellement que gr{\^a}ce au foncteur $\gd$. \subsection{Structure du cube dans la cat{\'e}gorie des faisceaux inversibles gradu{\'e}s} Pour montrer l'existence {\'e}ventuelle d'un structure du $p$-cube sur le foncteur $\delta ^{\prime} : \mathcal{C} \longrightarrow \mathcal{D}^{\prime}$ entre cat{\'e}gories de Picard strictement commutatives, il sera n{\'e}cessaire de passer par l'interm{\'e}diaire du foncteur $\gd : \mathcal{C} \longrightarrow \mathcal{D}$ {\`a} valeurs dans une cat{\'e}gorie de Picard non strictement commutative. Examinons ici comment les axiomes d'une structure de $p$-cube sur $\delta ^{\prime}$ se traduisent en termes de $\gd$. \begin{nota} Si $K$ est un $p$-cube dans $\mathcal{C}$, pour tout couple d'indices distincts $i$ et $j$, les isomorphismes de commutativit{\'e} dans $\mathcal{D}$ induisent un isomorphisme $\gd (\sigma _{ij}^{\ast } K)\longrightarrow \gd (K)$ dans $\mathcal{D}$, qui induit, par oubli de la graduation, un isomorphisme $\delta ^{\prime} (\sigma _{ij}^{\ast } K)\longrightarrow \delta ^{\prime} (K)$ dans $\mathcal{D}^{\prime}$. Celui ci diff{\`e}re de celui induit par les isomorphismes de commutativit{\'e} dans $\mathcal{D}^{\prime}$ par un signe, que l'on notera $\ge _{ij}(K)$. \end{nota} \begin{rem} Si $X\longrightarrow S$ est {\`a} fibres de dimension $n$, pour tout $(n+2)$-cube \begin{equation} \label{ecriture-cube} K=\left( \begin{array}{c} \xymatrix{C\ar@{-}[d]_-j\ar@{-}[r] &D \ar@{-}[d] \\ A\ar@{-}[r]_-i &B } \end{array} \right) , \end{equation} on a $\chi (A) = \chi (B) = \chi (C) = \chi (D)$ (d'apr{\`e}s l'exemple \ref{caracteristique2}) et de plus $\ge _{ij}(K)=(-1)^{\chi (A)}$. \end{rem} \addtocounter{subsubsection}{2} \subsubsection{} \addtocounter{theo}{1} Soit $K$ un $p$-cube dans $\mathcal{C}$ et soit $t: \mathcal{O}_{S} \stackrel{\sim }{\F } \delta ^{\prime} (K)$ une trivialisation de $\delta ^{\prime} (K)$. Pour tout entier $1\leq i\leq p$, $t$ induit un isomorphisme $t_i: \delta ^{\prime} (\phi^{\prime\ast}_i K) \stackrel{\sim }{\F } \delta ^{\prime} (\phi^{\prime \prime\ast}_i K)$ dans $\mathcal{D}^{\prime}$. Le choix de l'ordre usuel sur les sommets de $\phi^{\prime\ast}_i K$ et $\phi^{\prime \prime\ast}_i K$ associe {\`a} $t_i$ un isomorphisme $s_i: \gd (\phi^{\prime\ast}_i K) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i K)$ dans $\mathcal{D}$. Exprimons alors, {\`a} l'int{\'e}rieur de la cat{\'e}gorie $\mathcal{D}^{\prime}$ les relations entre $s_i$ et $s_j$, pour $i\neq j$. On peut associer {\`a} $s_i$ le morphisme dans $\mathcal{D}$: \[ \begin{CD} \overline{s_i}:\mathcal{O} @>>> (\gd (\phi^{\prime\ast}_i K))^{-1} \otimes \gd ( \phi^{\prime\ast}_i K) @>{\text{Id}\otimes s_i}>> (\gd (\phi^{\prime\ast}_i K))^{-1} \otimes \gd ( \phi^{\prime \prime\ast}_i K) @>>> \gd (K) \end{CD} \] \begin{lemme} La donn{\'e}e d'une trivialisation $t$ de $\delta ^{\prime}(K)$ est {\'e}quivalente {\`a} la donn{\'e}e d'une collection $(s_i)_{1\leq i\leq p}$ d'isomorphismes $s_i: \gd (\phi^{\prime\ast}_i K) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i K)$ dans $\mathcal{D}$ telle que, pour deux indices $i$ et $j$ distincts, les trivialisations $\overline{s_i}$ et $\overline{s_j}$ de $\gd (K)$ induites par $s_i$ et $s_j$ diff{\`e}rent d'un signe $\ge _{ij}(K)$. \end{lemme} \addtocounter{subsubsection}{1} \begin{proof}[Preuve] Reprenons la notation (\ref{ecriture-cube}). Le diagramme, dont les fl{\`e}ches verticales sont donn{\'e}es par les isomorphismes les isomorphismes structuraux de $\mathcal{D}^{\prime}$ est commutatif: \[ \begin{CD} (\delta ^{\prime} (A) \otimes _{\os} \delta ^{\prime} (C)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (C)\otimes _{\os}\delta ^{\prime} (D)) @>{t_i\otimes \text{id} }>> (\delta ^{\prime} (B) \otimes _{\os} \delta ^{\prime} (D)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (C)\otimes _{\os}\delta ^{\prime} (D))\\ @VVV @VVV \\ (\delta ^{\prime} (A) \otimes _{\os} \delta ^{\prime} (D) ) \otimes _{\os} (\delta ^{\prime} (C)\otimes _{\os}\delta ^{\prime} (C)^{\vee } ) @. (\delta ^{\prime} (B) \otimes _{\os} \delta ^{\prime} (C)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (D)\otimes _{\os}\delta ^{\prime} (D)^{\vee } )\\ @VVV @VVV \\ (\delta ^{\prime} (A) \otimes _{\os} \delta ^{\prime} (D) ) @>t>> (\delta ^{\prime} (B) \otimes _{\os} \delta ^{\prime} (C) ) \\ @AAA @AAA \\ (\delta ^{\prime} (A) \otimes _{\os} \delta ^{\prime} (D) ) \otimes _{\os} (\delta ^{\prime} (B)\otimes _{\os}\delta ^{\prime} (B)^{\vee } ) @. (\delta ^{\prime} (C) \otimes _{\os} \delta ^{\prime} (B)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (D)\otimes _{\os}\delta ^{\prime} (D)^{\vee } )\\ @AAA @AAA \\ (\delta ^{\prime} (A) \otimes _{\os} \delta ^{\prime} (B)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (B)\otimes _{\os}\delta ^{\prime} (D)) @>{t_j\otimes \text{id} }>> (\delta ^{\prime} (C) \otimes _{\os} \delta ^{\prime} (D)^{\vee } ) \otimes _{\os} (\delta ^{\prime} (B)\otimes _{\os}\delta ^{\prime} (D)) \end{CD} \] Consid{\'e}rons le diagramme obtenu en rempla\c{c}ant dans le pr{\'e}c{\'e}dent $\delta ^{\prime}$ par $\gd$ et en utilisant les morphismes structuraux de $\mathcal{D}^{\prime}$. Son circuit ext{\'e}rieur est donc commutatif {\`a} un signe $(\ge _{ij})^k$ pr{\`e}s, o{\`u} $k$ est le nombre de transpositions apparaissant dans le diagramme. On constate alors que ce nombre est impair.\\ R{\'e}ciproquement, si $(s_i)_{1\leq i\leq p}$ est une telle collection de morphismes, chaque $s_i$ induit un isomorphisme $t_i:\delta ^{\prime} \left(\phi^{\prime\ast}_i K \right) \stackrel{\sim }{\F } \delta ^{\prime}\left( \phi^{\prime \prime\ast}_i K \right)$ dans $\mathcal{D}^{\prime}$. La trivialisation de $\delta ^{\prime} (K)$ dans $\mathcal{D}^{\prime}$ est alors ind{\'e}pendante de $i$. \end{proof} \subsubsection{Recollements de cubes} \addtocounter{theo}{1} Soient $K$ et $K^{\prime}$ deux $p$-cubes dans $\mathcal{C}$ et $i$ un indice tel que $\phi^{\prime\ast}_i K = \phi^{\prime \prime\ast}_i K$. Si $t , t^{\prime}$ sont des trivialisations de $\delta ^{\prime} (K)$ et $\delta ^{\prime} (K^{\prime} )$, elles induisent une trivialisation $t \ast_i t^{\prime}: \mathcal{O}_{X} \stackrel{\sim }{\F } \mathcal{O}_{X} \otimes \mathcal{O}_{X} \stackrel{\sim }{\F } \delta ^{\prime} (K) \otimes \delta ^{\prime} (K^{\prime} ) \stackrel{\sim }{\F } \delta ^{\prime} (K\ast_i K^{\prime} )$. Notons $s_i: \gd (\phi^{\prime\ast}_i K) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i K)$, $s^{\prime}_i: \gd (\phi^{\prime\ast}_i K^{\prime} ) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i K^{\prime} )$ et $s\sec_i: \gd (\phi^{\prime\ast}_i (K\ast _i K^{\prime} )) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i (K\ast _i K^{\prime} ))$ les isomorphismes dans $\mathcal{D}^{\prime}$ associ{\'e}s {\`a} $t$, $t^{\prime}$ et $t \ast_i t^{\prime}$. On a alors $s\sec_i =s^{\prime}_i \circ s_i$. On peut alors regrouper les r{\'e}sultats de ce paragraphe dans la \begin{prop} \label{cubebis} Une structure de $p$-cube sur $\delta ^{\prime}$ est {\'e}quivalente {\`a} la donn{\'e}e, pour tout $p$-cube $K$ dans $\mathcal{C}$, de $p$ isomorphismes dans $\mathcal{D}$: \[ s_{K,i}:\gd \left(\phi^{\prime\ast}_i K \right) \stackrel{\sim }{\F } \gd\left( \phi^{\prime \prime\ast}_i K \right) \] tels que: \begin{enumerate} \item Pour tout isomorphisme de $p$-cubes $f: K \stackrel{\sim }{\F } K^{\prime}$, le diagramme induit \[ \begin{CD} \gd \left( \phi^{\prime\ast}_i K \right) @>>> \gd\left( \phi^{\prime \prime\ast}_i K \right)\\ @VVV @VVV\\ \gd \left( \phi^{\prime\ast}_i K^{\prime} \right) @>>> \gd\left( \phi^{\prime \prime\ast}_i K^{\prime} \right) \end{CD} \] est commutatif. \item Soient $K$ et $K^{\prime}$ deux $p$-cubes ayant leur i-{\`e}me $(p-1)$-face en commun, on a l'{\'e}galit{\'e}: \[ s_{K\ast _i K^{\prime} ,i} = s_{ K^{\prime} ,i} \circ s_{K ,i} \] \item Les trivialisations de $\gd (K)$ induites par $s_{K,i}$ et $s_{K,j}$ diff{\`e}rent d'un signe $\ge _{ij}(K)$. \item Pour toute permutation $\sigma _{ij} \in S_p$ de deux indices distincts $i$ et $j$, on a: \[ s_{\sigma _{ij}^{\ast } (K),i} = s_{K,i} \] o{\`u} d{\'e}signe la permutation des indices $i$ et $j$. \end{enumerate} \end{prop} \section{Constructions de structures du cube} \subsection{Cas de la dimension 0} \label{norme} Si $\pi : X \longrightarrow S$ est un morphisme fini et plat, le foncteur $\gd =\det \text{R}\pi \lst $ se r{\'e}duit au foncteur: \[ \gd : PIC (X) \longrightarrow PICgr (S)\; ,\; L \mapsto \left(\det (\pi _{\ast } L), \deg \pi \right) . \] et $\gd ^{\prime} : PIC (X) \longrightarrow PIC (S)$ est simplement le foncteur: $ L\mapsto \det (\pi _{\ast } L)$. \subsubsection{Norme } Rappellons quelques propri{\'e}t{\'e}s de la norme (cf \cite{EGA2},6.5 et \cite{FD1},3.1) pour un morphisme fini et plat. Notons d'abord que, comme $X$ est fini sur $S$, pour tout faisceau inversible $L$ sur $X$ et tout $s\in S$, il existe un ouvert $U$ de $S$ contenant $s$ tel que $L$ est trivial sur $X_U$. Si $\ga$ est une section inversible de $\mathcal{O}_{X}$, c'est {\`a} dire un automorphisme de $\mathcal{O}_{X}$, la norme $N_{X/S} (\ga )$ est le d{\'e}terminant de l'automorphisme de $\det \pi _{\ast } \mathcal{O}_{X}$ induit par $\ga$. Enfin la norme du faisceau inversible $L$ est par d{\'e}finition le $\mathcal{O}_{S}$-module inversible \[ N_{X/S} (L) = (\det \pi _{\ast } L) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} )^{-1} \] Si $(U_i)_{i\in I}$ est un recouvrement ouvert de $S$ tel que $L$ est trivial sur $X_{U_i}$ avec des fonctions de transition $(g_{ij})_{i,j \in I}$, alors $N_{X/S} (L)$ est trivial sur chaque $U_i$ et a pour fonctions de transition les $N_{X/S}(g_{ij})$. \subsubsection{} Construisons une structure du 2-cube (carr{\'e}), qui traduit les propri{\'e}t{\'e}s de la norme pour un morphisme fini et plat: Il s'agit de construire, pour tous faisceaux inversiblest $L$, $M$, $N$ et $P$ sur $X$ et tout isomorphisme $\phi : L \otimes _{\ox} M \stackrel{\sim }{\F } N \otimes _{\ox} P$, un isomorphisme \[ (\det \pi _{\ast } L) \otimes _{\os} (\det \pi _{\ast } M) \stackrel{\sim }{\F } (\det \pi _{\ast } N) \otimes _{\os} (\det \pi _{\ast } P) . \] D'apr{\`e}s les remarques de la section pr{\'e}c{\'e}dente, il suffit de construire pour tout isomorphisme $\phi : \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X} \stackrel{\sim }{\F } \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X}$ un isomorphisme $ \gd _{\phi}:\det (\pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} \det (\pi _{\ast } \mathcal{O}_{X} ) \stackrel{\sim }{\F } \det (\pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} \det (\pi _{\ast } \mathcal{O}_{X} )$ tel que si $\phi$ et $\psi$ sont deux tels isomorphismes et $\ga$, $\gb$, $\gc$ et $\gd$ des automorphismes de $\mathcal{O}_{X}$ tels que le diagramme \[ \begin{CD} \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X} @>\phi >> \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X} \\ @V{\ga \otimes \gb}VV @VV{\gc \otimes \gd}V\\ \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X} @>\psi >> \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X} \end{CD} \] est commutatif, alors le diagramme induit: \[ \begin{CD} (\det \pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} ) @>{\det \pi _{\ast }\phi}>> (\det \pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} ) \\ @V{N_{X/S} (\ga ) \otimes N_{X/S}(\gb )}VV @VV{N_{X/S}(\gc ) \otimes N_{X/S}(\gd )}V \\ (\det \pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} ) @>{\det \pi _{\ast }\psi}>> (\det \pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} ) \end{CD} \] est commutatif. $\phi$ est un automorphisme de $ \mathcal{O}_{X} \otimes _{\ox} \mathcal{O}_{X}$, c'est {\`a} dire une section inversible de $\mathcal{O}_{X}$. On peut consid{\'e}rer la section inversible $N_{X/S}(\phi )$ de $\mathcal{O}_{S}$, qui d{\'e}finit donc l'automorphisme recherch{\'e} de $(\det \pi _{\ast } \mathcal{O}_{X} ) \otimes _{\os} (\det \pi _{\ast } \mathcal{O}_{X} )$. La condition de fonctorialit{\'e} se traduit en disant que, si $\ga , \gb , \gc ,\gd , \phi ,\psi$ sont des sections inversibles de $\mathcal{O}_{X}$ telles que $\ga\gb\psi =\gc\gd\phi$, alors \[ N_{X/S}(\ga ) N_{X/S}(\gb) N_{X/S}(\psi ) = N_{X/S}(\gc ) N_{X/S}(\gd ) N_{X/S}(\phi )\; , \] ce qui est simplement la mutiplicativit{\'e} de la norme.\\ Le lemme suivant est une cons{\'e}quence imm{\'e}diate de la d{\'e}finition de la norme d'un faisceau inversible et des propri{\'e}t{\'e}s de multiplicativit{\'e} de la norme. \addtocounter{theo}{2} \begin{lemme} \begin{enumerate} \item La construction pr{\'e}c{\'e}dente d{\'e}termine une structure du carr{\'e} sym{\'e}trique sur $\delta ^{\prime}$. \item Le foncteur lin{\'e}aire $PIC(X) \longrightarrow PIC (S)$ d{\'e}duit de cette structure est la norme relativement au morphisme fini et plat $\pi$. \end{enumerate} \end{lemme} \begin{rem} \label{carre} Consid{\'e}rons un carr{\'e} $ \left( \begin{array}{c} \xymatrix{N\ar@{-}[r]\ar@{-}[d]&P\ar@{-}[d]\\ L\ar@{-}[r]&M } \end{array} \right) $ dans $PIC(X)$ correspondant {\`a} un isomorphisme $\phi: L\otimes P \stackrel{\sim }{\F } M \otimes N$ et soient des isomorphismes $\ga : L \stackrel{\sim }{\F } M$ et $\gb : N \stackrel{\sim }{\F } P$ tels que le diagramme \[ \begin{CD} L\otimes P @>{\phi}>> M \otimes N \\ @V{\ga \otimes \text{id}}VV @V{\text{id}\otimes \gb}VV \\ M \otimes P @= M \otimes P \end{CD} \] soit commutatif, alors le diagramme suivant, obtenu par application de $\det \pi _{\ast }$ l'est aussi: \[ \begin{CD} \det\pi_{\ast } L\otimes \det\pi_{\ast } P @>{\gd_{\phi}}>> \det\pi_{\ast } M \otimes \det\pi_{\ast } N \\ @V{\det (\ga )\otimes \text{id}}VV @V{\text{id}\otimes \det (\gb )}VV \\ \det\pi_{\ast } M \otimes \det\pi_{\ast } P @= \det\pi_{\ast } M \otimes \det\pi_{\ast } P \end{CD} \] \end{rem} \subsection{Restriction {\`a} un diviseur effectif} \label{restriction} Si $\pi : X\longrightarrow S$ un morphisme projectif et plat sur $S$ localement noeth{\'e}rien et $D$ est un diviseur relatif sur $X$, si $K$ est un p-cube dans $\mathcal{C} = PIC(X)$, consid{\'e}rons le (p+1)-cube $A =( \xymatrix{K \ar@{-}[r]_-i &K\otimes \mathcal{O} (D) })$. On a alors un isomorphisme canonique $r_i: \gd (K^{\prime} ) \stackrel{\sim }{\F } \gd (K\otimes \mathcal{O} (D)|_D)$ dans $\mathcal{D} = PICgr (S)$, qu'on appellera {\em isomorphisme de restriction}. Il est donn{\'e} en appliquant, pour chaque sommet $L$ de $K$, le foncteur $\gd$ {\`a} la suite exacte: \[ 0 \longrightarrow L \longrightarrow L(D) \longrightarrow L(D)|_D \longrightarrow 0 \] \subsubsection{Application au cas des courbes} \label{cas-des-courbes} Soit $\pi :X\longrightarrow S$ un morphisme projectif et plat, {\`a} fibres de dimension 1, sur un sch{\'e}ma localement noeth{\'e}rien $S$. Soient $D$ et $E$ deux diviseurs de Cartier relatifs effectifs et $L$ un faisceau inversible sur $X$, consid{\'e}rons alors le carr{\'e} $A=K_L(\mathcal{O}_{X} (D),\mathcal{O}_{X} (E))$. Les consid{\'e}rations pr{\'e}c{\'e}dentes nous donnent un isomorphisme $ r_1:\gd(A) \stackrel{\sim }{\F } \gd( \xymatrix{ L(D)|_D \ar@{-}[r]_-1&L(D+E)|_D }) $ et, comme $D$ est fini et plat sur $S$, on obtient un isomorphisme \begin{equation} \label{isom-dim1} \gd (A)\stackrel{\sim }{\F } (\det \pi _{\ast } L(D)|_D)^{-1} \otimes (\det \pi _{\ast } L(D+E)|_D). \end{equation} La section canonique de $\mathcal{O}_{X}(E)$ induit un morphisme $\pi _{\ast } L(D)|_D \longrightarrow \pi _{\ast } L(D+E)|_D$, dont le d{\'e}terminant d{\'e}finit donc une section $s_L(D,E)$ de $\gd(A)$. \addtocounter{theo}{1} \begin{lemme} \label{indep2} L'isomorphisme canonique $\gd(K_L(\mathcal{O}_{X} (D),\mathcal{O}_{X} (E))) \stackrel{\sim }{\F } \gd(K_L(\mathcal{O}_{X} (E),\mathcal{O}_{X} (D)))$, donn{\'e} par les isomorphismes de commutativit{\'e} dans $PICgr(S)$, {\'e}change les sections $s_L(D,E)$ et $s_L(E,D)$ de ces deux faisceaux. \end{lemme} \begin{proof}[Preuve] D{\'e}crivons l'isomorphisme (\ref{isom-dim1}). Consid{\'e}rons le diagramme commutatif de suites exactes: \[ \xymatrix{ &L(E)|_E \ar@{^{(}->}[rr]\ar@{}[rd]|{B} &&L(D+E)|_{D+E} \ar@{->>}[rr] &&L(D+E)|_D \\ 0\ar[ru]\ar[rr]&&L(D)|_D\ar[ru]\ar@{=}[rr]&&L(D)|_D \ar[ru]\\ &L(E)\ar@{^{(}->}'[r][rr]\ar@{->>}'[u][uu]\ar@{}[rd]|{A} &&L(D+E)\ar@{->>}'[r][rr]\ar@{->>}'[u][uu] &&L(D+E)|_D\ar[uu] \\ L\ar@{^{(}->}[rr]\ar[ru]\ar[uu] &&L(D)\ar@{->>}[rr]\ar[ru]\ar@{->>}[uu] &&L(D)|_D\ar[uu]\ar[ru] \\ &L\ar@{=}'[r][rr]\ar@{^{(}->}'[u][uu]&&L\ar@{^{(}->}'[u][uu]\\ L\ar@{=}[rr]\ar@{=}[ru]\ar@{=}[uu]&&L\ar@{^{(}->}[uu] \ar@{=}[ru] } \] On en d{\'e}duit que l'isomorphisme (\ref{isom-dim1}) se d{\'e}compose en $\gd(A)\stackrel{\sim }{\F }\gd(B)\stackrel{\sim }{\F } (\det \pi _{\ast } L(D)|_D)^{-1} \otimes (\det \pi _{\ast } L(D+E)|_D)$, o{\`u} le premier isomorphisme est donn{\'e} par les suites exactes verticales et le second isomorphisme est obtenu en appliquant le foncteur d{\'e}terminant au diagramme suivant de $\mathcal{O}_{S}$-modules localement libres \[ \begin{CD} \pi _{\ast } L(E)|_E @>{\gb}>> \pi _{\ast } L(D+E)|_E @. \\ @| @AAA @. \\ \pi _{\ast } L(E)|_E @>>> \pi _{\ast } L(D+E)|_{D+E} @>>> \pi _{\ast } L(D+E)|_D\\ @AAA @AAA @A{\ga}AA\\ 0@>>> \pi _{\ast } L(D)|_D @= \pi _{\ast } L(D)|_D \end{CD} \] On traduit l'assertion du lemme en disant que la suite d'isomorphismes dans $PICgr(S)$ \begin{multline*} \det( \pi _{\ast } L(D)|_D)^{\vee } \otimes \det( \pi _{\ast } L(D+E)|_D)\stackrel{\sim }{\F } \\ \det( \pi _{\ast } L(D)|_D)^{\vee } \otimes \det( \pi _{\ast } L(E)|_E)^{\vee } \otimes\det( \pi _{\ast } L(D+E)|_{D+E}) \stackrel{\sim }{\F } \\ \det( \pi _{\ast } L(E)|_E)^{\vee } \otimes\det( \pi _{\ast } L(D)|_D)^{\vee } \otimes \det( \pi _{\ast } L(D+E)|_{D+E}) \stackrel{\sim }{\F } \\ \det( \pi _{\ast } L(E)|_E)^{\vee } \otimes \det( \pi _{\ast } L(D+E)|_E) \end{multline*} fait correspondre les sections $\det(\ga)$ et $\det(\gb)$ des deux termes extr{\`e}mes, ce qui provient des propri{\'e}t{\'e}s du d{\'e}terminant des $\mathcal{O}_{S}$-modules localement libres. \end{proof} \addtocounter{ptheo}{2} \begin{ptheo} \addtocounter{subsection}{1} \label{th-prin} Soit $S$ un sch{\'e}ma localement noeth{\'e}rien. Pour tout morphisme projectif et plat $\pi : X \longrightarrow S$ {\`a} fibres de dimension n, il existe une structure du (n+2)-cube canonique sur le foncteur $\delta ^{\prime} : PIC(X) \longrightarrow PIC (S)$ telle que: \begin{enumerate} \item Si $\pi$ est un morphisme fini, la structure du carr{\'e} correspondante est simplement celle donn{\'e}e par la norme (\ref{norme}). \item Si $D$ est un diviseur relatif sur $X$, les structures du (n+2)-cube sur $\delta ^{\prime} _X: PIC(X)\longrightarrow PIC(S)$ et du (n+1)-cube sur $\delta ^{\prime} _D: PIC(D) \longrightarrow PIC(S)$ sont compatibles aux isomorphismes de restriction. \end{enumerate} \end{ptheo} \begin{proof}[Preuve] La preuve s'effectue par r{\'e}currence sur $n$. Le cas de la dimension 0 a {\'e}t{\'e} trait{\'e} dans (\ref{norme}). Soit n un entier strictement positif, supposons que, pour tout morphisme projectif et plat $Y\longrightarrow S$ {\`a} fibres de dimension $p<n$, on sache construire une structure du (p+2)-cube sur $\delta ^{\prime}_Y$ verifiant les propri{\'e}t{\'e}s 1 et 2 du th{\'e}or{\`e}me. Soit $\pi :X\longrightarrow S$ projectif et plat {\`a} fibres de dimension n, construisons une structure du (n+2)-cube sur $\delta ^{\prime}_X$. D'apr{\`e}s la proposition (\ref{cubebis}), on doit donc construire, pour tout $(n+2)$-cube $A$ dans $PIC(X)$, des isomorphismes $ s_{A,i} : \gd ((\phi ^{\prime}_i)^{\ast } A) \stackrel{\sim }{\F } \gd ((\phi \sec_i)^{\ast } A $ v{\'e}rifiant des propri{\'e}t{\'e}s de sym{\'e}trie et de compatibilit{\'e} aux recollement de $(n+2)$-cubes. Cette construction, qui occupe les paragraphes suivants sera d{\'e}coup{\'e}e de la fa\c{c}on suivante: \begin{itemize} \item Si un $(n+2)$-cube $A$ poss{\`e}de une ar{\^e}te $L_j$ ayant une section r{\'e}guli{\`e}re $\sigma_j$, on construit un isomorphisme $ s_{A,i,\sigma_j} : \gd ((\phi ^{\prime}_i)^{\ast } A) \stackrel{\sim }{\F } \gd ((\phi \sec_i)^{\ast } A $ d{\'e}pendant du choix de $\sigma_j$. \item On s'affranchit de la d{\'e}pendance en la section $\sigma_j$, sous l'hypoth{\`e}se que $A$ poss{\`e}de deux ar{\^e}tes $L_j$ et $L_k$ suffisamment positives (on dira dans ce cas que $A$ est suffisamment positif dans les directions $i$ et $j$). \item On {\'e}limine enfin cette hypoth{\`e}se de positivit{\'e}; \end{itemize} \subsection{Construction et propri{\'e}t{\'e}s de $s_{A,i,\sigma_j}$ } \label{constr-avec-div} \begin{constr} \label{constr1} Une section r{\'e}guli{\`e}re $\sigma_j$ de la $j$-i{\`e}me ar{\^e}te $L_j$ de $A$ d{\'e}finit un diviseur de Cartier relatif $D$ et $A$ est isomorphe {\`a} un $(n+2)$-cube $( \xymatrix{K \ar@{-}[r]_-j& K\otimes \mathcal{O}_{X} (D)})$. On dispose donc, d'apr{\`e}s (\ref{restriction}), d'isomorphismes de restriction $ \gd ((\phi ^{\prime} _i)^{\ast } A) \stackrel{\sim }{\F } \gd ((\phi ^{\prime} _i)^{\ast } K(D)|_D) $ et $ \gd ((\phi \sec _i)^{\ast } A) \stackrel{\sim }{\F } \gd ((\phi \sec _i)^{\ast } K(D)|_D) $, provenant du diagramme de suites exactes \[ \begin{CD} 0 @>>> (\phi ^{\prime} _i)^{\ast } K @>>> (\phi ^{\prime} _i)^{\ast } K(D) @>>> (\phi ^{\prime} _i)^{\ast } K(D)|_D @>>> 0 \end{CD} \] et du diagramme analogue faisant intervenir $(\phi \sec _i)^{\ast } K$. L'hypoth{\`e}se de r{\'e}currence, appliqu{\'e}e au $(n+1)$-cube $K(D)|_D$ sur le sch{\'e}ma relatif $D/S$ de dimension $n-1$ entra{\^\i}ne l'existence d'un isomorphisme \[ s_{K(D)|_D,i} : \gd ((\phi ^{\prime} _i)^{\ast } K(D)|_D ) \stackrel{\sim }{\F } \gd ((\phi \sec _i)^{\ast } K(D)|_D). \] En composant $s_{K(D)|_D,i}$ avec les isomorphismes pr{\'e}c{\'e}dents, on obtient $s_{A,i,\sigma _j}$. \end{constr} \addtocounter{subsubsection}{1} \subsubsection{Comportement de $s_{A,i,\sigma _j}$ par recollement.} \label{recol_div} Si deux $(n+2)$-cubes $A$ et $B$ sont recollables le long de leurs i-{\`e}me face, ils ont alors m{\^e}me j-i{\`e}me ar{\^e}te $L_j$. Si $L_i$ poss{\`e}de une section r{\'e}guli{\`e}re $\sigma_j$ d{\'e}finissant un diviseur relatif $D$, on peut alors {\'e}crire $A\simeq(\xymatrix{K\ar@{-}[r]_-j &K\otimes\mathcal{O}(D)})$ et $B=\simeq(\xymatrix{K^{\prime} \ar@{-}[r]_-j &K^{\prime}\otimes\mathcal{O}(D)})$, avec $K=( \xymatrix{K_1 \ar@{-}[r]_-i &K_2 })$ et $K^{\prime} =( \xymatrix{K_2 \ar@{-}[r]_-i &K_3 })$. En {\'e}crivant le diagramme suivant, dont chaque ligne est exacte: \[ \xymatrix{ 0 \ar[r] & K_3 \ar@{-}[d] \ar[r] \ar@{}[dr]|{B} & K_3(D) \ar@{-}[d] \ar[r] & K_3(D)|_D \ar@{-}[d] \ar[r] &0 \\ 0 \ar[r] & K_2 \ar@{-}[d]_-i \ar[r] \ar@{}[dr]|{A}& K_2(D) \ar@{-}[d] \ar[r] & K_2(D)|_D \ar@{-}[d] \ar[r] &0 \\ 0 \ar[r] & K_1 \ar[r]_-j & K_1(D) \ar[r] & K_1(D)|_D \ar[r] &0 } \] et en utilisant que, d'apr{\`e}s l'hypoth{\`e}se de r{\'e}currence, \[ s_{(K(D)|_D \ast _i K^{\prime}(D)|_D),i} = s_{ K^{\prime}(D)|_D,i} \circ s_{K(D)|_D ,i}\; , \] on obtient: \[ s_{A \ast _i B,i,\sigma _j} = s_{B,i,D} \circ s_{A,i,\sigma _j}. \] \begin{lemme}[Lien entre $s_{A,i,\sigma _j}$ et $s_{A,i,\sigma _k}$] \label{indep} Supposons que les ar{\^e}tes $L_j$ et $L_k$ du $n+2$-cube $A$ poss{\`e}dent des sections r{\'e}guli{\`e}res $\sigma _j$ et $\sigma _k$ et soit $i\neq j,k$. Si l'une des deux hypoth{\`e}ses suivantes est v{\'e}rifi{\'e}e: \begin{enumerate} \item $X/S$ est {\`a} fibres de dimension 1 et $Z(\sigma_j)\cap Z(\sigma_k)=\emptyset$. \item $X/S$ est {\`a} fibres de dimension $n>1$ et les suites $(\sigma _j,\sigma _k)$ et $(\sigma _k,\sigma _j)$ sont $\pi$-r{\'e}guli{\`e}res \end{enumerate} On a alors $s_{A,i,\sigma _j}= s_{A,i,\sigma _k}$. \end{lemme} \begin{proof}[Preuve] Soient $D$ et $E$ les diviseurs de Cartier relatifs effectifs d{\'e}finis par $s_{A,i,\sigma _j}$ et $s_{A,i,\sigma _k}$, on {\'e}crit alors $A$ sous la forme $ A=\left( \begin{array}{c} \xymatrix{ K(E) \ar@{-}[r]\ar@{-}[d]_-k &K (D+E)\ar@{-}[d] \\ K \ar@{-}[r]_-j &K (D) } \end{array} \right) $, o{\`u} $K$ est un $n$-cube. Ecrivons le diagramme suivant, dont les lignes et les colonnes sont des suites exactes courtes, et qui relie le $(n+2)$-cube $A$ sur X aux $(n+1)$-cubes $A^{\prime}$ et $A\sec$ sur $E$ et $D$ et au $n$-cube $A ^{\prime\prime\prime}$ sur $D\cap E$: \begin{equation} \label{double-restriction} \xymatrix{A\sec & K(E)|_E \ar@{^{(}->}[r]_-j&K(D+E)|_E\ar@{->>}[r] &K(E)|_{D\cap E} &A ^{\prime\prime\prime} \\ A& K(E)\ar@{^{(}->}[r]\ar@{->>}[u] &K(D+E)\ar@{->>}[r]\ar@{->>}[u] &K(D+E)|_D\ar@{->>}[u] &A^{\prime}\\ &K\ar@{^{(}->}[r]_-j\ar@{^{(}->}[u]_-k & \hspace{4mm} K(D) \hspace{4mm} \ar@{^{(}->}[u]\ar@{->>}[r]&K(D)|_D\ar@{^{(}->}[u]_-k \save "2,2"."3,3"*[F-]\frm{}\ar@{.}"2,1"\restore \save "1,2"."1,3"*[F-]\frm{}\ar@{.}"1,1"\restore \save "2,4"."3,4"*[F-]\frm{}\ar@{.}"2,5"\restore \save "1,4".*[F-]\frm{}\ar@{.}"1,5"\restore } \end{equation} {\em Cas 1}:\\ $K$ est ici un 1-cube $(\xymatrix{L\ar@{-}[r]&M})$ et le diagramme (\ref{double-restriction}) se r{\'e}duit alors {\`a}: \[ \xymatrix{ &M(E)|_E\ar[rr] \ar @{} [dr] |{A\sec} && M(D+E)|_E \\ L(E)|_E\ar[rr]\ar@{-}[ur] && L(D+E)|_E\ar@{-}[ur]&\\ &M(E)\ar@{^{(}->}'[r][rr]\ar@{->>}'[u][uu] && M(D+E)\ar@{->>}[uu]\ar@{->>}[rr]&&M(D+E)|_D \\ L(E)\ar@{^{(}->}[rr]\ar@{->>}[uu]\ar@{-}[ur] && L(D+E)\ar@{->>}[uu]\ar@{-}[ur]\ar@{->>}[rr] &&L(D+E)|_D\ar@{-}[ur]\ar @{} [rd] |{A^{\prime}}&\\ &M\ar@{^{(}->}'[r][rr]\ar@{^{(}->}'[u][uu] \ar @{} [ur] |{A} && M(D)\ar@{->>}'[r][rr]\ar@{^{(}->}'[u][uu] &&M(D)|_D\ar[uu] \\ L\ar@{^{(}->}[rr]_-j\ar@{-}[ur]_-i\ar@{^{(}->}[uu]_-k && L(D)\ar@{-}[ur]\ar@{->>}[rr]\ar@{^{(}->}[uu]&&L(D)|_D\ar[uu]\ar@{-}[ur]& } \] o{\`u} $\{i,j,k\} =\{1,2,3\}$. On peut alors {\'e}crire un diagramme d'isomorphismes: \[ \xymatrix{ \gd(\phi^{\prime\ast}_i(A\sec)) \ar[dd]^-{s_{A\sec,i}} &\gd(\phi^{\prime\ast}_i(A)) \ar[l] \ar[r] &\gd(\phi^{\prime\ast}_i(A^{\prime})) \ar[dd]_-{s_{A^{\prime},i}}\\ & {\mathcal{O}_{S}}\ar[ru] \ar[rd]\ar[lu] \ar[ld] \\ \gd(\phi^{\prime \prime\ast}_i(A\sec)) &\gd(\phi^{\prime \prime\ast}_i(A)) \ar[l] \ar[r] &\gd(\phi^{\prime \prime\ast}_i(A^{\prime})) } \] dont les lignes sup{\'e}rieures et inf{\'e}rieures proviennent respectivement des faces avant et arri{\`e}res du diagramme pr{\'e}c{\'e}dent et les fl{\`e}ches obliques sont les isomorphismes d{\'e}crits dans (\ref{cas-des-courbes}). Il s'agit de montrer que le trac{\'e} ext{\'e}rieur est commutatif, ce qui provient du fait que les deux triangles lat{\'e}raux sont commutatifs par la remarque (\ref{carre}) et que les deux autres triangles sont commutatis par le lemme (\ref{indep}).\\ {\em Cas 2}\\ Le diagramme (\ref{double-restriction}) induit un diagramme commutatif d'isomorphismes \[ \begin{CD} \gd (A\sec ) @>{\gd}>>\gd (A ^{\prime\prime\prime} )\\ @A{\gc}AA @A{\gb}AA \\ \gd (A) @>{\ga}>>\gd (A^{\prime} ) \end{CD} \] Par construction, $\ga $ identifie $s_{A^{\prime} ,i}$ {\`a} $s_{A,i,\sigma_j}$ et $\gc $ identifie $s_{A\sec ,i}$ {\`a} $s_{A,i,\sigma_k}$. Par l'hypoth{\`e}se de r{\'e}currence appliqu{\'e}e {\`a} $\gd _D$, $\gd _E$ et $\gd _{D\cap E}$, les isomorphismes $\gb$ et $\gd$ identifient respectivement $s_{A^{\prime} ,i}$ et $s_{A\sec ,i}$ {\`a} $s_{A ^{\prime \prime\prime} ,i}$, ce qui prouve l'assertion dans le cas 2. \end{proof} \subsubsection{Lien entre $s_{A,i,\sigma_j}$ et $s_{A,k,\sigma_j}$. } \label{signe_div} Soient $A$ un $(n+2)$-cube sur $X$ dont la $j$-i{\`e}me ar{\^e}te poss{\`e}de une section $\pi$-r{\'e}guli{\`e}re $\sigma_j$ et soent $i,k$ deux indices distincts et distincts de $j$, montrons que les trivialisations de $\gd (A)$ induites par $s_{A,i,\sigma_j}$ et $s_{A,k,\sigma_j}$ diff{\`e}rent d'un signe {\'e}gal {\`a} $\ge _{ik} (A)$ (introduit dans la proposition (\ref{cubebis})).\\ A cet effet, {\'e}crivons $K$ sous la forme $\left( \begin{array}{c} \xymatrix{ K_2 \ar@{-}[r]\ar@{-}[d]_-k &K_3\ar@{-}[d] \\ K_0 \ar@{-}[r]_-i &K_1 } \end{array} \right) $ et notons $D$ le diviseur de Cartier relatif d{\'e}fini par $\sigma_j$. Consid{\'e}rons le diagramme suivant dont les lignes sont des suites exactes courtes: \[ \xymatrix{ &K_3\ar@{^{(}->}[rr]\ar@{-}'[d][dd]\ar@{-}[dl]&& K_3(D)\ar@{->>}[rr]\ar@{-}'[d][dd]\ar@{-}[dl] &&K_3(D)|_D \ar@{-}[dl]\ar@{-}[dd]\\ K_2\ar@{^{(}->}[rr] \ar@{-}[dd]&& K_2(D)\ar@{->>}[rr]\ar@{-}[dd]&&K_2(D)|_D\ar@{-}[dd]& \\ &K_1\ar@{^{(}->}'[r][rr]\ar@{-}[dl]&&K_1(D)\ar@{-}[dl]\ar@{->>}'[r][rr] &&K_1(D)|_D\ar@{-}[dl] \\ K_0\ar@{^{(}->}[rr]&&K_0(D)\ar@{->>}[rr]&&K_0(D)|_D& } \] le cube de gauche de ce diagramme est pr{\'e}cis{\'e}ment $A$ et les suites exactes donnent un isomorphisme entre $\gd (A) $ et $\gd \left( \begin{array}{c} \xymatrix{ K_2(D)|_D \ar@{-}[r]\ar@{-}[d]_-j &K_3(D)|_D\ar@{-}[d] \\ K_0(D)|_D \ar@{-}[r]_-i &K_1(D)|_D } \end{array} \right) $. Cet isomorphisme identifie les trivialisations de $\gd (A)$ induites par $s_{A,i,\sigma_j}$ et $s_{A,k,\sigma_j}$ aux trivialisations de $\gd (K(D)|_D)$ induites par $s_{K(D)|_D,i}$ et $s_{K(D)|_D,k}$. Or, par l'hypoth{\`e}se de r{\'e}currence, interpr{\'e}t{\'e}e {\`a} la lumi{\`e}re de la proposition (\ref{cubebis}), ces deux trivialisations diff{\`e}rent d'un signe $\ge _{ik}(K(D)|_D)$. On conclut en affirmant que $\ge _{ik}(A) = \ge _{ik}(K(D)|_D)$. En effet, par l'additivit{\'e} de la caract{\'e}ristique d'Euler-Poincar{\'e}, on a: \[ \chi_{D/S} (K_0 (D)|_D) =\chi_{X/S} (K_0(D))-\chi_{X/S} (K_0) =\chi_{X/S}(\xymatrix{K_0 \ar@{-}[r] &K_0 (D)}) \] et donc: \[ \ge _{ik}(A) = (-1)^{\chi_{X/S}(K_0)-\chi_{X/S}(K_0 (D))} =(-1)^{\chi_{D/S}(K_0(D)|_D)} = \ge_{ik} (K(D)|_D) \] \subsection{Elimination des hypoth{\`e}ses de diviseurs effectifs} Pour tout $(n+2)$-cube $A$ suffisamment positif dans les directions $i$ et $j$ et tout indice $i\neq j,k$, on construit dans cette section un isomorphisme \[ s_{A,i}: \gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A), \] fonctoriel en les isomorphismes de $(n+2)$-cubes et compatible aux empilements de cubes dans la direction $i$ (conditions 1 et 2 de la proposition (\ref{cubebis})). \subsubsection{Premi{\`e}re r{\'e}duction} \label{reduction} Soit $A$ un $(n+2)$-cube suffisamment positif dans les directions $i$ et $j$. $A$ est alors isomorphe un $(n+2)$-cube $ \left( \begin{array}{c} \xymatrix{ K\otimes M\ar@{-}[r]\ar@{-}[d]_-k &K\otimes (L\otimes M)\ar@{-}[d]\\ K \ar@{-}[r]_-j &K\otimes L } \end{array} \right) $ avec $L,M \gg 0$. Consid{\'e}rons les fibr{\'e}s projectifs $P_L$ et $P_M$ sur $S$ et effectuons le changement de base \[ \begin{CD} X_{P_L \times P_M} @>g>> X\\ @V{\pi}VV @V{\pi}VV \\ P_L \times P_M @>f>> S \end{CD} \] Notons encore $K$, $L$ et $M$ les images r{\'e}ciproques de $K$, $L$ et $M$ par $g$ et introduisons les faisceaux inversibles $L^{\prime} = L\otimes \pi_{P_L}^{\ast } \mathcal{O}_{P_L}(1)$ et $M^{\prime} = M\otimes \pi_{P_M}^{\ast } \mathcal{O}_{P_M}(1)$ et consid{\'e}rons le $(n+2)$-cube $A^{\prime} = \left( \begin{array}{c} \xymatrix{ K\otimes M^{\prime}\ar@{-}[r]\ar@{-}[d]_-k &K\otimes (L^{\prime}\otimes M^{\prime})\ar@{-}[d]\\ K \ar@{-}[r]_-j &K\otimes L^{\prime} } \end{array} \right) $ sur $X_{P_L \times P_M}$. On a des isomorphismes canoniques: $ \gd(K\otimes L^{\prime}) \stackrel{\sim }{\F } f^{\ast }\gd(K\otimes L) \otimes (\mathcal{O}_{P_L}(1))^{\chi_{X/S}(K\otimes L)} $, $ \gd(K\otimes M^{\prime}) \stackrel{\sim }{\F } f^{\ast }\gd(K\otimes M) \otimes (\mathcal{O}_{P_M}(1))^{\chi_{X/S}(K\otimes M)} $ et $ \gd(K\otimes L^{\prime}\otimes M^{\prime}) \stackrel{\sim }{\F } f^{\ast }\gd(K\otimes L\otimes M) \otimes (\mathcal{O}_{P_L}(1)\otimes\mathcal{O}_{P_M}(1))^{\chi_{X/S}(K\otimes L\otimes M)} $. L'{\'e}galit{\'e} $\chi_{X/S}(K\otimes L\otimes M)= \chi_{X/S}(K\otimes L)=\chi_{X/S}(K\otimes M)$, d{\'e}duite de (\ref{caracteristique2}), entraine alors l'existence d'un isomorphisme canonique \[ \gd(A^{\prime}) \stackrel{\sim }{\F } f^{\ast } \gd(A) \] Comme on a $f_{\ast } f^{\ast } \mathcal{O} _{P_L \times _S P_M} = \mathcal{O} _S$, on en d{\'e}duit que la donn{\'e}e d'un isomorphisme $s_{A,i}: \gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A)$ sur $S$ est {\'e}quivalente {\`a} la donn{\'e}e d'un isomorphisme $s_{A^{\prime},i}: \gd (\phi^{\prime\ast}_i A^{\prime}) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A^{\prime})$ sur $P_L\times _S P_M$. \subsubsection{Construction de $s_{A^{\prime},i}$ sur des ouverts} Consid{\'e}rons les ouverts $U_L = P_L \setminus Z_L$ et $U_M = P_M \setminus Z_M$ introduits en (\ref{div2}). Sur $X_{U_L \times P_M}$, on dispose d'un diviseur relatif $D_L$ et d'un isomorphisme canonique $L^{\prime} \stackrel{\sim }{\F } \mathcal{O} (D_L)$. Le cube $A^{\prime}$ est donc canoniquement isomorphe sur $X_{U_L \times P_M}$ {\`a} un cube de la forme $(\xymatrix{ K_0 \ar@{-}[r] & K_0 (D_L)})$. En appliquant la construction (\ref{constr1}) {\`a} cette situation, on obtient un isomorphisme canonique $s^{\prime} :\gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A)$ d{\'e}fini sur $U_L\times _S P_M$. La m{\^e}me construction, effectu{\'e}e en intervertissant les r{\^o}les de $L$ et $M$, nous donne un autre isomorphisme $s\sec$ entre les m{\^e}mes faisceaux et d{\'e}fini sur $P_L\times _S U_M$. \addtocounter{theo}{2} \addtocounter{subsubsection}{1} \begin{lemme} \label{coincid} $s^{\prime}$ et $s\sec$ coincident sur $U_L \times _S U_M$. \end{lemme} \addtocounter{subsubsection}{1} \begin{proof}[Preuve] Soit $p$ la projection $U_L \times _S U_M \longrightarrow U_M$. Le lemme (\ref{genericite2}) montre l'existence d'un ferm{\'e} $Z$ de $U_L \times _S U_M$ tel que \begin{enumerate} \item $\forall u \in U_M, \forall x \in Z_u, \text{Prof}(\mathcal{O}_{p ^{-1}(u),x}) \geq 1$. \item Au dessus de $V=(U_L \times _S U_M) \setminus Z$, les diviseurs $D_L$ et $D_M$ sont en position d'intersection compl{\`e}te et $(D_L \cap D_M)_V$ est plat sur $V$. \end{enumerate} Au dessus de $V$, $s^{\prime}$ et $s\sec$ coincident d'apr{\`e}s (\ref{indep}). On en d{\'e}duit donc, d'apr{\`e}s (\cite{EGA4},19.9.8), en utilisant la propri{\'e}t{\'e} 1 que $s^{\prime}$ et $s\sec$ coincident sur $U_L \times _S U_M$. \end{proof} \subsubsection{Extension de $s^{\prime}$ et $s\sec$ {\`a} $P_L \times _S P_M$.} Consid{\'e}rons le ferm{\'e} $Y=Z_L \times _S Z_M$ du $s$-sch{\'e}ma $P_L\times _S P_M$. Le lemme (\ref{genericite1}) entra{\^\i}ne que pour tout $s\in S$ et tout $y\in Y_s$, on a $\text{Prof}(\mathcal{O}_{(P_L \times _S P_M)_s,y}) \geq 2$. les isomorphismes $s^{\prime}$ et $s\sec$ sont d{\'e}finis respectivement sur $U_L \times _S P_M$ et $P_L \times _S U_M$ et coincident sur $U_L \times _S U_M$. Ils d{\'e}finissent donc un isomorphisme $s: \gd ((\phi_i^{\prime})^{\ast } A) \stackrel{\sim }{\F } \gd ((\phi_i\sec ) ^{\ast } A)$ sur l'ouvert $U= (P_L \times _S P_M)\setminus Y$. Par (\cite{EGA4},19.9.8), $s$ se prolonge alors (de mani{\`e}re unique) {\`a} $P_L \times _S P_M$ en un isomorphisme $s_{A^{\prime},i}^{j,k} : \gd (\phi^{\prime\ast}_i A^{\prime}) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A^{\prime})$ d{\'e}pendant a priori du choix des directions $j$ et $k$.\\ Par la r{\'e}duction (\ref{reduction}), on d{\'e}duit de cette construction un isomorphisme \[ s_{A^{\prime},i}^{j,k} : \gd (\phi^{\prime\ast}_i A^{\prime}) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A^{\prime}) \] d{\'e}fini maintenant sur $S$. Il reste {\`a} montrer le \addtocounter{theo}{1} \begin{lemme} $s_{A,i}^{j,k}$ est ind{\'e}pendant du choix des directions $j$ et $k$. \end{lemme} \begin{proof}[Preuve] Supposons en effet que $A$ soit suffisamment positif dans trois directions diff{\'e}rentes $j,k,l \neq i$. $A$ s'{\'e}crit donc de trois fa\c{c}ons diff{\'e}rentes $(\xymatrix{K_j \ar@{-}[r]_-j & K_j \otimes L})$, $(\xymatrix{K_k \ar@{-}[r]_-k & K_k \otimes M})$ et $(\xymatrix{K_l \ar@{-}[r]_-l & K_l \otimes N})$, avec $L,M,N \gg 0$. Montrons alors $s_{A,i}^{j,k}=s_{A,i}^{k,l}$.\\ Il suffit de le montrer apr{\`e}s changement de base par $P_M \longrightarrow S$, et donc apr{\`e}s changement de base par $U_M \longrightarrow S$ (par le m{\^e}me argument que dans la preuve du lemme (\ref{coincid}) ). Mais, par construction, au dessus de $U_M$ les isomorphismes $s_{A,i}^{j,k}$ et $s_{A,i}^{k,l}$ proviennent tous deux de la structure du cube associ{\'e}e au sch{\'e}ma relatif $D_M/S$, et sont donc {\'e}gaux. \end{proof} \addtocounter{subsubsection}{1} \subsubsection{Propri{\'e}t{\'e}s des isomorphismes ainsi construits} On a donc construit, pour tout $(n+2)$-cube $A$, qui est suffisamment positif dans deux directions et pour toute direction $i$ diff{\'e}rente des pr{\'e}c{\'e}dentes, un isomorphisme \[ s_{A,i} : \gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A) \] les $s_{A,i}$ v{\'e}rifient les propri{\'e}t{\'e}s suivantes: \begin{enumerate} \item Si les deux $(n+2)$-cubes $A$ et $B$ sont suffisamment positifs dans deux directions diff{\'e}rentes de $i$ et ont leur i-{\`e}me face en commun, alors: \[ s_{B,i} \circ s_{A,i} = s_{A\ast _i B,i} \] \item Si $s_{A,i}$ et $s_{A,j}$ sont tous les deux d{\'e}finis, alors les trivialisations de $\gd (A)$ qu'elles induisent diff{\`e}rent d'un signe {\'e}gal {\`a} $\ge _{ij} (A)$. \item Si $A$ est suffisamment positif dans deux directions diff{\'e}rentes de $i$, pour toute permutation $\sigma \in S_n$ on a: \[ s_{\sigma ^{\ast } A,i} = s_{A, \sigma (i)} \] \end{enumerate} Pour montrer les propri{\'e}t{\'e}s 1 et 2, on choisit une direction $k$ dans laquelle $A$ est suffisamment positif, et on {\'e}crit $A=(\xymatrix{K \ar@{-}[r]_-k &K \otimes L})$ avec $L\gg 0$ et on effectue, comme pr{\'e}c{\'e}demment, un changement de base $U_L \subset P_L \longrightarrow S$, ce qui permet de disposer d'un diviseur effectif $D_L$. La propri{\'e}t{\'e} 1 provient alors de la propri{\'e}t{\'e} de recollement (\ref{recol_div}) et la propri{\'e}t{\'e} 2 provient du lien entre $s_{A,i,D}$ et $s_{A,k,D}$ (\ref{signe_div}), montr{\'e}s tous les deux sous l'hypoth{\`e}se de l'existence d'un diviseur effectif.\\ Pour la propri{\'e}t{\'e} 3, il suffit de montrer l'{\'e}galit{\'e} $s_{\sigma ^{\ast } A,i} = s_{A, \sigma (i)}$ dans le cas o{\`u} $\sigma$ est une transposition $\sigma _{rs}$ et o{\`u} $A$ est suffisamment positif dans au moins une direction $k$ diff{\'e}rente de $r$ et de $s$. On {\'e}crit alors $A=(\xymatrix{K \ar@{-}[r]_-k &K \otimes L})$ avec $L\gg 0$ et on effectue le changement de base $U_L \subset P_L \longrightarrow S$. L'{\'e}galit{\'e} provient alors de l'{\'e}galit{\'e} analogue pour le $(n+1)$-cube $K\otimes L |_{D_L}$, qui est v{\'e}rifi{\'e}e d'apr{\`e}s l'hypoth{\`e}se de r{\'e}currence. \subsection{Elimination des hypoth{\`e}ses de positivit{\'e}.} On veut {\'e}tendre ici la d{\'e}finition de $s_{A,i}$ {\`a} un $(n+2)$-cube quelconque $A$ dans $PIC(X)$. \begin{rem} \label{compar-sign} Si $A$ et $B$ sont deux $(n+2)$-cubes dans $PIC(X)$, recollables le long de leur j-i{\`e}me face (notons $C= A\ast _j B$) et tels que $s_{A,i}$ et $s_{B,i}$ sont bien d{\'e}finis. On peut alors d{\'e}finir un isomorphisme $s_{A,i}\otimes s_{B,i}:\gd (\phi^{\prime\ast}_i C) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i C)$ par: \[ \gd (\phi^{\prime\ast}_i C) \stackrel{\sim }{\F } \gd (\phi^{\prime\ast}_i A) \otimes \gd (\phi^{\prime\ast}_i B) \longrightarrow \gd (\phi^{\prime \prime\ast}_i A) \otimes \gd (\phi^{\prime \prime\ast}_i B) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i C) \] Notons que si $s_{C,i}$ est aussi d{\'e}fini, on a la relation: \[ s_{C,i} =\ge _{ij}(A) s_{A,i}\otimes s_{B,i}\; , \] r{\'e}sultant de la comparaison entre $s_{A,i}$ et $s_{A,j}$ et de la relation $s_{C,j} =s_{A,j}\otimes s_{B,j}$. \end{rem} On peut alors {\'e}noncer le \begin{lemme} Soit $i$ un indice compris entre 1 et $n+2$ et soit $A$ un $(n+2)$-cube suffisamment positif dans {\em une} direction $j\neq i$. \begin{enumerate} \item Pour toute direction $k\neq i,j$, il existe un $(n+2)$-cube $B$ recollable avec $A$ dans la direction $k$ tel que $B$ et $A\ast _k B$ soient tous les deux suffisamment positifs dans la direction $k$. \item Soient $k$ et $l$ deux directions eventuellement {\'e}gales mais distinctes de $i$ et $j$ et soient $B$ (resp. $B^{\prime}$) deux $(n+2)$-cubes recollables avec $A$ dans les directions $k$ (resp. $l$) tels que $B$ et $A\ast _k B$ sont suffisament positifs dans la direction $k$ et $B^{\prime}$ et $A\ast _k B^{\prime}$ sont suffisamment positifs dans la direction $l$. Alors les deux isomorphismes: \[ s_{A\ast_k B ,i} \otimes s_{B ,i}^{-1} \; \text{et} \; s_{A\ast_l B^{\prime} ,i}\otimes s_{B^{\prime} ,i}^{-1}: \gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A) \] sont {\'e}gaux. \end{enumerate} \end{lemme} \begin{proof}[Preuve] Le premier point r{\'e}sulte du fait que pour tout faisceau inversible $L$ sur $X$, on peut trouver un faisceau inversible $M\gg 0$ tel que $L\otimes M \gg 0$.\\ Montrons d'abord le second point quand $k=l$. Dans ce cas, on peut trouver des $(n+2)$-cubes $C$ et $C^{\prime}$, suffisamment positifs dans la direction $k$ , tels que $A\ast _k B \ast _k C = A\ast _k B^{\prime} \ast _k C^{\prime}$ et $ B \ast _k C$, $B^{\prime} \ast _k C^{\prime}$ et $A\ast _k B \ast _k C$ sont suffisamment positifs dans la direction $k$. On a alors en utilisant la remarque (\ref{compar-sign}: \[ s_{A\ast _kB,i} \otimes s_{B,i}^{-1} = s_{A\ast _kB \ast _k C,i} \otimes s_{B \ast _k C,i}^{-1} = s_{A\ast _kB^{\prime} \ast _k C^{\prime} ,i} \otimes s_{B^{\prime} \ast _k C^{\prime} ,i}^{-1} = s_{A\ast _kB^{\prime} ,i} \otimes s_{B^{\prime} ,i}^{-1} \] Dans le cas o{\`u} $k\neq l$, on introduit un $(n+2)$-cube $C$ d{\'e}termin{\'e} uniquement par les conditions suivantes: $C$ est recollable avec $B$ dans la direction $l$ et avec $B^{\prime}$ dans la direction $k$, comme le d{\'e}crit le diagramme suivant: \begin{center} \begin{tabular}{cc} $l \uparrow$ &\begin{tabular}{|c|c|} \hline $B^{\prime}$&$C$\\ \hline $A$&$B$ \\ \hline \end{tabular} \\ &$\xrightarrow{k}$ \end{tabular} \end{center} En notant $D$ le cube total, on {\'e}crit alors: \[ s_{A\ast _kB,i} \otimes s_{B,i}^{-1} = s_{C,i} \otimes s_{B^{\prime} \ast _kC,i}^{-1} \otimes s_{B,i}^{-1}= s_{C,i} \otimes s_{B \ast _lC,i}^{-1} \otimes s_{B^{\prime} ,i}^{-1}= s_{A\ast _lB^{\prime} ,i} \otimes s_{B^{\prime} ,i}^{-1} \] \end{proof} Ce lemme permet donc de d{\'e}finir $s_{A,i}:\gd (\phi^{\prime\ast}_i A) \stackrel{\sim }{\F } \gd (\phi^{\prime \prime\ast}_i A)$ d{\'e}s que $A$ est suffisamment positif dans {\em une} direction $j\neq i$ par la formule: \[ s_{A,i}=s_{B ,i}^{-1} \circ s_{A\ast_k B ,i} \] En r{\'e}it{\'e}rant cette argument, on construit $s_{A,i}$ pour tout $(n+2)$-cube $A$ dans $PIC(X)$, sans hypoth{\`e}se de positivit{\'e} sur $A$. \begin{lemme} Les isomorphismes $s_{A,i}$ ainsi construits v{\'e}rifient les propri{\'e}t{\'e}s de la proposition (\ref{cubebis}). \end{lemme} Ce qui conclut la preuve du th{\'e}or{\`e}me principal. \end{proof} \section{Fibr{\'e} d'intersection et r{\'e}sultant} \subsection{Fibr{\'e} d'intersection} Pour tout morphisme projectif et plat $\pi:X \longrightarrow S$, {\`a} fibres de dimension $n$, sur un sch{\'e}ma $S$ localement noeth{\'e}rien, d{\'e}finisssons le $(n+1)$-foncteur: \[ \begin{array}{rlcl} I_{X/S}\; :& PIC^{n+1}(X) &\longrightarrow &PIC(S)\\ & (L_1, \cdots , L_{n+1}) &\mapsto & {\displaystyle \bigotimes_{k=0}^{n+1} \left( \bigotimes_{i_1<\cdots <i_k} \gd (L_{i_1} \otimes \cdots \otimes L_{i_k}) \right) ^{(-1)^{n+1-k}}} \end{array}\; , \] appel{\'e} {\em foncteur fibr{\'e} d'intersection} pour $X/S$. \begin{prop} \label{propr-inter} Le foncteur $I_{X/S}$ est un $(n+1)$-foncteur additif en chaque variable, v{\'e}rifiant les propri{\'e}t{\'e}s suivantes: \begin{enumerate} \item La formation de $I_{X/S}$ commute aux changements de bases. \item $I_{X/S}$ est muni de donn{\'e}es de sym{\'e}trie, compatibles avec les donn{\'e}es d'additivit{\'e} en chaque variable. \item Si $\pi : X\longrightarrow S$ est fini et plat, le foncteur $I_{X/S}:PIC(X)\longrightarrow PIC(S) $ est simplement le foncteur norme $N_{X/S}$ et les contraintes d'additivit{\'e} pour $I_{X/S}$ sont les isomorphismes usuels $N_{X/S} (L\otimes L^{\prime} ) \stackrel{\sim }{\F } N_{X/S} (L) \otimes N_{X/S} (L^{\prime} )$. \item Soient $L_1,\cdots ,L_{n+1}$ des faisceaux inversibles sur $X$ et $\sigma_{n+1}$ une section $\pi$-r{\'e}guli{\`e}re de $L_{n+1}$ d{\'e}finissant un diviseur de Cartier relatif effectif $D$. Il existe un isomorphisme canonique: \[ \rho _D: I_{X/S} (L_1, \cdots ,L_{n+1}) \stackrel{\sim }{\F } I_{D/S} (L_1|_D, \cdots ,L_n|_D) \] qui est fonctoriel en les isomorphismes $L_i \stackrel{\sim }{\F } L_i^{\prime}$ pour $1\leq i\leq n$ et compatible avec les donn{\'e}es d'additivit{\'e} et de sym{\'e}trie de $I_{X/S}$ en les $L_1,\cdots ,L_n$ et de $I_{D/S}$ en les $L_1|_D,\cdots ,L_n|_D$. \item Si, en plus des donn{\'e}es de 4, on dispose d'une section $\sigma_n$ de $L_n$ telle que $(\sigma_n,\sigma_{n+1})$ et $(\sigma_{n+1},\sigma_n)$ sont des suites $\pi$-r{\'e}guli{\`e}res, notons $E$ le diviseur d{\'e}fini par $\sigma_n$. Le diagramme suivant est alors commutatif: \[ \begin{CD} I_{X/S} (L_1, \cdots ,L_{n+1}) @>{\rho_D}>> I_{D/S} (L_1|_D, \cdots ,L_n|_D)\\ @V{\rho_E}VV @VV{\rho_E}V\\ I_{E/S} (L_1|_E, \cdots ,L_{n-1}|_E,L_{n+1}|_E) @>{\rho_D}>> I_{D\cap E/S} (L_1|_{D\cap E}, \cdots ,L_{n-1}|_{D\cap E}) \end{CD} \] \end{enumerate} \end{prop} \begin{proof}[Preuve] Notons que, par d{\'e}finition, \[ I_{X/S}(L_1,\cdots ,L_n) = \gd _{X/S}( K_{\mathcal{O}_{X}}(L_1,\cdots ,L_n)) \; . \] C'est donc le $(n+1)$-foncteur multi-additif et sym{\'e}trique associ{\'e} {\`a} la structure du cube sur le foncteur $\gd :PIC(X) \longrightarrow PIC (S)$. La propri{\'e}t{\'e} 1 provient de la compatibilit{\'e} du foncteur $\gd$ aux changements de bases. L'existence de donn{\'e}es d'additivit{\'e} en chaque variable et de donn{\'e}es de sym{\'e}trie compatibles {\`a} l'additivit{\'e} provient de l'existence d'une structure de $(n+2)$-cube sur $\gd$. L'identification 3 de $I_{X/S}$ et $N_{X/S}$ dans le cas d'un sch{\'e}ma relatif fini et plat est simplement la d{\'e}finition de la structure du carr{\'e} dans le cas de dimension relative 0.\\ Il reste {\`a} construire l'isomorphisme $\rho _D$: Les isomorphismes de restrictions d{\'e}crits en (\ref{th-prin}) induisent un isomorphisme canonique: \[ \gd _{X/S} (K_{\mathcal{O}_{X}} (\mathcal{O}_{X} (D),L_1,\cdots ,L_n)) \stackrel{\sim }{\F } \gd _{D/S} (K_{\mathcal{O}_{X} (D)|_D}(L_1 (D)|_D, \cdots ,L_n (D)|_D)) \] Par ailleurs, l'{\'e}galit{\'e} \[ K_{\mathcal{O}_{X} (D)|_D}(L_1(D)|_D,\cdots ,L_n(D)|_D) = K_{\mathcal{O} _D}(L_1|_D,\cdots ,L_n|_D) \otimes_{\mathcal{O}_D} \mathcal{O}_{X} (D)|_D \] induit, d'apr{\`e}s (\ref{def-multifonct}), un isomorphisme \[ \gd _{D/S} (K_{\mathcal{O}_{X} (D)|_D}(L_1 (D)|_D, \cdots ,L_n (D)|_D)) \stackrel{\sim }{\F } I_{D/S} (L_1 |_D, \cdots ,L_n |_D). \] D'apr{\`e}s les propri{\'e}t{\'e}s des structures du cube pour $\gd _{D/S}$ et $\gd _{X/S}$, ces deux isomorphismes sont fonctoriels et additifs en chaque $L_i$ et compatibles aux donn{\'e}es de sym{\'e}tries. Notons $\rho _D$ leur compos{\'e}; il v{\'e}rifie donc les propri{\'e}t{\'e}s demand{\'e}es. \end{proof} \begin{cor} Soient $L_1,\cdots ,L_n$ des faisceaux inversibles sur $X$ et $(\sigma_1,\cdots ,\sigma_n)$ une suite $\pi$-r{\'e}guli{\`e}re de sections des $L_i$. Notons $D_i$ le lieu des z{\'e}ros de $\sigma_i$. Pour tout faisceau inversible $L$ sur $X$, les isomorphismes de restriction successifs de $\cap_{i=1}^p D_i$ {\`a} $\cap_{i=1}^{p+1} D_i$ induisent un isomorphisme canonique: \[ I_{X/S} (L_1, \cdots ,L_n,L) \stackrel{\sim }{\F } N_{D_1\cap \cdots \cap D_n/S} (L|_{D_1\cap \cdots \cap D_n})\; , \] fonctoriel en les isomorphismes $L\stackrel{\sim }{\F } L^{\prime}$ et additif en $L$. \end{cor} \begin{proof}[Preuve] Remarquons que, par hypoth{\`e}se, pour tout entier $p$ tel que $1\leq p \leq n$, $D_1 \cap \cdots \cap D_p \longrightarrow S$ est un morphisme projectif et plat {\`a} fibres de dimension $n-p$ et $D_1 \cap \cdots \cap D_p\cap D_{p+1}$ est un diviseur de Cartier relatif effectif de $D_1 \cap \cdots \cap D_p /S$. On peut donc it{\'e}rer la construction pr{\'e}c{\'e}dente, et on obtient un isomorphisme \[ I_{X/S} (\mathcal{O}_{X} (D_1), \cdots , \mathcal{O}_{X} (D_n),L) \stackrel{\sim }{\F } I_{D_1\cap \cdots \cap D_n/S} (L|_{D_1\cap \cdots \cap D_n})\; . \] On conclut alors en utilisant le fait que $D_1\cap \cdots \cap D_n \longrightarrow S$ est fini et plat, ce qui entra{\^\i}ne que $I_{D_1\cap \cdots \cap D_n/S} (L|_{D_1\cap \cdots \cap D_n}) \simeq N_{D_1\cap \cdots \cap D_n/S} (L|_{D_1\cap \cdots \cap D_n})$. La fonctorialit{\'e} et l'additivit{\'e} de l'isomorphisme obtenu se v{\'e}rifient {\`a} chaque {\'e}tape de l'it{\'e}ration. \end{proof} \subsection{Sections du fibr{\'e} d'intersection} \subsubsection{Cas de la dimension 0} Soit $\pi :Y\longrightarrow T$ est un morphisme fini et plat, et soit $L$ un faisceau inversible sur $Y$, muni d'une section $\sigma$ qui ne s'annulle pas au dessus d'un ouvert $U=T\setminus Z$ de $T$. La section $\sigma$ induit un isomorphisme $\mathcal{O} |_{Y_U} \stackrel{\sim }{\F } L|_{Y_U}$, qui induit donc par fonctorialit{\'e} un isomorphisme $N_{Y_U/U} (\sigma ):\mathcal{O} _U \stackrel{\sim }{\F } (N_{Y/T} L)|_U$. Celui ci se prolonge en une section $s:\mathcal{O} _T \stackrel{\sim }{\F } N_{Y/T} L$, non nulle en dehors de $Z$ et d{\'e}finie par le morphisme \[ \det \pi _{\ast } s: \det \pi _{\ast } \mathcal{O} _Y \longrightarrow \det \pi _{\ast } L \; . \] \subsubsection{Construction d'une section sur un ouvert} Soit $\pi : X\longrightarrow S$ un morphisme projectif et plat, {\`a} fibres de dimension $n$ et $L_1,\cdots ,L_{n+1}$ des faisceaux inversibles sur $X$. Supposons que l'on dispose de sections $\sigma_i$ de chaque $L_i$ telles que $(\sigma_1,\cdots ,\sigma_n)$ est une suite $\pi$-r{\'e}guli{\`e}re. Notons $D_i$ le lieu des z{\'e}ros de $\sigma_i$ et $U$ l'ouvert de $S$ au dessus duquel $D_1\cap \cdots \cap D_{n+1} = \emptyset$ , on peut appliquer ce qui pr{\'e}c{\`e}de au morphisme fini et plat $D_1\cap \cdots \cap D_n \longrightarrow S$, au faisceau inversible $L_{n+1}|_{D_1\cap \cdots \cap D_n}$ et {\`a} sa section $\sigma_{n+1}|_{D_1\cap \cdots \cap D_n}$. On obtient ainsi une section $\mathcal{O}_{S} \longrightarrow I_{X/S}(L_1,\cdots ,L_{n+1})$, dont la restriction {\`a} $U$ est un isomorphisme. \subsection{Construction du r{\'e}sultant.} Soit $\pi : X\longrightarrow S$ un morphisme projectif et plat, {\`a} fibres de dimension $n$ et $L_1,\cdots ,L_{n+1}$ des faisceaux inversibles sur $X$. Effectuons le changement de base: \[ \xymatrix{X_P \ar[r]^-g\ar[d]_-{\pi} &X \ar[d]^-{\pi}\\ P= P_{L_1}\times _S \cdots \times _S P_{L_n} \ar[r]^-f & S } \] Pour tout entier $i$, notons $p_i$ la projection de $P$ sur $P_i$ et $\pi_i$ la compos{\'e}e $X\longrightarrow P\longrightarrow P_i$, et consid{\'e}rons le faisceau inversible $L^{\prime}_i=g^{\ast } L_i \otimes \pi_i^{\ast } \mathcal{O}_{P_i}(1)$. Consid{\'e}rons alors le faisceau inversible sur $P$: \[ Res (L_1,\cdots ,L_{n+1}) = I_{X_P/P}(L_1^{\prime},\cdots ,L_{n+1}^{\prime}) \] \begin{lemme} \label{isom-resultant} $Res (L_1,\cdots ,L_{n+1})$ est canoniquement isomorphe {\`a} $ f^{\ast } I_{X/S}(L_1,\cdots ,L_{n+1}) \otimes {\displaystyle \bigotimes_{i=1}^{n+1}p_i^{\ast } \mathcal{O}_{P_i}(k_i)} $, o{\`u} $k_i:S\longrightarrow\mathbb{Z}$ est la fonction localement constante qui {\`a} $s\in S$ associe le nombre d'intersection sur $X_s$ des restrictions {\`a} $X_s$ des faisceaux $L_j$ pour $j\neq i$. \end{lemme} \begin{proof}[Preuve] Rappelons que, si $K$ est un $p$-cube dans $PIC(X)$, pour tout faisceau inversible $L$ sur $S$, on a un isomorphisme canonique $\gd(K\otimes\pi^{\ast } L) \stackrel{\sim }{\F } \gd(K)\otimes L^{\otimes\chi_{X/S}(K)}$. Soit $1\leq i \leq n+1$, notons $i_1,\cdots,i_n$ les $n$ entiers compris entre 1 et $n+1$ et distincts de $i$, class{\'e}s par ordre croissant. Si $M_{i_1},\cdots,M_{i_n}$ sont des faisceaux inversibles sur $P$, on a alors: \begin{multline*} \chi_{X_P/P}(K_{\mathcal{O}_{X}}(g^{\ast } L_{i_1}\otimes \pi^{\ast } M_{i_1} , \cdots ,g^{\ast } L_{i_n}\otimes \pi^{\ast } M_{i_n})=\\ \chi_{X_P/P}(K_{\mathcal{O}_{X}}(g^{\ast } L_{i_1},\cdots ,g^{\ast } L_{i_n})= \chi_{X/S}(K_{\mathcal{O}_{X}}(L_{i_1},\cdots ,L_{i_n}))=k_i \end{multline*} On peut alors {\'e}crire: \begin{multline*} Res (L_1,\cdots ,L_{n+1}) = I_{X_P/P}(L^{\prime}_1,\cdots ,L^{\prime}_n, g^{\ast } L_{n+1}\otimes \pi_i^{\ast } \mathcal{O}_{P_{n+1}}(1))\\ \simeq \gd(K)^{-1}\otimes \gd(K\otimes g^{\ast } L_{n+1}\otimes \pi_i^{\ast } \mathcal{O}_{P_{n+1}}(1)) \simeq \gd(K)^{-1}\otimes \gd(K\otimes g^{\ast } L_{n+1}) \otimes( p_i^{\ast } \mathcal{O}_{P_{n+1}}(1))^{k_{n+1}}\\ \simeq\gd( \xymatrix{K\ar@{-}[r]_-{n+1}&K\otimes g^{\ast } L_{n+1}})\otimes p_i^{\ast } \mathcal{O}_{P_{n+1}}(k_{n+1}), \end{multline*} o{\`u} l'on a pos{\'e} $K=K_{\mathcal{O}_{X}}(L_1^{\prime},\cdots,L_n^{\prime})$. En effectuant la m{\^e}me op{\'e}ration sur chacun des $L_i$, on obtient finalement: \begin{align*} Res (L_1,\cdots ,L_{n+1})& \simeq I_{X_P/P}(g^{\ast } L_1,\cdots,g^{\ast } L_{n+1}) \otimes \bigotimes_{i=1}^{n+1} p_i^{\ast } \mathcal{O}_{P_i}(k_i)\\ &\simeq f^{\ast } I_{X/S}(L_1,\cdots,L_{n+1}) \otimes \bigotimes_{i=1}^{n+1} p_i^{\ast } \mathcal{O}_{P_i}(k_i) \end{align*} \end{proof} Pour tout $i$, consid{\'e}rons la section canonique $\sigma_i$ des $L_i$ sur $X_P$ et notons $D_i$ le lieu des z{\'e}ros de $\sigma_i$. On va construire une section canonique $Res(\sigma_1,\cdots ,\sigma_{n+1})$ de $Res(L_1^{\prime}, \cdots , L_{n+1}^{\prime})$, qu'on appellera {\em r{\'e}sultant} des $\sigma_i$ .\\ Pour toute permutation $\phi\in S_{n+1}$, notons $U_{\phi}$ l'ouvert de $P$ au dessus duquel $(\sigma _{\phi (1)},\cdots ,\sigma _{\phi (n)})$ est une suite $\pi$-r{\'e}guli{\`e}re. \begin{lemme} \label{genericite-suite} Soient $U=\bigcup_{\phi\in S_{n+1}}U_{\phi}$ et $V=\bigcap_{\phi\in S_{n+1}}U_{\phi}$ et soient $Z$ et $Z^{\prime}$ les ferm{\'e}s compl{\'e}mentaires. Si $f:P\longrightarrow S$ d{\'e}signe le morphisme structural, on a: \begin{enumerate} \item Pour tout $z\in Z$, on a: $\text{Prof} (\mathcal{O}_{f^{-1}(f(z)),z}) \geq 2$. \item Pour tout $z\in Z^{\prime}$, on a: $\text{Prof}(\mathcal{O}_{f^{-1}(f(z)),z})\geq 1$. \end{enumerate} \end{lemme} \begin{proof}[Preuve] Soit $z\in Z$, soit $k<n$ la longueur maximale d'une suite $\pi$-r{\'e}guli{\`e}re prise parmi les $\sigma_i$ au dessus de $z$, soit $(\sigma_{\phi (1)},\cdots ,\sigma_{\phi (k)})$ une telle suite et soit $P^{\prime}= P_{\phi (1)}\times _S\cdots\times _SP_{\phi (k)}$. D{\'e}composons $f:P\longrightarrow S$ en \[ \xymatrix{ &P^{\prime}\times _SP_{\phi (k+1)}\ar[rd]^-{q^{\prime}} \\ P\ar[rr]^-p \ar[ru]^-{p^{\prime}} \ar[rd]^-{p\sec} &&P^{\prime} \ar[r]^-r&S\\ &P^{\prime}\times _SP_{\phi (k+2)}\ar[ru]^-{q\sec} } \] Par le lemme (\ref{genericite2}), $p^{\prime} (z)$ est contenu dans un hyperplan de la fibre $(q^{\prime} )^{-1} (p(z))$ et, de m{\^e}me, $p\sec (z)$ est contenu dans un hyperplan de la fibre $(q\sec )^{-1} (p(z))$. Donc $z$ est contenu dans l'intersection de deux hyperplans de la fibre $p ^{-1}(p(z))$ et donc dans l'intersection de deux hyperplans de $f ^{-1}(f(z))$, ce qui montre le r{\'e}sultat.\\ La deuxi{\`e}me assertion se montre de mani{\`e}re analogue. \end{proof} Pour tout $\phi\in S_{n+1}$, on a un isomorphisme d{\'e}fini sur l'ouvert $U_{\phi}$: \begin{equation} \label{isom-recur} I_{X/S} (L_1,\cdots ,L_{n+1}) \stackrel{\sim }{\F } N_{D_{\phi(1)}\cap\cdots\cap D_{\phi(n)}/S} (L_{\phi(n+1)}) \end{equation} et la section $N_{D_{\phi(1)}\cap\cdots\cap D_{\phi(n)}/S} (\sigma_{\phi(n+1)})$ d{\'e}finit donc une section $s_\phi$ de $I_{X/S}|_{U_\phi}$. \begin{lemme} Pour tous $\phi,\psi\in S_n$, on a $s_\phi|_V=s_\psi|_V$. \end{lemme} \begin{proof}[Preuve] Si $\phi(n+1)=\psi(n+1)$, ceci provient de l'assertion 5 de la proposition (\ref{propr-inter}). Il suffit donc de montrer l'assertion dans le cas o{\`u} $\phi(n)=\psi(n+1)$ et $\phi(n+1)=\psi(n)$. Notons alors $Y=D_{\phi(1)}\cap\cdots\cap D_{\phi(n-1)}$. C'est un sch{\'e}ma relatif sur $S$ {\`a} fibres de dimension 1 et les isomorphismes (\ref{isom-recur}) correspondant {\`a} $\phi$ et $\psi$ se factorisent en \[ I_{X/S} (L_1,\cdots ,L_{n+1}) \stackrel{\sim }{\F } I_{Y/S}(L_{\phi(n)},L_{\phi(n+1)}) \stackrel{\sim }{\F } N_{Y\cap D_{\phi(n)}/S} (L_{\phi(n+1)}) \] et \[ I_{X/S} (L_1,\cdots ,L_{n+1}) \stackrel{\sim }{\F } I_{Y/S}(L_{\phi(n)},L_{\phi(n+1)}) \stackrel{\sim }{\F } N_{Y\cap D_{\phi(n+1)}/S} (L_{\phi(n)}) \] Le lemme (\ref{indep2}) entraine que le diagramme \[ \xymatrix{ &N_{Y\cap D_{\phi(n)}/S} (L_{\phi(n+1)})\ar[r]^-{\sim}& I_{Y/S}(L_{\phi(n)},L_{\phi(n+1)})\ar[dd] \\ {\mathcal{O}_{S}} \ar[ur]^-{N_{Y\cap D_{\phi(n)}/S} (\sigma_{\phi(n+1)})} \ar[dr]_-{N_{Y\cap D_{\phi(n+1)}/S} (L_{\phi(n)})} \\ &N_{Y\cap D_{\phi(n+1)}/S} (L_{\phi(n)})\ar[r]^-{\sim}& I_{Y/S}(L_{\phi(n)},L_{\phi(n+1)}) } \] est commutatif, ce qui montre l'assertion. \end{proof} En utilisant l'assertion 2 du lemme (\ref{genericite-suite}), on en d{\'e}duit que $s_\phi$ et $s_\psi$ coincident sur $U_\phi\cap U_\psi$. Les $s_\phi$ d{\'e}finissent donc une section de $I_{X_P/P}$ sur l'ouvert $U$ de $P$. D'apr{\`e}s l'assertion 1 du lemme (\ref{genericite-suite}), une telle section se prolonge de mani{\`e}re unique en une section de $I_{X_P/P}$ sur $P$. \begin{Def} On appelle r{\'e}sultant de $\sigma_1,\cdots ,\sigma_{n+1}$ et on note $Res(\sigma_1,\cdots ,\sigma_{n+1})$ l'uni\-que section de $I_{X_P/P}$ qui coincide avec $s_\phi$ sur l'ouvert $U_\phi$ pour tout $\phi \in S_{n+1}$. \end{Def} Les r{\'e}sultats suivants justifient la terminologie "r{\'e}sultant": \begin{lemme} Soit $R\subset P$ l'image par $\pi$ du lieu des z{\'e}ros de la section $\sigma =\bigoplus\sigma_i$ de $\bigoplus L^{\prime}_i$. $R$ est un ferm{\'e} de $P$ et pour tout point $p\in P$ tel que $\text{Prof}(\mathcal{O}_{P_{f(p)},p})\leq 1$, $p$ est {\'e}l{\'e}ment de $R$ si et seulement si $Res(\sigma_1,\cdots ,\sigma_{n+1})(p)=0$. \end{lemme} \begin{proof}[Preuve] D'apr{\`e}s la partie 2 du lemme (\ref{genericite-suite}), il suffit de montrer que, pour tout point $p$ de l'ouvert $U=\bigcup_{\phi\in S_{n+1}}U_{\phi}$, on a $Res(\sigma_1,\cdots ,\sigma_{n+1})(p)=0$ si et seulement si $p\in R$. Soit donc $p\in U$ et $\phi\in S_{n+1}$ tel que $p\in U_\phi$. Sur $U_\phi$, $Res(\sigma_1,\cdots ,\sigma_{n+1})$ coincide avec $N_{D_{\phi(1)}\cap\cdots\cap D_{\phi(n)}} (\sigma_{\phi(n+1)})$ et cette section s'annulle si et seulement si $\sigma_{\phi(n+1)}$ s'annulle en point de la fibre $(D_{\phi(1)}\cap\cdots\cap D_{\phi(n)})_p$, ce qui montre le r{\'e}sultat. \end{proof} \begin{lemme} \label{sym-resultant} Pour toute permutation $\phi\in S_{n+1}$, l'isomorphisme de sym{\'e}trie \[ Res(L_{\phi(1)},\cdots,L_{\phi(n+1)}) \stackrel{\sim }{\F } Res(L_1,\cdots,L_{n+1}) , \] provenant des isomorphismes de sym{\'e}trie du fibr{\'e} d'intersection, {\'e}change leurs sections respectives $Res(\sigma_{\phi(1)},\cdots,\sigma_{\phi(n+1)})$ et $Res(\sigma_1,\cdots,\sigma_{n+1})$. \end{lemme} \begin{proof}[Preuve] Il suffit de montrer cette {\'e}galit{\'e} sur l'ouvert $V=\bigcap_{\phi\in S_{n+1}}U_{\phi}$ et de plus il suffit de consid{\'e}rer le cas o{\`u} $\phi$ est une transposition de deux indices $i$ et $j$. Si $n=1$, il suffit alors d'appliquer le lemme (\ref{indep2}). Si $n>1$, choisissons une permutation $\psi\in S_n$ telle que $\psi(n+1)\notin \{ i,j\} $. Comme, sur $V$, la section $Res(\sigma_1,\cdots,\sigma_{n+1})$ se factorise en: \[ \mathcal{O}_{S} \longrightarrow N_{D_{\psi(1)}\cap\cdots\cap D_{\psi(n)}/S} (L_{\psi(n+1)}) \stackrel{\sim }{\F } I_{X/S}(L_1,\cdots,L_{n+1})\; , \] le r{\'e}sultat en r{\'e}sulte, puisque $D_{\psi(1)}\cap\cdots\cap D_{\psi(n)}$ est invariant par la transposition $\phi$. \end{proof} \addtocounter{subsubsection}{6} \subsubsection{Multiplicativit{\'e} du r{\'e}sultant} \addtocounter{theo}{1} Soient $L_1^{\prime},L_1\sec,L_2,\cdots,L_{n+1}$ des faisceaux inversibles suf\-fisamment positifs sur $X$ et supposons de plus que $L_1^{\prime}\otimes L_1\sec\gg 0$. On notera alors $P^{\prime}=P_{L_1^{\prime}}\times_S\cdots\times_SP_{L_{n+1}}$, $P\sec =P_{L_1\sec}\times_S\cdots\times_SP_{L_{n+1}}$, $P=P^{\prime} \times_S P\sec$ et $Q=P_{L_1^{\prime}\otimes L_1\sec}\times_S\cdots\times_SP_{L_{n+1}}$. Le morphisme canonique de multiplication des sections $m:P_{L_1^{\prime}}\times_SP_{L_1\sec}\longrightarrow P_{L_1^{\prime}\otimes L_1\sec}$ induit un morphisme not{\'e} toujours $m: P\longrightarrow Q$. Consid{\'e}rons alors le diagramme suivant \begin{equation} \label{diagramme-resultant} \xymatrix{ &X_P\ar[d]_-{\pi}\ar[rr]_-n \ar@/_/[ddl]_-{\pi^{\prime}} \ar@/^/[ddr]^-{\pi\sec} \ar@/^1pc/[rrr]^g &&X_Q\ar[d]_-{\pi}\ar[r]_-f\ar@/^1pc/[ddd]^{\pi_1} &X\ar[d]_-{\pi} \\ &P \ar[rr]_-m \ar[dl]^-{p^{\prime}}\ar[dr]_-{p\sec} &&Q\ar[r] \ar[dd]_-{p_1}&S\\ P^{\prime} \ar[d]_-{p_1^{\prime}}&&P\sec \ar[d]_-{p_1\sec}\\ P_{L_1^{\prime}} &&P_{L_1\sec}& P_{L_1^{\prime}\otimes L_1\sec} } \end{equation} On notera enfin $p_i$ les diff{\'e}rentes projections de $P$, $P^{\prime}$ ou $Q$ sur $P_{L_i}$ et $\pi_i$ leur compos{\'e}e avec $\pi$. \begin{lemme} \label{mult-resultant} Il existe un isomorphisme canonique de faisceaux inversibles sur $P^{\prime}\times_SP\sec$ \[ m^{\ast } Res(L_1^{\prime}\otimes L_1\sec,L_2,\cdots,L_{n+1}) \stackrel{\sim }{\F } (p_1^{\prime})^{\ast } Res(L_1^{\prime},L_2,\cdots,L_{n+1}) \otimes (p_1\sec)^{\ast } Res(L_1\sec,L_2,\cdots,L_{n+1}) \] qui induit une {\'e}galit{\'e} sur les sections: \[ m^{\ast } Res(\tau,\sigma_2,\cdots,\sigma_{n+1}) = (p_1^{\prime})^{\ast } Res(\sigma_1\sec,\sigma_2,\cdots,\sigma_{n+1}) \otimes (p_1\sec)^{\ast } Res(\sigma_1^{\prime},\sigma_2,\cdots,\sigma_{n+1}) \] o{\`u} $\tau$ est la section canonique de $f^{\ast } (L_1\otimes L_1^{\prime})\otimes \pi^{\ast }_1 \mathcal{O}_{P_{L_1^{\prime} \otimes L_1\sec}}(1)$. \end{lemme} \begin{proof}[Preuve] Il existe un isomorphisme canonique de faisceau inversibles sur $P^{\prime}\times_SP\sec$: \[ m^{\ast } p_1^{\ast }\mathcal{O}_{P_{L_1^{\prime}\otimes L_1\sec}}(1) \simeq (p_1^{\prime})^{\ast }\mathcal{O}_{P_{L_1^{\prime}}}(1) \otimes (p_1\sec)^{\ast }\mathcal{O}_{P_{L_1\sec}} \] En prenant l'image inverse par $\pi$ et en tensorisant avec $g^{\ast }(L_1^{\prime}\otimes L_1\sec)$, on obtient sur $X_{P^{\prime}\times_SP\sec}$ un isomorphisme: \[ n^{\ast } \pi^{\ast }\mathcal{O}_{P_{L\otimes M}}(1) \otimes g^{\ast }(L_1^{\prime}\otimes L_1\sec) \simeq \left( (\pi_1^{\prime})^{\ast } \mathcal{O}_{P_{L_1^{\prime}}}(1) \otimes g^{\ast } L_1^{\prime} \right) \otimes \left( (\pi_1\sec)^{\ast } \mathcal{O}_{P_{L_1\sec}}(1) \otimes g^{\ast } L_1\sec \right) \] qui identifie $n^{\ast }\tau$ avec $\sigma_1^{\prime} \otimes \sigma_1\sec$. La prori{\'e}t{\'e} de multiplicativit{\'e} du fibr{\'e} d'intersection $I_{X_Q/Q}$ induit donc un isomorphisme: \begin{multline*} m^{\ast } I_{X_Q/Q} (f^{\ast } (L_1^{\prime}\otimes L_1\sec) \otimes \pi_1^{\ast } \mathcal{O}_{P_{L_1^{\prime}\otimes L_1\sec}}(1),\cdots, f^{\ast } L_{n+1}\otimes \pi_n^{\ast } \mathcal{O}_{P_{L_{n+1}}}(1)) \stackrel{\sim }{\F } \\ I_{X_Q/Q}(g^{\ast } L_1^{\prime}\otimes g^{\ast } L_1\sec \otimes (\pi_1^{\prime})^{\ast } \mathcal{O}_{P_{L_1^{\prime}}}(1) \otimes (\pi_1\sec)^{\ast } \mathcal{O}_{P_{L_1\sec}}(1) ,\cdots, g^{\ast } L_{n+1}\otimes \pi_n^{\ast } \mathcal{O}_{P_{L_{n+1}}}(1)) \\ \stackrel{\sim }{\F } (p_1^{\prime})^{\ast } Res(L_1^{\prime},\cdots,L_{n+1}) \otimes (p_1\sec)^{\ast } Res(L_1\sec,\cdots,L_{n+1}) \end{multline*} qui identifie les sections correspondantes. \end{proof} \subsection{Lien avec la d{\'e}finition classique du r{\'e}sultant} Consid{\'e}rons le sch{\'e}ma $X=\mathbb{P}^n_k$ sur $S=Spec(k)$. Dans tout ce qui suit, on {\'e}crira explicitement $X$ sous la forme $Proj(k[X_1,\cdots,X_{n+1}]$, et on consid{\`e}rera les $n+1$ hyperplans $H_i$ de $X$ d'{\'e}quations $X_i=0$. Fixons des entiers strictement positifs $d_1,\cdots,d_{n+1}$ et consid{\'e}rons les faisceaux $L_i=\mathcal{O}_{X}(d_i)$, qu'on identifiera {\`a} $\mathcal{O}_{X}(d_iH_i)$, en notant $\tau_i$ la section canonique correspondante. Les $L_i$ sont suffisamment positifs et on regarde les espaces projectifs $P_i$ correspondants sont de dimension $\binom{d_i+n}{n}-1$. Notons enfin $V_i=H^0(X,L_i)$.\\ L'isomorphisme canonique (\ref{isom-resultant}) s'{\'e}crit ici: \[ Res (L_1,\cdots,L_{n+1}) \stackrel{\sim }{\F } \bigotimes_{i=1}^{n+1} p_i^{\ast } \mathcal{O}_{P_i}(k_i) \otimes I_{X/S}(L_1,\cdots,L_{n+1}) \] avec $k_i=\prod_{k\neq i}d_k$. Comme $\bigcap_{i=1}^{n+1}H_i=\emptyset$, $I_{X/S}(\tau_1,\cdots,\tau_{n+1})$ d{\'e}finit une trivialisation de $ I_{X/S}(L_1,\cdots,L_{n+1})$. On en d{\'e}duit donc un isomorphisme, associ{\'e} canoniquement au choix des coordonn{\'e}es $X_i$: \[ Res (L_1,\cdots,L_{n+1}) \stackrel{\sim }{\F } \bigotimes_{i=1}^{n+1} p_i^{\ast } \mathcal{O}_{P_i}(k_i) \] La section canonique $Res (\sigma_1,\cdots,\sigma_{n+1})$ de $Res (L_1,\cdots,L_{n+1})$ d{\'e}finit donc une section de $\bigotimes_{i=1}^{n+1} p_i^{\ast } \mathcal{O}_{P_i}(k_i)$ sur le produit $P=P_1\times\cdots\times P_{n+1}$, c'est {\`a} dire un polyn{\^o}me quasi-homog{\`e}ne sur $V=\bigoplus_{i=1}^{n+1}V_i$, de degr{\'e} $k_i$ relativement {\`a} $V_i$. C'est le r{\'e}sultant {\'e}tudi{\'e} par \textsc{Jouanolou} dans \cite{jouanolou1}. On le notera alors $\underline{Res}(\sigma_1,\cdots,\sigma_{n+1})$. Notons alors que les lemmes (\ref{sym-resultant}) et (\ref{mult-resultant}) entrainent que ce polyn{\^o}me est sym{\'e}trique etmultiplicatif en chaque groupe de variables. \subsubsection{La formule de Poisson} Dans cette section, on interpr{\`e}te la formule de Poisson pour les polyn{\^o}mes r{\'e}sultants (\cite{jouanolou1}, Prop 2.7,p124), qui permet de calculer le r{\'e}sultant $\underline{Res}(\sigma_1,\cdots,\sigma_{n+1})$ par r{\'e}currence sur la dimension, en l'exprimant en fonction d'un r{\'e}sultant $\underline{Res}(\sigma_1|_H,\cdots,\sigma_n|_H)$ associ{\'e} {\`a} un hyperplan $H$, et de la norme d'une fonction d{\'e}finie sur l'ouvert affine, compl{\'e}mentaire de $H$.\\ On note ici $H$ l'hyperplan $H_{n+1}$ de $X$ et $Q=P_1\times\cdots\times P_n$ de telle sorte que $P=Q\times P_{n+1}$. Soit $Z\subset X_P$ le lieu des z{\'e}ros communs de $\sigma_1,\cdots ,\sigma_n$, consid{\`e}rons alors les ouverts $U_0=\{ q\in Q| Z\cap H_{n+1}\cap X_q=\emptyset\}$ et $U=U_0\times P_{n+1}\subset P=Q\times P_{n+1}$. $H$ est un espace projectif de dimension $n-1$, muni de coordonn{\'e}es $X_1,\cdots,X_n$, et on peut donc d{\'e}finir un r{\'e}sultant $Res(L_1|_H,\cdots,L_n|_H)$ sur $P_1^{\prime}\times\cdots P_n^{\prime}$, o{\`u} $P_i^{\prime}=\mathbb{P} (H^0(H,L_i)^{\vee })$. R{\'e}sumons d'abord dans un diagramme la situation: \[ \xymatrix{&&Res(\sigma_1,\cdots,\sigma_{n+1}) \ar@{.}[dr] &X_U\subset X_P \ar[d] \\ Res(\sigma_1|_H,\cdots,\sigma_n|_H)\ar@{.}[dr] & H_{Q^{\prime}}\ar[d] &X_{U_0}\subset X_Q \ar[d] & P \ar[dl]_-{\pi_Q}\ar[dr]_-{\pi_{n+1}} & X_{P_{n+1}}\ar[d]\ar[r]_-p &X \\ &Q^{\prime} \ar[d]_-{\pi^{\prime}_i} &U_0\subset Q\ar[l]_-{\pi_{Q^{\prime}}}\ar[d]_-{\pi_i} &&P_{n+1} \\ &P_i^{\prime} &P_i } \] Sur l'ouvert $X\setminus H$, on dispose d'un isomorphisme $L_{n+1} \stackrel{\sim }{\F } \mathcal{O}_{X\setminus H}$. La restriction $\Tilde{\sigma}_{n+1}$ de $\sigma_{n+1}$ {\`a} $P\times (X\setminus H)$ est donc une section de $p_n^{\ast }(\mathcal{O}(1))$. Comme $Z_U$ est contenu dans $(X\setminus H)\times U$, la restriction de $\Tilde{\sigma}_{n+1}$ {\`a} $Z_U$ est bien d{\'e}finie et sa norme $N_{Z_U/U}(\Tilde{\sigma}_{n+1})$ est une section de $\pi_{n+1}^{\ast } \mathcal{O}(k_{n+1})$. Notons que ces consid{\'e}rations montrent l'existence d'un isomorphisme canonique \begin{equation} \label{poisson} Res(L_1,\cdots,L_{n+1})|_U \stackrel{\sim }{\F } \pi_{n+1}^{\ast } \mathcal{O}_{P_{n+1}}(k_{n+1}) \end{equation} qui identifie les sections $Res(\sigma_1,\cdots,\sigma_{n+1}|_U$ et $N_{Z_U/U}(\Tilde{\sigma}_{n+1})$. \begin{prop}[Formule de Poisson] Sur l'ouvert $U$, on a l'{\'e}galit{\'e} entre polyn{\^o}mes de $\Gamma(U_0,\mathcal{O}_{U_0})[V_{n+1}]$ : \[ \underline{Res}(\sigma_1,\cdots,\sigma_{n+1})|_U = \pi_{Q^{\prime}}^{\ast } \underline{Res}(\sigma_1|_H,\cdots,\sigma_n|_H)|_{U_0} . N_{Z_U/U}(\Tilde{\sigma}_{n+1}) \] \end{prop} \begin{proof}[Preuve] Utilisons la d{\'e}finition du faisceau r{\'e}sultant et la multiplicativit{\'e} du fibr{\'e} d'inter\-sec\-tion pour {\'e}crire le diagramme d'isomorphismes: \[ \begin{array}{ccccc} Res(L_1,\cdots,L_{n+1}) & \stackrel{\sim }{\F } & I_{X/S}(L^{\prime}_1,\cdots,L_n^{\prime},p^{\ast } \mathcal{O}_{X}(d_{n+1}H))& \otimes & I_{X/S}(L^{\prime}_1,\cdots,L_n^{\prime},\pi_{n+1}^{\ast }\mathcal{O}_{P_{n+1}}(1))\\ \| &&\downarrow\wr && \downarrow\wr\\ Res(L_1,\cdots,L_{n+1})&& \left( I_{H/S}(L^{\prime}_1|_H,\cdots,L^{\prime}_n|_H)\right)^{d_{n+1}} && N_{Z/P}(\pi_{n+1}^{\ast }\mathcal{O}_{P_{n+1}}(1))\\ \downarrow\wr&&\downarrow\wr && \downarrow\wr\\ \begin{array}{c} \bigotimes_{i=1}^{n+1} \pi_i^{\ast }\mathcal{O}_{P_i}(k_i)\\ \otimes\\ I_{X/S}(L_1,\cdots,L_{n+1}) \end{array} && \begin{array}{c} \bigotimes_{i=1}^n \pi_i^{\ast }\mathcal{O}_{P_i}(k_i)\\ \otimes \\ \left( I_{H/S}(L_1|_H,\cdots,L_n|_H)\right)^{d_{n+1}} \end{array} && \begin{array}{ccc} \pi_{n+1}^{\ast } \mathcal{O}_{P_{n+1}}(k_{n+1})\\ {}\\ {} \end{array} \end{array} \] La deuxi{\`e}me ligne de ce diagramme donne un isomorphisme \begin{equation} \label{poisson2} Res(L_1,\cdots,L_{n+1})\stackrel{\sim }{\F } \left( I_{H/S}(L^{\prime}_1|_H,\cdots,L^{\prime}_n|_H)\right)^{d_{n+1}} \otimes N_{Z/P}(\pi_{n+1}^{\ast }\mathcal{O}_{P_{n+1}}(1)) \end{equation} En combinant la restriction de (\ref{poisson2}) {\`a} $U$ et l'isomorphisme $Res(\sigma_1|_H,\cdots,\sigma_n|_H)|_U: \mathcal{O}_U\stackrel{\sim }{\F } I_{H/S}(L^{\prime}_1|_H,\cdots,L^{\prime}_n|_H)|_U$, on obtient l'isomorphisme (\ref{poisson}). La restriction de (\ref{poisson2}) {\`a} $U$ identifie donc $Res(\sigma_1,\cdots,\sigma_{n+1})|_U$ {\`a} = $\pi_{Q^{\prime}}^{\ast } Res(\sigma_1|_H,\cdots,\sigma_n|_H)|_{U_0} \otimes N_{Z_U/U}(\Tilde{\sigma}_{n+1})$. L'{\'e}galit{\'e} polyn{\^o}miale recherch{\'e}e s'en d{\'e}duit en utilisant la d{\'e}finition des polyn{\^o}mes r{\'e}sultants et les isomorphismes de la partie inf{\'e}rieure du diagramme. \end{proof}
1997-12-21T20:47:45
9712
alg-geom/9712027
en
https://arxiv.org/abs/alg-geom/9712027
[ "alg-geom", "math.AG" ]
alg-geom/9712027
Ron Donagi
Ron Donagi and Ron Livne
The arithmetic-geometric mean and isogenies for curves of higher genus
Latex, 18 pages
null
null
null
null
Computation of Gauss's arithmetic-geometric mean involves iteration of a simple step, whose algebro-geometric interpretation is the construction of an elliptic curve isogenous to a given one, specifically one whose period is double the original period. A higher genus analogue should involve the explicit construction of a curve whose jacobian is isogenous to the jacobian of a given curve. The doubling of the period matrix means that the kernel of the isogeny should be a lagrangian subgroup of the group of points of order 2 in the jacobian. In genus 2 such a construction was given classically by Humbert and was studied more recently by Bost and Mestre. In this article we give such a construction for general curves of genus 3. We also give a similar but simpler construction for hyperelliptic curves of genus 3. We show that the hyperelliptic construction is a degeneration of the general one, and we prove that the kernel of the induced isogeny on jacobians is a lagrangian subgroup of the points of order 2. We show that for g at least 4 no similar construction exists, and we also reinterpret the genus 2 case in our setup. Our construction of these correspondences uses the bigonal and the trigonal constructions, familiar in the theory of Prym varieties.
[ { "version": "v1", "created": "Sun, 21 Dec 1997 19:47:44 GMT" } ]
2007-05-23T00:00:00
[ [ "Donagi", "Ron", "" ], [ "Livne", "Ron", "" ] ]
alg-geom
\section{Introduction} It is well-known that computation of Gauss's arithmetic-geometric mean involves iteration of a simple step, whose algebro-geometric interpretation is the construction of an elliptic curve isogenous to a given one, specifically one whose period is double the original period (for a modern survey see \cite{cox}). A higher genus analogue should involve the explicit construction of a curve whose jacobian is isogenous to the jacobian of a given curve. The doubling of the period matrix means that the kernel of the isogeny should be a lagrangian subgroup of the group of points of order $2$ in the jacobian. In genus $2$ such a construction was given classically by Humbert \cite{hum} and was studied more recently by Bost and Mestre \cite{bome}. In this article we give such a construction for general curves of genus $3$. We also give a similar but simpler construction for hyperelliptic curves of genus $3$. We show that the hyperelliptic construction is a degeneration of the general one, and we prove that the kernel of the induced isogeny on jacobians is a lagrangian subgroup of the points of order $2$. We show that for $g \geq 4$ no similar construction exists, and we also reinterpret the genus $2$ case in our setup. To construct these correspondences we use the bigonal and the trigonal constructions, familiar in the theory of Prym varieties (\cite{don}). In genus $2$ Bost and Mestre note that Humbert's construction induces on jacobians an isogeny whose kernel is of type $(\ZZ/2\ZZ)^2$. We show that Humbert's construction is an instance of the bigonal construction, and prove that the above kernel is a lagrangian subgroup of the points of order $2$. In fact Bost and Mestre use Humbert's construction to give a variant of Richelot's genus $2$ arithmetic-geometric mean. In light of the clear analogy, in particular the fact that a generic principally polarized abelian threefold is a jacobian, one might hope that our construction could be used in a similar way. We work throughout over an algebraically closed field of characteristic $0$. However, our methods clearly extend more generally. For example, the results of Section~\ref{genustwo} hold if the characteristic is not $2$, and those of Sections~\ref{hyperelliptic} and \ref{genusthree} if it is $>3$. The first author thanks the Hebrew University of Jerusalem and the Institute for Advanced Studies in Princeton for their hospitality during the time this work was done. The second author thanks the University of Pennsylvania for its hospitality while this article was being written. \section{Preliminaries} \label{prel} \noindent {\bf Polarizations.} For an abelian variety $A$ denote by $A[n]$ the kernel of multiplication by $n$. In the sequel we will need the following standard facts and notation. \begin{enumerate} \item A polarization $\Theta$ on an abelian variety $A$ induces by restriction a polarization $\Theta_B$ on any abelian subvariety $B$ of $A$. \item Recall that the type of a polarization $\Theta$ on a $g$\/-dimensional abelian variety is a g\/-tuple of positive integers $d_g|\dots|d_2|d_1$. We say that $\Theta$ is a principal polarization if it is of type $1^g = (1,\dots,1,1)$ ($g$\/ times). In that case, suppose that $p$ is a prime, and that $K$ is a subgroup of $A[p]$ isomorphic to $({\bbB Z}/p{\bbB Z})^r$ and isotropic for the Weil pairing ${\rm w}_p$. Then $\Theta$ induces a polarization on $A/K$, characterized by the property that its pull back to $A$ is $p\Theta$. Its type is then $p^{g-r}\cdot 1^r$. In this situation we will say that $K$ is a lagrangian subgroup of $A[p]$ if $r=g$. \item The type of a polarization is preserved under continuous deformations. \end{enumerate} \vspace*{0.1cm} \noindent {\bf Double covers.} Given a double cover, i.e. a finite morphism $\pi: \tilde{C} \rightarrow C$ of degree $2$ between smooth projective curves, the Prym variety ${\rm Prym}\,(\tilde{C}/C)$ is defined to be the connected component of the kernel of the norm map \[ \pi_*: {\rm Jac}\,(\tilde{C}) \rightarrow {\rm Jac}\,(C). \] It is an abelian variety, and it has a natural principal polarization when $\pi$ is unramified, namely one half of the polarization induced on it as an abelian subvariety of ${\rm Jac}\,(\tilde{C})$ (\cite{mum2}). This definition extends to singular curves $C$, $\tilde{C}$, if we interpret ${\rm Jac}\,$ as the (not necessarily compact) generalized jacobian. This was studied by Beauville \cite{bea}. Particularly important for us will be the cases when 1. $C$, $\tilde{C}$ have only ordinary double points, 2. $\pi^{-1}(C_{\rm sing}) = \tilde{C}_{\rm sing}$, and 3. for each $x\in C_{\rm sing}$ the inverse image $\pi^{-1}(x)$ consists of a single point, and each branch of $\pi^{-1}(x)$ maps to a different branch of $x$ and is ramified over it. (We shall then say that $\pi$ is of Beauville type at $x$.) In such cases ${\rm Prym}\,(\tilde{C}/C)$ is compact, and the following three conditions are equivalent: \begin{enumerate} \item $\pi$ is unramified away from $C_{\rm sing}$. \item The arithmetic genera satisfy $g(\tilde{C}) = 2g(C)-1$. \item The cover $\tilde{C}/C$ is a flat limit of smooth unramified double covers. \end{enumerate} We shall call a cover satisfying these conditions allowable; from the third condition we see that the ${\rm Prym}\,$ is prinicipally polarized in such a case. Let $C$ be a curve having only ordinary double points as singularities, and let $\nu_x: N_x \rightarrow C$ be the normalization map of exactly one such singular point $x$. We denote by $L(x)$ the line bundle of order $2$ in ${\rm Ker}\,\nu_x^\ast$. (It is obtained from the trivial line bundle on $N_x$ by gluing the fibers over the two inverse images of $x$ with a twist of $-1$ relative to the natural identification.) \begin{lemma} \label{polar} Let $\pi: \tilde{C} \rightarrow C$ be an allowable double cover, $\nu \pi: \nu \tilde{C} \rightarrow \nu C$ its (partial) normalization at $r \geq 1$ ordinary double points $x_1,\dots,x_r$. Let $g$ be the (arithmetic) genus of the partial normalization $\nu C$, so the arithmetic genus of $C$ is $g+r$. Then ${\rm Prym}\, (\tilde{C}/C)$ has a principal polarization, ${\rm Prym}\,(\nu \tilde{C}/ \nu C)$ has a polarization of type $2^{g}1^{r-1}$, and the pullback map $\nu^*:{\rm Prym}\,(\tilde{C}/C) \rightarrow {\rm Prym}\,(\nu\tilde{C}/\nu C)$ is an isogeny of degree $2^{r-1}$. The kernel of $\nu^*$ is the subgroup of ${\rm Prym}\,(\tilde{C}/C)[2]$ generated by the pairwise differences of the line bundles $L(x_i)$ defined above. This subgroup is isotropic for the mod $2$ Weil pairing $w_2$. \end{lemma} \begin{pf} The generalized jacobians fit in short exact sequences \[\begin{array}{ccccccccc} 0 & \rightarrow & \GG_m^r & \rightarrow & {\rm Jac}\,(\tilde{C}) & \rightarrow & {\rm Jac}\,(\nu \tilde{C}) &\rightarrow & 0 \\ & & \downarrow & & \downarrow & & \downarrow & & \\ 0 & \rightarrow & \GG_m^r & \rightarrow & {\rm Jac}\,(C) & \rightarrow & {\rm Jac}\,(\nu C) &\rightarrow & 0 \end{array} \] where the vertical maps are the norm maps induced by $\pi$ and by $\nu\pi$. We compare the kernels: to begin with, the kernel of the norm map is connected for ramified double covers (in particular for $\nu \tilde{C}/\nu C$), and has two components for unramified covers. This is shown in \cite{mum2} in the nonsingular case, and so by continuity this holds also for allowable singular covers (in particular for $\tilde{C}/C$). The multiplicative groups parametrize extension data and the norm is the squaring map. So the short exact sequence of kernels gives \[0 \rightarrow (\ZZ/2\ZZ)^r \rightarrow \Prym(\Ctil/C) \times \ZZ/2\ZZ \stackrel{\nu^*}{\rightarrow} \Prym(\nu \Ctil/\nu C) \rightarrow 0,\] and the first part of the lemma follows. To prove that the subgroup $(\ZZ/2\ZZ)^{r-1}$ of $\Prym(\Ctil/C)[2]$ is isotropic for ${\rm w}_2$, notice that its generators are reductions modulo $2$ of the vanishing cycles for $\tilde{C}$, and vanishing cycles for distinct ordinary double points are disjoint. Therefore these vanishing cycles have $0$ intersection number in ${\bbB Z}$\/- (or ${\bbB Q}$\/-) homology. By the definition of the polarization of $\Prym(\Ctil/C)$ in Section~\ref{prel}, and the well-known expression for the Weil pairing in terms of the intersection (or cup product) pairing (see e.g. \cite[theorem 1, Ch. 23]{mum}), the rest of the lemma follows. \end{pf} \section{The bigonal and the trigonal constructions} \label{bigonal} There are several elementary constructions which associate a double cover of some special kind with another cover (or curve) with related Prym (of Jacobian). We now review the bigonal and the trigonal constructions, following (\cite{don}). Assume we are given smooth projective curves $\tilde{C}$, $C$ and $K$ and surjective maps $f:C\rightarrow K$ and $\pi:\tilde{C}\rightarrow C$, so that $\deg\pi = \deg f=2$ over any component. The bigonal construction associates new curves and maps of the same type $\tilde{{C'}}\sra{\pi'} C'\sra{f'}K$ as follows. Let $U\subset K$ be the maximal open subset over which $f\pi$ is unramified. Then $\tilde{{C'}}$ represents over $U$ the sheaf of sections, in the complex or the \'etale topology, of $\pi:(f\pi)^{-1}U\rightarrow f^{-1}U$. It is a $4$\/-sheeted cover of $U$. We then view $\tilde{{C'}}_{|U}$ as a locally closed subvariety of $\tilde{C} \times \tilde{C}$ and define $\tilde{{C'}}$ as the closure. The projection to $U$ extends to a morphism $\tilde{{C'}}\rightarrow K$, and the involution $\iota$ of $\tilde{{C'}}_{|U}$ which sends a section to the complementary section extends to $\tilde{{C'}}$. We define $C'=\tilde{{C'}}/\iota$ and $f'$ and $\pi'$ as the quotient maps. We will need to extend this construction to allowable covers of curves with ordinary double points; however in a family acquiring a singularity of Beauville type the arithmetic genus of the resulting $\tilde{{C'}}$ is not locally constant. More technically, the naive construction as the closure of $\tilde{{C'}}_{|U}$ in $\tilde{C}\times\tilde{C}$ is not flat in families, which is not adequate for our purposes: for example, we want the bigonal construction to be symmetric. To achieve this, we define the bigonal construction for singular allowable covers by {\em choosing} a flat family of smooth covers whose limit is our allowable cover, and {\em defining} the construction to be the limit of the construction for the nonsingular fibers. Beauville's results imply that this is well defined, and does give a symmetric construction: this is more or less clear except at a singularity of Beauville type. There the problem reduces to a local calculation whose answer, which we record in \ref{type}. below, is visibly symmetric. We will need a few properties of this construction (see \cite[Section 2.3]{don}) \begin{enumerate} \item As we said, the construction over $U$ is symmetric: starting with $\tilde{{C'}},\dots,f'$ gives back $\tilde{C},\dots,f$. \item \label{type} Denote the type of $\tilde{C}/C$ at a point $k\in K$ by \begin{itemize} \item $\ssub \seq \sslash \smin\smin $ if $C$ is unramified over $k$ and $\tilde{C}$ is ramified over exactly one point in $f^{-1}(k)$; \item $\stackrel{\subset}{\ssub} \sslash \ssub $ if $C$ is ramified over $k$ but $\tilde{C}$ is unramified over the point $f^{-1}(k)$; \item $\ssub \ssub \sslash \smin\smin $ if $C$ is unramified over $k$ and $\tilde{C}$ is ramified over both branches of $C$ over $k$; \item $\btf$ if both $C$, $\tilde{C}$ have ordinary double points above it, and $\tilde{C}/C$ is of Beauville type there. \end{itemize} If $\tilde{C}/C$ is of type $\ssub \seq \sslash \smin\smin $, $\stackrel{\subset}{\ssub} \sslash \ssub $, $\ssub \ssub \sslash \smin\smin $, $\btf$ at $k$ then $\tilde{{C'}}/C'$ is respectively of type $\stackrel{\subset}{\ssub} \sslash \ssub $, $\ssub \seq \sslash \smin\smin $, $\btf$,$\ssub \ssub \sslash \smin\smin $ there. Notice that normalization takes type $\btf$ to type $\ssub \ssub \sslash \smin\smin $. \item The natural $2$-$2$ correspondence between $\tilde{C}$ and $\tilde{{C'}}$ induces an isogeny $\Prym(\Ctil/C)\rightarrow\Prym(\tilde{{C'}}/C')$, whose kernel is the same as the kernel of the natural isogeny $\Prym(\Ctil/C)\rightarrow\Prym(\Ctil/C)^\vee$ induced by the polarization from $\Prym(\Ctil/C)$ to its dual abelian variety $\Prym(\Ctil/C)^{\vee}$. In other words we get an isomorphism $\Prym(\Ctil/C)^\vee\sra{\sim}\Prym(\tilde{{C'}}/C')$ (cf. Pantazis \cite{pan}, at least when $K={\bbB P}^1$ which is all we need). As a check, let $a$, $b$, $c$, and $d$ be the numbers of points where $\tilde{C}/C$ is of type $\ssub \seq \sslash \smin\smin $, $\stackrel{\subset}{\ssub} \sslash \ssub $, $\ssub \ssub \sslash \smin\smin $, and $\btf$ respectively. Then by Lemma~\ref{polar} the polarization type for $\Prym(\Ctil/C)$ is $1^{\frac{a+2c}{2}-1}2^{\frac{b+2d}{2}-1}$. Similarly the polarization type for $\Prym(\tilde{{C'}}/C')$ is obtained by interchanging $a$ with $b$ and $c$ with $d$, and this gives exactly the type dual to the one of $\Prym(\Ctil/C)$. \end{enumerate} For Recillas's trigonal construction start with $K$, $C$, $\tilde{C}$, $\pi$, and $f$ as before except that $f$ now has degree $3$. We get a cover $g:X\rightarrow K$ of degree $4$ by making over the smooth unramified part $U$, defined as before, a construction analogous to what we previously did to get $C'$. Namely, let $\tilde{X}/U$ represent the sheaf of sections of $\pi: (f\pi)^{-1}U\rightarrow f^{-1}U$, and define $X/U$ as the quotient of $\tilde{X}$ divided by $\iota$ (which is defined as before). In the nonsingular case we define $X$ as the closure of $X/U$ in $X\times X\times X$, and in the general allowable case by taking a flat limit of the construction for smooth, unramified covers. Here we have (\cite[Section 2.4]{don}) \begin{enumerate} \item Over $U$ the construction is reversible: $C_{|U}$ represents the sheaf of partitions of $X_{|U}$ to two pairs of sections, and $\tilde{{C'}}$ represents the choice of one of these pairs. \item Denote the type of $\tilde{C}/C$ at a point $k\in K$ by \begin{itemize} \item $\tto$ for $C$, $\tilde{C}$ if $C$ has exactly one simple branch point over $K$ and $\pi$ is unramified over $f^{-1}(k)$; \item $\ssub\ssub\seq \sslash \smin\smin\smin $ if $f$ is unramified at $k$ and $\pi$ is branched over two of the branches of $f$ and unramified over the third; \item $\tttp$ if two branches of $C$ over $K$ cross normally, the third is unramified, and moreover, if $\tilde{C}/C$ is of Beauville type over the double point and unramified over the unramified branch. \end{itemize} Then $X$ has exactly one simple branch point at a point $k\in K$ of type $\tto$ for $\tilde{C}/C$, and we denote by $\ssub \smin\smin $ the type of $X$ over $k$. Conversely, if $X$ is of type $\ssub \smin\smin $ at $k$ then $\tilde{C}/C$ is of type $\tto$ there. If $\tilde{C}/C$ is of type $\ssub\ssub\seq \sslash \smin\smin\smin $ at $k$ then $X$ has two simple branch points over $k$, which we denote by type $\ssub \ssub $. Here the situation is not reversible: if $X$ is of type $\ssub \ssub $ at $k$ then $\tilde{C}/C$ is of type $\tttp$ there. Notice that normalization takes type $\tttp$ to type $\ssub\ssub\seq \sslash \smin\smin\smin $. \item If $K\simeq{\bbB P}^1$ and $\tilde{C}/C$ is allowable, then $X$ is smooth and ${\rm Jac}\,(X)\simeq{\rm Prym}\,(\tilde{C}/C)$. This is due to Recillas when $\tilde{C}/C$ is smooth unramified, and again limiting arguments imply this in general. \end{enumerate} \section{The genus $2$ case} \label{genustwo} Humbert's correspondence of curves of genus $2$ was studied by Bost and Mestre (see \cite{hum}, \cite{bome}). We shall show how to make this correspondence via the bigonal construction, and use this to determine the type of the isogeny. Humbert's construction starts with a conic $C$ in ${\bbB P}^2$ with $6$ general points on it (see Remark~\ref{gen} below), which are given as $3$ unordered pairs $\{P'_i,P_i''\}$, $i=1,2,3$. It associates to these $3$ new unordered pairs of points, all distinct, on $C$ as follows. Let $\overline{P'_iP_i''}$ be the $3$ lines joining paired points, and let $l_k$ be the intersection of $\overline{P'_iP_i''}$ and $\overline{P'_jP_j''}$ if $\{i,j,k\}=\{1,2,3\}$. The new $3$ unordered pairs of points on $C$ are then the pairs of points of tangency to $C$ from the $l_k$\/'s. For our purposes it is more convenient to view the new points as lying on the conic $C^*$ dual to $C$ in the dual plane ${\bbB P}^{2*}$. A point of ${\bbB P}^{2*}$ is a line in ${\bbB P}^2$; it is in $C^*$ if and only if this line is tangent to $C$. Let $\phi:{\bbB P}^2\rightarrow{\bbB P}^{2*}$ be the isomorphism defined by $C$; namely, for $P\not\in C$ there are two tangents to $C$ through $P$, and $\phi(P)$ is the line joining their points of tangency. For $P\in C$, $\phi(P)$ is the tangent to $C$ at $P$. Under the isomorphism $\phi_{|C}:C\sra{\sim} C^*$, Humbert's new pairs go to the pairs $L'_k,L''_k$ of tangents to $C$ through $l_k$. \begin{theorem} Let $\pi:H\rightarrow C$ and $\pi^*:H^*\rightarrow C^*$ be double covers branched over the old and new sets of points respectively. Then there is an isogeny ${\rm Jac}\,(H)\rightarrow{\rm Jac}\,(H^*)$ whose kernel is a lagrangian subgroup of ${\rm Jac}\,(H)[2]$. \end{theorem} \begin{pf} Choose some $k\in\{1,2,3\}$. The set $L^* = L_k^*$ of lines through $l_k$ is the line in ${\bbB P}^{2*}$ dual to $l_k$. Let $f:C\rightarrow L^*$ be the ``projection'' sending each point of $C$ to the line joining it to $l_k$. Dually, let $f^*:C^*\rightarrow L = \overline{P_k'P_k''}$ be the ``projection'' sending each tangent line of $C$ to its intersection with $L$. Let $\psi:L\rightarrow L^*$ be the isomorphism sending a line through $l_k$ to its intersection with $L$. The maps $f$ and $f^*$ have degree $2$, and hence also $g=\psi f^*$ has degree $2$. Both coverings $H\sra{\pi}C\sra{f}L^*$ and $H^*\sra{\pi^*}C^*\sra{g}L^*$ are unramified over the complement in $L^*$ of the six points \begin{itemize} \item The tangents $L'_k,L''_k$ to $C$ through $l_k$; there $H/C$ is of type $\stackrel{\subset}{\ssub} \sslash \ssub $ and $H^*/C^*$ is of type $\ssub \seq \sslash \smin\smin $. \item The lines $\overline{P_i'P''_i}$ and $\overline{P_j'P''_j}$ whose intersection defines $l_k$; there both $H/C$ and $H^*/C^*$ are of type $\ssub \ssub \sslash \smin\smin $. \item The lines $\overline{P'_kl_k}$ and $\overline{P''_kl_k}$; there $H/C$ is of type $\ssub \seq \sslash \smin\smin $ and $H^*/C^*$ is of type $\stackrel{\subset}{\ssub} \sslash \ssub $. \end{itemize} It follows that if we perform the bigonal construction on $H\rightarrow C\rightarrow L^*$, the two points $\overline{P_i'P''_i}$ and $\overline{P_j'P''_j}$ of $L^*$ are of Beauville type for the resulting cover $H'\rightarrow C' \rightarrow L^*$ and there are no other singularities (see Section~\ref{bigonal}). The preceding analysis of the ramification of $H^*/C^*$, combined with the one for the bigonal construction $H'/C'$ in Section~\ref{bigonal} shows that the normalization of $H'/C'$ is isomorphic to $H^*/C^*$. It remains to determine the kernel of the induced isogeny on jacobians; by Pantazis's result recalled above, it factors as \[ \begin{array}{rcl} {\rm Jac}\, H & \simeq & {\rm Prym}\,(H/C) \sra{\sim} {\rm Prym}\,(H/C)^{\vee} \sra{\sim} {\rm Prym}\,(H'/C') \\ & \sra{\nu^{\ast}} & {\rm Prym}\,(H^*/C^*) \simeq {\rm Jac}\, H^* \,. \end{array}\] To compute the kernel of $\nu^{\ast}$ we cannot use Lemma~\ref{polar} directly, since $H'/C'$ is not allowable, being ramified over two points $x'\in g^{-1}(L'_k)$, $x''\in g^{-1}(L''_k)$. Instead glue $x'$ to $x''$ to obtain a curve with one more double point $C''$ and glue their inverse images in $H'$ to get a curve $H''$, which is now an allowable cover of $C''$ ($H''$ is obtained from $H$ by gluing the Weierstrass points in pairs). We have maps of covers $H^*/C^*\rightarrow H'/C'\rightarrow H''/C''$ inducing maps of Prym varieties. Applying Lemma~\ref{polar} twice now gives that the kernel of ${\rm Prym}\,(H''/C'') \rightarrow {\rm Prym}\,(H/C)$ is an isotropic subgroup isomorphic to $(\ZZ/2\ZZ)^2$ and that ${\rm Prym}\,(H''/C'')$ is isomorphic to ${\rm Prym}\,(H^*/C^*)$. This implies that ${\rm Ker}\,\nu^{\ast}$ is as asserted, completing the proof of the Theorem. \end{pf} \begin{remark} \label{gen} {\rm The points $\{P'_i,P''_i\}$ are assumed general only to guarantee that they are distinct and that the resulting new $6$ points are also distinct (for which it suffices that the tangents to $C$ from $l_k$ in the proof do not touch $C$ at $P'_k$ nor at $P''_k$). In the case considered in \cite{bome} this holds, because they assume that $C$ and the points are real and satisfy some ordering relations. } \end{remark} \section{The hyperelliptic genus $3$ case} \label{hyperelliptic} In this section we will solve our problem in the hyperelliptic case: we will construct a correspondence between the generic hyperelliptic curve of genus $3$ and a certain non-generic curve of genus $3$ (which is not hyperelliptic). Let $H$ be a hyperelliptic curve of genus $3$ and let $\pi_1:H\rightarrow{\bbB P}^1$ be the hyperelliptic double cover. Choose a grouping in pairs of the $8$ branch points $w_1,\dots,w_8\in{\bbB P}^1$ of $\pi_1$. We claim that there exists a map $g_1:{\bbB P}^1\rightarrow{\bbB P}^1$, of degree $3$, which identifies paired points. This can be seen in several ways. Firstly, let $T$ be the curve obtained from ${\bbB P}^1$ by identifying paired points to ordinary double points. We think of $T$ as a curve of genus $4$ and take its canonical embedding to ${\bbB P}^3$. As in the nonsingular case, the canonical map is well behaved, and in particular the canonical image of $T$ lies on a unique, generically nonsingular quadric by the Riemann-Roch theorem. Projecting via either of the two ruling of this quadric will give the desired map $g_1$. Notice that by its construction $g_1$ factors as ${\bbB P}^1\sra{\nu} T \sra{g} {\bbB P}^1$, where $\nu$ is a normalization map. Another way to get $g_1$ is to embed ${\bbB P}^1$ in ${\bbB P}^3$ as a rational normal curve. We look for a projection from ${\bbB P}^3$ to ${\bbB P}^1$ which identifies paired points. The center of this projection is a line $L$ which must meet the $4$ lines joining the pairs. The grassmanian $G(1,{\bbB P}^3)$ of lines in ${\bbB P}^3$ is naturally a quadric in ${\bbB P}^5$ and the condition to meet a line is a linear condition. We see again that there is always at least one such $L$, and generically two. We now perform the trigonal construction. This gives a map of degree $4$ $f:C\rightarrow{\bbB P}^1$ sitting in a diagram \begin{equation} \label{hyp1} \begin{array}{rcl} &&H \\ &&\downarrow \! \mbox{$\scriptstyle{\pi_1}$} \\ C&&{\bbB P}^1\\ &\sf\!\!\searrow\;\;\swarrow\!\!\mbox{$\scriptstyle{g_1}$}&\\ &{\bbB P}^1& \end{array} \end{equation} Let $w_{12},\dots,w_{78}$ be the $4$ images of the $w_i$\/'s under $g_1$, with the indices indicating the grouping. By the Riemann-Hurwitz formula there are generically $4$ points $a_1,\dots,a_4$ in ${\bbB P}^1$ over which $g_1$ is branched, with a simple branch point over each. Hence $H/{\bbB P}^1$ is of type $\tto$ at each $a_i$ and of type $\ssub\ssub\seq \sslash \smin\smin\smin $ at each $w_{2i-1,2i}$. >From the properties of the trigonal construction we get $2-2g(C)=8-8-4$, so that $C$ has genus $3$. The trigonal construction gives a birational correspondence between \begin{itemize} \item The moduli of the data $(H \sra{\pi_1}{\bbB P}^1 \sra{g_1} {\bbB P}^1)$ with $4$ points of type $\tto$ and $4$ points of type $\ssub\ssub\seq \sslash \smin\smin\smin $. \item A component of the Hurwitz scheme parametrizing $4$-sheeted covers $f:C\rightarrow {\bbB P}^1$ with $4$ simple branch points and $4$ double branch points. \end{itemize} Each of these moduli spaces is $5$ dimensional. (Another component of this Hurwitz scheme parametrizes bielliptics, namely maps $f:C\rightarrow {\bbB P}^1$ which factor through a double cover $E \rightarrow {\bbB P}^1$ where $E$ is elliptic. Curves in this latter component are taken by the trigonal construction to towers $H\rightarrow \overline{H} \rightarrow {\bbB P}^1$ where $\overline{H}=A \cup B$ is reducible, with $A$, $B$ of degrees $1$, $2$ respectively over ${\bbB P}^1$. We shall not need this component in what follows.) The key point for us is that the trigonal construction induces an isogeny ${\rm Jac}\,(C)\rightarrow{\rm Jac}\,(H)$ whose kernel is lagrangian in ${\rm Jac}\,(C)[2]$. More precisely we have the following \begin{proposition} Let ${\bbB P}^1\sra{\nu}T\sra{g}{\bbB P}^1$ be as before, and let $\tilde{T}$ be the curve obtained by identifying the Weierstrass points in $H$ to ordinary double points with the same grouping as the one we chose to get $T$. Then \begin{enumerate} \item Diagram (\ref{hyp1}) extends to \[ \begin{array}{rcccl} &&\tilde{T}&\sla{\tilde{\nu}}&H\\ &&\mbox{$\scriptstyle{\pi}$}\!\downarrow\;&&\;\downarrow\!\mbox{$\scriptstyle{\nu}$}\mbox{$\scriptstyle{\pi}$}\\ C&&T&\sla{\nu}&{\bbB P}^1\\ &\sf\!\!\searrow\;\;\swarrow\!\!\mbox{$\scriptstyle{g}$}&\\ &{\bbB P}^1& \end{array}\,. \] Here $\tilde{\nu}:H\rightarrow\tilde{T}$ is the normalization map, and we view $\nu\pi:=\pi_1:H\rightarrow {\bbB P}^1$ as the normalization of $\pi:\tilde{T}\rightarrow T$. \item $\tilde{\nu}$ induces an isogeny of polarized abelian varieties $\tilde{\nu}^*:{\rm Prym}\,(\tilde{T}/T)\rightarrow{\rm Jac}\,(H)$ whose kernel is lagrangian in ${\rm Prym}\,(\tilde{T}/T)[2]$. \item Let $\phi:{\rm Jac}\,(C)\rightarrow{\rm Jac}\,(H)$ be the isogeny obtained by composing $\tilde{\nu}^*$ with the isomorphism ${\rm Jac}\,(C)\simeq{\rm Prym}\,(\tilde{T}/T)$. Then the kernel of $\phi$ is lagrangian in ${\rm Jac}\,(C)[2]$, and the kernel of the dual isogeny $\phi^*:{\rm Jac}\,(H)\rightarrow{\rm Jac}\,(C)$ is the lagrangian subgroup of ${\rm Jac}\,(H)[2]$ generated by the differences of identified Weierstrass points. \end{enumerate} \end{proposition} \begin{pf} Part 1. holds because $\tilde{T}/T$ is allowable. The pairs of points of $H$ identified by $\nu$ lie over points of type $\tttp$ for $T$, $\tilde{T}$. Hence they are branch points for $\pi$, namely Weierstrass points. The rest follows from Lemma~\ref{polar}. \end{pf} \section{The generic genus $3$ case} \label{genusthree} Let $C$ be a generic curve of genus $3$. In this section we shall give a construction of a curve $C'$ of genus $3$ and an isomorphism ${\rm Jac}\, (C)/L \simeq {\rm Jac}\, (C')$ where $L$ is a lagrangian subgroup of ${\rm Jac}\, (C)[2]$. Let $f:C\rightarrow{\bbB P}^1$ be a map of degree $4$, and let $b_1$, $b_2$ be points in ${\bbB P}^1$ such that $f$ has two simple branch points over each $b_i$. It is easy to show such $f$, $b_1,b_2$ exist, and in fact we will parametrize the space of such $f$\/'s in the end of this section. We perform the trigonal construction on $f$. This gives curves $T$, $\tilde{T}$ and maps $g:T\rightarrow{\bbB P}^1$ and $\pi:\tilde{T}\rightarrow T$, with $\deg g = 3$ and $\deg \pi=2$. Let $\tilde{\nu}:\nu \tilde{T}\rightarrow\tilde{T}$ and $\nu:\nu T\rightarrow T$ be normalization maps and let $\nu\pi:\nu\tilde{T}\rightarrow\nu T$ be the map induced by $\pi$. The properties of the trigonal construction show the following. Firstly, $T$ and $\tilde{T}$ have each two ordinary double points, one over each $b_i$, and no other singularities. Next, the map $g\nu:\nu T\rightarrow{\bbB P}^1$ has exactly $8$ branch points, all simple, one over each $a_i$. It follows that the genus $g(\nu T)$ is $2$ and therefore the arithmetic genus $g(T)$ is $4$. The map $\nu\pi$ has exactly $4$ ramification points $P_i$, $Q_i$, two over each $b_i$ for $i = 1,2$, and hence $g(\nu\tilde{T})=5$ and $g(\tilde{T})=7$. Since $\nu T$ has genus $2$, it is hyperelliptic. Let $h:\nu T\rightarrow {\bbB P}^1$ be the hyperelliptic double cover, and let $w_1,\dots,w_6\in{\bbB P}^1$ be the branch points of $h$. The bigonal construction gives curves and maps of degree $2$ $\nu\tilde{T}'\sra{\nu\pi'}\nu T'\sra{h'}{\bbB P}^1$. The points in ${\bbB P}^1$ over which $h\nu\pi$ is not \'etale are the $6$ $w_i$\/'s, which are of type $\stackrel{\subset}{\ssub} \sslash \ssub $ for $\nu T$ and $\nu\tilde{T}$, and the $4$ points $h(P_i)$\/, $h(Q_i)$\/, $i=1,2$, which are of type $\ssub \seq \sslash \smin\smin $. The types get reversed for $\nu T'$ and $\nu\tilde{T}'$, and in particular $\nu T'$ is ramified exactly over the $h(P_i)$\/'s and the $h(Q_i)$\/'s. It follows that $g(\nu T')=1$. We also see that $\nu\pi'$ has $6$ branch points, say $w'_1,\dots,w'_6$, one over each of the $w_i$\/'s, and hence $g(\nu\tilde{T}')=4$. The curves $\nu T'$ and $\nu\tilde{T}'$ are nonsingular. Choose a grouping of the $w_i$\/'s in $3$ pairs. Identify the corresponding $w'_i$\/'s in $\nu T'$ to get a curve $T'$ with $3$ ordinary double points, say $w'_{12}$, $w'_{34}$, $w'_{56}$, the indices indicating the groupings. $T'$ has arithmetic genus $4$. Likewise identify the corresponding points above the $w'_i$\/'s on $\nu\tilde{T}'$ to obtain a curve $\tilde{T}'$ with $3$ ordinary double points and arithmetic genus $7$. As in the nonsingular case, the canonical embedding sends $T'$ to ${\bbB P}^3$ and the image sits on a unique, generically smooth quadric. Choosing one of the two rulings of this quadric gives a map $g':T'\rightarrow{\bbB P}^1$. This map is of degree $3$, because the canonical curve is a curve of type $(3,3)$ on the quadric. The map $g'\nu':\nu T'\rightarrow{\bbB P}^1$ is ramified over $n=6$ points, since $2-2g(\nu T')=0=3(2-2g({\bbB P}^1))-n$. Over these the pair $\nu T'$, $\nu\tilde{T}'$ is of type $\tto$. There are also $3$ points of type $\ssub\ssub\seq \sslash \smin\smin\smin $, the images under $g'$ of the identified pairs $w'_{12}$, $w'_{34}$, $w'_{56}$. The trigonal construction performed on $\nu\tilde{T}'\slra{\nu\pi'}\nu T'\slra{g'\nu'}{\bbB P}^1$ gives a curve $C'$ and a map $f':C'\rightarrow{\bbB P}^1$ of degree $4$. We readily see it has genus $3$. The following diagram summarizes the procedure: \[ \begin{array}{rcccccccccl} &&{}_7\tilde{T}&\sla{\mbox{$\scriptstyle{\tilde{\nu}}$}}&{}_5\nu\tilde{T}&&{}_4\nu\tilde{T}'& \sra{\mbox{$\scriptstyle{\tilde{\nu}}$}'}& {}_7\tilde{T}'&&\\ &&\mbox{$\scriptstyle{\pi}$}\!\downarrow&&\downarrow\mbox{$\scriptstyle{\nu}$}\mbox{$\scriptstyle{\pi}$}&&\mbox{$\scriptstyle{\nu}$}\mbox{$\scriptstyle{\pi}$}'\!\downarrow&&\downarrow\mbox{$\scriptstyle{\pi}$}'&&\\ {}_3 C&&{}_4 T&\sla{\nu}&{}_2\nu T&&{}_1\nu T'&\sra{\nu'}& {}_4 T'&&{}_3 C'\\ &\sf\!\searrow\;\swarrow\!\mbox{$\scriptstyle{g}$}&&&& \mbox{$\scriptstyle{h}$}\!\searrow\;\swarrow\!\mbox{$\scriptstyle{h}$}'&&&& \mbox{$\scriptstyle{g}$}'\!\searrow\;\swarrow\!\sf'&\\ &{\bbB P}^1&&&&{\bbB P}^1&&&&{\bbB P}^1& \end{array}\,. \] Before stating our main result we need to discuss the choices made in the construction. Writing $f^{-1}(b_i) = 2(P_i+Q_i)$, we obtain a point of order $2$ \[ \alpha = \alpha(f) = P_1 + Q_1 - P_2 - Q_2. \] in ${\rm Jac}\,(C)$. The trigonal isomorphism ${\rm Jac}\,(C) \simeq {\rm Prym}\,(\tilde{T}/T)$ maps $\alpha$ to the difference $L(b_1) -L(b_2)$ (defined in the discussion preceeding Lemma \ref{polar}), which is the nontrivial element in ${\rm Ker}\,\nu^\ast$. Moreover the only choice made other than $f$ is the grouping of $w_1,\dots,w_6$ under $\nu'$. The differences of the corresponding paired points in $\nu T$ are the nonzero elements of a lagrangian subgroup $L_0$ of ${\rm Jac}\,(\nu T)[2]$. Now observe that the pullback to $\nu\tilde{T}$ by $\nu\pi$ of a line bundle of order $2$ on $\nu T$ is in the kernel of the norm map to $\nu T$. This gives a symplectic embedding \[ \iota:{\rm Jac}\,(\nu T)[2] \hookrightarrow {\rm Prym}\,(\nu\tilde{T}/\nu T)[2]\,. \] The image ${\rm Im}\,(\iota)$ of $\iota$ can be described in two ways. On the one hand, it is the kernel of the polarization map ${\rm Prym}\,(\nu\tilde{T}/\nu T) \rightarrow {\rm Prym}\,(\nu\tilde{T}/\nu T)^\vee$. (Observe that ${\rm Prym}\,(\nu\tilde{T}/\nu T)$ has a polarization of type $221$ by Lemma~\ref{polar}, whose kernel is then isomorphic to $({\bbB Z}/2{\bbB Z})^4$.) On the other hand, let $\alpha^\perp$ denote the orthogonal complement to (the image of) $\alpha$ in ${\rm Prym}\,(\tilde{T}/T)$ for the Weil pairing $w_2$. Then ${\rm Im}\,(\iota)$ is also the pullback of $\alpha^\perp$ by the normalization map. Indeed, for $u\in {\rm Jac}\,(\nu T)[2]$ we have $\iota(u)\in\alpha^\perp$ because \[ w_2(\iota(u),\alpha) = w_2(u,{\rm Nm}\,_{\tilde{T}/T}(\alpha)) = 0\,, \] and as both groups have cardinality $16$ they coincide. Hence this image is isomorphic to $\alpha^\perp / \langle\alpha\rangle$. In particular, the inverse image $L$ of $L_0$ in ${\rm Jac}\,(C)$ is a lagrangian subgroup of ${\rm Jac}\,(C)[2]$ containing $\alpha$. Conversely, let $L\subset {\rm Jac}\,(C)[2]$ be a lagrangian subgroup. We will say that the choices $f$, $\nu'$ made in the course of the construction are compatible with $L$ if $\alpha = \alpha(f)$ is in $L$ and $\nu'$ corresponds to $L/\langle \alpha \rangle$ as above. We can now formulate our main theorem, to which we shall give two proofs: \begin{theorem} \label{main} Let $C'$ be the result of the construction applied to a curve $C$ of genus $3$ compatibly with a lagrangian subgroup $L\subset {\rm Jac}\,(C)[2]$. Then there is an induced isomorphism $Jac(C)/L \sra{\sim} {\rm Jac}\,(C')$. In particular $C'$ is independent of the (compatible) choices made in the construction. \end{theorem} \begin{pf} The construction induces isogenies whose degrees are marked below: \[ \begin{array}{rcl} {\rm Jac}\,(C) & \simeq & {\rm Prym}\,(\tilde{T}/T) \sbth{\nu^\ast}{\longrightarrow}{2} {\rm Prym}\,(\nu\tilde{T}/\nu T) \sbth{\delta}{\longleftarrow}{4} {{\rm Prym}\,(\nu \tilde{T}'/\nu T')}\\ &\sbth{{\nu'}^{\ast}}{\longleftarrow}{4}& {\rm Prym}\,(\tilde{T}'/T') \simeq {\rm Jac}\,(C')\,. \end{array} \] Here the middle step $\delta$ is identified with the polarization map from an abelian variety of polarization type $211$ to its dual. As before we identify $\nu^\ast$ with the quotient by $\alpha$, so to construct our isomorphism ${\rm Jac}\,(C)/L \simeq {\rm Jac}\,(C')$ it would suffice to produce a natural map $\epsilon: {\rm Prym}\,(\nu\tilde{T}/\nu T) \rightarrow {\rm Prym}\,(\tilde{T}{}'/ T')$ whose kernel is the subgroup $L/\langle \alpha \rangle$ of ${\rm Prym}\,(\nu\tilde{T}/\nu T)$. One way to do this is to define $\epsilon$ as the dual map of $\nu'{}^\ast$, using $\delta$ to identify the dual of ${\rm Prym}\,(\nu\tilde{T}'/\nu T')$ with ${\rm Prym}\,(\nu\tilde{T}/\nu T)$, and using the principal polarization on ${\rm Prym}\,(\tilde{T}'/T')$ to view it as its own dual. Tracing through the definitions one verifies that ${\rm Ker}\,(\epsilon)$ is indeed $L/\langle \alpha \rangle$ as asserted. An alternative, and more geometric approach, is to show that the hyperelliptic case treated in Section~\ref{hyperelliptic} is a specialization of our present general construction. In fact the hyperelliptic case is obtained when the $4$-sheeted cover $f: C\rightarrow {\bbB P}^1$ happens to have $4$, rather than the generic $2$, double branch points. We shall see that this determines a preferred gluing $\nu'$. In going to this special case we have to note that the limits of the curves $\nu T$, $\nu \tilde{T}$ (which we continue to denote with the same symbols) are no longer non-singular: they are now only partial normalizations of $T$, $\tilde{T}$, and the map $\nu \tilde{T} \rightarrow \nu T$ now has $2$ points of Beauville type, at the singularities which were not normalized. The full normalizations, say $\nu \nu T$ and $\nu \nu \tilde{T}$, now have genera $0$ and $3$ respectively, and the resulting diagram \[\begin{array}{rcl} && \nu\nu\tilde{T} \\ && \mbox{$\scriptstyle{\pi}$}\!\downarrow \\ C && \nu\nu T \\ &\sf\!\!\searrow\;\!\swarrow\!\!\mbox{$\scriptstyle{g}$}' &\\ & {\bbB P}^1 & \end{array} \] clearly coincides with diagram~(\ref{hyp1}). Reagardless of the singularities of the intermediate curves, we will see that each of the abelian varieties in the diagram specializes to an abelian variety. In particular, the limit of ${\rm Prym}\,(\nu\tilde{T}/\nu T)$ is, by Lemma~\ref{polar}, a $4$\/-sheeted cover of ${\rm Prym}\,(\nu\nu\tilde{T}/\nu\nu T) \simeq {\rm Jac}\,(H)$, whose kernel is $L/\langle \alpha \rangle$. Below we will also identify the limit of ${\rm Prym}\,(\tilde{T}{}'/T')$ with ${\rm Jac}\,(H)$. This will produce the desired map $\epsilon$ in this special case, and hence in general. For this, we note that the bigonal data $\nu \tilde{T} \rightarrow \nu T \rightarrow {\bbB P}^1$ has $4$, $2$, $0$ and $2$ points of types $\ssub \seq \sslash \smin\smin $, $\stackrel{\subset}{\ssub} \sslash \ssub $, $\ssub \ssub \sslash \smin\smin $, and $\btf$ respectively, which turn into points of types $\stackrel{\subset}{\ssub} \sslash \ssub $, $\ssub \seq \sslash \smin\smin $, $\btf$ and $\ssub \ssub \sslash \smin\smin $, respectively, for $\nu \tilde{{T'}} \rightarrow \nu T' \rightarrow {\bbB P}^1$. To obtain $\tilde{{T'}} \rightarrow T'$ we need to pair the $6$ ramification points of $\nu \tilde{{T'}} \rightarrow \nu T'$. There are $15$ ways to do this, of which one is distinguished: each pair of Beauville branches gets paired, as do the remaining two ramification points. Let $T'_{h}$ be the intermediate object, obtained by gluing only the Beauville branches but not the remaining pair. It is a singular hyperelliptic curve of genus $3$, and is a partial normalization of $T'$ at the double point $p$. Let $\tilde{\Tph}$ be the corresponding $1$-point partial normalization of $\tilde{{T'}}$, of arithmetic genus $6$. To continue our construction, we need to identify the two $g_3^1$'s on $T'$: these two turn out to coincide, and the unique $g_3^1$ is in fact given by the $g_2^1$ on $T'_{h}$ plus a base point at the double point $p$. To see this we examine what happens to our general construction of the $g_3^1$ in this case. The unique quadric surface through the canonical model of $T'$ is now a quadric cone, with vertex at (the image of) $p$, because projection from $p$ gives the canonical image of the hyperelliptic $T'_{h}$, which is the double cover of a conic. Therefore the two rulings, hence the two $g_3^1$'s, coincide and have a base point at $p$, as asserted. At this point we need to turn the $g_3^1$ into a morphism, which requires us to blow up the point $p$. This results in a reducible trigonal curve $T'_{t} := T'_{h} \cup P$, where $P$ is a copy of ${\bbB P}^1$ intersecting $T'_{h}$ in the two inverse images $p_1$, $p_2$ of $p$ in $T'_{h}$. The trigonal map has degrees $2$ and $1$ respectively on the two components $T'_{h}$ and $P$. This curve is indeed a flat limit, in the family of triple covers of ${\bbB P}^1$, of the trigonal curves encountered in the non-hyperelliptic situation. The corresponding double cover $\tilde{\Tpt} \rightarrow T'_{t}$ is of Beauville type at all $4$ of the singular points (the two singularities of $T'_{h}$ plus $p_1$, $p_2$). Here $\tilde{\Tpt} = \tilde{\Tph} \cup \tilde{P}$, where $\tilde{P}$ is another copy of ${\bbB P}^1$, double cover of $P$ branched at the points glued to $p_1$ and $p_2$. Now that we have identified the trigonal data, we can complete the construction. By example~2.10(iii) of \cite{don}, or by inspection, we see that the result $C'$ of applying the trigonal construction to the reducible trigonal data $(\tilde{\Tph} \cup \tilde{P}) \rightarrow (T'_{h} \cup P) \rightarrow {\bbB P}^1$ is the $4$-sheeted cover of ${\bbB P}^1$ obtained by applying the bigonal construction to $\tilde{\Tph} \rightarrow T'_{h} \rightarrow {\bbB P}^1$. But since the bigonal construction is reversible this is nothing but the hyperelliptic curve $H= \nu\nu \tilde{T}$ which resulted from the construction of Section~\ref{hyperelliptic}, as claimed. The degeneration just described involves a flat family of abelian varieties, so the polarization type and the type of the kernel of the isogeny on jacobians remain constant. From the hyperelliptic case we now see that $L$ is the kernel of our isogeny in the general case. By Torelli's theorem, $C'$ is determined by its polarized jacobian, which is ${\rm Jac}\,(C)/L$. Hence $C'$ is indeed independent of the choices (compatible with $L$) made during the construction. This concludes the proof of Theorem~\ref{main}. \end{pf} We now make some further comments on the choices we made in the course of the construction. Starting on the left, we fix the curve C, the Lagrangian subgroup L and an element $\alpha \in L$. Our $g^1_4$\/'s $f: C \rightarrow {\bbB P}^1$ with $\alpha =\alpha(f)$ are determined by a divisor class in the intersection \[Z = Z_\alpha = \Theta_C \cap(\alpha + \Theta_C) \subset {\rm Pic}\,^2(C)\,.\] More accurately $Z$ parametrizes the family of $g^1_4$'s (with the specified $\alpha$), together with a marking of the two singular points $b_1$, $b_2$. Interchanging these two points gives an involution $i$ of $Z$ induced by the involution $x \mapsto x+ \alpha$ of ${\rm Pic}\,^2(C)$, and it is the quotient of $Z$ by $i$ which parametrizes the $g^1_4$'s alone. $Z$ is also invariant under the involution $j: x \mapsto K_C- x$ (where $K_C$ is the canonical class) and $i$ and $j$ commute. In addition, since $\Theta_C$ is an ample divisor, $Z$ is connected. Counting fixed points shows that the respective quotients of $Z$ by $i$, $j$, $k=ij$ have genera $4$, $1$, $4$, and that the common quotient $\overline{\overline{Z}}:=Z/\langle i,j\rangle$ has genus $1$. These quotients clearly have the following interpretations as parameter spaces: \begin{enumerate} \item $Z$ parametrizes the $g^1_4$\/'s $f:C\rightarrow {\bbB P}^1$ (equivalently, via the trigonal construction, towers of double covers $\tilde{T} \rightarrow T \rightarrow {\bbB P}^1$ of the indicated type), {\rm with} a choice of a double branch point $b_1$. \item $Z/i$ parametrizes the $g^1_4$\/'s $f:C\rightarrow {\bbB P}^1$ (equivalently, towers of double covers $\tilde{T} \rightarrow T\rightarrow{\bbB P}^1$ of the indicated type). \item $Z/j$ parametrizes the double covers $\tilde{T}\rightarrow T$ of the indicated type together with a singular point of $T$. \item $\overline{\overline{Z}}$ parametrizes the double covers $\tilde{T}\rightarrow T$ of the indicated type, hence it also parametrizes their normalizations, as well as the maps $\nu\pi': \nu\tilde{T}' \rightarrow \nu T'$. \end{enumerate} We will now discuss what choices we make when we perform the construction in reverse order, and how the choices from the two directions are related. Starting with the genus $3$ curve $C'$, we now assume given a lagrangian subgroup $L'$\/($={\rm Jac}\,(C)[2]/L)$ of ${\rm Jac}\,(C')[2]$, and a subgroup $G \subset L'$ of order 4 (which corresponds to $\alpha^\perp/L$). A marking of the three double branch points $b'_i$ of $f'$ is equivalent to a choice of a basis $\beta' = P'_2 + Q'_2 - P'_1 -Q'_1$ and $\gamma' = P'_3 + Q'_3 - P'_1 -Q'_1$ of $G$. Let $\Theta' \subset {\rm Pic}\,^2(C')$ be the theta divisor of $C'$, and for a class $u\in{\rm Jac}\,(C')$ let $\Theta'_u$ denote the translation of $\Theta'$ by $u$. Consider a line bundle $L$ in the intersection $S = \Theta \cap \Theta_{\beta'} \cap \Theta_{\gamma'}$. Since the canonical bundle is the only degree 4 bundle on $C'$ with $h^0>2$, there are only two possibilities: either $L^{\otimes 2}$ gives a $g^1_4$\/ $f':C'\rightarrow {\bbB P}^1$ with three marked double branch points $b'_i\in {\bbB P}^1$, $i=1,\dots,3$, or else $L$ must be a theta characteristic on $C'$. We claim the following:\\ \noindent (1) $S$ consists of six points \\ \noindent (2) $S$ is closed under $v\rightarrow K_{C'} - v$.\\ \noindent (3) Four of the points of $S$ are theta characteristics, and two are not.\\ \begin{pf} (1) holds because $6=g!$. For (2), suppose that $f',f'':C' \rightarrow {\bbB P}^1$ correspond to $2v$, $2K_{C'} - 2v$ respectively. Then for each double ramification point $P'_i$, $Q'_i$ of $f'$ we get a unique double ramification point $P''_i$, $Q''_i$ for $f''$ by imposing the condition $P''_i + Q''_i + P'_i + Q'_i = K_{C'}$. For (3) , one checks that there is a unique coset $G'$ of $G$ in the set of odd theta characteristics on $C'$: indeed, in coordinates we may take the set of theta characteristics to be $V = ({\bbB Z}/2{\bbB Z})^6$ with coordinates $x_1,\dots,x_6$, and we may suppose that $h^0(C',O_{C'}(x))$ mod 2 for $x = (x_1,\dots,x_6)$ is given by $q(x) = x_1x_2 + x_3x_4 + x_5x_6$. Also we may simultaneously identify ${\rm Jac}\,(C')[2]$ with $V$, with the Weil pairing given by $w_2(x,y) = q(x+y)-q(x) -q(y)$. Without loss of generality we can also take $\beta' = e_1$ and $\gamma' = e_3$, with $e_i$ the standard $i$\/th unit vector. Then $G' = \{(a,0,b,0,1,1)\}$. Part (3) is now clear: $G'$ is contained in $S$, and no other theta characteristics appear in $S$. This establishes our claim. \end{pf} We can now describe all the choices made when we start from the right side. Our data $C',L',G$ determines a complementary pair of maps $f',f"$. These determine the data $\tilde{T}'\sra{\pi'} T'$ uniquely (the two resulting maps $g',g"$ are the usual two $g^1_3$'s on the genus 4 curve $T'$). The normalization $\nu\pi': \nu\tilde{T}' \rightarrow \nu T'$ is therefore also uniquely determined. So the {\em only} choice made is that of $h'$, given by an arbitrary point of $ Pic^2(\nu T') \approx \nu T'$. Comparing with what we found starting from the left, we discover that $\nu T'$ is precisely identified with the double quotient $\overline{\overline{Z}}$. \section{The case of genus$\geq 4$.} One might try to generalize our construction to higher genus by finding, for a generic curve $C$ of genus $g$, a correspondence with another generic curve $C'$ of genus $g$ such that ${\rm Jac}\,(C')\simeq{\rm Jac}\,(C)/K$, with $K$ a lagrangian subgroup of ${\rm Jac}\,(C)[2]$. We shall show that this is not possible. \begin{theorem} Let $K$ be a lagrangian subgroup in ${\rm Jac}\,(C)[p]$, where $C$ is a generic curve of genus $g\geq 4$ and $p$ is a prime. Then ${\rm Jac}\,(C)/K$ with its induced principal polarization is not a jacobian. \end{theorem} \begin{pf} Let ${\cal T}$, ${\cal S}$, ${\cal M}$ and ${\cal A}$ denote respectively the Teichm\"uller space, the Siegel space, the moduli space of curves and the moduli space of principally polarized abelian varieties, all of genus $g$. The mapping class group $M=M(g)$ acts on ${\cal T}$ with quotient ${\cal M}$ and the modular group $\Gamma=\Sp(2g,{\bbB Z})$ acts on ${\cal S}$ with quotient ${\cal A}$. Moreover $\Gamma$ is naturally a quotient of $M$, because $M$ acts on symplectic bases for $H_1(C,{\bbB Z})$ through its action on $\pi_1(C)$, and the period map $\tau:{\cal T}\rightarrow{\cal S}$ is $M$-equivariant for these actions. Passing to the quotient, we get Torelli's map $\ov{\tau}:{\cal M}\rightarrow{\cal A}$, which is injective (Torelli's theorem) and exhibits ${\cal M}$ as a locally closed subvariety of ${\cal A}$. Since ${\cal T}$ is irreducible it follows that the {\em Torelli space} $\ov{{\cal T}}=\tau({\cal T})$ is a locally closed irreducible analytic subvariety of ${\cal S}$. \[ \begin{array}{rcl} {\cal T} & \slra{\tau} & {\cal S} \supset \overline{{{\cal T}}}=\tau ({\cal T}) \\ M \downarrow && \downarrow \Gamma \\ {\cal M} & \slra{\overline{{\tau}}} & {\cal A} \end{array} \] Let $W$ be the finite cover of ${\cal S}$ obtained by taking over each marked abelian variety $A$ the lagrangian subgroups of $A[p]$. Since $W$ is unramified over the contractible space ${\cal S}$, it is in fact a union of copies of ${\cal S}$. Our generic isogeny ${\rm Jac}\,(C)\rightarrow{\rm Jac}\,(C)/K$ translates to the following data. The curve $C$ lives over an open subset of ${\cal M}$, hence of $\ov{{\cal T}}$. The subgroup $K$ corresponds to a sheet of $W_{|\ov{{\cal T}}}$. Therefore our isogeny extends to the quotient map by the subgroup, still denoted $K$, corresponding to the ``same'' sheet over all of ${\cal S}$. Now recall that ${\cal S}$ is the space of symmetric $g\times g$ complex matrices $\Omega$ with positive imaginary part, and the abelian variety over $\Omega$ is $A_\Omega = {\bbB C}^g/({\bbB Z}^g+\Omega{\bbB Z}^g)$. Since monodromy (i.e. $\Gamma$) acts transitively on the lagrangian subgroups of $A_\Omega$, we may take $K = (\frac{1}{p}{\bbB Z}/{\bbB Z})^g$ for convenience. Then $A_\Omega/K \simeq A_{p\Omega} = A_{s\Omega}$, with \[ s = \matr{pI_{g\times g}}{0}{0}{I_{g\times g}} \in\Sp(2g,{\bbB R})\,.\] If ${\rm Jac}\,(C)/K$, with its principal polarization, were a jacobian, it would follow that the Torelli locus $\ov{{\cal T}}$ was invariant under the subgroup $\Delta$ of $\Sp(2g,{\bbB R})$ generated by $\Gamma$ and by $s$. We claim that $\Delta$ is dense in $\Sp(2g,{\bbB R})$. Indeed, consider the subgroup $N_+$ of $\Sp(2g, {\bbB R})$ consisting of the matrices $n(x) = \matr{I_{g\times g}}{x}{0}{I_{g\times g}}$, where $x$ runs over the real symmetric $g\times g$ matrices. Then $\Delta$ contains $s^i n(x) s^{-i}=n(p^{-i}x)$ for all integral symmetric matrices $x$ and integers $i$. These are dense in $N_+$, and $\Delta$ likewise contains a dense subgroup of $N_-={}^t N_+$. It is well-known (and easy) that $N_+$ and $N_-$ generate $\Sp(2g,{\bbB R})$, so $\Delta$ is indeed dense in $\Sp(2g, {\bbB R})$. Therefore, under our assumption, $\ov{{\cal T}}$ would be dense in ${\cal S}$ (in the complex topology), so that ${\cal M}$ would be dense in ${\cal A}$. This is a contradiction when $g>3$, because for dimension reasons ${\cal M}$ is not dense in $A$ even for the Zariski topology then, and the theorem follows. \end{pf}
1998-01-01T01:35:47
9712
alg-geom/9712035
en
https://arxiv.org/abs/alg-geom/9712035
[ "alg-geom", "math.AG" ]
alg-geom/9712035
Jun Li
Jun Li and Gang Tian
Comparison of the algebraic and the symplectic Gromov-Witten invariants
45 pages, Latex
null
null
null
null
We show that the algebraic and the symplectic GW-inivariants of smooth projective varieties are equivalent.
[ { "version": "v1", "created": "Thu, 1 Jan 1998 00:35:47 GMT" } ]
2007-05-23T00:00:00
[ [ "Li", "Jun", "" ], [ "Tian", "Gang", "" ] ]
alg-geom
\section{Introduction} As Witten suggested in [W1], [W2], the GW-invariants for a symplectic manifold $X$ are multi-linear maps \begin{equation} \gamma_{A,g,n}^X: H^{\ast}(X;{\mathbb Q})^{\times n}\times H^{\ast}(\overline{\M}_{g,n};{\mathbb Q}) \lra {\mathbb Q}, \label{eq:0.1} \end{equation} where $A\in H_2(X,{\mathbb Z})$ is any homology class, $n$, $g$ are two non-negative integers, and $\overline{\M}_{g,n}$ is the Deligne-Mumford compactification of $\M_{g,n}$, the space of smooth $n$-pointed genus $g$ curves. The basic idea of defining these invariants is to enumerate holomorphic maps from Riemann surfaces to the manifolds. To illustrate this, we let $X$ be a smooth projective manifold and form the moduli space $\M_{g,n}(X,A)$ of all holomorphic maps $f\!:\!\Sigma\to X$ from smooth $n$-pointed Riemann surfaces $(\Sigma;x_1,\ldots,x_n)$ to $X$ such that $f_{\ast}([\Sigma])=A$. $\M_{g,n}(X,A)$ is a quasi-projective scheme and its expected dimension can be calculated using the Riemann-Roch theorem. We will further elaborate the notion of expected dimension later, and for the moment we will denote it by $r_{\rm exp }$. Note that it depends implicitly on the choice of $X$, $A$, $g$ and $n$. When $r_{\rm exp }=0$, then $\M_{g,n}(X,A)$ is expected to be discrete. If $\M_{g,n}(X,A)$ is discrete, then the degree of $\M_{g,n}(X,A)$, considered as a $0$-cycle, is a GW-invariant of $X$. We remark that we have and will ignore the issue of non-trivial automorphism groups of maps in $\M_{g,n}(X,A)$ in the introduction. When $r_{\rm exp }>0$, then $\M_{g,n}(X,A)$ is expected to have pure dimension $r_{\rm exp }$. If it does, then we pick $n$ subvarieties of $X$, say $V_1,\ldots,V_n$, so that their total codimension is $r_{\rm exp }$. We then form a subscheme of $\M_{g,n}(X,A)$ consisting of maps $f$ so that $f(x_i)\in V_i$. This subscheme is expected to be discrete. It it does, then its degree is the GW-invariant of $X$. Put them together, we can define the GW-invariants $\gamma_{A,g,n}^X$ of $X$. This is similar to construction of the Donaldson polynomial invariants for 4-manifolds. Here are the two big {\sl ifs} in carry out this program are {\bf Question I}: Whether the moduli scheme $\M_{g,n}(X,A)$ has pure dimension $r_{\rm exp }$. {\bf Question II}: Whether the subschemes of $\M_{g,n}(X,A)$ that satisfy certain incidence relations have the expected dimensions. Similar to Donaldson polynomial invariants, the affirmative answer to the above two questions are in general not guaranteed. One approach to overcome this difficulty, beginning with Donaldson's invariants of 4-manifolds, is to ``deform'' the moduli problems and hope that the answers to the ``deformed'' moduli problems are affirmative. In the case of GW-invariants, one can deform the complex structure of the smooth variety $X$ to not necessary integrable almost complex structure $J$ and study the same moduli problem by replacing holomorphic maps with pseudo-holomorphic maps. This was investigated by Gromov in [Gr], Ruan [Ru], in which he constructed certain GW-invariants of rational type for semi-positive symplectic manifolds. The first mathematical theory of GW-invariants came from the work of Ruan and the second author, in which they found that the right set up of GW-invariants for semi-positive manifolds can be provided by using the moduli of maps satisfying non-homogeneous Cauchy-Riemann equations. In this set up, they constructed the GW-invariants of all semi-positive symplectic manifolds and proved fundemantal properties of these invariants. All Fano-manifolds and Calabi-Yau manifolds are special examples of semi-positive symplectic manifolds. Also any symplectic manifold of complex dimension less than $4$ is semi-positive. Attempts to push this to cover general symplectic manifolds so far have failed. New approaches are needed in order to get a hold on the GW-invariants of general varieties (or symplectic manifolds). The first step is to convert the problem of counting mappings, which essentially is homology in nature, into the frame work of cohomology theory of the moduli problem. More precisely, we first compactify the moduli space $\M_{g,n}(X,A)$ to, say, $\overline{\M}_{g,n}(X,A)$. We require that the obvious evaluation map $$e: \M_{g,n}(X,A)\lra X^n $$ that sends $(f;\Sigma;x_1,\ldots,x_n)$ to $(f(x_1),\ldots,f(x_n))$ extends to $$\bar e: \overline{\M}_{g,n}(X,A)\lra X^n. $$ We further require that if $\M_{g,n}(X,A)$ has pure dimension $r_{\rm exp }$, then $\overline{\M}_{g,n}(X,A)$ supports a fundamental class $$[\overline{\M}_{g,n}(X,A)]\in H_{2r_{\rm exp }}(\overline{\M}_{g,n}(X,A);{\mathbb Q}). $$ Then the GW-invariants of $X$ are multi-linear maps \begin{equation} \gamma^X_{A,g,n}: H^{\ast}(X)^{\times n}\times H^{\ast}(\overline{\M}_{g,n})\lra {\mathbb Q} \label{eq:0.4} \end{equation} that send $(\alpha,\beta)$ to $$\gamma^X_{A,g,n}(\alpha,\beta)=\int_{[\overline{\M}_{g,n}(X,A)]}{\bar e}^{\ast}(\alpha) \cup\pi^{\ast}(\beta). $$ where $\pi\!:\!\overline{\M}_{g,n}(X,A)\to\overline{\M}_{g,n}$ is the forgetful map. Note that in such cases the GW-invariants are defined without reference to the answer to {\sl question II}. Even when the answer to {\sl question I} is negative, we can still define the GW-invariants if a virtual moduli cycle $$[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}\in H_{2r_{\rm exp }}(\overline{\M}_{g,n}(X,A);{\mathbb Q}) $$ can be found that function as the fundamental cycle $[\overline{\M}_{g,n}(X,A)]$ should the dimension of $\M_{g,n}(X,A)$ is $r_{\rm exp }$. In this case, we simply define $\gamma^X_{A,g,n}$ as before with $[\overline{\M}_{g,n}(X,A)]$ replaced by $[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}$. The standard compactification of $\M_{g,n}(X,A)$ is the moduli space of stable morphisms from $n$-pointed genus $g$ curves, possibly nodal, to $X$ of the prescribed fundamental class. This was first studied for pseudo-holomorphic maps by T. Parker and J. Wolfson \cite{PW} and in algebraic geometry by Kontsevich \cite{Ko}. Because points of the compactification $\overline{\M}_{g,n}(X,A)$ are maps $f$ whose domains have $n$-marked points $x_1,\ldots,x_n$, the evaluation map $e$ extends canonically to $\bar e$ that sends such map $f$ to $(f(x_1),\ldots,f(x_n))$. The virtual moduli cycles $[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}$ for projective variety $X$ were first constructed by the authors. Their idea is to construct a virtual normal cone embedded in a vector bundle based on the obstruction theory of stable morphisms \cite{LT1}. An alternative construction of such cones was achieved by Behrend and Fantechi \cite{BF, Be}. For general symplectic manifolds, such virtual moduli cycles were constructed by the authors, and independently, by Fukaya and Ono \cite{FO, LT2}. Shortly after them, B. Siebert [Si] and later, Y. Ruan [Ru2] gave different constructions of such virtual moduli cycles. Both Siebert and Ruan's approach needs to construct global, finite-dimensional resolutions of so called cokernel bundles (cf. \cite{Si} and \cite{Ru2}, Appendix). However, one question remains to be investigated. Namely, if $X$ is a smooth projective variety then on one hand we have the algebraically constructed GW-invariants, and on the other hand, by viewing $X$ as a symplectic manifold using the K\"ahler form on $X$, we have the GW-invariants constructed using analytic method. These two approaches are drastically different. One may expect, although far from clear, that for smooth projective varieties the algebraic GW-invariants and their symplectic counterparts are identical. The main goal of this paper is to prove what was expected is indeed true. \begin{theo} Let $X$ be any smooth projective variety with a K\"ahler form $\omega$. Then the algebraically constructed GW-invariants of $X$ coincide with the analytically constructed GW-invariants of the symplectic manifold $(X^{{\rm top}},\omega)$. \end{theo} This result was first announced in [LT2]. Its proof was outlined in [LT3]. During the preparation of the paper, we learned from B. Siebert that he was able to prove a similar result. We now outline the proof of our Comparison Theorem. We begin with a few words on the algebraic construction of the virtual moduli cycle. Let $w\in\overline{\M}_{g,n}(X,A)$ be any point associated to the stable morphism $f\!:\! \Sigma\to X$. It follows from the deformation theory of stable morphisms that there is a complex ${\mathcal C}_w$, canonical up to quasi-isomorphisms, such that its first cohomology ${\mathcal H}^1({\mathcal C}_w)$ is the space of the first order deformations of the map $w$, and its second cohomology ${\mathcal H}^2({\mathcal C}_w)$ is the obstruction space to deformations of the map $w$. Let $\varphi_w$ be a Kuranishi map of the obstruction theory of $w$. Note that $\varphi_w$ is the germ of a holomorphic map from a neighborhood of the origin $o\in{\mathbb C}^{m_1}$ to ${\mathbb C}^{m_2}$, where $m_i=\dim{\mathcal H}^i({\mathcal C}_w)$. Let $\hat o$ be the formal completion of ${\mathbb C}^{m_1}$ along $o$ and let $\hat w$ be the subscheme of $\hat o$ defined by the vanishing of $\varphi_w$. Note that $\hat w$ is isomorphic to the formal completion of $\overline{\M}_{g,n}(X,A)$ along $w$ (Here as before we will ignore the issue of non-trivial automorphism groups of maps in $\overline{\M}_{g,n}(X,A)$). This says that ``near'' $w$, the scheme $\overline{\M}_{g,n}(X,A)$ is a ``subset'' of ${\mathbb C}^{m_1}$ defined by the vanishing of $m_2$-equations. Henceforth, it these equations are in general position, them $\dim\hat w=m_1-m_2$, which is the expected dimension $r_{\rm exp }$ we mentioned before. The case where $\overline{\M}_{g,n}(X,A)$ has dimension bigger than $r_{\rm exp }$ is exactly when the vanishing locus of these $m_2$- equations in $\varphi_w$ do not meet properly near $o$. Following the excess intersection theory of Fulton and MacPherson \cite{Fu}, the ``correct'' cycle should come from first constructing the normal cone $C_{\hat w/\hat o}$ to $\hat w$ in $\hat o$, which is canonically a subcone of $\hat w\times {\mathbb C}^{m_2}$, and then intersect the cone with the zero section of $\hat w\times{\mathbb C}^{m_2}\to\hat w$. The next step is to patch these cones together to form a global cone over $\M_{g,n}(X,A)$. The main difficulty in doing so comes from the fact that the dimensions ${\mathcal H}^2({\mathcal C}_w)$ can and do vary as $w$ vary, only $\dim{\mathcal H}^1({\mathcal C}_w)-\dim{\mathcal H}^2({\mathcal C}_w)$ is a topological number. This makes the cones $C_{\hat w/\hat o}$ to sit inside bundles of varying ranks. To overcome this difficulty, the authors came with the idea of finding a global ${\mathbb Q}$-vector bundle $E_2$ over $\overline{\M}_{g,n}(X,A)$ and a subcone $N$ of $E_2$ such that near fibers over $w$, the cone $N$ is a fattening of the cone $C_{\hat w/\hat o}$ (See section 3 or \cite{LT1} for more details). In the end, we let $j$ be the zero section of $E_2$ and let $j^{\ast}$ be the Gysin map $$ A_{\ast} E_2\lra A_{\ast}\overline{\M}_{g,n}(X,A), $$ where $A_{\ast}$ denote the Chow-cohomology group (see \cite{Fu}). Then the algebraic virtual moduli cycle is $$[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}=j^{\ast}([N])\in A_{r_{\rm exp }}\overline{\M}_{g,n}(X,A). $$ Now let us recall briefly the analytic construction of GW-invariants of symplectic manifolds. Let $(X,\omega)$ be any smooth symplectic manifold with $J$ a tamed almost complex structure. For $A$, $g$ and $n$ as before, we can form the moduli space of $J$-holomorphic maps $f\!:\! \Sigma\to X$ where $\Sigma$ are $n$-pointed smooth Riemann surfaces such that $f_{\ast}([\Sigma])=A$. We denote this space by $\M_{g,n}(X,A)^{J}$. It is a finite dimensional topological space. As before, we compactify it to include all $J$-holomorphic maps whose domains are possibly with nodal singularities. We denote the compactified space by $\overline{\M}_{g,n}(X,A)^{J}$. To proceed, we will embed $\overline{\M}_{g,n}(X,A)^{J}$ inside an ambient space ${\mathbf B}$ and realize it as the vanishing locus of a section of a ``vector bundle''. Without being precise, the space ${\mathbf B}$ is the space of all {\sl smooth} maps $f \in{\mathbf B}$ from possibly nodal $n$-pointed Riemann surfaces to $X$, the fiber of the bundle over $f$ are all $(0,1)$-forms over domain$(f)$ with values in $f^{\ast} T_X$ and the section is the one that sends $f$ to $\bar\partial f$. We denote this bundle by ${\mathbf E}$ and the section by $\Phi$. Clearly, $\Phi^{-1}(0)$ is homeomorphic to $\overline{\M}_{g,n}(X,A)^{J}$. Defining the GW-invariants of $(X,\omega)$ is essentially about constructing the Euler class of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$. This does not make much sense since ${\mathbf B}$ is an infinite dimensional topological space. Although at each $w\in\Phi^{-1}(0)$ the formal differential $d\Phi(w)\!:\! T_w{\mathbf B}\to{\mathbf E}_w$ is Fredholm, which has real index $2r_{\rm exp }$, the conventional perturbation scheme does not apply directly since near maps in ${\mathbf B}$ whose domains are singular the space ${\mathbf B}$ is not smooth and ${\mathbf E}$ does not admit local trivializations. To overcome this difficulty, the authors introduced the notion of weakly ${\mathbb Q}$-Fredholm bundles, and showed in \cite{LT2} that $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ is a weakly ${\mathbb Q}$-Fredholm bundle and that any weakly ${\mathbb Q}$-Fredholm bundle admits an Euler class. Let $$ e[\Phi\!:\!{\mathbf B}\to{\mathbf E}]\in H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q}) $$ be the the Euler class of $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$. Since the evaluation map of $\M_{g,n}(X,A)^{J}$ extends to an evaluation map ${\mathbf e}\!:\!{\mathbf B}\to X^n$, the Euler class, which will also be referred to as the symplectic virtual cycle of $\overline{\M}_{g,n}(X,A)^{J}$, defines a multi-linear map $\gamma_{A,g,n}^{X,J}$ as in \eqref{eq:0.1}. We will review the notion of weakly smooth Fredholm bundles in section 2. Here to say the least, $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ is weakly Fredholm means that near each point of $\Phi^{-1}(0)$ we can find a finite rank subbundle $V$ of ${\mathbf E}$ such that $W=\Phi^{-1}(V)$ is a smooth finite dimensional manifold, $V|_W$ is a smooth vector bundle and the lift $\phi\!:\! W\to V|_W$ of $\Phi$ is smooth. (Note that the rank of $V$ may vary but $\dim_{{\mathbb R}}W-\rank_{{\mathbb R}} V=2r_{\rm exp }$). For such finite models $[\phi\!:\! W\to V|_W]$, which are called weakly smooth approximations, we can perturb $\phi$ slightly to obtain $\phi^{\prime}$ so that $\phi^{\prime-1}(0)$ are smooth manifolds in $W$. To construct the Euler class, we first cover a neighborhood of $\Phi^{-1}(0)$ in ${\mathbf B}$ by finitely many such approximations that satisfy certain compatibility condition. We then perturb each section in the approximation and obtain a collection of locally closed ${\mathbb Q}$-submanifolds of ${\mathbf B}$ of dimension $2r_{\rm exp }$. By imposing certain compatibility condition on the perturbations, this collection of ${\mathbb Q}$-submanfolds patch together to form a $2r_{\rm exp }$-dimensional cycle in ${\mathbf B}$, which represents a homology class in $H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q})$. This is the Euler class of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$. Now we assume that $X$ is a smooth projective variety and $\omega$ is a K\"ahler form of $X$. Let $J$ be the complex structure of $X$. Then $\overline{\M}_{g,n}(X,A)$ is homeomorphic to $\overline{\M}_{g,n}(X,A)^{J}$. Hence the two GW-invariants $\gamma_{A,g,n}^X$ and $\gamma_{A,g,n}^{X,J}$ are identical if the homology classes $[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}$ and $e[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ will be identical. Here we view $[\overline{\M}_{g,n}(X,A)]^{{\rm vir}}$ as a class in $H_{\ast}({\mathbf B};{\mathbb Q})$ using $$\overline{\M}_{g,n}(X,A)\sim_{{\rm homeo}}\overline{\M}_{g,n}(X,A)^{J}\subset{\mathbf B}. $$ To illustrate why these two classes are equal, let us first look at the following simple model. Let $Z$ be a compact smooth variety and let $E$ be a holomorphic vector bundle over $Z$ with a holomorphic section $s$. There are two ways to construct the Euler classes of $E$. One is to perturb $s$ to a smooth section $r$ so that the graph of $r$ is transversal to the zero section of $E$, and then define the Euler class of $E$ to be the homology class in $H_{\ast}(Z;{\mathbb Q})$ of $r^{-1}(0)$. This is the topological construction of the Euler class of $E$. The algebraic construction is as follows. Let $t$ be a large scalar and let $\Gamma_{ts}$ be the graph of $ts$ in the total space of $E$. Since $s$ is an algebraic section, it follows that the limit $$\Gamma_{\infty s} =\lim_{t\to\infty}\Gamma_{ts} $$ is a complex dimension $\dim Z$ cycle supported on union of subvarieties of $E$. We then let $r$ be a smooth section of $E$ in general position and let $\Gamma_{\infty s}\cap \Gamma_r$ be their intersection. Its image in $Z$ defines a homology class, which is the image of the Gysin map $j^{\ast}([C])$, where $j$ is the zero section of $E$. The reason that $$e(E)=[r^{-1}(0)]=j^{\ast}([\Gamma_{\infty s}])\in H_{\ast}(Z;{\mathbb Q}) $$ is that if we choose $r$ to be in general position, then $$[r^{-1}(0)]=[\Gamma_r\cap\Gamma_0]=[\Gamma_r\cap\Gamma_s], $$ and the family $\{\Gamma_{ts}\cap\Gamma_r\}_{t\in[1,\infty]}$ forms a homotopy of the cycles $\Gamma_s\cap\Gamma_r$ and $\Gamma_{\infty s}\cap \Gamma_r$. One important remark is that the cone $\Gamma_{\infty s}$ is contained in $E|_{s^{-1}(0)}$ and the intersection of $\Gamma_{\infty s}$ and $\Gamma_r$ in $E$ is the same as their intersection in $E|_{s^{-1}(0)}$. Back to our construction of GW-invariants, the analytic construction of GW-invariants, which was based on perturbations of sections in the finite models (weakly smooth approximations) $[\phi\!:\! W\to V|_W]$, is clearly a generalization of the topological construction of the Euler classes of vector bundles. As to the algebraic construction of GW-invariants, it is based on a cone in a ${\mathbb Q}$-vector bundle over $\overline{\M}_{g,n}(X,A)$. Comparing to the algebraic construction of the Euler class of $E\to Z$, what is missing is the section $s$ and that the cone is the limit of the graphs of the dilations of $s$. Following \cite{LT1}, the cone $\Gamma_{\infty s}$ only relies on the restriction of $s$ to an ``infinitesimal'' neighborhood of $s^{-1}(0)$ in $Z$, and can also be reconstructed using the Kuranishi maps of the obstruction theory to deformations of points in $s^{-1}(0)$ induced by the defining equation $s=0$. Along this line, to each finite model $[\phi\!:\! W\to V|_W]$ we can form a cone $\Gamma_{\infty \phi}=\lim \Gamma_{t\phi}$ in $V|_{\phi^{-1}(0)}$. Hence to show that the two virtual moduli cycles coincide, it suffices to establish a relation, similar to quasi-isomorphism of complexes, between the cone $N$ constructed based the obstruction theory of $\overline{\M}_{g,n}(X,A)$ and the collection $\{\Gamma_{\infty\phi}\}$. In the end, this is reduced to showing that the obstruction theory to deformations of maps in $\overline{\M}_{g,n}(X,A)$ is identical to the obstruction theory to deformations of elements in $\phi^{-1}(0)$ induced by the defining equation $\phi$. This identification of two obstruction theories follows from the canonical isomorphism of the C\v{e}ch cohomology and the Dolbeault cohomology of vector bundles. The layout of the paper is as follows. In section two, we will recall the analytic construction of the GW-invariants of symplectic manifolds. We will construct the Euler class of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ in details using the weakly smooth approximations constructed in \cite{LT2}. In section three, we will construct a collection of holomorphic weakly smooth approximations for projective manifolds. The proof of the Comparisom Theorem will occupy the last section of this paper. \section{Symplectic construction of GW invariants} The goal of this section is to review the symplectic construction of the GW-invariants of algebraic varieties. We will emphasize on those parts that are relevant to our proof of the Comparison Theorem. In this section,we will work mainly with real manifolds and will use the standard notation in real differential geometry. We begin with the symplectic construction of GW-invariants. Let $X$ be a smooth complex projective variety, and let $A\in H_2(X,{\mathbb Z})$ and let $g,\, n\in{\mathbb Z}$ be fixed once and for all. We recall the notion of stable $C^l$-maps \cite[Definition 2.1]{LT2}. \begin{defi} \label{1.1} An $n$-pointed stable map is a collection $(f;\Sigma;x_1\ldots,x_n)$ satisfying the following property: First, $(\Sigma;x_1,\ldots,x_n)$ is an $n$-pointed connected prestable complex curve with normal crossing singularity; Secondly, $f\!:\!\Sigma\to X$ is continuous, and the composite $f\circ\pi$ is smooth, where $\pi\!:\!\tilde\Sigma\to\Sigma$ is the normalization of $\Sigma$; And thirdly, if we let $S\subset\Sigma$ be the union of singular locus of $\Sigma$ with its marked points, then any rational component $R\subset\tilde{\Sigma}$ satisfying $(f\circ \pi)_{\ast}([R])=0\in H_2(X,{\mathbb Z})$ must contains at least three points in $\pi^{-1}(S)$. \end{defi} For convenience, we will abbreviate $(f;\Sigma;x_1,\ldots,x_n)$ to $(f;\Sigma;\{x_i\})$. Later, we will use ${\mathcal C}$ to denote an arbitrary stable map and use $f_{{\mathcal C}}$ and $\Sigma_{{\mathcal C}}$ to denote its corresponding mapping and domain. Two stable maps $(f;\Sigma;\{x_i\})$ and $(f^{\prime};\Sigma^{\prime};\{x^{\prime}_i\})$ are said to be equivalent if there is an isomorphism $\rho\!:\!\Sigma\to\Sigma^{\prime}$ such that $f^{\prime}\circ\rho=f$ and $x_i^{\prime}=\rho(x_i)$. When $(f;\Sigma;\{x_i\})\equiv(f^{\prime};\Sigma^{\prime};\{x^{\prime}_i\})$, such a $\rho$ is called an automorphism of $(f;\Sigma;\{x_i\})$. We let ${\mathbf B}$ be the space of equivalence classes $[{\mathcal C}]$ of $C^l$-stable maps ${\mathcal C}$ such that the arithmetic genus of $\Sigma_{{\mathcal C}}$ is $g$ and $f_{{\mathcal C}\ast}([\Sigma])=A\in H_2(X;{\mathbb Z})$. Note that ${\mathbf B}$ was denoted by $\bar{{\mathcal F}}^l_A(X,g,n)$ in~\cite{LT2}. Over ${\mathbf B}$ there is a generalized bundle ${\mathbf E}$ defined as follows. Let ${\mathcal C}$ be any stable map and let $\tilde f_{{\mathcal C}}\!:\!\tilde \Sigma_{{\mathcal C}}\to X$ be the composite of $f_{{\mathcal C}}$ with $\pi\!:\!\tilde\Sigma_{{\mathcal C}}\to\Sigma_{{\mathcal C}}$. We define $\Lambda^{0,1}_{{\mathcal C}}$ to be the space of all $C^{l-1}$-smooth sections of $(0,1)$-forms of $\tilde\Sigma$ with values in ${\tilde f}^{\ast} TX$. Assume ${\mathcal C}$ and ${\mathcal C}^{\prime}$ are two equivalent stable maps with $\rho\!:\!\Sigma_{{\mathcal C}} \to\Sigma_{{\mathcal C}^{\prime}}$ the associated isomorphism, then there is a canonical isomorphism $\Lambda^{0,1}_{{\mathcal C}^{\prime}}\cong\Lambda^{0,1}_{{\mathcal C}}$. We let $\Lambda^{0,1}_{[{\mathcal C}]}$ be $\Lambda^{0,1}_{{\mathcal C}}/\Aut({\mathcal C})$. Then the union $${\mathbf E}=\bigcup_{[{\mathcal C}]\in{\mathbf B}} \Lambda^{0,1}_{[{\mathcal C}]} $$ is a fibration over ${\mathbf B}$ whose fibers are finite quotients of infinite dimensional linear spaces. There is a natural section $$\Phi: {\mathbf B}\lra{\mathbf E} $$ defined as follows. For any stable map ${\mathcal C}$, we define $\Phi({\mathcal C})$ to be the image of $\bar\partial f_{[{\mathcal C}]}\in\Lambda^{0,1}_{{\mathcal C}}$ in $\Lambda^{0,1}_{[{\mathcal C}]}$. Obviously, for ${\mathcal C}\sim{\mathcal C}^{\prime}$ we have $\Phi({\mathcal C})=\Phi({\mathcal C}^{\prime})$. Thus $\Phi$ descends to a map ${\mathbf B}\to{\mathbf E}$, which we still denote by $\Phi$. >From now on, we will denote by $\M_{g,n}(X,A)$ the moduli scheme of stable moprhisms $f\!:\! C\to X$ with $n$-marked points such that $C$ is (possibly with nodal singularities) has arithmetic genus $g$ with $f_{\ast}([C])=A$. \begin{lemm} \label{1.2} The vanishing locus of $\Phi$ is canonically homeomorphic to the underlying topological space of $\M_{g,n}(X,A)$. \end{lemm} \begin{proof} A stable $C^l$-stable map ${{\mathcal C}}$ in ${\mathbf B}$ belongs to the vanishing locus of $\Phi$ if and only if $f_{{\mathcal C}}$ is holomorphic. Since $\Sigma_{{\mathcal C}}$ is compact, ${\mathcal C}$ is the underlying analytic map of a stable morphism. Hence there is a canonical map $\Phi^{-1}(0)\to \M_{g,n}(X,A)^{\rm top}$, which is one-to-one and onto. This proves the lemma. \end{proof} To discuss the smoothness of $\Phi$, we need the local uniformizing charts of $\Phi\!:\! {\mathbf B}\to{\mathbf E}$ near $\Phi^{-1}(0)$. Let $w\in{\mathbf B}$ be any point represented by the stable map $(f_0;\Sigma_0;\{x_i\})$ with automorphism group $G_{w}$. We pick integers $r_1, r_2>0$ and smooth ample divisors $H_1,\ldots,H_{r_2}$ with $[H_i]\cdot[A]=r_1$ such that all $f_0^{-1}(H_i)$ are contained in the smooth locus of $\Sigma_0$ and that for any $x\in f_0^{-1}(H_i)$ we have \begin{equation} \image(df_0(x))+T_{f(x)}H_i= T_{f(x)}X. \label{eq:1.1} \end{equation} Now let $U\subset {\mathbf B}$ be a sufficiently small neighborhood of $w\in{\mathbf B}$ and let $\tilde U$ be the collection of all $({\mathcal C}; z_{n+1},\ldots,z_{n+r_1r_2})$ such that ${\mathcal C}\in U$ and the $z_i$'s is a collection of smooth points of $\Sigma_{{\mathcal C}}$ such that for each $1\leq j\leq r_2$ the subcollection $(z_{n+(j-1)r_1+1},\ldots,z_{n+jr_1})$ contains distinct points and is exactly $f_{{\mathcal C}}^{-1}(H_j)$. Note that we do not require $(z_{n+1},\ldots,z_{n+r_1r_2})$ to be distinct. \footnote{In case $X$ is a symplectic manifold, then we should use locally closed real codimension 2 submanifold instead of $H_i$, as did in \cite{LT2}. Here we use this construction of uniformizing charts because it is compatible to the construction of atlas of the stack $\M_{g,n}(X,A)$ in algebraic geometry.} Let $\pi_U\!:\!\tilde U\to U$ be the projection that sends $({\mathcal C};z_{n+1},\ldots,z_{n+r_1r_2})$ to ${\mathcal C}$. Clearly, $G_w$ acts on $\pi_U^{-1}(w)$ canonically by permuting their $(n+r_1r_2)$-marked points. Namely, for any $\sigma\in G_w$ and ${\mathcal C}\in\pi_U^{-1}(w)$ with marked points $z_1,\ldots,z_{n+r_1r_2}$, $\sigma({\mathcal C})$ is the same map with the marked points $\sigma(z_1),\ldots,\sigma(z_{n+r_1r_2})$. In particular, we can view $G_w$ as a subgroup of the permutation group $S_{n+r_1r_2}$. Hence $G_w$ acts on $\tilde U$ by permuting the marked points of ${\mathcal C}\in\tilde U$ according to the inclusion $G_w\subset S_{n+r_1r_2}$. Note that if $H_i$'s are in general position then elements in $\tilde U$ has no automorphisms and have distinct marked points. Let $G_{\tilde U}=G_w$. Since fibers of $\pi_U$ are invariant under $G_{\tilde U}$, $\pi_U$ induces a map $\tilde U/G_{\tilde U}\to U$, which is obviously a covering\! \footnote{In this paper we call $p\!:\! A\to B$ a covering if $p$ is a covering projection~\cite{Sp} and $\#(p^{-1}(x))$ is independent of $x\in B$. We call $p\!:\! A\to B$ a local covering if $p(A)$ is open in $B$ and $p\!:\! A\to p(A)$ is a covering.} if $U$ is sufficiently small. Further, if we let $${\mathbf E}_{\tilde U}=\bigcup_{{\mathcal C}\in\tilde U} \Lambda^{0,1}_{{\mathcal C}} $$ and let $\Phi_{\tilde U}\!:\!\tilde U\to{\mathbf E}_{\tilde U}$ be the section that sends ${\mathcal C}$ to $\bar\partial f_{{\mathcal C}}$, then $\Phi_{\tilde U}$ is $G_{\tilde U}$-equivariant and $\Phi|_U\!:\! U\to {\mathbf E}|_U$ is the descent of $\Phi_{\tilde U}/G_{\tilde U}\!:\!\tilde U/G_{\tilde U}\to{\mathbf E}_{\tilde U}/G_{\tilde U}$. Note that fibers of ${\mathbf E}_{\tilde U}$ over $\tilde U$ are linear spaces. Following the convention, we will call $\Lambda=(\tilde U,\bE_{\tilde U},\Phi\ltilu,G_{\tilde U})$ a uniformizing chart of $({\mathbf B},{\mathbf E},\Phi)$ over $U$. Let $V\subset U$ be an open subset and let $\tilde V=\pi_U^{-1}(V)$, let $G_{\tilde V}=G_{\tilde U}$, let ${\mathbf E}_{\tilde V}={\mathbf E}_{\tilde U}|_{_{\tilde V}}$ and let $\Phi_{\tilde V}=\Phi_{\tilde U}|_{\tilde V}$. We will call $\Lambda^{\prime}=(\tilde V,{\mathbf E}_{\tilde V},\Phi_{\tilde V}, G_{\tilde V})$ a uniformizing chart of $({\mathbf B},{\mathbf E},\Phi)$ that is the {\it restriction} of the original chart to $V$, and denoted by $\Lambda|_V$. We can also construct uniformizing charts by pull back. Let $G_{\tilde V}$ be a finite group acting effectively on a topological space $\tilde V$, let $G_{\tilde V}\to G_{\tilde U}$ be a homomorphism and $\varphi\!:\! \tilde V\to\tilde U$ be a $G_{\tilde V}$-equivariant map so that $\tilde V/G_{\tilde V}\to \tilde U/G_{\tilde U}$ is a local covering map. Then we set ${\mathbf E}_{\tilde V}=\varphi^{\ast}{\mathbf E}_{\tilde U}$ and $\Phi_{\tilde V}=\varphi^{\ast}\Phi_{\tilde U}$. The data $\Lambda^{\prime}=(\tilde V,{\mathbf E}_{\tilde V},\Phi_{\tilde V}, G_{\tilde V})$ is also a uniformizing chart. We will call $\Lambda^{\prime}$ the {\sl pull back} of $\Lambda$, and denoted by $\varphi^{\ast}\Lambda$. In the following, we will denote the collection of all uniformizing charts of $({\mathbf B},{\mathbf E},\Phi)$ by ${\mathfrak C}$. The collection ${\mathfrak C}$ has the following compatibility property. Let $$\Lambda_i=(\tilde U_i,{\mathbf E}_{\tilde U_i},\Phi_{\tilde U_i}, G_{\tilde U_i}), $$ where $i=1,\ldots,k$, be a collection of uniformizing charts in ${\mathfrak C}$ over $U_i\subset{\mathbf B}$ respectively. Let $p\in\cap_{i=1}^k U_i$ be any point. Then there is a uniformizing chart $\Lambda=(\tilde V,{\mathbf E}_{\tilde V},\Phi_{\tilde V},G_{\tilde V})$ over $V\subset\cap^k U_i$ with $p\in V$ such that there are homomorphisms $G_{\tilde V}\to G_{\tilde U_i}$ and equivariant local covering maps $\varphi_i\!:\!\tilde V\to\pi_{U_i}^{-1}(V) \subset\tilde U_i$ compatible with $\tilde V\to V$ and $\pi_{U_i}^{-1}(V)\to V\subset U_i$, such that $\varphi_i^{\ast}({\mathbf E}_{\tilde U_i},\Phi_{\tilde U_i}) \cong ({\mathbf E}_{\tilde V},\Phi_{\tilde V})$. In this case, we say $\Lambda$ is {\it finer than} $\Lambda_i|_{V}$. The main difficulty in constructing the GW invariants in this setting is that the smoothness of $(\tilde U,\bE_{\tilde U},\Phi\ltilu)$ is unclear when $U$ contains maps whose domains are singular. To overcome this difficulty, the authors introduced the notion of generalized Fredholm bundles in~\cite{LT2}. The main result of~\cite{LT2} is the following theorems, which enable them to construct the GW invariants for all symplectic manifolds. \begin{theo} \label{1.3} The data $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ is a generalized oriented Fredholm V-bundle of relative index $2r_{\rm exp }$, where $r_{\rm exp }= c_1(X)\cdot A+n+(n-3)(1-g)$ is half of the virtual (real) dimension of $\Phi^{-1}(0)$. \end{theo} \begin{theo} \label{1.4} For any generalized oriented Fredholm V-bundle $[ \Phi\!:\! {\mathbf B}\to{\mathbf E}]$ of relative index $r$, we can assign to it an Euler class $e([\Phi\!:\! {\mathbf B}\to{\mathbf E}])$ in $H_r({\mathbf B};{\mathbb Q})$ that satisfies all the expected properties of the Euler classes. \end{theo} As explained in the introduction, the pairing of the Euler class of $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ with the tautological topological class will give rise to the symplectic version of the GW invariants of $X$. Further, the Comparison Theorem we set out to prove amounts to compare this Euler class with the image of the virtual moduli cycle $[\M_{g,n}(X,A)]^{{\rm vir}}$ in $H_r({\mathbf B};{\mathbb Q})$ via the inclusion $\M_{g,n}(X,A)^{\rm top}\subset{\mathbf B}$. In the remainder part of this section, we will list all properties of $ [\Phi\!:\! {\mathbf B}\to{\mathbf E}] $ that are relevant to the construction of its Euler class. This list is essentially equivalent to saying that $ [\Phi\!:\! {\mathbf B}\to{\mathbf E}] $ is a generalized oriented Fredholm V-bundle. After that, we will construct the Euler class of $ [\Phi\!:\! {\mathbf B}\to{\mathbf E}] $ in details. We begin with the notion of weakly smooth structure. A local smooth approximation of $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ over $U\subset {\mathbf B}$ is a pair $(\Lambda,V)$, where $\Lambda=(\tilde U,\bE_{\tilde U},\Phi\ltilu,G_{\tilde U})$ is a uniformizing chart over $U$ and $V$ is a finite equi-rank $G_{\tilde U}$-vector bundle over $\tilde U$ that is a $G_{\tilde U}$-equivariant subbundle of $\bE_{\tilde U}$ such that $R_V: = \Phi\ltilu^{-1}(V)\subset\tilde U$ is an equi-dimensional smooth manifold, $V|_{R_V}$ is a smooth vector bundle and the lifting $\phi_V\!:\! R_V\to V|_{R_V}$ of $\Phi\ltilu|_{R_V}$ is a smooth section. An orientation of $(\Lambda,V)$ is a $G_{\tilde V}$-invariant orientation of the real line bundle $\wedge^{\rm top}(TR_V)\otimes\wedge^{\rm top} (V|_{R_V})^{-1}$ over $R_V$. We call $\rank V-\dim R_V$ the index of $(\Lambda,V)$ (We remind that all ranks and dimensions in this section are over reals). Now assume that $(\Lambda^{\prime},V^{\prime})$ is another weakly smooth structure of identical index over $W\subset{\mathbf B}$. We say that $(\Lambda^{\prime},V^{\prime})$ is finer than $(\Lambda,V)$ if the following holds. First, the restriction $\Lambda^{\prime}|_{W\cap U}$ is finer than $\Lambda|_{W\cap U}$; Secondly, if we let $\varphi\!:\! \pi_W^{-1}(W\cap U)\to\pi_U^{-1}(W\cap U)$ be the covering map then $\varphi^{\ast} V\subset \varphi^{\ast}\bE_{\tilde U}\equiv {\mathbf E}_{\tilde W}|_{\pi_W^{-1}(W\cap U)}$ is a subbundle of $V^{\prime}|_{\pi_W^{-1}(W\cap U)}$; Thirdly, for any $w\in\tilde W$ the homomorphism $ T_wR_{V^{\prime}}\to \bigl( V^{\prime}/\varphi^{\ast} V\bigr) |_w$ induced by $d\phi_{V^{\prime}}(w)\!:\! T_wR_V\to V^{\prime}|_w$ is surjective, and the map $\phi_{V^{\prime}}^{-1}(\varphi^{\ast} V)\to R_V$ induced by $\varphi$ is a local diffeomorphism between smooth manifolds. Note that the last condition implies that if we identify $T_{\varphi(w)}R_V$ with $T_w\phi_{V^{\prime}}^{-1}(\varphi^{\ast} V) \subset T_w R_{V^{\prime}}$, then the induced homomorphism \begin{equation} \label{eq:1.2} T_wR_{V^{\prime}} /T_{\varphi(w)}R_V\lra \bigl( V^{\prime}/\varphi^{\ast} V)|_w \end{equation} is an isomorphism. In case both $(\Lambda,V)$ and $(\Lambda^{\prime},V^{\prime})$ are oriented, then we require that the orientation of $(\Lambda,V)$ coincides with that of $(\Lambda^{\prime},V^{\prime})$ based on the isomorphism \begin{equation} \label{eq:1.3} \wedge^{\rm top}(T_w R_{V^{\prime}})\otimes\wedge^{\rm top}(V^{\prime}|_w)^{-1}\cong \wedge^{\rm top}(T_{\varphi(w)}R_V)\otimes\wedge^{\rm top}(V|_{\varphi(w)})^{-1} \end{equation} induced by~\eqref{eq:1.2}. Now let ${\mathfrak A}=\{(\Lambda_i,V_i)\}_{i\in{\mathcal K}}$ be a collection of oriented smooth approximations of $({\mathbf B},{\mathbf E},\Phi)$. In the following, we will denote by $U_i$ the open subsets of ${\mathbf B}$ such that $\wedge_i$ is a smooth chart over $U_i$. We say ${\mathfrak A}$ covers $\phi^{-1}(0)$ if $\phi^{-1}(0)$ is contained in the union of the images of $U_i$ in ${\mathbf B}$. \begin{defi}\label{1.5} An index $r$ oriented weakly smooth structure of $({\mathbf B},{\mathbf E},\Phi)$ is a collection ${\mathfrak A}=\{(\Lambda_i,V_i)\}_{i\in{\mathcal K}}$ of index $r$ oriented smooth approximations such that ${\mathfrak A}$ covers $\Phi^{-1}(0)$ and that for any $(\Lambda_i,V_i)$ and $(\Lambda_j,V_j)$ in ${\mathfrak A}$ with $p\in U_i\cap U_j$, there is a $(\Lambda_k,V_k)\in{\mathfrak A}$ such that $p\in U_k$ and $(\Lambda_k,V_k)$ is finer than $(\Lambda_i,V_i)$ and $(\Lambda_j,V_j)$. \end{defi} Let ${\mathfrak A}^{\prime}$ be another index $r$ oriented weakly smooth structure of $({\mathbf B},{\mathbf E},\Phi)$. We say ${\mathfrak A}^{\prime}$ is finer than ${\mathfrak A}$ if for any $(\Lambda,V)\in{\mathfrak A}$ over $U\subset{\mathbf B}$ and $p\in U\cap\Phi^{-1}(0)$, there is a $(\Lambda^{\prime},V^{\prime})\in{\mathfrak A}^{\prime}$ over $U^{\prime}$ such that $p\in U^{\prime}$ and $(\Lambda^{\prime},V^{\prime})$ is finer than $(\Lambda,V)$. We say that two weakly smooth structures ${\mathfrak A}_1$ and ${\mathfrak A}_2$ are equivalent if there is a third weakly smooth structure that is finer than both ${\mathfrak A}_1$ and ${\mathfrak A}_2$. \begin{prop}[\cite{LT2}]\label{1.6} The tuple $({\mathbf B},{\mathbf E},\Phi)$ constructed at the beginning of this section admits a canonical oriented weakly smooth structure of index $2r_{\rm exp }$. \end{prop} We remark that the construction of such a weakly smooth structure is the core of the analytic part of [LT2]. In the following, we will use the weakly smooth structure of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ to construct its Euler class. The idea of the construction is as follows. Given a local smooth approximation $(\Lambda,V)$ over $U\subset{\mathbf B}$, we obtain a smooth manifold $R_V$, a vector bundle $V|_{R_V}$ and a smooth section $\phi_V\!:\! R_V\to V|_{R_V}$. Following the topological construction of the Euler classes, we shall perturb $\phi_V$ to a new section $\tilde\phi_V\!:\! R_V\to V|_{R_V}$ so that $\tilde\phi_V$ is transversal to the zero section of $V|_{R_V}$. Here by a section transversal to the zero section, we mean that the graph of this section is transversal to the zero section in the total space of the vector bundle. Hence the Euler class will be the cycle represented by ${\tilde\phi_V}^{-1}(0)$ near $U$. Since the weakly smooth structure of $\Phi\!:\! {\mathbf B}\to{\mathbf E}$ is given by a collection of compatible by not necessary matching local smooth approximations, we need to work out this perturbation scheme with special care so that $\{\tilde\phi_V^{-1}(0)\}$ patch together to form a well-defined cycle. Let ${\mathfrak A}=\{(\Lambda_{\alpha},V_{\alpha})\}_{\alpha\in{\mathcal K}}$ be the weakly smooth structure provided by Proposition~\ref{1.6}. For convenience, for any $\alpha\in{\mathcal K}$ we will denote the corresponding uniformizing chart $\Lambda_{\alpha}$ by $(\tilde U_{\alpha},\tilde{\bB}_{\alpha},\tilde{\Phi}_{\alpha},G_{\alpha})$ and will denote its descent by $(U_{\alpha},E_{\alpha},\Phi_{\alpha})$. Accordingly, we will denote the projection $\pi_{U_{\alpha}}\!:\!\tilde U_{\alpha}\to U_{\alpha}$ by $\pi_{\alpha}$, denote ${\tilde\Phi_{\alpha}}^{-1}(V_{\alpha})$ by $R_{\alpha}$, denote $V_{\alpha}|_{R_{\alpha}}$ by $W_{\alpha}$ and denote the lifting of $\tilde\Phi_{\alpha}|_{R_{\alpha}} \!:\! R_{\alpha}\to\tilde{{\mathbf E}}_{\alpha}|_{R_{\alpha}}$ by $\phi_{\alpha}\!:\! R_{\alpha}\to W_{\alpha}$. Without loss of generality, we can assume that for any approximation $(\Lambda_{\alpha},V_{\alpha})\in{\mathfrak A}$ over $U_{\alpha}$ and any $U^{\prime}\subset U_{\alpha}$, the restriction $(\Lambda_{\alpha},V_{\alpha})|_{U^{\prime}}$ is also a member in ${\mathfrak A}$. In the following, we call $S\subset R_{\alpha}$ {\it symmetric} if $S=\pi_{\alpha}^{-1}(\pi_{\alpha}(S))$. Next, we pick a covering data for $\Phi^{-1}(0)\subset{\mathbf B}$ provided by the following covering lemma. \begin{lemm}[\cite{LT2}]\label{1.20} There is a finite collection ${\mathcal L}\subset{\mathcal K}$ and a total ordering of ${\mathcal L}$ of which the following holds. the set $\Phi^{-1}(0)$ is contained in the union of $\{R_{\alpha}\}_{\alpha\in{\mathcal L}}$ and for any $\alpha$ and $\beta\in{\mathcal L}$ such that $\alpha<\beta$ then approximation $(\Lambda_{\beta},V_{\beta})$ is finer than the approximation $(\Lambda_{\alpha},V_{\alpha})$. \end{lemm} \begin{proof} The lemma is part of Proposition 2.2 in \cite{LT2}. It is proved there by using the stratified structures of $({\mathbf B},{\mathbf E},\Phi)$. Here we will give a direct proof of this by using the definition of smooth approximations, when $\Phi^{-1}(0)$ is triangulable, which is true when $X$ is projective. Let $k$ be the real dimension of $\Phi^{-1}(0)$. To prove the lemma, we will show that there are $k+1$ subsets ${\mathcal L}_k,\ldots,{\mathcal L}_0\subset{\mathcal K}$ and that for each $\alpha\in\cup_{i=0}^k{\mathcal L}_i$ there is an open symmetric subset $U_{\alpha}^{\prime}\Subset U_{\alpha}$ such that $R_{\alpha}^{\prime}=R_{\alpha}\cap\pi_{\alpha}^{-1}(U_{\alpha}^{\prime})\Subset R_{\alpha}$ of which the following holds: first, for each $i\leq k$ the set $Z_i=\Phi^{-1}(0)-\cup_{j\geq i}\cup_{\alpha\in{\mathcal L}_j} U_{\alpha}^{\prime}$ is a triangulable space whose dimension is at most $i-1$, and secondly, for any pair of distinct $(\alpha,\beta)\in{\mathcal L}_i\times{\mathcal L}_j$ with $i\leq j$, the restriction $(\Lambda_{\alpha},V_{\alpha})|_{U_{\alpha}^{\prime}\cap U_{\beta}^{\prime}}$ is finer than $(\Lambda_{\beta},V_{\beta})|_{U_{\alpha}^{\prime}\cap U_{\beta}^{\prime}}$. We will construct ${\mathcal L}_i$ inductively, starting from ${\mathcal L}_k$. We first pick a finite ${\mathcal L}_k\subset{\mathcal K}$ so that $\cup_{\alpha\in{\mathcal L}_k} U_{\alpha} \supset \Phi^{-1}(0)$. This is possible since $\Phi^{-1}(0)$ is compact. Since it is also triangulable, we can find a symmetric $U_{\alpha}^{\prime}\subset U_{\alpha}$ for each $\alpha\in{\mathcal L}_k$ so that $\{U^{\prime}_{\alpha}\}_{\alpha\in{\mathcal L}_k}$ is disjoint, $R_{\alpha}^{\prime}\Subset R_{\alpha}$ and $Z_k$ is trangulable with dimension at most $k-1$. Now we assume that we have found ${\mathcal L}_k,\ldots,{\mathcal L}_i$ as desired. Then for each $x\in Z_i$ we can find a neighborhood $O$ of $x\in{\mathbf B}$ such that for any $\alpha\in\cup_{j\geq i}{\mathcal L}_j$ either $x\in U_{\alpha}$ or $O\cap U_{\alpha}^{\prime}=\emptyset$. Let ${\mathcal I}_x$ be those $\alpha$ in $\cup_{j\geq i}{\mathcal L}_{\alpha}$ such that $x\in U_{\alpha}$. Then by the property of ${\mathfrak A}$ there is a $\beta\in{\mathcal K}$ so that $(\Lambda_{\beta},V_{\beta})$ is finer than $(\Lambda_{\alpha},V_{\alpha})|_{U_{\alpha}^{\prime}}$ for all $\alpha\in{\mathcal I}_x$. Without loss of generality, we can assume that $U_{\beta}\subset O$. Then $(\Lambda_{\beta},V_{\beta})$ is finer than $(\Lambda_{\alpha},V_{\alpha})$ for all $\alpha\in\cup_{j\geq i}{\mathcal L}_j$. Since $Z_i$ is compact, we can cover it by finitely many such $(\Lambda_{\beta},V_{\beta})$'s, say indexed by ${\mathcal L}_{i-1}\subset{\mathcal K}$. On the other hand, since $Z_i$ is triangulable with dimension at most $i-1$, we can find symmetric $U_{\alpha}^{\prime}\subset U_{\alpha}$ for each $\alpha\in{\mathcal L}_{i-1}$ so that $R^{\prime}_{\alpha}\Subset R_{\alpha}$ for $\alpha\in{\mathcal L}_{i-1}$ and $Z_i-\cup_{\alpha\in{\mathcal L}_{i-1}} U_{\alpha}^{\prime}$ is trianglable with dimension at most $i-2$. This way, we can find the set ${\mathcal L}_k,\ldots, {\mathcal L}_0$ as desired. In the end, we simply put ${\mathcal L}=\cup_{i=0}^k{\mathcal L}_i$. We give it a total ordering so that whenever $\alpha\in{\mathcal L}_i$, $i\geq j$ and $\beta\in{\mathcal L}_j$ then $\alpha\leq\beta$. This proves the Lemma. \end{proof} We now fix such a collection ${\mathcal L}$ once and for all. Since ${\mathcal L}$ is totally ordered, in the following we will replace the index by integers that range from 1 to $\#({\mathcal L})$ and use $k$ to denote an arbitrary member of ${\mathcal L}$. We first build the comparison data into the collection $\{R_k\}_{k\in{\mathcal L}}$ and $\{W_k\}_{k\in{\mathcal L}}$. To distinguish the projection $\pi_k\!:\!\tilde U_k\to U_k$ from the composite $\tilde U_k\to U_k\to{\mathbf B}$, we will denote the later by $\iota_k$. For any pair $k\geq l$, we set $R_{k,l}=\iota_k^{-1}(\iota_l(R_l))$. Then there is a canonical map and a canonical vector bundle inclusion \begin{equation} \label{eq:1.4} f^l_k\!:\! R_{k,l}\to R_l \quad{\rm and}\quad (f^l_k)^{\ast}(W_l)\xrightarrow{\subset} W_k|_{R_{k,l}}, \end{equation} that is part of the data making $(\Lambda_k,V_k)$ finer than $(\Lambda_l,V_l)$. Note that $R_{k,l}\subset R_k$ is a locally closed submanifold, $f^l_k(R_{k,l})$ is open in $R_l$ and $f^l_k\!:\! R_{k,l}\to f^l_k(R_{k,l})$ is a covering map. Because of the compatibility condition, for any $k>l>m$ if $R_{k,l}\cap R_{k,m}\ne\emptyset$ then $f^l_k(R_{k,l}\cap R_{k,m})\subset R_{l,m}$ and \begin{equation} f^m_l\circ f^l_k=f^m_k: R_{k,l}\cap R_{k,m}\lra R_m. \label{eq:1.5} \end{equation} Further, restricting to $R_{k,l}\cap R_{k,m}$, the pull backs \begin{equation} \label{eq:1.6} (f_k^m)^{\ast}(W_m)|_{R_{k,l}\cap R_{k,m}} =(f^l_k)^{\ast}(f^m_l)^{\ast}(W_m)|_{R_{k,l}\cap R_{k,m}} \subset W_k|_{R_{k,l}\cap R_{k,m}}. \end{equation} In the following, we will use ${\mathfrak R}$ to denote the collection of data $\{(R_{k,l},f^l_k)\}$ and use ${\mathfrak W}$ to denote the data $\{(W_k,(f^l_k)^{\ast})\}$. We will call the pair $({\mathfrak R},{\mathfrak W})$ a good atlas of the weakly smooth structure ${\mathfrak A}$ of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$. For technical reason, later we need to shrink each $R_k$ slightly. More precisely, let $\{ S_k\}_{k\in{\mathcal L}}$ be a collection of symmetric open subsets $S_k\Subset R_k$ such that $\{ S_k\}$ still covers $\Phi^{-1}(0)$. We then let $S_{k,l}=(f^l_k)^{-1}(S_l)\cap S_k$, let $W_k^{\prime}=W_k|_{S_k}$ and let $g^l_k$ and $(g^l_k)^{\ast}$ be the restriction to $S_{k,l}$ of $f^l_k$ and $(f^l_k)^{\ast}$ respectively. Then $({\mathfrak S},{\mathfrak W}^{\prime})$, where ${\mathfrak S}=\{(S_{k,l},g^l_k)\}$ and ${\mathfrak W}^{\prime}= \{(W_k^{\prime},(g^l_k)^{\ast})\}$, is also a good atlas of $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$. We call it a precompact sub-atlas of $({\mathfrak R},{\mathfrak W})$, and denote it in short by ${\mathfrak S}\Subset{\mathfrak R}$. To describe the collection $\{\phi_k\}$, we need to introduce the notion of regular extension. Let $M$ be a manifold and $M_0\subset M$ be a locally closed submanifold. Let $V\to M$ be a smooth vector bundle and $V_0\to M_0$ a subbundle of $V|_{M_0}$. We assume that both $(M,V)$ and $(M_0,V_0)$ are oriented. We say that a section $h\!:\! M\to V$ is a smooth extension of $h_0\!:\! M_0\to V_0$ if both $h_0$ and $h$ are smooth and if the induced section $M_0\xrightarrow{h_0} V_0\to V|_{M_0}$ is identical to the restriction $h|_{X_0}\!:\! X_0\to V_0$. We say $h$ is a regular extension of $h_0$ if in addition to $h$ being a smooth extension of $h_0$ we have that for any $x\in X_0$ the homomorphism \begin{equation} dh(x): T_xM/T_xM_0\lra (V/V_0)|_x \label{eq:1.7} \end{equation} is an isomorphism and the orientation of $(M,V)$ and $(M_0,V_0)$ are compatible over $M_0$ based on the isomorphism~\eqref{eq:1.7}. \begin{defi} A collection $\{h_k\}_{k\in{\mathcal L}}$ is called a smooth section of ${\mathfrak W}$ if $h_k$ is a smooth section of $W_k$ for each $k\in{\mathcal L}$ and $h_k$ is a smooth extension of $h_l$ for any pair $k\geq l$ in ${\mathcal L}$. If in addition that $h_k$ is a regular extension of $h_l$ for all $k\geq l$, then we call $\{h_k\}$ a regular section of ${\mathfrak W}$. \end{defi} In the following, we will use ${\mathfrak h}\!:\!{\mathfrak R}\to{\mathfrak W}$ to denote a smooth section with ${\mathfrak h}$ understood to be $\{h_k\}_{k\in{\mathcal L}}$. We set ${\mathfrak h}^{-1}(0)$ to be the collection $\{h_k^{-1}(0)\}$ and set $\iota({\mathfrak h}^{-1}(0))$ to be the union of $\iota_k(h_k^{-1}(0))$ in ${\mathbf B}$. We say ${\mathfrak h}^{-1}(0)$ is proper if $\iota({\mathfrak h}^{-1}(0))$ is compact. Without loss of generality, we can assume that $\dim R_k>0$ for all $k\in{\mathcal L}$. We say that ${\mathfrak h}$ is transversal to the zero section $\mathbf 0\!:\! {\mathfrak R}\to{\mathfrak W}$ if ${\mathfrak h}$ is a regular section and if for any $k\in{\mathcal L}$ the graph $\Gamma_{h_k}$ of $h_k$ is transversal to the 0 section of $W_k$ in the total space of $W_k$. \begin{lemm} \label{1.9} Let the notation be as before. Then ${\mathfrak h}^{-1}(0)$ is proper if and only if there is a symmetric open subsets $R_k^{\prime}\Subset R_k$ for each $k\in{\mathcal L}$ such that $\cup_{k\in{\mathcal L}}\iota_k(h_k^{-1}(0))\subset \cup_{k\in{\mathcal L}}\iota_k(R^{\prime}_k)$ and such that for each $k\in{\mathcal L}$, \begin{equation} \label{eq:1.8} h_k^{-1}(0)\cap (R_k- R_k^{\prime}) \subset\bigl(\bigcup_{l<k}(f^l_k)^{-1}(R_l^{\prime})\bigr) \cup\bigl( \bigcup_{l>k}f^l_k(R_{k,l}^{\prime})\bigr). \end{equation} \end{lemm} \begin{proof} We first assume that $Z=\cup_{k\in{\mathcal L}} \iota_k(h_k^{-1}(0))$ is compact. Then since $\{R_k\}_{k\in{\mathcal L}}$ covers $Z$ and since $\dim R_k>0$, for each $k\in{\mathcal L}$ we can find symmetric $R_k^{\prime}\Subset R_k$ so that $\{R_k^{\prime}\}_{k\in{\mathcal L}}$ still covers $Z$. Obviously, this implies~\eqref{eq:1.8}. Conversely, if we have found $R_k^{\prime}\subset R_k$ as stated in the lemma, then $\{\mathop{\rm cl\hspace{1pt}}(\iota_k(R_k^{\prime}))\cap Z\}$ will cover $Z$, where $\mathop{\rm cl\hspace{1pt}}(A)$ is the closure of $A$. Since $\mathop{\rm cl\hspace{1pt}}(\iota_k(R_k^{\prime}))$ are compact and since $Z\cap \mathop{\rm cl\hspace{1pt}}(\iota_k(R_k^{\prime}))$ is closed in $\mathop{\rm cl\hspace{1pt}}(\iota_k(R_k^{\prime}))$, $Z$ is compact as well. This proves the lemma. \end{proof} \begin{lemm} \label{1.10} Let ${\boldsymbol \phi}\!:\! {\mathfrak R}\to{\mathfrak W}$ be the collection $\{\phi_k\}$ induced by $\{\tilde\Phi_k\}_{k\in{\mathcal L}}$. Then ${\boldsymbol \phi}$ is a regular section with proper vanishing locus. \end{lemm} \begin{proof} This is equivalent to the fact that $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ is a weakly Fredholm V-bundle, which was introduced and proved in \cite{LT2}. \end{proof} Now let ${\mathfrak h}\!:\!{\mathfrak R}\to{\mathfrak W}$ be a regular section such that ${\mathfrak h}$ is transversal to the zero section and ${\mathfrak h}^{-1}(0)$ is proper. We claim that the data $\{h_k^{-1}(0)\}$ descends to an oriented current in ${\mathbf B}$ with rational coefficients supported on a stratified set whose boundary is empty. In particular, it defines a singular homology class in $H_{\ast}({\mathbf B},{\mathbb Q})$. Recall that for each $k\in{\mathcal L}$ the associated group $G_k$ acts on $R_k$ such that $R_k/G_k$ is a covering of $\iota_k(R_k)$. We let $m_k$ be the product of the order of $G_k$ with the number of the sheets of the covering $R_k/G_k\to\iota_k(R_k)$. Note that then the covering $R_{k,l}\to f^l_k(R_{k,l})$ is an $m_k/m_l$-fold covering. Because $h_k$ is a regular extension of $(f^l_k)^{\ast}(h_l)$, $(f^l_k)^{\ast}(h_l)^{-1}(0)$ is an open submanifold of $h_k^{-1}(0)$ with identical orientations. Hence $\iota_k(h^{-1}_k(0))$ and $\iota_l(h_l^{-1}(0))$ patch together to form a stratified subset, and consequently the collection $\{\iota_k(h_k^{-1}(0))\}_{k\in{\mathcal L}}$ patch together to form a stratified subset, say $Z$, in ${\mathbf B}$. Now we assign multiplicities to open strata of $Z$. Let $O_k=\iota_k(h_k^{-1}(0))$. Since $O_k\subset Z$ is an open subset, we can assign multiplicities to $O_k$ so that as oriented current $[O_k]=\iota_{\ast}(\frac{1}{m_k}[ h_k^{-1}(0)])$, where $[h_k^{-1}(0)]$ is the current of the oriented manifold $h_k^{-1}(0)$ with multiplicity one. Here the orientation of $h_k^{-1}(0)$ is the one induced by the orientation of $(R_k,W_k)$. Using the fact that $R_{k,l}\to f^l_k(R_{k,l})$ is a covering with $m_k/m_l$ sheets, the assignments of the multiplicities of $O_k$ and $O_l$ over $\iota_k(R_k)\cap\iota_l(R_l)$ coincide. Therefore $Z$ is an oriented stratified set of pure dimension with rational multiplicities. We let $[Z]$ be the corresponding current. It remains to check that $\partial[Z]=0$ as current. Clearly, $\partial[Z\cap O_k]\subset\mathop{\rm cl\hspace{1pt}}(O_k)-O_k$. Since $\{O_k\cap Z\}$ is an open covering of $Z$, $\partial[Z]=0$ if $Z$ is compact. But this is what we have assumed in the first place. Later, we will denote the so constructed cycle by $$[{\mathfrak h}^{-1}(0)]\in H_{\ast}({\mathbf B},{\mathbb Q}). $$ In the remainder of this section, we will perturb the section ${\boldsymbol \phi}\!:\!{\mathfrak R}\to{\mathfrak W}$ to a new section so that it is transversal to the zero section and so that its vanishing locus is compact. The current defined by the vanishing locus of the perturbed section will define the Euler class of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$. We begin with a collection ${\mathcal S}=\{S_l\}_{l\in{\mathcal L}}$ of symmetric open $S_k\Subset R_k$ such that $\{\iota_l(S_l)\}$ cover $\iota({\boldsymbol \phi}^{-1}(0))$. For technical reason, we assume that for each $k\in{\mathcal L}$ the boundary $\partial S_k$, which is defined to be $\mathop{\rm cl\hspace{1pt}}(S_k)-S_k$ in $R_k$, is a smooth manifold of dimension $\dim S_k-1$. By slightly altering $S_k$ if necessarily, we can and do assume that $\partial S_k$ is transversal to $R_{k,l}$ along $\partial S_k\cap (f^l_k)^{-1}(\mathop{\rm cl\hspace{1pt}}(S_l))$ for all $l<k$. (We will call such ${\mathcal S}$ satisfying the transversality condition on its boundary.) Following the convention, we set $S_{k,l}=(f^l_k)^{-1}(S_l)\cap S_k$. We now construct a collection of (closed) tubular neighborhoods of $S_{k,l}$ in $R_k$. We fix the index $k$ and consider the closed submanifold (with boundary) $\Sigma_l:=\mathop{\rm cl\hspace{1pt}}(S_{k,l})\subset R_k$. Because of the transversality condition on $\partial S_l$ and on $\partial S_k$, we can find a $D^h$-bundle $p_l\!:\! T_l\to \Sigma_l$, where $D^h$ is the closed unit ball in ${\mathbb R}^h$ and $h=\dim R_k-\dim R_{k,l}$, and a smooth embedding $\eta_l\!:\! T_l\to R_k$ of which the following two conditions holds. First, the restriction of $\eta_l$ to the zero section $\Sigma_l\subset T_l$ is the original embedding $\Sigma_l\subset R_k$, and secondly \begin{equation} \label{eq:1.9} \eta_l(p_l^{-1}(\Sigma_l\cap \partial S_k ))\subset \partial S_k \quad{\rm and}\quad \eta_l(p_l^{-1}(S_{k,l} ))\subset S_k. \end{equation} For any $0<\epsilon<1$, we let $T_l^{\epsilon}\subset T_l$ be the closed $\epsilon$-ball subbundle of $T_l$. By abuse of notation, in the following we will not distinguish $T^{\epsilon}_l$ from its image $\eta_l(T_l^{\epsilon})$ in $R_k$. We will call $T^{\epsilon}_l$ the $\epsilon$-tubular neighborhood of $\Sigma_l$ in $R_k$. One property we will use later is that if $R_{k,l}\cap R_{k,l^{\prime}}\ne\emptyset$ for $l^{\prime}<l<k$, then $R_{k,l}\cap R_{k,l^{\prime}}$ is an open subset of $R_{k,l^{\prime}}$, and hence for $0<\epsilon\ll 1$ we have $\Sigma_{l^{\prime}}\cap T^{\epsilon}_l\subset \Sigma_{l^{\prime}}\cap \Sigma_l$. Now consider $\Sigma_l\subset R_k$. Since $T_l$ is a disk bundle over $\Sigma_l$, it follows that we can extend the subbundle $(f^l_k)^{\ast}(W_l)|_{\Sigma_l}\subset W_k|_{\Sigma_l}$ to a smooth subbundle of $W_k|_{T_l}$, denoted by $F_l\subset W_k|_{T_l}$. We then fix an isomorphism and the inclusion \begin{equation} p_l^{\ast}\bigl( (f_k^l)^{\ast}(W_l)|_{\Sigma_l}\bigr) \cong F_l. \label{eq:1.10} \end{equation} In this way, we can extend any section $\zeta$ of $(f_k^l)^{\ast}(W_l)|_{\Sigma_l}$ to a section of $W_k|_{T_l}$ as follows. We first let $\zeta^{\prime}\!:\! T_l\to F_l$ be the obvious extension using the isomorphism \eqref{1.10}. We then let $\zeta_{{\rm ex}}\!:\! T_l\to W_k|_{T_l}$ be the induced section using the inclusion $F_l\subset W_k|_{T_l}$. We will call $\zeta_{{\rm ex}}$ the standard extension of $\zeta$ to $T_l$. We fix a Riemannian metric on $R_k$ and a metric on $W_k$. For any section $\zeta$ as before, we say $\zeta$ is sufficiently small if its $C^2$-norm is sufficiently small. We now state a simple but important observation. \begin{lemm} \label{1.11} Let the notation be as before. Then there is an $\epsilon>0$ such that for any section $g\!:\! T_l\to F_l\subset W_k|_{T_l}$ such that $\parallel\! g\!\parallel_{C^2}<\epsilon$, the section $h_k|_{T_l}+g$ is non-zero over $T^{\epsilon}_l-\Sigma_l$. \end{lemm} \begin{proof} This follows immediately from the fact that $\Sigma_l$ is compact and that for any $x\in R_{k,l}$ the differential $$dh_k: T_xR_k/T_x R_{k,l}\lra \bigl( W_k / (f^l_k)^{\ast}(W_l)\bigr)|_x $$ is an isomorphism. \end{proof} We now state and prove the main proposition of this section. \begin{prop} \label{1.12} Let ${\mathfrak h}\!:\! {\mathfrak R}\to{\mathfrak W}$ be a regular section with ${\mathfrak h}^{-1}(0)$ proper, let ${\mathfrak R}^{\prime}\Subset{\mathfrak R}$ be a good sub-atlas and let ${\mathfrak h}^{\prime}$ be the the restriction of ${\mathfrak h}$ to ${\mathfrak R}^{\prime}$. We assume that the vanishing locus of ${\mathfrak h}^{\prime}$ is still proper. Then there is a smooth family of regular sections $\mathfrak g(t)\!:\! {\mathfrak R}^{\prime}\to{\mathfrak W}^{\prime}$, where ${\mathfrak W}^{\prime}$ be the restriction of ${\mathfrak W}$ to ${\mathfrak R}^{\prime}$, parameterized by $t\in [0,1]$ such that \begin{equation} \bigcup_{t\in[0,1]}\iota\bigl(\mathfrak g(t)^{-1}(0)\bigr)\times\{t\} \subset {\mathbf B}\times [0,1] \label{eq:1.11} \end{equation} is compact, that $\mathfrak g(0)={\mathfrak h}^{\prime}$ and that $\mathfrak g(1)$ is transversal to the zero section of ${\mathfrak W}^{\prime}$. \end{prop} \begin{proof} We will construct the perturbation over $R_1^{\prime}$ and then successively extends it to the remainder of $\{R^{\prime}_k\}$. We first fix a collection of symmetric open subsets $\{S_k\}_{k\in{\mathcal L}}$ such that $R_k^{\prime}\Subset S_k\Subset R_k$ and that $S_k$ satisfies the transversality condition on its boundary. Let $k$ be any positive integer no bigger than $\#({\mathcal L})+1$. The induction hypothesis ${\mathcal H}_k$ states that for each integer $l<k$ we have constructed a symmetric open $S_l^{\prime}$ satisfying $R_l^{\prime}\Subset S_l^{\prime}\Subset S_l$ and a smooth family of small enough sections $e_l(t)\!:\! R_l\to W_l$ such that $e_l(0)\equiv0$ of which the following holds. First, let ${\mathbf h}_l(t)=h_l+e_l(t)$, then for any $l<m<k$ the section ${\mathbf h}_{m}(t)|_{S_m^{\prime}}$ is a regular extension of $(f^l_{m})^{\ast}({\mathbf h}_l(t))|_{S_{m,l}^{\prime}}$; Secondly, for any $l<k$, the section ${\mathbf h}_l(1)$ is transversal to the zero section of $W_l$ over $S_l^{\prime}$, and finally, for any $l<k$ and $t\in[0,1]$, \begin{equation} {\mathbf h}_l(t)^{-1}(0)\cap\bigl( S_l^{\prime}-R_l^{\prime}\bigr)\subset \bigl(\bigcup_{i\leq l}(f_l^i)^{-1}(R_i^{\prime})\bigr)\bigcup \bigl(\bigcup_{m\geq l}f_m^l(R_{m,l}^{\prime})\bigr). \label{eq:1.12} \end{equation} Clearly, the condition ${\mathcal H}_1$ is automatically satisfied. Now assume that we have found $\{ S_l^{\prime}\}_{l<k}$ and $\{e_l\}_{l<k}$ required by the condition ${\mathcal H}_k$. We will demonstrate how to find $e_k$ and a new sequence of open subsets $\{S_l^{\prime}\}_{l\leq k}$ so that the condition ${\mathcal H}_{k+1}$ will hold for $\{e_l\}_{l\leq k}$ and $\{ S^{\prime}_l\}_{l\leq k}$. We continue to use the notation developed earlier. In particular, we let $\Sigma_l$ be the closure of $S_{k,l}$, let $T_l$ be the (closed) tubular neighborhood of $\Sigma_l\subset R_k$ with the projection $p_l\!:\! T_l\to \Sigma_l$ and let $F_l$ be the subbundle of $W_k|_{T_l}$ with the isomorphism \eqref{1.10}. Let $\zeta_l(t)$ be the standard extension of $(f_k^l)^{\ast}(e_l(t))|_{\Sigma_l}$ to $T_l$. Note that $h_k|_{T_l}+\zeta_l(t)$ is a regular extension of $(f_k^l)^{\ast}({\mathbf h}_l(t))|_{\Sigma_l}$. Because $\{{\mathbf h}_l\}_{l<k}$ satisfies condition ${\mathcal H}_k$, for $l<m<k$ and $x\in \Sigma_l\cap \Sigma_{m}$ we have $(f_k^l)^{\ast}({\mathbf h}_l(t))(x)=(f^{m}_k)^{\ast}({\mathbf h}_{m}(t))(x)$. Now let \begin{equation*} A_l=p_l^{-1} \bigl((f_k^l)^{-1}(S^{\prime}_l)\bigr)- \bigcup_{l<m<k} p_{m}^{-1}\bigl((f_k^{m})^{-1}(\mathop{\rm cl\hspace{1pt}}(R^{\prime}_{m}))\bigr) \end{equation*} and let \begin{equation*} B_l=\mathop{\rm cl\hspace{1pt}} (R_{k,l}^{\prime})- \bigcup_{k>m>l}(f_k^{m})^{-1}(S_{m}^{\prime}). \end{equation*} Note that $\{A_l\}_{l<k}$ covers $\Int\bigl(\cup_{l<k} T_l\bigr)$, that $B_l\Subset A_l$ and that $\{B_l\}_{l<k}$ \ is a collection of compact subsets of $R_k$. Now let $\epsilon>0$ be sufficiently small. We choose a collection of non-negative smooth functions $\{\rho_l\}_{l<k}$ that obeys the requirement that $\supp(\rho_l)\Subset\Int(A_l \cap T_l^{\epsilon})$, that $\rho_l\equiv1$ in a neighborhood of $B_l$ and that $\sum_{l<k}\rho_l\equiv1$ in a neighborhood of $\cup_{l<k}\mathop{\rm cl\hspace{1pt}}(R_{k,l}^{\prime})$. This is possible because the last set is compact and is contained in $\Int( \cup_{l<k}A_l)$. We set $$\zeta(t)=\sum_{l<k}\rho_l\cdot \zeta_l(t). $$ Now we check that for each $l<k$ the section $h_k+\zeta(t)$ is a regular extension of $(f_k^l)^{\ast}({\mathbf h}_l(t))$ in a neighborhood of $\mathop{\rm cl\hspace{1pt}}(R_{k,l}^{\prime})$. Let $x$ be any point in $\mathop{\rm cl\hspace{1pt}}(R_{k,l}^{\prime})$. We first consider the case where $x$ is contained in $B_{m}$ for some $m\geq l$. Let $y=f_k^{m}(x)$. Note that $y\in S_{m}^{\prime}$. Then restricting to a sufficiently small neighborhood of $x$ the section $h_k+\zeta(t)$ is equal to $h_k+\zeta_{m}(t)$. Since $h_k+\zeta(t)$ is a regular extension of $(f_k^{m})^{\ast}({\mathbf h}_{m}(t))$ near $x$ and since ${\mathbf h}_{m}(t)$ is a regular extension of $(f^l_{m})^{\ast} ({\mathbf h}_l(t))$ in a neighborhood of $y\in S_{m}^{\prime}$, $h_k+\zeta(t)$ is a regular extension of $(f^l_k)^{\ast}({\mathbf h}_l(t))$ near $x$. We next consider the case where $x$ is not contained in any of the $B_{m}$'s. Let $\Lambda$ be the set of all $m>l$ such that $x\in (f_k^{m})^{-1}(S_{m}^{\prime})$. Then for any $m<k$ that is not in $\Lambda$, $\rho_{m}\equiv 0$ in a neighborhood of $x$. Here we have used the fact that $\Sigma_m\cap T_l^{\epsilon}\subset\Sigma_m\cap\Sigma_l$ for $0<\epsilon\ll 1$. On the other hand, by induction hypothesis for each $m\in\Lambda$ the section $h_k+\zeta_{m}(t)$ is a regular extension of $(f_k^l)^{\ast}({\mathbf h}_l(t))$ near $x$. Therefore since $\sum_{m\in\Lambda}\rho_{m}\equiv1$ near $x$, in a small neighborhood of $x$ $$h_k+\zeta(t)=\sum_{m\in\Lambda}\rho_{m}\cdot (h_k+\zeta_{m}(t)) $$ is also a regular extension of $(f_k^l)^{\ast}({\mathbf h}_l(t))$. Our last step is to extend $\zeta(t)$ to $R_k$. We let $e_k(t)$ be a smooth family of sufficiently small sections of $W_k$ such that $e_k(0)\equiv0$, that the restriction of $e_k(t)$ to a neighborhood of $\cup_{l<k}\mathop{\rm cl\hspace{1pt}}\bigl( (f_k^l)^{-1}(R_l^{\prime})\bigr)$ is $\zeta(t)$ and such that the section ${\mathbf h}_k(1)$ is transversal to the zero section in a neighborhood of $\mathop{\rm cl\hspace{1pt}}(R_k^{\prime})$ in $S_k$. The last condition is possible because $h_k+\zeta(1)$ is transversal to the zero section in a neighborhood of $\cup_{l<k}\mathop{\rm cl\hspace{1pt}}(R_{k,l}^{\prime})$. Therefore, by possibly shrinking $S_l^{\prime}$ while still keeping $R_l^{\prime}\Subset S_l^{\prime}$ for $l<k$ if necessary, we can find an $S_k^{\prime}\Subset S_k$ satisfying $R_k^{\prime}\Subset S_k^{\prime}$ such that the induction hypothesis ${\mathcal H}_k$ holds for $\{e_l\}_{l\leq k}$ and $\{S_l^{\prime}\}_{l\leq k}$, except possibly the third condition. We now show that the third condition of ${\mathcal H}_k$ holds as well. We only need to check the inclusion \eqref{eq:1.12} for $l=k$. First, by Lemma \ref{1.9} we can find an open $S\Subset S_k$ such that $R_k^{\prime}\Subset S$ and that \begin{equation} \label{eq:1.13} {\mathbf h}_k^{-1}(0)\cap (\mathop{\rm cl\hspace{1pt}}(S)-R^{\prime}_k)\subset \bigl(\bigcup_{i< k}(f^i_k)^{-1}(R_i^{\prime})\bigr) \bigcup\bigl(\bigcup_{i> k}f^k_i(R_{i,k}^{\prime})\bigr). \end{equation} Now let $$D_1={\mathbf h}_k^{-1}(0)\bigcap (\mathop{\rm cl\hspace{1pt}}(S)-R_k^{\prime})\bigcap \bigl(\bigcup_{i< k}(f^i_k)^{-1}(R_i^{\prime})\bigr) $$ and let $$D_2={\mathbf h}_k^{-1}(0)\bigcap \bigl(\mathop{\rm cl\hspace{1pt}}(S)-R_k^{\prime}\bigr)\bigcap \bigl(\bigcup_{i> k}f^k_i(R_{i,k} ^{\prime})\bigr). $$ Since ${\mathbf h}_{k}(t)$ are small perturbations of $h_k$, we can assume that ${\mathbf h}_k(t)$ are chosen so that for any $t\in[0,1]$ the left hand side of \eqref{eq:1.13} is contained in the union of neighborhood $V_1$ of $D_1$ and a neighborhood $V_2$ of $D_2$. We remark that if we choose $\{e_l\}_{l\leq k}$ so that their $C^2$-norms are sufficiently small, then we can make $V_1$ and $V_2$ arbitrary small. Then by Lemma \ref{1.11} the vanishing locus of ${\mathbf h}_k(t)$ inside $V_1$ is contained in $\cup_{i\leq k}(f^i_k)^{-1}(S_i^{\prime})$. On the other hand, since $\cup_{i\geq k}f_i^k(R_{i,k}^{\prime})$ is open, it contains $V_2$ since $D_2$ is compact and $V_2\supset D_2$ is sufficiently small. This proves the inclusion \eqref{eq:1.12}. Therefore, by induction we have found $\{S_k^{\prime}\}_{k\in{\mathcal L}}$ and $\{e_k(t)\}_{k\in{\mathcal L}}$ that satisfy the condition ${\mathcal H}_k$ for $k=\#({\mathcal L})+1$. Now let ${\mathbf g}_l(t)={\mathbf h}_l(t)|_{R_l^{\prime}}$. Then $\mathfrak g(t)=\{{\mathbf g}_l(t)\}_{l\in{\mathcal L}}$ satisfies the condition of the proposition. Note that the left hand side of \eqref{eq:1.11} is compact because it is contained in the union of compact sets $\{\iota_k(\mathop{\rm cl\hspace{1pt}}(R_k^{\prime}))\}_{k\in{\mathcal L}}$. This proves the proposition. \end{proof} Let $\mathfrak g(t)$ be the perturbation constructed by Proposition \ref{1.12} with ${\mathfrak h}={\boldsymbol \phi}$. We define the Euler class of $[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ to be the homology class in $H_{\ast}({\mathbf B};{\mathbb Q})$ represented by the current $[\mathfrak g(1)^{-1}(0)]$. In the remainder of this section, we will sketch the argument that shows that this class is independent of the choice of the chart ${\mathfrak R}$ and the perturbation $\mathfrak g$. \begin{prop} Let the notation be as before. Then the homology class $[\mathfrak g(1)^{-1}(0)] \in H_{\ast}({\mathbf B};{\mathbb Q})$ so constructed is independent of the choice of perturbations. \end{prop} \begin{proof} First, we show that if we choose two perturbations $\mathfrak g_1(t)$ and $\mathfrak g_2(t)$ based on identical sub-atlas ${\mathfrak R}^{\prime}\Subset{\mathfrak R}$ as stated in Proposition \ref{1.12}, then we have $[\mathfrak g_1(1)^{-1}(0)] =[\mathfrak g_2(1)^{-1}(0)]$. To prove this, all we need is to construct a family of perturbations $\mathfrak g_s(t)$, where $s\in [0,1]$, that satisfies conditions similar to that of the perturbations constructed in Proposition \ref{1.12}. Since then we obtain a current $$\bigcup_{s\in [0,1]}\iota(\mathfrak g_s(1)^{-1}(0))\times\{s\}\subset{\mathbf B}\times [0,1] $$ is a homotopy between the currents $\mathfrak g_0(1)^{-1}(0)_{\text{cur}}$ and $\mathfrak g_1(1)^{-1}(0)_{\text{cur}}$. The construction of $\mathfrak g_s(t)$ is parallel to the construction of $\mathfrak g(t)$ in Proposition \ref{1.12} by considering the data over $\{ R_k\times [0,1]\}_{k\in{\mathcal L}}$. Next, we show that the cycle $[\mathfrak g(1)^{-1}(0)]$ does not depend on the choice of ${\mathfrak R}_1\Subset {\mathfrak R}$. Let ${\mathfrak R}_1\Subset{\mathfrak R}$ and ${\mathfrak R}_2\Subset{\mathfrak R}$ be two good sub-atlas and let $\mathfrak g_1(t)$ and $\mathfrak g_2(t)$ are two perturbations subordinate to ${\mathfrak R}_1$ and ${\mathfrak R}_2$ respectively. Clearly, we can choose a sub-atlas ${\mathfrak R}_0\Subset{\mathfrak R}$ such that ${\mathfrak R}_1\subset{\mathfrak R}_0$ and ${\mathfrak R}_2\subset{\mathfrak R}_0$. Let $\mathfrak g_0(t)$ be a perturbation given by Proposition \ref{1.12} subordinate to ${\mathfrak R}_0$. Then $\mathfrak g_0(t)$ is also subordinate to ${\mathfrak R}_1$ and ${\mathfrak R}_2$. Hence by the previous argument $$[\mathfrak g_1(1)^{-1}(0)]=[\mathfrak g_0(1)^{-1}(0)]=[\mathfrak g_2(1)^{-1}(0)]. $$ It remains to show that the class $[\mathfrak g(1)^{-1}(0)]$ does not depend on the choice of the good atlas ${\mathfrak R}$. For this, it suffices to show that for any two good atlas ${\mathfrak R}$ and ${\mathfrak R}^{\prime}$ so that ${\mathfrak R}$ is finer than ${\mathfrak R}^{\prime}$, the respective perturbations $\mathfrak g(t)$ and $\mathfrak g^{\prime}(t)$ gives rise to identical homology classes $[\mathfrak g(1)^{-1}(0)]=[\mathfrak g^{\prime}(1)^{-1}(0)]$. Let ${\mathfrak R}=\{R_k\}_{k\in{\mathcal K}}$ and ${\mathfrak R}^{\prime}=\{ R_k\}_{k\in{\mathcal L}}$, and let $U_k\subset{\mathbf B}$ (resp. $U_l\subset{\mathbf B}$) be the open subsets so that $(R_k,W_k,\phi_k)$ (resp. $(R_l,W_l,\phi_l)$) are the smooth approximations of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ over $U_k$ (resp. $U_l$) for $k\in{\mathcal K}$ (resp. $l\in{\mathcal L}$). As before, we denote $\iota_k\!:\! R_k\to U_k$ be the tautological map with $k\in{\mathcal K}$ or $k\in{\mathcal L}$. Let $U_{kl}=U_k\cap U_l$. We consider the good atlas ${\mathfrak R}_0$ with charts $R_{kl}=\iota_k^{-1}(U_{kl})$, where $(k,l)\in{\mathcal K}\times{\mathcal L}$, with bundles $W_{kl}=W_k|_{R_{kl}}$ and $\phi_{kl}$ the restriction of $\phi_k$, where $R_{kl}$ is considered to be an open subset of $R_k$. Using the extension technique in the proof of Proposition \ref{1.12}, we can construct a perturbation $\mathfrak g_0(t)$ that is a regular extension of $\mathfrak g^{\prime}(1)^{-1}(0)$ under the obvious $\iota_l^{-1}(U_{kl})\to R_{kl}$ and $W_l|_{\iota_l^{-1}(U_{kl})} \subset W_{kl}$. Therefore, $[\mathfrak g_0(1)^{-1}(0)]=[\mathfrak g^{\prime}(1)^{-1}(0)]$. On the other hand, since $W_{kl}=W_k|_{R_{kl}}$, $\mathfrak g(t)$ induces a perturbation $\mathfrak g_0^{\prime}(t)$ subordinate to ${\mathfrak R}_0$. Hence, $[\mathfrak g_0(1)^{-1}(0)]=[\mathfrak g_0^{\prime}(1)^{-1}(0)]=[\mathfrak g(1)^{-1}(0)]$. This proves the proposition. \end{proof} \section{Analytic charts} The goal of this section is to construct a collection of local smooth approximations $(\Lambda,V)$ so that the data $\phi_V\!:\! R_V\to V|_{R_V}$ are analytic. Namely, $R_V$ are complex manifolds, $V|_{R_V}$ are holomorphic vector bundles and $\phi_V$ are holomorphic sections. In the next section we will show that such $\phi_V$'s are Kuranishi maps, and hence the cones $\lim_{t\to\infty}\Gamma_{t\phi}$ are the virtual cones constructed in \cite{LT1}. We will use the standard notation in complex geometry in this section. For instance, if $M$ is a complex manifold, we will denote by $T_xM$ the complex tangent space of $M$ at $x$ unless otherwise is mentioned. We will use complex dimension throughout this section, unless otherwise is mentioned. Accordingly the complex dimension of a set is half of its real dimension. We will use the words analytic and holomorphic interchangably in this section as well. We begin with the construction of such local smooth approximations. Let $w\in {\mathbf B}$ be any point representing a holomorphic stable map $f\!:\!\Sigma\to X$ with $n$-marked points. We pick a uniformizing chart $\Lambda=(\tilde U,{\mathbf E}_{\tilde U},\Phi_{\tilde U},G_{\tilde U})$ of $w$ over $U\subset{\mathbf B}$ such that the elements of $\tilde U$ are stable maps $f_1\!:\!\Sigma_1\to X$ with (distinct) $(n+k)$-marked points $\{x_i\}$ so that $\{f_1(x_m)\}_{m=n+1}^{n+k}$ are the $k$-distinct points of $f_1^{-1}(H)$, where $H$ is a smooth complex hypersurface of $X$ in general position of degree $k=[H]\cdot [A]$ and $A=f_{\ast}([\Sigma])$, and that the stable maps resulting from discarding the last $k$ marked points of $f_1$ are in $U$. Here as usual we assume that $U$ is sufficiently small so that all stable maps in $U$ intersect $H$ transversally and positively. Note that the later correspondence is the projection $\pi_U\!:\! \tilde U\to U$. Let ${\mathcal Y}$ over $\tilde U$ be the universal (continuous) family of curves with $(n+k)$ marked sections and let ${\mathcal F}\!:\!{\mathcal Y}\to X$ be the universal map. We let $\pi\!:\! \tilde U\to\M_{g,n+k}$ be the tautological map induced by the family ${\mathcal Y}$ with its marked sections. Here $\M_{g,n+k}$ is the moduli space of $(n+k)$-pointed stable curves of genus $g$. Without loss of generality, we can assume that no fibers of ${\mathcal Y}$ with the marked points have non-trivial automorphisms. It follows that $\M_{g,n+k}$ is smooth near $\pi(\tilde U)$. As in section 1, we view $G_{\tilde U}$ as a subgroup of $S_{n+k}$. Then $G_{\tilde U}$ acts on $\M_{g,n+k}$ by permuting the $(n+k)$-marked points of the curves in $\M_{g,n+k}$, and the map $\pi\!:\!\tilde U\to\M_{g,n+k}$ is $G_{\tilde U}$-equivariant. Now let $O\subset\M_{g,n+k}$ be a smooth $G_{\tilde U}$-invariant open neighborhood of $\pi(\tilde U)\subset\M_{g,n+k}$ and let $p\!:\!{\mathcal X}\to O$ be the universal family of stable curves over $O$ with $(n+k)$ marked sections (In this section we will work with the analytic category unless otherwise is mentioned). It follows that the $G_{\tilde U}$-action on $O$ lifts to ${\mathcal X}$ that permutes its marked sections. For convenience, we let ${\mathcal X}\times_O\tilde U$ be the topological subspace of ${\mathcal X}\times\tilde U$ that is the preimage of $\Gamma_{\pi}\subset O\times\tilde U$ under ${\mathcal X}\times\tilde U\to O\times\tilde U$, where $\Gamma_{\pi}\subset O\times\tilde U$ is the graph of $\pi\!:\!\tilde U\to O$. Since no fibers of ${\mathcal Y}$ (with marked points) have non-trivial automorphisms, there is a canonical $G_{\tilde U}$-equivariant isomorphism ${\mathcal Y}\cong {\mathcal X}\times_O\tilde U$ as family of pointed curves. Let $\pi_{{\mathcal X}}$ and $\pi_{\tilde U}$ be the first and the second projection of ${\mathcal X}\times_O\tilde U$. Next, we let $({\mathcal X}_n,O_n;\Sigma,p_n,\varphi_n)$ be a semi-universal family of the $n$-pointed curve $\Sigma$. Namely, ${\mathcal X}_n$ is a (holomorphic) family of pointed prestable curves over the pointed smooth complex manifold $p_n\in O_n$ whose dimension is equal to $\dim_{{\mathbb C}}\Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})$, where $D\subset\Sigma$ is the divisor of the $n$-marked points of $\Sigma$, $\varphi_n\!:\! \Sigma\to{\mathcal X}_n|_{p_n}$ is an isomorphism of $\Sigma$ with the fiber of ${\mathcal X}_n$ over $p_n$ as $n$-pointed curve, and the Kuranishi map $T_{p_n}O_n\to \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})$ of the family ${\mathcal X}_n$ is an isomorphism. Note that $G_{\tilde U}$ acts canonically on $\Sigma$. For convenience, we let $\Pi_n({\mathcal X})$ be the family of curves over $O$ that is derived from ${\mathcal X}$ by discarding its last $k$-marked sections. We now let $B=\pi_U^{-1}(w)$ and fix a $G_{\tilde U}$-equivariant isomorphism \begin{equation} \coprod_{z\in B}\Pi_n({\mathcal X})|_z\lra B\times \Sigma \label{eq:2.50} \end{equation} over $B$. Let $\aut_{p_n}({\mathcal X}_n)$ be the group of those biholomorphisms of ${\mathcal X}_n$ that keep the fiber ${\mathcal X}_n|_{p_n}$ invariant, that send fibers of ${\mathcal X}_n$ to fibers of ${\mathcal X}_n$ and that fix the $n$-sections of ${\mathcal X}_n$. Possibly after shrinking $O_n$ if necessary, we can assume that there is a homomorphism $\rho\!:\! G_{\tilde U}\to \aut_{p_n}({\mathcal X}_n)$ such that for any $\sigma\in G_{\tilde U}$ the $\rho(\sigma)$ action on ${\mathcal X}_n|_{p_n}$ is exactly the $\sigma$ action on $\Sigma$ via the isomorphism $\varphi_n$. Finally, possibly after shrinking $U$ and $O$, we can pick a $G_{\tilde U}$-equivariant holomorphic map $\varphi\!:\! O\to O_n$ such that $\varphi(B)=p_n$ and that there is a $G_{\tilde U}$-equivariant isomorphism of $n$-pointed curves $\tilde{\varphi}\!:\!{\mathcal X}\to O\times_{O_n}{\mathcal X}_n$ that extends the isomorphism \eqref{eq:2.50}. We remark that the reason for doing this is to ensure that the smooth approximation we are about to construct is $G_{\tilde U}$-equivariant. Next, we let $l$ be an integer to be specified later and let $U_i\subset\Sigma$, $i=1,\ldots,l$, be $l$ disjoint open disks away from the marked points and the nodal points of $\Sigma$. We assume that $\cup_{i=1}^l U_i$ is $G_{\tilde U}$-invariant and that for any $\sigma\in G_{\tilde U}$ whenever $\sigma(U_i)=U_i$ then $\sigma|_{U_i}={\mathbf 1}_{U_i}$. By shrinking $U$, $O$ and $O_n$ if necessary, we can find disjoint open subsets ${\mathcal U}_{n,i}\subset{\mathcal X}_n$ such that $\cup_{i=1}^l{\mathcal U}_{n,i}$ is $G_{\tilde U}$-invariant, that $\cup_{i=1}^l{\mathcal U}_{n,i}$ is $G_{\tilde U}$-equivariantly biholomorphic to $O\times\cup_{i=1}^l U_i$, that ${\mathcal U}_{n,i}\cap\Sigma=U_i$ and that the projections ${\mathcal U}_{n,i}\to O$ induced by the projection ${\mathcal X}\to O$ is the first projection of $O\times U_i$ ($={\mathcal U}_{n,i}$). For convenience, for each $i$ we will fix a biholomorphism between $U_i$ and the unit disk in ${\mathbb C}$, and will denote by $U_i^{\hf}$ the open disk in $U_i$ of radius $1/2$. We let ${\mathcal U}_i$ be the disjoint open subsets of ${\mathcal Y}$ defined by $${\mathcal U}_i={\mathcal U}_{n,i}\times_{O_n}\tilde U\subset{\mathcal X}_n\times_{O_n}\tilde U \cong{\mathcal X}\times_O\tilde U\cong{\mathcal Y}. $$ We will call $U_i$ and ${\mathcal U}_i$ the distinguished open subsets of $\Sigma$ and ${\mathcal Y}$ respectively. Without loss of generality, we can assume that $\cup_{i=1}^l{\mathcal U}_i$ is disjoint from the $(n+k)$-sections of ${\mathcal Y}$. We also assume that there are holomorphic coordinate charts $V_i\subset X$ so that ${\mathcal F}({\mathcal U}_i)\subset V_i$. We let $(w_{i,1},\ldots,w_{i,m})$, where $m=\dim X$, be the coordinate variable of $V_i$ and let ${\mathbf v}_i=\partial/\partial w_{i,1}$. For each $i$ we pick a nontrivial $(0,1)$-form $\gamma_i$ on $U_i$ with $\supp(\gamma_i)\Subset U_i^{\hf}$. We demand further that if there is a $\sigma\in G_{\tilde U}$ so that $\sigma(U_i)=U_j$ then $V_i=V_j$ as coordinate chart and $\sigma^{\ast}(\gamma_i)=\gamma_j$. We then let $\sigma_i$ be the $(0,1)$-form over ${\mathcal U}_i$ with values in ${\mathcal F}^{\ast}(T_X)|_{{\mathcal U}_i}$ that is the product of the pull back of $\gamma_i$ via ${\mathcal U}_i\times_O\tilde U\to U_i$ with ${\mathcal F}^{\ast}({\mathbf v}_i)|_{{\mathcal U}_i}$, and let $\tilde{\sigma}_i$ be the section over ${\mathcal Y}$ that is the extension of ${\sigma}_i$ by zero. Obviously, $\tilde{\sigma}_i$ is a section of ${\mathbf E}_{\tilde U}$, and $(\tilde{\sigma}_1,\ldots,\tilde{\sigma}_l)$ is linearly independent fiberwise. Hence it spans a complex subbundle of ${\mathbf E}_{\tilde U}$, denoted by $V$. It follows from the construction that $V$ is $G_{\tilde U}$-equivariant. As in the previous section, we let $R=\Phi_{\tilde U}^{-1}(V)$, let $W=V|_R$ and let $\phi\!:\! R\to W$ be the lifting of $\Phi_{\tilde U}|_R\!:\! R\to {\mathbf E}_{\tilde U}|R$. The main task of this section is to show that we can choose $U_i$, $\gamma_i$ and $V_i$ so that $R$ admits a canonical complex structure and that the section $\phi$ is holomorphic when $W$ is endowed with the holomorphic structure so that the basis $\tilde{\sigma}_1|_R,\ldots,\tilde{\sigma}_l|_R$ is holomorphic. To specify our choice of $U_i$, $\gamma_i$ and $V_i$, we need first to define the Dolbeault cohomology of holomorphic vector bundles over singular curves. Let ${\mathcal E}$ be a locally free sheaf of ${\mathcal O}_{\Sigma}$-modules and let $E$ be the associated vector bundle, namely, ${\mathcal O}_{\Sigma}(E)={\mathcal E}$. We let $\Omega^0_{{\rm cpt}}(E)$ be the sheaf of smooth sections of $E$ that are holomorphic in a neighborhood of $\sing(\Sigma)$ and let $\Omega_{{\rm cpt}}^{0,1}(E)$ be the sheaf of smooth sections of $(0,1)$-forms with values in $E$ that vanish in a neighborhood of $\sing(\Sigma)$. Let $$\bar\partial: \Gamma(\Omega^0_{{\rm cpt}}(E))\lra \Gamma( \Omega_{{\rm cpt}}^{0,1}(E)) $$ be the complex that send $\varphi\in \Omega^0_{{\rm cpt}}(E))$ to $\bar\partial(\varphi)$. Since $\varphi$ is holomorphic near nodes of $\Sigma$, $\bar\partial(\varphi)$ vanishes near nodes of $\Sigma$ as well, and hence the above complex is well defined. We define the Dolbeault cohomology $H^0_{\bar\partial}(E)$ and $H^1_{\bar\partial}(E)$ to be the kernel and the cokernel of $\bar\partial$. \begin{lemm} \label{2.0} Let $H^i({\mathcal E})$ be the C\v{e}ch cohomology of the sheaf ${\mathcal E}$. Then there are canonical isomorphisms $H^0_{\bar\partial}(E)\cong H^0({\mathcal E})$ and $\Psi\!:\! H^{0,1}_{\bar\partial}(E)\cong H^1({\mathcal E})$. \end{lemm} \begin{proof} The proof is identical to the proof of the classical result that the Dolbeault cohomology is isomorphic to the C\v{e}ch cohomology for smooth complex manifolds. Obviously, $H^0_{\bar\partial}(E)$ is canonically isomorphic to $H^0({\mathcal E})$. We now construct $\Psi$. We first cover $\Sigma$ by open subsets $\{W_i\}$ so that the intersection of any of its subcollection is contractible. Now let $\varphi$ be any global section in $\Omega^{0,1}_{{\rm cpt}}(E)$. Then over each $W_i$ we can find a smooth function $\eta_i\in\Gamma_{W_i}(\Omega_{{\rm cpt}}^0(E))$ such that $\bar\partial\eta_i=\varphi|_{W_i}$. Clearly, the class in $H^1({\mathcal E})$ represented by the cocycle $[\eta_{ij}]$, where $\eta_{ij}=\eta_i|_{W_i\cap W_j}- \eta_j|_{W_i\cap W_j}$, is independent of the choice of $\eta_i$, and thus defines a homomorphism $\Gamma(\Omega_{{\rm cpt}}^{0,1}(E))\to H^1({\mathcal E})$. It is routine to check that it is surjctive and its kernel is exactly $\image(\overline{{\mathbf\partial}})$. Therefore, we have $H^{0,1}_{\bar\partial}(E)\cong H^1({\mathcal E})$. Also, it is direct to check that this isomorphism does not depend on the choice of the covering $\{W_i\}$. This proves the lemma. \end{proof} For any $z\in\tilde U$, we denote by $\tilde{\sigma}_i(z)$ the restriction of $\tilde{\sigma}_i$ to the fiber of ${\mathcal Y}$ over $z$. We now choose the $l$ open disks $U_i\subset \Sigma$, the $(0,1)$-forms $\gamma_i$ on $U_i$ and the coordinate charts $V_i\subset X$ such that for any $\tilde w\in\pi_U^{-1}(w)$ the collection $\tilde{\sigma}_1(\tilde w),\ldots, \tilde{\sigma}_l(\tilde w)$ spans $H^{0,1}_{\bar\partial}(f^{\ast} T_X)$. This is certainly possible if we choose $l$ large because the locus of $U_i$ are arbitrary as long as they are away from the nodal points of $\Sigma$ and the marked points, and the charts $V_i$ can also be chosen with a lot of choice. We fix once and for all such choices of $U_i$, $V_i$ and $\gamma_i$. We then let ${\mathcal U}_i\subset{\mathcal Y}$, $V\to{\mathbf E}_{\tilde U}$ and $R=\Phi_{\tilde U}^{-1}(V)$ be the objects constructed before according to this choice of $U_i$, $\gamma_i$ and $V_i$. Let ${\mathcal Y}_R\to R$ be the restriction to $R\subset\tilde U$ of the family ${\mathcal Y}\to\tilde U$ with the marked sections and let $F\!:\! {\mathcal Y}_R\to X$ be the associated map. We also fix a smooth function $\eta_i$ over $U_i$ so that $\bar\partial\eta_i=\gamma_i$. We next extend the collection $\{ U_i\}_{i=1}^l$ to an open covering $\{U_i\}_{i=1}^L$ so that the intersection of any subcollection of $\{U_i\}$ are contractible, and that for any $i\leq l$ and $j\geq l+1$ the sets $U_i^{\hf}$ and $U_j$ are disjoint. For convenience, we agree that $\eta_j=0$ for $j>l$ From now on, we will fix an $\tilde w\in R$ over $w$. \begin{lemm}\label{2.1} There is a constant $A$ such that for any C\v{e}ch 1-cocycle $[\tau_{ij}]$, where $\tau_{ij}\in\Gamma_{U_i\cap U_j}(f^{\ast} {\mathcal T}_X)$, there are constants $a_i$ and holomorphic sections $\zeta_i\in\Gamma_{U_i}(f^{\ast} {\mathcal T}_X)$ for $i=1,\ldots,L$ such that $$(\zeta_j+a_j\eta_j)|_{U_j\cap U_i}-(\zeta_i+a_i\eta_i)|_{U_j\cap U_i} =\tau_{ji} $$ and $$\sum_{i=1}^L\bigl(\parallel\!\zeta_i\!\parallel_{L_2}+|a_i|\bigr)\leq A\bigl(\sum_{i,j}^L\parallel\!\tau_{ij}\!\parallel_{L_2}\bigr). $$ \end{lemm} \begin{proof} The existence of $\{a_i\}$ and $\{\zeta_i\}$ follows from the fact that the images of $\tilde\sigma_1(\tilde w), \ldots,\tilde\sigma_l(\tilde w)$ spans $H_{\bar\partial}^{0,1}(f^{\ast} T_X)$ and that $H_{\bar\partial}^{0,1}(f^{\ast} T_X)$ is isomorphic to $H^1(f^{\ast}{\mathcal T}_X)$. The elliptic estimate is routine, using the harmonic theory on the normalization of $\Sigma$. We will leave the details to the readers. \end{proof} We let $\Gamma(\Omega_{{\rm cpt}}^{0,1}(f^{\ast} T_X))^{\dag}$ be the quotient of $\Gamma(\Omega_{{\rm cpt}}^{0,1}(f^{\ast} T_X))$ by the linear span of $\tilde{\sigma}_1(\tilde w),\ldots,\tilde{\sigma}_l(\tilde w)$. Because $\{\tilde\sigma_i(\tilde w)\}_{i=1}^L$ is invariant under the automorphism group of the stable map $f$, $\Gamma(\Omega_{{\rm cpt}}^{0,1}(f^{\ast} T_X))^{\dag}$ is independent of the choice of $\tilde w\in\pi_U^{-1}(w)$. We let $${\bar\partial}^{\dag}\!:\! \Gamma(\Omega_{{\rm cpt}}^0(f^{\ast} T_X))\to \Gamma(\Omega_{{\rm cpt}}^{0,1}(f^{\ast} T_X))^{\dag} $$ be the induced complex. We define $H^0_{\bar\partial}(f^{\ast})^{\dag}$ and $H^{0,1}_{\bar\partial}(f^{\ast} T_X)^{\dag}$ be the kernel and the cokernel of the above complex. \begin{coro}\label{2.2} Let the notation be as before. Then $H^{0,1}_{\bar\partial}(f^{\ast} T_X)^{\dag}=0$. Further, the complex dimension of $H^0_{\bar\partial}(f^{\ast} T_X)^{\dag}$ is $\deg(f^{\ast} T_X)+m(1-g)+l$. \end{coro} \begin{proof} The vanishing of $H_{\bar\partial}^{0,1}(f^{\ast} T_X)^{\dag}$ follows from the surjectivity of ${\bar\partial}^{\dag}$. The second part follows from $$\begin{array}{ll} &\dim H^0_{\bar\partial}(f^{\ast} T_X)^{\dag}-\dim H^{0,1}_{\bar\partial}(f^{\ast} T_X)^{\dag}\\ =&\dim H^0_{\bar\partial}(f^{\ast} T_X)-\dim H^{0,1}_{\bar\partial}(f^{\ast} T_X)+l =\chi(f^{\ast} T_X)+l \end{array} $$ and the Riemann-Roch theorem. \end{proof} Next, we will describe the tangent space of $R$ at $\tilde w$. By the smoothness result of \cite{LT2}, we know that $R$ is a smooth manifold of (complex) dimension $r_{\rm exp }$. As before, we let $D\subset \Sigma$ be the divisor of the first $n$-marked points of $\tilde w$. Since $f$ is holomorphic, $df^{\vee}$ is a homomorphism of sheaves $f^{\ast}\Omega_X\to\Omega_{\Sigma}$. We let $${\mathcal D}_{\tilde w}^{\bullet}=[f^{\ast}\Omega_X \xrightarrow{\alpha}\Omega_{\Sigma}(D)] $$ be the induced complex indexed at $-1$ and $0$. We will first define the extension space $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$ and then show that it is canonically isomorphic to $T_{\tilde w}R$. We begin with some more notations. Let ${\mathcal E}$ be a sheaf of ${\mathcal O}_{\Sigma}$-modules that is locally free away from the nodal points of $\Sigma$. Then there is a holomorphic vector bundle $E$ over $\Sigma^0$, where $\Sigma^0$ is the smooth locus of $\Sigma$, such that ${\mathcal O}_{\Sigma^0}(E) ={\mathcal E}|_{\Sigma^0}$. We define ${\mathcal E}^{\cA}$ to be the sheaf so that the germs of ${\mathcal E}^{\cA}$ at nodal points $p\in\Sigma$ (resp. smooth points $p\in\Sigma^0$) are isomorphic to the germs of ${\mathcal E}$ at $p$ (resp. germs of $\Omega_{\Sigma^0}^0(E)$ at $p$). The set $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$ is the set of equivalence classes of pairs $(v_1,v_2)$ as follows. The data $v_1$ is an element in $\Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})$, which defines an exact sequence \begin{equation} \begin{CD} 0@>>>{\mathcal O}_{\Sigma}@>{\varphi_1}>> {\mathcal B}@>{\varphi_2}>>\Omega_{\Sigma}(D)@>>> 0, \end{CD} \label{eq:2.2} \end{equation} and equivalently a family of $n$-pointed nodal curves over $T=\spec {\mathbb C}[t]/(t^2)$, say ${\mathcal C}_T$ with $n$-marked sections $\tilde{x}_i$ (See \cite[section 1]{LT1}). Note that ${\mathcal B}$ is locally free over $\Sigma^0$. The data $v_2$ is a homomorphism $f^{\ast}\Omega_X\to {\mathcal B}^{\cA}$ such that, first of all, the diagram \begin{equation} \begin{CD} @. @. f^{\ast}\Omega_X @= f^{\ast}\Omega_X\\ @. @. @Vv_2VV @V{df^{\vee}}VV\\ 0 @>>> {\mathcal O}_{\Sigma}^{\cA} @>{\varphi_1}>> {\mathcal B}^{\cA} @>{\varphi_2}>> \Omega_{\Sigma}(D)^{\cA} @>>> 0 \end{CD} \label{eq:2.3} \end{equation} is commutative, where the lower sequence is induced by \eqref{eq:2.2}. Secondly, since $v_2$ is holomorphic near nodes of $\Sigma$, the differential $\bar\partial v_2$ vanishes near nodes of $\Sigma$, and since $df^{\vee}$ is holomorphic, $\bar\partial v_2$ lifts to a global section $\beta$ of $\Omega_{\rm cpt}^{0,1}(f^{\ast} T_X)$. We require that there are constants $a_1,\ldots,a_l$ such that $\beta=\sum a_i\tilde\sigma_i(\tilde w)$. The equivalence relation of such pairs are the usual equivalence relation of the diagrams \eqref{eq:2.3}. Namely, two pairs $(v_1,v_2)$ and $(v^{\prime}_1,v_2^{\prime})$ with the associated data $\{{\mathcal B},\varphi_i\}$ and $\{{\mathcal B}^{\prime},\varphi_i^{\prime}\}$ are equivalent if there is an isomorphism $\eta\!:\!{\mathcal B}\to{\mathcal B}^{\prime}$ so that $\eta\circ\varphi_1=\varphi_1^{\prime}$, $\varphi_2=\varphi_2^{\prime}\circ\eta$ and $\eta\circ v_2=v_2^{\prime}$. \begin{lemm}\label{2.3} Let the notation be as before. Then $\Ext^1_{\Sigma}({\mathcal D}_{\tilde w}^{\bullet},{\mathcal O}_{\Sigma}^{\cA})^{\dag}$ is canonically a complex vector space of complex dimension $r_{\rm exp }$. \end{lemm} \begin{proof} The fact that $\Ext^1_{\Sigma}({\mathcal D}_{\tilde w}^{\bullet},{\mathcal O}_{\Sigma}^{\cA})^{\dag}$ forms a complex vector space can be established using the usual technique in homological algebra. For instance, if $r\in \Ext^1_{\Sigma}({\mathcal D}_{\tilde w}^{\bullet},{\mathcal O}_{\Sigma}^{\cA})^{\dag}$ is represented by $\{{\mathcal B},\varphi_i,v_2\}$ shown in the diagram \eqref{eq:2.3}, then for any complex number $a$ the element $ar$ is represented by the same diagram with $\varphi_1$ replaced by $a\varphi_1$. We now prove that \begin{equation} \dim\Ext^1_{\Sigma}({\mathcal D}_{\tilde w}^{\bullet},{\mathcal O}_{\Sigma}^{\cA})^{\dag}=r_{\rm exp }. \label{eq:2.4} \end{equation} Clearly, the following familiar sequence is still exact in this case: $$ \Ext^0(\mathcal D^{\bullet}_f,{\mathcal O}_{\Sigma}^{\cA})\lra \Ext^0(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma}) \lra H^0_{\bar\partial}(f^{\ast} T_X)^{\dag}\qquad\qquad\qquad $$ $$ \qquad \lra \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})\lra \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})^{\dag} \lra H^{0,1}_{\bar\partial}(f^{\ast} T_X)^{\dag}. $$ Since $f$ is stable, $\Ext^0(\mathcal D^{\bullet}_f,{\mathcal O}_{\Sigma}^{\cA})=0$. Hence \eqref{eq:2.4} follows from Corollary \ref{2.2} and the Riemann-Roch theorem. This proves the lemma. \end{proof} Recall that should $R$ be a scheme then the Zariski tangent space of $T_{\tilde w}R$ would be the space of morphisms $\spec {\mathbb C}[t]/(t^2)\to R$ that send their only closed points to $\tilde w$ modulo certain equivalence relation. In the following, we will imitate this construction and construct the space of pre-${\mathbb C}$-tangents of $R$ at $\tilde w$. We still denote by $U_1,\ldots, U_l$ the $l$-distinguished open subsets of $\Sigma$ and let $\{U_i\}_{i=1}^L$ be an extension of $\{U_i\}_{i=1}^l$ to an open covering of $\Sigma$ such that the intersection of any of its subcollection are contractible. Without loss of generality, we assume $U_j\cap U_i^{\hf}=\emptyset$ for $j>l$ and $i\leq l$. We also assume that there are coordinate charts $V_i$ of $X$ such that $f(U_i)\subset V_i$. By abuse of notation, we will fix the embedding $V_i\subset {\mathbb C}^m$ and view any map to $ V_i$ as a map to ${\mathbb C}^m$. We let $\iota_i\!:\! V_i\to X$ be the tautological inclusion and let $$g_{ij}\!:\! \iota_j^{-1}(\iota_i(V_i))\to \iota_i^{-1}(\iota_j(V_j)) \subset V_i\subset {\mathbb C}^m $$ be the transition functions of $X$. We define a pre-${\mathbb C}$-tangent $\xi$ of $R$ at $\tilde w$ to be a collection of data as follows: First, there is a flat analytic family of $n$-pointed pre-stable curves $C_T$ over an open neighborhood $T$ of $0\in {\mathbb C}$ such that the fiber of $C_T$ over $0$, denoted by $C_0$, is isomorphic to $\Sigma$ as $n$-pointed curve; Secondly, there is an open covering $\{\tilde U_i\}_{i=1}^L$ of $C_T$ such that $\tilde U_i\cap C_0=U_i$, and that for each $i\leq l$, there is a biholomorphism $\tilde U_i\cong U_i\times T$ such that its restriction to $U_i=\tilde U_i\cap C_0$ is compatible to the identity map of $U_i$; Thirdly, there is a collection of smooth maps $\tilde f_i\!:\! \tilde U_i\to V_i$ such that for $i>l$, all $\tilde f_i$ are holomorphic and that for each $i\leq l$ we have ${\bar\partial}_0(\tilde f_i)=0$ and \begin{equation} {\bar\partial}_i(\tilde f_i)= \pi_T^{\ast}\varphi_i \cdot\pi_{U_i}^{\ast}(\gamma_i\cdot f^{\ast}({\mathbf v}_i)), \label{eq:2.22} \end{equation} where $\pi_{U_i}$ and $\pi_T$ are the first and the second projection of $U_i\times T$, $\varphi_i$ are holomorphic functions over $T$ and $\bar\partial_0$ (resp. $\bar\partial_i$) is the $\bar\partial$-differential with respect to the holomorphic variable of $T$ (resp. $U_i$) using $\tilde U_i\cong U_i\times T$ and the $\gamma_i$ and ${\mathbf v}_i$ are the $(0,1)$-form and the vector field chosen before; Forthly, if we let $z_0$ be the holomorphic variable of ${\mathbb C}\supset T$, then we require that \begin{equation} \label{eq:2.27} \tilde f_{ji}=\tilde f_i- g_{ij}\circ \tilde f_j: \tilde U_{ij}\lra {\mathbb C}^m, \end{equation} where $\tilde U_{ij}$ is a neighborhood of $U_i\cap U_j$ in $\tilde U_i\cap \tilde U_j$ over which $\tilde f_{ji}$ is well-defined, is divisible by $z^2_0$ (Namely, $\tilde f_{ji}$ has the form $\pi_T^{\ast}(z^2_0)\cdot h_{ji}$ for some smooth function $h_{ji}\!:\! \tilde U_{ij}\to {\mathbb C}^m$). Intuitively, a pre-${\mathbb C}$-tangent is a scheme analogue of a morphism $\spec {\mathbb C}[t]/(t^2)\to R$ should $R$ be a scheme. We denote the set of all pre-${\mathbb C}$-tangents by $T_{\tilde w}^{{\rm pre}} R$. Note that $T_{\tilde w}^{{\rm pre}} R$ is merely a collection of all pre-${\mathbb C}$-tangents. We next define a canonical map \begin{equation} T_{\tilde w}^{{\rm pre}} R\lra \Ext^1({\mathcal D}_{\tilde w}^{\bullet},{\mathcal O}_{\Sigma})^{\dag}. \label{eq:2.26} \end{equation} Let $\xi$ be any pre-${\mathbb C}$-tangent given by the data above. By the theory of deformation of $n$-pointed curves \cite[section 1]{LT1}, the analytic family $C_T$ defines canonically an exact sequence \begin{equation} 0\lra{\mathcal O}_{\Sigma}\lra {\mathcal B}\lra \Omega_{\Sigma}(D)\lra 0, \label{eq:2.23} \end{equation} associated to an extension class $v_1(\xi)\in\Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})$, where away from the nodes of $\Sigma$ and the suport of $D$ the sheaf ${\mathcal B}$ is canonically isomorphic to $\Omega_{C_T}\otimes_{{\mathcal O}_{C_T}}{\mathcal O}_{C_0}$. Because $\tilde f_i\!:\! \tilde U_i\to V_i$ are holomorphic for $i>l$, it follows from \cite{LT1} that there is a canonical homomorphism of sheaves $u_i\!:\! f^{\ast}\Omega_X|_{U_i} \to {\mathcal B}|_{U_i}$ such that \begin{equation} \begin{CD} @. @. f^{\ast}\Omega_X|_{U_i} @= f^{\ast}\Omega_X|_{U_i}\\ @. @. @VV{u_i}V @VV{df^{\vee}|_{U_i}}V\\ 0 @>>> {\mathcal O}_{U_i}^{\cA} @>>> {\mathcal B}^{\cA}|_{U_i} @>>> \Omega_{U_i}(D)^{\cA} @>>> 0\\ \end{CD} \end{equation} is commutative, where the lower sequence is induced by \eqref{eq:2.23}. Indeed, at smooth point $p\in U_i$ away from the support of $D$ the dual of $u_i\otimes k(p)$ is the differential $$d\tilde f_i(p) : T_p C_T={\mathcal B}^{\vee}\otimes k(p)\lra f^{\ast} T_X|_p. $$ Note that by our choice of $U_i$, for $i\leq l$ the distinguished open subsets $U_i$ are disjoint from the support of the $(n+k)$-marked points of $\tilde w$. Hence ${\mathcal B}^{\cA}|_{U_i}$ are canonically isomorphic to $\Omega_{U_i}^0(TC_T|_{U_i})$, and the dual of $d\tilde f_i$ define canonical homomorphisms $u_i\!:\! f^{\ast}\Omega_X|_{U_i}\to{\mathcal B}|_{U_i}$ that make the above diagrams commutative. Because of the condition \eqref{eq:2.22}, the lift of $\bar\partial u_i$ is a constant multiple of $\tilde\sigma_i(\tilde w)|_{U_i}$. Further, because of the condition that $\tilde f_{ji}$ is divisible by $z_0^2$, the collection $\{ u_i\}_{i=1}^L$ patch together to form a homomorphism $v_2(\xi)\!:\! f^{\ast}\Omega_X\to{\mathcal B}^{\cA}$ that makes the diagram \eqref{eq:2.3} commutative. Hence $(v_1(\xi),v_2(\xi))$ defines an element in $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$, which is defined to be the image of $\xi$. We remark that in this construction we have only used the fact that the stable map associated to $\tilde w$ is holomorphic, that the domain $\Sigma$ of $\tilde w$ has $l$ distinguished open subsets $U_i$ with $(0,1)$-forms $\tilde\sigma_i(\tilde w)$. Because for any $z\in R$ its domain $\Sigma_z$ also has $l$ distinguished open subsets, namely ${\mathcal U}_i\cap \Sigma_z\cong U_i$, and the forms $\tilde\sigma(z)$, we can define the extension group $\Ext^1({\mathcal D}^{\bullet}_z,{\mathcal O}_{\Sigma_z})^{\dag}$, the space of pre-${\mathbb C}$-tangents of $R$ at $z$ and the analogut canonical map as in \eqref{eq:2.26} if the map $f_z$ of $z$ is holomorphic. To justify our choice of $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$, we will construct, to each $v\in\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$, a pre-${\mathbb C}$-tanget $\xi^v\in T_{\tilde w}^{{\rm pre}} R$ so that the image of $\xi^v$ under \eqref{eq:2.26} is $v$. Let $v=(v_1,v_2)$ be any element in $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$ defined by the diagram \eqref{eq:2.3}. Let $T\subset {\mathbb C}$ be a neighborhood of $0$ and let $C_T$ be an analytic family of $n$-pointed curves so that $C_0\cong \Sigma$ and the Kuranishi map $T_0{\mathbb C}\to\Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma})$ send $1$ to $v_1$. For instance, we can take $C_T$ be the pull back of ${\mathcal X}_n$ via an analytic map $(T,0)\to (O_n,p_n)$. We let $\{U_i\}_{i=1}^L$ be a covering of $\Sigma$ as before and let $\{\tilde U_i\}_{i=1}^L$ be a covering of $C_T$ that are the pull back of ${\mathcal U}_{n,i}$. Note that for $i\leq l$, they come with biholomorphisms $\tilde U_i\cong U_i\times T$. Let $V_i$ be open charts of $X$ as before with $f(U_i)\Subset V_i$. For $i>l$, since the restriction of \eqref{eq:2.3} to $U_i$ is analytic, we can find analytic $\tilde f_i\!:\! \tilde U_i\to V_i$, possibly after shrinking $T$ if necessary, such that $\tilde f_i$ are related to $v_2|_{U_i}$ as to how $u_i$ are related to $v_2(\xi)|_{U_i}$ before. By analytic analogue of deformation theory (see \cite{LT1}) such $\tilde f_i$ do exist. For $i\leq l$, since $U_i$ are smooth and ${\mathcal B}^{\cA}|_{U_i}$ are the sheaves $\Omega_{U_i}^0(T^{\ast} C_T|_{U_i})$, we simply let $\tilde f_i$ be smooth so that in addition to $\tilde f_i$ satisfying the condition on pre-${\mathbb C}$-tangents we require that $v_2|_{U_i}$ coincide with the dual of $d\tilde f_1|_{U_i}$. Note that $(C_T,\{\tilde f_i\})$ will be a pre-${\mathbb C}$-tangent if $\tilde f_{ji}$ in \eqref{eq:2.27} is divisible by $\pi_T^{\ast}(z_0^2)$. But this is true because for any $p\in U_i\cap U_j$, the differential $d\tilde f_i(p)$ and $d\tilde f_j(p)$ from $T_p C_T$ to $T_{f(p)}X$ are identical. We let the so constructed pre-${\mathbb C}$-tangent be $\xi^v$. Of course $\xi^v$ are not unique. It is obvious from the construction that the image of $\xi^v$ under \eqref{eq:2.26} is $v$. We remark that it follows from the construction that for any complex number $c\ne 0$ the pull back of $(C_T,\{\tilde f_i\})$ under $L_c \!:\! {\mathbb C}\to {\mathbb C}$ defined by $L_c(z_0)=cz_0$ is a pre-${\mathbb C}$-tangent, say $\xi^{cv}$, whose image under \eqref{eq:2.26} is $cv$. We next construct a holomorphic coordinate chart of $R$ at $\tilde w$. Let $r=\dim R$, which is $r_{\rm exp }+l=\dim\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$. We fix a ${\mathbb C}$-isomorphism $T_0{\mathbb C}^r\cong\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$. Composed with the canonical $$\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}\to \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma}), $$ we obtain \begin{equation} T_0{\mathbb C}^r\lra \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma}). \label{eq:2.20} \end{equation} Let ${\mathcal X}_n$ over $O_n$ be the semi-universal family of the $n$-pointed curve $\Sigma$ given before. We let $S$ be a neighborhood of $0\in{\mathbb C}^r$ and let $\varphi\!:\! S\to O_n$ be a holomorphic map with $\varphi(0)=0$ such that $$d\varphi(0): T_0S\equiv T_0{\mathbb C}^r\lra T_{p_n}O_n\cong \Ext^1(\Omega_{\Sigma}(D),{\mathcal O}_{\Sigma}) $$ is the homomorphism \eqref{eq:2.20}. We let $\pi_S\!:\! C_S\to S$ be the family of $n$-pointed curves over $S$ that is the pull back of ${\mathcal X}_n$. Note that $C_S|_0$, denoted by $C_0$, is canonically isomorphic to $\Sigma$. We keep the open covering $\{U_i\}^L_{i=1}$ of $\Sigma$ ($\cong C_0$) chosen before. We let $\{W_i\}_{i=1}^L$ be an open covering of a neighborhood of $C_0\subset C_S$ so that $W_i\cap C_0=U_i$. For $i\leq l$, we let $W_i$ be the pull back of ${\mathcal U}_{n,i}\subset {\mathcal X}_n$. For $i>l$ and $U_i$ smooth, we choose $W_i$ so that there is a holomorphic map $\pi_i\!:\! W_i\to U_i$ so that the restriction of $\pi_i$ to $U_i$ is the identity map. For $i>l$ and $U_i$ contains a nodal point, we assume that $W_i$ is biholomorphic to the unit ball in ${\mathbb C}^{r+1}$ so that $U_i\subset W_i$ is defined by $w_1w_2=0$ and $w_i=0$ for $i\geq 3$, where $(w_i)$ are the coordinate variables of ${\mathbb C}^{r+1}$, and the restriction of $\pi_S$ to $W_i$ is given by $$z_1=w_1w_2, \ z_2=w_3,\ldots,z_r=w_{r+1}, $$ where $(z_i)$ are the coordinate variables of ${\mathbb C}^r$. The upshot of this is that if $h$ is a holomorphic function on $U_i$, then we can extend it canonically to $W_i$ as follows. In case $U_i$ is smooth, then the extension of $h$ is the composite of $W_i\to U_i$ with $h$; In case $U_i$ is singular, then $\varphi$ has a unique expression $$a+w_1h_1(w_1)+w_2h_2(w_2), $$ where $a\in{\mathbb C}$ and $h_1, h_2$ are holomophic. We then let its extension be the holomorphic function on $W_i$ that has the same expression. We fix the choice of $\{U_i\}$ and $\{W_i\}$. Without loss of generality, we can assume that there are coordinate charts $V_i\subset X$ so that $f(U_i)\Subset V_i$. Of course, for $i\leq l$ the charts $V_i$ are the charts we have chosen before. Our construction of the local holomorphic chart of $R$ is parallel to the original construction of Kodaira-Spencer of semi-universal family of deformation of holomorphic structures without obstructions. To begin with, possibly after shrinking $W_i$ if necessary we can assume that the maps $f|_{U_i}\!:\! U_i\to V_i$ can be extended to a holomorphic $F_{0,i}\!:\! W_i\to V_i$ (Recall $f$ is holomorphic). We now let ${\mathcal A}(W_i,V_i)$ be the space of smooth maps from $W_i$ to ${\mathbb C}^m$ defined as follows. If $i>l$, then ${\mathcal A}(W_i,V_i)$ consists of holomorphic maps from $W_i$ to ${\mathbb C}^m$; If $i\leq l$, then using the isomorphism $W_i\cong U_i\times S$ and holomorphic coordinate $z=(z_i)$ of $S$ and holomorphic coordinate $\xi$ of $U_i$, any smooth function $\varphi\!:\! W_i\to {\mathbb C}^m$ can be expressed in terms of its $m$ components $\varphi_j(z, \xi)$, $j=1,\ldots,m$. We define ${\mathcal A}(W_i,V_i)$ to be the set of those smooth maps $\varphi\!:\! W_i\to{\mathbb C}^m$ so that $$\left\{ \begin{array}{l} \bar\partial_{z_k}\varphi_j=0\quad \text{for}\quad k=1,\ldots,r \quad{\rm and}\quad j=1,\ldots,m;\\ \bar\partial_{\xi}\varphi_j=0\quad \text{for}\quad j\geq 2 \quad{\rm and}\quad \bar\partial_{\xi}\varphi_1=c\sigma_i^{\prime}\quad \text{for some}\quad c\in{\mathbb C}, \end{array} \right. $$ where $\sigma^{\prime}_i$ is a $(0,1)$-form taking values in $\varphi^{\ast} {\mathbb C}^n$ corresponding to the form $\sigma_i$ using the canonical embedding $V_i\subset{\mathbb C}^n$. Note that ${\mathcal A}(W_i,V_i)$ are ${\mathcal O}_S$-modules. In particular, if we let ${\mathcal I}\subset {\mathcal O}_S $ be the ideal sheaf of $0\in S$, then we denote by ${\mathcal I}^q {\mathcal A}(W_i,V_i)$ the image of ${\mathcal I}^q\otimes_{{\mathcal O}_S }{\mathcal A}(W_i,V_i)$ in ${\mathcal A}(W_i,V_i)$. In the following, we will construct a sequence of maps $F_{s,i}\in{\mathcal A}(W_i, V_i)$ indexed by $s\geq 1$ and $1\leq i\leq L$ of which the following holds: \noindent 1. For each $i$, $F_{s+1,i}-F_{s,i}\in {\mathcal I}^{s}{\mathcal A}(W_i, V_i)$; \noindent 2. The restrictions $F_{1,i}|_{U_i}\!:\! U_i\to{\mathbb C}^m$ factor through $V_i\subset{\mathbb C}^m$ and $\iota_i\circ (F_{1,i}|_{U_i})\!:\! U_i\to X$ is identical to $f|_{U_i}\!:\! U_i\to X$; \noindent 3. In a neighborhood $W_{ij}$ of $U_i\cap U_j$ in $W_i\cap W_j$ over which the map \begin{equation} F_{s,ij}=g_{ij}\circ F_{s,j}- F_{s,i}\!:\! W_{ij}\to {\mathbb C}^m \label{eq:2.21} \end{equation} is well defined, $F_{s,ij}\in {\mathcal I}^{s}{\mathcal H}(W_{ij},{\mathbb C}^m)$, where ${\mathcal H}(W_{ij},{\mathbb C}^m)$ is the ${\mathcal O}_S $-module of holomorphic maps from $W_{ij}$ to ${\mathbb C}^m$; \noindent 4. For any vector $\eta\in{\mathbb C}^r$, we let $L_{\eta}\!:\!{\mathbb C}\to{\mathbb C}^r$ be the unique ${\mathbb C}$-linear map so that $L_{\eta}(1)=\eta$, and let $\eta^{{\rm pre}}$ be the pre-${\mathbb C}$-tangent associated to the pull back of $(C_S,\{ F_{2,i}\})$ under $L_{\eta}$. Using the standard isomorphism $T_0 S\equiv T_0{\mathbb C}^r\cong {\mathbb C}^r$, we obtain a map \begin{equation} T_0S\lra \Ext^1(\ddotw,\cO_{\Sigma})^{\dag} \end{equation} that send $\eta\in T_0 S$ to the image of $\eta^{{\rm pre}}$ under \eqref{eq:2.26}. We require that this map is the isomorphism \eqref{eq:2.20}. For $s=1$ we simply let $F_{1,i}$ be the standard extension of $f|_{U_i}\!:\! U_i\to V_i$ to $W_i\to {\mathbb C}^m$. We now show that we can construct $\{ F_{2,i}\}$ as required. We let $\pi_1$ and $\pi_2$ be the first and the second projection of ${\mathbb C}^r\times \Sigma$, where we view ${\mathbb C}^r$ as the total space of $\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$. It follows from the definition of the extension group that there is a universal diagram \begin{equation} \begin{CD} @.@.\pi_2^{\ast} f^{\ast}\Omega_X @= \pi_2^{\ast} f^{\ast}\Omega_X\\ @.@. @VV{{\mathcal V}_2}V @VV{\pi_2^{\ast}(df^{\vee})}V \\ 0 @>>> \pi_2^{\ast}{\mathcal O}_{\Sigma}^{\cA} @>>> {\tilde{\mathcal B}}^{\cA} @>>> \pi_2^{\ast}\Omega_{\Sigma}(D)^{\cA} @>>> 0 \end{CD} \label{eq:2.28} \end{equation} such that its restriction to fibers of ${\mathbb C}^r\times\Sigma$ over $\xi\in{\mathbb C}^r$ are the diagrams \eqref{eq:2.3} associated to $\xi \in\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$. By deformation theory of pointed curves, for any smooth point $p\in\Sigma$ the vector space $\tilde{\mathcal B}\otimes k(p)$ is canonically isomorphic to the cotangent space $T_p^{\ast} C_S$. By applying the construction of $\xi^v\in T_{\tilde w}^{{\rm pre}} R$ from $v\in\Ext^1(\ddotw,\cO_{\Sigma})^{\dag}$ to the family version, we can construct the family $\{ F_{2,i}\}$ as required. We will leave the details to the readers. Now we show that we can successively construct $F_{s,i}$ that satisfies the four conditions above. Assume that for some $s\geq 2$ we have constructed $\{F_{s,i}\}$ that satisfies the four conditions above. Let $W_{ij}$ be the neighborhood of $U_{ij}=U_i\cap U_j\subset W_i\cap W_j$ so that \eqref{eq:2.21} is well-defined. Then by the condition 3 above, $F_{s,ij}\in {\mathcal I}^{s} {\mathcal H}(W_{ij},{\mathbb C}^m)$. Let $I=(i_1,\ldots,i_r)$ be any length $s$ mulptiple index, namely, $i_j\geq 0$ and $\sum i_j=s$. As usual, we will denote by $\partial^I$ the symbol $\partial^{i_1}/\partial z_1^{i_1}\cdots\partial^{i_r}/\partial z_r^{i_r}$ and by $z^I$ the term $z_1^{i_1}\cdots z_r^{i_r}$. Then because of the condition 3 above, $\varphi_{I,ij}={\partial^I} F_{s,ij}|_{U_{ij}}$ is a holomorphic section of $f^{\ast} T_X|_{U_{ij}}$ using the standard isomorphism $$TX|_{V_i}\cong TV_i\cong V_i\times{\mathbb C}^m, $$ and the collection $[\varphi_{I,ij}]$ defines a C\v{e}ch 1-cocycle of $f^{\ast}{\mathcal T}_X$. We let $\{\phi_{I,i}\}$, where $\phi_{I,i}=\zeta_i+a_i\eta_i$, be the collection provided by Lemma \ref{2.1}. Using the standard isomorphism $TX|_{V_i}\cong V_i\times{\mathbb C}^m$, we can view $\phi_{I,i}$ as a map $V_i\to {\mathbb C}^m$. We let $\tilde\phi_{I,i}\!:\! W_i\to{\mathbb C}^m$ be the standard extension of $\phi_{I,i}$ and let $G_{I,i}= \pi_S^{\ast}(z^I)\tilde\phi_{I,i}$. Clearly, $\partial^I G_{I,i}=\phi_{I,i}$. Now we let $$F_{s+1,i}=F_{s,i}+\sum_{\ell(I)=s} G_{I,i}. $$ It is direct to check that the collection $\{F_{s+1,i}\}$ satisfies the condition 1-4 before. Finally, by the estimate in Lemma \ref{2.1}, there is a neighborhood of $U_i\subset W_i$, say $W_i^0$, such that $\lim_s F_{s,i}$ converges over $W_i^0$. Let $F_{\infty,i}$ be its limit. Because $f(U_i)\Subset V_i$, there is a neighborhood $\tilde{W}_i$ of $U_i\subset W_i^0$ such that $F_{\infty,i}(\tilde{W}_i) \subset V_i\subset{\mathbb C}^m$. It follows that we can find a neighborhood $S^0\subset S$ of $0\in S$ such that $\pi_S^{-1}(S^0)\subset \cup \tilde{W}_i$. Finally, because $F_{\infty,i}$ is analytic near $U_i$ for $i>l$ and is analytic in $S$ direction using $W_i\cong U_i\times S$ otherwise, the condition 3 implies that the collection $F_{\infty,i}|_{{W}_i\cap\pi_S^{-1}(S^0)}$ defines a map $$F_S: \pi_S^{-1}(S^0)\lra X. $$ Clearly, $F_S$ is holomorphic away from the union of $W_1,\ldots, W_l$. Further, for each $i\leq l$ if we let $\xi_i$ be a holomorphic variable of $U_i$ and let $\pi_{U_i}$ and $\pi_{S^0}$ be the first and the second projection of $W_i\cap \pi_S^{-1}(S^0)\cong U_i\times S^0$, then \begin{equation} \frac{\partial}{\partial\bar\xi_i} F_S|_{W_i\cap\pi_S^{-1}(S^0)}d\bar\xi_i= \pi_{S^0}^{\ast}(\varphi_i)\pi_{U_i}^{\ast}(\gamma_i) F_S^{\ast}({\mathbf v}_i)|_{W_i} \label{eq:2.30} \end{equation} where $\varphi_i$ is a holomorphic function over $S^0$. Finally, we let $Z$ be the subset of $$\pi_S^{-1}(S^0)\times_S\cdots\times_S\pi^{-1}_S(S^0)\qquad (\text{$k$ times}) $$ consisting of $(s;x_{n+1},\ldots,x_{n+k})$ such that $s\in S^0$ and that $ x_{n+1},\ldots,x_{n+k}$ are distinct points in $\pi_S^{-1}(s)$ that lie in $F_s^{-1}(H)$. Note that if we choose $U$ to be small enough, then $F_s^{-1}(H)$ has exactly $k$ points. Let $C_Z$ be the family of $(n+k)$-pointed curves over $Z$ so that its domain is the pull back of $C_S$ via $Z\to S$, its first $n$-marked sections is the pull back of the $n$-marked sections of $C_S$ and its last $k$-sections of the fiber of $C_Z$ over $(s;x_{n+1},\ldots, x_{n+k})$ is $x_{n+1},\ldots,x_{n+k}$. Coupled with the pull back of $F_S$, say $F_Z\!:\! C_Z\to X$, we obtain a family of stable (continuous) maps from $(n+k)$-pointed curves to $X$. Let $\eta\!:\! Z\to U$ be the tautological map. We claim that $\eta(Z)\subset R$. Indeed, let $z\in Z$ be any point and let $C_z$ be the domain of $z$. It follows from our construction that $C_z$ has $l$ distinguished open subsets, denoted by $U_1,\ldots, U_l$, such that $f_z=F_Z|_{C_z}$ is holomorphic away from $\cup_{i=1}^l U_i$ and $\bar\partial f_z|_{U_i}$ is a constant multiple of $\gamma_i\cdot f_z^{\ast}({\mathbf v}_i)$. Hence the value of the section $\Phi_{\tilde U}\!:\! \tilde U\to {\mathbf E}_{\tilde U}$ at $\eta(z)$ is contained in the subspace $V|_{\eta(z)}\subset {\mathbf E}_{\tilde U}|_{\eta(z)}$. This shows that $\eta(z)\in R$. \begin{prop} The induced map $\eta\!:\! Z\to R$ is a local diffeomorphism near those $z\in R$ whose associated map $f_z\!:\! C_z\to X$ are holomorphic. \end{prop} \begin{proof} This follows immediately from the proof of the basic Lemma in \cite{LT2}. We will omit the details here. \end{proof} By shrinking $S^0$ if necesary, we can assume that $\eta\!:\! Z\to R$ is a local diffeomorphism. We can further assume that there is an open subset $Z^{\prime}\subset Z$ containing $\tilde w$ such that $\eta^{\prime}=\eta|_{Z^{\prime}}\!:\! Z^{\prime}\to R$ is one-to-one and the image $\eta(Z^{\prime})\subset R$ is invariant under $G_{\tilde U}$. $\eta^{\prime}\!:\! Z^{\prime}\to R$ is the analytic coordinate of $\tilde w\in R$ we want. For convenience, we will view $Z^{\prime}$ as an open subset of $R$. \begin{prop} Let $V^{\prime}$ be the restriction of $W$ to $Z^{\prime}$ endowed with the holomorphic structure so that $\tilde\sigma_1|_{Z^{\prime}},\ldots, \tilde\sigma_l|_{Z^{\prime}}$ is a holomorphic frame. Then $\phi^{\prime} \equiv \phi_V|_{Z^{\prime}} \!:\! Z^{\prime}\to V^{\prime}$ is holomorphic. \end{prop} \begin{proof} This follows immediately from \eqref{eq:2.30}. \end{proof} Let ${\phi^{\prime}}^{-1}(0)$ be any point and let $f_z\!:\! C_z\to X$ be the associated (analytic) stable map with $D_z$ the divisor of its first $n$-marked points. Then there is a caonical exact sequence of vector spaces $$ \Ext^1(\Omega_{C_z}(D_z),{\mathcal O}_{C_z})\lra H^1(f_z^{\ast}{\mathcal T}_X)\lra \Ext^2({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z})\lra 0 $$ induced by the short exact sequence of complexes $$ 0\lra [0\to\Omega_{C_z}(D_z)]\lra [f_z^{\ast}\Omega_{X}\to\Omega_{C_z}(D_z)] \lra[f_z^{\ast}\Omega_X\to 0]\lra 0. $$ Similarly, the differentil $d\phi_V(z)\!:\! T_z R\to W_z$ induces an exact sequence of vector spaces $$ \Ext^1({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z}^{\cA})^{!}\xrightarrow{d\phi_V(z)} W|_z \lra \coker(d\phi_V(z))\lra 0. $$ Note that there are canonical homomorphisms $\Ext^1({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z} ^{\cA})^{!}\to\Ext^1({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z})$ and $W|_z\to H^{0,1}_{\bar\partial}( f_z^{\ast} T_X)\cong H^1(f_z^{\ast}{\mathcal T}_X)$. \begin{lemm} \label{2.22} There is a canonical isomorphism $\xi$ (as shown below) that fits into the diagram \begin{equation*} \begin{CD} \Ext^1({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z})^{!} @>{d\phi_V(z)}>> W|_z @>>> \coker(d\phi_V(z)) @>>> 0\\ @VVV @VVV @VV{\xi}V\\ \Ext^1(\Omega_{C_z}(D_z),{\mathcal O}_{C_z}) @>>> H^1(f_z^{\ast}{\mathcal T}_X) @>>> \Ext^2({\mathcal D}^{\bullet}_z,{\mathcal O}_{C_z}) @>>> 0. \end{CD} \end{equation*} \end{lemm} \begin{proof} This is obvious and will be left to the readers. \end{proof} \section{The proof of the comparison theorem} In this section, we will prove that the algebraic and the symplectic construction of GW-invariants yield identical invariants. We will work with the category of algebraic schemes as well as the category of analytic schemes. Specifically, we will use the words schemes, morphisms and \`etale neighborhoods to mean the corresponding objects in algebraic category and use the word analytic maps and open subsets to mean the corresponding objects in analytic category. As before, the words analytic and holomorphic are interchangable. Also, we will use ${\mathcal O}_S$ to mean the sheaf of algebraic sections or the sheaf of analytic sections depending on whether $S$ is an algebraic scheme or an analytic scheme. We will continue to use the complex dimension through out this section. We now clarify our usage of the notions of cycles and currents. Let $W$ be a scheme. We denote by $Z_k^{\text{alg}} W$ the group of formal sums of finitely many $k$-dimensional irreducible subvarieties of $W$ with rational coefficients. We call elements of $Z_k^{\text{alg}} W$ $k$-cycles of $W$. Now let $W$ be any stratified topological space with stratification ${\mathcal S}$. We say that a (complex) $k$-dimensional current $C$ is stratifiable if there is a refinement of ${\mathcal S}$, say ${\mathcal S}^{\prime}$, such that there are finitely many $k$-dimensional strata $S_i$ and rationals $a_i\in{\mathbb Q}$ such that $C=\sum a_i[S_i]$ (All currents in this paper are oriented). Here we assume that each stratum of ${\mathcal S}^{\prime}$ was given an orientation a priori and $[S_i]$ is the oriented current defined by $S_i$. We identify two currents if they define identical measures in the sense of rectifiable currents. We denote the set of all stratifiable $k$-dimensional currents modulo the equivalence relation by $Z_k W$. Clearly, if $W$ is a scheme then any $k$-cycle has an associated current in $Z_k W$, which defines a map $Z_k^{\text{alg}} W\to Z_k W$. In the following, we will not distinguish a cycle from its associated current. Hence for $C\in Z_k^{\text{alg}} W$ we will view it as an element of $Z_{k} W$. Note that if $C\in Z_k W$ has zero boundary in the sense of current and $C$ has compact support, then $C$ defines canonically an element in $H_{2k}(W,{\mathbb Q})$. Finally, if $C=\sum a_i[S_i]\in Z_k W$ and $F\subset W$ is a stratifiable subset, we say that $C$ intersects $F$ transversally if $F$ intersects each $S_i$ transversally as stratified sets (See \cite{GM} for topics on stratifications). In such cases, we can define the intersection current $C\cap F$ if the orientation of the intersection can be defined according to the geometry of $W$ and $F$. We begin with a quick review of the algebraic construction of GW-invariants. Let $X$ be any smooth projective variety and let $A\in H_2(X,{\mathbb Z})$ and $g,n\in{\mathbb Z}$ as before be fixed once and for all. We let $\M_{g,n}(X,A)$ be the moduli scheme of stable morphisms defined before. $\M_{g,n}(X,A)$ is projective. The GW-invariants of $X$ is defined using the virtual moduli cycle $$[\M_{g,n}(X,A)]^{{\rm vir}}\in A_{\ast}\M_{g,n}(X,A). $$ To review such a construction, a few words on the obstruction theory of deformations of morphisms are in order. Let $w\in\M_{g,n}(X,A)$ be any point associated to the stable morphism ${\mathcal X}$. Let $(B,I,{\mathcal X}_{B/I})$ be any collection where $B$ is an Artin ring, $I\subset B$ is an ideal annihilated by the maximal ideal ${\mathfrak m}_B$ of $B$ and ${\mathcal X}_{B/I}$ is a flat family of stable morphisms over $\spec B/I$ whose restriction to the closed fiber of ${\mathcal X}_{B/I}$ is isomorphic to ${\mathcal X}$. An obstruction theory to deformation of ${\mathcal X}$ consists of a ${\mathbb C}$-vector space $V$, called the obstruction space, and an assignment that assigns any data $(B,I,{\mathcal X}_{B/I})$ as before to an obstruction class $$\ob(B,B/I,{\mathcal X}_{B/I})\in I\otimes_{{\mathbb C}} V $$ to extending ${\mathcal X}_{B/I}$ to $\spec B$. Here by an obstruction class, we mean that its vanishing is the necessary and sufficient condition for ${\mathcal X}_{B/I}$ to be extendable to a family over $\spec B$. We also require that such an assignment satisfies the obvious base change property (For reference on obstruction theory please consult \cite{Ob}). In case ${\mathcal X}$ is the map $f\!:\! C\to X$ with $D\subset C$ the divisor of its $n$ marked points, the space of the first order deformations of ${\mathcal X}$ is parameterized by $\Ext^1({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C)$, where ${\mathcal D}^{\bullet}_{{\mathcal X}}= [f^{\ast}\Omega_X\to\Omega_C(D)]$ is the complex as before, and the standard obstruction theory to deformation of ${\mathcal X}$ takes values in $\Ext^2({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C)$. An example of obstruction theories is the following. Let $R$ be the ring of formal power series in $m$ variables and let ${\mathfrak m}_R\subset R$ be its maximal ideal. Let $F$ be a vector space and let $f\in {\mathfrak m}_R\otimes_{{\mathbb C}}F$. We let $(f)\subset R$ be the ideal generated by components of $f$. Then there is a standard obstruction theory to deformations of $0$ in $\spec R/(f)$ taking values in $V$, where $V$ is the cokernel of $df\!:\! ({\mathfrak m}_R/{\mathfrak m}_R^2)^{\vee}\to F$, defined as follows. Let $I\subset B$ be an ideal of an Artin ring as before and let $\varphi_0\!:\!\spec B/I\to\spec R/(f)$ be any morphism. To extend $\varphi_0$ to $\spec B$, we first pick a homomorphism $\sigma\!:\! R\to B$ extending the induced $R\to B/I$, and hence a morphism $\varphi_{\text{pre}}\!:\! \spec B \to \spec R$. The image $\sigma(f)\in B\otimes F$ is in $I\otimes F$, and is the obstruction to $\varphi_{\text{pre}}$ factor through $\spec R/(f)\subset\spec R$. Let $\ob(B,B/I,\varphi_0)$ be the image of $\sigma(f)$ in $I\otimes V$ via $F\to V$. It is direct to check that $\ob(B,B/I,\varphi_0)=0$ if and only if there is an extension $\varphi\!:\!\spec B\to\spec R/(f)$ of $\varphi_0$. This assignment \begin{equation} \label{eq:3.0} (B,B/I,\varphi_0)\mapsto \ob(B,B/I,\varphi_0)\in I\otimes V \end{equation} is the induced obstruction theory of $\spec R/(f)$. \begin{defi} \label{3.1} A Kuranishi family of the standard obstruction theory of ${\mathcal X}$ consists of a vector space $F$, a ring of formal power series $R$ with ${\mathfrak m}_R$ its maximal ideal, an $f\in{\mathfrak m}_R\otimes F$, a family ${\mathcal X}_{R/(f)}$ of stable morphisms over $\spec R/(f)$ whose closed fiber over $0\in \spec R/(f)$ is isomorphic to ${\mathcal X}$ and an exact sequence \begin{equation} 0\lra \Ext^1({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C)\xrightarrow{\alpha} ({\mathfrak m}_R/{\mathfrak m}_R^2)^{\vee} \xrightarrow{df} F\lra \Ext^2({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C) \lra 0 \label{eq:3.1} \end{equation} of which the following holds: First, the composite $$\Ext^1({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C)\xrightarrow{\alpha} \ker(df)\equiv T_0\spec R/(f) \xrightarrow{} \Ext^1({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C), $$ where the second arrow is the Kodaira-Spencer map of the family ${\mathcal X}_{R/(f)}$, is the identity homomorphism; Secondly, let $I\subset B$ and $\varphi_0\!:\!\spec B/I\to \spec R/(f)$ be as before and let $$\ob(B,B/I,\varphi_0^{\ast}{\mathcal X}_{R/(f)})\in I\otimes\Ext^2({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C) $$ be the obstruction to extending $\varphi_0^{\ast}{\mathcal X}_{B/I}$ to $\spec B$. Then it is identical to $\ob(B,B/I,\varphi_0)$ under the isomorphism $$\coker(df)\cong \Ext^2({\mathcal D}^{\bullet}_{{\mathcal X}},{\mathcal O}_C), $$ where $\ob(B,B/I,\varphi_0)$ is the obstruction class in \eqref{eq:3.0}. \end{defi} We now sketch how the virtual moduli cycle $[\M_{g,n}(X,A)]^{{\rm vir}}$ was constructed. Similar to the situation of the moduli of stable smooth maps, we need to treat $\M_{g,n}(X,A)$ either as a ${\mathbb Q}$-scheme or as a Deligne-Mumford stack. The key ingredient here is the notion of atlas, which is a collection of charts of $\M_{g,n}(X,A)$. A chart of $\M_{g,n}(X,A)$ is a tuple $(S, G, {\mathcal X}_S)$, where $G$ is a finite group, $S$ is a $G$-scheme (with effective $G$-action) and ${\mathcal X}_S$ is a $G$-equivariant family of stable morphisms so that the tautological morphism $\iota \!:\! S/G\to\M_{g,n}(X,A)$ induced by the family ${\mathcal X}_S$ is an \'etale neighborhood. For details of such an notion, please consult \cite{DM, Vi, LT1}. We now let $f\!:\! C\to X$ be the representative of ${\mathcal X}_S$ with $D\subset C$ the divisor of the $n$-marked sections of ${\mathcal X}_S$. Let $\pi\!:\! C\to S$ be the projection. We consider the relative extension sheaves $\mathop{{\mathcal E} xt\hspace{1pt}}^i_{\pi}({\mathcal D}^{\bullet}_{{\mathcal X}_S},{\mathcal O}_C)$, where ${\mathcal D}^{\bullet}_{{\mathcal X}_S}=[f^{\ast}\Omega_X\to\Omega_{C/S}(D)]$ as before. For short, we denote the sheaves $\mathop{{\mathcal E} xt\hspace{1pt}}^i_{\pi}({\mathcal D}^{\bullet}_{{\mathcal X}_S},{\mathcal O}_C)$ by ${\mathcal T}^i_S$. Because they vanish for $i=0$ and $i> 2$, for any $w\in S$, the Zariski-tangent space $T_wS$ is ${\mathcal T}_S^1\otimes_{{\mathcal O}_S}k(w)$ and the obstruction space to deformations of $w$ in $S$ is $V_w={\mathcal T}_S^2\otimes _{{\mathcal O}_S}k(w)$. Now we choose a complex of locally free sheaves of ${\mathcal O}_S$-modules $\cE^{\bullet}=[{\mathcal E}_1\to{\mathcal E}_2]$ so that it fits into the exact sequence \begin{equation} 0\lra {\mathcal T}_S^1\lra{\mathcal E}_1\lra{\mathcal E}_2\lra {\mathcal T}_S^2\lra 0. \label{eq:3.2} \end{equation} We let $F_i(w)={\mathcal E}_i\otimes_{{\mathcal O}_S}k(w)$. Then we have the exact sequence of vector spaces \begin{equation} 0\lra T_wS\lra F_1(w)\lra F_2(w) \lra V_w\lra 0. \label{3.31} \end{equation} We let $K_w\in R(w)$ be a Kuranishi map of the obstruction theory to deformations of $w$, where $R(w)= \varprojlim\oplus _{k=0}^N \text{Sym}^k(F_1(w)^{\vee})$, so that \eqref{3.31} is part of the data of the Kuranishi family specified in Definition \ref{3.1}. Let $(K_w)\subset R(w)$ be the ideal generated by the components of $K_w$ and let $\spec R_w/(K_w)\subset\spec R_w$ be the corresponding subscheme. It follows that $\spec R_w/(K_w)$ is isomorphic to the formal completion of $S$ along $w$, denoted $\hat w$. We let $N_w$ be the normal cone to $\spec R_w/(K_w)$ in $\spec R_w$. Then $N_w$ is canonically a subcone of $F_2(w)\times\hat w$. Here, by abuse of notation we will use $F_2(w)$ to denote the total space of the vector space $F_2(w)$. Note that $N_w$ is the infinitesimal normal cone to $S$ in its obstruction theory at $w$. To obtain a global cone over $S$, we need the following existence and uniqueness theorem, which is the main result of [LT1]. In this paper, we will call a vector bundle $E$ the associated vector bundle of a locally free sheaf ${\mathcal E}$ if ${\mathcal O}(E)\cong {\mathcal E}$. For notational simplicity, we will not distinguish a vector bundle from the total space (scheme) of this vector bundle. \begin{theo}[\cite{LT1}] \label{3.15} Let $E$ be the associated vector bundle of ${\mathcal E}_2$. Then there is a cone scheme $N_S\subset E$ such that for each $w\in S$ there is an isomorphism \begin{equation} F_2(w)\times\hat w\cong E \times_S\hat w \label{eq:3.3} \end{equation} of cones over $\hat w$ extending the canonical isomorphism $F_2(w) \cong E\times_S w$ such that under the above isomorphism $N_w$ is isomorphic to $N_S\times_S\hat w$. In particular, the cycle defined by the scheme $N_S$ is uniquely characterized by this condition. \end{theo} In the previous discussion, if we replace $F_1(w)$ and $F_2(w)$ by $T_wS$ and $V_w$ respectively, we obtain a Kuranishi map and correspondingly a cone scheme in $V_w\times \hat w$, denoted by $N_w^0$. \begin{theo}[\cite{LT1}] \label{3.16} Let the notation be as before. Then there is a vector bundle homomorphism $r\!:\! E\times_S\hat w\to V_w\times \hat w$ extending the canonical homomorphism $E|_w\to V_w$ induced by \eqref{3.31} such that $$N_w^0\times_{V_w\times\hat w} E\times_S\hat w= N_S\times_S \hat w. $$ \end{theo} To construct the virtual cycle $[\M_{g,n}(X,A)]^{{\rm vir}}$, we need to find a global complex over $\M_{g,n}(X,A)$ analogous to $\cE^{\bullet}$. For the purpose of comparing with the analytic construction of the virtual cycles, we will use atlas of analytic charts. We let $\{(R_i,W_i,\phi_i)\}_{i\in\Lambda}$ be the good atlas of the smooth approximation of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ chosen in section 2. Then the collection $Z_i=\phi_i^{-1}(0)$ with the tautological family of stable analytic maps (with the last $k$-marked points discarded) form an atlas of the underlying analytic scheme of $\M_{g,n}(X,A)\cong\Phi^{-1}(0)$. Since we are only interested in constructing and working with cone cycles in ${\mathbb Q}$-bundles (known as V-bundles) over $\M_{g,n}(X,A)$, there is no loss of generality that we work with $\M_{g,n}(X,A)$ with the reduced scheme structure. Hence, for simplicity we will endow $Z_i=\phi_i^{-1}(0)$ with the reduced analytic scheme structure. We let ${\mathcal X}_i$ be the tautological family of the $n$-pointed stable analytic maps over $Z_i$ that is derived by discarding the last $k$ marked points of the restriction to $Z_i$ of the tautological family over $\tilde U_i$. We let $G_i$ be the finite group associated to the chart $(R_i,W_i,\phi_i)$, and let ${\mathcal X}_i$ be represented by $f_i\!:\! C_i\to X$ with $D_i\subset C_i$ be the divisor of the $n$-marked sections of $C_i$ and let $\pi_i\!:\! C_i\to Z_i$ be the projection. In \cite{LT1}, to each $i$, we have constructed a $G_i$-equivariant complex of locally free sheaves of ${\mathcal O}_{Z_i}$-modules $\cE^{\bullet}_i=[{\mathcal E}_{i,1} \to{\mathcal E}_{i,2}]$ such that $\mathop{{\mathcal E} xt\hspace{1pt}}_{\pi_i}^{\bullet}( {\mathcal D}^{\bullet}_{{\mathcal X}_i},{\mathcal O}_{C_i})$ is the sheaf cohomology of $\cE^{\bullet}_i$. It follows from the algebraic and the analytic constructions of charts that each $(Z_i,{\mathcal X}_i)$ can be realized as an analytic open subset of an algebraic chart, say $(S, G, {\mathcal X}_S)$, and the complex $\cE^{\bullet}_i$ is the restriction to this open subset of an algebraic complex $\cE^{\bullet}$, as in \eqref{3.2}. Therefore we can apply Theorem \ref{3.15} to obtain a unique analytic cone cycle $M_i^{\text{alg}}\in Z_{\ast} E_i$, where $E_i$ is the associated vector bundle of ${\mathcal E}_{i,2}$. Let $\iota_i\!:\! Z_i/G_i\to\M_{g,n}(X,A) $ be the tautological map induced by the family ${\mathcal X}_i$. One property that follows from the construction of the complexes $\cE^{\bullet}_i$ which we did not mention is that to each $i$, the cone bundle $E_i/G_i$ over $Z_i/G_i$ descends to a cone bundle over $\iota_i(Z_i/G_i)$, denoted by $\tilde E_i$, and $\{\tilde E_i\}_{i\in\Lambda}$ patch together to form a global cone bundle over $\M_{g,n}(X,A)$, denoted by $\tilde E$. Further, by the uniqueness of the cone cycles $M_i^{\text{alg}}\in Z_{\ast} E_i$ in Theorem 3.2 and 3.3, to each $i$ the cone cycle $M_i^{\text{alg}}/G_i$ in $E_i /G_i$ descends to a cone cycle ${\mathcal M}_i^{\text{alg}}\in Z_{\ast} \tilde E_i$, and $\{{\mathcal M}_i^{\text{alg}}\}_{i\in\Lambda}$ patch together to form a cone cycle in $Z_{\ast}\tilde E$, denoted by ${\mathcal M}^{\text{alg}}$. It follows from \cite{LT1} that $\tilde E$ is an algebraic cone over $\M_{g,n}(X,A)$ and ${\mathcal M}^{\text{alg}}$ is an algebraic cone cycle in $\tilde E$. In the end, we let $\eta_E\!:\! \M_{g,n}(X,A)\to \tilde E$ be the zero section and let $$\eta_E^{\ast}\!:\! \{\text{algebraic cycles in}\ Z_{\ast}\tilde E\}\lra H_{\ast}(\M_{g,n}(X,A);{\mathbb Q}) $$ be the Gysin homomorphism. Then the virtual moduli cycle is $$[\M_{g,n}(X,A)]^{{\rm vir}}=\eta_E^{\ast}[{\mathcal M}^{\text{alg}}]\in A_{\ast}\M_{g,n}(X,A). $$ There is an analogous way to construct the $GW$-invariants of algebraic varieties using analytic method. We continue to use the notion developed in section 1. Let $(R,W,\phi)$ be a smooth approximation of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ constructed in Lemma \ref{1.20}. Then we can construct a cone current in the total space of $W$ as follows. Let $\Gamma_{t\phi}$ be the graph of $t\phi$ in $W$ and let $N_{0/\phi}$ be the limit current $\lim_{t\to\infty}\Gamma_{t\phi}$, when it exists. Clearly, if such a limit does exist, then it is contained in $W|_{\phi^{-1}(0)}$. In general, though $\phi$ is smooth there is no guarantee that such a limit will exist. However, if the approximation is analytic, then we will show that such limit does exist as an stratifiable current. Indeed, assume $(R,W,\phi)$ is an analytic smooth approximation. Since the existence of $\lim\Gamma_{t\phi}$ is a local problem, we can assume that there is a holomorphic basis of $W$, say $e_1,\ldots,e_r$. Then $\phi$ can be expressed in terms of $r$ holomorphic functions $\phi_1,\ldots,\phi_r$. Now let ${\mathbf C}$ be the complex line with complex variable $t$, let $w_i$ be the dual of $e_i$ and let $\Theta\subset W\times {\mathbf C}$ be the analytic subscheme defined by the vanishing of $tw_i-\phi_i$, $i=1,\ldots,r$. We let $\Theta_0$ be the smallest closed analytic subscheme of $\Theta$ that contains $\Theta\cap (W\times{\mathbf C}^{\ast})$, where ${\mathbf C}^{\ast} ={\mathbf C}-\{0\}$. By the Weierstrass preparation theorem, such $\Theta_0$ does exist. Then we define $N_{0/\phi}$ to be the associated cycle of the intersection of the scheme $\Theta_0$ with $W\times\{0\}$. By \cite{Fu}, $N_{0/\phi}$ is the limit of $\Gamma_{t\phi}$. Obviously, $N_{0/\phi}$ is stratifiable. This shows that for any analytic approximation $(R,W,\phi)$ the limit $\lim\Gamma_{t\phi}$ does exist. We now state a simple lemma which implies that if $(R^{\prime},W^{\prime}, \phi^{\prime})$ is a smooth approximation that is finer than the analytic approximation $(R,W,\phi)$, then $\lim\Gamma_{t\phi}$ exists as well. We begin with the following situation. Let $V$ be a smooth oriented vector bundle over a smooth oriented manifold $M$ and let $\varphi\!:\! M\to V$ be a smooth section. Let $V^{\prime}\subset V$ be a smooth submanifold such that for any $x\in\varphi^{-1}(0)$ we have $\image(d\varphi(x))+V^{\prime}_x=V_x$. Then $M_0=\varphi^{-1}(V^{\prime})$ is a smooth submanifold of $M$ near $\varphi^{-1}(0)$. Let $V_0$ be the restriction of $V^{\prime}$ to $M_0$ and let $\varphi_0\!:\! M_0\to V_0$ be the induced section. We next let $N\subset TM|_{\varphi^{-1}(0)}$ be a subbundle complement to $TM_0|_{\varphi^{-1}(0)}$ in $TM|_{\varphi^{-1}(0)}$. Then the union of $d\varphi(x)(N_x)$ for all $x\in\varphi^{-1}(0)$ forms a subbundle of $V|_{\varphi^{-1}(0)}$. We denote this bundle by $d\varphi(N)$. Since $V|_{\varphi^{-1}(0)}\equiv V_0|_{\varphi^{-1}(0)} \oplus d\varphi(N)$, there is a unique projection $P\!:\! V|_{\varphi^{-1}(0)}\to V_0|_{\varphi^{-1}(0)}$ such that whose kernel is $d\varphi(N)$ and the composite of the inclusion $V_0|_{\varphi^{-1}(0)}\to V|_{\varphi^{-1}(0)}$ with $P$ is the identity map. \begin{lemm} \label{3.2} Let the notation be as before and let $l=\dim M$ and $l_0=\dim M_0$. Then $\lim\Gamma_{t\varphi}$ exists as an $l$-dimensional current in $V|_{\varphi^{-1}(0)}$ if and only if $\lim\Gamma_{t\varphi_0}$ exists as an $l_0$-dimensional oriented current in $V_0|_{\varphi^{-1}(0)}$. Further, if they do exist then $$\lim\Gamma_{t\varphi}=P^{\ast}(\lim\Gamma_{t\varphi_0}). $$ Hence $\lim\Gamma_{t\phi}$ is stratifiable if $\lim\Gamma_{ t\phi_0}$ is stratifiable. \end{lemm} \begin{proof} This is obvious and will be left to the readers. \end{proof} Now let $\{(R_{\alpha},W_{\alpha},\phi_{\alpha})\}_{\alpha\in\Xi}$ be a collection of analytic smooth approximations of $[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ such that the images of $Z_{\alpha}=\phi_{\alpha}^{-1}(0)$ (in $\Phi^{-1}(0)$) covers $\Phi^{-1}(0)$. It follows that we can choose a good atlas $\{(R_i,W_i,\phi_i)\}_{k\in\Lambda}$ constructed in Lemma \ref{1.20} so that all approximations in $\Lambda$ are finer than approximations in $\Xi$. Now let $i\in\Lambda$ and let $x\in Z_i=\phi_i^{-1}(0)\subset R_i$ be any point. Because charts in $\Xi$ cover $\Phi^{-1}(0)$, there is an $\alpha\in\Xi$ such that the image of $R_{\alpha}$ in ${\mathbf B}$ contains the image of $x$ in ${\mathbf B}$. Then because $(R_i,W_i,\phi_i)$ is finer than $(R_{\alpha},W_{\alpha},\phi_{\alpha})$, by definition, there is a locally closed submanifold $R_{i,\alpha}\subset R_i$, a local diffeomorphism $f_i^{\alpha}\!:\! R_{i,\alpha}\to R_{\alpha}$ and a vector bundle inclusion $(f_i^{\alpha})^{\ast} W_{\alpha} \subset W_i|_{R_{i,\alpha}}$ such that $(f_i^{\alpha})^{\ast}(\phi_{\alpha})=\phi_i$, as in \eqref{eq:1.4}. This is exactly the situation studied in Lemma \ref{3.2}. Hence $\lim\Gamma_{t\phi_i}$ exists near fibers of $W$ over $x$. Because $\{Z_{\alpha}\}$ covers $\Phi^{-1}(0)$, $\lim\Gamma_{t\phi_i}$ exists and is a pure dimensional stratifiable current of dimension $\dim R_i$. We denote this current by $N_{0/\phi_i}$. Now it is clear how to construct the GW-invariants of algebraic varieties using these analytically constructed cones. By the property of good coverings, for $j\leq i\in\Lambda$ the approximation $(R_i,W_i,\phi_i)$ is finer than $(R_j,W_j,\phi_j)$. We let $Z_i=\phi_i^{-1}(0)$ be as before and let $Z_{i,j}= Z_i\cap R_{i,j}\subset Z_i$, where $R_{i,j}$ is defined before \eqref{eq:1.4}. Let $\rho^j_i\!:\! Z_{i,j}\to Z_j$ be the restriction of $f^j_i$ to $Z_{i,j}$. Note that $Z_{i,j}$ is an open subset of $Z_i$ and $\rho^j_i\!:\! Z_{i,j}\to \rho^j_i(Z_{i,j})$ is a covering. Let $F_i$ be the restriction of $W_i$ to $Z_i$ and let $p_i\!:\! F_i\to Z_i$ be the projection. Hence, $(\rho_i^j)^{\ast}(F_j)$ is canonically a subvector bundle of $F_i|_{Z_{i,j}}$. By Lemma \ref{3.2}, $(\rho^j_i)^{\ast}(F_j)$ intersects transversally with $N_{0/\phi_i}\cap p_i^{-1}(Z_{i,j})$ and as currents, $N_{0/\phi_i}\cap (\rho_i^j)^{\ast}(F_j)=(p_i^j)^{\ast}(N_{0/\phi_j})$. For convenience, in the following we will call the collection $\{F_i\}$ with transition functions $f^j_i$ a semi-${\mathbb Q}$-bundle and denote it by ${\mathcal F}$, and will denote $\{ N_{0/\phi_i}\}$ by ${\mathcal N}^{\text{an}}$. As in section two, we call a collection ${\mathbf s}=\{s_i\}_{i\in\Lambda}$ of smooth sections $s_i\!:\! Z_i\to F_i$ a global section of ${\mathcal F}$ if for $j\leq i\in\Lambda$ the restriction $s_i|_{Z_{i,j}} \!:\! Z_{i,j}\to F_i|_{Z_{i,j}}$ coincides with the pull back section $(\rho_i^j)^{\ast} s_j\!:\! Z_{i,j}\to (\rho_i^j)^{\ast} F_j$ under the canonical inclusion $(\rho_i^j)^{\ast} F_j\subset F_i|_{Z_{i,j}}$. We say that the section ${\mathbf s}$ is transversal to ${\mathcal N}^{\text{an}}$ if for each $i\in\Lambda$, the graph of the section $s_i$ is transversal to $N_{0/\phi_i}$ in $F_i$. Obviously, if ${\mathbf s}$ is a global section of $F$ that is transversal to ${\mathcal N}^{\text{an}}$, then following the argument after Lemma \ref{1.10}, currents $$\frac{1}{m_i}\iota^{\prime}_{i\ast}\pi_{i\ast}(N_{0/\phi_i}\cap \Gamma_{s_i}),\quad {i\in\Lambda}, $$ where $\iota^{\prime}_i\!:\! Z_i\to{\mathbf B}$ is the restriction of $\iota_i\!:\! R_i\to{\mathbf B}$ to $Z_i\subset R_i$ and $m_i$ is the number of sheets of the branched covering $\iota^{\prime}_i\!:\! Z_i\to\iota_i^{\prime}(Z_i)$, patch together to form an oriented current in ${\mathbf B}$ without boundary. We denote this current by ${\mathbf s}^{\ast}({\mathcal N}^{\text{an}})$. It has pure dimension $r_{\rm exp }$ since the currents $N_{0/\phi_i}$ has dimension $\dim R_i=\rank F_i+r_{\rm exp }$. Hence it defines a homology class $[{\mathbf s}^{\ast}({\mathcal N}^{\text{an}})]$ in $H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q})$. \begin{prop} $[{\mathbf s}^{\ast}({\mathcal N}^{\text{an}})]$ is the Euler class $e[\Phi\!:\! {\mathbf B}\to{\mathbf E}]$ constructed in section one. \end{prop} \begin{proof} Recall that the class $e[\Phi\!:\!{\mathbf B}\to{\mathbf E}]$ was constructed by first selecting a collection of perturbations $h_i(s)\!:\! R_i\to W_i$ of $\phi_i$ parameterized by $s\in [0,1]$ satisfying certain property and then form the current that is the patch together of the currents $\frac{1}{ m_i}\iota_{k\ast}(\Gamma_{h_i(1)} \cap\Gamma_0)$, where $\Gamma_{h_i(1)}$ and $\Gamma_0$ are the graph of $h_i(1)$ and $0\!:\! R_i\to W_i$. Alternatively, we can perturb the $0$-section instead of $\{\phi_i\}$ to obtain the same cycle. Namely, we let $h_i^{\prime}(s)\!:\! R_i\to W_i$ be a collection of perturbations of the zero section $0\!:\! R_i\to W_i$, such that it satisfies the obvious compatibility and properness property similar to that of $h_i(s)$ in section two. Moreover, we require that the graph $\Gamma_{h^{\prime}_i(1)}$ is transversal to $N_{0/\phi_i}$ and transversal to the graph $\Gamma_{t\phi_i}$ for sufficiently large $t$. Of course such perturbations do exist following the proof of Proposition \ref{1.12}. Let $C_t$ be the current in ${\mathbf B}$ that is the result of patching together the currents $\frac{1}{ m_i}\iota_{i\ast}p_{i\ast}(\Gamma_{h_i^{\prime}(1)} \cap\Gamma_{t\phi_i})$, where $p_i$ is the projection $W_i\to R_i$. Clearly, for $t\gg 0$, we have $\partial C_t=0$ and $\supp(C_t)$ is compact. Hence $C_t$ defines a homology class in $H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q})$, denoted by $[C_t]$. It follows from the uniqueness argument in the end of section two that for sufficiently large $t$, the homology class $[C_t]$ in $H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q})$ is exactly the Euler class. On the other hand, we let $C_{\infty}$ be the current in ${\mathbf B}$ that is the patch together of the currents $\frac{1}{ m_i}\iota_{i\ast}p_{i\ast}(\Gamma_{h_i^{\prime}(1)}\cap N_{0/\phi_0})$. Because $N_{0/\phi_i}$ is the limit of $\Gamma_{t\phi_i}$, and because $\Gamma_{h_i^{\prime}(1)}$ intersects transversally with $\Gamma_{t\phi_i}$ for $t\gg 0$ and with $N_{0/\phi_i}$, the union $$\bigcup_{t\in [0,\epsilon]} \{t\}\times C_{1/t}\subset [0,\epsilon]\times{\mathbf B}, $$ where $1\gg\epsilon>0$, is a current whose boundary is $C_{1/\epsilon}- C_{\infty}$. This implies that $$[C_{\infty}]=[C_t]\in H_{2r_{\rm exp }}({\mathbf B};{\mathbb Q})\qquad \text{for}\ t\gg 0. $$ Further, because the currents $N_{0/\phi_i}$ are contained in $F_i=W_i|_{Z_i}$, $p_{i\ast}(N_{0/\phi_i}\cap \Gamma_{h_i^{\prime}(1)})$ as current is identical to $\pi_{i\ast}(N_{0/\phi_i}\cap \Gamma_{r_i})$, where $r_i\!:\! Z_i\to F_i$ is the restriction of $h_i^{\prime}(1)$ to $Z_i$. Hence $C_{\infty}={\mathbf r}^{\ast}({\mathcal N}^{\text{an}})$ with ${\mathbf r}=\{r_i\}$. Finally, it is direct to check that the homology classes $[{\mathbf s}^{\ast}({\mathcal N}^{\text{an}})]$ do not depend on the choices of the sections ${\mathbf s}$ of ${\mathcal F}=\{F_i\}$ so long as they satisfy the obvious transversality conditions. Therefore, $$[{\mathbf s}^{\ast}({\mathcal N}^{\text{an}})]=[{\mathbf r}^{\ast}({\mathcal N}^{\text{an}})]= [C_{1/\epsilon}]=e[\Phi:{\mathbf B}\to{\mathbf E}]. $$ This proves the Proposition. \end{proof} In the end, we will compare the algebraic normal cones with the analytic normal cones to demonstrate that the algebraic and analytic construction of the GW-invariants give rise to the identical invariants. Here is our strategy. Taking the good atlas $\{(Z_i,{\mathcal X}_i)\}_{i\in\Lambda}$ of $\M_{g,n}(X,A)$ as before, we have two collections of semi-${\mathbb Q}$-vector bundles, namely ${\mathcal E}=\{E_i\}$ and ${\mathcal F}=\{F_i\}$, and two collections of cone currents ${\mathcal M}^{\text{alg}}=\{M_i^{\text{alg}}\}$ and ${\mathcal N}^{\text{an}}=\{N_{0/\phi_i}\}$ such that $[\eta_E^{\ast}({\mathcal M}^{\text{alg}})]$ and $[\eta_F^{\ast}({\mathcal N}^{\text{an}})]$ are the algebraic and the symplectic virtual moduli cycles of $\M_{g,n}(X,A)$ respectively. Here $\eta_E$ and $\eta_F$ are generic sections of ${\mathcal E}$ and ${\mathcal F}$ respectively. To compare these two classes, we will form a new semi-${\mathbb Q}$-vector bundle ${\mathcal V}=\{V_i\}$, where $V_i=E_i\oplus F_i$, and construct a stratifiable cone current ${\mathcal P}$ in ${\mathcal V}$ such that the cycle ${\mathcal P}$ intersect ${\mathcal E}\subset{\mathcal V}$ and ${\mathcal F}\subset{\mathcal V}$ transversally and the intersection ${\mathcal P}\cap {\mathcal E}$ and ${\mathcal P}\cap {\mathcal F}$ are ${\mathcal M}^{\text{alg}}$ and ${\mathcal N}^{\text{an}}$ respectively. Therefore, if we let $\eta_V$ be a generic section of ${\mathcal V}$, then $$[\eta_E^{\ast}({\mathcal M}^{\text{alg}})]=[\eta^{\ast}_V({\mathcal P})]=[\eta^{\ast}_F({\mathcal N}^{\text{an}})] \in H_{\ast}(\M_{g,n}(X,A);{\mathbb Q}). $$ This will prove the Comparison Theorem. We now provide the details of this argument. We begin with any index $i\in\Lambda$ and an open subset $S\subset Z_i$. Let $f\!:\! C\to X$ be the restriction to $S$ of the tautological family ${\mathcal X}_i$ of stable maps over $Z_i$, with $D\subset C$ the divisor of its $n$-marked sections and $\pi\!:\! C\to S$ the projection. Note that $f$ is the restriction of a family of stable morphisms over a scheme to an analytic open subset of the base scheme. Following the construction in \cite[section 3]{LT1}, after fixing a sufficiently ample line bundle over $X$, we canonically construct a locally free sheaf of ${\mathcal O}_{C}$-modules ${\mathcal K}$ so that $f^{\ast}\Omega_X$ is canonically a quotient sheaf of ${\mathcal K}$. Let ${\mathcal L}$ be the kernel of ${\mathcal K}\to f^{\ast}\Omega_X$. Then the restriction to $S$ of the sheaf ${\mathcal E}_{i,1}$ (resp. ${\mathcal E}_{i,2}$) mentioned before is the the relative extension sheaf $\mathop{{\mathcal E} xt\hspace{1pt}}_{\pi}^1([{\mathcal K}\to\Omega_{C/S}(D)],{\mathcal O}_{C})$ (resp. $R\pi_{\ast}({\mathcal L}^{\vee})$). We denote them by ${\mathcal E}_{S,1}$ and ${\mathcal E}_{S,2}$ respectively. As usual, we let $E_{S,1}$ and $E_{S,2}$ be the associated vector bundle of ${\mathcal E}_{S,1}$ and ${\mathcal E}_{S,2}$ respectively. Following the notation in \cite{LT1}, the tangent-obstruction complex $[{\mathcal T}_S^1\to{\mathcal T}_S^2]$ of ${\mathcal X}_i|_S$ is $$\bigl[\sideset{}{^1_{\pi}}\mathop{{\mathcal E} xt\hspace{1pt}}([f^{\ast}\Omega_X\to\Omega_{C/S}(D)],{\mathcal O}_{C}) \xrightarrow{\times0} \sideset{}{^2_{\pi}}\mathop{{\mathcal E} xt\hspace{1pt}}([f^{\ast}\Omega_X\to\Omega_{C/S}(D)],{\mathcal O}_{C})\bigr], $$ and that there is a canonical homomorphism $\epsilon: {\mathcal E}_{S,1}\lra {\mathcal E}_{S,2}$ so that the kernel and the cokernel of $\epsilon$ are ${\mathcal T}_S^1$ and ${\mathcal T}_S^2$ respectively. The homomorphism $\epsilon$ is the middle arrow in the sequence \eqref{eq:3.2}. We now assume that there is an analytic approximation $\alpha\in\Xi$ so that $(R_i,W_i,\phi_i)$ is finer than $\alpha$ and $\iota_i(S)\subset{\mathbf B}$ is contained in $\iota_{\alpha}(Z_{\alpha})$. Let $\rho_{\alpha}\!:\! Z_i\to Z_{\alpha}$ be induced by $f_i^{\alpha}\!:\! R_{i,\alpha}\to R_{\alpha}$ (see \eqref{eq:1.4}). Let $F_{S,\alpha}$ be the vector bundle over $Z_i$ that is the pull back of $F_{\alpha}$. Note that $F_{S,\alpha}$ is a smooth vector bundle. Let $G_{S,\alpha,2}= E_{S,2}\oplus F_{S,\alpha}$. In the following, we will construct a holomorphic vector bundle $G_{S,\alpha,1}$ and a possibly degenerate vector bundle homomorphism $\beta$ and non-degenerate vector bundle inclusions $\tau_{\alpha,j}$ as shown below so that \begin{equation} \begin{CD} E_{S,1} @>{\epsilon}>> E_{S,2}\\ @VV{\tau_{\alpha,1}}V @VV{\tau_{\alpha,2}}V\\ G_{S,\alpha,1} @>{\beta}>>G_{S,\alpha,2} \end{CD} \label{eq:3.5} \end{equation} is commutative. Let $w$ be any point in $S$. We denote by $C_w$ the fiber of $C$ over $w$ and let $f_w$ (resp. ${\mathcal K}_w$, resp. ${\mathcal L}_w$) be the restriction of the respective objects to $C_w$. As before, for any locally free sheaf of ${\mathcal O}_{C_w}$-modules ${\mathcal W}$ that is locally free away from the nodal points of $C_w$, we denote by ${\mathcal W}^{\cA}$ the sheaf whose stalk at nodal points $z$ of $C_w$ are ${\mathcal W}_z$ and its stalks at smooth points $z$ of $C_w$ are germs of smooth sections of the associated vector bundle of ${\mathcal W}$ at $z$. We let $G_{S,\alpha,1}|_w$ be the vector space of the equivalence classes of commutative diagrams \begin{equation} \begin{CD} @. @. {\mathcal K}_w @>>> f_w^{\ast}\Omega_X\\ @.@. @VV{h}V @VV{df_w^{\vee}}V\\ 0@>>> {\mathcal O}_{C_w}^{\cA} @>>> {\mathcal B}_w^{\cA} @>>> \Omega_{C_w}(D_w)^{\cA} @>>> 0 \end{CD} \label{eq:3.6} \end{equation} such that the lower exact sequences are induced by the exact sequences of sheaves of ${\mathcal O}_{C_w}$-modules \begin{equation*} \begin{CD} 0 @>>> {\mathcal O}_{C_w} @>>> {\mathcal B}_w @>>> \Omega_{C_w}(D_w) @>>> 0 \end{CD} \end{equation*} and that $h$ satisfies the following two requirements. First, let $c\!:\! {\mathcal L}_w\to {\mathcal B}_w^{\cA}$ be the composite of ${\mathcal L}_w\to{\mathcal K}_w$ with $h$. Since ${\mathcal L}_w$ is the kernel of ${\mathcal K}_w\to f_w ^{\ast}\Omega_X$, $c$ automatically lifts to $h_{E}\!:\! {\mathcal L}_w \to {\mathcal O}_{C_w}^{\cA}$. The first requirement is that $h_{E}$ is holomorphic. Secondly, since both ${\mathcal K}_w$ and ${\mathcal L}_w$ are sheaves of ${\mathcal O}_{C_w}$-modules and since $h$ is analytic near nodal points of $C_w$, $\bar\partial h$ is a $(0,1)$-form with compact support \footnote{By which we mean that $\bar\partial h$ vanishes in a neighborhood of the nodal points of $C_w$.} taking values in the associated vector bundle of ${\mathcal K}_w^{\vee}\otimes_{{\mathcal O} _{C_w}}{\mathcal B}_w$. Because of the first requirement, it factors through a section $h_{F}$ of $\Omega^{0,1}_{\rm cpt}(f_w^{\ast} T_X)$. We require that $h_{ F}$ is an element in $\rho_{\alpha}^{\ast} W_{\alpha}|_w$. Using Lemma \ref{2.1} and Corollary \ref{2.2} and the fact that ${\mathcal K}^{\vee}$ is sufficiently ample which was the precondition on our choice of ${\mathcal K}$, it is direct to check that the collection $\{G_{S,\alpha,1}|_w\mid w\in S\}$ forms a smooth vector bundle, denoted $G_{S,\alpha,1}$, and the correspondence that sends \eqref{eq:3.6} to $h_{E}- h_{F}$ form a possibly degenerate vector bundle homomorphism $\beta \!:\! G_{S,\alpha,1}\to G_{S,\alpha,2}$. We next define the homomorphisms $\tau_{\alpha,j}$. The homomorphism $\tau_{\alpha,2}\!:\! E_{S,2}\to G_{S,\alpha,2}$ is the obvious homomorphism based on the definition $G_{S,\alpha,2}=E_{S,2}\oplus F_{S,\alpha}$. For $\tau_{\alpha,1}$, we recall that for any $w\in S$ the vector space $E_{S,1}|_w$ is the set of equivalence classes of the diagrams \eqref{eq:3.6} of which the $h$ are holomorphic. Namely, $h$ are induced by homomorphisms $f_w^{\ast}\Omega_X\to{\mathcal B}$. Hence $E_{S,1}$ is canonically a subbundle of $G_{S,\alpha,1}$. This shows that both $\tau_{\alpha,1}$ and $\tau_{\alpha,2}$ are inclusions of vector bundles. Finally, let $\xi\in E_{S,1}|_w$ be any element associated to the diagram \eqref{eq:3.6}, then $\epsilon(\xi)$ is the section of ${\mathcal L}_w^{\vee}$ that is the lift of ${\mathcal L}_w\to{\mathcal K}_w\xrightarrow{h}{\mathcal B}_w$ to ${\mathcal L}_w\to{\mathcal O}_{C_w}$. It follows that the square of \eqref{eq:3.5} is commutative. We now show that $\coker(\tau_{\alpha,1})= \coker(\tau_{\alpha,2})$. It suffices to show that the sequence \begin{equation} \begin{CD} 0 @>>> E_{S,1}@>{\tau_{\alpha,1}}>> G_{S,\alpha,1} @>{c}>> F_{S,\alpha} @>>> 0, \end{CD} \label{eq:3.9} \end{equation} where $c$ is the composite of $\beta$ with $G_{S,\alpha,2}\to F_{S,\alpha}$, is an exact sequence. But this follows directly from the definition of $G_{S,\alpha,1}$ and Lemma \ref{2.1} and Corollary \ref{2.2}. This proves that $\coker(\tau_{\alpha,1})= \coker(\tau_{\alpha,2})$, and consequently \begin{equation} \coker(\beta|_w)=\coker(\epsilon|_w)={\mathcal T}^2_S|_w \label{eq:3.31} \end{equation} for any $w\in S$. In the following, we will construct the cone current $Q_{S,\alpha}\in Z_{\ast} V_{S,\alpha,2}$. We first pick a subbundle $H_{\alpha}\subset G_{S,\alpha,2}$ such that $H_{\alpha}\to G_{S,\alpha,2}\to\coker(\tau_{\alpha,1})$ is an isomorphism. We let $P_{\alpha}\!:\! G_{S,\alpha,2}\to E_{S,2}$ be the projection so that $\ker(P_{\alpha})=\beta(H_{\alpha})$ and $P_{\alpha}\circ \tau_{\alpha,2}={\mathbf 1}_{E_{S,2}}$. We then take $Q_{S,\alpha}$ to be the flat pull back current $P_{\alpha}^{\ast}(M_i^{\text{alg}})\in Z_{\ast} G_{S,\alpha,2}$. It follows that $Q_{S,\alpha}$ intersects the subbundle $E_{S,2} \subset G_{S,\alpha,2}$ transversally and the intersection $Q_{S,\alpha}\cap E_{S,2}$ is exactly $M^{\text{alg}}_S=M_i^{\text{alg}}|_S$. In the following, we will demonstrate that $Q_{S,\alpha}$ intersects the subbundle $F_{S,\alpha}\subset G_{S,\alpha,2}$ transversally as well and that the intersection $Q_{S,\alpha}\cap F_{S,\alpha}$ is the current $\rho_{\alpha}^{\ast} (N_{\alpha}^{\text{an}})\in Z_{\ast} F_{S,\alpha}$. Let $w\in S$ ($\subset Z_i$) be any point. Since $T_{w^{\prime}}R_{\alpha}$, where $w^{\prime}=\rho_{\alpha}(w)$, is the vector space $\Ext^1({\mathcal D}^{\bullet}_w,{\mathcal O}_{C_w}^{\cA})^{\dag}$, there is a canonical injective homomorphism $\sigma_w \!:\! T_{w^{\prime}}R_{\alpha}\to G_{S,\alpha,1}|_w$ of vector spaces that send the diagram \eqref{eq:2.3} to \eqref{eq:3.6} with ${\mathcal K}_w\to {\mathcal B}_w^{\cA}$ the composite of ${\mathcal K}|_w\to f_w^{\ast}\Omega_X$ with the $v_2$ in \eqref{eq:2.3}. It is easy to see that the collection $\{\sigma_w \}_{w\in S}$ forms a smooth non-degenerate vector bundle homomorphism $\sigma \!:\! \rho_{\alpha}^{\ast}(TR_{\alpha})\to G_{S,\alpha,1}$. If follows from the description of $$\rho_{\alpha}^{\ast}(d \phi_{\alpha}): \rho_{\alpha}^{\ast}(TR_{\alpha})\lra F_{S,\alpha} $$ that the diagram of vector bundle homomorphisms \begin{equation} \begin{CD} G_{S,\alpha,1} @>{\beta}>> G_{S,\alpha,2}\\ @AAA @AAA\\ \rho_{\alpha}^{\ast}(TR_{\alpha}) @>{\rho_{\alpha}^{\ast}(d\phi_{\alpha})}>> F_{S,\alpha} \end{CD} \end{equation} is commutative, where the second vertical arrow is the obvious inclusion. To compare $Q_{S,\alpha}$ with $\rho_{\alpha}^{\ast}(N_{\alpha}^{\text{an}})$, we need the following two lemmas. \begin{lemm} \label{3.10} Let $w\in S$ be any point and let $w^{\prime}=\rho_{\alpha}(w)$. Let $d_2\!:\! G_{S,\alpha,2}|_w\to {\mathcal T}_S^2|_w$ be the homomorphism induced by \eqref{eq:3.31} and let $F_{S,\alpha}|_w\to {\mathcal T}_S^2|_w$ be the canonical homomorphism given in Lemma \ref{2.22}. Then the following squares are commutative: \begin{equation} \begin{CD} F_{S,\alpha}|_w @>{\subset}>> G_{S,\alpha,2}|_w @<{\tau_{\alpha,2}}<< E_{S,2}\\ @V{d_3}VV @V{d_2}VV @V{d_1}VV\\ {\mathcal T}_S^2|_w @= {\mathcal T}_S^2|_w @={\mathcal T}_S^2|_w \end{CD} \label{eq:3.40} \end{equation} \end{lemm} \begin{lemm} \label{3.11} For any point $w\in Z_{\alpha}$, the germ of $\phi_{\alpha}\!:\! R_{\alpha}\to W_{\alpha}$ at $w$ is a Kuranishi map of the standard obstruction theory of the deformation of stable morphisms associated to the exact sequence $$ 0\lra {\mathcal T}_{\alpha}^1|_w\lra T_wR_{\alpha} \lra F_{\alpha}|_w\lra {\mathcal T}_{\alpha}^2|_w\lra 0. $$ \end{lemm} \begin{proof} We first prove Lemma \ref{3.10}. Since $G_{S,\alpha,2}\equiv E_{S,2}\oplus F_{S,\alpha}$, $d_1$ and $d_3$ induces a homomorphism $G_{S,\alpha,2}|_w\to {\mathcal T}^2_S|_w$. To prove the lemma, it suffices to show that $d_2=d_1\oplus d_3$. To accomplish this, we only need to show that for any $\xi\in G_{S,\alpha,1}|_w$ with $\xi_E$ and $-\xi_F$ its two components of $\beta(\xi)$ according to the direct sum decomposition $G_{S,\alpha,2}|_w=E_{S,2}|_w\oplus F_{S,\alpha}|_w$, then $d_1(\xi_E)=d_3(\xi_F)$. To prove this, we first pick an $h_0\!:\! f_w^{\ast}\Omega_X\to {\mathcal B}^{\cA}_w$ such that \begin{equation} \begin{CD} f_w^{\ast}\Omega_X @= f_w^{\ast}\Omega_X\\ @VV{h_0}V @VV{df_w^{\vee}}V\\\ {\mathcal B}^{\cA}_w @>>> \Omega_{C_w}(D_w)^{\cA} \end{CD} \end{equation} is commutative. Let $h_0^{\prime}$ be the composite of ${\mathcal K}_w\to f_w^{\ast}\Omega_X$ with $h_0$. Then $h^{\prime}-h_0$ factor through ${\mathcal O}_{C_w}^{\cA}\to{\mathcal B}_w^{\cA}$, say $\tilde h\!:\! {\mathcal K}_w\to{\mathcal O}_{C_w}^{\cA}$. Clearly, $\tilde h$ composed with ${\mathcal L}_w\to{\mathcal K}_w$ is the section $\xi_E\in H^0({\mathcal L}_w^{\vee})$. On the other hand, the lift of $\bar\partial \tilde h$ to $\Omega^{0,1}_{\rm cpt}(f_w^{\ast} T_X)$ is $\xi_F-(\bar\partial h_0)^{\text{lift}}$. By the definition of the connecting homomorphism $\delta\!:\! H^0({\mathcal L}_w^{\vee})\to H^1(f_w^{\ast}\Omega_X^{\vee})$, $$\delta(\xi_E)=\text{the image of}\ (\xi_F -(\bar\partial h_0)^{\text{lift}})\ \text{in}\ H^{0,1}_{\bar\partial}(f_w^{\ast} T_X)\cong H^1(f_w^{\ast}\Omega_X^{\vee}). $$ However, the image of $(\bar\partial h_0)^{\text{lift}}$ is contained in the image of the connecting homomorphism $$ \Ext^1(\Omega_{C_w}(D_w),{\mathcal O}_{C_w})\lra \Ext^2([f_w^{\ast}\Omega_X\to 0],{\mathcal O}_{C_w})\equiv H^1(f_w^{\ast}\Omega_X^{\vee}). $$ Hence $d_1(\xi_E)=d_3(\xi_F)$. This proves Lemma \ref{3.10}. \end{proof} \begin{proof} We now prove Lemma \ref{3.11}. Let $I\subset B$ be an ideal of an Artin ring annihilated by the maximal ideal ${\mathfrak m}_B$ and let $\varphi\!:\!\spec B/I \to R_{\alpha}$ be a morphism that sends the closed point of $\spec B/I$ to $w$ and such that $\varphi^{\ast}(\phi_{\alpha})=0$. By the description of the tautological family ${\mathcal X}_{\alpha}$ over $R_{\alpha}$, the pull back $\varphi^{\ast}({\mathcal X}_{\alpha})$ forms an algebraic family of stable morphisms over $\spec B/I$. We continue to use the open covering of the domain ${\mathcal X}_{\alpha}$ used before. Since $R_{\alpha}$ is smooth, we can extend $\varphi$ to $\tilde\varphi\!:\!\spec B\to R_{\alpha}$. Let $C_B$ over $\spec B$ be the domain of the pull back of the domain of ${\mathcal X}_{\alpha}$ via $\tilde\varphi$ and let $C_{B/I}$ be the domain of $C_B$ over $\spec B/I$. We let $\{U_i\}$ (resp. $\{\tilde U_i\}$) be the induced open covering of $C_{B/I}$ (resp. $C_B$) and let $f_i\!:\! U_i\to X$ be the restriction to $U_i$ of the pull back of the stable maps in ${\mathcal X}_{\alpha}$. Because $\varphi^{\ast}(\phi_{\alpha})=0$, $f_i$ are holomorphic. Hence they define a morphism $f\!:\! C_{B/I}\to X$. Now we describe the obstruction to extending $f$ to $\spec B$. Let $C_0$ be the closed fiber of $C_B$ and let $f_0\!:\! C_0\to X$ be the restriction of $f$. For each $i$, we pick a holomorphic extension $\tilde f_i\!:\! \tilde U_i\to X$ of $f_i$. Then over $\tilde U_{ij}=\tilde U_i\cap \tilde U_j$, $\tilde f_j-\tilde f_i$ is canonically an element in $\Gamma(f_0^{\ast}{\mathcal T}_X|_{U_i\cap U_j})\otimes I$, denoted by $f_{ij}$. Further, the collection $\{f_{ij}\}$ is a cocycle and hence defines an element $[f_{ij}]\in H^1(f_0^{\ast}{\mathcal T}_X)\otimes I$. The obstruction to extending $f$ to $\spec B$ is the image of $[f_{ij}]$ in $\Ext^2(\mathcal D^{\bullet}_w,{\mathcal O}_{C_0})\otimes I$ under the homomorphism in the statement in Lemma \ref{2.22} with $z$ replaced by $w$. We denote the image by $\text{ob}^{\text{alg}}$. The obstruction to extending $\varphi$ to $\tilde\varphi\!:\!\spec B\to R_{\alpha}$ so that $\tilde\varphi^{\ast}(\phi_{\alpha})=0$ can be constructed as follows. Let $g_i\!:\! \tilde U_i\to X$ be the pull back of the maps in ${\mathcal X}_{\alpha}$. Note that $g_i$ are well defined since maps in ${\mathcal X}_{\alpha}$ depend analytically on the base manifold $R_{\alpha}$. By the construction of $R_{\alpha}$, for each $i>l$ the map $g_i$ is holomorphic. For $i<l$, we have canonical biholomorphism $\tilde U_i\cong \spec B\times (U_i\cap C_0)$. Because $\varphi^{\ast}(\phi_{\alpha})\equiv0$, if we let $\xi_i$ be a holomorphic variable of $U_i\cap C_0$, then $\frac{\partial}{\partial\bar\xi_i}g_i\cdot d\bar\xi_i$, denoted in short $\bar\partial g_i$, vanishes over $U_i\subset \tilde U_i$. Hence $\bar\partial h$ is a section of $\Gamma(\Omega^{0,1}_{\rm cpt}(f_0^{\ast} T_X)|_{U_i\cap C_0})\otimes I$. Clearly they patch together to form a global section $\gamma$ of $\Omega_{\rm cpt}^{0,1}(f_0^{\ast} T_X)\otimes I$. The element $\gamma$ can be also defined as follows. Let ${\tilde\varphi}^{\ast} \!:\! {\mathcal O}_{R_{\alpha}}\to B$ be the induced homomorphism on rings. Then since the image of ${\tilde\varphi_{\alpha}}^{\ast}(\phi_{\alpha})\in B\otimes_{{\mathcal O}_{R_{\alpha}}}\!{\mathcal O}_{R_{\alpha}}(W_{\alpha})$ in $B/I\otimes_{{\mathcal O}_{R_{\alpha}}}\!{\mathcal O}_{R_{\alpha}}(W_{\alpha})$ vanishes, it induces an element $\gamma^{\prime}\in I\otimes W_{\alpha}|_w$. By our construction of $R_{\alpha}$ and $\phi_{\alpha}$, $\gamma$ coincides with $\gamma^{\prime}$ under the inclusion $W_{\alpha}|_w\subset \Gamma_{C_0}(\Omega_{\rm cpt}^{0,1}(f_0^{\ast} T_X))$. Let $\text{ob}^{\text{an}}$ be the image of $\gamma$ in the cokernel of $d\phi_{\alpha}(w)\!:\! T_w R_{\alpha}\to W_{\alpha}|_w$. By definition, $\text{ob}^{\text{an}}$ is the obstruction to extending $\varphi$ to $\tilde\varphi\!:\! \spec B\to \{\phi_{\alpha}=0\}$. To finish the proof of the lemma, we need to show that $\text{ab}^{\text{alg}}=\text{ob}^{\text{an}}$ under the isomorphism $$\coker\{d\phi_{\alpha}(w)\}\cong \Ext^1({\mathcal D}^{\bullet}_w,{\mathcal O}_{C_0}) $$ given in Lemma \ref{2.22}. For this, it suffices to show that the Dolbeault cohomology class of $\gamma$, denoted $[\gamma]\in H^{0,1}_{\bar\partial}(f_0^{\ast} T_X)\otimes I$, coincides with the C\v{e}ch cohomology class $[f_{ij}]\in H^1(f^{\ast}_0{\mathcal T}_X) \otimes I$ under the canonical isomorphism $H^{0,1}_{\bar\partial}(f_0^{\ast} T_X)\cong H^1(f^{\ast}_0{\mathcal T}_X)$. But this is obvious since $\varphi_i =\tilde f_i-g_i$ is in $\Gamma_{U_i\cap C_0}( \Omega_{\rm cpt}^0(f_0^{\ast} T_X))\otimes I$ such that $\varphi_j-\varphi_i=f_{ij}$ and $\bar\partial \varphi_i=-\bar\partial g_i$. Hence, $[f_{ij}]=[\gamma]$ under the given isomorphism. This proves the lemma. \end{proof} Now we come back to $Q_{S,\alpha}\in Z_{\ast} G_{S,\alpha,2}$. Let $w\in S$ be any point, let $\hat w$ be the formal completion of $S$ along $w$, let $V_w$ be the total space of ${\mathcal T}^2_S|_w$ and let $N_w^0\subset V_w\times \hat w$ be the the cone in Theorem \ref{3.16}. We let $M_S^{\text{alg}}$, $N_{S,\alpha}^{\text{an}}=\rho_{\alpha}^{\ast}(N^{\text{an}}_i)$ and $Q_{S,\alpha}$ be the cone currents in $E_{S,2}$, $F_{S,\alpha}$ and $G_{S,\alpha,2}$ respectively as before. Note that they are supported on union of closed subsets each diffeomorphic to analytic variety. By Theorem \ref{3.15}, we have vector bundle homomorphisms $$e_1: E_{S,2}\times_S\hat w\lra V_w\times \hat w \quad{\rm and}\quad e_3: F_{S,\alpha}\times_S\hat w\lra V_w\times \hat w $$ extending $E_{S,2}|_w\to{\mathcal T}^2_S|_w$ and $F_{S,\alpha}|_w\to{\mathcal T}^2_S|_w$ such that $e_1^{\ast}(N_w^0)$ and $e_3^{\ast}(N_w^0)$ are the restrictions of $M_S^{\text{alg}}$ and $N_{S,2}^{\text{an}}$ to fibers over $\hat w$ in $S$ respectively. Let $e_2\!:\! G_{S,\alpha,2}\times_S\hat w \to V_w\times\hat w$ be induced by $P_{\alpha} \!:\! G_{S,\alpha,2}\to E_{S,2}$ and $e_1$. Then $e_2^{\ast}(N_w^0)$ is the restriction of $Q_{S,\alpha}$ to $G_{S,\alpha,2}\times_S\hat w$. Because the squares in \eqref{eq:3.40} are commutative, \begin{equation*} \begin{CD} e_2\!:\! F_{S,\alpha}\times_S\hat w @>{\subset}>> G_{S,\alpha,2}\times_S\hat w @>{e_2|_{\hat w}}>> V_w\times\hat w \end{CD} \end{equation*} is surjective. Hence $F_{S,\alpha}\times_S\hat w$ intersects $Q_{S,\alpha}$ transversally along fiber over $w$. Let $e_3^{\prime}\!:\! F_{S,\alpha}\times_S\hat w\to V_w\times\hat w$ be induced by $F_{S,\alpha}\to G_{S,\alpha,2}$ and $e_2$, then the intersection of $Q_{S,\alpha}$ with $F_{S,\alpha}\times_S\hat w$ is $(e_3^{\prime})^{\ast}(N_w^0)$. However, by the choice of $P_{\alpha}$, we have $e_3^{\prime}\equiv e_3|_w$, therefore the support of $Q_{S,\alpha}\cap F_{S,\alpha}|_w$ is identical to the support of $N_{S,\alpha}^{\text{an}}|_w$. Because $w\in S$ is arbitrary, the support of $Q_{S,\alpha}\cap F_{S,\alpha}$ is identical to the support of $N_{S,\alpha}^{\text{an}}$. Further, for the same reason, for general point $p$ in $N_{S,\alpha}^{\text{an}}$ the multiplicity of $N_{S,\alpha}^{\text{an}}$ at $p$ is identical to the multiplicity of the corresponding point in $Q_{S,\alpha}\cap F_{S,\alpha}$. This proves that the cycles (or currents) $Q_{S,\alpha}$ intersect $F_{S,\alpha}\subset G_{S,\alpha,2}$ transversally and $Q_{S,\alpha}\cap F_{S,\alpha}=N_{S,\alpha}^{\text{an}}$. We remark that for the same reason, the current $Q_{S,\alpha}$ is independent of the choice of the subbundles $H_{\alpha}\subset G_{S,\alpha,2}$. We now let $F_S=F_i|_S$ and let $G_{S,2}=E_{S,2}\oplus F_S$. Note that $G_{S,\alpha,2}\subset G_{S,2}$. Because $R_i$ is finer than $R_{\alpha}$, $\rho_{\alpha}^{\ast} TR_{\alpha}$ is a subbundle of $TR_i|_S$. Let $K_{\alpha}\subset TR_i|_S$ be a complement of $\rho_{\alpha}^{\ast} TR_{\alpha}\subset TR_i|_S$ and let $d\phi_i(K_{\alpha})\subset F_S$ be the image of this subbundle. Let $P_{S,\alpha}\!:\! F_S\to F_{S,\alpha}$ be the projection so that $\ker P_{S,\alpha}=d\phi_i(K_{\alpha})$ and the composite of $F_{S,\alpha}\subset F_S$ with $P_{S,\alpha}$ is ${\mathbf 1}_{F_{S,\alpha}}$. By Lemma \ref{3.2}, $N_i^{\text{an}}|_S=P_{S,\alpha}^{\ast}(N_{S,\alpha}^{\text{an}})$. Now let $P_S$ be the projection $$P_S= P_{\alpha}\circ ({\mathbf 1}_{E_{S,2}}\oplus P_{S,\alpha}): G_{S,2}\lra G_{S,\alpha,2}\lra E_{S,2} $$ and let $Q_S=P_S^{\ast}(M_i^{\text{alg}})$ be the pull back cone. Let $\tilde d_3$ be \begin{equation*} \begin{CD} \tilde d_3: F_S|_w @>{P_{S,\alpha}|_w}>> F_{S,\alpha}|_w @>{d_3}>> {\mathcal T}_S^2|_w, \end{CD} \end{equation*} then clearly we have a commutative diagram of vector spaces \begin{equation} \begin{CD} F_S|_w @>>> G_{S,2}|_w @<<< E_{S,2}|_w\\ @VV{\tilde d_3}V @VV{P_S|_w}V @VV{d_1}V\\ {\mathcal T}^2_S|_w @= {\mathcal T}^2_S|_w @= {\mathcal T}^2_S|_w. \end{CD} \end{equation} Because $w$ is arbitrary, similar to the previous case, we have that $F_S$ intersects $Q_S$ transversally and $F_S\cap Q_S=N_i^{\text{an}}|_S$, as stratifiable currents. To enable us to patch $Q_S$, where $S\subset Z_i$, to form a current in $G_{i,2}= E_{i,2}\oplus F_i$, we need to show that $Q_S$ is independent of the choice of analytic chart $\alpha$. Namely if we let $\beta\in\Xi$ be another analytic chart so that $\iota_i(S)\subset\iota_{\beta}(Z_{\beta})$, then the cone current $Q_S^{\prime}\subset G_{S,2}$ constructed using $F_{\beta}$, etc., is identical to $Q_S$. Again, following the same argument before, it suffices to show that the homomorphism $\tilde d_3\!:\! F_S|_w\to{\mathcal T}^2_S|_w$ does not depend on the choice of $\alpha$. Note that $\tilde d_3$ also fits into the commutative diagram of exact sequences \begin{equation} \begin{CD} T_{\rho_{\alpha}(w)}R_{\alpha} @>{d\phi_{\alpha}(\rho_{\alpha}(w))}>> F_{\alpha}|_{\rho_{\alpha}(w)} @>>> {\mathcal T}_{\alpha}^2|_{\rho_{\alpha}(w)}@>>> 0\\ @VVV @VVV @| \\ T_w R_i @>{d\phi_i(w)}>> F_S|_w @>>> {\mathcal T}_S^2|_w @>>> 0. \end{CD} \label{eq:3.33} \end{equation} Now assume $\beta\in\Xi$ as before. Without loss of generality, we can assume that near $w$, the vector subbundles $\rho_{\alpha}^{\ast} F_{\alpha}$ and $\rho_{\beta}^{\ast} F_{\beta}$ span a $2l$-dimensional subvector bundle of $F_i$. Now let $V_{\alpha}\to \tilde U_{\alpha}$ and $V_{\beta}\to\tilde U_{\beta}$ be the vector bundles that define $R_{\alpha}$ and $R_{\beta}$ as in section 2 and let $V_{\alpha\beta}\to \tilde U_i$ be the direct sum of the pull back of $V_{\alpha}$ and $V_{\beta}$ via the tautological map $\tilde U_i\to \tilde U_{\alpha}$ and $\tilde U_i\to\tilde U_{\beta}$. Then near a neighborhood of $w\in\tilde U_i$, the set $\tilde\Phi^{-1}(V_{\alpha\beta})$ will form a base of a smooth approximation containing $w$. We denote $R_{\alpha\beta}=\tilde\Phi_i^{-1}(V_{\alpha\beta})$ and let $\phi_{\alpha\beta}\!:\! R_{\alpha\beta}\to V_{\alpha\beta}|_{R_{\alpha\beta}}$ be the lift of $\tilde\Phi_i$. Clearly, $R_i$ is still finer than $R_{\alpha\beta}$. Hence we have commutative diagrams \begin{equation} \begin{CD} T_{\rho_{\alpha}(w)}R_{\alpha} @>{d\phi_{\alpha}(\rho_{\alpha}(w))}>> V_{\alpha}|_{w} @>>> {\mathcal T}^2_{\alpha}|_{\rho_{\alpha}(w)} @>>> 0\\ @VVV @VVV @|\\ T_wR_{\alpha\beta} @>{d\phi_{\alpha\beta}(w)}>> V_{\alpha\beta}|_w @>>> {\mathcal T}^2_i|_{w} @>>> 0\\ @VVV@VVV@|\\ T_w R_i @>{d\phi_i(w)}>> F_i|_w @>>> {\mathcal T}^2_i|_w @>>> 0\\ \end{CD} \label{eq:3.34} \end{equation} with exact rows. Note that $V_{\alpha\beta}|_w\to {\mathcal T}^2_i|_w$ is equal to $$V_{\alpha}|_{\rho_{\alpha}(w)}\oplus V_{\beta}|_{\rho_{\beta}(w)}\lra \Gamma(\Omega_{\rm cpt}^{0,1}(f_w^{\ast} T_X))\lra H^{0,1}_{\bar\partial}(f_w^{\ast} T_X) \lra {\mathcal T}^2_i|_w. $$ (Here that $V_{\alpha\beta}|_w\to {\mathcal T}_i^2|_w$ is defined apriori but not $F_i|_w\to{\mathcal T}^2_i|_w$ because elements of $V_{\alpha}|_w$ and $V_{\beta}|_w$ are $(0,1)$-forms with compact support.) Therefore, the homomorphism $\tilde d_3$ defined earlier is independent of the choice of $\alpha$. Now we are ready to prove the theorem. Let $i\in\Lambda$ be any approximation and let $\{S_a\}$ be an open covering of $Z_i$ so that to each $a$ there is an $\alpha_a\in\Xi$ so that $\iota_i(S_a)\subset\iota_{\alpha_a}(Z_{\alpha_a})$. We let $G_{i,2}=E_{i,2}\oplus F_i$ and let $Q_{S_a}$ be the cone in $G_{i,2}|_{S_a}$ constructed before using the analytic chart $\alpha$. We know that over $G_{i,2}|_{S_a\cap S_b}$, the currents $Q_{S_a}$ and $Q_{S_b}$ coincide. Hence $\{Q_{S_a}\}$ patchs together to form a stratifiable current, denoted $Q_i$. Assume that $j<i\in\Lambda$ be any two indices. Let $Z_{i,j}\subset Z_i$ be the open subset $\iota_i^{-1}(\iota_j(Z_j))$ and let $f^j_i\!:\! Z_{i,j}\to Z_j$ be the map induced by $Z_i$ being finer than $Z_j$. Then $(f_i^j)^{\ast}(F_j)$ is canonically a subbundle of $F_i|_{Z_{i,j}}$, and $(f_i^j)^{\ast}(E_{j,2})$ is canonically isomorphic to $E_{i,2}|_{Z_{i,j}}$. Let $(f_i^j)^{\ast}(G_{j,2})\to G_{i,2}|_{Z_{i,j}}$ be the induced homomorphism. It follows from the previous argument that $Q_i$ intersects $(f_i^j)^{\ast}(G_{j,2})$ transversally and the intersection $Q_i\cap(f_i^j)^{\ast}(G_{j,2})$ is $(f_i^j)^{\ast}(Q_j)$. Finally, by our construction, $Q_i$ intersects transversally with $E_{i,2}$ and $F_i\subset G_{i,2}$, and $E_{i,2}\cap Q_i=M_i^{\text{alg}}$ and $F_i\cap G_i=N_i^{\text{an}}$. Let ${\mathcal G}$ be the semi-${\mathbb Q}$-vector bundle $\{G_{i,2}\}$, which is ${\mathcal E}\oplus{\mathcal F}$, and let ${\mathcal Q}$ be the cone $\{Q_i\}$. It follows from the perturbation argument in section two that for generic sections $\eta_E$, $\eta_F$ and $\eta_G$ of ${\mathcal E}$, ${\mathcal F}$ and ${\mathcal G}$ respectively, we have $$[\M_{g,n}(X,A)]^{{\rm vir}}=[\eta_E^{\ast}{\mathcal M}^{\text{alg}}]=[\eta_G^{\ast}{\mathcal Q}] =[\eta_F^{\ast}{\mathcal N}^{\text{an}}]=e[\Phi\!:\!{\mathbf B}\to{\mathbf E}]. $$ This proves the comparison theorem.
1997-12-19T04:51:34
9712
alg-geom/9712021
en
https://arxiv.org/abs/alg-geom/9712021
[ "alg-geom", "math.AG" ]
alg-geom/9712021
Alexander Polishchuk
Alexander Polishchuk
Analogue of Weil representation for abelian schemes
39 pages, AMSLatex
null
null
null
null
In this paper we construct a projective action of certain arithmetic group on the derived category of coherent sheaves on an abelian scheme $A$, which is analogous to Weil representation of the symplectic group. More precisely, the arithmetic group in question is a congruence subgroup in the group of "symplectic" automorphisms of $A\times\hat{A}$ where $\hat{A}$ is the dual abelian scheme. The "projectivity" of this action refers to shifts in the derived category and tensorings with line bundles pulled from the base. In particular, if $A$ is an abelian scheme over $S$ equipped with an ample line bundle $L$ of degree 1 then we construct an action of a central extension of $Sp_{2n}(\Bbb Z)$ by $\Bbb Z\times Pic(S)$ on the derived category of coherent sheaves on $A^n$ (the $n$-th fibered power of $A$ over $S$). We describe the corresponding central extension explicitly using the the canonical torsion line bundle on $S$ associated with $L$. As a main technical result we prove the existence of a representation of rank $d$ for a symmetric finite Heisenberg group scheme of odd order $d^2$.
[ { "version": "v1", "created": "Fri, 19 Dec 1997 03:51:33 GMT" } ]
2007-05-23T00:00:00
[ [ "Polishchuk", "Alexander", "" ] ]
alg-geom
\section{Heisenberg group schemes} Let $K$ be a finite flat group scheme over a base scheme $S$. A {\it finite Heisenberg group scheme} is a central extension of group schemes \begin{equation}\label{ext} 0\rightarrow{\Bbb G}_m\rightarrow G\stackrel{p}{\rightarrow} K\rightarrow 0 \end{equation} such that the corresponding commutator form $e:K\times K\rightarrow{\Bbb G}_m$ is a perfect pairing. Let $A$ be an abelian scheme over $S$, $L$ be a line bundle on $A$ trivialized along the zero section. Then the group scheme $K(L)=\{x\in A\ |\ t_x^*L\simeq L\}$ has a canonical central extension $G(L)$ by ${\Bbb G}_m$ (see \cite{Mum}). When $K(L)$ is finite, $G(L)$ is a finite Heisenberg group scheme. A {\it symmetric} Heisenberg group scheme is an extension $0\rightarrow{\Bbb G}_m\rightarrow G\rightarrow K\ra0$ as above together with an isomorphism of central extensions $G\widetilde{\rightarrow} [-1]^*G$ (identical on ${\Bbb G}_m$), where $[-1]^*G$ is the pull-back of $G$ with respect to the inversion morphism $[-1]:K\rightarrow K$. For example, if $L$ is a symmetric line bundle on an abelian scheme $A$ (i.e. $[-1]^*L\simeq L$) with a symmetric trivialization along the zero section then $G(L)$ is a symmetric Heisenberg group scheme. For any integer $n$ we denote by $G^n$ the push-forward of $G$ with respect to the morphism $[n]:{\Bbb G}_m\rightarrow{\Bbb G}_m$. For any pair of central extensions $(G_1,G_2)$ of the same group $K$ we denote by $G_1\otimes G_2$ their sum (given by the sum of the corresponding ${\Bbb G}_m$-torsors). Thus, $G^n\simeq G^{\otimes n}$. Note that we have a canonical isomorphism of central extensions \begin{equation}\label{inver} G^{-1}\simeq [-1]^*G^{op} \end{equation} where $[-1]^*G^{op}$ is the pull-back of the opposite group to $G$ by the inversion morphism $[-1]:K\rightarrow K$. In particular, a symmetric extension $G$ is commutative if and only if $G^2$ is trivial. \begin{lem}\label{mult} For any integer $n$ there is a canonical isomorphism of central extensions $$[n]^*G\simeq G^{\frac{n(n+1)}{2}}\otimes [-1]^*G^{\frac{n(n-1)}{2}}$$ where $[n]^*G$ is the pull-back of $G$ with respect to the multiplication by $n$ morphism $[n]:K\rightarrow K$. In particular, if $G$ is symmetric then $[n]^*G\simeq G^{n^2}$. \end{lem} \noindent {\it Proof} . The structure of the central extension $G$ of $K$ by ${\Bbb G}_m$ is equivalent to the following data (see e.g. \cite{Breen}): a cube structure on ${\Bbb G}_m$-torsor $G$ over $K$ and a trivialization of the corresponding biextension $\Lambda(G)=(p_1+p_2)^*G\otimes p_1^*G^{-1}\otimes p_2^*G^{-1}$ of $K^2$. Now for any cube structure there is a canonical isomorphism (see \cite{Breen}) $$[n]^*G\simeq G^{\frac{n(n+1)}{2}}\otimes [-1]^*G^{\frac{n(n-1)}{2}}$$ which is compatible with the natural isomorphism of biextensions $$([n]\times [n])^*\Lambda(G)\simeq\Lambda(G)^{n^2}\simeq \Lambda(G)^{\frac{n(n+1)}{2}}\otimes ([-1]\times [-1])^*\Lambda(G)^{\frac{n(n-1)}{2}}.$$ The latter isomorphism is compatible with the trivializations of both sides when $G$ arises from a central extension. \qed\vspace{3mm} \begin{rem} Locally one can choose a splitting $K\rightarrow G$ so that the central extension is given by a 2-cocycle $f:K\times K\rightarrow{\Bbb G}_m$. The previous lemma says that for any 2-cocycle $f$ the functions $f(nk,nk')$ and $f(k,k')^{\frac{n(n+1)}{2}}f(-k,-k')^{\frac{n(n-1)}{2}}$ differ by a canonical coboundary. In fact this coboundary can be written explicitly in terms of the functions $f(mk,k)$ for various $m\in{\Bbb Z}$. \end{rem} \begin{prop}\label{order} Assume that $K$ is annihilated by an integer $N$. If $N$ is odd then for any Heisenberg group $G\rightarrow K$ the central extension $G^N$ is canonically trivial, otherwise $G^{2N}$ is trivial. If $G$ is symmetric and $N$ is odd then $G^N$ (resp. $G^{2N}$ if $N$ is even) is trivial as a symmetric extension. \end{prop} \noindent {\it Proof} . Combining the previous lemma with (\ref{inver}) we get the following isomorphism: $$[n]^*G\simeq G^{\frac{n(n+1)}{2}}\otimes (G^{op})^{-\frac{n(n-1)}{2}} \simeq G^n\otimes (G\otimes G^{op -1})^{\frac{n(n-1)}{2}}.$$ Now we remark that $G\otimes G^{op -1}$ is given by a trivial ${\Bbb G}_m$-torsor over $K$ with the group law induced by the commutator form $e:K\times K\rightarrow{\Bbb G}_m$ considered as 2-cocycle. It remains to note that $e^{\frac{n(n-1)}{2}}=1$ for $n=2N$ (resp. for $n=N$ if $N$ is odd). Hence, the triviality of $G^n$ in these cases. \qed\vspace{3mm} \begin{cor} Let $G\rightarrow K$ be a symmetric Heisenberg group such that the order of $K$ over $S$ is odd. Then the ${\Bbb G}_m$-torsor over $K$ underlying $G$ is trivial. \end{cor} \noindent {\it Proof} . The isomorphism (\ref{inver}) implies that the ${\Bbb G}_m$-torsor over $K$ underlying $G^2$ is trivial. Together with the previous proposition this gives the result. \qed\vspace{3mm} If $G\rightarrow K$ is a (symmetric) Heisenberg group scheme, such that $K$ is annihilated by an integer $N$, $n$ is an integer prime to $N$ then $G^n$ is also a (symmetric) Heisenberg group. When $N$ is odd this group depends only on the residue of $n$ modulo $N$ (due to the triviality of $G^N$). We call a flat subgroup scheme $I\subset K$ $G$-{\it isotropic} if the central extension (\ref{ext}) splits over $I$ (in particular, $e|_{I\times I}=1$). If $\sigma:I\rightarrow G$ is the corresponding lifting, then we have the reduced Heisenberg group scheme $$0\rightarrow{\Bbb G}_m\rightarrow p^{-1}(I^\perp)/\sigma(I)\rightarrow I^\perp/I\rightarrow 0$$ where $I^\perp\subset K$ is the orthogonal complement to $I$ with respect to $e$. If $G$ is a symmetric Heisneberg group, then $I\subset K$ is called {\it symmetrically} $G$-isotropic if the restriction of the central extension (\ref{ext}) to $I$ can be trivialized as a symmetric extension. If $\sigma:I\rightarrow G$ is the corresponding symmetric lifting them the reduced Heisenberg group $p^{-1}(I^\perp)/\sigma(I)$ is also symmetric. Let us define the Witt group $\operatorname{WH}_{\operatorname{sym}}(S)$ as the group of isomorphism classes of finite symmetric Heisenberg groups over $S$ modulo the equivalence relation generated by $[G]\sim [p^{-1}(I^\perp)/\sigma(I)]$ for a symmetrically $G$-isotropic subgroup scheme $I\subset K$. The (commutative) addition in $\operatorname{WH}_{\operatorname{sym}}(S)$ is defined as follows: if $G_i\rightarrow K_i$ ($i=1,2$) are Heisenberg groups with commutator forms $e_i$ then their sum is the central extension $$0\rightarrow {\Bbb G}_m\rightarrow G_1\times_{{\Bbb G}_m} G_2\rightarrow K_1\times K_2\rightarrow 0$$ so that the corresponding commutator form on $K_1\times K_2$ is $e_1\oplus e_2$. The neutral element is the class of ${\Bbb G}_m$ considered as an extension of the trivial group. The inverse element to $[G]$ is $[G^{-1}]$. Indeed, there is a canonical splitting of $G\times_{{\Bbb G}_m} G^{-1}\rightarrow K\times K$ over the diagonal $K\subset K\times K$, hence the triviality of $[G]+[G^{-1}]$. We define the order of a finite Heisenberg group scheme $G\rightarrow K$ over $S$ to be the order of $K$ over $S$ (specializing to a geometric point of $S$ one can see easily that this number has form $d^2$). Let us denote by $\operatorname{WH}'_{\operatorname{sym}}(S)$ the analogous Witt group of finite Heisenberg group schemes $G$ over $S$ of odd order. Let also $\operatorname{WH}(S)$ and $\operatorname{WH}'(S)$ be the analogous groups defined for all (not necessarily symmetric) finite Heisenberg groups over $S$ (with equivalence relation given by $G$-isotropic subgroups). \begin{rem} Let us denote by $\operatorname{W}(S)$ the Witt group of finite flat group schemes over $S$ with non-degenerate symplectic ${\Bbb G}_m$-valued forms (modulo the equivalence relation given by global isotropic flat subgroup schemes). Let also $\operatorname{W}'(S)$ be the analogous group for group schemes of odd order. Then we have a natural homomorphism $\operatorname{WH}(S)\rightarrow \operatorname{W}(S)$ and one can show that the induced map $\operatorname{WH}'_{\operatorname{sym}}\rightarrow \operatorname{W}'(S)$ is an isomorphism. This follows essentially from the fact that a finite symmetric Heisenberg group of odd order is determined up to an isomorphism by the corresponding commutator form, also if $G\rightarrow K$ is a symmetric finite Heisenberg group with the commutator form $e$, $I\subset K$ is an isotropic flat subgroup scheme of odd order, then there is a unique symmetric lifting $I\rightarrow G$. \end{rem} \begin{thm}\label{annih} The group $\operatorname{WH}_{\operatorname{sym}}(S)$ (resp. $\operatorname{WH}'_{\operatorname{sym}}(S)$) is annihilated by $8$ (resp. $4$). \end{thm} \noindent {\it Proof} . Let $G\rightarrow K$ be a symmetric finite Heisenberg group. Assume first that the order $N$ of $G$ is odd. Then we can find integers $m$ and $n$ such that $m^2+n^2\equiv -1\mod(N)$. Let $\a$ be an automorphism of $K\times K$ given by a matrix $\left( \matrix m & -n\\ n & m \endmatrix \right)$. Let $G_1=G\times_{{\Bbb G}_m} G$ be a Heisenberg extension of $K\times K$ representing the class $2[G]\in \operatorname{WH}'_{\operatorname{sym}}(S)$. Then from Lemma \ref{mult} and Proposition \ref{order} we get $\a^*G_1\simeq G_1^{-1}$, hence $2[G]=-2[G]$, i.e. $4[G]=0$ in $\operatorname{WH}'(S)$. If $N$ is even we can apply the similar argument to the 4-th cartesian power of $G$ and the automorphism of $K^4$ given by an integer $4\times 4$-matrix $Z$ such that $Z^t Z=(2N-1)\operatorname{id}$. Such a matrix can be found by considering the left multiplication by a quaternion $a+bi+cj+dk$ where $a^2+b^2+c^2+d^2=2N-1$. \qed\vspace{3mm} \section{Schr\"odinger representations}\label{Schr} Let $G$ be a finite Heisenberg group scheme of order $d^2$ over $S$. A representation of $G$ of weight 1 is a locally free $\O_S$-module together with the action of $G$ such that ${\Bbb G}_m\subset G$ acts by the identity character. We refer to chapter V of \cite{MB} for basic facts about such representations. In this section we study the problem of existence of a {\it Schr\"odinger representation} for $G$, i.~e. a weight-1 representation of $G$ of rank $d$ (the minimal possible rank). It is well known that such a representation exists if $S$ is the spectrum of an algebraically closed field (see e.g. \cite{MB}, V, 2.5.5). Another example is the following. As we already mentioned one can associate a finite Heisenberg group scheme $G(L)$ (called the Mumford group) to a line bundle $L$ on an abelian scheme $\pi:A\rightarrow S$ such that $K(L)$ is finite. Assume that the base scheme $S$ is connected. Then $R^i\pi_*(L)=0$ for $i\neq i(L)$ for some integer $i(L)$ (called the {\it index} of $L$) and $R^{i(L)}\pi_*(L)$ is a Schr\"odinger representation for $G(L)$ (this follows from \cite{Mum} III, 16 and \cite{Muk2}, prop.1.7). In general, L. Moret-Bailly showed in \cite{MB} that a Schr\"odinger representation exists after some smooth base change. The main result of this section is that for symmetric Heisenberg group schemes of odd order a Schr\"odinger representation always exists. Let $G$ be a symmetric finite Heisenberg group scheme of order $d^2$ over $S$. Then locally (in {\it fppf} topology) we can choose a Schr\"odinger representation $V$ of $G$. According to Theorem V, 2.4.2 of \cite{MB} for any weight-1 representation $W$ of $G$ there is a canonical isomorphism $V\otimes\underline{\operatorname{Hom}}_G(V,W)\widetilde{\rightarrow} W$. In particular, locally $V$ is unique up to an isomorphism and $\underline{\operatorname{Hom}}_G(V,V)\simeq\O$. Choose an open covering $U_i$ such that there exist Schr\"odinger representations $V_i$ for $G$ over $U_i$. For a sufficently fine covering we have $G$-isomorphisms $\phi_{ij}:V_i\rightarrow V_j$ on the intersections $U_i\cap U_j$, and $\phi_{jk}\phi_{ij}=\a_{ijk}\phi_{ik}$ on the triple intersections $U_i\cap U_j\cap U_k$ for some functions $\a_{ijk}\in\O^*(U_i\cap U_j\cap U_k)$. Then $(\a_{ijk})$ is a Cech 2-cocycle with values in ${\Bbb G}_m$ whose cohomology class $e(G)\in H^2(S,{\Bbb G}_m)$ doesn't depend on the choices made. Furthermore, by definition $e(G)$ is trivial if and only if there exists a global weight-1 representation we are looking for. Using the language of gerbs (see e.g. \cite{Gir}) we can rephrase the construction above without fixing an open covering. Namely, to each finite Heisenberg group $G$ we can associate the ${\Bbb G}_m$-gerb $\operatorname{Schr}_G$ on $S$ such that $\operatorname{Schr}_G(U)$ for an open set $U\subset S$ is the category of Schr\"odinger representations for $G$ over $U$. Then $\operatorname{Schr}_G$ represents the cohomology class $e(G)\in H^2(S,{\Bbb G}_m)$. Notice that the class $e(G)$ is actually represented by an Azumaya algebra $\AA(G)$ which is defined as follows. Locally, we can choose a Schr\'odinger representation $V$ for $G$ and put $\AA(G)=\underline{\operatorname{End}}(V)$. Now for two such representations $V$ and $V'$ there is a canonical isomorphism of algebras $\underline{\operatorname{End}}(V)\simeq\underline{\operatorname{End}}(V')$ induced by any $G$-isomorphism $f:V\rightarrow V'$ (since any other $G$-isomorphism differs from $f$ by a scalar), hence these local algebras glue together into a global Azumaya algebra $\AA(G)$ of rank $d^2$. In particular, $d\cdot e(G)=0$ (see e.g. \cite{Groth1}, prop. 1.4). Now let $W$ be a {\it global} weight-1 representation of $G$ which is locally free of rank $l\cdot d$ over $S$. Then we claim that $\underline{\operatorname{End}}_G(W)$ is an Azumaya algebra with the class $-e(G)$. Indeed, locally we can choose a representation $V$ of rank $d$ as above and a $G$-isomorphism $W\simeq V^l$ which induces a local isomorphism $\underline{\operatorname{End}}_G(W)\simeq\operatorname{Mat}_l(\O)$. Now we claim that there is a global algebra isomorphism $$\AA(G)\otimes\underline{\operatorname{End}}_G(W)\simeq\underline{\operatorname{End}}(W).$$ Indeed, we have canonical isomorphism of $G$-modules of weight 1 (resp. $-1$) $V\otimes\underline{\operatorname{Hom}}_G(V,W)\widetilde{\rightarrow}W$ (resp. $V^*\otimes\underline{\operatorname{Hom}}_G(V^*,W^*)\widetilde{\rightarrow}W^*$). Hence, we have a sequence of natural morphisms \begin{align*} &\underline{\operatorname{End}}(W)\simeq W^*\otimes W\simeq V^*\otimes V\otimes\underline{\operatorname{Hom}}_G(V^*,W^*)\otimes\underline{\operatorname{Hom}}_G(V,W)\rightarrow\\ &\rightarrow\underline{\operatorname{End}}(V)\otimes\operatorname{Hom}_{G\times G}(V^*\otimes V,W^*\otimes W)\rightarrow \underline{\operatorname{End}}(V)\otimes\underline{\operatorname{End}}_G(W) \end{align*} --- the latter map is obtained by taking the image of the identity section $\operatorname{id}\in V^*\otimes V$ under a $G\times G$-morphism $V^*\otimes V\rightarrow W^*\otimes W$. It is easy to see that the composition morphism gives the required isomorphism. This leads to the following statement. \begin{prop}\label{obst} For any finite Heisenberg group scheme $G$ over $S$ a canonical element $e(G)\in\operatorname{Br}(S)$ is defined such that $e(G)$ is trivial if and only if a Schr\"odinger representation for $G$ exists. Furthermore, $d\cdot e(G)=0$ where the order of $G$ is $d^2$, and if there exists a weight-1 $G$ representation which is locally free of rank $l\cdot d$ over $S$ then $l\cdot e(G)=0$. \end{prop} \begin{prop}\label{hom} The map $[G]\mapsto e(G)$ defines a homomorphism $\operatorname{WH}(S)\rightarrow\operatorname{Br}(S)$. \end{prop} \noindent {\it Proof} . First we have to check that if $I\subset K$ is a $G$-isotropic subgroup, $\widetilde{I}\subset G$ its lifting, and $\overline{G}=p^{-1}(I^{\perp})/\widetilde{I}$ then $e(\overline{G})=e(G)$. Indeed, there is a canonical equivalence of ${\Bbb G}_m$-gerbs $\operatorname{Schr}_G\rightarrow\operatorname{Schr}_{\overline{G}}$ given by the functor $V\mapsto V^{\widetilde{I}}$ where $V$ is a (local) Schr\"odinger representation of $G$. Next if $G=G_1\times_{{\Bbb G}_m} G_2$, then for every pair $(V_1,V_2)$ of weight-1 representations of $G_1$ and $G_2$ there is a natural structure of weight-1 $G$-representation on $V_1\otimes V_2$, hence we get an equivalence of ${\Bbb G}_m$-gerbs $\operatorname{Schr}_{G_1}+\operatorname{Schr}_{G_2}\rightarrow\operatorname{Schr}_G$ which implies the equality $e(G)=e(G_1)+e(G_2)$. At last, the map $V\rightarrow V^*$ induces an equivalence $\operatorname{Schr}_{G}^{op}\rightarrow\operatorname{Schr}_{G^{-1}}$ so that $e(G^{-1})=-e(G)$. \qed\vspace{3mm} \begin{thm}\label{odd} Let $G$ be a symmetric finite Heisenberg group scheme of odd order. Then $e(G)=0$, that is there exists a global Schr\"odinger representation for $G$. \end{thm} \noindent {\it Proof} . Let $[G]\in\operatorname{WH}'_{\operatorname{sym}}(S)$ be a class of $G$ in the Witt group. Then $4[G]=0$ by Theorem \ref{annih}, hence $4e(G)=0$ by Proposition \ref{hom}. On the other hand, $d\cdot e(G)=0$ by Proposition \ref{obst} where $d$ is odd, therefore, $e(G)=0$. \qed\vspace{3mm} Let us give an example of a symmetric finite Heisenberg group scheme of {\it even} order without a Schr\"odinger representation. First let us recall the construction from \cite{sympl} which associates to a group scheme $G$ over $S$ which is a central extension of a finite commutative group scheme $K$ by ${\Bbb G}_m$, and a $K$-torsor $E$ over $S$ a class $e(G,E)\in H^2(S,{\Bbb G}_m)$. Morally, the map $$H^1(S,K)\rightarrow H^2(E,{\Bbb G}_m): E\mapsto e(G,E)$$ is the boundary homomorphism corresponding to the exact sequence $$0\rightarrow {\Bbb G}_m\rightarrow G\rightarrow K\rightarrow 0.$$ To define it consider the category ${\cal C}$ of liftings of $E$ to to a $G$-torsor. Locally such a lifting always exists and any two such liftings differ by a ${\Bbb G}_m$-torsor. Thus, ${\cal C}$ is a ${\Bbb G}_2$-gerb over $S$, and by definition $e(G,E)$ is the class of ${\cal C}$ in $H^2(S,{\Bbb G}_m)$ Note that $e(G,E)=0$ if and only if there exists a $G$-equivariant line bundle $L$ over $E$, such that ${\Bbb G}_m\subset G$ acts on $L$ via the identity character. A $K$-torsor $E$ defines a commutative group extension $G_E$ of $K$ by ${\Bbb G}_m$ as follows. Choose local trivializations of $E$ over some covering $(U_i)$ and let $\a_{ij}\in K(U_i\cap U_j)$ be the corresponding 1-cocycle with values in $K$. Now we glue $G_E$ from the trivial extensions ${\Bbb G}_m\times K$ over $U_i$ by the following transition isomorphisms over $U_i\cap U_j$: $$f_{ij}:{\Bbb G}_m\times K\rightarrow{\Bbb G}_m\times K:(\lambda,x)\mapsto (\lambda e(x,\a_{ij}),x)$$ where $e:K\times K\rightarrow{\Bbb G}_m$ is the commutator form corresponding to $G$. It is easy to see that $G_E$ doesn't depend on a choice of trivializations. Now we claim that if $G$ is a Heisenberg group then \begin{equation}\label{diff} e(G,E)=e(G\otimes G_E)-e(G). \end{equation} This is checked by a direct computation with Cech cocycles. Notice that if $E^2$ is a trivial $K$-torsor then $G_E^2$ is a trivial central extension of $K$, hence $G_E$ is a symmetric extension. Thus, if $G$ is a symmetric Heisenberg group, then $G\otimes G_E$ is also symmetric. As was shown in \cite{sympl} the left hand side of (\ref{diff}) can be non-trivial. Namely, consider the case when $S=A$ is a principally polarized abelian variety over an algebraically closed field $k$ of characteristic $\neq2$. Let $K=A_2\times A$ considered as a (constant) finite group scheme over $A$. Then we can consider $E=A$ as a $K$-torsor over $A$ via the morphism $[2]:A\rightarrow A$. Now if $G\rightarrow A_2$ is a Heisenberg extension of $A_2$ (defined over $k$) then we can consider $G$ as a constant group scheme over $A$ and the class $e(G,E)$ is trivial if and only if $G$ embeds into the Mumford group $G(L)$ of some line bundle $L$ over $A$ (this embedding should be the identity on ${\Bbb G}_m$). When $\operatorname{NS}(A)={\Bbb Z}$ this means, in particular, that the commutator form $A_2\times A_2\rightarrow{\Bbb G}_m$ induced by $G$ is proportional to the symplectic form given by the principal polarization. When $\dim A\ge 2$ there is a plenty of other symplectic forms on $A_2$, hence, $e(G,E)$ can be non-trivial. Now we are going to show that one can replace $A$ by its general point in this example. In other words, we consider the base $S=\operatorname{Spec}(k(A))$ where $k(A)$ is the field of rational functions on $A$. Then $E$ gets replaced by $\operatorname{Spec}(k(A))$ considered as a $A_2$-torsor over itself corresponding to the Galois extension $$[2]^*:k(A)\rightarrow k(A): f\mapsto f(2\cdot)$$ with the Galois group $A_2$. Note that the class $e(G,E)$ for any Heisenberg extension $G$ of $A_2$ by $k^*$ is annihilated by the pull-back to $E$, hence, it is represented by the class of Galois cohomology $H^2(A_2,k(A)^*)\subset\operatorname{Br}(k(A))$ where $A_2$ acts on $k(A)$ by translation of argument. It is easy to see that this class is the image of the class $e_G\in H^2(A_2,k^*)$ of the central extension $G$ under the natural homomorphism $H^2(A_2,k^*)\rightarrow H^2(A_2,k(A)^*)$. From the exact sequence of groups $$0\rightarrow k^*\rightarrow k(A)^*\rightarrow k(A)^*/k^*\rightarrow 0$$ we get the exact sequence of cohomologies $$0\rightarrow H^1(A_2,k(A)^*/k^*)\rightarrow H^2(A_2,k^*)\rightarrow H^2(A_2,k(A)^*)$$ (note that $H^1(A_2,k(A)^*)=0$ by Hilbert theorem 90). It follows that central extensions $G$ of $A_2$ by $k^*$ with trivial $e(G,E)$ are classified by elements of $H^1(A_2,k(A)^*/k^*)$. \begin{lem} Let $A$ be a principally polarized abelian variety over an algebraically closed field $k$ of characteristic $\neq 2$. Assume that $\operatorname{NS}(A)={\Bbb Z}$. Then $H^1(A_2,k(A)^*/k^*)={\Bbb Z}/2{\Bbb Z}$. \end{lem} \noindent {\it Proof} . Interpreting $k(A)^*/k^*$ as the group of divisors linearly equivalent to zero we obtain the exact sequence $$0\rightarrow k(A)^*/k^*\rightarrow\operatorname{Div}(A)\rightarrow\operatorname{Pic}(A)\rightarrow 0,$$ where $\operatorname{Div}(A)$ is the group of all divisors on $A$. Note that as $A_2$-module $\operatorname{Div}(A)$ is decomposed into a direct sum of modules of the form ${\Bbb Z}^{A_2/H}$ where $H\subset A_2$ is a subgroup. Now by Shapiro lemma we have $H^1(A_2,{\Bbb Z}^{A_2/H})\simeq H^1(H,{\Bbb Z})$, and the latter group is zero since $H$ is a torsion group. Hence, $H^1(A_2,\operatorname{Div}(A))=0$. Thus, from the above exact sequence we get the identification $$H^1(A_2,k(A)^*/k^*)\simeq \operatorname{coker}(\operatorname{Div}(A)^{A_2}\rightarrow\operatorname{Pic}(A)^{A_2}).$$ Now we use the exact sequence $$0\rightarrow\operatorname{Pic}^0(A)\rightarrow\operatorname{Pic}(A)\rightarrow\operatorname{NS}(A)\rightarrow 0,$$ where $\operatorname{Pic}^0(A)=\hat{A}(k)$. Since the actions of $A_2$ on $\operatorname{Pic}^0(A)$ and $\operatorname{NS}(A)$ are trivial we have the induced exact sequence $$0\rightarrow\operatorname{Pic}^0(A)\rightarrow\operatorname{Pic}(A)^{A_2}\rightarrow\operatorname{NS}(A).$$ The image of the right arrow is the subgroup $2\operatorname{NS}(A)\subset\operatorname{NS}(A)$. Note that $\operatorname{Pic}^0(A)=[2]^*\operatorname{Pic}^0(A)$, hence this subgroup belongs to the image of $[2]^*\operatorname{Div}(A)\subset\operatorname{Div}(A)^{A_2}$. Thus, we deduce that $$H^1(A_2,k(A)^*/k^*)\simeq\operatorname{coker}(\operatorname{Div}(A)^{A_2}\rightarrow 2\operatorname{NS}(A)).$$ Let $[L]\subset\operatorname{NS}(A)$ be the generator corresponding to a line bundle $L$ of degree 1 on $A$. Then $L^4=[2]^*L$, hence $4\cdot [L]=[L^4]$ belongs to the image of $\operatorname{Div}(A)^{A_2}$. On the other hand, it is easy to see that there is no $A_2$-invariant divisor representing $[L^2]$, hence $$H^1(A_2,k(A)^*/k^*)\simeq{\Bbb Z}/2{\Bbb Z}.$$ \qed\vspace{3mm} It follows that under the assumptions of this lemma there is a unique Heisenberg extensions $G$ of $A_2$ by $k^*$ with the trivial class $e(G,E)$ (the Mumford extension corresponding to $L^2$, where $L$ is a line bundle of degree 1 on $A$). Hence, for $g\ge 2$ there exists a Heisenberg extension with a non-trivial class $e(G,E)\in\operatorname{Br}(k(A))$. \section{Representations of the Heisenberg groupoid} Recall that the Heisenberg group $H(W)$ associated with a symplectic vector space $W$ is a central extension $$0\rightarrow T\rightarrow H(W)\rightarrow W\rightarrow 0$$ of $W$ by the 1-dimensional torus $T$ with the commutator form $\exp(B(\cdot,\cdot))$ where $B$ is the symplectic form. In this section we consider an analogue of this extension in the context of abelian schemes (see \cite{Weilrep} , sect. 7, \cite{sympl}). Namely, we replace a vector space $W$ by an abelian scheme $X/S$. Bilinear forms on $W$ get replaced by biextensions of $X^2$. Recall that a {\it biextension} of $X^2$ is a line bundle ${\cal L}$ on $X^2$ together with isomorphisms \begin{align*} &{\cal L}_{x+x',y}\simeq {\cal L}_{x,y}\otimes {\cal L}_{x',y},\\ &{\cal L}_{x,y+y'}\simeq {\cal L}_{x,y}\otimes {\cal L}_{x,y'} \end{align*} --- this is a symbolic notation for isomorphisms $(p_1+p_2,p_3)^*{\cal L}\simeq p_{13}^*{\cal L}\otimes p_{23}^*{\cal L}$ and $(p_1,p_2+p_3)^*{\cal L}\simeq p_{12}^*{\cal L}\otimes p_{13}^*{\cal L}$ on $X^3$, satisfying some natural cocycle conditions (see e.g. \cite{Breen}). The parallel notion to the skew-symmetric form on $W$ is that of a {\it skew-symmetric biextension} of $X^2$ which is a biextension ${\cal L}$ of $X^2$ together with an isomorphism of biextensions $\phi:\sigma^*{\cal L}\widetilde{\rightarrow} {\cal L}^{-1}$, where $\sigma:X^2\rightarrow X^2$ is the permutation of factors, and a trivialization $\Delta^*{\cal L}\simeq\O_X$ of ${\cal L}$ over the diagonal $\Delta:X\rightarrow X^2$ compatible with $\phi$. A skew-symmetric biextension ${\cal L}$ is called {\it symplectic} if the corresponding homomorphism $\psi_{{\cal L}}:X\rightarrow\hat{X}$ (where $\hat{X}$ is the dual abelian scheme) is an isomorphism. An {\it isotropic} subscheme (with respect to ${\cal L}$) is an abelian subscheme $Y\subset X$ such that there is an isomorphism of skew-symmetric biextensions ${\cal L}|_{Y\times Y}\simeq\O_{Y\times Y}$. This is equivalent to the condition that the composition $Y\stackrel{i}{\rightarrow} X\stackrel{\psi_{{\cal L}}}{\rightarrow}\hat{X} \stackrel{\hat{i}}{\rightarrow} \hat{Y}$ is zero. An isotropic subscheme $Y\subset X$ is called {\it lagrangian} if the morphism $Y\rightarrow \ker(\hat{i})$ induced by $\psi_{{\cal L}}$ is an isomorphism. In particular, for such a subscheme the quotient $X/Y$ exists and is isomorphic to $\hat{Y}$. Note that to define the Heisenberg group extension it is not sufficient to have a symplectic form $B$ on $W$: one needs a bilinear form $B_1$ such that $B(x,y)=B_1(x,y)-B_1(y,x)$. In the case of the real symplectic space one can just take $B_1=B/2$, however in our situation we have to simply add necessary data. An {\it enhanced} symplectic biextension $(X,{\cal B})$ is a biextension ${\cal B}$ of $X^2$ such that ${\cal L}:={\cal B}\otimes\sigma^*{\cal B}^{-1}$ is a symplectic biextension. The standard enhanced symplectic biextension for $X=\hat{A}\times A$, where $A$ is any abelian scheme, is obtained by setting $${\cal B}=p_{14}^*\cal P\in\operatorname{Pic}(\hat{A}\times A\times \hat{A}\times A),$$ where $\cal P$ is the normalized Poincar\'e line bundle on $A\times\hat{A}$. Given an enhanced symplectic biextension $(X,{\cal B})$ one defines the {\it Heisenberg groupoid} $H(X)=H(X,{\cal B})$ as the stack of monoidal groupoids such that $H(X)(S')$ for an $S$-scheme $S'$ is the monoidal groupoid generated by the central subgroupoid ${\cal P}ic(S')$ of ${\Bbb G}_m$-torsors on $S'$ and the symbols $T_x$, $x\in X(S')$ with the composition law $$T_x\circ T_{x'}= {\cal B}_{x,x'} T_{x+x'}.$$ The Heisenberg groupoid is a central extension of $X$ by the stack of line bundles on $S$ in the sense of Deligne \cite{Des}. In \cite{Weilrep} we considered the action of $H(\hat{A}\times A)$ on ${\cal D}^b(A)$ which is similar to the standard representation of the Heisenberg group $H(W)$ on functions on a lagrangian subspace of $W$. Below we construct similar representations of the Heisenberg groupoid $H(X)$ associated with lagrangian subschemes in $X$. Further, we construct intertwining functors for two such representations corresponding to a pair of lagrangian subschemes, and consider the analogue of Maslov index for a triple of lagrangian subschemes that arises when composing these intertwining functors. To define an action of $H(X)$ associated with a lagrangian subscheme one needs some auxilary data described as follows. An {\it enhanced} lagrangian subscheme (with respect to ${\cal B}$) is a pair $(Y,\a)$ where $Y\subset X$ is a lagrangian subscheme with respect to $X$, $\a$ is a line bundle on $Y$ with a rigidification along the zero section such that an isomorphism of symmetric biextensions $\Lambda(\a)\simeq {\cal B}|_{Y\times Y}$ is given, where $\Lambda(\a)=(p_1+p_2)^*\a\otimes p_1^*\a^{-1}\otimes p_2^*\a^{-1}$. Note that an enhanced lagrangian subscheme is a particular case of an {\it isotropic pair} as defined in \cite{Weilrep} II, 7.3. With every enhanced lagrangian subscheme $(Y,\a)$ one can associate a representation of $H(X)(S)$ as follows (see \cite{Weilrep},\cite{sympl}). Let ${\cal D}(Y,\a)$ be the category of pairs $({\cal F},a)$ where ${\cal F}\in{\cal D}^b(X)$, $a$ is an isomorphism in ${\cal D}^b(Y\times X)$: \begin{equation}\label{Schrsp} a:(i_Yp_1+p_2)^*{\cal F}\widetilde{\rightarrow} {\cal B}^{-1}|_{Y\times X} \otimes p_1^*\a^{-1}\otimes p_2^*{\cal F} \end{equation} where $i_Y:Y\hookrightarrow X$ is the embedding, such that $(e\times\operatorname{id})^*a=\operatorname{id}$. These data should satisfy the following cocycle condition: $$(p_1+p_2,p_3)^*a=(p_2,p_3)^*a\circ (p_1,i_Yp_2+p_3)^*a$$ in ${\cal D}^b(Y\times Y\times X)$. Then there is a natural action of the Heisenberg groupoid $H(X)(S)$ on the category ${\cal D}(Y,\a)$ such that a line bundle $M$ on $S$ acts by tensoring with $p^*M$ and a generator $T_x$ acts by the functor \begin{equation}\label{act} {\cal F}\mapsto {\cal B}|_{X\times x}\otimes t_x^*({\cal F}). \end{equation} If $S'$ is an $S$-scheme then this action is compatible with the action of $H(X)(S')$ on ${\cal D}(Y_{S'},\a_{S'})$ via pull-back functors. Let $\delta_{Y,\a}\in{\cal D}(Y,\a)$ be the following object (delta-function at $(Y,\a)$): \begin{equation}\label{delta} \delta_{Y,\a}=i_{Y*}(\a^{-1}) \end{equation} where $i_Y:Y\rightarrow X$ is the embedding. It is easy to see that $\delta_{Y,\a}$ has a canonical structure of an object of ${\cal D}(Y,\a)$ and for $y\in Y$ one has $T_y(\delta_{Y,\a})\simeq\a_y^{-1}\delta_{Y,\a}$. Let $(Y,\a)$, $(Z,\b)$ be a pair of enhanced lagrangian subschemes in $X$, such that $Y\cap Z$ is finite over $S$. Then the natural morphism $Y\rightarrow X/Z\simeq\hat{Z}$ is an isogeny, hence, $Y\cap Z$ is flat over $S$. Note that we have isomorphisms of biextensions $\Lambda(\a|_{Y\cap Z})\simeq\Lambda(\b|_{Y\cap Z})\simeq {\cal B}|_{(Y\cap Z)^2}$, hence the trivialization of $\Lambda(\b|_{Y\cap Z}\otimes \a^{-1}|_{Y\cap Z})$. Thus, the ${\Bbb G}_m$-torsor $G_{Y,Z}=\b|_{Y\cap Z}\otimes \a^{-1}|_{Y\cap Z}$ has a natural structure of a central extension of $Y\cap Z$ by ${\Bbb G}_m$. Furthermore, the corresponding commutator form $(Y\cap Z)^2\rightarrow{\Bbb G}_m$ is non-degenerate since it corresponds to the canonical duality between $Y\cap Z=\ker(Y\rightarrow\hat{Z})$ and $Y\cap Z=\ker(Z\rightarrow\hat{Y})$ (see \cite{sympl}, remark after Prop. 3.1). Thus, $G_{Y,Z}$ is a finite Heisenberg group scheme over $S$. If the line bundles $\a$ and $\b$ are symmetric then so is $G_{Y,Z}$. Let $V$ be a Schr\"odinger representation of $G_{Y,Z}$. Generalizing the construction of \cite{sympl} we define the $H(X)(S)$-intertwining operator $$R(V):{\cal D}(Y,\a)\rightarrow{\cal D}(Z,\b): {\cal F}\mapsto \underline{\operatorname{Hom}}_{G_{Y,Z}}(V,p_{2*}({\cal B}|_{Z\times X}\otimes p_1^*\b\otimes (i_Zp_1+p_2)^*{\cal F})).$$ Here $p_1$ and $p_2$ are the projections of the product $Z\times_S X$ onto its factors. The $G_{Y,Z}$-module structure on $p_{2*}({\cal B}|_{Z\times X}\otimes p_1^*\b\otimes(i_Zp_1+p_2)^*{\cal F})$ comes from the natural $G_{Y,Z}$-action on $I({\cal F})={\cal B}|_{Z\times X}\otimes p_1^*\b\otimes(i_Zp_1+p_2)^*{\cal F}$ which is compatible with the action of $Y\cap Z$ on $Z\times X$ by the translation of the first argument and arises from the canonical isomorphism \begin{equation}\label{integrand} I({\cal F})_{(z+u,x)}\simeq \b_u\a_u^{-1} I({\cal F})_{(z,x)} \end{equation} where $z\in Z$, $x\in X$, $u\in Y\cap Z$ (one should consider this as an isomorphism in ${\cal D}^b((Y\cap Z)\times Z\times X)$). When $V$ is the representation associated with a lagrangian subgroup scheme $H\subset G_{Y,Z}$ this functor coincides with the one defined in \cite{sympl}. Let us call an enhanced lagrangian subscheme $(Y,\a)$ {\it admissible} if the projection $X\rightarrow X/Y$ splits. For such a subscheme we have an equivalence ${\cal D}(Y,\a)\simeq{\cal D}^b(X/Y)$. Namely, let $s_{X/Y}:X/Y\rightarrow X$ be a splitting of the canonical projection $q_{X/Y}:X\rightarrow X/Y$. Let $q_Y=\operatorname{id}-s_{X/Y}q_{X/Y}:X\rightarrow Y$ be the corresponding projection to $Y$. Then the functors ${\cal F}\mapsto s_{X/Y}^*{\cal F}$ and ${\cal G}\mapsto (q_Y,s_{X/Y}q_{X/Y})^*{\cal B}^{-1}\otimes q_Y^*\a^{-1}\otimes q_{X/Y}^*{\cal G}$ where ${\cal F}\in{\cal D}(Y,\a)$, ${\cal G}\in{\cal D}^b(X/Y)$ give the required equivalence. When $(Y,\a)$ and $(Z,\b)$ are both admissible we can represent the above functor $R(V):{\cal D}^b(X/Y)\rightarrow{\cal D}^b(X/Z)$ in the standard "integral" form. \begin{lem}\label{compker} Assume that $(Y,\a)$ and $(Z,\b)$ are admissible, $Y\cap Z$ is finite. Then $R(V)({\cal G})\simeq p_{2*}(p_1^*{\cal G}\otimes {\cal K}(V))$ where $p_i$ are the projections of $X/Y\times X/Z$ on its factors, ${\cal K}(V)$ is the following vector bundle on $X/Y\times X/Z$: \begin{eqnarray}\label{kernel} {\cal K}(V)=(p_1-q_{X/Y}s_{X/Z}p_2)^*E(V)\otimes (s_{X/Y}p_1-s_{X/Z}p_2,s_{X/Y}p_2)^*{\cal B}\otimes \nonumber\\ (s_{X/Y}(p_1-q_{X/Y}s_{X/Z}p_2),q_Ys_{X/Z}p_2)^*{\cal B}\otimes (q_Ys_{X/Z}p_2)^*\a^{-1} \end{eqnarray} where $s_{X/Y}:X/Y\rightarrow X$ (resp. $s_{X/Z}:X/Z\rightarrow X$) is the splitting of the projection $q_{X/Y}:X\rightarrow X/Y$ (resp. $q_{X/Z}:X\rightarrow X/Z$), $q_Y=\operatorname{id}-s_{X/Y}q_{X/Y}$, $E(V)$ is the following bundle on $X/Y$: \begin{equation}\label{kernelaux} E(V)=\underline{\operatorname{Hom}}_{G_{Y,Z}}(V,(q_{X/Y}i_Z)_*(\b\otimes (q_Yi_Z)^*\a^{-1}\otimes (i_Z,s_{X/Y}q_{X/Y}i_Z)^*{\cal B}^{-1})) \end{equation} where $i_Y:Y\rightarrow X$, $i_Z:Z\rightarrow X$ are the embeddings. \end{lem} \noindent {\it Proof} . By definition we have \begin{align*} &R(V)({\cal G})_{\bar{x}}\simeq\\ &\underline{\operatorname{Hom}}(V,\int_Z \b_z {\cal B}_{z,s_{X/Z}(\bar{x})} {\cal B}^{-1}_{q_Y(z+s_{X/Z}(\bar{x})),s_{X/Y}q_{X/Y}(z+s_{X/Z} (\bar{x}))}\a^{-1}_{q_Y(z+s_{X/Z}(\bar{x}))}{\cal G}_{q_{X/Y}(z+s_{X/Z} (\bar{x}))} dz) \end{align*} where $\bar{x}\in X/Z$, $z\in Z$. Using the isomorphism $\a_{q_Y(z+s_{X/Z}(\bar{x}))}\simeq \a_{q_Y(z)}\a_{q_Ys_{X/Z}(\bar{x})} {\cal B}_{q_Y(z),q_Ys_{X/Z}(\bar{x})}$ and collecting together terms depending only on $\bar{z}=q_{X/Y}(z)$ we get \begin{align*} &R(V)({\cal G})_{\bar{x}}\simeq \underline{\operatorname{Hom}}(V,\int_Z \b_z \a_{q_Y(z)}^{-1} {\cal B}^{-1}_{z, s_{X/Y}\bar{z}} \\ &\Bigl( {\cal B}_{s_{X/Y}(\bar{z}+q_{X/Y}s_{X/Z}(\bar{x}))-s_{X/Z}(\bar{x}), s_{X/Y}(\bar{z}+q_{X/Y}s_{X/Z}(\bar{x}))} {\cal B}_{s_{X/Y}(\bar{z}),q_Ys_{X/Z}(\bar{x})} \a^{-1}_{q_Ys_{X/Z}(\bar{x})} {\cal G}_{\bar{z}+q_{X/Y}s_{X/Z}(\bar{x})}\Bigr) dz)\simeq\\ &\int_{X/Y} E(V)_{\bar{z}} {\cal B}_{s_{X/Y}(\bar{z}+q_{X/Y}s_{X/Z}(\bar{x}))-s_{X/Z}(\bar{x}), s_{X/Y}(\bar{z}+q_{X/Y}s_{X/Z}(\bar{x}))} {\cal B}_{s_{X/Y}(\bar{z}),q_Ys_{X/Z}(\bar{x})}\\ &\a^{-1}_{q_Ys_{X/Z}(\bar{x})} {\cal G}_{\bar{z}+q_{X/Y}s_{X/Z}(\bar{x})} d\bar{z} \end{align*} where $\bar{z}$ is now considered as a variable on $X/Y$. Making the change of variables $\bar{z}\mapsto \bar{z}-q_{X/Y}s_{X/Z}(\bar{x})$ we arrive to the formula (\ref{kernel}). \qed\vspace{3mm} \begin{thm}\label{eq} Assume that $(Y,\a)$ and $(Z,\b)$ are admissible and $Y\cap Z$ is finite, then $R(V)$ is an equivalence of categories. Let $(T,\gamma)$ be an admissible enhanced lagrangian subscheme such that $Y\cap T$ and $Z\cap T$ are finite, $W$ (resp. $U$) be a Schr\"odinger representation for $G_{Y,T}$ (resp. $G_{Z,T}$). Then $$R(U)\circ R(V)\simeq R(W)\otimes M[n]$$ for some line bundle $M$ on $S$ and some integer $n$. \end{thm} \noindent {\it Proof} . The direct computation shows that the kernel ${\cal K}(V)\in{\cal D}^b(X/Y\times X/Z)$ constructed above satisfies the "uniform" intertwining property (with respect to $H(X)$-action) defined in \cite{Weilrep}. Hence, the analogue of Schur lemma for the action of $H(X)$ on ${\cal D}^b(X/Y)$ where $Y$ is an admissible lagrangian subscheme (see \cite{Weilrep} Thm 7.9) implies that $$p_{13*}(p_{12}^*{\cal K}(V)\otimes p_{23}^*{\cal K}(V^*))\simeq\Delta_*(F)$$ where ${\cal K}(V^*)$ is the similar kernel on $X/Z\times X/Y$ giving rise to the functor $R(V^*):{\cal D}^b(X/Z)\rightarrow{\cal D}^b(X/Y)$, $p_{ij}$ are the projections of $X/Y\times X/Z\times X/Y$ on the pairwise products, $\Delta:X/Y\rightarrow (X/Y)^2$ is the diagonal embedding. In the case when $S$ is the spectrum of a field we know that $F\simeq N[n]$ for some line bundle $N$ on $S$ and some integer $n$ (see \cite{Weilrep}). By Prop.1.7 of \cite{Muk2} this implies that the same is true when $S$ is connected. Therefore, in this case the composition $R(V^*)\circ R(V)$ is isomorphic to the tensoring with $N[n]$. Repeating this for the composition $R(V)\circ R(V^*)$ we conclude that $R(V)$ is an equivalence. Similar argument works for the proof of the second assertion. \qed\vspace{3mm} \begin{rems} 1. Most probably, one can extend this theorem to the case of arbitrary enhanced lagrangian subschemes. However, it seems that the definition of ${\cal D}(Y,\a)$ should be modified in this case (one should start with appropriate category of complexes and then localize it). \noindent 2. An integer $n$ and a line bundle $M$ on $S$ appearing in the above theorem should be considered as analogues of the Maslov index (see \cite{LV}) for a triple $(Y,Z,T)$. Note that different choices of Schr\"odinger representations $V$, $W$, and $U$ above affect $M$ but not $n$, hence the function $n(Y,Z,T)$ behaves very much like the classical Maslov index (cf. \cite{Orlov}). \end{rems} Let $(Y,\a)$, $(Z,\b)$ and $(T,\gamma)$ be a triple of enhanced lagrangian subschemes in $X$. Let us denote by $K=K(Y,Z,T)$ the kernel of the homomorphism $Y\times Z\times T\rightarrow X:(y,z,t)\mapsto y+z+t$. Let $p_Y:K\rightarrow Y$, $p_Z:K\rightarrow Z$ and $p_T:K\rightarrow T$ be the restrictions to $K$ of the natural projections from $Y\times Z\times T$ to its factors. Consider the following line bundle on $K$: \begin{equation}\label{MYZT} M(Y,Z,T)=(-p_Y)^*\a^{-1}\otimes p_Z^*\b\otimes p_T^*\gamma\otimes (p_Z,p_T)^*{\cal B}|_{Z\times T}. \end{equation} Then $M(Y,Z,T)$ has a canonical cube structure induced by that of $\a$, $\b$, $\gamma$ and ${\cal B}$. \begin{lem} There are canonical isomorphisms of line bundles with cube structures on $K$ $$M(Y,Z,T)\simeq M(Z,T,Y)\simeq M(T,Y,Z).$$ There is a canonical isomorphism of biextensions of $K\times K$: $$\Lambda(M(Y,Z,T))\simeq (p_Zp_2,p_Tp_1)^*{\cal L}$$ where $p_i$ are the projections of $K\times K$ on its factors. \end{lem} \noindent {\it Proof} . We have $$(M(Z,T,Y)\otimes M(Y,Z,T)^{-1})_{y,z,t}= \a_y\a_{-y}\b_z^{-1}\b_{-z}^{-1}{\cal B}_{t,y}{\cal B}_{z,t}^{-1}\simeq {\cal B}_{y,y}{\cal B}_{z,z}^{-1}{\cal B}_{t,y}{\cal B}_{z,-t}$$ where $y+z+t=0$ (here we used the isomorphism $\a_y\a_{-y}\simeq {\cal B}_{y,-y}\simeq {\cal B}_{y,y}^{-1}$ and the similar isomorphism for $\b$). It is easy to see that when we substitute $t=-y-z$ the right hand side becomes trivial. The second isomorphism is obtained as follows: $$\Lambda(M(Y,Z,T))_{(y,z,t),(y',z',t')}\simeq {\cal B}_{-y,-y'}^{-1}{\cal B}_{z,z'}{\cal B}_{t,t'}{\cal B}_{z,t'}{\cal B}_{z',t}.$$ If we substitute $-y=z+t$, $-y'=z'+t'$ the right hand side becomes ${\cal B}_{t,z'}^{-1}{\cal B}_{z',t}\simeq {\cal L}_{z',t}$. \qed\vspace{3mm} Consider the embedding $Z\cap T\hookrightarrow K:u\mapsto (0,-u,u)$. Then the previous lemma implies that $\Lambda(M(Y,Z,T))$ is trivial over $(Z\cap T)\times K$. Hence, $M(Y,Z,T)|_{Z\cap T}$ has a structure of central extension and the action of $Z\cap T$ on $K$ by translations lifts to an action of this central extension on $M(Y,Z,T)$. Moreover, we have a canonical isomorphism of central extensions $$M(Y,Z,T)_{(0,-u,u)}\simeq\gamma_u\b_{-u}{\cal B}_{-u,u}\simeq \gamma_u\b_u^{-1}=(G_{Z,T})_u.$$ Hence, there is an action of $G_{Z,T}$ on $M(Y,Z,T)$ compatible with the action of $Z\cap T$ on $K$ by translations. Using cyclic permutation we get embeddings of $Y\cap T$ and $Y\cap Z$ into $K$ and it is easy to see that the images of the three embeddings are independent so that we get an embedding $(Y\cap Z)\times_S (Z\cap T)\times_S (Y\cap T)\hookrightarrow K$ and the compatible action of $G_{Y,Z}\times_{{\Bbb G}_m}G_{Z,T}\times_{{\Bbb G}_m} G_{T,Y}$ on $M(Y,Z,T)$. \begin{thm}\label{intert} With the notation and assumptions of Theorem \ref{eq} we have $$M[n]\simeq\underline{\operatorname{Hom}}_{G_{Y,Z,T}}(V_{Y,Z,T},p_*M(Y,Z,T))$$ where $G_{Y,Z,T}=G_{Y,Z}\times_{{\Bbb G}_m} G_{Z,T}\times_{{\Bbb G}_m} \times G_{T,Y}$, $V_{Y,Z,T}=V\otimes U\otimes W^*$, $p:K\rightarrow S$ is the projection. \end{thm} \noindent {\it Proof} . Let us compare the restrictions of $R(U)\circ R(V)(\delta)$ and $R(W)(\delta)$ to the zero section, where $\delta=\delta_{Y,\a}\in{\cal D}(Y,\a)$ is the delta-function at $Y$ defined by (\ref{delta}). On the one hand, we have \begin{align*} &R(U)\circ R(V)(\delta)_0\simeq\underline{\operatorname{Hom}}_{G_{Y,Z}\times_{{\Bbb G}_m} G_{Z,T}} (V\otimes U,\int_{Z\times T} \gamma_t {\cal B}_{z,t}\b_z\delta_{z+t} dz dt)\simeq\\ &\simeq\underline{\operatorname{Hom}}_{G_{Y,Z}\times_{{\Bbb G}_m} G_{Z,T}}(V\otimes U, \int_{K}M(Y,Z,T)). \end{align*} On the other hand, $$R(W)(\delta)_0\simeq\underline{\operatorname{Hom}}_{G_{Y,T}}(W, \int_{Y\cap T} \gamma_u\a_u^{-1} du)\simeq W^*$$ since $\int_{Y\cap T}G_{Y,T}\simeq W^*\otimes W$ by \cite{MB} V 2.4.2. Therefore, $$\int_{K} M(Y,Z,T)\simeq V\otimes U\otimes W^*\otimes M[n]$$ as a representation of $G_{Y,Z,T}$. \qed\vspace{3mm} Consider the following example. Let $X=\hat{A}\times A$, ${\cal B}=p_{14}^*\cal P$, $(Y,\a)=(A,\O_A)$, $(T,\gamma)=(\hat{A},\O_{\hat{A}})$, and $(Z,\b)=(Z_{\phi,m},\b)$ where $\phi=\phi_L:A\rightarrow\hat{A}$ is the symmetric isogeny associated with a rigidified line bundle $L$ on $A$, $Z_{\phi,m}=(\phi,m\operatorname{id}_A)(A)\simeq A/\ker(\phi_m)$ where $\phi_m=\phi|_{A_m}$, $\b$ is obtained from $L^m$ by descent (such $\b$ always exists if $m$ is odd, since $\ker(\phi_m)$ is isotropic with respect to $e^{L^m}$). Then $K(Y,Z,T)\simeq Z$, $M(Y,Z,T)\simeq\b$, $Y\cap T=0$, $Y\cap Z\simeq \ker(\phi)/\ker(\phi_n)$, and $Z\cap T\simeq A_n/\ker(\phi_n)$. Hence, if we take $W=\O_S$ we get $$M[n]\simeq\underline{\operatorname{Hom}}_{G_{Y,Z}\times_{{\Bbb G}_m}G_{Z,T}}(V\otimes U,p_*\b).$$ In particular, when $m=1$ we have $Z\simeq A$, $\b=L$, $Z\cap T=0$, and $G_{Y,Z}=G(L)$. Thus, if we take $U=W=\O_S$ we obtain $M[n]\simeq\underline{\operatorname{Hom}}_{G(L)}(V,p_*L)$. Note that if one of the pairwise intersections of $Y$, $Z$ and $T$ is trivial then $K(Y,Z,T)$ is an abelian scheme over $S$. More precisely, if say $Y\cap Z=0$ then $K(Y,Z,T)\simeq T$ and it is easy to see from the above considerations that in this case we have an isomorphism of Heisenberg groups \begin{equation}\label{GYZT} G(M(Y,Z,T))\simeq G_{Y,Z}\times_{{\Bbb G}_m}G_{Z,T}\times_{{\Bbb G}_m}G_{T,Y}. \end{equation} \section{Weil representation on the derived category of an abelian scheme} In this section the base scheme $S$ is always assumed to noetherian, normal and connected. Let $K$ denotes the field of rational functions on $S$. \begin{lem}\label{extend} Let $A$ and $A'$ be abelian schemes over $S$, $A_K$ and $A'_K$ be their general fibers which are abelian varieties over $K$. Then the restriction map $$\operatorname{Hom}_S(A,A')\rightarrow\operatorname{Hom}_K(A_K,A'_K):f\mapsto f|_K$$ is an isomorphism. The morphism $f$ is an isogeny if and only if $f|_K$ is an isogeny. \end{lem} \noindent {\it Proof} . The proof of the first assertion is similar to the proof of the fact that an abelian scheme over a Dedekind scheme is a N\'eron model of its generic fiber (see \cite{Neron} 1.2.8). We have to check that any homomorphism $f_K:A_K\rightarrow A'_{K}$ extends to a homomorphism $f:A\rightarrow A'$. Let $\phi:A\rightarrow A'$ be the rational map defined by $f_K$. Since $A$ is normal, by a valuative criterion of properness $\phi$ is defined in codimension $\leq 1$. Let $V\subset A$ be a non-empty subscheme over which $\phi$ is defined. Then since the projection $p:A\rightarrow S$ is flat and of finite presentation it is open. Thus, $U=p(V)$ is open, and $\phi_U:A_U\rightarrow A'_U$ is a $U$-rational map in the terminology of \cite{Neron}. By Weil's theorem (see \cite{Neron} 4.4.1) $\phi_U$ is defined everywhere, hence we get a homomorphism $f_U:A_U\rightarrow A'_U$ extending $f_K$. It remains to invoke Prop. I 2.9 of \cite{FC} to finish the proof. The part concerning isogenies can be proven by exactly the same argument as in \cite{Neron} 7.3 prop. 6: starting with an isogeny $f|_K:A_K\rightarrow A'_K$ we can find another isogeny $g|_K:A'_K\rightarrow A_K$ such that the composition $g|_Kf|_K$ is the multiplication by an integer $l$ on $A_K$. By the first part we can extend $g_K$ to a homomorphism $g:A'\rightarrow A$. This implies that $gf=l_A$ --- the multiplication by $l$ morphism on $A$. It follows that the restriction of $f$ to each fiber is an isogeny, hence $f$ is an isogeny itself. \qed\vspace{3mm} For an abelian scheme $A$ the group $\operatorname{SL}_2(A)$ is defined as the subgroup of automorphisms of $\hat{A}\times A$ preserving the line bundle $p_{14}^*\cal P\otimes p_{23}^*\cal P^{-1}$ on $(\hat{A}\times A)^2$. More explicitly, if we write an automorphism of $\hat{A}\times A$ as a matrix $g=\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)$ where $a_{11}\in\operatorname{Hom}(\hat{A},\hat{A})$, $a_{12}\in\operatorname{Hom}(A,\hat{A})$ etc., then $g\in\operatorname{SL}_2(A)$ if and only if the inverse automorphism $g^{-1}$ is given by the matrix $\left(\matrix \hat{a}_{22} & -\hat{a}_{12}\\ -\hat{a}_{21} & \hat{a}_{11} \endmatrix \right)$. It follows from Lemma \ref{extend} that when the base $S$ is normal we have $\operatorname{SL}_2(A)\simeq\operatorname{SL}_2(A_K)$. Now similarly to the classical picture one has to consider the group of automorphisms of the Heisenberg extension $H(X)$ corresponding to $X=\hat{A}\times A$ with the structure of enhanced symplectic biextension given by ${\cal B}=p_{14}^*\cal P$. Namely, we define $\widetilde{\operatorname{SL}}_2(A)$ as the group of triples $g=(\bar{g},L^g,M^g)$ where $\overline{g}= \left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right) \in\operatorname{SL}_2(A)$, $L^g$ (resp. $M^g$) is a line bundle on $\hat{A}$ (resp. $A$) rigidified along the zero section, such that \begin{equation} \phi_{L^g}=\hat{a}_{11}a_{21},\ \phi_{M^g}=\hat{a}_{22}a_{12}, \end{equation} where for a line bundle $L$ on an abelian scheme $B$ we denote by $\phi_L:B\rightarrow\hat{B}$ the symmetric homomorphism corresponding to the symmetric biextension $\Lambda(L)$. The group law on $\widetilde{\operatorname{SL}}_2(A)$ is defined uniquely from the condition that the there is an action of $\widetilde{\operatorname{SL}}_2(A)$ on the stack of Picard groupoids ${\cal H}(A)$ such that an element $g=(\bar{g},L^g,M^g)$ acts by the functor which is identical on ${\cal P}ic$ and sends the generator $T_{(x,y)}$ (where $(x,y)\in\hat{A}\times A$) to $L_x\otimes M_y\otimes \cal P_{(a_{12}y,a_{21}x)}\otimes T_{\bar{g}(x,y)}$. We refer to \cite{Weilrep} for explicit formulas for the group law in $\widetilde{\operatorname{SL}}_2(A)$. It is easy to see that the natural projection $\widetilde{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$ is a homomorphism with the kernel isomorphic to $A(S)\times\hat{A}(S)$. Consider the subgroup $\widehat{\operatorname{SL}}_2(A)\subset\widetilde{\operatorname{SL}}_2(A)$ consisting of triples with symmetric $L^g$ and $M^g$. Then we have an isomorphism $\widehat{\operatorname{SL}}_2(A)\simeq\widehat{\operatorname{SL}}_2(A_K)$ since any symmetric line bundle on $A_K$ extends to a symmetric line bundle on $A$ (see \cite{MB}, II.3.3). Let $\Gamma(A)=\Gamma(A_K)$ be the image of the projection $\widehat{\operatorname{SL}}_2(A_K)\rightarrow\operatorname{SL}_2(A_K)$. Then $\Gamma(A)$ has finite index in $\operatorname{SL}_2(A_K)$ since it contains the subgroup $\Gamma(A,2)=\Gamma(A_K,2)\subset\operatorname{SL}_2(A_K)$ consisting of matrices with $a_{12}$ and $a_{21}$ divisible by 2. In the case when $S$ is the spectrum of an algebraically closed field it was shown in \cite{Weilrep} that there exist intertwining functors between representations of Heisenberg groupoid corresponding to the natural action of $\widetilde{\operatorname{SL}}_2(A)$ on the Heisenberg groupoid $H(\hat{A}\times A)$ which are analogous to the operators of Weil-Shale representation. We are going to extend this construction to the case of a normal base scheme. Recall (see \cite{Weilrep}, sect. 10) that there is a natural action of $\widetilde{\operatorname{SL}}_2(A)$ on the set of enhanced lagrangian subvarieties in $X=\hat{A}\times A$ such that a triple $g=(\bar{g},L^g,M^g)\in\widetilde{\operatorname{SL}}_2(A)$ maps $(\hat{A},\O_{\hat{A}})$ to $\bar{g}(\hat{A})= (a_{11},a_{21})(\hat{A})$ with the line bundle corresponding to $L^g\in\operatorname{Pic}(\hat{A})$. Furthermore, there is a natural equivalence of categories $$\overline{g}_*:{\cal D}(\hat{A},\O_{\hat{A}})\rightarrow {\cal D}(\overline{g}(\hat{A}),L^g):{\cal F}\mapsto \overline{g}_*{\cal F}$$ such that the standard $H(X)$-action on ${\cal D}(\hat{A},O_{\hat{A}})$ corresponds to the $g$-twisted $H(X)$-action on ${\cal D}(\overline{g}(\hat{A}),L^g)$. On the other hand, if $\hat{A}\cap \overline{g}(\hat{A})=\ker(a_{21})$ is finite (hence, flat) over $S$ and there exists a Schr\"odinger representation $V$ for the corresponding Heisenberg extension $G_g:=G_{\hat{A},\overline{g}(\hat{A})}$ of $\ker(a_{21})$ then the construction of the previous section gives another equivalence $${\cal D}(\overline{g}(\hat{A}),L^g))\rightarrow{\cal D}(\hat{A},\O_{\hat{A}})$$ compatible with the standard $H(X)$-actions. Composing it with the previous equivalence we get an equivalence $\rho\widetilde{\rightarrow}\rho^g$ where $\rho$ is the representation of $H(X)$ on ${\cal D}(\hat{A},\O_{\hat{A}})\simeq{\cal D}(A)$ given by (\ref{act}), $\rho^g=\rho\circ g$ is the same representation twisted by $g$. Using (\ref{kernel}) it is easy to compute that the kernel on $A\times_S A$ corresponding to this equivalence has form \begin{equation}\label{kernelnew} {\cal K}(g,V)=(p_2-a_{22}p_1)^*E\otimes (a_{12}\times\operatorname{id})^*\cal P^{-1}\otimes (-p_1)^*M^g \end{equation} where $$E=\underline{\operatorname{Hom}}_{G_g^{-1}}(V^*,a_{21*}((L^g)^{-1})).$$ Note that here $G_g^{-1}$ is the restriction of the Mumford's extension $G((L^g)^{-1})\rightarrow\ker(\hat{a}_{11}a_{21})$ to $\ker(a_{21})$. Hence, if $\ker(a_{21})$ is finite over $S$ we get a functor from the gerb of Schr\"odinger representations for $G_g$ to the stack of $H(X)$-equivalences $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$, More precisely, the category of intertwining operators between $\rho$ and $\rho^g$ is defined in terms of kernels in ${\cal D}^b(A\times_S A)$ (see \cite{Weilrep}) and the glueing property is satisfied because the kernels corresponding to equivalences are actually vector bundles (perhaps, shifted). Indeed, the latter property is local with respect to the {\it fppf} topology on the base $S$ and locally a Schr\"odinger representation for $G_g$ exists and give rise to the kernel (\ref{kernelnew}) in $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$ which is a vector bundle up to shift. Now any other object of $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$ is obtained from a given one by tensoring with a line bundle on $S$ and a shift. Thus, when $\ker(a_{21})$ is finite the obstacle for the existence of a global equivalence between $\rho$ and $\rho^g$ is given by the class $e(G_g)\in\operatorname{Br}(S)$. Let $U\subset\operatorname{SL}_2(A)$ be the subset of matrices such that $a_{21}$ is an isogeny. It turns out that similarly to the case of real groups one can deal with $U$ instead of the entire group when defining representation of $\operatorname{SL}_2(A)$. This observation can be formalized as follows. Let us call a subset $B$ of a group $G$ {\it big} if for any triple of elements $g_1, g_2, g_3\in G$ the intersection $B^{-1}\cap Bg_1\cap Bg_2\cap Bg_3$ is non-empty. This condition first appeared in \cite{Weil} IV. 42, while the term is due to D.~Kazhdan. The reason for introducing this notion is the following lemma. \begin{lem} Let $B\subset G$ be a big subset. Then $G$ is isomorphic to the abstract group generated by elements $[b]$ for $b\in B$ modulo the relations $[b_1][b_2]=[b]$ when $b,b_1,b_2\in B$ and $b=b_1b_2$. If $c:G\times G\rightarrow C$ is a 2-cocycle (where $C$ is an abelian group with the trivial $G$-action) such that $c(b_1,b_2)=0$ whenever $b_1, b_2, b_1b_2\in B$ then $c$ is a coboundary. \end{lem} \noindent {\it Proof} . For the proof of the first statement we refer to \cite{Weil}, IV, 42, Lem. 6. Let $c:G\times G\rightarrow C$ be a 2-cocycle, $H$ be the corresponding central extension of $G$ by $C$. Consider the group $\widetilde{H}$ generated by the central subgroup $C$ and generators $[b]$ for $b\in B$ subject to relations $[b_1][b_2]=c(b_1,b_2)[b_1b_2]$, where $b_1, b_2, b_1b_2\in B$. Then $\widetilde{H}/C\simeq G$, hence the natural homomorphism $\widetilde{H}\rightarrow H$ is an isomorphism. If $c(b_1,b_2)=0$ whenever $b_1, b_2, b_1b_2\in B$, then the extension $\widetilde{H}\simeq H\rightarrow G$ splits, hence $c$ is a coboundary. \qed\vspace{3mm} At this point we need to recall some results from \cite{Weilrep}, sect. 9 concerning the group $\operatorname{SL}_2(A)$. Since $\operatorname{SL}_2(A)=\operatorname{SL}_2(A_K)$ we can work with abelian varieties over a field. First note that this group can be considered as a group of ${\Bbb Z}$-points of an group scheme over ${\Bbb Z}$. It turns out that the corresponding algebraic group $\operatorname{SL}_{2,A,{\Bbb Q}}$ over ${\Bbb Q}$ is very close to be semi-simple. Namely, if we fix a polarization on $A$ then the latter group is completely determined by the algebra $\operatorname{End}(A)\otimes{\Bbb Q}$ and the Rosati involution on it. Decomposing (up to isogeny) $A$ into a product $A_1^{n_1}\times\ldots A_l^{n_l}$ where $A_i$ are different simple abelian varieties and choosing a polarization compatible with this decomposition it is easy to see that $$\operatorname{SL}_{2,A,{\Bbb Q}}\simeq\prod_i R_{K_{i,0}/{\Bbb Q}}\operatorname{U}^*_{2n_i,F_i}$$ where $F_i=\operatorname{End}(A_i)\otimes{\Bbb Q}$, $K_i$ is the center of $F_i$, $K_{i,0}\subset K_i$ is the subfield of elements stable under the Rosati involution, $\operatorname{U}^*_{2n_i,F_i}$ is the group of $F_i$-automorphisms of $F_i^{2n_i}$ preserving the standard skew-hermitian form, $R_{K_{i,0}/{\Bbb Q}}$ denotes the restriction of scalars from $K_{i,0}$ to ${\Bbb Q}$. Thus, the only case when the group $\operatorname{U}^*_{2n_i,F_i}$ is not semi-simple is when the Rosati involution on $F_i$ is of the second kind, i.~e. $K_{i,0}\neq K_i$. In the latter case, $\operatorname{U}^*_{2n_i,F_i}$ is a product of the semi-simple subgroup $\operatorname{SU}^*_{2n_i,F_i}$ (defined using the determinant with values in $K_i$) and the central subgroup $K^1_i=\{ x\in K_i\ |\ N_{K_i/K_{i,0}}=1\}$ consisting of diagonal matrices. Furthermore, the intersections of these two subgroup is finite. It follows that the group $\operatorname{SL}_{2,A,{\Bbb Q}}$ always has an almost direct decomposition into a product of the semi-simple subgroup $H=\prod_i R_{K_{i,0}/{\Bbb Q}}\operatorname{SU}^*_{2n_i,F_i}$ and a central subgroup $Z$ consisting of diagonal matrices. Now we can prove the following result. \begin{lem}\label{big} Let $\Gamma\subset\operatorname{SL}_2(A)$ be a subgroup of finite index. Then the subset $\Gamma\cap U$ is big. \end{lem} \noindent {\it Proof} . By definition $U$ is an intersection of a Zariski open subset $\underline{U}$ in the irreducible algebraic group $\operatorname{SL}_{2,A_K,{\Bbb Q}}$ with $\Gamma$. If $\Gamma$ were Zariski dense in $\operatorname{SL}_{2,A_K,{\Bbb Q}}$ the proof would be finished. This is not always true, however, we claim that $\Gamma\cap H$ is dense in $H$ where $H$ is the subgroup of $\operatorname{SL}_{2,A_K,{\Bbb Q}}$ introduced above. Indeed, this follows from the fact that $H$ is semi-simple and the corresponding real groups have no compact factors (see \cite{Weilrep} 9.4). On the other hand, the set $\underline{U}$ is $Z$-invariant (since $Z$ consists of diagonal matrices). It follows that for any $g\in\Gamma$ the intersection $\underline{U}g\cap H$ is a non-empty Zariski open subset of $H$ (as a preimage of a non-empty Zariski set under the isogeny $H\rightarrow\operatorname{SL}_{2,A_K,{\Bbb Q}}/Z$). Therefore, for any triple of elements $g_1,g_2,g_3\in\Gamma$ the intersection $\underline{U}\cap\underline{U}g_1\cap\underline{U}g_2\cap\underline{U}g_3\cap H$ is a non-empty open subset in $H$, hence it contains an element of $\Gamma\cap H$. As $U=U^{-1}$ this shows that $U$ is big. \qed\vspace{3mm} \begin{prop}\label{obshom} There exists a homomorphism $\delta_A:\Gamma(A)\rightarrow\operatorname{Br}(S)$ such that for any $g\in\widehat{\operatorname{SL}}_2(A)$ lying over $\overline{g}\in\Gamma(A)$ there exists a global object in $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$ if and only if $\delta_A(\overline{g})=0$. If the $a_{21}$-entry of $\overline{g}$ is an isogeny then $\delta_A(\overline{g})=e(G_g)$. \end{prop} \noindent {\it Proof} . For any $g\in\widehat{\operatorname{SL}}_2(A)$ let us denote by $\operatorname{Isom}_{H(X)}^0(\rho,\rho^g)\subset\operatorname{Isom}_{H(X)}(\rho,\rho^g)$ the full subcategory of kernels that belong to the core of the standard $t$-structure on ${\cal D}^b(A\times_S A)$. We claim that $\operatorname{Isom}_{H(X)}^0(\rho,\rho^g)$ is a gerb. Indeed, we know this when $\overline{g}\in\Gamma(A)\cap U$. Now by Lemma \ref{big} any element in $\widehat{\operatorname{SL}}_2(A)$ can be written as a product $gg'$ where $g$ and $g'$ lie over $\Gamma(A)\cap U$. Locally over $S$ there exist Schr\"odinger representations for $G_g$ and $G_{g'}$, hence (\ref{kernelnew}) defines the corresponding kernels ${\cal K}(g)\in\operatorname{Isom}_{H(X)}(\rho,\rho^g)$ and ${\cal K}(g')\in\operatorname{Isom}_{H(X)}(\rho,\rho^{g'})$. Now their composition ${\cal K}(gg')=p_{13*}(p_{12}^*{\cal K}(g')\otimes p_{23}^*{\cal K}(g))$ is an element of $\operatorname{Isom}_{H(X)}(\rho,\rho^{gg'})$ and Prop. 1.7 of \cite{Muk2} implies that ${\cal K}(gg')$ has only one non-zero sheaf cohomology which is flat over $S$ (since this is so in the case of an algebraically closed field considered in \cite{Weilrep}). It follows that any object of $\operatorname{Isom}_{H(X)}(\rho,\rho^{gg'})$ over any open subset of $U$ is a pure $S$-flat sheaf (perhaps shifted), hence, the gluing axiom is satisfied. Now we can define $\delta_A(\overline{g})\in H^2(S,{\Bbb G}_m)$ for an element $\overline{g}\in\Gamma(A)$ as the class of the gerb $\operatorname{Isom}_{H(X)}^0(\rho,\rho^g)$ where $g\in\widehat{\operatorname{SL}}_2(A)$ is any element lying over $\overline{g}$. When $\overline{g}\in U$ this class is equal to $e(G_g)\in\operatorname{Br}(S)$. Clearly, $\delta_A$ is a homomorphism $\Gamma(A)\rightarrow H^2(S,{\Bbb G}_m)$. Since $\Gamma(A)$ is generated by $\Gamma(A)\cap U$ we have $\delta_A(\overline{g})\in\operatorname{Br}(S)$ for any $\overline{g}\in\Gamma(A)$. \qed\vspace{3mm} \begin{lem}\label{neutrcomp} Let $g=(\overline{g},L^g,M^g)$ be an element of $\widehat{\operatorname{SL}}_2(A)$ such that $\ker({a_{21}})$ is flat over $S$, its neutral component $\ker(a_{21})^0$ is an abelian subscheme of $\hat{A}$, and $L^g|_{\ker(a_{21})^0}$ is trivial. Then $\delta_A(\overline{g})=0$ if and only if there exists a Schr\"odinger representation for the Heisenberg extension $G_g$ of $\pi_0(\ker(a_{21}))=\ker(a_{21})/(\ker(a_{21}))^0$ induced by $G(L^g)|_{\ker(a_{21})}$. \end{lem} \noindent {\it Proof} . Let $V$ be a Schr\"odinger representation for $G_g$. Then we can define ${\cal K}(g,V)$ by the formula (\ref{kernelnew}) where $E$ is defined as follows. First we descend $L^g$ to a line bundle $\overline{L}$ on $\hat{A}/\ker(a_{21})^0$, then we set $$E=\underline{\operatorname{Hom}}_{G_g^{-1}}(V^*,\overline{a}_{21*}(\overline{L}^{-1}))$$ where $\overline{a}_{21}:\hat{A}/\ker(a_{21})^0\rightarrow A$ is the finite map induced by $a_{21}$. Note that when $S$ is the spectrum of an algebraically closed field this kernel ${\cal K}(g,V)$ coincides with the one defined in \cite{Weilrep}, 12.3. By definition ${\cal K}(g,V)$ is the direct image of a bundle on an abelian subscheme $$\operatorname{supp} {\cal K}(g,V)=\{(x_1,x_2)\in A^2\ | x_2-a_{22}x_1\in a_{21}(\hat{A})\}$$ (note that $a_{21}(\hat{A})\subset A$ is an abelian subscheme since $\ker(a_{21})$ is flat). Applying this to $g^{-1}$ we get $$\operatorname{supp} {\cal K}(g^{-1},V^*)=\{(x_1,x_2)\in A^2\ |\ x_2-\hat{a}_{11}x_1\in \hat{a}_{21}(\hat{A})\}.$$ Hence, the sheaf $p_{12}^*{\cal K}(g^{-1},V^*)\otimes p_{23}^*{\cal K}(g,V)$ on $A^3$ is supported on the abelian subscheme $$X(g)=\{(x_1,x_2,x_3)\in A^3\ |\ x_3-a_{22}x_2\in a_{21}(\hat{A}), x_2-\hat{a}_{11}x_1\in\hat{a}_{21}(\hat{A})\}.$$ Note that for $(x_1,x_2,x_3)\in X(g)$ we have $$x_3-a_{22}\hat{a}_{11}x_1\in a_{21}(\hat{A})+ a_{22}\hat{a}_{21}(\hat{A})=a_{21}(\hat{A})$$ since $a_{22}\hat{a}_{21}=a_{21}\hat{a}_{22}$. Now $a_{22}\hat{a}_{11}x_1\equiv x_1\mod(a_{21}(\hat{A}))$, hence $x_3\equiv x_1\mod(a_{21}(\hat{A}))$. Thus, we have an isomorphism \begin{align*} &X(g)\rightarrow\{(x,x')\ |\ x-x'\in a_{21}(\hat{A})\} \times\hat{a}_{21}(\hat{A}):\\ &(x_1,x_2,x_3)\mapsto ((x_1,x_3), x_2-\hat{a}_{11}x_1). \end{align*} It follows that the restriction of the projection $p_{13}$ to $X(g)$ is flat and surjective. Therefore, applying Prop. 1.7 of \cite{Muk2} we conclude as in the proof of Theorem \ref{eq} that $\overline{R}(V)$ is an equivalence. Thus, the map $V\mapsto {\cal K}(g,V)$ gives a functor from $\operatorname{Schr}_{G_g}$ to $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$. \qed\vspace{3mm} \begin{prop}\label{compat} Let $A$ and $B$ be abelian schemes over the same base $S$, then there is a natural embedding $i_A:\Gamma(A)\rightarrow\Gamma(A\times B)$ such that $$\delta_A=\delta_{A\times B}\circ i_A.$$ \end{prop} \noindent {\it Proof} . It is sufficient to check this identity on elements of $\Gamma(A)\cap U$, in which case this follows immediately from Lemma \ref{neutrcomp}. \qed\vspace{3mm} Consider the subgroup of finite index $\Gamma_0(A)\subset\Gamma(A)$ consisting of matrices $\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)$ in $\Gamma(A)$ for which $a_{21}$ is divisible by 2. Note that $\Gamma_0(A)$ contains $\Gamma(A,2)$. \begin{lem}\label{oddisog} Let $A$ be an abelian variety over a field such that there exists a symmetric line bundle $L$ on $A$ which induces an isogeny $f:A\rightarrow\hat{A}$ of odd degree. Then the subgroup of $\Gamma(A)$ generated by its elements $\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)$, for which $a_{21}$ is an isogeny of odd degree, contains $\Gamma_0(A)$. \end{lem} \noindent {\it Proof}. Recall that $U\subset\operatorname{SL}_2(A)$ is the subset defined by the condition that $a_{21}$ is an isogeny. Consider the matrix $\gamma_f=\left( \matrix \operatorname{id} & 0 \\ f & \operatorname{id} \endmatrix \right)\in\Gamma(A)$. Let $U_1=U\cap U\gamma_f^{-1}$. Then the argument similar to that of Lemma \ref{big} shows that $\Gamma_0(A)\cap U_1$ is a big subset in $\Gamma_0(A)$, in particular, $\Gamma_0(A)$ is generated by $\Gamma_0(A)\cap U_1$. Now let $\gamma\in\Gamma_0(A)\cap U_1$, then $\gamma \gamma_f\in U$ and its $a_{21}$-entry is an isogeny of odd degree which implies the statement. \qed\vspace{3mm} \begin{prop}\label{triv1} The restriction of the homomorphism $\delta_A:\Gamma(A)\rightarrow\operatorname{Br}(S)$ to $\Gamma_0(A)$ is trivial. \end{prop} \noindent {\it Proof} . Recall that if $a_{21}$ is an isogeny then $\delta_A(g)$ is defined as an obstacle for the existence of a Schr\"odinger representation of the (symmetric) Heisenberg extension $G_g$ of $\ker(a_{21})$ attached to $g$, where $g$ projects to a matrix $\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)\in\operatorname{SL}_2(A)$. Hence, by Theorem \ref{odd} $\delta_A(g)=0$ if $a_{21}$ is an isogeny of odd degree. In particular, if $A$ is principally polarized then Lemma \ref{oddisog} implies that the restriction of $\delta_A$ to $\Gamma_0(A)$ is trivial. By Zarhin's trick (see \cite{Zarh}) for any abelian scheme $A$ over $S$ there exists an abelian scheme $B$ over $S$ such that $A\times B$ admits a principal polarization. Now by Proposition \ref{compat} we have $\delta_A=\delta_{A\times B}\circ i_A$. Therefore, the restriction of $\delta_A$ to $\Gamma_0(A)$ is trivial. \qed\vspace{3mm} \begin{rem} It is easy to see that the kernel of $\delta_A$ is in general bigger than $\Gamma_0(A)$. Namely, it contains also matrices for which $a_{21}$ (or $a_{12}$) is an isogeny of odd degree, those for which $a_{12}$ is divisible by 2, and those for which $a_{11},a_{22}\in{\Bbb Z}$. Sometimes, these elements together with $\Gamma_0(A)$ generate the entire group $\Gamma(A)$, however, it is not clear whether $\delta_A$ is always trivial. In the section \ref{real} we will prove it in some special cases. \end{rem} Let $\widehat{\Gamma}_0(A)$ be the preimage of $\Gamma_0(A)$ in $\widehat{\operatorname{SL}}_2(A)$. In other words, this is the subgroup of elements $g\in\widehat{\operatorname{SL}}_2(A)$ such that for the corresponding matrix in $\operatorname{SL}_2(A)$ the $a_{21}$-entry is divisible by 2. We say that there is a faithful action of a group $G$ on a category ${\cal C}$ if there is an embedding of $G$ into a group of autoequivalences of ${\cal C}$ (considered up to isomorphism). \begin{thm} For any abelian scheme $A$ over a normal connected noetherian base $S$ there is a faithful action of a central extension of the group $\widehat{\Gamma}_0(A)$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ on ${\cal D}^b(A)$. \end{thm} \noindent {\it Proof} . According to Proposition \ref{triv1} for every $g\in\widehat{\Gamma}_0(A)$ there exists a global object in $\operatorname{Isom}_{H(X)}(\rho,\rho^g)$. It is defined uniquely up to a shift and tensoring with a line bundle on $S$. Hence, the required action of a central extension. The fact that this action is faithful is clear in the case when the base is a field: for example, one can use explicit formulas for these functors from Lemma \ref{neutrcomp}. Since the action of $\operatorname{Pic}(S)$ on ${\cal D}^b(A)$ is obviously faithful the general case follows. \qed\vspace{3mm} \begin{cor} With the assumptions of the above theorem, there is a faithful action of a central extension of $\Gamma(A,2)$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ on ${\cal D}^b(A)$. \end{cor} \noindent {\it Proof} . This action is obtained from the canonical homomorphism $\Gamma(A,2)\rightarrow\widehat{\operatorname{SL}}_2(A)$ splitting the projection $\widehat{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$. Namely, under this splitting the matrix $\overline{g}=\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)\in\Gamma(A,2)$ maps to the element $(\overline{g},(\hat{a}_{11}a_{21}/2,\operatorname{id})^*\cal P, (\operatorname{id},\hat{a}_{22}a_{12}/2)^*\cal P)$ of $\widehat{\operatorname{SL}}_2(A)$, where $\cal P$ is the Poincar\'e line bundle on $A\times\hat{A}$. \qed\vspace{3mm} \section{The induced action on a Chow motive} In this section we will construct a projection action of the algebraic group $\operatorname{SL}_{2,A,{\Bbb Q}}$ on the relative Chow motive of an abelian scheme $\pi:A\rightarrow S$ with rational coefficients. Let us denote by $\operatorname{Cor}(A)$ the Chow group $\operatorname{CH}^*(A\times_S A)\otimes{\Bbb Q}$ considered as a ${\Bbb Q}$-algebra with multiplication given by the composition of correspondences: \begin{equation}\label{compcorr} \beta\circ\a=p_{13*}(p_{12}^*(\a)\cdot p_{23}^*(\beta)) \end{equation} where $\a,\beta\in\operatorname{CH}^*(A\times_S A)\otimes {\Bbb Q}$, $p_{ij}$ are the projections from $A^3$ to $A^2$. The unit of this algebra is $[\Delta]\in\operatorname{CH}^g(A\times_S A)$ where $\Delta\subset A\times_S A$ is the relative diagonal, $g=\dim A$. Using the Riemann-Roch theorem it is easy to see that the multiplication (\ref{compcorr}) is compatible with the composition law on $K^0(A\times_S A)$ arising from the interpretation of ${\cal D}^b(A\times_S A)$ as the category of functors from ${\cal D}^b(A)$ to itself considered above, via the map \begin{equation}\label{ch} K^0(A\times_S A)\otimes {\Bbb Q}\widetilde{\rightarrow}\operatorname{CH}^*(A\times_S A)\otimes{\Bbb Q}: x\mapsto \operatorname{ch}(x)\cdot\pi^*\operatorname{Td}(e^*T_{A/S}) \end{equation} where $\operatorname{ch}$ is the Chern character. Let us consider the embedding of algebras $$\operatorname{CH}(S)_{{\Bbb Q}}\rightarrow\operatorname{Cor}(A):x\mapsto \pi^*(x)\cdot [\Delta]$$ where $\operatorname{CH}(S)_{{\Bbb Q}}=\operatorname{CH}(S)\otimes{\Bbb Q}$ is equipped with the usual multiplication. In particular, we have an embedding of groups of invertible elements $\operatorname{CH}(S)_{{\Bbb Q}}^*\subset\operatorname{Cor}(A)^*$. Applying the map (\ref{ch}) to the kernels giving projective action of $\Gamma(A,2)$ on ${\cal D}^b(A)$ we obtain a homomorphism $$\widetilde{\phi}:\Gamma(A,2)\rightarrow(\operatorname{Cor}(A))^*/\pm\operatorname{Pic}(S),$$ where $\operatorname{Pic}(S)$ is embedded into $\operatorname{CH}(S)_{{\Bbb Q}}^*$ by Chern character (multiplication by $\pm$ arises from shifts in derived category). Our aim is to approximate this homomorphism by a morphism of algebraic groups over ${\Bbb Q}$. More precisely, we have to replace $\widetilde{\phi}$ by the induced homomorphism $$\phi:\Gamma(A,2)\rightarrow(\operatorname{Cor}(A))^*/\operatorname{CH}(S)_{{\Bbb Q}}^*.$$ Now we claim that one can replace here source and target by some algebraic groups over ${\Bbb Q}$ such that $\phi$ will be induced by an algebraic homomorphism. Naturally, the source should be replaced by $\operatorname{SL}_{2,A,{\Bbb Q}}$ (see the previous section). To approximate the target we have to replace algebras $\operatorname{Cor}(A)$ and $\operatorname{CH}(S)_{{\Bbb Q}}$ by their finite-dimensional subalgebras. \begin{thm}\label{actmotmain} There exists a finite-dimensional ${\Bbb Q}$-subalgebra $D\subset\operatorname{Cor}(A)$ and a morphism of algebraic ${\Bbb Q}$-groups $\rho:\operatorname{SL}_{2,A,{\Bbb Q}}\rightarrow D^*/(D\cap\operatorname{CH}(S)_{{\Bbb Q}})^*$ inducing $\phi$ on $\Gamma(A,2)$. \end{thm} \noindent {\it Proof} . For a pair of abelian schemes $A$ and $B$ over $S$ let us consider the map $$\gamma_{A,B}:\operatorname{Hom}(A,B)\rightarrow\operatorname{CH}(A\times_S B)$$ that sends an $S$-morphism $f:A\rightarrow B$ to the class $[\Gamma_f]$ of the (relative) graph of $f$. One can extend naturally $\gamma$ to a map $$\gamma:\operatorname{Hom}(A)\otimes{\Bbb Q}\rightarrow\operatorname{CH}(A\times_S B)\otimes{\Bbb Q}$$ by sending $f/n$ to $\gamma([n]_A)^{-1}\circ\gamma(f)$, where $f\in\operatorname{End}(A)$, $n\neq0$ --- here we take the inverse to $\gamma([n]_A)$ in the algebra $\operatorname{Cor}(A)$ and use its natural action on $\operatorname{CH}(A\times_S B)\otimes{\Bbb Q}$. It is easy to check that $\gamma$ is a polynomial map (see \cite{Weilrep}, Lemma 13.3, for the case $A=B$). Now let $\overline{g}=\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)$ be an element of $\Gamma(A,2)\cap U$ (recall that $U$ is defined by the condition that $a_{21}$ is an isogeny). From the formula (\ref{kernelnew}) we get the following expression for $\phi(\overline{g})$: \begin{equation}\label{chker} \phi(\overline{g})=(p_2-a_{22}p_1)^*(a_{21*}\operatorname{ch}(L^g)^{-1})\cdot (a_{12}\times\operatorname{id})^*(\operatorname{ch}(\cal P))\cdot p_1^*(\operatorname{ch}(M^g))\mod\operatorname{CH}(S)_{{\Bbb Q}}^*. \end{equation} Note that $\operatorname{ch}(L^g)$ (resp. $\operatorname{ch}(M^g)$) is a polynomial function of $\hat{a}_{11}a_{21}$ (resp. $\hat{a}_{22}a_{12}$). Also the functors $f^*$ and $f_*$ can be expressed as compositions with the correspondence given by the graph of $f$. It follows from the above remarks that the right hand side of (\ref{chker}) is obtained by evaluating at $\overline{g}$ of a polynomial map $\psi:\operatorname{End}(\hat{A}\times A)\otimes{\Bbb Q}\rightarrow\operatorname{Cor}(A)$. In particular, the image of $\psi$ belongs to a finite-dimensional ${\Bbb Q}$-subspace of $\operatorname{Cor}(A)$. Let $\underline{U}\subset\operatorname{SL}_{2,A,{\Bbb Q}}$ be the Zariski open subset defined by $\deg(a_{21})\neq0$. Note that $\underline{U}$ is stable under the inversion morphism $g\mapsto g^{-1}$. Let us also denote $\underline{U}^{(2)}= \mu^{-1}(\underline{U})\cap(\underline{U}\times\underline{U})\subset\underline{U}\times \underline{U}$ where $\mu$ is the group law. Consider two polynomial maps \begin{align*} &a_1:\underline{U}\rightarrow\operatorname{Cor}(A):u\mapsto a_1(u)=\psi(u^{-1})\circ\psi(u),\\ &a_2:\underline{U}^{(2)}\rightarrow\operatorname{Cor}(A): (u_1,u_2)\mapsto a_2(u_1,u_2)=\psi(u_1u_2)\circ\psi(u_2^{-1})\circ \psi(u_1^{-1}). \end{align*} It is easy to see that the images of both maps belong to the subalgebra $\operatorname{CH}(S)_{{\Bbb Q}}\subset\operatorname{Cor}(A)$. This can be done either by direct computation using (\ref{chker}) or using the density of $\Gamma$ in $H\subset\operatorname{SL}_{2,A,{\Bbb Q}}$ (see the previous section). Also an easy direct computation shows that $a_1(u)$ is invertible in the algebra $\operatorname{CH}(S)_{{\Bbb Q}}$ for all $u\in\underline{U}$. This immediately implies that $a_2(u_1,u_2)$ is invertible for any $(u_1,u_2)\in\underline{U}^{(2)}$: indeed, $a_2(u_1,u_2)$ is a divisor of $a_1(u_1)a_1(u_2)a_1(u_2^{-1}u_1^{-1})$. Note that the components of images of $a_1$ and $a_2$ span finite-dimensional subspaces in $\operatorname{CH}(S)^i_{{\Bbb Q}}$ for any $i$. It follows that there exists a finite dimensional subalgebra $D_S\subset\operatorname{CH}(S)_{{\Bbb Q}}$ such that the images of $a_1$ and $a_2$ belong to $D_S^*$. Now we have \begin{equation}\label{projmot} \psi(u_1)\circ\psi(u_2)=a_2(u_1,u_2)^{-1}a_1(u_1)a_1(u_2) \psi(u_1u_2) \end{equation} for $(u_1,u_2)\in\underline{U}^{(2)}$. Let $D\subset\operatorname{Cor}(A)$ be the $D_S$-submodule generated by $\psi(u)$ with $u\in\underline{U}$. Then $D$ is finite-dimensional as a ${\Bbb Q}$-vector space and (\ref{projmot}) shows that $\psi(u_1)\circ\psi(u_2)\in D$ for any $(u_1,u_2)\in\underline{U}^{(2)}$. Since $\underline{U}^{(2)}$ is dense in $\underline{U}\times\underline{U}$ it follows that $D$ is a subalgebra. Now (\ref{projmot}) implies that $\psi$ uniquely extends to a homomorphism $\operatorname{SL}_{2,A,{\Bbb Q}}\rightarrow D^*/D_S^*$. \qed\vspace{3mm} \section{Splittings of the extension $\widetilde{SL}_2(A)\rightarrow SL_2(A)$} Let $A/S$ be an abelian scheme with a principal polarization $\phi:A\rightarrow\hat{A}$. Then we have the Rosati involution $$\varepsilon_{\phi}:\operatorname{End}(A)\rightarrow\operatorname{End}(A): f\mapsto \phi^{-1}\circ \hat{f}\circ \phi.$$ The group $\operatorname{SL}_2(A)$ is completely determined by the algebra $\operatorname{End}(A)$ with involution $\varepsilon_{\phi}$. The definition of the group $\widetilde{\operatorname{SL}}_2(A)$ requires in addition the knowledge of the extension \begin{equation}\label{extA} 0\rightarrow \hat{A}(S)\rightarrow\operatorname{Pic}(A)\rightarrow\operatorname{Hom}^{\operatorname{sym}}(A,\hat{A}) \end{equation} together with the action of the multiplicative monoid of $\operatorname{End}(A)$ on it. Thus, splittings of the homomorphism $\widetilde{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$ should be related to splittings of (\ref{extA}). More precisely, it's natural to consider splittings compatible with the $\operatorname{End}(A)$-action. We'll show that such splittings of (\ref{extA}) correspond to simultaneous splittings of homomorphisms $\widetilde{\operatorname{SL}}_2(A^n)\rightarrow\operatorname{SL}_2(A^n)$ for all $n$, where $A^n/S$ is the $n$-th relative cartesian power of $A/S$. More generally, we start with arbitrary subring $R\subset\operatorname{End}(A)$ stable under the Rosati involution. Let us denote by $\varepsilon:R\rightarrow R$ the restriction of the Rosati involution to $R$, let $R^+\subset R$ be the subring of elements stable under $\varepsilon$. Then for any $n\ge 1$ we can consider the subgroup $\operatorname{SL}_2(A^n,R)\subset\operatorname{SL}_2(A^n)$ consisting of $2n\times 2n$ matrices with all entries belonging to $R$ (we identify $\hat{A}$ with $A$ via $\phi$). Let $\widetilde{\operatorname{SL}}_2(A^n,R)\subset\widetilde{\operatorname{SL}}_2(A^n)$ be the preimage of $\operatorname{SL}_2(A^n,R)$. We are interested in splittings of the natural homomorphisms \begin{equation}\label{homSL} \widetilde{\operatorname{SL}}_2(A^n,R)\rightarrow\operatorname{SL}_2(A^n,R) \end{equation} It turns out that the following structure on $A$ is relevant for this. \begin{defi} A $\Sigma_{R,\varepsilon}$-structure for $\phi$ is a homomorphism $R^+\rightarrow\operatorname{Pic}(A):r_0\mapsto L(r_0)$ such that \begin{equation}\label{CM1} \phi_{L(r_0)}=\phi\circ [r_0]_A \end{equation} for any $r_0\in R^+$ and \begin{equation}\label{CM2} [r]^*L(r_0)\simeq L(\varepsilon(r)r_0r) \end{equation} for any $r\in R$, $r_0\in R^+$. \end{defi} Note that (\ref{CM2}) for $r=-1$ implies that all line bundles $L(r_0)$ are symmetric. In \cite{thetaid} we studied the question of existence of $\Sigma_{R,\varepsilon}$-structure for an abelian variety. For example, in the case of a complex elliptic curve $E$ with complex multiplication such a structure for $R=\operatorname{End}(E)$ exists if and only if $R$ is unramified at $2$. Another example is the case when $R$ is a ring of integers in a totally real number field unramified at 2 and $\varepsilon=\operatorname{id}$ (see next section). \begin{thm}\label{Sig} A $\Sigma_{R,\varepsilon}$-structure on $A$ induces canonical splittings of the homomorphisms (\ref{homSL}) for all $n$. \end{thm} \noindent {\it Proof} . It is easy to see that a $\Sigma_{R,\varepsilon}$-structure on $A$ induces a similar structure on $A^n$ with $R$ replaced by the matrix algebra $\operatorname{Mat}_n(R)$ and $\varepsilon$ replaced by the corresponding involution $(a_{ij})\mapsto (\varepsilon(a_{ji}))$ of $\operatorname{Mat}_n(R)$. Hence, it is sufficient to consider the case $n=1$. In this case we define the splitting $$\operatorname{SL}_{2}(A,R)\rightarrow\widehat{\operatorname{SL}}_2(A,R): g=\left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)\mapsto (g, (\phi^{-1})^*L(\varepsilon(a_{11})a_{21}), L(\varepsilon(a_{22})a_{12}))$$ \qed\vspace{3mm} Now we are going to prove that conversely the existence of splitting of (\ref{homSL}) for $n=2$ implies the existence of $\Sigma_{R,\varepsilon}$-structure. We use the following observation. For any abelian scheme $A$ there is a natural embedding of the semi-direct product $\operatorname{Aut}(A)\ltimes\operatorname{Hom}^{\operatorname{sym}}(A,\hat{A})$ into $\operatorname{SL}_2(A)$ as the subgroup of matrices with $a_{21}=0$. Thus, a splitting of the homomorphism $\widetilde{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$ restricts to a splitting of the homomorphism of $\operatorname{Aut}(A)$-modules $\operatorname{Pic}(A)\rightarrow\operatorname{Hom}^{\operatorname{sym}}(A,\hat{A})$. In the situation considered above we have similar subgroups in $\operatorname{SL}_n(A^n,R)$. Note that the subgroup of $\operatorname{Hom}^{\operatorname{sym}}(A^n,\hat{A}^n)$ consisting of matrices with entries in $R$ can be identified with the group of hermitian matrices $\operatorname{Mat}^{\operatorname{herm}}_n(R)$ and the natural right action of $\operatorname{GL}_n(R)$ on it induced by its action on $A^n$ is given by the formula $$B\mapsto \overline{C}BC$$ where $C=(c_{ij})\in\operatorname{GL}_n(R)$, $B\in\operatorname{Mat}^{\operatorname{herm}}_n(R)$, $\overline{C}=(\varepsilon(c_{ji}))$. Thus, a splitting of the homomorphism (\ref{homSL}) induces a splitting of the homomorphism of $\operatorname{GL}_n(R)$-modules $$\operatorname{Pic}(A^n,R)\rightarrow\operatorname{Mat}^{\operatorname{herm}}_n(R)$$ where $\operatorname{Pic}(A^n,R)\subset\operatorname{Pic}(A^n)$ is the subgroup of line bundles $L$ such that $\phi_L\in\operatorname{Hom}^{\operatorname{sym}}(A^n,\hat{A}^n)$ has entries in $R$. Now we claim that such a splitting for $n=2$ leads to a $\Sigma_{R,\varepsilon}$-structure on $A$. \begin{thm} Any splitting of the homomorphism of $\operatorname{GL}_2(R)$-modules $\operatorname{Pic}(A^2,R)\rightarrow\operatorname{Mat}^{\operatorname{herm}}_2(R)$ is induced by a unique $\Sigma_{R,\varepsilon}$-structure. \end{thm} \noindent {\it Proof} . Let $$s:\operatorname{Mat}^{\operatorname{herm}}_2(R)\rightarrow\operatorname{Pic}(A^2,R)$$ be such a splitting. Then for $r_0\in R^+$ one has \begin{equation}\label{s1} s\left( \matrix r_0 & 0 \\ 0 & 0 \endmatrix \right)=p_1^*L(r_0)\otimes p_2^*\eta(r_0) \end{equation} for some line bundle $L(r_0)$ and $\eta(r_0)$ on $A$ such that $\phi_{L(r_0)}=\phi\circ [r_0]$, $\eta(r_0)\in\operatorname{Pic}^0(A)$. The compatibility of $s$ with the action of $\operatorname{GL}_2(R)$ means that $$[C]^*s(B)=s(\overline{C}BC)$$ where $B\in\operatorname{Mat}^{\operatorname{herm}}_2(R)$, $C=(c_{ij})\in\operatorname{GL}_2(R)$, $\overline{C}=(\varepsilon(c_{ji}))$. Applying this to $C=\left( \matrix 0 & 1 \\ 1 & 0 \endmatrix \right)$ we deduce from (\ref{s1}) the equality \begin{equation}\label{s2} s\left( \matrix 0 & 0 \\ 0 & r_0 \endmatrix \right)=p_2^*L(r_0)\otimes p_1^*\eta(r_0). \end{equation} Also using the identity $$\left( \matrix 1 & 0 \\ \varepsilon(r) & 1 \endmatrix \right)\cdot \left( \matrix 1 & 0 \\ 0 & 0 \endmatrix \right)\cdot \left( \matrix 1 & r \\ 0 & 1 \endmatrix \right)= \left( \matrix 1 & r \\ \varepsilon(r) & \varepsilon(r)r \endmatrix \right)$$ for any $r\in R$ we deduce that $$s\left( \matrix 1 & r \\ \varepsilon(r) & \varepsilon(r)r \endmatrix \right)= \left( \matrix \operatorname{id} & [r] \\ 0 & \operatorname{id} \endmatrix \right)^*(p_1^*L(1)\otimes p_2^*\eta(1))= (p_1+[r]p_2)^*L(1)\otimes p_2^*\eta(1).$$ Combining this with (\ref{s1}) and (\ref{s2}) one can easily compute that \begin{equation}\label{s3} s\left( \matrix 0 & r \\ \varepsilon(r) & 0 \endmatrix \right)=(\phi\times [r])^*\cal P\otimes p_1^*\eta(-\varepsilon(r)r)\otimes p_2^*([r]^*L(1)\otimes L(-\varepsilon(r)r)). \end{equation} Note that the formulas (\ref{s1}), (\ref{s2}), and (\ref{s3}) completely determine $s$. Now the identity $$\left( \matrix 1 & \varepsilon(r) \\ 0 & 1 \endmatrix \right)\cdot \left( \matrix 0 & 0 \\ 0 & r_0 \endmatrix \right)\cdot \left( \matrix 1 & 0 \\ r & 1 \endmatrix \right)= \left( \matrix \varepsilon(r)r_0r & \varepsilon(r)r_0 \\ r_0r & r_0 \endmatrix \right)$$ which holds for any $r\in R$, $r_0\in R^+$ implies the equality \begin{equation}\label{s4} \left( \matrix {\operatorname{id}} & 0 \\ {[r]} & {\operatorname{id}} \endmatrix \right)^*(p_2^*L(r_0)\otimes p_1^*\eta(r_0))= s\left( \matrix {\varepsilon(r)r_0r} & {\varepsilon(r)r_0} \\ {r_0r} & {r_0} \endmatrix \right). \end{equation} Computing the right hand side using (\ref{s1})--(\ref{s3}) and restricting to $A\times 0$ we obtain the identity \begin{equation}\label{s5} [r]^*L(r_0)=L(\varepsilon(r)r_0r)\otimes\eta(-r_0r\varepsilon(r)r_0). \end{equation} On the other hand, setting $r=1$ and restricting (\ref{s4}) to $0\times A$ we get \begin{equation}\label{s6} L(r_0^2)=[r_0]^*L(1)\otimes\eta(r_0). \end{equation} Setting $r=1$ in (\ref{s5}) we obtain the triviality of $\eta(r_0^2)$. Then taking $r_0=1$ and $r\in R^+$ in (\ref{s5}) we obtain that $$[r]^*L(1)=L(r^2)$$ for $r\in R^+$. Comparing this with (\ref{s6}) we deduce the triviality of $\eta(r_0)$ for all $r_0\in R^+$. Now (\ref{s5}) implies that $L(\cdot)$ gives a $\Sigma_{R,\varepsilon}$-structure. \qed\vspace{3mm} \begin{cor} A $\Sigma_{R,\varepsilon}$-structure for $\phi$ exists if and only if a splitting of the homomorphism (\ref{homSL}) for $n=2$ exists. \end{cor} \begin{ex} Let $E={\Bbb C}/{\Bbb Z}[i]$ be an elliptic curve with complex multiplication by the ring of Gaussian numbers $R={\Bbb Z}[i]$, so that the corresponding Rosati involution $\varepsilon$ is just the complex conjugation. In this situation there is no $\Sigma_{R,\varepsilon}$-structure corresponding to the standard polarization of $E$. Indeed, the corresponding line bundle $L(1)$ should be of the form $\O(x)$ where $x$ is a point of order 2 on $E$. Now the identity $[1+i]^*L(1)=L(2)$ leads to a contradiction (see \cite{thetaid} for details). \end{ex} \section{Abelian schemes with real multiplication}\label{real} Let $F$ be a totally real number field, $R$ be its ring of integers. Let $A\rightarrow S$ be an abelian scheme with real multiplication by $R$, i.e. a ring homomorphism $R\rightarrow\operatorname{End}_S(A):r\mapsto [r]_A$ is given. Then the dual abelian scheme $\hat{A}$ also has a natural real multiplication by $R$. Let $J, \hat{J}\subset F$ be fractional ideals for $R$ (= non-zero finitely generated $R$-submodules of $F$) such that $J\hat{J}\subset R$. \begin{defi} An $(J, \hat{J})$-polarization on $A$ is a pair of $R$-module homomorphisms \begin{align*} &\lambda_J:J\rightarrow\operatorname{Hom}_{R}^{\operatorname{sym}}(A,\hat{A}),\\ &\lambda_{\hat{J}}:\hat{J}\rightarrow\operatorname{Hom}_{R}^{\operatorname{sym}}(\hat{A},A) \end{align*} where $\operatorname{Hom}_{R}^{\operatorname{sym}}(A,\hat{A})$ is an $R$-module of symmetric $R$-linear homomorphisms $f:A\rightarrow\hat{A}$ (i.e. $\hat{f}=f$ and $f\circ [r]_A=[r]_{\hat{A}}\circ f$ for any $r\in R$), such that $\lambda_{\hat{J}}(m)\circ\lambda_J(l)=[lm]_A$ and $\lambda_J(l)\lambda_{\hat{J}}(m)=[lm]_{\hat{A}}$ for any $l\in J$, $m\in \hat{J}$. \end{defi} \begin{rem} Usually one also imposes some positivity condition on a polarization. In the case of $(J,\hat{J})$-polarizations one can fix an ordering on $J$: this means that for each embedding $\sigma:F\rightarrow{\Bbb R}$ an orientation of the line $J\otimes_{R,\sigma}{\Bbb R}$ is chosen. Then one should require that if an element $l\in J$ is totally positive then the homomorphism $\lambda_J(l):A\rightarrow\hat{A}$ is positive (i.e. $\lambda_J(l)$ is a polarization in the classical sense). \end{rem} Note that the notion of $(J,\hat{J})$-polarization is equivalent to that of $(Jx, \hat{J}x^{-1})$-polarization for any $x\in F^*$. Also an $(J,\hat{J})$-polarization of $A$ is the same as an $(\hat{J},J)$-polarization of $\hat{A}$. When $\hat{J}=J^{-1}$ we recover the notion of $J$-polarization in the sense of P.~Deligne and G.~Pappas \cite{DePa} (except for the positivity condition). Recall that they define an $J$-polarization of an abelian scheme $A$ with real multiplication by $R$ as an $R$-linear homomorphism $\lambda:J\rightarrow\operatorname{Hom}_{R}^{\operatorname{sym}}(A,\hat{A})$ such that the image of a totally positive element of $J$ under $\lambda$ is positive, and the induced morphism $A\otimes_{R} J\rightarrow\hat{A}$ is an isomorphism. In this case we have also an isomorphism $\hat{A}\otimes_{R} J^{-1}\widetilde{\rightarrow} A$ which induces an $R$-linear homomorphism $\lambda':J^{-1}\rightarrow\operatorname{Hom}_{R}^{\operatorname{sym}}(\hat{A},A)$, hence we get an $(J,J^{-1})$-polarization in our sense. Conversely, given an $(J,J^{-1})$-polarization as above then the morphism $\mu:A\otimes_R J\rightarrow\hat{A}$ induced by $\lambda_J$ and the morphism $\mu':\hat{A}\rightarrow A\otimes_R J$ induced by $\lambda_{J^{-1}}$ are inverse to each other, so that $\lambda_J$ gives an $J$-polarization of $A$ (except for the positivity condition). For a pair $(J,\hat{J})$ as above we define the subgroup $\Gamma(J,\hat{J})\subset\operatorname{SL}_2(F)$ as follows: $$\Gamma(J,\hat{J})=\{ \left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)\in\operatorname{SL}_2(F): a_{11}, a_{22}\in R, a_{12}\in J, a_{21}\in \hat{J} \}.$$ Note that for $x\in F^*$ the homomorphisms $\rho_{J,\hat{J}}$ and $\rho_{Jx,\hat{J}x^{-1}}$ are compatible with the natural isomorphism $\Gamma(J,\hat{J})\simeq \Gamma(Jx,\hat{J}x^{-1})$ (induced by the conjugation by $\left( \matrix x^{\frac{1}{2}} & 0\\ 0 & x^{-\frac{1}{2}} \endmatrix \right)$). In particular, in the case $R={\Bbb Z}$ the group $\Gamma(J,\hat{J})$ is always isomorphic to the principal congruenz-subgroup $\Gamma_0(N)=\Gamma({\Bbb Z},N{\Bbb Z})\subset \operatorname{SL}_2({\Bbb Z})$ (for some $N>0$). If $A$ is $(J,\hat{J})$-polarized then using $\lambda_J$ and $\lambda_{\hat{J}}$ we can define a homomorphism $$\rho_{J,\hat{J}}:\Gamma(J,\hat{J})\rightarrow\operatorname{SL}_2(A): \left( \matrix a_{11} & a_{12}\\ a_{21} & a_{22} \endmatrix \right)\mapsto \left( \matrix [a_{11}]_{\hat{A}} & \lambda_J(a_{12})\\ \lambda_{\hat{J}}(a_{21}) & [a_{22}]_A \endmatrix \right)$$ \begin{lem} For any non-zero element $r\in R$ (resp. $l\in J$, $m\in \hat{J}$) the corresponding morphism $[r]_A$ (resp. $\lambda_J(l)$, $\lambda_{\hat{J}}(m)$) is an isogeny. \end{lem} \noindent {\it Proof} . There exists a non-zero integer $N$ such that $r'=N/r\in R$, so that $[r']_A\circ [r]_A=[N]_A$. Hence, $[r]_A$ is an isogeny on each fiber, therefore, it is an isogeny. Similar argument works for $\lambda_J(l)$ and $\lambda_{\hat{J}}(m)$. \qed\vspace{3mm} Recall that we denote by $\Gamma(A)$ the image of the homomorphism $\widehat{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$. Let assume for simplicity that the image $\rho_{J,\hat{J}}$ is contained in $\Gamma(A)$. Otherwise, we can consider a finite flat base change $S'\rightarrow S$ such that the image of $\Gamma(J,\hat{J})$ is contained in $\Gamma(A_{S'})$ where $A_{S'}$ is the induced abelian scheme over $S'$. Indeed, recall that one has an exact sequence $$0\rightarrow\hat{A}_2\rightarrow\operatorname{Pic}^{\operatorname{sym}}(A)\rightarrow\underline{\operatorname{Hom}}^{\operatorname{sym}}(A,\hat{A})\ra0$$ of sheaves in fppf topology, hence the boundary homomorphism $J\stackrel{\lambda_J}{\rightarrow}\operatorname{Hom}^{\operatorname{sym}}(A,\hat{A})\rightarrow H^1(S,\hat{A}_2)$ which can be considered as a $J^*\otimes\hat{A}_2$-torsor over $S$ where $J^*=\operatorname{Hom}_{\Bbb Z}(J,{\Bbb Z})$ (similarly, for $\lambda_{\hat{J}}$ we get an $\hat{J}^*\otimes A_2$-torsor). Now we can take $S'$ to be the corresponding $J^*\otimes\hat{A}_2\times \hat{J}^*\otimes A_2$-torsor over $S$. By the above lemma we can define an obstruction homomorphism $\delta_{J,\hat{J}}:\Gamma(J,\hat{J})\rightarrow\operatorname{Br}(S)$ as follows: if $a_{21}$-entry of the matrix $h\in\Gamma(J,\hat{J})$ is non-zero then the same entry of $\overline{g}=\rho_{J,\hat{J}}(h)$ is an isogeny and we can put $\delta_{J,\hat{J}}(h)=e(G_g)$ where $g\in\widetilde{\operatorname{SL}}_2(A)$ lies above $\overline{g}$. Otherwise, $\delta_{J,\hat{J}}(h)=0$. As in proposition \ref{obshom} one can check that $\delta_{J,\hat{J}}$ is a homomorphism. Let $I\subset R$ be a non-zero ideal. Let us denote by $\overline{\Gamma}_I(J,\hat{J})$ the group of matrices $\left( \matrix \bar{a}_{11} & \bar{a}_{12}\\ \bar{a}_{21} & \bar{a}_{22} \endmatrix\right)$ where $\bar{a}_{11},\bar{a}_{22}\in R/I$, $\bar{a}_{12}\in J/IJ$, $\bar{a}_{21}\in \hat{J}/I\hat{J}$, such that $\bar{a}_{11}\bar{a}_{22}-\bar{a}_{12}\bar{a}_{21}=1$. Here we use the natural homomorphism of $R/I$-modules $J/IJ\otimes \hat{J}/I\hat{J}\rightarrow R/IR$ induced by $J\otimes \hat{J}\rightarrow R$. \begin{lem}\label{surj} Assume that $I\subset J\hat{J}$. Then the natural reduction homomorphism $\pi_I:\Gamma(J,\hat{J})\rightarrow\overline{\Gamma}_I(J,\hat{J})$ is surjective. \end{lem} \noindent {\it Proof} . Let $\left( \matrix \bar{a}_{11} & \bar{a}_{12}\\ \bar{a}_{21} & \bar{a}_{22} \endmatrix\right)\in \overline{\Gamma}_I(J,\hat{J})$ be any element. Choose any non-zero liftings $a_{12}\in J$, $a'_{21}\in \hat{J}$ and $a'_{22}\in R$ of $\bar{a}_{12}$, $\bar{a}_{21}$ and $\bar{a}_{22}$. For an element $r\in R$ and a finite $R$-module $Q$ we say that $r$ is relatively prime to $Q$ if $Q=rQ$, or equivalently, $r\not\in{\goth p}$ for any prime ideal ${\goth p}$ associated with $Q$. By the Chinese remainder theorem we can lift $\bar{a}_{11}$ to an element $a_{11}\in R$ which is relatively prime to $J/Ra_{12}$. On the other hand, $a_{11}$ is relatively prime to $R/J\hat{J}$ since $a_{11}a'_{22}\equiv 1\mod(I)$ and $I\subset J\hat{J}$. Therefore, $a_{11}$ is relatively prime to $J/J\hat{J}a_{12}$ (since $\operatorname{supp}(J/J\hat{J}a_{12})\subset\operatorname{supp}(R/J\hat{J})\cup\operatorname{supp}(J/Ra_{12})$). It follows that $J=a_{11}J+J\hat{J}a_{12}=(Ra_{11}+\hat{J}a_{12})J$ which implies the equality $R=Ra_{11}+\hat{J}a_{12}$. Thus, we can write $1=ra_{11}+ma_{12}$ where $r\in R$, $m\in \hat{J}$. Let $x=a_{11}a'_{22}-a_{12}a'_{21}-1\in I$. Then replacing $a'_{22}$ and $a'_{21}$ by $a_{22}=a'_{22}-xr$, $a_{21}=a'_{21}-xm$ we achieve $a_{11}a_{22}-a_{12}a_{21}=1$. \qed\vspace{3mm} \begin{prop}\label{triv2} The homomorphism $\delta_{J,\hat{J}}:\Gamma(J,\hat{J})\rightarrow\operatorname{Br}(S)$ is trivial. \end{prop} \noindent {\it Proof} . Consider the reduction homomorphism $\pi_I:\Gamma(J,\hat{J})\rightarrow\overline{\Gamma}_I(J,\hat{J})$ where $I\subset 2J\hat{J}$. Then by Lemma \ref{surj} $\pi_I$ is surjective. On the other hand, the kernel of $\pi_I$ is contained in the subgroup $\Gamma(J,2\hat{J})\subset\Gamma(J,\hat{J})$. By Proposition \ref{triv1} the restriction of $\delta_{J,\hat{J}}$ to the subgroup $\Gamma(J,2\hat{J})$ is trivial. Hence, $\delta_{J,\hat{J}}(\ker(\pi_I))=0$, so that $\delta_{J,\hat{J}}=\overline{\delta}\circ\pi_I$ for some homomorphism $\overline{\delta}:\overline{\Gamma}_I(J,\hat{J})\rightarrow\operatorname{Br}(S)$. Moreover, since by Lemma \ref{surj} the homomorphism $\Gamma(J,2\hat{J})\rightarrow\overline{\Gamma}_I(J,2\hat{J})$ is surjective, it follows that $\overline{\delta}$ is trivial on matrices with $\bar{a}_{21}\in 2\hat{J}/I\hat{J}$. In particular, $\overline{\delta}$ vanishes on any diagonal matrix. Let $h=\left(\matrix \bar{a}_{11} & \bar{a}_{12}\\ \bar{a}_{21} & \bar{a}_{22} \endmatrix\right)$ be any element of $\overline{\Gamma}_I(J,\hat{J})$. Then $\bar{a}_{11}\bar{a}_{22}\equiv 1\mod (J\hat{J})$, hence $\bar{a}_{11}\mod (J\hat{J})$ is a unit in $R/J\hat{J}$. Let $u\in (R/I)^*$ be any unit such that $u\equiv a_{11}\mod (J\hat{J})$ (such a unit always exists since $R/I$ is an artinian ring). Then replacing $h$ by $h\cdot\left(\matrix u & 0\\ 0 & u^{-1} \endmatrix\right)$ we reduce the problem of showing that $\overline{\delta}(h)=0$ to the case when $\bar{a}_{11}\equiv 1\mod(J\hat{J})$. Now we use the result of L.~Vaserstein \cite{Vas} which asserts that if $F\neq {\Bbb Q}$ then the subgroup of $\Gamma(J,\hat{J})$ consisting of matrices with $a_{11}\equiv 1\mod(J\hat{J})$ is generated by elementary matrices, i.e. matrices of the form $\left( \matrix 1 & l \\ 0 & 1 \endmatrix\right)$ and $\left( \matrix 1 & 0 \\ m & 1 \endmatrix\right)$, where $l\in L$, $m\in M$. In the case $F={\Bbb Q}$ we may assume that $J={\Bbb Z}$, $\hat{J}=N{\Bbb Z}$ for some $N\in{\Bbb Z}$ and the corresponding assertion for $\Gamma({\Bbb Z},N{\Bbb Z})=\Gamma_0(N)$ is trivial. Note that $\delta_{J,\hat{J}}$ vanishes on elementary matrices (the corresponding Heisenberg groups are either Mumford groups of line bundles or trivial, so they admit Shr\"odinger representations), hence it vanishes on any matrix with $a_{11}\equiv 1\mod(J\hat{J})$ and we are done. \qed\vspace{3mm} Let $\widehat{\Gamma}(J,\hat{J})$ be the preimage of $\Gamma(J,\hat{J})$ under the homomorphism $\widehat{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$. \begin{thm} For every $(J,\hat{J})$-polarized abelian scheme $A$ over $S$ with $R$-multiplication such that the image of $\rho_{J,\hat{J}}$ is contained in $\Gamma(A)$ there exists a faithful action of a central extension of the group $\widehat{\Gamma}(J,\hat{J})$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ on ${\cal D}^b(A)$. Without this assumption we always have compatible faithful projective actions of $\Gamma(2J,2\hat{J})$ on ${\cal D}^b(A)$ and of $\widehat{\Gamma}(J,\hat{J})$ on $A_{S'}$ for some finite flat base change $S'\rightarrow S$. \end{thm} \begin{cor} Let $A$ be an abelian scheme over a normal noetherian connected base $S$. Assume that the projection $\widehat{\operatorname{SL}}_2(A)\rightarrow\operatorname{SL}_2(A)$ is surjective and $\operatorname{End}_S(A)\simeq R$ is a totally real field. Then there is a faithful action of a central extension of $\widehat{\operatorname{SL}}_2(A)$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ on ${\cal D}^b(A)$. \end{cor} \noindent {\it Proof} . By Prop. X 1.5 of \cite{Geer} the general fiber $A_K$ admits an $R$-linear polarization $\lambda:A_K\rightarrow\hat{A}_K$. Hence, $J=\operatorname{Hom}_K(A_K,\hat{A}_K)$ and $\hat{J}=\operatorname{Hom}_K(\hat{A}_K,A_K)$ can be considered as fractional ideals for $R$, such that $J\hat{J}\subset R$. By definition $\operatorname{SL}_2(A_K)=\Gamma(J,\hat{J})$ and by Lemma \ref{extend} we have an $(J,\hat{J})$-polarization on $A$. \qed\vspace{3mm} Now let us consider the case of abelian scheme $A$ with $R$-linear principal polarization $\phi:A\widetilde{\rightarrow}\hat{A}$. In this case we have a natural inclusion $$i_{\phi}:\Sp_{2n}(R)\rightarrow\operatorname{SL}_2(A^n): \left( \matrix M_{11} & M_{12}\\ M_{21} & M_{22} \endmatrix \right)\mapsto \left( \matrix [M_{11}]_{\hat{A}} & \phi_{(n)}[M_{12}]_A\\ {[M_{21}]_{\hat{A}}\phi_{(n)}^{-1}} & [M_{22}]_A \endmatrix \right)$$ where $A^n$ is the relative $n$-th cartesian power of $A$ with the induced polarization $\phi_{(n)}$, $M_{ij}\in\operatorname{Mat}_n(R)$, for every abelian scheme $A$ with multiplication by $R$ we denote the natural map $\operatorname{Mat}_n(R)\rightarrow\operatorname{End}(A^n)$ by $M\mapsto [M]_A$. Now we claim that if $R$ is unramified at $2$ then one can split the extension $\widehat{\operatorname{SL}}_2(A^n)\rightarrow\operatorname{SL}_2(A)$ over $\Sp_{2n}(R)$ provided that a symmetric line bundle $L(1)$ on $A$ is given such that $\phi_{L(1)}=\phi$. According to Theorem \ref{Sig} it is sufficient to construct a $\Sigma_{R,\operatorname{id}}$-structure for $\phi$. Note that since $R$ is unramified at $2$ every element $r\in R$ can be represented in the form $r=a^2+2b$ with $a,b\in R$. Now we define $L(r)=[a]^*L(1)\otimes (\phi,[b]_A)^*\cal P$ where $\cal P$ is the Poincar\'e line bundle. It is easy to see that $L(r)$ doesn't depend on a choice of $a$ and $b$, and satisfies (\ref{CM1}) and (\ref{CM2}) with $\varepsilon=\operatorname{id}$. The induced structure for $A^n$ and $\operatorname{Mat}_n(R)$ is given by the homomorphism $$\operatorname{Mat}^{\operatorname{sym}}_n(R)\rightarrow\operatorname{Pic}^{\operatorname{sym}}(A^n): B=(b_{ij})\mapsto L(B)= \bigotimes_{i<j}(\phi p_i, [b_{ij}]_A p_j)^*\cal P\otimes \bigotimes_i p_i^*L(b_{ii})$$ where $\operatorname{Mat}^{\operatorname{sym}}_n(R)$ denotes symmetric matrices with entries in $R$, $p_i:A^n\rightarrow A$ is the projection on the $i$-th factor. It is easy to see that $\phi_{L(B)}=[B]_A$ and that $[C]_A^*L(B)\simeq L(\sideset{^t}{}{C} B C)$ for any $C\in\operatorname{Mat}_n(R)$. Now we can write the required splitting $$\Sp_{2n}(R)\rightarrow\widehat{\operatorname{SL}}_2(A^n): M=\left( \matrix M_{11} & M_{12}\\ M_{21} & M_{22} \endmatrix \right)\mapsto (i_{\phi}(M), (\phi^{-1}_{(n)})^*L(\sideset{^t}{_{11}}{M}M_{21}), L(\sideset{^t}{_{22}}{M}M_{12})).$$ Using the above splitting we can construct a projective action of $\Sp_{2n}(R)$ on $D^b(A^n)$. The vanishing of the obstacle follows in this case from the fact that $\Sp_{2n}(R)$ is generated by elementary matrices established in \cite{BMS}. \begin{thm} Let $A$ be an abelian scheme with real multiplication by $R$ over $S$, $L(1)$ be a symmetric line bundle on $A$ rigidified along the zero section such that $\phi_{L(1)}:A\rightarrow\hat{A}$ is an $R$-linear isomorphism. Assume that $R$ is unramified at $2$. Then there is a canonical faithful action of a central extension of $\Sp_{2n}(R)$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ on ${\cal D}^b(A^n)$ where $A^n$ is the relative cartesian power of $A$, $n\ge 1$. These actions are compatible via the natural embeddings ${\cal D}^b(A^n)\rightarrow{\cal D}^b(A^{n+1})$ and $\Sp_{2n}(R)\rightarrow\Sp_{2n+2}(R)$. \end{thm} \noindent {\it Proof} . The same argument as in Proposition \ref{obshom} allows to define an obstacle homomorphism $\delta:\Sp_{2n}(R)\rightarrow\operatorname{Br}(S)$ such that $\delta(h)=0$ if and only if there exists a global object in $\operatorname{Isom}_{H(X)}(\rho,\rho^h)$ where $X=\hat{A}^n\times_S A^n$, $\rho$ is the representation of the Heisenberg groupoid on ${\cal D}^b(A^n)$. It is easy to check that $\delta$ vanishes on elementary matrices, hence it is zero. \qed\vspace{3mm} \section{The central extension} In this section we describe explicitly the central extension of $\Sp_{2n}({\Bbb Z})$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ corresponding to the projective action defined in the previous section. We are going to use a presentation of $\Sp_{2n}({\Bbb Z})$ by generators and relations borrowed from \cite{Cl}. We always use the standard symplectic basis $e_1,\ldots,e_n,f_1,\ldots f_n$ in ${\Bbb Z}^{2n}$ such that $(e_i,f_j)=\delta_{i,j}$). First of all let us introduce the relevant elementary matrices following the notation of \cite{Cl} 5.3.1. Let $\S_{2n}$ be the set of pairs $(i,j)$ where $1\le i,j\le 2n$ which are not of the form $(2k-1,2k)$ or $(2k,2k-1)$. Then for for every $(i,j)\in \S_{2n}$ we define an elementary matrix $E_{ij}$ as follows: $$E_{2k,2l}= \left(\matrix 1 & 0 \\ \gamma_{k,l} & 1\endmatrix\right),$$ $$E_{2k-1,2l-1}= \left(\matrix 1 & -\gamma_{k,l} \\ 0 & 1\endmatrix\right),$$ $$E_{2k-1,2l}= \left(\matrix e_{kl} & 0 \\ 0 & e_{lk}^{-1}\endmatrix\right),$$ $$E_{2l,2k-1}=E_{2k-1,2l}$$ where $\gamma_{kl}$ has zero $(\a,\b)$-entry unless $(\a,\b)=(k,l)$ or $(\a,\b)=(l,k)$, in the latter case $(\a,\b)$-entry is 1; $e_{kl}$ for $k\neq l$ is the usual elementary matrix with units on the diagonal and at $(k,l)$-entry and zeros elsewhere. Now theorem 9.2.13 of \cite{Cl} asserts that for $n\ge 3$ the group $\Sp_{2n}({\Bbb Z})$ has a presentation consisting of the generators $E_{ij}=E_{ji}$ (where $(i,j)\in\S_{2n}$) subject to the relations \begin{enumerate} \item $[E_{ij},E_{kl}]=1$, if $(i,k), (i,l), (j,k), (j,l)$ are in $\S_{2n}$ \item $[E_{ij},E_{kl}]=E_{il}$, if $(j,k)\not\in\S_{2n}$, $j$ is even, and $i$, $j$, $k$, and $l$ are distinct \item $[E_{ij},E_{ki}]=E_{ii}^2$, if $(j,k)\not\in\S_{2n}$, $j$ is even, and $i$, $j$, and $k$ are distinct \item $[E_{ii},E_{kl}]=E_{il}E_{ll}^{-1}$ if $(i,k)\not\in\S_{2n}$, $i$ is even, and $i$, $k$, and $l$ are distinct \item $[E_{ii},E_{kl}]=E_{il}^{-1}E_{ll}^{-1}$ if $(i,k)\not\in\S_{2n}$, $i$ is odd, and $i$, $k$, and $l$ are distinct \item $(E_{11}E_{22}E_{11})^4=1$. \end{enumerate} It is convenient to introduce also the symplectic matrix $$\varphi= \left(\matrix 0 & -1 \\ 1 & 0\endmatrix\right).$$ Then one has the following relations \begin{equation}\label{conj1} \varphi^{-1}E_{2k-1,2l-1}\varphi=E_{2k,2l} \end{equation} for all $1\le k,l\le n$, \begin{equation}\label{conj2} \varphi^{-1}E_{2k-1,2l}\varphi=E_{2l-1,2k}^{-1} \end{equation} for all $k\neq l$. In particular, the group $\Sp_{2n}({\Bbb Z})$ is generated by $\varphi$ and $E_{ij}$ with $i$ odd. The latter set of generators is more convenient from the point of view of our projective representation on ${\cal D}^b(A^n)$ since the functors corresponding to $E_{ij}$ with $i$ odd are very easy to describe (see the proof of the theorem below). Let us denote by $\widetilde{\Sp}_{2n}({\Bbb Z})$ the group with generators $E_{ij}=E_{ji}$ ($(i,j)\in\S_{2n}$) and one more generator $\epsilon$ subject to relations (1)--(5) above, the commutativity relation $[\epsilon,E_{ij}]=1$ for all $(i,j)\in\S_{2n}$, and the modified relation (6) $$(E_{11}E_{22}E_{11})^4=\epsilon.$$ Let $\pi:A\rightarrow S$ be an abelian scheme with a symmetric line bundle $L$ (rigidified along the zero section) which induces a principal polarization $\phi:A\widetilde{\rightarrow}\hat{A}$. Let us also denote $\Delta(L)=2\pi_*L+e^*\omega_{A/S}\in\operatorname{Pic}(S)$ where $\omega_{A/S}$ is the relative canonical bundle. It is known that $4\cdot\Delta(L)=0$ (see e.~g. \cite{FC}, I, 5.1). \begin{thm}\label{centrext} Let $n\ge 3$. The group $\widetilde{\Sp}_{2n}({\Bbb Z})$ is a central extension of $\Sp_{2n}({\Bbb Z})$ by ${\Bbb Z}$. The central extension of $\Sp_{2n}({\Bbb Z})$ by ${\Bbb Z}\times\operatorname{Pic}(S)$ corresponding to the projective action on ${\cal D}^b(A^n)$ is obtained from the $\widetilde{\Sp}_{2n}({\Bbb Z})$ by the push-forward with respect to the homomorphism ${\Bbb Z}\rightarrow{\Bbb Z}\times\operatorname{Pic}(S): 1\mapsto (2g,2\Delta(L))$ \end{thm} \noindent {\it Proof} . Let us choose the intertwining functors corresponding to the generators $\varphi$ and $E_{ij}$ (with $i$ odd) in the following way. The functor corresponding to $\varphi$ is the composition $\phi_{(n)}^{-1}\circ F_{A^n}$ where $F_{A^n}:{\cal D}^b(A^n)\rightarrow{\cal D}^b(\hat{A}^n)$ is the Fourier-Mukai transform. The functor corresponding to $E_{2k-1,2l-1}$ is simply tensor multiplication with the line bundle $L(\gamma_{k,l})$. Note that $L(\gamma_{k,l})=(\phi p_k,p_l)^*\cal P$ if $k\neq l$ while $L(\gamma_{kk})=p_k^*L(1)$. At last the functor corresponding to $E_{2k-1,2l}$ is $[e_{lk}]_A^*$. We claim that these functors satisfy all the relations (1)-(5). Let $P\subset\Sp_{2n}({\Bbb Z})$ be the subgroup of matrices of the form $\left(\matrix * & * \\ 0 & *\endmatrix\right)$. Then there is an obvious action of $P$ on ${\cal D}^b(A^n)$ such that the element $\left(\matrix\sideset{^t}{}C^{-1} & 0 \\ 0 & C \endmatrix\right) \left(\matrix 1 & B \\ 0 & 1\endmatrix\right)$ where $C\in\operatorname{GL}_n({\Bbb Z})$, $B\in\operatorname{Mat}^{\operatorname{sym}}(n,{\Bbb Z})$, acts by the functor $[C^{-1}]_A^*\circ(\cdot \otimes L(-B))$. It is easy to see that our definition of the functors corresponding to the generators $E_{ij}$ for $i$ odd is compatible with this action. This means that all the relations out of (1)--(5) which contain only these generators are satisfied by the corresponding functors. Furthermore, using the relations (\ref{conj1}), (\ref{conj2}) one can see that all the relations out of (1)--(5) containing a generator $E_{ij}$ with $i$ and $j$ of opposite parity in the left hand side, are satisfied by our functors. Similarly, the relation (5) follows from (4) using the relations (\ref{conj1}) and (\ref{conj2}). It remains to check the relation (1) for $i$ and $j$ even, and $k$ and $l$ odd, the relations (2), (3) for $i$ even and $l$ odd, and the relation (4) for $l$ odd. This can be done directly applying the both sides of a relation to the object $e_*\O_S\in{\cal D}^b(A^n)$. Thus, the relations (1)--(5) hold for our functors. Now using that $F_A^2\simeq [-1]^*(\cdot)\otimes\omega_{A/S}^{-1}[-g]$, where $g=\dim A/S$, and that $F_A(L)\simeq L^{-1}\otimes\pi^*\pi_*L$ one can easily compute that the functor corresponding to $(E_{11}E_{22}E_{11})^4$ is $(\cdot)\otimes\Delta(L)^{\otimes 2}[2g]$. \qed\vspace{3mm} The ${\Bbb Z}$-part of the central extension of $\Sp_{2n}({\Bbb Z})$ acting on ${\cal D}^b(A^n)$ was computed in \cite{Orlov}. Namely, for $g=1$ the corresponding class in $H^2(\Sp_{2n}({\Bbb Z}),{\Bbb Z})$ is a half of the class of the cocycle given by the Malsov index. On the other hand, it is easy to see that the class of the central extension $\widetilde{\Sp}_{2n}({\Bbb Z})$ is a generator of $H^2(\Sp_{2n}({\Bbb Z}),{\Bbb Z})$ for sufficiently large $n$. Indeed, it is known that $H^2(\Sp_{2n}({\Bbb Z}),{\Bbb Z})={\Bbb Z}$ for large $n$ while $\Sp_{2n}({\Bbb Z})=[\Sp_{2n}({\Bbb Z}),\Sp_{2n}({\Bbb Z})]$ for $n\ge 3$. Moreover, the relations (2) and (4) easily imply that the element $\epsilon$ belongs to $[\widetilde{\Sp}_{2n}({\Bbb Z}),\widetilde{\Sp}_{2n}({\Bbb Z})]$, hence $\widetilde{\Sp}_{2n}({\Bbb Z})=[\widetilde{\Sp}_{2n}({\Bbb Z}),\widetilde{\Sp}_{2n}({\Bbb Z})]$. It follows, that the central extension of $\Sp_{2n}({\Bbb Z})$ by ${\Bbb Z}/p{\Bbb Z}$ obtained from $\widetilde{\Sp}_{2n}({\Bbb Z})$ is non-trivial for every prime $p$, so our claim follows. Let $\Gamma_{1,2}\subset\Sp_{2n}({\Bbb Z})$ be the subgroup of matrices $\left(\matrix M_{11} & M_{12} \\ M_{21} & M_{22}\endmatrix\right)$ such that $\sideset{^t}{_{11}}{M}M_{12}$ and $\sideset{^t}{_{22}}{M}M_{21}$ have even diagonal entries. Let $A$ be a principally polarized abelian scheme over $S$. Then we have a canonical splitting of the projection $\widehat{\operatorname{SL}}_2(A^n)\rightarrow\operatorname{SL}_2(A^n)$ over $\Gamma_{1,2}$ which is constructed as in the previous section using line bundles $$L(B)=\bigotimes_{i<j}(\phi p_i, [b_{ij}]_A p_j)^*\cal P\otimes \bigotimes_i (\phi p_i, [b_{ii}/2] p_i)^*\cal P$$ associated with symmetric integer even-diagonal matrices $B=(b_{ij})$ (note that this time we don't need any additional data on $A$). It is known (see \cite{ThetaII} A.4) that $\Gamma_{1,2}$ is generated by elements $$\varphi, \left(\matrix \sideset{^t}{^{-1}}{C} & 0 \\ 0 & C \endmatrix\right), \left(\matrix 1 & B \\ 0 & 1 \endmatrix\right)$$ where $C\in\operatorname{GL}_n({\Bbb Z})$, $B$ is symmetric integer with even diagonal. Obviously, this implies vanishing of the obstruction for the projective action of $\Gamma_{1,2}$ on ${\cal D}^b(A^n)$ by intertwining operators, hence this leads to a central extension of $\Gamma_{1,2}$ by ${\Bbb Z}\times\operatorname{Pic}(S)$. \begin{prop} Let $A/S$ be a principally polarized abelian scheme of dimension $g\ge 3$. Then the central extension of $\Gamma_{1,2}$ by $\operatorname{Pic}(S)$ acting on ${\cal D}^b(A^n)$ up to shifts is trivial. \end{prop} \noindent {\it Proof} . Considering a finite flat covering of $S$ corresponding to a choice of a symmetric line bundle inducing a principal polarization and using Theorem \ref{centrext} one can see that the central extension in question is induced by a central extension of $\Gamma_{1,2}$ by the torsion subgroup $\operatorname{Pic}(S)^{\operatorname{tors}}\subset\operatorname{Pic}(S)$. Note that it is sufficient to prove our assertion in the case when $A$ is the universal abelian scheme over the moduli stack $\AA_g$ of principally polarized abelian schemes. It remains to notice that $\operatorname{Pic}(\AA_g)^{\operatorname{tors}}=0$ since $\Sp_{2g}({\Bbb Z})$ has no abelian quotients for $g\ge 3$ (this is deduced using the Kummer exact sequence --- see \cite{Mu}). \qed\vspace{3mm} \begin{cor} The central extension of $\Sp_{2n}({\Bbb Z})$ by ${\Bbb Z}/2{\Bbb Z}$ obtained by push-forward from $\widetilde{\Sp}_{2n}({\Bbb Z})$ has a splitting over $\Gamma_{1,2}$. \end{cor}
1997-12-03T21:03:27
9712
alg-geom/9712006
en
https://arxiv.org/abs/alg-geom/9712006
[ "alg-geom", "math.AG", "math.DG", "q-alg" ]
alg-geom/9712006
Toshiyuki Akita
Toshiyuki Akita
Homological infiniteness of Torelli groups
AMSLaTeX v.1.2, 6 pages
null
null
Fukuoka University, Department of Applied Mathematics, preprint 1997/12/01
null
We prove that rational homology of the Torelli group of genus g is infinite dimensional, provided g>6. This means that rational homology of the Torelli space of genus g>6 is infinite dimensional. The Torelli groups with marked points are also considered. In addition, we prove that rational homology of the subgroup of the Torelli group of genus g generated by all the Dehn twists along separating simple closed curves is infinite dimensional for g>2.
[ { "version": "v1", "created": "Wed, 3 Dec 1997 20:03:26 GMT" } ]
2007-05-23T00:00:00
[ [ "Akita", "Toshiyuki", "" ] ]
alg-geom
\section{Introduction} Let $\Sigma_g$ be a closed orientable surface of genus $g\geq 2$. Let $\mathcal{M}_{g,r}^n$ be the mapping class group of $\Sigma_g$ relative to $n$ distinguished points and $r$ fixed embedded disks. The action of $\mathcal{M}_{g,r}^n$ on the homology of $\Sigma_g$ induces a surjective homomorphism \[ \mathcal{M}_{g,r}^n\rightarrow Sp(2g,\mathbb{Z}), \] where $Sp(2g,\mathbb{Z})$ is the Siegel modular group. The Torelli group $\mathcal{I}_{g,r}^n$ is defined to be its kernel so that we have an extension \[ 1\rightarrow\mathcal{I}_{g,r}^n\rightarrow\mathcal{M}_{g,r}^n\rightarrow Sp(2g,\mathbb{Z})\rightarrow 1. \] We omit the decorations $n$ and $r$ when they are zero. In a series of papers \cite{johnson0,johnson1,johnson2,johnson3}, D. Johnson obtained several fundamental results concerning the structure of $\mathcal{I}_g$ and $\mathcal{I}_{g,1}$ (see also \cite{johnson-survey,hain-survey}). In particular, he proved that $\mathcal{I}_g$ and $\mathcal{I}_{g,1}$ are finitely generated for all $g\geq 3$. On the contrary, A. Miller and D. McCullough \cite{miller} showed that $\mathcal{I}_2$ (and hence $\mathcal{I}_{2,r}^n$ for all $n,r\geq0$) is not finitely generated. G. Mess \cite{mess} showed that $\mathcal{I}_2$ is a free group on infinitely many generators. Johnson and J. Millson showed that $H_3(\mathcal{I}_3,\mathbb{Z})$ contains a free abelian group of infinite rank (cf. \cite{mess}). It is not known whether $\mathcal{I}_{g,r}^n$ is finitely presented for $g\geq 3$. In this paper, we will prove: \begin{thm}\label{thm-torelli} For all $n\geq 0$, the rational homology $H_*(\mathcal{I}_g^n,\mathbb{Q})$ of the Torelli group $\mathcal{I}_g^n$ is infinite dimensional over $\mathbb{Q}$ if $g$ is sufficiently large compared with $n$. In particular, $H_*(\mathcal{I}_g,\mathbb{Q})$ and $H_*(\mathcal{I}_{g}^1,\mathbb{Q})$ are infinite dimensional for $g\geq 7$. \end{thm}\noindent This theorem yields the negative answer to the question posed by Johnson \cite{johnson-survey} which asks whether the Torelli space $\mathbf{T}_{g}^1$ (see \S 2 for the definition) is homotopy equivalent to a finite complex, provided $g\geq 7$. For $n+r\leq 1$, let $\mathcal{K}_{g,r}^n$ be the subgroup of $\mathcal{M}_{g,r}^n$ generated by all the Dehn twists along separating simple closed curves. The groups $\mathcal{K}_{g,1}$ and $\mathcal{K}_g$ are related to the Casson invariants of homologly 3-spheres through the work of S. Morita \cite{morita1989,morita1991}. For $g=2$, $\mathcal{K}_2$ is equal to $\mathcal{I}_2$ so that it is a free group on infinitely many generators. In contrast, the group $\mathcal{K}_g$ is not free for $g\geq 3$ and almost nothing is known about the structure of this group, however, we can prove: \begin{thm}\label{thm-kg} For all $g\geq 3$, $H_*(\mathcal{K}_g,\mathbb{Q})$ and $H_*(\mathcal{K}_{g}^1,\mathbb{Q})$ are infinite dimensional over $\mathbb{Q}$. \end{thm}\noindent \begin{remark} There is no general agreement on the definition of $\mathcal{I}_{g,r}^n$ when $r+n>1$. We employ the one which was used in \cite{hain-survey}. It differs from that which was given in \cite{johnson2}. \end{remark} \section{Preliminaries} In this section, we recall relevant definitions and facts concerning of Torelli groups which will be used later. The reader should refer to \cite{hain-survey,harer} for further detail. Let $\mathcal{T}_g^n$ be the Teichm\"uller space with $n$ marked points. $\mathcal{M}_g^n$ acts on $\mathcal{T}_g^n$ properly discontinuously and the quotient space $\mathbf{M}_g^n=\mathcal{T}_g^n/\mathcal{M}_g^n$ is, by definition, the moduli space of curves of genus $g$ with $n$ marked points. The Torelli group $\mathcal{I}_g^n$ is torsion-free and hence it acts on $\mathcal{T}_g^n$ freely so that the quotient space $\mathbf{T}_g^n=\mathcal{T}_g^n/\mathcal{I}_g^n$ is a complex manifold. Moreover, since $\mathcal{T}_g^n$ is contractible, $\mathbf{T}_g^n$ is the classifying space of $\mathcal{I}_g^n$ so that there is a canonical isomorphism \[ H_*(\mathcal{I}_{g}^n,\mathbb{Z})\cong H_*(\mathbf{T}_{g}^n,\mathbb{Z}). \] $\mathbf{T}_g^n$ is called the {\em Torelli space} and is important in algebraic geometry (cf. \cite{hain-survey}). The action of $\mathcal{M}_g^n$ on $\mathcal{T}_g^n$ induces the properly discontinuous action of $Sp(2g,\mathbb{Z})=\mathcal{M}_g^n/\mathcal{I}_g^n$ on $\mathbf{T}_g^n$ so that the quotient space $\mathbf{T}_g^n/Sp(2g,\mathbb{Z})$ coincides with $\mathbf{M}_g^n$. Let $\mathfrak{S}_g$ be the Siegel upper half space of degree $g$. The group $Sp(2g,\mathbb{Z})$ acts on $\mathfrak{S}_g$ properly discontinuously and the quotient space $Sp(2g,\mathbb{Z})\backslash\mathfrak{S}_g$ is identified with the moduli space of principally polarized abelian varieties of dimension $g$. For an integer $L\geq 3$, let $\Gamma(L)$ be the principal congruence subgroup of $Sp(2g,\mathbb{Z})$ of level $L$ defined to be the kernel of the canonical homomorphism \[ Sp(2g,\mathbb{Z})\rightarrow Sp(2g,\mathbb{Z}/L\mathbb{Z}). \] $\Gamma(L)$ is a torsion-free subgroup of finite index of $Sp(2g,\mathbb{Z})$ and hence acts on $\mathfrak{S}_g$ freely. The quotient space $\Gamma(L)\backslash\mathfrak{S}_g$ is identified with the moduli space of principally polarized abelian varieties of dimension $g$ with level $L$ structure. It is known that $\Gamma(L)\backslash\mathfrak{S}_g$ is homotopy equivalent to a finite complex (see \cite{borel-serre,serre-arithmetic} for instance). Let $\mathcal{M}_g^n(L)\subset\mathcal{M}_g^n$ be the full inverse image of $\Gamma(L)$ under the homomorphism $\mathcal{M}_g^n\rightarrow Sp(2g,\mathbb{Z})$ so that it fits into the extension \begin{equation}\label{mg-level-l} 1\rightarrow \mathcal{I}_g^n\rightarrow \mathcal{M}_g^n(L)\rightarrow \Gamma(L)\rightarrow 1. \end{equation} $\mathcal{M}_g^n(L)$ is a torsion-free subgroup of finite index of $\mathcal{M}_g^n$ and hence acts on $\mathcal{T}_g^n$ freely. The quotient space $\mathbf{M}_g^n(L)=\mathcal{T}_g^n/\mathcal{M}_g^n(L)$ is, by definition, the moduli space of curves of genus $g$ with $n$ marked points and level $L$ structure. The action of $\mathcal{M}_g^n(L)$ on $\mathcal{T}_g^n$ induces the free action of $\Gamma(L)=\mathcal{M}_g^n(L)/\mathcal{I}_g^n$ on $\mathbf{T}_g^n$ so that the quotient space $\mathbf{T}_g^n/\Gamma(L)$ coincides with $\mathbf{M}_g^n(L)$. According to the work of W. Harvey \cite{harvey}, $\mathbf{M}_g^n(L)$ is homotopy equivalent to a finite complex. \section{Proof of Theorem \ref{thm-torelli}} Fix an integer $L\geq 3$. Since $\mathfrak{S}_g$ is contractible, the projection $\mathfrak{S}_g\rightarrow\Gamma(L)\backslash\mathfrak{S}_g$ is the universal principal $\Gamma(L)$-bundle. The associated bundle \begin{equation}\label{borel-const} \mathbf{T}_g^n\rightarrow\mathfrak{S}_g\times_{\Gamma(L)}\mathbf{T}_g^n \rightarrow \Gamma(L)\backslash\mathfrak{S}_g. \end{equation} is nothing but the Borel construction of the $\Gamma(L)$-space $\mathbf{T}_g^n$. Since $\Gamma(L)$ acts freely on $\mathbf{T}_g^n$, the total space $\mathfrak{S}_g\times_{\Gamma(L)}\mathbf{T}_g^n$ is homotopy equivalent to $\mathbf{T}_g^n/\Gamma(L)=\mathbf{M}_g^n(L)$. Note that the associated bundle (\ref{borel-const}) is identified, up to homotopy, with the fibration $B\mathcal{I}_g^n\rightarrow B\mathcal{M}_g^n(L)\rightarrow B\Gamma(L)$ of classifying spaces induced from the exact sequence (\ref{mg-level-l}). Now suppose that $H_*(\mathcal{I}_g^n,\mathbb{Q})\cong H_*(\mathbf{T}_g^n,\mathbb{Q})$ is finite dimensional. As $\Gamma(L)\backslash\mathfrak{S}_g$ is homotopy equivalent to a finite complex, we may apply the following lemma to the associated bundle (\ref{borel-const}): \begin{lem}\label{lem-q-euler-char} Let $F\rightarrow E\rightarrow B$ be a fibration such that $B$ is a finite complex and $\dim_{\mathbb{Q}}H_*(F,\mathbb{Q})<\infty$. Then $\dim_{\mathbb{Q}}H_*(E,\mathbb{Q})<\infty$ and \[ \chi_{\mathbb{Q}}(E)=\chi_{\mathbb{Q}}(F)\cdot\chi(B), \] where $\chi_{\mathbb{Q}}$ is defined by $\chi_{\mathbb{Q}}(-)=\sum_i(-1)^i\dim_{\mathbb{Q}}H_i(-,\mathbb{Q})$. \end{lem}\noindent This lemma is a direct consequence of the Serre spectral sequence applied to the fibration $F\rightarrow E\rightarrow B$. As a result, one has \[ \chi(\mathbf{M}_g^n(L)) =\chi_{\mathbb{Q}}(\mathbf{T}_g^n)\cdot\chi(\Gamma(L)\backslash\mathfrak{S}_g). \] Since both of the projections $\Gamma(L)\backslash\mathfrak{S}_g\rightarrow Sp(2g,\mathbb{Z})\backslash\mathfrak{S}_g$ and $\mathbf{M}_g^n(L)\rightarrow\mathbf{M}_g^n$ are $|Sp(2g,\mathbb{Z}/L\mathbb{Z})|$-fold branched coverings, one has \begin{align*}\label{prod-formula} \chi(\Gamma(L)\backslash\mathfrak{S}_g) &=|Sp(2g,\mathbb{Z}/L\mathbb{Z})|\cdot e(Sp(2g,\mathbb{Z})\backslash\mathfrak{S}_g) \\ \chi(\mathbf{M}_g^n(L)) &=|Sp(2g,\mathbb{Z}/L\mathbb{Z})|\cdot e(\mathbf{M}_g^n), \end{align*} and hence \begin{equation}\label{prod-formula} e(\mathbf{M}_g^n)=\chi_{\mathbb{Q}}(\mathbf{T}_g^n)\cdot e(Sp(2g,\mathbb{Z})\backslash\mathfrak{S}_g), \end{equation} where $e$ denotes the orbifold Euler characteristics. According to G. Harder \cite{harder}, one has \[ e(Sp(2g,\mathbb{Z})\backslash\mathfrak{S}_g)=\prod_{k=1}^g\zeta(1-2k), \] while according to J. Harer and D. Zagier \cite{harer-zagier}, one has \[ e(\mathbf{M}_g^n)= \begin{cases}{\displaystyle\frac{1}{2-2g}\zeta(1-2g)}&\mbox{ if }n=0\\ {\displaystyle (-1)^{n-1}\frac{(2g+n-3)!}{(2g-2)!}\zeta(1-2g)} &\mbox{ if }n>0, \end{cases} \] where $\zeta$ is the Riemman $\zeta$-function. See also \cite{penner,kont}. Hence the equality (\ref{prod-formula}) leads to \[ \chi_{\mathbb{Q}}(\mathbf{T}_g^n)= \begin{cases} {\displaystyle\frac{1}{2-2g}\prod_{k=1}^{g-1} \frac{1}{\zeta(1-2k)}}&\mbox{ if }n=0\\ {\displaystyle (-1)^{n-1}\frac{(2g+n-3)!}{(2g-2)!} \prod_{k=1}^{g-1}\frac{1}{\zeta(1-2k)}}&\mbox{ if }n>0 \end{cases} \] By definition, $\chi_{\mathbb{Q}}(\mathbf{T}_g^n)$ must be an integer and the proof of Theorem \ref{thm-torelli} is then completed by virtue of the following lemma which will be proven in the next section. \begin{lem}\label{lem-zeta} For positive integers $m,n$, set \[ e(m,n)= \frac{(2m+n-1)!}{(2m)!} \prod_{k=1}^{m}\frac{1}{|\zeta(1-2k)|}. \] Then, for each $n\geq 1$, $e(m,n)$ is not an integer for sufficiently large $m$ compared with $n$. In particular, $e(m,1)$ is not an integer for all $m\geq 6$. \end{lem}\noindent \section{The proof of Lemma \ref{lem-zeta}} Recall that the Riemann $\zeta$-function is defined for ${\operatorname{Re}}\ s>1$ by \[ \zeta(s)=\sum_{n=1}^{\infty}\frac{1}{n^s}, \] On the other hand, the equality \[ \zeta(1-2k)=(-1)^k\frac{2\cdot (2k-1)!}{(2\pi)^{2k}}\zeta(2k). \] holds for any integer $k\geq 1$. It follows that \[ |\zeta(1-2k)| =\frac{2\cdot (2k-1)!}{(2\pi)^{2k}} \sum_{n=1}^{\infty}\frac{1}{n^{2k}} >\frac{2\cdot (2k-1)!}{(2\pi)^{2k}} \] for any integer $k\geq 1$, and hence \[ e(m,n)< \frac{(2m+n-1)!}{(2m)!}\prod_{k=1}^{m} \frac{(2\pi)^{2k}}{2\cdot (2k-1)!} \] for any integer $m\geq 1$. We claim that the right hand side of the inequality converges to $0$ as $m\rightarrow\infty$. Indeed, regarding the right hand side as a numerical sequence with respect to $m$, the ratio of the $(m+1)$-th term to the $m$-th term is given by \[ \frac{(2m+n+1)(2m+n)}{(2m+2)(2m+1)}\cdot \frac{(2\pi)^{2m+2}}{2\cdot (2m+1)!}. \] This converges to $0$ as $m\rightarrow\infty$, hence verifying the claim. We conclude that $e(m,n)<1$ and hence $e(m,n)$ is not an integer for $m$ sufficiently large compared with $n$. The first assertion is proved. To prove the second assertion, observe that ${(2\pi)^{2k}}/({2\cdot (2k-1)!})<1$ for $k\geq 9$. It follows that $e(m,1)$ is strictly decreasing with respect to $m$ for $m \geq 9$. On the other hand, with the help of a computer, one has \[ \prod_{k=1}^{14}\zeta(1-2k)=-297203.11\cdots . \] We see that, for all $m\geq 14$, $e(m,1)<1$ and hence $e(m,1)$ is not an integer. It remains to be proven that $e(m,1)$ is not an integer for $6\leq m\leq 13$. However, this can be verified by direct calculations and we omit the detail. \begin{remark} Actually, $e(m,n)$ is not an integer for $n<678$ and $m\geq 6$ and hence $H_*(\mathcal{I}_g^n,\mathbb{Q})$ is infinite dimensional for $n<678$ and $g\geq 7$. We describe briefly how this can be proven. Recall that $\zeta(1-2k)$ for positive integer $k$ is given by \[ \zeta(1-2k)=-\frac{B_{2k}}{2k}\in\mathbb{Q}, \] where $B_{2k}$ is the $2k$-th Bernoulli number defined as the coefficient of $z^{2k}/(2k)!$ in the power series expansion of $z/(e^z-1)$. On the other hand, von Staudt's theorem asserts that the denominator of $B_{2k}$ is not divisible by a prime $p$ if $2k<p-1$. By applying von Staudt's theorem to the primes 691 and 3617, the numerators of $B_{12}$ and $B_{16}$ respectively, we see that $e(m,n)$ is not an integer for $n<678$ and $6\leq m\leq1470$. Now the assertion follows from the inequality \[ e(m,n)<\frac{(2m+677)!}{(2m)!}\cdot\prod_{k=1}^m \frac{(2\pi)^{2m+2}}{2\cdot (2m+1)!}<1 \] which holds for $n<678$ and $m\geq 37$. \end{remark} \section{Proof of Theorem \ref{thm-kg}} To prove Theorem \ref{thm-kg}, we first recall some of Johnson's results concerning the Torelli groups. Suppose $g\geq 3$ and $[\Sigma_g]\in\wedge^2 H_1(\Sigma_g,\mathbb{Z})$ corresponds to the fundamental class of $\Sigma_g$. Under these conditions, Johnson constructed in \cite{johnson0} natural $Sp(2g,\mathbb{Z})$-equivariant surjective homomorphisms \[ \tau_{2,1}:\mathcal{I}_{g,1}\rightarrow \wedge^3 H_1(\Sigma_g,\mathbb{Z}) \] and \[ \tau_{2}:\mathcal{I}_g\rightarrow \wedge^3 H_1(\Sigma_g,\mathbb{Z})/([\Sigma_g]\wedge H_1(\Sigma_g,\mathbb{Z})) \] and proved in \cite{johnson2} that $\operatorname{ker}\tau_{2,1}=\mathcal{K}_{g,1}$ and $\operatorname{ker}\tau_{2}=\mathcal{K}_g$. The homomorphisms $\tau_{2,1}$ and $\tau_2$ are called {\em Johnson homomorphisms}. For simplicity, we abbreviate $\wedge^3 H_1(\Sigma_g,\mathbb{Z})/([\Sigma_g]\wedge H_1(\Sigma_g,\mathbb{Z}))$ by $\wedge^3 H/H$ and $\wedge^3 H_1(\Sigma_g,\mathbb{Z})$ by $\wedge^3 H$. As a consequence, $\mathcal{K}_g$ fits into the extension \[ 1\rightarrow\mathcal{K}_g\rightarrow\mathcal{I}_g\stackrel{\tau_2}{\rightarrow} \wedge^3 H/H\rightarrow 1 \] Take the classifying space of each group in the extension to yield a fibration \begin{equation}\label{fibre-kg} B\mathcal{K}_{g}\rightarrow B\mathcal{I}_{g}\rightarrow B(\wedge^3 H/H). \end{equation} Observe that $B\mathcal{I}_{g}$ is homotopy equivalent to $\mathbf{T}_{g}$ and $B(\wedge^3 H/H)$ is homotopy equivalent to the $\binom{2g}{3}-2g$-dimensional torus since $\wedge^3 H/H$ is a free abelian group of rank $\binom{2g}{3}-2g$. Now suppose $H_*(\mathcal{K}_g,\mathbb{Q})\cong H_*(B\mathcal{K}_g,\mathbb{Q})$ is finite dimensional. It follows from Lemma \ref{lem-q-euler-char} that $\dim_{\mathbb{Q}}H_*(\mathbf{T}_g,\mathbb{Q})<\infty$. If $\dim_{\mathbb{Q}}H_*(\mathbf{T}_g,\mathbb{Q})<\infty$ (and hence $g\leq 6$), then, as in the proof of Theorem \ref{thm-torelli}, $\chi_{\mathbb{Q}}(\mathbf{T}_g)$ is defined and satisfies \[ \chi_{\mathbb{Q}}(\mathbf{T}_g)=\frac{1}{2-2g}\prod_{k=1}^{g-1} \frac{1}{\zeta(1-2k)}\not=0. \] On the other hand, by applying Lemma \ref{lem-q-euler-char} to the fibration (\ref{fibre-kg}), one has $\chi_{\mathbb{Q}}(\mathbf{T}_g)=\chi_{\mathbb{Q}}(B\mathcal{K}_g)\cdot \chi(B(\wedge^3 H/H))=0$ since $\chi(B(\wedge^3 H/H))=0$. A contradiction. To prove Theorem \ref{thm-kg} for the group $\mathcal{K}_g^1$, we will identify $\mathcal{K}_g^1$ with the kernel of a variant of the Johnson homomorphism. Recall that Torelli groups $\mathcal{I}_g^1$ and $\mathcal{I}_{g,1}$ fit into the central extension \[ 1\rightarrow\mathbb{Z}\rightarrow\mathcal{I}_{g,1}\rightarrow\mathcal{I}_{g}^1\rightarrow 1, \] where the center $\mathbb{Z}$ is generated by the Dehn twist $\xi$ along a simple closed curve parallel to the boundary of a fixed embedded disk $D\subset\Sigma_g$. Now the Dehn twist $\xi$ is contained in $\mathcal{K}_{g,1}$ and hence $\tau_{2,1}$ induces a homomorphism $\tau_2^1:\mathcal{I}_g^1\rightarrow \wedge^3 H$. We claim that $\operatorname{ker}\tau_2^1=\mathcal{K}_{g}^1$. Indeed, $\operatorname{ker}\tau_2^1$ coincides with the image of $\mathcal{K}_{g,1}$ under the homomorphism $\mathcal{I}_{g,1}\rightarrow\mathcal{I}_g^1$. But the image of $\mathcal{K}_{g,1}$ is nothing but $\mathcal{K}_g^1$ since any Dehn twist along separating simple closed curve is isotopic to that which fixes the embedded disk pointwise. In summary, we have the extension \[ 1\rightarrow\mathcal{K}_{g}^1\rightarrow\mathcal{I}_{g}^1 \stackrel{\tau_2^1}{\rightarrow}\wedge^3 H\rightarrow 1. \] Take the classifying space for each group in the extension to yield a fibration \[ B\mathcal{K}_{g}^1\rightarrow B\mathcal{I}_{g}^1\rightarrow B(\wedge^3 H). \] Now $B\mathcal{I}_{g}^1$ is homotopy equivalent to $\mathbf{T}_{g}^1$ and $B(\wedge^3 H)$ is homotopy equivalent to the $\binom{2g}{3}$-dimensional torus since $\wedge^3 H$ is a free abelian group of rank $\binom{2g}{3}$. The rest of the proof is similar to that of $\mathcal{K}_g$. \vspace{3mm}\\ {\em Acknowledgement.} The author thanks to Professor Shigeyuki Morita for calling author's attention to the groups $\mathcal{K}_{g,r}^n$. \ifx\undefined\leavevmode\hbox to3em{\hrulefill}\, \newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,} \fi
1997-12-30T17:24:30
9712
alg-geom/9712034
en
https://arxiv.org/abs/alg-geom/9712034
[ "alg-geom", "hep-th", "math.AG" ]
alg-geom/9712034
null
Victor V. Batyrev
Toric Degenerations of Fano Varieties and Constructing Mirror Manifolds
13 pages, AMS-Latex. This is an extended version of the author's talk given during the Summer Symposium on Algebra at University of Niigata, July 22-25, 1997
null
null
null
null
For an arbitrary smooth n-dimensional Fano variety $X$ we introduce the notion of a small toric degeneration. Using small toric degenerations of Fano n-folds $X$, we propose a general method for constructing mirrors of Calabi-Yau complete intersections in $X$. Our mirror construction is based on a generalized monomial-divisor mirror correspondence which can be used for computing Gromov-Witten invariants of rational curves via specializations of GKZ-hypergeometric series.
[ { "version": "v1", "created": "Tue, 30 Dec 1997 16:24:30 GMT" } ]
2007-05-23T00:00:00
[ [ "Batyrev", "Victor V.", "" ] ]
alg-geom
\section{Introduction} Recent progress in understanding the mirror symmetry phenomenon using explicit mirror constructions for Calabi-Yau hypersurfaces and complete intersections in toric varieties \cite{BA,BS,BB0,Bo} leads to the following natural question: \medskip {\em Is it possible to extend the mirror constructions for Calabi-Yau complete intersections in toric Fano varieties to the case of Calabi-Yau complete intersections in nontoric Fano varieties? } \medskip The first progress in this direction has been obtained for Grassmannians \cite{BCKS1} and, more generally, for partial flag manifolds \cite{BCKS2}. The key idea in both examples is based on a degeneration of Grassmannians (resp. partial flag manifolds) to some singular Gorenstein toric Fano varieties. These degenerations have been introduced and investigated by Sturmfels, Gonciulea and Lakshmibai in \cite{GL1,GL2,L,S1,S2}. The present paper is aimed to give a short systematic overview of our method for constructing mirror manifolds and to formulate some naturally arising questions and open problems. In Section 2 we start with a review of a method for constructing degenerations of unirational varieties $X$ to toric varieties $Y$ using canonical subalgebra bases. This method has been discovered by Kapur \& Madlener \cite{KM} and independently by Robbiano \& Sweedler \cite{RS}. Further results on this topic have been obtained in \cite{O,M,S1} (see also \cite{S2} for more details). In Section 3 we introduce the notion of a {\em small toric degeneration} of a Fano manifold and discuss some examples. Finally, in Section 4 we explain our generalized mirror construction which uses small toric degenerations. \section{Canonical subalgebra bases} Let $A$ be a finitely generated subalgebra of the polynomial ring $$K[{\bf u}]:=K[u_1, \ldots, u_n],$$ i.e., $X= Spec\, A$ is an unirational affine algebraic variety together with a dominant morhism ${\Bbb A}^n \to X$. We choose a weight vector $\omega = (\omega_1, \ldots, \omega_d) \in {\Bbb R}^n$ and set $$wt({\bf u}^{\bf a}) = wt(u_1^{a_1} \cdots u_n^{a_n}) := \sum_{i=1}^n a_i \omega_i.$$ The number $wt({\bf u}^{\bf a})$ will be called the {\bf weight} of the monomial ${\bf u}^{\bf a}$. We define a partial order on the set of all monomials in $K[{\bf u}]$ as follows: \[ {\bf u}^{\bf a} \prec {\bf u}^{\bf a'} \Leftrightarrow wt({\bf u}^{\bf a}) \leq wt({\bf u}^{\bf a'}). \] If $f \in K[{\bf u}]$ is a polynomial, then $in_{\prec}(f)$ denotes the {\bf initial part of $f$}, i.e., the sum of those monomials in $f$ whose weight is maximal. By definition, one has $in_{\prec}(fg) = in_{\prec}(f) in_{\prec}(g)$. For suficiently general choice of the weight vector $\omega \in {\Bbb R}^n$ the initial part of a polynomial $f \in K[{\bf u}]$ is a single monomial. \begin{dfn} {\rm The $K$-vector space spanned by initial terms of elements $f \in A$ is called the {\bf initial algebra} and is denoted by \[ in_{\prec}(A) : = \{ in_{\prec}(f)\; : \; f \in A \}. \]} \end{dfn} \begin{dfn} {\rm A subset ${\cal F} \subset A $ is called a {\bf canonical basis of the subalgebra} $A \subset K[{\bf u}]$, if the initial subalgebra $in_{\prec}(A)$ is generated by the elements $$\{ in_{\prec}(f)\; : \; f \in {\cal F} \}.$$ } \end{dfn} Fix a set of polynomials ${\cal F} = \{f_1, \ldots, f_m\} \subset A$. We set $K[{\bf v}]:=K[v_1, \ldots, v_m]$. Let $I$ be the kernel of the canonical epimorphism $$\varphi \; : \; K[{\bf v}] \to A$$ $$v_i \mapsto f_i$$ and $I_{\prec}$ the kernel of the canonical epimorphism $$\varphi_0 \; : \; K[{\bf v}] \to in_{\prec}(A)$$ $$v_i \mapsto in_{\prec}(f_i)$$ \begin{rem} {\rm It is easy to show that the ideal $I_{\prec}$ is generated by binomials (see \cite{ES} for general theory of binomial ideals). Hence, the spectrum of $in_{\prec}(A)$ is an affine toric variety (possibly not normal).} \end{rem} Now we assume that $\omega = (\omega_1, \ldots, \omega_d) \in {\Bbb Z}^n$ an integral weight vector. If the set of polynomials ${\cal F} = \{f_1, \ldots, f_m\} \subset A$ form a canonical basis of the subalgebra $A \subset K[{\bf u}]$ with respect to the partial order defined by $\omega$, then we can define a $1$-parameter family of subalgebras \[ A_t := \{ f(t^{-\omega_1}u_1, \ldots, t^{-\omega_n} u_n) \;\; |\; \; f(u_1, \ldots, u_n) \in A \},\; \;\; t \in K \setminus \{ 0 \} \}. \] Setting $A_0: = in_{\prec}(A)$, we obtain a flat family of subalgebras $A_t \subset K[{\bf u}]$ such that $A_t \cong A$ for $ t\neq 0$ and $A_0 \cong K[{\bf v}]/I_{\prec}$. This allows us to consider the affine toric variety $Spec\, A_0$ as a flat degeneration of $Spec\, A$. \begin{rem} {\rm It is important to remark that the above method for constructing toric degenerations strongly depends on the choice of the coordinates $u_1, \ldots, u_n$ on ${\Bbb A}^n$ and on the choice of a weight vector $\omega$.} \end{rem} \begin{exam} {\rm Let $A(r,s) \subset K[{\bf X}]: = K[X_{ij}]$ $(1 \leq i \leq r, \; 1 \leq j \leq s)$ be the subalgebra of the polynomial algebra $K[{\bf X}]$ generated by all $r \times r$ minors of a generic $r \times s$ matrix $( r \leq s)$, i.e., $A(r,s)$ is the homogeneous cooordinate ring of the Pl\"ucker embedded Grassmannian $G(r,s) \subset {\Bbb P}^{ { s \choose r } -1}$. Define the weights of monomials as follows \[ wt(X_{ij}) := (j-1) s^{i-1}, \;\; i,j \geq 1.\] In particular, one has \[ wt(X_{1,i_1} \cdots X_{r,i_r}) = (i_1-1) + (i_2-1)s + \cdots + (i_r-1)s^{r-1} \] and therefore the initial term of each $(i_1, \ldots, i_r)$-minor $(1 \leq i_1 < \cdots < i_r \leq s)$ is exactly the product of terms on the main diagonal: \[X_{1,i_1} \cdots X_{r,i_r}. \] The following result is due to Sturmfels \cite{S1,S2}: \begin{theo} The set of all $s \times s$-minors form a canonical base of the subalgebra $A(r,s) \subset K[{\bf X}]$ with respect to the partial order defined by the above weight vector. In particuar, one obtains a natural toric degeneration of the Grassmanninan $G(r,s)$. \end{theo} \label{grass} } \end{exam} \section{Small toric degenerations of Fano varieties} \begin{dfn} {\rm Let $X \subset {\Bbb P}^m$ be a smooth Fano variety of dimension $n$. A normal Gorenstein toric Fano variety $Y \subset {\Bbb P}^m$ is called a {\bf small toric degeneration} of $X$, if there exists a Zariski open neighbourhood $U$ of $0 \in {\Bbb A}^1$ and an irreducible subvariety ${\frak X} \subset {\Bbb P}^m \times U$ such that the morphism $\pi\; : \; {\frak X} \to U$ is flat and the following conditions hold: {(i)} the fiber $X_t := \pi^{-1}(t) \subset {\Bbb P}^m$ is smooth for all $t \in U \setminus \{ 0 \}$; {(ii)} the special fiber $X_0 := \pi^{-1}(0) \subset {\Bbb P}^m$ has at worst Gorenstein terminal singularities (see \cite{KMM}) and $X_0$ is isomorphic to $Y \subset {\Bbb P}^m$; {(iii)} the canonical homomorphism \[ Pic({\frak X}/U) \to Pic(X_t) \] is an isomorphism for all $t \in U$. } \label{def-small} \end{dfn} \begin{rem} {\rm It is weill-known that if $Y$ has at worst terminal singularities, then the codimension of the singular locus of $Y$ is at least $3$. On the other hand, it is easy to show that the only possible toric Gorenstein terminal singularities in dimension $3$ are ordinary double points (or nodes): $x_1x_2 - x_3x_4=0$. So, if $Y$ is a small toric degeneration of $X$, then the singular locus of $Y$ in codimension $3$ must consist of nodes.} \label{codim3} \end{rem} \begin{exam} {\rm Let $Y:= P(r,s) \subset {\Bbb P}^{ { s \choose r } -1}$ be the toric degeneration of the Grassmannian $X:= Gr(r,s) \subset {\Bbb P}^{ { s \choose r } -1}$ (see Example \ref{grass}). Then $Y$ is a small toric degeneration of $X$ \cite{BCKS1}. } \end{exam} \begin{exam} {\rm Let $X:= F(n_1, \ldots,n_k ,n) \subset {\Bbb P}^{m}$ be the partial flag manifold it is Pl\"ucker embedding. It is proved in \cite{BCKS2} that the toric degenerations introduced and investigated by Gonciulea and Lakshmibai in \cite{GL1,GL2,L} are small toric degenerations of $X$. } \end{exam} \begin{exam} {\rm Let $V_{d,n} \subset {\Bbb P}^{n+1}$ be a Gorenstein toric Fano hypersurface of degree $d$ $(d \geq 2)$ in projective space of dimension $n \geq 2d -2$ defined by the homogeneous equation \[ z_1 \cdots z_d = z_{d+1} \cdots z_{2d}. \] It is easy to check that irreducible components of the singular locus of $V_{d,n}$ are \[ \frac{d^2(d-1)^2}{4} \] codimension-3 linear subspaces \[ z_i=z_j=z_k =z_l =0, \] \[ \; \; \{i,j \} \subset \{1, \ldots, d\},\; \{k,l \} \subset \{d+1, \ldots, 2d \}, \;i \neq j, \; k \neq l. \] consisting of nodes.} \end{exam} \begin{theo} $V_{d,n} \subset {\Bbb P}^{n+1}$ is a small toric degeneration of a smooth Fano hypersurface $X_{d,n} \subset {\Bbb P}^n$ of degree $d$. \label{sm-hyp} \end{theo} \noindent {\em Proof.} Let us first consider the case $n = 2d-2$. In this case the $2(d-1)$-dimensional fan $\Sigma_d$ defining the toric variety $V_{d,2(d-1)}$ can be constructed as follows: Let $e_1, \ldots, e_{d-1}, f_1, \ldots, f_{d-1}$ be a ${\Bbb Z}$-basis of the lattice ${\Bbb Z}^{2(d-1)}$. We set $e_{d} := -e_1 - \cdots - e_{d-1}$ and $f_{d} := -e_1 - \cdots - f_{d-1}$. We denote by $h_{i,j}$ the sum $e_i + f_j$ ($i, j \in \{1, \ldots, n\})$. If $\Delta_d^*$ denotes the convex hull of $d^2$ points $h_{i,j}$, then the fan $\Sigma_d \subset N_{{\Bbb R}}$ consists of cones over faces of the reflexive polyhedron $\Delta_d^*$, where the integral lattice $N \subset {\Bbb Z}^{2(d-1)}$ is generated by all $d^2$ lattice vectors $h_{i,j}$ (the sublattice $N \subset {\Bbb Z}^{2(d-1)}$ coincides with $ {\Bbb Z}^{2(d-1)}$ unless $d =2$). Using the combinatorial characterisations of terminal toric singularities \cite{KMM}, one immediately obtains that all singularities of $V_{d,2(d-1)}$ are terminal, since the only $N$-lattice points on the faces of $\Delta_d^*$ are their vertices. If $d \geq 3$, then the Picard group of $V_{d,2(d-1)}$ is generated by the class of the hyperplane section, i.e., $Pic(V_{d,2(d-1)})\cong {\Bbb Z}$ and the anticanonical class of $V_{d,2(d-1)}$ is $d$-th multiple of the generator of $Pic(V_{d,2(d-1)})$. The latter can be show as follows: Consider a $(2d-3)$-dimensional face of $\Delta_d^*$ having vertices \[ h_{i,j}, \;\; i \in \{1, \ldots, d-1\}, \; j \in \{1, \ldots, d\}. \] Then every $\Sigma_d$-piecewise linear function $\varphi\, : \, N_{{\Bbb R}} \to {\Bbb R}$, up to summing a linear function, can be normalized by the condition \[ \varphi(h_{i,j}) = 0, \;\; \forall i \in \{1, \ldots, d-1\}, \; \forall j \in \{1, \ldots, d\}. \] On the other hand, for any $j \neq j'$, $j, j' \in \{1, \ldots, d\}$ four lattice points \[ h_{d,j}, h_{1,j}, h_{d,j'}, h_{1,j'} \] generate a $3$-dimensional cone in $\Sigma_d$. Hence \[ \varphi(h_{d,j}) = \varphi(h_{d,j'}) \;\; \forall j, j' \in \{1, \ldots, d\}. \] This means that the space of all $\Sigma_d$-piecewise linear functions modulo linear functions is $1$-dimensional. The anticanonical class is represented by the $\Sigma_d$-piecewise linear function $\varphi_1$ taking values $1$ on each vector $h_{i,j}$ $i,j \in \{1, \ldots, d\}$. Considering the difference \[ \varphi'_1 := \varphi_1 - \lambda, \] where $\lambda$ is a linear function on $N_{{\Bbb R}}$ satisfying the conditions \[ \lambda(e_1) = \cdots =\lambda(e_{d-1}) =1, \; \lambda(e_d) = -(d-1), \; \lambda(f_1) = \cdots = \lambda(f_d) = 0, \] we obtain a $\Sigma_d$-piecewise linear function having the properties \[ \varphi_1'(h_{i,j}) = 0, \;\; \forall i \in \{1, \ldots, d-1\}, \; \forall j \in \{1, \ldots, d\} \] and \[ \varphi(h_{d,j}) = d\;\; \forall j \in \{1, \ldots, d\}. \] So the class of $\varphi_1$ modulo linear functions is a $d$-th multiple of a generator of $Pic(V_{d,2(d-1)})$. The general case $n > 2(d-1)$ can be obtained by similar arguments using the fact that $V_{d,n}$ is a projective cone over $V_{d,2(d-1)}$. In order to construct the required flat $1$-parameter family ${\frak X}$ (cf. \ref{def-small}), it suffices to consider a pencil of hypersurfaces of degree $d$ in ${\Bbb P}^{n+1}$ joining $X_{d,n}$ and $V_{d,n}$. \hfill $\Box$ \begin{theo} Let $X_d \subset {\Bbb P}^{n+1}$ be a smooth Fano hypersurface of degree $d$. Then $X_d$ admits a small toric degeneration if and only if $n \geq 2d -2$. \label{hypersur} \end{theo} \noindent {\em Proof.} By \ref{sm-hyp}, it suffices to show that $X_d$ does not admit a small toric degeneration if $n< 2d-2$. Assume that $X_d$ admits a small toric degeneration $Y_d$. Then $Y_d$ is a toric hypersurface defined by a binomial equation $M_1 =M_2$ where $M_1$ and $M_2$ are monomials in $z_0, \ldots, z_{n+1}$ of degree $d$ $(z_0, \ldots, z_{n+1}$ are homogeneous coordinates on ${\Bbb P}^{n+1}$). If $n < 2d-2$, then at least one of the monomials $M_1$ and $M_2$ must be divisible by $z_i^2$ for some $i \in \{0, \ldots, n+1\}$. We can assume that for instance $z_0^2$ divides $M_1$. If $z_k$ and $z_l$ are two variables appearing in $M_2$, then $n-2$-dimensional linear subspace \[ z_0 = z_k = z_l = 0 \] is contained in $Sing(Y_d)$. This contradicts the fact that terminal singularities on $Y_d$ could appear only in codimension $\geq 3$ (see \ref{codim3}). \hfill $\Box$ Using \ref{codim3}, one immediately obtains: \begin{prop} If $X$ is a smooth Del Pezzo surface, then $X$ admits a small toric degeneration if and only if $X$ is itself a toric variety (i.e. $K_X^2 \geq 6$). \end{prop} As we have seen from \ref{sm-hyp}, a smooth quadric $3$-fold in ${\Bbb P}^4$ is an example of nontoric smooth Fano variety which admits a small toric degeneration. By \ref{hypersur}, cubic and quartic $3$-folds do not admit small toric degenerations. The compltete classification of smooth Fano $3$-folds has been obtained in \cite{C,I,MM1,MM2,MU}. It is natural to ask the following: \begin{ques} Which $3$-dimensional nontoric smooth Fano varieties do admit small toric degenerations? \end{ques} \section{The mirror construction} For our convenience, we assume $K= {\Bbb C}$. Let $X$ be a smooth Fano $n$-fold over ${\Bbb C}$ and $Y$ is its small toric degeneration. The toric variety $Y$ is defined by some complete rational polyhedral fan $\Sigma \subset N_{{\Bbb R}}$, where $N_{{\Bbb R}} = N \otimes {\Bbb R}$ is the real scalar extension of a $N \cong {\Bbb Z}^n$. We denote by $Cl(Y)$ (resp. by $Pic(Y)$) the group of Weil (resp. Cartier) divisors on $Y$ modulo the rational equivalence. One has a canonical embedding $$\alpha\; : \; Pic(Y) \hookrightarrow Cl(Y).$$ If $\{ e_1, \ldots, e_k \} \subset N$ is the set of integral generators of $1$-dimensional cones in $\Sigma$, then $Cl(Y)$ is a finitely generated abelian group of rank $k-n$ and the convex hull of $ e_1, \ldots, e_k$ is a reflexive polyhedron $\Delta^*$ (for definition of reflexive polyhedra see \cite{BA}). Assume that there exists a partition of the set $I = \{ e_1, \ldots, e_k \}$ into $r$ disjoint subsets $J_1, \ldots, J_r$ such that the union $D_i$ of toric strata in $Y$ corresponding to elements of $J_i$ is a semiample Cartier divisor on $Y$ for each $i \in \{1, \ldots, r\}$. Denote by $Z \subset Y$ a Calabi-Yau complete intersection of $r$ hypersurfaces $Z_i \subset Y$ defined by vanishing of generic global sections of ${\cal O}_Y(D_i)$. By \cite{BS} (see also \cite{Bo}), the mirrors $Z^*$ of Calabi-Yau complete intersections $Z \subset Y$ are birationally isomorphic to affine complete intersections in $({\Bbb C}^*)^n = Spec\, {\Bbb C}[t_1^{\pm 1}, \ldots, t_n^{\pm 1}]$ defined by $r$ equations \[ 1 = \sum_{e_j \in J_i}^k a_j{\bf t}^{e_j}, \;\; i \in \{1, \ldots, r\},\] where $(a_1, \ldots, a_k) \in {\Bbb C}^k$ is a general complex vector and ${\bf t}^{e_1}, \ldots, {\bf t}^{e_k}$ are Laurent monomials in variables $t_1, \ldots, t_n$ with the exponents $e_1, \ldots, e_k$. \begin{dfn} {\rm A complex vector $(a_1, \ldots, a_k ) \in {\Bbb C}^k$ is called $\Sigma$-{\bf admissible}, if there exists a $\Sigma$-piecewise linear function \[ \varphi\; : \; N_{{\Bbb R}} \to {\Bbb R}, \] (i.e., a continuous function such that $\varphi|_{\sigma}$ is linear for every $\sigma \in \Sigma$) having the property \[ \varphi(e_i) = \log|a_i|,\;\;\forall i \in \{1, \ldots, k\}. \] The set of all $\Sigma$-admissible vectors will be denoted by $A(\Sigma)$. } \end{dfn} \begin{rem} {\rm It is easy to show that $A(\Sigma) \subset {\Bbb C}^k$ is an irreducible closed subvariety which is isomorphic to an affine toric variety of dimension $rk\, Pic(Y) + n \leq k$.} \end{rem} Now our generalization of the mirror construction from \cite{BS} to the case of Calabi-Yau complete intersections in a nontoric Fano variety $X$ can be formulated as follows: \medskip \noindent {\bf Generalized mirror construction:} {\em Mirrors $W^*$ of generic Calabi-Yau hypersurfaces $W \subset X$ are birationally isomorphic to the affine complete intersections \[ 1 = \sum_{i=1}^k a_i {\bf t}^{e_i}, \] where ${\bf a}:= (a_1, \ldots, a_k)$ is a general point of $A(\Sigma)$.} \medskip \noindent {\bf Monomial-divisor correspondence:} Let us explain the monomial-divisor mirror correspondence for this mirror construction (cf. \cite{AGM}). By \ref{def-small}(iii), the group $Pic(Y)$ can be canonically identified with $Pic(X)$. The image of the restriction homomorphism $Pic(X) \to Pic(W)$ defines a subgroup $G \subset Pic(W)$, whose elements correspond to monomial deformations of the complex structure on mirrors: {\em if $\psi$ is an integral $\Sigma$-piecewise linear function representing an element $\gamma \in G$, then the $1$-parameter family of hypersurfaces \[ 1 = \sum_{i=1}^k t_0^{\varphi(e_i)} {\bf t}^{e_i},\;\; t_0 \in {\Bbb C} \] defines the corresponding $1$-parameter deformation of the complex structure on $W^*$ via the deformation of the coefficients $a_i = t_0^{\varphi(e_i)}$.} \medskip \noindent {\bf The main period:} Let $R(\Sigma)$ the group of all vectors $(l_1, \ldots, l_k) \in {{\Bbb Z}}^k$ satisfying the condition $\sum_{i =1}^k l_i e_i = 0$ and $L(\Sigma) \subset R(\Sigma)$ be the semigroup consisiting of vectors $(l_1, \ldots, l_k) \in R(\Sigma)$ with nonnegative coordinates $l_i$ $(i =1, \ldots, k)$. There exists a canonical pairing $ \langle *, * \rangle \; : \; R(\Sigma) \times Pic(Y) \to {\Bbb Z} $ which is the intersection pairing between $1$-dimensional cycles and Cartier divisors on $Y$. According to \cite{BS}, we can compute the main period in the family of mirrors $W^*$ in our generalized mirror construction as follows \[ \Phi_0({\bf a}) = \sum_{ {\bf l}= (l_1, \ldots, l_k) \in L(\Sigma)} \frac{ \langle l, D_1 + \cdots + D_r \rangle!}{ \langle l, D_1\rangle! \cdots \langle l, D_r\rangle !} \prod_{i=1}^k a_i^{l_i}, \; \; {\bf a} \in A(\Sigma). \] The condition ${\bf a} \in A(\Sigma)$ can be interpreted as a specialization of $GKZ$-hypergeometric series from \cite{BS}. \medskip Some evidences in favor of our generalized mirror construction were presented in \cite{BCKS1,BCKS2}. For our next examples confirming the proposed generalized mirror construction we use the following simple combinatorial statement: \begin{prop} Let $S_d(m)$ be the set of all $d \times d$-matrices $K = (k_{ij})$ with nonnegative integral coefficients $k_{ij}$ satisfying the equations \[ \begin{pmatrix} 1 & \cdots & 1 \end{pmatrix} \begin{pmatrix} k_{11} & \cdots & k_{1d} \\ \cdot & \cdots & \cdot \\ \cdot & \cdots & \cdot \\ \cdot & \cdots & \cdot \\ k_{d1} & \cdots & k_{dd} \end{pmatrix} = \begin{pmatrix} m & \cdots & m \end{pmatrix} \] and \[ \begin{pmatrix} k_{11} & \cdots & k_{1d} \\ \cdot & \cdots & \cdot \\ \cdot & \cdots & \cdot \\ \cdot & \cdots & \cdot \\ k_{d1} & \cdots & k_{dd} \end{pmatrix} \begin{pmatrix} 1 \\ \cdot \\ \cdot \\ \cdot \\ 1 \end{pmatrix}= \begin{pmatrix} m \\ \cdot \\ \cdot \\ \cdot \\ m \end{pmatrix}. \] Then \[ \sum_{K \in S_d(m)} \frac{(m!)^d}{\prod_{i,j =1}^{d} (k_{ij})!} = \frac{(dm)!}{(m!)^d}. \] \label{comb-f} \end{prop} \noindent {\em Proof.} Let $A$ be the set $\{1, 2, \ldots, dm \}$ of first $dm$ natural numbers. We fix a splitting $A$ into the disjoint union of $d$ subsets \[ A_i := \{ (i-1)m +1, (i-1)m +2, \ldots, im \}, \;\; i =1, \ldots, d \] consising of $m$ elements. Let $\beta : A = B_1 \cup \cdots \cup B_d$ be an arbitrary representation of $A$ as a disjoint union of the subsets $B_1, \ldots, B_d$ with the property $|B_1| = \cdots = |B_d| =m$. Then every such a representation defines a matrix $K(\beta) =(k_{ij}(\beta)) \in S_d(m)$ as follows: \[ k_{ij}(\beta) := |A_i \cap B_j|, \;\; i, j \in \{1, \ldots, d\}. \] For a fixed matrix $K \in S_d(m)$ there exist exactly \[ \prod_{j=1}^d \frac{(m!)}{\prod_{i =1}^{d} (k_{ij})!} \] ways to construct a representation $\beta$ of $A$ as a dusjoint union of $m$-element subsets $B_1, \ldots, B_d$ such that $K = K(\beta)$. Therefore, \[ \sum_{K \in S_d(m)} \frac{(m!)^d}{\prod_{i,j =1}^{d} (k_{ij})!} \] is the total number of ways to split $A$ into a disjoint union of $m$-element subsets $B_1, \ldots, B_d$. On the other hand, this number is equal to the multinomial \[ \frac{(dm)!}{(m!)^d} .\] \hfill $\Box$ \begin{exam} {\em Let $W$ be a generic Calabi-Yau complete intersection of two hypersurfaces $V_d, V_d'$ in ${\Bbb P}^{2d-1}$. By \ref{sm-hyp}, we can construct a small toric degeneration of one smooth hypersurface $V_d'$ to the $2(d-1)$-dimensional toric variety $Y_d \subset {\Bbb P}^{2d-1}$ \[ z_0z_1 \cdots z_{d-1} = z_d z_{d+1} \cdots z_{2d-1}. \] Using an explicit description of the Picard group $Pic(Y_d)$ from the proof of \ref{sm-hyp}, our generalized mirror construction suggests that mirrors $W^*$ for $W$ are birationally isomorphic to the affine hypersurfaces $Z_F$ in the algebraic torus $$Spec\, {\Bbb C}[ t_1^{\pm1},\ldots, t_{d-1}^{\pm1}, u_1^{\pm1},\ldots, u_{d-1}^{\pm1}]$$ defined by the $1$-parameter family of the equations \[ 1 = F(t_1,\ldots, t_{d-1},u_1, \ldots, u_{d-1},z) = \sum_{i=1}^{d-1} \sum_{j=1}^{d-1}t_iu_j + (u_1 \cdots u_{d-1})^{-1} \left( \sum_{i=1}^{d-1} t_i \right) + \] \[ + z(t_1 \cdots t_{d-1})^{-1} \left( (u_1 \cdots u_{d-1})^{-1} + \sum_{i=1}^{d-1} u_j \right) ,\; \;\;\; z \in {\Bbb C} \] On the other hand, it is known via a toric mirror construction for Calabi-Yau complete intersection $W = V_d \cap V_d'$ (see \cite{BS}) that the power series \[ \Phi_0(z) = \sum_{m \geq 0} \frac{(dm!)^2}{(m!)^{2d}} z^m \] generates the Picard-Fuchs $D$-module discribing the quantum differential system. Now we compare our generalized mirror construction with the known one from \cite{BS} computing the main period of the family $Z_F$ by the Cauchy residue formula: \[ \Psi_F(z) := \frac{1}{(2\pi\sqrt{-1})^{2(d-1)}} \int_{\Gamma} \frac{1}{1 - F({\bf t}, {\bf u}, z)} \frac{{\bf dt}}{{\bf t}} \wedge \frac{{\bf du}}{\bf u} = 1 + a_1z + a_2z^2 + \cdots, \] \[ \frac{{\bf dt}}{{\bf t}}:= \frac{dt_1}{t_1} \wedge \cdots \wedge \frac{dt_{d-1}}{t_{d-1}}, \; \; \; \frac{{\bf du}}{{\bf u}}:= \frac{du_1}{u_1} \wedge \cdots \wedge \frac{du_{d-1}}{u_{d-1}}, \] where the coefficients $a_m$ of the power series $\Psi_F(z)$ can be computed by the formula \[ a_m = \sum_{K \in S_d(m)} \frac{(dm)!}{\prod_{i,j =1}^{d} (k_{ij})!}. \] Using \ref{comb-f}, we obtain that \[ a_m = \frac{(dm!)^2}{(m!)^{2d}}, \] i.e., the power series $\Psi_F(z)$ coincides with $\Phi_0(z)$ and therefore our generalized mirror construction agrees with the already known one from \cite{BS}. For the special case $d=3$, we obtain a description for mirrors $W^*$ of complete intersections $W$ of two cubics in ${\Bbb P}^5$ as smooth compactifications of hypersurfaces in the $4$-dimensional algebraic torus $$Spec\, {\Bbb C}[ t_1^{\pm1},t_2^{\pm1}, u_1^{\pm1},u_2^{\pm1}]$$ defined by the $1$-parameter family of the equations \[ 1 = F(t_1,t_2,u_1,u_2, \lambda) = t_1u_1 + t_1u_2 + t_1(u_1u_2)^{-1} + t_2u_1 + t_2u_2 + t_2(u_1u_2)^{-1} + \] \[ + z(t_1t_2)^{-1}(u_1 + u_2 + (u_1u_2)^{-1}),\; \;\;\; z \in {\Bbb C}. \] This discription of mirrors is different from the one proposed by Libgober and Teilelbaum in \cite{LT}, but it seems that both constructions are equivalent to each other.} \end{exam} Now we want to suggest some problem which naturally arise from the proposed generalized mirror construction. \begin{prob} Check the topological mirror duality test \[ E_{\rm st}(W^*; u,v) = (-u)^n E_{\rm st}(W; u^{-1},v) \] for the above generalized mirror construction. Here $E_{\rm st}$ is the stringy $E$-function introduced in \cite{B1}. \end{prob} \begin{rem} {\rm The main difficulty of this checking arises from the fact that the affine complete intersections in the above mirror construction are not {\em generic}. For $\Delta^*$-regular affine hypersurfaces there exists explicit combinatorial formula for their $E$-polynomials (see \cite{BB1}). However, the affine hypersurfaces in our mirror construction are not $\Delta^*$-regular and no explicit formula for their $E$-polynomials (or Hodge-Deligne numbers) is known so far. } \end{rem} \begin{prob} Generalize the method of Givental \cite{G1,G2,G3} for computing Gromov-Witten invariants of complete intersections in smooth Fano varieties $X$ admitting small toric degenerations. \end{prob} \begin{rem} {\rm If $X$ is a smooth Fano $n$-fold admitting a small toric degeneration $Y$, then one can not expect that there exists a ${\Bbb C}^*$-action on $X$. So the equivariant arguments from \cite{G1} can not be applied directly to $X$. However, one could try to use equivariant Gromov-Witten theory for the ambient projective space ${\Bbb P}^m$ containing both $X$ and $Y$ and to show that the virtual fundamental classes corresponding to $Y$ and $X$ are the same. It seems that small quantum cohomology of $Y$ carry complete information about the subring in the small quantum cohomology ring $QH^*(X)$ generated by the classes of divisors. This would give an explicit description of such a subring (see \cite{ST}) as well as of its gravitational version via Lax operators (see \cite{EHX}). } \end{rem}
1997-12-19T13:34:51
9712
alg-geom/9712024
en
https://arxiv.org/abs/alg-geom/9712024
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9712024
Maxim Braverman
Maxim Braverman (Hebrew University)
Symplectic cutting of Kaehler manifolds
11 pages, LaTeX 2e
null
null
Warwick preprint
null
We obtain estimates on the character of the cohomology of an $S^1$-equivariant holomorphic vector bundle over a Kaehler manifold $M$ in terms of the cohomology of the Lerman symplectic cuts and the symplectic reduction of $M$. In particular, we prove and extend inequalities conjectured by Wu and Zhang. The proof is based on constructing a flat family of complex spaces $M_t$ such that $M_t$ is isomorphic to $M$ for $t\not=0$, while $M_0$ is a singular reducible complex space, whose irreducible components are the Lerman symplectic cuts.
[ { "version": "v1", "created": "Fri, 19 Dec 1997 12:34:50 GMT" } ]
2016-08-30T00:00:00
[ [ "Braverman", "Maxim", "", "Hebrew University" ] ]
alg-geom
\section{#2}\label{S:#1}% \ifShowLabels \TeXref{{S:#1}} \fi} \newcommand{\ssec}[2]{\subsection{#2}\label{SS:#1}% \ifShowLabels \TeXref{{SS:#1}} \fi} \newcommand{\refs}[1]{Section ~\ref{S:#1}} \newcommand{\refss}[1]{Section ~\ref{SS:#1}} \newcommand{\reft}[1]{Theorem ~\ref{T:#1}} \newcommand{\refl}[1]{Lemma ~\ref{L:#1}} \newcommand{\refp}[1]{Proposition ~\ref{P:#1}} \newcommand{\refc}[1]{Corollary ~\ref{C:#1}} \newcommand{\refd}[1]{Definition ~\ref{D:#1}} \newcommand{\refr}[1]{Remark ~\ref{R:#1}} \newcommand{\refe}[1]{\eqref{E:#1}} \newenvironment{thm}[1]% { \begin{Thm} \label{T:#1} \ifShowLabels \TeXref{T:#1} \fi }% { \end{Thm} } \renewcommand{\th}[1]{\begin{thm}{#1} \sl } \renewcommand{\eth}{\end{thm} } \newenvironment{lemma}[1]% { \begin{Lem} \label{L:#1} \ifShowLabels \TeXref{L:#1} \fi }% { \end{Lem} } \newcommand{\lem}[1]{\begin{lemma}{#1} \sl} \newcommand{\end{lemma}}{\end{lemma}} \newenvironment{propos}[1]% { \begin{Prop} \label{P:#1} \ifShowLabels \TeXref{P:#1} \fi }% { \end{Prop} } \newcommand{\prop}[1]{\begin{propos}{#1}\sl } \newcommand{\end{propos}}{\end{propos}} \newenvironment{corol}[1]% { \begin{Cor} \label{C:#1} \ifShowLabels \TeXref{C:#1} \fi }% { \end{Cor} } \newcommand{\cor}[1]{\begin{corol}{#1} \sl } \newcommand{\end{corol}}{\end{corol}} \newenvironment{defeni}[1]% { \begin{Def} \label{D:#1} \ifShowLabels \TeXref{D:#1} \fi }% { \end{Def} } \newcommand{\defe}[1]{\begin{defeni}{#1} \sl } \newcommand{\end{defeni}}{\end{defeni}} \newenvironment{remark}[1]% { \begin{Rem} \label{R:#1} \ifShowLabels \TeXref{R:#1} \fi }% { \end{Rem} } \newcommand{\rem}[1]{\begin{remark}{#1}} \newcommand{\end{remark}}{\end{remark}} \newcommand{\eq}[1]% { \ifShowLabels \TeXref{E:#1} \fi \begin{equation} \label{E:#1} } \newcommand{\end{equation}}{\end{equation}} \newcommand{ \begin{proof} }{ \begin{proof} } \newcommand{ \end{proof} }{ \end{proof} } \newcommand\alp{\alpha} \newcommand\Alp{\Alpha} \newcommand\bet{\beta} \newcommand\gam{\gamma} \newcommand\Gam{\Gamma} \newcommand\del{\delta} \newcommand\Del{\Delta} \newcommand\eps{\varepsilon} \newcommand\zet{\zeta} \newcommand\tet{\theta} \newcommand\Tet{\Theta} \newcommand\iot{\iota} \newcommand\kap{\kappa} \newcommand\lam{\lambda} \newcommand\Lam{\Lambda} \newcommand\sig{\sigma} \newcommand\Sig{\Sigma} \newcommand\vphi{\varphi} \newcommand\ome{\omega} \newcommand\Ome{\Omega} \newcommand\calA{{\mathcal{A}}} \newcommand\calB{{\mathcal{B}}} \newcommand\calC{{\mathcal{C}}} \newcommand\calD{{\mathcal{D}}} \newcommand\calE{{\mathcal{E}}} \newcommand\calF{{\mathcal{F}}} \newcommand\calG{{\mathcal{G}}} \newcommand\calH{{\mathcal{H}}} \newcommand\calI{{\mathcal{I}}} \newcommand\calJ{{\mathcal{J}}} \newcommand\calK{{\mathcal{K}}} \newcommand\calL{{\mathcal{L}}} \newcommand\calM{{\mathcal{M}}} \newcommand\calN{{\mathcal{N}}} \newcommand\calO{{\mathcal{O}}} \newcommand\calP{{\mathcal{P}}} \newcommand\calQ{{\mathcal{Q}}} \newcommand\calR{{\mathcal{R}}} \newcommand\calS{{\mathcal{S}}} \newcommand\calT{{\mathcal{T}}} \newcommand\calU{{\mathcal{U}}} \newcommand\calV{{\mathcal{V}}} \newcommand\calW{{\mathcal{W}}} \newcommand\calX{{\mathcal{X}}} \newcommand\calY{{\mathcal{Y}}} \newcommand\calZ{{\mathcal{Z}}} \newcommand\bfa{{\mathbf a}} \newcommand\bfA{{\mathbf A}} \newcommand\bfb{{\mathbf b}} \newcommand\bfB{{\mathbf B}} \newcommand\bfc{{\mathbf c}} \newcommand\bfC{{\mathbf C}} \newcommand\bfd{{\mathbf d}} \newcommand\bfD{{\mathbf D}} \newcommand\bfe{{\mathbf e}} \newcommand\bfE{{\mathbf E}} \newcommand\bff{{\mathbf f}} \newcommand\bfF{{\mathbf F}} \newcommand\bfg{{\mathbf g}} \newcommand\bfG{{\mathbf G}} \newcommand\bfh{{\mathbf h}} \newcommand\bfH{{\mathbf H}} \newcommand\bfi{{\mathbf i}} \newcommand\bfI{{\mathbf I}} \newcommand\bfj{{\mathbf j}} \newcommand\bfJ{{\mathbf J}} \newcommand\bfk{{\mathbf k}} \newcommand\bfK{{\mathbf K}} \newcommand\bfl{{\mathbf l}} \newcommand\bfL{{\mathbf L}} \newcommand\bfm{{\mathbf m}} \newcommand\bfM{{\mathbf M}} \newcommand\bfn{{\mathbf n}} \newcommand\bfN{{\mathbf N}} \newcommand\bfo{{\mathbf o}} \newcommand\bfO{{\mathbf O}} \newcommand\bfp{{\mathbf p}} \newcommand\bfP{{\mathbf P}} \newcommand\bfq{{\mathbf q}} \newcommand\bfQ{{\mathbf Q}} \newcommand\bfr{{\mathbf r}} \newcommand\bfR{{\mathbf R}} \newcommand\bfs{{\mathbf s}} \newcommand\bfS{{\mathbf S}} \newcommand\bft{{\mathbf t}} \newcommand\bfT{{\mathbf T}} \newcommand\bfu{{\mathbf u}} \newcommand\bfU{{\mathbf U}} \newcommand\bfv{{\mathbf v}} \newcommand\bfV{{\mathbf V}} \newcommand\bfw{{\mathbf w}} \newcommand\bfW{{\mathbf W}} \newcommand\bfx{{\mathbf x}} \newcommand\bfX{{\mathbf X}} \newcommand\bfy{{\mathbf y}} \newcommand\bfY{{\mathbf Y}} \newcommand\bfz{{\mathbf z}} \newcommand\bfZ{{\mathbf Z}} \newcommand\QQ{\mathbb{Q}} \newcommand\WW{\mathbb{W}} \newcommand\EE{\mathbb{E}} \newcommand\RR{\mathbb{R}} \newcommand\TT{\mathbb{T}} \newcommand\YY{\mathbb{Y}} \newcommand\UU{\mathbb{U}} \newcommand\II{\mathbb{I}} \newcommand\OO{\mathbb{O}} \newcommand\PP{\mathbb{P}} \renewcommand\AA{\mathbb{A}} \renewcommand\SS{\mathbb{S}} \newcommand\DD{\mathbb{D}} \newcommand\FF{\mathbb{F}} \newcommand\GG{\mathbb{G}} \newcommand\HH{\mathbb{H}} \newcommand\JJ{\mathbb{J}} \newcommand\KK{\mathbb{K}} \newcommand\LL{\mathbb{L}} \newcommand\ZZ{\mathbb{Z}} \newcommand\XX{\mathbb{X}} \newcommand\CC{\mathbb{C}} \newcommand\VV{\mathbb{V}} \newcommand\BB{\mathbb{B}} \newcommand\NN{\mathbb{N}} \newcommand\MM{\mathbb{M}} \newcommand\grA{{\mathfrak{A}}} \newcommand\gra{{\mathfrak{a}}} \newcommand\grB{{\mathfrak{B}}} \newcommand\grb{{\mathfrak{b}}} \newcommand\grC{{\mathfrak{C}}} \newcommand\grc{{\mathfrak{c}}} \newcommand\grD{{\mathfrak{D}}} \newcommand\grd{{\mathfrak{d}}} \newcommand\grE{{\mathfrak{E}}} \newcommand\gre{{\mathfrak{e}}} \newcommand\grF{{\mathfrak{F}}} \newcommand\grf{{\mathfrak{f}}} \newcommand\grG{{\mathfrak{G}}} \newcommand\grg{{\mathfrak{g}}} \newcommand\grH{{\mathfrak{H}}} \newcommand\grh{{\mathfrak{h}}} \newcommand\grI{{\mathfrak{I}}} \newcommand\gri{{\mathfrak{i}}} \newcommand\grJ{{\mathfrak{J}}} \newcommand\grj{{\mathfrak{j}}} \newcommand\grK{{\mathfrak{K}}} \newcommand\grk{{\mathfrak{k}}} \newcommand\grL{{\mathfrak{L}}} \newcommand\grl{{\mathfrak{l}}} \newcommand\grM{{\mathfrak{M}}} \newcommand\grm{{\mathfrak{m}}} \newcommand\grN{{\mathfrak{N}}} \newcommand\grn{{\mathfrak{n}}} \newcommand\grO{{\mathfrak{O}}} \newcommand\gro{{\mathfrak{o}}} \newcommand\grP{{\mathfrak{P}}} \newcommand\grp{{\mathfrak{p}}} \newcommand\grQ{{\mathfrak{Q}}} \newcommand\grq{{\mathfrak{q}}} \newcommand\grR{{\mathfrak{R}}} \newcommand\grr{{\mathfrak{r}}} \newcommand\grS{{\mathfrak{S}}} \newcommand\grs{{\mathfrak{s}}} \newcommand\grT{{\mathfrak{T}}} \newcommand\grt{{\mathfrak{t}}} \newcommand\grU{{\mathfrak{U}}} \newcommand\gru{{\mathfrak{u}}} \newcommand\grV{{\mathfrak{V}}} \newcommand\grv{{\mathfrak{v}}} \newcommand\grW{{\mathfrak{W}}} \newcommand\grw{{\mathfrak{w}}} \newcommand\grX{{\mathfrak{X}}} \newcommand\grx{{\mathfrak{x}}} \newcommand\grY{{\mathfrak{Y}}} \newcommand\gry{{\mathfrak{y}}} \newcommand\grZ{{\mathfrak{Z}}} \newcommand\grz{{\mathfrak{z}}} \newcommand\nek{,\ldots,} \newcommand\sdp{\times \hskip -0.3em {\raise 0.3ex \hbox{$\scriptscriptstyle |$}}} \newcommand\area{\operatorname{area}} \newcommand\Aug{\operatorname{Aug}} \newcommand\Aut{\operatorname{Aut}} \newcommand\Char{\operatorname{Char}} \newcommand\Cl{\operatorname{Cl}} \newcommand\cf{{\rm \,cf\,}} \newcommand\Cone{\operatorname{Cone}} \newcommand\cont{\operatorname{cont}} \newcommand\codim{\operatorname{codim}} \newcommand\conv{\operatorname{conv}} \newcommand\Conv{\operatorname{Conv}} \newcommand\const{\operatorname{const}} \newcommand\Const{\operatorname{Const}} \newcommand\Det{\operatorname{Det}} \newcommand\diag{\operatorname{diag}} \newcommand\diam{\operatorname{diam}} \newcommand\Diam{\operatorname{Diam}} \newcommand\dist{\operatorname{dist}} \newcommand\Dom{\operatorname{Dom}} \newcommand\dom{\operatorname{dom}} \newcommand\End{\operatorname{End\,}} \newcommand\Ext{\operatorname{Ext}} \newcommand\esssup{\operatorname{ess\ sup}} \newcommand\Ran{{\rm Ran}} \newcommand\RANK{\operatorname{rank}} \newcommand\Geo{\operatorname{Geo}} \newcommand\GL{\operatorname{GL}} \newcommand\Gr{\operatorname{Gr}} \newcommand\gl{\operatorname{gl}} \newcommand\grad{\operatorname {\rm grad}} \newcommand\Hom{\operatorname {Hom}} \newcommand\RHom{\operatorname {RHom}} \newcommand\im{\operatorname {im}} \newcommand\IM{\operatorname{Im}} \newcommand\Ind{\operatorname{Ind}} \newcommand\ind{\operatorname{ind}} \newcommand\Inf{\operatorname{Inf}} \newcommand\Int{\operatorname{Int}} \newcommand\Min{\operatorname{Min}} \newcommand\MOD{\operatorname{mod}} \newcommand{\operatorname{Mor}}{\operatorname{Mor}} \newcommand{\operatorname{Ob\,}}{\operatorname{Ob\,}} \newcommand\ord{\operatorname{ord}} \newcommand\Ka{\operatorname{Ka}} \newcommand\Ker{\operatorname{Ker}} \newcommand\PGL{{\rm PGL}} \newcommand\PGSp{{\rm PGSp}} \newcommand\Plt{\operatorname{Plt}} \newcommand\proj{\operatorname{proj}} \newcommand\Proj{\operatorname{Proj}} \newcommand\res{\rm res} \newcommand\rk{\operatorname{rk}} \newcommand\Range{\operatorname{Range}} \newcommand\RE{\operatorname{Re}} \newcommand\Res{\operatorname{Res}} \newcommand\rot{\operatorname{rot}} \newcommand\Max{\operatorname{Max}} \newcommand\Maximum{\operatorname{Maximum}} \newcommand\Minimum{\operatorname{Minimum}} \newcommand\Minimize{\operatorname{Minimize}} \newcommand\Prob{\operatorname{Prob}} \newcommand\sech{\rm sech} \newcommand\sgn{\operatorname{sgn}} \newcommand{\operatorname{sign}}{\operatorname{sign}} \newcommand\SL{{\rm SL}} \newcommand\Sbm{\operatorname{Sbm}} \newcommand\SO{{\rm SO}} \newcommand\SPAN{\operatorname{span}} \newcommand\spin{\operatorname{spin}} \newcommand\spec{{\rm spec}} \newcommand\supess{\operatorname{sup\ ess}} \newcommand\supp{\operatorname{supp}} \newcommand\Supp{\operatorname{Supp}} \newcommand\Sup{\operatorname{Sup}} \newcommand\Sym{\operatorname{Sym}} \newcommand\tr{\operatorname{tr}} \newcommand\Tr{\operatorname{Tr}} \newcommand\Tor{\operatorname{Tor}} \newcommand\Var{\operatorname{Var}} \newcommand\Vol{\operatorname{Vol}} \newcommand\oa{{\overline{a}}} \newcommand\oA{{\overline{A}}} \newcommand\ob{{\overline{b}}} \newcommand\oB{{\overline{B}}} \newcommand\oc{{\overline{c}}} \newcommand\oC{{\overline{C}}} \newcommand\oD{{\overline{D}}} \newcommand\od{{\overline{d}}} \newcommand\oE{{\overline{E}}} \renewcommand\oe{{\overline{e}}} \newcommand\of{{\overline{f}}} \newcommand\oF{{\overline{F}}} \newcommand\og{{\overline{g}}} \newcommand\oG{{\overline{G}}} \newcommand\oh{{\overline{h}}} \newcommand\oH{{\overline{H}}} \newcommand\oI{{\overline{I}}} \newcommand\oj{{\overline{j}}} \newcommand\oJ{{\overline{J}}} \newcommand\ok{{\overline{k}}} \newcommand\oK{{\overline{K}}} \newcommand\oL{{\overline{L}}} \newcommand\om{{\overline{m}}} \newcommand\oM{{\overline{M}}} \newcommand\oN{{\overline{N}}} \newcommand\oO{{\overline{O}}} \newcommand\oo{{\overline{o}}} \newcommand\op{{\overline{p}}} \newcommand\oP{{\overline{P}}} \newcommand\oq{{\overline{q}}} \newcommand\oQ{{\overline{Q}}} \newcommand\OR{{\overline{r}}} \newcommand\oS{{\overline{S}}} \newcommand\os{{\overline{s}}} \newcommand\ot{{\overline{t}}} \newcommand\oT{{\overline{T}}} \newcommand\ou{{\overline{u}}} \newcommand\oU{{\overline{U}}} \newcommand\ov{{\overline{v}}} \newcommand\oV{{\overline{V}}} \newcommand\ow{{\overline{w}}} \newcommand\oW{{\overline{W}}} \newcommand\ox{{\overline{x}}} \newcommand\oX{{\overline{X}}} \newcommand\oy{{\overline{y}}} \newcommand\oY{{\overline{Y}}} \newcommand\oz{{\overline{z}}} \newcommand\oZ{{\overline{Z}}} \newcommand\oalp{{\overline{\alpha}}} \newcommand\obet{{\overline{\bet}}} \newcommand\odel{{\overline{\del}}} \newcommand\oDel{{\overline{\Del}}} \newcommand\ocup{{\overline{\cup}}} \newcommand\ovarphi{{\overline{\varphi}}} \newcommand\ochi{{\overline{\chi}}} \newcommand\oeps{{\overline{\eps}}} \newcommand\oeta{{\overline{\eta}}} \newcommand\ogam{{\overline{\gam}}} \newcommand\okap{{\overline{\kap}}} \newcommand\olam{{\overline{\lambda}}} \newcommand\oLam{{\overline{\Lambda}}} \newcommand\omu{{\overline{\mu}}} \newcommand\onu{{\overline{\nu}}} \newcommand\oOme{{\overline{\Ome}}} \newcommand\ophi{\overline{\phi}} \newcommand\oPhi{{\overline{\Phi}}} \newcommand\opi{{\overline{\pi}}} \newcommand\oPsi{{\overline{\Psi}}} \newcommand\opsi{{\overline{\psi}}} \newcommand\orho{{\overline{\rho}}} \newcommand\osig{{\overline{\sig}}} \newcommand\otau{{\overline{\tau}}} \newcommand\otet{{\overline{\theta}}} \newcommand\oxi{{\overline{\xi}}} \newcommand\oome{\overline{\ome}} \newcommand\opart{{\overline{\partial}}} \newcommand\ua{{\underline{a}}} \newcommand\ub{{\underline{b}}} \newcommand\uc{{\underline{c}}} \newcommand\uD{{\underline{D}}} \newcommand\uk{{\underline{k}}} \newcommand\ue{{\underline{e}}} \newcommand\uj{{\underline{j}}} \newcommand\ul{{\underline{l}}} \newcommand\uL{{\underline{L}}} \newcommand\uo{{\underline{o}}} \newcommand\uO{{\underline{O}}} \newcommand\uP{{\underline{P}}} \newcommand\uQ{{\underline{Q}}} \newcommand\um{{\underline{m}}} \newcommand\uM{{\underline{M}}} \newcommand\un{{\underline{n}}} \newcommand\us{{\underline{s}}} \newcommand\ut{{\underline{t}}} \newcommand\uu{{\underline{u}}} \newcommand\uv{{\underline{v}}} \newcommand\uV{{\underline{V}}} \newcommand\ux{{\underline{x}}} \newcommand\uX{{\underline{X}}} \newcommand\uy{{\underline{y}}} \newcommand\uz{{\underline{z}}} \newcommand\ualp{{\underline{\alp}}} \newcommand\ubet{{\underline{\bet}}} \newcommand\uchi{{\underline{\chi}}} \newcommand\udel{{\underline{\del}}} \newcommand\uell{{\underline{\ell}}} \newcommand\ueps{{\underline{\eps}}} \newcommand\ueta{{\underline{\eta}}} \newcommand\uGam{{\underline{\Gamma}}} \newcommand\unu{{\underline{\nu}}} \newcommand\uome{{\underline{\omega}}} \newcommand\utet{{\underline{\tet}}} \newcommand\ulam{{\underline{\lam}}} \newcommand\hata{{\widehat{a}}} \newcommand\hatA{{\widehat{A}}} \newcommand\hatb{{\widehat{b}}} \newcommand\hatc{{\widehat{c}}} \newcommand\hatC{{\widehat{C}}} \newcommand\hatB{{\widehat{B}}} \newcommand\hatD{{\widehat{D}}} \newcommand\hate{{\widehat{e}}} \newcommand\hatE{{\widehat{E}}} \newcommand\hatf{{\widehat{f}}} \newcommand\hatF{{\widehat{F}}} \newcommand\hatg{{\widehat{g}}} \newcommand\hatG{{\widehat{G}}} \newcommand\hath{{\widehat{h}}} \newcommand\hatH{{\widehat{H}}} \newcommand\hati{{\hat{i}}} \newcommand\hatI{{\hat{I}}} \newcommand\hatj{{\widehat{j}}} \newcommand\hatJ{{\widehat{J}}} \newcommand\hatk{{\widehat{k}}} \newcommand\hatK{{\widehat{K}}} \newcommand\hatL{{\widehat{L}}} \newcommand\hatm{{\widehat{m}}} \newcommand\hatM{{\widehat{M}}} \newcommand\hatn{{\widehat{n}}} \newcommand\hatN{{\widehat{N}}} \newcommand\hatp{{\widehat{p}}} \newcommand\hatP{{\widehat{P}}} \newcommand\hatr{{\widehat{r}}} \newcommand\hatR{{\widehat{R}}} \newcommand\hatq{{\widehat{q}}} \newcommand\hatQ{{\widehat{Q}}} \newcommand\hatT{{\widehat{T}}} \newcommand\hatu{{\widehat{u}}} \newcommand\hatU{{\widehat{U}}} \newcommand\hatV{{\widehat{V}}} \newcommand\hatv{{\widehat{v}}} \newcommand\hatw{{\widehat{w}}} \newcommand\hatW{{\widehat{W}}} \newcommand\hatx{{\widehat{x}}} \newcommand\hatX{{\widehat{X}}} \newcommand\haty{{\widehat{y}}} \newcommand\hatY{{\widehat{Y}}} \newcommand\hatZ{{\widehat{Z}}} \newcommand\hatz{{\widehat{z}}} \newcommand\hatalp{{\widehat{\alpha}}} \newcommand\hatdel{{\widehat{\delta}}} \newcommand\hatDel{{\widehat{\Delta}}} \newcommand\hatbet{{\widehat{\beta}}} \newcommand\hateps{{\hat{\eps}}} \newcommand\hatgam{{\widehat{\gamma}}} \newcommand\hatGam{{\widehat{\Gamma}}} \newcommand\hatlam{{\widehat{\lambda}}} \newcommand\hatmu{{\widehat{\mu}}} \newcommand\hatnu{{\widehat{\nu}}} \newcommand\hatOme{{\widehat{\Ome}}} \newcommand\hatphi{{\widehat{\phi}}} \newcommand\hatPhi{{\widehat{\Phi}}} \newcommand\hatpi{{\widehat{\pi}}} \newcommand\hatpsi{{\widehat{\psi}}} \newcommand\hatPsi{{\widehat{\Psi}}} \newcommand\hatrho{{\widehat{\rho}}} \newcommand\hatsig{{\widehat{\sig}}} \newcommand\hatSig{{\widehat{\Sig}}} \newcommand\hattau{{\widehat{\tau}}} \newcommand\hattet{{\widehat{\theta}}} \newcommand\hatvarphi{{\widehat{\varphi}}} \newcommand\hatZZ{{\widehat{\ZZ}}} \newcommand\tilA{{\widetilde{A}}} \newcommand\tila{{\widetilde{a}}} \newcommand\tilB{{\widetilde{B}}} \newcommand\tilb{{\widetilde{b}}} \newcommand\tilc{{\widetilde{c}}} \newcommand\tilC{{\widetilde{C}}} \newcommand\tild{{\widetilde{d}}} \newcommand\tilD{{\widetilde{D}}} \newcommand\tilE{{\widetilde{E}}} \newcommand\tilf{{\widetilde{f}}} \newcommand\tilF{{\widetilde{F}}} \newcommand\tilg{{\widetilde{g}}} \newcommand\tilG{{\widetilde{G}}} \newcommand\tilh{{\widetilde{h}}} \newcommand\tilk{{\widetilde{k}}} \newcommand\tilK{{\widetilde{K}}} \newcommand\tilj{{\widetilde{j}}} \newcommand\tilm{{\widetilde{m}}} \newcommand\tilM{{\widetilde{M}}} \newcommand\tilH{{\widetilde{H}}} \newcommand\tilL{{\widetilde{L}}} \newcommand\tilN{{\widetilde{N}}} \newcommand\tiln{{\widetilde{n}}} \newcommand\tilO{{\widetilde{O}}} \newcommand\tilP{{\widetilde{P}}} \newcommand\tilp{{\widetilde{p}}} \newcommand\tilq{{\widetilde{q}}} \newcommand\tilQ{{\widetilde{Q}}} \newcommand\tilR{{\widetilde{R}}} \newcommand\tilr{{\widetilde{r}}} \newcommand\tilS{{\widetilde{S}}} \newcommand\tils{{\widetilde{s}}} \newcommand\tilT{{\widetilde{T}}} \newcommand\tilt{{\widetilde{t}}} \newcommand\tilu{{\widetilde{u}}} \newcommand\tilU{{\widetilde{U}}} \newcommand\tilv{{\widetilde{v}}} \newcommand\tilV{{\widetilde{V}}} \newcommand\tilw{{\widetilde{w}}} \newcommand\tilW{{\widetilde{W}}} \newcommand\tilX{{\widetilde{X}}} \newcommand\tilx{{\widetilde{x}}} \newcommand\tily{{\widetilde{y}}} \newcommand\tilY{{\widetilde{Y}}} \newcommand\tilZ{{\widetilde{Z}}} \newcommand\tilz{{\widetilde{z}}} \newcommand\tilalp{{\widetilde{\alpha}}} \newcommand\tilbet{{\widetilde{\beta}}} \newcommand\tildel{{\widetilde{\delta}}} \newcommand\tilDel{{\widetilde{\Delta}}} \newcommand\tilchi{{\widetilde{\chi}}} \newcommand\tileta{{\widetilde{\eta}}} \newcommand\tilgam{{\widetilde{\gamma}}} \newcommand\tilGam{{\widetilde{\Gamma}}} \newcommand\tilome{{\widetilde{\ome}}} \newcommand\tillam{{\widetilde{\lam}}} \newcommand\tilmu{{\widetilde{\mu}}} \newcommand\tilphi{{\widetilde{\phi}}} \newcommand\tilpi{{\widetilde{\pi}}} \newcommand\tilpsi{{\widetilde{\psi}}} \renewcommand\tilome{{\widetilde{\ome}}} \newcommand\tilOme{{\widetilde{\Ome}}} \newcommand\tilPhi{{\widetilde{\Phi}}} \newcommand\tilQQ{{\widetilde{\QQ}}} \newcommand\tilrho{{\widetilde{\rho}}} \newcommand\tilsig{{\widetilde{\sig}}} \newcommand\tiltau{{\widetilde{\tau}}} \newcommand\tiltet{{\widetilde{\theta}}} \newcommand\tilvarphi{{\widetilde{\varphi}}} \newcommand\tilxi{{\widetilde{\xi}}} \newcommand\twolongrightarrow{\ \hbox{$\longrightarrow\hskip -17pt \longrightarrow$}\ } \renewcommand{\>}{\rangle} \newcommand{\<}{\langle} \renewcommand{\d}{\text{\( \partial\)}}\newcommand{\bar{\d}}{\bar{\d}} \renewcommand{\b}{\bullet} \newcommand{\omega}{\omega} \newcommand{\calF}{\calF} \renewcommand{\O}{\calO} \newcommand{\calF_a^k}{\calF_a^k} \newcommand{\nabla}{\nabla} \newcommand{\hm}[2]{{H^{#2}(M_{#1},\calO(E_{#1}))}} \newcommand{\hmk}[2]{{H_k^{#2}(M_{#1},\calO(E_{#1}))}} \newcommand{\hma}[1]{{H^{*}(M_{#1},\calO(E_{#1}))}} \newcommand{\hmak}[1]{{H_k^{*}(M_{#1},\calO(E_{#1}))}} \newcommand{\hmo}[2]{{H^{#2}_0(M_{#1},\calO(E_{#1}))}} \newcommand{\backslash}{\backslash} \newcommand{{\Ome^{0,k}}}{{\Ome^{0,k}}} \newcommand{^p_k}{^p_k} \newcommand{{(1,0)}}{{(1,0)}} \newcommand{{(0,1)}}{{(0,1)}} \newcommand\ch{\operatorname{char}} \newcommand\mult{\operatorname{mult}} \newcommand{\preccurlyeq}{\preccurlyeq} \newcommand{\bar\square_t}{\bar\square_t} \newcommand{\frac12}{\frac12} \newcommand{K\"ahler }{K\"ahler } \begin{document} \title{Symplectic cutting of K\"ahler manifolds} \author{Maxim Braverman} \address{Institute of Mathematics\\ The Hebrew University \\ Jerusalem 91904 \\ Israel } \email{[email protected]} \thanks{This research was partially supported by grant No. 96-00210/1 from the United States-Israel Binational Science Foundation (BSF), Jerusalem, Israel} \begin{abstract} We obtain estimates on the character of the cohomology of an $S^1$-equivariant holomorphic vector bundle over a K\"ahler manifold $M$ in terms of the cohomology of the Lerman symplectic cuts and the symplectic reduction of $M$. In particular, we prove and extend inequalities conjectured by Wu and Zhang \cite{WuZhang}. The proof is based on constructing a flat family of complex spaces $M_t \ (t\in\CC)$ such that $M_t$ is isomorphic to $M$ for $t\not=0$, while $M_0$ is a singular reducible complex space, whose irreducible components are the Lerman symplectic cuts. \end{abstract} \maketitle \sec{introd}{Introduction} Let $M$ be a smooth K\"ahler manifold of complex dimension $n$ endowed with a holomorphic Hamiltonian action of the circle group $S^1$. Let $\mu:M\to \RR$ denote the moment map for this action. Assume that 0 is a regular value of $\mu$ and that $S^1$ acts freely on $\mu^{-1}(0)$. Using a construction of E.~Lerman, \cite{Lerman-cut}, one can "cut" $M$ into two smooth K\"ahler manifolds $M_+$ and $M_-$ endowed with a holomorphic circle action. The symplectic reduction $M_{red}=\mu^{-1}(0)/S^1$ of $M$ is embedded into $M_\pm$ as a connected component of the fixed point set. Let $E$ be an equivariant holomorphic vector bundle over $M$. Then $E$ induces a holomorphic vector bundle $E_{red}$ over $M_{red}$ and equivariant holomorphic bundles $E_\pm$ over $M_\pm$. In this paper we show that there are Morse-type inequalities which estimate the character of the $S^1$-action on the cohomology $\hm{}{*}$ of the sheaf of holomorphic sections of $E$ in terms of the cohomology \/ $\hm{\pm}{*}, \ \hm{red}{*}$ \/ of the sheaves of holomorphic sections of the bundles $E_\pm$ and $E_{red}$ respectively. These inequalities were conjectured by Wu and Zhang \cite{WuZhang} for the case when $E$ is a pre-quantum line bundle. As a consequence, we obtain a new geometric proof of the "gluing formula" for the index of $E$ (\cite{DGMW,Meinr-GS}, see \refe{glue}) for the case when the manifold $M$ is K\"ahler. Our proof is based on the following geometric construction which, we believe, is interesting by itself. We consider the union $M_{cut}$ of $M_\pm$ along $M_{red}$. Thus $M_{cut}$ is a singular reducible complex space, whose smooth irreducible components $M_\pm$ intersect by the symplectic reduction $M_{red}$. We show that $M_{cut}$ may be considered as a deformation of $M$. More precisely, we construct a family $M_t$ of complex spaces parameterized by a complex parameter $t$, such that $M_t$ is complex isomorphic to $M$ for any $t\not=0$ while $M_0$ is complex isomorphic to $M_{cut}$. It turns out that $M_t$ is {\em a flat family of complex spaces}. That implies that the dimension (and also the character) of the cohomology of $M_t$ is an upper semi-continuous function of $t$. In particular, the character of $\hm{cut}{*}$ is greater than the character of $\hm{}{*}$ (a partial order on the ring of characters is introduced in \refd{polyn}). Moreover, there are Morse-type inequalities (cf. \reft{Mcut>M}) which estimate the character of $\hm{}{*}$ in terms of the character of $\hm{cut}{*}$. The cohomology $\hm{cut}{*}$ of the space $M_{cut}$ can be, in turn, calculated by means of a Mayer-Vietoris-type long exact sequence via the cohomology of $M_\pm$ and $M_{red}$ (cf. \refss{Mcut}). That leads to estimates for the cohomology of $M$ in terms of the cohomology of $M_\pm$ and $M_{red}$. The paper is organized as follows. In \refs{main}, we formulate our main results. In \refs{family}, we present our geometric construction of the family of complex spaces and prove some important properties of this family. Finally, in \refs{proof}, we present the proof of \reft{Mcut>M}. \subsection*{Acknowledgments} I would like to thank I.~Zakharevich for very useful and inspiring discussions. It was I.~Zakharevich who suggested to consider the union of the Lerman symplectic cuts $M_\pm$ as a singular complex space. I would like to thank the University of Warwick, where this work was completed, for hospitality. \sec{main}{Main results} In this section we formulate the main results of the paper. All these results are consequences of \reft{Mcut>M}, which will be proved in \refs{proof}. \ssec{char}{Weights and formal characters} Irreducible representation of the circle group $S^1=\{e^{i\tet}:\, \tet\in\RR\}$ are classified by integer {\em weights} (here we use the identification of the Lie algebra of $S^1$ with $\RR$ which takes the {\em negative} primitive lattice element, $-2\pi i\in i\RR= Lie(S^1)$, to $1$). A representation of weight $k\in \ZZ$ is isomorphic to the complex line $\CC$ on which the element $e^{i\tet}\in S^1$ acts by multiplications by $e^{-ik\tet}$. If $W$ is a finite dimensional representation of \/ $S^1$ \/ we denote by $\mult_k(W)$ the multiplicity of the weight $k\in\ZZ$ in $W$. The {\em formal character} of $W$ is the formal sum $$ \ch(W) \ = \ \sum_{k\in\ZZ}\mult_k(W)e^{-ik\tet}. $$ It lies in the ring \/ $\calL=\ZZ[e^{i\tet},e^{-i\tet}]$ \/ of Laurent polynomials in $e^{i\tet}$ with integer coefficients. This ring is called the {\em ring of formal characters} of the circle group. \ssec{mom-red}{Momentum map and symplectic reduction} Let $V$ denote the vector field on $M$ that generates the $S^1$-action and let $\omega$ denote the K\"ahler form on $M$. We will assume that $S^1$-action is {\em Hamiltonian}, i.e. there is a moment map $\mu: M\to\RR$ such that $\iot_V\ome=d\mu$. Note (\cite{Frankel}) that it is always the case if the fixed-point set of $S^1$ on $M$ is non-empty. Assume that $0\in \RR$ is a regular value of the moment map $\mu$. Then $\mu^{-1}(0)\subset M$ is a smooth submanifold endowed with a locally free action of $S^1$. We will assume that this action is free. Then the quotient space $M_{red}=\mu^{-1}(0)/S^1$ is a smooth symplectic manifold called the {\em symplectic reduction of $M$ at level $0$}. The symplectic form $\omega_{red}$ on $M_{red}$ is defined by the condition that its lift on $\mu^{-1}(0)$ coincides with the restriction of $\omega$ on $\mu^{-1}(0)$. Recall now that our manifold $M$ is K\"ahler and that the K\"ahler structure on $M$ is preserved by the circle action. In this case, {\em the $S^1$ action can be canonically extended to a holomorphic action of the group of nonzero complex numbers $\CC^*$} (cf. \cite[Lemma~3.3]{GuiSter82}). The set $$ M_s \ = \ \big\{z\cdot x: \ z\in \CC^*, x\in \mu^{-1}(0)\subset M \big\}, $$ called {\em the set of stable points} for the $\CC^*$ action, is an open submanifold of $M$, \cite[Lemma~4.5]{GuiSter82}, and the $\CC^*$ action on $M_s$ is free. Obviously, the quotient of $M_s$ by this action is diffeomorphic to the reduced space: \eq{M/C} M_{red} \ \cong \ M_s/\CC^*. \end{equation} The equation \refe{M/C}, defines {\em a canonical complex structure on $M_{red}$}. This structure is, in fact, K\"ahler, and the corresponding K\"ahler form coincides with the form $\omega_{red}$ defined above. Let now $E$ be a holomorphic vector bundle over $M$ which is equivariant for the $S^1$ action. Then the $\CC^*$ action on $M$ can be also lifted on $E$. There is a unique holomorphic vector bundle $E_{red}$ over $M_{red}$ such that its pullback under the projection \/ $M_s\to M_{red}$ \/ is isomorphic to the restriction of $E$ on $M_s$. \ssec{cut}{Symplecting cuttings} We now recall the Lerman symplectic cutting construction, \cite{Lerman-cut}. Let $\CC_{\pm}$ denote the complex one-dimensional representations of the circle group of weights $\pm 1$ respectively. We endow both $\CC_+$ and $\CC_-$ with the standard K\"ahler form $\ome=\frac{i}{2}d z\wedge d\oz$. The diagonal actions of $S^1$ on $M\times\CC_\pm$ are Hamiltonian and the corresponding moment maps are $\mu\mp\frac12|z|^2$. One checks easily that 0 is a regular value for each one of these moment maps. Let us denote by $M_\pm$ the symplectic quotients of $M\times\CC_\pm$ at level $0$. The action of $S^1$ on the first factor of $M\times \CC_\pm$ reduces to a Hamiltonian action on $M_\pm$. Thus, $(M_\pm,\ome_\pm)$ are smooth symplectic manifolds with Hamiltonian $S^1$-actions. The reduced space $M_{red}$ is embedded into $M_\pm$ as one of the connected components (still denoted by $M_{red}$) of the fixed points set; the compliments $M_\pm\backslash M_{red}$ are $S^1$-equivariantly symplectomorphic to $\mu^{-1}(\RR^\pm)\subset M$, respectively. We refer to $M_\pm$ as {\em symplectic cuts of $M$}. The pull-back of the bundle $E$ under the natural projection $M\times \CC_\pm\to M$ is an equivariant vector bundle over $M\times\CC_\pm$. Hence (cf. \refss{mom-red}), it induces holomorphic vector bundles $E_\pm$ over $M_\pm$. One of the most important facts about the cohomology of the symplectic cuts is the {\em gluing formula} (cf. \cite{DGMW,Meinr-GS}) \begin{multline}\label{E:glue} \sum_{p=0}^n (-1)^p\ch\hm{}{p} \ = \ \sum_{p=0}^n (-1)^p\ch\hm{+}{p} \\ \ + \ \sum_{p=0}^n (-1)^p\ch\hm{-}{p} \ - \ \sum_{p=0}^{n-1} (-1)^p\dim_{\CC} \hm{red}{p}. \end{multline} \rem{symplectic} Though the individual cohomology $\hm{}{p}$ has sense only for complex manifold $M$, the alternating sums which appear in \refe{glue} may be defined in the case when $M$ is an almost complex manifold. The formula \refe{glue} remains true for this more general case \cite{DGMW,Meinr-GS} (see also \cite{SiKaTo} were the gluing formula is obtained in a still more general situation). \end{remark} Let us return to the situation when $M$ is K\"ahler. The aim of this paper is to strengthen the gluing formula \refe{glue} in order to obtain an information about individual cohomology $\hm{}{p}$ of $M$ in terms of the cohomology of $M_\pm$ and $M_{red}$. \ssec{sing}{Symplectig cutting as a singular space} Both manifolds $M_\pm$ contain the symplectic reduction $M_{red}$ as a submanifold. Consider the union $$ M_{cut}= M_+\cup_{M_{red}} M_- $$ along $M_{red}$. Then $M_{cut}$ is a singular reducible complex space whose irreducible components are $M_\pm$ and whose only singularities are the "double points" in $M_{red}$. Let $E_{cut}$ denote the vector bundle over $M_{cut}$ whose restriction onto $M_\pm$ is equal to $E_\pm$. The advantage of considering the singular space $M_{cut}$ rather then two disconnected manifolds $M_\pm$ is that $M_{cut}$ may be considered as a deformation of $M$ (cf. \refs{family}). This implies (cf. \reft{Mcut>M}) estimates on \/ $\ch H^*(M,\O(E))$ \/ in terms of the character \/ $\ch H^*(M_{cut},\O(E_{cut}))$ \/ of the cohomology of the sheaf of holomorphic sections of $E_{cut}$. The cohomology \/ $H^*(M_{cut},\O(E_{cut}))$ \/ my be, in turn, calculated in terms of the cohomology of the sheaves \/ $E_\pm$ \/ and \/ $E_{red}$ \/ (cf. \refss{Mcut} bellow). That gives an estimate on \/ $\ch H^*(M,\O(E))$ \/ in terms of the spaces \/ $M_{\pm},M_{red}$. To formulate the result we need the following \defe{polyn} Let $q(\tet)= \sum_{k\in\ZZ}q_ke^{-ik\tet}\in \calL$ \/ be a formal character of $S^1$, we say $q(\tet)\ge 0$ if $q_k\ge 0$ for all $k\in\ZZ$. For two characters $p,q\in \calL$, we say that $p\ge q$ if $p-q\ge 0$. Let $Q(\tet,t)= \sum_{m=0}^n q_m(\tet)t^m\in \ \calL[t]$ be a polynomial of degree $n$ with coefficients in $\calL$, we say $Q(\tet,t)\ge0$ if $q_m(\tet)\ge0$ for all $m$. \end{defeni} Our first result is the following Morse-type inequalities between the cohomology of $M$ and $M_{cut}$. \th{Mcut>M} There exists a polynomial $Q(\tet,t)\in \calL[t]$, such that $Q\ge 0$ and \eq{Mcut>M} \sum_{p=0}^n t^p\ch\hm{cut}{p} \ = \ \sum_{p=0}^n t^p\ch\hm{}{p} \ + \ (1+t)Q(\tet,t). \end{equation} \eth \reft{Mcut>M} is proven in \refs{proof}. \rem{Mcut>M} \ 1. \ The Morse-type inequalities \refe{Mcut>M} imply $$ \ch\hm{cut}{p} \ \ge \ \ch\hm{}{p} \quad \mbox{for any}\quad p=0\nek n. $$ 2. \ Substituting $t=-1$ into \refe{Mcut>M} we obtain the following index formula \eq{Mcut=M} \sum_{p=0}^n (-1)^p\ch\hm{}{p} \ = \ \sum_{p=0}^n (-1)^p\ch\hm{cut}{p}. \end{equation} \end{remark} \ssec{Mcut}{Cohomology of $M_{cut}$} To calculate the cohomology of $M_{cut}$ with coefficients in $\O(E_{cut})$ consider the equivariant short exact sequence of sheaves $$\begin{CD} 0\to \O(E_{cut}) \ @>\alp>> \ \O(E_+)\oplus \O(E_-) \ @>\bet>> \O(M_{red}) \ \to 0, \end{CD} $$ Here the map \/ $\alp$ \/ sends the section $s$ of \/ $\O(E_{cut})$ \/ to the pair \/ $(s|_{M_+},s|_{M_-})$ \/ and the map \/ $\bet$ \/ sends the pair of sections \/ $(s_+,s_-)\in \O(E_+)\oplus \O(E_-)$ \/ to the section \/ $s_+|_{M_{red}}-s_-|_{M_{red}}\in \O(M_{red})$. By standard arguments, the above short sequence leads to an equivariant long exact sequence in cohomology \begin{multline}\label{E:HMcut} \cdots\to H^p(M_{cut},\O(E_{cut}))\to H^p(M_+,\O(E_+))\oplus H^p(M_-,\O(E_-)) \\ \to H^p(M_{red},\O(E_{red}))\to H^{p+1}(M_{cut},\O(E_{cut}))\to\cdots \end{multline} We think about $M_{cut}$ as being glued from $M_\pm$ along $M_{red}$. So we refer to \refe{HMcut} as Mayer-Vietoris-type sequence. \rem{WuZhang} Wu and Zhang \cite[Remark~4.10]{WuZhang} conjectured that, if $E$ is a {\em pre-quantum line bundle}, then (for a proper choice of the moment map) the cohomology $\hm{}{p}$ may be calculated by a Mayer-Vietoris-type exact sequence of type \refe{HMcut}. In our terms, that would mean that, in this case, $\hm{}{p}$ is isomorphic to $\hm{cut}{p}$. \end{remark} The long exact sequence \refe{HMcut} leads to the following Morse-type inequalities \begin{multline}\label{E:MorseMcut} \sum_{p=0}^n t^p\ch\hm{+}{p} \\ + \sum_{p=0}^n t^p\ch\hm{-}{p} \ + \ \sum_{p=0}^{n-1} t^{p+1}\dim\hm{red}{p} \\ = \ \sum_{p=0}^n t^p\ch\hm{cut}{p} \ + \ (1+t)Q(\tet,t) \end{multline} for some $Q(\tet,t)\ge0$. Combining with \reft{Mcut>M} we obtain the following estimate on $\hm{}{*}$ in terms of the cohomology of $M_\pm$ and $M_{red}$ \th{morse} There exists a polynomial $Q'(\tet,t)\in \calL[t]$, such that $Q\ge 0$ and \begin{multline}\label{E:MorseM} \sum_{p=0}^n t^p\ch\hm{+}{p} \ + \ \sum_{p=0}^n t^p\ch\hm{-}{p} \\ \ + \ \sum_{p=0}^{n-1} t^{p+1}\dim\hm{red}{p} = \ \sum_{p=0}^n t^p\ch\hm{}{p} \ + \ (1+t)Q'(\tet,t) \end{multline} \eth In the case when $E$ is a pre-quantum line bundle \reft{morse} was conjectured by Wu and Zhang \cite[Remark~4.10]{WuZhang}. \rem{gluing} \ 1. \ The inequalities \refe{MorseMcut} a far from being exact. Hence, \reft{Mcut>M} together with the Mayer-Vietoris sequence \refe{HMcut} give much more information about the cohomology $\hm{}{*}$ than \reft{morse}. 2. \ The simplest consequence of \reft{morse} is the inequalities \eq{simple} \begin{aligned} \ch\hm{+}{0} \ &+ \ \ch\hm{-}{0} \ \ge \ \ch\hm{}{0}; \\ \ch\hm{+}{p} \ &+ \ \ch\hm{-}{p} \ + \ \dim\hm{red}{p-1} \\ &\ge \ \ch\hm{}{p}, \quad\ \mbox{for any} \ \quad p=1\nek n. \end{aligned} \end{equation} 3. \ Substituting $t=-1$ into \refe{MorseM} and using \refe{Mcut=M}, we get the gluing formula \refe{glue}. So we obtain a new proof of the gluing formula for K\"ahler manifolds, which is based on the geometric construction described in \refs{family}. Note that the standard proof of the gluing formula, \cite{DGMW,Meinr-GS,SiKaTo}, uses the Atiyah-Segal-Singer equivariant index theorem. Dietmar Salamon pointed out that the gluing formula for general symplectic manifold can also be proved using our geometric construction by a method similar to \cite[Appendix~A]{McDSal2}. \end{remark} \rem{combine} In the situation considered in this paper many other Morse-type inequalities may be obtained (cf. \cite{WuZhang,TianZhang1,Br-HM}). It would be very interesting to compare those inequalities. \end{remark} \ssec{example}{Example} We finish this section with a very simple but typical example illustrating \reft{morse}. Let $M=\CC P^1$. We identify $M$ with the 2-dimensional sphere $S^2\subset\RR^3$ and we let $S^1$ act on $M$ by rotations around the $z$-axis. This action has two fixed points $P$ and $Q$ (the poles of the sphere). We normalize the K\"ahler structure on $M$ and the moment map $\mu$ so that $\mu(P)=1, \ \mu(Q)=-1$. Then the image of $\mu$ is the interval $[-1,1]$ and all the internal points of this interval are regular values of $\mu$. Let $E$ be an equivariant line bundle over $M$. Then $S^1$ acts on the fibers of this bundle over the fixed points $P$ and $Q$. Denote by $r_Q, r_P$ the weights of these actions. It is well known that $E$ is defined up to an equivariant isomorphism by these weights. In particular (cf., for example, \cite[p.~330]{Witten84}) the character of the representation of $S^1$ on the cohomology $\hm{}{p}$ is given by \footnote{Note that our signs in the definition of weights and characters (cf. \refss{char}) are different from \cite{Witten84} but agree with \cite{WuZhang,SiKaTo,Br-HM}.} \eq{cohom} \begin{aligned} \ch \hm{}{0} &= \begin{cases} \sum_{m=r_Q}^{r_P}e^{-im\tet}, \ \ \quad&\mbox{if}\quad r_Q\le r_P;\\ 0, \ \ \quad&\text{if}\quad r_Q> r_P; \end{cases} \\ % \ch \hm{}{1} &= \begin{cases} 0, \quad&\text{if}\quad r_Q\le r_P;\\ \sum_{m=r_P-1}^{r_Q-1}e^{-im\tet}, \quad&\mbox{if}\quad r_Q> r_P. \end{cases} \end{aligned} \end{equation} These formulas allow us to calculate the right hand side of \refe{MorseM}. Let us calculate the left hand side of \refe{MorseM}. Since, $M_{red}$ is a point, \/ $\dim\hm{red}{0}=1$ \/ and \/ $\dim\hm{red}{1}=0$. Both manifolds $M_+$ and $M_-$ are isomorphic to $\CC P^1$. The weight of the fiber of the bundle $E_+$ over $P$ is still equal to $r_P$ while the weight of the fiber over $M_{red}$ is equal to zero. Similarly, the weight of the fiber of the bundle $E_-$ over $Q$ is equal to $r_Q$ while the weight of the fiber over $M_{red}$ is equal to zero. Using \refe{cohom}, one can now verify \reft{morse} in this simple case. For example, if $r_Q=r_P=r>0$ (this corresponds to the trivial bundle $E=M\times\CC$ with a nontrivial circle action), then the left hand side of \refe{MorseM} is equal to $$ \sum_{m=0}^re^{-im\tet} + t\sum_{m=0}^{r-1}e^{-im\tet} $$ while $$ \sum_{p=0}^n t^p \ch\hm{}{p} = e^{-ir\tet}. $$ It follows that in this case the polynomial $Q'$ of \reft{morse} does not depend on $t$ and equals to \/ $\sum_{m=0}^{r-1}e^{-im\tet}$. Note also, that {\em the inequalities \refe{MorseM}, \refe{simple} become equalities if and only if $r_Q\le0$ and $r_P\ge0$}. This verifies the conjecture of Wu and Zhang \cite{WuZhang} (cf. \refr{WuZhang}) in the case $M=\CC P^1$. \sec{family}{The geometric construction} In this section we present a geometric construction of a complex manifolds $\Phi$ and a holomorphic map $p:\Phi\to \CC$ such that $p^{-1}(t)$ is isomorphic to $M$ for $t\not=0$ while $p^{-1}(0)=M_{cut}$. \ssec{family}{} The idea of the construction is the same as in \refss{cut}. In fact, we just combine the constructions of $M_+$ and $M_-$ together and consider the diagonal action of $S^1$ on $M\times\CC_+\times\CC_-$. The moment map for this action is given by $$ \tilmu(x,z_+,z_-) \ = \ \mu(x) \ - \ \frac12|z_+|^2+\frac12|z_-|^2, \qquad x\in M, \ z_\pm\in\CC_{\pm}. $$ Zero is a regular value of $\tilmu$ and we define $\Phi= \tilmu^{-1}(0)/S^1$ to be the symplectic reduction of $M\times\CC_+\times\CC_-$ at zero level. Clearly, the map $\tilp: \, M\times\CC_+\times\CC_-\to \CC$ defined by the formula $$ \tilp: \ (x,z_+,z_-) \ \mapsto \ z_+z_-, \qquad x\in M, \ z_\pm\in\CC_{\pm}. $$ is $S^1$ invariant and, hence, descends to a map $p:\Phi\to \CC$. We think about $\Phi$ as family of complex manifolds $M_t=p^{-1}(t)$ parameterized by a complex parameter $t\in\CC$. We endow $\Phi$ with an $S^1$-action induced by the action of the circle on the first factor of $M\times\CC_+\times\CC_-$. Note that this action preserves the fibers $M_t=p^{-1}(t)$ of $p$. Thus $M_t \ (t\in\CC)$ are also endowed with a holomorphic circle action. As in \refss{cut} the bundle $E$ induces a bundle $\tilE$ over $\Phi$. We denote the restriction of $\tilE$ on $M_t$ by $E_t$. The following lemmas describe the fibers of the projection $p$. \lem{MtS} For any $t\not=0$, the fiber $M_t=p^{-1}(t)$ is a smooth manifold which is equivariantly symplectomorphic to $M$. \end{lemma} \begin{proof} Fix a nonzero number $t\in\CC$. For any $x\in M$ set $$ r(x) \ = \ \sqrt{\mu(x)+\sqrt{\mu(x)^2+|t|^2}} $$ and define an embedding \eq{it} i_t: \, M\to \tilmu^{-1}(0)\cap\tilp^{-1}(t) \ \subset M\times\CC_+\times\CC_-, \qquad i_t: \ x \ \mapsto \Big(x,r(x),\frac{t}{r(x)}\Big). \end{equation} Clearly, the composition $q\circ i_t$ of the above embedding with the natural projection $q: \, \tilmu^{-1}(0)\to \Phi$ is an equivariant diffeomorphism $M\to M_t=p^{-1}(t)$. Since the map $i_t:M\to M\times\CC_+\times\CC_-$ is symplectic, so is the composition $q\circ i_t$. \end{proof} \lem{MtC} For any $t\not=0$, the fiber $M_t=p^{-1}(t)$ is a smooth K\"ahler manifold which is equivariantly complex isomorphic to $M$. In other words, there exists an equivariant biholomorphic map $\phi_t:M\overset{~}{\to} M_t$. The pullback $\phi^*_tE_t$ of the bundle $E_t=\tilE|_{M_t}$ is equivariantly isomorphic to $E$. \end{lemma} \rem{isom} The manifolds $M$ and $M_t$ are not isomorphic as K\"ahler manifolds. In particular, the map $\phi_t$ of \refl{MtC} is different from the symplectomorphism of \refl{MtS}. \end{remark} \begin{proof} Recall from \refss{mom-red} that the action of the circle group on $M$ extends canonically to a holomorphic action of $\CC^*$ and consider the diagonal action of $\CC^*$ on $M\times\CC_+\times\CC_-$ (here $z\in \CC^*$ acts on the second factor by multiplication by $z$ and on the third factor by multiplication by $1/z$). Let $U\subset M\times\CC_+\times\CC_-$ denote the set of stable points for this action (cf. \refss{mom-red}). Then $\tilp^{-1}(t)\subset U$ for any $t\not=0$. Indeed, if $v\in \tilp^{-1}(t)$ and if the absolute value of the number $z\in\CC^*$ is large enough, then $\tilmu(z\cdot v) < 0$, while $\tilmu(\frac1z\cdot v) > 0$. Hence, one can find $z'\in\CC^*$ such that $z'\in \tilmu^{-1}(0)$. Fix $t\not=0$ and consider the complex map $$ j_t:M\to M\times\CC_+\times\CC_-, \qquad j_t: x \ \mapsto \big(x,1,t\big). $$ Clearly, the image of $j_t$ belongs to $\tilp^{-1}(t)$ and, by the previous paragraph, it belongs also to the set $U$ of stable points for $\CC^*$ action. Hence, the composition of $j_t$ with the quotient map $q:U\to U/\CC^*$ defines an equivariant holomorphic map $\phi_t:M\to M_t$. Clearly, $\phi_t$ is injective. We claim that $\phi_t$ is an isomorphism. By \refl{MtS}, it suffice to show that the image of $j_t$ contains the image of the map \refe{it}. But this follows from the obvious inclusion \/ $\frac1r\cdot(x,r,\frac{t}r)\in \IM(j_t)$. The first statement of the lemma is proven. Consider the commutative diagram \eq{diagr} \begin{CD} M\times\CC_+\times\CC_- &&\ &\ \hookleftarrow&\quad &U\cap \tilp^{-1}(t)\\ @V{\pi}VV \ &{\nearrow}&\quad &@VV{q}V\\ M &\ & @>\phi_t>>\quad &M_t \end{CD} \end{equation} By definition of the bundle $E_t$ we have $q^*E_t=\pi^*E|_{U\cap \tilp^{-1}(t)}$. Hence, using $\phi_t=q\circ j_t$ and $\pi\circ j_t=id$, we obtain $$ \phi_t^*E_t \ = \ j_t^*q^*E_t \ = \ \phi_t^*\pi^*E \ = \ E. $$ \end{proof} \lem{M0} The fiber $p^{-1}(0)$ is equivariantly complex isomorphic to the space $M_{cut}$. If we identify $p^{-1}(0)$ with $M_{cut}$ using this isomorphism then the restriction $E_0$ of $\tilE$ to $p^{-1}(0)$ is isomorphic to $E_{cut}$. \end{lemma} \begin{proof} The lemma is an obvious consequence of the equality $$ \tilp^{-1}(0) \ = \ (M\times\CC_+\times\{0\}) \, \cup \, (M\times\{0\}\times\CC_-). $$ \end{proof} For us the most important is the following consequence of the above lemmas \cor{Phi} The cohomology $H^*(M_t,\O(E_t))$ of the sheaf of holomorphic sections of the bundle $E_t$ is equivariantly isomorphic to $\hm{}{*}$ if $t\not=0$ and is equivariantly isomorphic to $\hm{cut}{*}$ if $t=0$. \end{corol} \sec{proof}{Flat morphisms. Proof of \reft{Mcut>M}} We are in a position now to prove \reft{Mcut>M}. The proof is based on the properties of flat morphisms in complex analysis. \ssec{flat}{Flat morphisms} First, we recall some basic facts about flat morphisms. For the details we refer the reader to \cite[Sections~II.2,III.4]{GraPetRem}. If \/ $X$ \/ is a complex space and \/ $x\in X$ \/ we denote by \/ $\O(X)$ \/ the sheaf of holomorphic functions on \/ $X$ \/ and by \/ $\O_x(X)$ \/ the ring of germs of holomorphic functions at \/ $x$. Let \/ $f:X\to Y$ \/ be a holomorphic map of complex spaces. For any \/ $y\in Y$, we denote by \/ $X_y=f^{-1}(y)$ \/ the fiber of \/ $f$ \/ over \/ $y$. A holomorphic map \/ $f:X\to Y$ \/ is called {\em flat at a point \/ $x\in X$} \/ if \/ $\O_x(X)$ \/ is a flat module over \/ $\O_{f(x)}(Y)$. Here \/ $\O_x(X)$ \/ is considered as \/ $\O_{f(x)}(Y)$-module via the canonical map \/ $f^*:\O_{f(x)}(Y)\to \O_x(X)$. The map \/ $f$ \/ is called {\em flat} if it is flat at any point \/ $x\in X$. Let \/ $f:X\to Y$ \/ be a morphism of complex spaces and suppose \/ $V$ \/ is a holomorphic vector bundle over \/ $X$. For any \/ $y\in Y$, let \/ $V_y=V|_{X_y}$ \/ denote the restriction of \/ $V$ \/ on the fiber \/ $X_y$ \/ and let \/ $\O(V_y)$ \/ denote the locally free sheaf of holomorphic sections of \/ $V_y$. If \/ $f$ \/ is a flat morphism, then, for any \/ $y\in Y$ \/ and for any pint \/ $\eta\in Y$ \/ closed to $y$, there exists a polynomial \/ $Q(t)$ \/ with nonnegative integer coefficients, such that \eq{family} \sum_p t^p\dim H^p(X_y,\O(V_y)) \ = \ \sum_p t^p\dim H^p(X_\eta,\O(V_\eta)) \ + \ (1+t)Q(t). \end{equation} The equation \refe{family} implies, in particular, that the function \/ $y\mapsto \dim H^p(X_y,\O(V_y))$ \/ is upper semi-continuous, while the {\em holomorphic index} of the fibers $$ \ind(V_y) \ = \ \sum_p (-1)^p\dim H^p(X_y,\O(V_y)) $$ is locally constant on \/ $Y$. Suppose that in the situation described above a compact Lie group \/ $G$ \/ acts holomorphically on \/ $X$ \/ and \/ $Y$ \/ and that this action commutes with \/ $f$. If the vector bundle \/ $E$ \/ is equivariant with respect to this action then \/ $G$ \/ acts on the cohomology of the fibers. Let \/ $\ch H^p(X_y,\O(V_y)), \ y\in Y$ \/ denote the character of this action. Then, for any point \/ $\eta\in Y$ \/ closed enough to \/ $y$, there exists a polynomial \/ $Q(t,\tet)\in \calL[t]$ \/ (cf. \refd{polyn}) such that \/ $Q\ge 0$ \/ and \eq{eq-family} \sum_p t^p\ch H^p(X_y,\O(V_y)) \ = \ \sum_p t^p\ch H^p(X_\eta,\O(V_\eta)) \ + \ (1+t)Q(t,\tet). \end{equation} \ssec{proof}{Proof of \reft{Mcut>M}} It follows now from \refe{eq-family} and \refc{Phi}, that in order to prove \reft{Mcut>M} it suffices to show that the projection \/ $p:\Phi\to \CC$ is flat. But, by a theorem of Kaup and Kerner, \cite[Ch.~II, Theorem~2.13]{GraPetRem}, {\em any open holomorphic map of smooth complex manifolds is flat}. Since, \/ $p$ \/ is open the theorem is proven. \hfill $\square$ \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1996-07-04T16:35:05
9607
alg-geom/9607007
en
https://arxiv.org/abs/alg-geom/9607007
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9607007
Bernd Kreussler
Bernd Kreussler
On the algebraic dimension of twistor spaces over the connected sum of four complex projective planes
23 pages LaTeX 2e
null
null
null
null
We study the algebraic dimension of twistor spaces of positive type over $4\bbfP^2$. We show that such a twistor space is Moishezon if and only if its anticanonical class is not nef. More precisely, we show the equivalence of being Moishezon with the existence of a smooth rational curve having negative intersection number with the anticanonical class. Furthermore, we give precise information on the dimension and base locus of the fundamental linear system $|{-1/2}K|$. This implies, for example, $\dim|{-1/2}K|\leq a(Z)$. We characterize those twistor spaces over $4\bbfP^2$, which contain a pencil of divisors of degree one by the property $\dim|{-1/2}K| = 3$.
[ { "version": "v1", "created": "Thu, 4 Jul 1996 14:24:33 GMT" } ]
2008-02-03T00:00:00
[ [ "Kreussler", "Bernd", "" ] ]
alg-geom
\section{Introduction} \label{intro} Twistor spaces usually arise in four--dimensional conformal geometry. Their construction reflects the impossibility to equip in general a four--dimensional conformal manifold $M$ with a compatible complex structure. It was shown in \cite{AHS} that the conformal metric on $M$ is self--dual if and only if the twistor space $Z$ associated to $M$ carries, in a natural way, the structure of a complex manifold. Therefore, the conformal geometry of $M$ is closely related to the holomorphic geometry of $Z$. Since we shall only work with methods of complex geometry, we can use the following definition: \\ A twistor space $Z$ is a complex three--manifold with the following additional structure: \begin{itemize} \item a proper differentiable submersion $\pi:Z\rightarrow M$ onto a real differentiable four--manifold $M$. The fibres of $\pi$ are holomorphic curves in $Z$ being isomorphic to ${\Bbb C}{\Bbb P}^1$ and having normal bundle in $Z$ isomorphic to ${\mathcal O}(1)\oplus{\mathcal O}(1)$; \item an anti--holomorphic fixed point free involution $\sigma:Z\rightarrow Z$ with $\pi\sigma=\pi$. \end{itemize} The fibres of $\pi$ are called ``real twistor lines'' and the involution $\sigma$ is called the ``real structure''. A geometric object will be called ``real'' if it is $\sigma$--invariant. For example, a line bundle ${\mathcal L}$ on $Z$ is real if $\sigma^\ast\bar{{\mathcal L}}\cong {\mathcal L}$, and a complex subvariety $D\subset Z$ is real if $\sigma(D) = D$. Instead of $\sigma(D)$ we shall often write $\bar{D}$. We only consider compact and simply connected twistor spaces. At the beginning of the 80's, the first classification result emerged in \cite{FK}, \cite{Hit2}:\\ There exist exactly two compact K{\"a}hlerian twistor spaces. They are automatically projective algebraic. The corresponding Riemannian four--manifolds are the $4$--sphere $S^4$ and the complex projective plane $\bbfC\bbfP^2$ (with Fubini--Study metric). This was generalized in \cite{Camsc} to the result that a twistor space which is bimeromorphic to a compact K{\"a}hler manifold must be Moishezon and simply connected. This implies (see \cite{Don}, \cite{F}) that $M$ is homeomorphic to the connected sum $n\bbfC\bbfP^2$ for some $n\geq 0$. New examples of Moishezon twistor spaces were constructed by Y.S.\ Poon \cite{Po1} (case $n = 2$) and C.\ LeBrun \cite{LeB}, H.\ Kurke \cite{Ku} (case $n\geq 3$). Nowadays the situation is well understood for $n\leq 3$ (cf. \cite{Hit2}, \cite{FK}, \cite{Po1}, \cite{KK}, \cite{Po2}). To become more precise, we have to introduce the notion of the ``type'' of a twistor space. By a result of R.~Schoen \cite{Sch}, every conformal class of a compact Riemannian four--manifold contains a metric of constant scalar curvature. Its sign will be called the {\sl type\ } of the twistor space. This is an invariant of the conformal class, hence of the twistor space. It was shown in \cite{Po3} that a Moishezon twistor space is always of positive type. If $n \leq 3$, the converse is also true. In this paper we focus on the positive type case, for two reasons. One reason is that we can then apply Hitchin's vanishing theorem (\ref{Hit}). The other reason is the following: a result of P.\ Gauduchon \cite{Gau} implies that any twistor space of negative type has algebraic dimension zero. From the results of M.\ Pontecorvo \cite{Pon} we easily derive that a twistor space of type zero over $n\bbfC\bbfP^2$ must also have algebraic dimension zero. It is not clear whether there exist twistor spaces of non--positive type over $n\bbfC\bbfP^2$. Computation of algebraic dimension is, therefore, interesting only in the case of positive type. A very important tool to compute the algebraic dimension of twistor spaces is the result of Y.S.\ Poon \cite{Po3} (see also \cite{Pon}) stating that the algebraic dimension is equal to the Iitaka dimension of the anticanonical bundle (cf. Section \ref{prelim}). From \cite{DonF} and \cite{Cam}, \cite{LeBP} it is known that the generic twistor space over $n\bbfC\bbfP^2$ has algebraic dimension one (if $n = 4$), respectively zero (if $n\geq 5$). For the case $n=4$, the characterizing property $c_1^3 = 0$ is of central importance. In this paper we study the following \begin{main}{Problem:} Compute the algebraic dimension $a(Z)$ of a twistor space $Z$ over $4\bbfC\bbfP^2$ in terms of geometric or numeric properties of certain divisors on $Z$. \end{main} A first attempt to tackle this problem was made by Y.S.\ Poon \cite[Section 7]{Po2}. He assumes, additionally, the existence of a divisor $D$ of degree one on $Z$. He studies a birational map $D\rightarrow{\Bbb P}^2$, which is the blow--up of four points. He seems to assume that these four points are actually in ${\Bbb P}^2$ (no infinitesimally near blown--up points). If these four points are in a special position he obtains $a(Z) = 3$. In the case of general position he can only show: $a(Z) \leq 2$. We shall, in general, not assume the existence of divisors of degree one. Because in case $n=4$ there exists at least a pencil of so--called fundamental divisors, we shall study their geometry to obtain our results. If $S\subset Z$ is a real fundamental divisor, we have a birational map $S\rightarrow\bbfP^1\times\bbfP^1$ which is the blow--up of eight points. We shall study in detail the possible positions for these points. We take into account that some of these points can be infinitesimally near each other. We are able to derive the algebraic dimension $a(Z)$ from the knowledge of the positions of these eight points. Similar considerations were made for general $n\geq 4$ in the paper \cite{PP}. But the authors of that paper are intersted in a study of small deformations of well--known Moishezon twistor spaces, and so they investigate only the case without infinitely near blown--up points. As a consequence of our results, we give a new characterization of the twistor spaces over $4\bbfC\bbfP^2$ which are first described by C.\ LeBrun \cite{LeB} (with methods from differential geometry). From the point of view of complex geometry the twistor space structure on these complex manifolds was found by H.\ Kurke \cite{Ku}. Following the literature, we call them {\em LeBrun twistor spaces}. These twistor spaces are characterized in \cite{Ku} and \cite{Po2} by the property to contain a pencil of divisors of degree one. In the case $n=4$ we show (Theorem (\ref{cb})) that they can also be characterized by the property $h^0(\fdb) = 4$ or by the structure of the base locus of $|\fund|$. Besides this, our main results are a precise description of the set of irreducible curves intersecting $\fdb$ negatively (Theorem \ref{ncurves}) and the following theorems, where $Z$ denotes always a simply connected compact twistor space of positive type over $4\bbfC\bbfP^2$: \setcounter{section}{6} \setcounter{thm}{1} \begin{thm} $a(Z) = 3 \iff \fdb$ is not nef;\\[-0.8mm] $a(Z) = 2 \iff \fdb$ is nef and $\exists m\geq 1: h^1(\fb{m}) \ne 0$;% \\[-0.8mm] $a(Z) = 1 \iff \forall m\geq 1: h^1(\fb{m}) = 0$. \end{thm} \begin{thm} The following conditions are equivalent: \begin{enumerate} \item $a(Z) = 3$;\vspace*{-1mm} \item $\fdb$ is not nef;\vspace*{-1mm} \item there exists a smooth rational curve $C\subset Z$ with $C.(\fund) < 0$. \end{enumerate} \end{thm} \setcounter{thm}{5} \begin{thm} $a(Z) \geq \dim|\fund|$. \end{thm} \begin{thm} If $\dim|\fund|\geq 2$, then:\\ $a(Z) = 2 \iff \fdb$ is nef $\iff |\fund|$ does not have base points. \end{thm} \setcounter{section}{1} \setcounter{thm}{0} This paper is organized as follows:\\ In Section \ref{prelim} well--known but necessary facts about simply connected compact twistor spaces of positive type are collected. Also Section \ref{eins} has preparatory character. We study there the structure of fundamental divisors for general $n$, using results and techniques contained in \cite{PP}. Technically important for the following sections will be Proposition \ref{types} where the structure of effective anticanonical curves on real fundamental divisors is described in detail. In the remaining three sections we assume $n=4$. In Section \ref{zwei} we study the case where the anti--canonical bundle $K_Z^{-1}$ is nef (in the sense of Mori theory). We shall prove that the algebraic dimension is, in this case, at most two. We also see how to distinguish between algebraic dimension one and two. This generalizes results of \cite{CK}. In Section \ref{drei} we assume $K_Z^{-1}$ to be not nef. We collect detailed information on the fundamental linear system $|\fund|$ and on the set of curves which intersect $\fdb$ negatively. In this cases the algebraic dimension is three. The final Section \ref{vier} combines the results of the previous part to prove the main theorems stated above. \begin{main}{Acknowledgement} I thank Fr\'ed\'eric Campana for encouragement and stimulating discussions. \end{main} \section{Preliminaries} \label{prelim} We briefly collect well--known facts which will be frequently used later. We refer the reader to \cite{AHS}, \cite{ES}, \cite{Hit2}, \cite{Kr}, \cite{Ku} and \cite{Po1}. For brevity, we assume, throughout this section, $Z$ to be a simply connected compact twistor space of positive type. As mentioned in Section \ref{intro} the corresponding Riemannian four--manifold $M$ is homeomorphic to $n\bbfC\bbfP^2$. \subsubsection*{Cohomology ring of $Z$} $H^i(Z,{\Bbb Z})$ is a free ${\Bbb Z}$--module.\\ $H^1(Z,{\Bbb Z}) = H^3(Z,{\Bbb Z}) = H^5(Z,{\Bbb Z}) = 0$ and $H^0(Z,{\Bbb Z}) \cong H^6(Z,{\Bbb Z}) \cong {\Bbb Z}$. \\ $H^2(Z,{\Bbb Z})$ and $H^4(Z,{\Bbb Z})$ are free modules of rank $n+1$. There exists a basis $x_1,\dots, x_n, w$ of $H^2(Z,{\Bbb Z})$ such that the pull--back $H^2(Z,{\Bbb Z}) \rightarrow H^2(F,{\Bbb Z})\cong {\Bbb Z}$ (for any real twistor line $F \subset Z$) sends $x_i$ to $0$ and $w$ to the positive generator. The cohomology ring $H^\ast(Z,{\Bbb Z})$ is isomorphic to the graded ring ${\Bbb Z} [x_1,\dots, x_n, w]/R$ where $R$ is the ideal generated by \[ x_i^2 - x_j^2,\quad x_ix_j\; (i\ne j),\quad w^2 + w\sum_{i=1}^n x_i + x_1^2.\] The grading is given by $\deg x_i = \deg w = 2$.\\ $H^4(Z,{\Bbb Z})$ is a free ${\Bbb Z}$--module with generators $wx_1,\dots, wx_n, w^2$. The dual class of a real twistor fibre $F\subset Z$ is $-x_i^2 \in H^4(Z,{\Bbb Z})$. $c_1(Z) = 4w + 2\sum_{i=1}^n x_i, \quad c_2(Z) = -6x_1^2 = 6F, \quad c_3(Z) = 2(n + 2)$. This yields the following Chern numbers: $c_1^3 = 16(4 - n), \quad c_1c_2 = 24, \quad c_3 = 2(n + 2)$. \subsubsection*{Cohomology of sheaves} The main reason to assume $Z$ to be of positive type is Hitchin's vanishing theorem. We shall only use the following special case: \begin{thm}[Hitchin \cite{Hit}]\label{Hit} If $Z$ is of positive type then we have for any ${\mathcal L}\in \Pic(Z)$ \begin{eqnarray*} \deg({\mathcal L})\le -2 &\Rightarrow& H^1(Z,{\mathcal L})=0. \end{eqnarray*} \end{thm} On the other hand, since the twistor lines cover $Z$, we obtain: \begin{eqnarray*} \deg({\mathcal L})\le -1 &\Rightarrow& H^0(Z,{\mathcal L})=0. \end{eqnarray*} By Serre duality this gives the following important vanishing results: \begin{eqnarray} \deg({\mathcal L})\ge -2 &\Rightarrow& H^2(Z,{\mathcal L})=0,\\ \deg({\mathcal L})\ge -3 &\Rightarrow& H^3(Z,{\mathcal L})=0. \end{eqnarray} In particular, we obtain $h^2({\mathcal O}_Z) = h^3({\mathcal O}_Z) = 0$. Because $Z$ is simply connected, we also have $h^1({\mathcal O}_Z) = 0$. Hence, we obtain an isomorphism of abelian groups, given by the first Chern class: \[\Pic(Z)\stackrel{\sim}{\rightarrow}H^2(Z,{\Bbb Z}).\] There exists a unique line bundle whose first Chern class is $\frac{1}{2}c_1$. We shall denote it by $\fdb$. Following Poon, we call it the {\sl fundamental} line bundle. The divisors in the linear system $|\fund|$ will be called {\sl fundamental divisors}. The description of the cohomology ring gives $(\fund)^3 = 2(4 - n)$. If $S\in|\fund|$ is a smooth fundamental divisor, we obtain by the adjunction formula $\canS\cong\fdb\otimes{\mathcal O}_S$. If $n\leq 4$, there exist smooth real fundamental divisors (cf.\ \cite[Lemma 3.1]{CK}). The degree of a line bundle ${\mathcal L}\in \Pic(Z)$ will be by definition the degree of its restriction to a real twistor line. For example, $\deg(\fdb)=2$. We obtain in this way a {\sl surjective\/} degree map \[\deg :\Pic(Z)\twoheadrightarrow {\Bbb Z}.\] From the above equations on Chern numbers we obtain, by applying the Riemann--Roch theorem, \begin{equation}\label{RR} \chi(Z,\fb{m})=m+1+2(4-n){\binom{m+2}{3}}. \end{equation} \subsubsection*{Algebraic dimension} We denote by $a(Z)$ the algebraic dimension of $Z$, which is by definition the transcendence degree of the field of meromorphic functions of $Z$ over ${\Bbb C}$. If $\dim Z = a(Z)$, then $Z$ is called Moishezon. To compute the algebraic dimension of twistor spaces we shall frequently use, without further reference, the following theorem of Y.S.\ Poon: \begin{thm}{\cite{Po3}, \cite[Prop.\ 3.1]{Pon}}\\ $ \begin{array}[t]{lcl} \kappa(Z, K^{-1}) \geq 0& \Rightarrow& a(Z) = \kappa(Z, K^{-1})\\ \kappa(Z, K^{-1}) = -\infty& \Rightarrow& a(Z) = 0. \end{array}$ \end{thm} The number $\kappa(Z, K^{-1})$ is usually called the Iitaka dimension (or L--dimension = line bundle dimension) of the line bundle $K^{-1}$. Its definition generalizes the well--known notion of Kodaira dimension. For details, including the following facts, we refer the reader to \cite{U}. \\ For any line bundle ${\mathcal L}\in\Pic(Z)$ there holds: $\dim Z\geq a(Z)\geq \kappa(Z, {\mathcal L}) $.\\ If $f:Z\rightarrow Y$ is a dominant morphism, then $a(Z) \geq a(Y)$. Particularly, if $f:Z\rightarrow {\Bbb P}^N$ is a meromorphic map, then $a(Z) \geq \dim f(Z)$, because any projective variety is Moishezon.\\ If we define $g:=\gcd\{m\in {\Bbb Z}\; |\; m > 0, h^0(Z, L^m) \ne 0\}$ and denote by $\Phi_{|L^m|}$ the meromorphic map given by the linear system $|L^m|$, then $\kappa(Z, L) = \max\{\dim \Phi_{|L^m|}(Z)\;|\;m \in g{\Bbb Z}, \; m>0\}$. If there exists a polynomial $P(X)$ such that for all large positive $m\in {\Bbb Z}$ we have $h^0(Z, L^{mg}) \leq P(m)$, then $\kappa(Z, L)\leq \deg P$. We apply these basic facts to obtain our first result on the algebraic dimension in case $n = 4$. The following proposition is a generalization of a result contained in \cite{CK}. For convenience we introduce the following \begin{main}{Definition:} If there exists an integer $m\ge 1$ with $h^1(\fb{m})\ne 0$ then we define $\tau := \min\{m | m\ge 1, h^1(\fb{m})\ne 0\}$. Otherwise we set $\tau := \infty$. \end{main} \begin{prop} \label{tau} Let $Z$ be a simply connected compact twistor space of positive type with $c_1^3 = 0$. Then:\\ \begin{tabular}[t]{rl} (i) & $a(Z)\ge 1$\\ (ii) & $a(Z) = 1 \iff \forall m\ge 1\quad h^1(\fb{m})=0$. \end{tabular} \end{prop} {\sc Proof:\quad} From Riemann--Roch and Hitchin's vanishing theorem we know: $h^0(\fb{m})=m+1+h^1(\fb{m})$. Therefore, $a(Z)=\kappa(Z,\fdb)\ge 1$ and if $\tau=\infty$ we have $\kappa(Z,\fdb)=1$. Assume $\tau<\infty$ and $a(Z)=1$. Let $S\in|\fund|$ be smooth and real. (Such a divisor exists, because we assume $n=4$, cf.\ \cite{CK}.) Since $h^1(\fb{\tau-1})=0$ the exact sequence \[0 \rightarrow\fb{\tau-1} \rightarrow \fb{\tau} \rightarrow \can{\tau} \rightarrow 0\] gives an exact sequence \[ 0 \rightarrow H^0(\fb{\tau-1}) \rightarrow H^0(\fb{\tau}) \rightarrow H^0(\can{\tau}) \rightarrow 0. \] Since $h^1(\fb{\tau})\ge1$ we have, furthermore, $h^0(\fb{\tau})=\tau+1+h^1(\fb{\tau})\ge\tau+2$. The linear system $|\fun{\tau}|$ cannot have a fixed component since $\tau S\in |\fun{\tau}|$ and $\dim |S| = \dim |\fund| \ge 1$. If necessary blow up $Z$ to obtain a morphism $\Phi_\tau:\tilde{Z}\rightarrow {\Bbb P}^d$ defined by $|\fun{\tau}|$. Here $d:=\dim |\fun{\tau}| \ge \tau+1\ge 2$. By assumption $\dim \Phi_\tau(\tilde{Z})=1$. Since the curve $\Phi_\tau(\tilde{Z})$ is not contained in a linear subspace of ${\Bbb P}^d$, its degree must be at least $d$. Hence, a generic member of the linear system $|\fun{\tau}|$ is the sum of $\lambda$ algebraically equivalent divisors and so it is linearly equivalent to $\lambda S_0$ with $\lambda\ge d\ge\tau+1$. This gives $2\tau=\deg(\fun{\tau})=\lambda\deg(S_0),$ which is only possible if $\lambda = 2\tau$ and $\deg(S_0)=1$. But then we have infinitely many divisors of degree one in $Z$. This implies $a(Z)=3$ by the Theorem of Kurke--Poon (see \cite{Ku}, \cite{Po2}). This contradiction proves the proposition.\qed \begin{rem} If $|-K_S|$ contains a smooth curve $C$, then we computed in \cite{CK} that $\tau$ is the order of $N:=\canS\otimes{\mathcal O}_C$ in the Picard group $\Pic{C}$ of the elliptic curve $C$. Under this additional assumption Proposition \ref{tau} was shown in \cite{CK}. \end{rem} \section{The structure of fundamental divisors} \label{eins} In this section $Z$ always denotes a simply connected compact twistor space of positive type. \begin{lemma}\label{pic} Let $S\in|\fund|$ be a smooth surface. Then the restriction map $\Pic{Z} \rightarrow \Pic{S}$ is injective. \end{lemma} {\sc Proof:\quad} By assumption we have $h^1({\mathcal O}_Z)=h^2({\mathcal O}_Z)=0$. Since $S$ is a rational surface \cite{Po1}, we also have $h^1({\mathcal O}_S)=h^2({\mathcal O}_S)=0$. Therefore, taking the first Chern class defines isomorphisms $\Pic{Z} \stackrel{\sim}{\rightarrow} H^2(Z,{\Bbb Z})$ and $\Pic{S} \stackrel{\sim}{\rightarrow} H^2(S,{\Bbb Z})$. Let us denote the inclusion of $S$ into $Z$ by $i$. The above isomorphisms transform then the restriction morphism $\Pic{Z} \rightarrow \Pic{S}$ into the map $i^\ast$ on cohomology groups. We shall apply standard facts from algebraic topology to verify the injectivity of $i^\ast$. Let $\mbox{\rm o}_S \in H_4(Z,{\Bbb Z})$ and $\mbox{\rm o}_Z \in H_6(Z,{\Bbb Z})$ be the fundamental classes of $S$ and $Z$ respectively. By $d_Z(S) \in H^2(Z,{\Bbb Z})$ we denote the Poincar\'e dual of $i_\ast(\mbox{\rm o}_S) \in H_4(Z,{\Bbb Z})$, this means $i_\ast(\mbox{\rm o}_S) = d_Z(S) \smallfrown \mbox{\rm o}_Z$ (cap--product). For any cohomology class $\alpha \in H^2(Z,{\Bbb Z})$ we obtain by the associativity of cap--product $\alpha \smallfrown i_\ast(\mbox{\rm o}_S) = \alpha \smallfrown (d_Z(S) \smallfrown \mbox{\rm o}_Z) = (\alpha \smallsmile d_Z(S)) \smallfrown \mbox{\rm o}_Z$. The naturalness of cap--product implies $\alpha \smallfrown i_\ast(\mbox{\rm o}_S) = i_\ast(i^\ast(\alpha) \smallfrown \mbox{\rm o}_S)$. Therefore, we obtain a commutative diagram: \centerline{ $\begin{array}[t]{ccc} H^2(Z,{\Bbb Z}) & \stackrel{{\scriptstyle i^\ast}}{\longrightarrow} & H^2(S,{\Bbb Z})\\ & & \downarrow {\scriptstyle\smallfrown \mbox{\rm o}_S}\\ \Bigg\downarrow {\scriptstyle \smallsmile d_Z(S)} & & H_2(S,{\Bbb Z})\\ & & \downarrow {\scriptstyle i_\ast}\\ H^4(Z,{\Bbb Z}) & \stackrel{{\scriptstyle \smallfrown \mbox{\rm o}_Z}}{\longrightarrow} & H_2(Z,{\Bbb Z}) \end{array}$} Since, by Poincar\'e duality, the cap--product with $\mbox{\rm o}_Z$ is an isomorphism, we obtain $\ker(i^\ast) \subset \ker( \smallsmile d_Z(S))$. The description of the cohomology ring given above allows us to compute the kernel of the cup--product with the dual class $d_Z(S)$ of S. With the notation of Section \ref{prelim} the elements $x_1,\dots, x_n,\omega$ form a basis of the free ${\Bbb Z}$--module $H^2(Z,{\Bbb Z})$. The dual class of $S$ is $d_Z(S) = c_1(\fdb) = 2\omega + x_1 + \dots + x_n$. If we use $\omega x_1,\dots,\omega x_n,x_1^2$ as basis of $H^4(Z,{\Bbb Z})$ then the cup--product with $d_Z(S)$ is described by the $(n+1)\times(n+1)$--matrix: \[ \left( \begin{array}{rrrrr} 2&0&\dots&0&1\\ 0&2&&0&1\\ \vdots& & &\vdots &\vdots\\ 0& &2 &0&1\\ 0&\dots&0&2&1\\ -1&\dots&-1&-1&-2 \end{array} \right)\] whose determinant is equal to $2^{n-1}(n-4)$. If $n\ne 4$ we obtain the injectivity of the map $\alpha \mapsto \alpha \smallsmile d_Z(S)$ and thus of the restriction map $\Pic{Z} \hookrightarrow \Pic{S}$. If $n = 4$ it is easy to see that $\alpha \smallsmile d_Z(S) = 0$ if and only if $\alpha \in {\Bbb Z}\cdot d_Z(S) \subset H^2(Z,{\Bbb Z})$. To prove the injectivity of $\Pic{Z} \rightarrow \Pic{S}$ it remains, therefore, to show that $\fb{m}\otimes{\mathcal O}_S \cong {\mathcal O}_S$ implies $m = 0$. By adjunction we have $\fb{m}\otimes{\mathcal O}_S \cong \can{m}$. But $S$ is rational, hence $\Pic{S}$ is torsion free and $K_S\ncong {\mathcal O}_S$. Thus, $\can{m}\cong{\mathcal O}_S$ if and only if $m=0$. This proves the Lemma. \qed \begin{lemma}\label{conn} Let $Z$ be a simply connected compact twistor space of positive type and $D\subset Z$ a divisor of degree one. If $S\in|\fund|$ is a smooth surface, then $C:=D\cap S$ is connected. \end{lemma} {\sc Proof:\quad} We shall show $h^0({\mathcal O}_C)=1$, which implies connectedness of $C$. Consider first the exact sequence \begin{equation} \label{seqelm} 0 \rightarrow {\mathcal O}_Z(-\bar{D}) \rightarrow {\mathcal O}_Z \rightarrow {\mathcal O}_{\bar{D}} \rightarrow 0. \end{equation} From $h^1({\mathcal O}_Z)=0$ we obtain $h^1({\mathcal O}(-\bar{D}))=h^0({\mathcal O}_{\bar{D}}) - h^0({\mathcal O}_Z)=0$ since $Z$ and $\bar{D}$ are connected. As $D+\bar{D}\in |\fund|$ we obtain an exact sequence \[ 0 \rightarrow K(\bar{D}) \rightarrow \fdbd \rightarrow {\mathcal O}_D(-C) \rightarrow 0. \] But the degree of $\fdbd$ is $-2$ and, therefore, Hitchin's vanishing theorem gives $h^i(\fdbd)=0$ for all $i$. Therefore, using Serre duality, $h^1({\mathcal O}_D(-C))= h^2(K(\bar{D})) = h^1({\mathcal O}_Z(-\bar{D})) = 0$. Consider finally the exact sequence \[ 0 \rightarrow {\mathcal O}_D(-C) \rightarrow {\mathcal O}_D \rightarrow {\mathcal O}_C \rightarrow 0. \] We have $h^0({\mathcal O}_D)=1$ since any divisor of degree one is connected \cite{Po1}. Because $C$ is effective, $h^0({\mathcal O}_D(-C))$ must vanish. Hence, $h^0({\mathcal O}_C)=h^0({\mathcal O}_D)=1$. \qed \begin{lemma}[cf.\ \cite{PP}, p.\ 693]\label{noreal} Let $Z$ be as above and $S\in|\fund|$ an irreducible real divisor. Then $S$ is smooth and contains a real twistor fibre $F\subset S$. The linear system $|F|$ is one--dimensional and its real elements are precisely the real twistor fibres contained in $S$. \end{lemma} {\sc Proof:\quad} The smoothness of $S$ was shown in \cite[Lemma 2.1]{PP1}. If $S$ does not contain a real twistor fibre, the restriction of the twistor fibration to $S$ would give an unramified double cover over a simply connected manifold, since $Z$ does not contain real points. But $S$ is connected and must, therefore, contain a real twistor fibre $F$. From the adjunction formula we obtain $F^2 = 0$ on $S$. Hence, we have an exact sequence $0 \rightarrow {\mathcal O}_S \rightarrow {\mathcal O}_S(F) \rightarrow {\mathcal O}_F \rightarrow 0$. From $h^1({\mathcal O}_S) = 0$ we infer, therefore, $\dim |F| = 1$. Since the linear system $|F|$ defines a flat family of curves in $S$, its elements form a curve in the Douady space ${\mathcal D}$ of curves on $Z$ (cf.\ \cite{Dou}). Since $h^0({\Bbb P}^1,{\mathcal O}(1)\oplus{\mathcal O}(1))=4$ and $h^1({\Bbb P}^1,{\mathcal O}(1)\oplus{\mathcal O}(1))=0,\;\;{\mathcal D}$ is a four--dimensional complex manifold near points which correspond to smooth rational curves on $Z$ with normal bundle ${\mathcal O}(1)\oplus{\mathcal O}(1)$. The real structure of $Z$ induces one on ${\mathcal D}$. If the set of real points ${\mathcal D}({\Bbb R})$ is non--empty, then it is a four--dimensional real manifold near points as before. Since the real twistor lines are smooth rational curves with the above normal bundle, the real manifold $M=4\bbfC\bbfP^2$ is a submanifold of ${\mathcal D}({\Bbb R})$. Since $M$ is compact and has the same dimension as ${\mathcal D}({\Bbb R})$, it must be a connected component. The set $U$ of members of $|F|$ which are smooth rational curves with normal bundle ${\mathcal O}(1)\oplus{\mathcal O}(1)$ is open and dense in ${\Bbb P}^1\cong|F|$ with respect to the Zariski topology. Therefore, the set $U({\Bbb R})$ of real points in $U$ is open and dense in the one--sphere of real members of $|F|$. Since ${\mathcal D}({\Bbb R})$ is smooth near $M$ and $M$ is a component of ${\mathcal D}({\Bbb R})$, we have $U({\Bbb R})\subset M$. But $M$ is compact and must, therefore, contain the closure of $U({\Bbb R})$ in ${\mathcal D}({\Bbb R})$, which is the set of all real members of $|F|$. Therefore, any real member of $|F|$ is a real twistor fibre and, in particular, smooth and ireducible. This proves the claim.\qed To obtain more information on the structure of real irreducible fundamental divisors $S\in|\fund|$ one can study the morphism $S\rightarrow {\Bbb P}^1$ given by $|F|$ (cf.\ \cite[p.\ 693]{PP}). Since the general fibre of this morphism is a smooth rational curve it factors through a rational ruled surface. Since $(-K_S)^2 = (\fund)^3 = 8-2n$, the surface $S$ is a blow--up of a ruled surface at $2n$ points. The exceptional curves of these blow--ups are contained in fibres of the morphism $S\rightarrow {\Bbb P}^1$. By Lemma \ref{noreal} none of the exceptional curves is real and none of the blown--up points lie on a real fibre of the ruled surface. Using this, in \cite[Lemma 3.5]{PP} it has been shown that the ruled surface is isomorphic to $\bbfP^1\times\bbfP^1$. Therefore, we obtain a morphism $\sigma:S\rightarrow\bbfP^1\times\bbfP^1$ which is a succession of blow--ups. Let us equip $\bbfP^1\times\bbfP^1$ with the real structure given by the antipodal map on the first factor and the usual real structure on the second. Then $\sigma$ is equivariant (or ``real''). Since we can always contract a conjugate pair of disjoint $(-1)$--curves, $\sigma$ is the succession of $n$ blow--ups. At each step a conjugate pair of points is blown--up to give a surface without real points. We should bear in mind that it is possible to have infinitesimally near blown--up points. As in \cite{PP} we shall call curves of type $(1,0)$ on $\bbfP^1\times\bbfP^1$ ``lines'' and curves of type $(0,1)$ ``fibres''. Then there do not exist real lines. But the images of real twistor fibres in $|F|$ are exactly the real ``fibres''. \begin{lemma}\label{smocomp} Equip $\bbfP^1\times\bbfP^1$ with the real structure $((a_0:a_1),(b_0:b_1))\mapsto ((\bar{a}_1:-\bar{a}_0),(\bar{b}_0:\bar{b}_1))$ as described above. Then the reduced components of any \underline{real} member of $|{\mathcal O}(2,2)| = |-K_{\bbfP^1\times\bbfP^1}|$ are smooth. A non--reduced component of a real member of $|{\mathcal O}(2,2)|$ can only be of the form $2F$ with a real curve $F\in|{\mathcal O}(0,1)|$. \end{lemma} {\sc Proof:\quad} As usual, ${\mathcal O}(k,l)$ denotes the locally free sheaf $p_1^\ast{\mathcal O}_{{\Bbb P}^1}(k) \otimes p_2^\ast{\mathcal O}_{{\Bbb P}^1}(l)$ on the smooth rational surface $\bbfP^1\times\bbfP^1$, where $p_i : \bbfP^1\times\bbfP^1 \rightarrow {\Bbb P}^1$ ($i =1,2$) are the projections and $k, l$ are integers. The Picard group $\Pic(\bbfP^1\times\bbfP^1)$ is free abelian of rank two with generators ${\mathcal O}(1,0)$ and ${\mathcal O}(0,1)$. In the proof we shall use the well--known fact that, if $k<0$ or $l<0$, then the linear system $|{\mathcal O}(k,l)|$ is empty. Let $C\in|{\mathcal O}(2,2)|$ be a real curve and $C_0\in |{\mathcal O}(a,b)|$ an {\sl irreducible\/} component (with {\sl reduced} scheme structure) of $C$. Let $\lambda\geq 1$ be the multiplicity of $C_0$ in $C$, that is the largest integer with $\lambda C_0\subset C$. Then we must have $0\leq \lambda a\leq 2$ and $0\leq \lambda b\leq 2$. The case $\lambda a = \lambda b = 2$ can only occur if $\lambda C_0 = C$. Assume first $\lambda C_0\ne C$, hence $\lambda a \leq 1$ or $\lambda b \leq 1$. If $\lambda\ge 2$ or $C_0$ singular, there exists a point $y\in C_0$ such that any curve not contained in $C_0$ but containing $y$ has intersection number at least two with $\lambda C_0$. If $F\in |{\mathcal O}(0,1)|$ and $G\in|{\mathcal O}(1,0)|$ are the unique curves in these linear systems containing the point $y$, we obtain (as $C_0$ is irreducible) $F.(\lambda C_0) = \lambda a\ge 2$ or $G.(\lambda C_0) = \lambda b\ge 2$. By the above inequalities, this means $\lambda a =2$ and $0\leq\lambda b\leq 1$ or $0\leq\lambda a\leq 1$ and $\lambda b =2$. If $\lambda = 1$, the curve $C_0$ is, by assumption, irreducible, reduced, singular and a member of $|{\mathcal O}(2,0)|$, $|{\mathcal O}(2,1)|$, $|{\mathcal O}(0,2)|$, or $|{\mathcal O}(1,2)|$. But these linear systems do not contain such a curve. Hence, we must have $\lambda = 2$ and, therefore, $C_0\in|{\mathcal O}(0,1)|$ or $C_0\in|{\mathcal O}(1,0)|$. In particular, $C_0$ is smooth. If $C_0\in|{\mathcal O}(1,0)|$, this curve is not real, since by definition of the real structure on $\bbfP^1\times\bbfP^1$ this linear system does not contain real members. Hence, the component $2C_0$ of $C$ is not real, which implies $2C_0 + 2\bar C_0 \subset C$, since $C$ is real. But $\bar C_0 \in|{\mathcal O}(1,0)|$ and so $2C_0 + 2\bar C_0 \in |{\mathcal O}(4,0)|$. Such a curve can never be contained in $C\in|{\mathcal O}(2,2)|$. So we obtain $\lambda C_0 = 2F$ with some $F\in|{\mathcal O}(0,1)|$. Again, since a curve of type $(0,4)$ can never be a component of $C$, the fibre $F$ is necessarily real. This proves the lemma in the case $\lambda C_0 \ne C$. Assume now $\lambda C_0=C$. Then $\lambda = 1$ or $\lambda = 2$. If $\lambda=2$, we have, by reality of $C = 2C_0$, that $C_0\in|{\mathcal O}(1,1)|$ is a real curve. Because $C_0$ is irreducible and $C_0.F =1$ for any $F\in|{\mathcal O}(0,1)|$, the curve $C_0$ would intersect each real fibre $F\in|{\mathcal O}(0,1)|$ at a real point. But on $\bbfP^1\times\bbfP^1$ real points do not exist. Hence $\lambda=1$, which means $C = C_0$ is irreducible and real. It remains to see that $C$ must be reducible if it is not smooth. Let $x\in C$ be a singular point of $C$. Since $C$ is real and $\bbfP^1\times\bbfP^1$ does not contain real points, $\bar{x}\ne x$ is also a singular point on $C$. If we embed $\bbfP^1\times\bbfP^1$ by $|{\mathcal O}(1,1)|$ as a smooth quadric into ${\Bbb P}^3,$ we easily see that the linear system of curves of type $(1,1)$ on $\bbfP^1\times\bbfP^1$ containing $x$ and $\bar{x}$ is one--dimensional. It is cut out by the pencil of planes in ${\Bbb P}^3$ containing the line connecting $x$ and $\bar{x}$. Therefore, any point of $\bbfP^1\times\bbfP^1$ is contained in such a curve. The intersection number of $C$ with a curve of type $(1,1)$ is four. Since $x$ and $\bar{x}$ are singular points on $C$, any curve of type $(1,1)$ containing $x$ and $\bar{x}$ and a third point of $C$ must have a common component with $C$. Therefore, $C$ cannot be irreducible and reduced.\qed \begin{defi} A reduced curve $C$ on a compact complex surface $S$ will be called a ``cycle of rational curves'', if the irreducible components $C_1,\dots, C_m$ of $C$ are smooth rational curves with the following properties: (We use the convention $C_{m+1} = C_1$.)\\ $m = 2$ and $C_1$ intersects $C_2$ transversally at two distinct points, $C_1.C_2 = 2$, or\\ $m\geq 3$, $C_i.C_{i+1} = 1$ and $C_i\cap C_j \ne \emptyset$ implies $j\in\{i-1, i, i+1\}$. \end{defi} \begin{prop}\label{types} Assume $\dim |\fund|\geq 1$ and let $S\in|\fund|$ be smooth and real. Then there exists a blow-down $S\rightarrow \bbfP^1\times\bbfP^1$ and a connected real member $C\in|-K_S|$, such that: \begin{itemize} \item $\sigma$ is compatible with real structures, where we use the real structure of Lemma \ref{smocomp} on $\bbfP^1\times\bbfP^1$, \item the composition $\mbox{pr}_2\circ \sigma$ of $\sigma$ with the second projection $\bbfP^1\times\bbfP^1\longrightarrow {\Bbb P}^1$ is the morphism given by the linear system $|\widetilde{F}|$, where $\widetilde{F}\subset S$ is a real twistor fibre, \item the curve $C$ is reduced and \item if $C$ is not smooth, it is a ``cycle of rational curves'' and its image $C'$ in $\bbfP^1\times\bbfP^1$ has one of the following structures: \end{itemize} \renewcommand{\labelenumi}{(\Roman{enumi})} \begin{enumerate} \item \mbox{ \begin{minipage}[t]{10.5cm} $C'$ has four components $C'=F+\bar{F}+G+\bar{G}$ where $F\in|{\mathcal O}(0,1)|$ is a non--real fibre and $G\in|{\mathcal O}(1,0)|$ is a line. \end{minipage} \begin{picture}(60,-60)(-15,-5) \put(0,-40){\line(1,0){60}} \put(30,-42){\makebox(0,0)[t]{$\bar{G}$}} \put(0,0){\line(1,0){60}} \put(30,2){\makebox(0,0)[b]{$G$}} \put(10,10){\line(0,-1){60}} \put(12,-20){\makebox(0,0)[l]{$F$}} \put(50,10){\line(0,-1){60}} \put(52,-20){\makebox(0,0)[l]{$\bar{F}$}} \end{picture}\rule[-60pt]{0pt}{60pt} } \item \begin{minipage}[t]{10.5cm} $C'$ has two components $C'=F+C_0$ where $F\in|{\mathcal O}(0,1)|$ is a real fibre and $C_0\in|{\mathcal O}(2,1)|$ is real, smooth and rational. \end{minipage} \begin{picture}(60,-60)(-15,-5) \put(45,10){\line(0,-1){60}} \put(47,-20){\makebox(0,0)[l]{$F$}} \put(50,-20){\oval(55,40)[l]} \put(18,-20){\makebox(0,0)[r]{$C_0$}} \end{picture}\rule[-60pt]{0pt}{60pt} \item \begin{minipage}[t]{13.8cm} $C'$ has two distinct components $C'=A'+\bar{A}'$ where $A',\bar{A}'\in |{\mathcal O}(1,1)|.$\\ \mbox{\rm [By Corollary \ref{omit} this item can be omitted if $n=4$ and $\fdb$ is not nef!]} \end{minipage} \end{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})} \end{prop} {\sc Proof:\quad} As $\dim|-K_S| = \dim|\fund| - 1$, we have by assumption $|-K_S|\ne\emptyset$. As seen above, we can choose a real blow--down map $\sigma:S\rightarrow\bbfP^1\times\bbfP^1$ such that $(\mbox{pr}_2\circ\sigma)^\ast{\mathcal O}_{{\Bbb P}^1}(1) ={\mathcal O}_S(\widetilde{F})$. If $|-K_S|$ contains a smooth member, we are done. Otherwise, take a reducible real $C\in|-K_S|$ and let $C'\subset\bbfP^1\times\bbfP^1$ be the image of $C$. Since $\sigma$ is a blow--up, $C'$ is a real member of $|{\mathcal O}(2,2)|$. By Lemma \ref{smocomp} the components of $C'$ are smooth and a multiple component can only be the multiple $2F$ of a real fibre $F\in|{\mathcal O}(0,1)|$. But Lemma \ref{noreal} shows that no point on such a real fibre is blown--up. Therefore, any other member of $|2F|$ missing the $2n$ blown--up points, defines a divisor in $|-K_S|$. Choosing, for example, a conjugate pair of appropriate fibres, we obtain a real member in $|-K_S|$ whose image in $\bbfP^1\times\bbfP^1$ has only reduced components. We assume for the rest of the proof that $C$ is chosen in this way. If $C'$ would be irreducible, it would be smooth by Lemma \ref{smocomp}. In this case $C$ is smooth, too. Assume $C'$ is reducible. A component of type $(1,a)$ with $a\in\{0,1,2\}$ cannot be real, since, otherwise, it would intersect real fibres at real points. Therefore, such components appear in conjugate pairs, hence $a\le 1$. If $a=0$ we are in case (I). If $a\ne 0$ then $a=1$ and $C'=A'+\bar{A}'$ with two distinct curves $A'$, $\bar{A}'$ of type $(1,1)$. This is case (III). Assume now that there is no component of type $(1,a)$. Then we must have a component $C_0$ of $C'$ which has type $(2,a)$. If $a=2$ it must be smooth and we are done. Therefore, $a=1$, because $|{\mathcal O}(2,0)|$ does not contain irreducible reduced elements. Then we have $C'=C_0+F$ with $F\in|{\mathcal O}(0,1)|$ and $C_0\in|{\mathcal O}(2,1)|$. $F$ and $C_0$ must be real, since they have different types. This is case (II) of our statement. It remains to show that $C$ is a ``cycle of rational curves''. We have seen this for the image $C'$ in $\bbfP^1\times\bbfP^1$. Exceptional components of $C$ are always rational. Furthermore, $C'$ has at most ordinary nodes as singularities. To obtain $C$ from $C'$, at every step of blowing--up, we have to subtract the exceptional locus from the total transform of $C'$. At every step we blow up either a conjugate pair of singular points or of smooth points. We obtain a curve which has again, at most, singularities of multiplicity two and is a ``cycle of rational curves''. So we obtain this property for $C$, too.\qed Using this structure result and assuming that the fundamental linear system $|\fund|$ is a pencil, we can show that the structure of its base locus is closely related to the effective divisors of degree one on $Z$. \begin{prop}\label{basel} Assume $\dim|\fund| = 1$ and denote by $C$ the base locus of the fundamental linear system.\\ If $C$ is smooth, then $Z$ does not contain effective divisors of degree one.\\ If $C$ is not smooth, then the number of effective divisors of degree one on $Z$ is equal to the number of components of $C$. \end{prop} {\sc Proof:\quad} Let $S\in|\fund|$ be a smooth real member. Then $|-K_S| = \{C\}$ and by Proposition \ref{types} $C$ is smooth or a ``cycle'' of smooth rational curves. If $C$ is smooth, there does not exist an effective divisor $D$ of degree one, because $D + \bar{D}\in|\fund|$ would produce a reducible member in $|-K_S|$. Let now $C$ be singular, hence reducible. The rest of the proof is an adaption of an idea of Pedersen and Poon \cite[p.\ 700]{PP}. Now let $\{P,\bar{P}\}$ be any pair of singular points on $C$. The image $C'$ of $C$ in $\bbfP^1\times\bbfP^1$ does not contain a real fibre, because, otherwise, by Proposition \ref{types} and Lemma \ref{noreal} the linear system $|-K_S|$ would be at least one--dimensional. Hence, the real twistor line $L_P$ containing $P$ and $\bar{P}$ is not contained in $S$. Hence, $L_P$ meets $S$ transversally at $P$ and $\bar{P}$. If $Q$ is a point on $L_P$ distinct from $P$ and$\bar{P}$, then there exists a divisor $S_0\in |\fund|$ containing $Q$. Since $S_0$ contains also $C$ it contains three points of $L_P$. Hence, $L_P\subset S_0$. Therefore, the real linear system of fundamental divisors containing $L_P$ is non--empty. This implies that we can choose a real $S_0 \in |\fund|$ containing $L_P$. Since $S_0$ contains also $C$ and $P$ is a singular point of $C$, the surface $S_0$ contains three curves meeting at $P$, namely $L_P$ and two components (call them $A$ and $B$) of $C$. On the other hand, $L_P$ intersects $S$ precisely at $P$ and $\bar{P}$ as we have seen above. From $L_P.S=2$ we infer that this is a transversal intersection. But $A$ and $B$ are contained in $S$ and are transversally there. We can conclude that the tangent space of $Z$ at $P$ is generated by the tangent directions of $A, B$ and $L_P$ at $P$. Hence, the real surface $S_0$ is singular at $P$. This implies that $S_0$ is singular along $L_P$ (cf. \cite[p.141]{Hit2}) and by \cite[Lemma 2.1]{PP1} such a divisor splits into the sum of two divisors of degree one. Therefore, we have at least as many pairs of conjugate divisors of degree one as we have pairs of conjugate singular points on $C$. In other words, the number of distinct divisors of degree one is at least equal to the number of components of $C$. Let $D$ and $\bar{D}$ be a conjugate pair of divisors of degree one on $Z$. Then $C\subset D\cup\bar{D}$. $D\cap \bar{D}$ is a real twistor line (cf.\ \cite[Prop. 2.1]{Ku}), and no component of $C$ is a real twistor line. Hence, every component of $C$ lies on exactly one of the surfaces $D$ and $\bar{D}$. By Lemma \ref{conn} $C\cap D$ is connected. The same is true for the conjugate curve $C\cap \bar{D}$. Since $C$ is a cycle of rational curves, $(C\cap D)\cap(C\cap\bar{D})$ consists of a conjugate pair $\{P,\bar{P}\}$ of singularities of $C$. Since $D$ and $\bar{D}$ are of degree one, the real twistor line $L_P$ containing $P$ and $\bar{P}$ must be contained in $D$ and in $\bar{D}$. Therefore, $D\cap\bar{D}=L_P$. Let $D'$ be an arbitrary divisor of degree one containing $L_P$. Then $D'\cap\bar{D}'=L_P$ and without loss of generality we may assume $D'\cap C=D\cap C$, since the decomposition of $C$ into two conjugate connected curves is determined by $\{P, \bar{P}\}=L_P\cap C$. By Lemma \ref{pic} the restriction map $\Pic{Z}\rightarrow\Pic{S}$ is injective. Since $(D'+\bar{D}')\cap S=C$ we have $D'\cap S=D'\cap C$ and $D\cap S=D\cap C$. Hence, we have ${\mathcal O}_Z(D)\cong {\mathcal O}_Z(D')$, which means that $D$ and $D'$ are linearly equivalent. If $D\ne D'$, then $\dim|D|\ge 1$ and $Z$ would contain infinitely many divisors of degree one and by Kurke \cite{Ku} and Poon \cite{Po2} it must be a conic--bundle twistor space. But then we should have $\dim|\fund|=3$ in contradiction to our assumption. Hence, $D=D'$ and we have exactly as many divisors of degree one on $Z$ as $C$ has components.\qed For technical reasons we state here the following lemma needed in Section \ref{drei}. \begin{lemma}\label{normal} Let $S$ be a smooth complex surface and $C\subset S$ a reduced curve. Assume $C =\sum_{i=1}^{m} C_i$ is a ``cycle of rational curves'' as defined above. If $L\in\Pic(S)$ is a line bundle, we define $l_i := L.C_i$. Let $I_\pm := \{i | \pm l_i>0\}$ and $C_\pm := \sum_{i\in I_\pm} C_i$. Let $\gamma$ denote the number of connected components of $C\setminus C_-$. Assume $|I_-|\geq 2$ and each connected component of $C\setminus C_-$ contains a component of $C_+$. Then we have: \[h^0(C,L) = \sum_{i\in I_+} l_i - \gamma.\] \end{lemma} {\sc Proof:\quad} Let $\eta:\tilde C = \sqcup_i C_i \rightarrow C$ be the normalization of $C$. By $P_i$ we denote the intersection point of $C_i$ with $C_{i+1}$ $(1\leq i \leq m)$. By assumption $m\geq 3$. Tensoring the exact sequence $0\rightarrow {\mathcal O}_C \rightarrow \eta_\ast {\mathcal O}_{\tilde C} \rightarrow \oplus_i {\Bbb C}_{P_i} \rightarrow 0$ with $L$ yields the exact sequence $0 \rightarrow L\otimes {\mathcal O}_C \rightarrow \eta_\ast\eta^\ast(L\otimes{\mathcal O}_C) \rightarrow \oplus_i {\Bbb C}_{P_i} \rightarrow 0$. Hence, $H^0(C, L\otimes{\mathcal O}_C) \cong \ker(\oplus_i H^0(C_i, L_i) \stackrel{\rho}{\rightarrow} \oplus_i {\Bbb C}_{P_i})$. Here we denote $L_i:= L\otimes{\mathcal O}_{C_i} \cong {\mathcal O}_{C_i}(l_i)$. Let $P'_i\in C_i$ and $P''_i\in C_{i+1}$ be the two points on $\tilde C$ lying over $P_i$. To describe $\rho$ we observe that the map $\eta$ gives isomorphisms $L_i(P'_i) \stackrel{\sim}{\rightarrow} {\Bbb C}_{P_i}$ and $L_{i+1}(P''_i) \stackrel{\sim}{\rightarrow} {\Bbb C}_{P_i}$. If $s_i\in H^0(C_i, L_i)$ is a section, we denote by $s_i(P_i)$ the image of $s_i$ under the map $H^0(C_i, L_i) \rightarrow L_i(P'_i) \stackrel{\sim}{\rightarrow} {\Bbb C}_{P_i}$. Similarly, $s_i(P_{i-1})$ is the image of $s_i$ under $H^0(C_i, L_i) \rightarrow L_i(P''_{i-1}) \stackrel{\sim}{\rightarrow} {\Bbb C}_{P_{i-1}}$. With this notation we have: \[ \rho(s_1,\dots, s_m) = (s_1(P_1) - s_2(P_1), s_2(P_2) - s_3(P_2),\dots, s_m(P_m) - s_1(P_m)).\] Since $P'_i \ne P''_{i-1}$ on $C_i \cong {\Bbb P}^1$, the restriction of $\rho$ $H^0(C_i, L_i) \rightarrow {\Bbb C}_{P_i}\oplus{\Bbb C}_{P_{i-1}}$ is surjective if and only if $l_i > 0$. If $C_i + \cdots + C_{i+r}$ is a connected component of $C\setminus C_-$, then we obtain by induction on $r$ that the restriction of $\rho$ $\oplus_{\mu = 0}^r H^0(C_{i+\mu}, L_{i+\mu}) \rightarrow \oplus_{\mu = -1}^r {\Bbb C}_{P_{i+\mu}}$ is surjective. Because $H^0(C_i, L_i) = 0$ if and only if $l_i < 0$, we obtain $\im(\rho) = \sum_{P_\mu\in C_0+C_+} {\Bbb C}_{P_\mu}$. (Here, we denote $I_0 := I \setminus (I_- \cup I_+)$ and $C_0 := \sum_{\nu\in I_0} C_\nu$.) The number of points $P_\mu\in C_0+C_+$ is equal to $|I_0| + |I_+| + \gamma$. Therefore, we obtain \[ \dim\ker(\rho) = \sum_i h^0(C_i, L_i) - (|I_0| + |I_+| +\gamma) = \sum_{l_i\geq 0} (l_i + 1) - |I_0| - |I_+| -\gamma = \sum_{l_i > 0} l_i - \gamma.\] \qed \section{The nef case} \label{zwei} For the rest of the paper we assume $n = 4$. Remember that $(\fund)^3 = 0$, $\chi(\fb{m}) = m+1$ and $h^0(\fdb) \geq 2$ in this case. Remember from Mori's theory that a line bundle $L\in\Pic(Z)$ is called {\em nef}, if for each irreducible curve $C\subset Z$ there holds $L.C\geq 0$. \begin{thm}\label{nef} The following properties are equivalent: \begin{enumerate} \item $\fdb$ is nef; \item for all smooth and real $S\in|\fund|$ and all $C\in|-K_S|$, every component $C_0$ of $C$ has the property $C_0.(-K_S)=0$; \item there exists a smooth and real $S\in|\fund|$ and a divisor $C\in|-K_S|$, such that all components $C_0$ of $C$ have the property $C_0.(-K_S)=0$. \end{enumerate} If $\fdb$ is nef, then $a(Z)\le 2$ and $\dim|\fund|\leq 2$.\\ If $\fdb$ is nef and $\dim|\fund| = 2$, then $a(Z) = 2$, $|\fund|$ does not have base points and for any smooth real $S\in|\fund|$ the pencil $|-K_S|$ contains a smooth real member. \end{thm} {\sc Proof:\quad} (i)$\Rightarrow$(ii):\\ Take any smooth real $S\in|\fund|$ and an arbitrary curve $C\in|-K_S|$. Since $C.(\fund)=0$ and $\fdb$ is nef we obtain (ii). (ii)$\Rightarrow$(iii) is obvious. (iii)$\Rightarrow$(i):\\ If $\fdb$ were not nef, then there would exist an irreducible curve $C_0\subset Z$ with $C_0.(\fund)<0$. If $S\in|\fund|$ is smooth and real, then $C_0\subset S$ and $C_0.(-K_S)=C_0.(\fund)<0$. Therefore, $C_0$ is a component of any element of $|-K_S|$ in contradiction to (iii). Assume for the rest of the proof that $\fdb$ is nef. Let $S\in|\fund|$ be smooth and real and $C\in|-K_S|$ a real member. If $C$ is smooth, we have shown in \cite{CK} that $a(Z)\leq 2$ and $\dim|\fund|\leq2$. Assume $C$ is not smooth. To compute the algebraic dimension consider the exact sequences \[0 \rightarrow \fb{m-1} \rightarrow \fb{m} \rightarrow \can{m} \rightarrow 0\] and \[0 \rightarrow \can{(m-1)} \rightarrow \can{m} \rightarrow N^{\otimes m} \rightarrow 0\] with $N:=\canS\otimes{\mathcal O}_C$. Since $\fdb$ is nef, $(-K_S).C_i = 0$ for any component $C_i$ of $C$. But $C_i \cong {\Bbb P}^1$ and so $N\otimes{\mathcal O}_{C_i} \cong {\mathcal O}_{C_i}$. This does not imply in general $N \cong {\mathcal O}_C$, because $C$ is a ``cycle'' of rational curves. But we obtain $h^0(N^{\otimes m}) = 1$ if $N^{\otimes m} \cong {\mathcal O}_C$ and $h^0(N^{\otimes m}) = 0$ if $N^{\otimes m} \ncong {\mathcal O}_C$. As in \cite{CK} this implies $a(Z)\leq 2$ and $h^0(\fdb) = h^0({\mathcal O}_Z) + h^0(\canS) = 1 + h^0({\mathcal O}_S) + h^0(N) \leq 3$. Assume now $\dim|\fund| = 2$. Hence, using the Riemann--Roch formula we have $h^1(\fdb) = 1$. By Proposition \ref{tau} this implies $a(Z)\geq 2$. From the above considerations we obtain $h^0(N) = 1$, hence $\canS\otimes{\mathcal O}_C \cong N \cong {\mathcal O}_C$. (The same is true if $C$ is smooth, cf.\ \cite{CK}.) The exact sequence $0 \rightarrow {\mathcal O}_S \rightarrow \canS \rightarrow N \rightarrow 0$ and $h^1({\mathcal O}_S) = 0$ give a surjective restriction map $H^0(\canS) \twoheadrightarrow H^0({\mathcal O}_C) \cong {\Bbb C}$. Because $C\in|-K_S|$, this shows that $|-K_S|$ does not have base points. Since $\dim|-K_S| = \dim|\fund| -1 = 1$, Bertini's Theorem \cite[I \S 1]{GH} states the existence of a smooth member in $|-K_S|$. Hence, the generic divisor in $|-K_S|$ is smooth and so the generic real member, too. On the other hand, we know from $h^1({\mathcal O}_Z) = 0$ that the restriction map $|\fund| \twoheadrightarrow |-K_S|$ is surjective. From the freeness of $|-K_S|$ we conclude that $|\fund|$ does not have base points.\qed \begin{rem}\label{smnef} If there exists a smooth real $S\in|\fund|$ and a smooth curve $C\in|-K_S|$, then $\fdb$ is nef. This is clear from the theorem, because $C.(\fund) = (\fund)^3 = 0$. \end{rem} \begin{cor}\label{omit} In Proposition \ref{types} we can omit case (III) if $\fdb$ is not nef. \end{cor} {\sc Proof:\quad} Assume $C'=A'+\bar{A}'$ as in the proof of Proposition \ref{types}. $A'$ and $\bar{A}'$ intersect at a pair of conjugate points, say $P$ and $\bar{P}$. If $\sigma$ does not blow up $P$ and $\bar{P}$, then, by reality of the blown--up set, on (the strict transforms of) $A'$ and $\bar{A}'$ exactly four points are blown--up. If we denote by $A$ and $\bar{A}$ the strict transforms of $A'$ and $\bar{A}'$ in $S$, then we have $C=A+\bar{A}$ and $A^2=\bar{A}^2=-2$. Since $A$ and $\bar{A}$ are rational we obtain, by the adjunction formula, $A.(-K_S)=\bar{A}.(-K_S)=0$. By Theorem \ref{nef} this implies that $\fdb$ is nef. If $\sigma$ blows up $P$ and $\bar{P}$, then we perform an elementary transform to arrive at case (I) as follows. Let $\sigma_1:S^{(1)}\rightarrow\bbfP^1\times\bbfP^1$ be the blow--up of $P$ and $\bar{P}$, then we have an induced real structure on $S^{(1)}$. Since $A'$ intersects any fibre at exactly one point, $P$ and $\bar{P}$ lie on a conjugate pair of fibres. The strict transforms in $S^{(1)}$ of these fibres form a conjugate pair of disjoint $(-1)$--curves. Contracting them we obtain a blow--down map $\sigma_1':S^{(1)}\rightarrow\bbfP^1\times\bbfP^1$ which is again compatible with real structures. If we denote by $E$ and $\bar{E}$ the exceptional curves of $\sigma_1$, then the image of $C$ in $S^{(1)}$ is $\sigma_1^\ast(A'+\bar{A}')-E-\bar{E} = A^{(1)}+\bar{A}^{(1)}+E+\bar{E}$. Here $A^{(1)}$ and $\bar{A}^{(1)}$ are the strict transforms of $A'$ and $\bar{A}'$. The morphism $\sigma_1'$ maps this curve onto a curve of type (I).\qed \section{The non--nef case} \label{drei} Throughout this section we assume $\fdb$ to be not nef and $n=4$.\\ By Theorem \ref{nef}, Remark \ref{smnef} we know that in any smooth real $S\in|\fund|$ the anticanonical system $|-K_S|$ contains only reducible elements. By Proposition \ref{types} and Corollary \ref{omit} we can, therefore, choose a real blow--down $S\rightarrow\bbfP^1\times\bbfP^1$ and a real reduced curve $C\in|-K_S|$, whose image $C'$ in $\bbfP^1\times\bbfP^1$ is of type (I) or (II) as described there. Observe that $C'$ has type (I) if and only if $C$ contains a real irreducible component. \begin{prop}\label{notnefi} If there exists a real irreducible curve intersecting $|\fdb|$ negatively, then $h^0(\fdb)=3$ and $a(Z) = 3$.\\ There exists a unique irreducible curve $C_0$ with $C_0.(\fund) < 0$. This curve is real, smooth and rational and $C_0.(\fund)=-2$. The base locus of $|\fund|$ is exactly $C_0$. $Z$ does not contain divisors of degree one. \end{prop} {\sc Proof:\quad} Let $S\in|\fund|$ be smooth and real and choose $C\in|-K_S|$ and $\sigma: S\rightarrow \bbfP^1\times\bbfP^1$ with the properties of Proposition \ref{types}. Because any irreducible curve intersecting $\fdb$ negatively is contained in $C$, this curve has a real component. Therefore, the image $C'$ of $C$ in $\bbfP^1\times\bbfP^1$ is of type (II). Let $C'=C_0'+F'$ be the decomposition of $C'$. By Lemma \ref{noreal} none of the blown--up points lie on the real fibre $F'$. In particular, only smooth points of $C_0'$ are blown--up. Hence, $C=C_0+F$ where $C_0$ and $F$ are the strict transforms of $C_0'$ and $F'$ respectively. Therefore, the eight blown--up points lie on $C_0'$ which implies $C_0^2 = {C'}_0^2-8=-4$. By adjunction formula we obtain $-2=C_0.(-K_S)=C_0.(\fund)$. Hence, $|-K_S|=C_0+|F|$ and we obtain: $\dim|-K_S|=1$ and $C_0$ is the base locus of $|-K_S|$. Since $h^1({\mathcal O}_Z)=0$ the restriction map $H^0(\fdb)\twoheadrightarrow H^0(\canS)$ is surjective. Hence, the linear system $|\fund|$ has dimension two and its base locus is precisely $C_0$ (with multiplicity one). $C_0$ is the unique irreducible curve in $Z$ having negative intersection number with $\fund$, since any other such curve should be contained in the base locus of $|\fund|$. If $Z$ contains a divisor $D$ of degree one, then $D+\bar{D}\in|\fund|$. If $D_0$ ($\bar{D}_0$ respectively) denotes the restriction of $D$ to $S$, then $D_0+\bar{D}_0\in|-K_S|=C_0+F$. In the proof of Lemma \ref{noreal} we have seen that the real elements of $|F|$ are irreducible. Therefore, any real element of $|-K_S|$ consists of two distinct real irreducible curves with multiplicity one. This shows that $|-K_S|$ cannot contain a member of the form $D_0+\bar{D}_0$. It remains to show that the {\bf algebraic dimension} of Z must be three in this case. Let $\sigma:\tilde{Z}\rightarrow Z$ be the blow--up of the smooth rational curve $C_0$. By $E\subset \tilde{Z}$ we denote the exceptional divisor. Then we obtain a morphism $\pi:\tilde{Z}\rightarrow{\Bbb P}^2$ defined by the linear system $|\fund|$ such that $\pi^\ast{\mathcal O}(1)\cong \sigma^\ast\fdb\otimes{\mathcal O}_{\tilde{Z}}(-E)$. Since the restriction map $|\fund|\rightarrow|-K_S|$ is surjective, the restriction $\pi_{|S}$ is given by the linear system $|-K_S|=C_0+|F|$. This means that $\pi$ exibits $S$ as the blow--up of a ruled surface and $\pi(S)$ is a line in ${\Bbb P}^2$. Since $\pi(\tilde{Z})$ is not contained in a linear subspace, $\pi$ must be surjective. If we equip ${\Bbb P}^2$ with the usual real structure, $\pi$ becomes compatible with real structures since the linear system $|\fund|$ and the blown--up curve $C_0$ are real. Since $Z$ does not contain divisors of degree one, any real fundamental divisor $S$ is irreducible and, therefore, smooth. By $\tilde{S}\subset\tilde{Z}$ we denote the strict transform of $S\in |\fund|$. Since $C_0$ is a smooth curve in a smooth surface, $\sigma:\tilde{S}\rightarrow S$ is an isomorphism. Furthermore, $E\cap\tilde{S}$ will be mapped isomorphically onto $C_0\subset S$. Since $F.C_0=2$ and the restriction of $\pi$ onto $\tilde{S}$ is the map defined by the linear system $|F|$, the restriction of $\pi$ exibits $E\cap\tilde{S}$ as a double covering over $\pi(S)\cong{\Bbb P}^1$. Since real lines cover ${\Bbb P}^2$ the morphism $\pi:E\rightarrow{\Bbb P}^2$ does not contract curves and is of degree two. Since generic fibres of $\pi$ are smooth rational curves, the line bundle ${\mathcal O}_{\tilde{Z}}(E)$ restricts to ${\mathcal O}_{{\Bbb P}^1}(2)$ on such fibres. Hence, after replacing (if necessary) ${\Bbb P}^2$ by the open dense set $U$ of points having smooth fibre, the adjunction morphism $\pi^\ast\pi_\ast {\mathcal O}_{\tilde{Z}}(E) \rightarrow {\mathcal O}_{\tilde{Z}}(E)$ is surjective. This defines a $U$--morphism $\Phi:\tilde{Z} \rightarrow {\Bbb P}(\pi_\ast{\mathcal O}_{\tilde{Z}}(E))$, where $\pi_\ast{\mathcal O}_{\tilde{Z}}(E)$ is a locally free sheaf of rank three. The restriction of $\Phi$ to smooth fibres coincides with the Veronese embedding ${\Bbb P}^1\hookrightarrow{\Bbb P}^2$ of degree two. Therefore, the image of $\Phi$ is a three--dimensional subvariety of the ${\Bbb P}^2$--bundle ${\Bbb P}(\pi_\ast{\mathcal O}_{\tilde{Z}}(E))\rightarrow U$. Hence, $\tilde{Z}$ is bimeromorphically equivalent to a quasiprojective variety and has, therefore, algebraic dimension three.\qed For the rest of this section we assume that there does not exist a {\em real} irreducible curve contained in the base locus of $|\fund|$. We keep the assumptions $n=4$ and $\fdb$ is not nef. In this situation, we obtain: \begin{lemma}\label{sub} \begin{enumerate} \item[(a)] If $A\subset Z$ is an irreducible curve, then $A.(\fund) \geq -2$. \item[(b)] If $A\subset Z$ is an irreducible curve with $A.(\fund) <0$, then there exists at least a one--parameter family of real smooth divisors $S\in|\fund|$, containing a curve $C\in|-K_S|$ and possessing a birational morphism $\sigma: S\longrightarrow \bbfP^1\times\bbfP^1$ as in Proposition \ref{types}, such that moreover:\\ $A$ and $\overline{A}$ are components of $C$ and for twistor fibres $F\subset S$ we have $F.A = F.\overline{A} = 1$.\\ In particular, the image $A'$ of $A$ in $\bbfP^1\times\bbfP^1$ is a ``line'', that means $A'\in|{\mathcal O}(1,0)|$. \end{enumerate} \end{lemma} {\sc Proof:\quad} Our assumptions imply that $C'$ is a curve of type (I) in Proposition \ref{types}. The components of $C'$ are curves in $\bbfP^1\times\bbfP^1$ with self--intersection number zero. They are not real. Hence, after a succession of four blow--ups of a conjugate pair of points, each component $A$ of $C$ fulfills $A^2\geq-4$ in $S$. The adjunction formula, together with the rationality of $A$, implies $A.(\fund) = A.(-K_S) = A^2 + 2 \geq -2$. Because a curve $A$ with $A.(\fund) < 0$ must be a component of $C$, the assertion (a) is shown. Let now $A\subset Z$ be an irreducible curve with $A.(\fund)<0$. Then we have $A\subset C$. Let $x\in A\subset C$ be a smooth point of $C$ and $x\in F \subset Z$ a twistor fibre. Since $|\fund|$ is at least a pencil, there exists a divisor $S\in|\fund|$ containing a given point $y\in F\setminus\{x,\overline{x}\}$. Because $F.S = 2$ and $S\cap F \supset \{y, x, \overline{x}\}$ the twistor fibre $F$ is contained in $S$. So the real subsystem $|\fund|_F\subset |\fund|$ of divisors containing $F$ is not empty. Hence, we can choose a real smooth $S\in|\fund|$ containing $F$. By construction, we have $F.A = F.\overline{A} \geq 1$. But $F.B\geq 0$ for any curve $B\subset S$ together with $F.(-K_S)=2$ implies $F.A = F.\overline{A} = 1$. Because $S$ contains only a real one--parameter family of real twistor lines, the intersection points with real twistor fibres form only a real one--dimensional subset of points $z$ on $A$. Therefore, we obtain at least a one--parameter family of such surfaces $S$. Proposition \ref{types} implies now the claim. \qed \begin{prop}\label{notnefii} Assume the existence of an irreducible (non--real) curve $A\subset Z$ with $A.(\fund) = -2$. Then: $h^0(\fdb)=4$ and $a(Z) = 3$.\\ The curves $A$ and $\bar{A}$ are disjoint smooth and rational. $A$ and $\bar{A}$ are the unique irreducible reduced curves having negative intersection number with $\fund$. The base locus of $|\fund|$ is exactly the union of $A$ and $\bar{A}$. $Z$ contains infinitely many divisors of degree one and is one of the twistor spaces studied by LeBrun \cite{LeB} and Kurke \cite{Ku}. \end{prop} {\sc Proof:\quad} We choose $S\in|\fund|$ as in Lemma \ref{sub}(b). Then we have $C\in|-K_S|$ containing $A$ and $\overline{A}$ as smooth rational components which are mapped to ``lines'' $A'$ and $\overline{A}'\in|{\mathcal O}(1,0)|$ in $\bbfP^1\times\bbfP^1$. We have by the adjunction formula $A^2 = A.(-K_S) -2 = -4$, which implies that the eight blown--up points lie on $A'+\overline{A}'$ (or the strict transforms after partial blow--ups). Hence, any member of $A'+\overline{A}'+|{\mathcal O}(0,2)|$ contains the eight blown--up points and defines, therefore, a divisor in $|-K_S|$. On the other hand, any curve in $|-K_S|$ is mapped onto a curve of type $(2,2)$ on $\bbfP^1\times\bbfP^1$ containing the blown--up points. Such a curve must contain $A'$ and $\overline{A}'$ since the intersection number with $A$ is two, but four of the blown--up points lie on $A$. Hence, the image of $|-K_S|$ is precisely the two--dimensional system $A'+\overline{A}'+|{\mathcal O}(0,2)|$. So we have $\dim|-K_S|=2$. Using the exact sequence $0 \rightarrow {\mathcal O}_Z \rightarrow \fdb \rightarrow \canS \rightarrow 0$, this implies $h^0(\fdb)=4$. Furthermore, we see that $A+\bar{A}$ is in the base locus of $|-K_S|$ which coincides with the base locus of $|\fund|$. To see that $Z$ contains infinitely many divisors of degree one we modify an idea of Pedersen, Poon \cite[p.\ 700]{PP}: The above description of the image of $|-K_S|$ in $\bbfP^1\times\bbfP^1$ shows that there exist infinitely many real curves $C\in|-K_S|$ whose image $C'$ in $\bbfP^1\times\bbfP^1$ is of type (I) and the singular points of $C'$ are not blown--up. Then $C$ is the strict transform of such a curve and $C'$ consists of four irreducible components. Let $P$ and $\bar{P}$ denote a conjugate pair of singular points of $C$. Since $A$ and $\bar{A}$ are disjoint, both points $P$ and $\bar{P}$ are contained in $A+\bar{A}$, hence in the base locus of $|\fund|$. On the twistor space $Z$ there exists exactly one real twistor line $L_P$ connecting $P$ with $\bar{P}$. Let $Q\in L_P$ be a point different from $P$ and $\bar{P}$. The linear system $|\fund|_Q$ of all fundamental divisors containing $Q$ has at least dimension $\dim|\fund|-1=\dim|-K_S|=2$. Since fundamental divisors have degree two and any member of $|\fund|_Q$ contains $P,\,\bar{P}$ and $Q$ it also contains $L_P$. Hence, $|\fund|_Q$ coincides with the {\sl real} linear subsystem of divisors in $|\fund|$ containing $L_P$. Choose now a point $R\in C$ which is not on $A+\bar{A}$. Then the real linear system $|\fund|_{L_P,R,\bar{R}}$ is non--empty. Let $C$ be decomposed as $A+\bar{A}+B+\bar{B}$. Then, by our choice of $C$, $B.(\fund)=2$ and $A.(\fund)=-2$. These four curves intersect as indicated in the following picture: \begin{picture}(90,-90)(-30,0) \put(0,-40){\line(1,0){60}} \put(30,-42){\makebox(0,0)[t]{$\bar{A}$}} \put(0,0){\line(1,0){60}} \put(30,2){\makebox(0,0)[b]{$A$}} \put(10,10){\line(0,-1){60}} \put(12,-20){\makebox(0,0)[l]{$B$}} \put(50,10){\line(0,-1){60}} \put(52,-20){\makebox(0,0)[l]{$\bar{B}$}} \end{picture} \rule[-50pt]{0pt}{60pt} \vspace*{12pt} Any real member $S_0$ of $|\fund|_{L_P,R,\bar{R}}$ contains three distinct points of $B$, namely $R$ and the intersection of $B$ with $A+\bar{A}$. Hence, $B\subset S_0$. But this means that $S_0$ contains three curves, say $A, B$ and $L_P$, which meet at $P\in S_0$. As in the proof of Proposition \ref{basel} we obtain that $S_0$ is reducible, hence splits into the sum of two divisors of degree one. Since the intersection of a conjugate pair of divisors of degree one is a twistor line, we obtain in this way infinitely many divisors of degree one on $Z$. By the result of Kurke--Poon \cite{Ku}, \cite{Po2} we obtain that $Z$ is a LeBrun twistor space and $A+\bar{A}$ is precisely the base locus of $|\fund|$. Hence, no other curve can have negative intersection number with $(\fund)$. This proves that really all properties of the Propositon are fulfilled.\qed \begin{prop}\label{notnefiii} Assume $A.(\fund)\geq-1$ for all irreducible curves $A\subset Z$. Then: $h^0(\fdb)=2$ and $a(Z) = 3$.\\ There exists a conjugate pair of irreducible curves $A$ and $\bar{A}$ which are smooth and rational and $A.(\fund)=\bar{A}.(\fund)=-1$. The base locus $C$ of $|\fund|$ consists of a cycle of an even number of rational curves. The number of distinct divisors of degree one on $Z$ is equal to the number of components of $C$. \end{prop} {\sc Proof:\quad} By our assumption we obtain the existence of an irreducible curve $A\subset Z$ with $A.(\fund) = -1$. Now we choose $S\in|\fund|$ real and smooth and $C\in|-K_S|$ as in Lemma \ref{sub}(b). The curve $C\in|-K_S|$ has $A$ and $\overline{A}$ as components and the images $A'$ and $\overline{A}'$ of $A$ and $\overline{A}$ in $\bbfP^1\times\bbfP^1$ are members of $|{\mathcal O}(1,0)|$. But $A^2 = A.(-K_S) - 2 = -3$ implies that exactly one pair of blown--up points does not lie on $A' + \overline{A}'$. This implies that the two components of $C'$, which are members of $|{\mathcal O}(0,1)|$, are not movable. Hence, $|-K_S| = \{C\}$ and, as above, $h^0(\fdb)=2$. Because $C'$ is of type (I), the curve $C$ consists of 2, 3, 4, 5 or 6 pairs of conjugate rational curves. By Proposition \ref{basel} we have: the number of components of $C$ is equal to the number of effective divisors of degree one. It remains to be seen that the {\bf algebraic dimension} is three. Since $|\fund|$ is a pencil we study in more detail the linear system $|-K|$ on $Z$. We need to investigate the structure of $C$ before we can collect more information on the linear system $|-2K_S|$. We know that the blow--up $\sigma:S\rightarrow S^{(0)} := \bbfP^1\times\bbfP^1$ factors through a succession of four blow--ups $S=S^{(4)} \rightarrow S^{(3)} \rightarrow S^{(2)} \rightarrow S^{(1)} \rightarrow S^{(0)}$ such that at each step a conjugate pair of points is blown--up. The image of $C$ in $S^{(i)}$ will be denoted by $C^{(i)}$. The blown--up points in $S^{(i)}$ should lie on $C^{(i)}$. If they are smooth points of $C^{(i)}$ then $C^{(i+1)} \stackrel{\sim}{\rightarrow} C^{(i)}$. If we blow up a conjugate pair of singular points of $C^{(i)}$, the curve $C^{(i+1)}$ has two components more than $C^{(i)}$. By assumption, $C^{(0)} = C' \subset \bbfP^1\times\bbfP^1$ is of type (I). Each $C^{(i)}$ is a ``cycle of rational curves''. We can choose the factorization of $\sigma$ in such a way that at the first $k$ steps, only singular points of $C^{(i)}$ are blown--up and at the last $4-k$ steps, only smooth points of $C^{(i)}$ are blown--up. Then $C$ will have $2(2+k)$ components, where $0\leq k \leq 4$. If we would have a component $A$ of $C$ with $A^2 =0$, then the image $A^{(0)}$ of $A$ in $S^{(0)}$ would be a component of $C^{(0)}$ and none of the blown--up points would lie on $A^{(0)}$. But then four of the blown--up points must lie on a line or on a fibre in $S^{(0)}$, which implies that $C$ has a component $B$ with $B^2 = -4$. This was excluded by assumption. Therefore, for any component $A$ of $C$ we have $-1 \geq A^2 \geq -3$. Since $A$ is a smooth rational curve, this means $A.(-K_S) \in \{-1, 0, +1\}$. By assumption $C$ is reduced. Let $C = \sum\nolimits_{\nu = 1}^m C_\nu$ be the decomposition of $C$ into irreducible components. By Proposition \ref{types} we have $C_\nu \cong {\Bbb P}^1$, $C_\nu.C_{\nu+1} = 1$ and $C_\nu$ intersects only $C_{\nu-1}$ and $C_{\nu+1}$. (This means ``$C$ is a cycle of rational curves''. For convenience, we use cyclic subscripts, that is $C_\nu = C_{\nu+m}$.) For $\varepsilon\in\{-1, 0, +1\}$ we define $I_\varepsilon := \{\nu | C_\nu.(-K_S) = \varepsilon\}$ and $C_\varepsilon := \sum\nolimits_{\nu\in I_\varepsilon} C_\nu$. In this way we split $C$ into three parts $C = C_- + C_0 + C_+$. As $C.(-K_S) = 0$ we have $|I_-| = |I_+|$. The assumption that $\fdb$ is not nef implies $I_- \ne \emptyset$. The curve $C$ has no real component, hence $|I_-| = |I_+|\geq 2$. {\bf Claim:} Any two components of $C_+$ are disjoint.\\ Let $C_\alpha$ and $C_\beta$ be two distinct components of $C_+$. Then $C_\alpha^2 = C_\beta^2 = -1$. If $C_\alpha$ and $C_\beta$ are both contracted to a point on $S^{(0)}$, they are obviously disjoint. If $C_\alpha$ and $C_\beta$ are mapped to curves in $S^{(0)}$, by our choice of $S$ both must be members of $|{\mathcal O}(0,1)|$, because $A$ and $\overline{A}$ are components of $C_-$. Finally, we have to exclude the case where $C_\alpha$ is mapped to a curve $C'_\alpha\in|{\mathcal O}(0,1)|$ in $\bbfP^1\times\bbfP^1$ and $C_\beta$ is contracted to a point $P\in C'_\alpha$. This implies $C_\alpha \ne \overline{C}_\beta$. Since $C_\beta$ is a component of the anticanonical divisor $C\subset S$, the point $P$ must be singular on $C'$. Thus, we can take $S^{(1)} \longrightarrow S^{(0)} = \bbfP^1\times\bbfP^1$ to be the blow--up of $P$ and $\overline{P}$. The curve $C^{(1)}$ consists then of six $(-1)$--curves among which we find the images of $C_\alpha, C_\beta, \overline{C}_\alpha$ and $\overline{C}_\beta$. Because those are $(-1)$--curves on $S$, the remaining six blown--up points must lie on one pair of $(-1)$--curves giving rise to $(-4)$--curves in $C$ contradicting our assumption. Thus, the claim is proved and $C_+$ is the disjoint union of an even number of smooth rational $(-1)$--curves. Since $C$ is a cycle of rational curves, the curve $C\setminus C_+ = C_- + C_0$ has the same number of connected components as $C_+$. We claim that each connected component of $C_- + C_0$ contains exactly one component of $C_-$. Since $C_-$ and $C_+$ have the same number of components, this is equivalent to the statement that each connected component of $C_- + C_0$ contains at most one component of $C_-$.\\ Assume the contary, that is there is a connected component of $C_- + C_0$ containing two irreducible components of $C_-$. Then these two components of $C_-$ are not conjugate to each other. (Two conjugate components of $C$ are ``opposite'' in the cycle $C$.) Therefore, $C_-$ has at least four components and so $C_+$. Thus $C_- + C_0$ has at least four connected components, hence at least six irreducible ones. Therefore, $C$ contains at least ten irreducible components, that is $k\geq 3$. Therefore, the image $C^{(3)}$ of $C$ in $S^{(3)}$ consists of a cycle of ten rational curves with self--intersection numbers $-2, -1, -3, -1, -2, -2, -1, -3, -1, -2$ (in this order). By assumption, at the last step of blow--up, no point on a $(-3)$--curve is blown--up. To obtain a second pair of $(-3)$--curves, we have to blow up points on a conjugate pair of $(-2)$--curves. Such curves are not neighbours in the cycle $C^{(3)}$, they are opposite to each other. Thus, one easily sees that after the last step of blow--up, between two $(-3)$--curves on the cycle $C$ we always have a $(-1)$--curve. This means that no connected component of $C\setminus C_+$ contains two irreducible components of $C_-$, as claimed. We can now compute the dimension of $|-K|$ and $|-2K_S|$.\\ The Riemann--Roch formula and $h^0(\fdb) = 2$ imply $h^1(\fdb) = 0$. Hence, the exact sequence $0 \rightarrow \fdb \rightarrow K^{-1} \rightarrow \can{2} \rightarrow 0$ gives $h^0(K^{-1}) = h^0(\fdb) + h^0(\can{2}) = 2 + h^0(\can{2})$ and a surjective restriction map $|-K| \twoheadrightarrow |-2K_S|$. With $N:=\canS \otimes {\mathcal O}_C$ we obtain an exact sequence $0 \rightarrow \can{1} \rightarrow \can{2} \rightarrow N^{\otimes 2} \rightarrow 0$. Since $h^1(\canS) = h^1(\fdb) = 0$, this sequence yields $h^0(\can{2}) = h^0(\canS) + h^0(N^{\otimes 2}) = 1 + h^0(N^{\otimes 2})$. We can apply Lemma \ref{normal} with $L=\can{2}$, because we have seen that the components of $C_+$ are disjoint to each other and that each connected component of $C\setminus C_+$ contains exactly one irreducible component of $C_-$. We obtain $h^0(N^{\otimes 2}) = h^0(C,L) = \sum_{\nu\in I_+} C_\nu.(-2K_S) - |I_-| = 2|I_+| - |I_-| = |I_+|$. Hence, $\dim |-2K_S| = |I_+| \geq 2$ and $\dim |-K| = 2 + \dim |-2K_S| \geq 4$. Next we study the base locus of $|-2K_S|$.\\ Since any component $A$ of $C_-$ fulfills $A.(-K_S) = -1$, $C_-$ is in the base locus of $|-2K_S|$. Now let $A$ be an irreducible component of $C_0$. Since, as we have shown above, any connected component of $C_0 + C_-$ contains a curve $B\subset C_-$ there exists a finite chain of components $A_1,\dots,A_r = A$ of $C_0$ with $B\cap A_1 \ne \emptyset$ and $A_i \cap A_{i+1} \ne \emptyset\quad (1\leq i < r)$. But for a component $A_i$ of $C_0$ we have $A_i.(-2K_S) = 0$ which implies: if $A_i$ intersects the base locus of $|-2K_S|$, it must be contained in this base locus. Hence, by induction on $i$, we obtain that $A_i$ is contained in the base locus of $|-2K_S|$ for all $1\leq i \leq r$. This shows that $C_0 + C_-$ is contained in the base locus of the linear system $|-2K_S|$. Therefore, we have $|-2K_S| = C_0 + C_- + |-K_S + C_+|$ and the map defined by $|-2K_S|$ coincides with the map given by $|-K_S + C_+|$. We have seen above that $C_+$ is a disjoint union of smooth rational $(-1)$--curves. We can, therefore, contract $C_+$ to obtain a smooth surface $S'$. If $\sigma':S\rightarrow S'$ is this contraction, we have $\canS\otimes{\mathcal O}_S(C_+) \cong \sigma^{'\ast}(K^{-1}_{S'})$. Hence, $-K_{S'}$ is nef if and only if $-K_S + C_+$ is nef. First we deal with the case where $C_+ - K_S$ is nef. In this case, the generic member of the moving part of $|-2K_S|$ is irreducible. This can be seen as follows: Observe first that $S'$ can be blown down to ${\Bbb P}^2$. This follows from \cite[Prop. 3, p. 48]{Dem} because $-K_{S'}$ is nef and $K^2_{S'} = |I_+| > 0$. Because $C_+$ has 2, 4 or 6 components, $S'\rightarrow{\Bbb P}^2$ is a blow--up of 7, 5 or 3 points. Therefore, we can apply a theorem of Demazure \cite[p.\ 39 and p.\ 55]{Dem} stating that $|-K_{S'}|$ contains a smooth irreducible member and is base point free if $|-K_{S'}|$ is nef. Hence, there exists a smooth irreducible curve in $|-K_{S'}|$ avoiding the blown--up points. Its preimage in $S$ is a smooth irreducible member of $|\sigma^\ast K^{-1}_{S'}| = |C_+ - K_S| = |-2K_S - C_0 -C_-|$ which is, therefore, the moving part of $|-2K_S|$. Since $\dim |-2K_S| \geq 2$ and the generic member of the moving part of $|-2K_S|$ is a smooth irreducible curve, the image of the map defined by $|-2K_S|$ has dimension two. We have seen before that the restriction $|-K|\twoheadrightarrow |-2K_S|$ is surjective which implies that the map $\Phi_{|-K|}$ given by $|-K|$ on $Z$ coincides, after restriction to $S$, with the map given by $|-2K_S|$. Since $2S\in|-K|$, the $\Phi_{|-K|}$--image of $S$ is contained in a hyperplane. But the $\Phi_{|-K|}$--image of $Z$ cannot be contained in a hyperplane. Therefore, the image of $\Phi_{|-K|}$ has dimension three. This implies $a(Z)=3$. We are left with the case where $C_+ - K_S$ is not nef.\\ In this case we shall see that $\Phi_{|-K|}$ has only two--dimensional image but equips $Z$ with a conic--bundle structure. Under the assumption that $C_+ - K_S$ is not nef we study the structure of $C$. Let $A$ be an irreducible curve in $S$ with $A.(C_+ - K_S) < 0$. If $A\nsubseteq C$, then $A.C_+\geq 0$. But the base locus of $|-K_S|$ is contained in $C$ and this implies $A.(-K_S)\geq 0$. Hence, we have necessarily $A\subseteq C$. If $A\subseteq C_+$, then $A.(-K_S) = 1$ and $A.C_+ \geq A^2$, hence, $A.(C_+ - K_S) \geq A^2 + 1 = 0$. If $A\subseteq C_0$, then $A.(-K_S) = 0$ and $A.C_+\geq 0$, hence, $A.(C_+ - K_S) \geq 0$. If, finally, $A\subseteq C_-$, then $A.(C_+ - K_S) = A.C_+ - 1 \geq -1$. So, we obtain: the irreducible curves $A\subset S$ with $A.(C_+ - K_S) < 0$ are exactly those components of $C_-$ which are disjoint to $C_+$. They fulfill $A.(C_+ - K_S) = -1$. As we have seen above, each connected component of $C\setminus C_+$ contains exactly one irreducible component of $C_-$. Hence, if $A$ is a component of $C_-$ which does not meet $C_+$, then its connected component should contain at least two curves from $C_0$. Thus, using reality, we see that $C$ has at least eight components. This means, using the convention introduced above, the image $C^{(2)}$ of $C$ in $S^{(2)}$ (the surface obtained after two steps of blow--up) consists of eight curves whose self--intersection numbers are alternately $-1$ and $-2$:\vspace{2mm} \centerline{ \begin{picture}(90,90)(0,-90) \put(2,-32){\line(1,1){30}} \put(15,-17){\makebox(0,0)[br]{$\scriptstyle -2$}} \put(4,-70){\line(0,1){44}} \put(2,-48){\makebox(0,0)[r]{$\scriptstyle -1$}} \put(2,-64){\line(1,-1){30}} \put(15,-79){\makebox(0,0)[tr]{$\scriptstyle -2$}} \put(27,-92){\line(1,0){43}} \put(49,-94){\makebox(0,0)[t]{$\scriptstyle -1$}} \put(64,-94){\line(1,1){30}} \put(80,-79){\makebox(0,0)[tl]{$\scriptstyle -2$}} \put(92,-70){\line(0,1){44}} \put(94,-48){\makebox(0,0)[l]{$\scriptstyle -1$}} \put(94,-32){\line(-1,1){30}}\put(80,-17){\makebox(0,0)[bl]{$\scriptstyle -2$}} \put(27,-4){\line(1,0){43}} \put(49,-2){\makebox(0,0)[b]{$\scriptstyle -1$}} \end{picture} } \vspace{1pt} One now easily sees: if we were to blow up a pair of singular points on $C^{(2)}$, in the resulting curve $C^{(3)}$ any $(-2)$--curve would meet a $(-1)$--curve and any $(-3)$--curve would meet two $(-1)$--curves. Therfore, after the last step of blow--up, no $(-3)$--curve is disjoint to all $(-1)$--curves on $C$. Thus, we can only blow up smooth points in the last two steps. If we were to blow up a conjugate pair of points on $(-2)$--curves, the resulting $(-3)$--curve would intersect two $(-1)$--curves. Then, again, in $C$ there would be no $(-3)$--curve disjoint to all $(-1)$--curves. So we conclude that the last two pairs of conjugate blown--up points cannot lie on $(-2)$--components of $C^{(2)}$. If each of the four $(-1)$--curves in $C^{(2)}$ contains one of the blown--up points, then any component of $C$ has zero intersection number with $-K_S$. But, then by Theorem \ref{nef}, $\fdb$ would be nef which contradicts our general assumption. Thus, the four blown--up points lie on a pair of conjugate $(-1)$--curves. The structure of $C$ is, therefore, the following: \centerline{ \begin{picture}(90,90)(0,-90) \put(2,-32){\line(1,1){30}} \put(15,-17){\makebox(0,0)[br]{$\scriptstyle -2$}} \put(4,-70){\line(0,1){44}} \put(2,-48){\makebox(0,0)[r]{$\scriptstyle -3$}} \put(2,-64){\line(1,-1){30}} \put(15,-79){\makebox(0,0)[tr]{$\scriptstyle -2$}} \put(27,-92){\line(1,0){43}} \put(49,-94){\makebox(0,0)[t]{$\scriptstyle -1$}} \put(49,-89){\makebox(0,0)[b]{$\scriptstyle C_+$}} \put(64,-94){\line(1,1){30}} \put(80,-79){\makebox(0,0)[tl]{$\scriptstyle -2$}} \put(92,-70){\line(0,1){44}} \put(94,-48){\makebox(0,0)[l]{$\scriptstyle -3$}} \put(94,-32){\line(-1,1){30}}\put(80,-17){\makebox(0,0)[bl]{$\scriptstyle -2$}} \put(27,-4){\line(1,0){43}} \put(49,-2){\makebox(0,0)[b]{$\scriptstyle -1$}} \put(49,-6){\makebox(0,0)[t]{$\scriptstyle C_+$}} \end{picture} } \vspace{11pt} In particular, we obtain: $\dim |-2K_S| = |I_+| = 2$ and $\dim |-K| = 2 + \dim|-2K_S| = 4$. Furthermore, since both components of $C_-$ have negative intersection number with $C_+ - K_S$, the curve $C_-$ is contained in the base locus of $|C_+ - K_S|$. This means $|-2K_S| = |2C_+ + C_0| + C_0 + 2C_-$. By our choice of $S$, the two components $A$ and $\overline{A}$ of $C_-$ are mapped onto lines $A'$ and $\overline{A}'$ on $\bbfP^1\times\bbfP^1$. The above analysis of $C$ shows that we can decompose $\sigma : S \longrightarrow \bbfP^1\times\bbfP^1$ into the following steps: First we blow up a conjugate pair of singular points on the curve $C'$ (which is of type (I)). This produces precisely two singular fibres of the ruling (whose general fibre is the image of a twistor fibre). In the second step we blow up the two singular points of these singular fibres. The exceptional curves of this blow--up form the components of $C_+$. Because we blow up points of multiplicity two on the fibres, the total transform of the two singular fibres contains $2C_+$. In the remaining two steps we have to blow up smooth points on $A' + \overline{A}'$. Hence, we obtain $C_0 + 2C_+ \in |2F|$. So, we can write $|-2K_S| = C_0 + 2C_- + |2F|$. Since we have $\dim|\fund| = 1$, this is true for the generic real surface $S\in|\fund|$ by Lemma \ref{sub} (b). Let us denote by $\Phi = \Phi_{|-K|}$ the meromorphic map $Z\dashrightarrow{\Bbb P}^4$ defined by $|-K|$. If $\varphi:S\rightarrow{\Bbb P}^2$ is the restriction of $\Phi$ to a generic smooth real $S\in|\fund|$, then the image of $\varphi$ is a conic in ${\Bbb P}^2$. The general fibre of $\varphi$ is a twistor fibre, hence a smooth rational curve intersecting $C$ transversally at two points lying on $A$ and $\overline{A}$ respectively. Let $\tilde{Z}\rightarrow Z$ be a modification such that $\Phi$ becomes a morphism $\tilde{\Phi}:\tilde{Z}\rightarrow{\Bbb P}^4$. Because the smooth real fundamental divisors $S$ sweep out a Zariski dense subset of $Z$, the image of this set is also Zariski dense in $\tilde{\Phi}(\tilde{Z})\subset{\Bbb P}^4$. As the general fibre of $\Phi$, restricted to such surfaces $S$, is one--dimensional, we obtain $\dim \tilde{\Phi}(\tilde{Z}) = 2$. Since $\tilde{Z}\rightarrow Z$ is a modification, there exists an open Zariski dense subset $U\subset\tilde{\Phi}(\tilde{Z})$ such that the fibres of $\tilde{\Phi}$ are irreducible curves. Moreover, we can choose $U$ such that the fibres of $\tilde{\Phi}$ over $U$ are isomorphic to ${\Bbb P}^1$,because this is true over a Zariski dense subset of $\tilde{\Phi}(\tilde{Z})$. Let $\tilde{\Phi}:\tilde{Z}_U\rightarrow U$ denote the restriction of $\Phi$ over $U$. Then the preimage in $\tilde{Z}$ of the two components of $C_-$ defines a pair of divisors $\Sigma$ and $\bar{\Sigma}$ in $\tilde{Z}_U$ which are sections of $\tilde{Z}_U\rightarrow U$. Therefore, ${\mathcal E}:=\tilde{\Phi}_\ast{\mathcal O}_{\tilde{Z}_U}(\Sigma + \bar{\Sigma})$ is a vector bundle of rank three on $U$ and the canonical morphism $\tilde{\Phi}^\ast{\mathcal E}\rightarrow {\mathcal O}(\Sigma + \bar{\Sigma})$ is surjective. This means that we obtain a morphism $\tilde{Z}_U\rightarrow{\Bbb P}({\mathcal E})$ which is compatible with the projections to $U$. Restricted to each fibre this morphism is the Veronese embedding of degree two ${\Bbb P}^1\hookrightarrow{\Bbb P}^2$. Hence, the image of $\tilde{Z}_U$ in the quasi--projective variety ${\Bbb P}({\mathcal E})$ is three--dimensional. This implies $a(Z) = a(\tilde{Z}_U) = 3$, which completes the proof.\qed \section{Conclusions} \label{vier} In this section we collect the results of this paper to obtain a clear picture of the situation considered. By $Z$ we always denote a simply connected compact twistor space of positive type over $4\bbfC\bbfP^2$. We call ${\mathcal N}:= \{C\subset Z\mid C$ irreducible curve, $C.(\fund)<0\}$ the set of negative curves. By definition, $\fdb$ is nef if and only if ${\mathcal N}\ne \emptyset$. The structure of ${\mathcal N}$ is described by the following \begin{thm}\label{ncurves} If ${\mathcal N}\ne \emptyset$ this set consists of a finite number of smooth rational curves. More precisely, only the following cases are possible: \begin{itemize} \item[(a)] ${\mathcal N}$ contains a real member $C_0$. Then: ${\mathcal N} = \{C_0\}$ and $C_0(\fund) = -2$, $\dim|\fund| = 2$ and $a(Z) = 3$. \item[(b)] ${\mathcal N}$ contains a non--real member $A$ with $A.(\fund) = -2$. Then, ${\mathcal N} = \{ A, \overline{A}\}$, $\dim|\fund| = 3, a(Z) = 3$ and $Z$ is a LeBrun twistor space. \item[(c)] Each member $A\in {\mathcal N}$ fulfills $A.(\fund) = -1$. Then $|{\mathcal N}| \in \{2, 4, 6\}$, $\dim|\fund| = 1$ and $a(Z) =3$. \end{itemize} \end{thm} {\sc Proof:\quad} We have only to collect the results of Section \ref{drei}. \qed We can compute the algebraic dimension in the following way: \begin{thm}\label{main} $a(Z) = 3 \iff \fdb$ is not nef;\\ $a(Z) = 2 \iff \fdb$ is nef and $\exists m\geq 1: h^1(\fb{m}) \ne 0$;\\ $a(Z) = 1 \iff \forall m\geq 1: h^1(\fb{m}) = 0$. \end{thm} {\sc Proof:\quad} This results from Proposition \ref{tau} and Theorems \ref{nef} and \ref{ncurves}.\qed We can characterize Moishezon twistor spaces as follows: \begin{thm} The following conditions are equivalent: \begin{enumerate} \item $a(Z) = 3$; \item $\fdb$ is not nef; \item there exists a smooth rational curve $C\subset Z$ with $C.(\fund) < 0$. \end{enumerate} \end{thm} {\sc Proof:\quad} Apply Theorems \ref{ncurves} and \ref{main}.\qed \begin{rem} Remembering that, by Poon's theorem, $\fdb$ is big if and only if $Z$ is Moishezon, we obtain from the preceding theorem: the line bundle $\fdb$ is never nef and big (under our special assumptions). \end{rem} LeBrun twistor spaces are characterized (see \cite{Ku}, \cite{Po2}) by the property to contain a pencil of divisors of degree one. We can give (for the case $n=4$) two further characterizations: \begin{thm}\label{cb} The following properties are equivalent: \begin{enumerate} \item $Z$ contains a pencil of divisors of degree one; \item $\dim|\fund| = 3$; \item there exists a smooth rational curve $A\subset Z$ with $A \ne \bar{A}$ and $A.(\fund) = -2$. \end{enumerate} \end{thm} {\sc Proof:\quad} The implications (i)$\Rightarrow$(ii) and (i)$\Rightarrow$(iii) follow from the Kurke--Poon theorem. The reverse implications follow from Theorem \ref{ncurves}.\qed \begin{thm} $a(Z) \geq \dim|\fund|$. \end{thm} {\sc Proof:\quad} This follows directly from Proposition \ref{tau} and Theorems \ref{nef} and \ref{ncurves}. \qed If $|\fund|$ is not a pencil, we obtain the following nice result: \begin{thm} If $\dim|\fund|\geq 2$, then:\\ $a(Z) = 2 \iff \fdb$ is nef $\iff |\fund|$ does not have base points. \end{thm} {\sc Proof:\quad} The first equivalence results from the previous theorem and Theorem \ref{main}. If $|\fund|$ does not have base points, $\fdb$ is necessarily nef. If $\fdb$ is nef and $\dim|\fund|\geq 2$ we have seen in Theorem \ref{nef} that $|\fund|$ is base point free.\qed \begin{cor} $|\fund|$ is base point free $\Rightarrow a(Z) = 2$. \end{cor} {\sc Proof:\quad} This is immediate from the previous theorem, because a pencil $|\fund|$ has always base points. \qed \begin{rem} The reverse implication is not true, which follows from the existence theorem in \cite{CK}. There, twistor spaces with $a(Z) = 2$ and $\dim|\fund| = 1$ over $4\bbfC\bbfP^2$ were constructed. \end{rem}
1997-01-17T12:54:38
9607
alg-geom/9607015
en
https://arxiv.org/abs/alg-geom/9607015
[ "alg-geom", "dg-ga", "hep-th", "math.AG", "math.DG" ]
alg-geom/9607015
Teleman
Christian Okonek, Alexander Schmitt and Andrei Teleman
Master Spaces for stable pairs
26 pages. New introduction and applications LaTeX2e
null
null
null
null
We construct master spaces for oriented torsion free sheaves coupled with morphisms into a fixed reference sheaf. These spaces are projective varieties endowed with a natural $\C^*$-action. The fixed point set of this action contains the moduli space of semistable oriented torsion free sheaves and the quot scheme associated with the given data. In the case of curves with trivial reference sheaf, our master spaces compactify the moduli spaces constructed by Bertram, Daskalopoulos and Wentworth. In the 2-dimensional case with trivial rank 1 reference sheaf, master spaces provide algebraic analoga of compactified moduli spaces of twisted quaternionic monopoles.
[ { "version": "v1", "created": "Wed, 17 Jul 1996 09:10:39 GMT" }, { "version": "v2", "created": "Fri, 17 Jan 1997 11:20:20 GMT" } ]
2008-02-03T00:00:00
[ [ "Okonek", "Christian", "" ], [ "Schmitt", "Alexander", "" ], [ "Teleman", "Andrei", "" ] ]
alg-geom
\section*{Introduction} In this paper we construct master spaces for certain coupled vector bundle problems over a fixed projective variety $X$. From a technical point of view, master spaces classify oriented pairs $({\cal E},\varepsilon,\varphi)$ consisting of a torsion free coherent sheaf ${\cal E}$ with fixed Hilbert polynomial, an orientation $\varepsilon$ of the determinant of ${\cal E}$, and a framing $\varphi:{\cal E}\longrightarrow {\cal E}_0$ with values in a fixed reference sheaf ${\cal E}_0$, satisfying certain semistability conditions. The relevant stability concept is new and does not involve the choice of a parameter, but it can easily be compared to the older parameter-dependent stability concepts for (unoriented) pairs. The corresponding moduli spaces ${\cal M}$ have the structure of polarized projective varieties endowed with a natural ${\Bbb C}^*$-action which can be exploited in two interesting ways: 1. The fixed point set ${\cal M}^{{\Bbb C}^*}$ of the ${\Bbb C}^*$-action is a union $${\cal M}^{{\Bbb C}^*}={\cal M}_{source}\cup{\cal M}_{sink}\cup {\cal M}_R\ ,$$ where ${\cal M}_{source}$ is a Gieseker moduli space of semistable oriented sheaves, ${\cal M}_{sink}$ is a certain (possible empty) Grothendieck Quot-scheme, and the third term ${\cal M}_R:={\cal M}^{{\Bbb C}^*}\setminus({\cal M}_{source}\cup{\cal M}_{sink})$ is the so-called "variety of reductions", which consists essentially of lower rank objects. The structure as a ${\Bbb C}^*$-space can be used to relate "correlation functions" associated with the different parts of ${\cal M}^{{\Bbb C}^*}$ to each other [OT2]. 2. Master spaces are also useful for the investigation of the birational geometry of the moduli spaces ${\cal M}_\delta$ of $\delta$-semistable pairs in the sense of [HL2]. Indeed, each of the ${\cal M}_\delta$'s can be obtained as a suitable ${\Bbb C}^*$-quotient of the master space ${\cal M}$, and it can be shown that every two quotients ${\cal M}_\delta$, ${\cal M}_{\delta'}$ are related by a chain of generalized flips in the sense of [Th]. When $X$ is a projective curve with trivial reference sheaf ${\cal E}_0={\cal O}_X^{\oplus k}$, our master space can be considered as a natural compactification of the one described in [BDW]. Their space becomes an open subset of ours whose complement is the Quot-scheme ${\cal M}_{sink}$ alluded to above (${\cal M}_{sink}$ is empty iff $k<\mathop{\rm rk}({\cal E})$). Applying the ideas of 1.\ in this situation leads to formulas for volumina and characteristic numbers and to a new proof of the Verlinde formula when $k=1$, and allows to relate Gromov-Witten invariants for Grassmannians to simpler vector bundle data when $k>\mathop{\rm rk}({\cal E})$. In the case of an algebraic surface $X$, master spaces can be viewed as algebraic analoga of certain gauge theoretic moduli spaces of monopoles which can be used to relate Seiberg-Witten invariants and Donaldson polynomials [OT1], [T1]. The latter application was actually our original motivation for the construction of master spaces. The study of non-abelian monopoles on K\"ahler surfaces leads naturally to the investigation of a certain moment map on an infinite dimensional K\"ahler space. The associated stability concept, which is expected to exist on general grounds [MFK], is precisely the one which gave rise to the stability definition for oriented pairs [OT2]. Since the moduli space of non-abelian monopoles admit an Uhlenbeck type compactification [T1], it was natural to look for a corresponding Gieseker type compactification of their algebro-geometric analoga. These compactifications, the master spaces for stable pairs, provide very useful models for understanding the ends of monopole moduli spaces in the more difficult gauge theoretical context [T2]. Understanding these ends is the essential final step in our program for relating Donaldson polynomials and Seiberg-Witten invariants [OT1], [T1]. Let us now briefly describe the main ideas and results of this paper. The construction of master spaces requires the study of GIT-quotients for direct sums of representations, i.e.\ the construction of quotients $\P(A\oplus B)^{ss}/\hskip-3pt/ G$, where $G$ is a reductive group acting linearly on vector spaces $A$ and $B$. Since the Hilbert criterion is difficult to apply in this situation, we have chosen another approach instead. The idea is to use the natural ${\Bbb C}^*$-action $z\cdot\langle a,b\rangle:=\langle a,zb\rangle$ on $\P(A\oplus B)$ which commutes with the given action of $G$. Our first main result characterizes $G$-semistable points in $\P(A\oplus B)$ in terms of $G$-semistability of their images in all possible ${\Bbb C}^*$-quotients of $\P(A\oplus B)$. The proof is based on a commuting principle for actions of products of groups. These results, which we prove in the first section, explain in particular why chains of flips occur in GIT-problems for $G\times{\Bbb C}^*$-actions [DH], [Th]. In the second section of our paper, after defining stability for oriented pairs $({\cal E},\varepsilon,\varphi)$, we prove a crucial boundedness result and construct the corresponding parameter space ${\frak B}$. This space admits a morphism $\iota:{\frak B}\longrightarrow \P({\frak Z})$ into a certain Gieseker space $\P({\frak Z})$ which is equivariant w.r.t.\ a natural action of a product $\mathop{\rm SL}\times{\Bbb C}^*$ of ${\Bbb C}^*$ with a special linear group. The $\mathop{\rm SL}$-action on $\P({\frak Z})$ possesses a linearization in a suitable line bundle, and the preimage of the subset $\P({\frak Z})^{ss}$ of $\mathop{\rm SL}$-semistable points is precisely the open subspace ${\frak B}^{ss}\subset {\frak B}$ of points representing semistable oriented pairs. In order to prove this, we apply our GIT-Theorem from the first section to the $\mathop{\rm SL}\times{\Bbb C}^*$-action on $\P({\frak Z})$, and thereby reduce the proof to results in [G] and [HL1]. Then we show that the induced map $\iota|_{{\frak B}^{ss}}:{\frak B}^{ss}\longrightarrow \P({\frak Z})^{ss}$ is finite and hence descends to a finite map $\bar\iota:{\frak B}^{ss}/\hskip-3pt/ \mathop{\rm SL}\longrightarrow \P({\frak Z})^{ss}/\hskip-3pt/ \mathop{\rm SL}$. The quotient ${\frak B}^{ss}/\hskip-3pt/ \mathop{\rm SL}$, which is therefore a projective variety, is our master space. The ideas and techniques of this paper can also be applied to construct master spaces in other interesting situations, e.g.\ by coupling with sections in twisted endomorphism bundles. When $X$ is a curve and the twisting line bundle is the canonical bundle, one obtains a natural compactification of the moduli spaces of Higgs bundles [H], [S]. Similar ideas should also apply to coupling with singular objects like parabolic structures. We refer to [OT2] for a general description of the underlying coupling principle and its application to computations of correlation functions. \subsection*{Conventions} Our ground field is ${\Bbb C}$. A \it polarization \rm on a quasi-projective variety $X$ is an equivalence class $[L]$ of ample line bundles, where two line bundles $L_1$ and $L_2$ are \it equivalent\rm , if there exist positive integers $n_1$ and $n_2$ such that $L_1^{\otimes n_1}\cong {L_2}^{\otimes n_2}$. \par If $W$ is a finite dimensional vector space, we denote by $\P(W)$ its projectivization in the sense of Grothendieck, i.e., the closed points of $\P(W)$ correspond to lines in the dual space $W^\vee$. We do not distinguish notationally between a vector space $W$ and its associated scheme. \section{A theorem from Geometric Invariant Theory} \subsection{Background material from GIT} \label{BackGIT} Let $G$ be a reductive algebraic group and let $\gamma\colon G\longrightarrow \mathop{\rm GL}(W)$ be a rational representation in the finite dimensional vector space $W$. The map $\gamma$ defines an action of $G$ on the dual space $W^\vee$ given by $$g\cdot w:= w\circ \gamma(g^{-1})\qquad \forall g\in G; w\in W^\vee,$$ an action $\overline{\gamma}$ on the projective space $\P(W)$, and a linearization of this action in $\O_{\P(W)}(1)$. In the following we identify $H^0(\P(W),\O_{\P(W)}(k))$ with $S^kW$.\par Recall that a point $x\in \P(W)$ is \it $\gamma$-semistable \rm if and only if the orbit closure $\overline{G\cdot w}$ of any lift $w\in W^\vee\backslash\{0\}$ does not contain $0$. Denote by $\P(W)_{\gamma}^{ss}\subset\P(W)$ the open set of semistable points and by $\P(W)_\gamma^{ps}$ the set of \it $\gamma$-polystable \rm points, i.e.\ the semistable points whose orbit is closed in $\P(W)_{\gamma}^{ss}$. Equivalently, a point $x\in \P(W)$ is polystable if and only if the orbit $G\cdot w$ of any lift $w\in W^\vee\backslash\{0\}$ is closed in $W^\vee$. With this terminology, $x\in \P(W)$ is \it $\gamma$-stable \rm if and only if it is polystable and its stabilizer $G_x$ is finite. Let $\pi_\gamma\colon \P(W)_\gamma^{ss}\longrightarrow Q_\gamma:= \P(W)/\hskip-3pt/_\gamma G$ be the categorical quotient. For sufficiently large $n$, $Q_\gamma$ admits a projective embedding $j_n\colon Q_\gamma\hookrightarrow \P({S^nW}^G)$ such that the following diagram commutes: \begin{equation} \label{eqemb} \begin{array}{c} \unitlength=1mm \begin{picture}(70,24)(0,9) \put(0,29){$\P(W)^{ss}_\gamma\subset \P(W)$} \put(34,30){\vector(1,0){20}} \put(34,31){\oval(3,1.8)[l]} \put(40,32){${\scriptstyle v_n}$} \put(56,29){$\P(S^n W)\ $} \put(7,25){\vector(0,-1){10}} \multiput(64,24)(0,-2){4}{\line(0,1){1}} \put(64,16){\vector(0,-1){1}} \put(5,9){$Q_\gamma$} \put(14,11){\oval(3,1.8)[l]} \put(14,10){\vector(1,0){40}} \put(56,9){$\P(S^nW^G)$} \put(2,20){$\scriptstyle\pi_\gamma$} \put(38,12){$\scriptstyle j_n$} \put(66,20){${\scriptstyle p_G}$} \end{picture} \end{array} \end{equation} In this diagram, $v_n$ stands for the $n$-th Veronese embedding and $p_G$ is the projection induced by the inclusion ${S^nW}^G\subset S^nW$. The space $Q_\gamma$ comes with a natural polarization represented by $L_n:= j_n^*\O_{\P({S^nW}^G)}(1)$. Indeed, by (\ref{eqemb}) we have $\pi_\gamma^*L_n\cong\O_{\P(W)_\gamma^{ss}}(n)$, and from the commutative diagram \begin{equation} \label{eqemb} \begin{array}{c} \unitlength=1mm \begin{picture}(120,24)(0,9) \put(0,29){$\P(W)^{ss}_\gamma\subset \P(W)$} \put(34,30){\vector(1,0){10}} \put(34,31){\oval(3,1.8)[l]} \put(36,32){${\scriptstyle v_{n_1}}$} \put(46,29){$\P(S^{n_1} W)$} \put(7,25){\vector(0,-1){10}} \put(5,9){$Q_\gamma$} \put(14,11){\oval(3,1.8)[l]} \put(14,10){\vector(1,0){15}} \put(31,9){$\P(S^{n_1}W^G)$} \put(2,20){$\scriptstyle\pi_\gamma$} \put(18,12){$\scriptstyle j_{n_1}$} \put(110,20){${\scriptstyle p_G}$} \put(53,10){\vector(1,0){10}} \put(53,11){\oval(3,1.8)[l]} \put(55,12){$\scriptstyle v_{n_2}$} \put(66,9){$\P(S^{n_2}(S^{n_1}W^G))$} \multiput(116,24)(0,-2){4}{\line(0,1){1}} \put(116,16){\vector(0,-1){1}} \multiput(97,10)(2,0){3}{\line(1,0){1}} \put(104,10){\vector(1,0){1}} \put(107,9){$\P(S^{n_1n_2} W^G)$} \put(67,30){\vector(1,0){36}} \put(67,31){\oval(3,1.8)[l]} \put(80,32){${\scriptstyle v_{n_2}}$} \put(106,29){$\P(S^{n_1n_2} W)$} \end{picture} \end{array} \end{equation} we infer $L_{n_1}^{\otimes n_2}\cong L_{n_1n_2}$, hence \begin{equation} \label{polind} L_{n_1}^{\otimes n_2}\cong L_{n_2}^{\otimes n_1},\quad \forall n_1,n_2\quad \hbox{large enough}. \end{equation} \begin{Rem} \label{MoreGeneral} In the following, we will mainly consider actions on projective spaces. However, if $X$ is a quasi-projective variety with an action of an algebraic group $G$ which is linearized in an ample line bundle $L$, then $L^{\otimes n}$ induces, for $n$ large enough, a $G$-invariant embedding of $X$ into $\P:=\P(H^0(L^{\otimes n}))$ such that the semistable, polystable, and stable points of $X$ are mapped to the semistable, polystable, and stable points of $\P$. Hence all the results which we will prove hold also in this more general setting, and will be used in this generality in Section 2. \end{Rem} \subsection{Polarized ${\Bbb C}^*$-quotients} \label{PolC*Quot} Let $\lambda\colon{\Bbb C}^*\rightarrow \mathop{\rm GL}(W)$ be a rational representation of ${\Bbb C}^*$ in the finite dimensional vector space $W$ and let $\overline{\lambda}\colon {\Bbb C}^*\times\P(W)\longrightarrow \P(W)$ be the induced action. The space $W^\vee$ splits as a direct sum $$W^\vee=\bigoplus_{i=1}^m W^\vee_i,$$ where $W^\vee_i$ is the eigenspace of the character $\chi_{d_i}\colon{\Bbb C}^*\longrightarrow{\Bbb C}^*, z\longmapsto z^{d_i}$. We assume $d_1<d_2<\cdots<d_m$. Let $x\in \P(W)$ and choose a lift $w\in W^\vee\backslash\{0\}$ of $x$. Define \begin{eqnarray*} d^\lambda_{\min}(x)&:=&\min\bigl\{\, d_i\ \vert\ w\hbox{ has a non-trivial component in } W_i^\vee\,\bigr\}\\ d^\lambda_{\max}(x)&:=&\max\bigl\{\, d_i\ \vert\ w\hbox{ has a non-trivial component in } W_i^\vee\,\bigr\} . \end{eqnarray*} \begin{Prop} \label{C^*-ss}\hfill{\break} {\rm i)} A point $x\in\P(W)$ is $\lambda$-semistable if and only if $d^\lambda_{\min}(x)\le 0\le d^\lambda_{\max}(x)$.\\ {\rm ii)} A point $x\in\P(W)$ is $\lambda$-polystable if and only if either $d^\lambda_{\min}(x)=0=d^\lambda_{\max}(x)$ or $d^\lambda_{\min}(x)<0<d^\lambda_{\max}(x)$. \end{Prop} \begin{pf} Let $w=(w_1,...,w_n)\in W^\vee\backslash\{0\}$ be a lift of $x$, where we take coordinates with respect to a basis of eigenvectors. For $z\in {\Bbb C}^*$, we get $$z\cdot w=(0,...,0,z^{d^\lambda_{\min}(x)}\cdot w_{i_0},...,z^{d^\lambda_{\max}(x)}\cdot w_{i_r},0,...,0).$$ Using this description, the assertion becomes obvious. \end{pf} As remarked above, we can view $\lambda$ as a linearization of the action $\overline{\lambda}$. There are two natural ways of changing this linearization: \begin{enumerate} \item Multiplying $\lambda$ by a character: Let $d$ be an integer, and denote by $\lambda_d$ the representation $z\longmapsto z^d\cdot \lambda(z)$ of ${\Bbb C}^*$ in $\mathop{\rm GL}(W)$. This means that we change the $\O_{\P(W)}(1)$-linearization of $\overline{\lambda}$ by multiplying it with the character $\chi_{d}\colon {\Bbb C}^*\longrightarrow{\Bbb C}^*, z\longmapsto z^{d}$. \item Replacing $\lambda$ by a symmetric power: Let $\lambda^k\colon{\Bbb C}^*\longrightarrow \mathop{\rm GL}(S^kW)$ be the $k$-th symmetric power of $\lambda$. This induces an $\O_{\P(W)}(k)$-linearization of $\overline{\lambda}$. \end{enumerate} Now we combine both methods, i.e., we change $\lambda^k$ to the representation $\lambda^k_d$ of ${\Bbb C}^*$ in $\mathop{\rm GL}(S^kW)$. As above, this defines an $\O_{\P(W)}(k)$-linearization of $\overline{\lambda}$. Altogether, we have a family $\lambda_d^k$, $k\in{\Bbb Z}_{>0}$, $d\in{\Bbb Z}$, of linearizations of $\overline{\lambda}$. Since two $\O_{\P(W)}(k)$-linearizations of $\overline{\lambda}$ differ by a character of ${\Bbb C}^*$, these are indeed all possible linearizations.\par Every linearization $\lambda_d^k$ yields a polarized GIT-quotient $\bigl(Q_d^k:=\P(W)/\hskip-3pt/_{\lambda_d^k}{\Bbb C}^*, [L^k_d]\bigr)$, and $(Q_d^k,[L^k_d])$ and $(Q_{d^\prime}^{k^\prime},[L^{k^\prime}_{d^\prime}])$ are isomorphic as polarized varieties when the ratios $d/ k$ and $d^\prime/ k^\prime$ coincide. To see this, one just has to observe that, for any positve integer $t$, the linearization $\lambda_{t\cdot d}^{t\cdot k}$ is the $t$-th symmetric power of the linearization $\lambda_d^k$. \par Since for a point $x\in \P(W)$ we have $$d^{\lambda^k_d}_{\min}(x)=k\cdot d^\lambda_{\min}-d,\quad d^{\lambda_d^k}_{\max}(x)=k\cdot d^{\lambda}_{\max}-d,$$ we obtain the following corollary to Proposition~\ref{C^*-ss}: \begin{Prop} \label{C^*-ss2} \hfill{\break} {\rm i)} The point $x$ is $\lambda_d^k$-semistable if and only if $d^{\lambda}_{\min}(x)\le d/ k\le d^\lambda_{\max}(x)$. \\ {\rm ii)} The point $x$ is $\lambda_d^k$-polystable if and only if either $d^{\lambda}_{\min}(x)=d/ k= d^\lambda_{\max}(x)$ or $d^{\lambda}_{\min}(x)< d/ k< d^\lambda_{\max}(x)$. In particular, every point $x\in\P(W)$ is $\lambda_d^k$-polystable for suitable numbers $k\in{\Bbb Z}_{>0}$, $d\in{\Bbb Z}$. \end{Prop} For integers $i$ with $1\le i\le 2m$ we define the following intervals in $\P^1_{\Bbb Q}$: $$I_i:=\cases \P_{\Bbb Q}^1\setminus[d_m,d_1] &\hbox{if $i=2m$}\\ \{d_{{i+1\over 2}}\} & \hbox{if $i$ is odd}\\ (d_{{i\over 2}},d_{{i\over 2}+1}) & \hbox{if $i$ is even.} \endcases $$ \begin{Cor} $\P(W)^{ss}_{\lambda^k_d}=\P(W)^{ss}_{\lambda^{k^\prime}_{d^\prime}}$ if and only if there is an $i$ with $1\le i\le 2m$, such that $I_i$ contains both $d/ k$ and $d^\prime/ k^\prime$. \end{Cor} We see that for the given action $\overline{\lambda}$ there are exactly $2m$ notions of stability. Denote by $Q_i$, $i=1,...,2m$, the corresponding unpolarized GIT-quotients, where $Q_{2m}=\emptyset$. Then, for any $i=1,...,2m$, there is a $k$ with $Q_i=Q^k_2$.\par \begin{Rem} Bia\l ynicki-Birula and Sommese \cite{BS} investigated ${\Bbb C}^*$-actions in a more general context. Specialized to our situation, their main result is the following: Let $\lambda$ be a ${\Bbb C}^*$-action on $W$ with a decomposition of the dual space $W^\vee=\bigoplus_{i=1}^m W_i^\vee$ as above. The fixed point set of the induced ${\Bbb C}^*$-action on $\P(W)$ is given by $\bigcup_{i=1}^m \P(W_i)$. Set $F_i:=\P(W_i)$, and define for each index $i$: \begin{eqnarray*} X_i^+&:=&\bigl\{\,x\in \P(W)\ \vert\ \mathop{\rm lim}_{z\longrightarrow 0} z\cdot x\in F_i \,\bigr\}=\P(W_i\oplus\cdots\oplus W_m)\\ X_i^-&:=&\bigl\{\,x\in \P(W)\ \vert\ \lim_{z\longrightarrow \infty} z\cdot x\in F_i\,\bigr\}= \P(W_1\oplus\cdots \oplus W_i), \end{eqnarray*} and for $i\neq j$ set $C_{ij}:=(X_i^+\backslash F_i)\cap (X_j^-\setminus F_j)$. This means $C_{ij}$ is empty for $i\ge j$ and equal to $\P(W_i\oplus\cdots\oplus W_j)\setminus (\P(W_i)\cup\P(W_j))$ for $i<j$. We write $F_i<F_j$ when $C_{ij}\neq\emptyset$, i.e. $$F_1<F_2<\cdots<F_m.$$ In the terminology of \cite{BS}, $F_1$ is the \sl source \rm and $F_m$ is the \sl sink. \rm For each $i$ with $1\le i\le m-1$, one has a partition of $A:=\{1,...,m\}$: $$A=A_i^-\cup A_i^+,\quad \hbox{with } A_i^-:=\{1,...,i\} \hbox{ and } A_i^+=\{ i+1,...,m\},$$ and an associated open set $$U_i:=\bigcup_{\mu\in A_i^-, \nu\in A_i^+} C_{\mu\nu}.$$ The main theorem of \cite{BS} asserts that the $U_i$ are the only Zariski-open ${\Bbb C}^*$-invariant subsets of $\P(W)$ not intersecting the fixed point set whose quotients by the ${\Bbb C}^*$-action are compact. One checks directly that $U_i$ is the set of $\lambda^k_d$-semistable points for any pair $k,d$ with $d/ k\in(d_i,d_{i+1}).$ \end{Rem} \begin{Ex} \label{C^*-ex} Consider an action $\lambda$ of ${\Bbb C}^*$ on a finite dimensional vector space $W$ such that the dual space decomposes as $W^\vee=W_1^\vee\oplus W_2^\vee$ with weights $d_1<d_2$. If $d\in{\Bbb Z}$ and $k\in{\Bbb Z}_{>0}$ are such that $d_1<d/k <d_2$, then the set of $\lambda^k_d$-semistable points is $\P(W_1\oplus W_2)\backslash \left(\P(W_1)\cup\P(W_2)\right)$ and the quotient $Q_{\lambda^k_d}$ is naturally isomorphic to $\P(W_1)\times\P(W_2)$. The quotient map $$\pi\colon\P(W_1\oplus W_2)\backslash \left(\P(W_1)\cup\P(W_2)\right) \subset \P(W_1\oplus W_2)\dasharrow \P(W_1)\times\P(W_2)$$ is the obvious one. \end{Ex} \begin{Claim} The polarization induced by $\lambda^k_d$ on $\P(W_1)\times\P(W_2)$ is the equivalence class of the bundle $\O_{\P(W_1)\times\P(W_2)}(kd_2-d,-kd_1+d)$. In particular, for every $m,n\in {\Bbb Z}_{>0}$, the class $[\O_{\P(W_1)\times\P(W_2)}(m,n)]$ occurs as an induced polarization. \end{Claim} {\it Proof}. Let $L:=\O_{\P(W_1)\times\P(W_2)}(m,n)$ represent the induced polarization. From the description of $\pi$ it follows that $H^0(\pi^*L)^{{\lambda_d^k}}=\pi^*H^0(L)=S^mW_1\otimes S^nW_2$ is the set of bihomogenous polynomials of bidegree $(m,n)$, for some $m,n$. If $S^mW_1\otimes S^nW_2$ occurs as an eigenspace of the induced ${\Bbb C}^*$-action on the space $H^0(\O_{\P(W_1\oplus W_2)}(m\cdot n))$, then it must obviously be an eigenspace for the character $\chi_{-(md_1+nd_2)+((m+n)/ k) d}$. Now invariance implies $md_1+nd_2-((m+n)/ k) d=0$, which can be written as $m (kd_1-d) + n (kd_2-d)=0$. This yields the first assertion.\par To prove the second part of the claim one has to find positive integers $k$, $r$ and an integer $d$ such that the following equations hold \begin{eqnarray*} kd_2-d &=& r m\\ -kd_1+d &=& r n; \end{eqnarray*} this results from a straightforward computation. The other quotients are $\P(W_1)$, $\P(W_2)$ with the obvious polarizations, and $\emptyset$. \subsection{Stability for actions of products of groups} Consider now two reductive groups $G$, $H$ and a rational representation $\rho\colon G\times H\longrightarrow \mathop{\rm GL}(W)$ in the finite dimensional space $W$. We denote by $\gamma$ and $\lambda$ the induced representations of $G$ and $H$, respectively. Choose $n$ large enough in order to obtain an embedding $j_n\colon Q_\gamma\hookrightarrow \P({S^nW}^G)$. Since the actions of $G$ and $H$ commute, $\lambda$ induces actions of $H$ on $Q_\gamma$, on ${S^nW}^G$, and on $\P({S^nW}^G)$; for these actions $j_n$ is $H$-equivariant. The action of $H$ on $Q_\gamma$ possesses a natural linearization in $j_n^*\O_{\P({S^nW}^G)}(1)$. By \ref{BackGIT}(\ref{polind}), the corresponding concept of stability does not depend on the choice of $n$. Let us denote the set of semistable points by $Q_\gamma^{ss}$ and the set of polystable points by $Q_\gamma^{ps}$. \begin{Prop} \label{prodss} The set of $\rho$-semistable points in the projective space $\P(W)$ is given by $\P(W)^{ss}_\rho=\P(W)^{ss}_\gamma\cap \pi_\gamma^{-1}(Q^{ss}_\gamma)$, and there exists a natural isomorphism $Q_\gamma/\hskip-3pt/_\lambda H\cong Q_\rho$. \end{Prop} \begin{pf} Suppose $x\in\P(W)$ is $\gamma$-semistable and its image $\pi_\gamma(x)$ is $\lambda$-semistable in $Q_\gamma$. If $n$ is large, $j_n(\pi_\gamma(x))$ is semistable in $\P({S^nW}^G)$, so that there exists an integer $k\ge 1$ and a section $\overline{s}\in H^0(\P({S^nW}^G),\O_{\P({S^nW}^G)}(k))^H$ not vanishing at \ $j_n(\pi_\gamma(x))$. Identifying $\overline{s}\in {S^k({S^nW}^G)}^H$ with an element of ${S^{kn}W}^{G\times H}$, we obtain a $G\times H$-invariant section in $\O_{\P(W)}(kn)$ not vanishing at $x$, hence $x$ is $\rho$-semistable. \par Conversely, suppose $x\in \P(W)_\rho^{ss}$. Then there exists, for some $m\ge 1$, a section $s\in H^0(\P(W),\O_{\P(W)}(m))^{G\times H}$ with $s(x)\neq 0$. Viewing $s\in {S^mW}^{G\times H}$ as an $H$-invariant element of ${S^mW}^G$, we see that $x\in \P(W)^{ss}_\gamma\cap \pi_\gamma^{-1}(Q_\gamma^{ss}).$ This proves the first assertion. \par The second assertion follows immediately from the first one and the universal property of the categorical quotient. \end{pf} The corresponding result for the polystable points is \begin{Prop} The set of $\rho$-polystable points is $\P(W)^{ps}_\rho=\P(W)^{ps}_\gamma\cap \pi_\gamma^{-1}(Q^{ps}_\gamma)$. \end{Prop} \begin{pf} Let $x\in \P(W)$ be a $\gamma$-polystable point with $\pi_\gamma(x)\in Q_\gamma^{ps}$. By \ref{prodss}, $x$ is $\rho$-semistable. Choose a $\rho$-polystable point $y\in \overline{(G\times H)\cdot x}\cap \P(W)_\rho^{ss}$. Projecting onto $Q_\gamma$, it follows that $\pi_\gamma(y)$ is contained in $\overline{H\cdot \pi_\gamma(x)}$ and hence in $H\cdot \pi_\gamma(x)$, because $\pi_\gamma(x)$ is polystable by assumption. Therefore, there exists an $h\in H$ with $\pi_\gamma(x)=h\cdot \pi_\gamma(y)=\pi_\gamma(h\cdot y)$. But this means that the closures of the $G$-orbits of $x$ and $h\cdot y$ intersect, so that $G\cdot x\subset \overline{G\cdot (h\cdot y)}\cap \P(W)^{ss}_\gamma$, since $x$ is $\gamma$-polystable. In particular, $x\in \overline{(G\times H)\cdot y}\cap\P(W)_\rho^{ss}= (G\times H)\cdot y.$ Hence $x$ is also $\rho$-polystable. \par To prove the converse, suppose $x$ is a $\rho$-polystable point. We first show that $x$ is $\gamma$-polystable, too. Let $y\in \overline{G\cdot x}\cap\P(W)_\gamma^{ss}$ be a $\gamma$-polystable point. Since $\pi_\gamma(y)=\pi_\gamma(x)$, it follows from \ref{prodss} that $\pi_\gamma(y)\in Q_\gamma^{ss}$. Applying \ref{prodss} again, we see that $y\in\P(W)_\rho^{ss}$. The orbit $(G\times H)\cdot x$ being closed in $\P(W)_\rho^{ss}$, there exist $g\in G$ and $h\in H$ with $y=g\cdot h\cdot x$, i.e.\ $x= h^{-1}\cdot g^{-1}\cdot y$. Now $g^{-1}\cdot y$ is $\gamma$-polystable, hence $x$ is $\gamma$-polystable too, because $\gamma$ and $\lambda$ commute. Finally, we must show that $\pi_\gamma(x)\in Q_\gamma^{ps}$. Choose $y$ such that $\pi_\gamma(y)\in\overline{H\cdot\pi_\gamma(x)}\cap Q_\gamma^{ps}$. We may assume that $y$ is $\gamma$-polystable. By what we have already proved, $y$ is $\rho$-polystable. Now $\pi_\gamma(y)$ and $\pi_\gamma(x)$ are mapped to the same point in $Q_\gamma/\hskip-3pt/_\lambda G=Q_\rho$. But the projection $\pi_\rho\colon\P(W)_\rho^{ss}\longrightarrow Q_\rho$ separates closed $\rho$-orbits, thus $(G\times H)\cdot x=(G\times H)\cdot y$, and therefore $H\cdot \pi_\gamma(x)=H\cdot \pi_\gamma(y)$ is closed in $Q_\gamma^{ss}$. \end{pf} \subsection{Applications to $G\times{\Bbb C}^*$-actions} Let $G$ be a reductive algebraic group possessing only the trivial character, so that for any action of $G$ on a projective variety $V$ and any line bundle $L$ on $V$ there is at most one $L$-linearization of the given action. Consider a rational representation $\rho$ of $G\times{\Bbb C}^*$ in the finite dimensional vector space $W$. As above we denote by $\gamma$ and $\lambda$ the induced representations of $G$ and ${\Bbb C}^*$, respectively, and by $\overline{\rho}$, $\overline{\gamma}$, and $\overline{\lambda}$ the induced action of $G\times{\Bbb C}^*$, $G$, and ${\Bbb C}^*$ on $\P(W)$. Let $\P(W)_i^s\subset\P(W)_i^{ps}\subset\P(W)_i^{ss}$ be the set stable, polystable, or semistable points w.r.t.\ the $i$-th stability concept for the action $\overline{\lambda}$, and let $I_i$, $i=1,...,2m$, be the associated intervals of rational numbers. The representation $\rho$ induces an action of $G$ on $Q^k_d$ which is equipped with a natural linearization in the ample line bundle $L^k_d$, and there is no natural way to alter this linearization, because $G$ does not possess a non-trivial character. The corresponding concept of $G$-stability depends only on the rational parameter $d/ k$.\par Now fix a rational parameter $\eta:=d/ k\in I_i$ for some index $i$. A point $y\in Q_i$ is called \it $\eta$-stable ($\eta$-polystable, $\eta$-semistable) \rm if it is $G$-stable ($G$-polystable, $G$-semistable) w.r.t.\ the $G$-linearized line bundle $L^k_d$ on $Q_i=Q_d^k$. \par Recall that every point $x\in\P(W)$ lies in $\P(W)_i^{ps}$ for a suitable index $i$; let $\pi_i(x)\in Q_i$ be its image under $\pi_i\colon \P(W)_i^{ps}\longrightarrow Q_i$. \begin{Thm} \label{GITThm} Fix a point $x\in\P(W)$. Then the following conditions are equivalent: \par {\rm i)} The point $x$ is $G$-semistable ($G$-polystable).\par {\rm ii)} There exists an index $i$ and a parameter $\eta\in I_i$ such that $x\in \P(W)_i^{ss}$ ($x\in \P(W)_i^{ps}$) and $\pi_i(x)$ is $\eta$-semistable ($\eta$-polystable). \end{Thm} \begin{pf} We explain the semistable case; the arguments in the polystable case are similar. Suppose first that $x\in \P(W)$ is $G$-semistable. Choose $n$ large enough (cf.\ Section~\ref{BackGIT}) in order to obtain a commutative diagram as in \ref{BackGIT}(\ref{eqemb}). Since $\gamma$ and $\lambda$ commute, the representation $\lambda^n\colon {\Bbb C}^*\longrightarrow \mathop{\rm GL}(S^nW)$ induces a representation $\lambda^\prime\colon {\Bbb C}^*\longrightarrow \mathop{\rm GL}({S^nW}^G).$ By \ref{C^*-ss2}, we find $k\in {\Bbb Z}_{>0}$ and $d\in{\Bbb Z}$ such that $\pi_\gamma(x)$ is semistable w.r.t.\ the stability concept induced by $(\lambda^\prime)^k_d$ on $Q_\gamma$. Since $(\lambda^\prime)^k_d$ is induced by the representation $$\lambda^{nk}_d\colon {\Bbb C}^*\longrightarrow \mathop{\rm GL}\left(S^k(S^nW)\right),$$ we may replace $n$ by $kn$ and, therefore, assume that $\pi_\gamma(x)$ is semistable w.r.t.\ the stability concept induced by $(\lambda^\prime)_d$ on $Q_\gamma$, for some integer $d$. We now apply Proposition~\ref{prodss} to the representation $$(\gamma^n\times \lambda^n_d)\colon G\times{\Bbb C}^*\longrightarrow \mathop{\rm GL}(S^nW).$$ (Note that this representation induces the action $\overline{\rho}$ on $\P(W)$.) Since $x\in\P(W)$ is $\gamma$-semistable, it is also $\gamma^n$-semistable. By construction, $\pi_\gamma(x)$ is semistable w.r.t.\ the induced ${\Bbb C}^*$-action on $Q_\gamma$, and hence $x$ is $\gamma^n\times\lambda_d^n$-semistable by \ref{prodss}. Applying \ref{prodss} the other way round, setting $\eta:=d/ k$ and choosing $i$ with $\eta\in I_i$, it follows that $x\in\P(W)_i^{ss}$ and that $\pi_i(x)$ is $\eta$-semistable. This settles the implication i)$\Rightarrow$ii). \par To prove the other implication suppose $x\in \P(W)$ fulfills the assumptions of ii). By definition and by Proposition~\ref{prodss}, this means that there are $k\in{\Bbb Z}_{>0}$ and $d\in{\Bbb Z}$ with $\eta=d/ k$ such that $x\in\P(W)$ is $\gamma^k\times\lambda^k_d$-semistable. This implies that $x$ is $\gamma^k$- and hence $\gamma$-semistable. This concludes the proof. \end{pf} \begin{Rem} \label{ChainsofFlips} At this point it becomes clear why chains of flips appear: Let $G$, $\rho$, $\gamma$, and $\lambda$ be as above. We have constructed a family $(\gamma^k\times\lambda^k_d)$ of linearizations of the action $\overline{\rho}$ on $\P(W)$. Each of these linearizations yields a GIT-quotient of $\P(W)$ by the action $\overline{\rho}$. This family of quotients can be constructed in another manner: First take the $G$-quotient in order to obtain a polarized variety $(\tilde{Q}:=\P(W)/\hskip-3pt/_{\gamma} G,[L])$. The resulting ${\Bbb C}^*$-action on this variety yields a family of quotients $Q_i$, $i=1,...,2n$, where $2n$ is usually (much) larger than $2m$, the number of \sl unpolarized \rm ${\Bbb C}^*$-quotients of $\P(W)$ (see~\ref{FlipsEx}). But the family $Q_i$, $i=1,...,2n$, coincides with the family $\P(W)/\hskip-3pt/_{\gamma^k\times \lambda_d^k}G\times{\Bbb C}^*$, $k\in {\Bbb Z}_{>0}, d\in{\Bbb Z}$. This phenomenon is responsible for the occurence of chains of flips in these situations. It explains the question which was left open in \cite{R}, 2.4 Remark (2), 2.5. \end{Rem} \begin{Ex} \label{FlipsEx} Let $W^\vee:=S^3{{\Bbb C}^2}^\vee\oplus {{\Bbb C}^2}^\vee$ and let $\mathop{\rm SL}_2({\Bbb C})$ act on $W^\vee$ in the following way: Given $(f,p)\in W^\vee$ and $m\in \mathop{\rm SL}_2({\Bbb C})$, we interpret $f$ and $p$ as functions on ${\Bbb C}^2$ and set $(m\cdot f)(v):= f(m^t\cdot v)$ and $(m\cdot p)(v):=p(m^t\cdot v)$; then we define $m\cdot (f,p) := (m\cdot f, m\cdot p).$ Let ${\Bbb C}^*$ act on $W^\vee$ by multiplication with $z^{d_1}$ on the first factor and by multiplication with $z^{d_2}$ on the second one. The quotient $V:=W^\vee/\hskip-3pt/\mathop{\rm SL}_2({\Bbb C})$ is of the form $\mathop{\rm Spec}{\Bbb C}[I,J,D,R]$, where $I$, $J$, $D$, and $R$ are certain bihomogenous polynomials of bidegrees $(2,2)$, $(3,3)$, $(4,0)$, and $(1,3)$ in the coordinates of $S^3{{\Bbb C}^2}^\vee$ and ${{\Bbb C}^2}^\vee$. Furthermore, $I$, $D$, and $R$ are algebraically independent, and there is a relation $$27 J^2={1\over 256}DR^2+I^3.$$ We examine the $\mathop{\rm SL}_2({\Bbb C})\times{\Bbb C}^*$-action on $\P(W)$. The quotient $Q:=\P(W)/\hskip-3pt/\mathop{\rm SL}_2({\Bbb C})$ is given by $\mathop{\rm Proj}{\Bbb C}[I,J,D,R]$ where $I$, $J$, $D$, and $R$ have weights 4, 6, 4, and 4, respectively. The ring ${\Bbb C}[I,J,D,R]_{(12)}$ is generated by its elements in degree 1, i.e. by $I^3, I^2D, I^2R, ID^2, IR^2, IDR, J^2, D^3, D^2R, DR^2, R^3$; hence there is an embedding $Q\hookrightarrow \P(S^{12}W^{\mathop{\rm SL}_2({\Bbb C})})$. The ${\Bbb C}^*$-action on $Q$ can be extended to $\P(S^{12}W^{\mathop{\rm SL}_2({\Bbb C})})$ such that the weights of the corresponding action on ${S^{12}W^{\mathop{\rm SL}_2({\Bbb C})}}^\vee$ are $$6d_1+6d_2, 8d_1+4d_2, 5d_1+7d_2, 10 d_1+2d_2, 4d_1+8d_2, 7d_1+5d_2, 12 d_1, 9d_1+3d_2, 3d_1+9d_2.$$ For a point in $p\in Q$, $d_{\min}(p)$ and $d_{\max}(p)$ can take the values $6d_1+6d_2$, $12d_1$, and $3d_1+9d_2$. Hence, for $d_1\neq d_2$, there are 6 different notions of semistability on $Q$, hence 6 different notions of $\mathop{\rm SL}_2\times{\Bbb C}^*$-semistability on $\P(W)$, whereas there are only 4 different notions of ${\Bbb C}^*$-semistability on $\P(W)$. \end{Ex} \section{Oriented pairs and their moduli} Let $X$ be a smooth projective variety over the field of complex numbers and fix an ample divisor $H$ on $X$. All degrees will be taken with respect to $H$ and the corresponding line bundle will be denoted by $\O_X(1)$. Fix a torsion free coherent sheaf ${\cal E}_0$ and a Hilbert polynomial $P$. Finally, let $\mathop{\rm Pic}(X)$ be the Picard scheme of $X$ and choose a Poincar\' e line bundle $\L$ over $\mathop{\rm Pic}(X)\times X$. If $S$ is a scheme and ${\frak E}_S$ a flat family of coherent sheaves over $S\times X$, then there is a morphism $\det_S\colon S\longrightarrow \mathop{\rm Pic}(X)$ mapping a closed point $s$ to $[\det({\frak E}_{S\vert \{s\}\times X})]$. We set $\L[{\frak E}_S]:=(\det_S\times\mathop{\rm id})^*(\L)$; this line bundle depends only on the isomorphism class of the family ${\frak E}_S$. The Hilbert polynomial of a sheaf ${\cal F}$ will be denoted by $P_{\cal F}$. For any non-trivial torsion free coherent sheaf ${\cal F}$ there is a unique subsheaf ${\cal F}_{\max }$ for which $P_{\cal F}/\mathop{\rm rk}{\cal F}$ is maximal and whose rank is maximal among the subsheaves ${\cal F}^\prime$ with $P_{{\cal F}^\prime}/\mathop{\rm rk}{\cal F}^\prime$ maximal. Set $\mu_{\max }({\cal F}):=\mu({\cal F}_{\max })$. \subsection{Oriented pairs} An \it oriented pair of type $(P,\L,{\cal E}_0)$ \rm is a triple $({\cal E},\varepsilon,\phi)$ consisting of a torsion free coherent sheaf ${\cal E}$ with Hilbert polynomial $P_{\cal E}=P$, a homomorphism $\varepsilon\colon\det{\cal E}\longrightarrow\L[{\cal E}]$, and a homomorphism $\phi\colon {\cal E}\longrightarrow {\cal E}_0$. The homomorphisms $\varepsilon$ and $\phi$ will be called the \it orientation \rm and the \it framing \rm of the pair $({\cal E},\varepsilon,\phi)$. Two oriented pairs $({\cal E}_1,\varepsilon_1,\phi_1)$ and $({\cal E}_2,\varepsilon_2,\phi_2)$ are said to be \it equivalent\rm , if there is an isomorphism $\Psi\colon{\cal E}_1\longrightarrow {\cal E}_2$ with $\varepsilon_1 =\varepsilon_2\circ\det\Psi$ and $\phi_1=\phi_2\circ\Psi$. When $\ker(\phi)\neq 0$, we set $$\delta_{{\cal E},\phi}:=P_{\cal E}-{\mathop{\rm rk}{\cal E}\over\mathop{\rm rk}\ker(\phi)_{\max}}P_{\ker(\phi)_{\max }}.$$ An oriented pair $({\cal E},\varepsilon,\phi)$ of type $(P,\L,{\cal E}_0)$ is \it semistable\rm , if either $\phi$ is injective, or $\varepsilon$ is an isomorphism, $\ker(\phi)\neq 0$, $\delta_{{\cal E},\phi}\ge 0$, and for all non-trivial subsheaves ${\cal F}\subset{\cal E}$ $${P_{\cal F}\over \mathop{\rm rk}{\cal F}}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal F}}\quad\le\quad {P_{\cal E}\over\mathop{\rm rk}{\cal E}}- {\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal E}}.$$ \\ The corresponding stability concept is slightly more complicated: An oriented pair $({\cal E},\varepsilon,\phi)$ of type $(P,\L,{\cal E}_0)$ is \it stable\rm , if either $\phi$ is injective, or $\varepsilon$ is an isomorphism, $\ker(\phi)\neq 0$, $\delta_{{\cal E},\phi}> 0$, and one of the following conditions holds: \begin{enumerate} \item For all non-trivial proper subsheaves ${\cal F}\subset {\cal E}$: $${P_{\cal F}\over \mathop{\rm rk}{\cal F}}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal F}}\quad <\quad{P_{\cal E}\over\mathop{\rm rk}{\cal E}}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal E}}.$$ \item $\phi\neq 0$, $\ker(\phi)_{\max}$ is stable, and ${\cal E}\cong\ker(\phi)_{\max}\oplus{\cal E}^\prime$, where the pair $({\cal E}^\prime,\phi)$ satisfies \begin{eqnarray*} {P_{\cal F}\over\mathop{\rm rk}{\cal F}}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal F}}\quad <\quad{P_{{\cal E}^\prime}\over\mathop{\rm rk}{\cal E}^\prime} -{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal E}^\prime} &\ & \hbox{$\forall$ proper subsheaves $0\neq{\cal F}\subset{\cal E}^\prime\ ,$} \\ {P_{\cal F}\over\mathop{\rm rk}{\cal F}}\quad <\quad{P_{{\cal E}^\prime}\over\mathop{\rm rk}{\cal E}^\prime}-{\delta_{{\cal E},\phi} \over\mathop{\rm rk}{\cal E}^\prime} &\ & \hbox{$\forall$ proper subsheaves $0\neq {\cal F}\subset{\cal E}^\prime\cap\ker(\phi)$.} \end{eqnarray*} \end{enumerate} Our (semi)stability concept is related to the \sl parameter dependent \rm (semi)stability concept of \cite{HL1} and \cite{HL2} in the following way: Let $\delta$ be a polynomial over the rationals with positive leading coefficient. Recall that a pair $({\cal E},\phi)$ consisting of a torsion free coherent sheaf ${\cal E}$ with $P_{{\cal E}}=P$ and a non-zero homomorphism $\phi\colon {\cal E}\longrightarrow{\cal E}_0$ is called \it (semi)stable w.r.t.\ $\delta$\rm , if for any non-trivial proper subsheaf ${\cal F}\subset{\cal E}$ the following conditions hold: \begin{eqnarray*} {P_{\cal F}\over\mathop{\rm rk}{\cal F}}-{\delta\over\mathop{\rm rk}{\cal F}}&(\le)&{P_{\cal E}\over\mathop{\rm rk}{\cal E}}-{\delta\over\mathop{\rm rk}{\cal E} }\ ,\\ {P_{\cal F}\over\mathop{\rm rk}{\cal F}}&(\le)&{P_{\cal E}\over\mathop{\rm rk}{\cal E}}-{\delta\over\mathop{\rm rk}{\cal E}} ,\qquad\hbox{when ${\cal F}\subset\ker(\phi)$}. \end{eqnarray*} In this terminology, (semi)stable oriented pairs can be characterized as follows: \begin{Lem} \label{HLcharac} {\rm i)} An oriented pair $({\cal E},\varepsilon,\phi)$ is semistable if and only if it satisfies one of the following three conditions: \begin{enumerate} \item $\phi$ is injective. \item ${\cal E}$ is semistable and $\varepsilon$ is an isomorphism. \item $\phi\neq 0$, $\varepsilon$ is an isomorphism, and $({\cal E},\phi)$ is semistable w.r.t.\ some $\delta>0$. \end{enumerate}\par {\rm ii)} An oriented pair $({\cal E},\varepsilon,\phi)$ is stable if and only if it satisfies one of the following four conditions: \begin{enumerate} \item $\phi$ is injective. \item ${\cal E}$ is stable and $\varepsilon$ is an isomorphism. \item $\phi\neq 0$, $\varepsilon$ is an isomorphism, and $({\cal E},\phi)$ is stable w.r.t.\ some $\delta>0$. \item $\phi\neq 0$, $\delta_{{\cal E},\phi}>0$, $\varepsilon$ is an isomorphism, and ${\cal E}$ splits as $\ker(\phi)_{\max}\oplus{\cal E}^\prime$, where $\ker(\phi)_{\max}$ is stable and $({\cal E}^\prime,\phi)$ is stable w.r.t.\ $\delta_{{\cal E},\phi}$. \end{enumerate} \end{Lem} We note that the stable oriented pairs appearing in Lemma~\ref{HLcharac}.ii)4. are precisely those pairs $({\cal E},\varepsilon,\phi)$, for which $\varepsilon$ is isomorphic, $\phi\neq 0$, $\delta_{{\cal E},\phi}>0$, the pair $({\cal E},\phi)$ is polystable w.r.t.\ $\delta_{{\cal E},\phi}$, and which have \sl only finitely many automorphisms\rm . To see this, recall from \cite{HL2} that for a given $\delta\in{\Bbb Q}[x]$, $\delta>0$, the polystable pairs $({\cal E},\phi)$ are those for which ${\cal E}$ splits in the form $${\cal E}\cong {\cal E}_1\oplus\cdots\oplus{\cal E}_{s-1}\oplus{\cal E}_s\ ,$$ where the sheaves ${\cal E}_1,...,{\cal E}_{s-1}$ are stable subsheaves of $\ker(\varphi)$, $({\cal E}_s,\phi)$ is a stable pair w.r.t.\ $\delta$, and $P_{{\cal E}_1}/\mathop{\rm rk}{\cal E}_1=\cdots= P_{{\cal E}_{s-1}}/\mathop{\rm rk}{\cal E}_{s-1}= P_{{\cal E}_s}/\mathop{\rm rk}{\cal E}_s-\delta/\mathop{\rm rk}{\cal E}_s$. This makes our assertion obvious. \begin{Rem} \label{properlysemistablepairs} Let $({\cal E},\varepsilon,\phi)$ be a stable oriented pair of type 4. (see~\ref{HLcharac}.ii)). Then $\delta_{{\cal E},\phi}$ is the only rational polynomial with positive leading coefficient w.r.t.\ which the pair $({\cal E},\phi)$ is semistable. This follows from the equalities \begin{eqnarray*} {P_{{\cal E}^\prime}\over\mathop{\rm rk}{\cal E}^\prime}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal E}^\prime}&= &{P_{{\cal E}}\over\mathop{\rm rk}{\cal E}}-{\delta_{{\cal E},\phi}\over\mathop{\rm rk}{\cal E}}\ , \\ {P_{\ker(\phi)_{\max}}\over\mathop{\rm rk}\ker(\phi)_{\max}}&= &{P_{{\cal E}}\over\mathop{\rm rk}{\cal E}}-{{\delta_{{\cal E},\phi}}\over{\mathop{\rm rk}{\cal E}}}\stackrel{.}{} \end{eqnarray*} \end{Rem} For all stability concepts introduced so far, there are analogous notions of \it slope-(semi)stability\rm . As usual, slope-stability implies stability and semistability implies slope-semistability. \par Let $S$ be a noetherian scheme. \it A family of oriented pairs parametrized by $S$ \rm is a quadruple $({\frak E}_S,\varepsilon_S, \widehat{\phi}_S,{\frak M}_S)$ consisting of a flat family ${\frak E}_S$ of torsion free coherent sheaves over the product $S\times X$, an invertible sheaf ${\frak M}_S$ on $S$, a morphism $\varepsilon_S\colon \det{\frak E}_S\rightarrow \L[{\frak E}_S]\otimes \pi_S^*{\frak M}_S$, and a morphism $\widehat{\phi}_S\colon S^r{\frak E}_S\rightarrow \pi_X^*S^r{\cal E}_0\otimes \pi_S^*{\frak M}_S$ with $\widehat{\phi}_{S|\{s\}\times X}=S^r\phi_s$ for any closed point $s\in S$ and a suitable $\phi_s\in\mathop{\rm Hom}({\frak E}_{S|\{s\}\times X},{\cal E}_0)$, so that the pair $(\varepsilon_{S|\{s\}\times X}, \widehat{\phi}_{S|\{s\}\times X})$ is non-zero. Two families $({\frak E}^i_S,\varepsilon^i_S,\widehat{\phi}_S^i,{\frak M}_S^i)$, $i=1,2$, are called \it equivalent\rm , if there exist an isomorphism $\Psi_S\colon {\frak E}_S^1\longrightarrow {\frak E}_S^2$ and an isomorphism ${\frak m}\colon {\frak M}_S^1\longrightarrow {\frak M}_S^2$ such that $(\mathop{\rm id}_{\L[{\frak E}_S^1]}\otimes \pi_S^*{\frak m})\circ \varepsilon_S^1=\varepsilon_S^2\circ \det\Psi$ and $(\mathop{\rm id}_{\pi_X^*S^r{\cal E}_0}\otimes \pi_S^*{\frak m})\circ \widehat{\phi}_S^1=\widehat{\phi}_S^2\circ S^r\Psi$. \par With these notions, we define the functors $M^{ss}_{(P,\L,{\cal E}_0)}$ and $M^s_{(P,\L,{\cal E}_0)}$ of equivalence classes of families of semistable and stable oriented pairs of type $(P,\L,{\cal E}_0)$. \begin{Rem} Though the definition of a family may appear a little odd at first sight, it will become clear that families must be defined in this way for technical reasons. Families of the above type are precisely those which are locally induced by the universal family on the parameter space which we will construct in Section~\ref{ParSpace}. The functors defined above do depend on the choice of the Poincar\' e bundle and there is no natural way to compare functors associated to different Poincar\' e bundles. \end{Rem} \subsection{A boundedness result} \label{Bound} Here we show that the family of isomorphism classes of torsion free coherent sheaves occuring in oriented slope-semistable pairs of type $(P,\L,{\cal E}_0)$ is bounded. We use Maruyama's boundedness criterion: \begin{Thm}\cite{Ma} Let $C$ be some constant. The set of isomorphism classes of torsion free coherent sheaves with Hilbert polynomial $P$ and $\mu_{\max }\le C$ is bounded. \end{Thm} \begin{Prop} The set of isomorphism classes of torsion free sheaves occuring in a slope-semistable oriented pair of type $(P,\L,{\cal E}_0)$ is bounded. \end{Prop} \begin{pf} Set $C:=\max\{\,\mu_{\max}({\cal E}_0), \mu({\cal E})\,\}$. Let $({\cal E},\varepsilon,\phi)$ be a slope-semistable oriented pair of type $(P,\L,{\cal E}_0)$. We claim that $\mu_{\max}({\cal E})\le C;$ in view of Maruyama's theorem, this assertion proves the proposition. \par Write a given non-trivial subsheaf ${\cal F}$ of ${\cal E}$ as an extension $$0\longrightarrow {\cal F}\cap\ker(\phi)\longrightarrow {\cal F}\longrightarrow\phi({\cal F})\longrightarrow 0.$$ If ${\cal F}$ is entirely contained in the kernel of $\phi$, the definition of slope-semistability implies $\mu({\cal F})\le \mu({\cal E})\le C$. If ${\cal F}$ is isomorphic to $\phi({\cal F})$, then obviously $\mu({\cal F})\le\mu_{\max}({\cal E}_0)\le C$. In the remaining cases \begin{eqnarray*} \mu({\cal F}) &=& {\mu({\cal F}\cap\ker(\phi))\mathop{\rm rk}({\cal F}\cap\ker(\phi))+ \mu(\phi({\cal F}))\mathop{\rm rk}\phi({\cal F})\over\mathop{\rm rk}{\cal F}}\\ &\le& {\mathop{\rm rk}({\cal F}\cap \ker(\phi))\over\mathop{\rm rk}{\cal F}}\mu({\cal E})+ {\mathop{\rm rk}\phi({\cal F})\over\mathop{\rm rk}{\cal F}}\mu_{\max}({\cal E}_0) \le C. \end{eqnarray*} \end{pf} \subsection{The parameter space for semistable oriented pairs} \label{ParSpace} By the boundedness result of the previous paragraph, there is a natural number $m_0$ such that for all torsion free coherent sheaves ${\cal E}$ occuring in a semistable oriented pair, and for all $m\ge m_0$ the following properties hold true: ${\cal E}(m)$ is globally generated and $H^i(X,{\cal E}(m))=0$ for $i>0$. Let $V$ be a complex vector space of dimension $p:=P(m)$. There exists a quasi-projective scheme ${\frak Q}$, the $\mathop{\rm Quot}$-scheme of torsion free coherent quotient sheaves of $V\otimes\O_X(-m)$ with Hilbert polynomial $P$, and a universal quotient on ${\frak Q}\times X$: $$q_{\frak Q}\colon V\otimes\pi_X^*\O_X(-m) \longrightarrow {\frak E}_{\frak Q}.$$ Let ${\cal N} $ be the sheaf $\pi_{\frak Q *}(\det({\frak E}_{\frak Q})^\vee \otimes \L[{\frak E}_{\frak Q}])$. By the universal property of the Picard scheme, there is a line bundle ${\frak M}$ on ${\frak Q}$ such that $$\det({\frak E}_{\frak Q})^\vee\otimes \L[{\frak E}_{\frak Q}]\cong \pi_{\frak Q}^*\frak M.$$ This implies that ${\cal N}$ is invertible and $${\cal N}\langle [q]\rangle\cong H^0(X,\det({\frak E}_{{\frak Q}\vert\{[q]\}\times X}^\vee)\otimes\L[{\frak E}_{{\frak Q}\vert\{[q]\}\times X}])\ .$$ Let ${\frak N}\buildrel \over\longrightarrow {\frak Q}$ be the associated geometric line bundle. The space ${\frak N}$ is a parameter space for equivalence classes $[q\colon V\otimes\O_X(-m)\longrightarrow {\cal E},\varepsilon]$ consisting of a quotient \linebreak $q\colon V\otimes\O_X(-m)\longrightarrow {\cal E}$ and an orientation $\varepsilon\colon \det({\cal E})\longrightarrow \L[{\cal E}]$. Here two objects $(q_i\colon V\otimes \O_X(-m)\longrightarrow{\cal E}_i,\varepsilon_i)$, $i=1,2$, are \it equivalent\rm , if there is an isomorphism $\Psi\colon {\cal E}_1\longrightarrow{\cal E}_2$ with $\Psi\circ q_1=q_2$ and $\varepsilon_1=\varepsilon_2\circ \det(\Psi)$. \par Next we have to construct a parameter space for all oriented pairs. We choose $m\ge m_0$ so large that ${\cal E}_0(m)$ is also globally generated. Every oriented pair yields an element in $K:=\mathop{\rm Hom}(V,H^0({\cal E}_0(m)))$ and hence an element in $S^rK$. On the projective bundle ${\frak P}:=\P(({\frak N}\times S^rK)^\vee) \stackrel{\frak p} {\longrightarrow}{\frak Q}$ there is a (nowhere vanishing) tautological section $${\frak s}\colon \O_{\frak P}\longrightarrow {\frak p}^*({\cal N}\oplus (S^rK\otimes{\cal O}_{\frak Q}))\otimes\O_{\frak P}(1).$$ Let $$q_{\frak P}\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_{\frak P}$$ be the pullback of the universal quotient on ${\frak Q}\times X$ to ${\frak P}\times X$. We view the pullback $\pi_{\frak P}^*{\frak s}$ of ${\frak s}$ to ${\frak P}\times X$ as a pair consisting of a homomorphism $$\varepsilon_{\frak P}\colon \det({\frak E}_{\frak P})\longrightarrow \L[{\frak E}_{\frak P}]\otimes \pi_{\frak P}^*\O_{\frak P}(1)$$ and a homomorphism $$\kappa_{\frak P}\colon S^rV\otimes\O_{{\frak P}\times X}\longrightarrow S^rH^0({\cal E}_0(m))\otimes \pi_{\frak P}^*\O_{\frak P}(1).$$ \begin{Rem} \label{MorphtoN} For a scheme $S$, giving a morphism $f\colon S\longrightarrow {\frak P}$ is equivalent to giving a map $\overline{f}\colon S\longrightarrow {\frak Q}$ - which yields the family ${\frak E}_S:=(\overline{f}\times\mathop{\rm id}_X)^*{\frak E}_{\frak Q}$ - , a line bundle ${\frak M}_S$ on $S$, and homomorphisms $$\varepsilon_S\colon \det({\frak E}_S)\longrightarrow \L[{\frak E}_S]\otimes \pi_S^*{\frak M}_S\ ,$$ $$\kappa_S\colon S^rV\otimes\O_{S\times X}\longrightarrow S^rH^0({\cal E}_0(m))\otimes\pi_S^*{\frak M}_S\ $$ on $S\times X$ such that the pair $(\varepsilon_{S|\{s\}\times X},\kappa_{S|\{s\}\times X})$ is non-zero for every closed point $s\in S$. Of course, for the morphism $f$ determined by $\overline{f}$ and $(\varepsilon_S,\kappa_S,{\frak M}_S)$, we have $\overline{f}={\frak p}\circ f$, and there is an isomorphism ${\frak m}\colon {\frak M}_S\longrightarrow \overline{f}^*\O_{\frak P}(1)$ such that $$({\rm id}_{\L[{\frak E}_S]}\otimes \pi_S^*{\frak m})\circ\varepsilon_S=(f\times{\rm id}_X)^*(\varepsilon_{{\frak P}})\ ,$$ $$ ({\rm id}_{\pi_X^*S^rH^0({\cal E}_0(m))}\otimes\pi_S^*{\frak m})\circ \kappa_S=(f\times{\rm id}_X)^*(\kappa_{\frak P})\ .$$ \end{Rem} Our parameter space ${\frak B}$ will be a closed subscheme of ${\frak P}$ whose closed points are of the form $[[q\colon V\otimes\O_X(-m)\longrightarrow {\cal E},\varepsilon], S^rk]$, with $[q,\varepsilon]\in {\frak N}$ and $k\in K$, such that there is a map $\phi\colon {\cal E}\longrightarrow {\cal E}_0$ making the following diagramm commutative: \begin{center} \unitlength=1mm \begin{picture}(70,24)(0,8) \put(0,29){$V\otimes{\cal O}_X(-m)$} \put(28,30){\vector(1,0){30}} \put(42,32){${\scriptstyle q}$} \put(62,29){${\cal E}$} \put(7,25){\vector(0,-1){10}} \put(63,25){\vector(0,-1){10}} \put(-15,9){$H^0({\cal E}_0(m))\otimes{\cal O}_X(-m)$} \put(28,10){\vector(1,0){30}} \put(62,9){${\cal E}_0$} \put(2,20){$\scriptstyle k$} \put(42,12){$\scriptstyle ev$} \put(66,20){${\scriptstyle \varphi}$} \end{picture} \end{center} Scheme-theoretically, ${\frak B}$ is constructed as follows: On ${\frak P}\times X$, there is a homomorphism $$\overline{\phi}_{\frak P}\colon S^rV\otimes\pi_X^*\O_X(-rm)\longrightarrow \pi_X^*S^r{\cal E}_0\otimes\pi_{\frak P}^*\O_{\frak P}(1).$$ Set $\widehat{\cal G}:=\ker(S^rq_{\frak P})$, choose $n\ge m$ large enough so that $\widehat{\cal G}_{\vert \{b\}\times X}(n)$ is globally generated and without higher cohomology for any closed point $b\in {\frak P}$, and let $$\widehat{\gamma}\colon {\cal G}:=\widehat{\cal G}\otimes\pi_X^* \O_X(n)\longrightarrow \pi_X^* S^r{\cal E}_0(n)\otimes\pi_{\frak P}^*\O_{\frak P}(1)$$ be the induced homomorphism. We first define a scheme $\widehat{\frak B}$ whose closed points are those elements $b\in {\frak P}$ for which $\widehat{\gamma}_{|\{b\}\times X}$ is the zero map. Since ${\cal G}_{|\{b\}\times X}$ and $S^r{\cal E}_0(n)$ are globally generated for any closed point $b\in {\frak P}$, the scheme $\widehat{\frak B}$ is the zero locus of the following homomorphism between locally free sheaves: $$\gamma:=\pi_{{\frak P}*}(\widehat{\gamma})\colon \pi_{{\frak P}*}{\cal G}\longrightarrow \pi_{{\frak P}*}(\pi_X^*S^r{\cal E}_0(n)\otimes \pi_{\frak P}^*\O_{\frak P}(1))=H^0(S^r{\cal E}_0(n))\otimes \O_{\frak P}(1).$$ The scheme ${\frak B}$ we are looking for is the scheme-theoretic intersection of $\widehat{\frak B}$ with the image in ${\frak P}$ of the weighted projective bundle associated with the vector bundle ${\frak N}\times K$ over ${\frak Q}$. There exists a universal family $({\frak E}_{\frak B},\varepsilon_{\frak B},\widehat{\phi}_{\frak B},{\frak M}_{\frak B})$: ${\frak M}_{\frak B}$ is the restriction of $\O_{\frak P}(1)$ to ${\frak B}$, $q_{\frak B}$ and $\varepsilon_{\frak B}$ are the restrictions of $q_{\frak P}$ and $\varepsilon_{\frak P}$, and $\widehat{\phi}_{\frak B}$ is induced by the restriction of $\widehat{\phi}_{\frak P}$ which factorizes through $S^r{\frak E}_{\frak B}$ by definition. In the following, a closed point $b=[[q\colon V\otimes\longrightarrow \O_X(-m),\varepsilon], S^rk]\in {\frak B}$ will be denoted by $[q,\varepsilon,\phi]$; here $\phi$ is the unique framing on ${\cal E}$ induced by $k$. \begin{Rem} By construction, a morphism $\widehat{f}\colon S\longrightarrow {\frak P}$ factorizes through ${\frak B}$ if and only if it factorizes through the image of the associated weighted projective bundle of ${\frak N}\times K$, and $(\widehat{f}\times\mathop{\rm id}_X)^*(\widehat{\phi}_{\frak P})$ is identically zero on the kernel of the map $(\widehat{f}\times \mathop{\rm id}_X)^*(S^rq_{\frak P})$. \end{Rem} On the parameter space ${\frak B}$, there is a natural action (from the right) of the group $\mathop{\rm SL}(V)$. To define this action, it suffices to construct a $\mathop{\rm SL}(V)$-action on ${\frak P}$ which leaves the scheme ${\frak B}$ invariant. The standard representation of $\mathop{\rm SL}(V)$ on $V$ gives us the homomorphism $$\Gamma\colon V\otimes \O_{{\frak Q}\times\mathop{\rm SL}(V)\times X}\longrightarrow V\otimes \O_{{\frak Q}\times\mathop{\rm SL}(V)\times X}.$$ Moreover, on ${\frak Q}\times\mathop{\rm SL}(V)\times X$ there is the pullback of the universal quotient $$\pi_{{\frak Q}\times X}^*(q_{\frak Q})\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow \pi_{{\frak Q}\times X}^*{\frak E}_{\frak Q}.$$ By the universal property of the $\mathop{\rm Quot}$-scheme, $\pi_{{\frak Q}\times X}^* (q_{\frak Q})\circ \bigl(\Gamma\otimes\mathop{\rm id}_{\pi_X^*\O_X(-m)}\bigr)$ yields a morphism $\overline{f}\colon {\frak Q}\times\mathop{\rm SL}(V)\longrightarrow {\frak Q}$ such that there is a well-defined isomorphism $$\Psi_{{\frak Q}\times \mathop{\rm SL}(V)}\colon (\overline{f} \times\mathop{\rm id}{}_X)^* {\frak E}_{\frak Q}\longrightarrow \pi^*_{{\frak Q}\times X}{\frak E}_{\frak Q}$$ with $\Psi_{{\frak Q}\times\mathop{\rm SL}(V)}\circ (\overline{f}\times\mathop{\rm id}_X)^*(q_{\frak Q})= \pi_{{\frak Q}\times X}^*(q_{\frak Q})\circ \bigl(\Gamma\otimes\mathop{\rm id}_{\pi_X^*\O_X(-m)}\bigr)$. Let $\Psi_{{\frak P}\times \mathop{\rm SL}(V)}$ be the pullback of $\Psi_{{\frak Q}\times\mathop{\rm SL}(V)}$ to ${\frak P}\times \mathop{\rm SL}(V)\times X$, and set ${\frak M}_{{\frak P}\times\mathop{\rm SL}(V)}:= \pi_{\frak P}^*\O_{\frak P}(1)$, \begin{eqnarray*}\varepsilon_{{\frak P}\times\mathop{\rm SL}(V)}&:= &\pi_{{\frak P}\times X}^* (\varepsilon_{{\frak P}})\circ \det\Psi_{{\frak P}\times\mathop{\rm SL}(V)}\ ,\\ \kappa_{{\frak P}\times\mathop{\rm SL}(V)} &:=& \pi^*_{{\frak P}\times X} (\kappa_{\frak P})\circ S^r\left(({\frak p}\times\mathop{\rm id}{}_{\mathop{\rm SL}(V)\times X})^*\Gamma\right)\ . \end{eqnarray*} By Remark~\ref{MorphtoN}, the data $\overline{f}$ and ($\varepsilon_{{\frak P}\times\mathop{\rm SL}(V)}, \kappa_{{\frak P}\times \mathop{\rm SL}(V)}, {\frak M}_{{\frak P}\times\mathop{\rm SL}(V)})$ define an action $$f\colon {\frak P}\times \mathop{\rm SL}(V)\longrightarrow {\frak P}.$$ \begin{Prop} \label{LocUnivProp} Let $S$ be a noetherian scheme and let $({\frak E}_S,\varepsilon_S,\widehat{\phi}_S, {\frak M}_S)$ be a family of semistable oriented pairs parametrized by $S$. Then $S$ can be covered by open subschemes $S_i$ for which there exist morphisms $\beta_i\colon S_i\longrightarrow {\frak B}$ such that the restricted families $({\frak E}_{S|S_i},\varepsilon_{S|S_i},\widehat{\phi}_{S|S_i},{\frak M}_{S|S_i})$ are equivalent to the pullbacks of $({\frak E}_{\frak B}, \varepsilon_{\frak B}, \widehat{\phi}_{\frak B}, {\frak M}_{\frak B})$ via the maps $\beta_i\times\mathop{\rm id}_X$. \end{Prop} \begin{pf} The scheme $S$ can be covered by open subschemes $S_i$ such that the family ${\frak E}_{S\vert S_i}$ over $S_i\times X$ can be written as a family of quotients: $$q_{S_i}\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_{S|S_i}.$$ Each $q_{S_i}$ defines a morphism $\overline{f}_i\colon S_i\longrightarrow {\frak Q}$ such that there is a well defined isomorphism $\Psi_{S_i}\colon {\frak E}_{S_i}:=(\overline{f}_i\times\mathop{\rm id}_X)^* {\frak E}_{\frak Q}\longrightarrow {\frak E}_{S|S_i}$. Define ${\frak M}_{S_i}:={\frak M}_{S|S_i}$, $$\begin{array}{cl} \varepsilon_{S_i}\colon & \det({\frak E}_{S_i})\textmap{\det\Psi_{S_i}} \det{\frak E}_{S|S_i}\textmap{\varepsilon_{S|S_i}} \L[{\frak E}_{S\vert S_i}] \otimes \pi_{S_i}^*{\frak M}_{S_i}\ \hbox{,}\\ \widehat{\phi}_{S_i}\colon & S^r{\frak E}_{S_i}\textmap{S^r\Psi_{S_i}} S^r{\frak E}_{S|S_i}\textmap{\widehat{\phi}_{S|S_i}}\pi_X^*S^r{\cal E}_0\otimes\pi_{S_i}^* {\frak M}_{S_i}. \end{array}$$ The homomorphism $\widehat{\phi}_{S_i}$ yields a homomorphism $$\overline{\kappa}_{S_i}\colon S^rV\otimes\O_{S_i\times X}\longrightarrow \pi_X^*S^r{\cal E}_0(m)\otimes\pi_{S_i}^*{\frak M}_{S_i}$$ and hence a homomorphism $$\kappa_{S_i}:=\pi_{S_i}^*\pi_{S_i*}(\overline{\kappa}_{S_i})\colon S^rV \otimes\O_{S_i\times X} \longrightarrow S^rH^0({\cal E}_0(m))\otimes \pi^*_{S_i}{\frak M}_{S_i};$$ here we have used the fact that our definition of a family implies that the map $$\pi_{S_i*}(\overline{\kappa}_{S_i})\colon S^rV\otimes\O_{S_i}\longrightarrow H^0(S^r{\cal E}_0(m))\otimes{\frak M}_{S_i}$$ factorizes through $S^rH^0({\cal E}_0(m))\otimes {\frak M}_{S_i}$. By Remark~\ref{MorphtoN}, the quadruple $(\overline{f}_i,\varepsilon_{S_i},\kappa_{S_i},{\frak M}_{S_i})$ determines a morphism $\beta_i\colon S_i\longrightarrow {\frak P}$. It is clear that the morphism $\beta_i$ factorizes through ${\frak B}$ and that the family $({\frak E}_{S_i},\varepsilon_{S_i}, \widehat{\phi}_{S_i}, {\frak M}_{S_i})$ is the pullback of the universal family by $\beta_i\times \mathop{\rm id}_X$. The family $({\frak E}_{S_i},\varepsilon_{S_i}, \widehat{\phi}_{S_i},{\frak M}_{S_i})$ is equivalent to $({\frak E}_{S|S_i},\varepsilon_{S|S_i},\widehat{\phi}_{S|S_i}, {\frak M}_{S|S_i})$ by construction. \end{pf} Let ${\frak B}^{\mathop{\rm iso}}$ be the open subscheme of oriented pairs $[q,\varepsilon,\phi]$ for which $H^0(q(m))$ is an isomorphism. The maps constructed in the above proof factorize through ${\frak B}^{\mathop{\rm iso}}$. \begin{Prop} \label{GlueTog} Let $S$ be a noetherian scheme and let $\beta_i\colon S\longrightarrow {\frak B}^{\mathop{\rm iso}}$, $i=1,2$, be two morphisms such that the pullbacks of $({\frak E}_{\frak B}, \varepsilon_{\frak B}, \widehat{\phi}_{\frak B},{\frak M}_{\frak B})$ via the maps $(\beta_i\times\mathop{\rm id}_X)$ are equivalent families. Then there exists an \'etale cover $\eta\colon T\longrightarrow S$ and a morphism $g\colon T\longrightarrow \mathop{\rm SL}(V)$ such that $\beta_1\circ\eta=(\beta_2\circ\eta)\cdot g.$ \end{Prop} \begin{pf} Denote the two families by $({\frak E}_S^i,\varepsilon_S^i,\widehat{\phi}_S^i, {\frak M}_S^i)$, and let $\Psi_S\colon {\frak E}_S^1\longrightarrow {\frak E}_S^2$ be the corresponding isomorphism. The bundles ${\frak E}_S^i$ can be written as quotients $q_S^i\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_S^i$, and there is a morphism $g_S\colon S\longrightarrow \mathop{\rm GL}(V)$ making the following diagramm commutative: \begin{center} \unitlength=1mm \begin{picture}(70,24)(6,9) \put(0,29){$V\otimes\pi_X^*{\cal O}_X(-m)$} \put(32,30){\vector(1,0){20}} \put(37,32){${\scriptstyle g_S\otimes{\rm id}}$} \put(56,29){$V\otimes\pi_X^*{\cal O}_X(-m)$} \put(7,25){\vector(0,-1){10}} \put(64,25){\vector(0,-1){10}} \put(5,9){${\frak E}^1_S$} \put(16,10){\vector(1,0){42}} \put(62,9){${\frak E}^2_S$} \put(2,20){$\scriptstyle q^1_S$} \put(35,12){$\scriptstyle \Psi_S$} \put(66,20){${\scriptstyle q^2_S}$} \end{picture} \end{center} As in the proof of \cite{HL1}, Lemma 1.15, one constructs an \'etale cover $\eta\colon T\longrightarrow S$ such that there is a morphism ${\frak d}\colon T\longrightarrow {\Bbb C}^*$ with $({\frak d}(t))^p=\det(g_S(\eta(t)))$ for any closed point $t\in T$. Now define $g:={\frak d}\cdot (g_S\circ \eta)$. In view of the description of the $\mathop{\rm SL}(V)$-action at the beginning of this section, the assertion is obvious. \end{pf} \subsection{The GIT-construction} \label{GITconst} Let ${\frak A}$ be the union of the finitely many components of $\mathop{\rm Pic}(X)$ meeting the image of $\det_{\frak B}\colon {\frak B}\longrightarrow \mathop{\rm Pic}(X)$. We may choose $m$ so large that the restriction of the line bundle $\L_{\vert{\frak A}\times X}\otimes\pi_X^*\O_X(rm)$ to $\{a\}\times X$ is globally generated and without higher cohomology for any closed point $a\in {\frak A}$. The direct image sheaf $\pi_{{\frak A}_*}(\L_{\vert{\frak A}\times X}\otimes\pi_X^*\O_X(rm))$ is then locally free and commutes with base change. The same holds for ${\cal H}om(\bigwedge^r V\otimes\O_{\frak A},\pi_{{\frak A}*}(\L_{\vert{\frak A}\times X} \otimes\pi_X^*\O_X(rm)))$; let ${\frak H}$ be the geometric vector bundle associated to this locally free sheaf. Consider the homomorphism $$\sigma_{\frak N}\colon \bigwedge^rV\otimes \O_{{\frak N}\times X}\longrightarrow \det{\frak E}_{\frak N}\otimes \pi_X^*\O_X(rm)\stackrel{\varepsilon_{\frak N}}{\longrightarrow} \L[{\frak E}_{\frak N}] \otimes\pi_X^*\O_X(rm).$$ By the universal property of the scheme ${\frak H}$, the pushforward $\pi_{{\frak N}*} (\sigma_{\frak N})$ determines a morphism of schemes ${\frak N}\longrightarrow {\frak H}$ and hence a morphism ${\frak N}\times S^rK\longrightarrow {\frak H}\times S^rK$. Let ${\frak Z}$ be the vector bundle $({\frak H}\times S^rK)^\vee$ over ${\frak A}$, and denote by $\P({\frak Z})$ the associated projective bundle. $\P({\frak Z})$ can be polarized by tensorizing ${\cal O}_{\P({\frak Z})}(1)$ with the pull back of a very ample line bundle from ${\frak A}$. On $\P({\frak Z})$ there is a natural action of the group $\mathop{\rm SL}(V)$ from the right, which is trivial on the base ${\frak A}$ and admits a canonical linearization in the polarizing line bundle. We have a natural morphism $$\iota\colon {\frak B}\hookrightarrow {\frak P}\longrightarrow \P({\frak Z}) $$ which is equivariant w.r.t.\ the given actions. Let us describe the effect of $\iota$ on closed points: Given $b\in {\frak B}$, let $({\cal E}_b,\varepsilon_b,\phi_b)$ be the oriented pair induced by the restriction of $({\frak E}_{\frak B},\varepsilon_{\frak B}, \widehat{\phi}_{\frak B})$ to $\{b\}\times X$, i.e., ${\cal E}_b$ and $\varepsilon_b$ are the restrictions of ${\frak E}_{\frak B}$ and $\varepsilon_{\frak B}$ and $\phi_b$ is a framing with $S^r\phi_b=\widehat{\phi}_{{\frak B}\vert \{b\}\times X}$ ($\phi_b$ is unique up to an $r$-th root of unity). The point $b$ is mapped to $[\L[{\cal E}_b], h,S^rk]$ with $$h\colon \bigwedge^rV\longrightarrow H^0(\det({\cal E}_b)(rm))\textmap{H^0(\varepsilon_b(rm))} H^0(\L[{\cal E}_b](rm))$$ and $k=H^0((\phi_b\circ q)(m))$. A point in $\P({\frak Z})$ is $\mathop{\rm SL}(V)$-\it (semi)stable \rm if it is semistable in the projective space $\P((\mathop{\rm Hom}(\bigwedge^rV,H^0(\L_{\vert\{a\}\times X}(rm)))\oplus S^rK)^\vee),$ where $a$ is its image in ${\frak A}$. \par Let ${\frak B}^{ss}$ (${\frak B}^s$) be the open subscheme of points $[q,\varepsilon,\phi]$ such that the triple $({\cal E},\varepsilon,\phi)$ is a semistable (stable) oriented pair and such that the homomorphism $H^0(q(m))\colon V\longrightarrow H^0({\cal E}(m))$ is an isomorphism. \par \begin{Thm} \label{MyStabCrit} For $m$ large enough, ${\frak B}^{ss}=\iota^{-1}(\P({\frak Z})^{ss})$, and ${\frak B}^s=\iota^{-1}(\P({\frak Z})^{s})$. \end{Thm} Before we can start with the proof, we have to recall some definitions and results from \cite{HL1} and \cite{HL2}. Let $({\cal E},\phi)$ be a pair consisting of a torsion free coherent sheaf ${\cal E}$ with $P_{\cal E}=P$ and a non-trivial framing $\phi$. \par Let $\overline{\delta}$ be any positive rational number. The pair $({\cal E},\phi)$ is called \it sectional (semi)stable w.r.t.\ \rm $\overline{\delta}$, if there is a subspace $V\subset H^0({\cal E})$ of dimension $\chi({\cal E})=P(0)$ such that the following conditions are satisfied: \begin{enumerate} \item For all non-trivial submodules ${\cal F}$ of $\ker(\phi)$: $$(\mathop{\rm rk}{\cal E})\dim \left(H^0({\cal F})\cap V\right)(\le) \mathop{\rm rk}{\cal F}(\chi({\cal E})-\overline{\delta}).$$ \item For all non-trivial submodules ${\cal F}\neq {\cal E}$: $$(\mathop{\rm rk}{\cal E})\dim \left(H^0({\cal F})\cap V\right)(\le) \mathop{\rm rk}{\cal F}(\chi({\cal E})-\overline{\delta})+ (\mathop{\rm rk}{\cal E})\overline{\delta}.$$ \end{enumerate} \par Then we have the following result \cite{HL2}{, Th.\ 2.1}: \begin{Thm} \label{SecStab} For any polynomial $\delta$, there exists a natural number $m_1$ such that for all $m\ge m_1$ the following conditions are equivalent for a pair $({\cal E},\phi)$: \par {\rm i)} $({\cal E},\phi)$ is (semi)stable w.r.t.\ the polynomial $\delta$.\par {\rm ii)} $({\cal E},\phi)(m)$ is sectional (semi)stable w.r.t.\ $\delta(m)$. \end{Thm} Let $(q\colon V\otimes\O_X(-m)\longrightarrow{\cal E},\phi)$ be a pair consisting of a \sl generically \rm surjective map $q$ of $V\otimes \O_X(-m)$ to a torsion free sheaf ${\cal E}$ with $P_{\cal E}=P$ and a non-zero homomorphism $\phi\colon{\cal E}\longrightarrow{\cal E}_0$. We can associate to this pair an element $([h], [k])\in\P(H^\vee)\times\P(K^\vee)$, where $H:=\mathop{\rm Hom}(\bigwedge^rV,H^0(\L[{\cal E}](rm)))$. There is a natural $\mathop{\rm SL}(V)$-action on $\P(H^\vee)\times\P(K^\vee)$ which can be linearized in every sheaf $\O(a_1,a_2)$, where $a_1$ and $a_2$ are positive integers. Define $\nu:=a_2/a_1$ and $\overline{\delta}:= p\nu/(\mathop{\rm rk}{\cal E} +\nu)$. The proof of \cite{HL1}, Proposition 1.18 is valid in any dimension and yields the following \begin{Thm} \label{StabCrit} Let $(q\colon V\otimes \O_X(-m)\longrightarrow{\cal E},\phi)$ be as above. The associated element $([h], [k])$ is (semi)stable w.r.t.\ the linearization in $\O(a_1,a_2)$ if and only if the following two conditions are satisfied: \par {\rm i)} The homomorphism $H^0(q(m))$ is injective.\par {\rm ii)} The pair $({\cal E},\phi)(m)$ is sectional (semi)stable w.r.t.\ $\overline{\delta}$. \end{Thm} We also need the following obvious observation: \begin{Lem} \label{EasyStabCrit} Let $(q\colon V\otimes\O_X(-m)\longrightarrow {\cal E},\phi)$ be as above. The following conditions are equivalent: \par {\rm i)} The homomorphism $k=H^0((\phi\circ q)(m))$ is injective.\par {\rm ii)} The associated element $[k]\in \P(K^\vee)$ is stable. \end{Lem} After these preparations, we return to our situation. Let ${\frak B}_0\subset {\frak B}$ be the open set of all oriented pairs $[q,\varepsilon,\phi]$ for which ${\cal E}$ is semistable, and define for each polynomial $\delta$ the set ${\frak B}_\delta$ as the open set of oriented pairs $({\cal E},\varepsilon,\phi)$ with $\phi\neq 0$ such that $({\cal E},\phi)$ is semistable w.r.t.\ $\delta$. The union ${\frak B}^\prime:={\frak B}_0\cup\bigcup {\frak B}_\delta$ is quasi-projective, hence quasi-compact, so that there exist finitely many polynomials, say, $\delta_1$,...,$\delta_s$ with ${\frak B}^\prime={\frak B}_0\cup {\frak B}_{\delta_1}\cup\cdots\cup {\frak B}_{\delta_s}$. Let $M$ be some constant. By \cite{Ma}{, Theorem 1.7}, the set of points $b \in {\frak B}$ such that $\mu_{\max}({\frak E}_{{\frak B}\vert\{b\}\times X})\le M$ is open. Since ${\frak B}$ is quasi-compact, there is a constant $\mu_0$ such that $\mu_{\max}({\frak E}_{{\frak B}\vert\{b\}\times X})\le \mu_0$ for all $b\in {\frak B}$. We also know that the family ${\frak Ker}$ of kernels of framings of semistable oriented pairs is bounded. It follows that $\mu_{\max}(\ker(\phi))$, for $\ker(\phi)\in {\frak Ker}$, can only take finitely many values. As in \cite{HL2}, Lemma 2.7, this implies that there are only finitely many polynomials of the form $P_{\ker(\phi)_{\max}}$. In particular, there are only finitely many polynomials of the form $$P_{\cal E}-(\mathop{\rm rk}{\cal E}/\mathop{\rm rk}\ker(\phi)_{\max})P_{\ker(\phi)_{\max}}.$$ We assume in the following that these polynomials are among $\delta_1,...,\delta_s$, and that the chosen $m$ is large enough, so that Theorem~\ref{SecStab} holds for all $\delta_i$ and set $\overline{\delta}_i:=\delta_i(m)$. \begin{Thm} \label{TheProp} Suppose $m$ is sufficiently large. Let $[q,\varepsilon,\phi]\in {\frak B}$ be a pair with $\phi\neq 0$ which is not (semi)stable. Then there is no positive rational number $\overline{\delta}$ such that $({\cal E},\phi)(m)$ is sectional (semi)stable w.r.t.\ $\overline{\delta}$. \end{Thm} \begin{pf} Denote by ${\frak S}$ the bounded set of equivalence classes of pairs $({\cal E},\phi)$ for which there is an element $[q,\varepsilon,\phi]\in {\frak B}$. \par By the above, any pair $({\cal E},\phi)\in {\frak S}$ satisfies $\mu_{\max}({\cal E})\le\mu_0$. Let $\tilde\delta$ be a rational polynomial of degree $\dim X-1$ whose leading coefficient $\tilde{\delta}_0$ satisfies $\mu({\cal E})+\tilde{\delta}_0\ge \max\{\,0,\mu_0\,\}$. One can now copy the proof of \cite{HL2}{, page 305}, to show that there is a constant $C$ such that for any submodule $(\tilde{{\cal E}},\tilde{\phi})$ of a pair $({\cal E}, \phi)\in {\frak S}$ either $\vert \deg(\tilde{{\cal E}})-\mathop{\rm rk}\tilde{{\cal E}}\mu({\cal E})\vert <C$, or for all $m$ large enough \begin{eqnarray*} {h^0(\tilde{{\cal E}}(m))\over \mathop{\rm rk}\tilde{{\cal E}}}- {\tilde{\delta}(m)\over \mathop{\rm rk}\tilde{{\cal E}}}&<& {P_{\cal E}(m)\over \mathop{\rm rk}{\cal E}}- {\tilde{\delta}(m)\over \mathop{\rm rk}{\cal E}}\qquad \hbox{if $\tilde{{\cal E}}\not\subset \ker(\phi)\ ,$}\\ \\ {h^0(\tilde{{\cal E}}(m))\over \mathop{\rm rk}\tilde{{\cal E}}} &<&{P_{\cal E}(m)\over \mathop{\rm rk}{\cal E}}- {\tilde{\delta}(m)\over \mathop{\rm rk}{\cal E}}\qquad\hbox{otherwise.}\\ \end{eqnarray*} Recall that a submodule $\tilde{{\cal E}}\subset {\cal E}$ is called \it saturated\rm , if the quotient ${\cal E}/\tilde{{\cal E}}$ is torsion free. The family of saturated submodules $\tilde{{\cal E}}$ of modules ${\cal E}$ with $({\cal E},\phi)\in {\frak S}$ satisfying $\vert \deg(\tilde{{\cal E}})-\mathop{\rm rk}\tilde{{\cal E}}\mu({\cal E})\vert <C$ is bounded (\cite{HL2}{, Lemma 2.7}). Denote this family by $\tilde{\frak S}$. There are only finitely many possibilities for the Hilbert polynomials of those submodules. Let $\delta_j^\prime$ be the finite family of polynomials of the form $P_{\cal E}-(\mathop{\rm rk}{\cal E}/ \mathop{\rm rk}{\cal E}^\prime)P_{{\cal E}^\prime}$ where ${\cal E}^\prime$ is a saturated submodule of $\ker(\phi)$ for some $({\cal E},\phi)\in \tilde{\frak S}$, and $\delta_k^{\prime\p}$ be the finite family of polynomials of the form $(\mathop{\rm rk}{\cal E}^{\prime\p} P_{\cal E}- \mathop{\rm rk}{\cal E} P_{{\cal E}^{\prime\p}})/(\mathop{\rm rk}{\cal E}-\mathop{\rm rk}{\cal E}^{\prime\p})$ where ${\cal E}^{\prime\p}$ is a saturated submodule of a pair $({\cal E},\phi)\in \tilde{\frak S}$ not contained in the kernel of $\phi$. We may assume that $\tilde{\delta}$, the $\delta_j^\prime$'s and the $\delta_k^{\prime\p}$'s with positive leading coefficients are among $\delta_1,...,\delta_s$. Next, we choose $m$ large enough, so that $\tilde{{\cal E}}(m)$ is globally generated and has no higher cohomology for all $\tilde{{\cal E}}\in \tilde{\frak S}$. Let $({\cal E},\phi)$ be a pair which is not semistable w.r.t.\ any of the polynomials $\delta_1,...,\delta_s$. This is equivalent to $({\cal E},\phi)(m)$ not being sectional semistable w.r.t.\ any of the numbers $\overline{\delta}_1,...,\overline{\delta}_s$. Since $({\cal E},\phi)$ is not semistable w.r.t.\ $\tilde{\delta}$, there is either a saturated submodule ${\cal E}_0^\prime\subset\ker(\phi)$ with $\delta_{{\cal E}_0^\prime}:=P_{\cal E}-(\mathop{\rm rk}{\cal E}/ \mathop{\rm rk}{\cal E}_0^\prime)P_{{\cal E}_0^\prime}<\tilde{\delta}$, or there exists a saturated submodule ${\cal E}_0^{\prime\p}\not\subset\ker(\phi)$ such that $$\delta_{{\cal E}_0^{\prime\p}}:=(\mathop{\rm rk}{\cal E}_0^{\prime\p} P_{\cal E}- \mathop{\rm rk}{\cal E} P_{{\cal E}_0^{\prime\p}})/(\mathop{\rm rk}{\cal E}-\mathop{\rm rk}{\cal E}_0^{\prime\p})>\tilde{\delta}\ .$$ In the first case suppose that $\delta_{{\cal E}_0^\prime}$ is minimal and in the second that $\delta_{{\cal E}_0^{\prime\p}}$ is maximal. We consider only the first case, since the second can be treated similarly. If $\delta_{{\cal E}_0^\prime}\le 0$, then we are done. Otherwise, set $\delta^\prime_{i_0}:=\delta_{{\cal E}_0^\prime}$. By the minimality of $\delta_{i_0}^\prime$, any submodule ${\cal E}^\prime$ of $\ker(\phi)$ satisfies $$(\mathop{\rm rk}{\cal E})\dim H^0({\cal E}^\prime(m)) \le \mathop{\rm rk}{\cal E}^\prime (p-\overline{\delta^\prime}_{i_0}),$$ and for ${\cal E}^\prime={\cal E}_0^\prime$ we have equality. Since ${\cal E}$ is not sectional semistable w.r.t.\ $\overline{\delta^\prime}_{i_0}$, there must exist a submodule ${\cal E}^{\prime\p}\not\subset\ker(\phi)$ with $$(\mathop{\rm rk}{\cal E})\dim H^0({\cal E}^{\prime\p}(m))> \mathop{\rm rk}{\cal E}^{\prime\p} (p-\overline{\delta^\prime}_{i_0})+(\mathop{\rm rk}{\cal E})\overline{\delta^\prime}_{i_0}.$$ This makes it obvious that $({\cal E},\phi)$ cannot be sectional semistable w.r.t.\ to any positive rational number.\par We still have to prove the ``stable'' version of the proposition. For this we enlarge the constant $C$ such that $-C\le -\delta_i^0$, $i=1,...,s$, where $\delta_i^0$ is the leading coefficient of $\delta_i$. If $({\cal E},\phi)$ is a pair which is semistable w.r.t.\ the polynomial, say, $\delta_{i_0}$ but not stable w.r.t.\ any other polynomial $\delta$, then there must exist submodules ${\cal E}^\prime\subset\ker(\phi)$ and ${\cal E}^{\prime\p}$ belonging to $\tilde{\frak S}$ with $${P_{{\cal E}^\prime}\over\mathop{\rm rk}{\cal E}^\prime}={P_{{\cal E}}- \delta_{i_0}\over\mathop{\rm rk}{\cal E}}\quad\hbox{and}\quad {P_{{\cal E}^{\prime\p}}- \delta_{i_0}\over \mathop{\rm rk}{\cal E}^{\prime\p}}={P_{{\cal E}^\prime}-\delta_{i_0}\over\mathop{\rm rk}{\cal E}}.$$ Since $m$ was so large that all modules in $\tilde{\frak S}$ are globally generated and without higher cohomology, this gives \begin{eqnarray*} (\mathop{\rm rk}{\cal E})\dim H^0({\cal E}^\prime(m)) &=& (\mathop{\rm rk}{\cal E}^\prime)(p-\overline{\delta}_{i_0})\\ (\mathop{\rm rk}{\cal E})\dim H^0({\cal E}^{\prime\p}(m)) &=& (\mathop{\rm rk}{\cal E}^{\prime\p})(p-\overline{\delta}_{i_0})+ (\mathop{\rm rk}{\cal E})\overline{\delta}_{i_0},\\ \end{eqnarray*} and hence the assertion. \end{pf} \subsection{Proof of Theorem~\ref{MyStabCrit}} \label{PfMyStabCrit} For $b\in {\frak B}$, put $H_b:=\mathop{\rm Hom}(\bigwedge^rV,H^0(\L[{\cal E}_b](rm)))$ and $\P_b:=\P((H_b\oplus S^rK)^\vee)$. The space $\P_b$ admits the following natural ${\Bbb C}^*$-action: $$z\cdot [h,\widehat{k}]:= [h, z\widehat{k}]=[z^{-1} h,\widehat{k}].$$ By \ref{C^*-ex}, this ${\Bbb C}^*$-action can be linearized in such a way that the quotient is either $\P(H_b^\vee)$, $\P((S^rK)^\vee)$, or $\P(H_b^\vee)\times\P((S^rK)^\vee)$ equipped with the polarization $[\O(a_1,a_2)]$ for any prescribed ratio $a_2/a_1$. We are now able to apply our GIT-Theorem~\ref{GITThm} to reduce Theorem~\ref{MyStabCrit} to Theorem~\ref{StabCrit}. \par First we explain the assertion about semistability: Suppose that $b=[q,\varepsilon,\phi]$ lies in ${\frak B}^{ss}$. Then either $\phi$ is injective, or ${\cal E}$ is semistable, or $\phi\neq 0$ and the pair $({\cal E},\phi)$ is semistable w.r.t.\ some $\delta_i$. If $\phi$ is injective, we linearize in such a way that we obtain $\P((S^rK)^\vee)$ as the quotient. By \ref{EasyStabCrit}, the point $[k]$ is semistable in $\P(K^\vee)$ and hence $[S^rk]$ is semistable in $\P((S^rK)^\vee)$. This implies by~\ref{GITThm} that $[h, S^rk]$ is semistable in $\P_b$. If ${\cal E}$ is semistable, we linearize the ${\Bbb C}^*$-action in such a way that the quotient $\P_b/\hskip-3pt/{\Bbb C}^*$ is given by $\P(H_b^\vee)$. By \cite{Gi}, Theorem 0.7 (which does not depend on dimension 2), the point $[h]$ is then semistable in $\P(H_b^\vee)$, and hence $[h,S^rk]$ is $\mathop{\rm SL}(V)$-semistable in $\P_b$ by \ref{GITThm}. If $\phi\neq 0$, $\varepsilon\neq 0$ and $({\cal E},\phi)$ is semistable w.r.t.\ $\delta_i$, we choose the linearization of the ${\Bbb C}^*$-action in such a way that the quotient is $\P(H_b^\vee)\times \P((S^rK)^\vee)$, equipped with a polarization $[\O(ra_1,a_2)]$ satisfying $(a_2/a_1)= \mathop{\rm rk}{\cal E} \overline{\delta}_i/(p-\overline{\delta}_i)$. By Theorem~\ref{StabCrit}, $([h], [S^rk])$ is semistable and thus $[h,S^rk]$ is semistable. \par Conversely, suppose $[h,S^rk]$ is $\mathop{\rm SL}(V)$-semistable. By \ref{GITThm} there is a linearization of the ${\Bbb C}^*$-action such that the image of $[h,S^rk]$ is $\mathop{\rm SL}(V)$-semistable in the quotient $\P_b/\hskip-3pt/{\Bbb C}^*$. There are three possible quotients: If the quotient is $\P((S^rK)^\vee)$, then semistability implies that $[k]$ is semistable in $\P(K^\vee)$ and hence that $k$ is injective. It follows that ${\cal E}$ is a subsheaf of ${\cal E}_0$, since we may assume that $m$ is so large that $\ker(\phi(m))$ is globally generated. If the quotient is $\P(H_b^\vee)$, then ${\cal E}$ is semistable by \cite{Gi}, loc.\ cit.. If the quotient is $\P(H_b^\vee)\times\P((S^rK)^\vee)$ with polarization $[\O(a_1,a_2)]$, then $({\cal E},\phi)$ is sectional semistable w.r.t.\ $$\overline{\delta}:=p(ra_2/a_1)/(\mathop{\rm rk}{\cal E} +(ra_2/a_1))\ .$$ In view of \ref{SecStab} and \ref{TheProp}, $({\cal E},\phi)$ is semistable w.r.t.\ some $\delta$, hence $[q,\varepsilon,\phi]$ lies in ${\frak B}^{ss}$. \par We still have to identify the stable points. As the proof of \ref{GlueTog} shows, the oriented pair $({\cal E},\varepsilon,\phi)$ given by a point $b=[q,\varepsilon,\phi]\in {\frak B}$ has only finitely many automorphisms if and only if the associated point $[h,S^rk]\in\P_b$ has a finite $\mathop{\rm SL}(V)$-stabilizer. Let $b=[q,\varepsilon,\phi]$ be a point whose associated element $[h,S^rk]$ in $\P_b$ is stable. If $h=0$ or $k=0$, then it is easy to see that the corresponding element $[S^rk]\in\P((S^rK)^\vee)$ or $[h]\in\P(H_b^\vee)$ is stable. Hence $H^0(q(m))$ is an isomorphism and either $\phi$ is injective or ${\cal E}$ is a stable sheaf. In both cases, the oriented pair $({\cal E},\varepsilon,\phi)$ is stable and $H^0(q(m))$ is an isomorphism, in other words $b\in {\frak B}^s$. If both $h\neq 0$ and $k\neq 0$, then by \ref{GITThm} $([h],[S^rk])\in \P(H_b^\vee)\times\P((S^rK)^\vee)$ is a polystable point w.r.t.\ the polarization, say, $\O(a_1,a_2)$. By what we have already proved, $({\cal E},\phi)$ is a semistable pair. Remark~\ref{properlysemistablepairs} shows that either $({\cal E},\phi)$ is a stable pair or there is an $i\in\{\, 1,...,s\,\}$ such that $({\cal E},\phi)$ is polystable w.r.t.\ $\delta_i$. In the first case, we are done. In the second case, the finiteness of the stabilizer of $[h,S^rk]$ implies that the oriented pair $({\cal E},\varepsilon,\phi)$ has only finitely many automorphisms, hence it is a stable oriented pair. \par Suppose now that $b\in {\frak B}^s$. If $\phi=0$, then ${\cal E}$ must be a stable coherent sheaf and thus $[h]\in\P(H_b^\vee)$ is a stable point. It follows that $[h,0]$ is a polystable point. But as $[h,0]$ is a fixed point of the ${\Bbb C}^*$-action, the $\mathop{\rm SL}(V)$-stabilizer of $[h,0]\in\P_b$ can be identified with the $\mathop{\rm SL}(V)$-stabilizer of $[h]\in \P(H_b^\vee)$, so that $[h,0]$ is indeed a stable point. If $\varepsilon=0$, then $\phi$ must be injective and we may argue in the same manner. If both $\varepsilon\neq 0$ and $\phi\neq 0$, it suffices to show that $[h,S^rk]$ is a polystable point, since its stabilizer is finite by definition. By the stability of $({\cal E},\phi)$, by the ``stable'' version of \ref{TheProp}, and by the choice of the $\delta_i$, there exists an index $i\in \{\, 1,...,s\,\}$ such that $({\cal E},\phi)$ is polystable w.r.t.\ $\delta_i$. This in turn shows that $([h],[S^rk])\in \P(H_b^\vee)\times\P((S^rK)^\vee)$ is polystable w.r.t.\ the linearization in $\O(ra_1,a_2)$ satisfying $\overline{\delta}_i=p(a_2/a_1)/(\mathop{\rm rk}{\cal E} +(a_2/a_1))$. \subsection{Moduli spaces of stable oriented pairs} We need the following proposition \begin{Prop} \label{Proper} The map ${\iota}_{\vert {\frak B}^{ss}}\colon {\frak B}^{ss}\longrightarrow \P({\frak Z})^{ss}$ is finite. \end{Prop} \begin{pf} We claim that ${\iota}_{\vert {\frak B}^{ss}}$ is proper and injective. Injectivity follows by standard arguments. For the proof of properness, we will make use of the discrete valuative criterion. Let $C=\mathop{\rm Spec} R$ be the spectrum of a discrete valuation ring, $c_0\in C$ the closed point, and $C_0:=C\setminus \{c_0\}$. Suppose there is a commutative diagram: \begin{center} \unitlength=0.08mm \begin{picture}(400,400)(0,60) \put(0,390){$C_0$} \put(75,400){\vector(1,0){220}} \put(170,410){${\scriptstyle u}$} \put(320,390){${\frak B}^{ss}$} \put(20,350){\vector(0,-1){200}} \put(340,350){\vector(0,-1){200}} \put(0,90){$C$} \put(70,100){\vector(1,0){220}} \put(320,90){$\P({\frak Z})^{ss}$} \put(170,120){$\scriptstyle \bar u$} \put(380,240){$\scriptstyle \iota|_{{\frak B}^{ss}}$} \multiput(70,140)(80,80){3}{\line(1,1){45}} \put(250,320){\vector(1,1){36}} \put(160,280){$\scriptstyle\tilde u$} \end{picture} \end{center} We have to construct a lifting $\tilde{u}$ of the map $\overline{u}$. By assumption, we are given a family $({\frak E}_{C_0}, \varepsilon_{C_0},\widehat{\phi}_{C_0},\O_{C_0})$ of semistable oriented pairs over $C_0\times X$. Note that ${\frak E}_{C_0}$ is torsion free. We claim that we can extend the quotient map $$q_{C_0}\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_{C_0}$$ to a homomorphism $q_C\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_C$ over $C\times X$, where ${\frak E}_C$ is a flat family of torsion free coherent sheaves extending ${\frak E}_{C_0}$, $q_{C}$ extends $q_{C_0}$, and its restriction to $\{c_0\}\times X$ is generically surjective. In order to prove this claim, we first extend the family ${\frak E}_{C_0}$ to a flat family of quotients $$\tilde{q}_C\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow \tilde{\frak E}_C.$$ There is a locally free sheaf ${\cal H}$ on $X$ and an epimorphism $\pi_X^*{\cal H}\longrightarrow \tilde{\frak E}^\vee_C$. This yields a homomorphism $$\lambda\colon V\otimes\pi_X^*\O_X(-m)\longrightarrow \tilde{\frak E}_C\longrightarrow \tilde {\frak E}_C^{\vee\vee}\longrightarrow \pi_X^*{\cal H}^\vee.$$ Let ${\frak E}_C$ be the maximal subsheaf of $\pi_X^*{\cal H}^\vee$ with the following properties $${\frak E}_{C\vert C_0\times X}={\frak E}_{C_0};\qquad \Im\lambda\subset {\frak E}_C;\qquad \dim(\mathop{\rm supp}({\frak E}_C/\Im\lambda))<\dim X.$$ Note that the set of subsheaves of $\pi_X^*{\cal H}^\vee$ having the above properties contains $\Im\lambda$. One checks that ${\frak E}_{C\vert \{c_0\}\times X}$ is torsion free, using arguments as in \cite{HL1}, p.85. Let $t\in R$ be a generator of the maximal ideal. There is a well defined integer $\alpha$ such that $(t^{\alpha}\varepsilon_{C_0}, t^\alpha\widehat{\phi}_{C_0})$ extends to the family ${\frak E}_C$. The classifying map to $\P({\frak Z})^{ss}$ induced by the resulting family $$(q_C\colon V\otimes\pi_X^*\O_X(-m)\longrightarrow {\frak E}_C,\tilde{\varepsilon}_C,\tilde{\widehat{\phi}}_C,\O_C)$$ is the same as the one induced by $\overline{u}$. By the various stability criteria we have encountered so far, it follows that $H^0(q_{C\vert \{c_0\}\times X}(m))$ is injective and that the triple $({\frak E}_{C\vert \{c_0\}\times X}, \tilde{\varepsilon}_{C\vert \{c_0\}\times X}, \tilde{\phi}_{c_0})$, where $\phi_{c_0}$ is a framing induced by $\tilde{\widehat{\phi}}_{C|\{c_0\}\times X}$, is a semistable oriented pair. Thus, ${\frak E}_{C\vert \{c_0\}\times X}(m)$ is globally generated and without higher cohomology, the map $q_{C\vert \{c_0\}\times X}$ is surjective, and hence $q_C\colon V\otimes \pi_X^*\O_X(-m)\longrightarrow {\frak E}_C$ is a flat family of torsion free quotients. The family $(q_C\colon V\otimes\pi_X^*\O_X(-m)\longrightarrow {\frak E}_C,\tilde{\varepsilon}_C, \tilde{\widehat{\phi}}_C,\O_C)$ defines by \ref{LocUnivProp} a morphism $$\tilde{u}\colon C\longrightarrow {\frak B}^{ss}$$ which extends $u$ by construction. \end{pf} By Proposition 2.6.1. and \cite{Gi}, Lemma 4.6, the quotient ${\frak B}^{ss}/\hskip-3pt/\mathop{\rm SL}(V)$ exists as a projective scheme. We set $${\cal M}_{(P,\L,{\cal E}_0)}^{ss}:={\frak B}^{ss}/\hskip-3pt/\mathop{\rm SL}(V)\ ,$$ $${\cal M}_{(P,\L,{\cal E}_0)}^{s}:={\frak B}^s/\hskip-3pt/\mathop{\rm SL}(V)\ .$$ \begin{Thm} \label{ModuliSpaces} {\rm i)} There is a natural transformation of functors $$ \tau\colon M_{(P,\L,{\cal E}_0)}^{ss}\longrightarrow h_{{\cal M}_{(P,\L,{\cal E}_0)}^{ss}},$$ \hspace*{0.5cm} such that for any scheme $\tilde{\cal M}$ and any natural transformation of functors $$\tau^\prime\colon M_{(P,\L,{\cal E}_0)}^{ss}\longrightarrow h_{\tilde{\cal M}}$$ \hspace*{0.5cm}there is a unique morphism $\vartheta\colon {\cal M}_{(P,\L,{\cal E}_0)}^{ss}\longrightarrow \tilde{\cal M}$ such that $\tau^\prime=h(\vartheta)\circ \tau$.\\ \hspace*{0.5cm}{\rm ii)} The space ${\cal M}_{(P,\L,{\cal E}_0)}^s$ is a coarse moduli space for stable oriented pairs. \end{Thm} \begin{pf} The existence of the natural transformation is a direct consequence of Proposition~\ref{LocUnivProp} and \ref{GlueTog}. The minimality property of ${\cal M}_{(P,\L,{\cal E}_0)}^{ss}$ follows from the universal property of the categorical quotient.\par Since ${\frak B}^s$ is contained in the set of $\mathop{\rm SL}(V)$-stable points, the set of closed points of ${\cal M}_{(P,\L,{\cal E}_0)}^s$ is the set of equivalence classes of stable oriented pairs which means that ${\cal M}_{(P,\L,{\cal E}_0)}^s$ is a coarse moduli space. \end{pf} \vspace{0.3cm} In our applications [OT2] we shall also need a slightly modified version of the constructions and results above. We fix a line bundle ${\cal L}_0\in\mathop{\rm Pic} (X)$ and consider only torsion free sheaves of determinant isomorphic to ${\cal L}_0$. More precisely, an ${\cal L}_0$-{\it oriented pair of type $(P,{\cal E}_0)$} is a triple $({\cal E},\varepsilon,\varphi)$ consisting of a torsion free coherent sheaf ${\cal E}$ with Hilbert polynomial $P$ and with $\det{\cal E}$ isomorphic to ${\cal L}_0$, a homomorphism $\varepsilon:\det{\cal E}\longrightarrow {\cal L}_0$, and a homomorphism $\varphi:{\cal E}\longrightarrow {\cal E}_0$. Equivalence classes of such ${\cal L}_0$-oriented pairs, families, equivalence classes of families, (semi)stability and the corresponding functors $M^{ss}_{(P,{\cal L}_0,{\cal E}_0)}$ are defined as in 2.1. The same methods as above yield the following result: \begin{Thm} \label{ModuliSpaces} There exist moduli spaces ${\cal M}^{ss}_{(P,\L_0,{\cal E}_0)}$ and ${\cal M}^{s}_{(P,\L_0,{\cal E}_0)}$ with the following properties:\\ \hspace*{0.5cm} {\rm i)} There is a natural transformation of functors $$ \tau\colon M_{(P,\L_0,{\cal E}_0)}^{ss}\longrightarrow h_{{\cal M}_{(P,\L_0,{\cal E}_0)}^{ss}},$$ \hspace*{0.5cm} such that for any scheme $\tilde{\cal M}$ and any natural transformation of functors $$\tau^\prime\colon M_{(P,\L_0,{\cal E}_0)}^{ss}\longrightarrow h_{\tilde{\cal M}}$$ \hspace*{0.5cm} there is a unique morphism $\vartheta\colon {\cal M}_{(P,\L_0,{\cal E}_0)}^{ss}\longrightarrow \tilde{\cal M}$ such that $\tau^\prime=h(\vartheta)\circ \tau$.\\ \hspace*{0.5cm} {\rm ii)} The space ${\cal M}_{(P,\L_0,{\cal E}_0)}^s$ is a coarse moduli space for stable ${\cal L}_0$-oriented pairs. \end{Thm} \subsection{The closed points of ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$} Let $({\cal E},\varepsilon,\phi)$ be a semistable oriented pair of type $(P,\L,{\cal E}_0)$. If $({\cal E},\varepsilon,\phi)$ is stable, then it defines a closed point in ${\cal M}_{(P,\L,{\cal E}_0)}^{ss}$. If $({\cal E},\varepsilon,\phi)$ is not stable, then either ${\cal E}$ is a semistable but not stable coherent sheaf, or $\phi\neq 0$ and there exists a $\delta\in{\Bbb Q}[x]$, $\delta>0$, such that $({\cal E},\phi)$ is semistable but not stable w.r.t.\ $\delta$. In both cases, there is a Harder-Narasimhan filtration $$0= {\cal E}_0\subset {\cal E}_1\subset\cdots\subset{\cal E}_s={\cal E}$$ of ${\cal E}$, whose associated graded sheaf $\mathop{\rm gr}({\cal E}):=\bigoplus_{i=1}^s {\cal E}_i/{\cal E}_{i-1}$ inherits a well-defined orientation $\varepsilon_{\mathop{\rm gr}}$ and a well-defined framing $\phi_{\mathop{\rm gr}}$ from $({\cal E},\varepsilon,\phi)$. As usual, the resulting object $({\mathop{\rm gr}}({\cal E}),\varepsilon_{\mathop{\rm gr}}, \phi_{\mathop{\rm gr}})$ is determined up to equivalence. We call it the \it graded object associated to $({\cal E},\varepsilon,\phi)$\rm . Using the techniques of Section~\ref{PfMyStabCrit}, i.e., applying \ref{GITThm} in the ``polystable'' version, we reduce the polystability of $({\mathop{\rm gr}}({\cal E}),\varepsilon_{\mathop{\rm gr}},\phi_{\mathop{\rm gr}})$ to the respective results of \cite{HL1}, \cite{HL2}, \cite{Gi}, and \cite{Ma2}. Finally, one easily adapts the proof in \cite{HL2}, p.312, to show that a semistable oriented pair $({\cal E},\varepsilon,\phi)$ can be deformed into its graded object. If we call two semistable oriented pairs $({\cal E}_i,\varepsilon_i,\phi_i)$, $i=1,2$, \it gr-equivalent \rm if their associated graded objects are equivalent, then we see that the closed points of ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$ correspond to gr-equivalence classes of semistable oriented pairs of type $(P,\L,{\cal E}_0)$. \subsection{The ${\Bbb C}^*$-action on ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$} \label{Flips} The moduli space possesses a natural ${\Bbb C}^*$-action, given by $$z\cdot [{\cal E},\varepsilon,\phi]:=[{\cal E},\varepsilon, z\phi]=[{\cal E},z^{-r}\varepsilon,\phi].$$ The set of fixed points of this action can easily be described: It consists of classes $[{\cal E},0,\phi]$, $[{\cal E},\varepsilon,0]$, and of classes $[\ker(\phi)_{\max}\oplus{\cal E}^\prime, \varepsilon, \phi]$ with $0\neq \ker(\phi)_{\max}$. \par The ${\Bbb C}^*$-action is naturally linearized in an ample line bundle coming from the description of ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$ as GIT-quotient. This line bundle and the polarization which it represents may, however, depend on an integer $m$ chosen in the course of the construction. Nevertheless, we can state the following result which clarifies the birational geometry of the moduli spaces ${\cal M}^{ss}_{\delta}(X;{\cal E}_0,P)$ constructed in \cite{HL2}: \begin{Thm} Let $\delta_i\in{\Bbb Q}[x]$, $i=1,2$, be polynomials with positive leading coefficients. For a suitable choice of the polarization on ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$ the following properties hold true: \par {\rm i)} ${\cal M}^{ss}_{\delta_i}(X;{\cal E}_0,P)$, $i=1,2$, are ${\Bbb C}^*$-quotients of the master space ${\cal M}^{ss}_{(P,\L,{\cal E}_0)}$. \par {\rm ii)} ${\cal M}^{ss}_{\delta_1}(X;{\cal E}_0,P)$ and ${\cal M}^{ss}_{\delta_2}(X;{\cal E}_0,P)$ are related by a chain of generalized flips. \end{Thm} \begin{pf} Let $m$ be so large that a pair $({\cal E},\phi)$ is semistable w.r.t.\ $\delta_i$ if and only the pair $({\cal E}(m),\phi(m))$ is sectional semistable w.r.t.\ $\delta_i(m)$, $i=1,2$, and that all the other requirements needed in the constructions are met. Then our proof of Theorem~2.4.1 together with the results of Section~1 easily yields the assertions of the theorem. \end{pf} We note that the $\delta_i$ for which the corresponding set of ${\Bbb C}^*$-stable points meets the fixed point set of the ${\Bbb C}^*$-action, i.e., for which the corresponding set of ${\Bbb C}^*$-stable points contains stable oriented pairs of the type $[\ker(\phi)_{\max}\oplus{\cal E}^\prime, \varepsilon, \phi]$ with $0\neq \ker(\phi)_{\max}$ are uniquely determined. The corresponding polynomial is $\mathop{\rm rk}{\cal E}^\prime(P_{{\cal E}^\prime}-P_{\ker(\phi)_{\max}}/\mathop{\rm rk}\ker(\phi)_{\max})$. The associated moduli spaces ${\cal M}_{\delta_i}$ are those which show up ``at the top'' of the flips. \newpage % \vspace*{1.5cm}
1996-07-03T21:01:01
9607
alg-geom/9607003
en
https://arxiv.org/abs/alg-geom/9607003
[ "alg-geom", "math.AG" ]
alg-geom/9607003
Indranil Biswas
Indranil Biswas
A remark on the jet bundles over the projective line
AMS-Latex file, to appear in Mathematical Research Letters
null
null
null
null
This is a footnote of a recent interesting work of Cohen, Manin and Zagier, where they, among other things, produce a natural isomorphism between the sheaf of (n-1)-th order jets of the n-th tensor power of the tangent bundle of a Riemann surface equipped with a projective structure and the sheaf of differential operators of order n (on the trivial bundle) with vanishing 0-th order part. We give a different proof of this result without using the coordinates, and following the idea of this proof we prove: Take a line bundle L with $L^2 = T$ on a Riemann surface equipped with a projective structure. Then the jet bundle $J^n(L^n)$ has a natural flat connection with $J^n(L^n) = S^n(J^1(L))$. For any $m >n$ the obvious surjection $J^m(L^n) \rightarrow J^n(L^n)$ has a canonical splitting. In particular, taking $m = n+1$, one gets a natural differential operator of order $n+1$ from $L^n$ to $L^{-n-2}$.
[ { "version": "v1", "created": "Wed, 3 Jul 1996 18:56:39 GMT" } ]
2008-02-03T00:00:00
[ [ "Biswas", "Indranil", "" ] ]
alg-geom
\section{Introduction} Let $X$ be a Riemann surface equipped with a projective structure (i.e., a covering by coordinate charts such that the transition functions are of the form $z \longmapsto (az +b)/(cz+d)$). Let ${\cal L}$ be a line bundle on $X$ such that ${{\cal L}}^2 = T_X$. Let $J^m({{\cal L}}^{\otimes n}) \longrightarrow X$ denote the jet bundle of order $m$ for the line bundle ${{\cal L}}^{\otimes n}$. For $i \geq j$, there is a natural restriction homomorphism from $J^i({{\cal L}}^{n})$ onto $J^j({{\cal L}}^{n})$. We prove that for any $m\geq n$, the surjective homomorphism $$ J^m({{\cal L}}^{n}) \, \longrightarrow \, J^n({{\cal L}}^{n}) $$ admits a canonical splitting [Theorem 4.1]. As a consequence, for each $n\geq 0$ we construct a differential operator of order $n$ from ${{\cal L}}^{n -1}$ to ${{\cal L}}^{-n-1}$ whose symbol is the constant function $1$. Theorem 4.1 follows from the results on the jet bundles over the projective line established in Section 2. In \cite{CMZ} certain differential operators on a Riemann surface equipped with a projective structure are explicitly constructed (see (3.1)). As an application of the set-up we use to prove Theorem 4.1, in Section 3 we derive the differential operators constructed in \cite{CMZ}. The present work was inspired by \cite{CMZ}; in fact, it grew out of attempts to reconstruct the differential operators there without using the coordinates. \section{Constructions on the projective line} Let $V$ be a two dimensional vector space over ${\Bbb C}$. Let ${\Bbb P}(V)$ denote the projective space given by the space of all one dimensional quotients of $V$. Define the line bundle $$ L \, := \, {{\cal O}}_{{\Bbb P}(V)}(1) $$ on ${\Bbb P}(V)$, whose fiber over the quotient line $[q]$ is the line $[q]$ itself. We will recall the definition of the jet bundles for a line bundle. For a line bundle $\xi$ on a Riemann surface $X$, the $n$-th jet bundle, denoted by $J^n(\xi)$, is the rank $n+1$ vector bundle on $X$ whose fiber over $x \in X$ is $$ {\xi}_x {\otimes}_{{\Bbb C}} ({{\cal O}}_{X,x}/{\bf m}^{n+1}_x) $$ where ${\xi}_x$ is the fiber of $\xi$ over $x$, and ${{\cal O}}_{X,x}$ is the ring of functions defined around $x$ with ${\bf m}_x$ being the maximal ideal consisting of functions vanishing at $x$. The inclusion of ${\bf m}^{n+1}_x$ in ${\bf m}^n_x$ induces the following short exact sequence of vector bundle on $X$: $$ 0 \, \longrightarrow \, K^{\otimes n}_X\otimes \xi \, \longrightarrow \, J^n(\xi) \, \longrightarrow \, J^{n-1}(\xi) \, \longrightarrow \, 0 \leqno{(2.1)} $$ where $K_X$ is the canonical bundle of $X$. Let ${\cal V}$ denote the rank two trivial vector bundle on ${\Bbb P}(V)$ with $V$ as the fiber. For any $n \geq 0$, $S^n({\cal V})$ will denote the $n$-th symmetric power of ${\cal V}$, with $S^0({\cal V})$ being the trivial line bundle. \medskip \noindent {\bf Lemma 2.2.}\, {\it For any integer $n \geq 0$, the vector bundle $J^n(L^n)$ on ${\Bbb P}(V)$ is canonically isomorphic to the symmetric power $S^n({\cal V})$. For any $m \geq n$, the surjection $$ J^m(L^n) \, \longrightarrow \, J^n(L^n) \, \longrightarrow \, 0 $$ given by (2.1), admits a canonical splitting (i.e., a homomorphism from $J^n(L^n)$ to $J^m(L^n)$ such that the composition is identity on $J^n(L^n)$).} \medskip \noindent {\bf Proof.}\, Take any integer $n \geq 0$. Since $S^n(V) = H^0({\Bbb P}(V), L^n)$, for any $x \in {\Bbb P}(V)$ there is a natural homomorphism of $S^n(V)$ into the fiber $J^n(L^n)_x$ given by the restriction of sections to the $n$-th order infinitesimal neighborhood of $x$. Since for any integer $j$ with $j\leq n$, $$ \dim H^0({\Bbb P}(V), L^n \otimes {{\cal O}}_{{\Bbb P}(V)}(-jx)) \, - \, \dim H^0({\Bbb P}(V), L^n\otimes {{\cal O}}_{{\Bbb P}(V)}(-(j+1)x)) \, = \, 1 $$ the above obtained homomorphism must be an isomorphism. This proves the first part of the lemma. Take any integer $m$ such that $m\geq n$. We may restrict a section of $L^n$ to the $m$-th order infinitesimal neighborhood of $x$ to get a homomorphism from the vector space $S^n(V)$ (= $H^0({\Bbb P}(V), L^n)$) to the fiber $J^m(L^n)_x$. Now using the previous identification of $S^n(V)$ with $J^n(L^n)_x$ we get the required splitting. $\hfill{\Box}$ \medskip Setting $m = n+1$ in Lemma 2.2 we obtain the following: \medskip \noindent {\bf Corollary 2.3.}\, {\it For any integer $n \geq 0$, the exact sequence $$ 0\, \longrightarrow \, K^{n+1}_{{\Bbb P}(V)}\otimes L^n \, \longrightarrow \, J^{n+1}(L^n) \, \longrightarrow \, J^n(L^n)\, \longrightarrow \,0 $$ admits a canonical splitting.} \medskip Choose and fix a trivialization of ${\stackrel{2}{\wedge}V}$; this is equivalent to fixing a nonzero vector $\theta $ in ${\stackrel{2}{\wedge}}V$. The canonical bundle $K_{{\Bbb P}(V)} = L^{-2}\otimes {\rm det}\, {\cal V}$. Using the trivialization of ${\stackrel{2}{\wedge}}V$ we have, $K_{{\Bbb P}(V)} = L^{-2}$. The sheaf of differential operators of order $k$ from the sections of a line bundle $\xi$ to the sections of a line bundle $\eta$ is precisely the sheaf ${\rm Hom}(J^k(\xi), \eta)$. Consider the projection of $J^{n+1}(L^n)$ onto $K^{n+1}_{{\Bbb P}(V)}\otimes L^n = L^{-n-2}$ defining the splitting in Corollary 2.3. This gives a global differential operator of order $n+1$, $$ {{\cal D}}({n+1}) \, \in \, H^0({\Bbb P}(V), {\rm Diff}^{n+1}(L^n, L^{-n-2})) \leqno{(2.4)} $$ The symbol of a differential operator in ${\rm Diff}^{n+1}(L^n, L^{-n-2})$ is a section of of the line bundle $T^{n+1}_{{\Bbb P}(V)}\otimes L^{-2n-2} = {\cal O}_{{\Bbb P}(V)}$. Since the differential operator ${{\cal D}}(n+1)$ in (2.4) gives a splitting of the jet sequence -- it's symbol, which is a constant function, must be the constant function $1$. Let $SL(V)$ denote the subgroup of $GL(V) = {\rm Aut}(V)$ that acts trivially on ${\stackrel{2}{\wedge}}V$. The group $SL(V)$ has a natural action on ${\Bbb P}(V)$, and ${\rm Aut}({\Bbb P}(V)) = SL(V)/{{\Bbb Z}}_2$. There is a natural induced action of $SL(V)$ on any sheaf $J^m(L^n)$ that lifts the action on ${\Bbb P}(V)$. The isomorphism between $S^n({\cal V})$ and $J^n(L^n)$, and the splitting in Lemma 2.2, are both equivariant for this action. Indeed, this follows from the canonical nature of the construction in Lemma 2.2. So, in particular, the differential operator ${{\cal D}}(n)$ in (2.4) is an invariant for the action of $SL(V)$ on the space of all global sections of ${\rm Diff}^n(L^{n-1}, L^{-n-1})$. Note that ${{\cal D}}(n)$ is not an invariant for the action $GL(V)$ since the trivialization of ${\stackrel{2}{\wedge}}V$ was used in its construction. The identification between $K_{{\Bbb P}(V)}$ and $L^{-2}$ is not equivariant for the action of the center of $GL(V)$. Setting $n=2$ in Lemma 2.2 we get that $J^2(L^2) = S^2({\cal V})$. This implies that the homomorphism $$ \rho \, : \, H^0({\Bbb P}(V), J^2(L^2)) \, \longrightarrow \, H^0({\Bbb P}(V), L^2) $$ induced by the obvious projection, namely $J^2(L^2) \longrightarrow L^2$, is actually an isomorphism. Moreover, $\rho $ is the identity map of $S^2(V)$. Thus, after identifying the tangent bundle $T_{{\Bbb P}(V)}$ with $L^2$ using the trivialization of ${\stackrel{2}{\wedge}}V$, the Lemma 2.2 implies that $$ H^0({\Bbb P}(V) , T_{{\Bbb P}(V)}) \, = \, H^0({\Bbb P}(V), J^2(L^2)) \, = \, S^2(V) \leqno{(2.5)} $$ The Lie-bracket operation equips the vector space $H^0({\Bbb P}(V), T_{{\Bbb P}(V)})$ with the structure of a Lie algebra. The action of $SL(V)$ on ${\Bbb P}(V)$ gives a Lie algebra homomorphism from its Lie algebra, $sl(V)$, into $H^0({\Bbb P}(V), T_{{\Bbb P}(V)})$. This homomorphism is actually an isomorphism. The Lie algebra structure on $S^2(V)$ induced by the equality (2.5) can be seen directly as follows: using contraction, $S^2(V)$ maps $V^*$ into $V$; on the other hand, $\theta $ identifies $V^*$ with $V$ -- combining these, the resulting homomorphism from $S^2(V)$ into $sl(V)$ is an isomorphism. Let $C \in S^2(H^0({\Bbb P}(V), T_{{\Bbb P}(V)}))$ be the Casimir of the Lie algebra $H^0({\Bbb P}(V), T_{{\Bbb P}(V)})$. The section $C$ is evidently an invariant for the obvious action of $SL(V)$ on the vector space $S^2(H^0({\Bbb P}(V), T_{{\Bbb P}(V)}))$. For a section $s$ of $T_{{\Bbb P}(V)}$, let $L_{s}$ denote the Lie derivative with respect to $s$. The (second order) Lie derivative with respect to $s{\otimes} s$ is defined to be $L_{s}\circ L_{s}$. Thus $C$ acts as a differential operator, denoted by $L_C$, on all vector bundles associated to ${\Bbb P}(V)$. This differential operator is actually of order zero (i.e., a constant scalar multiplication). Using (2.5) and Lemma 2.2 we get that $$ S^2(H^0({\Bbb P}(V), T_{{\Bbb P}(V)})) \, = \, H^0({\Bbb P}(V) ,S^2(J^2(T_{{\Bbb P}(V)}))) $$ Let ${\bar C} \in H^0({\Bbb P}(V), S^2(J^2(T_{{\Bbb P}(V)})))$ be the element corresponding to the Casimir $C$; ${\bar C}$ is actually the Casimir for the Lie algebra $S^2(V)$ (which is the fiber of $J^2(T_{{\Bbb P}(V)})$). Let $p : J^2(T_{{\Bbb P}(V)}) \longrightarrow T_{{\Bbb P}(V)}$ be the obvious projection. If (locally) $$ {\bar C} \, = \, \sum_{i} A_i\otimes A_i \leqno{(2.6)} $$ where $A_i$ are local sections of $J^2(T_{{\Bbb P}(V)})$, consider the operator $$ L_{{\bar C}} \, = \, \sum L_{p(A_i)}\circ L_{p(A_i)} $$ with $L_{\phi (A_i)}$ being the Lie derivative with respect to the vector field $p(A_i)$. It is easy to check that the operator $L_{{\bar C}}$ does not depend upon the choice of the decomposition of ${\bar C}$, and that $L_{C} = L_{{\bar C}}$. \section{Jets of the trivial line bundle on the projective line} Let ${\rm Diff}^n({\cal O} , {\cal O}) = J^n({\cal O})^*$ be the sheaf of differential operators on the trivial line bundle over ${\Bbb P}(V)$. The symbol map, which is the dual of the injection in (2.1), gives a surjective homomorphism $$ \sigma \, : \, {\rm Diff}^n({\cal O}, {\cal O}) \, \longrightarrow \, T^n_{{\Bbb P}(V)} $$ Let $\gamma $ denote the obvious projection of $J^{n-1}(T^n_{{\Bbb P}(V)})$ onto $T^n_{{\Bbb P}(V)}$. For any $n \geq 1$, let $J^n_0({\cal O}) \subset J^n({\cal O})$ be the kernel of the obvious homomorphism from $J^n({\cal O})$ onto $J^0({\cal O}) = {\cal O}$. This subsheaf has a canonical splitting given by the constant functions. Define the subsheaf, ${\rm Diff}^n_0({\cal O}, {\cal O}) := J^n_0({\cal O})^*$, of ${\rm Diff}^n({\cal O},{\cal O})$. For a function $f$ on ${\Bbb C}$ and any integer $n \geq 1$, in Proposition 1 (page 4) of \cite{CMZ} (where it is called ${{\cal L}}_{-n}(f)$) the following differential operator of order $n$ is constructed: $$ D_{n}(f) \, := \, \sum_ {i=0}^{n-1} {{(2n-i)!}\over {i!(n-i)!(n-i-1)!}} f^{(i)}{\partial}^{n-i} \leqno{(3.1)} $$ with $f^{(i)} = {\partial}^{i}f$ being the $i$-th derivative of $f$. The operator ${D}_{n}$ has the property that for any M\"obius transformation, $M(z) = (az+b)/(cz+d)$, of ${\Bbb C}{\Bbb P}^1$, the following equality holds: $$ {D}_{n}(f)\circ M \, = \, {D}_{n}(M_n.(f \circ M)) \leqno{(3.2)} $$ where $M_n(z) =(cz+d)^{2n}$. Thus ${D}_{n}$ is a $SL(V)$ equivariant ${{\cal O}}_{{\Bbb P}(V)}$ linear isomorphism (in other words, a canonical isomorphism) $$ \phi \, : \, J^{n-1}(T^n_{{\Bbb P}(V)}) \, \longrightarrow \, {\rm Diff}^n_0({\cal O},{\cal O}) \leqno{(3.3)} $$ The operator ${D}_{n}$ has the further property that $\sigma \circ \phi = \gamma $. It is shown in \cite{CMZ} that this splitting condition together with the automorphic property (3.2) actually determine the operator ${D}_{n}$. In this section we want to deduce the above result of \cite{CMZ} in the set-up of Section 2. Take a point $x \in {\Bbb P}(V)$. The long exact sequence of cohomology for the exact sequence of sheaves on ${\Bbb P}(V)$ $$ 0 \, \longrightarrow \, {{\cal O}}(-(n+1).x) \, \longrightarrow \, {\cal O} \, \longrightarrow \, J^n({\cal O})_x \, \longrightarrow \, 0 $$ gives the equality $$ J^n_0({\cal O})_x \, = \, H^1({\Bbb P}(V) , {{\cal O}}(-(n+1).x)) $$ where $J^n_0({\cal O})_x$ is the fiber of $J^n_0({\cal O})$ over $x$. Choose and fix an isomorphism between the two line bundles ${{\cal O}}(x)$ and $L$. Since the fiber ${{\cal O}}(x)_x = T_{{\Bbb P}(V), x} = L^2_x$, fixing such an isomorphism is equivalent to fixing a nonzero vector $\omega $ in $L_x$. Using Serre duality for ${{\cal O}}(-(n+1).x)$, and then identifying $K_{{\Bbb P}(V)}$ with ${{\cal O}}(-2x)$ using $\omega $, we have $$ {\rm Diff}^n_0({\cal O}, {\cal O})_x \, = \, H^0({\Bbb P}(V), {\cal O}((n-1).x)) \, = \, H^0({\Bbb P}(V), L^{n-1}) \, = \, S^{n-1}(V) \leqno{(3.4)} $$ Consider the restriction of sections of $T^n_{{\Bbb P}(V)}$ to the $(n-1)$-th order infinitesimal neighborhood of $x$, namely $$ \beta \, : \, S^{2n}(V) \, = \, H^0({\Bbb P}(V), T^n_{{\Bbb P}(V)}) \, \longrightarrow \, J^{n-1}(T^n_{{\Bbb P}(V)})_x \leqno{(3.5)} $$ which is clearly a surjective homomorphism. Indeed, in the proof of Lemma 2.2 we saw that $S^{2n}(V)$ surjects onto $J^{2n}(L^{2n})_x = J^{2n}(T^n_{{\Bbb P}(V)})_x$. We want to identify the kernel of the homomorphism $\beta $. The symplectic form on $V$ given by the trivialization of ${\stackrel{2}{\wedge}}V$ identifies $V$ with $V^*$. Let $v$ be the vector in the kernel of the quotient homomorphism $V \longrightarrow L_x$ which corresponds to $\omega $ using the symplectic form on $V$. (This vector $v \in V$ corresponds to the section of the sheaf ${{\cal O}}(x)$ given by the constant function $1$.) Consider the homomorphism, $m_v : S^n(V) \longrightarrow S^{2n}(V)$, defined by multiplication with $v^{\otimes n}$. The inclusion $m_v$ corresponds to the natural inclusion of $H^0({\Bbb P}(V), {{\cal O}}(n.x))$ into $H^0({\Bbb P}(V), {\cal O}(2n.x))$. The image of $m_v$ is precisely the kernel of $\beta $ in (3.5). Consider the homomorphism $i_{\omega } : S^{2n}(V) \longrightarrow S^{n-1}(V)$ given by the contraction with ${\omega }^{\otimes (n+1)}$. (The vector $\omega $ is considered as an element of $V^*$.) This homomorphism vanishes on the image $m_v(S^n(V))$. Indeed, this follows from the fact that $\omega (v) = 0$. Thus using the equality (3.4) and $i_{\omega }$ we have the homomorphism $$ {\phi }_x \, : \, J^{n-1}(T^n_{{\Bbb P}(V)})_x \, \longrightarrow \, {\rm Diff}^n_0({\cal O}, {\cal O})_x $$ It is easy to check that the homomorphism ${\phi }_x$ does not depend upon the choice of the nonzero vector $\omega \in L_x$. The resulting homomorphism $\phi $ from $J^{n-1}(T^n_{{\Bbb P}(V)})$ to ${\rm Diff}^n_0({\cal O},{\cal O})$ satisfies the condition that $\sigma \circ \phi = \gamma $. The canonical nature of the construction of $\phi $ ensures that it is equivariant for the action of $SL(V)$. Since ${\phi }_x$ is an isomorphism, $\phi $ is an isomorphism. Let $U \subset SL(V)$ be the unipotent subgroup which fixes the vector $v$. Let ${\frak n}$ be the nilpotent part of the Lie algebra of $U$. Let $N$ denote the unique element in $\frak n$ which maps a preimage of $\omega $ (in $V$) to $v$. For any $0\leq i \leq 2n$, the image of $S^i(V)$ in $S^{2n}(V)$, for the homomorphism given by the multiplication with $v^{2n-i}$, is denoted by $S^{2n}_i(V)$. In this notation, $N$ maps $S^{2n}_{i+1}(V)$ onto $S^{2n}_i(V)$; the resulting homomorphism from $S^{i+1}(V)$ onto $S^i(V)$ is the contraction by $\omega $. From this it is easy to deduce that any homomorphism from $S^{2n}(V)/m_v(S^n(V))$ to $S^{n-1}(V)$, which is equivariant for the actions of $N$, must be a scalar multiple of $i_{\omega }$. Now the condition, $\sigma \circ \phi = \gamma $, uniquely determines the homomorphism $\phi $. \section{Jets on a Riemann surface with a projective structure} Let $X$ be a Riemann surface, not necessarily compact. A {\it projective structure} on $X$ is a maximal atlas of holomorphic coordinate charts, $\{U_{\alpha }, f_{\alpha }\}_{\alpha \in I}$, covering $X$, such that any $f_{\alpha }$ maps $U_{\alpha }$ biholomorphically onto some analytic open set in ${\Bbb P}(V)$ and the transition function $f_{\alpha }\circ f^{-1}_{\beta }$, for any $\alpha , \beta \in I$, is a restriction of an automorphism of ${\Bbb P}(V)$, \cite{G}, \cite{D}, \cite{T}. We note that any Riemann surface admits a projective structure, since, from the uniformization theorem, the universal cover has a natural projective structure. It is know that for any projective structure, it is possible to choose a sub-cover such that the transition functions have a compatible lift to $SL(V)$ (from ${\rm Aut}({\Bbb P}(V))$). Actually, more than one inequivalent lifts are possible. For a compact Riemann surface, the set of equivalence classes of lifts correspond to the set of square roots of the canonical bundle (called {\it theta characteristics}) \cite{G}, \cite{T}. Henceforth, by a projective structure we will always mean a lift of the structure group to $SL(V)$. Let $X$ be a Riemann surface equipped with a projective structure in the above sense. Since the natural action of $SL(V)$ on ${\Bbb P}(V)$ lifts to the bundle $L$, the projective structure gives a line bundle on $X$ associated to $L$. Let ${\cal L}$ denote this line bundle on $X$. Since the isomorphism between $L^2$ and $T_{{\Bbb P}(V)}$ is $SL(V)$ equivariant, we have ${{\cal L}}^{\otimes 2} = T_X$. Since $J^n(L^n)$ on ${\Bbb P}(V)$ is a trivial bundle (Lemma 2.2), it has a natural flat connection, which is equivariant under the action of $SL(V)$. We now have the following consequence of Lemma 2.2, Corollary 2.3 and (2.4): \medskip \noindent {\bf Theorem 4.1.}\, {\it For any $n \geq 0$, the jet bundle $J^n({{\cal L}}^n)$ on $X$ has a natural flat connection, and $S^n(J^1({\cal L})) = J^n({{\cal L}}^n)$, with the identification being compatible with the flat connections. For any $m\geq n$, the natural surjection $$ J^m({{\cal L}}^n) \, \longrightarrow \, J^n({{\cal L}}^n) \, \longrightarrow \, 0 $$ has a canonical splitting. Setting $m=n+1$, a global differential operator of order $n+1$ $$ {{\cal D}}_X (n+1) \, \in \, H^0(X, {\rm Diff}^n_X({{\cal L}}^{n}, {{\cal L}}^{-n-2})) $$ is obtained. The symbol of ${{\cal D}}_X(n)$ is the constant function $1$. The fibers of $J^2(T_X)$ have the structure of a Lie algebra compatible with the flat connection on $J^2(T_X)$. The Lie derivative action of the Casimir $$ C_X \, \in \, H^0(X, S^2(J^2(T_X))) $$ on any tensor power of ${\cal L}$ is a multiplication by a constant scalar.} \medskip Similarly, since the isomorphism $\phi $ in (3.3) is $SL(V)$ equivariant, we have an isomorphism of vector bundles on $X$ $$ {\phi }_X \, : \, J^{n-1}(T^n_X) \, \longrightarrow \, {\rm Diff}^n_0({\cal O}, {\cal O}) $$ (${\rm Diff}^n_0 ({\cal O},{\cal O}) \subset {\rm Diff}^n_X({\cal O},{\cal O})$ is the canonical complement of ${\rm Diff}^0_X({\cal O},{\cal O})$) such that the composition of the symbol map on ${\rm Diff}^n_X({\cal O}, {\cal O})$ with the isomorphism ${\phi }_X$ is the natural projection of $J^{n-1}(T^n_X)$ onto $T^n_X$.
1996-07-03T13:29:54
9607
alg-geom/9607002
en
https://arxiv.org/abs/alg-geom/9607002
[ "alg-geom", "math.AG" ]
alg-geom/9607002
Carlos Simpson
Carlos Simpson
A relative notion of algebraic Lie group and applications to $n$-stacks
83 pages Latex
null
null
null
null
If $S$ is a scheme of finite type over $k=\cc $, let $\Xx /S$ denote the big etale site of schemes over $S$. We introduce {\em presentable group sheaves}, a full subcategory of the category of sheaves of groups on $\Xx /S$ which is closed under kernel, quotient, and extension. Group sheaves which are representable by group schemes of finite type over $S$ are presentable; pullback and finite direct image preserve the notions of presentable group sheaves; over $S=Spec (k)$ then presentable group sheaves are just group schemes of finite type over $Spec(k)$; there is a notion of connectedness extending the usual notion over $Spec(k)$; and a presentable group sheaf $G$ has a Lie algebra object $Lie(G $. If $G$ is a connected presentable group sheaf then $G/Z(G)$ is determined up to isomorphism by the Lie algebra sheaf $Lie (G)$. We envision the category of presentable group sheaves as a generalisation relative to an arbitrary base scheme $S$, of the category of algebraic Lie groups over $Spec (k)$. The notion of presentable group sheaf is used in order to define {\em presentable $n$-stacks} over $\Xx$. Roughly, an $n$-stack is presentable if there is a surjection from a scheme of finite type to its $\pi_0$ (the actual condition on $\pi_0$ is slightly more subtle), and if its $\pi_i$ (which are sheaves on various $\Xx /S$) are presentable group sheaves. The notion of presentable $n$-stack is closed under homotopy fiber product and truncation. We propose the notion of presentable $n$-stack as an answer in characteristic zero for A. Grothendieck's search for what he called ``schematization of homotopy types''.
[ { "version": "v1", "created": "Wed, 3 Jul 1996 11:33:02 GMT" } ]
2008-02-03T00:00:00
[ [ "Simpson", "Carlos", "" ] ]
alg-geom
\section*{A relative notion of algebraic Lie group and applications to $n$-stacks} Carlos Simpson\newline {\small Laboratoire Emile Picard (UMR 5580 CNRS) \newline Universit\'e Paul Sabatier\newline 31062 Toulouse CEDEX, France} \bigskip Let ${\cal X}$ be the big etale site of schemes over $k={\bf C}$. If $S$ is a scheme of finite type over $k$, let ${\cal X} /S$ denote the big etale site of schemes over $S$. The goal of this paper is to introduce a full subcategory of the category of sheaves of groups on ${\cal X} /S$, which we will call {\em the category of presentable group sheaves} (\S 2), with the following properties. \newline 1. \, The category of presentable group sheaves contains those group sheaves which are representable by group schemes of finite type over $S$ (Corollary \ref{uvw}). \newline 2. \, The category of presentable group sheaves is closed under kernel, quotient (by a normal subgroup sheaf which is presentable), and extension (Theorem \ref{I.1.e}). \newline 3. \, If $S'\rightarrow S$ is a morphism then pullback takes presentable group sheaves on $S$ to presentable group sheaves on $S'$ (Lemma \ref{I.1.h}). \newline 4. \, If $S'\rightarrow S$ is a finite morphism then direct image takes presentable group sheaves on $S'$ to presentable group sheaves on $S$ (Lemma \ref{I.1.i}). \newline 5. \, If $S=Spec (k)$ then presentable group sheaves are just group schemes of finite type over $Spec (k)$ (Theorem \ref{I.1.m}). In particular if ${\cal G}$ is a presentable group sheaf over any $S$ then the pullback to each point $Spec (k )\rightarrow S$ is an algebraic group. \newline 6. \, There is a notion of connectedness extending the usual notion over $Spec(k )$ and compatible with quotients, extensions, pullbacks and finite direct images; and a presentable group sheaf ${\cal G}$ has a largest connected presentable subsheaf ${\cal G} ^0\subset {\cal G}$ which we call the {\em connected component} (Theorem \ref{I.1.o}). \newline 7. \, A presentable group sheaf ${\cal G}$ has a Lie algebra object $Lie({\cal G} )$ (Theorem \ref{lmn}) which is a vector sheaf with bracket operation (see below for a discussion of the notion of vector sheaf---in the case $S=Spec (k)$ it is the same thing as a finite dimensional $k$-vector space). \newline 8. \, If ${\cal G}$ is a connected presentable group sheaf then ${\cal G} /Z({\cal G} )$ is determined up to isomorphism by the Lie algebra sheaf $Lie ({\cal G} )$ (where $Z({\cal G} )$ denotes the center of ${\cal G}$). This is Theorem \ref{abc} below. \bigskip We envision the category of presentable group sheaves as a generalisation relative to an arbitrary base scheme $S$, of the category of algebraic Lie groups over $Spec ({\bf C} )$. We mention here a few questions related to the analogy with classical algebraic groups. Property 8 poses an obvious existence problem: given a Lie algebra object in the category of vector sheaves, does it come from a presentable group sheaf with vector sheaf center? I don't know the answer to this question. We do know, however, that $Aut(L)$ is a presentable group sheaf (Lemma \ref{AutLie}). Another question is the existence of a ``universal covering'', i.e. a morphism $\tilde{{\cal G} }\rightarrow {\cal G}$ surjective with finite kernel such that for any other such morphism ${\cal F} \rightarrow {\cal G}$ there is a factorization $\tilde{{\cal G} } \rightarrow {\cal F} \rightarrow {\cal G}$. There are obvious questions about the generalisation of the theory of representations to the case of presentable group sheaves. The first among these is whether there always exists a faithful representation into $Aut(V)$ for $V$ a vector sheaf. I suspect that the answer is no, but don't have a counterexample. For connected group sheaves this problem concerns only the center, because we always have the adjoint representation of ${\cal G}$ on $Lie ({\cal G} )$. Beyond the question of the description of the representations, there is also the question of whether a suitable tannakian theory exists, namely given a group ${\cal G} \subset Aut (V)$, is ${\cal G}$ defined as the stabilizer of some ${\cal G}$-invariant sub-vector-sheaf $U$ in a tensor power of $V$? The motivation for introducing presentable group sheaves comes from the theory of homotopy types over $Spec ({\bf C} )$, or what Grothendieck called ``schematization of homotopy types'' in \cite{Grothendieck}. We will discuss the application to this theory at the end of the paper---note also that it is explained in essentially the same way in \cite{kobe} where some applications to nonabelian de Rham cohomology are also announced. Briefly, the considerations are as follows. A homotopy type over ${\cal X}$ (which we call an ``$n$-stack'') is a presheaf of topological spaces on ${\cal X}$ satisfying a homotopic descent condition (``fibrant'' in the terminology of Jardine \cite{Jardine1}, cf \cite{kobe}). This condition is the generalisation of the descent condition that goes into the definition of $1$-stack. An $n$-stack or fibrant presheaf $T$ has homotopy sheaves as follows. First, $\pi _0(T)$ is a sheaf of sets on ${\cal X}$. Then for $i\geq 1$ if $S\in {\cal X}$ and $t\in T(S)$, $\pi _i (T|_{{\cal X} /S},t)$ is a sheaf of groups on ${\cal X} /S$ (abelian if $i\geq 2$). In the fibrant presheaf point of view, these homotopy sheaves are the sheafifications of the presheaves which one defines in the obvious way. These things satisfy the same sorts of properties as in the homotopy theory of spaces. In particular there are notions of homotopy fiber products and (as special cases) homotopy fibers and loop or path spaces. The homotopy groups of the homotopy fiber of a morphism fit into the usual long exact sequence (and there is a similar exact sequence for homotopy fiber products in general). There are also notions of morphism spaces $Hom (T,T')$ which are spaces or $n$-groupoids (depending on the point of view) and internal morphism objects $\underline{Hom}(T,T')$ which are $n$-stacks whose global sections are the morphism spaces. The main particularity of this situation is that $\pi _0(T)$ can be nontrivial and not just the union a set of points. Because of this, one must consider basepoints not only in $T(Spec (k ))$ but in $T(S)$ for any scheme $S$ (say, of finite type) in order to get the full picture of $T$. One is thus lead to consider sheaves of groups on ${\cal X} /S$. We would like to define a restricted class of $n$-stacks or fibrant presheaves of spaces which we will call {\em presentable}. We would like this category to be closed under homotopy fiber products and also under the truncation (or coskeleton) operations of eliminating the homotopy groups above a certain level. From these requirements it follows that the condition for inclusion in the class of presentable presheaves of spaces should be expressed solely in terms of the homotopy group sheaves. From the exact sequences for homotopy fibers or more generally fiber products, one can see that the category of group sheaves allowable as homotopy group sheaves of presentable spaces must be closed under kernel, cokernel and extension. We would like our allowable group sheaves to be the algebraic Lie groups when the base space is $Spec (k )$, and of course for doing anything useful we need notions of connectedness and an infinitesimal (Lie algebra) picture. These are the reasons which lead us to look for a notion of sheaf of groups on ${\cal X} /S$ with the properties listed above. I should add a note of caution about the terminology, for we propose the terminology {\em presentable group sheaf} and also {\em presentable $n$-stack}. If ${\cal G}$ is a sheaf of sets on ${\cal X} /S$ (i.e. a $0$-stack) which happens to have a group structure, then the condition that ${\cal G}$ be presentable as a $0$-stack is {\em not} the same as the condition that ${\cal G}$ be a presentable group sheaf on ${\cal X} /S$. The right way to think of a sheaf of groups is as corresponding to a $1$-stack which we can denote $K({\cal G} , 1)$ or $B{\cal G}$ (with a morphism to $S$). From this point of view the terminologies are compatible: ${\cal G}$ is a presentable group sheaf over $S$ if and only if $K({\cal G} , 1)$ is a presentable $1$-stack. Let's look more carefully at the reasoning that leads to our definition of presentable $n$-stack. What are we going to do with a presentable $n$-stack $T$? If $W$ is (the $n$-truncation of) a finite CW complex considered as a constant $n$-stack on ${\cal X}$ then we can look at the $n$-stack $Hom (W, T)$. This is the {\em nonabelian cohomology of $W$ with coefficients in $T$}. If $T= K({\cal O} , n)$ is the Eilenberg-MacLane presheaf with homotopy group sheaf equal to the structure sheaf of rings ${\cal O}$ on ${\cal X}$ in degree $n$, then $\pi _0Hom (W, T)$ is just the cohomology $H^n(W, {\bf C} )$---or rather the sheaf on ${\cal X}$ represented by this vector space. Similarly, if $G$ is a group scheme over ${\bf C}$ then for $T=K(G, 1)=BG$ we get that $Hom (W, G)$ is the moduli stack for flat principal $G$-bundles or equivalently representations $\pi _1(W)\rightarrow G$. We hope to obtain an appropriate mixture of these cases by considering a more general class of $n$-stacks $T$. In particular we would like to have a {\em Kunneth formula} for two CW complexes $V$ and $W$, $$ \underline{Hom} (U, \underline{Hom}(V,T))=\underline{Hom}(U\times V, T). $$ One can imagine for example the problem of trying to compute the moduli stack of flat principal $G$-bundles on $U\times V$ in terms of a Kunneth formula as above. One is forced to consider the cohomology of $U$ with coefficients in the moduli stack $T'=\underline{Hom}(V,BG)$, and this stack is not necessarily connected ($\pi _0(T')$ is roughly speaking the moduli space of principal $G$-bundles). The Kunneth formula is not an end in itself, as it is rare for a space to decompose into a product. It points the way to a ``Leray-Serre theory'' which could be more generally useful. If $W\rightarrow U$ is a morphism we would be led to consider a relative morphism stack $T'=\underline{Hom}(W/U, T) \rightarrow U$ and then try to take the $n$-stack of sections $U\rightarrow T'$, a sort of {\em nonabelian cohomology with twisted coefficients}. I haven't fully thought about this yet (and in particular not about the de Rham theory---see below---which seems to be significantly more complicated than that which is needed in the constant coefficient case, for example to make sense of the Kunneth formula). The motivation for all of this is to be able to do geometric versions of the nonabelian cohomology in the case where $W$ is, say, a smooth projective variety. It is announced with some sketches of proofs in \cite{kobe}, how to get a de Rham version of the morphism space $\underline{Hom}(W_{DR}, T)$ when $T$ is a presentable $n$-stack. We want of course to have the (analytic) isomorphism between de Rham and Betti cohomology. Needless to say, this will not work for an arbitrary $n$-stack $T$ on $X$ (for example if one takes $T=W$ to a constant stack associated to a CW complex which is an algebraic variety then there will probably be nothing in $Hom (W_{DR}, W)$ corresponding to the identity in $Hom (W, W)$). We need to impose conditions on $T$ which guarantee that it is reasonably close to the examples $K({\cal O} , n)$ or $K(G, 1)$ given above (in these cases, the de Rham-Betti isomorphism works as is already well known). As a first approach, the condition we want seems to be that the homotopy group sheaves should be representable by group schemes over the base $S$. In the case where $T$ is the moduli stack of flat principal $G$-bundles on a space $V$, encountered above when looking at the Kunneth formula, the $\pi _1$ sheaves are indeed representable (the moduli stack is an algebraic stack). Unfortunately the condition of being representable is not stable under cokernels, but as explained above this is important if we want our notion of good $n$-stack to be stable under homotopy fiber products. Before going directly to the conclusion that we need a category stable under kernels, cokernels and extensions, we can analyze a bit more precisely just what is needed. Notice first of all that the algebraic de Rham theory is not going to work well in the case of higher cohomology with coefficients in the multiplicative group scheme, i.e. when $T= K({\bf G}_m, n)$ for $n\geq 2$. I won't go into the explanation of that here! Thus, at least for the algebraic de Rham theory we would like to have an appropriate notion of unipotent abelian group sheaf. Not yet having come up with a reasonable general theory of this, we can replace this notion by the (possibly more restrictive) notion of {\em vector sheaf}. The notion of vector sheaf is explained in \S 4 below. The reader may actually wish to start by reading this section, since the theory of vector sheaves is in some sense a paradigm, applicable only for abelian group sheaves, of what we are trying to do in general. The notion of vector sheaf was introduced by A. Hirschowitz \cite{Hirschowitz} who called it ``U-coherent sheaf''. He defined the category of U-coherent sheaves as the smallest abelian category of sheaves of abelian groups containing the coherent sheaves (note that the category of coherent sheaves is not abelian on the big etale site or any big site). We take a more constructive approach, defining the notion of vector sheaf in terms of presentations, although in the end the two notions are equivalent. The notion of vector sheaf doesn't work too nicely in characteristic $p>0$, basically because the Frobenius automorphism of the sheaf ${\cal O}$ is not linear, so the linear structure is no longer encoded in the sheaf structure. As we try in the beginning of the paper to put off the hypothesis of characteristic zero as long as possible, and as the notion of vector sheaf comes into the analysis at a later stage (the infinitesima study related to properties 7 and 8 listed above), I have decided not to put the section on vector sheaves at the beginning. Still, it is essentially self-contained for the reader who wishes to start there. In considering the algebraic de Rham theory we will only be looking at $n$-stacks $T$ with $\pi _i(T,t)$ a vector sheaf on $S$ for $t\in T(S)$ and $i\geq 2$. What does this mean for our restriction on $\pi _1(T, t)$? Going back to the question of stability under fiber products, we see from looking at the long exact homotopy sheaf sequence that the minimum that is absolutely necessary is that our class of group sheaves $G$ be stable under central extension by a vector sheaf. On the other hand it also must be stable under taking kernels. One could thus hope to make good on a {\em minimalist approach} saying that we should look at the category of group sheaves generated by representable group sheaves (affine, say---this again would be needed to make the de Rham theory work), under the operations of kernel and central extension by a vector sheaf. A vector sheaf always has a presentation as the cokernel of a morphism of {\em vector schemes}, i.e. representable vector-space objects over the base $S$ (these are sometimes called {\em linear spaces} in the complex analytic category \cite{Grauert} \cite{Fischer} \cite{Axelsson-Magnusson}). The most natural approach then is to say, suppose a group sheaf $G$ has a presentation as the cokernel of a morphism $F_2 \rightarrow F_1$ of representable group sheaves, and suppose $E$ is a central extension of $G$ by a vector sheaf $U$ which is itself the cokernel of a morphism $V_2 \rightarrow V_1$ of vector schemes. Then try to combine these into a presentation of $E$ with, for example, a surjection $V_1\times F_1\rightarrow E$. The problem (which I was not able to solve although I don't claim that it is impossible) is then to lift the multiplication of $E$ to an associative multiplication on $V_1\times F_1$. As I didn't see how to do this, a slightly more general approach was needed, wherein we consider groups which have presentations by objects where the multiplication lifts but not necessarily to a multiplication satisfying the associativity property. This is the reasoning that leads to the definition of {\em $S$-vertical morphism:} \, a morphism where one can lift things such as multiplications in a nice way cf \S 2. We finally come to the definition of {\em presentable group sheaf} as a group sheaf $G$ which admits a vertical surjection $X\rightarrow G$ from a scheme of finite type over $S$, and such that there is a vertical surjection $R\rightarrow X\times _{G}X$ again from a scheme of finite type over $S$. One could, on the other hand, take a {\em maximalist approach} and try to include anything that seems vaguely algebraic. This would mean, for example, looking at sheaves $G$ such that there are surjections (in the etale sense, although not necessarily etale morphisms) $X\rightarrow G$ and $R\rightarrow X\times _GX$ with $X$ and $R$ schemes of finite type over the base $S$. We call this condition P2. This might also work (in fact it might even be the case that a P2 group sheaf is automatically presentable). However, I was not able to obtain a reasonable infinitesimal analysis which could lead, for example, to the notion of connected component---though again, I don't claim that this could never work. In a similar vein, one might point out that there is a fairly limited range of situations in which we use the lifting properties going into the definition of verticality. I have chosen to state the condition of verticality in what seems to be the most natural setting, but this leads to requiring that many more lifting properties be satisfied than what we actually use. One could rewrite the definition of verticality to include only those lifting properties that we use afterward. It might be interesting to see if this change makes any difference in which group sheaves are allowed as presentable. All in all, the definitions we give here of presentable group sheaf and of presentable $n$-stack are first attempts at giving useful and reasonable notions, but is is quite possible that they would need to be altered in the future in view of applications. A word about the characteristic of the ground field (or base scheme). While our aim is to work over a field of characteristic zero, there are certain parts of our discussion valid over any base scheme, namely those concerning the abstract method for defining conditions of presentability. When it comes down to finding conditions which result in a nice theory (and in particular which result in a theory having the required local structure) we must restrict ourselves to characteristic zero. It is possible that a variant could work nicely in positive characteristic, so we will present the first part of the argument concerning the definition of presentability (which is valid over any base scheme), in full generality (\S 1) before specifying in characteristic zero which morphisms we want to use in the presentations (\S 2). Actually the definition given in \S 2 works in any characteristic but we can only prove anything about local properties in characteristic zero (\S\S 4-9), so it is probably the ``right'' definition only in characteristic zero. With an appropriate different definition of verticality (certainly incorporating divided powers) what we do in these later sections might be made to work in any characteristic. \subnumero{Notations} We fix a noetherian ground ring $k$, for sections 1-3. From section 4 on, we assume that $k$ is an uncountable field of characteristic zero, and we may when necessary assume that the ground field is $k={\bf C}$. Let ${\cal X}$ denote the site of noetherian schemes over $k$ with the etale topology (this is known as the ``big etale site''). If $S\in {\cal X}$ then we denote by ${\cal X} /S$ the site of schemes over $S$ (again with the etale topology). A {\em sheaf} on ${\cal X}$ means (unless otherwise specified) a sheaf of sets. For a sheaf of groups, we sometimes use the terminology {\em group sheaf}. We will confuse notations between an object of ${\cal X}$ and the sheaf it represents. Denote by $\ast$ the sheaf on ${\cal X}$ with values equal to the one-point set; it is represented by $Spec (k)$ (and we can interchange these notations at will). If $S$ is a sheaf on ${\cal X}$ (most often represented by an object) then we have the site ${\cal X} /S$ of objects of ${\cal X}$ together with morphisms to $S$. There is an equivalence between the notions of sheaf on ${\cal X} /S$ and sheaf on ${\cal X}$ with morphism to $S$. Since we will sometimes need to distinguish these, we introduce the following notations. If ${\cal F}$ is a sheaf on ${\cal X}$ then its {\em restriction up} to ${\cal X} /S$ is denoted by ${\cal F} |_{{\cal X} /S}$, with the formula $$ {\cal F} |_{{\cal X} /S}(Y\rightarrow S)= {\cal F} (Y). $$ If ${\cal F}$ is a sheaf on ${\cal X} /S$ then we denote by $Res_{S/\ast}{\cal F}$ the corresponding sheaf on ${\cal X}$ together with its morphism $$ Res_{S/\ast}{\cal F} \rightarrow S. $$ It is defined by the statement that $Res_{S/\ast}{\cal F} (Y)$ is equal to the set of pairs $(a, f)$ where $a: Y\rightarrow S$ and $f\in {\cal F} ( Y\stackrel{a}{\rightarrow} S)$. We call this the {\em restriction of ${\cal F}$ from $S$ down to $\ast$}. More generally if $S'\rightarrow S$ is a morphism and if ${\cal F}$ is a sheaf on ${\cal X} /S'$ then we obtain a sheaf $Res _{S'/S}{\cal F}$ on ${\cal X}/S$ called the {\em restriction of ${\cal F}$ from $S'$ down to $S$}. The operations of restriction up and restriction down are not inverses: we have the formula, for a sheaf ${\cal F} $ on ${\cal X} /S$, $$ Res _{S'/S}({\cal F} |_{{\cal X} /S'}) = {\cal F} \times _SS' . $$ On the other hand, suppose $p:{\cal F} \rightarrow S'$ is a morphism of sheaves on ${\cal X}/S$. Then we denote by ${\cal F} /S'$ the corresponding sheaf on ${\cal X} /S'$ (the data of the morphism is implicit in the notation). It is defined by the statement that ${\cal F} /S' ( Y\rightarrow S')$ is equal to the set of $u \in {\cal F} (Y\rightarrow S)$ such that $p(u)\in S'(Y\rightarrow S)$ is equal to the given morphism $Y\rightarrow S'$. For another point of view note that there is a tautological section of $(S'/S)|_{ {\cal X} /S'}$, and ${\cal F} /S'$ is the preimage of this section in ${\cal F} |_{{\cal X} /S'}$. As a special case we get that if ${\cal F}$ is a sheaf on ${\cal X} = {\cal X} /\ast$ with a morphism ${\cal F} \rightarrow S$ then we obtain a sheaf ${\cal F} /S$ on ${\cal X} /S$. The operations $$ ({\cal F} \rightarrow S')\mapsto {\cal F} /S' $$ from sheaves on ${\cal X} /S$ with morphisms to $S'$ to sheaves on ${\cal X} /S'$, and $$ {\cal F} ' \mapsto (Res _{S'/S}{\cal F} ' \rightarrow S' $$ from sheaves on ${\cal X} /S'$ to sheaves on ${\cal X} /S$ with morphisms to $S'$, are inverses. For this reason it is often tempting to ignore the strict notational convention and simply use the same notations for the two objects. This is not too dangerous except in the last section of the paper where we will try to be careful. If a sheaf ${\cal F}$ on ${\cal X}$ is representable by an object $F\in {\cal X}$ and if $F\rightarrow S$ is a morphism then ${\cal F} /S$ is representable by the same object $F$ together with its morphism, considered as an object of ${\cal X} /S$. For this reason we will sometimes drop the notation ${\cal F} /S$ and just denote this as ${\cal F}$ when there is no risk of confusion (and in fact the attentive reader will notice that even in the definition two paragraphs ago we have written $S'$ when we should have written $S'/S$ in the first sentence...but the second version would have been impossible because not yet defined...!) Finally there is another natural operation: suppose $\pi : S'\rightarrow S$ is a morphism and ${\cal F}$ is a sheaf on ${\cal X} /S'$. Its {\em direct image} is the sheaf $\pi _{\ast}({\cal F} )$ defined by the statement that $$ \pi _{\ast}({\cal F} )(Y\rightarrow S):= {\cal F} (Y\times _SS' \rightarrow S'). $$ This is {\em not} the same thing as the restriction down from $S'$ to $S$. Think of the case where $S$ is one point and $S'$ is a collection of several points. The value of $Res_{S'/S}({\cal F} )$ at $S$ is the {\em union} of the values of ${\cal F}$ over the points in $S'$ whereas the value of $\pi _{\ast}({\cal F} )$ at $S$ is the {\em product} of the values of ${\cal F}$ at the points in $S'$. \numero{Presentability conditions for sheaves} We will define several conditions, numbered $P1$, $P2$, $P4({\cal M} )$, $P5({\cal M} )$ (whereas two other conditions $P3$ and $P3\frac{1}{2}$ will be defined later, in \S 2). The last two depend on a choice of a class ${\cal M}$ of morphisms in ${\cal X}$ subject to certain properties set out below. In the upcoming section we then specify which class ${\cal M}$ we are interested in (at least in characteristic zero), the class of {\em vertical morphisms}. Since the preliminary results depend only on the formal properties of ${\cal M}$ we thought it might be useful to state them in general rather than just for the class of vertical morphisms, this is why we have the seeming complication of introducing ${\cal M}$ into the notations for our properties. We also introduce {\em boundedness conditions} denoted $B1$ and $B2$. These conditions sum up what is necessary in order to be able to apply Artin approximation. Fix a base scheme $S\in {\cal X}$. In what follows, we work in the category of sheaves on ${\cal X} /S$. Thus a sheaf is supposed to be on ${\cal X} /S$ unless otherwise specified. \noindent {\bf P1.}\,\, We say that ${\cal F}$ is {\em P1} if there is a surjective morphism of sheaves $X\rightarrow {\cal F}$ where $X$ is represented by a scheme of finite type over $S$. We may assume that $X$ is affine. \noindent {\bf P2.}\,\, We say that ${\cal F}$ is {\em P2} if there are surjective morphisms of sheaves $X\rightarrow {\cal F}$ and $Y\rightarrow X\times _{{\cal F}}X$ where $X$ and $Y$ are represented by schemes of finite type over $S$. We may assume that $X$ and $Y$ are affine. \begin{lemma} \mylabel{I.t} If $G$ is a sheaf of groups which is P1, and $G$ acts on a sheaf $F$ which is P2, then the quotient sheaf $F/G$ is again P2. \end{lemma} {\em Proof:} Choose surjections $\varphi :X\rightarrow F$ and $(p_1,p_2):Y\rightarrow X\times _FX$. The action is a map $G\times F\rightarrow F$, and we can choose a surjection $(q_1,q_2):W\rightarrow G\times X$ (with $W$ an affine scheme, by condition P1 for $G$), such that the action lifts to a map $m:W\rightarrow X$. There is obviously a surjection $X\rightarrow F/G$. We have a map $$ W\times _X Y\rightarrow X\times X $$ (where the maps used in the fiber product are $m:W\rightarrow X$ and $p_1:Y\rightarrow X$), defined by $$ (w,y)\mapsto (q_2(w), p_2(y)). $$ This map surjects onto the fiber product $X\times _{F/G}X$. It clearly maps into this fiber product. The map is surjective because if $(x,x')\in X\times X$ with $g\varphi (x)=\varphi (x')$ then for a point $w$ of $W$ lying above $(g,x)$ we have $\varphi (m(w))= g\varphi (x)=\varphi (x')$; in particular there is a point $y$ of $Y$ with $p_1(y)=m(w)$ and $p_2(y)=x'$, so the point $(w,y)$ maps to $(x,x')$. Our surjection $$ W\times _X Y\rightarrow X\times _{F/G}X $$ now shows that $F/G$ is P2. \hfill $\Box$\vspace{.1in} {\em Remark:} These conditions are independent of base scheme $S$ for finite-type morphisms. More precisely if $S'\rightarrow S$ is a morphism of finite type and if ${\cal F} '$ is a sheaf on ${\cal X} /S'$ then denoting by ${\cal F} = Res _{S'/S}{\cal F} '$ its restriction down to $S$ we have that ${\cal F} $ is $P1$ (resp. $P2$) if and only if ${\cal F} '$ is $P1$ (resp. $P2$). \subnumero{Boundedness conditions} We consider the following boundedness conditions for a sheaves on ${\cal X}$. These two conditions are designed to contain exactly the information needed to apply the Artin approximation theorem \cite{Artin}. \newline {\bf B1.} \,\, We say that a sheaf ${\cal F}$ is {\em B1} if, for any $k$-algebra $B$, we have that $$ \lim _{\rightarrow}{\cal F} (Spec (B')) \rightarrow {\cal F} (Spec (B)) $$ is an isomorphism, where the limit is taken over the subalgebras $B'\subset B$ which are of finite type over $k$. This is equivalent to the local finite type condition of Artin \cite{Artin}. \noindent {\bf B2.} \,\, We say that a sheaf ${\cal F}$ is {\em B2} if, for any complete local ring $A$, we have that the morphism $$ {\cal F} (Spec (A)) \rightarrow \lim _{\leftarrow} {\cal F} (Spec (A/{\bf m} ^i) $$ is an isomorphism. The {\bf Artin approximation theorem} (\cite{Artin}) can now be stated as follows. {\em Suppose ${\cal F}$ is a sheaf of sets which is B1 and B2. If $S=Spec (A)$ is an affine scheme with point $P\in S$ corresponding to a maximal ideal ${\bf m} \subset A$ then for any $$ \eta \in \lim _{\leftarrow} {\cal F} (Spec (A/{\bf m} ^i)) $$ and for $i_0\geq 0$ there exists an etale neighborhood $P\in S' \rightarrow S$ and an element $\eta ' \in {\cal F} (S')$ agreeing with $\eta$ over $Spec (A/{\bf m} ^{i_0})$. } \begin{lemma} \mylabel{I.t.1} 1.\,\, If ${\cal F}$ and ${\cal G}$ are B1 (resp. B2) and $f,g$ are two morphisms from ${\cal F}$ to ${\cal G}$ then the equalizer is again B1 (resp. B2). \newline 2.\,\, Suppose ${\cal F}\rightarrow {\cal G}$ is a surjective morphism of sheaves. If ${\cal F}$ and ${\cal F} \times _{{\cal G}}{\cal F}$ are B1 then ${\cal G}$ is B1. \end{lemma} {\em Proof:} Fix $S=Spec (A)$ and $\{ B_i\}$ our directed system of $A$-algebras. Let $B= \lim _{\rightarrow}B_i$. Suppose $\eta \in {\cal G} (B)$. There is a natural morphism $$ \lim _{\rightarrow} {\cal G} (B_i)\rightarrow {\cal G} (B). $$ First we prove injectivity. Suppose $\varphi , \psi \in {\cal G} (B_i)$ map to the same element of ${\cal G} (B)$. We may choose an etale surjection of finite type $Spec (B'_i)\rightarrow Spec (B_i)$ such that the restrictions $\varphi '$ and $\psi '$ lift to elements $u,v\in {\cal F} (B_i)$. Their images in ${\cal F} (B')$ give a point $(u,v)_{B'}$ in ${\cal F} \times _{{\cal G}} {\cal F} (B')$ (here $B':= B\otimes _{B_i}B'_i$). By the condition B1 for the fiber product, there is a $j\geq i$ such that this point comes from a point $\eta \in {\cal F} \times _{{\cal G}}{\cal F} (B'_j)$. On the other hand, note that the product ${\cal F} \times {\cal F}$ is B1. The image of $\eta$ in ${\cal F} \times {\cal F} (B')$ is the same as that of $(u,v)$; and by the B1 condition for the product, there is $k\geq j$ such that the image of $\eta$ in ${\cal F} \times {\cal F} (B'_k)$ is equal to the image of $(u,v)$. In particular, $(u|_{Spec (B'_k)},v|_{Spec (B'_k)})$ lies in the fiber product ${\cal F} \times _{{\cal G}}{\cal F} (B'_k)$. In other words, $u$ and $v$ have the same images in ${\cal G} (B'_k)$. These images are the restrictions of the original $\varphi , \psi$. Since $Spec (B'_k)\rightarrow Spec (B_k)$ is an etale surjection, the images of $\varphi$ and $\psi $ in ${\cal G} (B_k)$ are the same. This proves the injectivity. Now we prove surjectivity. Then there exists an etale surjection of finite type $$ Spec (B')\rightarrow Spec (B) $$ such that $\eta |_{Spec (B')}$ comes from an element $\rho \in {\cal F} (B')$. The functor ``etale surjections of finite type'' is itself B1, so there is an etale $Spec (B'_i)\rightarrow Spec (B_i)$ inducing $B'$. Then $B'=\lim _{\rightarrow} B'_j$ where $B'_j= B_j\otimes _{B_i}B'_i$ for $j\geq i$. By the property B1 for ${\cal F}$ there is some $j$ such that $\rho$ comes from $\rho _j\in {\cal F} ( B'_j)$. We obtain an element $\eta '_j\in {\cal G} (B'_j)$ mapping to $\eta ':=\eta |_{Spec (B')}$. The two pullbacks of $ \eta '$ to $Spec (B'\otimes _BB')$ are equal. Note that $$ B'\otimes _BB' = \lim _{\rightarrow} B'_k\otimes _{B_k}B'_k, $$ so by the above injectivity, there is some $k$ such that the two pullbacks of $\eta _j|_{Spec (B'_k)}$ to $Spec (B'_k\otimes _{B_k}B'_k)$ are equal. Now the sheaf condition for ${\cal G}$ means that $\eta _j|_{Spec (B'_k)}$ descends to an element $\eta _k\in {\cal G} (B_k)$. The restriction of $\eta _k$ to $B'$ is equal to the restriction of $\eta$, so the sheaf condition for ${\cal G}$ implies that the restriction of $\eta _k$ to $Spec (B)$ is $\eta$. \hfill $\Box$\vspace{.1in} {\em Remark:} The direct product of a finite number of B1 (resp. B2) sheaves is again B1 (resp. B2) so part 1 of the lemma implies that the properties B1 and B2 are maintained under fiber products. \begin{theorem} \mylabel{I.t.2} Suppose ${\cal F}$ is a sheaf which is P2. Then ${\cal F}$ is B1. If the ground field is uncountable, then ${\cal F}$ is B2. \end{theorem} {\em Proof:} The condition B1 follows from the previous lemma. Indeed, let $X\rightarrow {\cal F}$ and $R\rightarrow X\times _{{\cal F}}X$ be the morphisms given by the property P2, with $X$ and $R$ of finite type (in particular, B1). Note that $R\times _{X\times _{{\cal F}}X}R= R\times _{X\times X}R$ is a scheme of finite type, so the lemma implies that $X\times _{{\cal F}}X$ is B1; another application of the lemma then shows that ${\cal F}$ is B1. For B2, let $S=Spec (A)$ with $A$ a complete local ring, and let $S_n:= Spec (A/{\bf m}^{n+1})$. Let $X\rightarrow {\cal F}$ and $R\rightarrow X\times _{{\cal F}}X$ be the morphisms given by the property P2, with $X$ and $R$ of finite type over $S$. Schemes of finite type are B2. We show surjectivity of the map $$ {\cal F} (S)\rightarrow \lim _{\leftarrow} {\cal F} (S_n). $$ Suppose $(\varphi _n )$ is a compatible system of elements of ${\cal F} (S_n)$. Let $$ E_n := \{ x_n \in X(S_n):\;\;\; x_n \mapsto \varphi _n \} . $$ Note that $E_n$ is a nonempty closed subset of the scheme $X(S_n )$ (that is, the scheme whose $Spec (k)$-valued points are $X(S_n)$). Let $$ E'_n:= \bigcap _{m\geq n} {\rm im}(E_m \rightarrow E_n); $$ this is an intersection of a decreasing family of nonempty constructible subsets of $E_n$. Since $k$ is uncountable, this intersection is nonempty. Indeed, the closures of the images form a decreasing family of closed sets, which stabilizes by the noetherian property of $E_n$; then within this closed subset, the dense constructible subsets contain open sets which are complements of proper closed subsets. The union of countably many proper closed subsets is a proper subset, so the intersection of the open complements is nonempty. (Note however that $E'_n$ is not necessarily constructible). The morphism $E'_{n+1} \rightarrow E' _n$ is surjective. To see this, suppose $u\in E' _n$. We can consider the subsets $$ D_m := \{ v\in E_m, \;\; v\mapsto u\} . $$ These are closed subsets of $E_m$, nonempty by the condition $u\in E'_n$. Let $D' _{n+1}= \bigcap _{m\geq n+1} {\rm im} (D_m \rightarrow D_{n+1})$. By the same proof as above, $D'_{n+1}$ is nonempty. But it is contained in $E'_{n+1}$ and maps to $u\in E'_n$. The surjectivity of the maps implies that the inverse limit $\lim _{\leftarrow} E'_n $ is nonempty. It is a subset of $\lim _{\leftarrow} X(S_n)=X(S)$, consisting of elements mapping to $(\varphi _n)$ in $\lim _{\leftarrow}{\cal F} (S_n)$. (In fact, this subset is equal to the inverse image of $(\varphi _n)$.) We obtain an element of $X(S)$, hence an element of ${\cal F} (S)$, mapping to $(\varphi _n)$. This proves surjectivity. Note that this part of the proof only used property P1 for ${\cal F}$. We now prove injectivity. Note that $X\times _{{\cal F}}X$ is P1, so by the proof above, we obtain surjectivity of the morphism $$ X\times _{{\cal F}}X(S)\rightarrow \lim _{\leftarrow}X\times _{{\cal F}}X(S_n). $$ Suppose two elements $u,v\in G(S)$ go to the same element of $G(S_n)$ for all $n$ (we write this $u_n=v_n$). We can lift them to elements $x,y\in X(S)$, and we obtain a compatible sequence of elements $(x_n, y_n)\in X\times_{{\cal F}}X (S_n)$. By the surjectivity of the above morphism, there is an element $(x',y')\in X\times _{{\cal F}}X(S)$ with $x'_n=x_n$ and $y'_n=y_n$. The images $u'$ and $v'$ of $x'$ and $y'$ in ${\cal F} (S)$ are equal. By the B2 property for $X$, this implies that $x'=x$ and $y'=y$, which shows that $u=v$. \hfill $\Box$\vspace{.1in} We have the following Krull-type property. \begin{lemma} \mylabel{Krull} Suppose ${\cal F}$ is a sheaf which is B1 and B2. Then for any scheme $S$ of finite type the natural morphism is an injection $$ {\cal F} (S ) \hookrightarrow \prod _{{\rm Art.} \, S'\rightarrow S} {\cal F} (S' ) $$ where the product is taken over $S'\rightarrow S$ which are artinian and of finite type. \end{lemma} {\em Proof:} Suppose $f,f'\in {\cal F} (S)$ agree over all artinian subschemes. Let ${\cal G} = S \times _{{\cal F} } S$ be the fiber product where $f$ and $f'$ provide the two morphisms from $S$ to ${\cal F}$. Then ${\cal G}$ is B1 and B2 (by the remark following Lemma \ref{I.t.1}). But ${\cal G}$ has a (unique) section over any artinian $S'\rightarrow S$ and applying B2, B1 and Artin approximation we obtain sections of ${\cal G}$ over an etale covering of $S$. \hfill $\Box$\vspace{.1in} \subnumero{Choice of a class of morphisms ${\cal M}$} Fix a base scheme $S\in {\cal X}$. We assume fixed for the rest of this section a subset ${\cal M}\subset Mor ({\cal X} /S)$ of morphisms in ${\cal X} /S$, containing the identities and closed under composition (i.e. ${\cal M}$ is the set of morphism of a subcategory of ${\cal X} /S$) subject to the following axioms: \newline {\bf M1}\,\, If $a$ and $b$ are composable morphisms such that $a$ and $ba$ are in ${\cal M}$, and $a$ is surjective, then $b$ is in ${\cal M}$. \newline {\bf M2}\,\, Compatibility with base change: if ${\cal F} \rightarrow {\cal G}$ is an ${\cal M}$-morphism and ${\cal H} \rightarrow {\cal G}$ any morphism, then ${\cal F} \times _{{\cal G}}{\cal H} \rightarrow {\cal H}$ is an ${\cal M}$-morphism; and conversely if $a:{\cal F}\rightarrow {\cal G}$ is a morphism such that ${\cal F} \times _{\cal G} Y\rightarrow Y$ is in ${\cal M}$ for every $S$-scheme and morphism $Y\rightarrow {\cal G}$, then $a$ is in ${\cal M}$. \newline {\bf M3}\,\, Etale morphisms between schemes are in ${\cal M}$. {\em Remark:} It follows from these axioms that the direct product of morphisms in ${\cal M}$ is again a morphism in ${\cal M}$. In the next section we will specify a certain such subcategory ${\cal M}$, the class of {\em vertical morphisms}, and show that it satisfies these axioms. But there may be other interesting examples of such a class of morphisms ${\cal M}$ to which the following definitions and lemmas could be applied. We can now extend our list of presentability properties which refer to the class ${\cal M}$. We use the notation ${\cal M}$-morphism for morphism lying in ${\cal M}$. The gap in the numbering is to leave a place for the property $P3$ later. This property (which is absolute rather than relative to a base scheme $S$) will come up only at the end of the paper, but it turns out to be more logical to number it in between $P2$ and $P4$ (this is the numbering used in \cite{kobe}). \noindent ${\bf P4({\cal M} )}$\,\, We say that a sheaf ${\cal F}$ is $P4({\cal M} )$ if there exist surjective ${\cal M}$-morphisms $$ X\rightarrow {\cal F} $$ and $$ R\rightarrow X\times _{{\cal F}}X $$ with $X$ and $R$ represented by affine schemes of finite type over $S$. \noindent ${\bf P5({\cal M} )}$\,\, We say that ${\cal F}$ is $P5({\cal M} )$ if it is $P5({\cal M} )$ and if, in addition, the structural morphism ${\cal F} \rightarrow S$ is in ${\cal M}$. \begin{lemma} \mylabel{I.z.1} If ${\cal F}$ and ${\cal G}$ are $P4({\cal M} )$ (resp. $P5({\cal M} )$) then so is ${\cal F} \times _S{\cal G}$. \end{lemma} {\em Proof:} The presentation is just the product of the presentations for ${\cal F}$ and ${\cal G}$. \hfill $\Box$\vspace{.1in} \subnumero{Kernels and extensions} \begin{lemma} \mylabel{I.1.a} If $f,g:{\cal G} \rightarrow {\cal H}$ are two morphisms, and if ${\cal G}$ and ${\cal H}$ are $P4({\cal M} )$, then the equalizer ${\cal F}$ is $P4({\cal M} )$. \end{lemma} {\em Proof:} Let $X\rightarrow {\cal H}$, $R\rightarrow X\times _{{\cal H}}X$, $Z\rightarrow {\cal G}$ and $T\rightarrow Z\times _{{\cal G}}Z$ be ${\cal M}$-morphisms with $X$, $R$, $Z$ and $T$ schemes of finite type over $S$. Assume that we have liftings $f',g': Z\rightarrow X$ of $f$ and $g$. Set $$ W:= Z\times _{X\times _SX}R. $$ It is a scheme of finite type over $S$. Note that the composed map $Z\times _{{\cal G}} {\cal F} \rightarrow Z\rightarrow X\times _SX$ factors through $X\times _{{\cal H}}X$, and we have $$ W= (Z\times _{{\cal G}} {\cal F} )\times _{X\times _{{\cal H}}X}R. $$ From this and property $M2$, it is clear that the morphism $W\rightarrow {\cal F}$ is surjective and in ${\cal M}$. Now set $$ V:= (W\times _SW)\times _{Z\times _SZ}T. $$ Again, this is of finite type over $S$. We have $$ W\times _{{\cal F}}W= W\times _{{\cal G}}W = (W\times _SW)\times _{Z\times _SZ}(Z\times _{{\cal G}}Z). $$ Therefore $$ V= (W\times _{{\cal F}} W)\times _{Z\times _{{\cal G}}Z}T. $$ From this and property $M2$ it is clear that the morphism $V\rightarrow W\times _{{\cal F}}W$ is surjective and in ${\cal M}$. We obtain the property $P4({\cal M} )$ for ${\cal F}$. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.1.a.1} If ${\cal F}\rightarrow {\cal H}$ and ${\cal G} \rightarrow {\cal H}$ are two morphisms between $P4({\cal M} )$ sheaves, then the fiber product ${\cal F} \times _{{\cal H}}{\cal G}$ is $P4({\cal M} )$. \end{corollary} {\em Proof:} The fiber product is the equalizer of the two morphisms ${\cal F} \times _S{\cal G} \rightarrow {\cal H}$. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.1.b} Suppose ${\cal H}$ is a group sheaf which is $P5({\cal M} )$. If ${\cal H}$ acts freely on a sheaf ${\cal G}$ with quotient ${\cal F} = {\cal G} /{\cal H}$, then the morphism ${\cal G} \rightarrow {\cal F}$ is in ${\cal M}$. \end{lemma} {\em Proof:} Make a base change by a scheme $Y\rightarrow {\cal F}$. Let ${\cal G} '':= {\cal G} \times _{{\cal F}}Y$. Then ${\cal H}$ acts freely on ${\cal G} ''$ with quotient $Y$. Since the morphism ${\cal G} '' \rightarrow Y$ is surjective in the etale topology, we may find an etale morphism (of finite type and surjective) $Y'\rightarrow Y$ such that the base change ${\cal G} ^3$ of ${\cal G} ''$ to $Y'$ admits a section. Then ${\cal G} ^3= Y'\times _S{\cal H}$. In particular, the morphism ${\cal G} ^3\rightarrow Y'$ is in ${\cal M}$, hence also the morphism ${\cal G} ^3 \rightarrow Y$. Finally, the morphism ${\cal G} ^3 \rightarrow {\cal G} ''$ is surjective, since $Y'\rightarrow Y$ is surjective, and is an ${\cal M}$-morphism because it becomes an etale morphism after base change to any scheme. By property $M1$, the morphism ${\cal G} '' \rightarrow Y$ is in ${\cal M}$; then by $M2$ the morphism ${\cal G} \rightarrow {\cal F}$ is in ${\cal M}$. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.1.c} Suppose ${\cal G}$ is a $P4({\cal M} )$ sheaf, and suppose $X\rightarrow {\cal G}$ is a morphism with $X$ a scheme of finite type over $S$. Then there exists a surjective ${\cal M}$-morphism $R\rightarrow X\times _{{\cal G}}X$ with $R$ a scheme of finite type over $S$. \end{lemma} {\em Proof:} Let $Y\rightarrow {\cal G}$ and $Q\rightarrow Y\times _{{\cal G}}Y$ be the surjective ${\cal M}$-morphisms. There is an etale surjection $X'\rightarrow X$ such that the lifting $X'\rightarrow Y$ exists. Note that $$ X'\times _{{\cal G}} X' = (X' \times _S X')\times _{Y\times _SY}(Y\times _{{\cal G}}Y). $$ We get that $$ R:= (X'\times _{{\cal G}}X')\times _{Y\times _{{\cal G}}Y}Q= (X' \times _S X')\times _{Y\times _SY}Q $$ is a scheme of finite type. But also the morphism $$ R= (X'\times _{{\cal G}}X')\times _{Y\times _{{\cal G}}Y}Q\rightarrow X'\times _{{\cal G}}X' $$ is in ${\cal M}$, by property $M2$. Finally, $$ X'\times _{{\cal G}} X' = (X\times _{{\cal G}}X) \times _{X\times _SX} X'\times _SX' $$ and $X'\times _SX'\rightarrow X\times _SX$ is an ${\cal M}$-morphism by $M3$ and the remark following the properties $M$. Thus $X'\times _{{\cal G}} X'\rightarrow X\times _{{\cal G}} X$ is in ${\cal M}$ (it is also surjective), so the surjection $R\rightarrow X\times _{{\cal G}}X$ is in ${\cal M}$. \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{I.1.d} Suppose ${\cal H}$ is a group sheaf which is $P5({\cal M} )$, and suppose that ${\cal H}$ acts freely on a sheaf ${\cal G}$ with quotient ${\cal F} = {\cal G} /{\cal H} $. Then ${\cal F}$ is $P4({\cal M} )$ (resp. $P5({\cal M} )$) if and only if ${\cal G}$ is $P4({\cal M} )$ (resp. $P5({\cal M} )$). \end{theorem} {\em Proof:} By the lemma, the morphism ${\cal G} \rightarrow {\cal F}$ is in ${\cal M}$. If ${\cal G}$ is $P4({\cal M} )$ then there is a surjective ${\cal M}$-morphism $X\rightarrow {\cal G}$ with $X$ a scheme of finite type over $S$. The morphism $X\rightarrow {\cal F}$ is then surjective and in ${\cal M}$. Let $Y\rightarrow {\cal H}$ be a surjective ${\cal M}$-morphism. Now we have a surjective ${\cal M}$-morphism $$ X\times _SY\rightarrow X\times _S{\cal H} = X\times _{{\cal F}}{\cal G} , $$ and another surjective ${\cal M}$-morphism $$ X\times _{{\cal F}}{\cal G} \rightarrow {\cal F} \times _{{\cal F}}{\cal G} ={\cal G} . $$ Apply the previous lemma to the composition of these two morphisms, using the property $P4({\cal M} )$ of ${\cal G}$. We obtain the existence of a surjective ${\cal M}$-morphism $$ T\rightarrow (X\times _SY)\times _{{\cal G}} (X\times _SY) $$ with $T$ a scheme of finite type over $S$. On the other hand, note that we have a surjective ${\cal M}$-morphism $$ X\times _{{\cal F}}X\times _S{\cal H}=(X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} )\rightarrow X\times _{{\cal F}} X, $$ and a surjective ${\cal M}$-morphism $$ (X\times _SY )\times _{{\cal G}} (X\times _SY )\rightarrow (X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} ). $$ Composing these three morphisms we obtain a surjective ${\cal M}$-morphism $$ T\rightarrow X\times _{{\cal F}}X. $$ This proves that ${\cal F}$ is $P4({\cal M} )$. Suppose now that ${\cal F}$ is $P4({\cal M} )$. Let $$ X\rightarrow {\cal F} , \;\;\; R\rightarrow X\times _{{\cal F}} X $$ be the presentation given by the property $P4({\cal M} )$. We may choose $X$ in such a way that there exists a lifting $X\rightarrow {\cal G}$ (the freedom to replace $X$ by an etale cover comes from Property $M3$ and Lemma \ref{I.1.c}). This gives an isomorphism $X\times _{{\cal F}}{\cal G}\cong X\times _S{\cal H}$. Let $$ Y\rightarrow {\cal H} , \;\;\; W\rightarrow Y\times _{{\cal H}}Y $$ be the presentation given by the property $P4({\cal M} )$ of ${\cal H}$. We obtain surjective ${\cal M}$-morphisms $$ X\times _SY\rightarrow X\times _S{\cal H} $$ and (defining $U:= X\times _SW$) $$ U:=X\times _SW \rightarrow (X\times _SY)\times _{X\times _S{\cal H} }(X\times _SY). $$ Put $Z:= X\times _SY$. Then we have surjections in ${\cal M}$ $$ Z\rightarrow X\times _{{\cal F}}{\cal G} \rightarrow {\cal G} $$ (giving the first part of property $P4({\cal M} )$), and $$ U\rightarrow Z\times _{X\times _{{\cal F}}{\cal G}}Z. $$ Now, $$ (X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} )= $$ $$ X\times _{{\cal F}}(X\times _{{\cal F}}{\cal G} )=(X\times _{{\cal F}}X)\times _{{\cal F}}{\cal G} , $$ and we have an ${\cal M}$-surjection $$ R\times _{{\cal F}}{\cal G} \rightarrow (X\times _{{\cal F}}X)\times _{{\cal F}} {\cal G} . $$ Since $R\rightarrow {\cal F}$ lifts to $R\rightarrow {\cal G}$ we have $R\times _{{\cal F}}{\cal G}=R\times _S{\cal H}$ and letting $V\rightarrow R\times _SY$ be an etale surjection (needed for a certain step below), we obtain ${\cal M}$-surjections $$ V\rightarrow R\times _SY \rightarrow R\times _S{\cal H} \rightarrow (X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} ). $$ On the other hand, $$ Z\times _{{\cal G}} Z= Z\times _{X\times _{{\cal F}}{\cal G} }((X\times _{{\cal F}}{\cal G} ) \times _{{\cal G}} (X\times _{{\cal F}}{\cal G} ))\times _{X\times _{{\cal F}}{\cal G} }Z $$ so we obtain a surjection in ${\cal M}$ $$ Z\times _{X\times _{{\cal F}}{\cal G} }V\times _{X\times _{{\cal F}}{\cal G} }Z\rightarrow Z\times _{{\cal G}}Z. $$ We can assume (by choosing $V$ appropriately) that the morphism $$ V\rightarrow (X\times _{{\cal F}}{\cal G} )\times _{{\cal F}} (X\times _{{\cal F}}{\cal G} ) $$ lifts to a morphism $$ V\rightarrow Z\times_{{\cal F}} Z. $$ We then have an ${\cal M}$-surjection $$ U\times _ZV\times _ZU \rightarrow (Z\times _{X\times _{{\cal F}}{\cal G} }Z)\times _ZV\times _Z (Z\times _{X\times _{{\cal F}}{\cal G} }Z) $$ (where the two maps from $V$ to $Z$ used in the fiber product are the two projections composed with $V\rightarrow Z\times _{{\cal F}}Z$). Note that the right hand side is equal to $$ Z\times _{X\times _{{\cal F}}{\cal G} }V\times _{X\times _{{\cal F}}{\cal G} }Z, $$ which admits, as we have seen above, an ${\cal M}$-surjection to $Z\times _{{\cal G}}Z$. Since $U\times _ZV\times _ZU$ is a scheme of finite type over $S$, this completes the verification of the property $P4({\cal M} )$ for ${\cal G}$. We have now shown the equivalence of the conditions $P4({\cal M} )$ for ${\cal F}$ and ${\cal G}$. By the lemma, the morphism ${\cal G} \rightarrow {\cal F}$ is in ${\cal M}$. By Property $M1$, the structural morphism ${\cal F} \rightarrow S$ is in ${\cal M}$ if and only if the structural morphism ${\cal G} \rightarrow S$ is. Given the equivalence of the conditions $P4({\cal M} )$, this gives equivalence of the conditions $P5({\cal M} )$. \hfill $\Box$\vspace{.1in} Finally we give a lemma which allows us some flexibility in specifying resolutions. \begin{lemma} \mylabel{I.1.j} Suppose that $F$ is a sheaf on $S$ with surjective ${\cal M}$-morphisms $X\rightarrow F$ and $R \rightarrow X\times _F X$ such that $X$ and $R$ are $P4({\cal M} )$. Then $F$ is $P4({\cal M} )$. \end{lemma} {\em Proof:} Let $X'\rightarrow X$ and $Q\rightarrow X'\times _XX'$, and $R'\rightarrow R$ be the ${\cal M}$-surjections given by the hypotheses. We obtain a surjection $X'\rightarrow F$ in ${\cal M}$. On the other hand, $R'\rightarrow X\times _FX$ is in ${\cal M}$ and surjective, so $$ X'\times _XR'\times _XX' = R'\times _{X\times _FX}(X'\times _FX')\rightarrow X'\times _FX' $$ is an ${\cal M}$-surjection. But the left side is equal to $$ (X'\times _XX')\times _{X'}R' \times _{X'}(X'\times _XX') $$ if we choose (as we may assume is possible) a lifting $R'\rightarrow X'\times _FX'$ over $X\times _FX$. There is thus a surjection in ${\cal M}$ $$ Q\times _{X'}R'\times _{X'}Q\rightarrow X'\times _XR'\times _XX'. $$ Composing we get the required $$ Q\times _{X'}R'\times _{X'}Q\rightarrow X'\times _FX'. $$ \hfill $\Box$\vspace{.1in} \subnumero{Stability of the condition $P5({\cal M} )$} In the following corollary and theorem we will make use of a supplementary condition on the class ${\cal M}$: \newline {\bf M4}\,\, If $f: {\cal F} \rightarrow {\cal G}$ is a surjective morphism of sheaves of groups, then $f$ is in ${\cal M}$. \begin{corollary} \mylabel{I.z} Suppose ${\cal M}$ satisfies condition M4 in addition to the conditions M1-3. If ${\cal G}$ is a $P4({\cal M} )$ group sheaf then it is also $P5({\cal M} )$. \end{corollary} Indeed, M4 applied with ${\cal G} = \{ 1\}=S$ gives that the structural morphism ${\cal F} \rightarrow S$ for any sheaf of groups, is in ${\cal M}$. \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{I.1.e} Suppose ${\cal M}$ satisfies condition M4 in addition to the conditions M1-3. Then if $$ 1\rightarrow {\cal F} \rightarrow {\cal E} \rightarrow {\cal G} \rightarrow 1 $$ is an extension of group sheaves and if any two of the elements are $P5({\cal M} )$, the third one is too. \end{theorem} {\em Proof:} Suppose that ${\cal F}$ is $P5({\cal M} )$. Then ${\cal E}$ is $P5({\cal M} )$ if and only if ${\cal G}$ is $P5({\cal M} )$ (by applying the previous theorem in view of the fact that ${\cal F}$ acts freely on ${\cal E}$ with quotient ${\cal G}$). The remaining case is if ${\cal E}$ and ${\cal G}$ are $P5({\cal M} )$. Then by Lemma \ref{I.1.a}, the kernel ${\cal F}$ (which is an equalizer of two maps ${\cal E} \rightarrow {\cal G}$) is $P4({\cal M} )$. By the above corollary, ${\cal F}$ is $P5({\cal M} )$. \hfill $\Box$\vspace{.1in} \numero{Lifting properties and verticality} We now fill in what class of morphisms ${\cal M}$ we would like to use in the theory sketched above. We could, of course, take ${\cal M} = {\cal X}$ to be the full set of morphisms of ${\cal X}$. This might well be a reasonable choice, but I don't see how to get a good infinitesimal theory in characteristic zero out of this choice. We could also try, for example, to take ${\cal M}$ as the class of flat (or maybe smooth) morphism s. But then any non-flat group scheme over $S$ would be a counterexample to property M4, and as we have seen this property is essential to be able to specify a class of presentable groups closed under kernels, cokernels and extensions. Thus we have to work a little harder to find an appropriate class of morphisms. We say that a morphism of sheaves $a:{\cal F} \rightarrow {\cal G}$, is {\em vertical} (or {\em $S$-vertical}, if the base needs to be specified), if it satisfies the following lifting properties for all $n\geq 1$: Suppose $Y$ is a scheme with $n$ closed subschemes $Y_i\subset Y$, with retractions $r_i:Y\rightarrow Y_i$---commuting pairwise ($r_ir_j=r_jr_i$)---such that for $j\leq i$, $r_i$ retracts $Y_i$ to $Y_j\cap Y_i$. Suppose given a morphism $Y\rightarrow {\cal G}$, and liftings $\lambda _i:Y_i\rightarrow {\cal F}$ such that $\lambda _i|_{Y_i\cap Y_j}= \lambda _j|_{Y_i\cap Y_j}$. Then for any $P\in Y$ lying on at least one of the $Y_i$ there exists an etale neighborhood $P\in Y' \rightarrow Y$ and a lifting $\lambda : Y' \rightarrow {\cal F}$ which agrees with the given liftings $\lambda _i|_{Y_i\times _YY'}$ on $Y_i\times _YY'$. For future reference we call this lifting property $Lift _n(Y; Y_i)$. \begin{lemma} \mylabel{I.u.1} Suppose $f:{\cal F} \rightarrow {\cal G}$ is a morphism of sheaves which are P2. Then $f$ is vertical if and only if $Lift _n(Y; Y_i)$ holds for all systems $(Y; Y_i)$ with $Y$ (and hence $Y_i$) artinian. \end{lemma} {\em Proof:} Suppose given a system $(Y, Y_i)$ which is not artinian. Choose a point $y_0$ (in one of the $Y_i$) and try to find a lifting in an etale neighborhood of $y_0$. We can find liftings on $Y^{(n)}$ (the infinitesimal neighborhoods of $y_0$) by hypothesis. Using the P2 property of ${\cal F}$ and an argument similar to that of Theorem \ref{I.t.2}, we can choose a compatible sequence of liftings. Since ${\cal F}$ is B2 we obtain a lifting over the spectrum of the complete local ring, then by Artin approximation (using B1) we obtain a lifting on an etale neighborhood of $y_0$. \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{I.u} We have the following statements: \newline 1. \,\, If ${\cal F} \rightarrow {\cal G}$ is vertical and if ${\cal H} \rightarrow {\cal G}$ is any morphism of sheaves, then ${\cal F} \times _{{\cal G}}{\cal H} \rightarrow {\cal H}$ is vertical. \newline 2. \,\, If $a:{\cal F} \rightarrow {\cal G}$ is a morphism of sheaves such that for any $S$-scheme $Y$ and morphism $Y\rightarrow {\cal G}$, we have that ${\cal F} \times _{{\cal G}}Y\rightarrow Y$ is vertical, then $a$ is vertical; \newline 3. \,\, If $a:{\cal F} \rightarrow {\cal G}$ and $b:{\cal G} \rightarrow {\cal H}$ are two morphisms which are vertical, then $ba$ is vertical (also the identity is vertical); and \newline 4. \,\, If $a:{\cal F} \rightarrow {\cal G}$ and $b:{\cal G} \rightarrow {\cal H}$ are two morphisms such that $a$ and $ba$ are vertical, and $a$ is surjective, then $b$ is vertical. \newline 5.\,\, The etale surjections between schemes are vertical. \newline 6.\,\, Any injective morphism ${\cal F} \hookrightarrow S$ is vertical. \newline 7.\,\, If $f: {\cal F} \rightarrow {\cal G}$ is a surjective morphism of sheaves of groups, then $f$ is vertical. \end{theorem} {\em Proof:} The lifting property concerns only maps from schemes to ${\cal G}$, so it obviously satisfies parts 1 and 2. For part 3, just lift two times successively (for the identity the lifting property is tautological). For part 4, the proof is by induction on $n$. Keep the notations $a$, $b$, ${\cal F}$, ${\cal G}$ and ${\cal H}$ of part 4. Suppose $n=1$. Then we just have to note that if we have a lifting $Y_1 \rightarrow {\cal G}$ for $b$, then since $a$ is surjective, we can lift further to $Y_1\rightarrow {\cal F}$ (locally in the etale topology). The lifting for $ba$ gives $Y\rightarrow {\cal F}$ and we just project back to ${\cal G}$ to get the lifting for ${\cal G}$. This gives the case $n=1$. We may assume that the present lemma is known when there are strictly fewer than $n$ subschemes. Suppose we have liftings $\lambda _i: Y_i\rightarrow {\cal G}$; in order to get a lifting $\lambda$, and using the lifting property for the morphism $ba$, it suffices to choose liftings $\mu _i : Y_i\rightarrow {\cal F}$ with $\mu _i|_{Y_i\cap Y_j}= \mu _j|_{Y_i\cap Y_j}$. We can do this by induction. Suppose we have chosen $\mu _1,\ldots , \mu _{k-1}$. Since $k-1<n$, we know the lemma when there are $k-1$ subschemes; apply the lifting property for the morphism $a$ with respect to the morphism $Y_k\rightarrow {\cal G}$, with respect to the subschemes $Y_k\cap Y_i$, $i=1,\ldots , k-1$, and with respect to the liftings $\mu _j|_{Y_k\cap Y_j}$. We obtain a lifting $\mu _k:Y_k\rightarrow {\cal F}$ such that $\mu _k|_{Y_k\cap Y_j}=\mu _j|_{Y_k\cap Y_j}$. By induction now we obtain all of the liftings $\mu _1,\ldots , \mu _n$. The lifting property for $ba$ gives a lifting $\mu$ and we can set $\lambda := a\mu$. This completes the verification of part 4. For the etale surjections (part 5), use the previous lemma. Suppose $i:{\cal F} \rightarrow S$ is injective (part 6). To verify the lifting property for $Y\rightarrow S$ we just have to verify that this morphism factors through $Y\rightarrow {\cal F}$. For this, use the facts that $Y$ retracts onto $Y_1$ (over $S$) and that the morphism $Y_1\rightarrow S$ factors through $Y_1\rightarrow {\cal F}$. Finally we verify $Lift _n(Y, Y_i)$ for the morphism $f:{\cal F} \rightarrow {\cal G}$ in part 7. Let $r_i: Y\rightarrow Y_i$ denote the retractions. Suppose given $\mu : Y\rightarrow {\cal G}$ and $\lambda _i : Y_i \rightarrow {\cal F}$ satisfying the necessary compatibility conditions. Since $f$ is surjective, we may suppose that there is a lifting $\sigma : Y\rightarrow {\cal F}$ of $\mu$ (by restricting to an etale neighborhood in $Y$). We construct inductively $\phi _i : Y\rightarrow {\cal F}$ lifting $\mu$, with $\phi _i |_{Y_j}=\lambda _j$ for $j\leq i$. Denote the multiplication operations in ${\cal F}$ or ${\cal G}$ by $\cdot$. Let $$ h_1:= (\lambda _1\cdot (\sigma |_{Y_1})^{-1})\circ r_1: Y\rightarrow \ker (f). $$ Put $\phi _1 := h_1\cdot \sigma $. Then $\phi _1$ restricts to $\lambda _1$ on $Y_1$, and lifts $\mu$. Suppose we have chosen $\phi _i$. Let $$ h_{i+1}:= (\lambda _{i+1}\cdot (\phi _i |_{Y_{i+1}})^{-1})\circ r_{i+1} :Y\rightarrow \ker (f), $$ and put $\phi _{i+1}:= \phi _i \cdot h_{i+1}$. This lifts $\mu$ because $h_{i+1}$ is a section of $\ker (f)$. For $j\leq i$, $r_{i+1}$ maps $Y_j$ to $Y_j\cap Y_{i+1}$, and there $\lambda _{i+1}=\lambda _j$ agrees with $\phi _i$ so $h_{i+1}|_{Y_j}=1$. We don't destroy the required property for $j\leq i$. On the other hand, we gain the required property for $j=i+1$, by construction. This completes the inductive step to construct $\phi _i$. Finally, the $\phi _n$ is the lifting required for property $Lift _n(Y,Y_i)$. This completes the proof of part 7. \hfill $\Box$\vspace{.1in} From the above results, the class ${\cal M}$ of vertical morphisms satisfies the axioms M1, M2, M3, {\em and} M4 of the previous section. This is the principal class ${\cal M}$ to which we will refer, in view of which we drop ${\cal M}$ from the notation when ${\cal M}$ is the class of vertical morphisms. Thus the conditions $P4$ and $P5$ refer respectively to $P4({\cal M} )$ and $P5({\cal M} )$ with ${\cal M}$ the class of vertical morphisms. In particular we obtain the results \ref{I.z.1}, \ref{I.1.a}, \ref{I.1.a.1}, \ref{I.1.b}, \ref{I.1.c}, \ref{I.1.d}, \ref{I.z}, and \ref{I.1.e} for the properties P4 and P5. We have some further results about $P4$ and $P5$. \begin{lemma} \mylabel{I.x} Suppose that ${\cal F}$ is $P4$. In the situation of the lifting property $Lift_n(Y; Y_i)$ for the morphism $X\rightarrow {\cal F}$ given by property $P4$, suppose that $Y$ is the scheme-theoretic union of $Y_1,\ldots , Y_n$. Then the lifting is unique. \end{lemma} {\em Proof:} In effect, for morphisms $Y\rightarrow X$ with $X$ a scheme, if $Y$ is the scheme theoretic union of the $Y_i$ then the morphism is determined by its restrictions to the $Y_i$. \hfill $\Box$\vspace{.1in} \begin{proposition} \mylabel{aaa} The property of being vertical is stable under base change of $S$: suppose $p:S'\rightarrow S$ is a morphism of schemes. If $f:{\cal F} \rightarrow {\cal G}$ is vertical then $p^{\ast}(f):p^{\ast}({\cal F})\rightarrow p^{\ast}({\cal G} )$ is vertical. Furthermore if ${\cal H} \rightarrow {\cal K}$ is an $S'$-vertical morphism of sheaves on ${\cal X} /S'$ then the restriction down to $S$, $$ Res _{S'/S}({\cal H} )\rightarrow Res _{S'/S}({\cal K} ) $$ is $S$-vertical. \end{proposition} {\em Proof:} This follows from the form of the lifting properties. \hfill $\Box$\vspace{.1in} {\em Remark:} We often ignore the notation of ``restriction down'', then the first part of the proposition states that if ${\cal F} \rightarrow {\cal G} \rightarrow S$ with the first morphism being $S$-vertical, then ${\cal F} \times _SS'\rightarrow {\cal G} \times _SS'$ is $S'$-vertical. The last part of the proposition states that if ${\cal H} \rightarrow {\cal K} \rightarrow S'$ with the first morphism being $S'$-vertical, then it is also $S$-vertical. \begin{corollary} \mylabel{I.1.j.1} Suppose ${\cal F}$ is a sheaf over $S$, and suppose $S'\rightarrow S$ is a surjective etale morphism such that ${\cal F} |_{S'}$ is $P4$ over $S'$. Then ${\cal F}$ is $P4$ over $S$. \end{corollary} {\em Proof:} If $(Y,Y_i)$ is a system for the lifting property over $S$, then their pullbacks $(Y',Y_i')$ form such a system over $S'$. If a morphism of sheaves over $S'$ satisfies the lifting property, then we can lift for the system $(Y', Y'_i)$. This gives a lifting over $Y'$ for the system $(Y,Y_i)$, that is a lifting etale locally, thus satisfying the lifting property over $S$. Thus a morphism which is $S'$-vertical is also $S$-vertical. It follows that ${\cal F} |_{S'}$ is $P4({\cal M} )$ over $S$. Now $$ ({\cal F} |_{S'})\times _{{\cal F}} ({\cal F} |_{S'})= {\cal F} |_{S'}\times _{S'} (S'\times _SS'), $$ and $S'\times _SS'$ is $P4({\cal M} )$ over $S$. Thus $({\cal F} |_{S'})\times _{{\cal F}} ({\cal F} |_{S'})$ is $P4({\cal M} )$ over $S$. We can now apply Lemma \ref{I.1.j} with $X={\cal F} |_{S'}$ and $R=({\cal F} |_{S'})\times _{{\cal F}} ({\cal F} |_{S'})$. \hfill $\Box$\vspace{.1in} \subnumero{Presentable group sheaves} In view of the nice properties of $P5$ group sheaves, we make the following change of notation. A sheaf of groups ${\cal G}$ over ${\cal X} /S$ is a {\em presentable group sheaf} if it is $P5$. Note that we use this terminology only for sheaves of groups. \begin{corollary} \mylabel{uvw} A sheaf of groups which is representable by a scheme of finite type $G$ over $S$, is presentable. \end{corollary} {\em Proof:} This is because we can take $X=G$ and $R$ equal to the diagonal $G$ in the definition of property $P4$; and property $P5$ is then Corollary \ref{I.z}. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{vwx} The category of presentable group sheaves contains the category generated by representable group sheaves under the operations of extensions, kernels, and division by normal subgroups. \end{corollary} {\em Proof:} Theorem \ref{I.1.e}. \hfill $\Box$\vspace{.1in} In particular, the category of presentable group sheaves is much bigger than the category of representable group sheaves. I believe that the category of presentable group sheaves is strictly larger than the category generated generated by representable group sheaves under the operations of kernel, cokernel and extension. For example the group sheaves $Aut (V)$ for a vector sheaf $V$, which are presentable as shown below, are probably not generated from representable group sheaves by kernels, extensions and quotients (although I don't have a counterexample). In an intuitive sense, however, the two categories are about the same. The two previous corollaries would also hold for the category of $P5({\cal M} )$ group sheaves for any class ${\cal M}$ satisfying $M1$ through $M4$. We now give the main argument where we use the lifting properties and the notion of verticality, i.e. the special definition of ${\cal M}$. \begin{lemma} If ${\cal G}$ is a sheaf of groups and $X\rightarrow {\cal G}$ is a vertical surjection, with identity section $e:S\rightarrow X$ then (choosing a point $P$ on $e(S)$) there is a lifting of the multiplication to a map of etale germs $$ \mu : (X,P)\times _S(X,P) \rightarrow (X,P) $$ such that $\mu (x,e)=\mu (e,x)=x$. \end{lemma} {\em Proof:} Let $Y=X\times _SX$ and $Y_1 = X\times _Se(S)\cong X$ and $Y_2 = e(S)\times _SX \cong X$. We have retractions $Y\rightarrow Y_1$ and $Y\rightarrow Y_2$ as in the lifting property. The multiplication map ${\cal G} \times _S{\cal G} \rightarrow {\cal G}$ composes to give a map $Y=X\times _SX\rightarrow {\cal G}$. The identity gives liftings $Y_1\rightarrow X$ and $Y_2\rightarrow X$ agreeing on $Y_1\cap Y_2 = e(S)\times _S e(S)$. By the definition of verticality of the morphism $X\rightarrow {\cal G}$, there is an etale neighborhood $P\in Y'\rightarrow Y$ and a lifting to a map $Y'\rightarrow X$ agreeing with our given lifts on $Y'_1$ and $Y'_2$. This gives the desired map (note that when we have written the product of two etale germs, this means the germ of the product rather than the product of the two spectra of henselian local rings). \hfill $\Box$\vspace{.1in} We use this result in the following way. A map $\mu : X\times X\rightarrow X$ (defined on germs at a point $P$) such that $\mu (e,x)=\mu (x,e) = x$, gives rise to an exponential map $T(X)_e^{\wedge}\rightarrow X$ where $T(X)_e$ is the tangent vector scheme (see \S\S 5-8 below) to $X$ along the identity section $e$ and $T(X)_e^{\wedge}$ denotes the formal completion at the zero section. To define this exponential map note that the multiplication takes tangent vectors at $e$ to tangent vector fields on $X$ which we can then exponentiate in the classical way. The formal exponential map is an isomorphism between $T(X)_e^{\wedge}$ and the completion $X^{\wedge}$ along $e$. This is a fairly strong condition on $X$ which we will exploit below, notably to get $Lie ({\cal G} )$ and to develop a theory of connectedness. In particular this technique allows us to prove directly (in \S 6 below) that when $k$ is a field of characteristic zero, presentable group sheaves over $Spec (k)$ are just algebraic Lie groups over $k$. It is possible that in characteristic $p$ there would be an appropriate notion of verticality taking into account divided powers, which would have the same effect of enabling a good infinitesimal theory. This is why we have left the class ${\cal M}$ as an indeterminate in the first part of our discussion above. \subnumero{The conditions $P3$ and $P3\frac{1}{2}$} We now add the following two conditions, which will be used as conditions on $\pi _0$ in the last section (in contrast to the condition $P5$ which is to be used on $\pi _1$ and even $\pi _i$, $i\geq 2$). These conditions depend on a functorial choice of class ${\cal M} (Y)$ of morphisms of sheaves over $Y$ for each $Y\in {\cal X}$. We will leave to the reader the (easy) job of stating these properties in this generality, and instead we will state them directly when ${\cal M} (Y)$ is taken as the class of $Y$-vertical morphisms. Note that the properties we are about to state are {\em absolute} properties of sheaves on ${\cal X}$ rather than relative properties of sheaves over some base $S$. \noindent {\bf P3.} \,\, A sheaf ${\cal F}$ on ${\cal X}$ is $P3$ if there is a surjection $X\rightarrow {\cal F}$ from a scheme $X$ of finite type over $Spec (k)$, and if there is a surjection $\varphi : R\rightarrow X\times _{{\cal F}}X$ from a scheme $R$ of finite type over $Spec (k)$ such that $\varphi$ is an $X\times X$-vertical morphism. \noindent ${\bf P3\frac{1}{2}.}$ \,\, A sheaf ${\cal F}$ on ${\cal X}$ is $P3\frac{1}{2}$ if there is a surjection $X\rightarrow {\cal F}$ from a scheme $X$ of finite type over $Spec (k)$, and if there is a surjection $\varphi : R\rightarrow X\times _{{\cal F}}X$ from a scheme $R$ of finite type over $Spec (k)$ such that $\varphi$ is an $X$-vertical morphism, where the map to $X$ is the first projection of $X\times _{{\cal F}}X$. {\em Remark:} These properties seem almost identical. The first was refered to in \cite{kobe} (already as property $P3$). However it will turn out that the second version (which I hadn't yet thought of at the time of writing \cite{kobe}) seems more useful---cf \S 10 below. The author apologizes for this complication of the notation! {\em Remark:} $$ P5 \Rightarrow P4\Rightarrow P3\frac{1}{2}\Rightarrow P3 \Rightarrow P2 \Rightarrow P1. $$ These properties will not come into our study of group sheaves over a base $S$. Rather, they come in as conditions on $\pi _0$ of $n$-stacks on ${\cal X}$, in our brief discussion at the end of the paper. In fact we could have put off stating these properties until \S 10, but the reader had probably been wondering for some time already why we are skipping number $3$ in our list of properties. We quickly give the analogues, for $P3\frac{1}{2}$, of some of the basic facts about our other properties. We leave to the reader the task of elicudating the corresponding properties for $P3$. \begin{lemma} \mylabel{P3a} Suppose ${\cal G}$ is $P3\frac{1}{2}$ and suppose $X$ is a scheme of finite type with a morphism $X\rightarrow {\cal G}$. Then there is a surjection from a scheme of finite type $R\rightarrow X\times _{{\cal G}} X$ which is vertical with respect to the first factor $X$. \end{lemma} {\em Proof:} Let $Y\rightarrow {\cal G}$ and $W\rightarrow Y\times _{{\cal G}}Y$ be the surjections with the second one being vertical with respect to the first factor $Y$. There is an etale covering $X' \rightarrow X$ and a lifting of our morphism to $X'\rightarrow Y$. Then $$ X'\times _{{\cal G}}X'=(X'\times X') \times _{Y\times Y} (Y\times _{{\cal G}}Y) $$ so $$ R:=(X' \times X' )\times _{Y\times Y} W \rightarrow X'\times _{{\cal G}}X' $$ is surjective. It is vertical with respect to the first factor $Y$ and hence vertical with respect to the first factor $X'$. Since $X'\rightarrow X$ is etale, this morphism is also vertical with respect to $X$ (via the first factor). The surjection $$ X'\times _{{\cal G}}X' \rightarrow X\times _{{\cal G}}X $$ is the pullback of the etale morphism $X'\times X'\rightarrow X\times X$ so it is also vertical with respect to the first factor $X$. Composing we obtain $$ R\rightarrow X\times _{{\cal G}}X $$ vertical with respect to the first factor. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{P3b} Suppose ${\cal G}$ is $P3\frac{1}{2}$ and suppose ${\cal F} \subset {\cal G}$. If ${\cal F}$ is $P1$ then it is $P3\frac{1}{2}$. \end{corollary} {\em Proof:} Let $X\rightarrow {\cal F}$ be a surjection from a scheme of finite type $Y$. From the above lemma we get a surjection $R\rightarrow X\times _{{\cal G}}X$ which is vertical with respect to the first factor, but since ${\cal F}\rightarrow {\cal G}$ is injective $X\times _{{\cal G}} X=X\times _{{\cal F}}X$ and we're done. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{P3c} Suppose ${\cal G}$ is $P3\frac{1}{2}$ and ${\cal H}$ is $P2$, then the equalizer ${\cal F}$ of any two morphisms $f,g: {\cal G} \rightarrow {\cal H}$ is again $P3\frac{1}{2}$. \end{corollary} {\em Proof:} By Lemma \ref{I.1.a} with ${\cal M}$ being the class of all morphisms, we obtain that ${\cal F}$ is $P2$ and in particular $P1$. Since it is a subsheaf of ${\cal G}$, the previous corollary applies to show that ${\cal F}$ is $P3\frac{1}{2}$. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{P3d} Suppose ${\cal F} \rightarrow {\cal H}$ and ${\cal G} \rightarrow {\cal H}$ are two morphisms such that ${\cal H}$ is $P2$ and ${\cal F}$ and ${\cal G}$ are $P3\frac{1}{2}$. Then the fiber product ${\cal F} \times _{{\cal H}} {\cal G}$ is $P3\frac{1}{2}$. \end{corollary} {\em Proof:} The fiber product is the equalizer of two morphisms ${\cal F} \times {\cal G} \rightarrow {\cal H}$. Note that the product of two $P3\frac{1}{2}$ sheaves is again $P3\frac{1}{2}$---this comes from the general statement that if ${\cal A} \rightarrow {\cal B}$ is $S$-vertical and if ${\cal A} '\rightarrow {\cal B}'$ is $S'$-vertical then ${\cal A} \times {\cal A}'\rightarrow {\cal B} \times {\cal B} '$ is $S\times S'$-vertical (a direct consequence of the form of the lifting properties). \hfill $\Box$\vspace{.1in} Finally we have the analogue of one half of Theorem \ref{I.1.d}. I didn't quite see how to do the other half. \begin{proposition} \mylabel{P3e} Suppose $S$ is a scheme of finite type, and suppose ${\cal H}$ is a group sheaf over $S$ which is $P5$. Suppose that ${\cal G}\rightarrow S$ is a sheaf and that ${\cal H}$ acts freely on ${\cal G}$ over $S$, with quotient ${\cal F} = {\cal G} /{\cal H}$. If ${\cal F}$ is $P3\frac{1}{2}$ then ${\cal G}$ is $P3\frac{1}{2}$ (here ${\cal F}$ and ${\cal G}$ are being considered as the restrictions down to $Spec (k)$ of the corresponding sheaves over $S$). \end{proposition} {\em Proof:} We follow the proof of the second half of Theorem \ref{I.1.d}. Let $$ X\rightarrow {\cal F} , \;\;\; R\rightarrow X\times _{{\cal F}} X $$ be the presentation given by the property $P3\frac{1}{2}$. We may choose $X$ in such a way that there exists a lifting $X\rightarrow {\cal G}$, giving an isomorphism $X\times _{{\cal F}}{\cal G}\cong X\times _S{\cal H}$. Let $$ Y\rightarrow {\cal H} , \;\;\; W\rightarrow Y\times _{{\cal H}}Y $$ be the presentation given by the property $P4({\cal M} )$ of ${\cal H}$. We obtain surjective $S$-vertical morphisms $$ X\times _SY\rightarrow X\times _S{\cal H} $$ and (defining $U:= X\times _SW$) $$ U:=X\times _SW \rightarrow (X\times _SY)\times _{X\times _S{\cal H} }(X\times _SY). $$ Put $Z:= X\times _SY$. Then we have a surjection $$ Z\rightarrow X\times _{{\cal F}}{\cal G} \rightarrow {\cal G} $$ and an $S$-vertical surjection $$ U\rightarrow Z\times _{X\times _{{\cal F}}{\cal G}}Z. $$ Now, $$ (X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} )= $$ $$ X\times _{{\cal F}}(X\times _{{\cal F}}{\cal G} )=(X\times _{{\cal F}}X)\times _{{\cal F}}{\cal G} , $$ and we have a surjection vertical with respect to the first factor $X$, $$ R\times _{{\cal F}}{\cal G} \rightarrow (X\times _{{\cal F}}X)\times _{{\cal F}} {\cal G} . $$ Since $R\rightarrow {\cal F}$ lifts to $R\rightarrow {\cal G}$ we have $R\times _{{\cal F}}{\cal G}=R\times _S{\cal H}$ and letting $V\rightarrow R\times _SY$ be an etale surjection, we obtain surjections $$ V\rightarrow R\times _SY \rightarrow R\times _S{\cal H} \rightarrow (X\times _{{\cal F}}{\cal G} )\times _{{\cal G}} (X\times _{{\cal F}}{\cal G} ). $$ The first is etale, the second is $S$-vertical, and the third is $X$-vertical for the first factor, so the composition is $X$-vertical. As before $$ Z\times _{{\cal G}} Z= Z\times _{X\times _{{\cal F}}{\cal G} }((X\times _{{\cal F}}{\cal G} ) \times _{{\cal G}} (X\times _{{\cal F}}{\cal G} ))\times _{X\times _{{\cal F}}{\cal G} }Z $$ so we obtain an $X$-vertical surjection $$ Z\times _{X\times _{{\cal F}}{\cal G} }V\times _{X\times _{{\cal F}}{\cal G} }Z\rightarrow Z\times _{{\cal G}}Z. $$ We can assume by choosing $V$ appropriately that the morphism $$ V\rightarrow (X\times _{{\cal F}}{\cal G} )\times _{{\cal F}} (X\times _{{\cal F}}{\cal G} ) $$ lifts to a morphism $$ V\rightarrow Z\times_{{\cal F}} Z. $$ We then have an $S$-vertical surjection $$ U\times _ZV\times _ZU \rightarrow (Z\times _{X\times _{{\cal F}}{\cal G} }Z)\times _ZV\times _Z (Z\times _{X\times _{{\cal F}}{\cal G} }Z). $$ The right hand side is equal to $$ Z\times _{X\times _{{\cal F}}{\cal G} }V\times _{X\times _{{\cal F}}{\cal G} }Z, $$ which admits, as we have seen above, an $X$-vertical surjection to $Z\times _{{\cal G}}Z$. By composing we obtain an $X$-vertical, and hence $Z$-vertical surjection $$ U\times _ZV\times _ZU\rightarrow Z\times _{{\cal G}}Z. $$ This completes the proof. \hfill $\Box$\vspace{.1in} \numero{Functoriality} Suppose $F$ is a sheaf over $S$, and suppose $\pi : S'\rightarrow S$ is a morphism. We denote by $\pi ^{\ast}(F)$ the restriction $F|_{{\cal X} /S'}$, which is the sheaf associated to the presheaf $Y\rightarrow S' \mapsto F(Y\rightarrow S)$. If $F$ is representable then $\pi ^{\ast}F$ is also representable by the fiber product $F\times _SS'$. In general, we allow ourselves to use the notations $\pi ^{\ast}F$, $F\times _SS'$ and $F|_{S'}$ interchangeably. We have defined, for a sheaf $G$ on $S'$, the {\em restriction down $Res _{S'/S}(G )$}. Suppose $G$ is a sheaf on $S'$. We defined the direct image by $$ \pi _{\ast}G(Y\rightarrow S):= G(Y\times _SS' \rightarrow S'). $$ The morphism $F(Y\rightarrow S) \rightarrow F(Y\times _SS' \rightarrow S)$ gives a natural morphism $$ F\rightarrow \pi _{\ast} \pi ^{\ast}(F), $$ and the morphism $G(Y\times _S S'\rightarrow S) \rightarrow G (Y\rightarrow S')$ (coming from the graph morphism $Y\rightarrow Y\times _S S'$) gives a natural morphism $$ \pi ^{\ast}\pi _{\ast} (G)\rightarrow G. $$ These functors are adjoints and the above are the adjunction morphisms. More precisely, we have a natural isomorphism $$ Hom (F, \pi _{\ast}G)\cong Hom (\pi ^{\ast}F,G). $$ This may be verified directly. {\em Remark:} If $f:A\rightarrow B$ is a vertical morphism over $S'$ then $\pi _{\ast}(f):\pi _{\ast}A\rightarrow \pi _{\ast}B$ is vertical over $S$. To see this, note that if $Y, Y^{(n)}$ is a collection of $S$-schemes with retractions etc. as in the definition of verticality, then $Y\times _S S', Y^{(n)} \times _SS'$ is a collection with retractions over $S'$. The verticality of $\pi _{\ast}(f)$ for the case of $Y, Y^{(n)}$ follows from the verticality of $f$ for the case of $Y\times _S S', Y^{(n)} \times _SS'$. {\em Remark:} Direct and inverse images are compatible with fiber products. For inverse images this is easy. For direct images, suppose we have morphisms $A\rightarrow C$ and $B\rightarrow C$ on $S'$. We obtain morphisms $A\times _CB\rightarrow A$ and $A\times _CB$ satisfying a universal property. These give morphisms $$ \pi _{\ast}(A\times _CB)\rightarrow \pi _{\ast} A \;\; (resp. \;\; \pi _{\ast}B \, ). $$ We show the universal property: suppose $$ (u,v)\in (\pi _{\ast}A\times _{\pi _{\ast}C}\pi _{\ast}B )(Y), $$ that is $u\in A(Y\times _SS') $ and $v\in B(Y\times _SS')$ with the same image in $C(Y\times _SS')$. We obtain a unique element of $(A\times _CB)(Y\times _SS')$ mapping to $(u,v)$. This gives the claim. \begin{lemma} \mylabel{I.1.g.2} if ${\cal F}$ is a coherent sheaf on $S'$ and $\pi :S'\rightarrow S$ is a finite morphism then $\pi _{\ast}({\cal F} )$ is a coherent sheaf on $S$. \end{lemma} {\em Proof:} We may assume $S$ and $S'$ affine, so that $S=Spec (A)$ and $S'=Spec (A')$ with $A'$ a finite $A$-algebra. The coherent sheaf ${\cal F}$ corresponds to an $A'$-module $M$. This implies that $$ \pi _{\ast}({\cal F} ) (Spec (B)\rightarrow Spec (A)) = {\cal F} ( Spec (B\otimes _AA')) $$ $$ = M\otimes _{A'}(B\otimes _AA') = M\otimes _AB. $$ This formula means that $\pi _{\ast}({\cal F} )$ corresponds to the same module $M$ considered as an $A$-module; in particular it is coherent. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.1.h} If $F$ is $P4$ (resp. $P5$) on $S$ then $\pi ^{\ast}F$ is $P4$ (resp. $P5$) on $S'$. \end{lemma} {\em Proof:} Note first of all that $S$-verticality of a morphism of sheaves over $S'$ implies $S'$-verticality. Now if $F$ is $P4$, let $X\rightarrow F$ and $R\rightarrow X\times _FX$ be the corresponding vertical surjections. We get $\pi ^{\ast}(X) \rightarrow \pi ^{\ast}(F)$ and $$ \pi ^{\ast}(R)\rightarrow \pi ^{\ast}(X\times _FX)= \pi ^{\ast}(X)\times _{\pi ^{\ast}(F)}\pi ^{\ast}(X), $$ surjective and $S$-vertical (hence $S'$-vertical) morphisms. Note that $\pi ^{\ast}(X)$ and $\pi ^{\ast}(R)$ are schemes of finite type over $S'$, so we obtain the proof for P3. For $P5$ note that $\pi ^{\ast}(F)=F\times _SS'$, so by Theorem \ref{I.u}, $\pi ^{\ast}(F)\rightarrow S'$ is $S$-vertical; hence it is $S'$-vertical as required. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.1.i} Suppose $\pi :S'\rightarrow S$ is a finite morphism and suppose $G'$ is a $P4$ sheaf on $S'$. Then $\pi _{\ast}(G')$ is a $P4$ sheaf on $S'$. \end{lemma} {\em Proof:} Let $X'\rightarrow G'$ and $R'\rightarrow X'\times_{G'}X'$ be the surjective vertical morphisms with $X'$ and $R'$ schemes of finite type over $S'$. Let $G:= \pi _{\ast}(G')$ and similarly for $X$ and $R$. By the above remark, we obtain vertical morphisms $X\rightarrow G$ and $R\rightarrow X\times _GX$. (Note that $X\times _GX= \pi _{\ast}(X'\times _{G'}X'$ by above.) In the case of a finite morphism $\pi : S'\rightarrow S$, note that if $f:A\rightarrow B$ is surjective over $S'$ then $\pi _{\ast}(f):\pi _{\ast}A\rightarrow \pi _{\ast}B$ is surjective. This is a general property of sheaves on the etale topology, for which we sketch the proof (an application of Artin approximation). If $\eta \in \pi _{\ast}(B)(Y)$, this means $\eta : Y\times _SS'\rightarrow B$. For $y'\in Y\times _SS'$ there is an etale neighborhood $U\rightarrow Y\times _SS'$ and a lifting $U\rightarrow A$. We need to find an etale neighborhood $V$ of the image $y\in Y$ and a lifting $V\times _SS' \rightarrow U$. Define a functor $L(V/Y)$ to be the set of liftings $V\times _SS' \rightarrow U$ over $Y\times _SS'$. It is B1, and a lifting exists on $\hat{V}= Spec ({\cal O} _{Y,y}^{\wedge})$, so by Artin approximation there is an etale neighborhood $V$ with a lifting. Applying this to our case, the morphisms $X\rightarrow G$ and $R\rightarrow X\times _GX$ are surjective. By Lemma \ref{I.1.j}, it suffices to prove that $X$ and $R$ are $P4$. Thus it suffices in general to show: if $Z$ is a scheme of finite type over $S'$ then $\pi _{\ast}(Z)$ is $P4$. We make a further reduction: a scheme of finite type can be presented as the kernel of a morphism ${\bf A}^n \rightarrow {\bf A}^m$; the direct image is then the kernel of $\pi _{\ast}{\bf A}^n \rightarrow \pi _{\ast}{\bf A}^m$. The kernel of a morphism of $P4$ sheaves is again $P4$ (Lemma \ref{I.1.a}) so it suffices to treat the case $Z={\bf A}^n$. But in this case, $Z$ is a coherent sheaf and its direct image is also a coherent sheaf. One can see directly that a coherent sheaf ${\cal F}$ is $P4$ by using the fact that it has a resolution of the form $$ {\cal O} ^a \rightarrow {\cal O} ^b \rightarrow {\cal F} \rightarrow 0 $$ (exact even on the big site ${\cal X} /S$), or by looking at Lemmas \ref{I.1.g.1} (below) and \ref{I.1.g.2}. This completes the proof. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{restrictionPreserves?} Suppose $S'\rightarrow S$ is a morphism of schemes of finite type. Suppose ${\cal F}$ is a sheaf on ${\cal X} /S'$. Then $Res _{S'/S}{\cal F}$ is $P3\frac{1}{2}$ if and only if ${\cal F}$ is $P3\frac{1}{2}$. \end{lemma} {\em Proof:} This is only a matter of terminology since, $P3\frac{1}{2}$ being a global property, the statement that ${\cal F}$ is $P3\frac{1}{2}$ really means that the restriction of ${\cal F}$ down to $Spec (k)$ is $P3\frac{1}{2}$. It is obviously equivalent to say this after first restricting down to $S$. \hfill $\Box$\vspace{.1in} \numero{Vector sheaves} With this section we begin the part of our study which requires working over a ground field $k$ of characteristic zero. From now on ${\cal X}$ denotes the big etale site of schemes over $Spec (k)$. Before returning to the definition of presentability and its infinitesimal study, we make a detour to discuss {\em vector sheaves}. These are objects which will be the linearizations of presentable group sheaves---we are also interested in vector sheaves as candidates for the $\pi _i (T,t)$ with $t\in T(S)$, for an $n$-stack $T$ on ${\cal X} /S$ for $i\geq 2$. To be slightly more precise, suppose $S\in {\cal X}$ is a scheme over $k$, and let ${\cal X} /S$ denote the category of schemes over $S$. We will define a notion of {\em vector sheaf} on ${\cal X} /S$. This notion is what was called ``U-coherent sheaf'' by Hirschowitz in \cite{Hirschowitz}. The particular case which we call ``vector scheme'' below has already been well known for some time as the ``linear spaces'' of Grauert \cite{Grauert}, appearing notably in Whitney's tangent cones \cite{Whitney}. We feel that the terminology ``vector sheaf'' is more suggestive. Many of the results below seem to be due to Hirschowitz \cite{Hirschowitz} (in particular, the observation that duality is involutive) although some parts of the theory are certainly due to \cite{Fischer}, \cite{Grauert}, \cite{Whitney}. We have integrated these results into our treatment for the reader's convenience. Essentially the only thing new in our treatment is the first lemma (and the analogous statement about extensions). Before starting in on the definition, I would like to make one note of caution. The category of vector sheaves will not satisfy any nice (ascending or descending) chain condition. This is one of the principal differences with vector spaces or modules over a noetherian ring, and could in the long run pose a major problem if one wants to consider an ``infinite dimensional'' version of the theory such as by looking at $ind$- or $pro$- objects. We have a sheaf of rings ${\cal O}$ on ${\cal X}$, defined by ${\cal O} (X):= \Gamma (X, {\cal O} _X)$. Note that it is represented by the affine line. \begin{lemma} \mylabel{I.a} Suppose $F$ is a sheaf of abelian groups on ${\cal X} /S$, representable by a scheme which is affine and of finite type over $S$. If there exists a structure of ${\cal O}$-module for $F$, then this structure is unique. If $F$ and $G$ are two such sheaves, and if $a:F\rightarrow G$ is a morphism of sheaves of groups, then $a$ is a morphism of ${\cal O}$-modules. \end{lemma} {\em Proof:} The first statement of the lemma follows from the second. For the second statement, suppose $u\in F_X$. Consider the element $tu\in F_{X\times {\bf A}^1}$. For any positive integer $n$ we have $tu|_{X\times \{ n\}}=u+\ldots +u$ ($n$ times). The same is true for the image $a(u)$. Therefore $$ a(tu)|_{X\times \{ n\}} =ta(u)|_{X\times \{ n\}} . $$ We obtain two morphisms $X\times {\bf A}^1\rightarrow G$ which are equal on the subschemes $X\times \{ n\}$; this implies that they are equal. (Here is a proof of this: we may suppose that $X$ and the base $S$ are affine, so $X=Spec (A)$ and $G=Spec (B)$ and a morphism $X\times {\bf A}^1\rightarrow G$ corresponds to a morphism $\phi : B\rightarrow A[t]$. Pick any $b\in B$ and write $$ \phi (b)= \sum _{j=1}^m p_{j}t^j; $$ but the matrix $a_{nj}= n^j$ for $n,j=1,\ldots , m$ is invertible as a matrix with coefficients in $k$, so there is a matrix $c_{nj}$ with $$ p_j= \sum _{n=1}^m c_{nj}\phi (b)(n) . $$ Thus $\phi (b)$ is determined by the values at positive integers $\phi (b)(n)$.) \hfill $\Box$\vspace{.1in} A {\em vector scheme over $S$} is a sheaf $V$ of abelian groups on ${\cal X} /S$ which is a sheaf of ${\cal O}$-modules and such that there exists an etale covering $\{ S_{\alpha}\rightarrow S\}$ such that each $V|_{S_{\alpha}}$ is representable by a scheme $F_{\alpha}$ which is affine of finite type over $S_{\alpha}$. The above lemma shows that the category of vector schemes is a full subcategory of the category of sheaves of abelian groups on ${\cal X}$. In the complex analytic category these objects were called ``linear spaces'' by Grauert and were studied in \cite{Grauert}, \cite{Fischer}. The first remark is that, in fact, the locality in the definition of vector scheme was extraneous. In effect, since the representing schemes $F_{\alpha}$ are unique up to unique isomorphism, they glue together to give a scheme $F$, affine and locally of finite type over $S$. \begin{lemma} \mylabel{I.b} Suppose $V$ is a vector scheme on ${\cal X} /S$, and suppose $S$ is affine. Then there is an exact sequence $$ 0\rightarrow V\rightarrow {\cal O} ^m \rightarrow {\cal O} ^n $$ of abelian sheaves on ${\cal X}$. \end{lemma} {\em Proof:} Write $S=Spec (A)$ and $V=Spec (B)$. The action of ${\bf G} _m$ gives a decomposition $$ B = \bigoplus B^{\lambda} $$ where $B^{\lambda}$ consists of functions $b$ such that $b(tv)= t^{\lambda}b(v)$. The sum is over $\lambda \geq 0$ (integers), since the action extends to an action of the multiplicative monoid ${\bf A}^1$. Furthermore, if $b\in B^0$ then $b(tv)=b(v)$ for all $t$ (including $t=0$), in particular $b(v)=b(0)$. Thus $B^0= A$. If $b\in B^{\lambda }$ for $\lambda >0$ then $b(0)=b(0\cdot O)= 0$. Thus the zero section corresponds to the projection onto $B^0=A$. The decomposition is compatible with multiplication in $B$. It is also compatible with the comultiplication $B\rightarrow B\otimes _A B$ corresponding to the addition law on $V$. The comultiplication is $$ B^{\lambda } \rightarrow \bigoplus _{\mu + \nu = \lambda} B^{\mu} \otimes _A B^{\nu}, $$ and furthermore the coefficients $B^{\lambda} \rightarrow B^{\lambda}\otimes _A B^0= B^{\lambda}$ and $B^{\lambda} \rightarrow B^0\otimes _A B^{\lambda}= B^{\lambda}$ are the identity (corresponding to the formula $v+0=v=0+v$). On the other hand, the composition $B\rightarrow B\otimes _A B \rightarrow B$ corresponds to the map $v\mapsto v+v=2v$, which is also scalar multiplication by $t=2$. Thus the composition $$ B^{\lambda } \rightarrow \bigoplus _{\mu + \nu = \lambda} B^{\mu} \otimes _A B^{\nu}\rightarrow B^{\lambda} $$ is equal to multiplication by $2^{\lambda}$. The first and last terms in the sum give a contribution of $b\mapsto 2b$ (by the observation $v+0=v=0+v$), so for $\lambda \geq 2$, the composition $$ B^{\lambda } \rightarrow \bigoplus _{\mu + \nu = \lambda , 0< \mu , \nu < \lambda } B^{\mu} \otimes _A B^{\nu}\rightarrow B^{\lambda} $$ is multiplication by $2^{\lambda}-2$, invertible. Hence every element of $B^{\lambda}$ is expressed as a sum of products of elements of $B^{\mu}$ and $B^{\nu}$ for $\mu , \nu < \lambda$. This proves that $B^1$ generates $B$ as an $A$-algebra. Since $B$ is of finite type over $A$ (a consequence of the fact that we have supposed all of our schemes noetherian), we can choose a finite number of elements of $x_1,\ldots , x_m \in B^1$ which generate $B$ as an $A$-algebra, and these elements give an embedding $V\subset {\cal O} ^m$. This embedding is linear, since the elements are elements of $B^1$ (from the above discussion one sees that for $b\in B^1$ we have $b(u+v)=b(u)+ b(v)$). Write $$ B= A[x_1,\ldots , x_m]/I $$ for a homogeneous ideal $I=\bigoplus I^{\lambda}$. We claim that $I$ is generated as an ideal by $I^1$. To see this, let $I'$ be the ideal generated by $I^1$ and put $B'= A[x]/I'$. Under the comultiplication of $A[x]$ we have $$ I^1 \rightarrow I^1 \otimes _A A \oplus A \otimes _A I^1 , $$ so $I^1$ maps to zero in $B'\otimes _A B'$. Thus so does $I'$. We obtain a comultiplication $$ B'= A[x]/I'\rightarrow B'\otimes _A B', $$ so $Spec (B')$ is a vector scheme too. But the map $B'\rightarrow B$ is surjective and an isomorphism on the pieces of degree $1$. It is compatible with the comultiplication. We claim that it is injective, showing this on the part of degree $\lambda$ by induction on $\lambda$ (starting at $\lambda =2$). If an element $b\in (B')^{\lambda}$ maps to zero in $B$, then by applying the process given above (in the algebra $B'$) we can write $b= \sum b_{\mu} b_{\nu}$ for $\mu , \nu < \lambda$. But $b_{\mu}$ and $b_{\nu}$ map to the elements in $B$ given by applying the same process to the image of $b$; as this image is $0$, so are the images of $b_{\mu}$ and $b_{\nu}$. By the induction hypothesis, the map is injective on the pieces of degrees $\mu , \nu$, so $b_{\mu}=b_{\nu}=0$, giving $b=0$. This induction shows that $B'\cong B$, so $I'=I$ is generated by $I^1$. Since $B$ is of finite type over $A$ (which is noetherian), $I$ is generated by a finite number of elements. This implies that it is generated by a finite number of elements $y_1, \ldots , y_n$ of $I^1$. These elements give a linear map ${\cal O} ^n \rightarrow {\cal O} ^m$, and $V$ is the kernel. \hfill $\Box$\vspace{.1in} We come now to the main definition of this section. A {\em vector sheaf on $S$} is a sheaf of abelian groups $F$ on ${\cal X} /S$ such thatthere exists an etale covering $\{ S_{\alpha}\rightarrow S\}$ such that for each $\alpha$ there exists an exact sequence $$ U_{\alpha}\rightarrow V_{\alpha}\rightarrow F|_{S_{\alpha}}\rightarrow 0 $$ of sheaves of abelian groups, with $U_{\alpha}$ and $V_{\alpha}$ vector schemes over $S_{\alpha}$. Denote by ${\cal V} (S)$ the category of vector sheaves over $S$. If $X\rightarrow S$ is an element of ${\cal X} /S$, we denote by $F|_X$ the restriction of $F$ to the category ${\cal X} /X$. It is a vector sheaf over $X$ (this is easy to see from the definitions). If $F$ is a vector sheaf and $f\in F(Y)$ and $a:X\rightarrow Y$ is a morphism, we denote the restriction of $f$ to $X$ by $a^{\ast}(f)$ or just $f|_X$. \begin{lemma} \mylabel{I.c} If $F$ is a vector sheaf, and $S$ is an affine variety, then the cohomology groups $H^i (S, F)$ vanish for $i>0$. If $$ F_1\rightarrow F_2\rightarrow F_3 $$ is an exact sequence of vector sheaves (that is, an exact sequence in the category of abelian sheaves on ${\cal X}$, where the elements are vector sheaves) then for any $X$ over $S$ which is itself an affine scheme, the sequence $$ F_1(X)\rightarrow F_2(X)\rightarrow F_3(X) $$ is exact. \end{lemma} {\em Proof:} Treat first the case where $F$ is a vector scheme. We have an exact sequence $$ 0\rightarrow F\rightarrow {\cal O} ^a \rightarrow {\cal O} ^b $$ by Lemma \ref{I.b}. Let $G$ be the kernel of the morphism ${\cal O} ^a \rightarrow {\cal O} ^b$ on the small etale site over $S$. It is a coherent sheaf. Let $F'$ be the sheaf on ${\cal X}$ whose value on $Y\rightarrow S$ is the space of sections of the pullback (of coherent sheaves) of $G$ to $Y$. There is a surjective morphism $F'\rightarrow F$, which induces $F'(U)\stackrel{\cong}{\rightarrow} F(U)$ for any $U$ etale over $S$ (or even any $U$ which is flat over $S$). Let $K$ denote the kernel of $F'\rightarrow F$. We claim that if $Y$ is any scheme etale over $S$, then $H^i(Y, K)=0$. Prove this by ascending induction on $i$. If the cohomology in degrees $<i$ of all fiber products of elements in all etale covering families of $Y$ vanishes, then the degree $i$ sheaf cohomology is equal to the degree $i$ \v{C}ech cohomology. But the \v{C}ech cohomology is calculated only in terms of the values of the sheaf on the fiber products, and here the values of $K$ are zero. Thus $H^i(Y,K)=\check{H}^i(Y, K)=0$, completing the induction. We obtain $H^i(S, F)= H^i(S, F')$. But the higher cohomology of a coherent sheaf on an affine scheme $S$ vanishes (even in the big etale site). We obtain the desired vanishing. For the second part, suppose that $X=S$ is affine. The restriction of the exact sequence to the small etale site (over $X$) remains exact. It can be completed to a $5$-term exact sequence where the first and last terms are also coherent sheaves; then broken down into short exact sequences. The vanishing of $H^1$ of coherent sheaves on the small etale site yields the desired exactness of all the short exact sequences of global sections, and hence the exactness of the sequence in question. \hfill $\Box$\vspace{.1in} {\em Remark:} One can show that a vector sheaf $V$ over an affine $S$ has a resolution by vector schemes, over $S$ rather than over an etale covering of $S$ \cite{Hirschowitz}. \begin{lemma} \mylabel{I.d} Suppose $F$ is a vector sheaf over $S$. Then for any $X\in {\cal X} /S$ and $Y$ a scheme of finite type over $k$, we have $$ F(X\times _{Spec (k)}Y )= F(X)\otimes _k {\cal O} (Y). $$ The isomorphism is given by the pullback $F(X)\rightarrow F(X\times _kY)$ and the scalar multiplication by the pullback of functions on $Y$. \end{lemma} {\em Proof:} We first prove this when $F$ is a vector scheme. There is an exact sequence $$ 0\rightarrow F\rightarrow {\cal O} ^a \stackrel{M}{\rightarrow} {\cal O} ^b . $$ We have $F(X)= \ker (M(X))$ and $F(X\times _kY)= \ker (M(X\times _kY))$. But ${\cal O} (X\times _kY)={\cal O} (X)\otimes _k{\cal O} (Y)$, and $M(X\times _kY)=M(X)\otimes 1$. Since tensoring over $k$ is exact, $$ \ker (M(X)\otimes 1) = \ker (M(X)) \otimes _k {\cal O} (Y) $$ as desired. Now suppose $F$ is a vector sheaf. There is an exact sequence $$ U\rightarrow V\rightarrow F\rightarrow 0. $$ If $Z$ is affine then the sequence $$ U(Z)\rightarrow V(Z)\rightarrow F(Z)\rightarrow 0 $$ remains exact. To see this, replace $F$ by a coherent sheaf $F'$ on the small etale site over $Z$. The restriction of $F$ to the small etale site over $Z$ is the quotient $F'$ of the restriction of $U\rightarrow V$ to the small etale site over $Z$, that is to say the sections of $F$ and $F'$ are the same on schemes etale over $Z$ (and in particular over $Y$). But if $Z$ is affine, then taking global sections preserves surjectivity of a morphism of coherent sheaves. This gives the desired exact sequence (proceed in a similar way for exactness at $V(Z)$). Suppose now that $X$ and $Y$ are affine. Then applying the above to $Z=X$ and $Z=X\times _k Y$ we get $$ U(X)\otimes _k {\cal O} (Y) \rightarrow V(X)\otimes _k{\cal O} (Y) \rightarrow F(X\times _kY)\rightarrow 0 . $$ The first morphism is the same as in the tensor product of $$ U(X)\rightarrow V(X)\rightarrow F(X)\rightarrow 0 $$ with ${\cal O} (Y)$, so the two quotients are isomorphic: $F(X\times _kY)\cong F(X)\otimes _k {\cal O} (Y)$. This completes the case where $X$ and $Y$ are affine. But both sides of the equation have the property that they are sheaves in each variable $X$ and $Y$ separately; thus we may first localize on $X$ and then localize on $Y$, to reduce to the case where $X$ and $Y$ are affine. Finally, suppose $F$ is a vector sheaf, and write $F= \bigcup _{i\in I} F_i$ as a directed union of vector sheaves. The tensor product of the union is equal to the union of the tensor products: $$ F(X)\otimes _k{\cal O} (Y) = \bigcup _{i\in I} F_i (X)\otimes _k{\cal O} (Y) = \bigcup _{i\in I} F(X\otimes _kY) = F(X\otimes _kY). $$ Note that the inclusion maps in the two directed unions are the same (since the isomorphisms established above are uniquely determined by compatibility with the morphisms $F_i (X)\rightarrow F_i (X\otimes _kY)$ and with scalar multiplication by elements of ${\cal O} (Y)$). This completes the proof. \hfill $\Box$\vspace{.1in} {\em Remark:} We will mostly use this lemma in the following two cases. Suppose $F$ is a vector sheaf over $S$. Then for any $X\in {\cal X} /S$ we have $F(X\times {\bf A}^1)= F(X)\otimes _k k[t]$. The isomorphism is given by the pullback $F(X)\rightarrow F(X\times {\bf A}^1)$ and the scalar multiplication by the pullback of the coordinate function $t$ on ${\bf A}^1$. Similarly, $F(X\times {\bf G} _m) = F(X)\otimes _k k[t,t^{-1}]$, with the isomorphism uniquely determined by compatibility with the previous one under the inclusion ${\bf G} _m \subset {\bf A}^1$. \begin{lemma} \mylabel{I.e} A vector sheaf has a unique structure of ${\cal O}$-module, and any morphism of vector sheaves is automatically compatible with the ${\cal O}$-module structure. \end{lemma} {\em Proof:} Suppose that $\phi :F\rightarrow G$ is a morphism of vector sheaves. Suppose $X\in {\cal X} /S$. Suppose $f\in F(X\otimes {\bf A}^1)$. The difference $g=\phi (tf)-t\phi (f )$ is an element of $G(X\times {\bf A}^1)$ which restricts to zero on $X\otimes \{ n\}$ for any integer $n$. We can write $$ g=\sum _{i=1}^pg_i t^i $$ with $g^i\in G(X)$ (by the previous lemma). We know that $$ g(n)=\sum _{i=1}^pg_i n^i = 0 $$ for any integer $n$. But in $k$ the matrix $(n^i)_{1\leq n, i\leq p}$ has an inverse $(c_{ni})$, and we have $$ g_i = \sum _{n=1}^p c_{ni}g(n) = 0. $$ Therefore $\phi (tf)-t\phi (f )=g=0$, for any $f$. Thus $\phi $ is compatible with multiplication by $t$. Now suppose $\lambda \in {\cal O} (X)$. This gives a morphism $\gamma : X\rightarrow X\times {\bf A} ^1$ such that $\gamma ^{\ast} (tp_1^{\ast}(f))= \lambda f$ for any $f\in F(X)$ or $G(X)$ (here $p_1:X\times {\bf A}^1\rightarrow X$ is the projection). The fact that $\phi$ is a morphism of sheaves means that it is compatible with $\gamma ^{\ast}$ and $p_1^{\ast}$, so we have $$ \phi (\lambda f)= \phi (\gamma ^{\ast} (tp_1^{\ast}(f)))= \gamma ^{\ast} (\phi (tp_1^{\ast}(f))) $$ $$ = \gamma ^{\ast}(t\phi (p_1^{\ast} (f)))= \gamma ^{\ast}(tp_1^{\ast}(\phi (f)))= \lambda \phi (f). $$ Thus $\phi$ is compatible with scalar multiplication. This fact, applied to the identity of $F$, implies that the scalar multiplication is unique if it exists. For existence, note that any morphism of vector schemes is automatically a morphism of ${\cal O}$-modules, so the quotient has a structure of ${\cal O}$-module. Thus any vector sheaf has a structure of ${\cal O}$-module. If $F$ is a vector sheaf expressed as a directed union $F= \bigcup _{i\in I} F_i$ of finite vector sheaves, then the inclusions in the directed union are compatible with the ${\cal O}$-module structures; thus the union has an ${\cal O}$-module structure. \hfill $\Box$\vspace{.1in} The conclusion of this lemma is that the category of vector sheaves, with morphisms equal to those morphisms of abelian sheaves compatible with the ${\cal O}$-module structure, is a full subcategory of the category of sheaves of abelian groups on ${\cal X} /S$. Next we establish a Krull-type property. \begin{lemma} \mylabel{I.f} Suppose that $F$ is a vector sheaf over $S$, with $f\in F(Y)$, and suppose that for every $X\rightarrow Y$ where $X$ is an artinian scheme, $f|_X=0$. Then $f=0$. Suppose $\phi : F\rightarrow G$ is a morphism of vector sheaves such that for every $X\rightarrow S$ with $X$ artinian, $\phi |_X=0$. Then $\phi =0$. \end{lemma} {\em Proof:} We work with vector schemes over base schemes which are not necessarily of finite type over $k$ (the definition is the same, but we require additionally that the vector scheme be of finite type over the base). If $U\rightarrow V$ is a morphism of vector schemes over a henselian local ring $A$, and if $v$ is a section of $V$ over $A$ such that for each $n$ there exists $u_n\in U(Spec (A/{\bf m}^n))$ with $u_n$ mapping to the restriction of $v$, then there exists a section $u$ of $U$ over $A$ which maps to $v$. This follows from the strong Artin approximation theorem at maximal ideals, applied to finding sections of the morphism $U\times _VSpec (A)\rightarrow Spec (A)$. Now onto the proof of the lemma. For the first statement, any section $f$ is contained in a vector subsheaf of $F$, so we may suppose that $F$ is a vector sheaf. Choose a presentation $$ U\rightarrow V \rightarrow F \rightarrow 0 $$ by vector schemes. We may replace $X$ by a covering, so we may suppose that our section $f$ comes from a section $v$ of $V$. From the previous paragraph, for every henselized local ring $A$ of $X$, there exists a section $u$ of $U (Spec (A)$ mapping to $v$. But any such $A$---henselization at a point $P$---is the direct limit of algebras $A_i$ etale of finite type over $X$ (which give etale neighborhoods of $P$), and the space of sections is the direct limit: $$ U(Spec (A))= \lim _{\rightarrow } U( Spec (A_i )). $$ Thus there is a section $u_i$ over some $Spec (A_i)$ mapping to $v$. Thus every point $P$ of $X$ has an etale neighborhood on which there is a lifting of $v$ to a section of $U$. This implies that the image of $v$ in the cokernel $F$ in the etale topology, is zero. This gives the first statement, and the second statement follows easily from this. \hfill $\Box$\vspace{.1in} {\em Remark:} An alternative to the above proof is to use Lemma \ref{Krull}. The utility of this property comes from the following fact. \begin{corollary} \mylabel{I.g} If $F$ is a vector scheme, and if $Y\rightarrow S$ is an element of ${\cal X} /S$ with $Y$ artinian, then the functor $F_Y: Z\mapsto F(Y\times _{Spec (k)}Z)$ from schemes over $Spec (k)$ to sets, is represented by an additive group scheme (that is, a finite dimensional vector space) over $k$. This vector space is the $k$-module $F(Y)$. \end{corollary} {\em Proof:} By Lemma \ref{I.d}, we have $F_Y(Z)= F(Y)\otimes _k {\cal O} (Z)$ which is the space of morphisms of schemes from $Z$ to the vector space $F(Y)$. Thus $F_Y$ is represented by the vector space $F(Y)$. Note that from the exact sequences used in the proof of Lemma \ref{I.d}, $F(Y)$ is a finite-dimensional $k$-vector space. \hfill $\Box$\vspace{.1in} The group scheme ${\bf G} _m$ acts on every vector sheaf, by scalar multiplication. This action may be thought of as an action of the functor ${\bf G} _m (X)$ on $F(X)$, or as an automorphism of $F(X\otimes {\bf G} _m )$ (multiplication by $t$) which is natural in $X$. We have seen above that if $F\rightarrow G$ is a morphism of sheaves of abelian groups between two vector sheaves, then it is compatible with the ${\bf G} _m$ action. Suppose $A$ is a vector scheme, and $F$ is a vector sheaf. We look at $F(A)$, the space of sections over the scheme $A$. Let $F(A)^{\lambda}$ denote the subgroup of elements $f\in F(A)$ such that $f(ta)=t^{\lambda} f(a)$. Here $a\mapsto ta$ is considered as a morphism $A\times {\bf G} _m\rightarrow A$ over $S$, and $f(a)\mapsto t^{\lambda}f(a)$ is the automorphism of $F(A\times {\bf G} _m)$ given by scalar multiplication by $t^{\lambda} \in k[t,t^{-1}]$; the notation $f$ in the second half of the formula actually denotes the pullback of $f$ to $A\times {\bf G} _m$. \begin{lemma} \mylabel{I.h} With the above notations, $F(A)$ decomposes as a direct sum $$ F(A) = \bigoplus _{\lambda \in {\bf Z} ,\lambda \geq 0} F(A)^{\lambda} . $$ This direct sum decomposition is natural with respect to morphisms $F\rightarrow G$, and the linear piece $F(A)^1$ is exactly the space of morphisms of vector sheaves $A\rightarrow F$. \end{lemma} {\em Proof:} Recall that $F(A\times {\bf A}^1)= F(A)\otimes _kk[t]$, which we will just write as $F(A)[t]$. The morphism of scalar multiplication $A\times{\bf A}^1\rightarrow A$ gives $\Psi _t: F(A)\rightarrow F(A)[t]$ defined by $(\Psi _tf)(a):= f(ta)$ (to be accurate, this should be defined in terms of restriction maps for the morphisms involved, but we keep this notation for simplicity). Then $F(A)^{\lambda}$ is the set of $f$ such that $\Psi _tf = t^{\lambda}f$ in $F(A)[t]$. Let $\Psi _s[t]: F(A)[t]\rightarrow F(A)[s,t]$ denote the extension of $\Psi _s: F(A)\rightarrow F(A)[s]$ to the polynomials in $t$. We have $$ (\Psi _s[t]\Psi _tf)(a)= f(tsa)= (\Psi _{st}f)(a). $$ Write $$ \Psi _t(f)= \sum _{i=0}^{\infty} \psi _i(f)t^i, $$ where $\psi _i(f)\in F(A)$ and for any $f$, there are only a finite number of nonzero $\psi _i (f)$. Our previous formula becomes $$ \sum _{i,j} \psi _i (\psi _j (f))s^it^j = \sum _k \psi _k (f)(st)^k. $$ Comparing terms we see that $\psi _i(\psi _j (f))=0$ for $i\neq j$ and $\psi _i(\psi _i(f))=\psi _i (f)$. But in general $f\in F(A)^{\lambda}$ if and only if $\psi _i(f)=0$ for $i\neq \lambda$ and $\psi _{\lambda}(f)=f$. Therefore $\psi _i (f)\in F(A)^i$. Restrict to $t=1$, and note that the composed morphism $a\mapsto (a,1)\mapsto a$ is the identity so $\Psi _1(f)=f$. We get $$ f= \sum _{i=0}^{\infty} \psi _i (f) , $$ and this sum is actually finite. Thus every element of $F(A)$ can be expressed as a finite sum of elements of the $F(A)^{\lambda}$. On the other hand, this expression is unique: if $f= \sum f_i$ with $f_i\in F(A)^i$ then $$ \sum \psi _i(f)t^i=\Psi _t(f)=\sum \Psi _t(f_i)= \sum \psi _i (f_i)t^i = \sum f_i t^i, $$ and comparing coefficients of $t^i$ we get $f_i = \psi _i (f)$. This completes the proof of the decomposition (note that in working with ${\bf A}^1$ instead of ${\bf G} _m$ we obtain automatically that the exponents are positive). We have to show that $F(A)^1$ is equal to the space of linear morphisms from $A$ to $F$. A linear morphism gives an element of $F(A)^1$ (since it is compatible with the action of ${\cal O}$ by Lemma \ref{I.e}), and the resulting map from the space of morphisms to $F(A)^1$ is injective, since $F(A)$ is the space of morphisms of functors $A\rightarrow F$. Finally, we show surjectivity. For this, suppose given an element $\phi \in F(A)^1$. Suppose $Y$ is artinian, and $F$ is a vector sheaf over $S$. Then the functor $Z\mapsto F(Y\times Z )$ is represented by a vector space $F_Y$ over $k$ (Lemma \ref{I.g}). Our element of $F(A)$ now gives a morphism of schemes $\phi _Y: A_Y \rightarrow F_Y$ between these two vector spaces. It is compatible with scalar multiplication, so it is linear. In particular, if $u,v\in A_Y(Spec (k))= A(Y)$ then $\phi (u+v)= \phi (u)+\phi (v)$ in $F_Y(Spec (k))= F(Y)$. Now suppose $X$ is any element of ${\cal X} /S$. We show that $\phi : A(X) \rightarrow F(X)$ is a morphism of abelian groups. Suppose $u,v\in A(X)$. Let $f=\phi (u+v)-\phi (u)-\phi (v)\in F(X)$. By the previous paragraph, for any $Y\rightarrow X$ with $Y$ artinian, we have $f|_Y=0$. But the Krull property of Lemma \ref{I.f} then implies that $f=0$. This shows that $\phi$ is a morphism of sheaves of abelian groups. \hfill $\Box$\vspace{.1in} If $F$ and $G$ are sheaves of abelian groups, we denote by $Hom (F,G)$ the internal $Hom$, that is the sheaf of homomorphisms of sheaves of abelian groups from $F$ to $G$. The value $Hom (F,G)(X)$ is the space of morphisms of sheaves of abelian groups from $F|_{{\cal X} /X}$ to $G|_{{\cal X} /X}$ (this is already a sheaf). \begin{corollary} \mylabel{I.i} If $F\rightarrow G $ is a surjection of vector sheaves, and if $A$ is a vector scheme, then the morphism sheaves $$ Hom (A, F) \rightarrow Hom (A, G) $$ is surjective. If $X$ is affine then $$ Hom (A,F)(X)\rightarrow Hom (A, G)(X) $$ is surjective. \end{corollary} {\em Proof:} It suffices to prove the second statement. We may assume that $X=S$. We have $$ Hom (A, G)(S)= G(A)^1 $$ by the last statement of the lemma. Since $A$ is affine, the morphism $F(A)\rightarrow G(A)$ is surjective, and by the previous lemma this implies that $Hom (A,F)(S)=F(A)^1\rightarrow G(A)^1$ is surjective, giving the corollary. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.j} If $\phi : F\rightarrow G$ is a morphism of vector sheaves, then ${\rm coker}(\phi )$ and ${\rm ker} (\phi )$ are vector sheaves. \end{corollary} {\em Proof:} We may suppose that $S$ is affine and small enough. Choose presentations by vector schemes (cf the remark before Lemma \ref{I.d}) $$ U\rightarrow V\rightarrow F\rightarrow 0 $$ and $$ 0\rightarrow P \rightarrow R\rightarrow T \rightarrow G \rightarrow 0 $$ (note that the kernel $P$ is automatically a vector scheme). The morphism $V\rightarrow G$ lifts to a morphism $V\rightarrow T$, by the previous corollary, and we obtain a presentation $$ R \oplus V \rightarrow T \rightarrow {\rm coker} (\phi ) \rightarrow 0. $$ The fiber products $V\times _T R$ and $V\times _T R$ are vector schemes, and we have a presentation $$ U\times _T R\rightarrow V\times _T R \rightarrow {\rm ker}(\phi )\rightarrow 0. $$ \hfill $\Box$\vspace{.1in} Now we have shown that the category of vector sheaves is an abelian subcategory of the category of sheaves of abelian groups on ${\cal X} /S$. Suppose $A$ is a vector scheme and $F$ is a vector sheaf. Let ${\bf 3}$ denote the automorphism of $A$ obtained by multiplication by the scalar $3$ (any integer $\neq 0, \pm 1$ will do). We have $$ F(A)=\bigoplus F(A)^{\lambda } $$ (the decomposition given by Lemma \ref{I.h}) where $F(A)^{\lambda}$ may be characterized as the subspace of elements $f$ such that ${\bf 3}^{\ast}(f)= 3^{\lambda}f$. In particular, the linear subspace $Hom (A, F)= F(A)^1$ is characterized as the subspace of elements $f$ such that ${\bf 3}^{\ast}(f)= 3f$. \begin{theorem} \mylabel{I.k} Suppose $E$ and $G$ are vector sheaves, and $$ 0\rightarrow E \rightarrow F\rightarrow G \rightarrow 0 $$ is an extension in the category of sheaves of abelian groups on ${\cal X} /S$. Then $F$ is a vector sheaf. \end{theorem} {\em Proof:} We proceed in several steps. We may assume that $S$ is affine and small enough. Let $$ \begin{array}{ccccccc} & V & & & & B & \\ & \downarrow & & & & \downarrow & \\ & U & & & & A & \\ & \downarrow & & & & \downarrow & \\ 0 \rightarrow & E &\rightarrow & F & \rightarrow & G & \rightarrow 0 \\ & \downarrow & & & & \downarrow & \\ & 0 & & & & 0 & \end{array} $$ be presentations for $E$ and $G$. {\em Step 1.} {\em There exists a lifting of the morphism $A\rightarrow G$ to an element $\phi \in F(A)$ with $({\bf 3}^{\ast} - 3)^2\phi =0$.} The cohomology of $E$ over the affine $S$ is zero, so $$ 0\rightarrow E(A)\rightarrow F(A)\rightarrow G(A)\rightarrow 0 $$ is exact. Let $\alpha : A\rightarrow G$ denote the morphism in the presentation above, and choose $f\in F(A)$ mapping to $\alpha$. Then write $$ ({\bf 3}^{\ast} - 3)f = \sum e_i $$ with $e_{\lambda} \in E(A)^{\lambda}$ (thus ${\bf 3}^{\ast} e_{\lambda} = 3^{\lambda}e_{\lambda}$). Let $$ \phi = f-\sum c_{\lambda} e_{\lambda} $$ for $c_{\lambda} = (3^{\lambda }-3)^{-1}$ when $\lambda \neq 1$ (and $c_1=0$). We then have \begin{eqnarray*} ({\bf 3}^{\ast} - 3)\phi &=& ({\bf 3}^{\ast} - 3)f -\sum c_{\lambda}({\bf 3}^{\ast} - 3) e_{\lambda} \\ &=&\sum e_{\lambda}-\sum c_{\lambda}(3^{\lambda }-3)e_{\lambda} \\ &=& e_1. \end{eqnarray*} On the other hand, $({\bf 3}^{\ast} - 3)e_1=0$, so we get $$ ({\bf 3}^{\ast} - 3)^2\phi =0. $$ In other words, $\phi$ is in the generalized eigenspace for the eigenvalue $3$ of the transformation ${\bf 3}^{\ast}$. {\em Step 2.} {\em The extension $F$ satisfies the Krull property of Lemma \ref{I.f}: if $f\in F(X)$ such that for any artinian $Y\rightarrow X$, $f|_Y=0$, then $f=0$.} Under these hypotheses, $f$ maps to an element $g\in G(X)$ satisfying the same vanishing, so by Lemma \ref{I.f} we have $g=0$; thus $f$ comes from an element $e\in E(X)$. This element again satisfies the same vanishing, so by Lemma \ref{I.f}, $e=0$. {\em Step 3.} {\em If $A$ is a vector scheme and $F$ is an extension of two vector sheaves, then any element $\phi \in F(A)$ with $({\bf 3}^{\ast} - 3)^2\phi =0$ is a morphism of sheaves of abelian groups from $A$ to $F$.} Suppose $Y$ is artinian, and $G$ is a vector sheaf over $S$. Then the functor $Z\mapsto G(Y\times Z )$ is represented by a vector space $G_Y$ over $k$. If $F$ is an extension of two finite vetor sheaves $E$ and $G$, then let $F_Y$ denote the functor $Z\mapsto F(Y\times Z)$. We obtain an extension $$ 0\rightarrow E_Y \rightarrow F_Y \rightarrow G_Y \rightarrow 0 $$ in the category of sheaves of abelian groups over $Spec (k)$. But since the cohomology of the affine space $G_Y$ with coefficients in the additive group $E_Y$ vanishes, there is a lifting of the identity to a section $u\in F_Y(G_Y)$. Using $u$ we obtain an isomorphism of functors $F_Y \cong E_Y \times G_Y$, so $F_Y$ is a scheme. Since $F_Y$ is a sheaf of abelian groups, $F_Y$ is an abelian group-scheme over $k$. Since it is an extension of two additive groups, it is additive. Our element of $F(A)$ now gives a morphism of schemes $\phi _Y: A_Y \rightarrow F_Y$ between these two vector spaces. We still have $({\bf 3}^{\ast}-3)^2)\phi _Y =0$. But $F_Y(A_Y)$ decomposes into eigenspaces $$ F_Y(A_Y)=\bigoplus F_Y(A_Y)^{\lambda} $$ where $f\in F_Y(A_Y)^{\lambda} \Leftrightarrow f(ta)= t^{\lambda}f(a)$. In particular, $F_Y(A_Y)$ is the $3^{\lambda}$-eigenspace for ${\bf 3}^{\ast}$. But since the space $F_Y(A_Y)$ is the direct sum of eigenspaces, the generalized eigenspaces are equal to the eigenspaces, so $\phi _Y \in F_Y(A_Y)^1= Hom (A_Y, F_Y)$. In particular, if $u,v\in A_Y(Spec (k))= A(Y)$ then $\phi (u+v)= \phi (u)+\phi (v)$ in $F_Y(Spec (k))= F(Y)$. Now suppose $X$ is any element of ${\cal X} /S$. We show that $\phi : A(X) \rightarrow F(X)$ is a morphism of abelian groups. Suppose $u,v\in A(X)$. Let $f=\phi (u+v)-\phi (u)-\phi (v)\in F(X)$. By the previous paragraph, for any $Y\rightarrow X$ with $Y$ artinian, we have $f|_Y=0$. But the Krull property of Step 2 then implies that $f=0$. This shows that $\phi$ is a morphism of sheaves of abelian groups. {\em Step 4.} {\em There is a surjection from a vector scheme to $F$.} The direct sum of the morphism $U\rightarrow F$ with our lifting $\phi : A\rightarrow F$ gives a surjection $U\oplus A \rightarrow F \rightarrow 0$. In fact, this fits into a diagram $$ \begin{array}{ccccccc} 0 \rightarrow &U& \rightarrow & U\oplus A& \rightarrow & A&\rightarrow 0\\ & \downarrow & & \downarrow & & \downarrow & \\ 0 \rightarrow & E &\rightarrow & F & \rightarrow & G & \rightarrow 0\\ & \downarrow & & \downarrow & & \downarrow & \\ & 0 & & 0 & & 0 . & \end{array} $$ {\em Step 5.} {\em There is a surjection from a vector scheme to the kernel of $U\oplus A \rightarrow F$ (proving the theorem).} Taking the kernels along the top row of the above diagram gives $$ \begin{array}{ccccccc} 0 \rightarrow & K & \rightarrow & L & \rightarrow & M & \rightarrow 0\\ & \downarrow & & \downarrow & & \downarrow & \\ 0 \rightarrow &U& \rightarrow & U\oplus A& \rightarrow & A&\rightarrow 0\\ & \downarrow & & \downarrow & & \downarrow & \\ 0 \rightarrow & E &\rightarrow & F & \rightarrow & G & \rightarrow 0\\ & \downarrow & & \downarrow & & \downarrow & \\ & 0 & & 0 & & 0 . & \end{array} $$ But $K$ and $M$ are vector sheaves, and we have surjections $V\rightarrow K \rightarrow 0$ and $B\rightarrow M \rightarrow 0$. By repeating the above argument in this case, we obtain a surjection $$ V\oplus B \rightarrow L \rightarrow 0, $$ finally giving our presentation $$ V\oplus B \rightarrow U\oplus A \rightarrow F \rightarrow 0. $$ Thus $F$ is a vector sheaf. \hfill $\Box$\vspace{.1in} Our abelian category ${\cal V}$ of vector sheaves is therefore closed under extensions of sheaves of abelian groups. \subnumero{Duality} Suppose $F, G$ are vector sheaves. We have defined $Hom (F,G)$ which is for now a sheaf of abelian groups. Put $$ F^{\ast} := Hom (F, {\cal O} ). $$ If $\phi :F\rightarrow G$ is a morphism of vector schemes, then we obtain a morphism $\phi ^t:G^{\ast} \rightarrow F^{\ast}$, and the construction $\phi \mapsto \phi ^t$ preserves composition (reversing the order, of course). \begin{lemma} \mylabel{I.l} {\rm (Hirschowitz \cite{Hirschowitz})} Suppose $$ 0\rightarrow U\rightarrow V\rightarrow W \rightarrow F \rightarrow 0 $$ is an exact sequence with $U$, $V$ and $W$ vector schemes. Then taking the dual gives an exact sequence $$ 0\rightarrow F^{\ast}\rightarrow W^{\ast}\rightarrow V^{\ast} \rightarrow U^{\ast} \rightarrow 0. $$ \end{lemma} {\em Proof:} Note first that the compositions are zero, since taking the dual is compatible with compositions (and the dual of the zero map is zero!). The map $F^{\ast}\rightarrow W^{\ast}$ is injective because $W\rightarrow F$ is surjective (so any morphism $F\rightarrow {\cal O}$ restricting to $0$ on $W$, must be zero). The morphism $V^{\ast }\rightarrow U^{\ast}$ is surjective: if $a:U\rightarrow {\cal O}$ is a morphism, it can be interpreted as a section of ${\cal O} (U)^1$; but since $U\subset V$ is a closed subscheme, we can extend this to a section $a'\in {\cal O} (V)$, then let $a''$ be the component of $a'$ in ${\cal O} (V)^1$; restriction from ${\cal O} (V)$ to ${\cal O} (U)$ is compatible with the ${\bf G} _m$ action, hence with the decomposition of Lemma \ref{I.h}, so $a''$ restricts to $a$. Suppose $b : W\rightarrow {\cal O}$ restricts to zero on $V$; then it factors through the quotient sheaf $F=W/V$, so it comes from $F^{\ast}$. Thus the sequence is exact at $W^{\ast}$. We still have to prove exactness at $V^{\ast}$. Choose embeddings $U\hookrightarrow {\cal O} ^m$ and $W\hookrightarrow {\cal O} ^n$. Then extend the first to a function $V\rightarrow {\cal O} ^m$; combining with the second we obtain $V\hookrightarrow {\cal O} ^{m+n}$, fitting into a diagram $$ \begin{array}{ccccccc} 0\rightarrow & {\cal O} ^m& \rightarrow & {\cal O} ^{m+n} & \rightarrow & {\cal O} ^n & \rightarrow 0 \\ &\downarrow &&\downarrow && \downarrow & \\ 0\rightarrow & U& \rightarrow & V & \rightarrow & W & . \end{array} $$ Furthermore, $U= {\cal O} ^m \cap V$ as subschemes of ${\cal O} ^{m+n}$ (by the injectivity of $W\rightarrow {\cal O} ^n$). Given a linear map $\lambda : V\rightarrow {\cal O}$ such that $\lambda |_{U}=0$, extend it to $\varphi : {\cal O} ^{m+n}\rightarrow {\cal O}$ such that $\varphi |_{{\cal O} ^{m}}=0$. Replace $\varphi$ by its linear part under the decomposition of Lemma \ref{I.h} (this will conserve the property $\varphi |_{{\cal O} ^{m}}=0$ as well as the property of restricting to $\lambda$). Our $\varphi$ now descends to a map ${\cal O} ^n\rightarrow {\cal O}$, restricting to $\varphi |_W$ which extends $\lambda$. Note in the previous paragraph, we have used the following general fact: if $X,Y\subset Z$ are closed subschemes of an affine scheme, and $\lambda \in {\cal O} (X)$ such that $\lambda |_{X\cap Y}=0$, then there exists $\varphi \in {\cal O} (Z)$ such that $\varphi |_X=\lambda$ and $\varphi |_Y=0$. To prove this, let $I_X$, $I_Y$ and $I_{X\cap Y}$ denote the ideals of $X$, $Y$ and $X\cap Y$ in the coordinate ring ${\cal O} (Z)$. The definition of the scheme-theoretic intersection $X\cap Y$ is that $I_{X\cap Y}= I_X+I_Y$, and our statement follows from the translation that $$ I_Y \rightarrow I_{X\cap Y}/I_X \subset {\cal O} (Z)/I_X $$ is surjective. We have completed the proof of the lemma. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.m} The functor $F\mapsto F^{\ast}$ is an exact functor from the category of finite vector sheaves, to the category of sheaves of abelian groups. \end{corollary} {\em Proof:} Suppose $$ 0\rightarrow F' \rightarrow F \rightarrow F'' \rightarrow 0 $$ is an exact sequence of vector schemes. Choose presentations $$ 0\rightarrow U'\rightarrow V'\rightarrow W'\rightarrow F' \rightarrow 0 $$ and $$ 0\rightarrow U''\rightarrow V''\rightarrow W''\rightarrow F \rightarrow 0, $$ and combine these into a presentation $$ 0\rightarrow U\rightarrow V\rightarrow W\rightarrow F \rightarrow 0 $$ with $U=U'\oplus U''$, $V=V'\oplus V''$ and $W=W'\oplus W''$ (using the method of Theorem \ref{I.k}, which is easier since we now have the required lifts automatically). These fit together into a diagram $$ \begin{array}{ccccccc} &0&&0&&0 & \\ & \downarrow & & \downarrow && \downarrow \\ 0\rightarrow &U'& \rightarrow &U& \rightarrow &U''& \rightarrow 0 \\ & \downarrow & & \downarrow && \downarrow \\ 0\rightarrow &V'& \rightarrow &V& \rightarrow &V''& \rightarrow 0 \\ & \downarrow & & \downarrow && \downarrow \\ 0\rightarrow &W'& \rightarrow &W& \rightarrow &W''& \rightarrow 0 \\ & \downarrow & & \downarrow && \downarrow \\ 0\rightarrow &F'& \rightarrow &F& \rightarrow &F''& \rightarrow 0 \\ & \downarrow & & \downarrow && \downarrow \\ &0&&0&&0 & \end{array} $$ where all the rows and columns are exact. Apply duality to this diagram; we obtain a diagram with the arrows reversed, with the columns exact, by the lemma. Furthermore, the same lemma shows that the upper three rows are exact (in fact, this is easier because the rows in the original diagram are split, by construction). This implies that the bottom row is exact, as desired. \hfill $\Box$\vspace{.1in} A {\em coherent sheaf} is a sheaf which (locally) has a presentation of the form $$ {\cal O} ^n\rightarrow {\cal O} ^m\rightarrow F\rightarrow 0. $$ In particular, note that it is a vector sheaf. This coincides with the usual definition: if $S$ is affine and $X\rightarrow S$ is a morphism, then $F(X)=F(S)\otimes _{{\cal O} (S)}{\cal O} (X)$ (this is because the same is true for ${\cal O}$, and the presentation remains exact on the right after tensoring). As usual, we can assume that a presentation as above exists globally over any affine base. \begin{corollary} \mylabel{I.n} The dual of a coherent sheaf is a vector scheme and vice-versa. \end{corollary} {\em Proof:} Note that ${\cal O} ^{\ast}={\cal O}$. Taking the dual of a presentation of a coherent sheaf gives $$ 0\rightarrow F^{\ast} \rightarrow {\cal O} ^m \rightarrow {\cal O} ^n, $$ so $F^{\ast}$ is a vector scheme (the kernel here is a closed subscheme of ${\cal O} ^n$). Conversely, if $V$ is a vector scheme, take an exact sequence such as given in Lemma \ref{I.b}, and apply the dual. We obtain a presentation for $V^{\ast}$ as a coherent sheaf. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.o} The dual of a vector sheaf is again a vector sheaf. \end{corollary} {\em Proof:} If $F$ is a vector sheaf, choose a presentation $$ U\rightarrow V\rightarrow F\rightarrow 0 $$ by vector schemes. Taking the dual gives $$ 0\rightarrow F^{\ast}\rightarrow U^{\ast}\rightarrow V^{\ast}. $$ By the previous corollary, $U^{\ast}$ and $V^{\ast}$ are coherent sheaves, in particular vector schemes. Thus $F^{\ast}$ is the kernel of a morphism of vector sheaves, so $F^{\ast}$ is a vector sheaf. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.p} If $F$ is a vector sheaf, then $F^{\ast\ast}=F$ (via the natural morphism). \end{lemma} {\em Proof:} If $F$ is a vector scheme, this follows from the construction given in Corollary \ref{I.n}: write $F=\ker (M)$ as the kernel of a matrix $M:{\cal O} ^m\rightarrow {\cal O} ^n$; then $F^{\ast} = {\rm coker} (M^t)$ is the cokernel of the transpose matrix (and this $M^t$ is really just the transpose, keeping the same coefficients as in $M$). Finally, $F^{\ast\ast}=\ker (M^{tt})$, but the transpose of the transpose is the same matrix $M=M^{tt}$, so $F=F^{\ast\ast}$. (The same argument works for coherent sheaves, of course). If $F$ is any vector scheme, choose a presentation $$ U\rightarrow V\rightarrow F\rightarrow 0 $$ and take the double dual. Since $U^{\ast\ast}=U$ and $V^{\ast\ast}=V$ we get $$ U\rightarrow V\rightarrow F^{\ast\ast}\rightarrow 0, $$ so $F^{\ast\ast}=F$. \hfill $\Box$\vspace{.1in} We have now shown that duality is an exact contravariant involution on the category ${\cal V} $ of vector sheaves, interchanging vector schemes and coherent sheaves. \begin{lemma} \mylabel{I.q} The vector schemes are projective objects in ${\cal V} $, and the coherent sheaves are injective objects. There exist enough projectives and injectives (assuming that $S$ is affine). \end{lemma} {\em Proof:} The argument given above shows that a vector scheme $A$ is a projective object: if $F\rightarrow G$ is a surjection of vector sheaves then, since $A$ is affine, $F(A)^1\rightarrow G(A)^1$ is surjective. By definition, every vector sheaf admits a surjection from a vector scheme, so there are enough projectives. By duality, the coherent sheaves are injective and there are enough injectives. \hfill $\Box$\vspace{.1in} Taking the dual of the three step resolution by vector schemes shows that every vector sheaf $F$ admits a resolution $$ 0\rightarrow F \rightarrow U\rightarrow V\rightarrow W\rightarrow 0, $$ with $U$, $V$ and $W$ coherent sheaves (in particular, injective). \subnumero{Internal $Hom$ and tensor products} We begin with a corollary to the last lemma. \begin{corollary} \mylabel{I.r} If $A$ is a vector scheme, then the functor $V\mapsto Hom (A,V)$ from ${\cal V} $ to the category of abelian sheaves, is exact. If $F$ is a coherent sheaf, then the functor $V\mapsto Hom (V,F)$ is exact. \end{corollary} {\em Proof:} If $S$ is affine, the functors $V\mapsto Hom (A,V)(S)$ and $V\mapsto Hom (V,F)(S)$ are exact, by the lemma. But the restriction of a vector scheme or a coherent sheaf, to any object $X\in {\cal X} /S$ is again a vector scheme or coherent sheaf over $X$, so we obtain exactness over every affine object; and since exactness is a local condition, we get exactness. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{I.s} If $F$ and $G$ are vector sheaves, then $Hom (F,G)$ is a vector sheaf. \end{lemma} {\em Proof:} Suppose $F$ and $G$ are vector schemes. Then the exact sequence $$ 0\rightarrow G\rightarrow {\cal O} ^a \rightarrow {\cal O} ^b $$ yields an exact sequence $$ 0\rightarrow Hom (F,G)\rightarrow Hom (F, {\cal O} ^a) \rightarrow Hom (F,{\cal O} ^b); $$ but the middle and right terms are direct sums of the dual $F^{\ast}$ which is a vector sheaf, so the kernel $Hom (F,G)$ is a vector sheaf. Now suppose $F$ is a vector scheme and $G$ is a vector sheaf; resolving $G$ by vector schemes we obtain a resolution of $Hom (F,G)$ by vector sheaves, from the previous sentence. Thus $Hom (F,G)$ is a vector sheaf in this case too. Now suppose $F$ is a vector scheme, and choose a resolution $$ U\rightarrow V\rightarrow F\rightarrow 0 $$ by vector schemes. The functor $W\mapsto Hom (W,G)$ is contravariant and left exact for any $G$, so we obtain an exact sequence $$ 0\rightarrow Hom (F,G)\rightarrow Hom (V,G)\rightarrow Hom (U,G). $$ The middle and right terms are vector sheaves by the previous arguments, so the kernel is also. This completes the proof in general. \hfill $\Box$\vspace{.1in} We now define the {\em tensor product} $F\otimes ^{{\cal V}} G$ of two vector sheaves to be $$ F\otimes _{{\cal O}}G:= (Hom (F, G^{\ast}))^{\ast}. $$ Beware that this is not just the tensor product of sheaves of ${\cal O}$-modules (although this will be the case if $F$ and $G$ are coherent sheaves). We can also define the {\em cotensor product} $$ F\otimes ^{{\cal O}} G:= Hom (F^{\ast} , G). $$ Again, beware here that this is not equal to the tensor product. The difference is seen in noting that the tensor product is right exact as usual, whereas the cotensor product is left exact. (These exactness statements hold in both variables since the tensor and cotensor products are commutative, as we see below). Duality permutes the tensor and cotensor products: $$ (F\otimes _{{\cal O}}G)^{\ast}= F^{\ast}\otimes ^{{\cal O}}G^{\ast} $$ and $$ (F\otimes ^{{\cal O}}G)^{\ast}= F^{\ast}\otimes _{{\cal O}}G^{\ast}. $$ Define recursively $$ V_1\otimes \ldots \otimes V_n := V_1\otimes (V_2\otimes \ldots \otimes V_{n}) $$ for either one of the tensor products. By {\em multilinear form} $V_1\times \ldots V_n\rightarrow W$ we mean simply a multilinear morphism of sheaves of groups. In the same way as above for the linear morphisms, we obtain a vector sheaf $Mult(V_1\times \ldots \times V_n , W)$ of multilinear forms (denoted $Bil (\;\;\; )$ when $n=2$). \begin{proposition} \mylabel{I.s.1} 1. \,\, There is a natural isomorphism $\alpha _{U,V}:Hom (U^{\ast}, V)\cong Hom (V^{\ast}, U)$ and $\alpha _{U,V}\alpha _{V,U}$ is the identity. \newline 2. \,\, There is a natural isomorphism $$ Multi (V_1\times \ldots \times V_n, W)\cong Hom (W^{\ast}, Multi (V_1\times \ldots \times V_n, {\cal O} ). $$ 3. \,\, There is a natural isomorphism $$ Multi (V_1\times \ldots \times V_n , W)\cong Hom (V_1, Multi (V_2\times\ldots \times V_n, W)). $$ \end{proposition} {\em Proof:} In each case one defines natural maps in both directions and checks that the two compositions are the identity. \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{I.s.2} Suppose $V_i$ are vector sheaves, $i=1,\ldots , n$. There is a multilinear form $$ \mu : V_1\times \ldots V_n \rightarrow V_1 \otimes _{{\cal O}} \ldots \otimes _{{\cal O}} V_n $$ which is universal in the sense that if $$ \phi: V_1\times \ldots \times V_n\rightarrow W $$ is a multilinear form then there is a unique morphism $$ \psi : V_1 \otimes _{{\cal O}} \ldots \otimes _{{\cal O}} V_n \rightarrow W $$ such that $\phi = \psi \circ \mu $. \end{theorem} {\em Proof:} Note first that for $n=2$ there is a natural bilinear map $U\times V \rightarrow Hom (U, V^{\ast})^{\ast}= U\otimes _{{\cal O}}V$. Inductively this gives the multilinear map for any $n$. The universal property says that the induced map $$ Hom (V_1\otimes _{{\cal O}}\ldots \otimes _{{\cal O}}V_n,W)\rightarrow Multi (V_1,\ldots , V_n , W) $$ should be an isomorphism. We prove this by induction on $n$, so we may suppose it is true for $n-1$. By the definition of the multiple tensor product, the quantity on the left is $$ Hom (Hom (V_1, (V_2\otimes _{{\cal O}}\ldots \otimes _{{\cal O}}V_n)^{\ast})^{\ast},W). $$ By part 1 of the proposition, this is equal to $$ Hom (W^{\ast}, Hom (V_1, (V_2\otimes _{{\cal O}}\ldots \otimes _{{\cal O}}V_n)^{\ast})). $$ By induction, $(V_2\otimes _{{\cal O}}\ldots \otimes _{{\cal O}}V_n)^{\ast}=Multi (V_2\times \ldots \times V_n,{\cal O} )$. Coupled with part 3 of the proposition we get $$ Hom (W^{\ast}, Multi (V_1\times \ldots \times V_n,{\cal O} ) $$ which then is equal to the right hand side above, by part 2 of the proposition. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.s.3} The tensor and cotensor products have natural commutativity and associativity isomorphisms satisfying the usual constraints. \end{corollary} {\em Proof:} For the tensor product this follows from the universal property and the fact that the notion of multilinear form is independent of the order of the variables. For the cotensor product this follows because it is the dual of the tensor product. \hfill $\Box$\vspace{.1in} We can define symmetric and exterior powers, either with respect to the tensor product or with respect to the cotensor product. Let $S_n$ denote the symmetric group on $n$ objects. Let $V^{\otimes _{{\cal O}}n}$ (resp. $V^{\otimes ^{{\cal O}}n}$) denote the tensor product (resp. cotensor product) of $n$ copies of a vector sheaf $V$. Then $S_n$ acts on $V^{\otimes _{{\cal O}}n}$ (resp. $V^{\otimes ^{{\cal O}}n}$) because of the commutativity and associativity. This representation is completely reducible (this can be see object-by-object). The components are vector sheaves; this can be seen by noting that the fixed part is the kernel of a morphism of vector sheaves (the direct sum of $1-\gamma $ for $\gamma \in S_n$); to get the other components, apply that to tensor products with the irreducible representations of $S_n$. The trivial component of $V^{\otimes _{{\cal O}}n}$ (resp. $V^{\otimes ^{{\cal O}}n}$) is denoted by $Sym _{{\cal O}}^n(V)$, the symmetric power (resp. $Sym ^{{\cal O}}_n(V)$, the symmetric copower). The component corresponding to the sign representation is denoted by $\bigwedge _{{\cal O}}^n(V)$, the exterior power (resp. $\bigwedge ^{{\cal O}}_n(V)$, the exterior copower). {\em Remark:} There is a natural morphism $U\otimes _{{\cal O}}V\rightarrow U\otimes ^{{\cal O}}V$, that is to say $$ Hom (U,V^{\ast})^{\ast}\rightarrow Hom (U^{\ast}, V). $$ However, this is not an isomorphism. A counterexample can be constructed by looking for a case where the cotensor product is left exact but not right exact (and noting that the tensor product is right exact), or vice-versa. We have the expression $U\otimes ^{{\cal O}}V= Bil (U^{\ast}\times V^{\ast},{\cal O})$. The inequality mentionned above implies in particular that there are bilinear functions on $U^{\ast}\times V^{\ast}$ which are not sums of tensors $u\otimes v$. This is a big difference from the case of schemes (for example if $U$ and $V$ are coherent so that $U^{\ast}$ and $V^{\ast}$ are vector schemes, then the bilinear functions {\em are} sums of tensor products. \subnumero{Automorphisms of vector sheaves} We end our discussion of vector sheaves by showing how they give examples of presentable group sheaves. \begin{lemma} \mylabel{I.1.g.1} If $V$ is a vector sheaf over $S$, then $V$ is presentable. \end{lemma} {\em Proof:} Suppose $V$ is a vector scheme. Then taking $X=V$ and $R= V\times _VV=V$ we obtain the required presentation (note that the identity morphisms are vertical)---so $V$ is $P4$, and then $P5$ by Corollary \ref{I.z}. It follows from Theorem \ref{I.1.d} that the quotient of one vector scheme by another is again $P5$; and finally that the quotient of a vector scheme by such a quotient is $P5$. In view of the 3-stage resolution of any vector sheaf by vector schemes, we obtain the lemma. \hfill $\Box$\vspace{.1in} One of the main examples of presentable group sheaves is given by the following theorem. \begin{theorem} \mylabel{I.1.g} Suppose $V$ is a vector sheaf over $S$. Then the group sheaf $Aut (V)$ is a presentable. \end{theorem} {\em Proof:} By the previous lemma and Lemma \ref{I.s}, $Hom (V,V)$ is $P5$. We can express $$ Aut (V)\subset Hom (V,V)\times Hom (V,V) $$ as the equalizer of the two morphisms $$ \begin{array}{ccc} Hom (V,V)\times Hom (V,V)&\rightarrow &Hom (V,V)\times Hom (V,V)\\ (a,b) & \mapsto & (ab,ba) \\ (a,b)&\mapsto & (1,1). \end{array} $$ Apply Lemma \ref{I.1.a} to obtain that $Aut (V)$ is $P4$, and then Corollary \ref{I.z} to obtain that it is $P5$. \hfill $\Box$\vspace{.1in} A particular case of this construction is when $V$ is a coherent sheaf which we denote by ${\cal F}$. There is a presentation $$ U_2 \stackrel{\phi}{\rightarrow} U_1 \rightarrow {\cal F} \rightarrow 0 $$ where $U_i = {\cal O} ^{a_i}$. Let $Aut (U_2, U_1, \phi )$ denote the group sheaf of automorphisms of the morphism $U_2 \rightarrow U_1$. Any such automorphism gives an automorphism of ${\cal F}$ so we have a morphism $$ Aut (U_2, U_1, \phi )\rightarrow Aut ({\cal F} ). $$ \begin{lemma} \mylabel{surjection} This morphism is a surjection onto $Aut ({\cal F} )$, and $Aut (U_2, U_1, \phi )$ is represesentable by a group scheme over $S$. \end{lemma} {\em Proof:} The representability by a group scheme is clear, since $Aut (U_i)$ are group schemes (isomorphic to $GL(a_i)$) and the condition of compatibility with $\phi$ is a closed condition so $Aut (U_2, U_1, \phi )$ is a closed subscheme of $Aut (U_1)\times Aut (U_2)$. Suppose $S' \rightarrow S$ is a scheme and $P\in S'$ is a point. Suppose $\eta : {\cal F} |_{S'}\rightarrow {\cal F} |_{S'}$ is an automorphism. Let $$ U'_2 \stackrel{\phi '}{\rightarrow }U'_1 \rightarrow {\cal F} |_{S'}\rightarrow 0 $$ be a minimal resolution of ${\cal F} |_{S'}$ at the point $P$ (that is to say that the value $\phi '(P)$ is identically zero and the rank of $U'_2$ is minimal). Then there are locally free $W_i\cong {\cal O} ^{b_i}$ on $S'$ such that $U_i|_{S'} \cong U'_i \oplus W$ and such that the map $\phi |_{S'}$ can be written in block form with respect to this decomposition, with a morphism $\psi '$ in the block of the $W_i$ and the map $\phi '$ in the block of the $U'_i$, such that $\psi '$ is surjective. Our morphism $\eta$ extends to a morphism of resolutions $U'_{\cdot} \rightarrow U'_{\cdot}$ which is an isomorphism near $P$ by the minimality of the resolution (in fact the values $U'_i(P)$ are the $Tor ^i_{{\cal O} _{S'}}({\cal F} |_{S'}, k_P)$ and an isomorphism of ${\cal F} |_{S'}$ induces an isomorphism on the $Tor ^i$). We can complete this with the identity in the block of the $W_i$ to get an isomorphism of resolutions $U_i |_{S'}$ inducing $\eta$. This gives the desired surjectivity. \hfill $\Box$\vspace{.1in} {\em Question:} Does a similar result hold for the automorphisms of any vector sheaf? \numero{Tangent sheaves of presentable sheaves} Suppose $S'\rightarrow S$ is an $S$-scheme. Put $$ Y:= S' \times Spec (k[\epsilon _1 ,\epsilon _2, \epsilon _3]/(\epsilon _i^2, \epsilon _i\epsilon _j )) $$ with the subschemes $$ Y_i:= S' \times Spec (k[\epsilon _i ]/(\epsilon _i^2)) $$ and $$ Y_{ij}:= S' \times Spec (k[\epsilon _i ,\epsilon _j]/(\epsilon _i^2, \epsilon _j^2, \epsilon _i\epsilon _j )). $$ Note that $Y=Y_1\cup Y_2\cup Y_3$, and $Y_{ij}=Y_i\cup Y_j$, as well as $Y_i\cap Y_k = S'\subset Y$ and $Y_{ij}\cap Y_{jk}=Y_j$ (for $i\neq k$). It should be stated explicitly that $Y_{ij}$ is the closed subscheme defined by the ideal $(\epsilon _k)$, $k\neq i,j$; and $Y_i$ is the closed subscheme defined by the ideal $(\epsilon _j,\epsilon _k)$, $j,k\neq i$. We need a weaker version of the notion of verticality. We say that a morphism ${\cal F} \rightarrow {\cal G}$ of sheaves is {\em $T$-vertical} if it satisfies the lifting property $Lift_2(Y_{ij};Y_i ,Y_j)$ and $Lift _3(Y; Y_{12},Y_{23}, Y_{13})$ (for any $S'$). Note that these systems satisfy the retraction hypotheses in the lifting property, so the property of $T$-verticality is weaker than the property of verticality. The result of Theorem \ref{I.u} holds also for $T$-verticality, so the class ${\cal T}$ of $T$-vertical morphisms satisfies the axioms M1-M4. In particular the properties $P4$ and $P5$ imply $P4({\cal T} )$ and $P5({\cal T})$ respectively. The advantage of th weaker property of $T$-verticality is that if $X\rightarrow Z$ is a morphism of schemes over $S$, then it is $T$-vertical. To prove this, note that the properties $Y=Y_1\cup Y_2\cup Y_3$, $Y_{ij}=Y_i\cup Y_j$, $Y_i\cap Y_k = S'\subset Y$ and $Y_{ij}\cap Y_{jk}=Y_j$ mean that for defining morphisms from $Y$ to a scheme (or from $Y_{ij}$ to a scheme) it suffices to have compatible morphisms on the $Y_{ij}$ or on the $Y_i$. ({\em Caution:} We did not include the lifting condition $Lift _1(Y_1; S')$ in the notion of $T$-verticality; morphisms of schemes do not necessarily satisfy this lifting property!) The conclusion of the previous paragraph and property $M1$ for $T$-verticality is that if ${\cal F}$ is a $P4({\cal T} )$ sheaf then the structural morphism $p:{\cal F} \rightarrow S$ is $T$-vertical; thus $P4({\cal T} )\Leftrightarrow P5({\cal T})$. \begin{lemma} \mylabel{I.1.e.1} Suppose $f:{\cal F}\rightarrow {\cal G} $ is a morphism of $P4$ sheaves. Then $f$ is $T$-vertical. Furthermore, the liftings in the lifting properties for $f$, for the systems $(Y_{ij};Y_i ,Y_j)$ and $(Y; Y_{12},Y_{23}, Y_{13})$, are unique. \end{lemma} {\em Proof:} For $T$-verticality, we can choose vertical surjections $X\rightarrow {\cal F}$ and $Y\rightarrow {\cal G}$ so that there is a lifting $X\rightarrow Y$. This lifting is $T$-vertical since it is a morphism between schemes (cf the above remark). Hence the composition $X\rightarrow {\cal G}$ is $T$-vertical. By Theorem \ref{I.u}, part 4 for $T$-verticality, applied to the composition $X\rightarrow {\cal F} \rightarrow {\cal G}$, we obtain $T$-verticality of the morphism $f$. To prove the uniqueness, note that liftings to schemes are unique since $Y=Y_1\cup Y_2\cup Y_3$ and $Y_{ij}=Y_i\cup Y_j$. Then descend the uniqueness down from $X$ to ${\cal F}$ where $X\rightarrow {\cal F}$ is the vertical (hence $T$-vertical) morphism provided by the property $P4$. This descent of the uniqueness property is immediate from the lifting property for $X\rightarrow {\cal F}$. \hfill $\Box$\vspace{.1in} In the statement of the following theorem, the condition is $P4$ and not $P4({\cal T} )$ (i.e. that isn't a misprint). \begin{theorem} \mylabel{I.1.f} Suppose ${\cal F}\rightarrow {\cal G} $ is a morphism of $P4$ sheaves on $S$. Suppose $u :S\rightarrow {\cal F}$ is a section. Then the relative tangent sheaf $T(f )_{u}$ over $S$, defined by $$ T({\cal F} )_{u} (b:S'\rightarrow S):= \{ \eta :S' \times Spec (k[\epsilon ]/(\epsilon ^2))\rightarrow {\cal F}\;\; :\;\;\;\; f \eta = fubp_1 \;\; \mbox{and} \;\; \eta |_{S'}= ub \} , $$ has a natural structure of sheaf of abelian groups making it a vector sheaf. \end{theorem} {\em Proof:} We first define the natural abelian group structure on this sheaf. Suppose $$ \eta _i: S'\times Spec(k[\epsilon _i]/(\epsilon _i^2))\rightarrow {\cal F} $$ are sections of $T(f )_u$ over $S'$ ($i=1,,\ldots , 3$). ({\em Nota:} for the definition of the group law we only need $i=1,2$; we need $i=1,2,3$ only to check that it is associative.) Here (and below) we attach various subscripts to the variables $\epsilon$. Use the notations established above: $$ Y:= S' \times Spec (k[\epsilon _1 ,\epsilon _2, \epsilon _3]/(\epsilon _i^2, \epsilon _i\epsilon _j )) $$ with the subschemes $$ Y_i:= S' \times Spec (k[\epsilon _i ]/(\epsilon _i^2)) $$ and $$ Y_{ij}:= S' \times Spec (k[\epsilon _i ,\epsilon _j]/(\epsilon _i^2, \epsilon _j^2, \epsilon _i\epsilon _j )). $$ Note that $Y=Y_1\cup Y_2\cup Y_3$, and $Y_{ij}=Y_i\cup Y_j$. Again $Y_{ij}$ is the closed subscheme defined by the ideal $(\epsilon _k)$, $k\neq i,j$; and $Y_i$ is the closed subscheme defined by the ideal $(\epsilon _j,\epsilon _k)$, $j,k\neq i$. The systems $(Y_{ij};Y_i ,Y_j)$ and $(Y; Y_{12},Y_{23}, Y_{13})$ satisfy a unique lifting property for the morphism $f$ (Lemma \ref{I.1.e.1}). Note that $Y_{ij}\cap Y_{jk}=Y_j$ (for $i\neq k$). We apply this first to the system $(Y_{ij}; Y_i,Y_j)$. There is a unique morphism $$ \eta _{ij}: Y_{ij}\rightarrow {\cal F} $$ over the base morphism $Y_{ij}\rightarrow S\rightarrow {\cal G}$ and agreeing with $\eta _i$ (resp. $\eta _j$) on $Y_i$ (resp. $Y_j$). Let $$ \delta _{ij}: S' \times Spec (k[\epsilon ]/(\epsilon ^2))\rightarrow Y_{ij} $$ be the diagonal and---for future use---let $$ \delta _{123}: S' \times Spec (k[\epsilon ]/(\epsilon ^2))\rightarrow Y $$ be the triple diagonal. Then we put $$ \eta _i+\eta _j:= \eta _{ij} \circ \delta _{ij} . $$ This gives a composition which is obviously commutative (the definition is symmetric in the two variables). To check that it is associative, apply unique lifting for $(Y,Y_{ij})$ to get a unique $\eta _{123}: Y\rightarrow {\cal F}$ restricting to the $\eta _{ij}$ on $Y_{ij}$. Next, note that the triple diagonal is equal to the composition of $1\times \delta _{23}$ with the diagonal $$ S' \times Spec (k[\epsilon _0]/(\epsilon _0^2))\rightarrow Spec (k[\epsilon _1,\epsilon ]/(\epsilon _1^2, \epsilon ^2, \epsilon _1\epsilon )). $$ Using this, we get $$ \epsilon _1+(\eta _2 +\eta _3)= \eta _{123}\circ \delta _{123}. $$ Similarly, we have $$ (\epsilon _1+\eta _2) +\eta _3= \eta _{123}\circ \delta _{123}, $$ giving associativity. The identity element (which we denote by $0$) is the composition $$ S'\times Spec (k[\epsilon ]/(\epsilon ^2))\rightarrow S \rightarrow {\cal F} . $$ This construction is natural: if $$ \begin{array}{ccc} {\cal F} & \rightarrow & {\cal F} ' \\ \downarrow &&\downarrow \\ {\cal G} & \rightarrow & {\cal G} ' \end{array} $$ is a diagram with vertical arrows vertical, and if $u:S\rightarrow {\cal F}$ is a section projecting to $u':S\rightarrow {\cal F} '$, then composition with the morphism ${\cal F} \rightarrow {\cal F} '$ respects the conditions in the definition of the tangent sheaves, and so it gives a morphism $T(f)_u\rightarrow T(f')_{u'}$. The addition we have defined is natural, so this morphism of tangent sheaves respects the addition (it also respects the identity). The inverse is obtained by applying the automorphism $\epsilon \mapsto -\epsilon$. This completes the construction of the natural structure of sheaf of abelian groups. Next, we show that if $$ {\cal F} \stackrel{a}{\rightarrow }{\cal G} \stackrel{b}{\rightarrow }{\cal H} $$ is a sequence of morphisms of $P4$ sheaves, and if $u:S\rightarrow {\cal F}$ is a section, then we have an exact sequence $$ 0\rightarrow T(a)_u\rightarrow T(ba)_u \rightarrow T(b)_{au} $$ We certainly get such a sequence with the composition being zero. Furthermore, $T(a)_u$ is the subsheaf of $T(ba)_u$ consisting of those elements projecting to zero in $T(b)_{au}$ (this follows immediately from the definition). Furthermore, if $a$ is vertical, then the sequence is exact on the right. This follows from the lifting property in the definition of vertical, in view of the fact that $S'$ is a retraction of $S'\times Spec (k[\epsilon ]/(\epsilon ^2))$. (Note that we have not required this lifting property in the definition of $T$-verticality.) Let $p: {\cal F} \rightarrow S$ denote the structural morphism for a $P4$ sheaf ${\cal F}$, and define the tangent sheaf $T({\cal F} )_u:= T(p)_u$. If $f:{\cal F} \rightarrow {\cal G}$ is a morphism of $P4$ sheaves, the exact sequence of the previous paragraph becomes $$ 0\rightarrow T(f)_u\rightarrow T({\cal F})_u \rightarrow T({\cal G} )_{fu}. $$ Again, if $f$ is vertical then this sequence is exact on the right also. Finally, we show that if ${\cal F}$ is $P4$ then $T({\cal F} )_u$ is a vector sheaf. The above exact sequence implies that if $f$ is a morphism of $P4$ sheaves then $T(f)_u$ is a vector sheaf. Let $f: X\rightarrow {\cal F}$ be the vertical morphism given by the property $P4$. Since the question is etale local on $S$, we may assume that our section $u: S\rightarrow {\cal F}$ lifts to a section $v: S\rightarrow X$. We have an exact sequence $$ 0\rightarrow T(f)_v \rightarrow T(X)_v\rightarrow T({\cal F} )_u\rightarrow 0. $$ Note that $T(X)_v$ is a vector scheme (an easy thing to see---it is given by the linear parts of the equations of $X$ at the section $v$). Let $g:R\rightarrow X\times _{{\cal F}} X$ be the other vertical morphism given by the property $P4$. We claim that we have an exact sequence $$ 0\rightarrow T(X\times _{{\cal F}}X)_{(v,v)} \rightarrow T(X)_v \oplus T(X)_v \rightarrow T({\cal F} )_u \rightarrow 0. $$ To see this, note that an element of $T(X\times _{{\cal F}}X)_{(v,v)}$ consists of an element of $T(X\times _SX)_{(v,v)}$ mapping to $T({\cal F} )_u\subset T({\cal F} \times _S{\cal F} )_{(u,u)}$. Note that $$ T(X\times _SX)_{(v,v)}=T(X)_v \oplus T(X)_v, $$ and $$ T({\cal F} \times _S{\cal F} )_{(u,u)}=T({\cal F} )_u\oplus T({\cal F} )_u $$ with the map from $T({\cal F} )_u$ being the diagonal. The quotient of $T({\cal F} \times _S{\cal F} )_{(u,u)}$ by the diagonal $T({\cal F} )_u$ is thus isomorphic to $T({\cal F} )_u$ and we obtain the exact sequence in question. The surjectivity on the right is from surjectivity of $T(X)_v\rightarrow T({\cal F} )_u$. Lift $(v,v)$ to a section $w:S\rightarrow R$. The exact sequence for $g$ gives a surjection $$ T(R)_w \rightarrow T(X\times _{{\cal F}}X)_{(v,v)}\rightarrow 0. $$ Combining this with the above exact sequence, we obtain the right exact sequence $$ T(R)_w \rightarrow T(X)_v \oplus T(X)_v \rightarrow T({\cal F} )_u \rightarrow 0. $$ Since $T(R)_w$ and $T(X)_v$ are vector schemes, this shows that $T({\cal F} )_u$ is a vector sheaf. \hfill $\Box$\vspace{.1in} \numero{The case $S=Spec (k)$} We now analyse the definitions of the previous sections in the case where the base scheme is $S=Spec (k)$ (a hypothesis we suppose for the rest of this section). {\em Caution:} We will use throughout this section certain properties of vertical morphisms etc. which hold only in the context $S=Spec (k)$. The reader should not extrapolate these properties to other cases. Our first lemma is a preliminary version of the next lemma which we include because the argument may be easier to understand in a simpler context. \begin{lemma} \mylabel{I.1.k} Suppose $f: X\rightarrow Spec (k)$ is morphism of finite type. Then $f$ is vertical if and only if $f$ is a smooth morphism. \end{lemma} {\em Proof:} Suppose $X$ is smooth. Then the required lifting properties hold. Indeed, $X$ is etale locally a vector space, and Theorem \ref{I.u} (part 7) implies that vector spaces are vertical over $Spec(k)$. Conversely, suppose $f$ is vertical, and suppose $x\in X$. The first claim is that for any $v\in T(X)_x$ there is a smooth germ of curve $(C,0)$ mapping to $(X,x)$ with tangent vector $v$ at the origin. Since $X$ is of finite type, and by Artin approximation, it suffices to construct a compatible family of morphisms $$ \gamma _n:Spec (k[t]/t^n)\rightarrow X $$ sending $Spec(k)$ to $x$ and with tangent vector $v$ (that is, the map $\gamma _2$ represents $v$). Before starting the construction, choose a morphism $$ \mu : X\times X\rightarrow X $$ with $\mu (x,y)=\mu (y,x)=y$ for any $y$ (the possibility of finding $\mu$ follows from the definition of verticality). We now construct $\gamma _n$ by induction, starting with $\gamma _2$ given by $v$. Suppose we have constructed $\gamma _{n}$ by the inductive procedure. Let $Y(n):= Spec (k[r]/r^{n})\times Spec (k[s]/s^2)$. The composition gives a morphism $$ \phi _n:= \mu \circ (\gamma _n , \gamma _2): Y(n)\rightarrow X. $$ We will show that $\phi _n$ factors through the morphism $$ d: Y(n)\rightarrow Spec (k[t]/t^{n+1}) $$ which is dual to the morphism $$ k[t]/t^{n+1} \rightarrow k[r,s]/(r^n,s^2) $$ $$ t\mapsto r+s. $$ We will then choose $\gamma _{n+1}$ equal to the resulting morphism $Spec (k[t]/t^{n+1})\rightarrow X$, that is with $\phi _n =\gamma _{n+1}d$. Since $\gamma _n$ restricts to $\gamma _{n-1}$, and since we have chosen $\gamma _n$ by the inductive procedure, we have that $$ \phi _n|_{Y(n-1)}=\phi _{n-1} = \gamma _n d. $$ Writing $X=Spec (A)$ (in a neighborhood of $x$) the morphism $\phi _n$ corresponds to $$ \phi _n^{\ast}: A\rightarrow k[r,s]/(r^n,s^2). $$ We have that $\phi _n^{\ast} (a)$ reduces modulo $r^{n-1}$ to $d^{\ast}\gamma _n^{\ast} (a)$. Writing $$ \gamma _n^{\ast}(a)= \sum _{j=0}^{n-1}b_jt^j $$ we have $$ \phi _n^{\ast} (a)= \sum _{j=0}^{n-1}b_j (r+s)^j + \alpha r^{n-1} + \beta r^{n-1}s. $$ Write, on the other hand, the equation $\phi _n |_{Spec (k[r]/r^n)} = \gamma _n$. We get that $$ \phi _n^{\ast} (a) \sim \sum _{j=0}^{n-1}b_j r^j \;\; \mbox{mod} (s). $$ This gives $\alpha = 0$ in the above equation. Finally, note that $(r+s)^n= nr^{n-1}s$ modulo $(r^n, s^2)$. Thus we may set $b_n:= \beta / n$ and get $$ \phi _n^{\ast} (a)= \sum _{j=0}^{n}b_j (r+s)^j . $$ Put $$ \gamma _{n+1}^{\ast} (a):= \sum _{j=0}^{n}b_j t^j , $$ and we get the desired factorization $\phi _n = \gamma _{n+1}d$. This completes the inductive step for the construction of the $\gamma _n$. We obtain the desired formal curve and hence a curve $(C,0)$ as claimed. {\em Remark:} Intuitively what we have done above is to integrate the vector field on $X$ given by the tangent vector $v$ and the multiplication $\mu$. Of course, the curve $C$ is an approximation to the integral curve, which might only exist formally. The next step in the proof of the lemma is to choose a collection of vectors $v_1,\ldots , v_m$ generating $T(X)_x$, and to choose resulting curves $C_1, \ldots , C_m$. Using the map $\mu$ in succession (or applying directly the definition of verticality) we obtain a map $$ \Phi : (U,0):=(C_1\times \ldots \times C_m , 0)\rightarrow (X,x), $$ inducing the given morphisms on the factors $C_i$ (considered as subspaces of the product by putting the origin in the other places). By construction the differential $d\Phi _0$ is given by the vectors $v_1,\ldots , v_m$, in particular it gives a surjection $$ d\Phi _0: T(U)_0 \rightarrow T(X)_x \rightarrow 0. $$ Note that $U$ is smooth of dimension $m$. We claim that this implies $dim _x (X) \geq dim T(X)_x$. To see this, let $d:= dim _x(X)$ and $n:= dim T(X)_x$. By semicontinuity, the dimension of the fiber $\Phi ^{-1}(x)$ at the origin is at least equal to $m-d$. In particular, the tangent space to the fiber has dimension at least $m-d$; but this gives a subspace of dimension $m-d$ of $T(U)_0$ which maps to zero in $T(X)_x$; by the surjectivity of $d\Phi _0$ we get $n \leq m-(m-d)=d$, the desired inequality. Finally, it follows from this inequality that $X$ is regular at $x$ and hence smooth at $x$ (and, of course, the inequality is an equality!). This proves the lemma. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{smooth} Suppose $X$ and $Y$ are schemes of finite type over $k$ and $f: X\rightarrow Y$ is a morphism. Then $f$ is $Spec (k)$-vertical if and only if $f$ is smooth. \end{lemma} {\em Proof:} Note first that if $f$ is smooth then it is etale-locally a product with affine space so we get all of the lifting properties. Suppose now that $f$ is vertical. If $Q\in Y$ and $P\in f^{-1}(Q)$ then $Lift _1(Y, Q)$ implies that, after replacing $Y$ by an etale neighborhood of $Q$ we may suppose that there is a section $\sigma : Y\rightarrow X$ with $\sigma (Q)=P$. Let $T(X/Y)_{\sigma}$ denote the relative tangent vector scheme along the section $\sigma$. It is easy to see that the morphism $T(X/Y)_{\sigma}\rightarrow Y$ is $Spec(k)$-vertical. We then obtain that the morphism $$ \Gamma (Y, T(X/Y)_{\sigma})\rightarrow (T(X/Y)_{\sigma})_Q=T(f^{-1}(Q))_P $$ is surjective, and this then implies that $T(X/Y)_{\sigma}$ is a vector bundle over $Y$. The same argument as in the previous lemma allows us to ``exponentiate'' in a formal neighborhood of $P$, to get a map $\varphi$ from $T(X/Y)_{\sigma}^{\wedge}$ (the formal completion in a neighborhood of $0(Q)$) to $X$, which sends the zero section $0$ to $\sigma$ and whose tangent map is the identity along $\sigma$. We claim that if $S'$ is artinian local with a morphism $S''\rightarrow X$ sending the origin to $P$, then the morphism factors via $\varphi$ through a map $S'\rightarrow T(X/Y)_{\sigma}^{\wedge}$ sending the origin to $0(Q)$. Prove this claim using the standard deformation theory argument by induction on the length of $S'$: suppose $S''\subset S'$ is defined by an ideal $I$ annihilated by the maximal ideal, and suppose we know the claim for $S''$. Then there exists a map $S'\rightarrow T(X/Y)_{\sigma}^{\wedge}$ extending the known map on $S''$ since $T(X/Y)_{\sigma}^{\wedge}$ is a vector bundle over $Y$. The space of such extensions is a principal homogeneous space over $I\otimes _k (T(X/Y)_{\sigma})_Q$ whereas the space of extensions of $S''\rightarrow X$ to morphisms $S'\rightarrow X$ is a principal homogeneous space over $I\otimes _kT(f^{-1}(Q))_P$. The map $\varphi$ induces an isomorphism $$ (T(X/Y)_{\sigma})_Q\cong T(f^{-1}(Q))_P $$ so there is an extension to a map $S' \rightarrow T(X/Y)_{\sigma}^{\wedge}$ which projects to our given map $S'\rightarrow X$. This proves the claim. Now we can prove that $X\rightarrow Y$ is formally smooth at $P$. If $S''\subset S'$ are artinian local and if $a:S'\rightarrow Y$ is a map lifting over $S''$ to a map $b:S'' \rightarrow X$ sending the origin to $P$, then we get (from the previous claim) that the map $b$ factors through a map $S'' \rightarrow T(X/Y)_{\sigma}^{\wedge}$. Since $T(X/Y)_{\sigma}^{\wedge}$ is a vector bundle and in particular smooth over $Y$, this extends to a map $S' \rightarrow T(X/Y)_{\sigma}^{\wedge}$. This extension projects into $X$ to an extension $S' \rightarrow X$ of the map $b$. This shows formal smoothness. Since $X$ and $Y$ are of finite type, $f$ is smooth. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.1.l} Suppose $G$ is a presentable sheaf of groups on ${\cal X} /Spec (k)$ (which is equal to ${\cal X}$ in this case), and suppose $f:X\rightarrow G$ is a vertical morphism. Then $X$ is smooth over $Spec (k)$. \end{corollary} {\em Proof:} The morphism $G\rightarrow Spec (k)$ is vertical by Theorem \ref{I.u} (7). The composed morphism $X\rightarrow Spec (k)$ is vertical hence smooth by Lemma \ref{I.1.k}. \hfill $\Box$\vspace{.1in} \begin{theorem} \mylabel{I.1.m} If $G$ is a presentable group sheaf on ${\cal X} /Spec (k)$ then it is represented by a smooth separated scheme of finite type over $k$ (in other words it is an algebraic Lie group over $k$). \end{theorem} {\em Proof:} We assume $k={\bf C}$ for this proof. Choose vertical surjections $f:X\rightarrow G$ and $R\rightarrow X\times _GX$. Note that $R\rightarrow G$ is vertical, so $X$ and $R$ are smooth schemes of finite type. By adding some factors of affine spaces we can assume that the components of $X$ and $R$ all have the same dimension. By the previous section, the morphism $df:T(X)\rightarrow f^{\ast}T(G)$ is a morphism of vector sheaves on $X$, hence it is a morphism of vector bundles. It is surjective, so the kernel is a strict sub-vector bundle ${\cal F} \subset T(X)$. For each $x\in X$ we have $$ {\cal F} _x:= \ker (T(X)_x \rightarrow T(G)_{f(x)}). $$ The morphism $p_1: R\rightarrow X$ is vertical (since $X\times _GX\rightarrow X$ is the pullback of the vertical $X\rightarrow G$ by the morphism $X\rightarrow G$, and $p_1$ is the composition of the vertical $R\rightarrow X\times _GX$ with this projection). Therefore, by Lemma \ref{smooth} $p_1$ is smooth. Suppose $r\in R$ maps to $(x,y)\in X\times X$. Let $g\in G$ denote the common image of $x$ and $y$. We have an exact sequence $$ T(R)_r \rightarrow T(X)_x \oplus T(X)_y \rightarrow T(G)_g \rightarrow 0. $$ From this we get that the image of the map on the left always has the same dimension; in particular this shows that the map $T(R)\rightarrow (p_1, p_2)^{\ast}T(X\times X)$ is strict. For any point $g$ in $G$ we can identify $T(G)_g\cong T(G)_1$ by left multiplication. The morphism on the right in the exact sequence then comes from a morphism of the form $p_1^{\ast}(\alpha )-p_2^{\ast}(\alpha )$ where $\alpha : T(X) \rightarrow T(G)_1$ is obtained from the differential of $f$ by the left-multiplication trivialization. This morphism is a morphism of vector bundles from the tangent bundle of $X\times X$ to the constant bundle $T(G)_1$, so its kernel is a distribution in the tangent bundle of $X\times X$. The image of $R$ is an integral leaf of this distribution. In particular, the image of $R$ is a smooth complex submanifold of $X\times X$ (note that the map from $R$ to the leaf is smooth since, by the above exact sequence, the differential is surjective at any point---this implies that the image is open in the leaf). Choose a subvariety $X'\subset X$ which is everywhere transverse to the distribution ${\cal F}$, and which meets every subvariety of $X$ of the form $p_2(p_1^{-1}(x))$ for $p_i$ denoting the projections $R\rightarrow X$. We may assume that $X'$ is of finite type. Let $R'$ be the intersection of $X'\times X'$ with the image of $R$ in $X\times X$. We claim that the morphism $X'\rightarrow G$ is surjective and vertical, and that $R'= X'\times _GX'$. To see this, note that by hypothesis $X'\times _X R\rightarrow X$ is surjective on closed points. By our transversality assumptions this morphism is also smooth. Thus any point in $X$ is equivalent via $R$ (etale-locally) to a point in $X'$. For verticality, it suffices to prove that $X'\times _GX \rightarrow X$ is vertical (Theorem \ref{I.u}, parts 3 and 4). And for this it suffices to note that $X'\times _X R \rightarrow X' \times _GX$ is surjective and vertical (being the pullback of $X\times _XR\rightarrow X\times _GX$ by $X'\times _GX\rightarrow X\times _GX$), that $X'\times _XR\rightarrow X$ is smooth and hence vertical, and to apply Theorem \ref{I.u}, part 4. We get $X'\rightarrow G$ surjective and vertical. If we put $R'' $ equal to the pullback of $R$ to $X'\times X'$ then $R'' \rightarrow X'\times _G X'$ is surjective and vertical (it being also the pullback of $R$ via $X'\times _GX' \rightarrow X\times _GX$). The previous proof applied to this case shows that $R''$ is smooth over its image $R'$, and that $R'$ is a smooth subvariety of $X'\times X'$. But now, by our previous transversality assumptions, the projections $R'\rightarrow X'$ are etale. We can now conclude that $G$, which is the quotient of $X'$ by the equivalence relation $R'$, is a smooth algebraic space. We will find an open subset $U\subset G$ which is a smooth variety over $k$. In order to do this, let $d$ be the maximum number of points in the fibers of $X'\rightarrow G$. The fiber through a point $x$ is equal to $p_2(p_1^{-1}(x))$ where $p_i: R' \rightarrow X'$ here denote the projections. Let $W\subset X$ be the set of points $x$ where the maximum number $d$ of points in the fiber $p_1^{-1}(x)$ is achieved. Since the morphism $p_1: R'\rightarrow X$ is etale, it is easy to see that $W$ is an open subset, and that if we let $R'_W $ denote $p_1^{-1}(W)$ then $R'_W\rightarrow W$ is a finite etale morphism of degree $d$. On the other hand, if $x\in W$ and $y$ is in the fiber through $x$ then $y$ is also in $W$. This means that $p_2(R'_W)\subset W$. The correspondence $$ x\mapsto p_2(p_1^{-1}(x)) $$ gives a morphism $\chi$ from $W$ to the symmetric product $W^{(d)}$ having image in the complement of the singular locus. Then $W\times _{W^{(d)}}W= R'_W$. In particular, the quotient of $W$ by the equivalence relation $R'_W$ is the image of $\chi$. Note that $\chi $ is etale over its image, which is thus a locally closed subscheme of $W^{(d)}$. This shows that the quotient of $W$ by the equivalence relation is a scheme $U$ of finite type. It is also smooth. The morphism $W\rightarrow G$ factors through $U\rightarrow G$. We claim that the morphism $U\rightarrow G$ is an open subfunctor, that is for any $Y\rightarrow G$ the fiber product $U\times _GY$ is an open subset of $Y$. The fiber product is the quotient of $W\times _GY$ by the induced equivalence relation; and the quotient of $X'\times _GY$ by the equivalence relation is equal to $Y$. Choosing local liftings $Y\rightarrow X'$ we find that $X'\times _GY$ is the image of $R'\times _{X'\times X'}(X'\times Y)\rightarrow X'\times Y$, that is it is the pullback of $R'$. In particular it is a subscheme of $X'\times Y$. This subscheme surjects to $Y$ by a vertical morphism, a morphism which is hence smooth. The image of the open subset $W\times _GY$ (which is the intersection of $X'\times _GY$ with $W\times Y$) is therefore an open set in $Y$. This shows that $U\subset G$ is an open subfunctor. We can choose a finite number of elements $g_i \in G(S)$ such that $g_i\cdot U$ cover $G$. For the finiteness use the surjection $X\rightarrow G$ with $X$ of finite type (in particular, quasi-compact). We now apply Grothendieck's theorem about representability which says that if a functor $G$ is a sheaf (in the Zariski topology, which is the case here since Zariski is coarser than etale), and if it is covered by a finite number of open subfunctors $G_i$ which are representable by schemes, then the functor $G$ is representable by a scheme (the union of the schemes $G_i$). In our case the $G_i$ are the $g_i\cdot U$, representable by $U$. Since $U$ is of finite type, the union of a finite number of copies is again of finite type. We obtain that $G$ is a scheme of finite type. Note that $U$ is smooth so $G$ is smooth (alternatively, use that any group scheme is smooth). To complete the proof we just have to show that $G$ is separated. Note first that all connected components of $G$ must have the same dimension, so we can speak of the dimension of $G$ without problem. Let $\Delta \subset G\times G$ denote the diagonal. It is preserved by the diagonal left action of $G(k)$ on $G\times G$ (that is, the action $g(a,b)=(ga, gb)$). The complement $K:=\overline{\Delta}-\Delta$ is a closed subset of $G\times G$, of dimension strictly smaller than the dimension of $G$. But $K$ is invariant under the diagonal left action of $G(k)$, so its image $pr_1(K)\subset G$ is invariant by the left action of $G(k)$. Since $dim (K)< dim (G)$ the image $pr _1(K)$ (which is a constructible subset of dimension $\leq dim (K)$) is not dense in $G$. On the other hand, if $K$ were nonempty then this image, being left invariant, would contain a right translate of $G(k)$ which is Zariski dense. This contradiction implies that $K$ is empty, in other words $G$ is separated. This completes the proof of the theorem. \hfill $\Box$\vspace{.1in} {\em Application:} Suppose $S$ is any base scheme of finite type over $Spec (k)$ now, and suppose $S'\rightarrow S$ is an artinian scheme of finite type. Let $\pi : S' \rightarrow Spec (k)$ denote the structural morphism. If $G$ is a presentable group sheaf over $S$ the pullback $G|_{S'}$ is presentable (Lemma \ref{I.1.h}) and the direct image $\pi _{\ast} (G|_{S'})$ is presentable over $Spec (k)$ (Lemma \ref{I.1.i}). By Theorem \ref{I.1.m}, $\pi _{\ast}(G|_{S'})$ is represented by a group scheme of finite type which we denote $G_{S'}$ over $k$. We have $$ G(S')= G_{S'}(Spec (k)). $$ Furthermore, if $X\rightarrow G$ is a vertical surjection then we obtain a scheme of finite type $X_{S'}= \pi _{\ast}(X|_{S'})$ with a morphism $X_{S'}\rightarrow G_{S'}$. This morphism is smooth. \numero{Local study of presentable group sheaves} In this section we return to the case of general base scheme $S$ (in particular, the hypothesis $S=Spec (k)$ is no longer in effect). First we establish some notations for formal completions. Suppose $G$ is a presentable group sheaf. Let $\widehat{G}$ denote the sheaf which associates to $Y\in {\cal X}$ the set of values in $G(Y)$ which restrict to the identity on $Y^{\rm red}$. More generally, use the same notation $\widehat{{\cal F}}$ whenever ${\cal F}$ is a sheaf with a given section playing the role of the identity section (usually the section in question is understood from the context). \subnumero{Local structure} \begin{lemma} \mylabel{I.1.n} Suppose $G$ is a presentable group over a base $S$. Suppose $Z\rightarrow G$ is a vertical surjection with $Z$ an affine scheme of finite type over $S$. Let $T(Z)_e\rightarrow S$ be the tangent vector scheme at a lift $e$ of the identity section. For any $s\in S$ there is an etale neighborhood $$ e(s)\in W \stackrel{p}{\rightarrow} Z $$ and an etale $S$-morphism $q:U\rightarrow TZ$, such that $q=p$ over the section $e$ (which maps to the zero section of $TZ$). \end{lemma} {\em Proof:} Verticality of $Z\rightarrow G$ means that we can choose a lifting of the multiplication of $G$ to $m: Z\times Z \rightarrow Z$ such that $m(x,e)=x$ and $m(e,y)=y$. Let $Q: Z\rightarrow Z$ be the automorphism $Q(x):= m(x,x)$. It has the effect of multiplication by $2$ on the tangent scheme $TZ$ at the identity section, because $$ \frac{\partial }{\partial x}m(x,x)(e)= \frac{\partial }{\partial x}m(x,e)+ \frac{\partial }{\partial x}m(e,x)(e)= 2 \frac{\partial x}{\partial x} =2. $$ If we embedd $Z\subset {\bf A}^N_S$ as a closed subscheme with the identity section going to the origin-section, then we may extend $Q$ to a morphism $Q': {\bf A}^N_S\rightarrow {\bf A}^N_S$ such that $Q'$ acts by multiplication by two on the tangent space at the origin. Let $\widehat{{\bf A}^N_S}$ denote the formal completion of the affine space along the origin-section. Then $Q'$ induces an automorphism of $\widehat{{\bf A}^N_S}$, and it is well known---and easy to see using power series---that such an automorphism is conjugate to its linear part (since the eigenvalues are different from $1$). We obtain an automorphism $F: \widehat{{\bf A}^N_S}\rightarrow \widehat{{\bf A}^N_S}$ such that $F^{-1}\circ Q'\circ F = 2$. Let $\widehat{Z}\subset \widehat{{\bf A}^N_S}$ be the closed formal subscheme obtained by completing $Z$ at the identity section. Note that $\widehat{Z}$ is preserved by $Q'$. Thus the image $F(\widehat{Z})$ is a formal subscheme which is preserved by multiplication by $2$. It follows that it is a cone, and in particular that the linear parts of the equations defining $F(\widehat{Z})$ vanish on $F(\widehat{Z})$. This means that $F(\widehat{Z})$ is included in its tangent scheme $T(F(\widehat{Z}))$ along the identity section. Translating back by $F$ we obtain an immersion $$ \widehat{Z}\hookrightarrow TZ $$ which is the identity on the tangent space at the identity section. The image is a closed formal subscheme preserved by scalar multiplication. For any artinian scheme $S'$ over $S$, $Z(S')$ is a smooth scheme over $Spec (k)$ and $\widehat{Z(S')}\subset TZ(S')$ is a closed formal subscheme at the origin, with the same Zariski tangent space, and which is formally preserved by scalar multiplication. Therefore $\widehat{Z(S')}\cong \widehat{TZ(S')}$. Now $\widehat{Z}(S')$ is the inverse image of $e\in \widehat{Z(Spec (k))}$ via the map $$ \widehat{Z(S')}\rightarrow \widehat{Z(Spec (k))}. $$ The same is true for the tangent scheme $TZ$. From these properties we get that $\widehat{Z}(S')\rightarrow \widehat{TZ}(S')$ is an isomorphism for any $S'$. As that holds true for all artinian schemes $S'$ over $S$ we get that the morphism $\widehat{Z} \rightarrow \widehat{TZ}$ is an isomorphism. Artin approximation now gives the existence of such an isomorphism (inducing the same map on tangent schemes along the identity section) over an etale neighborhood in $Z$, as required for the lemma. \hfill $\Box$\vspace{.1in} \subnumero{Theory of the connected component} We need to develop a suitable theory of the connected component of a presentable group sheaf $G$. \begin{theorem} \mylabel{I.1.o} If $G$ is a presentable group sheaf over $S$, then there is a unique subsheaf of groups $G^0\subset G$ such that $G^0$ is presentable and such that for any artinian $S$-scheme $S'$, we have $G^0(S')$ equal to the connected component of $G(S')$ (when these are considered as algebraic groups over the ground field of $S'$---cf the application at the end of the section on the situation over $Spec (k)$). \end{theorem} {\em Proof:} We first show existence. Let $Z\rightarrow G$ be a vertical surjection with $Z$ a scheme of finite type. Let $\sigma : S\hookrightarrow Z$ be the identity section. We claim that there is an open neighborhood $U\subset Z$ of $\sigma (S)$ such that for any artinian $S$-scheme $S'$, $U(S')$ is connected. By Lemma \ref{I.1.n}, there is an etale neighborhood of the zero section $W \rightarrow TZ$ and another etale morphism $W\rightarrow Z$ giving an etale neighborhood of the section $\sigma$. We claim that (possibly throwing out a closed subset of $W$ not meeting the section) we can assume that the $W(S')$ are connected. In what follows we refer to the lifting of the zero section of $TZ$ as the section $\sigma$ of $W$. For any given $S'$, artinian located at $s\in S$, there is a surjection of vector spaces $$ (TZ)(S')\rightarrow V_i \subset (TZ)(s), $$ for some subspace $V_i$ which depends on $S'$. If $W\rightarrow TZ$ is our etale morphism, then we have $$ W(S')=W(s)\times _{TZ(s)}(TZ)(S') = W(s)\times _{TZ(s)}V_i \times_{V_i}(TZ)(S'), $$ since a point $S'\rightarrow TZ$ has a unique lifting to $W$ once the lifting is specified on the closed point. Thus $W(S')$ is connected if and only if, for all subspaces $V_i \subset (TZ)(s)$ we have that $W(s)\times _{TZ(s)}V_i$ is connected. Let $Gr (TZ)\rightarrow S$ be the disjoint union of the grassmanian schemes of subspaces of different dimensions. It is proper over $S$. We have a universal subscheme $$ {\cal V} \subset Gr (TZ)\times _S TZ. $$ Note that the map ${\cal V} \rightarrow TZ$ is proper. Let $\tilde{W}:= W\times _{TZ} {\cal V} $; this is an etale covering of ${\cal V}$, and is proper over $W$. Let $\tilde{W}^N\subset \tilde{W}$ be the union of the connected components in fibers which do not pass through the section $\sigma$ (relative to $Gr (TZ)$). Note that $\tilde{W}^N$ is a constructible subset of $\tilde{W}$ (one can see this by noetherian induction). Let $W^N\subset W$ be the image of $\tilde{W}^N$. It is again a constructible subset. A point $w\in W$ is in $W^N$ if and only if there exists a vector subspace $V_i \subset (TZ)(s)$ such that $w$ is in a different connected component of $V_i \times _{TZ}W$ from $\sigma (s)$. In particular, if we choose an analytic neighborhood of the section $\sigma$ which is isomorphic to a tubular neighborhood of the zero-section of $TZ$, then this analytic neighborhood doesn't meet $W^N$. Thus there is a Zariski open neighborhood of $\sigma$ not meeting $W^N$. Since taking a Zariski open subset doesn't affect connectivity (the schemes $W_{S'}$ in question being smooth), we may replace $W$ by this open subset and hence assume that $W^N$ is empty. From the discussion of the previous paragraph, this implies that the $W_{S'}$ are connected, proving the first claim. Let $U$ be the image of $W$ in $Z$. Note that the set-theoretic image is an open set and is equal to the image of the functor, since $W\rightarrow Z$ is etale. Let $\eta : Z\times _SZ \rightarrow Z$ be a lifting of the multiplication map $(g,h)\mapsto gh$ such that $\eta (z, 1)= z$ and $\eta (1,z)=z$. We claim that the composition law $Z\times _SZ \rightarrow G$ is a vertical morphism. Note that $Z\times _SZ\rightarrow G\times _SG$ is vertical, so it suffices to prove that the composition $G\times _SG\rightarrow G$ is vertical. For this, notice that there is an isomorphism $G\times _SG\cong G\times _S G$ sending $(a,b)$ to $(ab,b)$, and which interchanges the multiplication and the first projection. Since the first projection is vertical (this comes from the fact that $G\rightarrow S$ is vertical), we obtain that the composition law is vertical, yielding the claim. By Lemma \ref{I.1.c}, there exists a vertical surjection $$ R\rightarrow (Z\times _SZ )\times _G (Z\times _S Z) $$ with $R$ a scheme of finite type. Let $G^0\subset G$ be the image of the morphism $U\times _SU\rightarrow G$. Then the morphism $U\times _S U\rightarrow G^0$ is a vertical surjection, and we have a vertical surjection $$ R'\rightarrow (U\times _SU )\times _{G^0} (U\times _S U) $$ obtained by letting $R'$ be the inverse image of $(U\times _SU )\times _{G^0} (U\times _S U)$ in $R$. Note that $R'$ is also equal to the fiber product $$ U\times _SU\times _SU\times _SU\times _{Z\times _SZ\times _S Z\times _S Z}R, $$ so $R'$ is a scheme of finite type over $S$. We claim that for any artinian $S'$, the $G^0(S')$ is equal to the connected component of $G(S')$. To see this, note first of all that $G^0(S')$ is connected (since it is the image of $U(S')\times U(S')$ which is connected). And secondly, note that the morphism $$ Z(S')\rightarrow G(S') $$ is an open map (this is a map of smooth varieties---cf the section on what happens over a field and in particular the application at the end). Therefore the image of $U(S')$ is an open subset $V\subset G(S')$. It is connected since $U(S')$ is connected. The image of $(U\times _SU)(S')$ is equal to the image of the multiplication map $V\times V\rightarrow G(S')$. It is easy to see that if $V$ is a connected Zariski open subset of an algebraic group over a field (containing the identity), then the image of the multiplication map is a subgroup. Thus $G^0(S')$ is a subgroup of $G(S')$. It contains an open neighborhood of the identity and it is connected, so it is equal to the connected component. We claim now that $G^0$ is a sheaf of subgroups of $G$. If $g,h\in G^0(S')$ then the product $gh$ restricts into $G^0(S'')$ for any artinian ring $S''$ over $S'$. The sheaf $G^0$ is P2, hence it is B1 and B2 (Theorem \ref{I.t.2}). The inverse image of the section $gh$ by the morphism $G^0\rightarrow G$ is again B1 and B2. This inverse image is nonempty artinian $S''$. By Artin approximation, the inverse image has a section locally over $S'$, and since this section is unique if it exists, it gives a section $gh\in G^0(S')$. We have now shown existence of $G^0$ as required by the theorem. For uniqueness, suppose that $G^1$ were another candidate. Then $G^0$ and $G^1$ are both B1 and B2 subsheaves of $G$ having the same points over artinian $S'$. Artin approximation implies that they are equal. \hfill $\Box$\vspace{.1in} We say that a presentable group sheaf $G$ is {\em connected} if $G= G^0$. The above theorem immediately gives the characterization that $G$ is connected if and only if $G(S')$ is connected for all artinian $S'$. \begin{corollary} \mylabel{connex} We have the following properties. \newline 1. \, If $G$ is connected then any quotient group of $G$ is connected; \newline 2.\, Of $G$ and $H$ are connected then any extension of $G$ by $H$ is connected; \newline 3. \, If $G$ is a connected group sheaf over a base $S$ and if $Y\rightarrow S$ is any morphism of schemes then $G|_{{\cal X} /Y}$ is a connected group sheaf over $Y$; and \newline 4. \, If $f:Y\rightarrow S$ is a finite morphism and if $G$ is a connected group sheaf over $Y$ then $f_{\ast}(G)$ is a connected group sheaf over $S$. \newline 5.\, If $G$ is any presentable group sheaf then the connected component $G^0$ is the largest connected presentable subgroup. \end{corollary} {\em Proof:} Items 1-3 are immediate from the characterization. To prove 4 note that if $S'\rightarrow S$ then $f_{\ast}(G)(S')= G(Y\times _SS')$ and $Y\times _SS'$ is artinian, so this latter group is connected, thus by the above characterization $f_{\ast}(G)$ is connected. To prove 5 note that if $H$ is any connected subgroup of $G$ then $H(S') \subset G^0(S')$ for all artinian $ S'$, hence $H\subset G^0$. \hfill $\Box$\vspace{.1in} \subnumero{Finite presentable group sheaves} We say that a presentable group sheaf $G$ is {\em finite} if $G^0=\{ 1\}$. If $G$ is any presentable group sheaf, then the connected component $G^0$ is a normal subgroup sheaf, and the quotient $C:=G/G^0$ is again presentable. Over artinian $S'$, this quotient is just the group of connected components, in particular the connected component is trivial. Thus $C$ is finite. \begin{lemma} \mylabel{I.1.p} If $G$ is a finite presentable group sheaf, then there is an integer $N$ such that for any henselian local $S$-scheme $S'$ (with algebraically closed residue field), the number of elements in $G(S')$ is less than or equal to $N$. \end{lemma} {\em Proof:} We first treat the case where $S'$ is artinian local with algebraically closed residue field. Let $Z\rightarrow G$ and $R\rightarrow Z\times _GZ$ be the vertical surjections given by the fact that $G$ is $P4$. There is an etale neighborhood $U\rightarrow Z\times _SZ$ of the diagonal such that $U$ is isomorphic to an etale neighborhood of the zero section in the total scheme $TZ$ (and this isomorphism is compatible with the first projection to $Z$). This is seen as in the argument above. Furthermore, as above we may assume that the fibers of the first projection $U\rightarrow Z$ are connected (over any artinian scheme). Then for any artinian scheme $S'\rightarrow U$, the two elements of $G(S')$ obtained from the two projections $U\rightarrow Z$ are the same, by the hypothesis that $G$ is finite. (To see this, compare $(a,b): S'\rightarrow U$ with $(a,a): S\rightarrow U$; they are in the same fiber over $a$, and this fiber is connected, so they have to have the same image in $G(S')$.) Thus, any artinian subscheme of $U$ lifts into $R$. This implies that there is (locally in the etale topology) a lifting $U\rightarrow R$. Let $V\subset Z\times _SZ$ be the image of $U$. It is a Zariski neighborhood of the diagonal, and locally there is a lifting from $V$ into $R$. Let $F\subset Z\times _SZ$ be the reduced closed subscheme corresponding to the closed subset which is the complement of $V$. Suppose $Y\rightarrow S$ is an artinian local scheme (with acrf). If $(\alpha _1,\ldots , \alpha _n)$ is an $n$-tuple of distinct points of $G(Y)$, then there is a lifting $(a_1,\ldots , a_n) \in Z\times _S \ldots \times _S Z(Y)$ such that for any $i,j$ we have that $(a_i, a_j): Y\rightarrow Z\times _SZ$ is not contained in $V$. In particular, the reduced point $(a_i,a_j)^{\rm red}$ is contained in $F$. Thus the reduced point $(a_1,\ldots , a_n)^{\rm red}$ is contained in the closed subscheme $$ F^{(n)}:= \bigcap _{i,j} pr_{ij}^{-1}(F)\subset X\times _S\ldots \times _SX. $$ We claim that there is an $n$ such that $F^{(n)}$ is empty. For any $(x_1,\ldots , x_k)\in F^{(k)}$, let $$ \Phi (x_1,\ldots , x_k):= \{ y\in X, \;\; \pi (y)= \pi (x_i)\in S,\;\; (y,x_1,\ldots , x_k)\in F^{(k+1)}\} . $$ Note that these are closed subschemes of $X$ with strict inclusions $$ \Phi (x_1,\ldots , x_k) \subset \Phi (x_1,\ldots , x_{k-1}). $$ Furthermore, $\Phi (x_1,\ldots , x_k)$ varies algebraically with $(x_1,\ldots , x_k)$. Let $d=dim (X)$ and let $\Lambda = {\bf N} ^d$ with the lexicographic ordering giving the most importance to the $d$th coordinate. For any algebraic set $Y$ of dimension $\leq d$, let $\lambda (Y)= (\lambda _1, \ldots , \lambda _d)$ be defined by setting $\lambda _d$ equal to the number of irreducible components of dimension $d$. Note that if $Y'\subset Y$ is a strict inclusion of a closed subset then $\lambda (Y')< \lambda (Y)$. Let $\Lambda ^{(k)}$ be the finite set of all $\lambda (\Phi (x_1,\ldots , x_k))$ for $(x_1,\ldots , x_k)\in F^{(k)}$ (it is finite because $\Phi (x_1,\ldots , x_k)$ varies algebraically with $(x_1,\ldots , x_k)$). Introduce an order relation on subsets $\Sigma \subset \Lambda$ by saying $$ \Sigma < \Sigma \; \Leftrightarrow \forall \sigma \in \Sigma ,\, \exists \sigma ' \in \Sigma ',\;\; \sigma < \sigma ' . $$ Then the sequence $\Lambda ^{(k)}$ is a sequence of finite subsets which is strictly decreasing for this order relation. We claim that this implies (by combinatorics) that one of the $\Lambda ^{(k)}$ is empty. To see this, assume that the combinatorial claim is true for $d-1$. We will show that the set of upper bounds for $\lambda _d$ on $\Lambda ^{(k)}$ doesn't stabilize. If it were to stabilize after $k_0$ at a certain $y$, then for $k\geq k_0$ we could let $A^k\subset {\bf N} ^{(d-1)}$ be the subset of elements $(a_1,\ldots, a_{d-1})$ such that $(a_1,\ldots , a_{d-1},y)\in \Lambda ^{(k)}$. We obtain a strictly decreasing sequence of subsets for the case of $d-1$, so it is eventually empty, meaning that in fact the upper bound for $\lambda _d$ didn't stabilize. A decreasing sequence which doesn't stabilize can't exist, so eventually there is no upper bound, in other words $\Lambda ^{(k)}$ becomes empty. This gives the claim. Since one of the $\Lambda ^{(k)}$ is empty, one of the $F^{(k)}$ is empty. Let $N$ be chosen so that $F^{(N)}$ is empty (and consequently $F^{(k)}$ is empty for $k\geq N$). Then by the above argument, if $(\alpha _1,\ldots , \alpha _n)$ is an $n$-tuple of distinct points of $G(Y)$, we must have $n<N$. This gives the theorem in the case of an artinian local $Y$. Now suppose $A$ is a henselian local ring and $S'= Spec (A)$. Let $S'_n:=Spec (A/{\bf m}_A^n)$. In the inverse system $\lim _{\leftarrow} G(S'_n)$ we have that all of the $G(S'_n)$ have cardinality bounded by $N$. In particular, the cardinality of the inverse limit is bounded by $N$. Now suppose that there are $N+1$ distinct points $y_i$ in $G(S')$. Two of the points go to the same point in $\lim _{\leftarrow} G(S'_n)$, which means that for two of the points, the liftings $z_i,z_j\in Z(S')$ give a point $(z_i,z_j)$ in $Z\times _SZ$ which lifts, over any $S'_n$, into $R$. By strong artin approximation (check here !!!), the point $(z_i,z_j)$ must lift into $R$ so the two points in $G(S')$ are equal, a contradiction. This completes the proof of the lemma. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{I.1.q} If $G$ is a presentable group sheaf, then $G$ is finite if and only if $G(S')$ is finite for any henselian (resp. artinian) $S'$. \end{corollary} {\em Proof:} The lemma provides one direction. For the other, note that if $G(S)$ is finite for artinian $S'$ then $G^0(S')=\{ 1\}$. By unicity in the characterization of $G^0$ we get $G^0= \{ 1\}$. \hfill $\Box$\vspace{.1in} \numero{Local study of presentable subgroups} In this section we show that if $H\subset G$ is a presentable subgroup of a presentable group $G$ then locally at the identity, in an appropriate sense, $H$ is defined by the vanishing of a section of a vector sheaf. This is a generalisation of the basic result that a subgroup of an algebraic group is smooth, and hence a local complete intersection---cut out by a section of its normal bundle. We obtain this result only in a ``neighborhood of the identity'', or more precisely upon pullback by a vertical morphism $X\rightarrow G$ such that $X$ admits a lift of the identity. If $Y$ is a scheme with morphism $Y\rightarrow G$ such that $P\in Y$ maps to the identity section in $G$, then there will be an etale neighborhood of $P\in Y$ lifting to $X$ (which is why we can think of $X$ as a neighborhood of the identity). This result will be used in a future study of de Rham cohomology (results announced in \cite{kobe}). There, it will be important to have a structure theory for presentable subgroups because of the general principle that if $G$ is a presentable group sheaf then $G/Z(G)\subset Aut ({\cal L} )$ where ${\cal L} = Lie (G)$ is the Lie algebra vector sheaf of $G$ (see \S 9 below). A good understanding of the structure of presentable subgroups will allow us to reduce to looking at de Rham cohomology with coefficients in $Aut ({\cal L} )$ for ${\cal L}$ a vector sheaf, and here we have a more concrete hold on what happens. \begin{theorem} \mylabel{D.1} Suppose $G$ is a connected presentable group sheaf over $S$, and suppose $H\subset G$ is a presentable subgroup sheaf. Suppose that $X_1\rightarrow G$ is a vertical morphism with lift of the identity section $e:S\rightarrow X_1$. Suppose $P\in S$. Then there is an etale neighborhood $X\rightarrow X_1$ of $e(P)$ with a lift of the identity $e: S\rightarrow X$ (possibly after localizing in the etale topology of $S$ here) and an etale morphism $\rho : X\rightarrow TX_e$ sending $e$ to the zero section, such that $$ X\times _GH = \rho ^{-1}(TX_e \times _{TG_e} TH_e). $$ In particular, there is a vector sheaf $V$ over $S$ and a section $\sigma : X\rightarrow V$ such that $X\times _GH= \sigma ^{-1}(0)$. \end{theorem} {\em Proof:} Let $X_1\rightarrow G$ be a surjective vertical morphism with $X_1(S')$ connected for all artinian $S'$ (with $X_1$ a scheme of finite type). Put $Y_1:= X_1\times _GH$. It is a subsheaf of $X_1$. We can choose a vertical surjection $Z_1\rightarrow Y_1$ (with $Z_1$ a scheme of finite type over $S$) together with a lift of $(e,e)$ also denoted by $e$. Note that the morphism $Z_1\rightarrow H$ is also vertical (using the composition property of vertical morphisms and the fact that the morphism $Y_1\rightarrow H$ is vertical by the pullback property). There is an etale neighborhood of $(e,e)\in X_1\times _SX_1$ denoted by $U_1\rightarrow X_1\times _SX_1$ together with a lifting $\psi : U_1\rightarrow X_1$ of the multiplication in $G$, such that $\psi $ restricted to the inverse images of $\{ e\} \times _SX_1$ or $X_1\times _S\{ e\} $ are the identity. We obtain a morphism $$ U_1\times _{X_1\times _SX_1}(Y_1\times _SY_1)\rightarrow Y_1 $$ compatible with the multiplication in $H$ and again having the property that the restrictions to the inverse images of the two ``coordinate axes'' are the identity. Now pull back our multiplication to $$ U_1\times _{X_1\times _SX_1}(Z_1\times _SZ_1) $$ and note that $Z_1\rightarrow Y_1$ being vertical, there is an etale neighborhood of the identity section (all of this is local on $S$!) $$ V_1\rightarrow U_1\times _{X_1\times _SX_1}(Z_1\times _SZ_1) $$ (which we can consider just as an etale neighborhood $V_1\rightarrow Z_1\times _SZ_1$) and a good lift of our multiplication $$ V_1\rightarrow Z_1 $$ restricting to the identity on the inverse images of the ``coordinate axes''. We obtain in this way morphisms on the etale germs $$ {\bf 2}_{Z_1}: (Z_1,e)\rightarrow (Z_1,e) $$ and $$ {\bf 2}_{X_1}: (X_1,e)\rightarrow (X_1,e) $$ compatible with the morphism $Z_1\rightarrow X_1$. These morphisms induce multiplication by $2$ on the tangent vector schemes. There are unique analytic isomorphisms of complex analytic germs $$ (X_1,e)^{\rm an}\cong (T(X_1)_e,0)^{\rm an} $$ and $$ (Z_1,e)^{\rm an}\cong (T(Z_1)_e,0)^{\rm an} $$ transforming the automorphisms ${\bf 2}$ into multiplication by $2$ and inducing the identity on tangent spaces at the identity section. (To see uniqueness, note that over artinian bases these are germs of vector spaces, and any germ of automorphism $f$ of a vector space, such that $f(2x)=2f(x)$, is linear; hence fixing it at the identity fixes it.) By uniqueness, these isomorphisms are compatible with the morphism $Z_1\rightarrow X_1$. On the formal level, we have an etale morphism of formal germs $$ \hat{\varphi}:\widehat{T(X_1)_e}\rightarrow X_1 $$ such that $\widehat{T(Z_1)_e}$ maps into $Y_1$. The {\em first claim} is that, in fact, this gives a map $$ Spec (\widehat{{\cal O}} _{T(X_1)_e,e(S)}) \rightarrow X_1 $$ such that $$ T(Z_1)_e \times _{T(X_1)_e}Spec (\widehat{{\cal O}} _{T(X_1)_e,e(S)}) $$ maps into $Y_1$. We can now apply Artin approximation to find an etale neighborhood $W_1\rightarrow T(X_1)_e$ of the identity section (of course locally on $S$) together with a morphism $W_1\rightarrow X_1$ inducing the identity on tangent vector schemes at the identity section, and sending $$ T(Z_1)_e\times _{T(X_1)_e}W_1\rightarrow Y_1. $$ We can suppose that the morphism $W_1\rightarrow X_1$ is etale. In particular the morphism $W_1\rightarrow G$ is vertical. We obtain two subsheaves $$ im (T(Z_1)_e\times _{T(X_1)_e}W_1\stackrel{pr_2}{\rightarrow} W_1) \subset W_1\times _{X_1} Y_1 \subset W_1. $$ They have the same tangent subsheaves at the identity. Our {\em main claim} is that by taking an open subset of $W_1$ (still a neighborhood of $e(P)$ for a given basepoint $P\in S$) we can assume that these two subsheaves are equal. The first subsheaf is given by the vanishing of the morphism $$ W_1\rightarrow T(X_1)_e /T(Z_1)_e = T(G)/T(H), $$ while the second subsheaf is equal to $W_1\times _GH$. Setting $X=W_1$ we obtain the result of the theorem. We just have to prove the {\em first claim} and the {\em main claim}. {\em Proof of the first claim:} By the sheaf condition and the finite type condition B1 and B2 for $Y_1$, it suffices to prove that for any artinian $S'$, we have $$ T(Z_1)_e \times _{T(X_1)_e}Spec (\widehat{{\cal O}} _{T(X_1)_e,e(S)})(S') $$ mapping into $Y_1(S')$. That is to say, we have to prove that for any point $S'\rightarrow T(Z_1)_e$ mapping to a point of $T(X_1)_e$ located near the origin (that is to say factoring through $Spec (\widehat{{\cal O}} _{T(X_1)_e,e(S)})$), this point maps into $Y_1(S')$. We change to an algebraic notation. We can suppose that $S=Spec(A)$, $T(X_1)_e=Spec(B)$ and $T(Z_1)_e=Spec(C)$. Further we can suppose that $S'=Spec (K)$ with $K$ artinian (although not necessarily of finite type). We have $C\rightarrow K$. Since $T(Z_1)_e$ is a vector scheme we have a map $C\rightarrow C[t]$ corresponding to multiplication by $t$ (and compatible with the same map on $B$). Let $\hat{B}$ denote the completion of $B$ around the zero section (which corresponds to an ideal ${\bf b}\subset B$). We are provided with a factorisation $B\rightarrow \hat{B}\rightarrow K$. We can assume that $K$ is of finite type over $\hat{B}$, and in particular that $K$ is the total fraction ring of a subring $R\subset K$ such that $R$ is finite over $\hat{B}$. Let ${\bf r}\subset R$ denote the ideal corresponding to ${\bf b}\subset B$ (note that $R$ is complete with respect to ${\bf r}$). Let $K\{ t\} \subset K[[ t]]$ denote the set of formal series of the form $\sum a_it^i$ such that there exists $\eta \in R$ such that $\eta a_i \in {\bf r}^i$. With the same notations for $B$, multiplication by $t$ provides a map $\hat{B}\rightarrow B\{ t\}$ compatible with the map $B\rightarrow B[t]$, hence we get a map $\hat{B} \rightarrow K\{ t\}$. On the other hand we get a map $C\rightarrow C[t]\rightarrow K[t]\rightarrow K\{ t\}$. Putting these together we get a map $$ \hat{B}\otimes _BC \rightarrow K\{ t\} $$ corresponding to multiplication by $t$. There is an evaluation at $t=1$ which is a map $K\{ t\} \rightarrow K$ (this summability of the formal series comes from the definition of $K\{ t\}$ and the completeness of $R$), and the above map is compatible with this and with the map $\hat{B}\otimes _BC\rightarrow K$ given at the start. All in all we obtain a map $$ Spec (K\{ t\}) \rightarrow T(Z_1)_e \times _{T(X_1)_e}Spec (\widehat{{\cal O}} _{T(X_1)_e,e(S)}) $$ which induces on the subscheme $Spec (K)\rightarrow Spec (K\{ t\})$ (evaluation at $t=1$) the original inclusion. Now compose with the projection into $G$. We obtain a morphism $$ Spec (K\{ t\} )\rightarrow G $$ which sends $Spec (K[[t ]])$ into $H$ (this comes from the condition that $\widehat{T(Z_1)}$ maps into $Y_1$ together with B1 and B2 for $Y_1$ or $H$) and we would like to show that it sends $Spec (K)$ (at $t=1$) into $H$. It suffices to show that $Spec (K\{ t\} )\rightarrow H$. By Noether normalization there is a morphism $R'\rightarrow R$ such that $R'$ is integral and $R$ is finite over $R'$. Let $K'$ be the total fraction ring of $R'$: it is a field, and $K$ is finite over $K'$. There is an ideal ${\bf r'}\subset R'$ which induces ${\bf r}$, and $K\{ t\}$ is finite over the ring $K'\{ t\}$ defined in the same way as above with respect to this ideal. Let $G'$ and $H'$ denote the direct images to $Spec (K')$ of the groups $G$ and $H$ pulled back to $K$. We have that $H'$ is a presentable subgroup of the presentable group $G'$ (Lemma \ref{I.1.i}), but since $K'$ is a field, $H'\subset G'$ is a closed subgroup of the algebraic group $G'$ over $K'$. Since $K$ is finite over $K'$ we have $$ K\{ t\} = K'\{ t\} \otimes _{K'}K, $$ whence our point $Spec (K\{ t\} )\rightarrow G$ gives a point $Spec (K'\{ t\} )\rightarrow G'$ sending $Spec (K'[[ t]])$ into $H'$. Now since $H'$ is a closed subgroup of $G'$ both of which are algebraic groups (of finite type) over $K'$, we get that $Spec (K' \{ t\} )\rightarrow H'$, meaning that $Spec (K\{ t\} )\rightarrow H$. This completes the proof of the first claim. \hfill $\Box$\vspace{.1in} {\em Proof of the main claim:} Suppose that the main claim is not true. Note that there is a scheme of finite type surjecting to $W_1\times _{X_1}Y_1$. The falsity of the main claim means that the morphism from this scheme to $T(X_1)_e/T(Z_1)_e$ is nonzero on any subset of the form pullback of an open subset of $W_1$ containing $P$. In particular we can find a (possibly nonreduced) curve inside this scheme, such that the section pulls back to something nonzero on the generic (artinian) point, but such that the image of the curve in $W_1$ contains $P$ in its closure. We get an $S$-scheme $S'$ with reduced scheme equal to a curve, and a morphism $\psi :S'\rightarrow W_1\times _GH$ such that the projection into $T(X_1)_e/T(Z_1)_e$ is nontrivial at the generic point of $S'$, such that $P$ is in the closure of the image of $S'$. Let $\overline{S}'$ be a closure of $S'$ relative to $W_1$ obtained by adding one point over $P$. Call this point $P'$. Then for any $n$ there is an etale neighborhood of $P\in W_1$ on which the squaring map $n$-times is defined. We obtain an etale $\overline{S}'_n\rightarrow \overline{S}'$ on which the squaring map $n$-times is defined. We may assume that $\overline{S}'_n$ consists of an etale morphism $S'_n\rightarrow S'$, union one point $P'_n$ over $P'$. Denote by $\psi _n: \overline{S}'_n\rightarrow W_1$ the result of the squaring operation iterated $n$ times. There is an analytic isomorphism of a neighborhood of $P'_n$ in $\overline{S}'_n$ with a neighborhood of $P'$ in $S'$, and an analytic trivialization of a neighborhood of $P$ in $W_1$ (isomorphism with the tangent vector scheme) such that $\psi _n= 2^n\psi$ as analytic germs around the point $P'_n$. {\em Step 1.} There is an $n_0$ such that for any $n\geq n_0$, the projection of $S'_n$ into $T(X_1)_e/T(Z_1)_e$ is nontrivial at the generic point of $S'_n$. In particular for any $m$ the projection of $S'_{mn_0}$ into $T(X_1)_e/T(Z_1)_e$ is nontrivial at the generic point of $S'_{mn_0}$. Let $v: W_1\rightarrow T(X_1)_e/T(Z_1)_e$ denote our section. With respect to our analytic trivialization of $W_1$ where the squaring map becomes multiplication by $2$, can take a Taylor expansion for $v$ around the identity section of $W_1$, $$ v= v_1 + v_2+ v_3 + \ldots + v_{i-1} + w_i, $$ with $v_j(2x)= 2^jv(x)$ and $w_i$ vanishes to order $i$ along $e$; this notion can be defined by considering $w_i$ as a section of a coherent sheaf ${\cal F}$ which contains $T(X_1)_e/T(Z_1)_e$. By hypothesis the restriction of $v$ to $S'$ is nonzero at the generic point of $S'$. Let ${\cal G} _{S'}$ be the quotient of ${\cal F} |_{S'}$ by the ``torsion'' subsheaf (i.e. the subsheaf of sections supported in dimension zero). That a section is nonzero at the generic point means that its projection into ${\cal G} _{S'}$ is nonzero. We may choose $i$ big enough so that $v$ is nonzero in ${\cal G} _{S'}$ modulo the image of sections which vanish to order $i$ along $e$. Let $\overline{v}_j$ denote the projection of $v_j$ into the space of sections of ${\cal G}_{S'}$ modulo the image of the sections vanishing to order $i$. At least one of the $\overline{v}_j$ is nonzero. Now notice that the projection of $v(2^nx)$ is equal to $$ \overline{v(2^nx)} = 2^n\overline{v}_1(x) + 2^{2n}\overline{v}_2(x) + \ldots + 2^{(i-1)n}\overline{v}_{i-1}(x). $$ A little $2$-adic argument shows that there is $n_0$ such that for $n\geq n_0$ this quantity must be nonzero. We obtain that $\overline{v(2^nx)}\neq 0$ and hence that $v(2^nx)=v(\psi_nx)$ is nonzero at the generic point of $S'_n$, as claimed for Step 1. {\em Step 2.} The Zariski closure of the union of the images of the $\psi _{mn_0}$ contains the zero-section. To prove this, note that in the formal completion at $P$, the union of the closures of the $S'_{mn_0}$ is a subset stable under multiplication by $2^{n_0}$, hence its Zariski closure is stable under (fiberwise) multiplication by $2$, hence it is fiberwise homogeneous and thus contains the zero-section. The completion of the Zariski closure contains the Zariski closure of the intersection with the completion, so the zero-section is in the closure. {\em Step 3.} Over the generic point of $S$, the zero section is in the Zariski closure of the $S'_{mn_0}$. Otherwise we would obtain a function nonvanishing on the zero section and vanishing on the $S'_{mn_0}$; clearing denominators this function can be assumed defined over $S$ rather than the generic point of $S$, and since (we may assume) the $S'_{mn_0}$ are all schemes of pure dimension $1$ dominating $S$, this function defined over $S$ which vanishes generically on the $S'_{mn_0}$, must vanish identically on the $S'_{mn_0}$. This would contradict the fact that the zero section is in the Zariski closure globally over $S$. {\em End of proof of claim:} Now we work over the generic geometric artinian point of $S$. Change notations now to suppose that $S$ is artinian and $S'=S$; we note the schemes $S'_{mn_0}$ by $S_{mn_0}$ (they are all isomorphic to $S$) with $S'=:S_1$. We have points $S_{mn_0}\rightarrow W_1\times _GH$ all mapping to something nonzero in $T(X_1)_e/T(Z_1)_e$. Note, as a bit of a detour, that the connected component of the identity in $W_1\times _GH (S)$, must map to zero in $T(X_1)_e/T(Z_1)_e(S)$. This is because $T(X_1)_e/T(Z_1)_e(S)= T(X_1)_e(S)/T(Y_1)_e(S) = T(G)_e(S)/T(H)_e(S)$, whereas verticality of $X_1\rightarrow G$ implies that $X_1(S)\rightarrow G(S)$ is smooth. In particular $W_1\times _GH (S)$ is a smooth local complete intersection so a morphism from $W_1(S)$ to the normal space $T(G)_e(S)/T(H)_e(S)$ of $W_1\times _GH (S)$, with zero set contained in the complete intersection, must have zero set which is a union of connected components of $W_1\times _GH (S)$. Containing the identity, it contains the connected component of the identity. In particular, our points $S_{mn_0}\rightarrow W_1\times _GH$ from before are never in the connected component of $W_1\times _GH (S)$ which contains the identity. On the other hand, these points all lift to $Z_2\rightarrow W_1$ (a scheme of finite type surjecting vertically to $W_1\times _GH$). Let $Z_2(S)'$ denote the union of components of $Z_2(S)$ which contain liftings of our points $S_{mn_0}\rightarrow W_1$. We have a morphism $Z_2(S)'\rightarrow W_1(Spec (k))$ whose image is a constructible set. But the image contains all of the points where the $S_{mn_0}$ are located, so the image must contain a generic point of any irreducible component of the Zariski closure of the $S_{mn_0}$. In particular, there is a component of $Z_2(S)'$ which maps to something in $W_1(Spec (k))$ containing the identity in its closure. Let $W_1(S)_e$ denote the inverse image of $e\in W_1(Spec (k))$ in $W_1(S)$. Let $N\subset G(S)$ denote the image of $W_1(S)_e\rightarrow G(S)$. We claim: that $N$ is a unipotent subgroup of $G(S)$, and that the morphism $W_1(S)_e\rightarrow N$ is a fibration with connected fibers. Assume this claim for the moment. The image of $W_1(S)\rightarrow W_1(Spec (k))$ is a closed subvariety $R\subset W_1(Spec (k))$ (this can be seen since $W_1$ is etale over the vector scheme $TX_1$). We have a morphism $R\rightarrow G(S)/N$. On the other hand, the above morphism $Z_1(S)' \rightarrow W_1(Spec (k))$ factors through a morphism $Z_1(S)'\rightarrow R$, and the image of this map contains $1\in R$ in its closure. The morphism $W_1(S)\rightarrow R$ is a fibration with fiber $W_1(S)_e$ in the etale topology. It suffices to prove that $TX_1 (S) \rightarrow TX_1(Spec (k))$ is a fibration over its image, since locally in the etale topology $W_1$ is isomorphic to $TX_1$. In fact if $V$ is any vector scheme then $V(S)$ and $V(Spec (k))$ are vector spaces so the morphism $V(S)\rightarrow V(Spec (k))$ is a fibration over its image, with fiber the inverse image of the origin. We now show that the morphism $$ W_1\times _GH(S) \rightarrow R \times _{G(S)/N} H(S) $$ is a fibration in the etale topology with fiber the kernel of $W_1(S)_e\rightarrow N$. Locally on $R$ we can choose a lifting $\lambda : R \rightarrow W_1(S)$ and then we have a morphism $$ R\times _{G(S)/N}H(S)\rightarrow N $$ given by $(r,h)\mapsto h^{-1}im(\lambda (r))$. We claim that (locally over $R$) $$ W_1(S)\times _{G(S)}H(S) = W_1(S)_e \times _N (R\times _{G(S)/N}H(S)). $$ The morphism from right to left associates to the point $(a,r,h)$ the point $(i(a)\ast \lambda (r) , h)$ where $i: W_1\rightarrow W_1$ is an etale-locally defined morphism covering the inverse. This shows that the morphism at the start of the paragraph is a fibration. Suppose $A$ is an algebraic group with connected algebraic subgroups $B\subset A$ and $N\subset A$. Then the morphism $$ B / (B\cap N) \rightarrow A/N $$ is proper over an open neighborhood of the class of the identity in $A/N$. To prove this, proceed as follows. Let $I\subset A/N$ denote the image. Let $Z\subset A/N$ denote the subset of points over which the map in question is not proper. This can be constructed as follows. Let $X:= B/(B\cap N)$, and let $\overline{X}$ be a relative completion with proper morphism $\overline{X}\rightarrow A/N$; and suppose that $X\subset \overline{X}$ is open and dense. Then $Z$ is the image of $\overline{X}-X$. Since the map $X\rightarrow A/N$ is injective, we have that the dimension of the image $Z$ is strictly less than the dimension of the image $I$ of $X$. In particular, there is a point $y\in I$ such that the morphism in question is proper over a neighborhood $U$ of $y$. But since $B$ acts on $X$ and compatibly on $A$ (by left multiplication) the morphism in question is proper over any translate of the form $bU$. Setting $b\in B$ equal to the inverse of a representative in for $y$ we obtain a neighborhood $bU$ of the identity over which the map is proper. Note that by the above claim that we are accepting for now, the fiber of the fibration $W_1(S)_e\rightarrow N$ is connected. On the other hand, by the previous paragraph the map $H(S)\rightarrow G(S)/N$ induces a map $H(S)/(H(S)\cap N)\rightarrow G(S)/N$ which is proper over a neighborhood of the class of the identity. Let $I\subset G(S)/N$ be the image of this map. It is a locally closed subset and the subset topology coincides with the topology of the base of the fibration, at least near the identity. Note that $H(S)$ is fibered over $I$ with fibers $H(S)\cap N$ which are connected because $N$ and hence $H(S)\cap N$ are unipotent groups (unipotent groups are always connected). Finally we have the following situation: $$ W_1\times _GH(S)\rightarrow R\times _{G(S)/N} I $$ is a fibration with connected fiber, whereas $I\subset G(S)/N$ is a locally closed subset. Since an etale fibration is an open map, the image of the connected component of the identity in $W_1\times _GH(S)$ is an open neighborhood of the identity in $R\times _{G(S)/N}I$. Since the fibers of the fibration are connected, the image of the complement of the identity component is the complement of the image of the identity component. In particular, there is an open neighborhood of the identity in $R\times _{G(S)/N}I\subset R$ (and hence an open neighborhood of the identity in $R$) whose inverse image doesn't meet any other connected component of $W_1\times _GH(S)$. Finally, since $R$ is a closed subset of $W_1(Spec(k))$, we obtain an open neighborhood of the identity in $W_1(Spec (k))$ whose inverse image in $W_1\times _GH(S)$ is contained in the connected component of the identity. This is a contradiction to our earlier situation where $Z_2(S)' \rightarrow W_1(Spec (k))$ has image a constructible set with the identity in its closure. This completes the proof modulo the following part. We have to show that $N$ is a unipotent subgroup of $G(S)$ and that the morphism $W_1(S)_e\rightarrow N$ is a fibration with connected fibers. Write $S=Spec (A)$ with $A$ artinian, and choose a sequence of ideals $I_j\subset A$ for example $I_j= {\bf m}^j$. Let $$ W_1(S)_j $$ be the set of points of $W_1(S)$ which restrict to the identity on $Spec (A/I_j)$. In particular $W_1(S)_1=W_1(S)_e$. Choose a good lift of the multiplication $$ W_1\times W_1 \rightarrow W_1 $$ in a formal neighborhood of our point $P$. We obtain $$ \ast : W_1(S)_e\times W_1 (S)_e\rightarrow W_1(S)_e. $$ This operation is not a group, however we have the following property. $$ \ast : W_1(S)_j\times W_1 (S)_j\rightarrow W_1(S)_j. $$ Next, note that the fact that $W_1$ is etale over a vector scheme gives another operation which we denote $$ + : W_1(S)_e\times W_1 (S)_e\rightarrow W_1(S)_e $$ which is an abelian group structure. We can write $$ a*b = a+b + F(a,b) $$ where $$ F: W_1(S)_j\times W_1 (S)_j\rightarrow W_1(S)_{j+1}. $$ This is because there is a unique good operation on the set of elements of $W_1(Spec (A/I_{j+1})$ which restrict to the identity in $W_1(Spec (A/I_j)$. (Also note that the $+$-quotient $W_1(S)_j/W_1(S)_{j+1}$ injects into this subset of $W_1(Spec(A/I_{j+1})$). Because of this formula, we can define the quotient $W_1(S)_j/W_1(S)_{j+1}$ with respect to the operation $\ast$ and it is the same as the quotient with respect to $+$. In particular note that the morphism $$ W_1(S)_j\rightarrow W_1(S)_j/W_1(S)_{j+1} $$ is a fibration with fiber a vector space. The morphism $W_1(S)_e\rightarrow G(S)$ is compatible with the operation $\ast$. Let $N_j$ denote the image of $W_1(S)_j$ in $G(S)$ (in particular $N_1=N$). The $N_j$ are constructible sets and subgroups so they are algebraic subgroups of $G(S)$. We obtain a surjective morphism $$ W_1(S)_j /W_1(S)_{j+1} \rightarrow N_j/N_{j+1}, $$ but from the previous formula the operation on the left is a unipotent algebraic group, this shows that $N_j/N_{j+1}$ is a unipotent group, and since extensions of unipotent groups are unipotent (and $N_j=\{ 1\}$ for $j$ large), $N=N_1$ is unipotent. We claim that $W_1(S)_e\rightarrow N$ is smooth. Suppose $R'\subset R$ is an inclusion of artinian schemes over $Spec(k)$. Look at the map $$ W_1(S)_e(R)\rightarrow W_1(S)_e(R')\times _{N(R')}N(R). $$ Suppose that we have a map $S\times R\rightarrow G$ and a lifting over $S\times R'$ to $W_1$, sending $Spec(k)\times R$ to $e$. We would like to find a lifting over $S\times R$ sending $Spec(k)\times R$ to $e$. We can do this whenever $R'$ is a union of $R_i$ and we have commuting retracts from $R'$ to $R_i$, just apply the verticality property to $R\times S$ with retracts to $R_i\times S$ and $R\times Spec( k)$. This proves that the morphism $W_1(S)_e\rightarrow G(S)$ is vertical with respect to $Spec(k)$. It is then immediate that the morphism $W_1(S)_e\rightarrow N$ is surjective (since $N$ is the image of the previous map). Note that $N$ is presentable, so by Lemma \ref{smooth}, the morphism $W_1(S)_e\rightarrow N$ is smooth. Let $K\subset W_1(S)_e$ be the inverse image of $1\in N$. Then for any two points $a,b$ with $f(a)=f(b)$ there is a unique element $k\in K$ such that $b=k\ast a$ (the existence and uniqueness of such an element $k\in W_1(S)_e$ can be seen using the above grading and the expression for $\ast$, and then it is immediate that $k\in K$ from compatibility of $f$ with $\ast$). Any point $n\in N$ has an etale neighborhood $n\in U\stackrel{p}{\rightarrow} N$ with a section $\sigma :U\rightarrow W_1(S)_e$. Then we obtain a morphism $$ K\times U \rightarrow W_1(S)_e\times _N U $$ obtained by sending $(k,u)$ to $(k\ast \sigma (u), p(u))$. This is an isomorphism on the level of points by the above property for $K$, and both sides are smooth, so it is an isomorphism. This proves that $W_1(S)_e\rightarrow N$ is a fibration in the etale topology. Finally, to show that the fibers are connected it suffices to show that $K$ is connected. But since $W_1(S)_e$ and $N$ are vector spaces and the morphism $f$ is a fibration in the etale topology, the associated analytic morphism is a fibration in the usual topology, so the fiber is contractible. \hfill $\Box$\vspace{.1in} \numero{The Lie algebra sheaf} \begin{theorem} \mylabel{lmn} If ${\cal G}$ is a presentable group sheaf, and if we set $Lie ({\cal G}) := T({\cal G} )_1$, then there is a unique bilinear form (Lie bracket) $$ [\cdot , \cdot ] : Lie ({\cal G} )\times Lie ({\cal G} )\rightarrow Lie ({\cal G} ) $$ which, over artinian base schemes, reduces to the usual Lie bracket. \end{theorem} {\em Proof:} A section of $Lie ({\cal G} )$ over $S'= Spec (A)$ is a morphism $Spec (A[\epsilon ])\rightarrow {\cal G}$ sending $Spec (A)$ to the identity section (in our notation here $\epsilon$ denotes an element with $\epsilon ^2=0$). Given two such morphisms which we denote $\alpha$ and $\beta$ we obtain $$ \alpha p_1, \, \beta p_2 : Spec (A[\epsilon , \epsilon '])\rightarrow {\cal G} $$ (where also $(\epsilon ')^2=0$) and we can form the morphism $$ \gamma := \alpha p_1 \cdot \beta p_2 \cdot (\alpha m p_1) \cdot (\beta m p_2) : Spec (A[\epsilon , \epsilon '])\rightarrow {\cal G} $$ where the $g\cdot h$ denotes composition in ${\cal G}$, and where $m= A[\epsilon ]\rightarrow A[\epsilon ]$ is the involution sending $\epsilon$ to $-\epsilon$. The morphism $\gamma$ restricts to the identity on $Spec (A[\epsilon , \epsilon ']/(\epsilon \epsilon '))$. Let $$ q: Spec (A[\epsilon , \epsilon '])\rightarrow Spec (A[\delta]) $$ denote the morphism sending $\delta$ to $\epsilon \epsilon '$ (here again $\delta ^2=0$). Our first claim is that if the morphism $\gamma$ factors as $\gamma = \varphi \circ q$ then $\varphi $ is unique. To see this suppose that $\phi$ and $\varphi$ were two morphisms $Spec (A[\delta])\rightarrow {\cal G}$ with $\phi \circ q = \varphi \circ q$. Let $X\rightarrow {\cal G}$ and $R\rightarrow X\times _{{\cal G}}X$ be the morphisms in a presentation of ${\cal G}$, with a chosen lift of the identity section into $X$. Choose liftings $\tilde{\varphi}$ and $\tilde{\phi}$ from $Spec (A[\delta ])$ into $X$ sending $Spec (A)$ to the identity section of $X$ (here we may have to localize on $S'=Spec (A)$ in the etale topology---but henceforth ignore this point, much as we have already ignored it in lifting the identity section into $X$\ldots ). The fact that the compositions with $q$ are the same means that the pair $(\tilde{\varphi}\circ q, \tilde{\phi}\circ q)$ defines a point which we denote $$ \eta :Spec (A[\epsilon , \epsilon '])\rightarrow X\times _{{\cal G}}X. $$ Note that $Spec (A[\epsilon , \epsilon ']/(\epsilon \epsilon '))$ projects by $q$ to $Spec (A)\subset Spec (A[\delta ])$ and both $\tilde{\varphi}$ and $\tilde{\phi}$ send $Spec (A)$ to the identity section (by hypothesis on our liftings) so in particular $\eta$ sends $Spec (A[\epsilon , \epsilon ']/(\epsilon \epsilon '))$ to the identity pair $(e,e)$ in $X\times _{{\cal G} }X$. On the other hand we can take $Y=Spec (A[\epsilon , \epsilon '])$ and $Y_1=Spec (A[\epsilon ])$ and $Y_2 = Spec (A[\epsilon '])$ and then apply the lifting property $Lift _2(Y, Y_i)$ which holds for the morphism $R\rightarrow X\times _{{\cal G}}X$ because (from the hypothesis in the property $P4$) this morphism is vertical. Fix a lifting $e_R: S\rightarrow R$ of the identity pair in $X\times _{{\cal G}}X$ and fix the values of the morphisms (denoted $\lambda _i$ in the definition of the lifting property) as being $e_R$ on $Y_1$ and $Y_2$. These are indeed liftings of our given morphisms $Y_i \rightarrow X\times _{{\cal G}}X$ since, as we have seen above, both $Y_1$ and $Y_2$ map to the identity pair (the subscheme defined by $(\epsilon \epsilon ')$ is the union of $Y_1$ and $Y_2$). We obtain by the lifting property a lifting $Y\rightarrow R$ which agrees with $e_R$ on $Y_1$ and $Y_2$. If we write (locally) $R=Spec (B)$ then this morphism corresponds to a morphism $a:B\rightarrow A[\epsilon , \epsilon ']$ such that the projection of $B$ modulo $\epsilon$ or modulo $\epsilon '$ is a constant morphism $B\rightarrow A$. It now follows that $a$ factors through $B\rightarrow A[\delta ]$. We obtain a morphism $Spec (A[\delta ])\rightarrow R$ whose projection into $X\times X$ is the pair $(\tilde{\varphi}, \tilde{\phi} )$ (that this is the case is easy to check directly again by supposing that $X$ is affine). This implies that $(\tilde{\varphi}, \tilde{\phi} )$ has image in $X\times _{{\cal G}}X$, in other words that the morphisms $\tilde{\varphi}$ and $\tilde{\phi}$ from $Spec (A[\delta ])$ into $X$ project to the same morphism into ${\cal G}$. Thus $\varphi = \phi$, completing the proof of uniqueness. Now we show existence of the factorization $\gamma = \varphi \circ q$. The preceding uniqueness result implies that it is sufficient to construct $\varphi$ after etale localization on $S'$. Thus we may assume that $\alpha$ and $\beta$ lift to points $\tilde{\alpha}, \tilde{\beta}: Spec (A[\epsilon ])\rightarrow X$ sending $Spec (A)$ to the identity section. There is a good lifting of the multiplication in ${\cal G}$ to a multiplication $X\times X \rightarrow X$ which we still denote $x\cdot y$, where goodness means the property $x\cdot e = e \cdot x = x$. We can now put $$ \tilde{\gamma }:= \tilde{\alpha }p_1 \cdot \tilde{\beta }p_2 \cdot (\tilde{\alpha }m p_1) \cdot (\tilde{\beta }m p_2) : Spec (A[\epsilon , \epsilon '])\rightarrow X. $$ We still have the formula that $$ \alpha \cdot (\alpha m) = e $$ (this is because the first order term of the composition is just addition of vectors) and from this formula it follows that $\tilde{\gamma }$ sends the subschemes $Spec (A[\epsilon ])$ and $Spec (A[\epsilon '])$ to $e$ (through their projections to $Spec (A)$). Since now $X$ is a scheme, this implies directly the existence of $\tilde{\varphi } : Spec (A[\delta ])\rightarrow X$ such that $\tilde{\gamma } = \tilde{\varphi } \circ q$. Projecting from $X$ to ${\cal G}$ we get the factorization $\varphi$ desired. Finally, we set $[\alpha , \beta ] := \varphi$ from the above construction. It is of course completely clear from the construction that if $S'$ is artinian, this gives the usual Lie bracket on the algebraic group ${\cal G} (S' )$. it remains to be seen that this morphism is bilinear and satisfies the Jacobi identity (i.e. that a certain deduced trilinear form vanishes). But these properties can be checked on values over artinian schemes $S'$, and there since the bracket we have defined coincides with the usual one, we get bilinearity and the Jacobi identity. \hfill $\Box$\vspace{.1in} {\em Remark:} The subtlety in our whole situation being essentially that the factorization, while immediate and obviously unique in the case where the target of the map is a scheme, does not necessarily exist and may not be unique even if it does exist, when the target of the map is just a sheaf. One can give examples of $P2$ sheaves ${\cal H}$ on ${\cal X} /S$ together with morphisms $Spec (A[\epsilon , \epsilon '])\rightarrow {\cal H}$ restricting to a given section $S\rightarrow {\cal H}$ over the subscheme defined by $(\epsilon \epsilon ')$, and where the morphism either doesn't factor through $Spec (A[\delta ])$ or else such that the factorization isn't unique. We indicate here a simpler example which shows the way toward the examples refered to in the above paragraph. Let $Y\rightarrow X$ be a degree $2$ morphism of smooth curves completely ramified above a point $x\in X$. Let ${\cal F}$ be the image of this morphism (considered as a sheaf on ${\cal X}$). Let $y$ be the point lying over $x$ and suppose $f: Spec (k[\epsilon]/\epsilon ^3)\rightarrow Y$ is a nonzero tangent vector located at $y$. Then the associated element of ${\cal F}( Spec (k[\tau ]/\tau ^3))$ is constant (equal to the constant point $y$) on the subscheme $$ Spec (k[\tau ]/\tau ^2)\subset Spec (k[\tau ]/\tau ^3). $$ Nevertheless there exists no factorization of the form $$ Spec (k[\tau ]/\tau ^3)\subset Spec (k[\tau ]/\tau ^2)\rightarrow {\cal F} $$ (this factorization would have existed had ${\cal F}$ been a scheme). We can obtain an example where a factorization of the type needed in the above theorem doesn't exist, simply by composing this example with the morphism $Spec (k[\epsilon , \epsilon ']) \rightarrow Spec(k[\tau ]/\tau ^3$ sending $\tau$ to $\epsilon + \epsilon '$. The sheaf ${\cal F}$ in this example is not $P4$ with respect to $Spec (k)$. Of course ${\cal F}$ is not a group sheaf. As stated elsewhere, I am not sure about whether a $P2$ group sheaf might not automatically have to be $P4$ for example (or at least satisfy some of the properties we use here). For example we have seen that an algebraic space (of finite type) which is a group is automatically a scheme. \subnumero{The adjoint representation} Suppose ${\cal G}$ is a presentable group sheaf. Then ${\cal G}$ acts on itself by conjugation, by the formula $$ Int (g)(h):= ghg^{-1}. $$ More precisely this action is a morphism ${\cal G} \times {\cal G} \rightarrow {\cal G}$ and if we put in the identity map on the first projection we obtain a morphism ${\cal G} \times{\cal G} \rightarrow {\cal G} \times {\cal G}$ which is a morphism of group objects (the second variable) over the first variable ${\cal G}$. From this and from the invariance of the above definition of the Lie algebra object, this action induces an action (the {\em adjoint action}) $$ {\cal G} \times Lie ({\cal G} )\rightarrow Lie ({\cal G} ) $$ which preserves the bracket. If $({\cal L} , [,])$ is a Lie algebra sheaf (that is to say, ${\cal L}$ is a vector sheaf with bilinear morphism $[,]: {\cal L} \times {\cal L} \rightarrow {\cal L}$ satisfying the Jacobi identity) then we obtain a group sheaf $Aut ({\cal L} , [,])$. \begin{lemma} \mylabel{AutLie} If $({\cal L} , [,])$ is a Lie algebra sheaf then $Aut ({\cal L} , [,])$ is a presentable group sheaf. \end{lemma} {\em Proof:} The group sheaf $Aut ({\cal L} )$ of automorphisms of the vector sheaf ${\cal L}$ is presentable and in particular $P4$ by Theorem \ref{I.1.g}. The Lie bracket can be considered as a morphism $$ {\cal L} \otimes _{{\cal O}} {\cal L} \rightarrow {\cal L} . $$ The subgroup $Aut ({\cal L} , [,])\subset Aut ({\cal L} )$ may thus be represented as the equalizer of two morphisms $$ Aut ({\cal L} ) \rightarrow Hom ({\cal L} \otimes _{{\cal O}} {\cal L} , {\cal L} ). $$ Note that $Hom ({\cal L} \otimes _{{\cal O}}{\cal L} , {\cal L} )$ is a vector sheaf by Lemma \ref{I.s} and the definition of tensor product following that lemma; and presentable by Theorem \ref{I.1.g}. In particular $Aut ({\cal L} , [,])$ is $P4$ by Lemma \ref{I.1.a} and presentable by Corollary \ref{I.z}. \hfill $\Box$\vspace{.1in} The adjoint action may be interpreted as a morphism of presentable group sheaves $$ Ad : {\cal G} \rightarrow Aut (Lie ({\cal G} ), [,] ). $$ We can of course forget about the bracket and compose this with the morphism into $Aut(Lie ({\cal G} ))$ which is just the automorphism sheaf of a vector sheaf. \begin{proposition} \mylabel{Adjoint} Suppose ${\cal G}$ is a connected presentable group sheaf. Then the kernel of the morphism $Ad$ is the center $Z({\cal G} )$ (that is to say the sheaf whose values are the centers of the values of ${\cal G}$). \end{proposition} {\em Proof:} The statement amounts to saying that a section $g$ of ${\cal G}$ acts trivially on ${\cal G}$ if and only if it acts trivially on $Lie ({\cal G} )$. This statement is true, in fact, of any automorphism (defined over any base scheme $S'\rightarrow S$). It suffices to prove this last statement for the values over artinian base schemes (if an automorphism agrees with the identity on the values over all artinian base schemes then it must be equal to the identity). In the case of values over artinian base schemes it is just the statement that an automorphism which acts trivially on the Lie algebra of a connected algebraic group must act trivially on the whole group. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{centerPres} If ${\cal G}$ is a connected presentable group sheaf, then the center $Z({\cal G} )$ is again presentable. \end{corollary} {\em Proof:} By Proposition \ref{Adjoint} the center is the kernel of a morphism of presentable group sheaves. By Theorem \ref{I.1.e}, this kernel is presentable. \hfill $\Box$\vspace{.1in} {\em Question} Suppose ${\cal G}$ is a presentable group sheaf, not necessarily connected. Is the center $Z({\cal G} )$ presentable? This is related to the following question. {\em Question} Suppose $H$ is a finite presentable group sheaf. Is $Aut (H)$ presentable? A positive response here would allow us to prove that the center $Z({\cal G} )$ is connected, because it is the kernel of the action of $Z({\cal G} ^o)$ on the group of connected components $H={\cal G} /{\cal G} ^o$. \subnumero{Determination of presentable group sheaves by their Lie algebras} The object of this section is to prove the following theorem, which is a generalization of the well known principle that a Lie group is determined by its Lie algebra, up to finite coverings, if the center is unipotent. \begin{lemma} \mylabel{123} Suppose $F, G \subset H$ are two presentable group subsheaves of a presentable group sheaf $H$, and suppose $F$ and $G$ are connected. If $Lie (F)=Lie (G)$ as subsheaves of $Lie (H)$ then $F=G$. \end{lemma} {\em Proof:} By the properties B1 and B2 and artin approximation, it suffices to show that for any artinian $S'$ we have $F(S')=G(S')$. But these two are connected Lie subgroups of $H(S')$ which by hypothesis have the same Lie algebras; thus they are equal. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{abc} Suppose $F$ and $G$ are connected presentable group sheaves on ${\cal X}$. Suppose $Lie (F)\rightarrow Lie(G)$ is an isomorphism of Lie algebras. Then this isomorphism lifts to a unique isomorphism $F/Z(F)\cong G/Z(G)$ where $Z()$ denotes the center. \end{corollary} {\em Proof:} Note that the center of a connected presentable group sheaf is presentable by \ref{centerPres}, so $F/Z(F)$ and $G/Z(G)$ are presentable. Let $L=Lie (F)=Lie (G)$ and let $A=Aut (L)$ (automorphisms of the vector sheaf or of the Lie algebra sheaf, we don't care). We get maps $F\rightarrow A$ and $G\rightarrow A$. Let $F_1$ and $G_1$ denote the images. We have $$ Lie (F_1) = im (L\rightarrow Lie (A)) = Lie (G_1) $$ as subsheaves of $Lie (A)$, so by Lemma \ref{123} we have $F_1=G_1$. On the other hand, note that $Z(F)$ is the kernel of the map $F\rightarrow A$ because if an element of $F$ acts trivially on $Lie (F)$ then by exponentiation and the fact that $F$ is connected, it acts trivially on all $F(S')$ for $S'$ artinian hence in fact it acts trivially on $F$. Thus $F_1 = F/Z(F)$ and similarly $G_1=G/Z(G)$. \hfill $\Box$\vspace{.1in} We have now finished verifying that the class of presentable group sheaves satisfies the properties set out in the introduction. In effect: \newline Property 1 is Corollary \ref{uvw}; \newline Property 2 is Theorem \ref{I.1.e}; \newline Property 3 is Lemma \ref{I.1.h}; \newline Property 4 is Lemma \ref{I.1.i}; \newline Property 5 is Theorem \ref{I.1.m}; \newline Property 6 is Theorem \ref{I.1.o} and Corollary \ref{connex}; \newline Property 7 is Theorem \ref{lmn}; and \newline Property 8 is Theorem \ref{abc}. \subnumero{Questions} We present in further detail some other questions analogous to well known properties of algebraic Lie groups, which seem to be more difficult here. {\bf 1.} \, (Existence) {\em If $({\cal L} , [,])$ is a Lie algebra sheaf (i.e. a vector sheaf with bilinear operation satisfying the Jacobi identity) then does there exist a presentable group sheaf ${\cal G}$ with $Lie ({\cal G} )= ({\cal L} , [,])$?} One has the following idea for a proof of existence in a formal sense. Take a resolution of ${\cal L}$ by vector schemes, and lift the bracket to a bracket (not necessarily satisfying the Jacobi identity) on the vector scheme $X$ surjecting to ${\cal L}$. Then use an explicit version of Baker-Campbell-Hausdorff to define a composition law on the formal completion of $X$ along the zero section. This composition law will not be associative, but one should be able to use the second part of the resolution of ${\cal L}$ to define a relation scheme $R$ (formally), such that when we set ${\cal G}$ to be the quotient of $X$ by $R$ we get a group sheaf. One would have to check that the maps are vertical. Of course this idea for a proof skirts the main question of how to integrate the formal structure out into an actual presentable group sheaf. {\bf 2.} \, {\em Does every (connected, say) presentable group sheaf have a faithful representation on a vector sheaf?} I guess that the answer is probably no, but I don't have a specific example in mind. {\bf 3.} \, {\em Suppose $Lie ({\cal F} )\rightarrow Lie ({\cal G} )$ is a morphism of vector Lie algebras. Under what conditions does this lift to a morphism ${\cal F} '\rightarrow {\cal G}$ where ${\cal F}' \rightarrow {\cal F}$ is a finite covering?} {\bf 4.} \, {\em What happens in Theorem \ref{abc} if we don't divide out by the centers?} {\bf 5.} \, {\em Suppose $G\subset Aut (V)$ is a presentable subgroup of the automorphisms of a vector sheaf. Is there a vector subsheaf $U\subset T^{a,b}(V)$ of a tensor power of $V$ (or possibly a cotensor power or a mixture\ldots ) such that $U$ is preserved by the action of $G$ and such that $G$ is characterized as the subgroup of $Aut (V)$ preserving $U$?} One of the main problems in trying to prove such a statement is that the vector sheaves (and similarly $P4$ or $P5$ sheaves) don't satisfy any nice chain condition. Note that in the situation of question 4, for any sub-vector sheaf $U$ of a tensor and cotensor combination of $V$, the subgroup of $Aut (V)$ of elements preserving $U$ is a presentable subgroup, so at least we obtain a way of constructing examples, even if we don't know whether we get everything this way. \numero{Presentable $n$-stacks} Recall that an $n$-groupoid in the sense of \cite{Tamsamani} is essentially the same thing as an $n$-truncated homotopy type \cite{Tamsamani2}. In view of this, we can approach the theory of $n$-stacks (we assume from here on that this means $n$-stack of $n$-groupoids and drop the word ``groupoid'' from the notation) via the theory of presheaves of topological spaces or equivalently simplicial presheaves\cite{Jardine1}. We adopt a working convention that by {\em $n$-stack} we mean the presheaf of $n$-groupoids associated to a fibrant presheaf of spaces \cite{Jardine1} \cite{kobe} or, a bit more generally, any presheaf of $n$-groupoids such that the associated simplicial presheaf (taking the diagonal of the nerve) is fibrant in the sense of \cite{kobe} which means that it satisfies the global part of the fibrant condition of Jardine \cite{Jardine1}. Some special cases are worth mentioning. A $0$-stack is simply a sheaf of sets. A $1$-stack is what is usually called a stack---it is a sort of sheaf of groupoids. The notions of $2$-stack and $3$-stack were explored heuristically from the category-theoretic point of view in \cite{Breen23}. We suppose given an adequate theory of morphism $n$-stacks $Hom (R,T)$; and of homotopy fiber products $T\times _RT'$ for $n$-stacks. These can be had, for example, within the realm of presheaves of spaces \cite{Jardine1} \cite{kobe} \cite{flexible}. The path-stack $P^{t_1,t_2}T$ on ${\cal X} /S$ between two basepoints (i.e. objects) $t_1,t_2\in T(S)$ is then well defined. We denote by $\pi _0(T)$ the truncation down to a sheaf of sets, and from this and the path space construction we obtain the homotopy group sheaves $\pi _i(T,t)$ over ${\cal X} /S$ for an $n$-stack $T$ and object $t\in T(S)$. In terms of the easier-to-understand version version of the theory involving presheaves of spaces, the homotopy group sheaves are defined as follows. If $t\in T(S)$ then for any $Y\rightarrow S$ we get a basepoint $t|_Y\in T(Y)$. The functor $$ Y/S\mapsto \pi _i (T(Y), t|_Y) $$ is a presheaf on ${\cal X} /S$ which we denote by $\pi _i^{\rm pre}(T,t)$. Then $\pi _i(T,t)$ is sheaf associated to this presheaf. There is probably a good extension of the theory to $\infty$-stacks which would correspond to presheaves of spaces which are not necessarily truncated (and I suppose that it again becomes equivalent to Jardine's theory but there may be a few subtleties hidden here). Generally below when we speak of $n$-stacks, $n$ will be indeterminate. There is probably not too much difference between the theory of $\infty$-stacks and the projective limit of the theories of $n$-stacks, so we will stick to the notation $n$-stack. For $t_1, t_2\in T(S)$ use the notation $\varpi _1(T,t_1,t_2)$ for the sheaf on ${\cal X} /S$ of paths in $T|_{{\cal X} /S}$ from $t_1$ to $t_2$ up to homotopy. Thus $$ \varpi _1(T,t_1,t_2)= \pi _0 (P^{t_1,t_2}T). $$ We make the following definition. \newline ---We say that an $n$-stack $T$ on ${\cal X}$ is {\em presentable} if it satisfies the following conditions: \begin{enumerate} \item The sheaf $\pi _0(T)$ is P1 over $k$. \item For any finite type morphism of schemes $Z\rightarrow Y$ and any two sections $\eta : Y\rightarrow T$ and $\eta ': Z\rightarrow T$ the sheaf $\varpi _1(T|_{{\cal Z} /Z}, \eta |_Z, \eta ' )$, when restricted down from $Z$ to $Y$, is $P4$ over $Y$. \item For any scheme $Y$ and section $\eta : Y\rightarrow T$, the higher homotopy group sheaves $\pi _i( (T|_{{\cal Z} /Y}), \eta )$, for $i\geq 1$, are presentable group sheaves ($P5$) over $Y$. \end{enumerate} (Recall that if ${\cal H}$ is a sheaf on ${\cal X} / Z$ then it can also be considered as a sheaf on ${\cal X}$ with a map to $Z$; the restriction down to $Y$ is the same sheaf taken with the composed map to $Y$, then considered as a sheaf on ${\cal X} /Y$. This shouldn't be confused with the direct image from $Z$ to $Y$. In heuristic topological terms the fiber over $y\in Y$ of the restriction is obtained by taking the direct union of the fibers of ${\cal H}$ over the points $z$ lying over $y$, whereas the fiber of the direct image is obtained by taking the direct product of the fibers of ${\cal H}$ over points $z$ lying over $y$.) {\bf Caution:} This definition of presentability is very slightly different from the definition given in \cite{kobe}. The older version of presentability for $T$ as defined in \cite{kobe} corresponds to the property $P3$ for $\pi _0$ (see Theorem \ref{I.1.q.1kobe} below); whereas the present definition corresponds to the property $P3\frac{1}{2}$ (see Theorem \ref{I.1.q.1} below). I hope that the present version corresponding to $P3\frac{1}{2}$ will be the most useful. The reason for changing the definition was to be able to state Theorem \ref{stability} in a nice way, i.e. to have a reasonable definition of {\em presentable morphism} of $n$-stacks. {\em Caution:} If $T$ is $0$-truncated, that is a sheaf of sets, and happens to have a group structure, then this notion is not the same as the notion that $T$ be a presentable group sheaf. The presentability in $T$ as defined here refers to the higher homotopy groups. In fact, presentability in this case corresponds to the property $P3\frac{1}{2}$ rather than $P4$ (see below). We can also reasonably use the notations {\em presentable homotopy sheaf}; {\em presentable space over ${\cal X}$} or just {\em presentable space}; or {\em presentable fibrant presheaf of spaces}, for the notion of presentable $n$-stack. Property $1$ implies the seemingly stronger statement that there is a section $f: Z\rightarrow T$ over a scheme $Z$ of finite type over $k$, such that the associated morphism $Z\rightarrow \pi _0(T)$ is surjective. The second condition reduces, in the case $\eta = \eta '$, to the statement that for any scheme $Y$ and section $\eta : Y\rightarrow T$, the fundamental group sheaf $\pi _1( (T|_{{\cal Z} /Y}), \eta )$ is a presentable group sheaf over $Y$. We can give an alternative characterization, from which it follows that any truncation $\tau _{\leq n}T$ of a presentable space is again presentable. Recall that we have defined a condition $P3\frac{1}{2}$ which is intermediate between $P2$ and $P4$. \begin{theorem} \mylabel{I.1.q.1} Suppose $T$ is an $n$-stack over $X$. Then $T$ is presentable if and only if the sheaf $\pi _0$ is $P3\frac{1}{2}$, and for any $Y\in {\cal X}$ and $t\in T(Y)$, the sheaves $\pi _i (T|_{{\cal X} /Y}, t)$ are presentable group sheaves ($P5$) over $Y$. \end{theorem} {\em Proof:} Suppose $T$ is presentable. Then we just have to show that $\pi _0$ is $P3\frac{1}{2}$. We know that it is P1, so there is a surjection $Y\rightarrow \pi _0$. By replacing $Y$ by an etale cover, we may assume that this comes from a point $t\in T(Y)$. The path space $P^{p_1^{\ast}t, p_2^{\ast}t}T$ maps to $Y\times Y$, and $$ \varpi _1(T, p_1^{\ast}t, p_2^{\ast}t )=\pi _0(P^{p_1^{\ast}t, p_2^{\ast}t}T)\rightarrow Y\times _{\pi _0} Y $$ is surjective. Let $G\rightarrow Y$ be the sheaf of groups $\pi _1(T|_Y,t)$. It is presentable by hypothesis, and $G$ acts freely on (the restriction from $Y\times Y$ down to $Y$ of) $\varpi _1(T, p_1^{\ast}t, p_2^{\ast}t )$ with quotient $Y\times _{\pi _0}Y$. Finally, we know that (the restriction from $Y\times Y$ down to $Y$ of) $\varpi _1(T, p_1^{\ast}t, p_2^{\ast}t )$ is $P4$ over $Y$; thus the quotient $Y\times _{\pi _0}Y$ is $P4$ over $Y$ by Theorem \ref{I.1.d}. Now by definition there exists a surjective morphism $Q\rightarrow Y\times _{\pi _0}Y$ which is $Y$-vertical. This is what is required to show that $\pi _0$ is $P3\frac{1}{2}$. Now suppose that $\pi _0$ is $P3\frac{1}{2}$ and that the other homotopy group sheaves are presentable. We obtain immediately that $\pi _0$ is P1. Let $X\rightarrow \pi _0$ be the surjection given by the property $P3\frac{1}{2}$. Then we have an $X$-vertical surjection $Q\rightarrow X\times _{\pi _0}X$ (where the morphism to $X$ is the first projection). Suppose $X'\rightarrow X$ is an etale surjection chosen so that the map $X\rightarrow \pi _0$ lifts to $t\in T(X')$. Let $Q'$ be the pullback of $X' \times X'$ to $Q$. Then $Q'= (X'\times _{\pi _0}X')\times _{X\times _{\pi _0}X}Q$ so $Q'\rightarrow X'\times _{\pi _0}X'$ is $X$-vertical, and hence $X'$-vertical. This implies that $X'\times _{\pi _0}X'$ is $P4$ over $X'$, because we can take as the relation scheme $$ Q'\times _{X'\times _{\pi _0}X'}Q'= Q'\times _{X'\times X'}Q' $$ which is already a scheme of finite type (and the identity is vertical). Now we have a sheaf of groups $G= \pi _1(T|_{X'}, t)$ over $X'$ which is by hypothesis presentable, and $G$ acts freely on $\varpi _1(T, p_1^{\ast}t, p_2^{\ast}t )$ with quotient $X'\times _{\pi _0}X'$. By Theorem \ref{I.1.d}, $\varpi _1(T, p_1^{\ast}t, p_2^{\ast}t )$ is $P4$ over $X'$. Now suppose that we have a finite type morphism $q:Z\rightarrow Y$ and two points $\eta _1 \in T(Y)$ and $\eta _2 \in T(Z)$, and we show that the restriction from $Z$ to $Y$ of the path space $\varpi _1(T, \eta _1 |_Z,\eta _2)$ is $P4$ over $Y$. There are etale surjections $ Y'\rightarrow Y$ and $Z'\rightarrow Z$ (of finite type) with $Z'\rightarrow Y'$ and there are morphisms $f_1:Y'\rightarrow X'$ and $f_2: Z'\rightarrow X'$ such that $f_1^{\ast} (t)$ is homotopic to $\eta _1|_{Y'}$ and $f_2^{\ast} (t)$ is homotopic to $\eta _2|_{Z'}$. Let $(f_1|_{Z'},f_2): Z'\rightarrow X'\times X'$ denote the resulting morphism (the first projection of which factors through $Y'$). Then $$ \varpi _1(T,\eta _1|_{Z}, \eta _2)|_{Z'}= \varpi _1(T,\eta _1|_{Z'}, \eta _2|_{Z'})= (f_1|_{Z'},f_2)^{\ast} \varpi _1(T,p_1^{\ast}t, p_2^{\ast}t) $$ $$ = (q, f_2)^{\ast}[\varpi _1(T,p_1^{\ast}t, p_2^{\ast}t)|_{Y'\times X'}]. $$ Note that $\varpi _1(T,p_1^{\ast}t, p_2^{\ast}t)|_{Y'\times X'}$ is $P4$ with respect to $Y'$, so by the appendix to the proof below, one gets that the restriction down to $Y'$ of $\varpi _1(T,\eta _1|_{Z}, \eta _2)|_{Z'}$ is $P4$ with respect to $Y'$. By Corollary \ref{I.1.j.1}, the restriction down to $Y$ of $\varpi _1(T,\eta _1|_{Z}, \eta _2)$ is $P4$ over $Y$. \hfill $\Box$\vspace{.1in} {\em Appendix to the proof:} Suppose $Z\rightarrow Y$ is a finite type morphism, and suppose ${\cal F}$ is a sheaf on $Y$. Then the restriction from $Z$ down to $Y$ of the pullback ${\cal F} |_Z$ is equal to the fiber product $Z\times _Y{\cal F}$. Note also that $Z$ is $P4$ over $Y$. Thus if ${\cal F}$ is $P4$ over $Y$ then the restriction of the pullback is again $P4$. \begin{corollary} \mylabel{truncation} If $T$ is a presentable $n$-stack and if $m<n$ then $\tau _{\leq m}T$ is a presentable $m$-stack. \end{corollary} {\em Proof:} Indeed the truncation operation preserves the homotopy group sheaves (and the homotopy sheaf $\pi _0$). By the theorem, presentability is expressed solely in terms of these sheaves so it is preserved by truncation. \hfill $\Box$\vspace{.1in} We have a similar theorem for the old version of presentability of $T$ \cite {kobe}. \begin{theorem} \mylabel{I.1.q.1kobe} Suppose $T$ is an $n$-stack over $X$. Then $T$ is presentable in the sense of \cite{kobe} if and only if the sheaf $\pi _0$ is $P3$, and for any $Y\in {\cal X}$ and $t\in T(Y)$, the sheaves $\pi _i (T|_{{\cal X} /Y}, t)$ are presentable group sheaves ($P5$) over $Y$. \end{theorem} {\em Proof:} The proof is the same as above only very slightly easier. The details are left to the reader. \hfill $\Box$\vspace{.1in} \subnumero{Very presentable $n$-stacks} We make the following more restrictive definition. Say that a presentable group sheaf $G$ on ${\cal X} /S$ is {\em affine} if, for any artinian $S$-scheme $S'$, the group scheme $G(S')$ over $Spec (k)$ is affine. A truncated homotopy sheaf $T$ is {\em very presentable} if $T$ is presentable and if for any $\eta \in T_Y$ we have that $\pi _1(T/Y,\eta )$ is affine, and $\pi _i(T/Y, \eta )$ are vector sheaves for $i\geq 2$. The idea behind the definition of ``very presentable'' is that we want to require the higher homotopy groups to be unipotent. Note that if we don't require $\pi _1$ to be affine, or $\pi _i$ to be unipotent $(i\geq 2$), then the comparison between algebraic and analytic de Rham cohomology (announced in \cite{kobe}) is no longer true, even over the base $S=Spec (k)$ when all of the groups are representable. This is the reason for making the definition of ``very presentable''. I make the following conjecture: \begin{conjecture} \mylabel{I.1.r} If $G$ is an abelian affine presentable group sheaf on ${\cal X} /S$ such that for any artinian $S'\rightarrow S$ the group scheme $G(S')$ over $k$ is a direct sum of additive groups, then $G$ is a vector sheaf. \end{conjecture} If we knew this conjecture, we could replace the condition of being a vector sheaf by the condition that the $G(S')$ are unipotent (hence additive) for $G=\pi _i$, $i\geq 2$; this would then be along the same lines as the affineness condition for $\pi _1$. As it is, we need to require the condition of $\pi _i$ being vector schemes ($i\geq 2$) for many of the arguments concerning de Rham cohomology sketched in \cite{kobe} to work. {\em Remark:} The categories of presentable and very presentable $n$-stacks are closed under weak equivalences and fiber products but not under cofiber products (push-outs); thus they are not closed model categories. {\em Remark:} We have the same statement as Corollary \ref{truncation} for very presentable stacks (if $T$ is very presentable then $\tau _{\leq m} T$ is very presentable). \subnumero{Other presentability conditions} Recall from \cite{kobe} that we used the notation $P6$ for affine presentable group sheaves and $P7$ for vector sheaves. An $n$-stack $T$ on ${\cal X}$ is {\em $(a_0,\ldots , a_n)$-presentable} (with $a_i \in \{ 0,1, 2 ,3, 3\frac{1}{2}, 4, 5,6, 7\}$) if $\pi _0(T)$ is $Pa_0$ and if for any scheme $Y$ and $t\in T(Y)$, $\pi _i (T, t)$ is $Pa_i$ over $Y$. Here by convention $P0$ means no condition at all. Thus a presentable $n$-stack in our previous notation becomes a $(3\frac{1}{2},5,5, \ldots )$-presentable $n$-stack in this notation. A very presentable $n$-stack is a $( 3\frac{1}{2}, 6, 7, 7, \ldots )$-presentable $n$-stack. The old notions of presentability and very presentability as defined in \cite{kobe} are respectively $(3,5,5,\ldots )$-presentability and $(3,6,7,7, \ldots )$ presentability. There may be some interest in considering, for example, the $(2,2,2,\ldots )$-presentable $n$-stacks, or the $(0,0, 7,7,7,\ldots )$-presentable $n$-stacks. Some other useful versions might be $(4, 5, 5, \ldots )$-presentable $n$-stacks, or $(4, 6, 7, 7, \ldots )$-presentable $n$-stacks for example. Here the condition $P4$ on $\pi _0$ would be with respect to $S=Spec (k)$. For example an algebraic stack with smooth morphisms from the morphism scheme to the object scheme (or even more strongly a Deligne-Mumford stack where these morphisms are etale) would be a $(4, 5)$-presentable stack. The converse is not true since in the condition of $(4,5)$-presentability, the morphism sheaves are not necessarily representable. In fact we will never see the condition of representability of the morphism sheaves in our context, since this is unnatural from the point of view of higher-order stacks (and even in the context of algebraic stacks, one may wonder why the morphism object itself was never allowed to be an algebraic space?). {\em Remark:} Again we have the statement of Corollary \ref{truncation}: if $T$ is an $(a_0,\ldots , a_n)$-presentable $n$-stack then $\tau _{\leq m}T$ is an $(a_0,\ldots , a_m)$-presentable $m$-stack. {\em Remark:} A good convention for using all of these different notions would be to chose some variables $A$, $B$, etc. and set them to be specific $(a_0, a_1, \ldots )$ at the start of a discussion, then to use the notation ``$A$-presentable'' or ``$B$-presentable'' throughout the discussion. \subnumero{A relative version of presentability} We can make a relative definition. In general, say that a morphism $T\rightarrow R$ of $n$-stacks is {\em $(a_0,\ldots , a_n)$-presentable} if for any scheme $Y\in {\cal X}$ and any morphism $Y\rightarrow R$, the fiber $T\times _RY$ is $(a_0,\ldots , a_n)$-presentable. In particular we obtain the notions of presentable and very presentable morphisms by taking $(3\frac{1}{2},5,5, \ldots )$ and $(3\frac{1}{2},6,7,7, \ldots )$ respectively. It is clear that if $T\rightarrow R$ is an $(a_0,\ldots , a_n)$-presentable morphism and if $R'\rightarrow R$ is any morphism of $n$-stacks then the morphism $T':= T\times _RR'\rightarrow R'$ is $(a_0,\ldots , a_n)$-presentable. \begin{lemma} \mylabel{structural} Suppose that $a_0 \leq 5$. An $n$-stack $T$ on ${\cal X}$ is $(a_0,\ldots , a_n)$-presentable if and only if the structural morphism $T\rightarrow \ast$ is $(a_0,\ldots , a_n)$-presentable. \end{lemma} {\em Proof:} Since $\ast$ is itself a scheme of finite type (it is $Spec (k)$) the structural morphism being $(a_0,\ldots , a_n)$-presentable implies that $T$ is $(a_0,\ldots , a_n)$-presentable. For the other implication, suppose $T$ is $(a_0,\ldots , a_n)$-presentable, then for any scheme of finite type $Y$ we have that $T\times Y = T\times _{\ast}Y$ is $(a_0,\ldots , a_n)$-presentable (since a scheme $Y$ is $a_0$-presentable for any $a_0 \leq 5$). \hfill $\Box$\vspace{.1in} {\em Remark:} If ${\cal G}$ is a sheaf of groups on ${\cal X} /S$ then ${\cal G}$ is a presentable group sheaf if and only if $K({\cal G} , 1)\rightarrow S$ is a presentable morphism of $1$-stacks. This is the correct point of view relating our terminologies ``presentable group sheaf'' and ``presentable morphism'' or ``presentable $n$-stack'', i.e. the answer to the terminological problem posed by the caution at the start of this section. \begin{theorem} \mylabel{stability} Suppose $R$ is a presentable (resp. very presentable) $n$-stack. Then a morphism $T\rightarrow R$ is presentable (resp. very presentable) if and only if $T$ itself is presentable (resp. very presentable). \end{theorem} The proof of this theorem will be given in the next subsection below. We first state a few corollaries. \begin{corollary} \mylabel{fiberprod} Suppose $T\rightarrow R$ and $S\rightarrow R$ are morphisms between presentable (resp. very presentable) $n$-stacks. Then the fiber product $T\times _RS$ is presentable (resp. very presentable). \end{corollary} {\em Proof:} From the theorem, the morphism $T \rightarrow R$ is presentable, hence the morphism $T\times _RS$ is presentable and since $S$ is presentable, again from the theorem we conclude that $T\times_RS$ is presentable. The same goes for very presentable. \hfill $\Box$\vspace{.1in} \begin{lemma} \mylabel{basechange} Suppose $R'\rightarrow R$ is a morphism inducing a surjection on $\pi _0$. Then a morphism $T\rightarrow R$ is presentable (resp. very presentable) if and only if the morphism $T':= T\times _RR'\rightarrow R'$ is presentable (resp. very presentable). \end{lemma} {\em Proof:} One direction follows directly from the first remark after the definition above. For the other direction, suppose that $T'\rightarrow R'$ is presentable (resp. very presentable). Then for any scheme $Y\rightarrow R$ there is an etale covering $Y' \rightarrow Y$ and a lifting $Y'\rightarrow R'$, and we have $$ (T\times _RY)\times _YY'=T\times _RY'=T' \times _{R'}Y', $$ which is presentable (resp. very presentable) by hypothesis. The conditions on homotopy sheaves for being presentable (resp. very presentable) are etale-local, so $T\times _RY$ is presentable (resp. very presentable). \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{composition} Suppose $R\rightarrow S$ and $S\rightarrow T$ are presentable (resp. very presentable) morphisms of $n$-stacks. Then the composition $R\rightarrow T$ is a presentable (resp. very presentable) morphism. \end{corollary} {\em Proof:} Suppose $X$ is a scheme of finite type with a morphism $X\rightarrow T$. Then $$ X\times _TR = (X\times _TS) \times _SR. $$ By hypothesis, $(X\times _TS)$ is presentable (resp. very presentable), and by the other hypothesis and the base change property given at the start of the subsection, the morphism $(X\times _TS) \times _SR\rightarrow (X\times _TS)$ is presentable (resp. very presentable). Theorem \ref{stability} now implies that $X\times _TR$ is presentable (resp. very presentable). By definition then, the morphism $R\rightarrow T$ is presentable (resp. very presentable). \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{check} Suppose $f:T\rightarrow R$ is a morphism such that $R$ is presentable (resp. very presentable), and suppose $X\rightarrow R$ is a morphism from a scheme of finite type $X$ which is surjective on $\pi _0$. Then $T$ and the morphism $f$ are presentable (resp. very presentable) if and only if $T\times _RX$ is presentable (resp. very presentable). \end{corollary} {\em Proof:} By Lemma \ref{basechange} the morphism $f$ is presentable if and only if the morphism $p_2: T\times _RX\rightarrow X$ is presentable. On the other hand, $T$ is presentable if and only if $f$ is presentable, from Theorem \ref{stability}. Similarly $T\times _RX$ is presentable if and only if $p_2$ is presentable again by \ref{stability}. This gives the desired statement (the same proof holds for very presentable). \hfill $\Box$\vspace{.1in} We now give some results that will be used in the proof of Theorem \ref{stability}. \begin{lemma} \mylabel{vector?} Suppose $V$ is a vector sheaf and $G$ is a presentable group sheaf on ${\cal X} /S$. If $f: V\rightarrow G$ is a morphism of group sheaves then the kernel of $f$ is a vector sheaf. \end{lemma} {\em Proof:} There is a natural isomorphism of vector sheaves $\varphi : V \cong Lie (V)$, such that $\varphi$ reduces to the exponential on the values over artinian $S'$. To construct $\varphi$ note that a section of $V$ may be interpreted as a map ${\cal O} \rightarrow V$. We have a tautological section of $Lie ({\cal O} )$ so for every section of $V$ the image of this tautological section is a section of $Lie (V)$. This map is an isomorphism on values over artinian schemes, so it is an isomorphism. Let $U \subset Lie (V)$ be the kernel of $$ Lie (f) : Lie (V)\rightarrow Lie (G). $$ Since $Lie (f)$ is a morphism of vector sheaves, its kernel $U$ is a vector sheaf. We claim that $\varphi ^{-1}(U)$ is the kernel of $f$. In order to prove this claim it suffices to prove it for the values over artinian $S'$ (since both are presentable and contained in $V$, and using \ref{Krull}). Here it reduces to the following statement about Lie groups: the kernel of an algebraic morphism from a vector space to a Lie group is the exponential of the kernel of the corresponding morphism of Lie algebras. To prove this notice first that this exponential is a subvector subspace; we can take the quotient and then we are reduced to the case where the map is injective on Lie algebras. The kernel is thus a finite subgroup, but a vector space contains no finite subgroups so we are done. \hfill $\Box$\vspace{.1in} \begin{proposition} \mylabel{I.1.s.3} Suppose $R$, $S$ and $T$ are $n$-stacks over ${\cal X}$, with morphisms $R\rightarrow T$ and $S\rightarrow T$. Suppose $Z\in {\cal X}$ and $(r,s)\in R\times _TS(Z)$. Let $t\in T(Z)$ be the common image of $r$ and $s$. Then we have the following long exact sequence of homotopy group sheaves on ${\cal X} /Z$: $$ \ldots \rightarrow \pi _i (R\times _TS|_{{\cal X} /Z},(r,s))\rightarrow \pi _i(R|_{{\cal X} /Z},r)\times \pi _i (S|_{{\cal X} /Z})\rightarrow $$ $$ \pi _i (T|_{{\cal X} /Z},t)\rightarrow \pi _{i-1}(R\times _TS|_{{\cal X} /Z},(r,s)) \rightarrow \ldots , $$ terminating with the sequence $$ \pi _2(R|_{{\cal X} /Z},r)\times \pi _2(S|_{{\cal X} /Z},s)\rightarrow \pi _2(T|_{{\cal X} /Z},t)\rightarrow \pi _1(R\times _TS|_{{\cal X} /Z}, (r,s))\rightarrow $$ $$ \pi _1(R|_{{\cal X} /Z},r)\times \pi _1(S|_{{\cal X} /Z},s) \stackrel{\displaystyle \rightarrow }{\rightarrow } \pi _1(T|_{{\cal X} /Z},t) $$ (the last part meaning that the image is equal to the equalizer of the two arrows). Furthermore, we have a similar sequence for the path spaces. Suppose $(r_1,s_1)$ and $(r_2,s_2)$ are two points, with images $t_1$ and $t_2$. We have the exact sequence $$ \pi _2(R|_{{\cal X} /Z},r_1)\times \pi _2(S|_{{\cal X} /Z},s_1)\rightarrow \pi _2(T|_{{\cal X} /Z},t_1)\stackrel{acts\; on}{\rightarrow} \varpi _1(R\times _TS|_{{\cal X} /Z}, (r_1,s_1),(r_2,s_2)) $$ $$ \mbox{with quotient the equalizer of} \varpi _1(R|_{{\cal X} /Z},r_1,r_2)\times \varpi _1(S|_{{\cal X} /Z},s_1,s_2) \stackrel{\displaystyle \rightarrow }{\rightarrow } \varpi _1(T|_{{\cal X} /Z},t_1,t_2). $$ \end{proposition} {\em Proof:} We show that we have similar exact sequences at the homotopy presheaf level; then the sequences for the homotopy sheaves follow by sheafification. To define the exact sequences at the presheaf level, we can work object by object, so it suffices to give functorial exact sequences for fibrations of topological spaces $R\rightarrow T$ and $S\rightarrow T$ with basepoints $(r,s)$ mapping to $t$. The morphisms are defined as follows. The morphism from $\pi _i (R\times _TS,(r,s))$ to $\pi _i(R,r)\times \pi _i (S,s)= \pi _i (R\times S, (r,s))$ comes from the inclusion $R\times _TS\rightarrow R\times S$. The morphism from the product to $\pi _i (T,t)$ is the difference of the two projection maps. The morphism from $\pi _i (T,t)$ to $\pi _{i-1}(R\times _TS,(r,s))$ is obtained as a composition $$ \pi _i (T,t)\stackrel{\delta}{\rightarrow} \pi _{i-1} (R_t,r) \stackrel{(1, 0_s)}{\rightarrow }\pi _{i-1}(R_t\times S_t,(r,s)) \stackrel{i}{\rightarrow}\pi _{i-1}(R\times _TS,(r,s)) $$ where $\delta$ is the connecting homomorphism for the fibration $R\rightarrow T$, $0_s$ is the constant class at the basepoint $s$, and $i$ is the inclusion of the fiber $i: R_t\times S_t\rightarrow R\times _TS$. If we took $(1,1)$ instead of $(1,0_s)$ we would get the connecting morphism for the fibration $R\times _TS\rightarrow T$, which goes to zero in the homotopy of the total space $R\times _TS$. Thus, our map is the same as the map which would be obtained by putting in $-(0_r,1)$ instead. From the equality of these two maps, one obtains that the composition of this map with the difference of projections, is equal to zero. That the other compositions are zero is easy to see. Exactness follows by making a diagram with this sequence on one horizontal row, with the sequence $$ \pi _i(R_t\times S_t, (r,s))= \pi _i(R_t,r)\times \pi _i (S_t,s)\rightarrow 0\rightarrow \ldots $$ on the row above, and the sequence $$ \pi _i (T,t)\rightarrow \pi _i (T,t)\times \pi _i (T,t)\rightarrow \pi _i (T,t) \stackrel{0}{\rightarrow} \pi _{i-1}(T,t) $$ on the row below. The vertical rows then have the exact fibration sequences going downwards. One obtains the exactness of the sequence of homotopy groups in question (this works at the end by using the extension of the homotopy sequence for a fibration, to the action of $\pi _1$ of the base on $\pi _0$ of the fiber, with the $\pi _1$ of the total space being the stabilizer of the component of $\pi _0$ of the fiber containing the basepoint. Finally, we treat the case of the path spaces. What is written on the left means, more precisely, that the cokernel of the first map acts freely on the middle sheaf, with quotient equal to the equalizer. The action in question is by the map to $\pi _1(R\times _TS, (r_1,s_1))$ which itself acts on the path space. Now if $\varpi _1(R\times _TS, (r_1,s_1), (r_2,s_2))$ is empty then the equalizer in question is also empty (any element of the equalizer can be realized as a pair of paths mapping to exactly the same path in $T$, giving a path in the fiber product). Note that we count an action on the empty set as free. So we may assume that $\varpi _1(R\times _TS, (r_1,s_1), (r_2,s_2))$ is nonempty, and choose an element. This choice gives compatible choices in all the other path spaces, so composing with the inverse of this path we reduce to the exact sequence for fundamental groups. \hfill $\Box$\vspace{.1in} {\em Remark:} We can extend this sequence to a statement involving $\pi _0$, specially in the case of a fibration sequence. This will be done as we need it below. \begin{lemma} \mylabel{kernel} Suppose $S$ is a base scheme and suppose ${\cal F}$ is a sheaf on ${\cal X} /S$ whose restriction down to ${\cal X}$ is $P3\frac{1}{2}$. Suppose that ${\cal G}$ is a $P4$ sheaf on ${\cal X} /S$ with morphism ${\cal G} \rightarrow {\cal F}$, and finally suppose that $\eta : S\rightarrow {\cal F}$ is a section. Then the inverse image ${\cal H} \subset {\cal G}$ of the section $\eta$ is a $P4$ sheaf. \end{lemma} {\em Proof:} Let $X\rightarrow {\cal G}$ and $W\rightarrow X\times _{{\cal G}}X$ be the $S$-vertical surjections for ${\cal G}$. Fix a surjection $Z\rightarrow {\cal F}$ and a surjection $W\rightarrow Z\times _{{\cal F}}Z$ which is vertical with respect to the first projection to $Z$. Fix a lifting $\eta '$ of the section to $Z$ (note that we are allowed to etale-localize on the base $S$). Let $U:= S\times _{Z}W$ where the morphism in the fiber product is the first projection from $W$ to $Z$ (note that $U$ is a scheme of finite type over $S$). The surjective morphism $$ U\rightarrow S\times _{Z}(Z\times _{{\cal F}}Z) = S\times _{{\cal F}} Z $$ is $S$-vertical since the morphism $W\rightarrow Z\times _{{\cal F}}Z$ was $Z$-vertical. We can choose a lifting $X\rightarrow Z$ of the morphism ${\cal G} \rightarrow {\cal F}$. Then $$ S\times _{{\cal F}} X= (S\times _{{\cal F}}Z)\times _ZX $$ so there is an $S$-vertical morphism $$ U\times _Z X \rightarrow S\times _{{\cal F}} X. $$ On the other hand the $S$-vertical morphism $X\rightarrow {\cal G}$ gives an $S$-vertical morphism $$ S\times _{{\cal F}} X\rightarrow S\times _{{\cal F}} {\cal G} = {\cal H} . $$ Note that $Y:= U\times _Z X$ is a scheme of finite type with a surjective vertical morphism to ${\cal H}$. Since ${\cal G}$ is $P4$ there exists a scheme of finite type $V$ and an $S$-vertical morphism $$ V\rightarrow Y\times _{{\cal G}}Y = Y\times _{{\cal H}} Y. $$ This gives the condition $P4$ for ${\cal H}$. \hfill $\Box$\vspace{.1in} The following lemma is a consequence of Corollary \ref{fiberprod}, but we need it in the proof of Theorem \ref{stability}. \begin{lemma} \mylabel{I.1.s.?} If $R$ and $S$ are presentable (resp. very presentable) $n$-stacks over ${\cal X}$ and $X$ a scheme of finite type, with morphisms $R\rightarrow S$ and $X\rightarrow S$, then the homotopy fiber product $X\times _SR$ is presentable (resp. very presentable). \end{lemma} {\em Proof:} Suppose $f:Y\rightarrow X\times _SR$ is a morphism. Let $r: Y\rightarrow R$ and $s: Y\rightarrow S$ be the composed morphisms. Then (since $X$ is zero-truncated) for $i\geq 1$ we have $$ \pi _i(X\times _SR |_{{\cal X} /Y}, f)= \pi _i (Y\times _SR/Y, r). $$ The latter fits into a homotopy exact sequence, which we can therefore write $$ \ldots \pi _{i+1}(S|_{{\cal X} /Y}, s)\rightarrow \pi _i(X\times _SR |_{{\cal X} /Y}, f)\rightarrow \pi _i(R|_{{\cal X} /Y}, r)\rightarrow \ldots . $$ In the presentable case we obtain immediately from Theorem \ref{I.1.e} that $\pi _i(X\times _SR |_{{\cal X} /Y}, f)$ is a presentable group sheaf over $Y$. In the very presentable case, for $i\geq 3$ we obtain immediately (from Corollary \ref{I.j} and Theorem \ref{I.k}) that $\pi _i(X\times _SR |_{{\cal X} /Y}, f)$ is a vector sheaf. For $i=2$ we obtain the same conclusion but must also use Lemma \ref{vector?}. For $i=1$ we obtain that $\pi _1(X\times _SR |_{{\cal X} /Y}, f)$ is $P5$. In fact it is an extension of the kernel of a morphism of $P6$ group sheaves, by a vector sheaf. Therefore it is also affine (since kernels and extensions by vector sheaves at least, preserve the affineness property). Thus it is $P6$. We just have to prove (in both the presentable and very presentable case) that $\pi _0(X\times _SR)$ is $P3\frac{1}{2}$. Let $a: X\rightarrow S$ denote the given morphism. Recall that $\pi _0(X\times _SR)/X$ denotes this sheaf considered as a sheaf on ${\cal X} /X$. We have an action of $\pi _0(S|_{{\cal X} /X}, a)$ (which is a $P5$ group sheaf over $X$) on $\pi _0(X\times _SR)/X$, and the quotient is the fiber product $X\times _{\pi _0(S)}\pi _0(R)/X$ (i.e. again considered as a sheaf over ${\cal X} /X$). This is the same thing as the inverse image of the given section $a$ via the map $\pi _0(R|_{{\cal X} /X})\rightarrow \pi _0(S|_{{\cal X} /X})$. By Corollary \ref{P3c} or \ref{P3d} the quotient by the action is $P3\frac{1}{2}$. Finally by Proposition \ref{P3e}, the sheaf $\pi _0(X\times _SR)$ is $P3\frac{1}{2}$. \hfill $\Box$\vspace{.1in} {\em Remark:} A similar technique allows one to directly prove Corollary \ref{fiberprod}, that if $R$, $S$ and $T$ are presentable (resp. very presentable) $n$-stacks with morphisms $R\rightarrow S$ and $T\rightarrow S$ then the fiber product $R\times _ST$ is presentable (resp. very presentable). This is left to the reader. Our technique is to use only the above special case to get Theorem \ref{stability}, and then to deduce Corollary \ref{fiberprod} as a consequence. \subnumero{The proof of Theorem \ref{stability}} Lemma \ref{I.1.s.?} immediately implies one direction in Theorem \ref{stability}, namely that if $R$ and $S$ are presentable then the morphism $f$ is presentable. We have to show the other direction: suppose $S$ is a presentable $n$-stack, $R$ is an $n$-stack, and $f:R\rightarrow S$ is a presentable morphism. Choose a scheme of finite type $X$ with a morphism $X\rightarrow S$ inducing a surjection on $\pi _0$. We will show that if $X\times _SR$ is presentable then $R$ is presentable. First of all the morphism $\pi _0(X\times _SR)\rightarrow \pi _0(R)$ is surjective so if $\pi _0(X\times _SR)$ is $P1$ then so is $\pi _0(R)$. For the higher homotopy groups, suppose that $s:Z\rightarrow R$ is a morphism. Lift the projection into $S$ (denoted by $s$) to a morphism $Z\rightarrow X$. This gives a point $f: Z\rightarrow X\times _SR$ and by composition $f_Z: Z\rightarrow Z\times _SR= Z\times _X(X\times _SR)$. Then we have the exact sequence $$ \ldots \rightarrow \pi _i(Z\times _SR |_{{\cal X} /Z}, f_Z) \rightarrow \pi _i(R|_{{\cal X} /Z}, r)\rightarrow \pi _i(S|_{{\cal X} /Z}, s)\rightarrow \ldots . $$ But since $Z$ and $X|_{{\cal X} /Z}$ are zero-truncated, and we have that $Z\times _SR = Z\times _X(X\times _SR)$, the higher homotopy groups $\pi _i(Z\times _SR |_{{\cal X} /Z}, f_Z)$ are the same as the $\pi _i(X\times _SR |_{{\cal X} /Z}, f)$. Thus we can write the exact sequence as $$ \ldots \rightarrow \pi _i(X\times _SR |_{{\cal X} /Z}, f) \rightarrow \pi _i(R|_{{\cal X} /Z}, r)\rightarrow \pi _i(S|_{{\cal X} /Z}, s)\rightarrow \ldots . $$ Note that (in the very presentable case) the kernel of the morphism $$ \pi _2(S|_{{\cal X} /Z}, s)\rightarrow \pi _1(X\times _SR |_{{\cal X} /Z}, f) $$ is a vector sheaf by Lemma \ref{vector?}. In the other cases the kernel (and the cokernel on the other end) are automatically vector sheaves by Corollary \ref{I.j}. Since the property of being a vector sheaf is preserved under extension we get the condition that the $\pi _i(R|_{{\cal X} /Z}, r)$ are vector sheaves ($i\geq 2$). In the presentable case the exact sequence immediately gives the property $P5$ for the $\pi _i(R|_{{\cal X} /Z}, r)$ for ($i\geq 2$). We have to treat the case of $\varpi _1$. Suppose $Z\rightarrow Y$ is a morphism of finite type and suppose $r, r': Z\rightarrow R$ are points such that $r$ factors through $Y$. Let $s,s'$ denote the images in $S$ and assume that they lift to points $f, f'$ and $f_Z, f'_Z$ as above (with $f$ or $f_Z$ factoring through $Y$). We first study everything on the level of sheaves on ${\cal X} /Z$. Note first that $$ Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z}) \rightarrow R|_{{\cal X} /Z}\rightarrow S|_{{\cal X} /Z} $$ is a fibration sequence (this should actually have been pointed out above in the treatment of the $\pi _i$, $i\geq 2$), over the basepoint $s\in S(Z)$. On the other hand note that $r': Z\rightarrow R$ is a point lying over $s'$. Consider the map $$ \varpi _1(S|_{{\cal X} /Z}, s, s')\rightarrow \pi _0(Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z})) $$ which sends a path to the point obtained by transporting $f'_Z$ along the path from $s'$ back to $s$. The fibration sequence gives the following statement: {\em The group $\pi _1(Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z}, f_Z)$ acts on $\varpi _1(R|_{{\cal X} /Z}, r, r')$ with quotient the inverse image in $\varpi _1(S|_{{\cal X} /Z}, s, s')$ of the section $f_Z: Z \rightarrow \pi _0( Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z})$. } Now we note that $$ \pi _1(Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z}, f_Z) = \pi _1(X\times _SR|_{{\cal X} /Z}, f), $$ and $$ \pi _0(Z \times _{S|_{{\cal X} /Z}}(R|_{{\cal X} /Z})\subset \pi _0(X\times _S R|_{{\cal X} /Z}). $$ The transport of $f'$ along the path from $s'$ to $s$ again gives a map $$ \varpi _1(S|_{{\cal X} /Z}, s, s')\rightarrow \pi _0(X\times _SR|_{{\cal X} /Z}) $$ and we obtain the following statement. {\em The group $\pi _1(X\times _SR|_{{\cal X} /Z}, f)$ acts on $\varpi _1(R|_{{\cal X} /Z}, r, r')$ with quotient the inverse image in \linebreak $\varpi _1(S|_{{\cal X} /Z}, s, s')$ of the section $f: Z \rightarrow \pi _0( X\times _SR|_{{\cal X} /Z})$. } Now we look at everything in terms of sheaves on ${\cal X} /Y$. Let $Res _{Z/Y}$ denote the restriction from $Z$ down to $Y$, and let $\tilde{f}$ denote the $Y$-valued point corresponding to $f$. Note that $$ Res _{Z/Y} \pi _0(X\times _SR|_{{\cal X} /Z}) = \pi _0(X\times _SR|_{{\cal X} /Y})\times _YZ. $$ In general if ${\cal A}$ is a sheaf over $Z$ and ${\cal B}$ a sheaf over $Y$ with a section $Y\rightarrow {\cal B}$ then $$ Res _{Z/Y}({\cal A} \times _{{\cal B} |_{{\cal X} /Z}}Z) = (Res _{Z/Y}{\cal A} )\times _{{\cal B}}Y. $$ In particular the inverse image in $\varpi _1(S|_{{\cal X} /Z}, s, s')$ of the section $f: Z \rightarrow \pi _0( X\times _SR|_{{\cal X} /Z})$ restricts down to $Y$ to the inverse image in $Res _{Z/Y}\varpi _1(S|_{{\cal X} /Z}, s, s')$ of the section $\tilde{f}: Y \rightarrow \pi _0(X\times _SR|_{{\cal X} /Y})$. Another general principal is that if ${\cal G}$ is a group sheaf on $Y$ such that ${\cal G} |_{{\cal X} /Z}$ acts on a sheaf ${\cal H}$ then ${\cal G}$ acts on $Res _{Z/Y}{\cal H}$ with quotient equal to $Res _{Z/Y}({\cal H} /({\cal G} |_{{\cal X} /Z}))$. With these things in mind, our above statement becomes: {\em The group $\pi _1((X\times _SR |_{{\cal X} /Y}, \tilde{f})$ acts on $Res _{Z/Y}\varpi _1(R|_{{\cal X} /Z}, r, r')$ with quotient the inverse image in $Res _{Z/Y}\varpi _1(S|_{{\cal X} /Z}, s, s')$ of the section $\tilde{f}: Y \rightarrow \pi _0(X\times _SR|_{{\cal X} /Y})$. } Now the facts that $\pi _0(X\times _SR|_{{\cal X} /Y})$ is $P3\frac{1}{2}$ and that $Res _{Z/Y}\varpi _1(S|_{{\cal X} /Z}, s, s')$ is $P4$ (which comes by hypothesis) imply that the inverse image in question is $P4$ (Lemma \ref{kernel}); then the theorem on quotients (Theorem \ref{I.1.d}) and the fact that the group $\pi _1((X\times _SR |_{{\cal X} /Y}, \tilde{f})$ is $P5$ over $Y$ gives the condition that $Res _{Z/Y}\varpi _1(R|_{{\cal X} /Z}, r, r')$ is $P4$ over $Y$. This is the condition on $\varpi _1$ needed to insure that $R$ is presentable. This completes the proof of Theorem \ref{stability}. \hfill $\Box$\vspace{.1in} We have the following characterization of presentable morphisms via the relative homotopy group sheaves. \begin{proposition} \mylabel{characterization} Suppose $f: R\rightarrow S$ is a morphism of $n$-stacks. Then $f$ is presentable (resp. very presentable) if and only if the following conditions are satisfied for any scheme $X$ of finite type: \newline ---for any morphism $X\rightarrow S$, the sheaf $\pi _0(X\times _SR)$ is $P3\frac{1}{2}$; and \newline ---for any morphism $r: X\rightarrow R$ the sheaves $\pi _i(X\times _SR/X, r)$ on ${\cal X} /X$ are presentable group sheaves over $X$ (resp. $\pi _1$ is affine presentable and $\pi _i$ are vector sheaves for $i\geq 2$). \end{proposition} {\em Proof:} This falls out of the proof of \ref{stability}. \hfill $\Box$\vspace{.1in} {\em Exercise:} For which values of $(a_0,a_1,\ldots )$ does Theorem \ref{I.1.s.?} hold for $(a_0,a_1,\ldots )$-presentable spaces? Place these conditions in Corollary \ref{I.1.u} below. \subnumero{Going to the base of a fibration} It is an interesting question to ask, if $R\rightarrow S$ is a morphism of $n$-stacks such that $R$ is presentable and such that for every scheme-valued point $X\rightarrow S$ the fiber product $X\times _SR$ is presentable, then is $S$ presentable? The answer is surely no in this generality. We need to make additional hypotheses. Directly from the fibration exact sequences, one can see that if $\pi _0(S)$ is assumed to be $P3\frac{1}{2}$ (a hypothesis which seems unavoidable) and if we suppose that for any point $a:X\rightarrow S$, the action of $\pi _1(S|_{{\cal X} /X}, a)$ on $\pi _0(X\times _SR)$ factors through a presentable group sheaf over $X$, then $S$ will be presentable. As a particular case, if the morphism $R\rightarrow S$ is relatively $0$-connected (i.e. the fibers are connected) and surjective on $\pi _0$, then presentability of $R$ implies presentability of $S$. One might look for other weaker conditions, for example that the fibers satisfy some sort of artinian condition (e.g. there is a surjection from a scheme finite over $X$, to $\pi _0(X\times _SR)$). I don't know if this can be made to work. \subnumero{Presentable shapes} We have a notion of internal $Hom$ for $n$-stacks. In the topological presheaf interpretation (\cite{kobe} \S 2), recall that $\underline{Hom}(R,T)$ is defined to be the presheaf $X\mapsto Hom (R'_X,T|_{{\cal X} /X})$ where $R'_X$ is a functorial replacement of $R|_{{\cal X} /X}$ by a cofibrant presheaf. \begin{corollary} \label{I.1.u} Suppose $W$ is a finite CW complex, and let $W_{{\cal X}}$ denote the constant $n$-stack with values $\Pi _n(W)$ (or in terms of presheaves of spaces, it is the fibrant presheaf associated to the constant presheaf with values $\tau _{\leq n}W$). If $T$ is a presentable (resp. very presentable) $n$-stack over $X$ then the $n$-stack $\underline{Hom}(W_{{\cal X}}, T)$ is presentable (resp. very presentable). \end{corollary} {\em Proof:} We first show this for $W=S^m$, the $m$-sphere. Do this by induction on $m$. It is clear for $m=0$ because then $W$ consists of two points and $\underline{Hom}(W_{{\cal X}}, T)=T\times T$. For any $m$, write $S^m$ as the union of two copies of $B^m$ joined along $S^{m-1}$. We get $$ \underline{Hom}(S^m_{{\cal X}}, T)=T\times _{ \underline{Hom}(S^{m-1}_{{\cal X}}, T)}T, $$ since $\underline{Hom}(B^m_{{\cal X}}, T)=T$. By Theorem \ref{I.1.s.?}, $\underline{Hom}(S^m_{{\cal X}}, T)$ is presentable (resp. very presentable). This shows the corollary for the spheres. We now treat the case of general $W$, by induction on the number of cells. We may thus write $W=W'\cup B^m$ with the cell $B^m$ attached over an attaching map $S^{m-1}\rightarrow W'$, and where we know the result for $W'$. Then $$ \underline{Hom}(W_{{\cal X}}, T)=\underline{Hom}(W'_{{\cal X}}, T)\times _{ \underline{Hom}(S^{m-1}_{{\cal X}}, T)}T. $$ Again by Theorem \ref{I.1.s.?}, we obtain the result for $W$. \hfill $\Box$\vspace{.1in} Let $Pres ^n/{\cal X}$ denote the $n+1$-category of presentable $n$-stacks. We define the {\em presentable shape} of $W$ to be the $n+1$-functor $$ Shape (W):T\mapsto \underline{Hom}(\underline{W}, T) $$ from $Pres ^n/{\cal X}$ to $Pres ^n/{\cal X}$. One can show (using the calculations of \cite{kobe} Corollary 3.9 over $S=Spec (k)$) that if $W$ is connected and simply connected then this functor is homotopy-representable by an object $Hull (W)\in Pres /{\cal X} $. On the other hand, in most cases where $W$ is not simply connected, the presentable shape is not representable. We could try to interpret the hull of $W$ as the inverse limit of $Shape (W)$, but this is not a standard kind of inverse limit. It is a question for further study, just what information is contained in $Shape (W)$. {\em Example:} Take $G=GL(n)$ and $T= K(G, 1)$. Fix a finite CW complex $U$. Then $M:=\underline{Hom}(U, T)$ is the moduli stack for flat principal $G$-bundles (i.e. flat vector bundles of rank $n$) on $U$. More generally it should be interesting to look at presentable or very presentable {\em connected} $T$, these are objects whose homotopy group sheaves are algebraic Lie groups over $Spec (k)$. Note that if $k$ is algebraically closed then there is an essentially unique choice of basepoint $t\in T(Spec (k))$. If $G= \pi _1(T, t)$ then we have a fibration $T\rightarrow K(G,1)$ and we get a morphism $$ \underline{Hom} (U, T) \rightarrow \underline{Hom}(U, K(G, 1)). $$ This expresses $\underline{Hom} (U, T)$ as a presentable $n$-stack over the moduli stack $M$ of flat principal $G$-bundles over $U$. One can see from this example that we should consider the notion of vector sheaf as a candidate for the higher homotopy group sheaves. \subnumero{Leray theory} We develop here a nonabelian Leray theory and K\"unneth formula. This is in some sense one of the principal reasons for going to nonconnected $n$-stacks, as they can intervene as intermediate steps even when the original coefficient stacks were connected. We give some notation for the stack of sections. If $T\rightarrow S$ is a morphism of $n$-stacks on ${\cal X}$ (or on any site) then we denote by $\underline{\Gamma}(S, T)$ the $n$-stack of sections, i.e. of diagrams $$ \begin{array}{ccc} S&\rightarrow &T\\ & {\displaystyle =}\searrow&\downarrow \\ & & S \end{array} $$ (with homotopy making the diagram commutative). We also have a notion of relative morphism stack. Suppose that $T\rightarrow S$ and $T' \rightarrow S$ are two morphisms of $n$-stacks. Then we obtain an $n$-stack together with morphism to $S$ $$ \underline{Hom}(T/S, T'/S) \rightarrow S. $$ In topological language this corresponds to the space whose fiber over $s$ is the space of morphisms from $T_s$ to $T'_s$. This should not be confused with another useful construction in the same situation, the space $$ \underline{Hom}_S(T, T') $$ which is the $n$-stack of diagrams $$ \begin{array}{ccc} T&\rightarrow &T'\\ & \searrow&\downarrow \\ & & S \end{array} $$ (again with homotopy making the diagram commutative). These things can be constructed using the point of view of simplicial presheaves or presheaves of spaces---cf for example \cite{flexible}. It remains to be seen how to give constructions of these things purely within the realm of stacks (and consequently to extend the same constructions to stacks of $n$-categories which are not necessarily $n$-groupoids). We have the following relationships among the above constructions. First of all, $\underline{\Gamma}(S, T) = \underline{Hom}_S(S, T)$. Then, \begin{lemma} Suppose $T\rightarrow S$ and $T' \rightarrow S$ are morphisms of $n$-stacks. There is a natural equivalence $$ \underline{\Gamma} (S, \underline{Hom}(T/S, T'/S)) \cong \underline{Hom}_S(T,T'). $$ \end{lemma} {\em Proof:} From the point of view of presheaves of spaces, see \cite{flexible}. \hfill $\Box$\vspace{.1in} Finally note that if $T$ is an $n$-stack and $R\rightarrow S$ is a morphism of $n$-stacks then $$ \underline{Hom}_S(R/S, T\times S/S) \cong \underline{Hom}(R, T). $$ From the above lemma we obtain a method of ``devissage'': \begin{corollary} Suppose $T$ is an $n$-stack and $R\rightarrow S$ is a morphism of $n$-stacks, then $$ \underline{Hom}(R, T) \cong \underline{\Gamma} (S, \underline{Hom}(R/S, T\times S/S)). $$ \end{corollary} \vspace*{-.5cm} \hfill $\Box$\vspace{.1in} In words this says that to calculate the stack of morphisms from $R$ to $T$ we first look at the fiberwise morphisms from $R/S$ to $T$, and then we take the sections over $S$. Rather than taking the internal morphism and section spaces we can take the external ones, removing the underline in the notation which means taking the sections over $\ast$ (which is $Spec (k)$ in our case). We get the statement $$ Hom(R, T) \cong \Gamma (S, \underline{Hom}(R/S, T\times S/S)). $$ Note that it is still essential to look at the internal $\underline{Hom}$ inside the space of sections. It might be worthwhile looking at how this works in the case of usual cohomology. Suppose ${\cal A}$ is a sheaf of abelian groups on ${\cal X}$. Let $T= K({\cal A} , n)$, so that $Hom(R, T)$ is an $n$-groupoid with homotopy groups $$ \pi _i = H^{n-i}(R, {\cal A} ). $$ Similarly $\underline{Hom}(R/S, T)$ is an $n$-stack over $S$ whose relative homotopy group sheaves over $S$ are the higher direct images $$ \pi _i = R^{n-i}f_{\ast} ({\cal A} |_R). $$ There is a spectral sequence for the $n$-stack of sections going from the cohomology of $S$ with coefficients in the relative homotopy sheaves to the homotopy groups of the space of sections, which turns out to be the Leray spectral sequence in this case. This version of Leray theory is due to Thomason \cite{Thomason}, who developed it mostly in the context of presheaves of spectra. We finally introduce one more bit of notation combining the previous notations, that is the {\em relative section stack}. Suppose $R\rightarrow S\rightarrow T$ are morphisms of $n$-stacks. Then we obtain the $n$-stack $$ \underline{\Gamma}(S/T, R/T)\rightarrow T $$ which is geometrically the ``fiberwise space of sections of the morphism $R\rightarrow S$ along the fibers of $S\rightarrow T$''. The above Leray theory can itself be presented in a relative context: \begin{lemma} \mylabel{RelativeLeray} Suppose $R\rightarrow S\rightarrow T\rightarrow U$ are morphisms of $n$-stacks. Then $$ \underline{\Gamma}(T/U, \underline{\Gamma}(S/T, R/T)/U) \cong \underline{\Gamma}(T/U, R/U). $$ \end{lemma} \hfill $\Box$\vspace{.1in} Of course, given four morphisms there should be a diagram expressing compatibility of these Leray equivalences (and further diagrams of homotopy between the homotopies). \subnumero{Leray theory for presentable and very presentable $n$-stacks} Now we get back to presentable and very presentable $n$-stacks. Our goal is to show that in certain cases the Leray theory stays within the world of presentable $n$-stacks. The first task is to generalize Corollary \ref{I.1.u} to the case of a local coefficient system, i.e. a presentable morphism of $n$-stacks to our given finite CW complex. \begin{lemma} \mylabel{Leray2} Suppose $U$ is a constant $n$-stack associated to the $n$-groupoids associated to a finite CW complex. Suppose $T\rightarrow U$ is a presentable (resp. very presentable) morphism of $n$-stacks. Then the $n$-groupoid of sections $\underline{\Gamma}(U, T)$ is a presentable (resp. very presentable) $n$-stack. \end{lemma} {\em Proof:} The proof is identical to that of Corollary \ref{I.1.u} but we repeat it here for the reader's convenience. As before, we first treat the case $U=S^m$ by induction on $m$. It is clear for $m=0$ because then $W$ consists of two points $a,b$ and $\underline{\Gamma }(W_{{\cal X}}, T)=T_a\times T_b$, with the fibers $T_a$ and $T_b$ being presentable (resp. very presentable). Now for any $m$, write $S^m$ as the union of two copies of $B^m$ joined along $S^{m-1}$ and let $T_a$ be the fiber of $T$ over a basepoint. This fiber is presentable (resp. very presentable). We get $$ \underline{\Gamma }(S^m_{{\cal X}}, T)=T_a\times _{ \underline{\Gamma }(S^{m-1}_{{\cal X}}, T)}T_a, $$ since $\underline{\Gamma }(B^m_{{\cal X}}, T)\cong T_a$. By the induction hypothesis and Theorem \ref{I.1.s.?}, $\underline{\Gamma }(S^m_{{\cal X}}, T)$ is presentable (resp. very presentable). This shows the lemma for the spheres. We now treat the case of general $U$, by induction on the number of cells. We may thus write $U=U'\cup B^m$ with the cell $B^m$ attached over an attaching map $S^{m-1}\rightarrow U'$, and where we know the result for $U'$. Again let $T_a$ be the fiber over a basepoint in the attached cell. Then $$ \underline{\Gamma }(U_{{\cal X}}, T)=\underline{\Gamma }(U'_{{\cal X}}, T)\times _{ \underline{\Gamma }(S^{m-1}_{{\cal X}}, T)}T_a. $$ By Theorem \ref{I.1.s.?} and the above result for spheres, we obtain the result for $U$. \hfill $\Box$\vspace{.1in} Say that a morphism $U\rightarrow V$ of $n$-stacks is {\em of finite CW type} if for any scheme of finite type $X$ with morphism $X\rightarrow V$ there is a covering family $\{ Y_{\alpha} \rightarrow X\}$ and finite CW complexes $W^{\alpha}$ such that $Y_{\alpha} \times _V U \cong Y_{\alpha} \times W^{\alpha}_{{\cal X}}$ (with $W^{\alpha}_{{\cal X}}$ being the constant $n$-stack associated to $\Pi _n(W^{\alpha})$ as defined previously). \begin{theorem} \mylabel{Leray} Suppose $U\rightarrow V$ is a morphism of $n$-stacks of finite CW type, and suppose $T\rightarrow U$ is a presentable (resp. very presentable) morphism of $n$-stacks. Then $\underline{\Gamma} (U/V, T/V)\rightarrow V$ is a presentable (resp. very presentable) morphism. \end{theorem} {\em Proof:} Suppose $X$ is a scheme of finite type with a morphism $X\rightarrow V$. Let $\{ Y^{\alpha} \rightarrow X\}$ be the covering family and $\{ W^{\alpha}\}$ the collection of finite CW complexes with isomorphisms $U\times _VY^{\alpha}\cong W^{\alpha}_{{\cal X}}$ given by the fact that $U\rightarrow V$ is a morphism of finite CW type. It suffices to prove that $$ \underline{\Gamma} (U/V, T/V)\times _V Y^{\alpha}= \underline{\Gamma} (U\times _VY^{\alpha}/Y^{\alpha}, T\times _VY^{\alpha}/Y^{\alpha}) $$ is presentable (resp. very presentable). Thus it suffices to prove the theorem in the case where $V$ is a scheme of finite type and $U=V\times W_{{\cal X}}$ for a finite CW complex $W$. With these hypotheses we return to the notations of the theorem. If $W$ is a finite union of components then the section space in question will be the product of the section spaces of each of the components. Thus we may assume that $W$ is connected. The $n$-stack of sections from $W_{{\cal X}}$ to $V\times W_{{\cal X}}$ is isomorphic to $V$. Thus the $n$-stack of sections of the morphism $T\rightarrow W_{{\cal X}}$ maps to $V$, and this $n$-stack of sections is the same as the relative section stack $\underline{\Gamma}(U/V, T/V)$. It suffices to prove that $\underline{\Gamma}(W_{{\cal X}}, T)$ is presentable (resp. very presentable). But the morphism $V\times W_{{\cal X}}\rightarrow W_{{\cal X}}$ is very presentable, so by Corollary \ref{composition} the morphism $T\rightarrow W_{{\cal X}}$ is presentable (resp. very presentable), and Lemma \ref{Leray2} applies to give that $\underline{\Gamma}(W_{{\cal X}}, T)$ is presentable (resp. very presentable) as needed. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{Leray1} Suppose $U\rightarrow V$ is a morphism of $n$-stacks of finite CW type and Suppose $T\rightarrow V$ is a presentable morphism of $n$-stacks. Then the morphism $$ \underline{Hom}(U/V, T/V)\rightarrow V $$ is a presentable morphism. \end{corollary} {\em Proof:} We have $$ \underline{Hom}(U/V, T/V) = \underline{\Gamma}(U/V, T\times _VU/V) $$ and $T\times _VU\rightarrow U$ is presentable by \ref{fiberprod}, so Theorem \ref{Leray} applies. \hfill $\Box$\vspace{.1in} \begin{corollary} \mylabel{Leray1a} Suppose $T$ is a presentable $n$-stack, and suppose $V\rightarrow U$ is a morphism whose fibers are finite CW complexes in the sense of the above theorem. Then $$ \underline{Hom}(V/U, T\times U/U)\rightarrow U $$ is a presentable morphism. \end{corollary} {\em Proof:} Indeed, the morphism $T\times U\rightarrow U$ is presentable. \hfill $\Box$\vspace{.1in} We look at the case of a morphism of $n$-groupoids $f:U\rightarrow V$ such that $U$ and $V$ are the $n$-groupoids associated to finite CW complexes. Suppose that the fibers of $f$ are the $n$-groupoids associated to finite CW complexes. This is the case for example if $f$ comes from a smooth morphism of manifolds. For a presentable $n$-stack $T$ we can calculate $$ \underline{Hom}(V, T) = \underline{\Gamma}(U, \underline{Hom}(V/U, T\times U/U)). $$ Corollary \ref{Leray1a} states that $\underline{Hom}(V/U, T\times U/U)\rightarrow U$ is a presentable morphism, and Lemma \ref{Leray2} (which is also a corollary of Theorem \ref{Leray}) states that for any presentable morphism $R\rightarrow U$ the space of sections is presentable. We obtain in particular the presentability of $\underline{Hom}(V, T)$ (which we already knew beforehand). The Leray devissage process thus stays within the realm of presentable $n$-stacks. {\em The K\"unneth formula:} We can apply the above discussion to the particular case where $V=U\times U'$ is a product. In this case the formula is simplified: $$ \underline{Hom}(U\times U', T) = \underline{Hom}(U, \underline{Hom}(U', T)) $$ and again (this time using only Corollary \ref{I.1.u}) this process of first taking $\underline{Hom}(U', T)$ and then $\underline{Hom}(U, -)$ stays within the realm of presentable $n$-stacks. Of course the entire discussion above works equally well if we replace ``presentable'' by ``very presentable''. {\em Example:} Take $G=GL(n)$ and $T= K(G, 1)$. Then $M':=\underline{Hom}(U', T)$ is the moduli stack for flat principal $G$-bundles (i.e. flat vector bundles of rank $n$) on $U'$. After that, assuming that $U$ is connected, $\underline{Hom}(U, M')$ is the moduli stack of flat $G$-bundles on $U\times U'$. More generally it should be interesting to look at presentable or very presentable {\em connected} $T$, these are objects whose homotopy group sheaves are algebraic Lie groups over $Spec (k)$. Note that if $k$ is algebraically closed then there is an essentially unique choice of basepoint $t\in T(Spec (k))$. If $G= \pi _1(T, t)$ then we have a fibration $T\rightarrow K(G,1)$ and we get a morphism $$ \underline{Hom} (U, T) \rightarrow \underline{Hom}(U, K(G, 1)). $$ This expresses $\underline{Hom} (U, T)$ as a presentable $n$-stack over the moduli stack $M$ of flat principal $G$-bundles over $U$.
1996-07-18T21:42:53
9607
alg-geom/9607018
en
https://arxiv.org/abs/alg-geom/9607018
[ "alg-geom", "math.AG" ]
alg-geom/9607018
Pablo Ares Gastesi
Pablo Ares Gastesi (Tata Institute, Bombay)
Torelli groups and Jacobian varieties of non-orientable compact Klein surfaces
AMSLaTeX, 18 pages, xypic, available from ftp://ftp.math.tifr.res.in/ with dvi file at http://www.math.tifr.res.in/~pablo/
null
null
null
null
The Torelli group of a compact non-orientable Klein surface is the subgroup of the modular group consisting of the mapping classes that act trivially on the first homology group of the surface. We prove that if a surface has genus at least $3$, then the Torelli group acts fixed points free on the Teichm\"{u}ller space of the surface. That gives an embedding of the Torelli space of a Klein surface in the Torelli space of its complex double. We also construct real tori associated to Klein surfaces, which we call the Jacobian of the surface. We prove that this Jacobian is isomorphic to a component of the real part of the Jacobian of the complex double.
[ { "version": "v1", "created": "Thu, 18 Jul 1996 19:42:55 GMT" } ]
2008-02-03T00:00:00
[ [ "Gastesi", "Pablo Ares", "", "Tata Institute, Bombay" ] ]
alg-geom
\section{Statement of results} Klein surfaces are the natural generalization of Riemann surfaces to the non-orientable situation: one considers holomorphic and anti-holomorphic changes of coordinates. One of the points of interest in the study of Klein surfaces is to determine which results of the theory of deformation of Riemann surfaces hold for the non-orientable case. A common approach to this problem is to consider a Klein surface $\Sigma$, as a Riemann surface $\Sigma^c$, with an anti-holomorphic involution $\sigma$, and thus one wants to find $\sigma$-invariant objects. In this paper we follow these two points of view to show two related results in the theory of Klein surfaces, that is, we will do some constructions on Klein surfaces, and then find the corresponding invariant objects related to the Riemann surface $\Sigma^c$. More precisely, we construct the {\bf Torelli space} $Tor(\Sigma)$, and prove that it can be identified with the set of fixed points of an involution on $Tor(\Sigma^c)$. We also construct the {\bf Jacobian variety} $J(\Sigma)$ of $\Sigma$ by integrating a basis of the space of real harmonic forms over the free part of ${\mathrm H}_1(\Sigma,{\Bbb Z})$. We prove that $J(\Sigma)$ is isomorphic to a component of the real part of the Jacobian $J(\Sigma^c)$ of the complex double $\Sigma^c$ of $\Sigma$. Given a compact smooth non-orientable surface $\Sigma$, the {Teichm\"{u}ller space} $T(\Sigma)$ of $\Sigma$ is defined as $T(\Sigma)={\cal M}(\Sigma)/Diff_0(\Sigma)$, where ${\cal M}(\Sigma)$ is the set of Klein surface structures on $\Sigma$ that agree with the given smooth structure, and $Diff_0(\Sigma)$ is the group of diffeomorphisms of $\Sigma$ homotopic to the identity \cite[pg. 145]{sep:book}. We will use $\Sigma$ for a Klein surface, if it is clear from the context what the structure is, or we will write $(\Sigma,X)$ if we need to specify more. The {\bf modular} or {\bf mapping class group}, $Mod(\Sigma)=Diff(\Sigma)/Diff_0(\Sigma)$, acts on $T(\Sigma)$ by pull-back of dianalytic structures (see \S $2$). The {\bf Torelli group} $U(\Sigma)$ is the subgroup of $Mod(\Sigma)$ consisting of the mapping classes that act trivially on ${\mathrm H}_1(\Sigma,{\Bbb Z})$. The parallel result to the following theorem is a classical fact on Riemann surfaces. \setcounter{section}{3} \setcounter{thm}{0} \begin{thm}Let $\Sigma$ be a compact non-orientable surface of genus $g \geq 3$. Let $[f]\in Mod(\Sigma)$, and suppose that there exists a Klein surface structure $X$ on $\Sigma$ such that $f:(\Sigma,X)\rightarrow (\Sigma,X)$ is dianalytic. Then $[f]=[id]$. Therefore, $U(\Sigma)$ acts fixed-points free on $T(\Sigma)$, and the Torelli space $Tor(\Sigma)=T(\Sigma)/U(\Sigma)$ is a smooth real manifold of dimension $3g-6$. \end{thm} Assume now that $\Sigma$ has a fixed Klein surface structure. Then there exists an unramified double covering of $\Sigma$ by a Riemann surface $\Sigma^c$, known as the {\bf complex double}. Moreover, $\Sigma$ is isomorphic to $\Sigma^c/<\!\sigma\!>$, where $\sigma$ is an anti-holomorphic involution. The mapping $\sigma$ induces involutions $\sigma^*$ and $\tilde\sigma$ on $T(\Sigma)$ and $Tor(\Sigma)$, respectively. It is a well known fact that $T(\Sigma)$ can be identified with the set of fixed points of $\sigma^*$. A similar result holds for Torelli spaces, as the next proposition shows. \setcounter{thm}{2} \begin{prop}The Torelli space $Tor(\Sigma)$ can be identified with the set of fixed points of $\tilde\sigma$ on $Tor(\Sigma^c)$. \end{prop} Torelli spaces are intimately related to the Jacobian variety of a compact Riemann surface. Recall that this variety $J(\Sigma^c)$, is a $g$-dimensional complex torus ($g$ is the genus of $\Sigma^c$) given by ${\Bbb C}^g/\Gamma$, where $\Gamma$ is the lattice generated by integration of a basis of holomorphic forms on $\Sigma^c$ over a basis of ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$. We can also construct the Jacobian by considering the lattice $\Gamma'$ generated by integration of harmonic forms and then taking the quotient ${\Bbb R}^{2g}/\Gamma'$, which is a real torus. The Hodge-$*$ operator gives a complex structure to this real torus in such a way that it becomes $J(\Sigma^c)$. This point of view can be generalized to construct a Jacobian variety $J(\Sigma)$ of a non-orientable Klein surface. \setcounter{section}{4} \setcounter{thm}{0} \begin{thm} Let $\Sigma$ be a compact non-orientable surface of genus $g\geq3$. Then we can associate to $\Sigma$ a real torus of dimension $g-1$, the {\bf Jacobian variety} $J(\Sigma)$ of $\Sigma$, such that $J(\Sigma)$ is isomorphic to any component of the real part of the Jacobian $J(\Sigma^c)$ of the complex double. This last set is defined as the set of fixed points of the symmetry $\sigma_1$ of $J(\Sigma^c)$ induced by $\sigma$. \end{thm} {\bf Acknowledgments}: the idea of using harmonic forms to construct $J(\Sigma)$ was suggested by S. Nag; I would like to thank him for useful conversations regarding this topic. I would like also to express my gratitude to D. S. Nagaraj and R. R. Simha for many helpful conversations while this paper was being written. \setcounter{section}{1} \setcounter{thm}{0} \section{Some general facts about Klein surfaces} A {\bf Klein surface} (or {\bf dianalytic}) structure $X$ on a surface without boundary $\Sigma$ is a covering by open sets $U_i$, and a collection of homeomorphisms $z_i:U_i\rightarrow V_i$, where $V_i\subset{\Bbb C}$ are open sets, such that $z_i\circ z_j^{-1}$ is holomorphic or anti-holomorphic, whenever $U_i\cap U_j\neq\emptyset$ \cite{all:klein}. Observe that a Klein surface structure on an orientable surface is just a pair of conjugate Riemann surface structures \cite{natan:klein}. A compact non-orientable surface $\Sigma$ is homeomorphic to the connected sum of $g\geq 1$ real projective planes \cite{blackett:topo}. The number $g$ is called the {\bf genus} of $\Sigma$. If $g=2n+1$, then the fundamental group of $\Sigma$ has a presentation given by generators $c$, $a_1,\ldots,a_n$, $b_1,\ldots,b_n$, satisfying $c^2\prod_{j=1}^n[a_j,b_j]=1$, where $[a,b]=aba^{-1}b^{-1}$. If the genus is even, $g=2n+2$, then we can choose generators $c$, $d,$ $a_1,\ldots,a_n$, $b_1,\ldots,b_n$, satisfying the relation $c^2d^2\prod_{j=1}^n[a_j,b_j]=1$. An alternative presentation for this latter case is given by generators $\gamma$, $\delta$, $a_1,\ldots,a_n,$ $b_1,\ldots,b_n,$ and the relation $\gamma\delta\gamma^{-1}\delta\prod_{j=1}^n[a_j,b_j]=1$. For the rest of this paper, we will assume that all surfaces are compact without boundary. We will further assume that non-orientable surfaces have genus $g\geq 3$, while orientable surfaces satisfy $g\geq 2$. The {\bf complex double} \cite{all:klein} of a Klein surface $\Sigma$ of genus $g$ is a triple $(\Sigma^c,\pi,\sigma)$, where:\\ \noindent (1) $\Sigma^c$ is a Riemann surface of genus $g-1$;\\ \noindent (2) $\pi:\Sigma^c\rightarrow\Sigma$ is an unramified double covering;\\ \noindent (3) there exist local coordinates $z$ and $w$ on $\Sigma^c$ and $\Sigma$, respectively, such that the function $w\circ\pi\circ z^{-1}$ is either holomorphic or anti-holomorphic (i.e. $\pi$ is a morphism of Klein surfaces);\\ \noindent (4) $\sigma:\Sigma^c\rightarrow\Sigma^c$ is a symmetry such that $\pi\circ\sigma=\pi$. Let $S$ be a compact orientable surface, with a fixed orientation and a smooth structure. The {\bf Teichm\"{u}ller space} $T(S)$ of $S$ is $T(S)={\cal M}(S)/Diff_0(S)$, where ${\cal M}(S)$ is the set of Riemann surface structures on $S$ that agree with the given orientation and smooth structure \cite{sep:book}. The classical definition of $T(S)$ involves quasiconformal mappings; to see that it is equivalent to the above definition, it suffices to observe that on a compact surface, any homeomorphism is homotopic to a smooth one, and diffeomorphisms are quasiconformal. The {\bf modular} or {\bf mapping class} group $Mod(S)$ is the group of homotopic classes of orientation preserving diffeomorphisms of $S$, that is $Mod(S)=Diff^+(S)/Diff_0(S)$. This group acts on $T(S)$ by pull-back of complex structures: if $[f]\in Mod(S)$, and $[X]\in T(S)$, then $[f]^*([X])=[f^*(X)]$, where $f^*(X)$ is the Riemann surface structure on $S$ that makes $f:(\Sigma,f^*(X))\rightarrow (\Sigma,X)$ biholomorphic. However, this action has fixed points; it is therefore interesting to find subgroups of $Mod(\Sigma)$ that act without fixed points on $T(S)$. A subgroup $G$ of $Mod(S)$ has the {\bf Hurwitz-Serre} property \cite{nag:teic} if $G$ satisfies that for any element $[g]\in G$ such that there exists an $[X]\in{\cal M}(S)$ with $g:(S,X)\rightarrow (S,X)$ biholomorphic, one has that $[g]=[id]$. A group with this property will act fixed-points free on Teichm\"{u}ller space and, therefore, the quotient $T(S)/G$ will be a smooth finite dimensional complex manifold. The {\bf Torelli group} $U(S)=\{[f]\in Mod(S);~ f~\mathrm{acts~trivially~on~H}_1(S,{\Bbb Z})\}$, is known to satisfy the Hurwitz-Serre property \cite{fk:book}. The quotient space $Tor(S)=T(S)/U(S)$ is called the {\bf Torelli space} of $S$. The Jacobian variety $J(S)$ of a compact Riemann surface is an abelian variety constructed as follows: let ${\cal B}^c=\{\alpha_1,\ldots,\alpha_g,\beta_1,\ldots,\beta_g\}$ be a symplectic basis of ${\mathrm H}_1(S,{\Bbb Z})$. Then we can find a dual basis for ${\mathrm H}^0(S,\Omega_S^1)$, the space of holomorphic forms on $S$, consisting of forms $\{\omega_1,\ldots,\omega_g\}$, satisfying $$\int_{\alpha_j}\omega_k=\begin{cases} 1 & \text{if } j=k,\\ 0 & \text{otherwise}.\end{cases}$$ Let $\Gamma^c$ be the lattice on ${\Bbb C}^g$ generated by the vectors $(\int_c\omega_1,\ldots,\int_c\omega_g)$, $c\in{\cal B}^c$; then we define $J(S)={\Bbb C}^g/\Gamma^c$. \section{Torelli groups of non-orientable compact surfaces} In this section we will show that the Torelli group of a compact non-orientable Klein surface $\Sigma$ has the Hurwitz-Serre property. The main idea of the proof is to see that, if a diffeomorphism of a Klein surface $\Sigma$ acts trivially on ${\mathrm H}_1(\Sigma,{\Bbb Z})$, its orientation preserving lift to the complex double will act trivially on ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$; then we use the fact that the Hurwitz-Serre property is satisfied for Riemann surfaces. The quotient space $Tor(\Sigma)=T(\Sigma)/U(\Sigma)$ is a smooth real manifold. We will show that $Tor(\Sigma)$ can be identified with the set of fixed points of a symmetry $\tilde\sigma$ on $Tor(\Sigma^c)$. Let us start with a smooth non-orientable surface $\Sigma$, of genus $g=2n+1$, and a diffeomorphism $f:\Sigma\rightarrow\Sigma$ such that $f$ acts trivially on ${\mathrm H}_1(\Sigma,{\Bbb Z})$. We can find a unique orientation preserving diffeomorphism $\tilde f$ of $\Sigma^c$ such that the following diagram commutes \cite{sep:spaces}: $$\diagram \Sigma^c\rto^{\tilde f}\dto_\pi & \Sigma^c\dto^\pi \\ \Sigma\rto^f & \Sigma.\enddiagram$$ We want to show that the mapping ${\tilde f}_\#$ induced by $\tilde f$ on ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ is trivial. For that purpose we need to recall the way $\Sigma^c$ is constructed, from the topological viewpoint. The reader can find more details in \cite{blackett:topo}. By the presentation of the fundamental group of $\Sigma$, we can identify this surface with a $(4n+2)$-polygon, whose sides are labeled to satisfy the relation of the fundamental group. Then $\Sigma^c$ is given by two polygons with boundary relations: $$c_1c_2\prod_{j=1}^n[a_{j,1},b_{j,1}]=1\hspace{5mm} \mathrm{and}\hspace{5mm} c_2c_1\prod_{j=1}^n[a_{j,2},b_{j,2}]=1.$$ To obtain a single relation, we find the value of $c_2$ on the right hand side equation and substitute it on the left hand one (equivalently, we glue the two polygons by the $c_2$ sides): $$c_2=(\prod_{j=1}^n[b_{n+1-j,2},a_{n+1-j,2}])c_1^{-1};$$ therefore $$c_1\big(\prod_{j=1}^n[b_{n+1-j,2},a_{n+1-j,2}]\big)c_1^{-1} \big(\prod_{j=1}^n[a_{j,1},b_{j,1}]\big)=$$ $$\big(\prod_{j=1}^n[c_1b_{n+1-j,2}c_1^{-1},c_1a_{n+1-j,2}c_1^{-1}]\big) \big(\prod_{j=1}^n[a_{j,1},b_{j,1}]\big)=1.$$ {}From this formula we see that $\Sigma^c$ is a compact surface of genus $g-1=2n$; we can choose the following paths as generators of the fundamental group of $\Sigma$: $$\alpha_1=c_1b_{n,2}c_1^{-1},\ldots,\alpha_n=c_1b_{n,2}c_1^{-1}, \alpha_{n+1}=a_{1,1}\ldots,\alpha_{2n}=a_{n,1},$$ $$\beta=c_1a_{n,2}c_1^{-1},\ldots,\beta_n=c_1a_{n,2}c_1^{-1}, \beta_{n+1}=b_{1,1}\ldots,\beta_{2n}=b_{n,1}.$$ These loops satisfy $\prod_{j=1}^{2n}[a_j,b_j]=1$. Let $\cal B$ and ${\cal B}^c$ denote the basis of ${\mathrm H}_1(\Sigma,{\Bbb Z})$ and ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ induced by the two given sets of generators of the corresponding fundamental groups. By an abuse of notation, we will use the same letters for the elements of the fundamental group and their classes in homology. We can see that ${\cal B}^c$ is a symplectic basis of ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$; that is, the intersection matrix is given by $J=\left(\begin{matrix} 0 & {\mathrm I} \\ -{\mathrm I} & 0 \end{matrix}\right) ,$ where ${\mathrm I}$ is the identity matrix. The covering map $\pi:\Sigma^c\rightarrow\Sigma$ induces a mapping on homology, with associated matrix $$\pi_\#=\left(\begin{matrix} 0 & 0 & 0 & 0 \\ 0 & {\mathrm I} & K & 0 \\ K & 0 & 0 & {\mathrm I} \end{matrix}\right),$$ with respect to ${\cal B}^c$ and $\cal B$. The matrix $K$ is given by $$K=\left(\begin{matrix} 0 & \cdots & \cdots & 1 \\ 0 & \cdots & 1 & 0 \\ \vdots & & & \vdots \\ 1 & 0 & \cdots & 0 \end{matrix}\right).$$ The symmetry $\sigma$ maps $a_{j1}$ (resp. $b_{j1}$) to $a_{j2}$ (resp. $b_{j2}$); it is not difficult to see that the map $\sigma_\#$ induced on ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ is given by $\sigma_\#=K$. Let ${\tilde f}_\#=\left(A_{jk}\right)_{j,k=1}^4$; since $f_\#$ acts trivially on ${\mathrm H}_1(\Sigma,{\Bbb Z})$, we have $\pi_\#{\tilde f}_\#=\pi_\#$. By the uniqueness of ${\tilde f}_\#$ we get ${\tilde f}_\#\sigma_\#=\sigma_\#{\tilde f}_\#$. Finally, ${\tilde f}^t_\#J{\tilde f}_\#=J$, where ${\tilde f}_\#^t$ is the transpose of ${\tilde f}_\#$, since $\tilde f$ preserves the intersection matrix (\cite[theorem N13, pg. 178]{mag:comb}). The condition $\pi_\#{\tilde f}_\# = \pi_\#$ is equivalent to the following set of equations: $$\left\{\begin{array}{lclccclcl} A_{21} + KA_{31} & = & 0 & & & & KA_{11} + A_{41} & = & K \\ A_{22} + KA_{32} & = & {\mathrm I} & & & & KA_{12} + A_{42} & = & 0 \\ A_{23} + KA_{33} & = & K & & & & KA_{13} + A_{43} & = & 0 \\ A_{24} + KA_{34} & = & 0 & & & & KA_{14} + A_{44} & = & {\mathrm I}. \end{array} \right .$$ Therefore, the matrix ${\tilde f}_\#$ can be written as $${\tilde f}_\# = \left(\begin{matrix} A_{11} & A_{12} & A_{13} & A_{14} \\ A_{21} & A_{22} & A_{23} & A_{24} \\ -KA_{21} & K-KA_{22} & {\mathrm I}-KA_{23} & -KA_{24} \\ K-KA_{11} & -KA_{12} & -KA_{13} & {\mathrm I}-KA_{14} \end{matrix}\right).$$ We now use the fact that $\tilde f$ and $\sigma$ commute, to obtain the following relations among the entries of the matrix ${\tilde f}_\#$: \begin{equation}\left\{ \begin{array}{lclccclcl} A_{14}K & = & K(K-KA_{11}) & & & & A_{24}K & = & K(-KA_{21}) \\ A_{13}K & = & K(-KA_{12}) & & & & A_{23}K & = & K(K-KA_{22}) \\ A_{12}K & = & K(KA_{13}) & & & & A_{22}K & = & K({\mathrm I}-KA_{23}) \\ A_{11}K & = & K({\mathrm I}-KA_{14})& & & & A_{21}K & = & K(-KA_{24}). \end{array}\right .\label{eq:f}\end{equation} Consider the equation ${\tilde f}_\#^t J {\tilde f}_\# = J$; looking at the first row of the matrices on both sides of the equality, we get, after using \eqref{eq:f} to simplify the result, $$\left\{\begin{array}{rcr} A_{21}^tK-KA_{21} & = & 0 \\ A_{11}^tK-KA_{22} & = & 0 \\ -2{\mathrm I}+A_{11}^t+KA_{22}K & = & 0 \\ A_{21}^t+KA_{21}K & = & 0 \end{array}\right .$$ Solving these equations, we get $A_{21}=0$ and $A_{11}={\mathrm I}$, which imply that $A_{24}=0$ and $A_{14}=0$. Using this, we now consider the equality between the second rows of the matrices ${\tilde f}_\#^t J {\tilde f}_\#$ and $J$, to obtain $A_{22}={\mathrm I}$, and $A_{12}=0$. By \eqref{eq:f}, we get $A_{23}=0$ and $A_{13}=0$. Therefore, we have that ${\tilde f}_\#$ is the identity matrix. {\bf Remark:} if we would have chosen the orientation reversing lift of $f$, say ${\tilde f}_1$, then ${\tilde f}_1={\tilde f}\sigma$, so $({\tilde f}_1)_\#={\tilde f}_\#\sigma_\#=\sigma_\#$. If $\Sigma$ has even genus $g=2n+2$, we use the first of the two presentations of its fundamental group given in \S $2$. We have that $\Sigma^c$ is given by two polygons with boundary relations: $$c_1c_2d_1d_2\prod_{j=1}^n[a_{j,1},b_{j,1}] = 1\hspace{5mm} \mathrm{and}\hspace{5mm} c_2c_1d_2d_1\prod_{j=1}^n[a_{j,2},b_{j,2}] = 1.$$ {}From the second equation we get $$d_2=c_1^{-1}c_2^{-1}\big(\prod_{j=1}^n[b_{n+1-j,2},a_{n+1-j,2}]\big) d_1^{-1},$$ which reduces the first equation to $$c_1c_2d_1c_1^{-1}c_2^{-1}\big(\prod_{j=1}^n[b_{n+1-j,2},a_{n+1-j,2}]\big) d_1^{-1}\big(\prod_{j=1}^n[a_{j,1},b_{j,1}]\big) = $$ $$c_1c_2d_1c_1^{-1}c_2^{-1}d_1^{-1} \big(\prod_{j=1}^n[d_1b_{n+1-j,2}d_1^{-1},d_1a_{n+1-j,2}d_1^{-1}]\big) \big(\prod_{j=1}^n[a_{j,1},b_{j,1}]\big) = 1.$$ We therefore obtain that the fundamental group of $\Sigma^c$ is generated by the loops $$\alpha_1=c_1d_1^{-1},~\alpha_2=d_1b_{n,2}d_1^{-1},\ldots, \alpha_{n+1}=d_1b_{1,2}d_1^{-1}, \alpha_{n+2}=a_{1,1},\ldots,\alpha_{2n+1}=a_{n,1},$$ $$\beta_1=d_1c_2,~\beta_2=d_1a_{n,2}d_1^{-1},\ldots, \beta_{n+1}=d_1\a_{1,2}d_1^{-1}, \beta{n+2}=a_{1,1},\ldots,\beta_{2n+1}=a_{n,1},$$ satisfying the relation $\prod_{j=1}^{2n+1}[\alpha_j,\beta_j]=1.$ The basis $\{\alpha_1, \alpha_2, \ldots, \alpha_{2n+1}, \beta_1, \beta_2, \ldots,$\newline \noindent $\beta_{2n+1} \}$ is symplectic, but computations are easier if we rearrange the basis as ${\cal B}^c=\{ \alpha_1, \beta_1, \alpha_2, \ldots,\alpha_{2n+1}, \beta_2, \ldots, \beta_{2n+1} \}$, whose intersection matrix is given by: $$J=\left(\begin{matrix} N & 0 & 0 & 0 & 0 & \\ 0 & 0 & 0 & {\mathrm I} & 0 & \\ 0 & 0 & 0 & 0 & {\mathrm I} & \\ -{\mathrm I} & 0 & 0 & 0 & 0 & \\ 0 & -{\mathrm I} & 0 & 0 & 0 & \end{matrix}\right) ,$$ where $N=\left(\begin{matrix} 0 & 1 \\ -1 & 0 \end{matrix}\right) .$ With respect to $\cal B^c$, the symmetry $\sigma$ has the following matrix representation, for its action on ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$: $$\sigma_\#=\left(\begin{matrix} M & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & K & 0 \\ 0 & 0 & K & 0 & 0 \\ 0 & K & 0 & 0 & 0 \\ K & 0 & 0 & 0 & 0 \end{matrix}\right) ,$$ where the matrix $M=\left(\begin{matrix} 1& 0 \\ 2 & -1 \end{matrix}\right) .$ To obtain the matrix $M$, observe that in homology one has $\alpha=c_1+d_1$ and $\beta=d_1+c_2$, so $\sigma(\alpha)=c_2+d_2$, $\sigma(\beta)=d_2+c_1.$ Substituting the value of $d_2$ obtained previously, we get $M$. Similarly we have that the expression for the action of the covering map $\pi$ on homology is given by: $$\pi_\#=\left(\begin{matrix} {\mathrm I} & 0 & 0 & 0 & 0 \\ 0 & 0 & {\mathrm I} & K & 0 \\ 0 & K & 0 & 0 & {\mathrm I} \end{matrix}\right) .$$ Computing in a way similar to the odd genus case, we get that if $f:\Sigma\rightarrow\Sigma$ is a diffeomorphism of $\Sigma$ that acts trivially on ${\mathrm H}_1(\Sigma,{\Bbb Z})$, and $\tilde f$ is a lift of $f$ to $\Sigma^c$, then the action of this last mapping on ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ is given by either the identity or $\sigma_\#$, depending on whether $\tilde f$ is orientation preserving or reversing. Recall that a map $f:\Sigma\rightarrow\Sigma$ on a Klein surface is {\bf dianalytic} if, when expressed in local coordinates $(U,z)$, $f\circ z^{-1}$ is either holomorphic or anti-holomorphic. The above computations show that $U(\Sigma)$ satisfies the equivalent of the Hurwitz-Serre property. \begin{thm}Let $\Sigma$ be a compact non-orientable surface of genus $g\geq 2$. Let $[f]\in Mod(\Sigma)$, and suppose that there exists a Klein surface structure $X$ on $\Sigma$ such that $f:(\Sigma,X)\rightarrow (\Sigma,X)$ is dianalytic. Then $f$ is homotopic to the identity. \end{thm} \begin{cor}The Torelli space $Tor(\Sigma)=T(\Sigma)/U(\Sigma)$ is a smooth manifold of real dimension $3g-6$. \end{cor} \begin{proof}{of the Proposition} Since $f$ is dianalytic on the Klein surface $(\Sigma,X)$, the orientation preserving lift $\tilde f$ is biholomorphic on the Riemann surface $\Sigma^c$. But then, since the genus of $\Sigma^c$ is at least $2$, we have that $\tilde f$ is the identity, which proves the result. \end{proof} The involution $\sigma$ of $\Sigma^c$ induces a symmetry $\sigma^*$ on $T(\sigma^c)$. The Teichm\"{u}ller space $T(\Sigma)$ can be identified with the set of fixed points of $\sigma^*$, which proves the corollary. It is clear that $\sigma^*$ descends to a symmetry $\tilde\sigma$ on $Tor(\Sigma^c)$. \begin{prop}The Torelli space $Tor(\Sigma)=T(\Sigma)/U(\Sigma)$ can be identified with the set of fixed points of $\tilde\sigma$ in $Tor(\Sigma^c)$.\end{prop} \begin{pf}The proof follows immediately from the definition of Torelli spaces. In fact, we have that two elements $[X_1]$ and $[X_2]$ of ${\cal M}(\Sigma)$ project to the same point in $Tor(\Sigma)$ {\it if and only if} there exists a diffeomorphism $h\in Diff(\Sigma)$ such that $h_\#:{\mathrm H}_1(\Sigma,{\Bbb Z})\rightarrow {\mathrm H}_1(\Sigma,{\Bbb Z})$ is the identity, and $h:(\Sigma,X_2)\rightarrow (\Sigma,X_1)$ is dianalytic. The rest of the proof is similar to the proof that $T(\Sigma)$ can be identified with the set of fixed points of $\sigma^*$; see \cite{sep:book} for more details.\end{pf} \section{Jacobi varieties of Klein surfaces} Throughout this section, $\Sigma$ will denote a fixed compact non-orientable Klein surface of genus $g\geq 3$, and $\Sigma^c$ its complex double. We can take $\Sigma^c$ to be defined by a polynomial, $p(z,w)=0$, with real coefficients (\cite{all:klein} and \cite{natan:gordon}). Then the involution $\sigma$ is given by $\sigma(z,w)=(\overline z,\overline w)$, and conjugation $z\mapsto\overline z$ in ${\Bbb C}^{g-1}$ induces an involution $\sigma_1$ on the Jacobian $J(\Sigma^c)$. The set of fixed points of $\sigma_1$, that is, the {\bf real part} of $J(\Sigma^c)$, is a real manifold of dimension $g-1$; the pair $J(\Sigma^c,\sigma_1)$ is usually considered as the Jacobian of $\Sigma$. On a Klein surface the concept of harmonic forms makes sense; it is not difficult to see that the space ${\cal H}_{\Bbb R}^1(\Sigma)$ of such forms has dimension precisely $g-1$. One can choose a basis of ${\mathrm H}_1(\Sigma,{\Bbb Z})_f$, the free part of ${\mathrm H}_1(\Sigma,{\Bbb Z})\cong{\Bbb Z}^{g-1}\oplus{\Bbb Z}/2Z$, and a dual basis for ${\cal H}_{\Bbb R}^1(\Sigma)$; these two basis generate a lattice $\Gamma$ in ${\Bbb R}^{g-1}$. We will call the real torus ${\Bbb R}^{g-1}/\Gamma$ the {\bf Jacobian variety} of $\Sigma$, and denote it by $J(\Sigma)$. On the other hand, the real part of a holomorphic form (on $\Sigma^c$) is a harmonic form, so one can expect some relationship between $J(\Sigma)$ and the real part of $J(\Sigma^c)$. We prove that, in fact, $J(\Sigma)$ is isomorphic to a component of the set of fixed points of $\sigma_1$ in $J(\Sigma^c)$. A continuous function $f:W\rightarrow{\Bbb R}$ defined on an open set of a Klein or Riemann surface is called {\bf harmonic} if for any local coordinate $(U,z)$, with $U\cap W\neq\emptyset$, the function $f\circ z^{-1}$ is harmonic. Since precomposition with holomorphic and anti-holomorphic functions preserves harmonicity, the above definition makes sense. Actually, a Klein surface is the most general surface in which the notion of harmonic function is well defined \cite{all:klein}. Similarly, a (real) form $\psi$ is {\bf harmonic} if it can be written locally as $\psi=df$, where $f$ is harmonic. We will denote by ${\cal H}_{\Bbb R}^1(\Sigma)$ the space of harmonic forms on $\Sigma$. Let $\sigma^*$ be the pull-back map induced by $\sigma$ on forms on $\Sigma^c$. If $\omega=gdz$, with $g$ holomorphic, we have that $\sigma^*(\omega) = g(\sigma)\sigma_{\overline z}d\overline z$, so $\sigma^*$ is anti-holomorphic. We also have that, for any holomorphic form $\omega\in {\mathrm H}^0(\Sigma^c,\Omega^1)$, and for any cycle $c$ on $\Sigma^c$, $\int_c\sigma^*(\omega)=\int_{\sigma(c)}\omega$. Observe that this last equality agrees with \cite{natan:gordon}, while it differs of \cite{silhol:comess} and \cite{all:klein}, since these two authors define $\sigma^*$ as the conjugate of our definition (in order to have that $\sigma^*$ preserves holomorphic forms). To compute the dimension of ${\cal H}_{\Bbb R}^1(\Sigma)$, it suffices to observe that $\sigma^*$ takes harmonic forms to harmonic forms; therefore, ${\cal H}_{\Bbb R}^1(\Sigma)$ will be isomorphic to the set of fixed points of $\sigma^*$ in ${\cal H}_{\Bbb R}^1(\Sigma^c)$. By Hodge theory, $\sigma^*$ acts like $\sigma_\#=K$; so $dim\,{\cal H}_{\Bbb R}^1(\Sigma)=g-1$. This result agrees with \cite{all:obst}. In order to justify later computations, we need the following lemma. \begin{lemma}[Duality Lemma]On a non-orientable compact Klein surface $\Sigma$, of genus $g \geq 3$, the space of harmonic forms, ${\cal H}_{\Bbb R}^1(\Sigma)$, and the dual space to the homology with real coefficients, ${\mathrm H}_1(\Sigma,{\Bbb R})^*$, are isomorphic.\end{lemma} \begin{pf}From Differential Topology \cite[Theorem 15.8]{bott:forms} we know that the \v{C}ech cohomology with coefficients in the constant presheaf $\Bbb Z$, ${\mathrm H}^1_{\Bbb Z}(\Sigma)^*,$ is isomorphic to the singular cohomology ${\mathrm H}^1(\Sigma)$. Furthermore, by the Universal Coefficients Theorem \cite[Corollary 15.14.1]{bott:forms}, we have that the space ${\mathrm H}^1(\Sigma)$ is isomorphic to the free part of ${\mathrm H}_1(\Sigma)$, which is just ${\Bbb Z}^{g-1}$. Tensoring with ${\Bbb R}$ we have that ${\mathrm H}_{\Bbb R}^1(\Sigma)$ is isomorphic to ${\Bbb R}^{g-1}$. On the other hand, ${\mathrm H}_{\Bbb R}^1(\Sigma)$ is isomorphic to the de Rham cohomology ${\mathrm H}_{DR}^1(\Sigma)$ \cite[8.9, 9.8 and 14.28]{bott:forms}. By the compactness of $\Sigma$, on each de Rham class there exists at most a harmonic form. A counting of dimensions shows that there exists exactly one harmonic form, which completes the proof. Therefore, ${\cal H}_{\Bbb R}^1(\Sigma)$ and ${\mathrm H}_{DR}(\Sigma)$ are isomorphic, and since this last space is isomorphic (by integration) to ${\mathrm H}_1(\Sigma,{\Bbb R})$, we are done. \end{pf} To make matters more clear, we will construct $J(\Sigma)$ on the cases of genus $3$ and $4$. The general case will follow easily from our examples. Let us start with a Klein surface $\Sigma$ of genus $g=3$. Then from \S $2$ we have that ${\cal B}=\{a,b\}$ is a basis of ${\mathrm H}_1(\Sigma,{\Bbb Z})_f$. The loops $\alpha_1$, $\alpha_2$, $\beta_1$, $\beta_2$ of \S $3$ give a basis of ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$; but for computational purposes, it is better to choose the basis $${\cal B}^c=\{\gamma_1 = -(\alpha_2+\beta_1), \gamma_2 = -(\alpha_1+\beta_2), \delta_1 = \alpha_1+\alpha_2+\beta_1, \delta_2 = \alpha_1+\alpha_2+\beta_2\}.$$ Observe that the change of basis in ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ is given by the matrix $$C=\left(\begin{matrix} 0 & -1 & 1 & 1 \\ -1 & 0 & 1 & 1 \\ -1 & 0 & 1 & 0 \\ 0 & -1 & 0 & 1 \end{matrix}\right) ,$$ which has determinant equal to $1$, and satisfies $C^tJC=J$. Therefore ${\cal B}^c$ is a symplectic basis. The mapping $\pi_\#$ gives $\pi_\#(\gamma_1) = -2b,$ $\pi_\#(\gamma_2) = -2a.$ This suggests that we should take a basis ${\cal B}_h=\{\phi_1, \phi_2\}$ of ${\cal H}_{\Bbb R}^1(\Sigma)$ normalized by $\int_a\phi_1=\int_b\phi_2=0$, and $\int_b\phi_1=\int_a\phi_2=-1/2.$ Observe that this normalization is possible because of the Duality Lemma, and the fact that $a$ and $b$ are not torsion classes in ${\mathrm H}_1(\Sigma,{\Bbb Z})$. We can use the pull-back mapping $\pi^*$ induced by $\pi$ to get forms $\psi_j=\pi^*(\phi_j)$ on $\Sigma^c$. These forms are real harmonic, so $\omega_j=\psi_j+i*\phi_j$ are holomorphic forms (where $*$ stands for the Hodge-$*$ operator). By the expression of $\sigma$ and the formula relating $\sigma^*$ with integrals, we have that $$\int_{\gamma_j}\overline{\omega}_k = \int_{\gamma_j}\sigma^*(\omega_k) = \int_{\sigma(\gamma_j)}\omega_k = \int_{\gamma_j}\omega_k .$$ In particular, we see that $\int_{\gamma_j}\omega_k$ is real. But since by \cite[Theorem 1.0.7, pg. 74]{all:klein} ${\mathrm Re}\int_{\gamma_j}\omega_k = \int_{\gamma_j}\psi_k = \int_{\pi(\gamma_j)}\phi_k,$ we have that ${\cal B}^*=\{\omega_1, \omega_2\}$ is normalized with respect to ${\cal B}^c$. Let $P$ denote the corresponding period matrix, that is the entries of this matrix are given by $p_{jk} = \int_{\delta_k}\omega_j$. The mapping $\sigma_\#:{\mathrm H}_1(\Sigma^c,{\Bbb Z})\rightarrow {\mathrm H}_1(\Sigma^c,{\Bbb Z})$ has the following expression with respect to the basis ${\cal B}^c$: $$\sigma_\#=\left(\begin{matrix} 1 & 0 & -2 & -1 \\ 0 & 1 & -1 & -2 \\ 0 & 0 & -1 & 0 \\ 0 & 0 & 0 & -1 \end{matrix}\right) .$$ This results agrees with the one obtained by Natanzon in \cite{natan:gordon} except in that he gets all sign positive. Nevertheless, the computation of the real part of $J(\Sigma^c)$ yields the same result. We have that $P$ satisfies $\overline{P}=A-P$, where $A=\left(\begin{matrix} -2 & -1 \\ -1 & -2 \end{matrix}\right) .$ The Jacobian variety of $\Sigma^c$ is then given by $J(\Sigma^c) = {\Bbb C}^2/\Gamma^c$, where $\Gamma^c={\Bbb Z}^2+P{\Bbb Z}^2$. To compute the real part of $J(\Sigma^c)$, we write, for any $z\in{\Bbb C}^2$, $z=P\alpha+\beta$, where $\alpha,~\beta\in{\Bbb R}^2$. Then $\sigma_1(z)=\overline z\equiv z$ is equivalent to $P\alpha+\beta = \overline{P}\alpha+\beta+Pn+m,$ for some $n,~m\in{\Bbb Z}^2$. The imaginary part of this equation gives $({\mathrm Im} P)\alpha = -({\mathrm Im} P)\alpha + ({\mathrm Im} P)n$. By Riemann bilinear relations (\cite{fk:book}; see also \cite{simha:rs} for a nice introduction to Jacobians and of Riemann surfaces from an algebro-geometric approach) we have that $({\mathrm Im} P)$ is invertible, so we obtain $\alpha=n/2$. On the other hand, taking real parts in the above equation we get $({\mathrm Re} P)\alpha+\beta = (A-{\mathrm Re} P)\alpha+\beta+({\mathrm Re} P)n+m$. Since $\overline P = A-P$, we have that $({\mathrm Re} P)=A-({\mathrm Re} P)$, so this equation reduces to $0=\frac{1}{2}An+m$, or equivalently $$\left\{\begin{array}{ccc} n_1+\frac{n_2}{2}+m_1 & = & 0 \\ \frac{n_1}{2}+n_2+m_2 & = & 0, \end{array}\right .$$ where notation the should be clear. This implies that $n_j\in 2{\Bbb Z}$, so $\alpha\in{\Bbb Z}^2$. Therefore, the set of fixed points of $\sigma_1$ is given by the real torus ${\mathrm Re}(J(\Sigma^c))=\{P{\Bbb Z}^2+\beta;~\beta\in{\Bbb R}^2\}/\Gamma^c$. which agrees with the results obtained by Silhol \cite[pgs. $349$ and 359]{silhol:comess} and Natanzon \cite{natan:klein}. By the form of the lattice $\Gamma^c$, it is clear that ${\mathrm Re}(J(\Sigma^c))\cong ({\Bbb R}/{\Bbb Z})^2$. In a similar way to the construction of $J(\Sigma^c)$, we can form a lattice in ${\Bbb R}^2$ using the basis ${\cal B}_h=\{\phi_1,\phi_2\}$ and ${\cal B}$ of ${\cal H}_{\Bbb R}^1(\Sigma)$ and ${\mathrm H}_1(\Sigma,{\Bbb Z})_f$, respectively. Let us denote this lattice by $\Gamma$. We define the {\bf Jacobian variety} of $\Sigma$ as the quotient $J(\Sigma)={\Bbb R}^2/\Gamma$. It is clear that $[z]\mapsto[-\frac{1}{2}z]$ induces an isomorphism between ${\mathrm Re}(J(\Sigma^c))$ and $J(\Sigma)$, which proves our result for $g=3$. The general case of a surface of odd genus is done in a similar way. To see the even genus case, we take a surface with $g=4$, and we choose the second of the two presentations of the fundamental group of $\Sigma$ given in \S $2$; i.e. the generators are the loops $c$, $d$, $a$ and $b$, and the relation is $cdc^{-1}d[ab] = 1$. To construct the complex double we proceed as in \S $3$; we do not include the computations here, since they are done as in \S $3$. We get that the fundamental group of $\Sigma^c$ is generated by the loops $$\alpha_1=c_2c_1, ~\alpha_2=(d_1^{-1}c_2)b_1(d_1^{-1}b_1c_2)^{-1}, {}~\alpha_3=a_2,$$ $$\beta_1=d_1^{-1}, ~\beta_2=(d_1^{-1}c_2)a_1(d_1^{-1}b_1c_2)^{-1}, {}~\beta_3=b_2,$$ We again change our basis of ${\mathrm H}_1(\Sigma^c,{\Bbb Z})$ to $${\cal B}^c=\{\gamma_1=\alpha_1,~ \gamma_2=-(\alpha_3+\beta_2),~ \gamma_3=-(\alpha_2+\beta_3),~ \delta_1=\alpha_1+\beta_1,~ \delta_2=\alpha_2+\alpha_3+\beta_2,~$$ $$\delta_3=\alpha_2+\alpha_3+\beta_3\}.$$ It is not hard to see that ${\cal B}^c$ is symplectic; one simply has to chech that the matrix $$C=\left(\begin{matrix} 1 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & -1 & 0 & 1 & 1 \\ 0 & -1 & 0 & 0 & 1 & 1 \\ 0 & 0 & 0 & 1 & 0 & 0 \\ 0 & 0 & -1 & 0 & 1 & 0 \\ 0 & -1 & 0 & 0 & 0 & 1 \end{matrix}\right) ,$$ satisfies $C^tJC=J$. The projection $\pi$ acts on these loops by $\pi_\#(\gamma_1)=2c$, $\pi_\#(\gamma_2)=-2a$, $\pi_\#(\gamma_1)=-2b\,$; so we take a basis $\{\phi_1, \phi_2, \phi_3\}$ of ${\cal H}_{\Bbb R}^1(\Sigma)$, and normalize it by requiring $$\int_c\phi_1 = -\int_a\phi_2 = -\int_b\phi_3 = \frac{1}{2}, $$ and the other integrals equal to $0$. As in the previous situation, we have that if $\psi_j=\pi^*(\phi_j)$, then ${\cal B}^* = \{ \omega_j=\psi_j+i*\psi_j;~ j=1,2,3\}$ is a basis of holomorphic $1$-forms dual to ${\cal B}^c$. It is not hard to see that $\sigma_\#$ is given by the following matrix, when computed with respect to ${\cal B}^c$: $$\sigma_\#=\left(\begin{matrix} 1 & 0 & 0 & 2 & 0 & 0 \\ 0 & 1 & 0 & 0 & -2 & -1 \\ 0 & 0 & 1 & 0 & -1 & -2 \\ 0 & 0 & 0 & -1 & 0 & 0 \\ 0 & 0 & 0 & 0 & -1 & 0 \\ 0 & 0 & 0 & 0 & 0 & -1 \end{matrix}\right) ,$$ and the period matrix $P$ satisfies $\overline P=A-P$, where $$A=\left(\begin{matrix} 2 & 0 & 0 \\ 0 & -2 & -1 \\ 0 & -1 & -2 \end{matrix}\right).$$ The above matrix of $\sigma_\#$ is different from the one given in \cite{natan:gordon}; we have not been able to obtain the matrix of that reference, but nevertheless, we obtain similar result in the computation of the real part of $J(\Sigma^c)$. In this case we have that the real part of $J(\Sigma^c)$ (which is found in the same way that the $g=3$ case) has two components, namely $T_1=\{{\Bbb Z}^3n+\beta;~\beta\in{\Bbb R}^3,~ n=(n_1,n_2,n_3),~ n_1,n_2,n_3\in{\Bbb Z}\}$ and $T_2=T_1+{\Bbb Z}^3(\frac{1}{2},0,0)^t$. We again obtained the results of \cite{silhol:comess} and \cite{natan:klein}. An isomorphism similar to the previous case holds in this situation, except that we have $J(\Sigma)$ is isomorphic to any of the two sets $T_1$ or $T_2$. {\bf Remark}: the results about the fixed points of $\sigma_1$ on $J(\Sigma^c)$ can also be obtained from the expression of the period matrix given in \cite[Proposition $4$, pg. 351]{silhol:comess}. The above results can be put together in the following theorem: \begin{thm} Let $\Sigma$ be a compact non-orientable surface of genus $g\geq3$. Then we can associate to $\Sigma$ a real torus of dimension $g-1$, the {\bf Jacobian variety} $J(\Sigma)$ of $\Sigma$, such that $J(\Sigma)$ is isomorphic to any component of the real part of the Jacobian $J(\Sigma^c)$ of the complex double. This last set is defined as the set of fixed points of the symmetry $\sigma_1$ of $J(\Sigma^c)$ induced by $\sigma$. \end{thm}
1996-07-23T14:52:20
9607
alg-geom/9607023
en
https://arxiv.org/abs/alg-geom/9607023
[ "alg-geom", "math.AG" ]
alg-geom/9607023
Stefan Mueller-Stach
Pedro Luis del Angel and Stefan M\"uller-Stach
Motives of uniruled 3-folds
Latex with amsfonts, 12 pages
null
null
null
null
We construct projectors in the ring of correspondences of a complex uniruled 3-fold $X$ which lift the Kuenneth components of the diagonal in singular cohomology and have other properties which were conjectured by J. Murre. Such Murre decompositions have been already obtained for curves, surfaces, abelian varieties and varieties with cell decompositions by the work of Manin, Shermenev, Beauville, Murre et.al.. In particular they define a natural filtration on the Chow groups of $X$ which was conjectured by Bloch and Beilinson. To do this we use Mori theory and construct projectors in the situation of a fiber space over a surface. These projectors may also be used in more general situations.
[ { "version": "v1", "created": "Tue, 23 Jul 1996 12:39:46 GMT" } ]
2014-10-24T00:00:00
[ [ "del Angel", "Pedro Luis", "" ], [ "Müller-Stach", "Stefan", "" ] ]
alg-geom
\section{Introduction} Let $F$ be a subfield of ${\Bbb C}$. We denote by $V(F)$ the category of smooth, projective varieties over $F$ with the usual morphisms. Let $CV(F)$ be the category with the same underlying object, but where the morphisms are replaced by correspondences of degree zero, i.e. for two irreducible varieties $X,Y$ we have $Mor(X,Y):=CH^{\dim(X)}(X \times Y)$. If $f \in Mor(X,Y)$ we view it as a homomorphism $f_\ast: CH^*(X) \to CH^*(Y)$, by defining $f_*(W)=(pr_2)_* ( (W \times X) \cap f) $. Given $X_1,X_2,X_3 \in V(k)$ the composition of correspondences $f \in Mor(X_1,X_2)$ and $g \in Mor(X_2,X_3)$ is defined by $$ g \circ f = (pr_{13})_*\{( pr_{12})^* f \cap (pr_{23})^* g \} $$ An element $p \in Mor(X,X)$ is called a {\bf projector} if $p \circ p = p$. A special example is the diagonal, denoted by $\Delta$. Finally denote by $M(F)$ the category of {\bf effective Chow motives}, where objects are pairs $(X,p)$ with $X \in V(F)$ and $p \in Mor(X,X)$ a projector. The morphisms are described by $Mor((X,p),(Y,q)):=q \circ Mor(X,Y) \circ p$.\\ \begin{definition} Let $M=(X,p) \in M(F)$. Define $$ CH^i(M):= p_*CH^i(X) \otimes {\Bbb Q} $$ \end{definition} \begin{definition} Let $X \in V(F)$ be a smooth projective variety of dimension $d$. We say that $X$ has a {\bf Murre decomposition}, if there exist projectors $p_0,p_1,...,p_{2d}$ in $CH^d(X \times X) \otimes {\Bbb Q}$ such that the following properties hold (modulo rational equivalence for (1) and (2)):\\ (1) $p_j \circ p_i = \delta_{i,j} \cdot p_i $\\ (2) $\Delta = \sum p_i $\\ (3) In cohomology the $p_i$ induce the $(2d-i,i)-$th K\"unneth component of the diagonal.\\ (4) $p_0,...,p_{j-1}$ and $p_{2j+1},..,p_{2d}$ act trivially on $CH^j(X) \otimes {\Bbb Q}$.\\ (5) If we put $F^0 CH^j(X) \otimes {\Bbb Q} = CH^j(X) \otimes {\Bbb Q}$ and inductively $F^k CH^j(X) \otimes {\Bbb Q} := Ker(p_{2j+1-k} \mid_{F^{k-1}}) $, then this descending filtration does not depend on the particular choice of the $p_i$.\\ (6) Always $F^1 CH^j(X) \otimes {\Bbb Q} = CH^j_{hom}(X) \otimes {\Bbb Q}$. \end{definition} The motives $(X,p_i)$ are traditionally denoted by $h^i(X)$ and we write $h(X)=h^0(X)+...+h^{2d}(X)$. In (6) one also wants to have that $F^2 CH^j(X) \otimes {\Bbb Q}$ is the kernel of the cycle class map in rational Deligne cohomology, but this is very hard to verify in general. \\ (1) - (6) have been proved for curves, surfaces (\cite{6}), abelian varieties (\cite{7}) and certain varieties close to projective varieties. Recently B. Gordon and J. Murre \cite{GM} computed the Chow motive of elliptic modular varieties using work of A. Scholl \cite{Tony}.\\ S. Saito has proposed a filtration in \cite{8} which has property (6). Manin (\cite{3}) and Murre (\cite{6}) have quite generally defined $p_0,p_1,p_{2d-1},p_{2d}$ for every $X$. A. Scholl has refined this in \cite{Tony} to have also the property that $p_1=p_{2d-1}^t$, where $p^t$ denotes a transpose of a projector $p$. Murre has formulated the following \\ \ \\ {\bf Conjecture:} {\sl Every smooth projective F-variety $X$ admits a Murre decomposition.}\\ J. Murre has also studied the case of a product of a curve with a surface where one in fact has a Murre decomposition. Inspired by this, we have tried to construct projectors in the following situation: Let $f:Y \to S$ be a morphism from a smooth 3-fold $Y$ to a smooth surface $S$ with connected fibers. Choose a smooth hyperplane section $i: Z \hookrightarrow Y$ and let $h=f|_Z$. Look the following cycles $$\pi_{i0}:={1 \over m} (i \times 1)_* (h\times f)^*\pi _i(S),$$ $$\pi_{i2}:={1 \over m} (1 \times i)_* (f\times h)^*\pi _i(S),$$ in $CH^3(Y\times Y)\otimes{\Bbb Q}$. Here the $\pi_i(S)$ are orthogonal projectors of a Murre decomposition of $S$ as constructed by Murre (\cite{6}) and $m$ is the number of points on a general fiber of $h$. These cycles are not orthogonal in general but we are able to construct orthogonal projectors $\pi_0,\cdots, \pi_6$ in the following way: $$\pi_0:=\pi_{00},\quad \pi_1:=\pi_{10}, \quad \pi_2:=\pi_{20}+\pi_{02}-\pi_{20}\cdot\pi_{02}$$ $$\pi_4:=\pi_{40}+\pi_{22}-\pi_{40}\cdot\pi_{22}, \quad \pi_5:=\pi_{32}, \quad \pi_6:=\pi_{42}, \quad \pi_3:=\Delta-\sum_{i\ne 3}\pi_i $$ The $\pi_j$ do not operate in the right way on cohomology, but if all higher direct images sheaves $R^if_*{\cal O}_Y$ vanish for $i \ge 1$, they can be modified to form a Murre decomposition. In particular a suitable blow up $Y$ of any smooth {\bf uniruled} 3-fold $X$ over a subfield of the complex numbers has this property. Recall that a 3-fold $X$ is called uniruled, if there exists a dominant rational map $\varphi: S \times {{\Bbb P}}^1 - - - \to X$ for some smooth projective surface $S$. By a theorem of Mori and Miyaoka (\cite{3}), this is equivalent to saying that $X$ has Kodaira dimension $-\infty$. There is no structure theorem for these varieties which is as simple as in the case of ruled surfaces, but there is a version in the category of 3-folds with ${\Bbb Q}$-factorial and terminal singularities (\cite{4}) stating that $X$ is birationally equivalent to a 3-fold $Y$ which has a fiber structure with rationally connected fibers over a base variety which can be a point, a smooth curve or a normal surface. Using this and suitable modifications of the projectors above we can therefore prove: \begin{thm} Let $X$ be a smooth uniruled complex projective 3-fold. Then $X$ admits a Murre decomposition. \end{thm} In the proof of this theorem, which makes heavy use of Fulton's machinery of intersection theory, the Murre decomposition provides the following description of the {\bf Chow motive } of a complex uniruled 3-fold $X$ (ignoring torsion):\\ \[ \begin{tabular}{|l|c|c|c|c|c|c|c|c|c|} \hline Motive $M$ & $h^0(X)$ & \ $h^1(X)$ \ & \ $h^2(X)$ \ & \ $h^3(X)$ \ & \ $h^4(X)$ \ & \ $h^5(X)$ \ & $h^6(X)$ \ \\ \hline $CH^0(M)$ & ${\Bbb Z}$ & \ 0 \ & \ 0 \ & \ 0 \ & \ 0 \ & \ 0 \ & $0$ \ \\ \hline $CH^1(M)$ & $0$ & ${\rm Pic}^0(X)$ & \ ${\rm NS}(X)$ \ & \ $0$ \ & \ $0$ \ & \ $0$ \ & $0$ \ \\ \hline $CH^2(M)$ & $0$ & \ $0$ \ & \ ${\rm Ker}(\psi)$ \ & \ ${\rm Im}(\psi)$ \ & $H^{2,2}(X,{\Bbb Z})$ & \ $0$ \ & $0$ \ \\ \hline $CH^3(M)$ & $0$ & \ $0$ \ & \ $0$ \ & \ $0$ \ & ${\rm Ker}(alb_X)$ & \ $Alb(X)$ \ & ${\Bbb Z}$ \ \\ \hline \end{tabular} \]\\ \ \\ Here $\psi: CH^2_{\rm hom}(X) \to J^2(X)$ is the Abel-Jacobi map. We hope that our approach may also be used to construct projectors in other situations. \\ {\bf Acknowledgements}: It is a pleasure to thank H. Esnault, B. Gordon, J. Murre and E. Viehweg for several discussions. The DFG and the universities of Essen and Leiden have supported the authors during this project. \newpage \section{\bf Projectors for special varieties} \bigskip The easiest case in which one has a Murre decomposition is the case of projective space, because there $H^{2k+1}(X,{\Bbb C})=0$ for all $k\ge 0$ and the other groups admit a basis represented by algebraic cycles. One has a more general theorem:\\ \begin{thm} \label{thm1} Let $X$ be a smooth variety of dimension $n$ and assume that for certain $1\le q\le n-1$ there is a basis $\{E_1,\cdots ,E_t\}$ of $H^{2q}(X,{\Bbb Q})$ and a basis $\{\ell_1, \cdots ,\ell_t\}$ of $H^{2(n-q)}(X,{\Bbb Q})$ represented by classes of algebraic cycles. Then: \\ a) There exists a matrix $B=(b_{ij})\in {\bf GL_n}({\Bbb Q})$ such that the cycle $p=\sum b_{ij}(\ell_i\times E_j)\in CH^n(X\times X)\otimes{\Bbb Q}$ operates as the identity on $H^{2q}(X,{\Bbb Q})$.\\ b) For the same choice of $b_{ij}$, $p^t=\sum b_{ij}(E_j\times \ell_i) \in CH^n(X\times X)\otimes{\Bbb Q}$ operates as the identity on $H^{2(n-q)}(X,{\Bbb Q})$.\\ c) Both cycles, $p$ and $p^t$ are idempotent and therefore projectors. \end{thm} {\noindent{\bf Proof. }} Let $A=(E_i\cdot \ell_j)$ be the intersection matrix, then take $B=A^{-1}$. {\hfill $\square$\\} Moreover, one can explicitely say how these projectors operate on cycles, namely:\\ \begin{prop} Let $p$ be as before and let $k\ne q$. Then, for all $Z\in CH^k(X)\otimes{\Bbb Q}$ one has $p(Z)=0$ as an element of $CH^k(X) \otimes {\Bbb Q}$. \end{prop} {\noindent{\bf Proof. }} By dimension reasons, as $p(Z)\in <E_i>\subset CH^q(X)\otimes{\Bbb Q}$. {\hfill $\square$\\} \begin{lemma} Let $p$ be as before and $Z\in CH^q(X)\otimes{\Bbb Q}$. If $[Z]$ denotes the homology class of $Z$ on $H^{2q}(X,{\Bbb Q})$, then $[p(Z)]=p([Z])=[Z]$. \end{lemma} {\noindent{\bf Proof. }} $p$ operates as the identity on $H^{2q}(X,{\Bbb Q})$ and $p(Z)=\sum b_{ij}(\ell_i\cdot Z)E_j$. {\hfill $\square$\\} \begin{cor} Let $p$ be as before, then $({\rm Ker}\; p)\cap CH^q(X)\otimes{\Bbb Q} =CH_{\rm hom}^q(X)\otimes{\Bbb Q}$. \end{cor} {\bf Examples:} Smooth Fano 3-folds and Calabi-Yau 3-folds have the property that the Hodge numbers $h^{i,0}$ are always zero for $i=1,2$ and therefore theorem \ref{thm1} applies. Another example is a del Pezzo fibration $f: X \to B$: to illustrate this, let $\ell$ be the extremal rational curve, $F$ a general fiber, $Y$ be a section of $|-mK_X|$, $C$ a twofold intersection in the linear system $|Y|$ and hence a multisection of $f$ over $B$, such that $C$ is a smooth curve dominating $B$. $H^2(X,{\Bbb Q})$ is free of rank two. Then theorem \ref{thm1} produces the following projector $$p_2:= {1 \over r}(C \times F)+{1 \over m }({\ell} \times Y)- {d \over {m \cdot r}}({\ell} \times F) $$ where $d=Y^3$ and $r:=(C.F)$. Note that $(-K_X.\ell)=1$. $p_2$ is unique as a cycle up to the choices of $Y,C,F$ and $\ell$. \section{\bf Murre decompositions of birational conic bundles} \bigskip Let $f:Y\longrightarrow S$ be a morphism from a smooth projective 3-fold $Y$ to a smooth projective surface $S$, such that every fiber of $f$ is rationally connected and the general fiber of $f$ is isomorphic to ${{\Bbb P}}^1$. Choose a smooth hyperplane section $i:Z\hookrightarrow Y$ such that $h:=f_{|Z}:Z\longrightarrow S$ is surjective and generically finite. Then define cycles $$\pi_{i0}:={1 \over m} (i \times 1)_* (h\times f)^*\pi _i(S),$$ $$\pi_{i2}:={1 \over m}(1 \times i)_* (f\times h)^*\pi _i(S),$$ in $CH^3(Y\times Y)\otimes{\Bbb Q}$ for $0\le i\le 4$. Here the $\pi_i(S)$ are the orthogonal projectors of a Murre decomposition of $S$ as constructed by Murre (\cite{6}) (and improved by A. Scholl in \cite{Tony} to have also the property that $\pi_1=\pi_3^t$) and $m$ is the number of points on a general fiber of $h$. The following is our {\bf key result} in some sense: \begin{lemma}\quad \\ a) $\pi_{i0}\circ \pi_{j0} = \delta_{ij}\pi_{i0}$ \\ b) $\pi_{i2}\circ \pi_{j2} = \delta_{ij}\pi_{i2}$ \\ c) $\pi_{j2}\circ \pi_{i0} = 0$ \end{lemma} {\noindent{\bf Proof. }} a) Using the projection formula and the theory of Gysin maps for l.c.i. morphisms from \cite[prop.6.6 (c)]{Fu} in the following diagram $$ \matrix{Y \times Y \times Y & \to & Y \times Y \cr \downarrow && \downarrow \cr Z \times Y \times Y & \to& Z \times Y \cr \downarrow && \downarrow \cr Z\times S \times Y &\to& Z \times Y \cr \downarrow && \downarrow \cr S \times S \times S & \to & S \times S } $$ where the vertical maps are canonical l.c.i. morphisms, one obtains: \\ \begin{small} $\pi_{i0}\circ \pi_{j0}$\\ $ ={1 \over m^2}(pr_{13}^{Y\times Y\times Y})_* ((i\times 1)_*((h\times f)^*(\pi_j(S))\times Y \cap Y\times (i\times 1)_*((h\times f)^*(\pi_i(S))))$\\ $={1 \over m^2}(pr_{13}^{Y\times Y\times Y})_* ((i\times 1\times 1)_*(h\times f\times f)^*(\pi_j(S)\times S) \cap (1\times i\times 1)_*(f\times h\times f)^*(S\times \pi_i(S)))$\\ $={1 \over m^2}(pr_{13}^{Y\times Y\times Y})_* (i\times 1\times 1)_*[(h\times f\times f)^*(\pi_j(S)\times S) \cap (i\times 1\times 1)^*(1\times i\times 1)_*(f\times h\times f)^* (S\times \pi_i(S))]$\\ $={1 \over m^2}(i\times 1)_*(pr_{13}^{Z\times Y\times Y})_* [(h\times f\times f)^*(\pi_j(S)\times S) \cap (1\times i\times 1)_*(i\times 1\times 1)^*(f\times h\times f)^* (S\times \pi_i(S))]$\\ $={1 \over m^2}(i\times 1)_*(pr_{13}^{Z\times Y\times Y})_* [(h\times f\times f)^*(\pi_j(S)\times S) \cap (1\times i\times 1)_*(h\times h\times f)^*(S\times \pi_i(S))]$\\ $={1 \over m^2}(i\times 1)_*(pr_{13}^{Z\times Y\times Y})_*(1\times i\times 1)_* [(1\times i\times 1)^*(h\times f\times f)^*(\pi_j(S)\times S) \cap (h\times h\times f)^*(S\times \pi_i(S))]$\\ $={1 \over m^2}(i\times 1)_*(pr_{13}^{Z\times S\times Y})_*(1\times h\times 1)_* (h\times h\times f)^*[(\pi_j(S)\times S) \cap (S\times \pi_i(S))]$\\ $={1 \over m}(i\times 1)_*(pr_{13}^{Z\times S\times Y})_* (h\times 1\times f)^*[(\pi_j(S)\times S) \cap (S\times \pi_i(S))]$ \hfill (\cite[prop. 6.6 (c)]{Fu}) \\ $={1 \over m}(i\times 1)_*(h\times f)^*(pr_{13}^{S\times S\times S})_* (\pi_j(S)\times S) \cap (S\times \pi_i(S))$\\ $={1 \over m}(i\times 1)_*(h\times f)^* (\pi_i(S) \circ \pi_j(S))=\delta_{ij}\pi_{i0}$.\\ \end{small} Similarly one proves b).\\ c) As before, one finds that\\ \begin{small} $\pi_{j2}\cdot \pi_{i0}$\\ $={1 \over m^2}(i\times i)_*(pr_{13}^{Z\times S\times Z})_* (1\times f\times 1)_* [(1\times f\times 1)^*(h\times 1\times h)^* (\pi_i(S)\times S\cap S\times \pi_j(S))\cap (Z\times Y\times Z)]$\\ $={1 \over m^2}(i\times i)_*(pr_{13}^{Z\times S\times Z})_* [(h\times 1\times h)^*(\pi_i(S)\times S\cap S\times \pi_j(S))\cap (1\times f\times 1)_*(Z\times Y\times Z)]= 0$\\ \end{small} because $(1\times f\times 1)_*(Z\times Y\times Z)= 0$ due to dimension reasons. {\hfill $\square$\\} \ \\ Define now a set of cycles $\pi_0,\cdots,\pi_6$ in the following way:\\ \begin{small} $\pi_0:=\pi_{00}$\\ $\pi_1:=\pi_{10}$\\ $\pi_2:=\pi_{20}+\pi_{02}-\pi_{20} \circ \pi_{02}$\\ $\pi_4:=\pi_{40}+\pi_{22}-\pi_{40} \circ \pi_{22}$\\ $\pi_5:=\pi_{32}$\\ $\pi_6:=\pi_{42}$\\ $\pi_3:=\Delta-\sum_{i\ne 3}\pi_i$ \end{small} \begin{cor} The $\pi_j$ defined above form a set of orthogonal projectors such that $\pi_k=\pi_{6-k}^t$ . \end{cor} \begin{thm} \label{thm2} $\pi_i=\delta_{ij}$ on $\cases{f^*H^j(S,{\Bbb Q}) \quad if \quad j = 0,1 \cr f^*H^j(S,{\Bbb Q}) \oplus {\Bbb Q}\cdot [Z] \quad if \quad j=2 \cr f^*H^j(S,{\Bbb Q}) \oplus [Z]\cdot f^*H^2(S,{\Bbb Q}) \quad if \quad j=4 \cr [Z]\cdot f^*H^3(S,{\Bbb Q}) \quad if \quad j=5 \cr [Z]\cdot f^*H^4(S,{\Bbb Q}) \quad if \quad j=6}$ \end{thm} {\noindent{\bf Proof. }} First note that one has the equation: $\pi_{i0}(f^*\alpha)= {1 \over m}(i \times 1)_* (h\times f)^*\pi _i(S)(f^*\alpha)$\\ $={1 \over m}(pr^{Y \times Y}_2)_*[(i \times 1)_* (h\times f)^*\pi _i(S)\cap (f^*\alpha\times Y)]$\\ $ ={1 \over m}(pr^{Y \times Y}_2)_*(i \times 1)_* [(h\times f)^*\pi _i(S)\cap (i\times 1)^*(f^*\alpha\times Y)]$\\ $ ={1 \over m}(pr^{Y \times Y}_2)_*(i \times 1)_* (h\times f)^*[\pi _i(S)\cap \alpha\times S]$\\ $ ={1 \over m}(pr^{Z \times Y}_2)_* (h\times f)^*[\pi _i(S) \cap \alpha\times S]$\\ $ ={1 \over m}(pr^{S \times Y}_2)_* (h\times 1)_*(h\times f)^* [\pi _i(S) \cap \alpha \times S)]$\\ $ =(pr^{S \times Y}_2)_*(1\times f)^*[\pi _i(S)\cap \alpha\times S)]$\\ $ =f^*(pr^{S \times S}_2)_*[\pi _i(S)\cap \alpha\times S] =f^*\pi_i(S)(\alpha)$.\\ Therefore $\pi_{i0}$ operates as $\delta_{ij}$ on $f^*H^j(S)$, proving the assertion for $\pi_0$ and $\pi_1$.\\ On the other hand, using projection formula, one gets\\ \noindent $\pi_{i2}(f^*\alpha)={1 \over m}(pr^{Y\times Y}_2)_* [ (1 \times i)_* (f\times h)^*\pi _i(S)\cap (f^* \alpha\times Y)$ \\ $\quad\quad\quad\quad = {1 \over m} i_* (pr^{S\times Z}_2)_* (f\times 1)_*[(f\times 1)^*(1\times h)^*(\pi_i(S)\cap (\alpha\times S)) \cap (Y\times Z)]$\\ $\quad\quad\quad\quad = {1 \over m} i_* (pr^{S\times Z}_2)_* [(1\times h)^*(\pi_i(S)\cap (\alpha\times S)) \cap (f\times 1)_*(Y\times Z)]=0$,\\ since $(f\times 1)_*(Y\times Z)=0.$\\ Take any $D\in H^k(S,{\Bbb Q})$ with $k=0,2,3,4$ and consider $C:=i_*h^*(D)$. Observe that $[C]=f^*(D)\cdot [Z]$. The same computation as above in cohomology shows that $$\pi_{i2}([C])=:{1 \over m}(pr_2^{Y\times Y})_* [(1\times i)_*(f\times h)^*\pi_i(S)\cap [C]\times [Y]]= i_*h^*(\pi_i(S)(D))$$ As the $\pi_i(S)$ induce the K\"unneth decomposition of $\Delta_S$ on cohomology, it follows that $\pi_i(S)([D])=\delta_{ik}([D])$ and therefore one gets $\pi_{i2}([C])=\delta_{ik}[C]$.\\ Moreover, a similar argument together with Chow's moving lemma shows that\\ $\pi_{i0}([C])={1 \over m}(pr_2^{Y\times Y})_* [(i \times 1)_*(h \times f)^*\pi_i(S)\cap [C] \times [Y]]$\\ $={1 \over m}(pr_2^{Y\times Y})_* (i \times 1)_*[(h \times f)^*\pi_i(S)\cap (i \times 1)^* [C] \times [Y]]$\\ $={1 \over m}(pr_2^{Z \times Y})_* [(h \times f)^*\pi_i(S)\cap [C \cap Z] \times [Y]]$\\ $={1 \over m}(pr_2^{S \times Y})_*(h \times 1)_* [(h \times 1)^*(1 \times f)^* \pi_i(S)\cap [C \cap Z] \times [Y]]$\\ $={1 \over m}(pr_2^{S \times Y})_*[(1 \times f)^* \pi_i(S)\cap h_* [C \cap Z] \times [Y]]$\\ $={1 \over m} f^* (pr_2^{S \times S})_*[ \pi_i(S) \cap h_* [C \cap Z] \times [S]]$\\ $={1 \over m} f^* \pi_i(S)(h_* [C \cap Z])=0$,\\ if $i \ne k+2$. As a consequence one also gets $\pi_{i0} \circ \pi_{j2}([C])=\delta_{jk}\pi_{i0}([C])$, which proves the assertion for $\pi_2,\pi_4,\pi_5$ and $\pi_6$ and the theorem. {\hfill $\square$\\} \ \\ Now assume additionally that $f: Y \to S$ is a desingularization of a conic bundle morphism $f': X' \to S' $ in the sense of \cite{4}, i.e. there is a commutative diagram $$\matrix{ Y & {\buildrel f \over \longrightarrow} & S \cr \quad \downarrow \sigma && \quad \downarrow \tau \cr X' & {\buildrel f' \over {\longrightarrow }} & S' } $$ with blow-up morphisms $\sigma, \tau$. Also we assume $Z \subset Y$ is a sufficiently general smooth hyperplane section of $Y$ that dominates $S$. Then we can choose irreducible divisors $H_1,...,H_r$ in $Y$ such that $H_1=Z$ and $$ H^{1,1}(Y,{\Bbb Q})= \bigoplus_{i=1}^r {\Bbb Q}[H_i] $$ form a basis of $H^{1,1}(Y,{\Bbb Q})$ and such that $f_*H_i=0$ in $CH^0(S)$ for $i \ge 2$, i.e. $H_i$ is exceptional with respect to $f$ for $i \ge 2$. \begin{lemma} \label{le1} For every cycle $W$ one has $\pi_{20}(W)={1 \over m} f^* \pi_2(S) (h_* (W \cap Z)) \in f^* CH^*(S) \otimes {\Bbb Q}$. Let $W$ be a cycle with $f_*(W) =0$. Then $\pi_{02}(W)=0$ already in the Chow group of $Y$. \end{lemma} {\noindent{\bf Proof. }} $\pi_{02}(W)={1 \over m}(pr_2^{Y\times Y})_* [(1\times i)_*(f\times h)^*\pi_0(S)\cap (W \times Y) ]$\\ $={1 \over m} i_* (pr_2^{S \times Z})_*[(1 \times h )^* \pi_0(S) \cap (f \times 1)_*(W \times Z)] =0 $ \\ by \cite[prop.6.6 (c)]{Fu} and since $f_*(W)=0 \in CH^*(S)$. \\ On the other hand\\ $\pi_{20}(W)={1 \over m}(pr_2^{Y\times Y})_* [(i \times 1)_* (h \times f)^*\pi_2(S) \cap (W \times Y) ]$\\ $={1 \over m}(pr_2^{Z \times Y})_* [(h \times f)^*\pi_2(S) \cap ((W \cap Z) \times Y) ]$\\ $={1 \over m} (pr_2^{S \times Y})_*[(1 \times f)^* \pi_2(S) \cap (h \times 1)_*((W \cap Z) \times Y)] $\\ $={1 \over m} (pr_2^{S \times Y})_*(1 \times f)^*[ \pi_2(S) \cap h_*(W \cap Z) \times S)] $\\ $={1 \over m} f^* (pr_2^{S \times S})_* [\pi_2(S) \cap h_*(W \cap Z) \times S ] $\\ $={1 \over m} f^* \pi_2(S)(h_*(W \cap Z)) \in f^* CH^*(S) \otimes {\Bbb Q}$. {\hfill $\square$\\} \begin{cor} $\pi_2(Y)(H_i)={1 \over m} f^*(h_* (H_i \cap Z)) \in f^* CH^1(S) \otimes {\Bbb Q} $ for $i \ge 2$. \end{cor} By theorem \ref{thm2} $\pi_2(Y)$ operates as zero on ${\rm Pic}^0(Y)$, therefore the image of $\pi_2(Y)$ in $CH^1(Y) \otimes {\Bbb Q}$ is a finite dimensional vector space. By changing our generators $H_i$ above modulo classes in ${\rm Pic}^0(Y)= f^*{\rm Pic}^0(S)$, we may assume that they generate ${\rm Im}(\pi_2) \subset CH^1(Y) \otimes {\Bbb Q}$. Then we write uniquely $$\pi_2(Y)(H_i)=\sum_k a_{i,k} H_k \in CH^1(Y) \otimes {\Bbb Q}$$ with a matrix $A=(a_{i,k}) \in {\rm Mat}(r \times r,{\Bbb Q})$. $\pi_2(Y)$ being a projector implies that $A^2=A$. Choose algebraic cycles $\ell_1,...,\ell_r $ such that $\ell_1=F$, a general fiber of $f$, and such that their cohomology classes form a basis of $H^{2,2}(Y,{\Bbb Q})$. By Poincar\'e duality the intersection matrix $M=(m_{i,j}):= (\ell_1,...,\ell_r)^{T}(H_1,...,H_r)$ has nonzero determinant.\\ We define $$ q_2:= \pi_2(Y) + \sum b_{i,j}(\ell_i \times H_j) - \sum b_{i,j}(\ell_i \times H_j) \circ \pi_2 $$ with some matrix $B=(b_{i,j}) \in {\rm Mat}(r \times r,{\Bbb Q})$.\\ \begin{lemma}{\label{le2}} If $B=M^{-1}({\bf 1} -A)$, then $q_2$ is a projector and operates as the identity on $H^2(Y,{\Bbb Q})$. \end{lemma} {\noindent{\bf Proof. }} $\pi_2$ acts as the identity on $f^*H^2(S,{\Bbb Q})$ by theorem \ref{thm2}. The higher direct images $R^if_*{\cal O}_Y$ vanish for $i \ge 1$ by \cite{4}. Therefore by the Leray spectral sequence $H^2(Y,{\cal O}_Y)=f^*H^2(S,{\cal O}_S)$ and it is enough to show that $q_2$ operates as the identity on $H^{1,1}(Y,{\Bbb Q})$ too. But $q_2$ acts via the matrix $MB + A +BA$ on $H^{1,1}(Y,{\Bbb Q})$ with respect to the basis $\{H_i\}$. Now $\pi_2^2=\pi_2$ and we get $A^2=A$ and therefore $BA=0$. By definition of $B$, we obtain that $MB + A +BA= M(M^{-1}({\bf 1}-A)) +A={\bf 1}$. \\ To show that $q_2$ is a projector, let us write $q_2=\pi_2+ \beta - \beta \pi_2$. Note that $\beta \beta = \beta$, since $BMB=B$. From $BA=0$ we deduce that $\pi_2 \beta=0$. Therefore\\ $q_2 \circ q_2= \pi_2^2 +\beta^2 + \beta \pi_2 \beta \pi_2 + \pi_2 \beta - \pi_2 \beta \pi_2 + \beta \pi_2 - \beta \beta \pi_2 - \beta \pi_2 \pi_2 -\beta \pi_2 \beta = \pi_2 +\beta -\beta \pi_2 = q_2$ is a projector. {\hfill $\square$\\} \begin{thm} \label{thm3} The following cycles $ p_0(Y):=\pi_0(Y), \ p_1(Y):=\pi_1(Y) $, \\ $ p_2(Y):= q_2 -\pi_1(Y) \circ \sum b_{i,j}(\ell_i \times H_j) -\pi_1(Y) \circ \sum b_{i,j}(\ell_i \times H_j) \circ \pi_2(Y) $\\ $ p_4:= p_2^{tr}(Y), \quad p_5(Y):= \pi_5(Y), \quad p_6(Y):=\pi_6(Y), \quad p_3(Y):=\Delta - \sum_{i \ne 3} p_i $\\ define orthogonal projectors, which satisfy properties (1)-(6) of a Murre decomposition. \end{thm} {\noindent{\bf Proof. }} By lemma \ref{le2} above, (1),(2) and (3) are straightforward. \\ To prove (4),(5) and (6) for $j=1$, note that ${\rm Pic}(Y) \otimes {\Bbb Q} =f^*{\rm Pic}^0(S) \otimes {\Bbb Q} \oplus \bigoplus_i {\Bbb Q} \cdot H_i$. By theorem \ref{thm2} above, $p_1$ operates on $Pic^0(Y) \otimes {\Bbb Q}=f^*Pic^0(S) \otimes {\Bbb Q}$ as the identity and trivially on $\bigoplus_i {\Bbb Q} \cdot H_i$. Vice versa $p_2$ is the identity on $\bigoplus_i {\Bbb Q} \cdot H_i$ and zero on $f^*Pic^0(S) \otimes {\Bbb Q}$, because it acts trivially on $f^*H^1(S,{\Bbb Q})$. All the other projectors are zero on $CH^1(Y) \otimes {\Bbb Q}$. Therefore we get (4)-(6) for $j=1$ with $F^2CH^1(Y) \otimes {\Bbb Q}=0$.\\ For $j=2$, property (4) follows from the analogous assertion for $S$. By construction $F^1 CH^2(Y) \otimes {\Bbb Q}={\rm Ker}(p_4)= CH^2_{\rm hom}(Y) \otimes {\Bbb Q}$. Then $F^2CH^2(Y) \otimes {\Bbb Q}={\rm Ker}(p_3)\cap{\rm Ker}(p_4)={\rm Im}(p_2)= {\rm Im}(\pi_2(Y))$.\\ Now we show that $F^2CH^2(Y) \otimes {\Bbb Q} \cong f^* F^2 CH^2(S) \otimes {\Bbb Q}$: $\pi_{02}$ operates as zero on $CH^2(Y)$ by Chow's moving lemma and if $C$ is any curve homologous to zero on $Y$, then by Lemma \ref{le1}, $\pi_{20}(C)= f^* h_*(C \cap Z) \in f^* F^2 CH^2(S) \otimes {\Bbb Q}$.\\ This proves that $F^2CH^2(Y) \otimes {\Bbb Q} \subset f^* F^2CH^2(S) \otimes {\Bbb Q}$, but since $\pi_2(Y)$ operates as the identity on every fiber of $f$, we get equality. This is then independent of all choices, because this is the case for $F^2 CH^2(S)$ by \cite{6}. Finally $F^3 CH^2(Y) \otimes {\Bbb Q}=0$, since $p_2$ acts as the identity on $F^2CH^2(Y) \otimes {\Bbb Q} ={\rm Im}(p_2)$. Hence we get (5) and (6) for $j=2$.\\ Finally consider $CH^3(Y)$: Clearly $F^1CH^3(Y) \otimes {\Bbb Q}=Ker(\pi_6)= CH^3_{\rm hom}(Y) \otimes {\Bbb Q}$. Further $F^2CH^3(Y) \otimes {\Bbb Q} = Ker (\pi_5|_{F^1CH^3(Y) \otimes {\Bbb Q}})$ and we claim that $F^2 CH^3(Y) \otimes {\Bbb Q} \cong Ker(alb_Y) \otimes {\Bbb Q}$, where $alb_Y: CH^3(Y)_{\rm hom} \to Alb(Y)$ is the Albanese map. But there is a commutative diagram $$\matrix{ CH^3(Y)_{\rm hom} & \to & Alb(Y) \cr f_* \downarrow && \downarrow f_* \cr CH^2(S)_{\rm hom} & \to & Alb(S) } $$ Both vertical maps are isomorphisms. To compute $F^2CH^3(Y) \otimes {\Bbb Q} $ we take any closed point $P$ in $Y$ and compute that $f_* \pi_5(P)= f_* {1 \over m} i_* h^* (\pi_3(S)(P))= \pi_3(S)(f_*(P))$.\\ This shows that $f_* F^2CH^3(Y) \otimes {\Bbb Q} \cong F^2 CH^2(S) \otimes {\Bbb Q} \cong Ker(alb_S) \otimes {\Bbb Q} $ by \cite{6}. Therefore $F^2CH^3(Y) \otimes {\Bbb Q} \cong Ker(alb_Y) \otimes {\Bbb Q}$, which is independent of all choices again by \cite{6}. Finally $F^3CH^3(Y) \otimes {\Bbb Q} = 0$, since if $P=\sum a_i P_i$ is a zero cycle on $Y$ with $\sum a_i=0$, then $f_* \pi_4(P)=f_* \pi_{20}^{t}(P) + f_* \pi_{02}^{t}(P) = f_*{1 \over m} (1 \times i)_*(f \times h)^*\pi_2(S) (P) + f_*{1 \over m} (i \times 1)_* (h \times f)^* \pi_4(S) (P)$. But $\pi_4(S)=S \times e$ , hence the last term is zero and the first term becomes $\pi_2(S)(f_*P)$. But $\pi_2(S)$ acts as the identity on $F^2 CH^2(S) \otimes {\Bbb Q}$. Thus $f_* F^3 CH^3(Y) \otimes {\Bbb Q} \subset F^3 CH^2(S) \otimes {\Bbb Q} = 0$. \\ This finishes the proof of the theorem. {\hfill $\square$\\} \newpage \section{Murre decompositions of uniruled 3-folds} Let $k={\Bbb C}$. By a 3-fold we just mean a normal 3-dimensional complex variety.\\ \begin{definition} A 3-fold $X$ is called {\bf uniruled}, if there exists a dominant rational map $\varphi: S \times {{\Bbb P}}^1 - - - \to X$ for some surface $S$. \end{definition} \begin{thm} (\cite{3}): A smooth projective 3-fold $X$ is uniruled if and only if it has Kodaira dimension $- \infty$, i.e. no multiple of $K_X$ has sections. \end{thm} \begin{thm} (\cite{4}): Let $X$ be a uniruled 3-fold with only ${\Bbb Q}$-factorial terminal singularities. Then there exists a birational mapping $r : X ---\to Y$ which is a composition of flips and divisorial contractions, such that $Y$ has an extremal ray $R$ whose extremal contraction map $f: Y \to Z$ satisfies one of the following cases:\\ (a) dim(Z)=0, $Y$ is a ${\Bbb Q}$-Fano 3-fold with $\rho(Y)=1$, i.e. $-mK_Y$ is an ample Cartier divisor for some $m \ge 1$ and the divisor class group is free with one generator.\\ (b) $Z$ is a smooth curve and $Y$ is a del Pezzo fibration over $Z$, i.e. the general fibre of $f$ is a del Pezzo surface.\\ (c) $Z$ is a surface with at most quotient singularities and $Y$ is a conic bundle over $Z$.\\ In cases (b) and (c) the reduced preimage of any irreducible divisor is again irreducible. \end{thm} \begin{thm} Let $X$ be a smooth complex uniruled 3-fold. Then $X$ admits a Murre decomposition. \end{thm} \bigskip {\noindent{\bf Proof. }} Since $X$ is uniruled, it is birational to one of the following varieties:\\ (a) A ${\Bbb Q}$-Fano 3-fold $Y$ with $\rho(Y)=1$, i.e. $-mK_Y$ is an ample Cartier divisor for some $m \ge 1$ and the divisor class group is free with one generator.\\ (b) A del Pezzo fibration over a smooth curve. \\ (c) A conic bundle over a normal surface with at most quotient singularities.\\ In cases (a), (b) $H^2(X,{\Bbb Q})$ and $H^4(X,{\Bbb Q})$ are generated by classes of algebraic cycles. Thus we define $p_0(X)=\{e\}\times X$ and $p_6(X)=X \times\{e\}$ for some rational point $e\in X$, $p_1(X)$ and $p_5(X)$ as in \cite{6} and $p_2(X)$ and $p_4(X)=p_2(X)^{tr}$ as in theorem \ref{thm1}. Then it is immediate to verify all properties (2)-(6) similar to the proof of \ref{thm3} while property (1) can be achieved like in \cite[remark 6.5.]{6}, by the non-commutative Gram-Schmidt process.\\ In case (c) we may assume that after blowing up $X$ along several smooth subvarieties, there is a situation as in the previous section:\\ Let $\varphi: Y \to X$ be the blow-up and assume that $f:Y \to S$ is a morphism to a smooth surface $S$ with rationally connected fibers. Take the projectors $p_0(Y),...,p_6(Y)$ as defined in the last section.\\ To define the projectors for $X$, consider the graph $\Gamma_\varphi \subset Y \times X$ of $\varphi$. Define $$p_i(X):=\Gamma_\varphi \circ p_i(Y) \circ \Gamma_\varphi^{tr}= (\varphi\times\varphi)_*(p_i(Y))$$ (by Liebermann's lemma \cite{2}) for $0 \le i \le 2$. We claim that all $p_i(X)$ are orthogonal projectors.\\ By induction on the number of blow-ups we may assume that there is just one blow-up along a smooth subvariety $W\subset X$.\\ Consider the canonical diagram $$\matrix{Y \times Y \times Y & {\buildrel pr_{13} \over \to} & Y \times Y \cr \downarrow && \downarrow \cr X \times Y \times X & {\buildrel pr_{13} \over \to} & X \times X} $$ where the vertical maps are $\varphi \times 1 \times \varphi$ and $\varphi \times \varphi$. Let $E$ be the exceptional divisor. Then we compute for $0 \le i,j \le 2$:\\ $p_i(X)\circ p_j(X)= (pr_{13})_*((\varphi\times id)_*p_j(Y)\times X\cap X\times (id\times \varphi)_*p_i(Y))=$\\ $= (\varphi\times\varphi)_* (pr_{13})_*(p_j(Y)\times Y\cap Y\times (id\times \varphi)^*(id\times \varphi)_*p_i(Y))$\\ $= (\varphi\times\varphi)_* (pr_{13})_*(p_j(Y)\times Y\cap Y\times (p_i(Y)+(id\times j)_*Q_{i,j}))$ \\ $=(\varphi\times\varphi)_* (pr_{13})_*(p_j(Y)\times Y\cap Y\times p_i(Y))+ (\varphi\times\varphi)_* (pr_{13})_*(p_j(Y)\times Y\cap Y \times (id\times j)_*Q_{i,j})$\\ $=(\varphi\times\varphi)_* (p_i(Y) \circ p_j(Y)+ (pr_{13})_*(p_j(Y)\times Y \cap Y \times (id\times j)_*Q_{i,j})$\\ where $Q_{i,j}\in CH_3(Y\times E)$ and $j: E \hookrightarrow Y$ is the inclusion. Hence \\ ${\cal C}_i:=p_i(X)\circ p_i(X)-p_i(X)= (\varphi\times id)_* (pr_{13})_*(p_i(Y)\times X \cap Y \times (id\times i)_*(id\times \varphi^E)_*Q_{i,i}))$.\\ $p_i(Y)={1 \over m} (i \times 1)_*(h \times f)^*\pi_i(S) + T_i$ with $T_0,T_1=0$ and $T_2=\sum c_{ij}(\ell_i \times H_j)- \sum b_{i,j}(\ell_i \times H_j) \circ \pi_2(Y)$ for some integers $c_{i,j}, b_{i,j}$ which is supported on $(Z \times Y) \cup (\ell_i \times Y)$. Therefore ${\cal C}_i$ is supported on $\varphi(Z) \times W$. Here $i:W\to X$ is the inclusion and $\varphi^E: E \to W$ is the restriction of $\varphi$ to $E$.\\ If $W$ is a point, ${\cal C}_i=0$ by dimension reasons. If $W$ is a curve, ${\cal C}_i=a(\varphi(Z)\times W)$ with $a \in {\Bbb Z}$. But ${\cal C}_i=p_i(X)\circ p_i(X)-p_i(X)$ operates as zero on the cohomology classes of every curve $T \in CH^2(X)$, since by Chow's moving lemma we can choose $T$ to be disjoint from $W$ and use that $p_i(Y)(T)=0$ in cohomology for $i=0,1,2$. Therefore $a=0$ and $p_i(X)$ is a projector.\\ For $ i \neq j$, $p_i(X)\circ p_j(X)= (\varphi\times\varphi)_* (pr_{13})_*(p_j(Y)\times Y \cap Y \times (id\times j)_*Q_{i,j})$\\ since $p_i(Y)$ and $p_j(Y)$ are orthogonal. As above this implies that $ p_i(X) \circ p_j(X)$ is supported on $\varphi(Z) \times W$ for all $j$. By the same argument with Chow's moving lemma for $CH^2(X)$ as before, $p_i(X)\circ p_j(X)=0$. \\ Now define $$p_4(X)=p_2(X)^{tr}, p_5(X)=p_1(X)^{tr}, p_6(X)=p_0^{tr} \quad {\rm and} \quad p_3(X)=\Delta-\sum_{i \ne 3} p_i(X)$$ Properties (3)-(6) follow from theorem \ref{thm3} together with the split exact sequences (\cite[prop. 6.7]{Fu}) $$ 0 \to CH_k(W) \to CH_k(E) \oplus CH_k(X) \to CH_k(Y) \to 0 $$ (1) and (2) can be obtained again via the Gram-Schmidt process. {\hfill $\square$\\} \bigskip
1997-03-21T17:15:49
9607
alg-geom/9607021
en
https://arxiv.org/abs/alg-geom/9607021
[ "alg-geom", "math.AG" ]
alg-geom/9607021
Miles Reid
F. Catanese, M. Franciosi, K. Hulek and M. Reid
Embeddings of curves and surfaces
LaTeX2e with packages: amstex, amssymb, theorem. 32 pp
null
null
To be issued as Univ. of Pisa preprint. Requests for copies to the second author at Pisa
null
We prove a general embedding theorem for Cohen--Macaulay curves (possibly nonreduced), and deduce a cheap proof of the standard results on pluricanonical embeddings of surfaces, assuming vanishing H^1(2K_X)=0.
[ { "version": "v1", "created": "Mon, 22 Jul 1996 10:30:14 GMT" } ]
2008-02-03T00:00:00
[ [ "Catanese", "F.", "" ], [ "Franciosi", "M.", "" ], [ "Hulek", "K.", "" ], [ "Reid", "M.", "" ] ]
alg-geom
\section{Introduction} Let $C$ be a curve over a field $k$ of characteristic $p\ge0$, and $H$ a Cartier divisor on $C$. We assume that $C$ is projective and Cohen--Macaulay (but possibly reducible or nonreduced). Write $HC=\deg\Oh_C(H)$ for the degree of $H$, $p_aC=1-\chi(\Oh_C)$ for the arithmetic genus of $C$, and $\om_C$ for the dualising sheaf (see \cite{Ha}, Chap.~III, \S7). Our first result is the following. (A {\em cluster} $Z$ of {\em degree} $\deg Z=r$ is simply a \hbox{$0$-dimensional} subscheme with $\length\Oh_Z=\dim_k\Oh_Z=r$; a curve $B$ is {\em generically Gorenstein} if, outside a finite set, $\om_B$ is locally isomorphic to $\Oh_B$. The remaining definitions and notation are explained below.) \begin{TEO}[Curve embedding theorem]\label{th:curve} $H$ is very ample on $C$ if for every generically Gorenstein subcurve $B\subset C$, either \begin{enumerate} \item $HB\ge2p_aB+1$, or \item $HB\ge2p_aB$, and there does not exist a cluster $Z\subset B$ of degree $2$ such that $\sI_Z\Oh_B(H)\iso\om_B$. \end{enumerate} More generally, suppose that $Z\subset C$ is a cluster (of any degree) such that the restriction \begin{equation} H^{0}(C,\Oh_C(H))\to\Oh_Z(H)=\Oh_C(H)\otimes\Oh_Z \label{eq:rest} \end{equation} is not onto. Then there exists a generically Gorenstein subcurve $B$ of $C$ and an inclusion $\fie\colon\sI_Z\Oh_B(H)\into\om_B$ not induced by a map $\Oh_B(H)\to\om_B$. In particular, (\ref{eq:rest}) is onto if \begin{equation} HB>2p_aB-2+\deg(Z\cap B) \nonumber \end{equation} for every generically Gorenstein subcurve $B\subset C$. \end{TEO} Theorem~\ref{th:curve} is well known for nonsingular curves $C$. Although particular cases were proved in \cite{Ca1}, \cite{Ba2}, \cite{C-F}, \cite{C-H}, it was clear that the result was more general. In discussion after a lecture on the Gorenstein case by the first author at the May 1994 Lisboa AGE meeting, the fourth author pointed out the above result, where $C$ is only assumed to be a pure 1-dimensional scheme. For divisors on a nonsingular surface, Mendes Lopes \cite{ML} has obtained results analogous to Theorem~\ref{th:curve} and to Theorem~\ref{th:hh}. We apply these ideas to the canonical map of a Gorenstein curve in \S\ref{sec:cc}. The proof of Theorem~\ref{th:curve} is based on two ideas from Serre and Grothendieck duality: \begin{enumerate} \renewcommand\labelenumi{(\alph{enumi})} \item we use Serre duality in its ``raw'' form \begin{equation} H^1(C,\sF)\dual\Hom(\sF,\om_C)\quad\text{for $\sF$ a coherent sheaf,} \nonumber \end{equation} where $\dual$ denotes duality of vector spaces. \item If $\Oh_C$ has nilpotents, a nonzero map $\fie\colon\sF\to\om_C$ is not necessarily generically onto; however (because we are $\Hom$'ming into $\om_C$), duality gives an automatic factorisation of $\fie$ of the form \begin{equation} \sF\onto\sF_{|B}\to\om_B\into\om_C, \nonumber \end{equation} via a purely 1-dimensional subscheme $B\subset C$, where $\sF_{|B}\to\om_B$ is generically onto. See Lemma~\ref{lem:adj} for details. \end{enumerate} Since our main result might otherwise seem somewhat abstract and useless, we motivate it by giving a short proof in \S\ref{sec:pluri}, following the methods of \cite{C-F}, of the following result essentially due to Bombieri (when $\chara k=0$) and to Ekedahl and Shepherd-Barron in general. Recall that a {\em canonical surface} (or canonical model of a surface of general type) is a surface with at worst Du Val singularities and $K_X$ ample. The remaining notation and definitions are explained below. \begin{TEO}[Canonical embeddings of surfaces]\label{th:surf} $X$ is a canonical surface. Assume that $H^1(2K_X)=0$. Then $mK_X$ is very ample if $m\ge5$, or if $m=4$ and $K_X^2\ge2$, or if $m=3$, $p_g\ge2$ and $K_X^2\ge3$. \end{TEO} Here $H^1(2K_X)=0$ follows at once in characteristic 0 from Kodaira vanishing or Mumford's vanishing theorem. One can also get around the assumption $H^1(2K_X)=0$ in characteristic $p>0$ (see \cite{Ek} or \cite{S-B}). In fact Ekedahl's analysis (see \cite{Ek}, Theorem~II.1.7) shows that $H^1(2K_S)\ne0$ is only possible in a very special case, when $p=2,\chi(\Oh_S)=1$ and $S$ is (birationally) an inseparable double cover of a K3 surface or a rational surface. In \S\ref{sec:tri} and \S\ref{sec:bi} we apply these ideas to prove the following theorems on tricanonical and bicanonical linear systems of a surface of general type. \begin{TEO}[Tricanonical embeddings]\label{th:tri} Suppose that $X$ is a canonical surface with $K_X^2\ge3$. Then $3K_X$ is very ample if either \begin{enumerate} \renewcommand\labelenumi{(\alph{enumi})} \item $q=h^1(\Oh_X)=0$; or \item $\chi(\Oh_X)\ge1$, $\dim\Pico X>0$ and $H^1(2K_X-L)=0$ for all $L\in\Pico X$. \end{enumerate} Note that (a) or (b) cover all cases with $\chara k=0$. Thus the cases {\em not} covered by our argument are in $\chara k=p>0$, with either $p_g<q$ or $\dim\Pico X=0$. \end{TEO} Theorem~\ref{th:tri} in characteristic 0 is a result of Reider \cite{Rei}, but see also \cite{Ca2}. Without the condition $K_X^2\ge3$, the double plane with branch curve of degree 8 (that is, $X_8\subset\proj(1,1,1,4)$) is a counter\-example. It follows from a result of Ekedahl (\cite{Ek}, Theorem~II.1.7) that if $\chi(\Oh_X)\ge1$ then $H^1(2K_X-L)=0$ for all $L\ne0$. The remaining assumption in Theorem~\ref{th:tri} is that $H^1(2K_X)=0$, and this can also be got around, as shown by Shepherd-Barron \cite{S-B}. \begin{TEO}[Bicanonical embeddings]\label{th:bi} We now assume that $q=0$ and $p_g\ge4$. \begin{enumerate} \renewcommand\labelenumi{(\alph{enumi})} \item $2K_X$ is very ample if every $C\in|K_X|$ is numerically $3$-connected (in the sense of Definition~\ref{def:m-conn}, see also Lemma~\ref{lem:n-conn}). More precisely, $|2K_X|$ separates a cluster $Z$ of degree $2$ provided that every curve $C\in|K_X|$ through $Z$ is $3$-connected. \item Assume in addition that $K_X^2\ge10$, and let $Z$ be a cluster of degree $2$ contained in $X$. Then $Z$ is contracted by $|2K_X|$ if and only if $Z$ is contained in a curve $B\subset X$ with \begin{equation} K_XB=p_aB=1\text{ or }2 \nonumber \end{equation} (a {\em Francia curve}, compare Definition~\ref{def:Frc}), and $\sI_Z\Oh_B(2K_X)\iso\om_B$. \item In particular, $|2K_X|$ defines a birational morphism unless $X$ has a pencil of curves of genus $2$. \end{enumerate} \end{TEO} \begin{REMS} (1) A cluster $Z$ of degree 2 is automatically contracted by $|2K_X|$ if it is contained in a curve $C\subset X$ for which $\sI_Z\Oh_C(2K_X)\iso\om_C$ (for a non\-singular curve, this reads ${2K_X}_{|C}=K_C+P+Q$). Thus (b) says in particular that if this happens for some $C$ then it also happens for a Francia curve. (2) The assumptions $q=0$ and $p_g\ge4$ are needed for the simple minded ``restriction method'' of this paper, but we conjecture that (b) holds without them (at least in characteristic zero, or assuming $q=0$); the case $Z=\{x,y\}$ with $x\ne y$ (that is, ``separating points'') follows in characteristic zero by Reider's method. We believe that the conjecture can be proved quite generally by a different argument based on Ramanujam--Francia vanishing, or by Reider's method applied to reflexive sheaves on $X$. Stay tuned! (3) In characteristic 0, Theorem~\ref{th:bi} (without the assumption $q=0$) is due essentially to Francia (unpublished, but see \cite{Fr1}--\cite{Fr2}) and Reider \cite{Rei}. Theorem~\ref{th:bi}, (a) is a consequence of Theorem~\ref{th:hh} on canonical embeddings of curves and the generalisation of hyper\-elliptic curves. The results in Theorem~\ref{th:bi} are only a modest novelty, in that there is no restriction on the characteristic of the ground field (see \cite{S-B}, Theorems~25, 26 and~27 for $\chara k\ge11$). Further results on the bicanonical map $\fie_{2K}$ for smaller values of $p_g$, $K_X^2$ (in characteristic 0) require a more intricate analysis, and we refer to recent or forthcoming articles (\cite{C-F-M}, \cite{C-C-M}). Other applications of our methods can be found in \cite{F}. \end{REMS} \subsection*{Acknowledgment} It is a pleasure to thank Ingrid Bauer for interesting discussions on linear systems on surfaces, out of which this paper originated. \subsection*{Conventions} This paper deals systematically with reducible and nonreduced curves and their subschemes $B\subset C$. A coherent sheaf $\sF$ on a curve $C$ is {\em torsion free} if there are no sections $s\in\sF$ supported at points; on a 1-dimensional scheme, this is obviously equivalent to $\sF$ {\em Cohen--Macaulay}. We say that $C$ is {\em purely \hbox{$1$-dimensional}\/} or {\em Cohen--Macaulay} if $\Oh_C$ is torsion free. A map $\fie\colon\sF\to\sG$ between coherent sheaves on $B$ is {\em generically injective} if it is injective at every generic point of $B$; if $\sF$ is torsion free then $\fie$ is automatically an inclusion $\sF\into\sG$. If we know that the generic stalks of $\sF$ and $\sG$ have the same length at every generic point of $C$ then a generically injective map $\fie\colon\sF\to\sG$ is an isomorphism at each generic point, and therefore $\ker\fie$ and $\coker\fie$ have finite length. Indeed, they are both coherent sheaves supported at a finite set, and by the Nullstellensatz, each stalk is killed by a power of the maximal ideal. This applies, for example, to the map $\fie\colon\sI_Z\Oh_B(H)\into\om_B$ of Theorem~\ref{th:curve}, see Lemma~\ref{lem:gg} below. A scheme $B$ is {\em Gorenstein in codimension $0$} or {\em generically Gorenstein} if $\om_B$ is locally isomorphic to $\Oh_B$ at every generic point of $B$. A {\em cluster} of degree $r$ is a 0-dimensional subscheme $Z\subset X$ supported at finitely many points, with ideal sheaf $\sI_Z$, structure sheaf $\Oh_Z=\Oh_X/\sI_Z$, and having $\deg Z=h^0(\Oh_Z)=\length\Oh_Z=r$. We sometimes write $Z=(x,y)$ for a cluster of degree 2, where $x,y$ are either 2 distinct points of $X$, or a point $x$ plus a tangent vector $y$ at $x$. We say that a linear system $|H|$ on $X$ {\em separates} $Z$ (or separates $x$ and $y$) if $H^0(X,\Oh_X(H))\to\Oh_Z(H)$ is onto, or {\em contracts} $Z$ if $Z$ does not meet the base locus $\Bs|H|$, and $\rank\{H^0(X,\Oh_X(H))\to\Oh_Z(H)\}=1$. \subsection*{Notation} \begin{enumerate} \item[$X$] A projective scheme over an arbitrary field $k$. We sometimes (not always consistently) write $k\subset\overline k$ for the algebraic closure, and $X_{\overline k}=X\otimes_k\overline k$. \item[$\om_X$] Dualising sheaf of $X$ (see \cite{Ha}, Chap.~III, \S7). \item[$|H|$] Linear system defined by a Cartier divisor $H$ on $X$. \item[$C$] A curve, that is, a projective scheme over $k$ which is purely 1-dimensional, in the sense that $\Oh_C$ is Cohen--Macaulay (torsion free). \item[$p_aC$] The arithmetic genus of $C$, $p_aC=1-\chi(\Oh_C)$. \item[$K_C$] A canonical divisor of a Gorenstein curve $C$, that is, a Cartier divisor such that $\Oh_C(K_C)\iso\om_C$ (only defined if $C$ is Gorenstein). \item[$\deg\sL$] The degree of a torsion free sheaf of rank 1 on $C$; it can be defined by \begin{equation} \deg\sL=\chi(\sL)-\chi(\Oh_C). \nonumber \end{equation} If $H$ is a Cartier divisor on $C$, we set $HC=\deg\Oh_C(H)$. \item[$S$] A nonsingular projective surface. \item[$DD'$] Intersection number of divisors $D,D'$ on a nonsingular projective surface. \item[$K_S$] A canonical divisor on $S$. \item[$K_X^2$] If $X$ is a Gorenstein surface, $K_X$ is a Cartier divisor with $\om_X=\Oh_X(K_X)$, and $K_X^2$ is the selfintersection number of the Cartier divisor $K_X$. If $X$ has only Du Val singularities and $\pi\colon S\to X$ is the minimal nonsingular model then $K_S=\pi^*K_X$ and $K_X^2=K_S^2$. \item[$p_g,q$] The geometric genus $p_g=h^0(S,K_S)=h^0(X,K_X)$ of $S$ or $X$ (respectively the irregularity $q=h^1(S,\Oh_S)=h^1(X,\Oh_X)$). \item[$P_n$] The $n$th plurigenus $P_n=h^0(S,nK_S)$ of $S$. \end{enumerate} \section{Embedding curves} We start with a useful remark. \begin{REM}\label{rem:1} Let $H$ be a Cartier divisor on a scheme $X$. Then $H$ is very ample if and only if the restriction map \begin{equation} H^0(\Oh_X(H))\to\Oh_Z(H) \label{eq1} \end{equation} is onto for every cluster $Z\subset X$ (more precisely, for every $Z\subset X_{\overline k}$) of degree $\le2$. \end{REM} \begin{pf} By the standard embedding criterion of \cite{Ha}, Chap.~II, Prop.~7.3, we have to prove that (\ref{eq1}) is onto for all the ideals $\sI_Z=m_x$ or $m_xm_y$ with $x,y\in X$. For $x\ne y$, we are done. By assumption $H^0(\Oh_C(H))\to\Oh_C/m_x$ is onto for every $x\in X$. Now if the image of $H^0(m_x\Oh_C(H))\to m_x/m_x^2$ is contained in a hyperplane $V\subset m_x/m_x^2$, then the inverse image of $V$ in $\Oh_{C,x}$ generates an ideal $\sI\subset\Oh_{X,x}$ defining a cluster $Z$ of degree 2 supported at $x$ such that $H^0(\Oh_C(H))\to\Oh_Z$ is not onto. \QED \end{pf} \begin{REM}\label{rem:2} The chain of reasoning we use below is that, by Remark~\ref{rem:1} and cohomology, $H$ is very ample if and only if $H^1(\sI_Z\Oh_X(H))\to H^1(\Oh_X(H))$ is injective for each cluster $Z$ of degree $2$, or dually (if $X=C$ is a curve), $\Hom(\Oh_C(H),\om_C)\to\Hom(\sI_Z(H),\om_C)$ is onto. \end{REM} \begin{LEM}\label{lem:gg} Let $C$ be a curve. Assume that there is a Cartier divisor $H$ on $C$ and a cluster $Z\subset C$ for which the sheaf $\sL=\sI_Z\Oh_C(H)$ has an inclusion $\sL\into\om_C$. Then $C$ is generically Gorenstein. \end{LEM} \begin{pf} By assumption, $\sL\iso\Oh_C$ at every generic point of $C$. We must prove that an inclusion $\sL\into\om_C$ maps onto every generic stalk $\om_{C,\eta}$, or equivalently, that the cokernel $\sN=\om_C/\sL$ has finite length. We give two slightly different proofs, one based on RR, and one using properties of dualising modules. Let $\Oh_C(1)$ be an ample line bundle on $C$. Then by Serre vanishing (see \cite{Se1}, n$^\circ$~66, Theorem~2 or \cite{Ha}, Chap.~III, Theorem~5.2), for $n\gg0$, the exact sequence \begin{equation} 0\to\sL(n)\to\om_C(n)\to\sN(n)\to0 \nonumber \end{equation} is exact on global sections, and all the $H^1$ vanish. Now by RR and duality, \begin{equation} h^0(\om_C(n))=h^1(\Oh_C(-n))=-\chi(\Oh_C)+n\deg\Oh_C(1)\quad\text{for $n\gg0$.} \nonumber \end{equation} On the other hand, RR also gives $h^0(\sL(n))=\chi(\Oh_C)+\deg\sL+n\deg\Oh_C(1)$ for $n\gg0$, since $\sL\iso\Oh_C$ at every generic point. Thus \begin{equation} h^0(\sN(n))=-2\chi(\Oh_C)+\deg\sL\quad\text{for all $n\gg0$,} \nonumber \end{equation} and therefore $\sN$ has finite length. The alternative proof of the lemma uses the ``well-known fact'' (see below) that the generic stalk $\om_{C,\eta}$ of the dualising sheaf at a generic point $\eta\in C$ is a dualising module for the Artinian local ring $\Oh_{C,\eta}$, so that they have the same length, and therefore an inclusion $\sL\into\om_C$ is generically an isomorphism. The above proof in effect deduces $\length\om_{C,\eta}=\length\Oh_{C,\eta}$ from RR together with Serre duality, the defining property of $\om_C$. \paragraph{Proof of the ``well-known fact''} This is proof {\em by incomprehensible reference}. First, if $\eta\in X$ is a generic point of a scheme, more-or-less by definition, a dualising module of the Artinian ring $\Oh_{X,\eta}$ is an injective hull of the residue field $\Oh_{X,\eta}/m_{X,\eta}=k(\eta)$ (see \cite{Gr-Ha}, Proposition~4.10); in simple-minded terms, $\Oh_{X,\eta}$ clearly contains a field $K_0$ such that $K_0\subset k(\eta)$ is a finite field extension, and the vector space dual $\Hom_{K_0}(\Oh_{X,\eta},K_0)$ is a dualising module. Next, if $\eta\in X$ is a generic point of a subscheme $X\subset\proj=\proj^N$ of pure codimension $r$, then by \cite{Ha}, Chap.~III, Prop.~7.5, the dualising sheaf of $X$ is $\om_X=\sExt^r_{\Oh_\proj}(\Oh_X,\om_\proj)$. On the other hand, the local ring $\Oh_{\proj,\eta}$ of projective space along $\eta$ is an \hbox{$r$-dimensional} regular local ring, and therefore Gorenstein, so that by \cite{Gr-Ha}, Prop.~4.13, $\Ext^r_{\Oh_{\proj,\eta}}(\Oh_{X,\eta},\om_{\proj,\eta})$ is a dualising module of $\Oh_{X,\eta}$ (an injective hull of the residue field $\Oh_{X,\eta}/m_{X,\eta}=k(\eta)$). \QED \end{pf} \begin{LEM}[Automatic adjunction]\label{lem:adj} Let $\sF$ be a coherent sheaf on $C$, and $\fie\colon\sF\to\om_C$ a map of $\Oh_C$-modules. Set $\sJ=\Ann\fie\subset\Oh_C$, and write $B\subset C$ for the subscheme defined by $\sJ$. Then $\fie$ has a canonical factorisation of the form \begin{equation} \sF\onto\sF_{|B}\to\om_B=\sHom_{\Oh_C}(\Oh_B,\om_C)\subset\om_C, \label{eq:adj} \end{equation} where $\sF_{|B}\to\om_B$ is generically onto. \end{LEM} \begin{pf} By construction of $\sJ$, the image of $\fie$ is contained in the submodule \begin{equation} \bigl\{s\in\om_C\bigm|\sJ s=0\bigr\}\subset\om_C \nonumber \end{equation} But this clearly coincides with $\sHom(\Oh_B,\om_C)$. Now the inclusion morphism $B\into C$ is finite, and $\om_B=\sHom_{\Oh_C}(\Oh_B,\om_C)$ is just the adjunction formula for a finite morphism (see, for example, \cite{Ha}, Chap.~III, \S7, Ex.~7.2, or \cite{Re}, Prop.~2.11). The factorisation (\ref{eq:adj}) goes like this: $\fie$ is killed by $\sJ$, so it factors via the quotient module $\sF/\sJ\sF=\sF_{|B}$. As just observed, it maps into $\om_B\subset\om_C$. Finally, it maps onto every generic stalk of $\om_B$, again by definition of $\sJ$: a submodule of the sum of generic stalks $\bigoplus\om_{B,\eta}$ is the dual to the generic stalk $\bigoplus\Oh_{B',\eta}$ of a purely 1-dimensional subscheme $B'\subset B$, and $\fie$ is not killed by the corresponding ideal sheaf $\sJ'$. \QED \end{pf} \begin{REM} We define $B$ to be the {\em scheme theoretic support} of $\fie$. Note that if $C=\sum n_i\Ga_i$ is a Weil divisor on a normal surface and $\sF$ a line bundle, the curve $B\subset C$ defines a splitting $C=A+B$ where $A$ is the {\em divisor of zeros} of $\fie$: at the generic point of $\Ga_i$, the map $\fie$ then looks like $y_i^{a_i}$, where $y_i$ is the local equation of $\Ga_i$, and $A=\sum a_i\Ga_i$. In the general case however, $A$ does not make sense. \end{REM} \begin{pfof}{Theorem~\ref{th:curve}} Let $H$ be a Cartier divisor, and $\sI$ the ideal sheaf of a cluster for which $H^1(\sI\Oh_C(H))\ne0$. Then $\Hom(\sI\Oh_C(H),\om_C)\ne0$ by Serre duality. First pick any nonzero map $\fie\colon\sI\Oh_C(H)\to\om_C$. By Lemma~\ref{lem:adj}, $\fie$ comes from an inclusion $\sI\Oh_B(H)\into\om_B$ for a subscheme $B\subset C$, and $B$ is generically Gorenstein by Lemma~\ref{lem:gg}. Finally, if $H^0(\Oh_C(H))\to\Oh_Z(H)$ is not onto, then the next arrow in the cohomology sequence \begin{equation} H^1(\sI\Oh_C(H))\to H^1(\Oh_C(H)) \nonumber \end{equation} is not injective, and dually, the restriction map \begin{equation} \Hom(\Oh_C(H),\om_C)\to\Hom(\sI\Oh_C(H),\om_C) \nonumber \end{equation} is not onto. Thus we can pick $\fie\colon\sI\Oh_C(H)\to\om_C$ which is not the restriction of a map $\Oh_C(H)\to\om_C$. Then also the map $\sI\Oh_B(H)\into\om_B$ given by Lemma~\ref{lem:adj} is not the restriction of a map $\Oh_B(H)\into\om_B$. For the final part, an inclusion $\sI\Oh_B(H)\into\om_B$ has cokernel of finite length, so that $\chi(\sI\Oh_B(H))\le\chi(\om_B)$. Plugging in the definition of degree gives \begin{equation} 1-p_aB+HB-\deg(Z\cap B)\le p_aB-1, \nonumber \end{equation} that is, \begin{equation} HB\le2p_aB-2+\deg(Z\cap B). \nonumber \end{equation} Thus, assuming the inequality (2) of Theorem~\ref{th:curve}, no such inclusion $\sI\Oh_B(H)\into\om_B$ can exist, so that $H^0(\Oh_C(H))\to\Oh_Z(H)$ is onto. \QED \end{pfof} \section{The canonical map of a Gorenstein curve}\label{sec:cc} We now discuss the canonical map $\fie_{K_C}$ of a Gorenstein curve, writing $K_C$ for a canonical divisor of $C$, that is, a Cartier divisor for which $\om_C\iso\Oh_C(K_C)$. Our approach is motivated in part by the examples and results in the reduced case treated in \cite{Ca1}. \begin{DEF}\label{def:m-conn} A Gorenstein curve $C$ over an algebraically closed field $k$ is {\em numerically $m$-connected} if \begin{equation} \deg\Oh_B(K_C)-\deg\om_B=\deg(\om_C\otimes\Oh_B)-(2p_aB-2)\ge m \nonumber \end{equation} for every generically Gorenstein strict subcurve $B\subset C$. For $C$ over any field, we say that $C$ is numerically $m$-connected if $C\otimes\overline k$ is numerically $m$-connected. \end{DEF} \begin{REM}\label{rem:m-conn} Note that for divisors on a nonsingular surface, \begin{equation} \deg\Oh_B(K_C)-\deg\om_B=(K_S+C)B-(K_S+B)B=(C-B)B. \nonumber \end{equation} In this context, Franchetta and Ramanujam define numerically connected in terms of the intersection numbers $AB=(C-B)B$ for all effective decompositions $C=A+B$. The point of our definition is to use the numbers $\deg\Oh_B(K_C)-\deg\om_B$ in the more general case as a substitute for $(C-B)B$. In effect, we think of the adjunction formula as defining the ``degree'' of the ``normal bundle'' to $B$ in $C$, in terms of the difference between $K_C{}_{|B}$ and $\om_B$. \end{REM} \begin{TEO}\label{th:free} Let $C$ be a Gorenstein curve over a field $k$. \begin{enumerate} \renewcommand\labelenumi{(\alph{enumi})} \item If $C$ is numerically $1$-connected then $H^0(\Oh_C)=k$ (the constant functions). \item If $C$ is numerically $2$-connected then either $|K_C|$ is free or $C\iso\proj^1$ (over the algebraic closure $\overline k$, of course). In particular, in this case $p_aC=0$ implies $C\iso\proj^1$. \end{enumerate} \end{TEO} \begin{pfof}{(a)} First, if $f\in H^0(\Oh_C)$ is a nonzero section vanishing along some reduced component of $C$, then applying Lemma~\ref{lem:adj} to the multiplication map $\mu_f\colon\Oh_C(K_C)\to\om_C$ gives an inclusion $\Oh_B(K_C)\into\om_B$, which is forbidden by numerically 1-connected (because $\deg\Oh_B(K_C)>\deg\om_B$). Now if $H^0(\Oh_C)\ne k$, there exists a nonzero section $f\in H^0(\Oh_{C\otimes\overline k})$ vanishing at any given point $x\in C\otimes\overline k$. An inclusion $\Oh_C\into m_x$ contradicts at once $0=\deg\Oh_C>\deg m_x=-1$, so that $f$ must vanish along some component of $C$, and we have seen that this is impossible. \QED \end{pfof} \begin{pfof}{(b)} As discussed in Remark~\ref{rem:2}, the standard chain of reasoning is as follows: \begin{enumerate} \item $x\in C$ is a base point of $|K_C|$ if and only if $H^0(\Oh_C(K_C))\to\Oh_x(K_C)$ is not onto, and then \item $H^1(m_x\Oh_C(K_C))\to H^1(\Oh_C(K_C))$ is not injective, \item dually, $\Hom(\Oh_C(K_C),\om_C)\to\Hom(m_x\Oh_C(K_C),\om_C)$ is not onto, \item therefore there exists a map $s\colon m_x\Oh_C(K_C)\to\om_C$ linearly independent of the identity inclusion. \end{enumerate} Now by Lemma~\ref{lem:adj}, the map $s$ factors via an inclusion $m_x\Oh_B(K_C)\into\om_B$ on a generically Gorenstein curve $B$. But then $B\subsetneq C$ is forbidden by the numerically 2-connected assumption $\deg m_x\Oh_B(K_C)-\deg\om_B\ge1$. Therefore $B=C$, that is, $s\colon m_x\Oh_C(K_C)\into\om_C$ is an inclusion. After tensoring down with $-K_C$, this gives an inclusion $i\colon m_x\into\Oh_C$ linearly independent of the identity. Write $\sF=i(m_x)\subset\Oh_C$. Then $\deg\sF=-1$, and therefore $\sF=m_z$ for some $z\in C$. Now for any point $y\in C\setminus\{x\}$, there exists a linear combination $s'=s+\la\id$ vanishing at $y$, which therefore defines an isomorphism $m_x\iso m_y$. This implies that every point $y\in C$ is a Cartier divisor, hence a nonsingular point. Since $C$ is clearly connected, and $\Oh_C(x-y)\iso\Oh_C$ for every $x,y\in C$, it follows that $C\iso\proj^1$. For the final statement, if $p_aC=0$ then $1=h^0(\Oh_C)=h^1(\om_C)$ by (a) and duality, hence $h^0(\om_C)=0$ by RR, so that $H^0(\Oh_C(K_C))\to\Oh_x$ is not onto for any $x\in C$. \QED \end{pfof} \begin{DEF} We say that a Gorenstein curve $C$ is {\em honestly hyperelliptic (\cite{Ca1}, Definition~3.18)} if there exists a finite morphism $\psi\colon C\to\proj^1$ of degree $2$ (that is, $\psi$ is finite and $\psi_*\Oh_C$ is locally free of rank $2$ on $\proj^1$). The linear system $\psi^*|\Oh_C(1)|$ defining $\psi$ is called an {\em honest $g^1_2$.} \end{DEF} We note the immediate consequences of the definition. \begin{LEM}\label{lem:hh} An honestly hyperelliptic curve $C$ of genus $p_aC=g\ge0$ is isomorphic to a divisor $C_{2g+2}$ in the weighted projective space $\proj(1,1,g+1)$, not passing through the vertex $(0,0,1)$, defined by an equation \begin{equation} w^2+a_{g+1}(x_1,x_2)w+b_{2g+2}(x_1,x_2)=0. \nonumber \end{equation} It follows that every point of $C$ is either nonsingular or a planar double point, and that $C$ is either irreducible, or of the form $C=D_1+D_2$ with $D_1D_2=g+1$. The projection $\fie\colon C\to\proj^1$ is a finite double cover, and the inverse image of any $x\in\proj^1$ is a Cartier divisor which is a cluster $Z\subset C$ of degree $2$. In other words, $Z$ is either $2$ distinct nonsingular points of $C$, a nonsingular point with multiplicity $2$, or a section through a planar double point of $C$. \qed\end{LEM} \begin{TEO}\label{th:hh} Let $C$ be a numerically $3$-connected Gorenstein curve. Then either $|K_C|$ is very ample or $C$ is honestly hyperelliptic. In particular, in this case if\/ $p_aC\ge2$ then $K_C$ is ample, and if\/ $p_aC=1$ then $C$ is honestly hyperelliptic (over the algebraic closure $\overline k$, of course). \end{TEO} \begin{pf} Let $Z$ be a cluster of degree 2 for which $H^0(\Oh_C(K_C))\to\Oh_Z(K_C)$ is not onto. The previous chain of reasoning gives a map $\sI_Z\Oh_C(K_C)\to\om_C$ linearly independent of the identity inclusion. An inclusion $\sI_Z\Oh_B(K_C)\into\om_B$ with $B\subsetneq C$ is forbidden as before by $C$ numerically 3-connected. Therefore we get an inclusion $s\colon\sI_Z\Oh_C(K_C)\into\om_C$ linearly independent of the identity inclusion. Note that any linear combination $s'=s+\la\id$ of the two sections is again generically injective, since an inclusion $\sI_Z\Oh_B(K_C)\into\om_B$ with $B\subsetneq C$ is forbidden by numerically 3-connected. The image $\sF=s(\sI_Z\Oh_C(K_C))\subset\om_C$ is a submodule of colength 2, therefore of the form $\sF=\sI_{Z'}\om_C$ for some cluster $Z'\subset C$. Tensoring down the iso\-morphism $s\colon\sI_Z\Oh_C(K_C)\to\sI_{Z'}\om_C$ gives an isomorphism $s\colon\sI_Z\iso\sI_{Z'}$, still linearly independent of the identity inclusion $\sI_Z\into\Oh_C$. Logically, there are 3 cases for $Z$ and $Z'$. The first of these corresponds to an honest $g^1_2$ on $C$; the other two, corresponding to a $g^1_2$ with one or two base points, lead either to $p_aC\le1$ or to a contradiction. The case division is as follows: \paragraph{Case $Z\cap Z'=\emptyset$} Then the isomorphism $\sI_Z\iso\sI_{Z'}$ implies that both $Z$ and $Z'$ are Cartier divisors, and the two linearly independent inclusions $\sI_Z\into\Oh_C$ define an honest $g^1_2$ on $C$. In more detail: $\Oh_C(Z)$ has 2 linearly independent sections with no common zeroes, and no linear combination of these vanishes on any component of $C$. Therefore $|Z|$ defines a finite 2-to-1 morphism $C\to\proj^1$. \paragraph{Case $Z=Z'$} This case leads to an immediate contradiction. Indeed, take any point $x\notin\Supp Z$; then some linear combination of the two isomorphisms $s,\id\colon\sI_Z\to\sI_Z$ vanishes at $x$, and therefore vanishes along any reduced component of $C$ containing $x$. But we have just said that this is forbidden. \paragraph{Case $Z\cap Z'=x$} Here the case assumption can be rewritten $\sI_Z+\sI_{Z'}=m_x$. This case is substantial, and it really happens in two examples: \begin{enumerate} \item if $C$ is an irreducible plane cubic with a node or cusp $P$, and $Q,Q'\in C\setminus P$ then $m_Pm_Q\iso m_Pm_{Q'}$; \item $\proj^1$ has an incomplete $g^1_2$ with a fixed point, of the form $P+|Q|$. \end{enumerate} We prove that we are in one of these cases. In either example, the curve $C$ has an honest $g^1_2$ (not directly given by our sections $s,\id$), so the theorem is correct. \begin{CLA}\label{cla:mov_y} For any point $y\in C\setminus\{x\}$, there exists a linear combination $s'=s+\la\id$ defining an isomorphism $\sI_Z\iso m_xm_y$. \end{CLA} \begin{pfof}{Claim} Since $\sI_Z,\sI_{Z'}\subset m_x$, we have two linearly independent maps $s,\id\colon\sI_Z\into m_x$, and some linear combination $s'=s+\la\id$ vanishes at $y$. Also, no map $\sI_Z\to m_x$ vanishes along a component of $C$. Thus $s'(\sI_Z)=m_xm_y$. \QED \end{pfof} It follows from the claim that $m_xm_y\iso m_xm_{y'}$ for any two points $y,y'\ne x$, so that $y,y'$ are nonsingular, and $C$ is reduced and irreducible. Now let $\si\colon C_1\to C$ be the blow up of $m_x$. Then, essentially by definition of the blow up, $m_x\Oh_{C_1}\iso\Oh_{C_1}(-E)$ where $E$ is a Cartier divisor on $C_1$. Then $m_{C_1,y}\iso m_{C_1,y'}$ for general points $y,y'\in C_1$, hence as usual $C_1\iso\proj^1$. If $C_1\iso C$ there is nothing more to prove. If $C_1\not\iso C$, the conductor ideal $\sC=\sHom_{\Oh_C}(\si_*\Oh_{C_1},\Oh_C)$ of $\si_*\Oh_{C_1}$ in $\Oh_C$ is $m_x$. Indeed, let $f\in k(C)$ be the rational function such that multiplication by $f$ gives $m_xm_y\iso m_xm_{y'}$; then $f$ is an affine parameter on $C_1=\proj^1$ outside $y$, so that all regular functions on $C_1$ are regular functions of $f$, and $fm_x=m_x$ implies $\si_*(m_x\Oh_{C_1})=m_x\subset\Oh_C$. Now it is known that the only Gorenstein curve singularity $x\in C$ with conductor ideal $m_x$ is a node or cusp (see \cite{Se2}, Chap.~IV, \S11 or \cite{Re}, Theorem~3.2): indeed, $m_x\subset\Oh_C\subset\si_*\Oh_{C_1}$, and the Gorenstein assumption $n=2\de$ gives $\length(\si_*\Oh_{C_1}/\Oh_C)=\length(\Oh_C/m_x)=1$. Therefore $p_aC=1$. For the final statement, if $p_aC=1$ then $1=h^0(\Oh_C)=h^1(\om_C)$ by Theorem~\ref{th:free}, (a) and duality, hence $h^0(\om_C)=1$ by RR, so that $H^0(\Oh_C(K_C))\to\Oh_Z$ is not onto for any cluster $Z\in C$ of degree 2. \QED \end{pf} \begin{REM} If $C$ is a numerically $3$-connected Gorenstein curve with $p_aC\ge2$, then Theorem~\ref{th:hh} says that $K_C$ is automatically ample, and the usual dichotomy holds: either $K_C$ is very ample, or $C$ is honestly hyperelliptic. Now assume instead that the dualising sheaf $\om_C=\Oh_C(K_C)$ is ample and generated by its $H^0$. Equivalently, that $|K_C|$ is a free linear system, defining a finite morphism (the {\em canonical morphism}) $\fie=\fie_{K_C}\colon C\to\proj^{p_a-1}$. In \cite{Ca1}, Definition~3.9, $C$ was defined to be {\em hyperelliptic} if $\fie_{K_C}$ is not birational on some component of $C$. Thus by Theorem~\ref{th:hh}, in the $3$-connected case, hyperelliptic and honestly hyperelliptic coincide. \end{REM} \section{Canonical maps of surfaces of general type}\label{sec:pluri} We give a slight refinement of a useful lemma due independently to J.~Alexander and I.~Bauer. \begin{LEM}[Alexander--Bauer]\label{lem:ab} Suppose that $H$ is a Cartier divisor on an irreducible projective scheme $X$. Assume given effective Cartier divisors $D_1,D_2$, $D_3$ such that \begin{enumerate} \renewcommand\labelenumi{(\roman{enumi})} \item $H^0(\Oh_X(H))\to H^0(\Oh_{D_i}(H))$ is onto. \item $H$ is very ample on every $\De\in|D_i|$ for $i=1,2,3$. \end{enumerate} Then $H$ is very ample on $X$ if either \begin{enumerate} \renewcommand\labelenumi{(\alph{enumi})} \item $H\lineq D_1+D_2$ and $\dim|D_2|\ge1$, or \item $H\lineq D_1+D_2+D_3$ and $\dim|D_i|\ge1$ for $i=1,2,3$. \end{enumerate} \end{LEM} \begin{pf} (a) is proved in \cite{Ba1}, Claim~2.19 and \cite{Ra}, Lemma~3.1, and also in \cite{C-F}, Prop.~5.1. We prove (b). By Remark~\ref{rem:1}, we need to prove that if $x$ is any point of $X$, and $y$ is either another point of $X$ or a tangent vector at $x$, then $|H|$ separates $x$ from $y$. If some $\De_i\in|D_i|$ contains both $x$ and $y$, we are done by the assumptions (i) and (ii). In particular, since $\dim|D_i|\ge1$, such a $\De_i$ exists if $x$ or $y$ belong to the base locus of $|D_i|$. Finally, if none of the above possibilities occurs, we can find $\De_1$ containing $x$ but not $y$, and $\De_2,\De_3$ containing neither $x$ nor $y$. Then $\De_1+\De_2+\De_3$ separates $x$ from $y$. \QED\end{pf} \begin{pfof}{Theorem~\ref{th:surf}} Let $\pi\colon S\to X$ be the natural birational morphism from a minimal surface of general type $S$ to its canonical model $X$; write $K_S$ and $K_X$ for the canonical divisors of $S$ and $X$. Then $\om_X$ is invertible and $\pi^*(\om_X)\iso\om_S$; in particular $H^0(X,mK_X)\iso H^0(X,mK_S)$ and$K_X^2=K_S^2$. \subparagraph{Step I} If $C\in|(m-2)K_X)|$, then $H^0(\Oh_X(mK_X))\to H^0(\Oh_C(mK_X))$ is onto. This follows from our assumption $H^1(\Oh_X(2K_X))=0$. \subparagraph{Step II} If $C\in|(m-2)K_X|$, then $\Oh_C(mK_X)$ is very ample. \begin{pf} By the curve embedding theorem Theorem~\ref{th:curve}, it is enough to prove that $mK_XB\ge2p_aB+1$ for every subcurve $B\subset C$. Note that by adjunction $K_C=(m-1){K_X}_{|C}$, so that we can write $m{K_X}_{|B}={K_X}_{|B}+{K_C}_{|B}$. Since $K_X$ is ample, $K_XB\ge1$, and therefore we need only prove that $K_XC\ge3$ and \begin{equation} \deg\Oh_B(K_C)-\deg\om_B\ge2\quad\text{for every strict subcurve $B\subset C$,} \nonumber \end{equation} that is, that $C$ is numerically 2-connected. The corresponding fact for the minimal nonsingular model $S\to X$ is easy and well known.\footnote{{\bf Tutorial}\enspace This is an easy consequence of the Hodge algebraic index theorem. If $D$ is nef and big and $D=A+B$ then $A^2+AB\ge0$, $AB+B^2\ge0$. The index theorem says that $A^2B^2\le(AB)^2$, with equality only if $A,B$ are numerically equivalent to rational multiples of one another. The reader should carry out the easy exercise of seeing that $AB\le0$ gives a contradiction, and proving all the connected assertions we need. Or see \cite{Bo}, \S4, Lemma~2 for details (the exceptional case $n=2$, $2K_S=A+B$, with $A\numeq B\numeq K_S$ and $K_S^2=1$ is excluded by the assumption $K_S^2\ge2$ if $m=4$ of Theorem~\ref{th:surf}).} Therefore $C$ numerically 2-connected follows from the next result, whose proof we relegate to an appendix. \begin{LEM}\label{lem:n-conn} Let $X$ be a surface with only Du Val singularities, and $\pi\colon S\to X$ the minimal resolution of singularities. Let $C\subset X$ be an effective Cartier divisor, and $C^*=\pi^*C$ the total transform of $C$ on $S$. Then \begin{equation} \text{$C^*$ numerically $k$-connected} \implies \text{so is $C$.} \nonumber \end{equation} Moreover, if $C^*$ is numerically $2$-connected, and is only $3$-disconnected by expressions $C^*=A+B$ where $A$ or $B$ is a $-2$-cycle exceptional for $\pi$ then $C$ is numerically $3$-connected. \end{LEM}\unskip\end{pf} \subparagraph{Step III} $h^0((m-2)K_X)\ge3$ if $m\ge5$, and $\ge2$ if $m=3$ or 4. \begin{pf} For $m=3$ this is just the assumption $p_g\ge2$. For $m\ge4$, if $p_g\ge2$, then clearly $h^0((m-2)K_X)\ge3$. Otherwise, in the case $p_g\le1$, we use the traditional numerical game of \cite{B-M}, based on Noether's formula $12\chi(\Oh_X)=(c_1^2+c_2)(X)$. It consists of writing out Noether's formula using Betti numbers for the etale cohomology, in the form \begin{equation} 10+12p_g=8h^1(\Oh_X)+2\De+b_2+K_X^2. \label{eq:No} \end{equation} Here the nonclassical term $\De=2h^1(\Oh_X)-b_1$ satisfies $\De\ge0$, and $\De=0$ if $\chara k=0$. Since all the terms on the right hand side of (\ref{eq:No}) are $\ge0$, it follows immediately that \begin{equation} \begin{aligned} &p_g\le1\implies h^1(\Oh_X)\le2\\ &p_g\le0\implies h^1(\Oh_X)\le1. \end{aligned} \nonumber \end{equation} Therefore, $p_g\le1$ implies $\chi(\Oh_X)\ge0$; hence, for $m\ge4$, by RR \begin{equation} h^0((m-2)K_X)\ge\chi(\Oh_X)+\binom{m-2}2K_X^2\quad \begin{cases} \ge3&\text{if $m\ge5$,}\\ \ge2&\text{if $m=4$.} \end{cases} \nonumber \end{equation} \end{pf} \unskip \subparagraph{Step IV} For $m=3$, we simply apply Lemma~\ref{lem:ab}, (b) to $3K\lineq K+K+K$. For $m=4$ we apply Lemma~\ref{lem:ab}, (a) to $4K\lineq 2K+2K$: the assumptions (i) and (ii) of the lemma hold by Steps~I, II and~III. For $m\ge5$, we want to show that $H^0(\Oh_X(mK_X))\to\Oh_Z$ is onto for any cluster $Z\subset X$ of degree 2. But by Step~III, there exists $C\in|(m-2)K_X|$ containing $Z$. The result then follows by Steps~I and~II. \QED \end{pfof} \subsection*{Appendix: Proof of Lemma~\ref{lem:n-conn}} Suppose that $B\subset C$ is a strict subcurve. Write $B'$ for the birational (=strict or proper) transform of $B$ in $S$ and $C^*=\pi^*C$ for the total transform of $C$. For the proof, we find a divisor $\Bh$ (the {\em hat transform}) with the properties \begin{enumerate} \renewcommand\labelenumi{(\roman{enumi})} \item $B'\le\Bh\le C^*$ and $\Bh-B'$ contains only exceptional curves; \item $p_a\Bh=p_aB$. \end{enumerate} Suppose first that we know $\Bh$ satisfying these conditions. Then \begin{equation} (C^*-\Bh)\Bh\ge k \nonumber \end{equation} by the assumption on $C^*$, which we write \begin{equation} (K_S+C^*)\Bh-(K_S+\Bh)\Bh\ge k. \nonumber \end{equation} Here the first term equals $(K_X+C)B=\deg\Oh_B(K_C)$, and the second $2p_a\Bh-2=2p_aB-2$. Thus \begin{equation} \deg\Oh_B(K_C)-(2p_aB-2)=(K_S+C^*)\Bh-(2p_a\Bh-2)\ge k. \nonumber \end{equation} So it is enough to find $\Bh$. For this, following the methods of \cite{Ar1}--\cite{Ar2}, let $\bigl\{\Ga_i\bigr\}$ be all the exceptional $-2$-curves. Define $\Bh=B'+\sum e_i\Ga_i$ with $e_i\in\Z$, $e_i\ge0$ minimal with respect to the property $\Bh\Ga_i\le0$; this exists, because $C^*-A'$ has the stated property (where $A'$ is the birational transform of the residual Weil divisor $C-B$). \begin{CLA}\label{cla:bhat} The curve $\Bh$ has the following properties: \begin{enumerate} \renewcommand\labelenumi{(\roman{enumi})} \setcounter{enumi}2 \item $\om_B=\pi_*\om_{\Bh}$; \item $R^1\pi_*\om_{\Bh}=0$. \end{enumerate} Therefore $p_a\Bh=p_aB$. \end{CLA} \begin{pfof}{Claim} Taking $\pi_*$ of the short exact sequence \begin{equation} 0\to\Oh_S(K_S)\to\Oh_S(K_S+\Bh)\to\om_{\Bh}\to0 \nonumber \end{equation} gives $0\to\Oh_X(K_X)\to\Oh_X(K_X+B)\to\pi_*\om_{\Bh}\to0$ and $R^1\pi_*\Oh_S(K_S+\Bh)=R^1\pi_*\om_{\Bh}$. The first of these implies that $\om_B=\pi_*\om_{\Bh}$. Indeed, if $B\subset X$ is an effective Weil divisor on any Cohen--Macaulay variety then the adjunction formula $\om_B=\sExt^1_{\Oh_X}(\Oh_B,\om_X)$ (see, for example, \cite{Re}, Theorem~2.12, (1)) boils down to an exact sequence $0\to\Oh_X(K_X)\to\Oh_X(K_X+B)\to\om_B\to0$. This proves (iii). By the method of \cite{Ar1}--\cite{Ar2}, \begin{equation} R^1\pi_*\Oh_S(K_S+\Bh)=\varprojlim H^1(D,\Oh_D(K_S+\Bh)), \nonumber \end{equation} where the inverse limit is taken over effective divisors $D=\sum a_j\Ga_j$. If all the $H^1=0$, the limit is zero, as required. Suppose then by contradiction that $D=\sum a_j\Ga_j$ has $H^1(\Oh_D(K_S+\Bh))\ne0$. Then dually, $\Hom(\Oh_D(K_S+\Bh),\om_D)\ne0$, and Lemma~\ref{lem:adj} gives an inclusion $\Oh_D(K_S+\Bh)\into\om_D$ (for a possibly smaller $D$). Writing out the adjunction formula for $\om_D$ and tensoring down by $K_S+\Bh$ gives $\Oh_D\into\Oh_D(D-\Bh)$. Therefore $(\Bh-D)\Ga_i\le0$ for every $\Ga_i\subset D$, and by construction of $\Bh$ for the other $\Ga_i$. Now $\Bh-D=B'+\sum e'_j\Ga_j$ contradicts the minimality of $\Bh$, provided we show that the $e_j'\ge0$. For this, note that \begin{equation} \bigl(\sum e'_j\Ga_j\bigr)\Ga_i=(\Bh-D)\Ga_i-B'\Ga_i\le0 \quad\text{for every $i$} \nonumber \end{equation} and the intersection form on the $\Ga_i$ is negative definite, so that the standard argument implies $\sum e'_j\Ga_j\ge0$ (write it as $A-B$ where $A,B\ge0$ have no common divisor, and calculate $B^2$). \QED \end{pfof} \section{The tricanonical map}\label{sec:tri} We state the following three points as independent lemmas in order to tidy up our proofs, and because they might be useful elsewhere. The first is a particular case of the numerical criterion for flatness, see \cite{Ha}, Chap.~III, Theorem~9.9. \begin{LEM}[Flat double covers]\label{lem:flat} If $\fie\colon X\to Y$ is a generically $2$-to-$1$ morphism (say with $Y$ integral), then for any $y\in Y$, the condition $\length\fie\1(y)=2$ implies that $\fie$ is flat over a neighbourhood of $y$. \qed \end{LEM} \begin{LEM}[Push-down of invariant linear systems]\label{lem:push} Let $\fie\colon X\to Y$ be a finite morphism of degree $2$, where $X$ and $Y$ are normal. Suppose that $L$ is a linear system of Cartier divisors on $X$ with the property that $\fie_{|D}\colon D\to\Ga_D=\fie(D)$ has degree $2$ for every $D\in L$. Then the $\Ga_D$ are linearly equivalent Weil divisors, that is, they are all members of one linear system. \end{LEM} \begin{pf} For any $D,D'\in L$, note that $2\Ga_D=\pi_*D$ is a Cartier divisor on $Y$, and $2\Ga_D\lineq 2\Ga_{D'}$, because if $D$ is locally defined by $f\in k(X)$ (or $D-D'=\div f$) then $2\Ga_D$ is locally defined by $\Norm(f)$, where $\Norm=\Norm_{k(X)/k(Y)}$. Thus the Weil divisor class $\Ga_D-\Ga_{D'}$ is a 2-torsion element of the Weil divisor class group $\WCl Y$ (modulo linear equivalence). The group of Weil divisors numerically equivalent to zero is an algebraic group of finite type, so that its 2-torsion subgroup is a finite algebraic group scheme $G$. Now for fixed $D_0\in L$, taking $D\mapsto\Ga_D-\Ga_{D_0}$ defines a morphism from the parameter space of the linear system $L$ to $G$, which must be the constant morphism to 0. This proves what we need. Assuming that $\fie$ is separable make this argument more intuitive, since then it is Galois, and $\fie_*\Oh_X$ splits into invariant and antiinvariant parts: $\fie_*\Oh_X=\Oh_Y\oplus\sL$, with $\sL$ a divisorial sheaf. Then $\Ga_D$ is locally either a Cartier divisor or in the local Weil divisor class of $\sL$, and $\Ga_D-\Ga_{D'}$ is in the kernel of $\fie^*$, which is a finite algebraic group scheme, etc. \QED \end{pf} \begin{LEM}\label{lem:sing} Let $\La$ be a linear system of Weil divisors through a point $P$ on a normal surface $Y$. Then the curves in $\La$ singular at $P$ form a projective linear subspace of codimension $\le2$. \end{LEM} \begin{pf} Easy exercise involving the resolution and birational transform. \qed \end{pf} \begin{pfof}{Theorem~\ref{th:tri}, Case~(a)} Since $q=0$, we have $\chi(\Oh_X)\ge1$, and $K_X^2\ge3$ gives $P_2=h^0(2K_X)\ge4$. Let $Z$ be a cluster of degree 2 on $X$. Since $P_2\ge4$, the linear subsystem $|2K_X-Z|$ consisting of curves $D\in|2K_X|$ through $Z$ has dimension $\ge1$, and any $D\in|2K_X|$ is 3-connected by the final part of Lemma~\ref{lem:n-conn} (whose assumptions are easily verified as in \cite{Bo}, \S4, Lemma~2). By $H^1(K_X)=0$, the sequence \begin{equation} 0\to H^0(X,\Oh_X(K_X))\to H^0(X,\Oh_X(3K_X))\to H^0(D,\om_D)\to0 \nonumber \end{equation} is exact. Since $|\om_D|$ is free by Theorem~\ref{th:free}, it follows that $\fie=\fie_{3K_X}$ is a finite morphism $\fie\colon X\to Y\subset\proj^N$, where $N=P_3-1$. Assume that $|3K_X|$ does not separate $Z$. Then, by Theorem~\ref{th:hh}, $D$ is honestly hyperelliptic. Since the same argument applies to any $D\in|2K_X-Z|$, it follows that $\deg\fie\ge2$. On the other hand, for any point $y\in Y$, if the scheme theoretic fibre $\fie\1(y)$ is a cluster of degree $\ge3$, then there is a curve $D'\in|2K_X|$ containing $\fie\1(y)$, and $\fie\1(y)$ is contained in a fibre of $\fie_{\om_{D'}}\colon D'\to\proj^1$, which contradicts Lemma~\ref{lem:hh}. Hence $\fie\colon X\to Y$ is of degree 2 (possibly inseparable if $\chara k=2$). In particular $2\mid9K^2$, so that $K^2$ is even and $K^2\ge4$; thus $P_2\ge5$, and $\dim|2K_X-Z|\ge2$ for any cluster $Z$ of degree 2. By changing $Z$ if necessary, we can assume that $\fie(Z)=y\in Y$ is a general point, and is thus nonsingular. We have just shown that every fibre $\fie\1(y)$ has degree exactly 2, so that $\fie$ is flat by Lemma~\ref{lem:flat}; it is easy to see that this implies that $Y$ is normal. Now for any $D\in|2K_X-Z|$, the image $\fie(D)=\Ga_D\subset Y$ is a curve through $y=\fie(Z)$ isomorphic to $\proj^1$, and $\deg\fie_{|D}=\deg\fie=2$. By Lemma~\ref{lem:push} the $\Ga_D\subset Y$ are linearly equivalent, so that they are all contained in a linear system. This contradicts Lemma~\ref{lem:sing}: in any linear system of curves through $y$, curves singular at $y$ form a linear subsystem of codimension $\le2$, whereas the $\Ga_D$ for $D\in|2K_X-Z|$ form an algebraic subfamily of nonsingular curves depending with a complete parameter space of dimension $\ge2$ made up of curves isomorphic to $\proj^1$. \QED \end{pfof} \begin{REM} Here we have assumed that $\fie(Z)=y\in Y$ is a general point only for simplicity (see Lemma~\ref{lem:sing}). \end{REM} \begin{pfof}{Theorem~\ref{th:tri}, Case~(b)} Let $Z$ be a cluster of degree 2 on $X$ and $x\in Z$ a reduced point; that is, $Z$ is either a pair $(x,y)$ of distinct points, or a point $x$ plus a tangent vector $y$ at $x$. We assume that $|3K_X|$ does not separate $Z$, and gather together a number of deductions concerning the curves \begin{equation} C_L\in|K_X+L|\quad\text{and}\quad D_L\in|2K_X-L|\quad\text{for all}\quad L\in\Pico X, \nonumber \end{equation} arriving eventually at a contradiction. \subparagraph{Step A} $h^0(K_X+L)\ge1$ for all $L\in\Pico X$. In fact if $L\ne0$ then $h^2(K_X+L)=0$, and hence $h^0(K_X+L)\ge\chi(K_X)\ge1$. \subparagraph{Step B} $Z\not\subset C_L$ for all $L\in\Pico X$ and all $C_L\in|K_X+L|$. Indeed \begin{equation} H^0(X,\Oh_X(3K))\to H^0(C_L,\Oh_{C_L}(3K_X)) \nonumber \end{equation} is onto by the assumption $H^1(\Oh_X(2K_X-L))=0$, and $\Oh_{C_L}(3K_X)$ very ample follows from Theorem~\ref{th:curve} exactly as in \S\ref{sec:pluri}, Step~II. Therefore if $Z\subset C_L$ then $|3K_X|$ separates $Z$, which we are assuming is not the case. \subparagraph{Step C} For general $L\in\Pico X$ and all $C_L\in|K_X+L|$ we have $x\in C_L$. First of all, since $\dim\Pico X\ge1$, there is an $L\in\Pico X$ and a curve $C_L\in|K_X+L|$ containing $x$, and $C_L$ does not contain $Z$ by Step~B. Now if $L_1,L_2\in\Pico X$ is a general solution of $L+L_1+L_2=0$, and $x\notin C_{L_1}$, $x\notin C_{L_2}$, then $C_L+C_{L_1}+C_{L_2}$ separates $x$ and $Z$, a contradiction. \subparagraph{Step D} $h^0(K_X+L)=1$ and $H^1(K_X+L)=0$ for general $L\in\Pico X$. By Step~C, every $s\in H^0(K_X+L)$ vanishes at $x$. If $h^0(K_X+L)\ge2$ then some nonzero section would vanish also at $y$. The statement about $H^1$ follows from RR: \begin{equation} 1=h^0(\Oh_X(K_X+L))\ge\chi(\Oh_X(K_X+L))=\chi(\Oh_X)\ge1. \nonumber \end{equation} \subparagraph{Step E} $x\in\Bs|2K_X-L|$ for general $L\in\Pico X$. For if $D_L\in|2K_X-L|$ does not contain $x$ then $D_L+C_L$ separates $x$ from $Z$ (since by Step~B already $C_L$ separates them). \subparagraph{Step F} For general $L,L_1\in\Pico X$, the point $x$ is a base point of the linear system $\bigl|(2K_X-L_1)_{\textstyle{|C_L}}\bigr|$ on $C_L$, and hence \begin{equation} H^1(m_x\Oh_{C_L}(2K_X-L_1))\ne0. \nonumber \end{equation} This follows from $x\in\Bs|2K_X-L_1|$ because by Step~D, restriction from $X$ maps onto $H^0(\Oh_{C_L}(2K_X-L_1))$. \subparagraph{Step G} We now observe that Step~B implies that $x$ is a singular point of $C_L$. If $x\in\Sing X$ then it is automatically singular on $C_L$. On the other hand, if $x$ is nonsingular on $X$ and on $C_L$, consider the blowup $\si\colon X_1\to X$ of $x$ and the algebraic system $C'_L=\si^*C_L-E$, where $E$ is the exceptional divisor. Let $y\in X_1$ be the point corresponding either to the other point or to the tangent vector of the cluster $Z$. Since the curves $C'_L$ move in a positive dimensional system, there is a curve $C'_L$ through $y$, and therefore a curve $C_L$ containing $Z$, contradicting Step~B. \subparagraph{Step H} For general elements $L,L_2\in\Pico X$, there is an isomorphism $m_x\Oh_{C_L}(L_2)\iso m_x$. This follows as usual by automatic adjunction (Lemma~\ref{lem:adj}) applied to the conclusion $H^1(m_x\Oh_{C_L}(2K_X-L_1))\ne0$ of Step~F, where $L_1=-L-L_2$. We first get a nonzero homomorphism \begin{equation} m_x\Oh_{C_L}(2K_X-L_1)\to\om_{C_L}=\Oh_{C_L}(2K_X+L), \nonumber \end{equation} that is, a map $m_x\Oh_{C_L}(L_2)\to\Oh_{C_L}$; since $C_L$ is 2-connected this must be an inclusion, and the image is the ideal of a point $m_z$. But $x$ is a singular point of $C_L$ (by Step~G), and thus $x=z$. \subparagraph{Step I} Let $\si\colon C'\to C=C_L$ be the blowup at $x$. Step~H implies that $L_2'=\si^*L_2$ is trivial on $C'$ for every general $L_2$, and hence for every $L_2\in\Pico X$ (by the group law). We derive a contradiction from this. Consider the diagram \begin{equation} \renewcommand\arraystretch{1.5} \begin{array}{rcl} \Pico X @>\res_C>> & \Pico C & @>\si^*>> \Pico(C')\\ & \uparrow \\ & G \end{array} \nonumber \end{equation} where $G$ is the kernel of $\si^*$. Now the key point (exactly as in Ramanujam and Francia vanishing) is that $G$ is an affine group scheme. Since the composite $\si^*\circ\res_C$ is zero, $\Pico X$ maps to $G$. Since $\Pico X$ is complete $\res_C$ is the constant morphism to zero. But this is obviously nonsense: for example, since $H^1(\Oh_X(2K_X+L))=0$ for all $L\in\Pico X$, the exact sequence \begin{equation} 0\to\Oh_X(-K_X-L+N)\to\Oh_X(N)\to\Oh_C(N)\to0 \nonumber \end{equation} is exact on global sections if $L\ne N$. Thus $H^0(\Oh_C(N))=0$ and the restriction of $N$ to $C$ is nontrivial. \QED \end{pfof} \section{The bicanonical map}\label{sec:bi} \subsection*{Preliminaries and the proof of Theorem~\ref{th:bi}, (a) and (c)} This section proves Theorem~\ref{th:bi}. We start by remarking that $|2K_X|$ is free. Indeed, for any $C\in|K_X|$, the restriction $\Oh_X(2K_X)\to\Oh_C(K_C)$ is surjective on $H^0$, and $|K_C|$ is free by Theorem~\ref{th:free}. For a cluster $Z$ of degree 2 in $X$, note the following obvious facts: \begin{enumerate} \renewcommand{\labelenumi}{(\roman{enumi})} \item If $Z$ is contracted by $|2K_X|$ then $|K_X|$ does not separate $Z$; thus \begin{equation} h^0(\sI_Z\Oh_X(K_X))\ge p_g-1\quad\text{or}\quad \dim|K_X-Z|\ge p_g-2. \nonumber \end{equation} \item If $|2K_X|$ contracts $Z$ then so does $|K_C|$ for any curve $C\in|K_X-Z|$. \end{enumerate} \begin{pfof}{Theorem~\ref{th:bi}, (a)} We suppose that every curve $C\in|K_X-Z|$ is \hbox{3-connected}, and derive a contradiction from the assumption that $|2K_X|$ contracts $Z$. By Theorem~\ref{th:hh}, every $C\in|K_X-Z|$ is honestly hyperelliptic. As in the proof of Theorem~\ref{th:tri}, Case~(a), it follows that $\fie_{2K}\colon X\to Y$ has degree 2, and maps every $C\in|K_X-Z|$ as a double cover of a curve $\Ga_C\subset Y$ isomorphic to $\proj^1$. Then $\Ga_C$ for $C\in|K_X-Z|$ form an algebraic subfamily of a linear system of curves through $y=\fie_{2K}(Z)$, with a complete parameter space of dimension $\ge2$. As before, this contradicts Lemma~\ref{lem:sing} (but $y\in Y$ may now be singular). \QED \end{pfof} \begin{DEF}\label{def:Frc} Let $X$ be a projective surface with at worst Du Val singularities and with $K_X$ nef. A {\em Francia curve} or {\em Francia cycle} is an effective Weil divisor $B$ on $X$ satisfying \begin{equation} K_XB=p_aB=1\text{ or }2. \nonumber \end{equation} If $K_X$ is ample and $B$ is Gorenstein (for example if $B$ is a Cartier divisor), it is clearly either an irreducible curve of genus 1, or a numerically 2-connected curve of arithmetic genus $p_a=2$. It would be interesting to know if $B$ is necessarily Gorenstein. \end{DEF} \begin{pfof}{Theorem~\ref{th:bi}, (b) $\implies$ (c)} The argument is standard and we omit some details. Suppose that the 2-canonical map $\fie=\fie_{2K}\colon X\to Y$ is not birational. Every point $x\in X$ is contained in a cluster $Z$ of degree 2 contracted by $\fie$; we choose $x\in\NonSing X$. Theorem~\ref{th:bi}, (b) gives a Francia curve $B_0\subset X$ through $Z$. Write $S\to X$ for the minimal nonsingular model of $X$ and $B=\Bh_0$ for the hat transform of $B_0$ (as in the proof of Lemma~\ref{lem:n-conn}). Then by Claim~\ref{cla:bhat}, $B$ is also a Francia cycle on $S$, that is, $1\le K_SB=p_aB\le2$. An easy argument in quadratic forms shows that there are at most finitely many effective divisors $B\subset S$ with $K_SB=1$ and $B^2=-1$ (compare \cite{Bo}, pp.~191--192 or \cite{BPV}, p.~224). Therefore every general point of $S$ is contained in a curve $B$ with $K_SB=p_aB=2$, and hence $B^2=0$. Now the same argument in quadratic forms shows that divisors with $K_SB=2$ and $B^2=0$ belong to finitely many numerical equivalence classes, so one class must contain an algebraic family of curves. This gives a genus 2 pencil on $S$, and therefore also on $X$. \QED \end{pfof} We use the following obvious lemma at several points in what follows. \begin{LEM}[Dimension lemma]\label{lem:dim} Let $\eta\subset X$ be a cluster of degree $d$ which is contracted by $|2K_X|$, and $C\in|K_X|$ a curve containing $\eta$. Then \begin{equation} h^1(\sI_\eta\Oh_C(K_C))=\dim\Hom(\sI_\eta,\Oh_C)=d. \nonumber \end{equation} In particular, for any $x\in C$, we have \begin{equation} h^1(m_x^2\Oh_C(K_C))=\dim\Hom(m_x^2\Oh_C,\Oh_C)=1+\dim T_{\fie,x}\le4, \nonumber \end{equation} where $T_{\fie,x}$ is the Zariski tangent space to the scheme theoretic fibre of $\fie_{2K_X}$ through $x$. \end{LEM} \begin{pf} Since $|K_C|$ is free and contracts $\eta$, the evaluation map $H^0(\Oh_C(K_C))\to\Oh_\eta(K_C)=k^d$ has rank 1, so that $h^1(\sI_\eta\Oh_C(K_C))=d$ comes from the exact sequence \begin{equation} \renewcommand\arraystretch{1.3} \begin{array}{l} 0\to H^0(\sI_\eta\Oh_C(K_C))\to H^0(\Oh_C(K_C))\to k^d \\ \hphantom{0}\to H^1(\sI_\eta\Oh_C(K_C))\to H^1(\Oh_C(K_C))=k. \end{array} \nonumber \end{equation} As usual, Serre duality gives \begin{equation} \Hom(\sI_\eta,\Oh_C)=\Hom(\sI_\eta\Oh_C(K_C),\om_C)\dual H^1(\sI_\eta\Oh_C(K_C)). \nonumber \end{equation} We obtain the last part by taking $\eta$ to be the intersection of the scheme theoretic fibre $\fie\1(\fie(x))$ with the subscheme $V(m_x^2)\subset C$ corresponding to the tangent space. \QED \end{pf} \subsection*{Case division and plan of proof of~(b)} Throughout this section, $Z$ is a cluster of degree 2, and we argue by restricting to a curve $C\in|K_X-Z|$, usually imposing singularities on $C$ at a point $x\in Z$. As usual, the assumption that $Z$ is contracted by $K_C$ gives a homomorphism $\sI_Z\to\Oh_C$ linearly independent of the identity inclusion. By passing to a suitable linear combination $s'=s+\la\id$ if necessary, we assume that $s\in\Hom(\sI_Z,\Oh_C)$ is injective, and hence $s(\sI_Z)=\sI_{Z'}$ for some cluster $Z'$ of degree 2; the family of clusters $Z'$ as $s$ runs through injective elements $s\in\Hom(\sI_Z,\Oh_C)$ is an analog of a $g^1_2$ on $C$. The argument is modelled on the proof of Theorem~\ref{th:hh}. As there, we use different arguments depending on how $Z$ and $Z'$ intersect, or, to put it another way, how $Z'$ moves as $s$ runs through injective elements $s\in\Hom(\sI_Z,\Oh_C)$. (In other words, how the $g^1_2$ corresponding to $\Hom(\sI_Z,\Oh_C)$ breaks up into a ``base locus'' plus a ``moving part''.) Let $s\in\Hom(\sI_Z,\Oh_C)$ be a general element, and $\sI_{Z'}=s(\sI_Z)$. Logically, there are 4 cases for $Z$ and $Z'$. \begin{enumerate} \item $\Supp Z\cap\Supp Z'=\emptyset$. \item $\Supp Z\cap\Supp Z'\ne\emptyset$, but $\Supp Z\ne\Supp Z'$. \item $Z=Z'$. \item $Z\ne Z'$ are nonreduced clusters supported at the same point $x\in X$. \end{enumerate} In Case~2, $|Z|$ has a fixed point plus a moving point; as we see in Lemma~\ref{lem:case2}, this contradicts $K_X$ ample. In Case~1, $|Z|$ is a free $g^1_2$, and the isomorphism $\sI_Z\iso\sI_{Z'}$ with $\Supp Z\cap\Supp Z'=\emptyset$ implies that $\sI_Z$ is locally free, so that $Z$ is a Cartier divisor on $C$. If $p_g\ge4$, it turns out that we can choose $C$ to be ``sufficiently singular'' at a point $x\in Z$ so that $Z\subset C$ is not Cartier, and Case~1 is excluded for such $C$ (see Lemma~\ref{lem:notCt}). In Cases~3--4, when the support of $Z$ does not move, we must find a map $s'\colon\sI_Z\to\Oh_C$ vanishing on a ``fairly large'' portion of $C$, so that its scheme theoretic support $B\subset C$ is ``fairly small''. The key idea is to look for $s'$ as a nilpotent or idempotent (see Lemma~\ref{lem:pot} and Corollary~\ref{cor:Art}). The assumption of Case~3 is $\Hom(\sI_Z,\Oh_C)=\End(\sI_Z)$, which is a 2-dimensional Artinian algebra; this makes it is rather easy to find a nilpotent or idempotent element, and to prove Theorem~\ref{th:bi}, (b). In Case~4, $Z'$ is $x$ plus a tangent vector $y$ which moves in $T_{C,x}$ as $s\in\Hom(\sI_Z,\Oh_C)$ runs through injective elements; this is an {\em infinitesimal} $g^1_2$, an interesting geometric phenomenon in its own right (see Remark~\ref{rem:g23} and the proof of Proposition~\ref{pro:m2}, Step~6 for more details). The key point in this case is to prove that the extra homomorphism $s\colon\sI_Z\to\Oh_C$ takes $m_x^2$ to itself, so that $\End(m_x^2)$ is a nontrivial Artinian algebra; see Proposition~\ref{pro:m2}. \begin{REM}\label{rem:g23} In Case~4, reversing the usual argument proves that $\fie_{K_C}$ also contracts $Z'$, and so it contracts a cluster $\eta$ of degree $\ge3$ contained in the first order tangent scheme $V(m_x^2)\subset C$. If $C$ is numerically 3-connected, this is of course impossible by Theorem~\ref{th:hh}. In this case, $\Hom(\sI_\eta,\Oh_C)$ is a certain analog of a $g^2_3$ or $g^3_4$ on $C$. Case~4 certainly happens on abstract numerically 2-connected Gorenstein curves, and more generally, the analog of a $g^{m-1}_m$. Example: let $C_i$ for $i=1,\dots,m$ be nonhyperelliptic curves of genus $g_i\ge3$ with marked points $x_i\in C_i$, and assemble the $C_i$ into a curve $C=\bigcup C_i$ by glueing together all the $x_i$ to one point $x$, at which the tangent directions are subject to a single nondegenerate linear relation, so that the singularity $x\in C$ is analytically equivalent to the cone over a frame of reference $\{P_1,P_2,\dots,P_m\}$ in $\proj^{m-2}$. Then $C$ is Gorenstein and $K_C$ restricted to each $C_i$ is $K_{C_i}+2x_i$ (see \cite{Ca1}, Proposition~1.18, (b), p.~64, or \cite{Re}, Theorem~3.7), so that $|K_C|$ contracts the whole $(m-1)$-dimensional tangent space $T_{C,x}$ to a point. A cluster $Z$ of degree 2 supported at $x$ corresponds to a point $Q\in\proj^{m-2}=\proj(T_{C,x})$. Since $Z$ is contracted by $K_C$ (together with the whole tangent space), by our usual argument, the group $\Hom(\sI_Z,\Oh_C)$ is 2-dimensional and a general $s\colon\sI_Z\to\Oh_C$ has image $\sI_{Z'}$ where $Z'$ is a moving cluster of degree 2 at $x$, corresponding to a moving point $Q'\in\proj^{m-2}$. It is an amusing exercise to see that if $Q$ is linearly in general position with respect to the frame of reference $\{P_1,P_2,\dots,P_m\}$ then $Q'$ moves around the unique rational normal curve of degree $m-2$ passing through $\{P_1,P_2,\dots,P_m,Q\}$. On the other hand, if $Z$ is in the tangent cone to $C$ (say, tangent to the branch $C_1$), then $\sI_Z$ is not isomorphic to any other cluster of degree 2, so that $\Hom(\sI_Z,\Oh_C)=\End(\sI_Z)$; this has 2 idempotents vanishing on $C_1$ and on $C_2+\cdots+C_m$. The following easy exercises may help to clarify things for the reader: \begin{enumerate} \item Let $x\in C$ be an ordinary triple point of a plane curve, say defined by an equation $f(u,v)=u^3+v^3+$ higher order terms; then for general $\la$, the ideals $(u+\la v,v^2)$ in $\Oh_{C,x}$ are all locally isomorphic. [Hint: Multiply by the rational function $(u+\mu v)/(u+\la v)$.] \item If $C$ is the planar curve defined by $vw=v^3+w^3$ then $m_x=(v,w)$ is locally isomorphic to $\sI_Z=(v,w^2)$ and to $\sI_{Z'}=(v^2,w)$. \item If $C$ is the planar curve locally defined by $v^2=w^3$ then $m_x=(v,w)$ is locally isomorphic to $\sI_Z=(v,w^2)$. \end{enumerate} (Compare the proof of Proposition~\ref{pro:m2}, Step~6.) \end{REM} \begin{LEM}\label{lem:case2} Case~2 is impossible. \end{LEM} \begin{pf} Since $x\in Z\cap Z'$ and $\Supp Z\ne\Supp Z'$, we can interchange $Z$ and $Z'$ if necessary and assume that $Z'=\{x,y\}$ with $x\ne y$. Consider the inclusion $s\colon\sI_Z\into\Oh_C$ with image $s(\sI_Z)=\sI_{Z'}=m_xm_y$ and the identity inclusion. One of these vanishes at $y$ and the other doesn't, so their restrictions to a component $\Ga$ containing $y$ are linearly independent on $\Ga$, and, as in Claim~\ref{cla:mov_y}, for any general point $y'\in\Ga$, some linear combination $s'=s+\la\id$ defines an isomorphism $s'\colon\sI_Z\iso m_xm_{y'}$. Reversing our usual argument shows that $x$ and $y'$ are contracted to the same point by $|K_C|$ or $|2K_X|$, so that the free linear system $|2K_X|$ contracts $\Ga$ to a point. This contradicts $K_X$ ample. \QED \end{pf} \subsection*{Clusters on singular curves} Our immediate aim is to exclude Case~1, but at the same time we introduce some ideas and notation used throughout the rest of this section. Choose a point $x\in Z$. Since $X$ has at worst hypersurface singularities and $C$ is a Cartier divisor in $X$, it is a local complete intersection, that is, locally defined by $F=G=0$. (Of course, $X$ may be nonsingular.) We think of $x\in Z\subset C\subset X\subset\aff^3$ as local, and write $\Oh_{\aff^3}$, $\Oh_C$, etc.\ for the local rings at $x$. We take local coordinates $u,v,w$ in $\aff^3$ so that $Z$ is defined by $u=v=w=0$ in the reduced case, or $u=v=w^2=0$ otherwise. \begin{LEM}\label{lem:notCt} \begin{enumerate} \renewcommand{\labelenumi}{{\rm(\arabic{enumi})}} \item The quotient $\sI_{\aff^3,Z}/m_{\aff^3,x}\sI_{\aff^3,Z}$ is a $3$-dimensional vector space, and $Z\subset C$ is a Cartier divisor at $x$ if and only if $F,G$ map to linearly independent elements of it. \item Suppose that $p_g\ge4$ and $Z$ is contracted by $|2K_X|$. Then the curve $C\in|K_X-Z|$ can be chosen such that $Z$ is not a Cartier divisor. For this $C$, Case~1 is excluded. \end{enumerate} \end{LEM} \begin{pf} (1) says that a minimal set of generators of the ideal $\sI_{\aff^3,Z}$ consists of 3 elements, which is obvious because $\sI_{\aff^3,Z}$ is locally generated at $x\in Z$ by the regular sequence $(u,v,w)$ or $(u,v,w^2)$. Now $Z$ is a Cartier divisor on $C$ if and only if $\sI_{C,Z}$ is generated by 1 element, that is, $F$ and $G$ provide two of the minimal generators of $\sI_{\aff^3,Z}$. This proves (1). For (2), suppose that $F=0$ is the local equation of $X\subset\aff^3$. If $F\in m_{\aff^3,x}\sI_{\aff^3,Z}$ then by (1), $Z$ is not a Cartier divisor on any curve $C\in|K_X-Z|$. Suppose then that $F\notin m_{\aff^3,x}\sI_{\aff^3,Z}$, so that $F$ provides one of the minimal generators of $\sI_{\aff^3,Z}$. Then the ideal $\sI_{X,Z}$ of $Z\subset X$ is generated by 2 elements, in other words, $\dim_k\sI_{X,Z}/m_{X,x}\sI_{X,Z}=2$. Therefore \begin{equation} h^0(m_x\sI_Z\Oh_X(K_X))\ge h^0(\sI_Z\Oh_X(K_X))-2\ge p_g-3\ge1 \nonumber \end{equation} (by remark (i) at the beginning of this section). Thus we can find a curve $C\in|K_X-Z|$ whose local equation at $x$ is $g\in m_{X,x}\sI_{X,Z}$. Then $g$ has a local lift $G\in m_{\aff^3,x}\sI_{\aff^3,Z}$, so that (1) applies to $C$. \QED \end{pf} \begin{REM}\label{rem:geom} The same argument can be expressed more geometrically. If $Z$ contains $x$ as a reduced point, that is, $\sI_{\aff^3,Z}=m_x$, then $x\in C$ is Cartier if and only if $C$ defined by $(F,G)$ is nonsingular at $x$, that is, $F,G$ map to linearly independent elements of $m_x/m_x^2$. To interpret the nonreduced case $\sI_{\aff^3,Z}=(u,v,w^2)$, note that \begin{equation} F\notin m_{\aff^3,x}\sI_{\aff^3,Z} \iff F=Pu+Qv+Rw^2 \quad\text{with one of $P,Q,R\notin m_x$.} \nonumber \end{equation} In other words, the surface $Y$ locally defined by $F=0$ is either nonsingular at $x$, or has a double point with $Z$ not in the tangent cone. In the opposite case $F\in m_{\aff^3,x}\sI_{\aff^3,Z}$, it is easy to see that $x\in C$ is either a complete intersection defined by two singular hypersurfaces, so has 3-dimensional tangent space $T_{C,x}$, or is a planar curve, which is either a double point with $Z$ in the tangent cone, or a point of multiplicity $\ge3$. \end{REM} \subsection*{The nilpotent--idempotent lemma} Our proof of Theorem~\ref{th:bi}, (b) in Cases~3--4 is based on the following result. Note first that $\Hom(\sI_Z,\Oh_C)\subset H^0(C\setminus\Supp Z,\Oh_C)$, and the latter is a ring. (We usually write $\sI_Z$ for $\sI_{C,Z}$ in what follows.) In other words, maps $\sI_Z\to\Oh_C$ can be viewed as rational sections of $\Oh_C$ that are regular outside $\Supp Z$, so that it is meaningful to multiply them (the product is again a rational section of $\Oh_C$ that is regular away from $Z$). \begin{LEM}\label{lem:pot} Assume that $K_X^2\ge10$, and let $C\in|K_X-Z|$. Suppose that $s\colon\sI_Z\to\Oh_C$ is a nonzero homomorphism which is either nilpotent with $s^4=0$, or a nontrivial idempotent with $s(1-s)=0$. Then the scheme theoretic support of $s$ (respectively, in the idempotent case, either $s$ or $1-s$) is a Francia curve $B$, and $\sI_Z\Oh_B(2K_X)\iso\om_B$. More generally, suppose that $s_i\colon\sI_Z\to\Oh_C$ for $i=1,\dots,4$ are nonzero homomorphisms such that $s_1s_2s_3s_4=0$. Then one of the $s_i$ has scheme theoretic support a Francia curve $B_i$ with $\sI_Z\Oh_{B_i}(2K_X)\iso\om_{B_i}$. \end{LEM} The final part is more general, because we allow some $s_i=\id$, or some of the $s_i$ to coincide. Notice that $\Oh_C$ has no sections supported at finitely many points, so we need only check the conditions $s^4=0$ etc.\ in each generic stalk of $\Oh_C$, that is, as rational functions on $C$. \begin{pf} If $s\colon\sI_Z\Oh_C(K_C)\to\om_C$ is not generically injective, the factorisation provided by automatic adjunction (Lemma~\ref{lem:adj}) gives a subcurve $B\subset C$ satisfying $\sI_Z\Oh_B(K_C)\iso\om_B$; we are in the limiting case of numerically 2-connected. Write $C=A+B$ for the decomposition of Weil divisors, so that $A$ is the divisor of zeros of $s$. Passing to the minimal nonsingular model $S$ and taking the hat transform $\Bh$ as in Lemma~\ref{lem:n-conn} and Claim~\ref{cla:bhat} gives a decomposition $K_S\lineq f^*C=A_1+\Bh$ such that $A_1\Bh=2$. Therefore by the Hodge algebraic index theorem, $A_1^2\Bh^2\le(A_1\Bh)^2=4$. If both $A_1^2$, $\Bh^2\ge1$, it follows that $K_S^2\le9$, a contradiction, so that either $A_1^2\le0$ or $\Bh^2\le0$. Then (because $K_S=A_1+\Bh$ and $A_1\Bh=2$), either $K_XA=K_SA_1\le2$ or $K_XB=K_S\Bh\le2$. Suppose for the moment that $K_S\Bh\le2$. Since $2p_a\Bh-2=\Bh^2+K_S\Bh$, it follows at once that we are in one of the two cases \begin{equation} \Bh^2=-1,K_S\Bh=1,p_a\Bh=1\quad\text{or}\quad \Bh^2=0,K_S\Bh=2,p_a\Bh=2. \nonumber \end{equation} But by Lemma~\ref{lem:n-conn} and Claim~\ref{cla:bhat} we have $K_S\Bh=K_XB$ and $p_a\Bh=p_aB$, so that $B$ is the required Francia curve. It remains to get rid of the possibility that $K_XA=K_SA_1\le2$ in the different cases. If $s$ is a nontrivial idempotent, we can swap $A\bij B$ by $s\bij1-s$ if necessary, so that $K_XB\le2$. In the nilpotent case, since $A$ equals the Weil divisor of zeros of $s$ and $s^4=0$, it follows that $C\le4A$. Then $K_XA\le2$ would imply $K_X^2\le8$, a contradiction. The last part is exactly the same: each $s_i$ (for $i=1,2,3,4$) is either injective or has scheme theoretic support a subcurve $B_i\subset C$ with $\sI_Z\Oh_{B_i}(K_C)\iso\om_{B_i}$, and divisor of zeros $A_i=C-B_i$. Since $\prod s_i=0$ it follows that $C\le\sum A_i$. Now arguing as above gives that one of $K_XA_i$ or $K_XB_i\le2$; if the first alternative holds for all $i$ then $K_X^2=K_XC\le \sum K_XA_i\le8$, a contradiction. This proves the lemma. \QED \end{pf} We apply Lemma~\ref{lem:pot} via a simple algebraic trick. \begin{COR}\label{cor:Art} If $A=\End_{\Oh_C}(\sI_{C,Z})$ is an Artinian algebra of length $\ge2$ then it has a nontrivial idempotent or a nonzero nilpotent with $s^2=0$. More generally, if\/ $\Hom(\sI_{C,Z},\Oh_C)$ is a $2$-dimensional vector space contained in an Artinian algebra $A\subset H^0(C\setminus\Supp Z,\Oh_C)$ of dimension $\le4$ then there exist nonzero elements $s_1,\dots,s_4\in\Hom(\sI_{C,Z},\Oh_C)$ with zero product. Under either assumption, Lemma~\ref{lem:pot} gives a Francia curve $B\subset C$ containing $Z$. \end{COR} This completes the proof of Theorem~\ref{th:bi}, (b) in Case~3, since the case assumption is that $s\colon\sI_Z\to\sI_Z\subset\Oh_C$, so that $\Hom(\sI_Z,\Oh_C)=\End(\sI_Z)$ is a 2-dimensional Artinian algebra. \begin{pf} In the main case $\dim A=2$, this is completely trivial: if $k\subset A$ is the constant subfield, any $s\in A\setminus k$ satisfies a quadratic equation over $k$ of the form \begin{equation} 0=s^2+as+b=(s-\al)(s-\be). \nonumber \end{equation} If $\al=\be$ then $s'=s-\al$ is nilpotent with $s'{}^2=0$; otherwise, $s'=(s-\al)/(\al-\be)$ and $1-s'=(s-\be)/(\be-\al)$ are orthogonal idempotents. More generally, an Artinian algebra is a product $A=A_1\times\cdots\times A_l$ with local Artinian rings $(A_i,n_i)$ as factors; the maximal ideals of $A$ are codimension 1 vector subspaces $m_i\subset A_i$ given by $n_1\times A_2\times\cdots\times A_l$ (say). The projection to the factors (if $l\ge2$) give nontrivial idempotents; if $l=1$ then $A$ is local, with nilpotent maximal ideal. This proves the first part. We now prove the more general statement: a 2-dimensional vector subspace $V\subset A$ in an Artinian algebra has nonzero intersection with every maximal ideal, say $s_i\in V\cap m_i$. If the local factors $(A_i,n_i)$ have dimension $d_i$ then $n_i^{d_i}=0$, and the product $\prod s_i^{d_i}$ maps to zero in each factor, so is zero in $A$. Taking $\sum d_i=\dim A\le4$ gives the final part of the claim. \QED \end{pf} \subsection*{Proof in Case~4} In the following proposition, $x\in C\subset\aff^3$ is a {\em local} curve which is a local complete intersection at $x$. We choose local coordinates $u,v,w$ on $\aff^3$ so that $\sI_{\aff^3,Z}\subset\Oh_{\aff^3}$ is generated at $x$ by the regular sequence $u,v,w^2$. As before, we write $\Oh_C$ for the local ring $\Oh_{C,x}$ and $\sI_Z=\sI_{C,Z}$ for the $\Oh_C$ module obtained as the stalk at $x$ of the corresponding ideal sheaf. (Thus the statement of the proposition only concerns homomorphisms $s\colon\sI_Z\to\Oh_C$ of modules over the local ring $\Oh_C$.) \begin{PROP}\label{pro:m2} Let $Z\subset C$ be a cluster of degree $2$ supported at $x$. We assume \begin{enumerate} \renewcommand{\labelenumi}{\rm(\roman{enumi})} \item $Z$ is not a Cartier divisor on $C$; \item there exists a homomorphism $s_0\colon\sI_Z\to\Oh_C$ such that for general $\la\in k$, $s_0+\la\id$ defines an isomorphism $\sI_Z\iso\sI_{Z_\la}$ with $Z_\la$ a cluster of degree $2$ supported at $x$, and $Z_0\ne Z$. \end{enumerate} Then any homomorphism $s\colon\sI_Z\to\Oh_C$ takes $m_{C,x}^2$ to $m_{C,x}^2$, that is, \begin{equation} \renewcommand\arraystretch{1.5} \begin{matrix} \sI_Z & @>{\quad s\quad}>> & \Oh_C\\ \bigcup && \bigcup\\ m_x^2 & @>{\hphantom{\quad s\quad}}>> & m_x^2 \end{matrix} \nonumber \end{equation} \end{PROP} \begin{pfof}{Theorem~\ref{th:bi}, (b) in Case~4} We apply the proposition to the {\em global} homomorphism $s\colon\sI_Z\to\Oh_C$, using the assumption of Case~4. We get \begin{equation} \Hom(\sI_Z,\Oh_C) \subset \End(m_x^2) \subset \Hom(m_x^2,\Oh_C). \nonumber \end{equation} Now Lemma~\ref{lem:dim} gives $\dim\Hom(\sI_Z,\Oh_C)=2$ and $\dim\Hom(m_x^2,\Oh_C)\le4$; but $A=\End(m_x^2)$ is a subring of $H^0(C\setminus\Supp Z,\Oh_C)$, so that Corollary~\ref{cor:Art} gives the result. \QED \end{pfof} \begin{pfof}{Proposition~\ref{pro:m2}, Step 1} If $s\in\Hom(\sI_Z,\Oh_C)$ is any element then $s(\sI_Z)\subset m_x$; for otherwise $s$ would be an isomorphism $\sI_Z\iso\Oh_C$ near $x$, contradicting the assumption that $Z\subset C$ is not Cartier. \subparagraph{Step 2} Note that $m_x^2\subset\sI_Z$, so that we can restrict $s\colon\sI_Z\to\Oh_C$ to $m_x^2$. Also, $m_x\sI_Z\subset m_x^2$, and obviously $s(\sI_Z)\subset m_x$ implies that $s(m_x\sI_Z)\subset m_x^2$. \subparagraph{Step 3} It is enough to prove that $s(w^2)\in m_x^2$. Indeed, \begin{equation} m_x\sI_Z=(u,v,w)\cdot(u,v,w^2)=(u^2,uv,v^2,uw,vw,w^3), \nonumber \end{equation} so that \begin{equation} m_x^2=(u,v,w)^2=(u^2,uv,v^2,uw,vw,w^2)=m_x\sI_Z+\Oh_Cw^2\subset\Oh_C. \nonumber \end{equation} \subparagraph{Step 4} Since $C$ is a local complete intersection, $\sI_{\aff^3,C}=(F,G)$, where $F,G\in\Oh_{\aff^3}$ is a regular sequence. Now $Z\subset C$ gives $F,G\in\sI_{\aff^3,Z}$, so that \begin{equation} \begin{aligned} F&=Pu+Qv+Rw^2,\\ G&=P'u+Q'v+R'w^2, \end{aligned} \quad\text{with}\quad P,Q,R,P',Q',R'\in\Oh_{\aff^3}. \end{equation} The set of local homomorphisms $\sI_Z\to\Oh_C$ is a module over $\Oh_C$; this is the stalk at $x$ of the sheaf $\sHom$. For the moment, we take on trust the following general fact (see Appendix to \S\ref{sec:bi} for a discussion and a detailed proof.) \begin{CLA}\label{cla:PQ-PQ} The $\Oh_C$ module $\sHom_{\Oh_C}(\sI_Z,\Oh_C)$ is generated by two elements, the identity inclusion $\id\colon\sI_{C,Z}\into\Oh_C$ and the map $t\colon\sI_{C,Z}\to\Oh_C$ determined by the minors of the $2\times3$ matrix of coefficients of $F,G$: \begin{equation} t(u)=QR'-RQ',\quad t(v)=-PR'+RP',\quad t(w^2)=PQ'-QP'. \end{equation} \end{CLA} \subparagraph{Step 5} According to Steps~3--4, to prove Proposition~\ref{pro:m2}, we need only prove that $PQ'-QP'\in m_{\aff^3,x}^2$. We are home if all four of $P,Q,P',Q'\in m_x$. Thus in what follows, we assume (say) that $P'\notin m_x$. Then $P'$ is a unit, and $G=0$ defines a nonsingular surface $Y$ containing $C$. Dividing by $P'$, we can rewrite $G$ in the form $u=-(Q'/P')v-(R'/P')w^2$. Then subtracting a multiple of this relation from $F$ gives $f=qv+rw^2$ as the local equation of $C\subset Y$ (where $q=Q-PQ'/P'$ and $r=R-PR'/P'$). Therefore it only remains to prove that if $C$ is the planar curve defined by $f=qv+rw^2$, the two assumptions of Proposition~\ref{pro:m2} imply that $q\in m_{Y,x}^2$. As in Lemma~\ref{lem:notCt}, assumption (i) implies that $q,r\in m_{Y,x}$, so that $q\in m_{Y,x}^2$ is equivalent to saying that $x\in C\subset Y$ has multiplicity $\ge3$ \subparagraph{Step 6} Consider the linear terms of the given isomorphism $s_0\colon\sI_Z\to\sI_{Z_0}$: \begin{equation} s_0(v)=av+bw\mod m_{Y,x}^2,\quad s_0(w^2)=cv+dw\mod m_{Y,x}^2. \nonumber \end{equation} Because $Z_0\ne Z$, it follows that $(b,d)\ne(0,0)$. However, if $b=0$ and $d\ne0$, then for general $\la$, the two generators of $\sI_{Z_\la}=(s_0(v)+\la v,s(w^2)+\la w^2)$ would have linearly independent linear terms, so that $\sI_{Z_\la}=m_{C,x}$. This contradicts assumption (ii). Therefore $b\ne0$, and $\sI_{Z_\la}$ has a generator with the {\em variable} linear term $(a+\la)v+bw$. It follows that $Z_\la$ runs linearly around the tangent space to $x$ in $C$. Now we claim that $x\in C\subset Y$ is a planar curve singularity of multiplicity $\ge3$. Indeed, the isomorphism $\sI_Z\iso\sI_{Z_\la}$ implies that $Z_\la\subset C$ cannot be a Cartier divisor; but if $x\in C\subset Y$ were a double point, this would restrict $Z_\la$ to be in the tangent cone, contradicting what we have just proved. This completes the proof of Proposition~\ref{pro:m2}. \QED \end{pfof} \subsection*{Appendix: Proof of Claim~\ref{cla:PQ-PQ}} We start by slightly generalising the set-up: let $\Oh_{\aff}$ be a local ring, assumed to be regular (for simplicity only), and $x,y,z$ a regular sequence generating a codimension 3 complete intersection ideal $\sI_Z=(x,y,z)$. Consider a regular sequence $F,G\in\sI_Z$. Note that \begin{equation} F=Px+Qy+Rz\quad\text{and}\quad G=P'x+Q'y+R'z \nonumber \end{equation} for some $P,\dots,R'\in\Oh_{\aff}$. Write $\Oh_C=\Oh_{\aff}/(F,G)$ and $\sI_{C,Z}=\sI_Z\Oh_C=(x,y,z)\subset\Oh_C$. (In the application, $Z\subset\aff=\aff^3$ was a nonreduced cluster defined by $(x,y,z)=(u,v,w^2)$ and $C\subset\aff^3$ a complete intersection curve through $Z$.) \begin{LEM} \begin{enumerate} \renewcommand{\labelenumi}{\rm(\arabic{enumi})} \item A presentation of $\sI_{C,Z}$ over $\Oh_C$ is given by \begin{equation} \Oh_C^{\oplus5}@>M>>\Oh_C^{\oplus3} @>\left(\begin{matrix} x\\y\\z\end{matrix}\right)>>\sI_{C,Z}\to0, \quad\text{where}\quad \renewcommand\arraystretch{1.2} M=\left(\matrix P&Q&R\\ P'&Q'&R'\\ 0&z&-y\\ -z&0&x\\ y&-x&0 \endmatrix\right). \nonumber \end{equation} \item $\sHom(\sI_{C,Z},\Oh_C)$ is generated over $\Oh_C$ by the two elements $\id$ and $t$, where \begin{equation} t\colon\left(\matrix x\\y\\z\endmatrix\right)\mapsto \left(\matrix QR'-RQ'\\-PR'+RP'\\PQ'-QP'\endmatrix\right). \label{eq:t} \end{equation} \end{enumerate} \end{LEM} \begin{pf} (1) An almost obvious calculation: because $\sI_{C,Z}=(x,y,z)$, there is a surjective map $\fie\colon\Oh_C^{\oplus3}\to\sI_{C,Z}$, such that $(h_1,h_2,h_3)\in\ker\fie$ if and only if $h_1x+h_2y+h_3z=0\in\Oh_C$. Write $H_1,H_2,H_3\in\Oh_\aff$ for lifts of the $h_i$. Then $H_1x+H_2y+H_3z\in\sI_{\aff^3,C}=(F,G)$. Subtracting off multiples of $F$ and $G$ means exactly subtracting multiples of the first two rows of $M$ from $(H_1,H_2,H_3)$, to give identities $H_1'x+H_2'y+H_3'z=0\in\Oh_\aff$. Now $x,y,z\in\Oh_C$ is a regular sequence, so it follows that $(H'_1,H'_2,H'_3)$ is in the image of the Koszul matrix given by the bottom 3 rows of $M$. This proves (1). (2) A homomorphism $s\colon\sI_{C,Z}\to\Oh_C$ is determined by $(x,y,z)\mapsto(a,b,c)$ where $a,b,c\in\Oh_C$ satisfy $M(a,b,c)^\mathrm{tr}=0$ (we write $(a,b,c)^\mathrm{tr}$ for the column vector). It is easy to check that (\ref{eq:t}) gives a map $t$ in this way. The condition $M(a,b,c)^\mathrm{tr}=0$ consists of 5 equalities in $\Oh_C=\Oh_{\aff}/(F,G)$. We choose lifts $A,B,C$ to $\Oh_{\aff}$, and write out the last 3 of these as identities in $\Oh_{\aff}$: \begin{equation} \begin{array}{cccl} &-zB&+yC&=\al F -\al'G \\ zA&&-xC&=\be F -\be'G \\ -yA&+xB&&=\ga F -\ga'G \end{array} \quad\text{for some $\al,\dots,\ga'\in\Oh_{\aff}$.} \label{eq:al} \end{equation} Taking $x$ times the first plus $y$ times the second plus $z$ times the third, the left-hand sides cancel, giving the identity \begin{equation} (\al x+\be y+\ga z)F=(\al'x+\be'y+\ga'z)G\in\Oh_{\aff}. \nonumber \end{equation} Now since $F,G$ is a regular sequence in $\Oh_{\aff}$, this implies that \begin{equation} \begin{aligned} \al x+\be y+\ga z&=DG=D(P'x+Q'y+R'z)\\ \al'x+\be'y+\ga'z&=DF=D(Px+Qy+Rz) \end{aligned} \label{eq:D} \end{equation} for some $D\in\Oh_{\aff}$. Now subtracting $D$ times the given generator $t$ changes \begin{equation} \left(\begin{matrix} A\\B\\C \end{matrix}\right) \mapsto \left(\begin{matrix} A\\B\\C \end{matrix}\right) -\left(\begin{matrix} QR'-Q'R\\-PR'+P'R\\PQ'-P'Q\\ \end{matrix}\right)D \nonumber \end{equation} and has the following effect on the quantities $\al,\dots,\ga'$ introduced in (\ref{eq:al}): \begin{align*} (\al,\be,\ga)&\mapsto(\al+DP',\be+DQ',\ga+DR'),\\ (\al',\be',\ga')&\mapsto(\al'+DP,\be'+DQ,\ga'+DR). \end{align*} To see this, note that the first equation of (\ref{eq:al}) is \begin{equation} -zB+yC=\al F -\al'G=\al(Px+Qy+Rz)-\al'(P'x+Q'y+R'z), \nonumber \end{equation} so that the effect of the two substitutions $\al\mapsto\al+DP'$ and $\al'\mapsto\al'+DP$ on the right exactly cancels out $B\mapsto B+D(PR'-P'R)$ and $C\mapsto C-D(PQ'-P'Q)$ on the left. The upshot is that we can assume $D=0$ in (\ref{eq:D}). But then since $(x,y,z)$ is a regular sequence, (\ref{eq:D}) with $D=0$ gives \begin{equation} \begin{array}{rccc} \al=&&ly&-mz\\ \be=&-lx&&+nz\\ \ga=&mx&-ny \end{array} \quad\text{and}\quad \begin{array}{rccc} \al'=&&l'y&-m'z\\ \be'=&-l'x&&+n'z\\ \ga'=&m'x&-n'y \end{array} \nonumber \end{equation} for some $l,\dots,n'\in\Oh_\aff$. Finally (\ref{eq:al}) can now be rearranged as \begin{equation} \begin{aligned} (C-lF+l'G)y&=(B-mG+m'G)z\\ (C-lF+l'G)x&=(A-nF+n'G)z\\ (B-mF+m'G)x&=(A-nF+n'G)y \end{aligned} \quad\text{therefore}\quad \begin{aligned} A-nF+n'G&=Ex\\ B-mF+m'G&=Ey\\ C-lF+l'G&=Ez \end{aligned} \nonumber \end{equation} for some $E\in\Oh_\aff$. This means that the map $s$ given by $(a,b,c)$ is a linear combination of $t$ and the identity, as required. \QED \end{pf} A less pedestrian method of arguing is to say that all three of $\Oh_\aff$, $\Oh_C$ and $\Oh_Z$ are Gorenstein, so that adjunction gives \begin{equation} 0\to\om_C\to\sHom(\sI_Z,\om_C)\to\om_Z=\Ext^1(\Oh_Z,\om_C)\to0. \nonumber \end{equation} The two generators $\id$ and $t$ correspond naturally to the generators of $\om_C$ and $\om_Z$.
1996-07-04T13:19:50
9607
alg-geom/9607005
en
https://arxiv.org/abs/alg-geom/9607005
[ "alg-geom", "math.AG" ]
alg-geom/9607005
Sandro Manfredini
Fabrizio Catanese, Sandro Manfredini
The orbifold fundamental group of Persson-Noether-Horikawa surfaces
LaTeX file, 19 pages with 1 figure
null
null
null
null
The Noether-Horikawa surfaces are the minimal surfaces S with K^2=2p_g-4. For 8 | K^2 they belong to two families of respective type C and N (connected, resp. non connected branch locus for the canonical map). For 16 | K^2 the two types are homeomorphic. Ulf Persson constructed surfaces of type N with a maximally singular canonical model X, whose topology encodes information on the differentiable structure of S. A similar analysis was done by the first author for type C. In this paper we study the genus 2 fibration on X and, in particular, our main result is (X^# being the nonsingular locus of X) \pi_1(X^#)= Z_4 x Z_4 if 8 | K^2 but 16 does not | K^2 \pi_1(X^#)= Z_4 x Z_2 if 16 | K^2.
[ { "version": "v1", "created": "Thu, 4 Jul 1996 11:18:05 GMT" } ]
2008-02-03T00:00:00
[ [ "Catanese", "Fabrizio", "" ], [ "Manfredini", "Sandro", "" ] ]
alg-geom
\section{Introduction.} Among the minimal surfaces of general type, the Noether surfaces are those for which the Noether inequality $K^2 \geq 2p_g -4$ is an equality ($K^2$ is the self intersection of a canonical divisor, $p_g$ is the dimension of the space of holomorphic 2-forms).\\ These surfaces were described by Noether (\cite{No}) and more recently by Horikawa (\cite{Ho}) who proved that if $8\ |\ K^2$ then there are two distinct deformation types, namely the Noether-Horikawa surfaces of connected type (for short, N-H surfaces of type C), and those of non connected type (for short, of type N). This notation refers to the fact that, the canonical map being a double covering of a rational ruled surface, for type C the branch locus is connected, whereas for type N it is not connected.\\ In particular Horikawa proved that the intersection forms are both of the same parity (in fact, both odd) if and only if $16\ |\ K^2$.\\ From M. Freedman's theorem (\cite{Fr}) follows that if $16\ |\ K^2$ type N and type C provide two orientedly homeomorphic compact 4-manifolds.\\ Horikawa posed the question whether type N and type C provide two orientedly diffeomorphic compact 4-manifolds.\\ It looked like a natural problem to try to see whether the two differentiable structures could be distinguished by means of the invariants introduced by S. Donaldson in \cite{Do}.\\ In the case of type C we have been able (\cite{Ca}) to calculate the constant Do\-nald\-son invariants (corresponding to zero-dimensional moduli spaces) using some singular canonical models of these surfaces with very many singularities, and an approach introduced by P. Kronheimer (\cite{Kr}) for the case of the Kummer surfaces. The number we obtained, namely $ 2^{2k}$ when $K^2=8k$, is the leading term of the Donaldson series (see \cite{K-M}), which was later fully calculated by Fintushel and Stern in the case of N-H surfaces of type C via the technique of rational blow-downs (\cite{F-S}).\\ The Donaldson series for N-H surfaces of type N has not yet, to our knowledge, been calculated; although, after the Seiberg-Witten theory (\cite{W}) has been introduced, and after Pidstrigach and Tyurin (\cite{P-T}) have announced the equality between Kronheimer-Mrowka and Seiberg-Witten classes, the two series should be equal.\\ Our original aim was to extend the application of the Kronheimer theory to the case of N-H surfaces of type N using a very singular model constructed by Ulf Persson (\cite{Per}), describing its orbifold fundamental group, its representations into $SO(3)$, and then trying to see which of those have virtual dimension zero.\\ In this article we consider the singular N-H surfaces of type N with maximal Picard number constructed by Persson, henceforth called Persson-Noether-Horikawa surfaces (P-N-H for short), and we determine their orbifold fundamental group.\\ This is our main result:\\ {\bf Theorem.} {\em The orbifold fundamental group of the P-N-H surfaces is $$\zeta_4\oplus\zeta_2$$ if $16\ |\ K^2,$ $$\zeta_4\oplus\zeta_4$$\nopagebreak in the other case where $8\ |\ K^2$ but $16$ does not divide $K^2$.}\\ It follows immediately that we have, for $16\ |\ K^2$, only six nontrivial classes of orbifold $SO(3)$-representations, and a result which we do not prove here is that we do not get anyone of virtual dimension zero.\\ This is not surprising in view of (\cite{P-T}), since if Kronheimer's approach would have worked, we would have had only a finite number of constant Donaldson invariants.\\ On the other hand, the algebro-geometric technique of studying canonical mo\-dels with many rational double points produces on the smooth model configurations of (-2)-projective lines (spheres) whose tubular neighborhood has a unique holomorphic structure and, in particular, a unique compatible $C^{\infty}$ structure. In this way one produces a decomposition of the 4-manifold in geometric pieces, one of which is the nonsingular part of the singular canonical model.\\ From this point of view, the calculation of the orbifold fundamental group leads to a better understanding of the differentiable structures of the smooth model.\\ Since our proof is rather involved technically we would like to give a brief geometrical "explanation" of our result.\\ Persson's construction starts with a plane nodal cubic $C$ meeting a conic $Q$ at only one point $P$. Moreover, $C$ and $Q$ have two common tangents $L_{-1}$ and $L_1$ which meet in a point $O$ collinear with $P$ and the node of $C$.\\ Blowing up $O$ we get a $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$-bundle $f':\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1\smash{\mathop{\longrightarrow}}\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ with a section $\Sigma_{\infty}$, a bisection $Q'$ and a 3-section $C'$ ($'$ denoting the proper transform under the blow up).\\ A cyclic cover of order $2k\piu2$ branched on $L'_{-1}$ and $L'_1$ yields a new $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$-bundle $f'':\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_{2k+2}\smash{\mathop{\longrightarrow}}\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ with a section $\Sigma''_{\infty}$ disjoint from a 3-section $C''$ and two sections $Q_1''$, $Q_2''$ (the inverse image $Q''$ of $Q'$ splits into two components).\\ The curve $B=C''\cup Q_1''\cup Q_2''\cup\Sigma''_{\infty}$ has many singular points, and our canonical model $X_{2k+2}$ is the double cover of $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_{2k+2}$ branched on $B$. By construction $X_{2k+2}$ has a genus 2 fibration onto $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$, whence the orbifold fundamental group $\pi_1(X_{2k+2}^{\#})$, $X_{2k+2}^{\#}$ being the nonsingular part of $X_{2k+2}$, is a quotient of $\pi_1(F)$, where $F$ is a fixed genus 2 fibre.\\ $F$ being a double cover of $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ branched in six points $P_0\hbox{\mat \char61}\hskip1pt\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\cap\Sigma''_{\infty}$, $P_1\hbox{\mat \char61}\hskip1pt\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\cap Q''_1$, $P_2\hbox{\mat \char61}\hskip1pt\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\cap Q''_2$, $\{P_3,P_4,P_5\}=\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\cap C''$, $\pi_1(F)$ is the subgroup of a free product $\hbox{$\cal F\!$}_5(2)$ of five copies of $\zeta_2$, given by words of even length.\\ $\hbox{$\cal F\!$}_5(2)$ is generated by elements $\unoenne{\hbox{$\varepsilon$}}{6}$ such that $\cunoenne{\hbox{$\varepsilon$}}{6}\ugu1$ ($\hbox{$\varepsilon$}_i$ corresponds to a loop in $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ around the point $P_{i-1}$).\\ The first main point (we must be rather vague here, else we must give the full proof) is that, since curve $C''$ is irreducible, when the fibre $F$ moves around, $\hbox{$\varepsilon$}_4,\hbox{$\varepsilon$}_5,\hbox{$\varepsilon$}_6$ become identified.\\ Thus we only have $\unoenne{\hbox{$\varepsilon$}}{4}$ with $\cunoenne{\hbox{$\varepsilon$}}{4}\ugu1$, and therefore we have "proved" that our group is abelian , being a quotient of the fundamental group $\Gamma$ of a curve of genus 1 obtained as the double cover of $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ branched in four points. More precisely, $\Gamma$ is an abelian group with generators $\hbox{$\varepsilon$}_1\hbox{$\varepsilon$}_2$, $\hbox{$\varepsilon$}_1\hbox{$\varepsilon$}_3$.\\ We must still take into account the fact that, when the fibre $F$ moves towards a singular point (corresponding to points of intersection $C''\cap Q_1''$, $C''\cap Q_2'', Q_1''\cap Q_2''$), further relations are introduced. These relations are hard to control globally but if we look locally around these points of intersection, and accordingly take a new basis $\unoenne{\hbox{$\varepsilon$}'}{4}$, the situation becomes simpler.\\ In fact, the local equation of the double cover is $z^2\hbox{\mat \char61}\hskip1pt y^2\hbox{\rmp \char123}\hskip1pt x^{2c}$, where $c\ugu6$ or $c\hbox{\mat \char61}\hskip1pt k+1$, and $x$ is the pullback of a local coordinate on $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$, so that the corresponding local braid yields the relation $(\hbox{$\varepsilon$}'_j\hbox{$\varepsilon$}'_i)^c=(\hbox{$\varepsilon$}'_i\hbox{$\varepsilon$}'_j)^c$. In turn, using $(\hbox{$\varepsilon$}'_i)^2\ugu1$, we obtain the relation $(\hbox{$\varepsilon$}'_j\hbox{$\varepsilon$}'_i)^{2c}\ugu1$.\\ That's how one shows that the two generators of the abelian group have period 2 or 4.\\ The paper is organized as follows:\\ In section two we take up Persson's construction using explicit equations showing that the surface is defined over a real quadratic field.\\ In the third section we describe the five steps leading to a presentation of our fundamental group in terms of the braid monodromy of the plane curve $D=C\cup Q$.\\ Finally, in section four we apply combinatorial group theory arguments in order to give the main result concerning the orbifold fundamental group.\\ Acknowledgements : Both authors acknowledge support from the AGE Project H.C.M. contract ERBCHRXCT 940557 and from 40\% M.U.R.S.T..\\ The first author would like to express his gratitude to the Max-Planck Institut in Bonn where this research was initiated (in 1993), and to the Accademia dei Lincei where he is currently Professore Distaccato. \section{Persson's configuration.} In this section we will provide explicit equations for the configuration constructed by Ulf Persson in \cite{Per}.\\This is the configuration formed by a smooth conic $Q$ and a nodal cubic $C$ intersecting in only one point $P$ which is smooth for $C$. Moreover $Q$ and $C$ have two common tangents $L_{1}$ and $L_{-1}$ meeting in a point $O$ lying on the line joining $P$ and the node of $C$.\\ Let $Q\subset\ci\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^2$ be the conic $\{(x,y,z)\app\ci\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^2\,|\,x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy\raise 1pt\hbox{\mat \char43}\hskip1pt z^2\hbox{\mat \char61}\hskip1pt 0\}$.\\ Since $$x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy\raise 1pt\hbox{\mat \char43}\hskip1pt z^2\hbox{\mat \char61}\hskip1pt(x\raise 1pt\hbox{\mat \char43}\hskip1pt z)^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2z(y\hbox{\rmp \char123}\hskip1pt x)\hbox{\mat \char61}\hskip1pt (x\hbox{\rmp \char123}\hskip1pt z)^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2z(y\raise 1pt\hbox{\mat \char43}\hskip1pt x)$$ $Q$ is tangent to the lines $L_1=\{x\hbox{\rmp \char123}\hskip1pt y\hbox{\mat \char61}\hskip1pt 0\}$ and $L_{-1}=\{x\raise 1pt\hbox{\mat \char43}\hskip1pt y\hbox{\mat \char61}\hskip1pt 0\}$.\\ The tangency points are: $$ x\hbox{\rmp \char123}\hskip1pt y\hbox{\mat \char61}\hskip1pt x\raise 1pt\hbox{\mat \char43}\hskip1pt z\hbox{\mat \char61}\hskip1pt 0 \hbox{$\,\Rightarrow\,$} (1,1,\men1)$$ $$x\raise 1pt\hbox{\mat \char43}\hskip1pt y\hbox{\mat \char61}\hskip1pt x\hbox{\rmp \char123}\hskip1pt z\hbox{\mat \char61}\hskip1pt 0\hbox{$\,\Rightarrow\,$} (1,\men1,1).$$ Note that $Q$ is also tangent to the line $z\hbox{\mat \char61}\hskip1pt 0$ at the point $(0,1,0)=P$.\\ We want to find an irreducible nodal cubic $C$ such that $C\cdot Q=6P$ and such that $C$ is tangent to the lines $x\hbox{\mat \char61}\hskip1pt\pm y$ in points different from those of $Q$.\\ Let $C$ be a cubic s.t. $P\app C$ and $C\cdot Q=6P$. Note that if $C$ were reducible, then the previous condition would imply that $z\ugu0$ is a component of $C$.\\ We then have ${\rm div}(C)={\rm div}(z^3)\ ({\rm mod}Q)$, so $C=z^3\raise 1pt\hbox{\mat \char43}\hskip1pt QL$ with $L$ a linear form, and thus $$C=z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)(ax\raise 1pt\hbox{\mat \char43}\hskip1pt by\raise 1pt\hbox{\mat \char43}\hskip1pt cz).$$ Since we want $C$ to be tangent to the two lines $L_{1}$ and $L_{-1}$ we obtain that the following homogeneous polynomials in $(x,z)$ \begin{equaz}z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (x\raise 1pt\hbox{\mat \char43}\hskip1pt z)^2((a\raise 1pt\hbox{\mat \char43}\hskip1pt b)(x\raise 1pt\hbox{\mat \char43}\hskip1pt z)\raise 1pt\hbox{\mat \char43}\hskip1pt z(c\hbox{\rmp \char123}\hskip1pt a\hbox{\rmp \char123}\hskip1pt b))\label{1}\end{equaz} \begin{equaz}z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (x\hbox{\rmp \char123}\hskip1pt z)^2((a\hbox{\rmp \char123}\hskip1pt b)(x\hbox{\rmp \char123}\hskip1pt z)\raise 1pt\hbox{\mat \char43}\hskip1pt z(c\raise 1pt\hbox{\mat \char43}\hskip1pt a\hbox{\rmp \char123}\hskip1pt b))\label{21}\end{equaz} must have a double root.\\ Set $\ze\hbox{\mat \char61}\hskip1pt (\frac{z}{x+z})$ and $\hat\ze\hbox{\mat \char61}\hskip1pt (\frac{z}{x-z})$ and rewrite \ref{1}, \ref{21} as: $$\ze^3\raise 1pt\hbox{\mat \char43}\hskip1pt \ze (c\hbox{\rmp \char123}\hskip1pt a\hbox{\rmp \char123}\hskip1pt b)\raise 1pt\hbox{\mat \char43}\hskip1pt (a\raise 1pt\hbox{\mat \char43}\hskip1pt b)=0\ \ \ \ \hat\ze^3\raise 1pt\hbox{\mat \char43}\hskip1pt\hat\ze (c\raise 1pt\hbox{\mat \char43}\hskip1pt a\hbox{\rmp \char123}\hskip1pt b)\raise 1pt\hbox{\mat \char43}\hskip1pt (a\hbox{\rmp \char123}\hskip1pt b)=0.$$ We recall that if $\ze$ is a double root of $z^3\raise 1pt\hbox{\mat \char43}\hskip1pt pz\raise 1pt\hbox{\mat \char43}\hskip1pt q\hbox{\mat \char61}\hskip1pt 0$ then $$3\ze^2+p=0{ \rm\ \ whence\ \ } \frac23\ze p\raise 1pt\hbox{\mat \char43}\hskip1pt q\hbox{\mat \char61}\hskip1pt 0$$ and this implies that $$\ze\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt \frac32\frac pq{\rm\ \ thus\ \ }27q^2\raise 1pt\hbox{\mat \char43}\hskip1pt 4p^3\hbox{\mat \char61}\hskip1pt 0.$$ Therefore we have a double root of \ref{1} if and only if $$\exists\, A\ :\ \ze\hbox{\mat \char61}\hskip1pt A,\ q\hbox{\mat \char61}\hskip1pt 2A^3,\ p\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt 3A^2,\ {\rm i.e.}\ \left\{\begin{array}{l} a\raise 1pt\hbox{\mat \char43}\hskip1pt b\hbox{\mat \char61}\hskip1pt 2A^3 \\ c\hbox{\rmp \char123}\hskip1pt (a\raise 1pt\hbox{\mat \char43}\hskip1pt b)\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt 3A^2.\end{array}\right.$$ Similarly if we set $\hat\ze\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt B$ we have $$\left\{\begin{array}{l} a\hbox{\rmp \char123}\hskip1pt b\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt 2B^3\\ c\hbox{\rmp \char123}\hskip1pt (b\hbox{\rmp \char123}\hskip1pt a)\hbox{\mat \char61}\hskip1pt\men3B^2\end{array}\right.$$ and so $$\left\{\begin{array}{l}a\hbox{\mat \char61}\hskip1pt A^3\hbox{\rmp \char123}\hskip1pt B^3 \\ b\hbox{\mat \char61}\hskip1pt A^3\raise 1pt\hbox{\mat \char43}\hskip1pt B^3 \\ c\hbox{\mat \char61}\hskip1pt 2A^3\hbox{\rmp \char123}\hskip1pt 3A^2\hbox{\mat \char61}\hskip1pt 2B^3\hbox{\rmp \char123}\hskip1pt 3B^2. \end{array}\right.$$ Then $A$ and $B$ must satisfy $2(A^3\hbox{\rmp \char123}\hskip1pt B^3)\hbox{\mat \char61}\hskip1pt 3(A^2\hbox{\rmp \char123}\hskip1pt B^2)$.\\ Recall that (we make no distinction between a curve and its equation) $$C=z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)((A^3\hbox{\rmp \char123}\hskip1pt B^3)x\raise 1pt\hbox{\mat \char43}\hskip1pt (A^3\raise 1pt\hbox{\mat \char43}\hskip1pt B^3)y\raise 1pt\hbox{\mat \char43}\hskip1pt (2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)z)$$ while $x\hbox{\rmp \char123}\hskip1pt y\hbox{\mat \char61}\hskip1pt 0$ is tangent to $C$ at the point where $$\ze\hbox{\mat \char61}\hskip1pt\frac{z}{x+z}\hbox{\mat \char61}\hskip1pt A.$$ Therefore the tangency point is $(1\hbox{\rmp \char123}\hskip1pt A,1\hbox{\rmp \char123}\hskip1pt A,A)$.\\ Similarly $x\raise 1pt\hbox{\mat \char43}\hskip1pt y\hbox{\mat \char61}\hskip1pt 0$ is tangent to $C$ at the point $(B\hbox{\rmp \char123}\hskip1pt 1,1\hbox{\rmp \char123}\hskip1pt B,B)$.\\ Let us now search for a cubic $C$ with a singular point on the line $x\ugu0$, as in Persson's construction.\\ Since $\frac{\partial C}{\partial x}$ on the line $x\ugu0$ equals $aQ$ and the singular point is different from $P$ it follows that $a\ugu0$. Whence $A^3\hbox{\rmp \char123}\hskip1pt B^3\hbox{\mat \char61}\hskip1pt A^2\hbox{\rmp \char123}\hskip1pt B^2\hbox{\mat \char61}\hskip1pt 0$ and so $A\hbox{\mat \char61}\hskip1pt B$.\\ If $A\hbox{\mat \char61}\hskip1pt B$ then $C$ contains only the monomial $x^2$ as a polynomial in $x$, so the involution $x{\longmapsto} \hbox{\rmp \char123}\hskip1pt x$ leaves the curve $C$ invariant. From this we deduce that a singular point of $C$ must have its $x$ coordinate equal to $0$ and $C$ has then a singularity on the line $x\hbox{\mat \char61}\hskip1pt 0$ if and only if $$z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (z^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy)(2A^3y\raise 1pt\hbox{\mat \char43}\hskip1pt (2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)z)\ \ {\rm has\ a\ double\ root.}$$ Remembering that it can't be $A\hbox{\mat \char61}\hskip1pt B\hbox{\mat \char61}\hskip1pt 0$, the double root cannot be $z\hbox{\mat \char61}\hskip1pt 0$ and we can write the above as $$z(z^2\raise 1pt\hbox{\mat \char43}\hskip1pt (z\raise 1pt\hbox{\mat \char43}\hskip1pt 2y)(2A^3y\raise 1pt\hbox{\mat \char43}\hskip1pt (2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)z)).$$ So we must check that $$z^2(1\raise 1pt\hbox{\mat \char43}\hskip1pt 2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy(A^3\raise 1pt\hbox{\mat \char43}\hskip1pt 2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)\raise 1pt\hbox{\mat \char43}\hskip1pt 4A^3y^2=$$ $$=z^2(1\raise 1pt\hbox{\mat \char43}\hskip1pt 2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)\raise 1pt\hbox{\mat \char43}\hskip1pt 2zy3A^2(A\hbox{\rmp \char123}\hskip1pt 1)\raise 1pt\hbox{\mat \char43}\hskip1pt 4A^3y^2$$ has a double root.\\ This is the case when $$9A^4(A\hbox{\rmp \char123}\hskip1pt 1)^2\hbox{\mat \char61}\hskip1pt 4A^3(1\raise 1pt\hbox{\mat \char43}\hskip1pt 2A^3\hbox{\rmp \char123}\hskip1pt 3A^2)\ \ {\rm\ i.e.}$$ $$9A^6\hbox{\rmp \char123}\hskip1pt 18A^5\raise 1pt\hbox{\mat \char43}\hskip1pt 9A^4\hbox{\mat \char61}\hskip1pt 4A^3\raise 1pt\hbox{\mat \char43}\hskip1pt 8A^6\hbox{\rmp \char123}\hskip1pt 12A^5.$$ Upon dividing by $A^3\ugu\hskip -7pt / \kern 2pt 0$ we get $$A^3\hbox{\rmp \char123}\hskip1pt 6A^2\raise 1pt\hbox{\mat \char43}\hskip1pt 9A\hbox{\rmp \char123}\hskip1pt 4\hbox{\mat \char61}\hskip1pt 0.$$ Observe that $1$ is a root of this equation, but if $A\hbox{\mat \char61}\hskip1pt 1$ then the singular point is $(0,0,1)$ and coincides with the point of tangency of $x\raise 1pt\hbox{\mat \char43}\hskip1pt y\hbox{\mat \char61}\hskip1pt 0$ so this root has to be discarded. Since $$A^3\hbox{\rmp \char123}\hskip1pt 6A^2\raise 1pt\hbox{\mat \char43}\hskip1pt 9A\hbox{\rmp \char123}\hskip1pt 4\hbox{\mat \char61}\hskip1pt (A\hbox{\rmp \char123}\hskip1pt 1)(A^2\hbox{\rmp \char123}\hskip1pt 5A\raise 1pt\hbox{\mat \char43}\hskip1pt 4)\hbox{\mat \char61}\hskip1pt (A\hbox{\rmp \char123}\hskip1pt 1)^2(A\hbox{\rmp \char123}\hskip1pt 4)$$ the other possible root is then $A\hbox{\mat \char61}\hskip1pt 4$, and in this case we have $B\hbox{\mat \char61}\hskip1pt A\hbox{\mat \char61}\hskip1pt 4$, $a\hbox{\mat \char61}\hskip1pt 0$, $b\hbox{\mat \char61}\hskip1pt 8\cdot4^2$, $c\hbox{\mat \char61}\hskip1pt 5\cdot4^2$.\\ Then $$C=z^3\raise 1pt\hbox{\mat \char43}\hskip1pt 4^2(x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2yz\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)(8y\raise 1pt\hbox{\mat \char43}\hskip1pt 5z)$$ The tangency points are $(\hbox{\rmp \char123}\hskip1pt 3,\hbox{\rmp \char123}\hskip1pt 3,4)$ and $(3,\hbox{\rmp \char123}\hskip1pt 3,4)$, while for the singular point we have $x\hbox{\mat \char61}\hskip1pt 0$ and a double root of $$z^2\raise 1pt\hbox{\mat \char43}\hskip1pt 4^2(2y\raise 1pt\hbox{\mat \char43}\hskip1pt z)(8y\raise 1pt\hbox{\mat \char43}\hskip1pt 5z)\hbox{\mat \char61}\hskip1pt 0\iff 81z^2\raise 1pt\hbox{\mat \char43}\hskip1pt 4^218zy\raise 1pt\hbox{\mat \char43}\hskip1pt 4^4y\hbox{\mat \char61}\hskip1pt 0\iff 9z\raise 1pt\hbox{\mat \char43}\hskip1pt 4^2y\hbox{\mat \char61}\hskip1pt 0$$ so the singular point is $(0,9,\hbox{\rmp \char123}\hskip1pt 16)$.\\ With this choice of $A$ and $B$, $C$ is irreducible (since $z\hbox{\mat \char61}\hskip1pt 0$ is not a component of $C$).\\ We want to find the lines through $(0,0,1)$ and tangent to $C$.\\ Let $A\hbox{\mat \char61}\hskip1pt B\hbox{\mat \char61}\hskip1pt\hbox{$\lambda$}$ and consider more generally the 1-parameter family of curves: $$C_{\hbox{$\lambda$}}=z^3\raise 1pt\hbox{\mat \char43}\hskip1pt (x^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2yz\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)(2\hbox{$\lambda$}^3y\raise 1pt\hbox{\mat \char43}\hskip1pt (2\hbox{$\lambda$}^3\hbox{\rmp \char123}\hskip1pt 3\hbox{$\lambda$}^2)z)\hbox{\mat \char61}\hskip1pt 0.$$ The tangency points on the two fixed lines $x\raise 1pt\hbox{\mat \char43}\hskip1pt y=0$, $x\hbox{\rmp \char123}\hskip1pt y=0$ are, as we know, $(1\hbox{\rmp \char123}\hskip1pt\hbox{$\lambda$},1\hbox{\rmp \char123}\hskip1pt\hbox{$\lambda$},\hbox{$\lambda$})$ and $(\hbox{$\lambda$}\hbox{\rmp \char123}\hskip1pt 1,1\hbox{\rmp \char123}\hskip1pt\hbox{$\lambda$},\hbox{$\lambda$})$.\\ Rewriting the last equation in powers of $z$ we obtain: $$z^3(1\raise 1pt\hbox{\mat \char43}\hskip1pt 2\hbox{$\lambda$}^3\hbox{\rmp \char123}\hskip1pt 3\hbox{$\lambda$}^2)\raise 1pt\hbox{\mat \char43}\hskip1pt z^26y\hbox{$\lambda$}^2(\hbox{$\lambda$}\hbox{\rmp \char123}\hskip1pt 1)\raise 1pt\hbox{\mat \char43}\hskip1pt z\hbox{$\lambda$}^2(4\hbox{$\lambda$} y^2\raise 1pt\hbox{\mat \char43}\hskip1pt (2\hbox{$\lambda$}\hbox{\rmp \char123}\hskip1pt 3)x^2)\raise 1pt\hbox{\mat \char43}\hskip1pt 2\hbox{$\lambda$}^3x^2y\hbox{\mat \char61}\hskip1pt 0.$$ Since we know what happens for $\hbox{$\lambda$}\hbox{\mat \char61}\hskip1pt 0$, we can divide by $\hbox{$\lambda$}^3$, set $w\hbox{\mat \char61}\hskip1pt \frac z{\lambda}$ and obtain: $$w^3(1\raise 1pt\hbox{\mat \char43}\hskip1pt 2\hbox{$\lambda$}^3\hbox{\rmp \char123}\hskip1pt 3\hbox{$\lambda$}^2)\raise 1pt\hbox{\mat \char43}\hskip1pt w^26y\hbox{$\lambda$}(\hbox{$\lambda$}\hbox{\rmp \char123}\hskip1pt 1)\raise 1pt\hbox{\mat \char43}\hskip1pt w(4\hbox{$\lambda$} y^2\raise 1pt\hbox{\mat \char43}\hskip1pt (2\hbox{$\lambda$}\hbox{\rmp \char123}\hskip1pt 3)x^2)\raise 1pt\hbox{\mat \char43}\hskip1pt 2x^2y\hbox{\mat \char61}\hskip1pt 0.$$ We let now $\hbox{$\Delta$}$ be the discriminant of $C_{\hbox{$\lambda$}}$ with respect to the variable $w$, and using a standard formula for $\hbox{$\Delta$}$, we find a degree 6 equation in $x$ and $y$ which is divisible by $x^2(x^2\hbox{\rmp \char123}\hskip1pt y^2)$.\\ Remembering that the discriminant of $a_0x^3\raise 1pt\hbox{\mat \char43}\hskip1pt a_1x^2\raise 1pt\hbox{\mat \char43}\hskip1pt a_2x\raise 1pt\hbox{\mat \char43}\hskip1pt a_3$ is: $$\hbox{$\Delta$}\hbox{\mat \char61}\hskip1pt a_1^2a_2^2\hbox{\rmp \char123}\hskip1pt 4a_0a_2^3\hbox{\rmp \char123}\hskip1pt 4a_1^3a_3\hbox{\rmp \char123}\hskip1pt 27a_0^2a_3^2\raise 1pt\hbox{\mat \char43}\hskip1pt 18a_0a_1a_2a_3$$ and applying this formula for simplicity when $\hbox{$\lambda$}\hbox{\mat \char61}\hskip1pt 4$, we obtain: $$y^22^63^4(16y^2\raise 1pt\hbox{\mat \char43}\hskip1pt 5x^2)^2\hbox{\rmp \char123}\hskip1pt 2^23^4(16y^2\raise 1pt\hbox{\mat \char43}\hskip1pt 5x^2)^3\hbox{\rmp \char123}\hskip1pt$$ $$\men2^{12}3^6x^2y^4 \hbox{\rmp \char123}\hskip1pt 2^23^{11}x^4y^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2^53^8(16y^2\raise 1pt\hbox{\mat \char43}\hskip1pt 5x^2)x^2y^2$$ and factoring this binary form we get: $$x^2(x^2\hbox{\rmp \char123}\hskip1pt y^2)2^23^4(2^7y^2\hbox{\rmp \char123}\hskip1pt 5^3x^2).$$ So we have that the tangent lines to $C$ passing through $(0,0,1)$ are $x\hbox{\mat \char61}\hskip1pt\pm y$, $x\hbox{\mat \char61}\hskip1pt\pm\sqrt{\frac{128}{125}}y$ while $x\hbox{\mat \char61}\hskip1pt 0$ passes through the node of $C$. We denote by $L_0$ the line $x\ugu0$ and by $L_+,L_-$ the two lines $x\hbox{\mat \char61}\hskip1pt\sqrt{\frac{128}{125}}y$, $x\hbox{\mat \char61}\hskip1pt-\sqrt{\frac{128}{125}}y$ respectively.\\ In order to find the tangency point on the lines $L_+,L_-$ we by symmetry may restrict to the line $L_+$.\\ Writing $x\hbox{\mat \char61}\hskip1pt 2^3\sqrt2\,a$, $y\hbox{\mat \char61}\hskip1pt 5\sqrt5\,a$ we have that \begin{equaz}\label{pol}z^3\raise 1pt\hbox{\mat \char43}\hskip1pt 2^4(2^7a^2\raise 1pt\hbox{\mat \char43}\hskip1pt 10\sqrt5\,az\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)(40\sqrt5\,a\raise 1pt\hbox{\mat \char43}\hskip1pt 5z)\hbox{\mat \char61}\hskip1pt 0\end{equaz} has a double root. Since for its derivative we have $$3z^2\raise 1pt\hbox{\mat \char43}\hskip1pt 2^4(10\sqrt5\,a\raise 1pt\hbox{\mat \char43}\hskip1pt 2z)(40\sqrt5\,a\raise 1pt\hbox{\mat \char43}\hskip1pt 5z)\raise 1pt\hbox{\mat \char43}\hskip1pt 2^45(2^7a^2\raise 1pt\hbox{\mat \char43}\hskip1pt 10\sqrt5\,az\raise 1pt\hbox{\mat \char43}\hskip1pt z^2)\hbox{\mat \char61}\hskip1pt 0$$ $$(15\raise 1pt\hbox{\mat \char43}\hskip1pt {3\over16})z^2\raise 1pt\hbox{\mat \char43}\hskip1pt 180\sqrt5\,az\raise 1pt\hbox{\mat \char43}\hskip1pt 2640a^2\ugu0$$ $${a\over z}\hbox{\mat \char61}\hskip1pt{-90\sqrt5\pm\sqrt{90^25- 2640(15+{3\over16})}\over2640}\hbox{\mat \char61}\hskip1pt\sqrt5{-30\pm3\over880}.$$ Thus $\frac{y}{z}={-25(30\pm3)\over880}$, $\frac{x}{z}={-8\sqrt{10}(30\pm3)\over880}$ and the point of tangency is one of the points $(\men33\cdot8\sqrt{10},\hbox{\rmp \char123}\hskip1pt 25\cdot 33,880)$, $(\men27\cdot8\sqrt{10},\hbox{\rmp \char123}\hskip1pt 25\cdot 27,880)$.\\ Upon substituting these values in the polynomial \ref{pol} we find that the correct choice is $(\men24\sqrt{10},\hbox{\rmp \char123}\hskip1pt 75,80)$.\\ By symmetry the point $(24\sqrt{10},\hbox{\rmp \char123}\hskip1pt 75,80)$ is the tangency point of the line $L_-$.\\ Let us write $$C=4^2(8y\raise 1pt\hbox{\mat \char43}\hskip1pt 5z)x^2\raise 1pt\hbox{\mat \char43}\hskip1pt z(16y\raise 1pt\hbox{\mat \char43}\hskip1pt 9z)^2\hbox{\mat \char61}\hskip1pt 0$$ and let us set $u\hbox{\mat \char61}\hskip1pt 16y\raise 1pt\hbox{\mat \char43}\hskip1pt 9z$. We have: $$C= zu^2\raise 1pt\hbox{\mat \char43}\hskip1pt 8x^2(u\raise 1pt\hbox{\mat \char43}\hskip1pt z)\hbox{\mat \char61}\hskip1pt 0$$ In these coordinates the singular point of $C$ is $(0,0,1)$, so the tangents at the singular point are given by: $$8x^2\raise 1pt\hbox{\mat \char43}\hskip1pt u^2\hbox{\mat \char61}\hskip1pt 0$$ whence they are complex and we have an isolated point.\\ In order to draw $C$, let's compute its flexes. Using the coordinates $x$, $u$, and $z$ the Hessian matrix is: $$\left(\begin{array}{ccc}16(u\raise 1pt\hbox{\mat \char43}\hskip1pt z)&16x&16x\\16x&2z&2u\\16x&2u&0\end{array} \right)$$ The Hessian curve is then given by the determinant of $$\left(\begin{array}{ccc}(u\raise 1pt\hbox{\mat \char43}\hskip1pt z)&0&x\\0&z\hbox{\rmp \char123}\hskip1pt 2u&u\\8x&u&0\end{array} \right)$$ which equals $$\hbox{\rmp \char123}\hskip1pt (u\raise 1pt\hbox{\mat \char43}\hskip1pt z)u^2\hbox{\rmp \char123}\hskip1pt 8x^2(z\hbox{\rmp \char123}\hskip1pt 2u)\ugu0.$$\\ Eliminating $8x^2$ from the two equations we get $$(u\raise 1pt\hbox{\mat \char43}\hskip1pt z)^2u^2\hbox{\rmp \char123}\hskip1pt zu^2(z\hbox{\rmp \char123}\hskip1pt 2u)\ugu0$$ so either $u=0$, and this implies either $x\hbox{\mat \char61}\hskip1pt 0$ (the singular point) or $z\hbox{\mat \char61}\hskip1pt 0$ that gives the point $(1,0,0)$, or $$(u\raise 1pt\hbox{\mat \char43}\hskip1pt z)^2\hbox{\rmp \char123}\hskip1pt z(z\hbox{\rmp \char123}\hskip1pt 2u)\hbox{\mat \char61}\hskip1pt u^2\raise 1pt\hbox{\mat \char43}\hskip1pt 4uz\hbox{\mat \char61}\hskip1pt 0$$ that gives ($u\diverso0$) $u\hbox{\mat \char61}\hskip1pt \hbox{\rmp \char123}\hskip1pt 4z$, that is $z\hbox{\mat \char61}\hskip1pt\hbox{\rmp \char123}\hskip1pt 1$, $u\hbox{\mat \char61}\hskip1pt 4$, $y\hbox{\mat \char61}\hskip1pt{13\over16}$, $x\hbox{\mat \char61}\hskip1pt\pm\sqrt{2/3}$.\\ For these points ${x\over y}=\pm\sqrt{2/3}{16\over13}$.\pagebreak \section{Fundamental groups.} In this section we are going to describe the five steps leading to the determination of the orbifold fundamental group of the Persson's surfaces.\\ {\bf Step 1.}\\ Let $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1$ be the blow up of $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^2$ at the point $(0,0,1)$ and let $\Sigma_{\infty}$ be the exceptional divisor.\\ We consider the fibre bundle $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1\mapright{f'}\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ and its restriction $f$ $$\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1\meno( C\cup Q\cup\Sigma_{\infty}\cup L_1\cup L_{-1} \cup L_+\cup L_-\cup L_0)=\tilde{\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}}_1$$ $$f\big\downarrow$$ $$\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\meno\{P_1,P_{-1},P_+,P_-,P_0\}=\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\meno\{5\ {\rm pts.}\}.$$ $f$ is again a fibre bundle and we have a corresponding homotopy exact sequence of fundamental groups $$1\smash{\mathop{\longrightarrow}} \hbox{$\cal F\!$}_5\smash{\mathop{\longrightarrow}}\tilde{\PI}\smash{\mathop{\longrightarrow}}\hbox{$\cal F\!$}_4\mapdestra1$$ \puntif where $\hbox{$\cal F\!$}_k$ denotes the free group with $k$ generators and $\tilde\PI=\pi_1(\tilde{\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}}_1)$.\\ Here we choose a small positive real number $\hbox{$\varepsilon$}\mag0$ and $x\hbox{\mat \char61}\hskip1pt\hbox{$\varepsilon$}$, $y\ugu1$ as base point on $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\meno\{5\ {\rm pts.}\}$ and $x\hbox{\mat \char61}\hskip1pt\hbox{$\varepsilon$}$, $y\ugu1$, $z\hbox{\mat \char61}\hskip1pt\men4\sqrt{-1}$ as base point on $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1$.\\ We let $\unoenne{\delta}{5}$ be a natural geometric basis of the free group $$\hbox{$\cal F\!$}_5=\pi_1(f^{-1}({\rm base\ pt.}))=\pi_1(L_{\hbox{$\varepsilon$}}\meno(C\cup Q\cup\Sigma_{\infty}))$$ where the five points $L_{\hbox{$\varepsilon$}}\cap C$, $L_{\hbox{$\varepsilon$}}\cap Q$ are ordered by lexicografic order on $\rm Re(\frac zy)$, $\rm Im(\frac zy)$.\\ $\hbox{$\cal F\!$}_4$ is generated by the five geometric paths $\gamma_i'$ around the five critical values described in figure \ref{figura} and whose product is the identity.\\ For these elements we choose lifts to $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_1$ using a $C^{\infty}$ section of a tubular neighborhood of $\Sigma_{\infty}$ meeting $\Sigma_{\infty}$ just in the point $\infty$ $(y\ugu0)$ with intersection number equal to $-1$.\\ Therefore such lifts give paths $\gamma_i$ such that $$\prod\gamma_i=\prod\delta_i$$ and more specifically $$\gamma_+\gamma_1\gamma_0\gamma_-\gamma_{-1}=\delta_1\cdots\delta_5= \gamma_{-1}\gamma_+\gamma_1\gamma_0\gamma_-.$$ We have that, indeed, $\tilde{\PI}$ occurs as a semidirect product described by the relations $$\gamma_j^{-1}\delta_i\gamma_j=(\delta_i)\beta_j$$ where the $\beta_j$'s are suitable braids in $$\hbox{\tengt B}_5 =\hbox{\matem \char60}\sigma_1,\ldots,\sigma_{4}|\:\:\:\, \sigma_i\sigma_j \hbox{\mat \char61}\hskip1pt\sigma_j\sigma_i\ \;\forall\: 1\hbox{\mpic \char20}\hskip1pt i\hbox{\mate \char60}\hskip1pt j\piu1\hbox{\mpic \char20}\hskip1pt 5 $$ \hspace*{15.4em}\immediate\vspace*{-1ex} $\sigma_i\sigma_{i+1}\sigma_i\hbox{\mat \char61}\hskip1pt \sigma_{i+1}\sigma_i\sigma_{i+1}\ \forall\: 1\hbox{\mpic \char20}\hskip1pt i\hbox{\mate \char60}\hskip1pt 4\ \hbox{\matem \char62}$ \vspace{\baselineskip}\\ the braid group on 5 strings which acts on the right on the free group $\hbox{$\cal F\!$}_5$ by the formulae \begin{eqnarray*}(\delta_h)\sigma_k&=&\delta_h\ \ \ \ {\rm if}\ h\ugu\hskip -7pt / \kern 2pt k,k\piu1\\ (\delta_k)\sigma_k&=&\delta_{k+1}\\ (\delta_{k+1})\sigma_k&=&\delta_{k+1}^{-1}\delta_k\delta_{k+1}. \end{eqnarray*} The braids $\beta_j$ are constructed by following the motion of the five points of the intersection of $f'^{-1}(P)$ with $C\cup Q$ while $P$ goes along $\gamma_j'$.\\ With our choice of the $\gamma'$'s we have, as the reader can easily verify, \begin{eqnarray*} \beta_0&=&\sigma_4^{12}\sigma_2^2\\ \beta_1&=&\sigma_1^{-1}\sigma_2\sigma_3\sigma_1\sigma_2^{-1}\sigma_1\\ \beta_{-1}&=&\sigma_4^{-6}\sigma_2^{-1}\beta_1\sigma_2\sigma_4^6\\ \beta_+&=&\sigma_1^{-2}\sigma_2\sigma_3\sigma_4\sigma_3^{-1}\sigma_2^{-1} \sigma_1^2\\ \beta_-&=&\sigma_4^{-6}\sigma_2^{-1}\beta_+\sigma_2\sigma_4^6. \end{eqnarray*} {\bf Step 2.}\\ By taking $\sqrt{\frac{x-y}{x+y}}$ we have a new fibre bundle $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_2\mapright{g'}\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1$ obtained by base change. Under this base change the inverse image $Q'$ of the conic $Q$ splits into two sections of $g'$ which we will denote by $Q_1'$ and $Q_2'$. Again, by restriction we have a fibre bundle $g$ $$\hat{\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}}_2=\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_2\meno (C'\cup Q_1'\cup Q_2'\cup \Sigma_{\infty}' \cup \{8\ {\rm fibres}\})\mapright{g}\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\meno \{8\ {\rm pts.}\}.$$ Correspondingly we get an exact sequence $$1\smash{\mathop{\longrightarrow}}\hbox{$\cal F\!$}_5=\hbox{\matem \char60}\unoenne{\delta}{5}\hbox{\matem \char62}\smash{\mathop{\longrightarrow}}\hat{\PI} \smash{\mathop{\longrightarrow}}\hbox{$\cal F\!$}_7= \hbox{\matem \char60}\gamma_0,\gamma_-,\gamma_+,\bar{\gamma}_0,\bar{\gamma}_-, \bar{\gamma}_+,\gamma_1^2\hbox{\matem \char62}\mapdestra1$$ where $\bar\gamma_i\hbox{\mat \char61}\hskip1pt\gamma_i^{\gamma_1}\hbox{\mat \char61}\hskip1pt\gamma_1\gamma_i\gamma_1^{-1}$ and $\hat\PI=\pi_1(\hat{\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}}_2)$.\\ The fact that $\hbox{$\cal F\!$}_7$ has seven generators as above follows since the double cover of $\hbox{{\uyu I}\hskip-.05truecm{\uyu P}}^1\meno\{5\ {\rm pts.}\}$ corresponds to the homomorphism $\hbox{$\cal F\!$}_4\rightarrow\zeta_{\!2}$ sending $\gamma_1',\gamma_{-1}'\mapsto\bar1$, and $\gamma_0',\gamma_+', \gamma_-'\mapsto\bar0$.\\ If we want to keep track of the eight critical values, we can also use $(\gamma_{-1}^2)^{\gamma_1}$ as a generator. In fact $$(\delta_1\cdots\delta_5)^2=(\gamma_+\gamma_1\gamma_0\gamma_-\gamma_{-1}) (\gamma_{-1}\gamma_+\gamma_1\gamma_0\gamma_-)$$ thus $$\gamma_+\gamma_0^{\gamma_1}\gamma_-^{\gamma_1}(\gamma_{-1}^2)^{\gamma_1} \gamma_+^{\gamma_1}\gamma_1^2\gamma_0\gamma_-=(\delta_1\cdots\delta_5)^2.$$ The geometric meaning of the above formula is related to the fact that $(\Sigma_{\infty}')^2=-2,$ and more precisely to the fact that the new generators of $\hbox{$\cal F\!$}_7$ lie in a $C^{\infty}$ section meeting $\Sigma_{\infty}'$ in one point with intersection number $(-2)$, and not meeting the other curves.\\ A presentation of $\hat{\PI}$ is thus given by $$\hbox{\matem \char60}\unoenne{\delta}{5},\gamma_0,\gamma_-,\gamma_+, \bar{\gamma}_0,\bar{\gamma}_-,\bar{\gamma}_+,\Gamma\hbox{\mat \char61}\hskip1pt\gamma_1^{-2}\ |\ \gamma_0^{-1}\delta_i\gamma_0=(\delta_i)\beta_0\ \ \ \ \ \ \ \ \ \ $$ $$\hspace*{16.5em}\vdots$$ $$\hspace*{18.5em}\bar{\gamma}_0^{-1}\delta_i\bar{\gamma}_0=(\delta_i) \beta_1\beta_0\beta_1^{-1}$$ $$\hspace*{16.5em}\vdots$$ $$\hspace*{17em}\Gamma\delta_i\Gamma^{-1}=(\delta_i)\beta_1^2\hbox{\matem \char62}$$ {\bf Step 3.}\\ The fundamental group $$\PI'=\pi_1(\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_2\meno (C'\cup Q_1'\cup Q_2'\cup \Sigma_{\infty}' \cup L_1'\cup L_{-1}'))$$ is a quotient of $\hat\PI$. The presentation of $\PI'$ is readily accomplished simply by introducing in the above presentation the further relations $$\gamma_0=\gamma_-=\gamma_+=\bar{\gamma}_0=\bar{\gamma}_-= \bar{\gamma}_+=1.$$ Then $\PI'$ is presented as $$\hbox{\matem \char60}\unoenne{\delta}{5},\Gamma\ |\ \delta_i=(\delta_i)\beta_0\ \ \ \delta_i=(\delta_i)\beta_-\ \ \ \delta_i=(\delta_i)\beta_+$$ $$\hspace*{7em}\delta_i=(\delta_i)\beta_1\beta_0\beta_1^{-1}\ \ \ \delta_i=(\delta_i)\beta_1\beta_-\beta_1^{-1}$$ $$\hspace*{8em}\delta_i=(\delta_i)\beta_1\beta_+\beta_1^{-1}\ \ \ \Gamma\delta_i\Gamma^{-1}=(\delta_i)\beta_1^2\hbox{\matem \char62}$$ {\bf Remark:} with the new relations we get, setting $\Gamma_{-1}=(\gamma_{-1}^2)^{\gamma_1},$ $$\Gamma_{-1}\Gamma=(\delta_1\cdots\delta_5)^2$$ {\bf Step 4'}.\\ We denote by $X_2^{\#}$ the non singular part of the double cover $X_2$ of $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_2$ (branched over $C'\cup Q_1'\cup Q_2' \cup\Sigma_{\infty}'$) and by $Z_2^{\#}$ the complement in $X_2^{\#}$ of $L_1'',L_{-1}''$, the respective inverse images of $L_1',L_{-1}'$.\\ We finally let $Y_2^{\#}$ be the double cover of $\hbox{{\uyu I}\hskip-.05truecm{\uyu F}$\!$}_2\meno (C'\cup Q_1'\cup Q_2'\cup \Sigma_{\infty}' \cup L_1'\cup L_{-1}')$.\\ Thus $Y_2^{\#}\subset Z_2^{\#}\subset X_2^{\#}$.\\ Clearly $\pi_1(Y_2^{\#})=\ker(\PI'\smash{\mathop{\longrightarrow}}\zeta_2)$, where $\delta_i{\longmapsto}\bar1$ and $\Gamma{\longmapsto}\bar0$, is generated by $\Gamma$, $\sigma=\delta_1\Gamma\delta_1^{-1}$, $A_i=\delta_1\delta_i$ $(i\ugu1,\ldots , 5)$ and $B_j=\delta_j\delta_1^{-1}$ $(j\ugu2,\ldots , 5)$.\\ To find the relations we apply the Reidemeister-Shreier rewriting process to the relations $R_{\alpha}$ of $\PI'$ and to the relations $\delta_1R_{\alpha}\delta_1^{-1}$.\\ {\bf Step 4''}.\\ Clearly, $\pi_1(Y_2^{\#})$ maps onto $\pi_1(Z_2^{\#})$ surjectively with kernel normally generated by $\delta_1^2$, $\delta_i^2=B_iA_i$ $(i\ugu2,\ldots , 5)$ and $(\delta_1\cdots\delta_5)^2$, thus $\pi_1(Z_2^{\#})$ is generated by $A_2,\ldots , A_5,\Gamma$ and has for relations the relations coming from the rewriting of $R_{\alpha}$, $\delta_1R_{\alpha}\delta_1^{-1}$, and the rewriting of $(\delta_1\cdots\delta_5)^2=1\!$, i.e. $A_2A_3^{-1}A_4A_5^{-1} A_2^{-1}A_3A_4^{-1}A_5=1$.\\ {\bf Remark:} This relation says that the four generators $A_2,\ldots , A_5$ are the generators of $\pi_1({\rm fibre})=\pi_1({\rm genus\ 2\ curve})$.\\ {\bf Step 5.}\\ Let $m\hbox{\mat \char61}\hskip1pt k\piu1$ and consider $X_{2m}^{\#}$, the non singular part of the m-fold cyclic cover of $X_2$ totally branched over $L_1''$ and $L_2''$.\\ To find a presentation of $X_{2m}^{\#}$ we first need a presentation of the kernel of the map $\pi_1(Z_{2}^{\#})\smash{\mathop{\longrightarrow}} \zeta_m$ such that $A_i{\longmapsto}\bar0$ and $\Gamma,\sigma{\longmapsto}\bar 1$, and then we add the relations $\Gamma^m=\sigma^m=1$.\\ Applying the Reidemeister-Shreier method, we find that the kernel is generated by $\Gamma^m$, $\Gamma^iA_j\Gamma^{-i}$ for $i\ugu1,\ldots , m\men1$ and $j\ugu2,\ldots , 5$, by $\Gamma^i\sigma\Gamma^{-i-1}$ for $i\ugu1,\ldots , m\men2$ and $\Gamma^{m-1}\sigma$; it has for relations the rewriting in the new generators of the relations $R_{\alpha}''$ of $\pi_1(Z_2^{\#})$ and the rewriting of $\Gamma^iR_{\alpha}''\Gamma^{-i}$ for $i\ugu1,\ldots , m\men1$. \section{Calculations.} {\bf Step 3.}\\ We have \begin{eqnarray*} \beta_0&=&\sigma_4^{12}\sigma_2^2\\ \beta_1&=&\sigma_1^{-1}\sigma_2\sigma_3\sigma_1\sigma_2^{-1}\sigma_1\\ \beta_{-1}&=&\sigma_4^{-6}\sigma_2^{-1}\beta_1\sigma_2\sigma_4^6\\ \beta_+&=&\sigma_1^{-2}\sigma_2\sigma_3\sigma_4\sigma_3^{-1}\sigma_2^{-1} \sigma_1^2\\ \beta_-&=&\sigma_4^{-6}\sigma_2^{-1}\beta_+\sigma_2\sigma_4^6 \end{eqnarray*} The relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_0$ are equivalent to the two relations \begin{equaz}\label{fofififo} (\delta_4\delta_5)^6=(\delta_5\delta_4)^6\end{equaz} \begin{equaz}\label{tttt} \delta_2\delta_3=\delta_3\delta_2.\end{equaz} The relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_+$ amount to \begin{equaz}\label{totot} \delta_5=\delta_2^{-1}\delta_1^{-1}\delta_2\delta_1\delta_2.\end{equaz} In fact, here and in the sequel, we use the following argument: $\beta_+$ is a conjugate $\sigma\sigma_4\sigma^{-1}$ of the braid $\sigma_4$ and the braid $\sigma_4$ yields the relation $\delta_4\hbox{\mat \char61}\hskip1pt\delta_5$. Therefore, if we set $\delta_4'\hbox{\mat \char61}\hskip1pt (\delta_4)\sigma^{-1}$, $\delta_5'\hbox{\mat \char61}\hskip1pt (\delta_5)\sigma^{-1}$, we get the relation $\delta_4'\hbox{\mat \char61}\hskip1pt\delta_5'$. By our particular choice of $\sigma$ \begin{eqnarray*} \delta_5'&\hbox{\mat \char61}\hskip1pt&\delta_5\\ \delta_4'&\hbox{\mat \char61}\hskip1pt&(\delta_4)\sigma_3^{-1}\sigma_2^{-1}\sigma_1^2\\ &\hbox{\mat \char61}\hskip1pt&(\delta_3)\sigma_2^{-1}\sigma_1^2\\ &\hbox{\mat \char61}\hskip1pt&(\delta_2)\sigma_1^2\\&\hbox{\mat \char61}\hskip1pt&(\delta_2^{-1}\delta_1\delta_2)\sigma_1\\ &\hbox{\mat \char61}\hskip1pt&\delta_2^{-1}\delta_1^{-1}\delta_2\delta_1\delta_2\end{eqnarray*} Similarly, the relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_-$ are equivalent to the relation \begin{equaz}\label{thothoth} \delta_3^{-1}\delta_1^{-1}\delta_3\delta_1\delta_3= (\delta_4\delta_5)^{-3}\delta_5(\delta_4\delta_5)^3. \end{equaz} We write down, for convenience of the reader, the action of the braid $\beta_1^{-1}$, since the new relations $\delta_i \hbox{\mat \char61}\hskip1pt(\delta_i)\beta_1\beta_j\beta_1^{-1}$ will be obtained from the relations equivalent to $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_j$ simply by applying the automorphism $\beta_1^{-1}$.\begin{eqnarray*} (\delta_1)\beta_1^{-1}&\hbox{\mat \char61}\hskip1pt& \delta_1\delta_2\delta_3\delta_2^{-1}\delta_1^{-1} \delta_2^{-1}\delta_1\delta_2\delta_4\delta_2^{-1}\delta_1^{-1} \delta_2\delta_1\delta_2\delta_3^{-1}\delta_2^{-1}\delta_1^{-1}\\ (\delta_2)\beta_1^{-1}&\hbox{\mat \char61}\hskip1pt& \delta_1\delta_2\delta_3\delta_2^{-1}\delta_1^{-1}\\ (\delta_3)\beta_1^{-1}&\hbox{\mat \char61}\hskip1pt&\delta_2^{-1}\delta_1\delta_2\delta_4^{-1} \delta_2^{-1}\delta_1^{-1}\delta_2\delta_1\delta_2\delta_4\delta_2^{-1} \delta_1^{-1}\delta_2\\ (\delta_4)\beta_1^{-1}&\hbox{\mat \char61}\hskip1pt&\delta_2^{-1}\delta_1\delta_2\\ (\delta_5)\beta_1^{-1}&\hbox{\mat \char61}\hskip1pt&\delta_5 \end{eqnarray*} Thus, the relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_1\beta_0\beta_1^{-1}$ are equivalent to the relations \begin{equaz}\label{otwtwo} (\delta_1\delta_2)^6=(\delta_2\delta_1)^6\end{equaz} \begin{equaz}\label{ftffff} \delta_5\delta_3\delta_5^{-1}\delta_4^{-1}\delta_5\delta_4= \delta_4^{-1}\delta_5\delta_4\delta_5\delta_3\delta_5^{-1} \end{equaz} where we have used \ref{totot}.\\ The relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_1\beta_+\beta_1^{-1}$ are equivalent to the relation \begin{equaz}\label{ftotfftfftot} \delta_5=\delta_2^{-1}\delta_1\delta_2\delta_4^{-1}\delta_5 \delta_3\delta_5^{-1}\delta_4\delta_2^{-1}\delta_1^{-1}\delta_2\end{equaz} where we have used \ref{totot}.\\ The relations $\delta_i\hbox{\mat \char61}\hskip1pt(\delta_i)\beta_1\beta_-\beta_1^{-1}$ pop up to the relation \begin{equaz}\label{thototh} \delta_3^{-1}\delta_1\delta_2\delta_1^{-1}\delta_3= (\delta_4\delta_5)^{-2}\delta_5(\delta_4\delta_5)^2. \end{equaz} In fact, using \ref{totot} we have $(\delta_4\delta_5)\beta_1^{-1}=\delta_1\delta_2$ and so $$((\delta_4\delta_5)^{-3}\delta_5(\delta_4\delta_5)^3) \beta_1^{-1}=(\delta_1\delta_2)^{-4}\delta_2(\delta_1\delta_2)^4$$ On the other side, $$(\delta_1\delta_3)\beta_1^{-1}=\delta_1\delta_2 \delta_3\delta_5^{-1}\delta_4\delta_5\delta_3^{-1}\delta_5^{-1}\delta_4^{-1} \delta_5\delta_4\delta_2^{-1}\delta_1^{-1}\delta_2$$ and using \ref{ftffff} and again \ref{totot} $$(\delta_1\delta_3)\beta_1^{-1}=\delta_1\delta_2 \delta_3\delta_4\delta_5\delta_3^{-1}\delta_5^{-1} \delta_2^{-1}\delta_1^{-1}\delta_2=\delta_1\delta_2 \delta_3\delta_4\delta_5\delta_3^{-1} \delta_2^{-1}\delta_1^{-1}.$$ With the same method we have \begin{eqnarray*} ((\delta_1\delta_3)^{-1}\delta_3\delta_1\delta_3)\beta_1^{-1}&=& \delta_1\delta_2\delta_3\delta_5^{-1}\delta_4^{-1}(\delta_3^{-1}\delta_5^{-1} \delta_4^{-1}\delta_5\delta_4\delta_5\delta_3)\delta_4\delta_5\delta_3^{-1} \delta_2^{-1}\delta_1^{-1}\\ &=&\delta_1\delta_2\delta_3\delta_5^{-1}\delta_4^{-1} (\delta_5^{-1}\delta_4^{-1} \delta_5\delta_4\delta_5)\delta_4\delta_5\delta_3^{-1} \delta_2^{-1}\delta_1^{-1}\\ &=&\delta_1\delta_2\delta_3(\delta_4\delta_5)^{-2} \delta_5(\delta_4\delta_5)^2\delta_3^{-1} \delta_2^{-1}\delta_1^{-1}\end{eqnarray*} So the relation is $$\delta_3(\delta_4\delta_5)^{-2}\delta_5(\delta_4\delta_5)^2\delta_3^{-1}= (\delta_1\delta_2)^{-5}\delta_2(\delta_1\delta_2)^5= \delta_1\delta_2\delta_1^{-1}$$ where we have used \ref{otwtwo}.\\ Finally we have to write the relations $\Gamma\delta_i\Gamma^{-1}= (\delta_i)\beta_1^2$, i.e. \begin{eqnarray*} \Gamma\delta_1\Gamma^{-1}&\!\!\!=\!\!\!&\Gamma\delta_2\Gamma^{-1}\delta_4^{-1} \delta_2^{-1}\delta_1\delta_2\delta_4\Gamma\delta_2^{-1}\Gamma^{-1} \hbox to 0pt{\hspace*{13.6em}\begin{numera}\label{GoG}\end{numera}\hss}\\ \Gamma\delta_2\Gamma^{-1}&\!\!\!=\!\!\!&\delta_2^{-1}\delta_1\delta_2 \delta_3^{-1}\delta_5\delta_3\delta_2^{-1}\delta_1^{-1}\delta_2 \hbox to 0pt{\hspace*{15.4em}\begin{numera}\label{GtwG}\end{numera}\hss}\\ \Gamma\delta_3\Gamma^{-1}&\!\!\!=\!\!\!& \delta_4^{-1}\delta_2^{-1}\delta_1^{-1}\delta_2\delta_4\delta_2^{-1} \delta_1\delta_2 \delta_3^{-1}\delta_5^{-1}\delta_3\delta_5\delta_3 \delta_2^{-1}\delta_1^{-1}\delta_2\delta_4^{-1}\delta_2^{-1}\delta_1\delta_2 \delta_4 \ \ \begin{numera}\label{GthG}\end{numera}\\ \Gamma\delta_4\Gamma^{-1}&\!\!\!=\!\!\!& \delta_4^{-1}\delta_2^{-1}\delta_1^{-1}\delta_2 \delta_4\delta_2^{-1}\delta_1\delta_2\delta_4 \hbox to 0pt{\hspace*{15em}\begin{numera}\label{GfoG} \end{numera}\hss}\\ \Gamma\delta_5\Gamma^{-1}&\!\!\!=\!\!\!&\delta_5 \hbox to 0pt{\hspace*{24.8em} \begin{numera}\label{GfiG} \end{numera}\hss} \end{eqnarray*} {\bf Step 4.}\\ We take as Shreier set for the left cosets of the kernel the set $\{S_0=1,S_1=\delta_1\}$, so applying the Reidemeister-Shreier method we get the generators $\Delta=\delta_1^2$, $\Gamma$, $\sigma=\delta_1\Gamma\delta_1^{-1}$, $A_i=\delta_1\delta_i$ and $B_i=\delta_i\delta_1^{-1}$ for $i\ugu2,3,4,5$. For the relations we must rewrite the relations \ref{fofififo},...,\ref{GfiG} and their conjugate by $\delta_1$ in terms of the new generators. The rewriting process goes as follows (cf. \cite{makaso}, pages 86-98): \begin{eqnarray*} S_0\delta_1=S_1&& S_1\delta_1=\Delta S_0\\ S_0\delta_i=B_iS_1&& S_1\delta_i=A_iS_0\ \ \ \ {\rm for}\ i\ugu2,3,4,5\\ S_0\Gamma=\Gamma S_0&& S_1\Gamma=\sigma S_1\end{eqnarray*} We want to show that it suffices to rewrite only the relations \ref{fofififo},...,\ref{GfiG}.\\ Observe that all our relations can be written in the form $W\delta_iW^{-1}=\delta_k$ for a suitable word $W$. Assume that $\Gamma$ doesn't appear in the relation and do the rewriting after moding out by the relations \begin{equaz}\hbox{$\Delta$}=B_iA_i=1.\label{canc1}\end{equaz} Since $S_0\delta_i\hbox{\mat \char61}\hskip1pt A_i^{-1}S_1$ and also $S_0\delta_i^{-1}\hbox{\mat \char61}\hskip1pt A_i^{-1}S_1$, if we write $\displaystyle W=\prod_{\lambda=1}^h\delta_{j_{\lambda}}^{\pm1}$, the rewriting of $W\delta_iW^{-1}\delta_k^{-1}$ is given by $$A_{j_1}^{-1}A_{j_2}\cdots A_i^{\pm1}\cdots A_{j_1}^{-1}A_k$$ (note that $A_1\ugu1$). The rewriting of the same relation conjugated by $\delta_1$ yields instead $$A_{j_1}A_{j_2}^{-1}\cdots A_i^{\mp1}\cdots A_{j_1}A_k^{-1}.$$ We get thus two relations of respective form $UA_k=1$, $U^{-1}A_k^{-1}=1$, which are obviously equivalent.\\ If instead $\Gamma$ appears in the relation, we have one of the \ref{GoG},...,\ref{GfiG} which are of the form $\Gamma\delta_i\Gamma^{-1}=W\delta_iW^{-1}$ where we can in fact assume that $\Gamma$ doesn't appear in the word $W$.\\ The rewriting of $\Gamma\delta_i\Gamma^{-1}W\delta_i^{-1}W^{-1}$ yields, again a relation of the form $$\Gamma A_i^{-1}\sigma^{-1}U^{-1}=1,$$ whereas the rewriting of the conjugate by $\delta_1$ gives a relation $$\sigma A_i\Gamma^{-1}U=1,$$ which is an equivalent relation.\\ For convenience of notation we shall keep the generators $B_i=A_i^{-1}$.\\ To calculate $\pi_1(Z_2^{\#})$ we must add the rewriting of $(\prod_{i=1}^5\delta_i)^2=1$ which gives $$A_2B_3A_4B_5B_2A_3B_4A_5=1.$$ We have thus that $\pi_1(Z_2^{\#})$ is generated by $A_2$, $A_3$, $A_4$, $A_5$, $\Gamma$ and $\sigma$ and has the following set of relations \begin{equaz}\label{3)'} (B_4A_5)^6=(B_5A_4)^3 \end{equaz} \begin{equaz}\label{1)'} B_3A_2=B_2A_3\end{equaz} \begin{equaz}\label{4)'} B_5=B_2^3 \end{equaz} \begin{equaz}\label{2)'} B_3^3=(B_5A_4)^6B_5 \end{equaz} \begin{equaz}\label{5)'} A_2^{12}=1 \end{equaz} \begin{equaz}\label{6)'} B_5A_3B_5A_4B_5A_4=B_4A_5B_4A_5B_3A_5 \end{equaz} \begin{equaz}\label{7)'} B_5=B_2^2A_4B_5A_3B_5A_4B_2^2 \end{equaz} \begin{equaz}\label{8)'} B_3B_2B_3=(B_5A_4)^4B_5 \end{equaz} \begin{equaz}\label{12)'} \sigma A_2^2\Gamma^{-1}=A_4B_2^2A_4 \end{equaz} \begin{equaz}\label{10)'}\Gamma B_2\sigma^{-1}=B_2^2A_3B_5A_3B_2^2\end{equaz} \begin{equaz}\label{13)'}\Gamma B_3\sigma^{-1}= B_4A_2^2B_4A_2^2B_3A_5B_3A_5B_3A_2^2B_4A_2^2B_4\end{equaz} \begin{equaz}\label{11)'}\Gamma B_4\sigma^{-1}=B_4A_2^2B_4A_2^2B_4\end{equaz} \begin{equaz}\label{9)'} \Gamma B_5\sigma^{-1}=B_5\end{equaz} \begin{equaz}\label{14)'} A_2B_3A_4B_5B_2A_3B_4A_5=1\end{equaz} where $B_i=A_i^{-1}$.\\ Let's reduce this presentation. Using \ref{4)'} relation \ref{7)'} becomes \begin{equaz}B_4A_2B_4=B_5A_3B_5\label{7)''}\end{equaz} and with this \ref{6)'} becomes $$B_2^4=1$$ which implies \ref{5)'}, and changes \ref{4)'} into $$B_5=A_2.$$ Moreover, using \ref{8)'} and the last equation, relation \ref{2)'} gives $$(A_2A_4)^2=A_3A_2B_3^2$$ and with this, using also \ref{1)'}, \ref{8)'} becomes \begin{equaz}\label{8)''}B_3B_2=A_2A_3\end{equaz} thus transforming \ref{7)''} into $$B_3=B_4A_2B_4$$ which allows us to delete the generator $A_3$. Upon substituting the expressions of $A_5$ and $A_3$ into \ref{14)'} and \ref{8)''} we have $$A_2A_4=A_4A_2$$ $$A_4^4=1.$$ We can then see that the relations \ref{1)'},...,\ref{14)'} are equivalent to the following $$A_5=A_2^{-1}\ \ \ \ \ \ \ A_3=A_2^{-1}A_4^2$$ $$A_2^4=A_4^4=1\ \ \ \ \ \ \ A_2A_4=A_4A_2$$ $$\sigma A_2^2\Gamma^{-1}=A_2^2A_4^2$$ $$\Gamma A_2^{-1}=A_2^{-1}\sigma$$ $$\Gamma A_2A_4^2=A_2A_4^2\sigma$$ $$\Gamma A_4^{-1}=A_4\sigma$$ $$\Gamma A_2=A_2\sigma.$$ {\bf Step 5.}\\ We take as Shreier set for the left cosets of the kernel the set $$\{R_i\hbox{\mat \char61}\hskip1pt\Gamma^i\ |\ i\ugu0,1,\ldots , m\men1\}$$ and we apply the Reidemeister-Shreier method.\\ The generators are $\hat\Gamma=\Gamma^m$, $A_{2,i}=\Gamma^iA_2\Gamma^{-i}$, $A_{4,i}=\Gamma^iA_4\Gamma^{-i}$ for $i\ugu0,\ldots , m\men1$ $\sigma_i=\Gamma^i\sigma\Gamma^{-(i+1)}$ for $i\ugu0,\ldots , m\men2$ and $\sigma_{m-1}=\Gamma^{m-1}\sigma$.\\ For the rewriting process we have \begin{eqnarray*}R_iA_j&=&A_{j,i}R_i\ \ \ \ {\rm for}\ j\ugu2,4\ i\ugu0,\ldots , m-1\\ R_i\Gamma&=&R_{i+1}\ \ \ \ \ {\rm for}\ i\ugu0,\ldots , m-2\\ R_{m-1}\Gamma&=&\hat\Gamma R_0\\ R_i\sigma&=&\sigma_i R_{i+1}\ \ \ \ {\rm for}\ i\ugu0,\ldots , m-2\\ R_{m-1}\sigma&=&\sigma_{m-1}R_0. \end{eqnarray*} Thus, taking indices $i$ (mod$m$) and adding (as we must) the relation $\hat\Gamma=1$, we obtain the relations $$A_{2,i}^4=A_{4,i}^4=1$$ $$A_{2,i}A_{4,i}=A_{4,i}A_{2,i}$$ $$\sigma_iA_{2,i+1}^2=A_{2,i}^2A_{4,i}^2$$ $$A_{2,i+1}^{-1}=A_{2,i}^{-1}\sigma_i$$ $$A_{2,i+1}A_{4,i+1}^2=A_{2,i}A_{4,i}^2\sigma_i$$ $$A_{4,i+1}^{-1}=A_{4,i}\sigma_i$$ $$A_{2,i+1}=A_{2,i}\sigma_i.$$ To simplify this presentation we write \begin{eqnarray*}\sigma_i&=&A_{2,i}^2A_{4,i}^2A_{2,i+1}^2\\ &=&A_{2,i}A_{2,i+1}^{-1}\\ &=&A_{4,i}^2A_{2,i}^{-1}A_{2,i+1}A_{4,i+1}^2= A_{2,i}^{-1}A_{4,i}^2A_{4,i+1}^2A_{2,i+1}\\ &=&A_{4,i}^{-1}A_{4,i+1}^{-1}\\ &=&A_{2,i}^{-1}A_{2,i+1}\end{eqnarray*} From the last and the second equations we get $$A_{2,i}^2=A_{2,i+1}^2=A_{2,0}^2$$ and from the first one, remembering that $A_{2,i}$ commutes with $A_{4,i}$ and that $A_{2,0}^4=1$, $$\sigma_i=A_{4,i}^2.$$ The fourth equation then gives $$A_{4,i}=A_{4,i+1}=A_{4,0}$$ which makes the last and the third relations equivalent. These two cancellation relations enable us to delete all the generators $\sigma_j$ and $A_{4,i}$ for $i\ugu1,\ldots , m-1$.\\ We may rewrite the five relations above as \begin{eqnarray*}\sigma_i&=&A_{4,0}^2\\ A_{4,i}&=&A_{4,0}\\ A_{4,0}^2&=&A_{2,i}^{-1}A_{2,i+1}\\ A_{4,0}^2&=&A_{2,i}A_{2,i+1}^{-1}. \end{eqnarray*} Clearly the last two equations are equivalent and give \begin{equaz}A_{2,2i}=A_{2,0}\ \ \ \ \ \ A_{2,2i+1}=A_{2,0}A_{4,0}^2\label{abv}.\end{equaz} Moreover, if we add the relation $\sigma^m=1$, that in the generators of $\pi_1(X_{2m}^{\#})$ reads out as $\sigma_0\sigma_1\cdots\sigma_{m-1}=1$, we get $A_{4,0}^{2m}=1,$ i.e., if $m$ is odd, $A_{4,0}^2=1$, while if $m$ is even we have no new relations. Observe that this is in accordance with the fact that in \ref{abv} the index is cyclic mod$(m)$.\\ Summing up, we have a commutative group with only two generators, namely $a=A_{2,0}$ and $b=A_{4,0}$, such that $a^4=1$ and $b^4=1$ if $m$ is even, $b^2=1$ if $m$ is odd, i.e. $$\pi_1(X^{\#}_{2k+2})=\zeta_4\times\zeta_4$$ if $k$ is odd and $$\pi_1(X^{\#}_{2k+2})=\zeta_4\times\zeta_2$$ if $k$ is even.
1996-07-23T20:36:16
9607
alg-geom/9607024
en
https://arxiv.org/abs/alg-geom/9607024
[ "alg-geom", "math.AG" ]
alg-geom/9607024
Rahul Pandharipande
R. Pandharipande
The Equivariant Chow Ring of SO(4)
10 pages, Latex2e
null
null
null
null
The integral equivariant Chow ring of S0(4) is computed via the geometry of ruled quadric surfaces in P^3.
[ { "version": "v1", "created": "Tue, 23 Jul 1996 18:32:10 GMT" } ]
2008-02-03T00:00:00
[ [ "Pandharipande", "R.", "" ] ]
alg-geom
\section{\bf{Introduction}} Let $\mathbf{G}$ be a reductive algebraic group. The algebraic analogue of $E\mathbf{G}$ is attained by approximation. Let $V$ be a $\mathbb{C}$-vector space. Let $\mathbf{G}\times V \rightarrow V$ be an algebraic representation of $\mathbf{G}$. Let $W\subset V$ be a $\mathbf{G}$-invariant open set satisfying: \begin{enumerate} \item[(i)] The complement of $W$ in $V$ is of codimension greater than $q$. \item[(ii)] $\mathbf{G}$ acts freely on $W$. \item[(iii)] There exists a geometric quotient $W\rightarrow W/\mathbf{G}$. \end{enumerate} $W$ is an approximation of $E\mathbf{G}$ up to codimension $q$. Let $e=dim( W/{\mathbf{G}})$. The equivariant Chow groups of $\mathbf{G}$ (acting on a point) are defined by: \begin{equation} \label{defff} A^{\mathbf{G}}_{-j}(\text{point})= A_{e-j}(W/\mathbf{G}) \end{equation} for $0\leq j \leq q.$ An argument is required to check the Chow groups are well-defined (see [EG1]). The basic properties of equivariant Chow groups are established in [EG1]. In particular, there is a natural intersection ring structure on $A_i^{\mathbf{G}}(\text{point})$. For notational convenience, a superscript will denote the Chow group codimension: $$A^{\mathbf{G}}_{-j}(\text{point}) = A^j_\mathbf{G}(\text{point}).$$ Equation (\ref{defff}) becomes: $$\forall\ 0\leq j \leq q, \ \ A_{\mathbf{G}}^{j}(\text{point})= A^j(W/\mathbf{G}).$$ $W/\mathbf{G}$ is an approximation of $B \mathbf{G}$. $A_{\mathbf{G}}^*(\text{point})$ is called the (equivariant) Chow ring of $\mathbf{G}$. $A_{\mathbf{G}}^*(\text{point})$ is naturally isomorphic to the ring of algebraic characteristic classes of (\'etale locally trivial) principal $\mathbf{G}$-bundles. The equivariant Chow ring of $\mathbf{G}$ was first defined by B. Totaro in [T]. Consider now the orthogonal and special orthogonal algebraic groups (over $\mathbb{C}$). The Chow ring of $\mathbf{O} (n)$ is generated by the Chern classes of the standard representation. The odd classes are 2-torsion: $$A^*_{\mathbf{O} (n)}(\text{point})= \mathbb{Z}[c_1, \ldots, c_n]/(2c_1, 2c_3, 2c_5, \ldots).$$ The Chow ring of $\mathbf{SO} (n=2k+1)$ is also generated by the Chern classes of the standard representation. The odd classes are 2-torsion and $c_1=0$: $$A^*_{\mathbf{SO} (n=2k+1)}(\text{point})= \mathbb{Z}[c_1, \ldots, c_n]/(c_1, 2c_3, 2c_5, \ldots).$$ $A^*_{\mathbf{O} (n)}(\text{point})$ was first computed by B. Totaro. Algebraic computations of $A^*_{\mathbf{O} (n)}(\text{point})$ and $A^*_{\mathbf{SO} (2k+1)}(\text{point})$ can be found in [P2]. The Chow ring of $\mathbf{SO}(n)$ has been computed with $\mathbb{Q}$-coefficients in [EG2]. The integral Chow ring of $\mathbf{SO}(n=2k)$ is not known in general. Since $B \mathbf{O} (n)$ is approximated by the set of non-degenerate quadratic forms in $Sym^2 S^*$ (where $S \rightarrow \mathbf G(n, \infty)$ is the tautological sub-bundle), the Chow ring can be analyzed by degeneracy calculations ([P2]). Algebraic $B \mathbf{SO} (n)$ double covers $B \mathbf{O}(n)$. If $n=2k+1$, there is a product decomposition $\mathbf{O}(n)\stackrel{\sim}{=} \mathbb{Z}/2\mathbb{Z} \times \mathbf{SO}(n)$. As a result, the double cover geometry for $n=2k+1$ is tractable and a computation of $A_{\mathbf{SO}(2k+1)}^*(\text{point})$ can be made. In case $n=2k$, the double cover geometry is more complicated. Since $\mathbf{SO}(2)\stackrel{\sim}{=} \mathbb{C}^*$, the first non-trivial even case is $\mathbf{SO}(4)$. In this paper, the ring $A^*_{\mathbf{SO}(4)}(\text{point})$ is determined. Let $c_1,c_2,c_3,c_4$ be Chern classes of the standard representation of $\mathbf{SO}(4)$. Let $F$ be one of the two distinct irreducible $3$-dimensional representations of $\mathbf{SO}(4)$, and let $f_2$ be the second Chern class $F$. \begin{tm} $A_{\mathbf{SO}(4)}^*(\text{\em point})$ is generated by the Chern classes $c_1, c_2, c_3,$ $c_4,$ and $f_2$. Define $x\in A_{\mathbf{SO}(4)}^{2}$ by $x=c_2-f_2$. $$A_{\mathbf{SO}(4)}^*(\text{\em point})= \mathbb{Z}[c_1,c_2,c_3,c_4,x]/ (c_1, 2c_3, xc_3, x^2-4c_4)$$ \end{tm} \noindent Let $\tilde{F}$ be the other irreducible $3$-dimensional representation of $\mathbf{SO}(4)$. Let $\tilde{f}_2$ be the second Chern class of $\tilde{F}$. Since (see [FH]) $$F \oplus \tilde{F} \stackrel{\sim}{=} \wedge^2 V,$$ the relation $\tilde{f}_2= 2c_2-f_2$ is obtained. Hence, $c_2-\tilde{f}_2= -x$. The presentation in Theorem 1 does not depend upon the choice of $3$-dimensional representation. Thanks are due to D. Edidin, W. Fulton, W. Graham, and B. Totaro for conversations about $B \mathbf{SO}(n)$. The $\mathbf{SO} (4)$ calculation presented here is similar in spirit to the $\mathbf{SO}(2k+1)$ calculations of [P2]. In [P1] and [P2], equivariant Chow rings are used to compute ordinary Chow rings of certain moduli spaces of maps and Hilbert schemes of rational curves. \section{\bf{Ruled Quadric Surfaces}} \label{rrr} Let $V\stackrel{\sim}{=} \mathbb{C}^4$ be equipped with a non-degenerate quadratic form $Q$. A {\em ruled quadric surface} in $\mathbf P(V)$ is a pair $(X,r)$ where $X\subset \mathbf P(V)$ is a nonsingular quadric surface and $r$ is a choice of ruling. Let $\mathcal{X}\subset \mathbf P(Sym^2 V^*)$ be the parameter space of nonsingular quadrics. The parameter space of ruled quadrics, $\mathcal{X}_{ruled}$, is an \'etale double cover of $\mathcal{X}$ via the natural map: $\mathcal{X}_{ruled} \rightarrow \mathcal{X}$. There are natural maps $\mathbf{SO}(V) \rightarrow \mathbf P\mathbf{SO}(V)\subset \mathbf{PGL}(V)$ and $\mathbf{O}(V) \rightarrow \mathbf P\mathbf{O}(V) \subset \mathbf{PGL}(V)$. Let $\mathbf{SO}(V)$ and $\mathbf{O}(V)$ act on $\mathbf{PGL}(V)$ on the right via these maps. There exist geometric quotients (see [P2]): $$\mathbf{PGL}(V)/\mathbf{SO}(V) \rightarrow \mathbf{PGL}(V)/\mathbf{O}(V).$$ Consider the quadric surface $(Q)\subset \mathbf P(V)$ obtained from the quadratic form. The standard left action $\mathbf{PGL}(V)\times \mathbf P(V) \rightarrow \mathbf P(V)$ yields a transitive $\mathbf{PGL}(V)$-action on the space of nonsingular quadric surfaces. The stabilizer of $(Q)$ for this action is exactly $\mathbf P\mathbf{O}(V) \subset \mathbf{PGL}(V)$. Hence, there is a canonical isomorphism $$\mathbf{PGL}(V)/\mathbf{O}(V) \stackrel{\sim}{=} \mathcal{X}.$$ For the entire paper, fix a ruling $r$ of $(Q)$. Since $\mathbf{PGL}(V)$ acts transitively on the space of ruled quadrics and the stabilizer of $((Q),r)$ is exactly $\mathbf P\mathbf{SO}(V) \subset \mathbf{PGL}(V)$, there is a canonical isomorphism determined by $((Q),r)$: $$\mathbf{PGL}(V)/\mathbf{SO}(V) \stackrel{\sim}{=} \mathcal{X}_{ruled}.$$ There is a canonical Pl\"ucker embedding $\mathbf G(2, V) \hookrightarrow \mathbf P(\wedge^2 V)$. Let $Z\subset \mathbf G(3, \wedge^2 V)$ be the open locus of $2$-planes in $\mathbf P(\wedge^2 V)$ which intersect $\mathbf G(2,V)$ transversely in a nonsingular conic curve. \begin{lm} \label{trick} There is a canonical isomorphism $Z \stackrel{\sim}{=} \mathcal{X}_{ruled}$. \end{lm} \noindent {\em Proof.} The family of lines determined by a nonsingular plane conic $C \subset \mathbf G(2,V) \subset \mathbf P(\wedge^2 V)$ sweeps out an irreducible degree $2$ surface in $\mathbf P(V)$. There are three possibilities for this degree $2$ surface: a double plane, a quadric cone, or a nonsingular quadric surface. If the conic $C$ sweeps out a a double plane $H\subset \mathbf P(V)$, then $C \subset P \subset \mathbf G(2,V)$ where $P$ is the plane of all lines contained in $H$. If $C$ sweeps out a quadric cone, then $C\subset P \subset \mathbf G(2,V)$ where $P$ is the plane of all lines passing through the vertex of the cone. Hence, if $C$ is the transverse intersection $P\cap \mathbf G(2,V)$ of a plane, then $C$ must correspond to a ruling of a unique nonsingular quadric surface. Conversely, a ruling of a nonsingular quadric surface yields a conic curve in $\mathbf G(2,V)$ which is the transverse intersection of a unique $2$-plane in $\mathbf P(V)$. These maps are easily seen to be algebraic. \qed \vspace{+10pt} By Lemma 1, the ruled quadric $((Q),r)$ corresponds to a $3$-dimensional subspace $F\subset \wedge^2 V$. $F$ is $\mathbf{SO}(V)$-invariant since $((Q),r)$ is stabilized by $\mathbf{SO}(V)$. $F$ is therefore a $3$-dimensional representation of $\mathbf{SO}(V)$. Let $s$ be the {\em other} ruling of $(Q)$. $((Q),s)$ similarly corresponds to an invariant $3$-dimensional subspace $\tilde{F} \subset \wedge^2 V$. The $\mathbf{SO}(V)$ representation $\wedge^2 V$ decomposes as $\wedge^2 V \stackrel{\sim}{=} F \oplus \tilde{F}$. \section{ $B\mathbf{SO}(V)$} \label{bsofour} Let $V\stackrel{\sim}{=} \mathbb{C}^4$ be equipped with a non-degenerate quadratic form as before. Approximations to $E\mathbf{SO}(V)$ and $B\mathbf{SO}(V)$ are obtained via direct sums of the representation $V^*$. Let $m>>0$ and let $$W_m \subset \oplus_{1}^{m} V^*$$ denote the spanning locus. $W_m$ is the locus of $m$-tuples of vectors of $ V^*$ which span $V^*$. The natural action of $\mathbf{SO} (V)$ on $W_m$ is free and has a geometric quotient (see [P2]). The codimension of the complement of $W_m$ in $\oplus_{1}^{m} V^*$ is $m-3$. $W_m$ is an approximation of $E\mathbf{SO} (V)$ up to codimension $m-4$. Therefore $$B\mathbf{SO} (V)=\ \stackrel{Lim}{m \rightarrow \infty} \ W_m/\mathbf{SO} (V),$$ $$A^*_{\mathbf{SO} (V)}(\text{point})=\ \stackrel{Lim}{m \rightarrow \infty} \ A^*(W_m/\mathbf{SO} (V)).$$ There is a scalar $\mathbb{C}^*$-action on $W_m$. Let $\mathbf P(W_m)=W_m/\mathbb{C}^*$. Since this $\mathbb{C}^*$-action commutes with the $\mathbf{SO}(V)$-action, there is diagram of quotients: \begin{equation} \label{diax} \begin{CD} W_m @>{\tau_1}>> W_m/ \mathbf{SO}(V) \\ @V{i_1}VV @V{i_2}VV \\ \mathbf P(W_m)= W_m/\mathbb{C}^* @>{\tau_2}>>\mathbf P( W_m)/ \mathbf{SO} (V)\\ \end{CD} \end{equation} All the maps in (\ref{diax}) are quotient maps: \begin{enumerate} \item[(i)] $i_1$ is a free $\mathbb{C}^*$-quotient. \item[(ii)] $i_2$ is a free $\mathbb{C}^*/(\pm)$-quotient. \item[(iii)] $\tau_1$ is a free $\mathbf{SO}(V)$-quotient. \item[(iv)] $\tau_2$ is a free $\mathbf P\mathbf{SO}(V)$-quotient. \end{enumerate} The existence of these quotients is easily deduced (see [P2]). First consider the space $\mathbf P(W_m)/ \mathbf{SO}(V)$. Let $Q$ be the quadratic form on $V$ and let $r$ be the ruling of the quadric surface $(Q)\subset \mathbf P(V)$ fixed in section \ref{rrr}. An element $f \in \mathbf P(W_m)$ yields a canonical embedding $$\mu_f: \mathbf P(V) \hookrightarrow \mathbf P(\mathbb{C}^m).$$ The image under $\mu_f$ of $((Q),r)$ is a ruled quadric surface in $\mathbf P(\mathbb{C}^m)$ associated canonically to $f \in \mathbf P(W_m)$. Since $\mathbf P\mathbf{SO}(V)\subset \mathbf{PGL}(V)$ is exactly the stabilizer of the ruled quadric $((Q),r)$, it follows that $\mathbf P(W_m)/ \mathbf{SO}(V)$ is isomorphic to the parameter space of ruled quadric surfaces in $\mathbf P(\mathbb{C}^m)$. Since a ruled quadric surface in $\mathbf P(\mathbb{C}^m)$ spans a unique $3$-plane in $\mathbf P(\mathbb{C}^m)$, the parameter space is fibered over $\mathbf G(4,m)$. By Lemma \ref{trick}, the parameter space of ruled quadric surface in $\mathbf P(\mathbb{C}^m)$ is canonically isomorphic to an open set $$Z \subset \mathbf G(3, \wedge^2 S)$$ where $S\rightarrow \mathbf G(4,m)$ is the tautological sub-bundle. The Chow computations in section \ref{calcc} will require two results about line bundles. We have seen $\mathbf P(W_m)/\mathbf{SO}(V)$ is canonically fibered over $\mathbf G(4,m)$. Let $c_1$ be the first Chern class of the tautological bundle $S$ on $\mathbf G(4,m)$. Let $c_1$ also denote the pull-back of this class to $\mathbf P(W_m)/ \mathbf{SO}(V)$. For $m>4$, $A^1(\mathbf P(W_m)) \stackrel{\sim}{=} \mathbb{Z}$ with generator $c_1({\mathcal{O}}_{\mathbf P}(-1))$ (which is the Chern class of the line bundle associated to the $\mathbb{C}^*$-bundle $i_1$.) \begin{lm} \label{lastl} $\tau_2^*(c_1)= c_1({\mathcal{O}}_{\mathbf P}(-4))$. \end{lm} \noindent {\em Proof.} Elements of $\mathbf P(W_m)$ correspond to embeddings of $\mathbf P(V)$ in $\mathbf P(\mathbb{C}^m)$. The class $\tau_2^*(-c_1)$ is the divisor class of embeddings that meet a fixed $(m-5)$-plane in $\mathbf P(\mathbb{C}^m)= \mathbf P^{m-1}$. This divisor class is determined by a $4\times 4$ determinant. Hence, $\tau_2^*(-c_1)=c_1({\mathcal{O}}_{\mathbf P}(4))$. \qed \vspace{+10pt} The map $i_2: W_m/ \mathbf{SO}(V) \rightarrow \mathbf P(W_m)/ \mathbf{SO}(V)$ is a $\mathbb{C}^*/(\pm)$-bundle. Since there is an abstract isomorphism $\mathbb{C}^*/(\pm)\stackrel{\sim}{=} \mathbb{C}^*$, $i_2$ is also a $\mathbb{C}^*$-bundle. Let $N$ be the line bundle on $\mathbf P(W_m)/\mathbf{SO}(V)$ canonically associated to $i_2$. \begin{lm} \label{twot} $\tau_2^*(N) \stackrel{\sim}{=} {\mathcal{O}}_{\mathbf P}(-2)$. \end{lm} \noindent {\em Proof.} Let $i_1/(\pm): W_m/(\pm) \rightarrow \mathbf P(W_m)$. The map $i_1/(\pm)$ is a free $\mathbb{C}^*/(\pm)$-quotient. The line bundle associated to the $\mathbb{C}^*/(\pm)$-bundle $i_1/(\pm)$ is ${\mathcal{O}}_{\mathbf P}(-2)$. The map $\tau_1/(\pm): W_m/(\pm) \rightarrow W_m/ \mathbf{SO}(V)$ is $\mathbb{C}^*/(\pm)$-equivariant. Hence, $\tau_2^*(N)\stackrel{\sim}{=} {\mathcal{O}}_{\mathbf P}(-2)$. \qed \vspace{+10pt} \section{\bf{ Chow Calculations}} \label{calcc} In this section, the Chow ring of $W_m/ \mathbf{SO}(V)$ is determined (up to codimension $m-4$). Consider the parameter space of ruled quadrics in $\mathbf P(\mathbb{C}^m)$: $$Z \subset \mathbf G(3,\wedge^2 S) \rightarrow \mathbf G(4,m).$$ Let $D$ be the complement of $Z$ in $\mathbf G(3, \wedge^2 S)$. Following the notation of section \ref{bsofour}, $W_m/\mathbf{SO}(V)$ is the $\mathbb{C}^*$-bundle associated to a line bundle $N\rightarrow Z$. Therefore, $$A^*(W_m/\mathbf{SO}(V)) \stackrel{\sim}{=} A^*(Z)/ (c_1(N)) \stackrel{\sim}{=} A^*(\mathbf G(3,\wedge^2 S))/(I_D, c_1(\overline{N}))$$ where $I_D \subset A^*(\mathbf G(3,\wedge^2 S))$ is the ideal generated by cycles supported on $D$ and $\overline{N}$ is any extension of $N$ to $\mathbf G(3,\wedge^2 S)$. The ideal $I_D$ is determined by constructing a well-behaved variety which surjects onto $D$. Let $\mathbf G(2,S) \hookrightarrow \mathbf P(\wedge^2 S)$ be the canonical relative Pl\"ucker embedding. $D$ is exactly the locus of $2$-planes in the the fibers of $\mathbf P(\wedge^2 S)$ which do not meet $\mathbf G(2,S)$ transversely in a nonsingular conic curve. Equivalently, $D$ is the locus of $2$-planes $P$ in the fibers of $\mathbf P(\wedge^2 S)$ which satisfy one of the following conditions: \begin{enumerate} \item[(i)] $P\cap \mathbf G(2,S)$ is a pair of distinct lines in $P$. \item[(ii)] $P \cap \mathbf G(2,S)$ is a double line in $P$. \item[(iii)] $P \cap \mathbf G(2,S)= P$. \end{enumerate} $D$ is dominated by a canonical Grassmannian bundle over $\mathbf G(2,S)$. Let $B\rightarrow \mathbf G(2,S)$ be the tautological sub-bundle. By wedging, there is canonical surjective bundle map on $\mathbf G(2,S)$: $$\wedge^2 S \otimes \wedge^2 B \rightarrow \wedge^4 S$$ which induces a canonical sequence on $\mathbf G(2,S)$: \begin{equation} \label{aaa} 0 \rightarrow K \rightarrow \wedge^2 S \rightarrow \wedge^4 S \otimes (\wedge^2 B)^*\rightarrow 0. \end{equation} There is a canonical inclusion $\wedge^2 B \subset K$ and a quotient sequence \begin{equation} \label{bbb} 0 \rightarrow \wedge^2 B \rightarrow K \rightarrow E \rightarrow 0 \end{equation} on $\mathbf G(2,S)$. The geometric interpretation of these sequences is as follows. Let $\xi\in \mathbf G(2,S)$. $\mathbf P(K_\xi)\subset\mathbf P( \wedge^2 S_\xi)$ is the projective tangent space to $\mathbf G(2, S_\xi)$ at $\xi$. $\mathbf P(\wedge^2 B_\xi)$ in $\mathbf P(\wedge^2 S_\xi)$ is the Pl\"ucker image of the point $\xi$. The fiber of the Grassmannian bundle $$\mathbf G(2,E) \rightarrow \mathbf G(2,S)$$ over $\xi$ corresponds to the $2$-planes $P$ of $\mathbf P(\wedge^2 S_{\xi})$ that are tangent to $\mathbf G(2,S)$ at $\xi$. There is a canonical map $$\rho: \mathbf G(2,E) \rightarrow D$$ which is a surjection of algebraic varieties. Let $[P]\in D$. The fiber of $\rho$ over $[P]$ is simply the set of points of $P \cap \mathbf G(2,S)$ where $P$ is tangent to $\mathbf G(2,S)$. In case (i) above, the fiber is a point. In case (ii), the fiber is a straight line in $\mathbf P(\wedge^2 S)$. In case (iii), the fiber is $2$-plane in $\mathbf P(\wedge^2 S)$. Hence, there is stratification of $D$ by intersection type (i-iii) where $\rho$ is a projective bundle over each stratum. The Chow groups of $\mathbf G(2,E)$ therefore surject upon the Chow groups of $D$. The ideal $I_D$ is determined by calculating the push-forwards of the Chow classes of $\mathbf G(2,E)$ to $\mathbf G(3, \wedge^2 S)$. Consider the projection $$\pi:\mathcal{G}= \mathbf G(3, \wedge^2 S) \times _{\mathbf G(4,m)} \mathbf G(2,S) \rightarrow \mathbf G(3,\wedge^2 S).$$ The sequences (\ref{aaa}) and (\ref{bbb}) pull-back to $\mathcal{G}$. Let $F\rightarrow \mathbf G(3, \wedge^2 S)$ denote the tautological sub-bundle (and also let $F$ denote the pull-back to $\mathcal{G}$ of this bundle). There is a canonical inclusion $$\iota: \mathbf G(2,E) \hookrightarrow \mathcal{G}$$ determined by the sequences (\ref{aaa}) and (\ref{bbb}). $\mathbf G(2,E)\subset \mathcal{G}$ is the closed subvariety of points $g \in \mathcal{G}$ where $$\wedge^2 B_g \subset F_g \subset K_g.$$ The class $[\mathbf G(2,E)]$ in Chow ring of $A^*(\mathcal{G})$ is easily found by degeneracy calculations. Let $c_1,c_2, c_3, c_4$ be the Chern classes of $S\rightarrow\mathbf G(4,m)$. Let $b_1, b_2$ be the Chern classes of $B \rightarrow \mathbf G(2,S)$. Let $f_1,f_2,f_3$ be the Chern classes of $F\rightarrow \mathbf G(3,\wedge^2 S)$. Since $\mathcal{G}$ is a tower of Grassmannian bundles, these Chern classes $c_i, b_j, f_k$ generate $A^*(\mathcal{G})$. Let $Y$ be the locus of points $g\in \mathcal{G}$ such that $F_g \subset K_g$. $Y$ is the nonsingular degeneracy locus of the canonical bundle map on $\mathcal{G}$, $$ F \rightarrow \wedge^4 S \otimes (\wedge^2 B)^*,$$ obtained from the inclusion $F \subset \wedge^2 S$ and sequence (\ref{aaa}). By the Thom-Porteous formula on $\mathcal{G}$ (see [F]), \begin{eqnarray*} A^*(\mathcal{G})\ni [Y] &= &c_3(F^*\otimes \wedge^4 S \otimes (\wedge^2 B)^*) \\ &=& -f_3 +(c_1-b_1)f_2- (c_1-b_1)^2 f_1+(c_1-b_1)^3. \end{eqnarray*} $Y$ is canonically isomorphic to the Grassmannian bundle $\mathbf G(3, K) \rightarrow \mathbf G(2,S)$. There is natural bundle quotient sequence on $Y$: \begin{equation} \label{ccc} 0 \rightarrow F \rightarrow K \rightarrow K/F \rightarrow 0. \end{equation} The locus $\mathbf G(2,E)\subset Y$ is the set of points $y\in Y$ such that $\wedge^2 B_y \subset F_y$. $\mathbf G(2,E)\subset Y$ is the nonsingular degeneracy locus of the canonical bundle map on $Y$, $$\wedge^2 B \rightarrow K/F,$$ obtained from the sequences (\ref{bbb}) and (\ref{ccc}). By the Thom-Porteous formula on $Y$, \begin{eqnarray*} A^*(Y)\ni [\mathbf G(2,E)] & = &c_2( (K/F)\otimes (\wedge^2 B)^*) \\ & =& b_1^2-c_1b_1+ c_1^2-2c_1 f_1+ f_1^2-f_2+2c_2. \end{eqnarray*} The class $[\mathbf G(2,E)] \in A^*(\mathcal{G})$ is there expressed by \begin{eqnarray*} A^*(\mathcal{G}) \ni [\mathbf G(2,E)] & = & \big(-f_3 +(c_1-b_1)f_2- (c_1-b_1)^2 f_1+(c_1-b_1)^3\big) \\ & & \cdot \big(b_1^2-c_1b_1+ c_1^2-2c_1 f_1+ f_1^2-f_2+2c_2\big). \end{eqnarray*} Since $\mathbf G(2,E)$ is a Grassmannian bundle over $\mathbf G(2,S)$, the Chow ring of $\mathbf G(2,E)$ is generated over $A^*(\mathbf G(2,S))$ by the Chern classes $h_1, h_2$ of the tautological sub-bundle $H\rightarrow \mathbf G(2,E)$. Via the embedding $\iota:\mathbf G(2,E) \hookrightarrow \mathcal{G}$, $H$ is isomorphic to $\iota^*(F)/ \iota^*(\wedge^2 B)$. The Chern classes $h_1$ and $h_2$ can be expressed via $\iota$ in terms of the classes $b_j$ and $f_k$. Therefore, the classes \begin{equation*} \mathbf G(2,E) \cap M(c_1,c_2,c_3,c_4,b_1,b_2,f_1,f_2,f_3) \end{equation*} (where $M$ is monomial in the Chern classes) span the Chow ring of $\mathbf G(2,E)$. The ideal $I_D \subset A^*(\mathbf G(3, \wedge^2 S))$ is generated by the $\pi$ push-forwards of the classes (\ref{clazz}): \begin{equation} \label{clazz} [\mathbf G(2,E)] \cdot M(c_1,c_2,c_3,c_4,b_1,b_2,f_1,f_2,f_3) \in A^*(\mathcal{G}) \end{equation} Since the classes $c_i, f_k$ in $A^*(\mathcal{G})$ are pull-backs from $\mathbf G(3,\wedge^2 S)$, $I_D$ is generated by the elements $$\pi_*([\mathbf G(2,E)] \cdot M(b_1,b_2)).$$ By the standard relations satisfied by the classes $b_1$ and $b_2$ over $A^*(\mathbf G(4,m))$, it follows that $I_D$ is generated by: $$\pi_*([\mathbf G(2,E)])$$ $$\pi_*([\mathbf G(2,E)]\cdot b_1)$$ $$\pi_*([\mathbf G(2,E)]\cdot b_1^2) \ \ \ \pi_*([\mathbf G(2,E)]\cdot b_2)$$ $$\pi_*([\mathbf G(2,E)]\cdot b_1b_2)$$ $$\pi_*([\mathbf G(2,E)]\cdot b^2_1b_2).$$ Since the class $[\mathbf G(2,E)]$ is determined explicitly in $A^*(\mathcal{G})$, these six push-forwards can be easily computed by hand or by the the Maple package Schubert ([KS]). The first push-forward is: $$\pi_*([\mathbf G(2,E)])=13c_1-2f_1.$$ \begin{lm} \label{killl} The pair $(13c_1-2f_1, c_1(\overline{N}))$ generates $A^1(\mathbf G(3,\wedge^2 S))$. \end{lm} \noindent {\em Proof.} Recall the notation of diagram (\ref{diax}): $$\tau_2: \mathbf P(W_m) \rightarrow Z \subset \mathbf G(3,\wedge^2 S).$$ Let $L=c_1({\mathcal{O}}_{\mathbf P}(-1))$ be a generator of $A^1(\mathbf P(W_m))$. By Lemma 2, $\tau_2^*(c_1)\stackrel{\sim}{=} 4L$. Since $\tau_2^*([D])=\tau_2^*(13c_1-2f_1)=0$, $\tau_2^*(f_1)=26 L$. Therefore the image of $\tau_2^*$ is the subgroup $\mathbb{Z}(2L)$. Since $[D]=13c_1-2f_1$ is not divisible in $A^1(\mathbf G(3,\wedge^2 S))$, $A^1(Z)\stackrel{\sim}{=} \mathbb{Z}$ and $\tau_2^*$ is an isomorphism: $$\tau_2^*: A^1(Z) \stackrel{\sim}{\rightarrow} \mathbb{Z}(2L).$$ By Lemma 3, $\tau_2^*(c_1(N))=2L$. Therefore, $c_1(N)$ generates $A^1(Z)$. It now follows that the pair $(13c_1-2f_1, c_1(\overline{N}))$ generates the group $A^1(\mathbf G(3,\wedge^2 S))$. \qed \vspace{+10pt} \noindent Therefore, $(I_D, c_1(\overline{N}))=(I_D,c_1, f_1).$ By Lemma \ref{killl}, it suffices to compute the five remaining push-forwards modulo the ideal $J=(c_1,f_1)$. The results are (modulo $J$): \begin{enumerate} \item[{}] $\pi_*([\mathbf G(2,E)]\cdot b_1)=0.$ \item[{}] $\pi_*([\mathbf G(2,E)]\cdot b_1^2)=-2f_3.$ \item[{}] $\pi_*([\mathbf G(2,E)]\cdot b_2)=c_3-f_3.$ \item[{}] $\pi_*([\mathbf G(2,E)]\cdot b_1b_2)=(c_2-f_2)^2-4c_4.$ \item[{}] $\pi_*([\mathbf G(2,E)]\cdot b_1^2 b_2)=c_2f_3+f_2c_3.$ \end{enumerate} Hence $(I_D, c_1(\overline{N}))=(c_1,f_1,2c_3, c_3-f_3, (c_2-f_2)^2-4c_4, (c_2-f_2)c_3)$. The ring $A^*(\mathbf G(4,m))$ is freely generated (up to codimension $m-4$) by $c_1,c_2,c_3,c_4$. The ring $A^*(\mathbf G(3,\wedge^2 S))$ has the following presentation (up to codimension $m-4$): $$A^*(\mathbf G(3,\wedge^2 S))\stackrel{\sim}{=} \mathbb{Z}[c_1,c_2,c_3,c_4, f_1,f_2,f_3]/ (t_4,t_5,t_6)$$ where the $t_4,t_5, t_6$ are the Chern classes of the tautological quotient bundle $T$: $0 \rightarrow F \rightarrow \wedge^2 S \rightarrow T \rightarrow 0.$ We find (modulo $J$): \begin{enumerate} \item[{}] $t_4= (c_2-f_2)^2-4c_4$. \item[{}] $t_5= 2f_2f_3-2c_2f_3$. \item[{}] $t_6= f_2(-(c_2-f_2)^2+4c4)+f_3^2-c_3^2$. \end{enumerate} There is a presentation: $A^*(\mathbf G(3,\wedge^2 S))/(I_D,c_1(\overline{N})) \stackrel{\sim}{=}$ \begin{equation}\label{prezz} \mathbb{Z}[c_1,c_2,c_3,c_4,f_1,f_2,f_3]/(I_D, c_1, f_1, t_4,t_5,t_6) \end{equation} (up to codimension $m-4$). Surprisingly, the relations $t_4,t_5,t_6$ are contained in the ideal $(I_D, c_1,f_1)$. By the limit procedure, $$A_{\mathbf{SO}(4)}^*(\text{point})\stackrel{\sim}{=} \mathbb{Z}[c_1,c_2,c_3,c_4, f_2]/(c_1,2c_3,(c_2-f_2)c_3, (c_2-f_2)^2-4c_4).$$ The vector bundles $S$, $F \subset \wedge^2 S$ on the approximation $W_m/\mathbf{SO}(V)$ are easily seen to be obtained from the principal $\mathbf{SO}(V)$-bundle $$W_m \rightarrow W_m/ \mathbf{SO}(V)$$ and the representations $V$, $F \subset \wedge^2 V$ defined in section \ref{rrr}. Define $$x=c_2-f_2.$$ Theorem 1 is proved.
1996-07-22T15:31:03
9607
alg-geom/9607022
en
https://arxiv.org/abs/alg-geom/9607022
[ "alg-geom", "math.AG" ]
alg-geom/9607022
Flavio. Angelini
Flavio Angelini
Ample divisors on the blow up of P^3 at points
AMS-LaTeX, 9 pages
null
null
null
null
We give a condition for certain divisors on the blow up of P^3 at points in general position to be ample. The result extends a theorem of G. Xu on the blow up of the projective plane.
[ { "version": "v1", "created": "Mon, 22 Jul 1996 14:26:14 GMT" } ]
2008-02-03T00:00:00
[ [ "Angelini", "Flavio", "" ] ]
alg-geom
\section{Introduction} In this note we will prove a theorem on divisors on the blow up of $\Bbb{P}^3$ at points which extends a theorem of G. Xu \cite{Xu} on the blow up of $\Bbb{P}^2$. The central idea of the proof works for any dimension and therefore opens the doors to a generalization of the theorem to higher dimension, once one overcomes certain technical difficulties that arise. Basically we will give a new proof of Xu's theorem that works also for $\Bbb{P}^3$ and so in most of this note $n$ will be either 2 or 3. \noindent {\bf Theorem.} Let $n=2$ or $3$. Let $\Bbb{P}^n$ be the projective space over the field of complex numbers. Let $p_1,\dots,p_k$ be k points in $\Bbb{P}^n$ in general position and let $\pi: X \longrightarrow \Bbb{P}^n$ be the blow up of $\Bbb{P}^n$ at $p_1,\dots,p_k$ with exceptional divisors $E_1,\dots,E_k$. Let $H=\pi^* \cal{O}_{\Bbb{P}^n}(1)$. Then, if $d\geq d_0(n)$, the divisor $L=dH- \displaystyle{\sum_{i=1}^k} E_i$ is ample if and only if $L^n >0$, i.e. $d^n>k$, where $d_0(2)=3$, $d_0(3)=5$. \noindent {\bf Remark 1.1.} In the case of $\Bbb{P}^2$ we obtain the same bound as in Xu's, which is sharp. For $\Bbb{P}^3$ we believe, following a conjecture, that the theorem holds for $d\geq3$. Xu's proof is based on an estimate for the self-intersection of moving singular curves in $\Bbb{P}^2$. The same estimate was obtained also by Ein and Lazarsfeld in the context of Seshadri constants on smooth surfaces \cite{EL} and used by K\"uchle to prove the above theorem in the case of a smooth surface \cite{Ku}. The basic idea of the proof comes from an example of R. Miranda regarding ample divisors on a smooth surface with arbitrarily small Seshadri constant \cite[\S5]{L}. The strategy is as follows: we will use the fact that ampleness is an open condition to reduce to the case when the points are part of the base locus of a $(n-1)$-dimensional general linear system of hypersurfaces of degree $d$. The corresponding line bundle $L$ on the blow up at the entire base locus is nef and this will imply that the divisor obtained by blowing down some of the exceptional loci is ample. For this last step we will use in a determinant way the fact that the fibres of the morphism to $\Bbb{P}^{n-1}$ determined by $L$ are irreducible, under the stated conditions on $d$. We will prove this fact in Lemma 2.1 which concerns curves in $\Bbb{P}^3$ which are complete intersection. The problem is to give a lower bound for the codimension, in the space of such curves, of the space of reducible ones. We emphasize that it is this technical lemma that gives, apart from the lower bound on $d$, that may be not sharp, but reasonable, the restriction for the dimension of the projective space for which the theorem holds. It is actually possible to prove Lemma 2.1, and therefore the theorem, also in the cases $n=4$ and $5$, in a very similar way and we will spend a couple of words about it at the end of the proof of the lemma. The rest of the proof works for any dimension and therefore we will present it as much as possible in its generality in section 3. This note is part of my Ph. D. thesis at UCLA, and I would like to thank Rob Lazarsfeld for his guidance and encouragement. \section{Preliminary material and lemmas} Let $\Bbb{P}^N$ be the projective space parametrizing hypersurfaces in $\Bbb{P}^n$ of degree $d$. So $N=N(d,n)=\binom{n+d}{n} -1$. If $F$ is such a hypersurface, we will denote by $\left[ F \right]$ the corresponding point in $\Bbb{P}^N$. Recall that there is an action of $PGL(N)$ on $\Bbb{P}^N$. If $\sigma$ is an element of $PGL(N)$ and $\Lambda$ is a linear subspace of $\Bbb{P}^N$, we will denote by $\Lambda^{\sigma}$ the linear subspace obtained by letting $\sigma$ act on $\Lambda$. We will say that a property holds for a general linear subspace $\Lambda$ of $\Bbb{P}^N$ if it holds for $\Lambda^{\sigma}$ for $\sigma$ outside a union (possibly countable) of proper subvarieties of $PGL(N)$. We will denote by $\Sigma_{d}$ the locus of singular hypersurfaces which is an irreducible subvariety of $\Bbb{P}^N$ of codimension one. We will also be dealing with curves which are complete intersection of hypersurfaces of same degree $d$ in $\Bbb{P}^3$. These are parametrized by an open subset of the Grassmannian $Gr(\Bbb{P}^1,\Bbb{P}^N)$ (of course for $n=2$ this is just $\Bbb{P}^{N(d,2)}$). We will denote by $\left[l_C\right] \in Gr(\Bbb{P}^1,\Bbb{P}^N)$ the point corresponding to a curve $C$. Also we have $dimGr(\Bbb{P}^k,\Bbb{P}^r)=(k+1)(r-k)$. Let now $$\begin{array}{rl} NL_{d,3}=&\lbrace \text{smooth surfaces}\phantom{.} S\subset\Bbb{P}^3 \phantom{.}\text{of degree}\phantom{.} d \\ &\phantom{.}: Pic(S)\phantom{.} \text{is not generated by the hyperplane class} \rbrace. \end{array}$$ $NL_{d,3}$ is called the {\em Noether-Lefschetz locus} and may be viewed as a subset of $\Bbb{P}^{N(d,3)}$. This locus is pretty well understood, at least as far as we are concerned here. $NL_{d,3}$ is a countable union of quasi-projective algebraic varieties. The Noether-Lefschetz theorem asserts that, for $d\geq 4$, the general surface of degree $d$ in $\Bbb{P}^3$ has Picard group generated by the hyperplane section, i.e. is not contained in $NL_{d,3}$. In other words the theorem says that the codimension of all the irreducible components $N_i$ of $NL_{d,3}$ is at least one. What we will need is an explicit Noether-Lefschetz theorem (see \cite{Gr} for a nice proof of it) giving a precise bound for the codimension of any of the $N_i$. \noindent {\bf Theorem (Green).} For $d\geq 4$, the codimension of any irreducible component of $NL_{d,3}$ in $\Bbb{P}^N$ is at least $(d-3)$. The main technical lemma we need is: \noindent {\bf Lemma 2.1.} Let $n=2,3$. Let $\Bbb{P}^N$ be the projective space parametrizing hypersurfaces of degree $d$ in $\Bbb{P}^n$ and assume $d\geq d_0(n)$, with $d_0(2)=3$ and $d_0(3)=5$. Let $\Lambda$ be a general linear subspace of $\Bbb{P}^N$ of dimension $(n-1)$ whose base locus consists of $d^n$ points. Then, for every $(n-1)$ linearly independent elements $\left[ F_1 \right] , \dots, \left[ F_{n-1} \right]$ of $\Lambda$, the intersection $F_1 \cap \dots \cap F_{n-1}$ is a curve and is irreducible. In the case $n=2$ this is just saying that every element of a general pencil of curves of degree $d$ is irreducible for $d\geq3$. \noindent {\bf Remark 2.1.} Lemma 2.1, and therefore the theorem, would be proven for any dimension if one found a good bound for the codimension, in the space of curves in $\Bbb{P}^n$ which are a complete intersection of hypersurfaces of degree $d$, of the space of reducible ones. It is easy to conjecture that this codimension should be at least $(n-1)(d-1)$, being this the codimension of such curves which have a line as a component. This conjecture would imply Lemma 2.1 with $d_0(n)=3$ and for any $n$. \noindent {\it Proof.} The proof in the case $n=2$ is rather easy. The set $\Sigma_d$ of singular curves of degree $d$ in $\Bbb{P}^2$ is an irreducible subvariety of codimension one. The reducible curves are union of closed subvarieties all lying in $\Sigma$. It is then enough to notice that, for $d\geq3$, there are irreducible singular curves to establish that the codimension of the reducible ones is at least two and therefore conclude that a general pencil does not contain reducible curves. For $n=3$ we take $d\geq5$ and we need to prove Lemma 2.1 for a general plane $\Lambda\subset \Bbb{P}^N$ with $N=\binom{d+3}{3}-1$. A curve $C$ which is a complete intersection of two elements of $\Lambda$ corresponds to a line $l_C \subset\Lambda$, i.e. to an element $\left[l_C\right] \in Gr(\Bbb{P}^1,\Lambda) \subset Gr(\Bbb{P}^1,\Bbb{P}^N)$. First we have the following: \noindent {\bf Claim 2.1.} For any $l_C$ in $\Lambda$, we can find a {\em smooth} surface $S$ with $\left[ S\right] \in l_C$ such that $\left[ S \right]$ is not in $NL_{d,3}$. For this we need the following sublemma: \noindent {\bf Lemma 2.2.} For $d\geq5$, the intersection of $\Lambda \subset\Bbb{P}^N$ with the Noether-Lefschetz locus consists of at most countably many points. This is to say that there are at most countably many smooth surfaces $S$ with $\left[ S\right] \in \Lambda$ and with $Pic(S)$ not generated by the hyperplane section. \noindent {\it Proof.} We use here Kleiman's Transversality Theorem \cite[Thm. III.10.8] {H} for the action of $PGL(N)$ on $\Bbb{P}^N$. We will apply the Theorem to a plane $\Lambda \subset \Bbb{P}^N$ and the closure $\overline{N_i}$ of one irreducible component $N_i$ of $NL_{d,3}$. The Theorem says that $\Lambda^{\sigma} \cap \overline{N_i}$ is either empty or of dimension $$dim\Lambda - codim\overline{N_i}$$ for $\sigma$ in a non-empty open subset $V_i \subset PGL(N)$. So, by the explicit Noether-Lefschetz theorem and for $d\geq5$, we have $$dim(\Lambda^{\sigma} \cap \overline{N_i}) \leq 2-(d-3) \leq 0$$ for $\sigma \in V_i$. This means that the intersection of a general plane with a component of $NL_{d,3}$ consists at most of a finite number of points. Therefore, for $\sigma$ outside a possibly countable union of closed proper subvarieties, the intersection of $\Lambda^{\sigma}$ with $NL_{d,3}$ consists of at most countably many points. $\hspace{1cm}\Box$ Also observe that the intersection of $\Lambda$ with $\Sigma_{d}$ does not contain any line $l_C \subset \Lambda$. In other words, for any curve $C$ which is a complete intersection of elements of $\Lambda$ there exist at most finitely many surfaces $S$ with $\left[ S\right] \in l_C$ that are singular. This is because $\Sigma_{d}$ is an irreducible subvariety of codimension one and high degree of $\Bbb{P}^N$, and hence, for a general $\Lambda$, $\Sigma_d \cap \Lambda$ is an irreducible curve of degree strictly greater than one. \noindent {\it Proof of Claim 2.1.} For any $l_C$ in $\Lambda$, by Lemma 2.2 there are at most countably many surfaces $S$ with $\left[ S\right] \in l_C$ and $Pic(S)$ not isomorphic to $\Bbb{Z}$ and by the observation above there are at most finitely many singular ones. It is then possible to find a smooth $S$ with $Pic(S)$ isomorphic to $\Bbb{Z}$. $\hspace{1cm}\Box$ Now we can prove Lemma 2.1 for $n=3$. \noindent {\it Proof of Lemma 2.1.} Let $C$ be any curve which is complete intersection of elements of $\Lambda$. We need to prove that $C$ is irreducible. Choose a surface $S$ as in Claim 2.1. Pick another surface $T$ with $\left[ T \right]$ in $l_C$ so that $C=S\cap T$. If $C$ were reducible, say $C=\cup C_i$, then any $C_i \subset S$ would be a complete intersection $S \cap T_i$, with $T_i$ a hypersurface in $\Bbb{P}^3$ of degree less than $d$, since $Pic(S)$ is generated by the hyperplane section. But then $$C=S \cap T=\cup(S \cap T_i)=S \cap (\cup T_i)$$ This means that $\left[ \cup T_i \right] \in l_C \subset \Lambda$. But $\Lambda$ misses reducible surfaces, so $C$ has to be irreducible. $\hspace{1cm}\Box$ \noindent {\bf Remark 2.2.} As mentioned in the introduction, it is possible to prove Lemma 2.1 in the case $n=4$ and $5$ along the same line as for $n=3$, using a generalized explicit Noether- Lefschetz theorem, due to S. Kim \cite{Kim}, regarding smooth surfaces which are complete intersection of hypersurfaces \cite{A}. The proof does not work in higher dimension due to the fact that the Noether- Lefschetz theorem does not give any information about singular surfaces and it is not possible anymore to ensure the existence of a smooth surface containing the curve. Another tool for the proof of the theorem is the fact that ampleness is an open condition. We will state this very well known fact in the form of: \noindent {\bf Proposition 2.1.} Let $\cal{L} \longrightarrow T$ be a flat family of line bundles over a flat family $\cal{X} \stackrel{f}\longrightarrow T$ of projective varieties of dimension $n$. Then the set $$\lbrace t\in T\phantom{.}:\phantom{.} \cal{L}_t=\cal{L}_{|_{X_t}}\text{is ample on}\phantom{.}X_t \rbrace$$ is open in $T$. (See \cite{A} for a proof of it). \section{Proof of Theorem} The central idea of the proof works for any dimension and we will present the argument as much as possible in its generality. The restriction on the dimension comes uniquely from Lemma 2.1, to ensure the irreducibility of the fibers of the morphism $\mu$ below. We are given $k$ points in general position in $\Bbb{P}^n$ and we need to prove that the divisor $dH-\displaystyle{\sum_{i=1}^k} E_i$ is ample when $d^n>k$ for $d\geq d_0(n)$. Clearly this condition is necessary. We consider a general linear system $\Lambda$ of hypersurfaces of degree $d$ and dimension $(n-1)$ and we let $p_1^\prime ,\dots , p_{d^n} ^\prime$ be the base locus of $\Lambda$. Let $\pi ^\prime : X^\prime \longrightarrow \Bbb{P}^n$ be the blow up at $k$ of these points $p_1 ^\prime , \dots , p_k ^\prime$, $H^\prime ={\pi ^\prime}^* \cal{O}_{\Bbb{P}^n}(1)$ and $E_i ^\prime, \dots ,E_k ^\prime$ the exceptional divisors. We will prove that $dH^\prime- \displaystyle{\sum_{i=1}^k} E_i^\prime$ is ample and the theorem will follow from Proposition 2.1. To do so we consider the blow up $\nu : Z \longrightarrow \Bbb{P}^n$ at the whole base locus $p_1 ^\prime, \dots , p_{d^n} ^\prime$ and set $H=\nu^* \cal{O}_{\Bbb{P}^n}(1)$, $E_i$ for $i=1, \dots ,k$, $F_i$ for $i=k+1, \dots , d^n$ the ecxeptional divisors of $\nu$ and $L=dH-\displaystyle{\sum_{i=1}^k} E_i -\displaystyle{\sum_{i=k+1}^{d^n}} F_i$. We have the following diagram: $$\begin{array}{rcl} Z\phantom{.} &\stackrel{\delta}\longrightarrow&X^\prime \\ &\stackrel{\nu}\searrow\phantom{.}\stackrel{\pi ^\prime}\swarrow& \\ &\phantom{.}\Bbb{P}^n& \end{array}$$ where $\delta$ is the blow down of $F_{k+1}, \dots ,F_{d^n}$. Now $L$ is globally generated and therefore is nef. We will show that, for $d\geq d_0(n)$, this implies that $dH^\prime-\displaystyle{\sum_{i=1}^k} E_i^\prime$ is ample. Now we have a morphism $\mu :Z \longrightarrow \Bbb{P}^{n-1}$ whose fibers are curves which are complete intersection of elements of $\Lambda$. Moreover, if $C$ is a curve, $$L\cdot C=0 \Longleftrightarrow \mu(C)\phantom{..} \mbox{\text{is a set of points}}.$$ By Lemma 2.1, for $n=2,3$ and $d\geq d_0(n)$, by taking $\Lambda$ sufficiently general, we may arrange for these fibers to be all irreducible. What we need is the following: \begin{align} \tag{3.1} &(L+\displaystyle{\sum_{i=k+1}^{d^n}} F_i)^m \cdot Y_m \geq0 \\ \tag{3.2} &(L+\displaystyle{\sum_{i=k+1}^{d^n}} F_i)^m \cdot Y_m =0 \Longleftrightarrow Y_m \subset F_i \phantom{...} \mbox{\text{for one}}\phantom{.}i=k+1, \dots ,d^n, \end{align} where $Y_m$ is any irreducible subvariety of dimension $m$ of $Z$, for any $0<m<n$. Notice that: $$L+\displaystyle{\sum_{i=k+1}^{d^n}} F_i=dH-\displaystyle{\sum_{i=1}^k} E_i=\delta^*(dH^\prime-\displaystyle{\sum_{i=1}^k} E_i^\prime).$$ (3.1) and (3.2) prove, by Nakai's criterion and the projection formula, that $dH^\prime-\displaystyle{\sum_{i=1}^k} E_i^\prime$ is ample on $X^\prime$. To simplify notation let $L^\prime=L+\displaystyle{\sum_{i=k+1}^{d^n}} F_i$. From here on we restrict to $n=2$ and 3 because we will use Lemma 2.1. If Lemma 2.1 were proven for any $n$ we would proceed by induction on $m$, using the same arguments. Here we just have to check (3.1) and (3.2) for $m=1$ in the case $n=2$ and for $m=1$ and 2 in the case $n=3$. Let $m=1$ so that $Y_1$ is a curve. If $Y_1 \subseteq F_i$ for one $i$, then, being $L^\prime$ trivial on $F_i$, $L^\prime \cdot Y_1 =0$ and, if $Y_1\not\subseteq F_i$ for all $i$, then, being $L \cdot Y_1 \geq 0$ ($L$ is nef) and $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot Y_1\geq 0$, $L^\prime \cdot Y_1 \geq 0$. This proves (3.1) and the easy direction of (3.2). For the other direction suppose $L^\prime \cdot Y_1=0$ but $Y_1 \not\subseteq F_i$ for all $i$. But then, being $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot Y_1 \geq 0$, $L \cdot Y_1$ has to be zero and therefore $Y_1$ is a component of a fiber of $\mu$. Since every fiber is irreducible, $Y_1$ is exactly a fiber and so it meets all the exceptional divisors of $\nu$. Therefore $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot Y_1$ is strictly positive which is a contradiction. In the case $n=3$ we need also to consider any irreducible subvariety $Y_m$ of $Z$ of dimension $m=2$. As before, if $Y_m \subseteq F_i$ for some $i$, then $L^\prime \cdot Y_m =0$. If $Y_m \not\subseteq F_i$ for all $i$, then we write: $${L^\prime}^m \cdot Y_m=L \cdot {L^\prime}^{m-1}\cdot Y_m +\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot {L^\prime}^{m-1} \cdot Y_m.$$ Now $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot Y_m$ has dimension strictly less than $m$ (or is empty) and $L \cdot Y_m$ can be represented by a cycle of dimension strictly less than $m$ (or empty) because $L$ is globally generated. So both $L \cdot {L^\prime}^{m-1}\cdot Y_m$ and $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot {L^\prime}^{m-1} \cdot Y_m$ are greater or equal than zero by the previous step and so is ${L^\prime}^m \cdot Y_m$. For the remaining direction of (3.2) suppose ${L^\prime}^m \cdot Y_m=0$ while $Y_m \not\subseteq F_i$ all $i$. Then, since as before $\displaystyle{\sum_{i=k+1}^{d^n}} F_i \cdot {L^\prime}^{m-1} \cdot Y_m \geq 0$, we have $L\cdot {L^\prime}^{m-1} \cdot Y_m=0$. By the previous step, this implies that $L \cdot Y_m \subseteq F_i$ for some $i$. We claim that this is a contradiction, i.e. we claim that, if $Y_m \not\subseteq F_i$ for any $i$, then $L \cdot Y_m \not\subseteq F_i$ for any $i$. Indeed, if $Y_m \not\subseteq F_i$ for any $i$, then $\nu_*(Y_m) \subseteq \Bbb{P}^3$ has dimension 2. Consider a divisor $D \in | \Lambda|$. By Bezout's Theorem $dim(\nu_*(Y_m) \cdot D) \geq 1$. Also $$dim(\nu_*(Y_m) \cdot D)=dim(\nu_*(Y_m \cdot \nu^*D))= dim(\nu_*(Y_m \cdot L)).$$ So $dim(\nu_*(Y_m\cdot L))\geq 1$ and therefore $Y_m \cdot L\not\subseteq F_i$ for any $i$. $\hspace{1cm}\Box$
1997-09-22T16:38:52
9607
alg-geom/9607020
en
https://arxiv.org/abs/alg-geom/9607020
[ "alg-geom", "math.AG" ]
alg-geom/9607020
Paul Bressler
P.Bressler, M.Saito, B.Youssin
Filtered Perverse Complexes
AMSLaTeX v 1.1. This version is a major revision. With the new co-author (M.Saito) it contains substantially new results, improvements and corrections
null
null
null
null
We introduce the notion of filtered perversity of a filtered differential complex on a complex analytic manifold $X$, without any assumptions of coherence, with the purpose of studying the connection between the pure Hodge modules and the \lt-complexes. We show that if a filtered differential complex $(\cM^\bullet,F_\bullet)$ is filtered perverse then $\aDR(\cM^\bullet,F_\bullet)$ is isomorphic to a filtered $\cD$-module; a coherence assumption on the cohomology of $(\cM^\bullet,F_\bullet)$ implies that, in addition, this $\cD$-module is holonomic. We show the converse: the de Rham complex of a holonomic Cohen-Macaulay filtered $\cD$-module is filtered perverse.
[ { "version": "v1", "created": "Fri, 19 Jul 1996 15:38:35 GMT" }, { "version": "v2", "created": "Sun, 22 Sep 1996 18:48:19 GMT" }, { "version": "v3", "created": "Wed, 9 Oct 1996 14:36:13 GMT" }, { "version": "v4", "created": "Mon, 22 Sep 1997 14:38:51 GMT" } ]
2008-02-03T00:00:00
[ [ "Bressler", "P.", "" ], [ "Saito", "M.", "" ], [ "Youssin", "B.", "" ] ]
alg-geom
\section{Introduction} \subsection{Cheeger---Goresky---MacPherson conjectures} J.~Cheeger, M.~Goresky and R.~MacPherson \cite{CGM} conjectured some fifteen years ago that the intersection cohomology of a singular complex projective algebraic variety is naturally isomorphic to its $L^2$ cohomology and the K\"ahler package holds for them. Their motivation was as follows. The intersection cohomology was discovered by M.~Goresky and R.~MacPherson \cite{GM1}, \cite{GM2} as an invariant of stratified spaces which for complex algebraic varieties might serve as a replacement of the usual cohomology: it had some properties that the usual cohomology of smooth projective varieties possessed but the usual cohomology of singular projective varieties did not. One of such properties was Poincar\'e duality which is a part of the ``K\"ahler package'' of properties that hold in the smooth case. At the same time, J.~Cheeger discovered that the $L^2$ cohomology groups of varieties with conical singularities have properties similar to those of intersection cohomology, and he proved in this case the Hodge---de Rham isomorphism between the $L^2$ cohomology that he defined and studied, and the intersection cohomology \cite{C}. The hope that underlied these conjectures was that it would be possible to use the $L^2$ K\"ahler methods to prove the K\"ahler package for intersection cohomology similarly to the way the K\"ahler package was proved for the usual cohomology of complex projective manifolds. The most important part of the K\"ahler package is the $(p,q)$-decomposition in the cohomology groups (the ``Hodge structure''). The definition of $L^2$ cohomology involves a metric (Riemannian or K\"ahler) defined almost everywhere on the variety (e.g.\ on its nonsingular part). The most important metric comes from a projective embedding of the variety and is induced by the Fubini---Studi metric on the projective space. (The $L^2$ cohomology is independent of the choice of the imbedding.) The isomorphism with intersection cohomology is known in case of surfaces \cite{HP}, \cite{Nag1} and in case of isolated singularities of any dimension both for Fubini---Studi metric \cite{O2}, \cite{O2a} and for a different, complete metric, introduced by L.~Saper, which is defined on the nonsingular part of the variety and blows up near the singularities \cite{Sap}. The $(p,q)$-decomposition is known for the case of Fubini---Studi metrics only in cases of dimension two \cite{Nag2} (except for the middle degree cohomology groups) while a classical result of Andreotti---Vesentini implies the $(p,q)$-decomposition for {\em any\/} complete metric. The general case is still open, despite the announcement of T.~Ohsawa \cite{O3}. In the meantime the second author \cite{S1}, \cite{S2} developed a theory of polarizable Hodge modules which implied the K\"ahler package for the intersection cohomology. His main tool was the theory of ${\cal D}$-modules and his methods were essentially algebraic, reducing the intersection cohomology to the intersection cohomology of a curve with coefficients in a polarised variation of Hodge structure \cite{Z1}. \subsection{The comparison between the Hodge structures} Assuming that the Cheeger---Goresky---MacPherson conjectures are true, one is faced with the question of comparison between the two Hodge structures on the intersection cohomology: one induced by the isomorphism with {$L^2$} cohomology, the other coming from the theory of polarised Hodge modules. In fact, different metrics give different $L^2$ cohomology theories and hence, pose different comparison problems. In case of isolated singularities, S.~Zucker \cite{Z} proved the coincidence between the Hodge structures coming from polarized Hodge modules and from $L^2$ cohomology with respect to the Saper metric (or arithmetic quotient metrics similar to it). Some partial results are also known in case of Fubini---Studi metric, see \cite{Z} and \cite{Nag2}. It is interesting to note that the original purpose of the conjectures was to construct the Hodge structure on the intersection cohomology. The $L^2$ methods, however, turned out to be so difficult that the Hodge structure was constructed by different, algebraic methods and now we are faced with the problem of comparison between the two Hodge structures. \subsection{The local comparison problem} A major component of a polarizable Hodge module is a regular holonomic ${\cal D}$-module $M$ with a good filtration $F_\bullet$. Suppose $M$ corresponds to the intersection cohomology complex of a complex projective subvariety $Z$; the correspondence is given by taking the de Rham complex $\operatorname{DR}(M)$ of $M$ (so that $\operatorname{DR}(M)$ is isomorphic to the intersection cohomology complex of $Z$). Then the filtration $F_\bullet$ induces a filtration on $\operatorname{DR}(M)$ which yields the Hodge structure on the intersection cohomology. The complex $\operatorname{DR}(M,F_\bullet)$ is a filtered differential complex~\cite{S1}: a complex of sheaves which are modules over the sheaf of analytic functions and the differentials are differential operators. This filtered differential complex completely determines $(M,F_\bullet)$ as there is an inverse functor $\operatorname{DR}^{-1}$ \cite{S1}. If the metric used in the construction of the {$L^2$} complex (it is a K\"ahler metric on the nonsingular part of $Z$) is bounded below with respect to Fubini---Studi metric, the {$L^2$} complex is a filtered differential complex (see, e.g.,~\ref{subsec:l2} below). The {\em local comparison problem\/} is as follows: is it true that the de Rham complex of the Hodge module is isomorphic to the {$L^2$} complex in the derived category of filtered differential complexes? The intersection cohomology can be taken with coefficients in a local system defined on the non-singular part of $Z$ or a Zariski-open subset of it. If this local system underlies a polarized variation of Hodge structures then a corresponding polarized Hodge module can be constructed and the intersection cohomology with coefficients in this local system has a Hodge structure. On the other hand, the $L^2$ cohomology can be taken with coefficients in the same polarized variation, and we can ask the same local comparison question in this situation. \subsection{Weak filtered perversity} A way to approach this problem is to try to identify the properties of a filtered differential complex $({\cal M}^\bullet,F_\bullet)$ which would imply that $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is isomorphic to the filtered ${\cal D}$-module which underlies a polarized Hodge module. In general, $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is a {\em complex\/} of filtered ${\cal D}$-modules; in this paper we study properties of $({\cal M}^\bullet,F_\bullet)$ which imply that this complex is isomorphic to one filtered ${\cal D}$-module in the filtered derived category. We call these properties {\em weak filtered perversity\/} (see Definition~\ref{cond:wfp}); it means that, first, the complex $({\cal M}^\bullet,F_\bullet)$ is locally trivial along the strata --- in a certain filtered sense --- with respect to some analytic stratification, and second, it satisfies certain local filtered cohomology vanishing which is similar to the local cohomology vanishing of perverse sheaves. No coherence assumption is being made on $({\cal M}^\bullet,F_\bullet)$. In case $({\cal M}^\bullet,F_\bullet)$ is the $L^2$\ complex, the cohomology that must vanish, turn out to be a version of the $L^2$-$\overline{\partial}$-cohomology, see~\S\ref{rem:bound-cond} below for the discussion. \subsection{The main results} We show (see Theorem~\ref{thm:main}) that if $({\cal M}^\bullet,F_\bullet)$ is weakly filtered perverse then, indeed, $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is isomorphic to a filtered ${\cal D}$-module. We show the converse (see Theorem~\ref{thm:converse}): if $(M^\bullet,F_\bullet)$ is a coherent filtered ${\cal D}$-module which is Cohen-Macaulay (i.e., its dual in the filtered sense is also a complex of filtered ${\cal D}$-modules isomorphic to one filtered ${\cal D}$-module) then $\operatorname{DR}(M,F_\bullet)$ is weakly filtered perverse. We show (see Proposition~\ref{cor:holon}) that a coherence assumption together with filtered perversity of $({\cal M}^\bullet,F_\bullet)$ implies that the filtered ${\cal D}$-module $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is holonomic. \subsection{Plan of the paper} In Section~\ref{sec:fdm-dc} we review the necessary background material from~\cite{S1}. In Section~\ref{sec:fil.perv} we introduce the notion of weak filtered perversity. In Sections \ref{sec:fpc-to-dm} and~\ref{sec:coh.case} we prove the results listed above. In Section~\ref{sec:appl} we give a modest application: we strengthen the results of \cite{KK2} and \cite{S1} and show (in the situation of \cite{KK2}) that filtered perversity of the $L^2$ complex implies the local filtered isomorphism (in the sense of derived category) between the $L^2$ complex and the de Rham complex of the ${\cal D}$-module that underlies the corresponding pure Hodge module. \subsection{Acknowledgements} It is our pleasant duty to express our heartful thanks to all people who helped us with their advice and helpful discussions: Daniel Barlet, Alexander Beilinson, Joseph Bernstein, Jean-Luc Brylinski, Michael Kapranov, Masaki Kashiwara, David Kazhdan, Takeo Ohsawa, Claude Sabbah. \section{Filtered ${\cal D}$-modules and differential complexes} \label{sec:fdm-dc} In this section we make a brief survey of the necessary parts of \cite{S1}. \subsection{General notation} Throughout this paper $X$ will denote a complex manifold, ${\cal O}_X$ the sheaf of holomorphic functions, $\omega_X$ the canonical sheaf of $X$, ${\cal D}_X$ the sheaf of differential operators. Unless specified otherwise, a ${\cal D}_X$-module will always refer to a sheaf of {\em right} modules over ${\cal D}_X$. For a complex of sheaves ${\cal F}^\bullet$, we shall denote by $H^j{\cal F}^\bullet$ its {\em sheaf\/} cohomology, and by $H^j_{\{x\}}{\cal F}^\bullet$ its (hyper)cohomology with supports in a one-point set $\{x\}$. \subsection{Filtered ${\cal D}$-modules} Recall that ${\cal D}_X$ is a filtered ring when equipped with the filtration $F_\bullet{\cal D}_X$, where $F_p{\cal D}_X$ is the ${\cal O}_X$-module of ${\cal D}_X$ of operators of order at most $p$. A filtered ${\cal D}_X$-module is a pair $(M,F_\bullet)$ consisting of a ${\cal D}_X$-module $M$ and a filtration $F_\bullet M$ of $M$ by ${\cal O}_X$ submodules compatible with the action of ${\cal D}_X$ and the filtration on the latter. We refer the reader to~\cite{S1}, \S2.1 for the precise definition of the category of filtered ${\cal D}_X$-modules and its derived category, in various flavors; what is important for us now, is that the derived category of filtered ${\cal D}_X$-modules is isomorphic to the derived category of the category whose objects are filtered ${\cal O}_X$-modules and whose morphisms are differential operators that agree with the filtration in a certain way. This equivalence is given by the two functors, $\operatorname{DR}_X$ and $\operatorname{DR}^{-1}_X$ (loc.~cit.) which act as follows. \subsection{The de Rham functor $\protect\operatorname{DR}_X$} For a filtered ${\cal D}_X$-module $(M,F_\bullet)$, the filtered differential complex $\operatorname{DR}_X(M,F_\bullet)$ is the usual de Rham complex of $M$, given by \[ \operatorname{DR}_XM= M\otimes_{{\cal D}_X}\left({\cal D}_X\otimes_{{\cal O}_X} \textstyle\bigwedge^{-\bullet}\Theta_X\right) =M\otimes_{{\cal O}_X}\textstyle\bigwedge^{-\bullet}\Theta_X \] where $\Theta_X$ is the tangent sheaf to $X$, $\bigwedge^{-\bullet}\Theta_X$ its exterior algebra with $p$-th exterior power placed in degree $-p$, the differential is given by \begin{equation*} \begin{split} {d}(P\otimes\xi_1\wedge\ldots\wedge\xi_p) & =\sum_{i=1}^{p} (-1)^{i-1}P\xi_i\otimes\xi_1\wedge\ldots\wedge\widehat{\xi_i} \wedge\ldots\wedge\xi_p \\ & +\sum_{1\leq i < j\leq p}(-1)^{i+j}P\otimes \lbrack\xi_i,\xi_j\rbrack\wedge\xi_1\wedge\ldots\wedge\widehat{\xi_i} \wedge\ldots\wedge\widehat{\xi_j}\wedge\ldots\wedge\xi_p \end{split} \end{equation*} (it corresponds to the differential in ${\cal D}_X\otimes_{{\cal O}_X}\bigwedge^{-\bullet}\Theta_X$ which makes it the standard Koszul resolution of ${\cal O}_X$ as a ${\cal D}_X$-module), and the filtration on $\operatorname{DR}_XM$ is given by \[ F_p\left(M\otimes_{{\cal O}_X}\textstyle\bigwedge^{-i}\Theta_X\right)= F_{p+i}M\otimes_{{\cal O}_X}\textstyle\bigwedge^{-i}\Theta_X\ . \] For a complex of filtered ${\cal D}_X$-modules $(M^\bullet,F_\bullet)$, the filtered differential complex $\operatorname{DR}_X(M^\bullet,F_\bullet)$ is the total complex of $\operatorname{DR}_X(M^q,F_\bullet)$ for all $q$. (Note that what we denote by $\operatorname{DR}_X$, was denoted by $\widetilde\operatorname{DR}$ in~\cite{S1}.) \subsection{The inverse de Rham functor $\protect\operatorname{DR}^{-1}_X$} For a filtered differential complex $({\cal M}^\bullet,F_\bullet)$, the complex of filtered ${\cal D}_X$-modules $\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ is described as the complex of differential operators from ${\cal O}_X$ into $({\cal M}^\bullet,F_\bullet)$ with the obvious differential and filtration. The action of ${\cal D}_X$, i.~e., differential operators ${\cal O}_X\to{\cal O}_X$, is by composition. The individual terms of this complex are simply ${\cal M}^j\otimes_{{\cal O}_X}{\cal D}_X$. The two functors $\operatorname{DR}_X$ and $\operatorname{DR}^{-1}_X$ are inverse to each other in the derived categories. \subsection{Duality} \label{subsec:duality} For a bounded complex of filtered ${\cal D}_X$-modules $(M^\bullet,F_\bullet)$, its dual ${\Bbb D}(M,F_\bullet)$ is is another complex of filtered ${\cal D}_X$-modules defined (\cite{S1}, 2.4.3) in such way that it agrees with various other duality functors, as follows. There is a duality functor (also denoted by ${\Bbb D}$) on filtered differential complexes which on an individual ${\cal O}_X$-module $L$ is defined as ${\Bbb D}(L)=\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L,\omega_X[n])$ where $n=\dim X$ (\cite{S1}, 2.4.11), and such that in the appropriate derived categories the functors ${\Bbb D}\circ\operatorname{DR}_X$ and $\operatorname{DR}_X\circ{\Bbb D}$ are isomorphic. For a filtered differential complex $({\cal M}^\bullet,F_\bullet)$ we have \begin{equation} \label{eqn:D.vs.Gr} \operatorname{Gr}^F_\bullet{\Bbb D}({\cal M}^\bullet,F_\bullet)\overset{\sim}{=} \operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(\operatorname{Gr}^F_\bullet{\cal M}^\bullet,\omega_X[n])\ . \end{equation} In case $(M^\bullet,F_\bullet)$ is a complex of filtered coherent ${\cal D}_X$-modules (i.e., coherent ${\cal D}_X$-modules with good filtrations), the complex of ${\cal D}_X$-modules which underlies ${\Bbb D}(M^\bullet,F_\bullet)$, is the usual dual of $M^\bullet$. In addition, under the same assumptions ${\Bbb D}\bbD(M^\bullet,F_\bullet)\overset{\sim}{=}(M^\bullet,F_\bullet)$. \subsection{Restriction to a noncharacteristic submanifold} Let $Y$ be a smooth submanifold of codimension $d$ in $X$; denote the embedding $i:Y\hookrightarrow X$ and $\omega_{Y/X}=\omega_Y\otimes_{{\cal O}_X}\omega_X^{-1}$. We say that a bounded complex of filtered ${\cal D}_X$-modules $(M^\bullet,F_\bullet)$ is {\em weakly noncharacteristic\/} with respect to $Y$ (or $Y$ with respect to $(M^\bullet,F_\bullet)$) if it satisfies the property \begin{equation} \label{eqn:tor-nonchar} {\underline{\operatorname{Tor}}}^{{\cal O}_X}_k(H^j(\operatorname{Gr}^F_p M^\bullet),{\cal O}_Y) = 0 \text{ \ \ for all $k\ne0$, $j$ and $p$.} \end{equation} Under this assumption, the noncharacteristic restriction $(M^\bullet,F_\bullet)_Y$ is defined as follows: \[ (M^\bullet,F_\bullet)_Y= (M^\bullet,F_\bullet)\otimes^{\Bbb L}_{{\cal D}_X}({\cal D}_{X\leftarrow Y},F_\bullet) \] where ${\cal D}_{X\leftarrow Y}={\cal D}_X\otimes_{{\cal O}_X}\omega_{Y/X}$ is the usual $({\cal D}_X,{\cal D}_Y)$-bimodule with the filtration $F_p{\cal D}_{X\leftarrow Y}= F_{p-d}{\cal D}_X\otimes_{{\cal O}_X}\omega_{Y/X}$; the restriction $(M^\bullet,F_\bullet)_Y$ thus defined, is a complex of right ${\cal D}_Y$-modules. As a complex of ${\cal O}_Y$-modules, it can be described as $(M^\bullet)_Y=M^\bullet\otimes^{\Bbb L}_{{\cal O}_X}\omega_{Y/X}$ with the filtration $F_p(M^\bullet)_Y=F_{p-d}M^\bullet\otimes^{\Bbb L}_{{\cal O}_X}\omega_{Y/X}$. We have \begin{equation} \label{eqn:HGr(restr)} H^j\operatorname{Gr}^F_p\left((M^\bullet,F_\bullet)_Y\right)\overset{\sim}{=} H^j(\operatorname{Gr}^F_{p-d} M^\bullet)\otimes_{{\cal O}_X}\omega_{Y/X}\ . \end{equation} Suppose that $(M^\bullet,F_\bullet)$ has the property that the complex $\operatorname{Gr}^F_\bullet M^\bullet$ has bounded $\operatorname{Gr}^F_\bullet{\cal D}_X$-coherent cohomology; in such case we say that $(M^\bullet,F_\bullet)$ is {\em noncharacteristic\/} with respect to $Y$ if, first, \eqref{eqn:tor-nonchar} is satisfied, and second, $\operatorname{Gr}^F_\bullet (M^\bullet)_Y$ also has bounded $\operatorname{Gr}^F_\bullet{\cal D}_Y$-coherent cohomology. In the particular case when the complex $(M^\bullet,F_\bullet)$ is actually a filtered coherent ${\cal D}_X$-module $(M,F_\bullet)$, this definition is equivalent to the definition in \cite{S1}, 3.5.1 because the condition of coherence of $(M^\bullet,F_\bullet)_Y$ is is equivalent to the finiteness of the projection $(Y\times_X T^\ast X)\cap\operatorname{Ch}(M)\to T^\ast Y$ where $\operatorname{Ch}(M)$ denotes the characteristic variety of $M$. In such case if $Y$ is noncharacteristic, we have $i^\ast(M,F_\bullet)=(M^\bullet,F_\bullet)_Y[d]$ and $i^!(M,F_\bullet)$ is isomorphic to $(M^\bullet,F_\bullet)_Y[-d]$ up to a shift of filtration. \begin{defn} A filtered coherent ${\cal D}_X$-module $(M,F_\bullet)$ is {\em Cohen---Macaulay\/} if $\operatorname{Gr}^F_\bullet M$ is a Cohen---Macaulay module over $\operatorname{Gr}^F_\bullet{\cal D}$. \end{defn} A Cohen---Macaulay ${\cal D}_X$-module $(M,F_\bullet)$ is holonomic iff the dimension of $\operatorname{Gr}^F_\bullet M$ over $\operatorname{Gr}^F_\bullet{\cal D}$ is equal to $\dim X$. \begin{lemma} \label{lem:dual-to-restriction} Suppose that $(M,F_\bullet)$ is a coherent holonomic filtered ${\cal D}_X$-module noncharacteristic with respect to $Y$. Then $(M,F_\bullet)$ is holonomic Cohen---Macaulay at a point $y\in Y$ if and only if $(M,F_\bullet)_Y$ is. \end{lemma} \begin{pf} We shall denote $(M,F_\bullet)_Y$ by $(M_Y,F)$. Let $\dim X=n$, $\operatorname{codim}_X Y=d$. Let $R=\operatorname{Gr}^F{\cal D}_{X,y}$ and $R'=\operatorname{Gr}^F{\cal D}_{Y,y}$, and let ${\frak m}$ (respectively, ${\frak m}'$) denote the maximal ideal in $R$ (respectively, $R'$) corresponding to the origin of $T^\ast_y X$ (respectively, $T^\ast_y Y$). Both $R$ and $R'$ are graded rings, $\operatorname{Gr}^F M_y$ and $\operatorname{Gr}^F M_{Y,y}$ are graded modules over them, and hence, the support of $\operatorname{Gr}^F M_y$ in $\operatorname{Spec} R$ corresponds to a homogeneous closed analytic subspace of $T^\ast U$ where $U$ is a sufficiently small open neighborhood of $y$ in $X$, and similarly for the support of $\operatorname{Gr}^F M_{Y,y}$ in $\operatorname{Spec} R'$. We need to show that $\operatorname{Gr}^F M_y$ is Cohen---Macaulay of dimension $n$ over $R$ iff $\operatorname{Gr}^F M_{Y,y}$ is Cohen---Macaulay of dimension $n-d$ over $R'$. The Cohen---Macaulay property of $\operatorname{Gr}^F M_y$ is equivalent to vanishing of $\operatorname{Ext}^j_R(\operatorname{Gr}^F M_y,R)$ for $j\ne n$; as the support of $\operatorname{Ext}^j_R(\operatorname{Gr}^F M_y,R)$ is homogeneous in $\operatorname{Spec} R$, this property holds at all points of $\operatorname{Spec} R$ iff it holds at the origin of $T^\ast_y X$, i.e., at the maximal ideal ${\frak m}$. In other words, $\operatorname{Gr}^F M_y$ is Cohen---Macaulay of dimension $n$ over $R$ iff $(\operatorname{Gr}^F M_y)_{\frak m}$ is Cohen---Macaulay of dimension $n$ over $R_{\frak m}$. Similarly, $\operatorname{Gr}^F M_{Y,y}$ is a Cohen---Macaulay $R'$-module of dimension $n-d$ iff $(\operatorname{Gr}^F M_{Y,y})_{{\frak m}'}$ is a Cohen---Macaulay $R'_{{\frak m}'}$-module of dimension $n-d$. It follows that we need to show that $(\operatorname{Gr}^F M_y)_{\frak m}$ is Cohen---Macaulay of dimension $n$ over $R_{\frak m}$ iff $(\operatorname{Gr}^F M_{Y,y})_{{\frak m}'}$ is Cohen---Macaulay of dimension $n-d$ over $R'_{{\frak m}'}$. Let $N=\operatorname{Gr}^F M_{Y,y}$. Let $x_1,\dots,x_n$ be a local coordinate system in $X$ at $y$ such that $x_1,\dots,x_d$ is a system of local equations of $Y$ in $X$. Since $Y$ is noncharacteristic, $x_1,\dots,x_d$ is a regular $\operatorname{Gr}^F M_y$-sequence in $R$, and $N\overset{\sim}{=}\operatorname{Gr}^F M_y/(\sum_{l=1}^d x_l\operatorname{Gr}^F M_y)$. Hence, $N$ is a module over $R/(\sum_{l=1}^d x_l R)$; its structure of an $R'$-module comes from the embedding $R'\hookrightarrow R/(\sum_{l=1}^d x_l R)$. Let $A$ and $A'$ be the quotients of $R/(\sum_{l=1}^d x_l R)$ and $R'$, respectively, by the annihilators of $N$. Then $A'\hookrightarrow A$; since $\operatorname{Gr}^F M_Y$ is $\operatorname{Gr}^F{\cal D}_Y$-coherent, $N$ is finite over $R'$ and hence, $A$ is a finite $A'$-module. Denote by $\tilde{\frak m}$ and $\tilde{\frak m}'$, respectively, the maximal ideals of $A$ and $A'$ that correspond to the maximal ideals ${\frak m}$ and ${\frak m}'$ of $R$ and $R'$, respectively. As $A$ is a finite $A'$-module, there are only finitely many ideals in $A$ lying over $\tilde{\frak m}'$, and clearly, $\tilde{\frak m}$ is one of them. By a homogeneity argument, $\tilde{\frak m}$ is the only ideal of $A$ lying over $\tilde{\frak m}'$. Hence, $A_{\tilde{\frak m}}$ is a finite $A'_{\tilde{\frak m}'}$-module. The localization $N_{{\frak m}'}=(\operatorname{Gr}^F M_{Y,y})_{{\frak m}'}$ of $N$ at ${\frak m}'$ as an $R'$-module is the same as the localization $N_{\tilde{\frak m}'}$ of $N$ at $\tilde{\frak m}'$ as an $A'$-module, and is isomorphic to the localization $N_{\tilde{\frak m}}$ of $N$ at $\tilde{\frak m}$ as an $A$-module (since $\tilde{\frak m}$ is the only ideal of $A$ lying over $\tilde{\frak m}'$). By~\cite{Se}, Ch.~IV, Proposition~12, $\operatorname{depth}_{A'_{\tilde{\frak m}'}}N_{\tilde{\frak m}'}=\operatorname{depth}_{A_{\tilde{\frak m}}}N_{\tilde{\frak m}}$ and $\dim_{A'_{\tilde{\frak m}'}}N_{\tilde{\frak m}'}=\dim_{A_{\tilde{\frak m}}}N_{\tilde{\frak m}}$. Clearly, $\operatorname{depth}_{R'_{{\frak m}'}}N_{{\frak m}'}=\operatorname{depth}_{A'_{\tilde{\frak m}'}}N_{\tilde{\frak m}'}$ and $\dim_{R'_{{\frak m}'}}N_{{\frak m}'}=\dim_{A'_{\tilde{\frak m}'}}N_{\tilde{\frak m}'}$. Since $x_1,\dots,x_d$ is a regular $\operatorname{Gr}^F M_y$-sequence in $R$, it is a regular $(\operatorname{Gr}^F M_y)_{\frak m}$-sequence in $R_{\frak m}$. We have $N_{\tilde{\frak m}}\overset{\sim}{=}(\operatorname{Gr}^F M_y)_{\frak m}/(\sum_{l=1}^d x_l(\operatorname{Gr}^F M_y)_{\frak m})$, and hence, $\operatorname{depth}_{A_{\tilde{\frak m}}}N_{\tilde{\frak m}}=\operatorname{depth}_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}-d$ and $\dim_{A_{\tilde{\frak m}}}N_{\tilde{\frak m}}=\dim_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}-d$. Altogether, we see that $\operatorname{depth}_{R'_{{\frak m}'}}N_{{\frak m}'}=\operatorname{depth}_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}-d$ and $\dim_{R'_{{\frak m}'}}N_{{\frak m}'}=\dim_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}-d$. It follows that $\operatorname{depth}_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}=\dim_{R_{\frak m}}(\operatorname{Gr}^F M_y)_{\frak m}=n$ iff $\operatorname{depth}_{R'_{{\frak m}'}}N_{{\frak m}'}=\dim_{R'_{{\frak m}'}}N_{{\frak m}'}=n-d$, i.e., $(\operatorname{Gr}^F M_y)_{\frak m}$ is Cohen---Macaulay of dimension $n$ over $R_{\frak m}$ iff $N_{{\frak m}'}$ is Cohen---Macaulay of dimension $n-d$ over $R'_{{\frak m}'}$. \end{pf} \begin{remark} \label{rem:right-exact} It is not hard to see from the definitions that the functors $\operatorname{DR}_X$ and $\operatorname{DR}^{-1}_X$ are right exact in the filtered sense: if a filtered complex $(M^\bullet,F_\bullet)$ has the property that $H^j\operatorname{Gr}^F_\bullet M^\bullet=0$ for $j>j_0$ then $H^j\operatorname{Gr}^F_\bullet\operatorname{DR}_X(M^\bullet,F_\bullet)=0$ for $j>j_0$, and vice versa, and the same holds for the functor of noncharacteristic restriction $(\bullet)_Y$. The functor $\operatorname{DR}^{-1}_X$ is also left exact in the similar sense. \end{remark} \section{Filtered perversity} \label{sec:fil.perv} In this section we introduce the notion of a {\em weakly filtered perverse} differential complex; its meaning is that the complex is ``locally trivial'' in a certain filtered sense made precise below, and satisfies filtered cohomology vanishing conditions which are similar to the cohomology vanishing conditions for perverse sheaves. The stratifications need to be defined only locally, which is made precise by the notion of {\em stratified chart.} This notion of {\em weak filtered perversity} is precisely the assumption that we need to use; we call it {\em weak} because we suspect that some stronger property of ``local triviality'' along the strata will appear eventually. We introduce also the notion of {\em coherent filtered perversity\/} which is somewhat stronger than coherence together with weak filtered perversity; we shall show in Proposition~\ref{cor:holon} that it implies holonomicity of the corresponding ${\cal D}_X$-module. \subsection{Stratified charts} \begin{defn} A {\em stratified chart} ${\cal U}$ on $X$ is the following collection of data: \begin{enumerate} \item an open subset $U$ of $X$; \item an analytic stratification of $U$; \item for every point $x$ of any stratum $S$ of this stratification, an open neighborhood $U_x$ of $x$ in $U$ and an analytic submersion $\pi_x : U_x\to U_x\cap S$ which restricts to the identity on $U_x\cap S$. \end{enumerate} \end{defn} \subsection{The definition of filtered perversity} In what follows we denote by $Y$ the fiber $\pi_x^{-1}(x)$. Given a filtered differential complex $({\cal M}^\bullet,F_\bullet)$ on $X$, we use the notation ${\cal F}^j_p=H^j\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$; this is a sheaf of ${\cal O}_X$-modules. \begin{defn} \label{cond:wfp} A filtered differential complex $({\cal M}^\bullet,F_\bullet)$ on $X$ is called {\em weakly filtered perverse} if $X$ can be covered by stratified charts ${\cal U}$ which satisfy the following properties for every point $x$ of any stratum $S$ of ${\cal U}$: \begin{itemize} \item[(i)] for all $j$ and $p$, the sheaf ${\cal F}^j_p$ has the property that for all $i>0$, and for all $y\in Y$, we have ${\operatorname{Tor}}^{{\cal O}_{X,y}}_i({\cal F}^j_{p,y},{\cal O}_{Y,y}) = 0$; \item[(ii)] for all $j$, $p$, if ${\cal F}^j_{p,x}\otimes_{{\cal O}_{X,x}}{\cal O}_{Y,x} = 0$ then ${\cal F}^j_{p,x}=0$; \item[(iii)] for all $p$, all $j<0$, we have $H^j_{\{x\}}\operatorname{Gr}^F_p\operatorname{DR}_Y\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right) = 0$; \item[(iv)] for all $p$, all $j>0$, we have $H^j\operatorname{Gr}^F_p{\cal M}^\bullet = 0$. \end{itemize} We say that $({\cal M}^\bullet,F_\bullet)$ is {\em coherent filtered perverse\/} if it is weakly filtered perverse, the complex $\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ has bounded $\operatorname{Gr}^F_\bullet{\cal D}_Y$-coherent cohomology, and for any point $x$, $({\cal M}^\bullet,F_\bullet)$ is noncharacteristic with respect to $Y$. \end{defn} Note that property (i) means that the complex $\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ satisfies the condition \eqref{eqn:tor-nonchar}, and hence, the noncharacteristic restriction $(\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y$ which appears in (iii), is defined. (We shall actually see that if $({\cal M}^\bullet,F_\bullet)$ is weakly filtered perverse then $\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ is isomorphic to one filtered ${\cal D}_X$-module $(M,F_\bullet)$, and the condition of coherent filtered perversity is equivalent to the condition that $(M,F_\bullet)$ coherent holonomic Cohen-Macaulay.) \subsection{Construction of stratified charts in the coherent case} Suppose that $({\cal M}^\bullet,F_\bullet)$ is a filtered differential complex such that $\operatorname{Gr}^F_\bullet{\cal M}^\bullet$ has bounded coherent cohomology. Then the condition (ii) is always satisfied. We shall see here that under this assumption, there always exist stratified charts satisfying also (i). \begin{prop} \label{prop:hom.shf.flat} Suppose that $p:E\to X$ is a holomorphic vector bundle on $X$ and $\left\{{\cal F}_i\right\}_I$ is a finite collection of homogeneous coherent sheaves on $E$. Then at any point $x_0$ of $X$ there exists a stratified chart ${\cal U}$ such that for every $x\in U$ and $i\in I$ the sheaf ${\cal F}_i\vert_{p^{-1}U_x}$ is $(\pi_xp)$-flat over $U_x\cap S$. \end{prop} \begin{pf} Let $n=\dim X$. We shall construct inductively a stratified chart ${\cal U}^k$ containing $x_0$ such that it satisfies the required flatness property at all points $x$ of all the strata $S$ of codimension smaller than $k$. We shall show how to construct ${\cal U}^{k+1}$ once ${\cal U}^k$ has been constructed. Let $U^k$ be the open set containing $x_0$ which underlies ${\cal U}^k$, and let $X^k$ be the union of the closures of all the strata of ${\cal U}^k$ of codimension $k$. Then $X^k$ is a closed analytic subset of $U^k$ of pure dimension $n-k$. We choose an open polydisc $\Delta^n$ embedded in $U^k$ which contains $x_0$ and such that there is a projection $q:\Delta^n\to\Delta^{n-k}$ with the property that the map $q|_{X^k\cap \Delta^n}:X^k\cap \Delta^n\to\Delta^{n-k}$ is finite. Consider the composite projection $qp:p^{-1}\Delta^n\to\Delta^{n-k}$. Let $Z$ be the set of points in $p^{-1}\Delta^n\subset E$ where one of the sheaves ${\cal F}_i$ is not $qp$-flat. By Frisch's theorem on the openness of the flat locus (\cite{F}, Theorem (IV,9) or \cite{BS}, Theorem V.4.5), $Z$ is a closed analytic subset in $p^{-1}\Delta^n$ such that its image in $\Delta^{n-k}$ is negligible. All the sheaves ${\cal F}_i$ are homogeneous, and hence, $Z$ is homogeneous; it follows that $p(Z)$ is a closed analytic subset of $\Delta^n$. The intersection $p(Z)\cap X^k$ is negligible in $X^k\cap \Delta^n$ since its image under $q$ is negligible and $q$ is finite on $X^k\cap \Delta^n$; it follows that $p(Z)\cap X^k$ is a proper closed analytic subset of $X^k\cap \Delta^n$. Construct ${\cal U}^{k+1}$ as follows. Take $U^{k+1}=\Delta^n$. All the strata of ${\cal U}^{k+1}$ of codimension less than $k$ are the intersections with $U^{k+1}$ of the strata of ${\cal U}^k$. Any $(n-k)$-stratum $S'$ is obtained from a $(n-k)$-stratum $S$ of ${\cal U}^k$ by intersecting with $U^{k+1}$ and then removing, first, all points where $q|_{S\cap U^{k+1}}:S\cap U^{k+1}\to\Delta^{n-k}$ is ramified, and second, the intersection with $p(Z)$. The complement of these strata in $U^{k+1}$ has codimension at least $k+1$; stratifying it, we complete the stratification of ${\cal U}^{k+1}$. Clearly, the stata of codimension less than $k$ satisfy the required flatness condition. Let $S'$ be any $(n-k)$-stratum constructed as above. At any point $x\in S'$ we take a neighborhood $U_x\subset U^{k+1}$ in such way that $q(U_x)=q(U_x\cap S')$ and $q$ is an isomorphism on $U_x\cap S'$. Take the projection $\pi_x:U_x\to U_x\cap S'$ such that $q\pi_x=q$, i.e., $\pi_x=(q|_{U_x\cap S'})^{-1}q$. Then all the sheaves ${\cal F}_i$ are $qp$-flat on $p^{-1}(U_x\cap S')$ since $S'$ does not intersect $p(Z)$, and hence, they are $\pi_xp$-flat. \end{pf} \begin{cor} \label{cor:strat} Suppose that $({\cal M}_i^\bullet,F_\bullet)$ is a finite collection of filtered differential complexes on $X$ such that $\operatorname{Gr}^F_\bullet{\cal M}_i^\bullet$ have bounded coherent cohomology. Then, locally at any point of $X$ there exists a stratified chart such that the properties (i) and (ii) of Definition~\ref{cond:wfp} hold for each $({\cal M}_i^\bullet,F_\bullet)$. \end{cor} \begin{pf} Consider the homogeneous coherent sheaves $\left(H^j\operatorname{Gr}^F_\bullet\operatorname{DR}^{-1}({\cal M}_i^\bullet,F_\bullet)\right)^\sim$ on $T^*X$ obtained by localizing the corresponding $\operatorname{Gr}^F_\bullet{\cal D}_X$-modules. Proposition~\ref{prop:hom.shf.flat} yields a stratified chart satifying the condition (i) for each $({\cal M}_i^\bullet,F_\bullet)$; the condition (ii) is satisfied by coherence. \end{pf} \section{Filtered perverse complexes correspond to filtered $\protect{\cal D}$-modules} \label{sec:fpc-to-dm} \subsection{The main theorem} Given a filtered complex $(M^\bullet,F_\bullet)$, the property that $H^j\operatorname{Gr}^F_pM^\bullet =0$ for all $p$ and all $j\neq 0$, means that $(M^\bullet,F_\bullet)$ is strict and $H^jM^\bullet = 0$ for $j\neq 0$. Another formulation of the same property is that $(M^\bullet,F_\bullet)$ is isomorphic to $H^0(M^\bullet,F_\bullet)$ in the filtered derived category, where $H^0(M^\bullet,F_\bullet)$ denotes $H^0M^\bullet$ equipped with the induced filtration. \begin{thm} \label{thm:main} Suppose that $({\cal M}^\bullet,F_\bullet)$ is weakly filtered perverse. Then, for all $p$, all $j\neq 0$, \begin{equation*} H^j\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet) = 0 \end{equation*} Consequently the filtered complex $\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ is strict and isomorphic in the filtered derived category to $H^0\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ equipped with the induced filtration. \end{thm} \begin{pf} The statement is local so we may assume that $X=U$ in the definition of weak filtered perversity, $x$ lies in the stratum $S$, $\pi_x: U_x\to S$ is an analytic submersion which restricts to the identity on $S$, and $Y=\pi_x^{-1}(x)$. We are going to show by induction on $\operatorname{codim} S$ that the conclusion holds for the stalk of $\operatorname{Gr}^F_p\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ at $x$. Thus we may assume that the conclusion holds on the complement of the stratum $S$. Condition (iv) of Definition \ref{cond:wfp} implies that, for all $p$, \begin{equation}\label{van:above} \text{$H^j\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet) = 0$ for $j>0$}\ . \end{equation} By~\eqref{eqn:HGr(restr)} we have \begin{equation}\label{iso:restr} H^j\operatorname{Gr}^F_p\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)\overset{\sim}{=} H^j\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)\otimes_{{\cal O}_X}\omega_{Y/X}\ . \end{equation} The induction hypothesis and \eqref{iso:restr} imply that \begin{equation}\label{van:on-Y-S} \text{ $H^j\operatorname{Gr}^F_p\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right) \vert_{Y\setminus\{x\}} = 0$ for $j\neq 0$}\ . \end{equation} Let $i:\{x\}\hookrightarrow Y$ be the embedding map. Condition (iii) of Definition \ref{cond:wfp} implies that $\operatorname{\bold R} i^!\operatorname{Gr}^F_p\operatorname{DR}_Y\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)$ is acyclic in negative degrees. Here $\operatorname{\bold R} i^!$ is the derived functor of the functor $i^!$ which assigns to a sheaf ${\cal F}$ its sections supported in $x$; if ${\cal F}$ is an ${\cal O}_Y$-module or a ${\cal D}_Y$-module then $\operatorname{\bold R} i^!{\cal F}$ is an an ${\cal O}_{Y,x}$-module or a ${\cal D}_{Y,x}$-module, and $\operatorname{\bold R} i^!$ commutes with the functors $\operatorname{Gr}^F_p$ and $\operatorname{DR}_Y$. Hence, $\operatorname{Gr}^F_p\operatorname{DR}_Y\operatorname{\bold R} i^!\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)$ is acyclic in negative degrees; as $\operatorname{DR}^{-1}_Y$ is left exact (Remark~\ref{rem:right-exact}), the complex $\operatorname{Gr}^F_p\operatorname{\bold R} i^!\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)$ is acyclic in negative degrees, so that \begin{equation}\label{van:loc-coh} H^j_{\{x\}}\operatorname{Gr}^F_p\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)= 0 \text{ for $j<0$ .} \end{equation} Examination of the long exact sequence in cohomology associated to the inclusion $Y\setminus\{x\}\subset Y$ in the light of \eqref{van:on-Y-S} and \eqref{van:loc-coh} shows that \[ \text{$H^j\operatorname{Gr}^F_p\left((\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet))_Y\right)= 0$ for $j<0$}\ . \] Together with \eqref{iso:restr} and the condition (ii) of Definition \ref{cond:wfp} this shows that \begin{equation} \label{eqn:van:j<0} \text{$H^j\operatorname{Gr}^F_p\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)_x = 0$ for $j<0$}\ . \end{equation} The statement of the Theorem is the combination of \eqref{van:above} and \eqref{eqn:van:j<0}. \end{pf} \section{The coherent case} \label{sec:coh.case} In this section we study the property of filtered perversity of a filtered differential complex $({\cal M}^\bullet,F_\bullet)$ under the assumption of coherence of $H^\bullet\operatorname{Gr}^F_\bullet{\cal M}^\bullet$; in particular, this implies that the cohomology of $\operatorname{Gr}^F_\bullet{\cal M}^\bullet$ is bounded. By~\cite{S1}, (2.2.10.5), this is equivalent to the $\operatorname{Gr}^F_\bullet{\cal D}_X$-coherence of $H^\bullet\operatorname{Gr}^F_\bullet\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$; in case $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is isomorphic to a single filtered ${\cal D}_X$-module, this property means that the module is ${\cal D}_X$-coherent and its filtration is good. \subsection{Duality for coherent complexes} The following technical lemma is a standard application of duality theory. \begin{lemma}\label{lemma:duality} Suppose that $X$ is a complex manifold of dimension $n$, $L^\bullet$ is a bounded complex of coherent ${\cal O}_X$-modules, and $x\in X$. Then for each $j$, there is a nondegenerate pairing between the spaces $H^{-j}_{\{x\}}L^\bullet$ and $\left(H^j\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])\right)_x$; the same is true for the spaces $H^jL^\bullet_x$ and $H^{-j}_{\{x\}}\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])$. \end{lemma} A nondegenerate pairing between two vector spaces is a pairing that induces a monomorphism from each of them into the (algebraic) dual of the other. (Actually, each of the vector spaces can be given a topology so that they become topologically dual. More precisely, the pairs of spaces indicated in the Lemma, are strong dual to each other with respect to certain natural FS and DFS topologies. In case the complex $L^\bullet$ is zero except in one degree, this statement is a particular case of a theorem of Harvey: take $K=\{x\}$ in Theorem~5.12 of~\cite{ST}. However, we do not need the topological duality; all we need is that these vector spaces are either both zero or both nonzero.) \begin{pf*}{Proof of Lemma \protect\ref{lemma:duality}} We shall establish the duality between the spaces $H^{-j}_{\{x\}}L^\bullet$ and\break $\left(H^j\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])\right)_x$; the other duality would follow by substituting the dual complex $\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])$ in place of $L^\bullet$. Replacing $L^\bullet$ by its bounded free resolution in a neighborhood of $x$, we may assume that all the sheaves $L^k$ are free. Then $\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])$ is represented by $\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])$, and $\left(H^j\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])\right)_x\overset{\sim}{=} H^j\operatorname{Hom}^\bullet_{{\cal O}_{X,x}}(L^\bullet_x,\omega_{X,x}[n])$. The complex $\operatorname{Hom}^\bullet_{{\cal O}_{X,x}}(L^\bullet_x,\omega_{X,x}[n])$ is a complex of free finitely generated ${\cal O}_{X,x}$-modules. Each of them has a canonical DFS topology and the differential is continuous with respect to it; moreover, the image of the differential is closed since it is closed with respect to the weaker topology of coefficientwise convergence of formal power series (Theorem~6.3.5 of~\cite{H}). Since each $L^k$ is free, by a theorem of Martineau (\cite{ST}, Theorem 5.9) we have that $H^j_{\{x\}}L^k$ is zero for $j\ne n$, and $H^n_{\{x\}} L^k$ can be given a natural Hausdorff FS topology in which it is a strong dual to $\operatorname{Hom}_{{\cal O}_{X,x}}(L_x^k,\omega_{X,x})$; in particular, it follows that $H^{-j}_{\{x\}} L^\bullet\overset{\sim}{=} H^{-j-n}(H^n_{\{x\}} L)^\bullet$ where we denote $(H^n_{\{x\}} L)^\bullet= \{\dots\to H^n_{\{x\}} L^k\to H^n_{\{x\}} L^{k+1}\to\dots\}$. The pairing between $H^n_{\{x\}} L^k$ and $\operatorname{Hom}_{{\cal O}_{X,x}}(L_x^k,\omega_{X,x})$, is given by the composition of the multiplication $H^n_{\{x\}}L^k\otimes\operatorname{Hom}_{{\cal O}_{X,x}}(L_x^k,\omega_{X,x})\to H^n_{\{x\}}\omega_X$ and the residue map $H^n_{\{x\}}\omega_X\to{\Bbb C}$, and hence, the complex $(H^n_{\{x\}} L)^\bullet$ is the strong dual to the complex $\operatorname{Hom}^\bullet_{{\cal O}_{X,x}}(L_x^\bullet,\omega_{X,x}[n])$. As the latter is a complex of DFS spaces with Hausdorff cohomology, the former is a complex of FS spaces with Hausdorff cohomology $H^{-j-n}(H^n_{\{x\}} L)^\bullet$ strong dual to $\left(H^j\operatorname{Hom}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])\right)_x$. This yields a nondegenerate pairing between $H^{-j}_{\{x\}} L^\bullet$ and $\left(H^j\operatorname{Hom}^\bullet_{{\cal O}_X}(L^\bullet,\omega_X[n])\right)_x$. \end{pf*} \subsection{Holonomicity} \begin{prop} \label{cor:holon} Suppose that $({\cal M}^\bullet,F_\bullet)$ is a coherent filtered perverse complex on a complex manifold $X$. Then $\operatorname{DR}^{-1}({\cal M}^\bullet,F_\bullet)$ is isomorphic in the filtered derived category to a filtered holonomic Cohen-Macauley ${\cal D}_X$-module. \end{prop} \begin{pf} The question is local and we need to prove it in a neighborhood of any point $x\in X$. The point $x$ is covered by a stratified chart ${\cal U}$ satisfying properties (i)--(iv) of Definition~\ref{cond:wfp}; we keep the notation introduced there. Our assumptions imply that $\operatorname{DR}^{-1}_X({\cal M}^\bullet,F_\bullet)$ is isomorphic to a coherent filtered ${\cal D}_X$-module; we shall denote this module by $(M,F_\bullet)$. We argue by induction by the codimension of the stratum $S$ containing $x$; by the inductive assumption, we may assume that $(M,F_\bullet)$ is holonomic Cohen-Macauley in the complement to $S$. By Lemma~\ref{lem:dual-to-restriction}, this implies that $(M,F_\bullet)_Y$ is holonomic Cohen-Macauley everywhere on $Y$ except possibly at $x$; hence, it is holonomic. Property (iii) implies that $H^j_{\{x\}}\operatorname{Gr}^F_p\operatorname{DR}_Y((M,F_\bullet)_Y)=0$ if $j<0$. By Lemma~\ref{lemma:duality} this yields $\left(H^j\operatorname{\bold R}\underline{\operatorname{Hom}}^\bullet_{{\cal O}_Y} (\operatorname{Gr}^F_p\operatorname{DR}_Y((M,F_\bullet)_Y),\omega_Y[\dim Y])\right)_x=0$ if $j>0$. By \S\ref{subsec:duality} this implies that $H^j\operatorname{Gr}^F_p\operatorname{DR}_Y{\Bbb D}((M,F_\bullet)_Y)=0$ if $j>0$, and by right exactness of $\operatorname{DR}^{-1}_Y$ we get $H^j\operatorname{Gr}^F_p{\Bbb D}((M,F_\bullet)_Y)=0$ if $j>0$. Since $(M,F_\bullet)_Y$ is a filtered ${\cal D}_Y$-module, we have $H^j\operatorname{Gr}^F_p\operatorname{DR}_Y((M,F_\bullet)_Y)=0$ for $j>0$. By Lemma~\ref{lemma:duality} and \S\ref{subsec:duality}, this implies $H^j_{\{x\}}\operatorname{Gr}^F_p\operatorname{DR}_Y{\Bbb D}((M,F_\bullet)_Y)=0$ for $j<0$. So we get $H^j_{\{x\}}\operatorname{Gr}^F_p{\Bbb D}((M,F_\bullet)_Y)=0$ for $j<0$ by the left exactness of $\operatorname{DR}^{-1}_Y$. The long exact sequence of the inclusion $Y\setminus\{x\}\subset Y$ (cf.\ the proof of Theorem~\ref{thm:main}) yields $H^j\operatorname{Gr}^F_p{\Bbb D}((M,F_\bullet)_Y)=0$ if $j<0$. (Actually, this vanishing also follows from the vanishing --- see, for example,~\cite{Borel}, V.2.2.2 --- of $\operatorname{Ext}^i_{\operatorname{Gr}^F{\cal D}_{Y,x}}(\operatorname{Gr}^F M_{Y,x},\operatorname{Gr}^F{\cal D}_{Y,x})$ for $i<d$ where $d$ is the codimension of the support of $\operatorname{Gr}^F M_{Y,x}$ in $\operatorname{Spec}\operatorname{Gr}^F{\cal D}_{Y,x}$; in our case $d=\dim Y$.) Altogether, we see that $H^j\operatorname{Gr}^F_p{\Bbb D}((M,F_\bullet)_Y)=0$ if $j\ne 0$, i.e., the filtered complex ${\Bbb D}((M,F_\bullet)_Y)$ is isomorphic to one filtered module. Hence, $(M,F_\bullet)_Y$ is Cohen-Macaulay at $x$. It follows by Lemma~\ref{lem:dual-to-restriction} that $(M,F_\bullet)$ is also Cohen-Macaulay at $x$. \end{pf} \subsection{The converse to the Main Theorem} \begin{thm} \label{thm:converse} If a coherent filtered ${\cal D}_X$-module $(M,F_\bullet)$ is holonomic and Cohen-Macauley, then $\operatorname{DR}_X(M,F_\bullet)$ is coherent filtered perverse. \end{thm} \begin{pf} Consider a point $x\in X$. The coherence of $(M,F_\bullet)$ implies that $H^\bullet\operatorname{Gr}^F_\bullet\operatorname{DR}(M,F_\bullet)$ is ${\cal O}_X$-coherent. Consequently, Corollary~\ref{cor:strat} implies that in a neighborhood of $x$ there exists a stratified chart ${\cal U}$ such that $\operatorname{DR}(M,F_\bullet)$ satisfies properties (i) and (ii) with respect to it. We shall keep the notation of Definition~\ref{cond:wfp}. As $M$ is holonomic, we may assume that that the characterisic variety $\operatorname{Ch}(M)$ is contained in the union of the conormal bundles to the strata of the Whitney stratification that underlies ${\cal U}$. It follows that the projection $(Y\times_X T^\ast X)\cap\operatorname{Ch}(M)\to T^\ast Y$ is finite (it is even an embedding), and hence, $(M,F_\bullet)_Y$ is a coherent filtered ${\cal D}_Y$-module. This implies that $Y$ is noncharacteristic with respect to $(M,F_\bullet)$. The property (iv) of Definition~\ref{cond:wfp} at $x$ is satisfied by $\operatorname{DR}(M,F_\bullet)$ since it is satisfied by the de Rham complex of any filtered ${\cal D}_X$-module. Let us show the property (iii) at $x$ with respect to this stratified chart. By Lemma~\ref{lem:dual-to-restriction}, $(M,F_\bullet)_Y$ is holonomic Cohen-Macaulay at $x$. Hence, the complex ${\Bbb D}((M,F_\bullet)_Y)$ is isomorphic to a filtered ${\cal D}_Y$-module, and consequently, the complex $\operatorname{DR}_Y{\Bbb D}((M,F_\bullet)_Y)$ satisfies (iv) at $x$: \[ H^j\operatorname{Gr}^F_\bullet\operatorname{DR}_Y{\Bbb D}((M,F_\bullet)_Y)=0\text{ \ at $x$ for all $j>0$.} \] By \S\ref{subsec:duality} and Lemma~\ref{lemma:duality} we get \[ H^j_{\{x\}}\operatorname{Gr}^F_\bullet\operatorname{DR}_Y((M,F_\bullet)_Y)=0\text{ \ for all $j<0$.} \] This is the property (iii) at $x$ for the filtered complex $\operatorname{DR}_X(M,F_\bullet)$. \end{pf} \section{An application to {\protect$L^2$} cohomology} \label{sec:appl} \renewcommand{F^\bullet}{{F_\bullet}} In this section we give an application to our results and show that in the situation of \cite{KK2} (and under the assumption of filtered perversity of the $L^2$\ complex) there is a local filtered isomorphism (in the sense of derived category) between the $L^2$ complex and the de Rham complex of the ${\cal D}$-module that underlies the corresponding pure Hodge module. \subsection{} \label{subsec:l2} Let $X$ denote a K\"ahler manifold of dimension $n$, let $j:{X^\circ}\hookrightarrow X$ be the inclusion map of the complement of a divisor with normal crossings, and ${\Bbb E} = ({\Bbb E}_{{\Bbb Q}},({\cal O}_{X^\circ}\otimes_{{\Bbb Q}}{\Bbb E}_{{\Bbb Q}},F^\bullet))$ a quasiunipotent polarised variation of pure Hodge structure of weight $w$ on $X^\circ$. Let $(N,F_\bullet)$ denote the filtered ${\cal D}_X$-module underlying the polarizable Hodge module~\cite{S1} which restricts to $(\omega_{X^\circ}\otimes_{{\Bbb Q}}{\Bbb E}_{{\Bbb Q}},F^\bullet)$ on $X^\circ$. Let $({\cal A}^{\bullet}_{(2)}({\Bbb E}),F^\bullet)$ denote the {$L^2$}-complex with coefficients in ${\Bbb E}$ constructed using the Hodge inner product in the fibers of ${\Bbb E}$ and a certain complete metric $\eta$ on $X^\circ$ as in~\cite{KK2}, \cite{CKS}; to keep up with our degree conventions, we shall assume that the grading of $({\cal A}^{\bullet}_{(2)}({\Bbb E}),F^\bullet)$ is chosen in such way that ${\cal A}^i_{(2)}({\Bbb E})$ contains forms of degree $i+n$. As the metric $\eta$ satisfies $\eta>C\eta_X$ locally in a neighborhood of any point of $X$, where $\eta_X$ is the metric on $X$ and $C$ a suitable positive constant, the holomorphic forms on $X$ are bounded in the pointwise norm with respect to $\eta$ and the {$L^2$}-complex is an ${\cal O}_X$-module: if $\omega$ is a section of ${\cal A}^{\bullet}_{(2)}({\Bbb E})$ and $f$ is a holomorphic function then $f\omega$ is also a section of ${\cal A}^{\bullet}_{(2)}({\Bbb E})$ since both $f\omega$ and $d(f\omega)=df\wedge\omega+fd\omega$ are $L^2$. We shall assume here that $({\cal A}^{\bullet}_{(2)}({\Bbb E}),F^\bullet)$ is weakly filtered perverse, and so is every direct summand of it (in the sense of derived category). By Theorem~\ref{thm:main} this implies that the complex $\operatorname{DR}^{-1}({\cal A}^{\bullet}_{(2)}({\Bbb E}),F^\bullet)$ is strict and isomorphic in the filtered derived category to its zeroeth cohomology with the induced filtration. By~\cite{KK2} and \cite{CKS}, ${\cal A}^{\bullet}_{(2)}({\Bbb E})$ is isomorphic in the derived category of complexes of sheaves on $X$ to the intersection complex with coefficients in ${\Bbb E}$. We shall assume, moreover, that for any cross-section $Y$ appearing in the definition of weak filtered perversity, the complex $\operatorname{DR}_Y\left(\left(\operatorname{DR}^{-1}_X({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}))\right)_Y\right)$ is isomorphic to the intersection cohomology complex on $Y$ with the coefficients in ${\Bbb E}|_{Y\capX^\circ}$ in the derived category of complexes (without filtration). (One would even expect that the filtered complex $\operatorname{DR}_Y\left(\left(\operatorname{DR}^{-1}_X({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}),F_\bullet) \right)_Y\right)$ is isomorphic in the filtered derived category to $({\cal A}^{\bullet}_{(2)}(Y,{\Bbb E}|_{Y\capX^\circ}),F_\bullet)$.) In case $X$ is compact, both complexes of global sections $\Gamma(X,{\cal A}^{\bullet}_{(2)}({\Bbb E}))$ and $\Gamma(X,\operatorname{DR}(N))$ are strict (\cite{KK2}, \cite{S1}), and their cohomology have pure Hodge structures. Their cohomology groups are isomorphic (\cite{KK2}) together with the Hodge filtrations (\cite{S2}, p.~294). Here we strengthen these results and show the isomorphism at the level of sheaves (without the assumption of compactness), in the filtered derived categories: \begin{prop} \label{thm:applic} Assume that \begin{enumerate} \item any direct summand of $({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}),F_\bullet)$ in the filtered derived category of filtered differential complexes (in particular, $({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}),F_\bullet)$ itself) is weakly filtered perverse; \item for any cross-section $Y$ as in Definition \ref{cond:wfp} (weak filtered perversity), the complex $\operatorname{DR}_Y\left(\left(\operatorname{DR}^{-1}_X({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}))\right)_Y\right)$ is isomorphic to the intersection cohomology complex on $Y$ with the coefficients in ${\Bbb E}|_{Y\capX^\circ}$. \end{enumerate} Then the filtered differential complexes $\operatorname{DR}_X(M,F_\bullet)$ and $({\cal A}^{\bullet}_{(2)}(X,{\Bbb E}),F_\bullet)$ are isomorphic in the filtered derived category. Equivalently, the filtered ${\cal D}_X$-modules $(M,F_\bullet)$ and $H^0\operatorname{DR}^{-1}({\cal A}^{\bullet}_{(2)}({\Bbb E}),F_\bullet)$ are isomorphic. \end{prop} \begin{pf} By Remark 3.15 of~\cite{S2} (the idea actually going back to~\cite{KK}), there is a direct sum decompostion in the derived category of filtered differential complexes \begin{equation*} ({\cal A}^{\bullet}_{(2)}({\Bbb E}),F^\bullet)\overset{\sim}{=} \operatorname{DR}_X(N,F_\bullet)\oplus ({\cal M}^\bullet,F_\bullet) \end{equation*} and we need to show that the second summand is trivial. Our assumptions imply that \begin{enumerate} \item $({\cal M}^\bullet,F_\bullet)$ is weakly filtered perverse, therefore, by Theorem~\ref{thm:main}, isomorphic to $\operatorname{DR}_X(M,F_\bullet)$ where $(M,F_\bullet)$ is a filtered ${\cal D}_X$-module; \item for any cross-section $Y$ as in Definition \ref{cond:wfp}, the complex $\operatorname{DR}_Y((\operatorname{DR}^{-1}_X{\cal M}^\bullet)_Y)\overset{\sim}{=}\operatorname{DR}_Y(M_Y)$ is acyclic. \end{enumerate} The weak filtered perversity of $({\cal M}^\bullet,F_\bullet)$ implies that $X$ is covered by stratified charts satisfying properties (i)--(iv) of Definition \ref{cond:wfp}. It is sufficient to show that the intersection of the support of $H^\bullet\operatorname{Gr}^F_\bullet({\cal M}^\bullet,F_\bullet)$ with any of the charts is empty. Assume to the contrary and consider a stratified chart and a point $x\in\operatorname{Supp} H^\bullet\operatorname{Gr}^F_\bullet({\cal M}^\bullet,F_\bullet)$ which lies on a stratum which is maximal among those which have a nonempty intersection with $\operatorname{Supp} H^\bullet\operatorname{Gr}^F_\bullet({\cal M}^\bullet,F_\bullet)$. Let $Y$ denote the cross-section at $x$. Then $\operatorname{Supp} M_Y\subseteq\{x\}$. In addition we have $H^j\operatorname{Gr}^F_\bullet\operatorname{DR}_Y((M,F_\bullet)_Y)=0$ for $j>0$ by the right exactness of $\operatorname{DR}_Y$, and for $j<0$ by property (iii) observing that $H^j\operatorname{Gr}^F_\bullet\operatorname{DR}_Y((M,F_\bullet)_Y)\overset{\sim}{=} H^j_{\{x\}}\operatorname{Gr}^F_\bullet\operatorname{DR}_Y((M,F_\bullet)_Y)$ since $\operatorname{Supp}\operatorname{Gr}^F_\bullet\operatorname{DR}_Y((M,F_\bullet)_Y)\subseteq\{x\}$. It follows that the filtered complex $\operatorname{DR}_Y((M,F_\bullet)_Y)$ is strict. Since the complex $\operatorname{DR}_Y(M_Y)$ is acyclic it follows that the complex $\operatorname{DR}_Y((M,F_\bullet)_Y)$ is filtered acyclic, and hence, $M_Y$ is trivial. The property (ii) of the weak filtered perversity shows that $M$ is trivial at $x$ (cf.\ the proof of Theorem~\ref{thm:main}), which contradicts our assumption. Hence, $M$ is trivial. \end{pf} \begin{cor} In the assumptions of Proposition~\ref{thm:applic} the sheaves $H^\bullet\operatorname{Gr}^F_\bullet{\cal A}^{\bullet}_{(2)}({\Bbb E})$ (the $L^2$-$\overline{\partial}$-cohomology, see below) are coherent. \end{cor} \subsection{Remarks on the $L^2$-$\overline\partial$-cohomology of a singular variety} \label{rem:bound-cond} Let $({\cal M}^\bullet,F_\bullet)$ be the $L^2$-complex of a singular subvariety $Z$ (the complex of sheaves of forms with locally summable coefficients on the nonsingular part $Z^\circ$ of $Z$ which are $L^2$\ together with their differentials near all points of $Z$, both smooth and singular). This is the sheafification of the presheaf assigning to each open set $U\subset Z$ the domain of the maximal closed extension of the differential $d$ on the Hilbert space of the $L^2$-forms on $U\cap Z^\circ$; there is another flavor of the $L^2$-complex constructed in a similar way by sheafification of the {\em minimal\/} closed extension, see details in~\cite{Y}, \S2.3. The complex $\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ consists of $(p,q)$-forms $\omega^{pq}$ on $Z^\circ$ (for any $q$) which have the following properties: $\omega^{pq}$ and $\overline{\partial}\omega^{pq}$ are $L^2$\ and moreover, there exists a form $\omega=\omega^{pq}+\omega^{p+1,q-1}+\dots$ such that both $\omega$ and $d\omega$ are $L^2$. The differential in the complex $\operatorname{Gr}^F_p{\cal M}^\bullet$ is the operator $\overline{\partial}$. Let us {\em assume\/} that this is a closed extension of $\overline{\partial}$. (To be precise, this means that the sections of this sheaf over an open set $U\subset Z$ form a closed extension of $\overline{\partial}$ in the Fr\'echet space of forms on $U\cap Z^\circ$ which are $L^2$\ locally in a neighborhood of any point of $U$; the topology on this Fr\'echet space is given by the seminorms $\|\bullet\|_K$ where $K$ is a relatively compact open subset of $U$; for a form $\omega$, the value $\|\omega\|_K$ is the $L^2$\ norm of $\omega$ on $K$.) In such case the complex $\operatorname{Gr}^F_p{\cal M}^\bullet$ can be viewed as an ``ideal boundary condition'' (the notion due to J. Cheeger) for the operator $\overline{\partial}$ at the singularities of $Z$; this complex contains the minimal closed extension of $\overline{\partial}$ and is contained in the maximal one (their sheafifications can be defined in a way similar to those of the operator $d$). If the operator $d$ on $L^2$\ forms has the property that its minimal extension coincides with the maximal one (this is called the {\em $L^2$ Stokes property\/}~\cite{C}, \cite{Y}; it is known for conical singularities~\cite{C} and seems from~\cite{O3} to be a reasonable conjecture in general) then, under our assumptions, it is not hard to see that the boundary condition of $\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ is dual to the boundary condition of $\operatorname{Gr}^F_{-p'}{\cal M}^\bullet$ if $p+p'=\dim Z$. (More precisely, this means the following. For any open $U\subset Z$, the dual to the Fr\'echet space of forms which are locally L2 on $U$ --- with the topology described above --- can be identified by the pairing $<\omega,\phi>=\int_{U\cap Z^\circ}\omega\wedge\phi$ with the DF space of forms $\phi$ on $U\cap Z^\circ$ such that the closure of $\operatorname{Supp}\phi$ in $U$ is compact. The $L^2$ Stokes property means that the adjoint of maximal extension of $d$ in the first space is, up to sign, the maximal extension of $d$ in the second one. The differential $\overline{\partial}$ on the sections of $\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ on $U$ is an unbounded operator on the subspace of $(p,q)$-forms of the above Fr\'echet space; similarly, the differential $\overline{\partial}$ on the sections of $\operatorname{Gr}^F_{-p'}{\cal M}^\bullet$ with compact support is an unbounded operator on the subspace of $(p',q)$-forms of the above DF space. The duality between the boundary conditions means that these two operators are adjoint up to sign.) It is easy to see that under our assumptions, the differential in $\operatorname{Gr}^F_{-\dim Z}{\cal M}^\bullet$ is the maximal closed extension of $\overline{\partial}$; it follows by duality that the differential in $\operatorname{Gr}^F_0{\cal M}^\bullet$ is the minimal closed extension of $\overline{\partial}$. In case the metric on $Z^\circ$ is complete (e.g., Saper metric), the minimal closed extension of $\overline{\partial}$ is known to coincide with the maximal one, and hence, $\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ is just the domain of $\overline{\partial}$ with any of the boundary conditions. In case the metric is incomplete (e.g., the restriction of the Fubini---Studi metric on the projective space to $Z^\circ$), it is known that the minimal closed extension of $\overline{\partial}$ may be different from the maximal one~\cite{P}. The results of~\cite{PS} and~\cite{FH} suggest that in case $p=0$ the ``correct'' boundary condition for $\overline{\partial}$ is the minimal (Dirichlet) one, and in case $p=\dim Z$ the ``correct'' boundary condition for $\overline{\partial}$ is the maximal (Neumann) one. This suggests that under our assumptions, the complex $\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ is the most natural boundary condition for the operator $\overline{\partial}$, and its cohomology $H^j\operatorname{Gr}^F_{-p}{\cal M}^\bullet$ is the most natural notion of the $L^2$-$\overline{\partial}$-cohomology sheaves.
1996-07-23T20:38:11
9607
alg-geom/9607025
en
https://arxiv.org/abs/alg-geom/9607025
[ "alg-geom", "math.AG" ]
alg-geom/9607025
Rahul Pandharipande
R. Pandharipande
The Chow Ring of the Hilbert Scheme of Rational Normal Curves
24 pages, Latex2e
null
null
null
null
Let H(d) be the (open) Hilbert scheme of rational normal curves of degree d in P^d. A presentation of the integral Chow ring of H(d) is given via equivariant Chow ring computations. Included also in the paper are algebraic computations of the integral equivariant Chow rings of the algebraic groups O(n), SO(2k+1). The results for S0(3)=PGL(2) are needed for the Hilbert scheme calculation.
[ { "version": "v1", "created": "Tue, 23 Jul 1996 18:30:38 GMT" } ]
2008-02-03T00:00:00
[ [ "Pandharipande", "R.", "" ] ]
alg-geom
\section{\bf{Introduction}} \subsection{Summary} Let $\mathbb{C}$ be the ground field of complex numbers. A rational normal curve in $\mathbf P^d$ is an irreducible, nonsingular, non-degenerate, degree $d$ rational curve. For $d\geq 1$, let $H(d)$ be the open Hilbert scheme of rational normal curves of degree $d$ in $\mathbf P^d$. $H(d)$ is a nonsingular, irreducible, quasi-projective, algebraic variety. Let $A^*(d)$ be the integral Chow ring of $H(d)$. In case $d=1$, there is a unique rational normal curve in $\mathbf P^1$. Hence, $H(1)$ is a point. $H(2)$ is the space of nonsingular plane conics. The dimension of $H(d)$ is $d^2+2d-3$. In this paper, a presentation of $A^*(d)$ is computed via the theory of equivariant Chow groups. The idea is to exhibit $H(d)$ as a quotient of an appropriate variety $X$ by a free $\mathbf{G}$-action. For free actions, the equivariant Chow ring $A^*_{\mathbf{G}}(X)$ is isomorphic to the ordinary Chow ring $A^*(d)$ of the quotient $X/\mathbf{G}\stackrel{\sim}{=} H(d)$. The equivariant Chow ring $A^*_{\mathbf{G}}(X)$ is then computed in the required cases via Chow rings of projective bundles and Chow ideals of degeneracy loci. The geometry of $H(d)$ depends significantly on the parity of $d$. The quotient approaches and the presentations of $A^*(d)$ differ for $d$ even and odd. In the even case, a $\mathbf{PGL}(2)$-quotient approach is taken. The geometry of algebraic $B\mathbf{PGL}(2)$ is studied as a necessary first step. There is an isomorphism of linear algebraic groups: $\mathbf{PGL}(2) \stackrel{\sim}{=} \mathbf{SO}(3)$. The space $B \mathbf{SO}(3)$ is analyzed via conic geometry in projective space. It is no more difficult to study algebraic $B \mathbf{O}(n)$ and $B \mathbf{SO} (n=2k+1)$ via higher dimensional quadrics. The equivariant Chow rings of these two series are computed. The $d$ odd case is simpler. In this case, the quotient group is taken to be a central extension of $\mathbf{SL}(2)$ in $\mathbf{GL}(2)$. Presentations of $A^*(d)$ in case $d$ is even and odd are determined in Theorems \ref{heven} and \ref{hodd} respectively. The equivariant Chow rings of the groups $\mathbf{O}(n)$ an $\mathbf{SO}(n=2k+1)$ are computed in Theorem \ref{chor}. The equivariant Chow ring of $\mathbf{O}(n)$ was first determined by B. Totaro using complex cobordism theory and topology. After the intial algebraic calculation of the ring of $\mathbf{SO}(3)$ presented here, it was realized both Totaro's methods and the $\mathbf{SO}(3)$ computation generalize to $\mathbf{SO}(2k+1)$. An algebraic approach to $B \mathbf{SO} (3)$ is required for the application to Theorem 1. In [P], Chow rings (with $\mathbb{Q}$-coefficients) of certain moduli spaces of maps are computed via equivariant Chow groups. The integral computations presented here were motivated by the calculations in [P]. These arguments show the equivariant constructions in [T] and [EG] can be used effectively to compute ordinary Chow rings of quotients. \subsection{Presentations of $A^*(d)$} \label{prezz} Equivariant Chow theory is reviewed in section \ref{chow}. Let $\mathbf{G}$ be a reductive algebraic group. Let $\mathbf{G}\times X \rightarrow X$ be a linearized algebraic group action on a nonsingular quasi-projective variety $X$. An equivariant Chow ring $A^*_\mathbf{G}(X)$ is defined via algebraic approximations to $E\mathbf{G}$ and $B \mathbf{G}$. Let $V$ be a fixed $2$-dimensional $\mathbb{C}$-vector space. Let $\mathbf P^1\stackrel{\sim}{=} \mathbf P(V)$. There is a canonical isomorphism $H^0(\mathbf P^1, {\mathcal{O}}_{\mathbf P^1}(d)) \stackrel{\sim}{=} Sym^{d}(V^*)$. Let $$U\subset \bigoplus_{0}^{d} Sym^d(V^*)$$ denote the non-degenerate locus (this is the open set consisting of linearly independent $(d+1)$-tuples of vectors of $Sym^d(V^*)$). $U$ parameterizes bases of the linear series of ${\mathcal{O}}_{\mathbf P^1}(d)$ on $\mathbf P^1$. There is a canonical $\mathbf{GL} (V)$-action on $U$ with geometric quotient $H(d)$. The required existence results for the algebraic quotient problems encountered in this paper are developed in the Appendix (section \ref{appx}). $\mathbf{GL} (V)$ acts with finite stabilizers on $U$ (the stabilizer of a point $u\in U$ is the subgroup of scalar $d^{th}$ roots of unity), By a theorem of D. Edidin and W. Graham ([EG]), there is a canonical isomorphism of graded rings $$A^*(d) \otimes_{\mathbb{Z}}{\mathbb{Q}} \stackrel{\sim}{=} A^*_{\mathbf{GL} (V)}(U) \otimes _{\mathbb{Z}}{\mathbb{Q}}.$$ The equivariant Chow ring $A^*_{\mathbf{GL} (V)}(U)$ is determined in section \ref{abe}. $A^*_{\mathbf{GL} (V)}(U)$ is generated (as a ring) in codimensions $1$, $2$ by elements $c_1$, $c_2$ respectively. There are $d+1$ relations given as follows. Let $S$ be a rank $2$ bundle with Chern classes $c_1$ and $c_2$. The $d+1$ Chern classes of $Sym^d(S)$ are the relations. It is not difficult to see $A^i_{\mathbf{GL} (V)}(U) \otimes_{\mathbb{Z}} \mathbb{Q}= 0$ for $i>0$. \begin{pr} \label{alltor} $A^*(d)$ is torsion in codimension $i>0$. \end{pr} \noindent Note that $\mathbf{GL}(Sym^d (V^*))$ acts transitively on $U$ and $H(d)$ is a homogeneous space for $\mathbf{GL}(Sym^d(V^*))=\mathbf{GL}(d+1)$. Let $\mathbf P(U) \subset \mathbf P(\bigoplus_{0}^{d} Sym^d(V^*))$ be the projective non-degenerate locus. $\mathbf P(U)$ is exactly the space of parameterized rational normal curves. There is a canonical $\mathbf{PGL} (V)$-action on $\mathbf P(U)$ with geometric quotient $H(d)$. This is a free action. Hence, there is a canonical isomorphism of graded rings (see [EG]): $$A^*(d) \stackrel{\sim}{=} A^*_{\mathbf{PGL} (V)}(\mathbf P(U)).$$ Assume $d\geq 2$ is even. Let $d=2n$ (where $n\geq 1$). $\mathbf P(U) \rightarrow H(d)$ is a principal $\mathbf{PGL}(V)$-bundle. Let $S$ be the rank $3$ algebraic vector bundle on $H(d)$ obtained from the principal bundle $\mathbf P(U) \rightarrow H(d)$ and the representation $Sym^2 (V)$ of $\mathbf{PGL}(V)$. A discussion of algebraic principal bundles can be found in the Appendix. For $1 \leq i \leq 3$, let $c_i\in A^*(d)$ be the Chern classes of $S$. Let $\mathcal{H}\in A^1(d)$ be the divisor class of curves meeting a fixed codimension $2$ linear space in $\mathbf P^d$. Let $\mathcal{L}= n \mathcal{H}$. In section \ref{evan}, the equivariant Chow ring $A^*_{\mathbf{PGL} (V)}(\mathbf P(U))$ is evaluated in the even case. \begin{tm} \label{heven} $A^*(d=2n)$ is generated by $c_1$, $c_2$, $c_3$, and $\mathcal{L}$. The first relations are: $$ c_1 =0$$ $$2c_3=0.$$ There are $d+1$ additional relations given by the first $d+1$ Chern classes of the formal expansion: $$ \frac{(1+\mathcal{L})^{d+1} \cdot c(Sym^{n-2} (S))} {c(Sym^{n} (S))}.$$ (If $n=1$ or $2$, then $c(Sym^{n-2}(S))= 1$.) \end{tm} \noindent It is easily seen from Theorem \ref{heven} that $A^1(d=2n) \stackrel{\sim}{=} \mathbb{Z}/(d+1)\mathbb{Z}$ with generator $\mathcal{L}$. The equation $\mathcal{L} = n\mathcal{H}$ can then be uniquely solved to obtain $\mathcal{H}=2d \mathcal{L} =-2 \mathcal{L}$. Now assume $d\geq 1$ is odd. Let $d=2n-1$ (where $n\geq 1$). Let $$det: \mathbf{GL} (V) \rightarrow \mathbb{C}^*$$ be the determinant homomorphism. Let $\mathbb{Z}/n\mathbb{Z} \subset \mathbb{C}^*$ be the subgroup of the $n^{th}$ roots of unity. Let $\mathbf{SL} (V,n)= det^{-1}(\mathbb{Z}/n\mathbb{Z})$. Consider again $\bigoplus_0^{2n-1} Sym^{2n-1}(V^*)$. There is a canonical, $\mathbf{GL} (V)$-equivariant, multilinear map $$\mu:\bigoplus_0^{2n-1} Sym^{2n-1}(V^*) \rightarrow \bigwedge^{2n} Sym^{2n-1}(V^*)$$ given by the exterior product: $$(\omega_0, \omega_1, \ldots, \omega_{2n-1}) \mapsto \omega_0 \wedge \omega_1 \wedge \ldots \wedge \omega_{2n-1}.$$ $\mathbf{SL} (V,n)$ acts trivially on the 1 dimensional space $\bigwedge^{2n} Sym^{2n-1}(V^*)$. Let $Y= \mu^{-1} (p)$ where $0 \neq p \in\bigwedge^{2n} Sym^{2n-1}(V^*)$. There is an $\mathbf{SL} (V,n)$-action on $Y$. In Lemma \ref{freeaq}, it is shown this is a free action with geometric quotient $H(d)$. Hence, there is a canonical isomorphism of graded rings $$A^*(d=2n-1) \stackrel{\sim}{=} A^*_{\mathbf{SL} (V,n)}(Y).$$ Let $S$ now denote the rank $2$ algebraic vector bundle obtained from the principal $\mathbf{SL}(V,n)$-bundle $Y \rightarrow H(d)$ and the standard representation $V$. For $1\leq i \leq 2$, let $c_i\in A^*(d)$ be the Chern classes of $S$. The equivariant Chow ring $A^*_{\mathbf{SL} (V,n)}(Y)$ is evaluated in section \ref{ode}. \begin{tm} $A^*(d=2n-1)$ is generated by $c_1$ and $c_2$. The first relation is $$ nc_1 =0.$$ There are $d+1$ additional relations given by the first $d+1$ Chern classes of $Sym^d(S)$. \label{hodd} \end{tm} \noindent It is easily seen that $A^1(d=2n-1) \stackrel{\sim}{=} \mathbb{Z}/n\mathbb{Z}$. \subsection{Chow rings of the Orthogonal Groups} The Chow ring of a reductive algebraic group $\mathbf{G}$ is, by definition, the equivariant Chow ring $A^*_\mathbf{G}( \text{point})$. Let $\mathbf{O} (n)$ and $\mathbf{SO} (n)$ denote the orthogonal and special orthogonal algebraic groups. The equivariant calculations of Theorem \ref{heven} require knowledge of $B\mathbf{PGL} (2)$. $\mathbf{PGL} (2)$ is isomorphic to $\mathbf{SO} (3)$. The following Theorem will be established: \begin{tm} \label{chor} The integral Chow ring of $\mathbf{O} (n)$ is generated by the Chern classes $c_1, \ldots, c_n$ of the standard representation. The odd classes are 2-torsion: $$A^*_{\mathbf{O} (n)}(\text{point})= \mathbb{Z}[c_1, \ldots, c_n]/(2c_1, 2c_3, 2c_5, \ldots).$$ \noindent The integral Chow ring of $\mathbf{SO} (n=2k+1)$ is generated by the Chern classes $c_1, \ldots, c_n$ of the standard representation. The odd classes are 2-torsion and $c_1=0$: $$A^*_{\mathbf{SO} (n)}(\text{point})= \mathbb{Z}[c_1, \ldots, c_n]/(c_1, 2c_3, 2c_5, \ldots).$$ \end{tm} \noindent The Chow ring of $\mathbf{SO} (2k)$ is not generated by the Chern classes of the standard representation. The main difference in the odd and even cases is that $\mathbf{SO} (2k+1) \stackrel{\sim}{=} \mathbb{Z}/2\mathbb{Z} \times \mathbf{O}(2k+1)$ while such a product decomposition does not hold for $\mathbf{SO} (2k)$. The methods of this paper do not yield a computation of $A^*_{\mathbf{SO} (2k)}( \text{point})$. The Chow ring of $\mathbf{SO}(n)$ has been computed with $\mathbb{Q}$-coefficients in [EG2]. \subsection{Acknowledgments} Equivariant Chow groups were first defined in [T]. Thanks are due to D. Edidin, W. Graham, and B. Totaro for conversations in which the theory of equivariant Chow groups was explained. The author particularly wishes to thank B. Totaro for his insights on $\mathbf{O}(n)$ and $\mathbf{SO}(n)$. Discussions with W. Fulton on many related issues have also been helpful. \section{\bf{Chow Ideals of Degeneracy Loci}} \subsection{Presentations} \label{idealz} For the Chow computations in this paper, presentations of four ideals associated to tautological degeneracy loci are needed. Let $E$ be a rank $e$ vector bundle on a nonsingular algebraic variety $M$. We will consider two affine and two projective fibrations over $M$: \begin{enumerate} \item[(i)] $\oplus_{1}^e E \rightarrow M$, \item[(ii)] $\mathbf P(\oplus_{1}^e E) \rightarrow M$, \item[(iii)] $Sym^2 E^* \rightarrow M$, \item[(iv)] $\mathbf P(Sym^2 E^*) \rightarrow M$. \end{enumerate} The subspace projectivization is taken in (ii) and (iv). Let $r=e^2$ denote the rank of $\oplus_{1}^{e} E$. Let $L$ in $A^1(\mathbf P(\oplus_{1}^{e} E))$ be the class of ${\mathcal{O}}_{\mathbf P}(1)$ obtained from the projectivization. The Chow ring of $\mathbf P(\oplus_{1}^{e} E)$ has a standard presentation: $$A^*(M)[L]\ /\ \big(L^{r}+ c_1(\oplus_{1}^{e} E)\cdot L^{r-1} + \ldots +c_{r}(\oplus_{1}^{e} E)\big).$$ Similarly, the Chow ring of $\mathbf P(Sym^2 E^*)$ has a presentation: $$A^*(M)[L]\ /\ \big(L^{s}+ c_1(Sym^2 E^*)\cdot L^{s-1} + \ldots +c_{s}(Sym^2 E^*)\big)$$ where $s=\frac{1}{2}(e^2+e)$ is the rank of $Sym^2 E^*$ and $L$ is again the class of ${\mathcal{O}}_{\mathbf P}(1)$ obtained from the projectivization. The Chow rings of the affine fibrations (i) and (iii) are canonically isomorphic to $A^*(M)$. There are intrinsic, fiberwise degeneracy loci in these fibrations. Let $D_1 \subset \oplus_{1}^{e} E$ and $\mathbf P(D_1) \subset \mathbf P(\oplus_{1}^{e} E)$ be the closed subvariety of linearly dependent $e$-tuples of vectors in the fibers of $E$. Let $D_2 \subset Sym^2 E^*$ and $\mathbf P(D_2) \subset \mathbf P(Sym^2 E^*)$ be the closed subvariety of degenerate quadratic forms on the fibers of $E$. Let $$I_1\subset A^*(\oplus_{1}^{e} E)\stackrel{\sim}{=} A^*(M), \ \ J_1\subset A^*(\mathbf P(\oplus_{1}^{e} E)),$$ $$I_2 \subset A^*(Sym^2 E^*)\stackrel{\sim}{=} A^*(M), \ \ J_2 \subset A^*(\mathbf P(Sym^2 E^*))$$ be the ideals generated by classes supported on the degeneracy loci $D_1$, $\mathbf P(D_1)$, $D_2$, and $\mathbf P(D_2)$ respectively. In this section, simple sets of generators of the ideals $I_1$, $J_1$, $I_2$, and $J_2$ are determined. The results of this section are essentially special cases of Pragacz's presentations of the ideals of Chow classes supported on degeneracy loci of bundle maps ([Pr]). Pragacz considers more general degeneracy loci and obtains presentations of their universal Chow ideals via Schur $S$-polynomials. Actual (not universal) Chow ideal presentations are needed here. Since the geometry of the cases (i)-(iv) is particularly simple, the actual and the universal presentations coincide. A full proof will be given here. For a rank $f$ bundle $F$, let $c(F)=1+c_1(F)+ \ldots +c_f(F)$. \begin{lm} \label{petey} $I_1\subset A^*(M)$ is generated by $(\alpha_1, \ldots, \alpha_e)$ where $$\frac{1}{c(E^*)}= 1+ \alpha_1 + \ldots + \alpha_e+ \ldots.$$ \end{lm} \begin{lm} \label{pete} $J_1\subset A^*(\mathbf P(\oplus_{1}^{e} E))$ is generated by $(\alpha'_1, \ldots, \alpha'_e)$ where $$\frac{c(\oplus_{1}^{e} {\mathcal{O}}_{\mathbf P}(1))}{c(E^*)}= 1+ \alpha'_1 + \ldots + \alpha'_e+ \ldots.$$ \end{lm} \begin{lm} \label{paul} $I_2\subset A^*(M)$ is generated by $(\beta_1, \ldots, \beta_e)$ where $$\frac{c(E^*)}{c(E)}= 1+ \beta_1 + \ldots + \beta_e+ \ldots.$$ \end{lm} \begin{lm} \label{pauly} $J_2\subset A^*(\mathbf P(Sym^2 E^*))$ is generated by $(\beta'_1, \ldots, \beta'_e)$ where $$\frac{c(E^*\otimes {\mathcal{O}}_{\mathbf P}(1))}{c(E)}= 1+ \beta'_1 + \ldots + \beta'_e+ \ldots.$$ \end{lm} \noindent The proofs of Lemmas \ref{petey} -- \ref{pauly} are essentially the same. The first step is to find a tower of bundles dominating the degeneracy loci $D_1$, $\mathbf P(D_1)$, $D_2$, and $\mathbf P(D_2)$. First consider $D_1$ and $\mathbf P(D_1)$. Let $\eta:\mathbf P(E^*) \rightarrow M$ be the projective bundle. A point $\xi \in \mathbf P(E^*)$ is a pair $(m,h)$ where $m\in M$ and $h \in \mathbf P(E^*_m)$. Let $B$ be the vector bundle on $\mathbf P(E^*)$ determined as follows. The fiber of $B$ at the point $(m,h)$ is the linear subspace of $\oplus_{1}^{e} E_m$ consisting of $e$-tuples of vectors annihilated by $h$. $B$ is a sub-bundle of $\eta^*(\oplus_{1}^{e} E)$. There are canonical, proper, surjective projections: $$\rho: B \rightarrow D_1 \subset \oplus_{1}^{e} E,$$ $$\mathbf P(\rho): \mathbf P(B) \rightarrow \mathbf P(D_1) \subset \mathbf P(\oplus_{1}^{e} E).$$ There are stratifications of $D_1$ and $\mathbf P(D_1)$ by the rank of the span of the $e$-tuple of vectors. Over these strata, $\rho$ and $\mathbf P(\rho)$ are projective bundles. Hence $\rho$ and $\mathbf P(\rho)$ induce {\em surjections} on the integral Chow rings via push-forward: $$\rho_*: A^*(B) \rightarrow A^*(D_1),$$ $$\mathbf P(\rho)_*: A^*(\mathbf P(B)) \rightarrow A^*(\mathbf P(D_1)).$$ Lemmas \ref{petey} and \ref{pete} are proven by computing the images of the generators of $A^*(B)$ and $A^*(\mathbf P(B))$ respectively. Consider the commuting diagrams: \begin{equation} \label{heyhey} \begin{CD} B @>{\rho}>> D_1 \subset \oplus_{1}^{e} E \\ @V{\pi}VV @VVV \\ \mathbf P(E^*) @>{\eta}>> M \\ \end{CD} \end{equation} \begin{equation} \label{heyheyy} \begin{CD} \mathbf P(B) @>{\mathbf P(\rho)}>> \mathbf P( D_1) \subset \mathbf P(\oplus_{1}^{e} E) \\ @V{\pi}VV @VVV \\ \mathbf P(E^*) @>{\eta}>> M \\ \end{CD} \end{equation} $A^*(B)$ is generated over $A^*(M)$ by the class corresponding to ${\mathcal{O}}_{\mathbf P(E^*)}(1)$. Let this class be denoted by $\zeta$. It follows that $$I_1=(\rho_{*}(1), \rho_{*}(\zeta^1), \rho_{*}(\zeta^2), \ldots, \rho_{*}(\zeta^{e-1})).$$ Similarly, there is a presentation of $J_1$: $$J_1=(\mathbf P(\rho)_*(1), \mathbf P(\rho)_{*}(\zeta^1), \mathbf P(\rho)_{*}(\zeta^2), \ldots, \mathbf P(\rho)_{*}(\zeta^{e-1})).$$ To prove Lemmas \ref{petey} and \ref{pete}, it is sufficient to establish the equalities \begin{equation} \label{eqql} \rho_{*}(\zeta^{i-1})= \alpha_i, \ \ \mathbf P(\rho)_{*}(\zeta^{i-1})= \alpha'_i \end{equation} for $1 \leq i \leq e$. First the equalities (\ref{eqql}) for Lemma \ref{petey} are proven. By definition, $B\subset \eta^*(\oplus_{1}^{e} E)$. In fact, there is a natural exact squence on $\mathbf P(E^*)$: \begin{equation} \label{yess1} 0 \rightarrow B \rightarrow \eta^*(\oplus_{1}^{e} E) \rightarrow \oplus_{1}^{e} {\mathcal{O}}_{\mathbf P(E^*)}(1) \rightarrow 0. \end{equation} As a first step, the class of $[B] \in A^*(\eta^*(\oplus_{1}^{e} E))$ is computed. Since $\eta^*(\oplus_{1}^{e} E)$ is a projective bundle over $\oplus_{1}^{e} E$, $A^*(\eta^*(\oplus_{1}^{e} E))$ is generated over $A^*(M)$ by $\zeta$ (which satisfies the Chern relation). By sequence (\ref{yess1}) and Lemma \ref{fullt} below, it follows that $[B] = \zeta ^{e} \in A^*(\eta^*(\oplus_{1}^{e} E))$. Denote the natural projection $\eta^*(\oplus_{1}^{e} E)) \rightarrow \oplus_{1}^{e} E$ by $\phi$. There is a fundamental equality: $$\rho_{*}(\zeta ^{i-1}) = \phi_{*} (\zeta^{i-1} \cap [B]) \in A^*(M).$$ The right side is easy to calculate. $$\phi_{*}(\zeta^{i-1}\cap [B])= \phi_{*}(\zeta^{e-1+i}).$$ For $1\leq i \leq e$, the latter is simply the $i^{th}$ Segre class of $E^*$. Lemma \ref{petey} is proved. Lemma \ref{pete} is only slightly more complicated. The class of $[\mathbf P(B)] \in A^*(\mathbf P(\eta^*(\oplus_{1}^{e} E)))$ is computed. Again $A^*(\mathbf P(\eta^*(\oplus_{1}^{e} E)))$ is generated over $A^*(\mathbf P(\oplus_{1}^{e} E))$ by $\zeta$. By sequence (\ref{yess1}) and Lemma \ref{fullt}, it follows that $[\mathbf P(B)] = (L+\zeta) ^{e} \in A^*(\mathbf P(\eta^*(\oplus_{1}^{e} E)))$ where $L$ is the class of ${\mathcal{O}}_{\mathbf P(\oplus_{1}^{r} E)}(1)$. Denote the natural projection $\mathbf P(\eta^*(\oplus_{1}^{e} E)) \rightarrow \mathbf P(\oplus_{1}^{e} E)$ by $\mathbf P(\phi)$. There are equalities: $$\mathbf P(\rho)_{*}(\zeta ^{i-1}) = \mathbf P(\phi)_{*} (\zeta^{i-1} \cap [B]) =\mathbf P(\phi)_{*} (\zeta^{i-1}\cap (L+\zeta)^e)$$ in $A^*(\mathbf P (\oplus_{1}^{e} E))$. Lemma \ref{ssegre} now yields Lemma \ref{pete}. The degeneracy loci $D_2$ and $\mathbf P(D_2)$ are considered next. The notation will parallel the notation used in the proofs of Lemmas \ref{petey} and \ref{pete}. Let $\eta:\mathbf P(E) \rightarrow M$ be the projective bundle. A point $\xi \in \mathbf P(E)$ is a pair $(m,p)$ where $m\in M$ and $p \in \mathbf P(E_m)$. Let $B$ be the vector bundle on $\mathbf P(E)$ determined as follows. The fiber of $B$ at the point $(m,p)$ is the linear subspace of quadratic forms on $E_m$ singular at $p$. $B$ is a sub-bundle of $\eta^*(Sym^2 E^*)$. There are canonical, proper, surjective projections: $$\rho: B \rightarrow D_2 \subset Sym^2 E^*,$$ $$\mathbf P(\rho): \mathbf P(B) \rightarrow \mathbf P(D_2) \subset Sym^2 E^*.$$ There are stratifications of $D_2$ and $\mathbf P(D_2)$ by the rank of the quadratic form. Over these strata, $\rho$ and $\mathbf P(\rho)$ are projective bundles. Hence $\rho$ and $\mathbf P(\rho)$ induce {\em surjections} on the integral Chow rings via push-forward: $$\rho_*: A^*(B) \rightarrow A^*(D_2),$$ $$\mathbf P(\rho)_*: A^*(\mathbf P(B)) \rightarrow A^*(\mathbf P(D_2)).$$ Lemmas \ref{paul} and \ref{pauly} are proven by computing the images of the generators of $A^*(B)$ and $A^*(\mathbf P(B))$ respectively. As before, $$I_2=(\rho_{*}(1), \rho_{*}(\zeta^1), \rho_{*}(\zeta^2), \ldots, \rho_{*}(\zeta^{e-1})),$$ $$J_2= (\mathbf P(\rho)_{*}(1), \mathbf P(\rho)_{*} (\zeta^1), \mathbf P(\rho)_{*}(\zeta^2), \ldots, \mathbf P(\rho)_{*}(\zeta^{e-1})).$$ To prove Lemma \ref{paul} and \ref{pauly}, it is sufficient to establish the equalities \begin{equation} \label{eqql2} \rho_{*}(\zeta^{i-1})= \beta_i, \ \ \mathbf P(\rho)_{*}(\zeta^{i-1})= \beta'_i \end{equation} for $1 \leq i \leq e$. First the equalities (\ref{eqql2}) for Lemma \ref{paul} are proven. There is an exact squence on $\mathbf P(E)$: \begin{equation} \label{yess2} 0 \rightarrow B \rightarrow \eta^*(Sym^2 E^*) \rightarrow E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1) \rightarrow 0. \end{equation} The class of $[B] \in A^*(\eta^*(\oplus_{1}^{e} E))$ is computed. $A^*(\eta^*(Sym^2 E^*))$ is generated over $A^*(M)$ by $\zeta$. By sequence (\ref{yess2}) and Lemma \ref{fullt} below, it follows that $$[B] = c_e (E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1) )\in A^*(\eta^*(Sym^2 E)).$$ Denote the natural projection $\eta^*(Sym^2 E^*)) \rightarrow Sym^2 E^*$ by $\phi$. There is an equality: $$\rho_{*}(\zeta ^{i-1}) = \phi_{*} (\zeta^{i-1} \cap [B]) \in A^*(M).$$ Lemma \ref{ssegre} now yields Lemma \ref{paul}. Lemma \ref{pauly} is established next. By sequence (\ref{yess2}) and Lemma \ref{fullt} below, it follows that $$[\mathbf P(B)] = c_e \bigg( \frac{c(E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1))} {c({\mathcal{O}}_{\mathbf P(Sym^2 E^*)}(-1))} \bigg)\in A^*(\mathbf P(\eta^*(Sym^2 E^*))).$$ There is an equality (since $E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1)$ is a rank $e$ bundle): $$c_e \bigg( \frac{c(E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1))} {c({\mathcal{O}}_{\mathbf P(Sym^2 E^*)}(-1))} \bigg)= c_e(E^*\otimes {\mathcal{O}}_{\mathbf P(E)}(1) \otimes {\mathcal{O}}_{\mathbf P(Sym^2 E^*)}(1)).$$ Denote the natural projection $\mathbf P(\eta^*(Sym^2 E^*)) \rightarrow \mathbf P(Sym^2 E^*)$ by $\mathbf P (\phi)$. There is an equality: $$\mathbf P(\rho)_{*}(\zeta ^{i-1}) = \mathbf P(\phi)_{*} (\zeta^{i-1} \cap [B]) \in A^*(\mathbf P(Sym^2 E^*)).$$ Lemma \ref{ssegre} now yields Lemma \ref{pauly}. \subsection{Lemmas} The following Lemmas were used in the proofs of Lemmas \ref{petey} -- \ref{pauly}. Let $F \rightarrow N$ be a vector bundle on a nonsingular algebraic variety $N$. \begin{lm} \label{fullt} Let $\ 0 \rightarrow B \rightarrow F \rightarrow Q \rightarrow 0$ be an exact sequence of bundles on $N$. Let $q$ be the rank of $Q$. The class $[B]\in A^*(F)\stackrel{\sim}{=} A^*(N)$ is determined by $$[B]= c_q(Q).$$ The class $[\mathbf P(B)] \in A^*(\mathbf P(F))$ is determined by $$[\mathbf P(B)]= c_q \bigg(\frac {c(Q)}{c({\mathcal{O}}_{\mathbf P(F)}(-1))} \bigg).$$ \end{lm} \noindent {\em Proof.} This is an application of the Thom-Porteous formulas for degeneracy loci of bundle maps (see [F]). \qed \vspace{+10pt} \noindent Let $f$ be the rank $F$. Let $\phi: \mathbf P(F) \rightarrow N$ be the projection. \begin{lm} \label{ssegre} Let $G$ be a bundle of rank $g=f$ on $N$. Let $\zeta= c_1({\mathcal{O}}_{\mathbf P(F)}(1))$. Let $\gamma_{i}$ be determined by $$\frac{c(G)}{c(F)} = 1 + \gamma_{1} + \ldots+ \gamma_{f} + \ldots.$$ Then, for $1 \leq i \leq f$, $\gamma_{i}= \phi_{*}\big( \zeta^{i-1} \cap c_f(G \otimes {\mathcal{O}}_{\mathbf P(F)}(1))\big).$ \end{lm} \noindent {\em Proof.} A simple Segre class argument yields the result. \qed \vspace{+10pt} \section{\bf{Equivariant Chow Groups}} \label{chow} Let $\mathbf{G}$ be a group. Let $\mathbf{G}\times X \rightarrow X$ be a left group action. In topology, the $\mathbf{G}$-equivariant cohomology of $X$ is defined as follows. Let $E\mathbf{G}$ be a contractible topological space equipped with a free left $\mathbf{G}$-action and quotient $E\mathbf{G}/\mathbf{G}=B\mathbf{G}$. Consider the left action of $\mathbf{G}$ on $X\times E\mathbf{G}$ defined by: $$g(x,b)= (g(x), g(b)).$$ $\mathbf{G}$ acts freely on $X\times E\mathbf{G}$. Let $X\times^{\mathbf{G}} E\mathbf{G}$ be the (topological) quotient. The $\mathbf{G}$-equivariant cohomology of of $X$, $H_\mathbf{G}^*(X)$, is defined by: $$H_\mathbf{G}^*(X) = H^*_{sing}(X\times^{\mathbf{G}} E\mathbf{G}).$$ If $X$ is a locally trivial principal $\mathbf{G}$-bundle, then $X\times^{\mathbf{G}} E\mathbf{G}$ is a locally trivial fibration of $E\mathbf{G}$ over the quotient $X/\mathbf{G}$. In this case, $X\times^{\mathbf{G}} E\mathbf{G}$ is homotopy equivalent to $X/\mathbf{G}$ and $$H_\mathbf{G}^*(X) = H^*_{sing}(X\times^{\mathbf{G}} E\mathbf{G}) \stackrel{\sim}{=} H^*_{sing}(X/\mathbf{G}).$$ For principal bundles, computing the equivariant cohomology ring is equivalent to computing the cohomology of the quotient. There is an analogous equivariant theory of Chow groups developed by B. Totaro in case $X$ is a point and generalized by D. Edidin and W. Graham to arbitrary $X$ ([T], [EG]). Let $\mathbf{G}$ be a reductive algebraic group. Let $\mathbf{G}\times X \rightarrow X$ be a linearized algebraic $\mathbf{G}$-action. The algebraic analogue of $E\mathbf{G}$ is attained by approximation. Let $V$ be a $\mathbb{C}$-vector space. Let $\mathbf{G}\times V \rightarrow V$ be an algebraic representation of $\mathbf{G}$. Let $W\subset V$ be a $\mathbf{G}$-invariant open set satisfying: \begin{enumerate} \item[(i)] The complement of $W$ in $V$ is of codimension greater than $q$. \item[(ii)] $\mathbf{G}$ acts freely on $W$ (see the Appendix for the definition). \item[(iii)] There exists a geometric quotient $W\rightarrow W/\mathbf{G}$. \end{enumerate} $W$ is an approximation of $E\mathbf{G}$ up to codimension $q$. By (iii) and the assumption of linearization, a geometric quotient $X\times ^{\mathbf{G}} W$ exists as an algebraic variety. Let $d=dim(X)$, $e=dim( X\times ^{\mathbf{G}} W)$. The equivariant Chow groups are defined by: \begin{equation} \label{defff} A^{\mathbf{G}}_{d-j}(X)= A_{e-j}(X\times ^{\mathbf{G}} W) \end{equation} for $0\leq j \leq q.$ An argument is required to check these equivariant Chow groups are well-defined (see [EG]). The basic functorial properties of equivariant Chow groups are established in [EG]. In particular, if $X$ is nonsingular, there is a natural intersection ring structure on $A_i^{\mathbf{G}}(X)$. Let $Z$ be a variety of dimension $z$. For notational convenience, a superscript will denote the Chow group codimension: $$A^{\mathbf{G}}_{z-j}(Z) = A^j_\mathbf{G}(Z), \ A_{z-j}(Z)=A^j(Z).$$ In particular, equation (\ref{defff}) becomes: $$\forall\ 0\leq j \leq q, \ \ A_{\mathbf{G}}^{j}(X)= A^j(X\times ^{\mathbf{G}} W).$$ The following result of [EG] will be used. \begin{pr} \label{dane} Let $\mathbb{C}$ be the ground field of complex numbers. Let $X$ be a quasi-projective variety. Let $\mathbf{G}$ be a reductive group. Let $\mathbf{G}\times X \rightarrow X$ be a linearized proper $\mathbf{G}$-action. Let $X\rightarrow X/\mathbf{G}$ be a quasi-projective geometric quotient. \begin{enumerate} \item[(i)] If the action is free, then there is a canonical isomorphism of graded rings: $$A^*_\mathbf{G}(X) \stackrel{\sim}{=} A^*(X/\mathbf{G}).$$ \item[(ii)] If $\mathbf{G}$ acts with finite stabilizers on $X$, then there is a canonical isomorphism of graded rings: $$ A^*_\mathbf{G}(X) \otimes \mathbb{Q} \stackrel{\sim}{=} A^*(X/\mathbf{G}) \otimes \mathbb{Q}.$$ \end{enumerate} \end{pr} \noindent Proposition \ref{dane} is a characteristic 0 specialization of Theorem 2 of [EG]. \section{\bf The Chow Rings of $\mathbf{O} (k)$ and $\mathbf{SO} (2k+1)$} \subsection{$B\mathbf{O} (V)$ and $B\mathbf{SO} (V)$} \label{orthoo} Let $V$ be a complex vector space equipped with a non-degenerate quadratic form. Let $\mathbf{O} (V)$, $\mathbf{SO} (V)$ be the orthogonal and special orthogonal groups respectively. Approximations to $E\mathbf{O} (V)$ and $E\mathbf{SO} (V)$ are obtained via direct sums of the representation $V^*$. Let $m>>0$ and let $$W_m \subset \oplus_{1}^{m} V^*$$ denote the spanning locus. $W_m$ is the locus of $m$-tuples of vectors of $ V^*$ which span $V^*$. The natural actions of $\mathbf{O} (V)$ and $\mathbf{SO} (V)$ on $W_m$ are free and have a geometric quotients (see section \ref{appx}). The codimension of the complement of $W_m$ in $\oplus_{1}^{m} V^*$ is $m-dim(V^*)+1$. $W_m$ is an approximation of $E\mathbf{O} (V)$ and $E\mathbf{SO} (V)$ up to codimension $m-dim(V^*)$. By the general theory of equivariant Chow groups (section \ref{chow}), we have approximations: $$B\mathbf{O} (V)= \stackrel{Lim}{m \rightarrow \infty} \ W_m/\mathbf{O} (V),$$ $$B\mathbf{SO} (V)= \stackrel{Lim}{m \rightarrow \infty} \ W_m/\mathbf{SO} (V).$$ In this section, equivariant Chow rings of $\mathbf{O} (k)$ and $\mathbf{SO} (2k+1)$ are computed via the approximations $$A^*_{\mathbf{O} (V)}(\text{point})= \stackrel{Lim}{m \rightarrow \infty} \ A^*(W_m/\mathbf{O} (V)),$$ $$A^*_{\mathbf{SO} (V)}(\text{point})= \stackrel{Lim}{m \rightarrow \infty} \ A^*(W_m/\mathbf{SO} (V)),$$ and the degeneracy loci results of section \ref{idealz}. \subsection{The Chow Ring of $\mathbf{O} (k)$} Let $k\geq 1$. Let $V\stackrel{\sim}{=} \mathbb{C}^k$ be equipped with a non-degenerate quadratic form $Q$ preserved by $\mathbf{O} (k)$. The quotient $W_m/ \mathbf{O} (k)$ can be explicitly realized as follows. Let $\mathbf G(k,m)$ be the Grassmannian of linear $k$-spaces in $\mathbb{C}^m$. Let $S \rightarrow \mathbf G(k,m)$ be the tautological sub-bundle. Let $Y_m \subset Sym^2 S^*$ be the open locus of non-degenerate quadratic forms on the fibers of $S$. \begin{lm} \label{fbb} There is canonical $\mathbf{O} (k)$-invariant map $\tau: W_m \rightarrow Y_m$ which induces an isomorphism $W_m/\mathbf{O} (k) \stackrel{\sim}{=} Y_m$. \end{lm} \noindent {\em Proof.} Let $w\in W_m$. By the definitions, $w$ naturally induces an injection $\iota_{w}:V \rightarrow \mathbb{C}^m$. The quadratic form $Q$ then induces a non-degenerate quadratic form $\iota_{w}(Q)$ on $\iota_{w}(V)$. Let $$\tau(w) = \iota_{w}(Q) \in Y_m.$$ It is easily checked that $\tau$ is an algebraic morphism. Let $g\in \mathbf{O} (k)$. Then, $$\iota_{g(w)}= \iota_{w} \circ g : V \rightarrow \mathbb{C}^m.$$ Hence, $\tau$ is $\mathbf{O} (k)$-invariant. Since the fibers of $\tau$ are exactly the $\mathbf{O} (k)$ orbits, the induced map $$W_m/ \mathbf{O} (k) \rightarrow Y_m$$ is a bijective morphism of nonsingular complex algebraic varieties and thus an algebraic isomorphism. \qed \vspace{+10pt} $Y_m$ is an approximation to $B\mathbf{O} (k)$ up to codimension $m-k$. $W_m\rightarrow Y_m$ is a principal $\mathbf{O}(k)$-bundle (see the Appendix). The pull-back of the tautological sub-bundle $S \rightarrow \mathbf G(k,m)$ to $Y_m$ is the vector bundle on $Y_m$ induced by the principal $\mathbf{O}(k)$-bundle $W_m\rightarrow Y_m$ and the representation $V$. The Chow ring of the Grassmannian $\mathbf G(k,m)$ is freely generated by the Chern classes $c_1, \ldots, c_k$ of $S$ up to codimension $m-k$ (the relations start in codimension $m-k+1$). By Lemma \ref{paul}, the Chow ring of $Y_m$ is isomorphic to $$\mathbb{Z}[c_1, \ldots, c_k] \ / \ (\beta_1, \ldots, \beta_k)$$ up to codimension $m-k$ where \begin{equation} \label{bbbt} \frac{c(S^*)}{c(S)}= 1+ \beta_1 + \ldots + \beta_k+ \ldots. \end{equation} Induction and simple algebra establishes: $$(\beta_1, \ldots, \beta_k)=(2c_1, 2c_3, 2c_5, \ldots).$$ The Chow ring limit $m\rightarrow \infty$ of $A^*(Y_m)$ is now easily seen to yield: $$A^*_{\mathbf{O} (k)}(\text{point})= \mathbb{Z}[c_1, \ldots, c_k] \ / \ (2c_1, 2c_3, 2c_5, \ldots). $$ Theorem \ref{chor} is proven for $\mathbf{O} (k)$. \subsection{The Chow ring of $\mathbf{SO} (2k+1)$} \label{ort} Let $k\geq 0$. Let $V\stackrel{\sim}{=} \mathbb{C}^{2k+1}$ be equipped with a non-degenerate quadratic form preserved by $$\mathbf{SO} (2k+1)\subset \mathbf{GL} (V).$$ Let $\mathbb{C}^* \subset \mathbf{GL} (V)$ be the scalars. Since $V$ is odd dimensional $$\mathbb{C}^* \cap \mathbf{SO} (2k+1)= \{1\}, \ \ \mathbb{C}^* \times \mathbf{SO}(2k+1) \subset \mathbf{GL}(V).$$ The approximations $W_m$ to $E\mathbf{SO} (2k+1)$ are used. There is a natural free $\mathbf{GL} (V)$-action on $W_m$ which induces a free $\mathbf{SO} (2k+1)$-action and a free scalar $\mathbb{C}^*$-action on $W_m$. The $\mathbf{SO} (2k+1)$-action and the $\mathbb{C}^*$-action commute. There is a commutative diagram: \begin{equation} \label{ffibb} \begin{CD} W_m @>>> W_m/ \mathbf{SO}(2k+1) \\ @VVV @VVV \\ W_m/\mathbb{C}^* @>{\tau}>> W_m/\ \mathbb{C}^* \times \mathbf{SO} (2k+1)\\ \end{CD} \end{equation} All morphisms are group quotients: the horizontal maps are free $\mathbf{SO} (2k+1)$-quotients, the vertical maps are free $\mathbb{C}^*$-quotients. See the Appendix for a discussion of these algebraic quotient problems. The quotients in diagram (\ref{ffibb}) are analyzed. Let $S \rightarrow \mathbf G(2k+1, m)$ be the tautological sub-bundle over the Grassmannian. By an argument identical to Lemma \ref{fbb}, it is seen that $$W_m/ \ \mathbb{C}^* \times \mathbf{SO} (2k+1) \stackrel{\sim}{=} Z_m$$ where $Z_m \subset \mathbf P(Sym^2 S^*)$ is the locus of non-degenerate quadratic forms on the fibers of $S$. Hence, $W_m/ \mathbf{SO} (2k+1) \rightarrow Z_m$ is a $\mathbb{C}^*$-bundle. Let $N \rightarrow Z_m$ be the line bundle associated to this $\mathbb{C}^*$-bundle. On the left side of the diagram, $$W_m/\mathbb{C}^* \subset \mathbf P (\oplus_{1}^m V^*)$$ is the projective spanning locus. $A^1( W_m/\mathbb{C}^*)= \mathbb{Z}$ and $W_m \rightarrow W_m/\mathbb{C}^*$ is the $\mathbb{C}^*$-bundle associated to the generator ${\mathcal{O}}_{\mathbf P}(-1)$ of $A^1(W_m/\mathbb{C}^*)$. \begin{lm} \label{ddd} $A^1(Z_m) \stackrel{\sim}{=} \mathbb{Z}$ and $c_1(N)$ is a generator. \end{lm} \noindent {\em Proof.} Consider the inclusion $Z_m \subset \mathbf P(Sym^2 S^*)$. Let $$\tau: W_m/\mathbb{C}^* \rightarrow Z_m \subset \mathbf P(Sym^2 S^*)$$ be the natural map. Let $\overline{N}$ denote an extension of $N$ to $A^1(\mathbf P(Sym^2 S^*))$. Since $\tau^*(\overline{N})= {\mathcal{O}}_{\mathbf P}(-1)$ generates $A^1(W_m/\mathbb{C}^*)\stackrel{\sim}{=} \mathbb{Z}$, the kernel $K$ of $$ \tau^*: A^1(\mathbf P(Sym^2 S^*)) \rightarrow A^1(W_m/ \mathbb{C}^*)$$ is isomorphic to $\mathbb{Z}$. The class $[D]$ of the locus of degenerate quadratic forms is in $K$. $A^1(\mathbf P(Sym^2 S^*))\stackrel{\sim}{=} \mathbb{Z} c_1 \oplus \mathbb{Z} L$ where $c_1=c_1(S)$ and $L$ is the canonical class ${\mathcal{O}}_{\mathbf P}(1)$. The class of $[D]$ is $-2c_1+ (2k+1) L$ which is not divisible in $A^1(\mathbf P(Sym^2 S^*))$. Hence $K$ is generated by $[D]$. Therefore $\tau^*: A^1(Z_m) \rightarrow A^1(W_m/ \mathbb{C}^*)$ is an isomorphism and $c_1(N)$ is a generator of $A^1(Z_m)$. \qed \vspace{+10pt} There is now enough information to compute the Chow ring of the approximation $W_m/\mathbf{SO} (2k+1)$ to $B\mathbf{SO} (2k+1)$. As before, $W_m/\mathbf{SO} (2k+1)$ is an approximation up to codimension $m-(2k+1)$. The Chow ring of $\mathbf G(2k+1, m)$ is freely generated by the Chern classes $c_1, \ldots, c_{2k+1}$ of the tautological sub-bundle $S$ up to codimension $m-(2k+1)$. By Lemma \ref{pauly}, the Chow ring of $Z_m$ (up to codimension $m-(2k+1)$) has a presentation: $$\mathbb{Z}[c_1, \ldots, c_{2k+1}, L]/ (p(L),\beta_1', \ldots, \beta_{2k+1}')$$ where $L$ is the class of ${\mathcal{O}}_{\mathbf P}(1)$, $p(L)$ is the Chern polynomial satisfied by $L$, and $$\frac{c(S^*\otimes {\mathcal{O}}_{\mathbf P}(1))}{c(S)}= 1+ \beta'_1 + \ldots + \beta'_{2k+1}+ \ldots.$$ Finally, since $W_m/\mathbf{SO} (2k+1)$ is the total space of the $\mathbb{C}^*$-bundle associated to the line bundle $N\rightarrow Z_m$, \begin{equation} \label{ttoott} \mathbb{Z}[c_1, \ldots, c_{2k+1}, L]/ (c_1(N),p(L),\beta_1', \ldots, \beta_{2k+1}') \end{equation} is a presentation of the Chow ring of $W_m/\mathbf{SO} (2k+1)$ (up to codimension $m-(2k+1)$). Since $c_1(N)$ generates $A^1(Z_m)$ and the pair $\{c_1, L\}$ also generate $A^1(Z_m)$, (\ref{ttoott}) is equivalent to: $$\mathbb{Z}[c_1, \ldots, c_{2k+1}, L]/ (c_1, L ,p(L),\beta_1', \ldots, \beta_{2k+1}').$$ By the defintions of $p(L)$ and the elements $\beta'_i$, there is an equality of ideals $$(c_1,L, p(L), \beta'_1,\ldots, \beta_{2k+1}') = (c_1,L, c_s(Sym^2 S^*), \beta_1, \ldots, \beta_{2k+1})$$ where $s= rank(Sym^2 S^*)$ and the $\beta_i$ are determined by (\ref{bbbt}). \begin{lm} $c_s(Sym^2 S^*) \in (\beta_1, \ldots, \beta_{2k+1})$. \end{lm} \noindent {\em Proof.} Consider the total space $Sym^2 S^*$. There is an isomorphism $A^*(Sym^2 S^*) \stackrel{\sim}{=} A^*(\mathbf G(2k+1,m))$. The pull-back of the bundle $$Sym^2 S^* \rightarrow \mathbf G(2k+1,m)$$ to the total space $Sym^2 S^*$ has a canonical section $\tau$. The zero scheme of $\tau$ is contained in the locus of degenerate quadratic forms $D\subset Sym^2 S^*$. Also, the zero scheme of $\tau$ represents the class $$c_s(Sym^2 S^*)\in A^*(Sym^2 S^*).$$ Therefore, $c_s(Sym^2 S^*) \in I_2$. The proof is complete by Lemma \ref{paul}. \qed \vspace{+10pt} \noindent As before, $(\beta_1, \ldots, \beta_{2k+1})= (2c_1, 2c_3,2c_5, \ldots, 2c_{2k+1}).$ Hence, the Chow ring of $W_m/ \mathbf{SO} (2k+1)$ up to codimension $m-(2k+1)$ has a presentation: $$\mathbb{Z}[c_1, \ldots, c_{2k+1}]/ (c_1, 2c_3, 2c_5,\ldots, 2c_{2k+1}).$$ The limit process yields Theorem \ref{chor} for $\mathbf{SO} (2k+1)$. \section{\bf The Proof of Proposition \ref{alltor}} \subsection{} We follow the notation of section \ref{prezz}. Let $V$ be a fixed $2$-dimensional $\mathbb{C}$-vector space. Let $\mathbf P^1\stackrel{\sim}{=} \mathbf P(V)$. Let $$U\subset \bigoplus_{0}^{d} Sym^d(V^*)$$ denote the non-degenerate locus parameterizing bases of the linear series of ${\mathcal{O}}_{\mathbf P^1}(d)$ on $\mathbf P^1$. $\mathbf{GL}(V)$ acts on $U$ properly with finite stabilizers and geometric quotient (see the Appendix) isomorphic to $H(d)$. By Proposition \ref{dane}, \begin{equation} \label{xeq} A^*(d) \otimes_{\mathbb{Z}}{\mathbb{Q}} \stackrel{\sim}{=} A^*_{\mathbf{GL} (V)}(U) \otimes _{\mathbb{Z}}{\mathbb{Q}}. \end{equation} By the definition of equivariant Chow groups, $$A^*_{\mathbf{GL}(V)}(U)= A^*(U \times^{\mathbf{GL}(V)} E \mathbf{GL} (V)).$$ \subsection{The Chow Rings of $\mathbf{GL}(V)$ and $\mathbf{SL}(V,n)$} \label{slvn} Algebraic approximations to $E\mathbf{GL} (V)$ are easily found. Since related results about the groups $\mathbf{SL}(V,n)$ are need in section \ref{ode}, a unified development is presented here. Recall $SL(V,n) \subset \mathbf{GL}(V)$ is defined to be $det^{-1}(\mathbb{Z}/n\mathbb{Z})$ where $det: \mathbf{GL} \rightarrow \mathbb{C}^*$ is the determinant homomorphism and $\mathbb{Z}/ n\mathbb{Z}$ is th group of $n^{th}$ roots of unity. As in the orthogonal cases, the easiest approach to $E \mathbf{GL} (V)$ and $E \mathbf{SL} (V,n)$ is via sums of the representation $V^*$. As before, let $m>>0$ and let $$W_m \subset \oplus_{1}^{m} V^*$$ be the spanning locus. The induced $\mathbf{GL} (V)$ and $\mathbf{SL} (V,n)$-actions on $W_m$ are free and have geometric quotients (see the Appendix) which approximate $B \mathbf{GL} (V)$ and $B \mathbf{SL} (V,n)$ up to codimension $m-2$. It is easily seen that $W_m / \mathbf{GL}(V) \stackrel{\sim}{=} \mathbf G(2,m)$. Since the Chow ring of this Grassmannian (up to codimension $m-2$) is freely generated by the Chern classes $c_1$ and $c_2$ of the tautological sub-bundle, $$A^*(W_m/ \mathbf{GL} (V)) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2]$$ up to codimension $m-2$. Taking the $m\rightarrow \infty $ limit, $$A^*_{\mathbf{GL} (V)}(\text{point}) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2].$$ Similarly, $W_m/ \mathbf{SL} (V, n)$ is the total space of the $n^{th}$ tensor power of the line bundle $\bigwedge^2 S$ over $\mathbf G(2,m)$. Hence up to codimension $m-2$, $$A^*(W_m/ \mathbf{SL} (V,n)) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2]/ (nc_1).$$ Taking the $m \rightarrow \infty$ limit, $$A^*_{\mathbf{SL} (V,n)}(\text{point}) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2]/(nc_1).$$ \subsection{Proposition \ref{alltor}} \label{abe} The quotient $U \times ^{\mathbf{GL} (V)} E\mathbf{GL} (V)$ is analyzed via approximation. $V \times ^{\mathbf{GL}(V)} W_m$ is the tautological sub-bundle $S$ over $\mathbf G(2,m)$. $$U \times ^{\mathbf{GL}(V)} W_m \subset \oplus_{0}^{d} Sym^d(V^*) \times ^{\mathbf{GL}(V)} E$$ is the non-degenerate open locus in the total space of the bundle $\oplus_{0}^{d} Sym^d (S^*)$ over $\mathbf G(2,m)$. By Lemma \ref{petey}, there is an isomorphism $$A^*(U \times ^{\mathbf{GL}(V)} W_m) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2]/ (\alpha_1, \ldots, \alpha_{d+1})$$ up to codimension $m-2$ where $$\frac{1}{c(Sym^d (S))}= 1+ \alpha_1+ \ldots + \alpha_{d+1} + \ldots.$$ The ideal generated by $(\alpha_1, \ldots, \alpha_{d+1})$ is equal to the ideal generated by the first $d+1$ Chern classes of $Sym^d (S)$. Taking the $m \rightarrow \infty$ limit, a presentation of $A^*_{\mathbf{GL} (V)} (U)$ is obtained. $A^*_{\mathbf{GL} (V)}(U)$ is generated (as a ring) in codimensions $1$, $2$ by elements $c_1$, $c_2$ respectively. There are $d+1$ relations given as follows. Let $S$ be a rank $2$ bundle with Chern classes $c_1$ and $c_2$. The $d+1$ Chern classes of $Sym^d(S)$ are the relations. \begin{lm} \label{ater} $A^*_{\mathbf{GL}(V)}(U) \otimes \mathbb{Q}$ is zero is positive codimension. \end{lm} \noindent {\em Proof.} A standard calculation yields: \begin{equation} \label{see1} c_1(Sym^d(S))=\frac{d(d+1)}{2} c_1, \end{equation} \begin{equation} \label{see2} c_2(Sym^d(S))= \frac{d(d-1)(d+1)(3d+2)}{24} c_1^2 + \frac{d(d+1)(d+2)}{6} c_2. \end{equation} Since the coefficients of $c_1$ and $c_2$ never vanish for positive $d$ in equations (\ref{see1}) and (\ref{see2}) respectively, the first two Chern classes of $Sym^d(S)$ generate the ideal $(c_1, c_2)$ in $\mathbb{Q}[c_1,c_2]$. \qed \vspace{+10pt} \noindent Lemma \ref{ater} and the isomorphism (\ref{xeq}) establish Proposition \ref{alltor}. \section{{\bf $A^*(d)$, $d$ Even}} \label{evan} The notation of section \ref{prezz} is used. Let $d=2n$ (where $n\geq 1$). Let $V\stackrel{\sim}{=} \mathbb{C}^2$. There is a free $\mathbf{PGL} (V)$-action on $\mathbf P(U) \subset \mathbf P(\oplus_{0}^{d} Sym^d V^*)$ with geometric quotient (see the Appendix) isomorphic to $H(d)$. By Proposition \ref{dane}, $$A^*(d) \stackrel{\sim}{=} A^*_{\mathbf{PGL} (V)} (\mathbf P (U)).$$ By the definition of equivariant Chow groups, $$A^*_{\mathbf{PGL} (V)} (\mathbf P (U)) \stackrel{\sim}{=} A^*( \mathbf P(U) \times ^{\mathbf{PGL} (V)} E \mathbf{PGL} (V)).$$ The Chow ring $A^*( \mathbf P(U) \times ^{\mathbf{PGL} (V)} E \mathbf{PGL} (V))$ is computed in this section for $d=2n$. Consider the $3$-dimensional representation $Sym^2(V)$ of $\mathbf{PGL}(V)$. This respresentation leaves invariant a unique (up to $\mathbb{C}^*$) quadratic form $Q$ on $Sym^2(V)$. A group isomorphism $\mathbf{PGL} (V) \stackrel{\sim}{=} \mathbf{SO} (3)$ is induced by this quadratic form. The dual of the standard $3$-dimensional representation of $\mathbf{SO}(3)$ corresponds to the representation $Sym^2 (V^*)$ of $\mathbf{PGL} (V)$. Let $$A_m \subset \oplus_{1}^{m} Sym^2 (V^*)$$ be the spanning locus. The approximations $A_m/ \mathbf{PGL} (V)$ to $B \mathbf{PGL} (V)$ correspond exactly to the approximations $W_m/ \mathbf{SO} (3)$ to $B \mathbf{SO} (3)$ defined in section \ref{orthoo}. $A_m/ \mathbf{PGL}(V)$ is therefore the total space of a $\mathbb{C}^*$-bundle $N \rightarrow Z_m$. $Z_m$ is the open set of non-degenerate quadratic forms in $\mathbf P(Sym^2 (S^*))$ over the Grassmannian $\mathbf G(3,m)$. Let $B_m$ denote this approximation to $B \mathbf{PGL} (V)$. $Sym^d(V^*)$ is a $\mathbf{PGL}(V)$ representation for $d$ even ({\em not} for $d$ odd). Hence, $$ Sym^d(V^*) \times^{\mathbf{PGL}(V)} A_m$$ is a rank $d+1$ vector bundle $F_d \rightarrow B_m$. The quotient $$\mathbf P(U) \times ^{\mathbf{PGL} (V)} A_m \subset \mathbf P(\oplus_{0}^{d} Sym^d(V^*)) \times ^{\mathbf{PGL} (V)} A_m$$ is simply the projective non-degenerate locus in $\mathbf P(\oplus_{0}^{d} F_d)$. The first step is to identify the bundle $F_d \rightarrow B_m$. There is a tautological sub-bundle $S \rightarrow B_m$ obtained from the Grassmannian. There is a tautological equivalence $S^* \stackrel{\sim}{=} F_2$. More generally, there is a tautological sequence on $\mathbf P(Sym^2 (S^*))$: \begin{equation} \label{toto} 0 \rightarrow {\mathcal{O}}_{\mathbf P}(-1) \otimes Sym^{n-2} (S^*) \rightarrow Sym^n (S^*) \rightarrow Q_n \rightarrow 0 \end{equation} for all $n\geq 2$. Let $([q], P)\in \mathbf P(Sym^2 (S^*))$ where $P\subset \mathbb{C}^m$ is a linear $3$-space and $0 \neq q \in Sym^2(P^*)$. The fiber of ${\mathcal{O}}_{\mathbf P}(-1)$ over $([q],P)$ is simply $\mathbb{C} \cdot q$. The left inclusion in sequence (\ref{toto}) is determined by the canonical multiplication map: $$0 \rightarrow \mathbb{C} \cdot q \otimes Sym^{n-2} (P^*) \rightarrow Sym^n(P^*).$$ Again, there is a tautological equivalence $F_{2n} \stackrel{\sim}{=} Q_n$ on $B_m$. Note $A^1(B_m)=0$ by Lemma \ref{ddd}. The Chern polynomial of $F_d$ on $B_m$ is therefore: $$c(F_d)= \frac{c(Sym^n (S^*))}{c(Sym^{n-2}(S^*))}.$$ Now, by Lemma \ref{pete}, a presentation of $A^*(\mathbf P(U) \times^{\mathbf{PGL}(V)} B_m)$ up to codimension $m-3$ is obtained by $$A^*(B_m)[\mathcal{L}]/ (p(\mathcal{L}), \alpha'_1, \ldots, \alpha'_{d+1})$$ where $\mathcal{L}$ is the class of ${\mathcal{O}}_{\mathbf P(\oplus_{0}^{d} F_d)}(1)$ and $$\frac{ (1+\mathcal{L})^{d+1}}{c(F^*_d)}= \frac{(1+\mathcal{L})^{d+1} \cdot c(Sym^{n-2} S)} {c(Sym^n S)} = 1+ \alpha'_1+ \ldots+ \alpha'_{d+1} \ldots .$$ By the presentation of $A^*(B_m)$ in section \ref{ort}, it follows $$A^*(\mathbf P(U) \times^{\mathbf{PGL}(V)} B_m) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2, c_3, \mathcal{L}]/ (p(\mathcal{L}), c_1, 2c_3, \alpha'_1, \ldots, \alpha'_{d+1})$$ up to codimension $m-3$. Taking the $m\rightarrow \infty$ limit, $$ A^*(d) \stackrel{\sim}{=} A^*_{\mathbf{PGL}(V)}(\mathbf P(U)) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2, c_3, \mathcal{L}]/ (p(\mathcal{L}), c_1, 2c_3, \alpha'_1, \ldots, \alpha'_{d+1}).$$ The relation $p(\mathcal{L})$ is of codimension $(d+1)^2$. Since the dimension of $H(d)$ is $d^2+ 2d-3=(d+1)^2-4$ and the generators $c_1,c_2, c_3$, and $\mathcal{L}$ have dimension at most 3, $p(\mathcal{L})$ is a relation among classes that are already zero. Hence, \begin{equation} \label{freedy} A^*(d) \stackrel{\sim}{=} A^*_{\mathbf{PGL}(V)}(\mathbf P(U)) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2, c_3, \mathcal{L}]/ (c_1, 2c_3, \alpha'_1, \ldots, \alpha'_{d+1}). \end{equation} Following [EG], there is natural map $$\mathbf P(U)\times^{\mathbf{PGL}(V)} A_m \rightarrow \mathbf P(U)/ \mathbf{PGL}(V) \stackrel{\sim}{=} H(d)$$ which expresses $\mathbf P(U)\times^{\mathbf{PGL}(V)} A_m$ as an open set of a vector bundle over $H(d)$. This fibration induces an isomorphism on Chow rings (up to codimension $m-3$). The classes $c_i\in A^*(d)$ are easily identified via this isomorphism (up to codimension $m-3$): \begin{equation} \label{ddod} A^*(\mathbf P(U)\times^{\mathbf{PGL}(V)} A_m) \stackrel{\sim}{=} A^*(d). \end{equation} They are the Chern classes of the vector bundle obtained from the the principal $\mathbf{PGL}(V)$-bundle $\mathbf P(U) \rightarrow H(d)$ and the representation $Sym^2(V)$. Let $\mathcal{H}\in A^1(d)$ be the class of curves meeting a fixed codimension 2 linear space $P$ of $\mathbf P^d$. $\mathcal{H}$ corresponds via the isomorphism (\ref{ddod}) to a resultant class in $A^1(\mathbf P(U)\times^{\mathbf{PGL}(V)} A_m)$. Routine calculations show $\mathcal{H}=2d \mathcal{L}$ where $2d$ is the degree of the resultant of degree $d$ polynomials. Since $(d+1)\mathcal{L}=0$ by the presentation (\ref{freedy}). $$n \mathcal{H} = d^2 \mathcal{L} = (d^2-1+1)\mathcal{L} = (d-1)(d+1)\mathcal{L} +\mathcal{L} = \mathcal{L}.$$ The proof of Theorem \ref{evan} is complete. \section{{\bf $A^*(d)$, $d$ Odd}} \label{ode} Let $d=2n-1$ (where $n\geq 1$). Let $V\stackrel{\sim}{=} \mathbb{C}^2$. There is a canonical, $\mathbf{GL} (V)$-equivariant, multilinear map $$\mu:\bigoplus_0^{2n-1} Sym^{2n-1}(V^*) \rightarrow \bigwedge^{2n} Sym^{2n-1}(V^*)$$ given by the exterior product (see section \ref{prezz}). \begin{lm} The $\mathbf{SL} (V,n)$-action on $\bigwedge^{2n} Sym^{2n-1}(V^*)$ is trivial. \end{lm} \noindent {\em Proof.} Since the $1$-dimensional representations of $\mathbf{SL}(V)$ are trivial, the action of $\mathbf{SL}(V)$ on $\bigwedge^{2n} Sym^{2n-1}(V^*)$ is certainly trivial. Let $H\subset\mathbf{SL}(V,n)$ be the subgroup of scalars. $H$ is the multiplicative group of scalar $2n^{th}$ roots of unity, Let $\xi \in H$ be a scalar. $\xi$ acts on $\bigwedge^{2n} Sym^{2n-1}(V^*)$ by the scalar $\xi^{(2n)(2n-1)}= 1$. It is easily checked that $\mathbf{SL}(V,n)$ is generated (as a group) by $H$ and $\mathbf{SL}(V)$. Hence, the $\mathbf{SL}(V,n)$-action is trivial. \qed \vspace{+10pt} \noindent Let $Y= \mu^{-1} (p)$ where $0 \neq p \in\bigwedge^{2n} Sym^{2n-1}(V^*)$. There is an $\mathbf{SL} (V,n)$-action on $Y$. \begin{lm} \label{freeaq} The $\mathbf{SL} (V,n)$-action on $Y$ is free with geometric quotient $H(d)$. \end{lm} \noindent {\em Proof.} Certainly $\mathbf{SL}(V,n)$ acts on $Y$ since the $\mathbf{SL}(V,n)$-action on $\bigwedge^{2n} Sym^{2n-1}(V^*)$ is trivial. Let $$U \subset \oplus_0^{2n-1} Sym^{2n-1} (V^*)$$ be the non-degenerate locus. First, it is shown that the $\mathbf{SL}(V,n)$-action on $U$ is free. Since $Y\subset U$, $\mathbf{SL}(V,n)$ acts freely on $Y$. Let $u\in U$. Suppose $g \in \mathbf{SL}(V,n)$ satisfies $g\cdot u = u$. $\mathbf{PGL}(V)$ acts freely on $\mathbf P(U)$. Let $\pi: \mathbf{SL}(V,n) \rightarrow \mathbf{PGL}(V)$. Then, $$\pi(g) \cdot \mathbf P(u)= \mathbf P(u).$$ Hence, $\pi(g) = 1 \in \mathbf{PGL} (V)$. The element $g$ is therefore a scalar in $\mathbf{SL}(V,n)$ equal to a $2n^{th}$ root of unity $\xi$. Then, $g$ acts on $u$ by the scalar $\xi^{2n-1}$. Since $g \cdot u=u$, $\xi^{2n-1} =1$. Since $(2n, 2n-1)=1$, $\xi^{2n-1}=1$ implies $\xi=1$. Therefore, $g=1 \in \mathbf{SL}(V,n)$. The $\mathbf{SL}(V,n)$-action on $U$ is free. It is now shown the quotient $Y/ \mathbf{SL}(V,n)$ is isomorphic to $H(d)$. There are natural, equivariant, algebraic projection maps: $$Y \rightarrow \mathbf P(U),$$ $$\pi: \mathbf{SL}(V,n) \rightarrow \mathbf{PGL}(V).$$ These maps induce a natural surjective map on quotients: $$\phi: Y/ \mathbf{SL}(V,n) \rightarrow \mathbf P(U)/ \mathbf{PGL}(V) \stackrel{\sim}{=} H(d).$$ It suffice to prove $\phi$ is injective. (A bijective map of nonsingular complex algebraic varieties is an algebraic isomorphism.) Let $y_1, y_2 \in Y$ be points. Let $[y_1], [y_2]\in \mathbf P(U)$ denote the corresponding points. Suppose there exists an element $\gamma \in \mathbf{PGL}(V)$ satisfying $\gamma \cdot [y_1]=[y_2]$. To prove $\phi$ is injective, it must be shown that $y_1$ and $y_2$ are in the same $\mathbf{SL}(V,n)$ orbit. Let $g\in \mathbf{SL}(V,n)$ satisfy $\pi(g)=\gamma$. Then, $[g \cdot y_1] =[y_2]$. Hence $g \cdot y_1= (\lambda y_2)$ where $\lambda \in \mathbb{C}^*$ is a scalar. By the conditions $g \cdot y_1, y_2 \in Y$, it follows $\lambda^{2n}=1$. Since $(2n, 2n-1)=1$, a $2n^{th}$ root of unity $\xi \in \mathbf{SL}(V,n)$ can be found satisfying $\xi^{2n-1}=\lambda^{-1}$. Let $h \in \mathbf{SL}(V,n)$ be determined by $h= \xi \cdot g$. $$h\cdot y_1= \xi \cdot g \cdot y_1= \xi \cdot (\lambda y_2) = \xi^{2n-1}(\lambda y_2)= y_2.$$ Therefore $y_1$ and $y_2$ are in the same $\mathbf{SL}(V,n)$ orbit. \qed \vspace{+10pt} \noindent There is a canonical isomorphism of graded rings $$A^*(d=2n-1) \stackrel{\sim}{=} A^*_{\mathbf{SL} (V,n)}(Y).$$ The equivariant Chow ring $A^*_{\mathbf{SL} (V,n)}(Y)$ is computed in this section. The approximations $W_m$ and $W_m/ \mathbf{SL}(V,n)$ to $E \mathbf{SL}(V,n)$ and $B \mathbf{SL}(V,n)$ determined in section \ref{slvn} are used here. Recall $W_m/ \mathbf{SL}(V,n) \rightarrow \mathbf G(2,m)$ is the $\mathbb{C}^*$-bundle associated to the $n^{th}$ tensor power of $\wedge^2 S$ (where $S$ is the tautological sub-bundle over $\mathbf G(2,m)$). Since $$Y \subset U \subset \oplus_{0}^{2n-1} Sym^{2n-1} (V^*),$$ there are inclusions: $$ Y \times^{\mathbf{SL}(V,n)} W_m \ \subset \ U \times ^{\mathbf{SL}(V,n)} W_m \ \subset \ \oplus_{0}^{2n-1} Sym^{2n-1}(V^*) \times ^{\mathbf{SL}(V,n)} W_m.$$ Let $F_n=V^* \times^{\mathbf{SL}(V,n)} W_m$. $F_n$ is an algebraic vector bundle over $W_m/ \mathbf{SL}(V,n)$. $F_n$ is easily identified as the pull-back of $S^*$ to $W_m/ \mathbf{SL}(V,n)$. $U \times ^{\mathbf{SL}(V,n)} W_m$ is the affine non-degenerate locus (i) of section \ref{idealz} associated to the bundle $Sym^{2n-1} F_n\stackrel{\sim}{=} Sym^{2n-1} S^*$. \begin{lm} \label{qwq} There is an isomorphism $$ \epsilon: \mathbb{C}^* \times (Y \times^{\mathbf{SL}(V,n)} W_m) \stackrel{\sim}{=} U \times ^{\mathbf{SL}(V,n)} W_m.$$ \end{lm} \noindent {\em Proof.} Let $\mathbf{SL}(V,n)$ act trivially on $\mathbb{C}^*$. Define a $\mathbf{SL}(V,n)$ equivariant isomorphism $$\delta: \mathbb{C}^* \times Y \rightarrow U$$ by the following: $$\delta\big( \lambda, (\omega_0, \omega_1\ldots, \omega_{2n-1})\big) = (\lambda \omega_0, \omega_1, \ldots, \omega_{2n-1}).$$ The isomorphism $\delta$ induces isomorphisms: $$\mathbb{C}^* \times (Y \times^{\mathbf{SL}(V,n)} W_m) \stackrel{\sim}{=} (\mathbb{C}^* \times Y) \times^{\mathbf{SL}(V,n)} W_m \stackrel{\sim}{=} U \times^{\mathbf{SL}(V,n)} W_m.$$ Let $\epsilon$ be the composition. \qed \vspace{+10pt} $W_m/ \mathbf{SL}(V,n)$ approximates $B \mathbf{SL}(V,n)$ up to codimension $m-2$. The Chow ring of $Y \times ^{\mathbf{SL}(V,n)} W_m$ is now computed (up to codimension $m-2$). By Lemma \ref{qwq}, there is an isomorphism: $$A^*(Y \times ^{\mathbf{SL}(V,n)} W_m) \stackrel{\sim}{=} A^*(U \times ^{\mathbf{SL}(V,n)} W_m).$$ Since $U$ is the affine non-degenerate locus associated to the bundle $$Sym^{2n-1} S^* \rightarrow (W_m / \mathbf{SL}(V,n)),$$ Lemma \ref{pete} can be applied. Recall the Chow ring of $W_m / \mathbf{SL}(V,n)$ (up to codimension $m-2$) has a presentation $\mathbb{Z}[c_1, c_2]/ (nc_1)$. Hence, there is an isomorphism (up to codimension $m-2$): $$A^*(Y \times ^{\mathbf{SL}(V,n)} W_m) \stackrel{\sim}{=} \mathbb{Z}[c_1, c_2]/ (nc_1, \alpha_1, \ldots, \alpha_{d+1})$$ where $$\frac{1}{c(Sym^d S)}= 1+ \alpha_1 + \ldots + \alpha_{d+1}+ \ldots.$$ The proof of Theorem \ref{hodd} is complete. \section {\bf Examples} Since $H(1)$ is a point, $A^*(1)$ is the trivial $\mathbb{Z}$-algebra (which agrees with the presentation of Theorem 2). $H(2)$ is the space of nonsingular plane conics. By Theorem 1, $A^*(2)$ is generated by $c_2$, $c_3$, and $\mathcal{L}=\mathcal{H}$ subject to $4$ relations: $$2c_3=0,$$ $$3\mathcal{H}=0, \ -c_2+3\mathcal{H}^2=0, \ -c_3+\mathcal{H}^3=0.$$ Since $\mathcal{H}$ is 3-torsion, $c_2=0$. Since $c_3$ is two torsion, the last equation can be reduced to $c_3=\mathcal{H}^3=0$. Therefore $A^*(2)$ is given by $$\mathbb{Z}[\mathcal{H}]/ (3\mathcal{H}, \mathcal{H}^3).$$ Since $H(2)$ is an open set of the projective space of plane conics, another approach to $A^*(2)$ is possible. The class $\mathcal{H}$ is simply the restriction of the hyperplane class which necessarily generates $H(2)$. The relation $3\mathcal{H}$ can be obtained from the degree $3$ degeneracy locus of singular plane conics. The relation $\mathcal{H}^3$ is a consequence of the fact that the locus of conics singular at a {\em fixed} point in $\mathbf P^2$ is a linear $\mathbf P^2$ in the $\mathbf P^5$ of conics. Let $d=3$, $n=2$, $d=2n-1$. By Theorem 2, $A^*(3)$ is generated by $c_1$ and $c_2$ with relations: $$ 2c_1=0,$$ $$6c_1=0, \ 11c_1^2+ 10c_2=0, \ 6c_1^3+30c_1c_2=0, \ 18c_1^2c_2+9c_2^2=0.$$ These relations simplify to yield the presentation: $$A^*(3)= \mathbb{Z}[c_1,c_2]/ (2c_1, c_1^2+10c_2, c_1^3, c_1^2c_2,c_2^2).$$ In particular, $A^i(3)=0$ for $i\geq 4$. \section {\bf Appendix On Algebraic Quotients} \label{appx} Let $\mathbb{C}$ be the ground field of complex numbers. The geometric invariant theory terminology of [MFK] is used here. Let $\mathbf{G}$ be a reductive linear algebraic group. A group action $\mathbf{G} \times X \rightarrow X$ is {\em proper} if the natural map $$\Psi: \mathbf{G} \times X \rightarrow X \times X$$ (given by the action and projection onto the second factor) is a proper morphism. The main result needed is the following: \begin{pr} Let $X$ be a quasi-projective variety with a linearized $\mathbf{G}$-action satisfying $X^{stable}_{(0)} =X$. \label{qquot} Then, the $\mathbf{G}$-action on $X$ is proper and there is a quasi-projective geometric quotient $X \rightarrow X/\mathbf{G}$. \end{pr} \noindent {\em Proof.} Properness of the action is exactly Corollary 2.5 of [MFK]. The geometric quotient is the main construction in geometric invariant theory (Theorem 1.10 of [MFK]). \qed \vspace{+10pt} \noindent The stable locus $X_{(0)}^{stable}$ is detected by the Numerical Criterion. Let $V\stackrel{\sim}{=} \mathbb{C}^k$ be a vector space equipped with a quadratic form. All of the linear algebraic groups considered in this paper are reductive: $\mathbf{GL}(V)$, $\mathbf{SL}(V)$, $\mathbf{PGL}(V)$, $\mathbf{SL}(V,n)$, $\mathbf{O}(V)$, $\mathbf{SO} (V)$, $\mathbb{C}^*$, $\mathbb{C}^* \times \mathbf{SO}(V)$. Let $$Span_m(V,d) \subset \oplus _{1}^{m} Sym^d(V*)$$ be the spanning locus (the locus of $m$-tuples of vector of $Sym^d(V^*)$ which span $Sym^d(V^*)$). The spanning loci $U \subset \oplus_{0}^{d} Sym^d(\mathbb{C}^{2*})$ and $W_m \subset \oplus_{1}^{m} V^*$ are special cases of $Span_m(V,d)$. The group actions considered in the paper are of three forms: \begin{enumerate} \item[(i)] The natural $\mathbf{G}$-action on $X=Span_m(V,d)$ where $\mathbf{G} \subset \mathbf{GL}(V)$ is a reductive subgroup. \item[(ii)] The $\mathbf{G}$-action on a $\mathbf{G}$-invariant subvariety $Y\subset Span_m(V,d)$ where $\mathbf{G} \subset \mathbf{GL}(V)$ is a reductive subgroup. \item[(iii)] The natural $\mathbf{PGL}(V)$-action on $X=\mathbf P(Span_m(V,d))$. \end{enumerate} For example, the $SL(V,n)$-action on $Y$ considered in section \ref{ode} is of form (ii). Consider first (i) and (ii). A linearization of the $\mathbf{GL}(V)$-action can be found on $X$ satisfying $X^{stable}_{(0)}=X$. Such a linearization is found in section 1 of [P]. Since the stable locus is detected by the Numerical Criterion, the result for $\mathbf{GL}(V)$ implies $X^{stable}_{(0)}=X$ for the induced action of any reductive subgroup $\mathbf{G} \subset \mathbf{GL}(V)$. It is similarly simple to find a linearization in case (iii) satisfying $X^{stable}_{(0)}=X$. Therefore, Proposition \ref{qquot} applies to all the quotient problems in the paper. In [MFK], the $\mathbf{G}$-action $\mathbf{G} \times X \rightarrow X$ is defined to be {\em free} if the natural map $\Psi$ is a closed embedding. An action is {\em set-theoretically free} if the stabilizers are trivial. For the set-theoretically free actions considered in this paper, the following Lemma is utilized. \begin{lm} Let $X$ be nonsingular. Let $\mathbf{G}\times X \rightarrow X$ be a proper action. In this case, set-theoretically free implies free. \end{lm} \noindent {\em Proof.} Let $I \subset X\times X$ be the image of $\Psi$. $I$ is a closed subvariety since $\Psi$ is proper. It must be shown that $\mathbf{G} \times X \rightarrow I$ is an isomorphism. First it is shown that $I$ is nonsingular. For this, it suffices to prove the differential of $\Psi$ is injective at each point of $\mathbf{G} \times X$. It is well known (over $\mathbb{C}$) that set-theoretically trivial stabilizers are also scheme-theoretically trivial. From this, the injectivity of the differential $d\Psi$ is easily deduced. Now $\Psi: \mathbf{G} \times X \rightarrow I$ is a bijective map of nonsingular complex algebraic varieties and thus an algebraic isomorphism. \qed \vspace{+10pt} \noindent A result of [MFK] (Proposition 0.9) relates free quotients to principal $\mathbf{G}$-bundles. \begin{pr} Let $\mathbf{G} \times X \rightarrow X$ be an algebraic group action with geometric quotient $X \rightarrow Y$. If the action is free, then $X \rightarrow Y$ is a (\'etale locally trivial) principal $\mathbf{G}$-bundle. \end{pr} \noindent A principal $\mathbf{G}$-bundle $X \rightarrow Y$ and a representation $\mathbf{G}\rightarrow\mathbf{GL}$ together yield a principal $\mathbf{GL}$-bundle $X \times^{\mathbf{G}} \mathbf{GL} \rightarrow Y$. Since every principal $\mathbf{GL}$-bundle is Zariski locally trivial ($\mathbf{GL}$ is a {\em special} group in the sense of Grothendieck (see {\em Anneau de Chow et Applications}, Seminaire Chevalley 1958)), an algebraic vector bundle over $Y$ is obtained.
1996-11-04T11:53:12
9607
alg-geom/9607004
en
https://arxiv.org/abs/alg-geom/9607004
[ "alg-geom", "math.AG" ]
alg-geom/9607004
E. Looijenga
Richard Hain and Eduard Looijenga
Mapping Class Groups and Moduli Spaces of Curves
We expanded section 7 and rewrote parts of section 10. We also did some editing and made some minor corrections. latex2e, 46 pages
null
null
null
null
This is a survey paper that also contains some new results. It will appear in the proceedings of the AMS summer research institute on Algebraic Geometry at Santa Cruz.
[ { "version": "v1", "created": "Thu, 4 Jul 1996 08:49:56 GMT" }, { "version": "v2", "created": "Mon, 4 Nov 1996 10:48:45 GMT" } ]
2008-02-03T00:00:00
[ [ "Hain", "Richard", "" ], [ "Looijenga", "Eduard", "" ] ]
alg-geom
\section{Introduction} \label{sec:intro} It is classical that there is a very strong relation between the topology of ${\mathcal M}_g$, the moduli space of smooth projective curves of genus $g$, and the structure of the mapping class group $\Gamma_g$, the group of homotopy classes of orientation preserving diffeomorphisms of a compact orientable surface of genus $g$. The geometry of ${\mathcal M}_g$, the topology of ${\mathcal M}_g$, and the structure of $\Gamma_g$ are all intimately related. Until recently, the principal tools for studying these topics were Teichm\"uller theory (complex analysis and hyperbolic geometry), algebraic geometry, and geometric topology. Recently, a fourth cornerstone has been added, and that is physics which enters through the theories of quantum gravity and conformal field theory. Already these new ideas have had a remarkable impact on the subject through the ideas of Witten and the work of Kontsevich. In this article, we survey some recent developments in the understanding of moduli spaces. Some of these are classical (do not use physical ideas), while others are modern. One message we would like to convey is that algebraic geometers, topologists, and physicists who work on moduli spaces of curves may have a lot to learn from each other. Having said this, we should immediately point out that, partly due to our own limitations, there are important developments that we have not included in this survey. Our most notable omission is the arithmetic aspect of the theory, much of which originates in Grothendieck's fundamental works \cite{groth:marche}, \cite{groth:esq}. We direct readers to the volume \cite{groth:dessins} and to the recent papers of Ihara, Nakamura and Oda for other recent developments (see Nakamura's survey \cite{nakamura:survey} for references). Other topics we have not covered include conformal field theory and recent work of Ivanov \cite{ivanov:rigid} and Ivanov and McCarthy \cite{ivanov-mccarthy} on homomorphisms from mapping class groups and arithmetic groups to mapping class groups. Of particular importance is Ivanov's version of Margulis rigidity for mapping class groups \cite{ivanov:rigid} which he obtains using some recent fundamental work of Kaimanovich and Masur \cite{masur} on the ergodic theory of Teichm\"uller space. We shall denote the moduli space of $n$ pointed smooth projective curves of genus $g$ by ${\mathcal M}_g^n$. Knudsen, Mumford and Deligne constructed a canonical compactification $\overline{\M}_g^n$ of it. It is the moduli space of stable $n$ pointed projective curves of genus $g$. It is a projective variety with only finite quotient singularities. Perhaps the most important developments of the decade concern the Chow rings% \footnote{All Chow rings and cohomology groups in this paper are with ${\mathbb Q}$ coefficients except when explicit coefficients are used.} of ${\mathcal M}_g^n$ and $\overline{\M}_g^n$. The first Chern class of the relative cotangent bundle of the universal curve associated to the $i$th point is a class $\overline{\tau}_i$ in $\CH^1(\overline{\M}_g^n)$. One can consider monomials in the $\overline{\tau}_i$'s of polynomial degree equal to the dimension of some $\overline{\M}_g^n$. For such a monomial, one can take the degree of the monomial as a zero cycle on $\overline{\M}_g^n$ to obtain a rational number. These can be assembled into a generating function. Witten conjectured that this formal power series satisfies a system of partial differential operators. Kontsevich proved this using topological arguments, and thereby provided inductive formulas for these intersection numbers. These developments are surveyed in Section~\ref{sec:ribbon}. For each positive integer $i$, Mumford defined a {\it tautological class} $\overline{\kappa} _i$ in $\CH^i(\overline{\M}_g)$. The restrictions $\kappa _i$ of these classes to $\CH^{\bullet}({\mathcal M}_g)$ generate a subalgebra of $\CH^{\bullet}({\mathcal M}_g)$ which is called the {\it tautological algebra} of ${\mathcal M}_g$. Faber has conjectured that this ring has the structure of the $(p,p)$ part of the cohomology ring of a smooth complex projective variety of complex dimension $g-2$. That is, it satisfies Poincar\'e duality and has the ``Hard Lefschetz Property'' with respect to $\kappa _1$. Considerable evidence now exists for this conjecture, much of which is presented in Section~\ref{sec:chow}. Other developments on the Chow ring, such as explicit computations in low genus, are also surveyed there. In the early 80s, Harer proved that the cohomology in a given degree of ${\mathcal M}_g$ is independent of the genus once the genus is sufficiently large relative to the degree. These stable cohomology groups form a graded commutative algebra which is known to be free. The tautological classes $\kappa_i$ freely generate a polynomial algebra inside the stable cohomology ring. Mumford and others have conjectured that the stable cohomology of ${\mathcal M}_g$ is generated by the $\kappa_i$'s. Some progress has been made towards this conjecture which we survey throughout the paper. In Section~\ref{sec:agstability} we consider the stabilization maps from an algebro-geometric point of view, and in Section~\ref{sec:algebras} we survey Kontsevich's methods for constructing classes in the cohomology of the ${\mathcal M}_g^n$. We have also tried to advertise the fecund work of Dennis Johnson on the Torelli groups. The Torelli group $T_g$ is the subgroup of the mapping class group $\Gamma_g$ consisting of those diffeomorphism classes that act trivially on the homology of the reference surface. This mysterious group, in some sense, measures the difference between curves and abelian varieties and appears to play a subtle role in the geometry of ${\mathcal M}_g$. Johnson proved that $T_g$ is finitely generated when $g\ge 3$ and computed its first integral homology group. These computations have direct geometric applications, especially when combined with M.~Saito's work in Hodge theory --- for example, they restrict the normal functions defined over ${\mathcal M}_g$ and its standard level covers. {}From this, one can give a computation of the Picard group of the generic curve with a level $l$ structure. Johnson's work and its applications is surveyed in Section~\ref{sec:torelli}. Since $\Gamma_g$ is the orbifold fundamental group of ${\mathcal M}_g$, an algebraic variety, one should be able to apply Hodge theory and Galois theory to study its structure. In Section~\ref{sec:hodgemap} we survey recent work on applications of Hodge theory to understanding the structure of the Torelli groups, mainly via Malcev completion. In Section~\ref{sec:algebras} we combine this Hodge theory with recent results of Kawazumi and Morita to show that the cohomology of ${\mathcal M}_g$ constructed by Kontsevich using graph cohomology are, after stabilization, polynomials in the $\kappa_i$'s. Thus Hodge theory provides some evidence for Mumford's conjecture that the stable cohomology of the mapping class group is generated by the $\kappa_i$'s. Some of the results we discuss have not yet appeared in the literature, at least not in the form in which we present them. Rather than mention all such results, we simply mention a few instances where we believe our presentation to be novel: the correspondences in Section~\ref{subsec:correspondences}, the r\^ole of the fundamental normal function for orbifold fundamental groups in Section~\ref{fundgroup}, Theorem~\ref{tautbound} and the contents of Section~\ref{subsec:relation}. \medskip \noindent{\it Notation and Conventions.} All varieties will be defined over ${\mathbb C}$ unless explicitly stated to the contrary. Unless explicit coefficients are used, all (co)homology groups are with rational coefficients. We will often abbreviate {\it mixed Hodge structure} by MHS. The sub- or superscript {\it pr} on a (co)homology group will denote the primitive part in both the context of the Hard Lefschetz Theorem and in the context of Hopf algebras. \medskip \noindent{\it Acknowledgements.} We would like to thank Carel Faber for his comments on part of an earlier version of this paper and Shigeyuki Morita for explaining to us some of his recent work. We also appreciate the useful comments by a referee. We gratefully acknowledge support by the AMS that enabled us to attend this conference. \section{Mapping Class Groups} \label{sec:groups} Fix a compact connected oriented reference surface $S_g$ of genus $g$, and a sequence of distinct points $(x_0,x_1,x_2,\dots )$ in $S_g$. Let us write $S_g^n$ for the open surface $S-\{x_1,\dots ,x_n\}$ and $\pi_g^n$ for its fundamental group $\pi_1(S_g^n ,x_0)$. This group admits a presentation with generators $\alpha _{\pm 1},\dots ,\alpha_{\pm g},\beta_1,\dots ,\beta_n$ and relation $$ (\alpha_1,\alpha_{-1})\cdots (\alpha_g,\alpha_{-g})=\beta_1\cdots\beta_n, $$ where $(x,y)$ denotes the commutator of $x$ and $y$.\footnote{For $n=0$ the righthand side is to be interpreted as the unit element.} The generators are represented by loops that do not meet outside the base point; $\beta_i$ is represented by a loop that follows an arc to a point close to $x_i$, makes a simple loop around $x_i$, and returns to the base point along the same arc. Let $\Diff^+(S)^n_r$ denote the group of orientation preserving diffeomorphisms of $S$ that fix the $x_i$ for $i=1,\dots ,n+r$, and are the identity on $T_{x_i}S$ for $i=n+1,\dots ,n+r$. Although not really necessary at this stage, it is convenient to assume that $2g-2+n+2r>0$. In other words, we do not consider the cases where $(g,n,r)$ is $(0,0,0)$, $(0,1,0)$, $(0,0,1)$, $(0,2,0)$ or $(1,0,0)$. We will keep this assumption throughout the paper. The {\it mapping class group} $\Gamma_{g,r}^n$ is defined to be the group of connected components of this group: $$ \Gamma_{g,r}^n = \pi_0 \Diff^+(S)^n_r. $$ We omit the decorations $n$ and $r$ when they are zero. The mapping class group $\Gamma_g^n$ acts on $\pi_g^n$ by outer automorphisms. A theorem that goes back to Baer (1928) and Nielsen (1927) \cite{nielsen} identifies $\Gamma_g$, via this representation, with the subgroup of $\Out (\pi_g)$ (of index two) that acts trivially on $H_2(\pi_g)\cong H_2(S_g)$. When $n\ge 1$ we can consider the diagonal action of $\Aut (\pi_g^n)$ on $(\pi_g^n)^n$. Clearly, $\Out (\pi_g^n)$ acts on the set of orbits of $\pi_g^n$ (which acts by inner automorphisms on each component) in $(\pi_g^n)^n$. Now $\Gamma_g^n$ can be identified with the group of outer automorphisms of $\pi_g^n$ that preserve the image of $(\beta_1,\dots ,\beta_n)$ in $\pi_g^n\backslash (\pi_g^n)^n$. If we choose $x_n$ as a base point, then a corresponding assertion holds: $\Gamma_g^n$ can be identified with a subgroup of $\Aut (\pi_1(S_g^{n-1} ,x_n))$ that is characterized in a similar way. The evident homomorphism $\Gamma_g^n\to \Gamma_g^{n-1}$ is surjective and its kernel can be identified with $\pi_1(S_g^{n-1} ,x_n)$ (acting by inner automorphisms). Ivanov and McCarthy \cite{ivanov-mccarthy} recently showed that the resulting exact sequence cannot be split. \subsection{Generators and basic properties}\label{subsec:basic} Although a lot is known about these groups they are still poorly understood. Let us quickly review some of their basic properties. Dehn proved in \cite{dehn} that the mapping class groups are generated by the `twists' that are now named after him: if $\alpha$ is a simple (unoriented) loop on $S_g^{n+r}$, then parameterize a regular neighborhood of $\alpha $ in $S_g^{n+r}$ by the cylinder $[0,1]\times S^1$ (preserving orientations) and define an automorphism of $S_g$ that on this neighborhood is given by $(t,z)\mapsto (t ,e^{2\pi it}z)$ and is the identity elsewhere. The isotopy class of this automorphism only depends on the isotopy class of $\alpha $ and is called the {\it Dehn twist} along $\alpha$. (Perhaps we should add that $\alpha$ is, in turn, already determined by its free homotopy class, in other words, by the associated conjugacy class in $\pi_g^n$.) The corresponding element of $\Gamma_{g,r}^n$ is the identity precisely when $\alpha$ bounds a disk in $S_g-\{ x_{n+1},\dots ,x_{n+r}\}$ which meets $\{ x_1,\dots ,x_{n}\}$ in at most one point. Several people have found a finite presentation for the mapping class groups. One with few generators was given by Waynryb \cite{waynryb}. From this presentation one sees that the mapping class groups considered here are perfect when $g \ge 3$ (a result due to Powell \cite{powell} in the undecorated case). There is an obvious homomorphism $\Gamma_{g,n}\to \Gamma_g^n$. It is easy to see that it is surjective and that the kernel is generated by the Dehn twists around the points $x_1,\dots ,x_n$. These Dehn twists generate a free abelian central subgroup of $\Gamma_{g,n}$ of rank $n$. Now recall that a central extension of a discrete group $G$ by ${\mathbb Z}$ determines an extension class in $H^2(G;{\mathbb Z} )$; it has a geometric interpretation as a first Chern class. In the present case we have $n$ such classes $\tau _i\in H^2(\Gamma_g^n;{\mathbb Z} )$, $i=1,\dots ,n$. Conversely, each subgroup of $H^2(G;{\mathbb Z} )$ determines a central extension of $G$ by that subgroup. Harer proved that $H^2(\Gamma _{g,r};{\mathbb Z} )$ is infinite cyclic if $g\ge 3$ \cite{harer:h2}, so that there is a corresponding central extension $$ 0\to{\mathbb Z}\to\widetilde\Gamma _{g,r}\to\Gamma _{g,r}\to 1. $$ Since $H_1(\Gamma_{g,r};{\mathbb Z} )$ vanishes, this central extension is perfect (and universal). A nice presentation of it was recently given by Gervais \cite{gervais}. The (imperfect) central extension by $\frac{1}{12}{\mathbb Z}$ containing this extension appears in the theory of conformal blocks; it has a simple geometric description which we will give in Section~\ref{sec:moduli}. \subsection{Stable cohomology}\label{subsec:stable} The mapping class groups $\Gamma_{g,r}^n$ turn up in a connected sum construction that we describe next. It is convenient to do this in a somewhat abstract setting. Suppose we are given a closed, oriented (but not necessarily connected) surface $S$, a finite subset $Y\subset S$, and a fixed point free involution $\iota $ of $Y$. Assume that $\iota $ has been lifted to an orientation reversing linear involution $\tilde\iota$ on the spaces of rays $\Ray (TS|Y)$. The {\it real oriented blow up} $S_Y\to S$ is a surface with boundary canonically isomorphic to $\Ray (TS|Y)$. So $\tilde\iota$ defines an orientation reversing involution of this boundary. Welding the boundary components of $S_Y$ by means of this involution produces a closed surface $S({\tilde\iota})$. Some care is needed to give it a differentiable structure inducing the given one on $S_Y$. Although there is no unique way to do this, all natural choices lie in the same isotopy class. If $S$ happens to have a complex structure, then each choice of a real ray $L$ in $T_pS\otimes_{{\mathbb C}} T_{\iota (p)}S$ determines a lift of $\iota$ over the pair $\{p, \iota (p)\}$: if $l$ is a ray in $T_pS$, then $\tilde\iota (l)$ is determined uniquely by the condition $l\otimes_{{\mathbb C}}\tilde\iota (l)=L$. If $S({\tilde\iota})$ is connected, then each finite subset $X$ of $S-Y$ determines a natural homomorphism from the mapping class group which is perhaps best denoted by $\Gamma (S)_Y^X$ (a product of groups of the type $\Gamma_{g,r}^n$) to the mapping class group $\Gamma (S(\tilde\iota))^X$. The image of this homomorphism is simply the stabilizer of the simple loops indexed by $Y/\iota $ that are images of boundary components of $S_Y$. Its kernel is a free abelian group whose generators can be labeled by a system of representatives $R$ of $\iota$ orbits in $Y$. Indeed, for each element $y$ of $R$, take the composite of the Dehn twist around $y$ and the inverse of the Dehn twist around $\iota (y)$. These maps appear in the stability theorems and are at the root of the recent operad theoretic approaches to the study of the cohomology of mapping class groups. \begin{theorem}[Stability theorem, Harer \cite{harer:stab}] There exists a positive constant $c$ with the following property. If $S(\tilde\iota )$ is connected and $S'$ is a connected component of $S$ and $X$ a finite subset of $S'\setminus Y$, then the homomorphism $$ \Gamma (S')_{Y\cap S'}^X\to \Gamma (S(\tilde\iota ))^X $$ induces an isomorphism on integral cohomology in degree $\le c.\text{genus}(S')$. \end{theorem} The constant $c$ appearing in this theorem was $1/3$ in Harer's original paper. It was later improved to $1/2$ by Ivanov in \cite{ivanov:teichm}. Most recently, Harer \cite{harer:imp_stab} has showed that we can take $c$ to be about $2/3$ and that this is the minimal possible value. There is also a version for twisted coefficients, due to Ivanov \cite{ivanov}. Harer's theorem says essentially that the $k$th cohomology group of $\Gamma_{g,r}^n$ depends only on $n$, provided that $g$ is large enough. These stable cohomology groups are the cohomology of a single group, namely the group $\Gamma_{\infty}^n$ of compactly supported mapping classes of a surface $S_{\infty}$ of infinite genus (with one end, say) that fix a given set of $n$ distinct points. Among the homomorphisms defined above are maps $\Gamma_{g,1}\times \Gamma_{g',1}\to \Gamma_{g+g'}$. These stabilize and define homomorphisms of ${\mathbb Q}$ algebras $$ \mu : H^{\bullet}(\Gamma_{\infty})\to H^{\bullet}(\Gamma_{\infty})\otimes H^{\bullet}(\Gamma_{\infty}). $$ This defines a coproduct on $H^{\bullet}(\Gamma_{\infty})$. Together with the cup product, this gives $H^{\bullet}(\Gamma_{\infty})$ the structure of a connected graded-bicommutative Hopf algebra. The classification of such Hopf algebras implies that $H^{\bullet}(\Gamma_{\infty})$ is free as a graded algebra and is generated by its set of primitive elements $$ H^{\bullet}_{{\rm pr}}(\Gamma_{\infty}) := \{x\in H^+(\Gamma_{\infty}): \mu (x)=x\otimes 1+1\otimes x\}. $$ For each $i>0$, Mumford \cite{mumford} and Morita \cite{morita:classes} independently found a class $\kappa_i$ in $H^{2i}_{{\rm pr}}(\Gamma_{\infty})$ (we shall recall the definition in Section~\ref{sec:agstability}) and Miller \cite{miller} and Morita \cite{morita:classes} independently showed that each $\kappa_i$ is nonzero. So the $\kappa_i$'s generate a polynomial subalgebra of the stable cohomology. Mumford conjectured that they span all of $H^{\bullet}_{{\rm pr}}(\Gamma_{\infty})$. This has been verified by Harer in a series of papers \cite{harer:h2}, \cite{harer:h3}, \cite{harer:h4} in degrees $\le 4$.\footnote{He also tells us that he has checked that there are no stable primitive classes in degree 5.} The first Chern class $\tau _i\in H^2(\Gamma_g^n;{\mathbb Z} )$ stabilizes also and we may think of it as an element of $H^{\bullet}(\Gamma_{\infty}^n;{\mathbb Z} )$ ($i=1,\dots ,n$). The forgetful map $\Gamma_{\infty}^n\to \Gamma_{\infty}$ gives $H^{\bullet}(\Gamma_{\infty}^n)$ the structure of a module over this Hopf algebra. From the stability theorem one can deduce: \begin{theorem}[Looijenga \cite{looijenga}] The algebra $H^{\bullet}(\Gamma_{\infty}^n;{\mathbb Z} )$ is freely generated by the classes $\tau _1,\dots ,\tau _n$ as a graded-commutative $H^{\bullet}(\Gamma_{\infty};{\mathbb Z} )$ algebra. \end{theorem} \section{Moduli Spaces} \label{sec:moduli} A conformal structure and an orientation on $S_g$ determine a complex structure on $S_g$. The {\it Teichm\"uller space} ${\mathcal X}_{g,r}^n$ is the space of conformal structures on $S_g$ (with some reasonable topology) up to isotopies that fix $\{ x_1,\dots ,x_{n+r}\}$ pointwise and act trivially on the tangent spaces $T_{x_i}S$ for $i=n+1,\dots, n+r$. It is, in a natural way, a complex manifold of dimension $3g-3+n+2r$. As a real manifold it is diffeomorphic to a cell. The group $\Gamma_{g,r}^n$ acts naturally on it. This action is properly discontinuous and a subgroup of finite index acts freely. If $\Gamma$ is any subgroup of $\Gamma_{g,r}^n$ that acts freely, then the orbit space $\Gamma\backslash {\mathcal X}_{g,r}^n$ is a classifying space for $\Gamma$ and so its singular integral cohomology coincides with $H^{{\bullet}}(\Gamma ;{\mathbb Z} )$. This is even true with twisted coefficients: if $V$ is a $\Gamma$ module, then the trivial sheaf over ${\mathcal X}_{g,r}^n$ with fiber $V$ comes with an obvious (diagonal) action of $\Gamma$. Passing to $\Gamma$ orbits yields a locally constant sheaf ${\mathbb V}$ on $\Gamma\backslash {\mathcal X}_{g,r}^n$. The cohomology of this sheaf equals $H^{{\bullet}}(\Gamma ;V)$. For an arbitrary subgroup $\Gamma$ of $\Gamma_{g,r}^n$ these statements still hold as long as we take our coefficients to be ${\mathbb Q}$ vector spaces (but ${\mathbb V}$ need no longer be locally constant). For $\Gamma =\Gamma_{g,r}^n$, we denote the orbit space by ${\mathcal M}_{g,r}^n$. The space ${\mathcal M}_{g,r}^n$ is, in a natural way, a normal analytic space and the obvious forgetful maps such as ${\mathcal M}_{g,r}^n\to {\mathcal M}_g^n$ are analytic. An interpretation as a coarse moduli space makes it possible to lift this analytic structure to the algebraic category. To see this, we first choose a nonzero vector in each tangent space $T_{x_i}S_g$. Each triple $(C;x,v)$, where $C$ is a connected nonsingular complex projective curve $C$ of genus $g$, $x$ an injective map $x:\{ 1,\dots ,n+r\} \to C$, and $v$ a nowhere zero section of $TC$ over $\{n+1,\dots ,n+r\}$, determines an element of ${\mathcal M}_{g,r}^n$. This point depends only on the isomorphism class of $(C,x,v)$ with respect to the obvious notion of isomorphism. Since each conformal structure on $S$ gives $S$ the structure of a nonsingular complex projective curve, ${\mathcal M}_{g,r}^n$ can be identified can be identified with the space of isomorphism classes of such triples. From the work of Knudsen, Mumford and Deligne, we know that ${\mathcal M}_g^n$ is, in a natural way, a quasi-projective orbifold. Recall that they also constructed a projective completion $\overline{\M}_g^n$ of ${\mathcal M}_g^n$, the {\it Deligne-Mumford completion} \cite{del_mum}, that also admits the interpretation of a coarse moduli space. Its points parameterize the connected stable $n$ pointed curves $(C,x)$ of arithmetic genus $g$, where we now allow $C$ to have ordinary double points, but still require $x$ to map to the smooth part of $C$ and the automorphism group of $(C,x)$ to be finite. The {\it Deligne-Mumford boundary} $\overline{\M}_g^n-{\mathcal M}_g^n$ is a normal crossing divisor in the orbifold sense. There is a projective morphism $\overline{\M}_g^{n+1}\to\overline{\M}_g^n$, defined by forgetting the last point. It comes with $n$ sections $x_1,\dots ,x_n$. The fibers of this morphism are stable $n$ pointed curves (modulo finite automorphism groups) and the morphism can be regarded as the universal stable $n$ pointed curve (in an orbifold sense). Let $\omega$ denote the relative dualizing sheaf of this morphism, considered as a line bundle in the orbifold sense. We can then think of ${\mathcal M}_{g,r}^n$ as the set of $(v_{n+1},\dots ,v_{n+r})$ in the total space of $x_{n+1}^*\omega \oplus\cdots \oplus x_{n+r}^*\omega$ restricted to ${\mathcal M}_g^{n+r}$ that have each component nonzero. So ${\mathcal M}_{g,r}^n$ is also quasi-projective. Each finite quotient group $G$ of $\Gamma_g^n$ determines, in an obvious way, a Galois cover ${\mathcal M}_g^n[G]\to {\mathcal M}_g^n$. The Deligne-Mumford completion $\overline{\M}_g^n[G]$ of this cover is, by definition, the normalization of $\overline{\M}_g^n$ in ${\mathcal M}_g^n[G]$. \begin{theorem}[Looijenga \cite{looijenga:cover}] There exists a finite group $G$ such that $\overline{\M}_g[G]$ is smooth with a normal crossing divisor as Deligne-Mumford boundary. \end{theorem} This has been extended by De Jong and Pikaart \cite{jong} to arbitrary characteristic, and by Boggi and Pikaart (independently) to the $n$-pointed case. (They show that it also can be arranged that each irreducible component of the Deligne-Mumford boundary of $\overline{\M}_g^n[G]$ is smooth.) This makes it relatively easy to define the Chow algebra of $\overline{\M}_g^n$: if $\overline{\M}_g^n[G]$ is smooth, then define $\CH^{{\bullet}} (\overline{\M} _g^n)$ to be the $G$ invariant part of $\CH^{{\bullet}} (\overline{\M} _g^n[G])$ (we take algebraic cycles modulo rational equivalence with coefficients in ${\mathbb Q}$). It is easy to see that this is independent of the choice of $G$. The central extension of $\Gamma _g$ by ${\mathbb Z}$ ($g\ge 3$) discussed in Section~\ref{subsec:basic} takes the geometric form of a complex line bundle over Teichm\"uller space with $\Gamma _g$ action and hence yields an orbifold line bundle over ${\mathcal M}_g$. Its twelfth tensor power has a concrete description: it is the determinant bundle of the direct image of the relative dualizing sheaf of ${\mathcal M}_g^1\to{\mathcal M}_g$ (this is a rank $g$ vector bundle). The orbifold fundamental group of the associated ${\mathbb C}^{\times}$ bundle is just the central extension of $\Gamma_g$ by $\frac{1}{12}{\mathbb Z}$ mentioned in Section~\ref{subsec:basic}. Since $\Gamma_g$ is perfect when $g\ge 3$, we have $H^1(\Gamma _g)=0$. Ivanov has asked the following question: \begin{question}[Ivanov] Is it true that $H^1(\Gamma)$ vanishes for all finite index subgroups $\Gamma$ of $\Gamma_g$, at least when $g$ is sufficiently large? \end{question} This would imply that the Picard group of each finite unramified cover of ${\mathcal M}_g$ (in the orbifold sense) is finitely generated. The answer to Ivanov's questions is affirmative, for example, for subgroups of finite index of $\Gamma _g$, $g\ge 3$, that contain the Torelli group --- see (\ref{van_h1}). \section{Algebro-Geometric Stability} \label{sec:agstability} The Deligne-Mumford completion $\overline{\M}_g^n$ comes with a natural stratification into orbifolds, with each stratum parameterizing stable $n$ pointed curves of a fixed topological type $T$. Denote this stratum by ${\mathcal M} (T)$. It has codimension equal to the number of singular points of $T$. The normalization of the topological type $T$ is an oriented closed surface $S$ that comes with $n$ distinct numbered points $X=\{ x_1,\dots ,x_n\}$ and a finite subset $Y$ of $S-X$ with a fixed point free involution $\iota$, so that $T$ is recovered by identifying the points of $Y$ according to $\iota$. These topological data define a moduli space ${\mathcal M} (S)^{X\cup Y}$ of the same type (we hope that the notation is self-explanatory) and there is a natural morphism ${\mathcal M} (S)^{X\cup Y}\to {\mathcal M}(S/\iota )^X$ that is a Galois cover of orbifolds. This morphism extends to a finite surjective morphism from the Deligne-Mumford completion $\overline{\M} (S)^{X\cup Y}$ to the closure of ${\mathcal M} (T)$ in $\overline{\M}_g^n$. The resulting morphism $\overline{\M} (S)^{X\cup Y}\to\overline{\M}_g^n$ has only self-intersections of normal crossing type and so carries a normal bundle in the orbifold sense. This normal bundle is a direct sum of line bundles with one summand for each $\iota$ orbit $\{ p,p'\}$, namely $p^*\omega^{-1}\otimes p'{}^*\omega^{-1}$. (To see this, notice that the restriction of the universal curve to $\overline{\M} (T)$ has a quadratic singularity along the locus defined by the pair $\{ p,p'\}$. Associating to a local defining equation its hessian determines a natural isomorphism between $p^*\omega^{-1}\otimes p'{}^*\omega^{-1}$ and the normal bundle of a divisor in the Deligne-Mumford boundary passing through $\overline{\M}(T)$.) We now see before us an algebro-geometric incarnation of the map that appears in the stability theorem: the set of normal vectors that point towards the interior ${\mathcal M}_g^n$ is the restriction to ${\mathcal M}(S)$ of the total space of the direct sum of ${\mathbb C}^{\times}$ bundles in this normal bundle. So ${\mathcal M} (S)_Y^X$ maps to the latter space, and although we do not have a morphism ${\mathcal M} (S)_Y^X\to {\mathcal M}_g^n$, the map on cohomology behaves as if there were. In particular, the map $H^{{\bullet}}({\mathcal M}_g^n)\to H^{{\bullet}}({\mathcal M} (S)_Y^X)$ is a MHS morphism. So the stability theorem implies: \begin{theorem} [Algebro-geometric stability] Suppose that the finite set $X$ is contained in a connected component $S'$ of $S$ of genus $g'$, so that ${\mathcal M} (S')^X_{Y\cap S'}$ appears as a factor of ${\mathcal M}(S)_Y^X$. Choose points in the remaining factors so that we have an inclusion of ${\mathcal M} (S')^X_{Y\cap S'}$ in ${\mathcal M} (S)_Y^X$. Then for $k\le cg'$ the composite map $$ H^k({\mathcal M}_g^n)\to H^k({\mathcal M} (S)_Y^X)\to H^k({\mathcal M} (S')^X_{Y\cap S'}) $$ is an isomorphism and so is the map $$ H^k({\mathcal M}_{g'}^n)\to H^k({\mathcal M} (S')^X_{Y\cap S'}) $$ induced by the forgetful morphism ${\mathcal M} (S')^X_{Y\cap S'}\to {\mathcal M} (S')^X\cong {\mathcal M} _{g'}^n$. These maps are also MHS morphisms. \end{theorem} So the stable rational cohomology $H^{{\bullet}}(\Gamma_{\infty}^n)$ comes with a natural MHS. A geometric consequence of this result is that each stable rational cohomology class of ${\mathcal M}_g^n$ (that is, a class whose degree is in the stability range) extends across the open part of the blow up of $\overline{\M} (T)$ parameterizing the normal directions pointing towards the interior. Pikaart showed that these partial extensions can be made to come from a single extension to $\overline{\M}_g^n$, at least if $g$ is large compared with $k$. But then it is not hard to show that if this is possible for large $g$, then it is possible in the stable range and so the conclusion is: \begin{theorem}[Pikaart \cite{pikaart}]\label{pikaart_purity} The restriction map $H^k(\overline{\M}_g^n)\to H^k({\mathcal M}_g^n)$ is surjective in the stable range. Consequently, the MHS on $H^k(\Gamma_{\infty}^n)$ is pure of weight $k$. \end{theorem} Mumford's Conjecture, if known, would imply this result, and so Pikaart's Theorem is evidence for the truth of this conjecture. We illustrate this theorem with the known stable classes. We have seen in the previous section that $\overline{\M}_g^n$ comes with $n$ orbifold line bundles $x_i^*\omega$, $i=1,\dots ,n$. Let $\overline{\tau}_{n,i}$ denote the first Chern class of this line bundle, regarded as an element of $\CH^1(\overline{\M}_g^n)$. The restriction of this class to $\CH^1({\mathcal M}_g^n)$ is a pull-back of the restriction of $\overline{\tau} _{n-1,i}$ to ${\mathcal M}_g^{n-1}$ (when $n\ge 1$) and so we denote that restriction simply by $\tau_i$. The underlying cohomology class of $\tau_i$ in $H^{2i}({\mathcal M}_g^n)\cong H^{2i}(\Gamma_g^n)$ is what we denoted earlier by that symbol, in particular, it is stable. For the definition of the tautological classes of ${\mathcal M} _g^n$, we shall not use Mumford's original definition, but a modification proposed by Arbarello-Cornalba. This might begin with the observation that the ``functor'' which associates to an $(n+1)$-pointed stable genus $g$ curve $(C;x_1,\dots ,x_n,x)$ the cotangent space $T_x^*C$ defines an orbifold line bundle over $\overline{\M}_g^{n+1}$. It is not quite the same as the relative dualizing sheaf $\omega$ of the forgetful map $\overline{\M}_g^{n+1}\to \overline{\M}_g^n$: a little computation shows that it is in fact $\omega (\sum_{i=1}^n (x_i))$. This is perhaps a more natural bundle to consider than $\omega$. In any case, we denote the direct image of $c_1(\omega (\sum_{i=1}^n (x_i))^{i+1}\in\CH ^{i+1}(\overline{\M}_g^{n+1})$ under the projection $\overline{\M}_g^{n+1}\to \overline{\M}_g^n$ by $\overline{\kappa}_{n,i}\in \CH^i({\mathcal M}_g^n)$ and its restriction to ${\mathcal M}_g^n$ by $\kappa_{n,i}$. The cohomology class underlying $\kappa_{n,i}$ can be regarded as an element of $H^{2i}(\Gamma^n_g)$ (of Hodge bidegree $(i,i)$). These cohomology classes stabilize and, for $n=0$, they define the nonzero primitive elements of degree $2i$ alluded to in \ref{subsec:stable}. We regard (for $k=0,1,\dots ,n$) $\CH ^{{\bullet}}(\overline{\M} _g^n)$ as a $\CH ^{{\bullet}}(\overline{\M}_g^k)$-algebra via the obvious forgetful morphism, and view the classes $\overline{\kappa}_{k,i}$ as elements of $\CH ^{{\bullet}}(\overline{\M}_g^n)$ when appropriate. The class $\overline{\kappa}_{n,i}$ is then not equal to $\overline{\kappa}_{n-1,i}$, but according to formula (1.10) of \cite{arb_cor} we have: $$ \overline{\kappa}_{n,i}=\overline{\kappa}_{n-1,i} +(\overline{\tau}_{n,n})^i. $$ As Arbarello-Cornalba explain, the classes $\overline{\kappa}_{n,i}$ possess a nice property not enjoyed by Mumford's classes. First recall that every stratum of $\overline{\M}_g^n$ is the image of a finite map $\overline{\M} (S)^{X\cup Y}\to\overline{\M}_g^n$ and that $\overline{\M} (S)^{X\cup Y}$ is a product of varieties of the type $\overline{\M}_{g_\alpha}^{n_\alpha}$. The pull-back of $\overline{\kappa}_{n,i}$ along this map is the sum of the classes $\overline{\kappa}_{n_\alpha ,i}$ (pulled back along the projection $\overline{\M} (S)^{X\cup Y} \to \overline{\M}_{g_\alpha }^{n_\alpha}$). Carel Faber pointed out to us that a similar property is enjoyed by the divisor class of the Deligne-Mumford boundary, but we know of no other examples. Since this behaviour is reminiscent of that of a primitive element in a Hopf algebra under the coproduct, we ask: \begin{question} What other collections $\left\{\mu_{g,n}\in \CH ^k({\mathcal M}_g^n)\right\}_{g,n}$ have this property? \end{question} \subsection{Correspondences between moduli spaces}\label{subsec:correspondences} There is an altogether different way to relate the cohomology of the moduli spaces ${\mathcal M}_g^n$ for different values of $g$. This involves certain Hecke type correspondences. For simplicity we shall restrict ourselves to the undecorated case $n=0$. We return to the reference surface $S_g$ and suppose that we are given a subgroup $\pi$ of $\pi_g$ of finite index $d$, say. (For what follows only its conjugacy class will matter.) This subgroup determines an unramified finite covering $\tilde S\to S$ of closed oriented surfaces. The genus $\tilde g$ of $\tilde S$ is then equal to $d(g-1)+1$. Consider the group of pairs $(\tilde h,h)\in \Diff^+(\tilde S)\times \Diff^+(S)$ such that $\tilde h$ is a lift of $h$. Let $\Gamma_g(\pi )$ be its group of connected components. The projection $\Gamma_g(\pi )\to \Gamma_g$ has as kernel the group of covering transformations of $\tilde S \to S$ (so is finite) and its image consists of the outer automorphisms of $\pi_g$ that come from an automorphism which preserves the subgroup $\pi$ (so is of finite index, $e$, say). There is a corresponding finite covering of moduli spaces $p_1:{\mathcal M}_g(\pi)\to {\mathcal M}_g$, where ${\mathcal M}_g(\pi )$ is simply the coarse moduli space of finite unramified coverings of nonsingular complex projective curves $\tilde C\to C$ topologically equivalent to $\tilde S\to S$. There is also a finite map $p_2:{\mathcal M}_g(\pi)\to {\mathcal M}_{\tilde g}$. Together they define a one-to-finite correspondence $p_2p_1^{-1}$ from ${\mathcal M}_g$ to ${\mathcal M}_{\tilde g}$. This extends over the Deligne-Mumford compactifications: if $p_1:\overline{\M}_g(\pi )\to \overline{\M}_g$ denotes the normalization of $\overline{\M}_g$ in ${\mathcal M}_g(\pi)$, then $p_2$ extends to a finite morphism $p_2 :\overline{\M}_g(\pi)\to\overline{\M}_{\tilde g}$. We have an induced map $$ T_{\pi}:=e^{-1}p_{1*}p_2^*: \CH^{{\bullet}}(\overline{\M}_{\tilde g})\to \CH^{{\bullet}}(\overline{\M}_g) $$ and likewise on cohomology. A computation shows that any monomial in the tautological classes is an ``eigen class'' for such correspondences: \begin{proposition} The map $T_{\pi}$ sends $\overline{\kappa}_{i_1}\overline{\kappa}_{i_2}\cdots \overline{\kappa}_{i_r}$ to $d^r\overline{\kappa}_{i_1}\overline{\kappa}_{i_2}\cdots \overline{\kappa}_{i_r}$. \end{proposition} This proposition suggests the consideration, for given positive integers $r$ and $s$, of sequences of classes $(x_g\in\CH^s(\overline{\M}_g))_{g\ge 2}$ of fixed degree that have the property that $T_{\pi}(x_g)=d^r x_{(d-1)g+1}$ for each index $d$ subgroup $\pi$ of $\pi_g$. \begin{question} Is for such a system the image of $x_g$ in $H^{{\bullet}}({\mathcal M}_g)$ stable? Is it in fact a polynomial of degree $r$ in primitive stable classes? \end{question} An affirmative answer would give us a notion of stability for the Chow groups of the moduli spaces ${\mathcal M}_g$. \section{Chow Algebras and the Tautological Classes} \label{sec:chow} We have already encountered some of the basic classes on $\overline{\M}_g^n$: the first Chern classes $\overline{\tau}_i\in \CH^1(\overline{\M}_g^n)$ ($i=1,\dots ,n$) and the tautological classes $\overline{\kappa}_i$ ($i=1,2,\dots $). More such classes come from the boundary: if $\prod _i\overline{\M} _{g_i}^{n_i}\to \overline{\M}_g$ is a Galois covering of a stratum of the boundary as in Section~\ref{sec:agstability}, then we can add to these the push-forwards along this map of the exterior products of the corresponding classes on the factors. Let us call the subalgebra of $\CH^{{\bullet}}(\overline{\M}_g^n)$ generated by all these classes the {\it tautological subalgebra} and denote it by ${\mathcal R}^{{\bullet}}(\overline{\M}_g^n)$. The image of this algebra in $\CH^{{\bullet}}({\mathcal M}_g^n)$ is denoted by ${\mathcal R}^{{\bullet}}({\mathcal M}_g^n)$; it is generated by the classes $\kappa_{n,i}$ ($i=1,2,\dots $) and $\tau _i$ ($i=1,\dots ,n$). It is possible that these classes generate the rational Chow ring of $\overline{\M}_g^n$ modulo homological equivalence, but this is of course unknown. In any case, these subalgebras are preserved under pull-back and push-forward along the natural maps that we have met so far. The first computations were done by Mumford \cite{mumford} who found a presentation of $\CH^{{\bullet}}(\overline{\M}_2)$. Subsequently Faber \cite{faber} calculated $\CH^{{\bullet}}(\overline{\M}_2^1)$, $\CH^{{\bullet}}(\overline{\M}_3)$ and obtained partial results on $\CH^{{\bullet}}(\overline{\M}_4)$. In all these cases the tautological algebra is the whole Chow algebra. This is also the case for $\overline{\M}_0^n$, whose Chow algebra was computed by Keel. This is a very remarkable algebra which appears in other contexts. Because of this, we describe it explicitly. We first introduce notation for the divisor classes on $\overline{\M}_0^n$. The boundary divisor $\overline{\M}_0^n - {\mathcal M}_0^n$ parameterizes all singular stable $n$ pointed rational curves. Its components correspond to the topological types of $n$ pointed stable rational curves with exactly one singular point. Such curves have exactly two irreducible components. By collecting the points $x_i$ lying on the same component, we obtain a partition $P$ of $\{1,\dots ,n\}$ into two subsets. The stability property implies that both members of $P$ have at least two elements. We denote the corresponding class in $\CH^1(\overline{\M}_0^n)$ by $D(P)$. \begin{theorem}[Keel \cite{keel}] The Chow algebra $\CH^{\bullet}(\overline{\M}_0^n)$ coincides with $H^{\bullet}(\overline{\M}_0^n)$ and, as a ${\mathbb Q}$ algebra, is generated by the $D(P)$'s subject to the following relations: \begin{enumerate} \item[(i)] If $\{i,j,k\}$ are distinct integers in $\{1,\dots ,n\}$, then the sum of the $D(P)$'s for which $P$ separates $i$ from $\{ j,k\}$ is independent of $j$ and $k$ (and equals $\overline{\tau}_i$). \item[(ii)] $D(P)\cdot D(P')=0$ if $P$ and $P'$ are independent in the sense that the partition they generate has four nonempty members. \end{enumerate} \end{theorem} The relations (ii) are geometrically obvious since the divisors $D(P)$ and $D(P')$ do not meet if $P$ and $P'$ are independent. The additive relations (i) are not difficult to see either: if $C$ is a stable $n$ pointed rational curve, then a moment of thought shows that there is a unique morphism $z:C\to {\mathbb P}^1$ that is an isomorphism on one irreducible component, constant on the other irreducible components, and is such that $z(x_i)=1$, $z(x_j)=0$ and $z(x_k)=\infty$. The differential $z^{-1}dz$ restricted to $x_i$ defines a section of $x^*_i\omega $. The image of $z^{-1}dz$ in $T^*_{x_i}C$ vanishes precisely when $z$ collapses the irreducible component containing $x_i$. In \cite{manin:trees}, Manin derives a formula for the Poincar\'e polynomial of $\overline{\M}_0^n$. Such a formula was independently found by Getzler \cite{getzler} who also obtained the $\mathcal{S}_n$ equivariant Poincar\'e polynomial of $H^{\bullet}(\overline{\M}_0^n)$. That is, he determined the character of the $\mathcal{S}_n$ representations $H^k(\overline{\M}_0^n)$, $k\ge 0$. Kaufmann \cite{k_m:product} recently gave a formula for the intersection number of classes of strata of complementary dimension. We now turn to the Chow and cohomology algebras of the moduli spaces ${\mathcal M}_g$. First we list some results about the Chow algebras. \begin{enumerate} \item[]$\CH^{{\bullet}}({\mathcal M}_1^n)={\mathbb Q}$ for $n=1,2$ (folklore)\par \item[]$\CH^{{\bullet}}({\mathcal M}_2 )={\mathbb Q}$ (folklore),\par \item[]$\CH^{{\bullet}}({\mathcal M}_2^1)={\mathbb Q} [\tau]/(\tau^2)$ (Mumford \cite{mumford})\par \item[]$\CH^{{\bullet}}({\mathcal M}_3 )={\mathbb Q} [\kappa_1]/(\kappa_1^2)$ (Faber \cite{faber}),\par \item[]$\CH^{{\bullet}}({\mathcal M}^1_3)={\mathbb Q} [\kappa_1,\tau ]/(\kappa_1^2, 4\tau ^2-\tau\kappa_1)$ (Faber\cite{faber}),\par \item[]$\CH^{{\bullet}}({\mathcal M}_4 )={\mathbb Q} [\kappa_1]/(\kappa_1^3)$ (Faber \cite{faber}),\par \item[]$\CH^{{\bullet}}({\mathcal M}_5 )={\mathbb Q} [\kappa_1]/(\kappa_1^4)$ (Izadi \cite{izadi} combined with Faber \cite{faber:hyp}). \end{enumerate} The reason that such computations can be made is that, when $g$ and $n$ are both small, the moduli space ${\mathcal M}_g^n$ has a concrete description. For example, when $g=2$, each curve is hyperelliptic and therefore given by configuration of $6$ points on the projective line. In the case $g=3$ a nonhyperelliptic curve is realized by its canonical system as a quartic curve in ${\mathbb P}^2$. The double cover of the projective plane along this curve is a Del Pezzo surface of degree $2$, i.e., is obtained by blowing up $7$ points in the plane in general position. General curves of genus 4 and 5 can be described as complete intersections of multidegrees $(2,3)$ (in ${\mathbb P}^3$) and $(2,2,2)$ (in ${\mathbb P}^4$), respectively. \subsection{The tautological algebra of ${\mathcal M}_g$ and Faber's Conjecture} \label{subsec:tautalg} On the basis of numerous calculations, Faber, around 1993, made the following conjecture. \begin{conjecture}[Faber\cite{seminar}] The tautological algebra ${\mathcal R}^{{\bullet}}({\mathcal M}_g)$ is a graded Frobenius algebra with socle in degree $g-2$. That is, $\dim {\mathcal R} ^{g-2}({\mathcal M}_g)=1$, and the intersection product defines a nondegenerate bilinear form ${\mathcal R} ^i({\mathcal M}_g)\times {\mathcal R} ^{g-2-i}({\mathcal M}_g)\to {\mathcal R} ^{g-2}({\mathcal M}_g)$ $(i=0,\dots ,g-2)$. Moreover, $\kappa_1$ has the Lefschetz property in ${\mathcal R}^{{\bullet}}({\mathcal M}_g)$ in the sense that multiplication by $(\kappa_1)^{g-2-2i}$ maps ${\mathcal R}^i({\mathcal M}_g)$ isomorphically onto ${\mathcal R}^{g-2-i}({\mathcal M}_g)$ for $0\le i\le (g-2)/2$. \end{conjecture} Since the conjecture was made, evidence for it has been growing. For example: \begin{theorem}[Looijenga \cite{looijenga:taut}] The algebra ${\mathcal R}^{{\bullet}}({\mathcal M}_g)$ is trivial in degree $>g-2$ and ${\mathcal R}^{g-2}({\mathcal M}_g)$ is generated by the class of the hyperelliptic locus (a closed irreducible variety of codimension $g-2$). \end{theorem} In particular $\kappa_1^{g-1}=0$. Since $\kappa_1$ is ample on ${\mathcal M}_g$, we recover a theorem of Diaz \cite{diaz} which asserts that every complete subvariety of ${\mathcal M}_g$ must be of $\dim \le g-2$. Actually, in \cite{looijenga:taut} a stronger result is proven, which, among other things, implies that ${\mathcal R}^k({\mathcal M}_g^n)=0$ for $k>g-2+n$. An induction argument then shows that ${\mathcal R}^{3g-3+n}(\overline{\M}_g^n)$ is spanned by the classes of the zero dimensional strata. But zero dimensional strata can be connected by one dimensional strata and the one dimensional strata are all rational. This shows that ${\mathcal R}^{3g-3+n}(\overline{\M}_g^n)\cong{\mathbb Q}$. Faber recently proved that the tautological class $\kappa_{g-2}$ is nonzero. To describe his result, we find it convenient to introduce a compactly supported version of the tautological algebra: let ${\mathcal R}^{{\bullet}}_c({\mathcal M}_g^n)$ be defined as the set of elements in ${\mathcal R}^{{\bullet}}(\overline{\M}_g^n)$ that restrict trivially to the Deligne-Mumford boundary. This is a graded ideal in ${\mathcal R}^{{\bullet}}(\overline{\M}_g^n)$ and the intersection product defines a map $$ {\mathcal R} ^{{\bullet}}({\mathcal M}_g^n)\times{\mathcal R} ^{{\bullet}}_c({\mathcal M}_g^n)\to{\mathcal R} _c^{{\bullet}}({\mathcal M}_g^n) $$ that makes ${\mathcal R} _c^{{\bullet}}({\mathcal M}_g^n)$ a ${\mathcal R} ^{{\bullet}}({\mathcal M}_g^n)$-module. Notice that every complete subvariety of ${\mathcal M}_g$ of codimension $d$ whose class is in ${\mathcal R} ^{{\bullet}}(\overline{\M}_g^n)$ defines a nonzero element of ${\mathcal R}^d_c({\mathcal M}_g^n)$ (but it is by no means clear that such elements span ${\mathcal R}^{{\bullet}}_c({\mathcal M}_g^n)$). A somewhat stronger form of the first part of Faber's Conjecture is: \begin{conjecture}\label{strongfaber} The intersection pairings $$ {\mathcal R}^k({\mathcal M}_g)\times {\mathcal R}^{3g-3-k}_c({\mathcal M}_g)\to {\mathcal R}^{3g-3}_c({\mathcal M}_g)\cong{\mathbb Q} , \quad k=0,1,2,\dots $$ are perfect (Poincar\'e duality) and ${\mathcal R} ^{{\bullet}}_c({\mathcal M}_g)$ is a free ${\mathcal R}^{{\bullet}}({\mathcal M}_g)$ module of rank one. \end{conjecture} Faber \cite{faber:hyp} finds a compactly supported class $I_g$ in ${\mathcal R}_c^{2g-1}({\mathcal M}_g)$ with $\kappa_{g-2}\cdot I_g\neq 0$. So ${\mathcal R}^{{\bullet}}_c({\mathcal M}_g)$ should be the ideal generated by this element. Faber verified his conjecture for genera $\le 15$ by writing down many relations in ${\mathcal R} ^{{\bullet}}({\mathcal M} _g)$ (this evidently gives an upper bound) and using the nonvanishing of $\kappa_{g-2}$ (this gives a surprisingly strong lower bound). A refined form of Conjecture~\ref{strongfaber} (which we shall not state here) also takes care of the Lefschetz property. \begin{question} Does the tautological ring of $\overline{\M} _g^n$ satisfy Poincar\'e duality? Does it have the Lefschetz property with respect to $\overline{\kappa}_{n,1}$? (It is known that $\overline{\kappa}_{n,1}$ is ample \cite{cor}.) \end{question} \subsection{Cohomology of some moduli spaces}\label{subsec:somecohom} As may be expected, even less is known about the cohomology algebras. Here is an incomplete list of special results. In genus 0 we have that the Chow algebra of $\overline{\M} _0^n$ maps isomorphically onto its rational cohomology algebra. The cohomology of ${\mathcal M}_0^n$ is easily computed if we start out from the observation that this space is the projective arrangement of type $A_{n-2}$. It then follows for instance, that its cohomology in degree $p$ is of type $(p,p)$. There are similar descriptions of the moduli spaces of $n$-pointed hyperelliptic curves of genus $g$ when $n=0,1,2$ that involve arrangements of type $A$ or $D$. These again should enable us to determine their rational cohomology ring, but it seems that this hasn't been done yet. In the same spirit arrangements of various types (among them $E_6$ and $E_7$) were used in \cite{looijenga:mthree} to prove that $$ H^{{\bullet}}({\mathcal M}_3 )=\CH^{{\bullet}}({\mathcal M}_3)+{\mathbb Q} u, $$ where $u$ is a class of degree $6$ of Hodge bidegree $(6,6)$ and $$ H^{{\bullet}}({\mathcal M}^1_3)=\CH^{{\bullet}}({\mathcal M}^1_3)+{\mathbb Q} u +{\mathbb Q} u\tau + {\mathbb Q} u\kappa _1+{\mathbb Q} v, $$ where $v$ is a class of degree $7$ and of Hodge bidegree $(6,6)$. \begin{question} The image of the tautological algebra in $H^{2p}({\mathcal M}_g^n)$ consists of classes of type $(p,p)$. Are all such classes of this form? \end{question} A version of the Hodge conjecture asserts that the rational classes in degree $2p$ of type $(p,p)$ are in the image of the Chow algebra, so modulo this conjecture we are asking whether every Chow class on ${\mathcal M}_g^n$ is homologically equivalent to a tautological class. \section{The Ribbon Graph Picture} \label{sec:ribbon} Around 1981 Thurston, Mumford and Harer observed that partial completions of the Teichm\"uller spaces ${\mathcal X} _g^n$ with $n>0$ possess two natural $\Gamma _g^n$ equivariant triangulations. One is based on the hyperbolic geometry of $S_g^n$ (Thurston) and the other based on the singular euclidean geometry of $S_g^n$ (Mumford, Harer). The last approach was actually a direct, but very powerful application of work that Jenkins and Strebel had done 10--20 years earlier. It is this approach that we shall explain. The basic notion is that of a {\it ribbon graph}. This is a finite graph\footnote{For us a graph is a cell complex of pure dimension one; its zero cells are called {\it vertices} and its one cells {\it edges}. So it has no isolated vertices.} $G$ together with a cyclic order on the set of oriented edges\footnote{An oriented edge of a graph is an edge together with an orientation of it.} emanating from each vertex. As we shall see, there is a canonical construction of a surface that contains $G$ and of which $G$ is a deformation retract. This construction should explain the name. We first give a somewhat more abstract characterization of ribbon graphs which is very useful in some applications. Let $X(G)$ be the set of oriented edges of $G$. Let $\sigma _1$ be the involution of $X(G)$ that reverses the orientation of each edge. The set $X_1(G)$ of $\sigma _1$ orbits can be identified with the set of edges of $G$. The cyclic orderings define another permutation $\sigma _0$ of $X(G)$ as follows. Each oriented edge $e$ has an initial vertex $\initial (e)$ and a terminal vertex $\term (e)$. Define $\sigma _0(e)$ to be the successor of $e$ with respect to the given cyclic order on the set of oriented edges that have $\initial (e)$ as their initial vertex. The set of orbits $X_0(G)$ of $\sigma _0$ can be identified with the set of vertices of $G$. Put $\sigma _{\infty}:= (\sigma _1\sigma _0)^{-1}=\sigma _0^{-1}\sigma _1$. Call an orbit of this permutation a {\it boundary cycle}. (Draw a picture to see why.) The set of boundary cycles wil be denoted $X_{\infty}(G)$. These data form a complete invariant of $G$, for we can reverse the construction and associate to a finite nonempty set $X$ endowed with a fixed point free involution $\sigma _1$ and a permutation $\sigma _0$ of $X$, a ribbon graph $G(X,\sigma _0,\sigma _1)$ whose oriented edges are indexed by $X$ and such that $\sigma_0$ and $\sigma_1$ are the permutations defined above. For every oriented edge $e$ of $G$ we form the one point compactification $\Delta _e$ of the half strip $e\times [0,\infty )$; this is just a $2$-simplex, parameterized in an unusual way. We make identifications along the boundaries of these simplices with the help of $\sigma _0$ and $\sigma _{\infty}$: $e\times \{0\}$ is identified with $\sigma _1e\times \{0\}$ and and $\{\term (e)\}\times [0,\infty )$ with $\{\initial (\sigma _{\infty}e)\}\times [0,\infty )$ (in either case, the identification map is essentially the identity). This is easily seen to be a compact, triangulated surface $S(G)$ that contains $G$ as a subcomplex. Its vertex set can be identified with the disjoint union of the vertex set of $G$ (so $X_0(G)$) and $X_{\infty}(G)$. We call vertices of the latter type {\it cusps}. Notice that $G$ is a deformation retract of $S(G)-X_{\infty}(G)$ and that the surface is canonically oriented if we insist that the cyclic orderings of the edges emanating from each vertex are induced by the orientation. Let us say that the ribbon graph $G$ is {\it $n$-pointed} if we are given a injection $y:\{ 1,\dots ,n\}\hookrightarrow X_{\infty}(G)\cup X_0(G)$ whose image contains $X_{\infty}(G)$ and the vertices of valency $\le 2$. Suppose that we are given a {\it metric} $l$ on $G$. That is, a function that assigns to every (unoriented) edge of $X$ a positive real number. Give $[0,\infty )$ the standard metric and every half strip $e\times [0,\infty )$ the product metric. This defines (at least locally) a metric on $S(G)$. This metric is euclidean except possibly at the vertices. However, it is not difficult to show that the underlying conformal structure extends across all the vertices of $S(G)$ so that we end up with a compact Riemann surface $C(G,l)$. Notice that each cusp has a ``circumference'' --- this is the length of the associated boundary cycle. It is clear that we get the same complex structure if $l$ is replaced by a positive multiple of it and so we may just as well assume that the total length of $G$ is $1$. With this convention, the sum of the circumferences of the cusps is $2$. The work of Jenkins and Strebel shows that all compact Riemann surfaces arise in this way: \begin{theorem}[Strebel \cite{strebel}] Let $(C;x:\{ 1,\dots,n\}\hookrightarrow C)$ be an $n$-pointed connected Riemann surface (so that the complement of the image of $x$ has negative Euler characteristic as usual) and let $c_1,\dots,c_n$ be nonnegative real numbers, not all zero. Then there exists an $n$-pointed metrized ribbon graph $(G,y,l)$, with $y(i)$ a cusp of $G$ of circumference $c_i$ when $c_i>0$ and a vertex of $G$ otherwise, such that $(C(G,l),y)$ and $(C,x)$ are isomorphic as $n$-pointed Riemann surfaces. Moreover, $(G,y,l)$ is unique up to the obvious notion of isomorphism. \end{theorem} The results of Strebel also include a continuity property: a continuous variation of the complex structure on $C$ corresponds to a continuous variation of $(G,y,l)$ in a sense that we make precise. Denote by ${\mathcal R}{\mathcal G} _g^n$ the set of isomorphism classes of $n$-pointed ribbon graphs $(G,y)$ that are marked in the sense that we are given an isotopy class of homeomorphisms $h: S_g\to S(G)$ with $h(x_i)=y(i)$, $i=1,\dots ,n$. On this set $\Gamma _g^n$ acts, and it is easy to see that the number of orbits of markings is finite. Suppose that $(G,y,[h])$ represents an element of ${\mathcal R}{\mathcal G} _g^n$. Denote the geometric realization of the abstract simplex on the set $X_1(G)$ by $\Delta (G)$. Notice that the metrics $l$ on $G$ that give $G$ unit length are parameterized by the interior of $\Delta (G)$. The circumferences of the cusps add up to two, so half the cicumferences are the barycentric coordinates of a simplicial projection $\lambda :\Delta (G)\to\Delta ^{n-1}$. Let $s$ be an edge of $G$ that is not a loop and does not connect two vertices in the image of $y$. Then collapsing that edge yields a member $(G/s ,y/s, [h]/s)$ of ${\mathcal R}{\mathcal G} _g^n$. We can regard $\Delta (G/s)$ as a face of $\Delta (G)$. Making these identifications produces a simplicial complex which we will denote by $\widehat{{\mathbb X}} _g^n$. It comes with a simplicial map $\lambda:\widehat{{\mathbb X}} _g^n\to \Delta ^{n-1}$. We have a simplicial action of $\Gamma _g^n$ on $\widehat{{\mathbb X}} _g^n$ which preserves the fibers of $\lambda$. The union of relative interiors of simplices of $\widehat{{\mathbb X}} _g^n$ indexed by the elements of ${\mathcal R}{\mathcal G} _g^n$ is an open subset ${\mathbb X} _g^n$ of $\widehat{{\mathbb X}} _g^n$. The results of Strebel can be strengthened to: \begin{proposition}[cf.\ \cite{looijenga:cell}] The above construction defines a $\Gamma _g^n$ equivariant homeomorphism of ${\mathbb X} _g^n$ onto ${\mathcal X} _g^n\times\Delta ^{n-1}$. \end{proposition} Now consider the quotient space $$ \widehat{{\mathbb M}} _g^n:=\Gamma _g^n\backslash\widehat{{\mathbb X}} _g^n. $$ This is a finite simplicial orbicomplex that is equipped with a simplicial map $\lambda :\widehat{{\mathbb M}} _g^n\to\Delta ^{n-1}$. We regard this complex as a compactification of its open subset ${\mathbb M} _g^n:= \Gamma _g^n\backslash{\mathbb X} _g^n$. According to the above theorem, the latter is canonically homeomorphic with ${\mathcal M} _g^n\times\Delta ^{n-1}$. This raises the question of how this compactification compares to that of Deligne-Mumford. The answer is essentially due to Kontsevich: \begin{theorem}[Kontsevich \cite{kontsevich:airy}, see also \cite{looijenga:cell}] The simplicial orbicomplex $\widehat{{\mathbb M}} _g^n$ is a quotient space of $\overline{\M} _g^n\times\Delta ^{n-1}$. Moreover, the part of $\widehat{{\mathbb M}} _g^n$ where $\lambda _i>0$ carries an oriented piecewise linear circle bundle (in the orbifold sense) whose pull-back to $\overline{\M} _g^n\times \{\lambda\in\Delta ^{n-1} | \lambda _i>0\}$ is the oriented circle bundle coming from the standard line bundle $\overline{\tau}_i$. In particular, the part of $\widehat{{\mathbb M}} _g^n$ lying over the interior of $\Delta ^{n-1}$ carries the tautological cohomology classes underlying $\overline{\tau} _{n,i}$, $i=1,\dots ,n$. \end{theorem} The defining equivalence relation on $\overline{\M} _g^n\times\Delta ^{n-1}$ is a little subtle and we refer to \cite{looijenga:cell} for details regarding both statement and proof.\footnote{The space used by Kontsevich is not quite $\widehat{{\mathbb M}} _g^n$, but basically the part lying over the interior of $\Delta ^{n-1}$ times a half line. In this case the circumference map $\lambda$ has image $(0,\infty)^n$.} This compactification of ${\mathcal M} _g^n\times\Delta ^{n-1}$ plays a crucial r\^ole in Kontsevich's proof of the Witten conjectures. There are however earlier applications. These include Harer's stability theorem we met before, the computation of the Euler characteristic of ${\mathcal M} _g^n$, and the proof that $\Gamma _g^n$ is a virtual duality group of dimension $4g-4+n$. We shall not explain the relation with stability here, but we will briefly touch on the other applications. There is also a remarkable arithmetic aspect of ribbon graphs that is presently under intense investigation, but which we merely mention in passing. This is the observation, made by Grothendieck in a research proposal \cite{groth:esq}, that for a metrized ribbon graph $(G,l)$ all of whose edges have equal length, the corresponding Riemann surface $C(G,l)$ is, in a canonical way, a ramified covering of the Riemann sphere ${\mathbb P} ^1$ with ramification locus contained in $\{ 0,1,\infty\}$. The graph $G$ appears here as the preimage of the interval $[0,1]$, its vertex set as the preimage of $0$ and the set of cusps as the preimage of $\infty$. The preimage of $1$ consists of the midpoints of the edges. At these points we have simple ramification. A covering of this type is naturally an algebraic curve defined over some number field. Conversely, every connected covering of the Riemann sphere of this type arises in this manner. The absolute Galois group of ${\mathbb Q}$ acts on the collection of isomorphism types of such coverings, and thus also on each finite set ${\mathcal R}{\mathcal G} _g^n$. It is very difficult to come to grips with this action. For more information we refer to the collection \cite{groth:dessins} and to Grothendieck's manuscripts \cite{groth:marche} and \cite{groth:esq}. Since these metrized ribbon graphs represent the barycenters of the simplices of ${\mathbb M} _g^n$, one can also think of this as an action of the absolute Galois group on the simplices of ${\mathbb M} _g^n$, but the significance of this is not clear to us. \subsection{Virtual duality and virtual Euler characteristic} \label{subsec:virtual} We first make some observations about simplicial complexes. Let $K$ be a simplicial complex, $L$ a subcomplex. Set $U:=K-L$. Then $U$ admits a canonical deformation retraction onto the union of the closed simplices of the barycentric subdivision of $K$ that lie in $U$. This is a subcomplex, called the {\it spine} of $U$, whose $k$-simplices correspond to strictly increasing chains $\sigma _0\subset\sigma _1\subset\cdots\subset\sigma _k$ of simplices of $K$ not in $L$. Further, if $\Gamma$ is a group of automorphisms of $K$ that preserves $L$, and if \begin{enumerate} \item[(i)] $U$ is contractible, \item[(ii)] a subgroup of $\Gamma$ of finite index acts freely on $U$, \end{enumerate} then $\Gamma\backslash U$ is a simplicial `orbicomplex' that is also a virtual classifying space for $\Gamma$.\footnote{This means there is a normal subgroup $\Gamma_1\subset\Gamma$ of finite index such that a $\Gamma /\Gamma_1$-cover of this space classifies $\Gamma_1$.} It has $\Gamma\backslash \spine (U)$ as deformation retract, and so the dimension of this spine is an upper bound for the virtual homological dimension of $\Gamma$. We apply this in the situation where $K$ is the preimage of the first vertex of $\Delta ^{n-1}$ in $\widehat{{\mathbb X}} _g^n$ under $\lambda$ and $U=K\cap {\mathbb X} _g^n$. Notice that $U\cong {\mathcal X} _g^n$. The simplices meeting $U$ are indexed by the elements of ${\mathcal R}{\mathcal G} _g^n$ with a single boundary cycle. A simple calculation shows that when $g\ge 1$, the number of edges of such a graph is at most $6g-5+2n$ and at least $2g-1+n$. For $g=0$ these numbers are $2n-5$, resp.\ $n-2$. So the spine of $U$ has dimension $\le 4g-4+n$, resp. $n-3$. One can verify that this is, in fact, an equality. It follows that $U$ admits a subcomplex of this dimension as an equivariant deformation retract. Hence: \begin{theorem}[Harer \cite{harer:virt}] If $n\ge 1$, then for every level structure, the moduli space ${\mathcal M} _g^n[G]$ contains a subcomplex of dimension $4g-4+n$ (when $g>0$) or $n-3$ (when $g=0$) as a deformation retract. \end{theorem} From this he deduces a similar result for the case when $n=0$: ${\mathcal M} _g[G]$ has the homotopy type of complex of dimension $4g-5$. \begin{problem} Is there a Lefschetz type of proof of this fact? For instance, the Lefschetz property would follow if one can find an orbifold stratification of ${\mathcal M}_g^n$ with all strata affine subvarieties of codimension $\le g$ ($n\ge 1$) or $\le g-1$ ($n=0$). That would also show that the cohomological dimension of ${\mathcal M}_g^n$ for quasicoherent sheaves is $\le g-1$ ($n\ge 1$) or $\le g-2$ ($n=0$). \end{problem} Let us return to the general situation considered earlier and suppose, in addition, that \begin{enumerate} \item[(iii)] $\Gamma\backslash K$ is a finite complex, and \item[(iv)] $U$ is a simplicial manifold of dimension $d$, say. \end{enumerate} These conditions are satisfied in the case at hand. It is then natural (and standard) to assign to each simplex of $K$ the weight that is the reciprocal of the order of its $\Gamma$ stabilizer. This weighting is constant on orbits. Wall's Euler characteristic of $\Gamma$ is simply the usual alternating sum of the number of $\Gamma$ orbits of simplices not in $L$, except that each is counted with its weight. Equivalently, it is the orbifold Euler characteristic of the quotient $\Gamma\backslash K$. In the present case, a ribbon graph $G$ defining a member of ${\mathcal R}{\mathcal G} _g^n$ gives a contribution $|\Aut (G)|^{-1}(-1)^{|X_1(G)|}$ to the virtual Euler characteristic. The computation of the resulting sum is a combinatorial problem that was first solved by Harer and Zagier. Kontsevich \cite{kontsevich:airy} later gave a shorter proof. The answer is: \begin{theorem}[Harer-Zagier \cite{harer_zagier}] The orbifold Euler characteristic of ${\mathcal M}_g^n$ equals $$ (-1)^{n-1}\,\frac{(2g+n-3)!}{(2g-2)!}\zeta (1-2g). $$ Here $\zeta$ denotes the Riemann zeta function. \end{theorem} Harer and Zagier also find formulae for the actual Euler characteristics of ${\mathcal M} _g^1$ and ${\mathcal M} _g$. These are often negative so that there must be lot of cohomology in odd degrees. For the discussion of virtual duality we go back to the general situation and assume that beyond the four conditions already imposed we have: \begin{enumerate} \item[(v)] $L$ has the homotopy type of a bouquet of $r$-spheres. \end{enumerate} Then the theory of Bieri-Eckmann can be invoked in a virtual setting: if we set $D:=\tilde H_r(L;{\mathbb Z} )$ and regard $D$ as a $\Gamma$ module in an obvious way, then $H_{d-r-1}(\Gamma ;D)$ is of rank one and for any $\Gamma$-module $V$ with rational coefficients the cap products $$ \cap :H^k(\Gamma ;V )\otimes H_{d-r-1}(\Gamma ;D )\to H_{d-r-1-k}(\Gamma ;V\otimes D ),\quad k=0,1,2,\dots $$ are isomorphisms. One calls $D$ the {\it Steinberg module} of $\Gamma$. Harer \cite{harer:virt} proves that in the present case hypothesis (v) is satisfied: $L$ is a subcomplex of dimension $2g-3-n$, resp.\ $n-4$ which is $(2g-4-n)$-connected, resp.\ $(n-5)$-connected when $g>0$, resp.\ $g=0$. We shall call the corresponding orbifold local system ${\mathbb D}$ over ${\mathcal M}_g^n$ the {\it Steinberg sheaf}. The homology group $H_{4g-4+n}({\mathcal M}_g^n ;{\mathbb D} )$ is of rank one. For every orbifold local system ${\mathbb V}$ of rational vector spaces on ${\mathcal M} _g^n$, cap product with a generator of this homology group defines isomorphisms $$ H^k({\mathcal M}_g^n ;{\mathbb V} )\stackrel{\sim}{\to} H_{4g-4+n-k}({\mathcal M} _g^n;{\mathbb V}\otimes{\mathbb D} ),\quad k=0,1,2,\dots $$ when $g>0$ and $n>0$ (and similar isomorphisms in the remaining cases). In particular, taking ${\mathbb V}$ to be ${\mathbb Q}$, we see that $H_{\bullet}({\mathcal M}_g^n;{\mathbb D})$ has a canonical MHS. This suggests that ${\mathbb D}$ has some Hodge theoretic significance. Unfortunately it is not of finite rank, yet we wonder: \begin{question} Is the Steinberg sheaf motivic? In particular, does it have natural completions that carry (compatible) Hodge and \'etale structures? \end{question} \subsection{Intersection numbers on the Deligne-Mumford completion} \label{subsec:intersection} The intersection numbers in question are those defined by monomials in the $\overline{\tau} _i$'s. To be precise, define for every such monomial $\overline{\tau} _1^{d_1}\dots \overline{\tau} _n^{d_n}$ (with all $d_i \ge 0$) the intersection number $\int _{\overline{\M}_g^n} \overline{\tau} _1^{d_1}\dots \overline{\tau} _n^{d_n}$ where $g$ is chosen in such a way that this has a possibility of being nonzero: $3g-3+n=d_1+\cdots +d_n$. A physics interpretation suggests that we should combine these numbers into the generating function \begin{equation*} \sum _{n=1}^{\infty} \frac{1}{n!}\sum _{g>1-{\frac{1}{2}} n}\, \sum_{d_1 + \cdots +d_n = 3g-3+n} t_{d_1}\dots t_{d_n} \int _{\overline{\M}_g^n} \overline{\tau} _1^{d_1}\cdots \overline{\tau} _n^{d_n}. \end{equation*} Now pass to a new set of variables $T_1,T_3,T_5,\dots$ by setting $$ t_i=1.3.5.\cdots(2i+1)T_{2i+1}. $$ The resulting expansion $F(T_1,T_3,T_5,\dots )$ encodes all these intersection numbers. Witten \cite{witten} conjectured two other characterizations of this function, both of which allow computation of its coefficients. These were proved by Kontsevich in his celebrated paper \cite{kontsevich:airy}. Perhaps the most useful characterization is the one which says that $F$ is killed by a Lie algebra of differential operators isomorphic to the Lie algebra of polynomial vector fields in one variable. This Lie algebra comes with a basis $(L_k)_{k\ge -1}$ corresponding to the vector fields $(z^k\partial /\partial z)_{k\ge -1}$ and Witten verified the identities $L_k(F)=0$ for $k=-1,0$ within the realm of algebraic geometry. However no such proof is known for $k\ge 1$. Kontsevich's strategy is to represent the classes $\overline{\tau} _1^{d_1}\cdots \overline{\tau} _n^{d_n}$ by piecewise differential forms on the ribbon graph model that can actually be integrated. This allows him to convert the intersection numbers into weighted sums over ribbon graphs. This leads to a new characterization of the generating function that is more manageable. Still a great deal of ingenuity is needed to complete the proof of Witten's Conjecture. \section{Torelli Groups and Moduli} \label{sec:torelli} In the early 80s, Dennis Johnson published a series of pioneering papers \cite{johnson:fg,johnson:ker,johnson:h1} on the Torelli groups. Although this work is in geometric topology, it has several interesting applications to algebraic geometry. Here we review some of his work. First a remark on notation. In the remainder of the paper we will write $V_g$ for the symplectic vector space $H_1(S_g)$ and $Sp_g({\mathbb Z})$ for the group $\Aut (H_1(S_g,{\mathbb Z}),\langle \phantom{x},\phantom{x} \rangle)$; this does not really clash with standard notation, since a choice of a symplectic basis of $H_1(S_g;{\mathbb Z})$ identifies this with the standard integral symplectic group of genus $g$. Likewise, $Sp_g$ will stand for the algebraic ${\mathbb Q}$-group defined by the symplectic transformations of $V_g$; so its group of ${\mathbb Q}$-points, $Sp_g({\mathbb Q} )$, is just the group of symplectic automorphisms of $V_g$. The mapping class group $\Gamma_{g,r}^n$ acts on the homology of the reference surface $S_g$. Since each of its elements preserves the orientation of $S_g$, we have a homomorphism \begin{equation}\label{homom} \Gamma_{g,r}^n \to Sp_g({\mathbb Z}). \end{equation} which is surjective. The {\it Torelli group} $T_{g,r}^n$ is defined to be its kernel\footnote{Note that there is no general agreement on the definition of $T_{g,r}^n$ when $r + n > 1$.} so that we have an extension $$ 1 \to T_{g,r}^n \to \Gamma_{g,r}^n \to Sp_g({\mathbb Z}) \to 1. $$ The homology groups of $T_{g,r}^n$ are therefore $Sp_g({\mathbb Z})$ modules. The simplest kind of element of $T_{g,r}^n$ is a Dehn twist along a simple loop in $S_g^{n+r}$ that separates $S$ into two connected components. We call such a loop a {\it separating simple loop}. Another type of element of $T_{g,r}^n$ is determined by a {\it separating pair of simple loops}. This is a pair of two disjoint nonisotopic loops $\alpha _1,\alpha _2$ on $S_g^{n+r}$ that together separate $S$ into two connected components. The Dehn twist along $\alpha _1$ composed with the inverse of the Dehn twist along $\alpha _2$ is in $T_{g,r}^n$. The first of Johnson's results is: \begin{theorem}[Johnson \cite{johnson:fg,johnson:ker,johnson:h1}] \label{johnson} When $g\ge 3$, $T_{g,r}^n$ is generated by elements associated to a finite number of separating simple loops and a finite number of separating pairs of simple loops. If $[S_g]\in \wedge ^2 H_1(S_g ;{\mathbb Z})$ corresponds to the fundamental class of $S_g$, then there are natural $Sp_g({\mathbb Z})$ equivariant surjective homomorphisms $$ \tau_g^1 : T_g^1 \to \wedge ^3 H_1(S_g;{\mathbb Z} ) \text{ and } \tau_g : T_g \to \wedge^3 H_1(S_g;{\mathbb Z} )/([S_g]\wedge H_1(S_g;{\mathbb Z} )). $$ In both cases, the kernel of $\tau$ is the subgroup generated by the elements associated to simple separating loops. Finally, the kernels of the induced homomorphisms $$ H_1(T_g^1;{\mathbb Z} ) \to\wedge ^3 H_1(S_g ;{\mathbb Z})\text{ and } H_1(T_g^1;{\mathbb Z} ) \to\wedge ^3 H_1(S_g ;{\mathbb Z} )/([S_g]\wedge H_1(S_g ;{\mathbb Z} )) $$ are both 2-torsion. \end{theorem} Johnson also finds an explicit description of this 2-torsion. We will give it in a moment, but first we want to point out an algebro-geometric consequence of this theorem. Let $\widetilde{{\mathcal M}}_g\subset\overline{\M} _g$ be the complement of the irreducible divisor whose generic point parametrizes irreducible singular stable curves, and let $\widetilde{{\mathcal M}} _g^1$ be its preimage in $\overline{\M} _g^1$. \begin{corollary}\label{cor:pioftildem} When $g\ge 3$, the orbifold fundamental group of $\widetilde{{\mathcal M}}_g$ (resp.\ $\widetilde{{\mathcal M}}_g^1$) is isomorphic to an extension of $Sp_g({\mathbb Z} )$ by $\wedge ^3H_1(S_g;{\mathbb Z})/([S_g]\wedge H_1(S_g;{\mathbb Z} ))$ (resp.\ $\wedge ^3H_1(S_g;{\mathbb Z} )$). \end{corollary} Johnson's theorem shows that the $Sp_g({\mathbb Z})$ action on $H_1(T_{g,r}^n)$ is the restriction of a representation of the algebraic group $Sp_g$. We shall see shortly the importance of this property. Let $\lambda_1, \lambda_2,\dots ,\lambda_g$ be a fundamental set of weights of $Sp_g$ so that $\lambda_j$ corresponds to the $j$th fundamental representation of $Sp_g$. This last representation can be realized as the natural $Sp_g$ action on the primitive part of $\wedge^j V_g$. The next result follows from Johnson's Theorem by standard arguments. \begin{corollary} For each $g\ge 3$, there is a natural $Sp_g({\mathbb Z})$ equivariant isomorphism $$ \tau_{g,r}^n : H^1(T_{g,r}^n) \stackrel{\sim}{\to} V(\lambda_3) \oplus V(\lambda_1)^{\oplus(r+n)}. $$ \end{corollary} A theorem of Ragunathan \cite{ragunathan} implies that when $g\ge 2$, the first cohomology of each finite index subgroup of $Sp_g({\mathbb Z})$ with coefficients in a rational representation of $Sp_g({\mathbb Q})$ vanishes. So Johnson's computation also gives: \begin{corollary}\label{van_h1} If $g\ge 3$, then every finite index subgroup of $\Gamma_{g,r}^n$ that contains $T_{g,r}^n$ has zero first Betti number. \end{corollary} The situation is very different when $g < 3$. The Torelli groups $T_1$ and $T_1^1$ are trivial, while Geoff Mess \cite{mess} proved that when $g=2$, $T_2$ is a countably generated free group. He also computed $H_1(T_2;{\mathbb Z})$. It is the $Sp_2({\mathbb Z})$ module obtained by inducing the trivial representation up to $Sp_2({\mathbb Z})$ from the stabilizer $({\mathbb Z}/2)\ltimes (SL_2({\mathbb Z})\times SL_2({\mathbb Z}))$ of a decomposition of $H_1(S_2;{\mathbb Z})$ into two symplectic modules each of rank 2. (We shall sketch a proof in the next subsection.) It is still unknown whether, for any $g\ge 3$, $T_g$ is finitely presented. \begin{problem} Determine whether $T_g$ is finitely presented when $g$ is sufficiently large. \end{problem} Next, we describe Johnson's computation of the torsion in $H_1(T_g;{\mathbb Z})$. Denote the field of two elements by ${\mathbb F}_2$. Recall that an ${\mathbb F}_2$ quadratic form on $H_1(S_g;{\mathbb F}_2)$ associated to the mod two symplectic form $\langle\phantom{x} ,\phantom{x}\rangle$ on $H_1(S_g;{\mathbb F}_2)$ is a function $\omega : H_1(S_g;{\mathbb F}_2) \to {\mathbb F}_2$ satisfying $$ \omega(a+b) = \omega(a) + \omega(b) + \langle a,b\rangle . $$ The difference between any two such is an element of $H^1(S_g;{\mathbb F}_2)$. This makes the set $\Omega_g$ of such quadratic forms an affine space over the ${\mathbb F} _2$ vector space $H^1(S_g;{\mathbb F}_2)$. Denote the algebra of ${\mathbb F} _2$ valued functions on $\Omega_g$ by $S\,\Omega_g$. All such functions are given by polynomials and so we have a filtration $$ {\mathbb F}_2 = S_0\Omega_g \subset S_1\Omega_g \subset S_2\Omega_g \subset \dots \subset S\,\Omega_g, $$ where $S_d\Omega_g$ denotes the space of polynomial functions of degree $\le d$. Since $f=f^2$ for each $f\in S\,\Omega_g$, the associated graded algebra is naturally isomorphic to the exterior algebra $\wedge^{\bullet} H_1(S_g;{\mathbb F}_2)$. The algebra $S\,\Omega_g$ has as a distinguished element which is called the {\it Arf invariant}, denoted here by ${\rm arf}$. If $a_1,\dots, a_g, b_1,\dots,b_g$ is a symplectic basis of $H^1(S;{\mathbb F}_2)$, then ${\rm arf}$ is defined by $$ {\rm arf} : \omega \mapsto \sum_i \omega(a_i)\omega(b_i). $$ It is an element of $S_2\Omega_g$, and its zero set $\Psi _g$ is an affine quadric in $\Omega_g$. Let $S_d\Psi _g$ denote the image of $S_d\Omega_g$ in the set of ${\mathbb F}_2$ valued functions on $\Psi_g$. \begin{theorem}[Johnson \cite{johnson:h1}] There are natural isomorphisms \begin{gather*} \sigma_{g,1} : H_1(T_{g,1};{\mathbb F}_2) \stackrel{\sim}{\to} S_3 \Omega_g,\quad \sigma_g^1 : H_1(T_g^1;{\mathbb F}_2)\stackrel{\sim}{\to} S_3\Omega_g/{\mathbb F}_2\,{\rm arf} ,\\ \sigma_g : H_1(T_g;{\mathbb F}_2)\stackrel{\sim}{\to} S_3\Psi_g \end{gather*} which are equivariant with respect to the $Sp_g({\mathbb F}_2)$-action. These induce natural isomorphisms $$ H_1(T_{g,1};{\mathbb Z})_{\rm tor}\cong S_2\Omega_g,\quad H_1(T_g^1;{\mathbb Z})_{\rm tor} \cong H_1(T_g;{\mathbb Z})_{\rm tor} \cong S_2\Psi_g. $$ Moreover, the natural isomorphisms $$ \phi_{g,r}^n : H_1(T_{g,r}^n;{\mathbb F}_2)/H_1(T_{g,r}^n;{\mathbb Z})_{\rm tor} \stackrel{\sim}{\to} \left[H_1(T_{g,r}^n;{\mathbb Z})/\text{\rm torsion}\right]\otimes{\mathbb F}_2 $$ correspond, under the isomorphisms $\sigma_{g,r}^n$ and $\tau_{g,r}^n$, to the obvious isomorphisms \begin{gather*} \phi_{g,1} : S_3\Omega_g/S_2\Omega_g\stackrel{\sim}{\to} \wedge^3H_1(S_g;{\mathbb F}_2),\quad \phi_g^1 : S_3\Omega_g/({\mathbb F}_2\,{\rm arf} + S_2\Omega_g) \stackrel{\sim}{\to} \wedge^3H_1(S_g;{\mathbb F}_2),\\ \phi_g : S_3\Psi_g/S_2\Psi_g \stackrel{\sim}{\to} \wedge^3H_1(S_g;{\mathbb F}_2)/([S_g]\wedge H_1(S_g;{\mathbb F}_2)). \end{gather*} \end{theorem} The homomorphisms $\tau _g^1$ and $\tau _g$ admit direct conceptual definitions that we will give later. Here we give a formula for the image of the standard generators of $T_{g,1}$ in $\wedge^3H_1(S_g;{\mathbb Z})$ and in $S_3\Omega _g$ under $\tau_{g,1}$ and $\sigma_{g,1}$, respectively. Let $(\alpha _1,\alpha _2)$ be a separating pair of simple loops. Let $t$ be the corresponding element of $T_{g,1}$ --- recall that this is the product of the Dehn twist about $\alpha_1$ and the {\em inverse} of the Dehn twist about $\alpha_2$. The two loops decompose $S_g$ into two pieces $S'$ and $S''$, say, where we suppose that $S'$ contains the point $x_1$. We orient $\alpha _1$ and $\alpha _2$ as boundary components of $S''$. The resulting cycles are opposite in $H_1(S'';{\mathbb Z})$: $[\alpha_2]=-[\alpha_1]$, and each spans the radical of the intersection pairing on this group. So there is a well-defined element in $\wedge ^2H_1(S'';{\mathbb Z})/[\alpha_1]\wedge H_1(S'';{\mathbb Z})$ representing the intersection pairing on $H_1(S'';{\mathbb Z})$. Its wedge with $[\alpha_1]$ can be regarded as an element of $\wedge ^3H_1(S'';{\mathbb Z})$. Since the inclusion $S''\subset S_g$ induces an injection on first homology, we can also view the latter as an element of $\wedge ^3H_1(S_g;{\mathbb Z})$. This is the element $\tau _{g,1}(t)$; it is clear that it only depends on the image of $t$ in $T_g^1$. Next we associate to $t$ a function $\sigma_t:\Omega_g\to {\mathbb F}_2$ as follows. If $\omega\in\Omega_g$ takes the value 1 on $[\alpha]$, then we put $\sigma_t(\omega )=0$; if it takes the value 0 on $[\alpha]$, then the restriction of $\omega$ to $H_1(S'';{\mathbb F}_2)$ factors through a nondegenerate quadratic function on $H_1(S'';{\mathbb F}_2)/{\mathbb F}_2[\alpha]$. Then $\sigma_{g,1}(t)(\omega)$ is its Arf invariant. It can be shown that $\sigma_{g,1}(t)$ lies in $S_3\Omega$. Now suppose that $t$ is the element of $T_{g,1}$ associated to a separating simple loop $\alpha$. Denote the pieces $S'$ and $S''$ as before. In this case, $\tau_{g,1}(t)$ is trivial and $\sigma_{g,1}(t)$ is the element of $S\Omega_g$ that assigns to $\omega$ the Arf invariant of its restriction to $H_1(S'';{\mathbb F}_2)$. Notice that if $\alpha$ is a simple loop around $x_1$, then $\sigma_{g,1}(t)$ is just the function ${\rm arf}$. (This explains why we mod out by this function when passing from $T_{g,1}$ to $T_g^1$.) Without a base point there is no way of telling $S'$ and $S''$ apart. It is because of this ambiguity that we have to restrict functions to $\Psi_g$ in order to obtain a well defined function. A diffeomorphism of $S_g$ onto a smooth projective curve $C$ determines a natural isomorphism between $\Omega_g$ and the space of theta characteristics of $C$ (i.e., square roots of the canonical bundle $K_C$; see for instance Appendix B of \cite{acgh}). This suggests that Johnson's computation should have an algebro-geometric interpretation, if not interesting applications to the geometry of curves. \begin{problem} Give an algebro-geometric construction of the epimorphism $T_g\to S_3\Omega$. \end{problem} Van Geemen has suggested such a construction (unpublished). \subsection{Torelli space and period space}\label{subsec:torelli} The group $T_g$ acts freely on ${\mathcal X}_g$. The quotient ${\mathcal T}_g$ is therefore a complex manifold. It is called {\it Torelli space}. According to the discussion at the beginning of Section~\ref{sec:moduli}, ${\mathcal T}_g$ is then a classifying space for $T_g$ so that there is a canonical isomorphism $H_{\bullet}(T_g;{\mathbb Z} ) \cong H_{\bullet}({\mathcal T}_g;{\mathbb Z} )$. Torelli space has a moduli interpretation; it is the moduli space of smooth projective curves $C$ of genus $g$ together with a symplectic isomorphism $$ {\mathbf \gamma} : H_1(S_g;{\mathbb Z}) \to H_1(C;{\mathbb Z}). $$ There are also decorated versions ${\mathcal T}_{g,r}^n$ of Torelli space. Their points are points of ${\mathcal M}_{g,r}^n$ together with a symplectic isomorphism ${\mathbf \gamma}$ of $H_1(S;{\mathbb Z})$ with the first homology of the curve corresponding to the point of ${\mathcal M}_g$. It is clear that the map ${\mathcal T}_{g,r}^n \to {\mathcal M}_{g,r}^n$ is Galois with Galois group $Sp_g({\mathbb Z})$. Denote the Siegel space associated to $V_g$ by ${\mathfrak h}_g$. To be precise, ${\mathfrak h}_g$ is the set of pure Hodge structures on $V_g$ with Hodge numbers $(-1,0)$ and $(0,-1)$, polarized by the intersection form. This is a contractible complex manifold of dimension $g(g+1)/2$ on which the group $Sp_g({\mathbb R})$ acts properly and transitively. We can also regard ${\mathfrak h}_g$ as the moduli space of pairs consisting of a $g$ dimensional principally polarized abelian variety $A$ plus a symplectic isomorphism $$ {\mathbf \gamma} : H_1(S;{\mathbb Z}) \to H_1(A;{\mathbb Z}). $$ This interprets the $Sp_g({\mathbb Z})$ orbit space of ${\mathfrak h}_g$ as the moduli space of principally polarized abelian varieties of dimension $g$, ${\mathcal A} _g$. We regard ${\mathcal A} _g$ as an orbifold with orbifold fundamental group $Sp_g({\mathbb Z})$, although $Sp_g({\mathbb Z})$ does not act faithfully on ${\mathfrak h}_g$. The kernel of this action is $\{\pm 1\}$. Assigning to a smooth projective curve the Hodge structure on its first homology group defines a map ${\mathcal T}_g \to {\mathfrak h}_g$, the {\it period map} for $T_g$. It is an isomorphism in genus 1, an open imbedding when $g=2$, and 2:1 with ramification along the hyperelliptic locus when $g\ge 3$.% \footnote{It is stated incorrectly in \cite{hain:normal} that ${\mathcal T}_2 \to {\mathfrak h}_2$ is an unramified 2:1 map onto its image.} The reason for this is that for all abelian varieties we have the equality $$ [A;{\mathbf \gamma}] = [A;-{\mathbf \gamma}] $$ of points of ${\mathfrak h}_g$ as $-\id$ is an automorphism of each abelian variety. On the other hand, we have the equality $$ [C;{\mathbf \gamma}] = [C;-{\mathbf \gamma}] $$ of points of ${\mathcal T}_g$ if and only if $C$ is hyperelliptic. Mess's result (mentioned at the beginning of the section) can now be deduced from this: ${\mathcal T}_2$ is the complement in ${\mathfrak h}_2$ of the locus of principally polarized abelian varieties that are products of two elliptic curves. The locus of such reducible abelian varieties is a countable disjoint union of copies of ${\mathfrak h}_1 \times {\mathfrak h}_1$. The group $Sp_2({\mathbb Z})$ permutes them transitively, and each is stabilized by a product of two copies of $SL_2({\mathbb Z})$ and an involution that switches the two copies of the upper half plane. Mess's result follows easily using the stratified Morse theory of Goresky and MacPherson --- use distance from a generic point of ${\mathfrak h}_2$ as the Morse function. Since each component of ${\mathfrak h}_2 -{\mathcal T}_2$ is a totally geodesic divisor, the distance function has a unique critical point (necessarily a minimum) on each stratum. It follows that ${\mathcal T}_2$ has the homotopy type of a wedge of circles, one for each component of ${\mathfrak h}_2 -{\mathcal T}_2$. The period map gives, after passage to $Sp _g({\mathbb Z} )$ orbit spaces, a morphism ${\mathcal M}_g\to{\mathcal A}_g$, the period mapping for ${\mathcal M} _g$. This period mapping extends to the partial completion $\widetilde{{\mathcal M}}_g$ of ${\mathcal M}_g$ and the resulting map $\widetilde{{\mathcal M}}_g\to {\mathcal A}_g$ is proper. Now assume $g\ge 3$ and denote the image of the period map ${\mathcal T}_g \to {\mathfrak h}_g$ by ${\mathcal S}_g$. This space is the quotient of ${\mathcal T} _g$ by the subgroup $\{\pm 1\}$ of $Sp_g({\mathbb Z})$. Consequently $$ H^{\bullet}({\mathcal S}_g) \cong H^{\bullet}(T_g)^{\{\pm 1\}}. $$ Observe that ${\mathcal S}_g$ is a locally closed analytic subvariety of ${\mathfrak h}_g$, but not closed. The $\{\pm 1\}$ cover ${\mathcal T}_g\to{\mathcal S}_g$ extends as a $\{\pm 1\}$ cover $\overline{{\mathcal T}}_g\to\overline{{\mathcal S}}_g$ over the closure of ${\mathcal S}_g$ in ${\mathfrak h} _g$, and the $\{\pm 1\}$ action on the total space is the restriction of an $Sp_g({\mathbb Z})$ action. Both $\overline{{\mathcal T}}_g$ and $\overline{{\mathcal S}}_g$ are rather singular along the added locus (which is of codimension $3$). If we pass to $Sp_g({\mathbb Z})$ orbit spaces, then the natural map $$ \widetilde{{\mathcal M}}_g\to Sp_g({\mathbb Z})\backslash \overline{{\mathcal T}}_g\cong Sp_g({\mathbb Z})\backslash \overline{{\mathcal S}}_g $$ resolves these singularities in an orbifold sense. A resolution of a normal analytic variety always induces a surjection on fundamental groups and so it follows from (\ref{johnson}) that the fundamental group of $\overline{{\mathcal T}}_g$ is abelian and is $Sp_g({\mathbb Z})$ equivariantly a quotient of $\wedge ^3 H_1(S_g ;{\mathbb Z} )/([S_g]\wedge H_1(S_g ;{\mathbb Z} ))$. \begin{problem} Understand the topology of ${\mathcal S}_g$ and its closure $\overline{{\mathcal S}}_g$ in ${\mathfrak h}_g$. In particular, how close is ${\mathcal S}_g$ to being a finite complex? (Observe that if it has a finite 2-skeleton, then $T_g$ is finitely presented.) \end{problem} Related, but formally independent of this problem, is the question of whether the cohomology of $T_g$ stabilizes in a suitable sense: \begin{question} Is $H^k(T_g)$ expressible as an $Sp_g({\mathbb Z} )$ module in a manner that is independent of $g$ if $g$ is large enough? For example, from Johnson's Theorem, we know that $H^1(T_g)$ is the third fundamental representation of $Sp_g$ for all $g\ge 3$. \end{question} \subsection{The Johnson homomorphism}\label{subsec:johnson} The proof of Johnson's Theorem is non-trivial and uses geometric topology, but the homomorphism $\tau_g^1$ is easily described. Since $T_g$ is torsion free, the projection ${\mathcal T} _g^1\to {\mathcal T}_g$ defines the universal curve over ${\mathcal T}_g$. Denote the corresponding bundle of jacobians by ${\mathcal J}_g \to {\mathcal T}_g$. Since the local system of first homology groups associated to the universal curve is canonically framed, this jacobian bundle ${\mathcal J}_g \to {\mathcal T}_g$ is analytically trivial as a bundle of Lie groups: we have a natural trivializing projection $p:{\mathcal J} _g \to \Jac S_g$, where $\Jac S_g:= H_1(S_g; {\mathbb R} /{\mathbb Z} )$ is the ``jacobian'' of the reference surface. The usual Abel-Jacobi map, which assigns to an ordered pair of points $(x,y)$ on a smooth curve $C$ the divisor class of $(x)-(y)$, induces a morphism $$ {\mathcal T}_g^1\times _{{\mathcal T} _g} {\mathcal T}_g^1\to {\mathcal J} _g. $$ over ${\mathcal T} _g$. This provides a correspondence $$ \begin{CD} {\mathcal T}_g^1\times _{{\mathcal T} _g} {\mathcal T}_g^1 @>>> {\mathcal J} _g @>p>> \Jac S_g \cr @VV{pr_2}V \cr {\mathcal T}_g^1 \cr \end{CD} $$ from ${\mathcal T}_g^1$ to $\Jac S_g $. It induces homomorphisms $$ H_k(T_g^1) \cong H_k({\mathcal T}_g^1) \to H_{k+2}(\Jac S_g ). $$ The first of these is the Johnson homomorphism $$ \tau_g^1 : H_1(T_g^1) \to H_3(\Jac S_g ) $$ for $T_g^1$. Since ${\mathcal T} _g^1\to {\mathcal T}_g$ is a fibration of Eilenberg-MacLane spaces, we have an exact sequence of fundamental groups: $$ 1 \to \pi_g \to T_g^1 \to T_g \to 1. $$ This induces an exact sequence $$ H_1(S_g;{\mathbb Z} ) \to H_1(T_g^1;{\mathbb Z} ) \to H_1(T_g;{\mathbb Z} ) \to 0 $$ on homology. Since $\Jac S_g $ is a topological group with torsion free homology, its integral homology has a product --- the {\it Pontrjagin product}. It is not difficult to check that the composite $$ H_1(S_g;{\mathbb Z} ) \to H_1(T_g^1;{\mathbb Z} ) \to H_3(\Jac S_g;{\mathbb Z} ) $$ is the map given by Pontrjagin product with the class $[S_g]$. It follows that there is a natural homomorphism $$ H_1(T_g;{\mathbb Z} ) \to H_3(\Jac S_g;{\mathbb Z} )/\left([S_g ]\times H_1(S_g;{\mathbb Z} )\right). $$ This is the Johnson homomorphism $\tau_g$ for $T_g$. Johnson's Theorem, alone and in concert with Saito's theory of Hodge modules, has several interesting applications to the geometry of moduli spaces of curves as we shall see in subsequent sections. \subsection{Monodromy of roots of the canonical bundle} In this subsection we assume that $g\ge 2$. Suppose that $C$ is a smooth projective curve of genus $g$. Since its canonical bundle $K_C$ is of degree $2g-2$ and since $\Pic^0 C$ is a divisible group, $K_C$ has $n$th roots whenever $n$ divides $2g-2$. Any two such $n$th roots will differ by an $n$ torsion point of $\Pic^0 C$. Because of this, $n$th roots of $K_C$ are rigid under deformation. It follows that they form a locally constant sheaf (in the orbifold sense) $\Rt^n$ over ${\mathcal M}_g$. The fiber over $C$, denoted $\Rt^n C$, is a principal homogenous space over $H_1(C;{\mathbb Z}/n)$, the group of $n$ torsion points of $\Pic^0 C$. Choose a conformal structure on $S_g$. Denote the corresponding algebraic curve by $C$. Sipe \cite{sipe} determined the monodromy representation $$ \rho^n : \Gamma_g \to \Aut \Rt^n(C). $$ of this sheaf. Before giving it, we make some remarks. Since the Torelli group acts trivially on the $n$ torsion of $\Pic^0 C$, it follows that the restriction of $\rho^n$ to $T_g$ factors through a representation $T_g \to H_1(S;{\mathbb Z}/(2g-2)) \to H_1(S;{\mathbb Z}/n)$. However, the action of $\Gamma_g$ on the set $\Rt^2 C$ of square roots of $K_C$ (the set of theta characteristics of $C$) factorizes through $Sp_g({\mathbb Z})$ also (even through $Sp_g({\mathbb F}_2)$) --- this is because there is a canonical correspondence between square roots of $K_C$ and ${\mathbb F}_2$ quadratic forms on $H_1(S;{\mathbb F}_2)$ associated to the intersection form. It follows that the image of the monodromy representation $\rho^n$ will be contained in an extension of $Sp_g({\mathbb Z}/n)$ by a subgroup of $2\cdot H^1(S_g;{\mathbb Z}/n)$. In fact, it is all of this group. \begin{theorem}[Sipe \cite{sipe}]\label{sipe} The monodromy group of $\Rt^n$ is an extension of $Sp_g({\mathbb Z}/n)$ by the subgroup $2\cdot H^1(S_g;{\mathbb Z}/n)$ of $H_1(S_g;{\mathbb Z}/n)$. \end{theorem} The subgroup $2\cdot H^1(S_g;{\mathbb Z}/n)$ appears as a quotient of the Torelli group $T_g$. In \cite{hain:normal} it is shown that the restriction of the monodromy representation to $T_g$ is the composite of the Johnson homomorphism with a natural surjection $$ \wedge^3H_1(S_g;{\mathbb Z})/([S_g]\wedge H_1(S_g;{\mathbb Z})\to H_1(S_g;{\mathbb Z}/(g-1)) \to 2\cdot H_1(S_g;{\mathbb Z}/n). $$ \subsection{Picard groups of level covers}\label{subsec:picard} Denote the moduli space of smooth projective genus $g$ curves with a level $l$ structure by ${\mathcal M}_g[l]$. This is convenient shorthand for the notation ${\mathcal M}_g[Sp_g({\mathbb Z}/l)]$ introduced in Section~\ref{sec:moduli}. Denote the kernel of the reduction mod $l$ map $$ Sp_g({\mathbb Z}) \to Sp_g({\mathbb Z}/l) $$ by $Sp_g({\mathbb Z})[l]$, and its full inverse image in $\Gamma_g$ by $\Gamma_g[l]$. Then ${\mathcal M}_g[l]$ is the quotient of Teichm\"uller space ${\mathcal X}_g$ by $\Gamma_g[l]$. As in the case of ${\mathcal M}_g$, there is a canonical isomorphism $$ H^{\bullet}({\mathcal M}_g[l]) \cong H^{\bullet}(\Gamma_g[l]). $$ This holds with rational coefficients for all $l$, and arbitrary coefficients whenever $\Gamma_g[l]$ is torsion free, which holds whenever $Sp_g({\mathbb Z})[l]$ is torsion free --- $l \ge 3$. We know from (\ref{van_h1}) that $$ H^1(\Gamma_g[l]) \cong H^1({\mathcal M}_g[l]) = 0 $$ when $g\ge 3$. By standard arguments (cf.\ \cite[\S 5]{hain:normal}), this implies that $$ c_1 : \Pic {\mathcal M}_g[l]\otimes{\mathbb Q} \to H^2({\mathcal M}_g[l]) $$ is injective, and therefore that $\Pic {\mathcal M}_g[l]$ is finitely generated when $g\ge 3$. The stable cohomology of an arithmetic group depends only on the ambient real algebraic group \cite{borel:triv}. Based on this, one might expect that the natural map $$ H^k(\Gamma_g) \to H^k(\Gamma_g[l]) $$ is an isomorphism for all $l\ge 0$, once the genus $g$ is sufficiently large compared to the degree $k$. It follows from Johnson's work that this is true when $k=1$ (cf.\ \cite{hain:normal}), but the only evidence for it when $k>1$ is Harer's computation of the second homology of the spin mapping class groups \cite{harer:spin}, and Foisy's theorem from which Harer's computation now follows: \begin{theorem}[Foisy \cite{foisy}] For all $g\ge 3$, the natural map $H^2(\Gamma_g) \to H^2(\Gamma_g[2])$ is an isomorphism. Consequently, $\Pic {\mathcal M}_g[2]$ is finitely generated of rank 1. \end{theorem} \begin{question} Is $\Pic {\mathcal M}_g[l]$ rank 1 for all $g\ge 3$ and all $l\ge 1$? \end{question} This would be the case if we knew that the $Sp_g({\mathbb Z})$ action on $H^2(T_g)$ extended to an algebraic action of $Sp_g$, for we could then invoke Borel's computation of the stable cohomology of arithmetic groups \cite{borel:triv}. \subsection{Normal Functions} \label{subsec:normal} Each rational representation $V$ of $Sp_g$ gives rise to an orbifold local system ${\mathbb V}$ over ${\mathcal M}_g[l]$. Such a local system underlies an admissible variation of Hodge structure. First, if $V$ is irreducible, then $V$ underlies a variation of Hodge structure unique up to Tate twist (\cite[(9.1)]{hain:normal}). Every polarized ${\mathbb Q}$ variation of Hodge structure whose monodromy representation comes from a rational representation of $Sp_g$ has the property that each of its isotypical components is an admissible variation of Hodge structure of the form $A_\lambda\otimes {\mathbb V}(\lambda)$, where $A_\lambda$ is a Hodge structure and ${\mathbb V}(\lambda)$ is a variation of Hodge structure corresponding to the $Sp_g$ module with highest weight $\lambda$ --- cf.\ \cite[(9.2)]{hain:normal}. For a Hodge structure $V$ of weight $-1$ one defines the corresponding {\it intermediate jacobian} $JV$ by $$ JV = V_{\mathbb C}/(F^0 V + V_{\mathbb Z}). $$ Its interest comes from the fact that it parametrizes the extensions of ${\mathbb Z}$ by $V$ in the MHS category: if $E$ is an extension of the ${\mathbb Z}$ (with its trivial Hodge structure of weight zero) by $V$, then choose an integral lift $e\in E$ of $1$ and consider the image of $e$ in $$ E_{\mathbb C}/(F^0 E + V_{\mathbb Z})\cong V_{\mathbb C}/(F^0 V + V_{\mathbb Z}). $$ This is independent of the lift and yields a complete invariant of the extension. There is an inverse construction that makes $JV$ support a variation of mixed Hodge structure $\E$ that is universal as an extension of the trivial Hodge structure ${\mathbb Z}$ by the constant Hodge structure $V$: $$ 0 \to {\mathbb V} _{JV}\to \E \to {\mathbb Z} _{JV}\to 0 $$ (see \cite{carlson}). This immediately generalizes to a relative setting: if ${\mathbb V}$ is an admissible variation of ${\mathbb Z}$ Hodge structure of weight $-1$ over a smooth variety $X$, then we have a corresponding bundle $\pi :{\mathcal J}{\mathbb V}\to X$ of intermediate jacobians over $X$ supporting a universal extension $$ 0 \to \pi ^*{\mathbb V} \to \E \to \pi ^*{\mathbb Z} _X\to 0. $$ A section $\sigma$ of ${\mathcal J}{\mathbb V}$ over $X$ determines an extension of Hodge structures: $$ 0 \to {\mathbb V} \to \sigma ^*\E \to {\mathbb Z} _X\to 0. $$ A {\it normal function} is a section of ${\mathcal J}{\mathbb V}$ such that the corresponding extension $\E$ is an admissible variation of mixed Hodge structure. The normal functions arising from algebraic cycles are normal functions in this sense --- cf.\ \cite[\S6]{hain:normal}. We briefly recall Griffiths' construction of a normal function associated to a family of homologically trivial algebraic cycles. First we consider the case where the base is a point. Suppose that $X$ is a smooth projective variety. A homologically trivial algebraic $d$-cycle $Z$ in $X$ canonically determines an extension of ${\mathbb Z}$ by $H_{2d+1}(X;{\mathbb Z}(-d))$ by pulling back the exact sequence $$ 0 \to H_{2d+1}(X;{\mathbb Z}(-d)) \to H_{2d+1}(X,|Z|;{\mathbb Z}(-d)) \to H_{2d}(|Z|;{\mathbb Z}(-d)) \to \cdots $$ of MHSs along the inclusion $$ {\mathbb Z} \to H_{2d}(|Z|;{\mathbb Z}(-d)) $$ that takes 1 to the class of $Z$. So an integral lift of $1$ is given by an integral singular $2d+1$ chain $W$ in $X$ whose boundary is $Z$. Integration identifies $JH_{2d+1}(X;{\mathbb Z} (-d))$ with the {\it Griffiths intermediate Jacobian} $$ J_d(X):=\Hom_{\mathbb C} (F^dH^{2d+1}(X);{\mathbb C} (-d))/H^{2d+1}(X;{\mathbb Z} (-d)), $$ and under this isomorphism the extension class in question is just given by integration over $W$. Families of homologically trivial cycles give rise to normal functions: Suppose that ${\mathcal X} \to T$ is a family of smooth projective varieties over a smooth base $T$ and that ${\mathcal Z}$ is an algebraic cycle in ${\mathcal X}$ which is proper over $T$ of relative dimension $d$. Then the local system whose fiber over $t\in T$ is $H_{2d+1}(X_t;{\mathbb Z}(-d))$ naturally underlies a variation of Hodge structure ${\mathbb V}$ over $T$ of weight $-1$ so that we can form the $d$th {\it relative intermediate jacobian} ${\mathcal J}_d ({\mathcal X} /T)\to T$, whose fiber over $t\in T$ is $J_d(X_t)$. The family of cycles ${\mathcal Z}$ defines a section of this bundle which is a normal function. \begin{theorem}[Hain \cite{hain:normal}]\label{norm_classn} Suppose that ${\mathbb V}$ is an admissible variation of Hodge structure of weight $-1$ over ${\mathcal M}_g[l]$ whose monodromy representation factors through a rational representation of $Sp_g$. If $g \ge 3$, then the space of normal functions associated to ${\mathbb V}$ is finitely generated of rank equal to the number of copies of the variation ${\mathbb V}(\lambda_3)$ of weight $-1$ that occur in ${\mathbb V}$. \end{theorem} The theorem implies that, up to torsion and multiples, there is only one normal function over ${\mathcal M}_g$ associated to a variation of Hodge structure whose monodromy factors through a rational representation of $Sp_g$. So what is the generator of these normal functions? To answer this question, recall that if $C$ is a smooth projective curve of genus $g$ and $x\in C$, we have the Abel-Jacobi morphism $$ C \to \Jac C, \quad y\mapsto (y)-(x). $$ Denote the image 1-cycle in $\Jac C$ by $C_x$ and the cycle $i_\ast C_x$ by $C_x^-$, where $i : \Jac C \to \Jac C$ takes $u$ to $-u$. The cycle $C_x - C_x^-$ is homologous to zero, and therefore defines a point $\nu ^1(C,x)$ in $J_1(\Jac C)$. Pontrjagin product with the class of $C$ induces a homomorphism $$ A:\Jac C \to J_1(\Jac C). $$ We call the cokernel of $A$ the {\it primitive first intermediate Jacobian} $J_1^{{\rm pr}}(\Jac C)$ of $\Jac C$. The family of such primitive intermediate jacobians over ${\mathcal M}_g$ is the unique one (up to isogeny) associated to the variation of Hodge structure of weight $-1$ over ${\mathcal M}_g$ whose associated $\Gamma_g$ module is $V(\lambda_3)$. It is not difficult to show that $$ \nu ^1(C,x) - \nu ^1(C,y) = 2A(x-y). $$ It follows that the image of $\nu ^1(C,x)$ in $J_1^{{\rm pr}}(\Jac C)$ is independent of $x$. This is the value of the normal function associated with $C-C^-$ over $[C]$. We can do better and realize half of this generator by a generalized normal function as follows. Let $A$ be a principally polarized abelian variety of dimension $g\ge 3$. The polarization determines a distinguished element $\omega$ of $H_2(A;{\mathbb Z})$. If $Z$ and $Z'$ are two piecewise smooth cycles representing $\omega$, then their difference is the boundary of a piecewise smooth $3$-chain $W$ on $A$. Represent the dual of $H_3(A;{\mathbb R})$ by translation invariant $3$-forms on $A$. Then integrating these forms over $A$ determines an element of $H_3(A;{\mathbb R} )$. Another choice of $W$ gives a class that differs from this one by an element of $H_3(A;{\mathbb Z} )$, and so we have a well-defined element $[Z-Z']$ of $H_3(A;{\mathbb R}/{\mathbb Z} )$. Notice that the latter torus is naturally identified with the first intermediate jacobian $J_1(A)$ of $A$. We declare $Z$ and $Z'$ to be equivalent if $[Z-Z']=0$ and denote the space of piecewise smooth cycles representing $\omega$ modulo this equivalence relation by $D(A)$. This is clearly a torsor of $J_1(A)$ and so it has a natural complex structure. In view of its connection with Deligne cohomology, we call it the {\it Deligne torsor} of $A$. This torsor contains naturally a subtorsor $D(A)[2]$ of the $2$-torsion in $J_1(A)$, $J_1(A)[2]\cong H_3(A;\frac{1}{2}{\mathbb Z}/{\mathbb Z} )$: Let $a=(a_1,a_{-1},\dots ,a_g,a_{-g})$ be a symplectic basis of $H_1(A;{\mathbb Z} )$. Each basis element $a_i$ is uniquely represented by a homomorphism $\alpha_i:S^1\to A$ and so $\omega$ is represented by the $2$-cycle $\sum_{i=1}^g \alpha_i\times\alpha_{-i}$. This cycle defines an element $z(a)\in D(A)$. It is easily verified that $z(a)$ only depends on the mod two reduction of $a$ and that if $a$ runs over all symplectic bases, $z(a)$ runs over an entire orbit $D(A)[2]$ of $J_1(A)[2]$. (So $J_1(A)[2]\backslash D(A)$ has a canonical point which identifies it with $J_1(A)$.) The group $Sp(H_1(A;{\mathbb Z}))$ acts on $D(A)$ as an affine transformation group in a way that is easily made explicit. The lifts of these transformations to a universal covering of $D(A)$ form a group of affine symplectic transformations. It is an extension of $Sp(H_1(A;{\mathbb Z}))$ by $H_3(A;{\mathbb Z} )$ which splits if we enlarge the extension to $H_3(A;\frac{1}{2}{\mathbb Z} )$. The Pontrjagin product with $\omega$ defines a homomorphism $A\to J_1(A)$ which gives rise to corresponding primitive notions: the {\it primitive Deligne torsor} $D^{{\rm pr}}(A):=A\backslash D(A)$ is a torsor of the primitive intermediate Jacobian $J^{\rm pr} _1(A):=A\backslash J_1(A)$. We have corresponding universal Deligne torsors over ${\mathcal A}_g$ which we denote ${\mathcal D} _g\to{\mathcal A}_g$ and ${\mathcal D} _g^{{\rm pr}}\to{\mathcal A}_g$. By the above argument, these torsors become trivial on the Galois cover of ${\mathcal A}_g$ representing principally polarized abelian varieties with a level $2$ structure. The torsors themselves are nontrivial, for it can be shown that the orbifold fundamental groups of these torsors are nonsplit extensions of the integral symplectic group of genus $g$ For $C$ a nonsingular projective curve of genus $g\ge 3$ and $x\in C$, the Abel-Jacobi morphism $C \to \Jac C$ defined by $y\mapsto (y)-(x)$ defines a cycle in the homology class of the natural polarization of $\Jac C$ and so we get an element $[(C,x)]$ of $D(\Jac C)$. Its image in $D^{{\rm pr}}(\Jac C)$ is independent of $x$ and so can be denoted by $[C]$. Universally this produces holomorphic lifts of the period map: $$ \nu_g^1: {\mathcal M}_g^1\to {\mathcal D}_g\text{ and } \nu_g: {\mathcal M}_g\to {\mathcal D}_g^{{\rm pr}}. $$ We call $\nu_g$ the {\it fundamental normal function} on ${\mathcal M}_g$. \subsection{Picard group of the generic curve with a level $l$ structure} \label{subsec:genericpicard} The classification of normal functions (\ref{norm_classn}) implies that there are no sections of $\Pic^0$ of infinite order defined over ${\mathcal M}_g[l]$ when $g\ge 3$. This, combined with Sipe's computation (\ref{sipe}) of the monodromy of roots of the canonical bundle allows one to determine the Picard group of the generic point of ${\mathcal M}_g[l]$. The case $l=1$ was the subject of the Franchetta Conjecture which was deduced from Harer's computation of $\Gamma_g$ by Beauville (unpublished) and by Arbarello and Cornalba \cite{arb_cor:pic}. \begin{theorem}[Hain \cite{hain:derham}] The Picard group of the generic curve of genus $g\ge 3$ with a level $l$ structure is of rank 1, has torsion subgroup isomorphic to the $l$ torsion points $H_1(\Jac S_g;{\mathbb Z}/l)$, and, modulo torsion, is generated by the canonical bundle if $l$ is odd, and a theta characteristic if $l$ is even. \end{theorem} \section{Relative Malcev Completion} \label{sec:malcev} Fundamental groups of smooth algebraic varieties are quite special as we know from the work of Morgan \cite{morgan} and others. The least trivial restrictions on these groups come from Hodge theory and Galois theory. Since $\Gamma_g$ is the (orbifold) fundamental group of ${\mathcal M}_g$, a smooth orbifold, Hodge theory and Galois theory should have something interesting to say about its structure. To put a MHS on a group one needs to linearize it. One way to do this is to replace the group by some kind of algebraic envelope and put a MHS on the coordinate ring of this (pro)algebraic group. In this section we introduce these linearizations and use them to establish a relation between the fundamental normal function and a remarkable central extension that is hidden in a quotient of the mapping class group. Here the impact of mixed Hodge theory is not yet felt, but we are setting the stage for Section~\ref{sec:hodgemap} where it is omnipresent. \subsection{Classical Malcev completion} \label{subsec:malcev} Suppose that $\pi$ is a finitely generated group. The classical Malcev (or unipotent) completion of $\pi$ consists of a prounipotent group ${\mathcal U}(\pi)$ (over ${\mathbb Q}$) and a homomorphism $\pi \to {\mathcal U}(\pi)$. It is characterized by the following universal mapping property: if $U$ is a unipotent group, and $\phi : \pi \to U$ is a homomorphism, there is a unique homomorphism of prounipotent groups ${\mathcal U}(\pi) \to U$ through which $\phi$ factors. There are several well known constructions of the unipotent completion, which can be found in \cite{hain:comp}, for example. Each (pro)unipotent group $U$ is isomorphic to its Lie algebra ${\mathfrak u}$, a (pro)nilpotent Lie algebra via the exponential map. Thus, to give the Malcev group ${\mathcal U}(\pi)$ associated to $\pi$ it suffices to give its associated pronilpotent Lie algebra ${\mathfrak u}(\pi)$. This Lie algebra is called the {\it Malcev Lie algebra associated to $\pi$}. It comes with a natural descending filtration whose $k$th term ${\mathfrak u} ^{(k)}(\pi)$ is the closed ideal of ${\mathfrak u} (\pi )$ generated by its $k$-fold commutators $(k=1,2,\dots )$ and it is complete with respect to this filtration. We will refer to this filtration as the {\it Malcev filtration}. When $\pi$ is the fundamental group $\pi_1(X,x)$ of a smooth complex algebraic variety, ${\mathfrak u}(\pi)$ has a canonical MHS which was first constructed by Morgan \cite{morgan}. If $X$ is also complete, or more generally, when $H_1(X)$ has a pure Hodge structure of weight $-1$, then the weight filtration is the Malcev filtration: $$ W_{-k}{\mathfrak u} (\pi _1(X,x))={\mathfrak u} ^{(k)}(\pi _1(X,x)). $$ Alternatively, this MHS determines and is determined by a MHS on the coordinate ring ${\mathcal O}({\mathcal U}(\pi))$ of the associated Malcev group. We shall denote the Malcev completion of $\pi_g^n=\pi_1(S_g^n,x_0)$ by ${\mathfrak p}_g^n$. \subsection{Relative Malcev completion} \label{subsec:relmalcev} The Malcev completion of a group $\pi$ is trivial when $H_1(\pi)$ vanishes, for then $\pi$ has no non-trivial unipotent quotients. Since the first homology of $\Gamma_g$ vanishes for all $g$, its Malcev completion will be trivial. Deligne has defined the notion of Malcev completion of a group $\pi$ relative to a Zariski dense homomorphism $\rho:\pi \to S$, where $S$ is a reductive algebraic group defined over a base field $F$ (that we assume to be of characteristic zero). The {\it Malcev completion of $\pi$ relative to $\rho : \pi \to S$} is a a proalgebraic $F$-group ${\mathcal G}(\pi,\rho)$, which is an extension $$ 1 \to {\mathcal U} \to {\mathcal G}(\pi,\rho) \to S \to 1 $$ of $S$ by a prounipotent group, together with a lift $\tilde{\rho} : \pi \to {\mathcal G}(\pi,\rho)$ of $\rho$.\footnote{In many cases the completion of $\pi$ over an algebraic closure $\bar F$ of $F$ is the set of $\bar F$ points of the completion of $\pi$ over $F$. This is the case for the mapping class groups when $g\ge 3$, but we do not know whether this is true in general, except when $S$ is trivial.} It is characterized by the following universal mapping property: if $G$ is an $F$-group which is an extension of $S$ by a unipotent group $U$, and if $\phi : \pi \to G$ is a homomorphism, then there is a unique homomorphism ${\mathcal G}(\pi,\rho) \to G$ through which $\phi$ factors: $$ \phi : \pi \stackrel{\tilde{\rho}}{\to} {\mathcal G}(\pi,\rho) \to G. $$ Since $S$ is reductive, we should think of ${\mathcal U}$ as the prounipotent radical of ${\mathcal G} (\pi ,\rho )$. One can show, for instance, that ${\mathcal U}$ has a Levi supplement so that ${\mathcal G} (\pi ,\rho )$ is a semidirect product of $S$ and ${\mathcal U}$. The Lie algebra ${\mathfrak g} (\pi ,\rho )$ of ${\mathcal G} (\pi ,\rho )$ also comes with a Malcev filtration with respect to which it is complete: ${\mathfrak g} (\pi ,\rho )^{(0)}={\mathfrak g} (\pi ,\rho )$, and for $k\ge 1$, ${\mathfrak g} (\pi ,\rho )^{(k)}$ is the closed ideal generated by $k$-fold commutators in the Lie algebra of ${\mathcal U}$. We will often write ${\mathcal G}(\pi)$ instead of ${\mathcal G}(\pi,\rho)$ when the representation $\rho$ is clear from the context. We shall denote the completion of the (orbifold) fundamental group of a pointed orbifold $(X,x)$ with respect to a Zariski dense reductive representation $\rho:\pi_1(X,x) \to S$ by ${\mathcal G}(X,x;\rho)$, or simply ${\mathcal G}(X,x)$ when $\rho$ is clear from the context. When $S$ is trivial, we recover the classical Malcev completion. The universal property of the Malcev completion of $\ker \rho$ yields a natural homomorphism of proalgebraic $F$-groups ${\mathcal U} (\ker \rho)(F)\to {\mathcal G} (\pi)$. In general, it is neither surjective nor injective as the following two examples show. \begin{example} The fundamental group of the symplectic Lie group $Sp_g({\mathbb R})$ is infinite cyclic and hence so is its universal cover $\widehat{Sp}_g({\mathbb R})\to Sp_g({\mathbb R} )$. This universal cover is not an algebraic group (which follows for instance from the fact that the complexification of $Sp_g({\mathbb R} )$, $Sp_g({\mathbb C} )$, is simply connected). The preimage $\widehat{Sp}_g({\mathbb Z})$ of $Sp_g({\mathbb Z})$ in this covering contains the universal central extension of $Sp_g({\mathbb Z})$ by ${\mathbb Z}$. Now take for $\pi$ this central extension and for $\rho$ its natural homomorphism to $Sp _g({\mathbb C} )$. The corresponding relative Malcev completion is then reduced to $Sp _g({\mathbb C} )$ itself, so that the homomorphism from ${\mathcal U} ({\mathbb Z})({\mathbb C})$ (which is just the abelian group ${\mathbb C}$) to ${\mathcal G} (\widehat{Sp}_g({\mathbb Z} ))$ is trivial. We will see that this example is realized inside a quotient of the mapping class group. \end{example} \begin{example} In this example, $\ker \rho$ is trivial, but ${\mathcal U}(\pi)$ is not. The basic fact we need (see \cite[(10.3)]{hain:derham}) is that there is always a natural $S$ equivariant isomorphism $$ H_1({\mathfrak u}(\pi)) \cong \prod_{\alpha \in \Check{S}} H_1(\pi;V_\alpha)\otimes V_\alpha^\ast, $$ where $V_\alpha$ denotes a representation with highest weight $\alpha$. For $\pi$ we take $\Gamma$, a finite index subgroup of $SL_2({\mathbb Z})$, for $S$ we take $SL_2({\mathbb Q})$, and for $\rho$ we take the natural inclusion. Denote the $n$th power of the fundamental representation of $SL_2$ by $S^nV$. For all such $\Gamma$, there is an infinite number of integers $n\ge 0$ such that $H^1(\Gamma;S^nV)$ is non-trivial.\footnote{This is easily seen when $\Gamma$ is free, for example. In general it is related to the theory of modular forms.} It follows that ${\mathcal U}(\Gamma)$ has an infinite dimensional $H_1$, even though $\ker \rho$ is trivial. \end{example} This example suggests the following problem: \begin{problem} Investigate the relationship between the theory of modular forms associated to a finite index subgroup $\Gamma$ of $SL_2({\mathbb Z})$ and the completion of $\Gamma$ relative to the inclusion $\Gamma \hookrightarrow SL_2({\mathbb Q})$. \end{problem} \subsection{The relative Malcev completion of $\Gamma_g$} \label{malcevofgamma} The natural homomorphism $\rho:\Gamma_{g,r}^n \to Sp_g$ has Zariski dense image. Denote the completion of $\Gamma_{g,r}^n$ relative to $\rho$ by ${\mathcal G}_{g,r}^n$, its prounipotent radical by ${\mathcal U}_{g,r}^n$ and their Lie algebras by ${\mathfrak g}_{g,r}^n$ and ${\mathfrak u}_{g,r}^n$. The following theorem indicates the presence of essentially one copy of the universal central extension of $Sp _g({\mathbb Z})$ in quotients of each mapping class group of genus $g$ when $g\ge 3$. \begin{theorem}[Hain \cite{hain:comp}] When $g\ge 2$, the homomorphism \begin{equation}\label{nat_map} {\mathcal U}(T_{g,r}^n) \to {\mathcal U}_{g,r}^n \end{equation} is surjective. When $g \ge 3$, its kernel is a central subgroup isomorphic to the additive group. \end{theorem} This phenomenon is intimately related to the cycle $C-C^-$ and its normal function as we shall now explain. \subsection{The central extension} \label{subsec:cent_extn} The existence of the central extension has both a group theoretic and a geometric explanation. It is also related to the Casson invariant through the work of Morita \cite{morita:casson,morita:cocycles}. We begin with the group theoretic one. The group analogue of the Malcev filtration for the Torelli group $T_g$ is the most rapidly descending central series of $T_g$ with torsion free quotients: $$ T_g = T_g^{(1)}\supset T_g^{(2)}\supset T_g^{(3)} \supset \cdots $$ Note that $T_g^{(1)}/T_g^{(2)}$ is the maximal torsion free abelian quotient of $T_g$, which is $$ V(\lambda _3)_g{\mathbb Z} := \wedge^3 H_1(S_g;{\mathbb Z})/\left([S_g]\times H_1(S_g;{\mathbb Z} )\right) $$ by Johnson's Theorem (\ref{johnson}). The group $\Gamma_g/T_g^{(3)}$ can be written as an extension \begin{equation} \label{ext1} 1 \to T_g^{(2)}/T_g^{(3)} \to \Gamma_g/T_g^{(3)}\to \Gamma_g/T_g^{(2)} \to 1. \end{equation} It turns out that this sequence contains a multiple of the universal central extension of $Sp_g({\mathbb Z} )$ by ${\mathbb Z}$. Since $V_g(\lambda _3)$ is a rational representation of $Sp_g$, and since the surjection $$ \wedge^2 V_g(\lambda _3)_{\mathbb Z} \to T_g^{(2)}/T_g^{(3)} $$ induced by the commutator is $Sp_g({\mathbb Z})$ equivariant, it follows that $T_g^{(2)}/T_g^{(3)}\otimes {\mathbb Q}$ is also a rational representation of $Sp_g$. Because $V_g(\lambda _3)$ is an irreducible symplectic representation, there is exactly one copy of the trivial representation in $\wedge^2 V_g(\lambda _3)$. This copy of the trivial representation survives in $T_g^{(2)}/T_g^{(3)}\otimes {\mathbb Q}$ \cite{hain:comp} so that there is an $Sp_g({\mathbb Z})$ equivariant projection $T_g^{(2)}/T_g^{(3)}\to {\mathbb Z}$. Pushing the extension (\ref{ext1}) out along this map gives an extension \begin{equation} \label{ext2} 0 \to {\mathbb Z} \to E \to \Gamma_g/T_g^{(2)}\to 1 \end{equation} Note that $E$ is a quotient of $\Gamma_g$. We will manufacture a multiple of the universal central extension of $Sp_g({\mathbb Z})$ from this group that turns out to be the obstruction to the map ${\mathcal U}(T_g) \to {\mathcal U}_g$ being injective. (Full details can be found in \cite{hain:comp}.) The group $\Gamma_g/T_g^{(2)}$ can be written as an extension \begin{equation} \label{ext3} 0 \to V_g(\lambda _3)_{\mathbb Z} \to \Gamma_g/T_g^{(2)}\to Sp_g({\mathbb Z}) \to 1. \end{equation} Morita \cite{morita:conj} showed that this extension is {\it semisplit}, that is, if we replace $V_g(\lambda _3)_{\mathbb Z}$ by $\frac{1}{2}V_g(\lambda _3)_{\mathbb Z}$, it splits. (This can also be seen using the normal function of $C-C^-$.) \begin{theorem}[Morita \cite{morita:cocycles}, Hain \cite{hain:comp}] \label{nosplit} The extension of $Sp_g({\mathbb Z})$ by ${\mathbb Z}$ obtained by pulling back the extension (\ref{ext2}) along a semisplitting of (\ref{ext3}) contains the universal central extension of $Sp_g({\mathbb Z})$. \end{theorem} The geometric picture uses the fundamental normal function $\nu_g$. The lifted period maps $\nu_g$ and $\nu_g^1$ to the Deligne torsors are easily seen to extend over the partial completions $\widetilde{{\mathcal M}}_g$ resp.\ $\widetilde{{\mathcal M}}^1_g$: $$ \tilde\nu_g: \widetilde{{\mathcal M}}_g\to {\mathcal D}_g^{{\rm pr}}, \quad \tilde\nu_g^1: \widetilde{{\mathcal M}}_g^1\to {\mathcal D}_g. $$ The orbifold fundamental group of ${\mathcal D}_g^{{\rm pr}}$ resp.\ ${\mathcal D}_{g,1}$ is an extension of $Sp_g({\mathbb Z} )$ by $V_g(\lambda _3)_{\mathbb Z}$ resp.\ $\wedge ^3V_{g,{\mathbb Z}}$ as both the base and fiber are Eilenberg MacLane spaces with these groups as orbifold fundamental groups. But by \ref{cor:pioftildem} the orbifold fundamental group of $\widetilde{{\mathcal M}}_g$, resp.\ $\widetilde{{\mathcal M}}^1_g$, also has such a structure. Indeed: \begin{theorem}\label{fundgroup} For $g\ge 3$, the normal functions $\tilde\nu_g: \widetilde{{\mathcal M}}_g\to {\mathcal D}_g^{{\rm pr}}$ and $\tilde\nu_g^1: \widetilde{{\mathcal M}}_g^1\to {\mathcal D}_g$ induce an isomorphism on orbifold fundamental groups. (The former can be identified with $\Gamma_g/T_g\!{}^{(2)}$ and the latter with $\Gamma_g^1/(T_g^1)^{(2)}$.) \end{theorem} {}From this theorem we recover the fact that (\ref{ext3}) is semisplit, not split. But we get more, since it should also lead to a description of that extension. The extension (\ref{ext2}) can also be realized geometrically. \begin{proposition}[Hain \cite{hain:comp}] There is a canonical (locally homogeneous) line bundle ${\mathcal B}_g$ over the bundle ${\mathcal D}_{g,1}^{{\rm pr}}\to{\mathcal A}_g$ that realizes the central extension (\ref{ext2}) via the isomorphism of the previous proposition as an extension of orbifold fundamental groups. In particular, both $\tilde\nu^\ast {\mathcal B}_g$ and $\nu^\ast {\mathcal B}_g$ have nonzero rational first Chern class. The bundle $\nu^\ast {\mathcal B}_g$ is canonically metrized and its square is isomorphic (as a metrized line bundle) to the metrized line bundle associated to the archimedean height of the cycle $C-C^-$. \end{proposition} \section{Hodge Theory of the Mapping Class Group}\label{sec:hodgemap} One reason that mixed Hodge theory is so powerful is that the MHS category is abelian. In many situations this turns out to have topological implications for algebraic varieties that are difficult, if not impossible, to obtain directly. A somewhat related (but less exploited) property is that a MHS is canonically split over ${\mathbb C}$. This implies that the weight filtration (which often has a topological interpretation) splits in a way that is compatible with all the algebraic structure naturally present. So, for many purposes, there is no loss of information regarding this algebraic structure if we pass to the corresponding weight graded object. For example, the Malcev filtration on the Malcev Lie algebra of a smooth projective variety is minus the weight filtration, and it therefore splits over ${\mathbb C}$ in a natural way. This splitting is natural in the sense that it respects the Lie algebra structure and is preserved under all base point preserving morphisms. But if we vary the complex structure on $X$ or the base point $x$, then the splitting will, in general, vary with it. A basic example is the Malcev Lie algebra ${\mathfrak p}_g^1$ of $\pi_g^1=\pi _1(S_g^1,x_0)$. The group $\pi _g^1$ is free on $2g$ generators and it is a classical fact that the graded of ${\mathfrak p}_g^1$ with respect to the Malcev filtration is just the free Lie algebra generated by $V_g$. If $S_g$ is given a conformal structure, then $V_g$ has a pure Hodge structure of weight $-1$ and the weight filtration of ${\mathfrak p}_g^1$ is minus the Malcev filtration. The splitting allows us to identify ${\mathfrak p}_g^1\otimes{\mathbb C}$ with the completion of $\Lie (V_g)\otimes{\mathbb C}$. We shall come back to this example in Section~\ref{subsec:relation}. But for now we will focus on the relative Malcev completions introduced in the previous section. \subsection{Hodge theory of ${\mathcal G}_{g,r}^n$} \label{subsec:hodge} A choice of a conformal structure on $S_g$ and nonzero tangent vectors at $x_{n+1},\dots ,x_{n+r}$, determines a point $x_o$ of the moduli space ${\mathcal M}_{g,r}^n$. We can thus identify $\Gamma_{g,r}^n$ with the orbifold fundamental group of $({\mathcal M}_{g,r}^n,x_o)$. This induces an isomorphism of ${\mathcal G}_{g,r}^n$ with ${\mathcal G}({\mathcal M}_{g,r}^n,x_o)$, the completion of $\pi_1({\mathcal M}_{g,r}^n,x_o)$ with respect to the standard symplectic representation. We shall write ${\mathcal G}_{g,r}^n(x_o)$ for ${\mathcal G}({\mathcal M}_{g,r}^n,x_o)$ and denote its prounipotent radical by ${\mathcal U}_{g,r}^n(x_o)$. There is a general Hodge de~Rham theory of relative Malcev completion \cite{hain:derham}. Applying it to $({\mathcal M}_{g,r}^n,x_o)$, one obtains the following result: \begin{theorem}[Hain \cite{hain:torelli}]\label{mhs} For each choice of a base point $x_o$ of ${\mathcal M}_{g,r}^n$, there is a canonical MHS on the coordinate ring ${\mathcal O}({\mathcal G}_{g,r}^n(x_o))$ which is compatible with its Hopf algebra structure. Consequently, the Lie algebra ${\mathfrak g}_{g,r}^n(x_o)$ of ${\mathcal G}_{g,r}^n(x_o)$ and the Lie algebra ${\mathfrak u}_{g,r}^n(x_o)$ of its prounipotent radical both have a natural MHS. \end{theorem} Denote the Malcev Lie algebra of the subgroup of $\pi_1({\mathcal M}_{g,r}^n,x_o)$ corresponding to the Torelli group $T_{g,r}^n$ by ${\mathfrak t}_{g,r}^n(x_o)$. The normal function of $C-C^-$ can be used to lift the MHS from ${\mathfrak u}_{g,r}^n(x_o)$ to ${\mathfrak t}_{g,r}^n(x_o)$. \begin{theorem}[Hain \cite{hain:torelli}]\label{cent_extn} For each $g\ge 3$ and for each choice of a base point $x_o$ of ${\mathcal M}_{g,r}^n$, there is a canonical MHS on ${\mathfrak t}_{g,r}^n(x_o)$ which is compatible with its bracket. Moreover, the canonical central extension $$ 0 \to {\mathbb Q}(1) \to {\mathfrak t}_{g,r}^n(x_o) \to {\mathfrak u}_{g,r}^n(x_o) \to 0 $$ is an extension of MHSs, and the weight filtration equals the Malcev filtration. \end{theorem} \subsection{A presentation of ${\mathfrak t}_g$} \label{subsec:torelli_presentn} We denote the Malcev Lie algebra of $T_{g,r}^n$ by ${\mathfrak t}_{g,r}^n$. The existence of a MHS on ${\mathfrak t}_{g,r}^n(x_o)$ implies that, after tensoring with ${\mathbb C}$, there is a canonical isomorphism $$ {\mathfrak t}_{g,r}^n(x_o)\otimes {\mathbb C} \cong \prod_m \Gr^W_{-m} {\mathfrak t}_{g,r}^n(x_o)\otimes {\mathbb C}. $$ Since the left hand side is (noncanonically) isomorphic to ${\mathfrak t}_{g,r}^n\otimes {\mathbb C}$, to give a presentation of ${\mathfrak t}_{g,r}^n\otimes{\mathbb C}$, it suffices to give a presentation of its associated graded. It follows from Johnson's Theorem (\ref{johnson}) that each graded quotient of the lower central series of ${\mathfrak t}_g$ is a representation of the algebraic group $Sp_g$. We will give a presentation of $\Gr^W_{\bullet} {\mathfrak t}_g$ in the category of representations of $Sp_g$. Recall that $\lambda_1,\dots, \lambda_g$ is a set of fundamental weights of $Sp_g$. For a nonnegative integral linear combination of the fundamental weights $\lambda =\sum _{i=1}^gn_i\lambda _i$ we denote by $V_g(\lambda)$ the representation of $Sp_g$ with highest weight $\lambda$. For all $g \ge 3$, the representation $\wedge^2 V_g(\lambda_3)$ contains a unique copy of $V_g(2\lambda_2) + V_g(0)$. Denote the $Sp_g$ invariant complement of this by $R_g$. Since the quadratic part of the free Lie algebra $\Lie (V_g)$ is $\wedge^2 V_g$, we can view $R_g$ as being a subspace of the quadratic elements of $\Lie (V_g(\lambda_3))$. As mentioned earlier, it is unknown whether any $T_g$ is finitely presented when $g\ge 3$. But the following theorem says that its de~Rham incarnation is: \begin{theorem}[Hain \cite{hain:torelli}]\label{presentation} For all $g \ge 3$, ${\mathfrak t}_g$ is isomorphic to the completion of its associated graded $\Gr^W_{\bullet}{\mathfrak t}_g$. When $g\ge 6$, this has presentation $$ \Gr^W_{\bullet} {\mathfrak t}_g = \Lie (V_g(\lambda_3))/(R_g), $$ where $R_g$ is the set of quadratic relations defined above. When $3 \le g < 6$, the relations in $\Gr^W_{\bullet}{\mathfrak t}_g$ are generated by the quadratic relations $R_g$, and possibly some cubic relations. In particular, ${\mathfrak t}_{g,r}^n$ is finitely presented whenever $g \ge 3$. \end{theorem} Note that this, combined with (\ref{cent_extn}) gives a presentation of $\Gr^W_{\bullet}{\mathfrak u}_g$ when $g\ge 6$: \begin{corollary} For all $g \ge 3$, ${\mathfrak u}_g$ is isomorphic to the completion of its associated graded $\Gr^W_{\bullet}{\mathfrak u}_g$. When $g\ge 6$, this has quadratic presentation $$ \Gr^W_{\bullet} {\mathfrak u}_g = \Lie (V_g(\lambda_3))/(R_g + V_g(0)), $$ where $R_g$ is the set of quadratic relations defined above and where $V_g(0)$ is the unique copy of the trivial representation in $\wedge^2 V_g(\lambda_3)$. When $3 \le g < 6$, the relations in $\Gr^W_{\bullet}{\mathfrak u}_g$ are generated by the quadratic relations $R_g+V_g(0)$, and possibly some cubic relations. In particular, ${\mathfrak u}_{g,r}^n$ is finitely presented whenever $g \ge 3$. \end{corollary} The proof that the relations in the presentation of ${\mathfrak t}_g$ are generated by quadratic relations when $g\ge 6$ and quadratic and cubic ones when $g \ge 3$ is not topological, but uses deep Hodge theory and, surprisingly, intersection homology. The key ingredients are a result of Kabanov, which we state below, and M.~Saito's theory of Hodge modules. We define the {\it Satake compactification} $\overline{\M}^{\rm sat}_g$ of ${\mathcal M}_g$ as the closure of ${\mathcal M}_g$ inside the (Baily-Borel-)Satake compactification of ${\mathcal A}_g$. \begin{theorem}[Kabanov \cite{kabanov}] For each irreducible representation $V$ of $Sp_g$, the natural map $$ IH^2(\overline{\M}^{\rm sat}_g;{\mathbb V}) \to H^2({\mathcal M}_g;{\mathbb V}) $$ is an isomorphism when $g\ge 6$. Here ${\mathbb V}$ denotes the generically defined local system corresponding to $V$. \end{theorem} Such a local system ${\mathbb V}$ is, up to a Tate twist, canonically a variation of Hodge structure. Saito's purity theorem then implies that $H^2({\mathcal M}_g;{\mathbb V})$ is pure of weight $2+$ the weight of ${\mathbb V}$ when $g\ge 6$. It is this purity result that forces $H^2({\mathfrak t}_g)$ to be of weight 2, and implies that no higher order relations are needed. \subsection{Understanding ${\mathfrak t}_g$}\label{subsec:understanding} Even though we have a presentation of ${\mathfrak t}_g$, we still do not have a good understanding of its graded quotients, either as vector spaces or as $Sp_g$ modules. There is an exact sequence $$ 0 \to {\mathfrak p}_g \to {\mathfrak t}_g^1 \to {\mathfrak t}_g \to 0 $$ of Lie algebras (recall that ${\mathfrak p}_g$ stands for the Malcev Lie algebra of $\pi _g=\pi_1(S,x_0)$). It is the de~Rham incarnation of the exact sequence of fundamental groups associated to the universal curve. Fix a conformal structure on $(S,x_0)$. Then this sequence is an exact sequence of MHSs. Since $\Gr^W$ is an exact functor, and since $\Gr^W_{\bullet} {\mathfrak p}_g$ is well understood, it suffices to understand $\Gr^W_{\bullet} {\mathfrak t}_g^1$. There is a natural representation \begin{equation}\label{rep} {\mathfrak t}_g^1 \to \Der {\mathfrak p}_g \end{equation} It is a morphism of MHS, and therefore determined by the graded Lie algebra homomorphism $$ \Gr^W_{\bullet}{\mathfrak t}_g^1 \to \Der \Gr^W_{\bullet} {\mathfrak p}_g. $$ One can ask how close it is to being an isomorphism. Since this map is induced by the natural homomorphism \begin{equation}\label{can_hom} \Gamma_g^1 \to \varprojlim \Aut {\mathbb C}\pi_g/I^m, \end{equation} the homomorphism (\ref{rep}) factors through the projection ${\mathfrak t}_g^1 \to {\mathfrak u}_g^1$, and therefore cannot be injective. On the other hand, we have the following (reformulated) result of Morita: \begin{theorem}[Morita \cite{morita:trace}] There is a natural Lie algebra surjection $$ Tr_M : W_{-1}\Der \Gr^W_{\bullet} {\mathfrak p}_g \to \oplus_{k \ge 1} S^{2k+1}H_1(S) $$ onto an abelian Lie algebra whose composition with (\ref{rep}) is trivial. Here $S^m$ denotes the $m^{\text{th}}$ symmetric power. \end{theorem} One may then hope that the sequence $$ 0 \to {\mathbb C} \to \Gr^W_{\bullet}{\mathfrak t}_g^1 \to W_{-1}\Der \Gr^W_{\bullet} {\mathfrak p}_g \to \oplus_{k \ge 1} S^{2k+1}H_1(S) \to 0 $$ is exact. However, there are further obstructions to exactness at $W_{-1}\Der \Gr^W_{\bullet} {\mathfrak p}_g $ which were discovered by Nakamura \cite{nakamura:obstn}. They come from Galois theory and use the fact that ${\mathcal M}_g^1$ is defined over ${\mathbb Q}$.\footnote{Actually, Nakamura proves his result for a corresponding sequence for ${\mathfrak t}_{g,1}$, but his obstructions most likely appear in this case too.} On the other hand, one can ask: \begin{question} Is the map ${\mathfrak u}_g^1 \to \Der {\mathfrak p}_g$ injective? Equivalently, is ${\mathcal G}_g^1$ the Zariski closure of the image of the representation (\ref{can_hom})? \end{question} A good understanding of ${\mathfrak t}_g$ may help in understanding the stable cohomology of $\Gamma_g$ as we shall explain in the next subsection. \subsection{Torelli Lie algebras and the cohomology of $\Gamma_g$} \label{subsec:torelli&coho} Each Malcev Lie algebra ${\mathfrak g}$ can be viewed as a complete topological Lie algebra. A basis for the neighbourhoods of 0 being the terms ${\mathfrak g}^{(k)}$ of the Malcev filtration. One can define the continuous cohomology of such a ${\mathfrak g}$ to be $$ H^{\bullet}({\mathfrak g}) := \varinjlim H^{\bullet}({\mathfrak g}/{\mathfrak g}^{(k)}). $$ If ${\mathfrak g}$ has a MHS, then so will $H^{\bullet}({\mathfrak g})$. The continuous cohomology of ${\mathfrak t}_{g,r}^n$, ${\mathfrak u}_{g,r}^n$, etc. each has an action of $Sp_g$. The general theory of relative Malcev completion \cite{hain:derham} gives a canonical homomorphism \begin{equation}\label{nat_homom} H^{\bullet}({\mathfrak u}_{g,r}^n)^{Sp_g} \to H^{\bullet}({\mathcal M}_{g,r}^n) \end{equation} One can ask how much of the cohomology of ${\mathcal M}_{g,r}^n$ is captured by this map. Fix a base point $x_o$ of ${\mathcal M}_{g,r}^n$. Then ${\mathfrak t}_{g,r}^n$, etc.\ all have compatible MHSs, and these induce MHSs on their continuous cohomology groups. These groups have the property that the weights on $H^k$ are $\ge k$. \begin{theorem}[Hain \cite{hain:torelli}]\label{morphism} The map~(\ref{nat_homom}) is a morphism of MHS. \end{theorem} Since $H_1(T_g)$ is a quotient of ${\mathfrak u}_g$, there is an induced map \begin{equation}\label{ext_map} H^{\bullet}(H_1(T_g)) \to H^{\bullet}({\mathfrak u}_g). \end{equation} This is also a morphism of MHS. The following result follows directly from \cite[(9.2)]{hain:cycles}, the presentation (\ref{presentation}) of ${\mathfrak t}_g$, and the existence of the MHS on ${\mathfrak u}_g$. \begin{proposition} If $g\ge 3$, the map~(\ref{ext_map}) surjects onto the lowest weight subring $$ \oplus_{k\ge 0} W_kH^k({\mathfrak u}_g) $$ of $H^{\bullet}({\mathfrak u}_g)$, and the kernel is generated by the ideal generated by the unique copy of $V_g(2\lambda_2)$ in $H^2(H_1(T_g))$. \end{proposition} Similar results hold when $r+n>0$ --- cf.\ \cite[\S 14.6]{hain:torelli}. The following result of Kawazumi and Morita tells us that the image of the lowest weight subring of $H^{\bullet}({\mathfrak u}_g)^{Sp_g}$ contains no new cohomology classes. \begin{theorem}[Kawazumi-Morita \cite{kawazumi-morita}] \label{kawazumi-morita} The image of the natural map $$ H^{\bullet}(H_1(T_g))^{Sp_g} \to H^{\bullet}({\mathcal M}_g) $$ is precisely the subring generated by the $\kappa_i$'s. \end{theorem} If we combine this with the previous two results and Pikaart's Purity Theorem (\ref{pikaart_purity}), we obtain the following strengthening of the theorem of Kawazumi and Morita (and obtained independently by Morita, building on our work): \begin{theorem}\label{tautbound} When $k\le g/2$, the image of $H^k({\mathfrak u}_g)^{Sp_g} \to H^k({\mathcal M}_g)$ is the degree $k$ part of the subring generated by the $\kappa_i$'s. \end{theorem} To continue the discussion further, it seems useful to consider cohomology with symplectic coefficients. \subsection{Cohomology with symplectic coefficients} The irreducible representations of $Sp_g$ are parametrized by Young diagrams with $\le g$ rows (and no indexing of the boxes), in other words, by nonincreasing sequences of nonnegative integers whose terms with index $>g$ are zero. So any such sequence $\alpha =(\alpha_1,\alpha_2,\dots)$ defines an irreducible representation of $Sp_h$ for all $h\ge g$. We will denote the representation of $Sp_g$ corresponding to $\alpha$ by $V_{g,\alpha}$, and the corresponding (orbifold) local system over ${\mathcal M}_g$ by ${\mathbb V}_{g,\alpha}$. A theorem of Ivanov \cite{ivanov} (that in fact pertains to more general local systems) implies that, when $r\ge 1$, the group $H^k(\Gamma_{g,r}^n;V_{g,\alpha})$ is independent of $g$ once $g$ is large enough. In the case at hand we have a more explicit result that we state here for the undecorated case (a case that Ivanov actually excludes). \begin{theorem}[Looijenga \cite{looijenga}] Let $\alpha =(\alpha_1,\alpha_2,\dots)$ be a nonincreasing sequence of nonnegative integers that is eventually zero, and let $c_1,c_2,\dots$ be weighted variables with $\deg (c_i)=2i$. Put $|\alpha |:=\sum _{i\ge 1} \alpha_i$. Then there exists a finitely generated, evenly graded ${\mathbb Q} [c_1,\dots ,c_{|\alpha |}]$-module $A^{{\bullet}}_{\alpha}$ (that can be described explicitly) and a graded homomorphism of $H^{\bullet} (\Gamma _{\infty})$ modules $$ A^{\bullet} _{\alpha}[-|\alpha |]\otimes H^{\bullet}(\Gamma _{\infty}) \to H^{\bullet} (\Gamma_g;V_{g,\alpha }) $$ that is an isomorphism in degree $\le cg-|\alpha |$. It is also a MHS morphism if we take $A^{2k}_{\alpha}$ to be pure of type $(k,k)$. In particular, we have $$ A^{\bullet}_{\alpha}[-|\alpha |]\otimes H^{\bullet}(\Gamma _{\infty})\cong H^{\bullet}(\Gamma_\infty;V_\alpha ) $$ both as MHSs and as graded $H^{\bullet}(\Gamma _{\infty})$ modules. So, by (\ref{pikaart_purity}), $H^k(\Gamma_\infty;V_\alpha)$ is pure of weight $k+|\alpha|$. \end{theorem} It is useful to try to understand all cohomology groups with symplectic coefficients at the same time. To do this we take a leaf out of the physicist's book and consider the ``generating function'' \begin{equation}\label{gen_fn} \oplus_\alpha H^{\bullet}(\Gamma_g;V_\alpha^\ast)\otimes V_\alpha \end{equation} where $\alpha$ ranges over all partitions with $\le g$ rows, and $\ast$ denotes dual. This is actually a graded commutative ring as the Peter-Weyl Theorem implies that the coordinate ring ${\mathcal O}_g$ of $Sp_g$ is $$ {\mathcal O}_g = \oplus_\alpha \left(\End V_\alpha\right)^\ast \cong \oplus_\alpha V_\alpha^\ast\otimes V_\alpha. $$ The mapping class group acts on ${\mathcal O}_g$ by composing the right translation action of $Sp_g$ on ${\mathcal O}_g$ with the canonical representation $\Gamma_g \to Sp_g$. The corresponding cohomology group $H^{\bullet}(\Gamma_g;{\mathcal O})$ is then the ``generating function''~(\ref{gen_fn}). Note that ${\mathcal O}_g$ is a variation of Hodge structure of weight 0, so the group $H^k(\Gamma_g;{\mathcal O}_g)$ is stably of weight $k$ by the above theorem. There is a canonical algebra homomorphism $$ H^{\bullet}({\mathfrak u}_g) \to H^{\bullet}(\Gamma_g;{\mathcal O}_g) $$ whose existence follows from the de~Rham theory of relative completion suggested by Deligne --- cf.\ \cite{hain:derham}. The map~(\ref{nat_homom}) of the previous subsection is just its invariant part. This map is a MHS morphism for each choice of complex structure on $S$. The $\alpha$ isotypical part of both sides stabilizes as $g$ increases. It is natural to ask: \begin{question} Is this map stably an isomorphism? \end{question} This has been verified by Hain and Kabanov (unpublished) in degrees $\le 2$ for all weights, and in degree 3 and weight 3. If the answer is yes, or even if one has surjectivity, then it will follow from the theorem of Kawazumi and Morita (\ref{kawazumi-morita}) that the stable cohomology of ${\mathcal M}_g$ is generated by the $\kappa_i$'s. A consequence of injectivity and Pikaart's Purity Theorem would be that for each $k$, $H^k({\mathfrak u}_g)$ is pure of weight $k$ once the genus is sufficiently large. This is equivalent to the answer to the following question being affirmative. \begin{question} Are $H^{\bullet}({\mathfrak u}_g)$ and $U\Gr^W_{\bullet} {\mathfrak u}_g$ stably Koszul dual? \end{question} Note that $U\Gr^W_{\bullet} {\mathfrak u}_g$ and the lowest weight subalgebra of $H^{\bullet}({\mathfrak u}_g)$ have dual quadratic presentations. \section{Algebras Related to the Cohomology of Moduli Spaces of Curves} \label{sec:algebras} The ribbon graph description is the root of a number of ways of constructing (co)homology classes on moduli spaces of curves from certain algebraic structures. These constructions have in common that they actually produce cellular (co)chains on ${\mathbb M} _g^n$, and so they are recipes that assign numbers to `oriented' ribbon graphs. The typical construction, due to Kontsevich \cite{kontsevich:feynman}, goes like this: assume that we are given a complex vector space $V$, a symmetric tensor $p\in V\otimes V$, and linear forms $T_k:V^{\otimes k}\to {\mathbb C}$ that are cyclically invariant. If $\Gamma$ is a ribbon graph, then the decomposition of $X(G)$ into $\sigma _1$ and $\sigma _0$-orbits gives isomorphisms $$ \otimes _{s\in X_1(\Gamma)}V^{\otimes\ori (s)}\cong V^{\otimes X(G)}\cong \otimes _{v\in X_0(\Gamma)}V^{\otimes\out (v)}, $$ where $\ori (s)$ stands for the two-element set of orientations of the edge $s$ and $\out (v)$ for the set of oriented edges that have $v$ as initial vertex. Now $p^{\otimes X_1(\Gamma)}$ defines a vector of the lefthand side and a tensor product of certain $T_k$'s defines a linear form on the righthand side. Evaluation of the linear form on the vector gives a number, which is clearly an invariant of the ribbon graph. Since this invariant does not depend on an orientation on the set of edges of $G$, it cannot be used directly to define a cochain on the combinatorial moduli spaces. To this end we need some sign rules so that, for instance, the displayed isomorphisms acquire a sign. The tensors $p$ and $T_k$ are sometimes referred to as the {\it propagator} and the {\it interactions}, respectively, to remind us of their physical origin. If $p$ is nondegenerate, then we may use it to identify $V$ with its dual. In this case $T_k$ defines a linear map $V^{\otimes (k-1)}\to V$. The properties one needs to impose on propagator and interactions in order that the above recipe produce a cocycle on ${\mathbb M} _g^n$ is that they define a ${\mathbb Z} /2$ graded $A_{\infty}$ algebra with inner product. A similar recipe assigns cycles on ${\mathbb M}_g^n$ to certain ${\mathbb Z} /2$ graded differential algebras. The cocycles can be evaluated on the cycles and this, in principle, gives a method of showing that some of the classes thus obtained are nonzero. We shall not be more precise, but instead refer to \cite{kontsevich:feynman} or \cite{seminar} for an overview. A simple example is to take $V={\mathbb C}$, $p:=1\otimes1$ and $T_k(z^k)$ arbitrary for $k\ge 3$ odd, and zero otherwise. Kontsevich asserts that the classes thus obtained are all tautological. \subsection{Outer space}\label{subsec:outer} In Section~\ref{sec:ribbon} we encountered a beautiful combinatorial model for a virtual classifying space of the mapping class group $\Gamma _g^n$. There is a similar, but simpler, combinatorial model that does the same job for the outer automorphism group of a free group. We fix an integer $r\ge 2$ and consider connected graphs $G$ with first Betti number equal to $r$ and where each vertex has degree $\ge 3$. Let us call these graphs {\it $r$-circular graphs}. The maximal number of edges (resp.\ vertices) such a graph can have is $3r-3$ (resp.\ $2r-2$). These bounds are realized by all trivalent graphs of this type. Notice that an $r$-circular graph $G$ has fundamental group isomorphic to the free group on $r$ generators, $F_r$. We say that $G$ is {\it marked} if we are given an isomorphism $\phi :F_r\to\pi _1(G,\text{base point})$ up to inner automorphism. The group $\Out (F_r)$ permutes these markings simply transitively. There is an obvious notion of isomorphism for marked $r$-circular graphs. We shall denote the collection of isomorphism classes by ${\mathcal G} _r$. Let $(G,[\phi ])$ represent an element of ${\mathcal G} _r$. The metrics on $G$ that give $G$ total length $1$ are parameterized by the interior of a simplex $\Delta (G)$. We fit these simplices together in a way analogous to the ribbon graph case: if $s$ is an edge of $G$ that is not a loop, then collapsing it defines another element $(G/s,[\phi ]/s)$ of ${\mathcal G} _r$. We may then identify $\Delta (G/s)$ with a face of $\Delta (G)$. After we have made these identifications we end up with a simplicial complex $\widehat{{\mathbb O}} _r$. The union of the interiors of the simplices $\Delta (G)$ (indexed by ${\mathcal G} _r$) will be denoted by ${\mathbb O} _r$; it is the complement of a closed subcomplex of $\widehat{{\mathbb O}} _r$. This construction is due to Culler-Vogtmann \cite{culler}. We call ${\mathbb O} _r$ the {\it outer space of order $r$} for reasons that will become apparent in a moment. Observe that $\widehat{{\mathbb O}} _r$ comes with a simplicial action of $\Out (F_r)$. We denote the quotient of $\widehat{{\mathbb O}} _r$ (resp.\ ${\mathbb O} _r$) by $\Out (F_r)$ by $\widehat{{\mathbb G}}_r$ (resp.\ ${\mathbb G} _r$). It is easy to see that $\widehat{{\mathbb G}} _r$ is a finite orbicomplex. The open subset ${\mathbb G} _r$ is the moduli space of metrized $r$-circular graphs. It has a spine of dimension $2r-3$. \begin{theorem}[Culler-Vogtmann \cite{culler}, Gersten] The outer space of order $r$ is contractible and a subgroup of finite index of $\Out (F_r)$ acts freely on it. Hence ${\mathbb G} _r$ is a virtual classifying space for $\Out (F_r)$ and $\Out (F_r)$ has virtual homological dimension $\le 2r-3$. \end{theorem} In contrast to the ribbon graph case, ${\mathbb O} _r$ is not piecewise smooth. If we choose $2g-1+n$ free generators for the fundamental group of our reference surface $S_g^n$, then each ribbon graph without vertices of degree $\le 2$ determines an element of ${\mathbb G} _{2g-1+n}$: simply forget the ribbon structure. The ribbon data is finite and it is therefore not surprising that forgetting the ribbons defines a finite map $$ \widehat{f} : {\mathcal S} _n\backslash\widehat{{\mathbb M}} _g^n\to \widehat{{\mathbb G}} _{2g-1+n} $$ of orbicomplexes. Here ${\mathcal S} _n$ stands for the symmetric group, which acts in the obvious way on $\widehat{{\mathbb M}} _g^n$. Following Strebel's theorem, the preimage of ${\mathbb G} _{2g-1+n}$ can be identified with ${\mathcal S} _n\backslash ({\mathcal M} _g^n\times \inn \Delta ^{n-1})$. We denote the resulting map by $$ f: {\mathcal S} _n\backslash ({\mathcal M} _g^n\times \inn\Delta ^{n-1})\to {\mathbb G} _{2g-1+n}. $$ It induces the evident map $$ f_*: H_k(\Gamma _g^n)_{{\mathcal S} _n}\to H_k(\Out F_{2g-1+n}) $$ on rational homology. It is unclear whether there is such an interpretation for the induced map on cohomology with compact supports. We remark that ${\mathcal M} _g^n\times \inn\Delta ^{n-1}$ is canonically oriented, but that its ${\mathcal S} _n$-orbit space is not (since transpositions reverse this orientation). Poincar\'e duality therefore takes the form $$ H_k({\mathcal M} _g^n)_{{\mathcal S} _n}\cong H_k({\mathcal S} _n\backslash ({\mathcal M} _g^n\times \inn\Delta ^{n-1});\epsilon)\cong H_c^{6g-7+3n-k}({\mathcal S} _n\backslash ({\mathcal M} _g^n\times \inn\Delta ^{n-1}); \epsilon ), $$ where $\epsilon$ is the signum representation of ${\mathcal S} _n$. If $\delta$ denotes the (signum) character of $\Out (F_r)$ on $\wedge ^rH_1(F_r)$, then the adjoint of $f_*$ is a map $$ H_c^{6g-7+3n-k}({\mathbb G} _{2g-1+n};\delta )\to H_c^{6g-7+3n-k}({\mathcal S} _n\backslash ({\mathcal M} _g^n\times \inn\Delta ^{n-1}); \epsilon)\cong H_k({\mathcal M} _g^n)_{{\mathcal S} _n}. $$ So, when $m\ge 1$, we have maps $$ H_c^{i+m-1}({\mathbb G} _{m+1};\delta )\to\oplus _{2g-2+n=m}H_{2m-i} ({\mathcal M} _g^n)_{{\mathcal S} _n}\to H_{2m-i}({\mathbb G} _{m+1}),\quad i=0,1,\dots . $$ There is a remarkable interpretation of this sequence that we will discuss next. \subsection{Three Lie algebras}\label{subsec:threeLie} We describe Kontsevich's three functors from the category of symplectic vector spaces to the category of Lie algebras and their relation with the cohomology of the moduli spaces ${\mathcal M} _g^n$. The basic references are \cite{kontsevich:symp} and \cite{kontsevich:feynman}. We start out with a finite dimensional ${\mathbb Q}$ vector space $V$ endowed with a nondegenerate antisymmetric tensor $\omega _V\in V\otimes V$. Let $\Ass (V)$ be the tensor algebra (i.e., the free associative algebra) generated by $V$. We grade it by giving $V$ degree $-1$. The Lie subalgebra generated by $V$ is free and so we denote it by $\Lie (V)$. It is well-known that $\Ass (V)$ may be identified with the universal enveloping algebra of $\Lie (V)$. If we mod out $\Ass (V)$ by the two-sided ideal generated by the degree $\le -2$ part of $\Lie (V)$, we obtain the symmetric algebra $\Com (V)$ of $V$. Define ${\mathfrak g} _{{\rm ass}} (V)$ (resp.\ ${\mathfrak g} _{{\rm lie}} (V)$) to be the Lie algebra of derivations of $\Ass (V)$ (resp.\ $\Lie (V)$) of degree $\le 0$ that kill $\omega _V$. Since each derivation of $\Lie (V)$ extends canonically to its universal enveloping algebra, we have an inclusion ${\mathfrak g} _{{\rm lie}} (V)\subset{\mathfrak g} _{{\rm ass}} (V)$. There is also a corresponding Lie algebra ${\mathfrak g} _{{\rm com}} (V)$ of derivations of degree $\le 0$ of $\Com (V)$ that kill $\omega _V$. Here we regard the latter as a two-form on the affine space $\spec \Com (V)$. This Lie algebra is a quotient of ${\mathfrak g} _{{\rm ass}} (V)$. All three Lie algebras are graded and have as degree zero summand the Lie algebra $\sp (V)$ of the group $\Symp (V)$ of symplectic transformations of $V$. A simple verification shows that the degree $-1$ summands have as $\sp (V)$ representations the following natural descriptions: $$ {\mathfrak g}_{{\rm com}}(V)_{-1}\cong S^3(V),\quad {\mathfrak g}_{{\rm ass}}(V)_{-1}\cong S^3(V)\oplus \wedge ^3V,\quad {\mathfrak g}_{{\rm lie}}(V)_{-1}\cong \wedge^3V. $$ These Lie algebras are functorial with respect to symplectic injections $(V,\omega _V)\hookrightarrow (W,\omega _W)$. Note that $\Symp (V)$ acts trivially on this cohomology of the Lie algebra in question because $\sp (V)\subset{\mathfrak g} _*(V)$. This implies that $H^k({\mathfrak g} _*(V))$, $\ast \in \{{\rm lie},{\rm ass},{\rm com}\}$, depends only on $\dim V$. We form the inverse limit: $$ H^k({\mathfrak g} _*):=\varprojlim_{V}H^k({\mathfrak g} _*(V)). $$ The sum over $k$, $H^{\bullet}({\mathfrak g} _*)$, has the structure of a connected graded bicommutative Hopf algebra; the coproduct comes from the direct sum operation on symplectic vector spaces. It is actually bigraded: apart from the cohomological grading there is another coming from the grading of the Lie algebras. Notice that the latter grading has all its degrees $\ge 0$. The primitive part $H^{{\bullet}}_{{\rm pr}}({\mathfrak g} _*)$ inherits this bigrading. Furthermore, the natural maps $$ H^{{\bullet}}({\mathfrak g} _{{\rm com}})\to H^{{\bullet}}({\mathfrak g} _{{\rm ass}})\to H^{{\bullet}}({\mathfrak g} _{{\rm lie}}) $$ are homomorphisms of bigraded Hopf algebras. Consequently, we have induced maps between the bigraded pieces of their primitive parts. \begin{theorem}[Kontsevich \cite{kontsevich:symp}, \cite{kontsevich:feynman}] \label{thm:lie} For $\ast \in \{{\rm lie},{\rm ass},{\rm com}\}$ we have $$ H^k_{{\rm pr}}({\mathfrak g} _*)_0=H^k_{{\rm pr}}(\sp_\infty )\cong \begin{cases} {\mathbb Q} &\text{for }k=3,7,11,\dots\, ;\\ 0 &\text{otherwise}. \end{cases} $$ Furthermore, $H^k_{{\rm pr}}({\mathfrak g} _*)_l=0$ when $l$ is odd and, when $m>0$, we have a natural diagram $$ \begin{CD} H^k_{{\rm pr}}({\mathfrak g} _{{\rm com}})_{2m} @>>> H^k_{{\rm pr}}({\mathfrak g} _{{\rm ass}})_{2m} @>>> H^k_{{\rm pr}}({\mathfrak g} _{{\rm lie}})_{2m} \cr @V{\cong }VV @V{\cong }VV @V{\cong }VV \cr H^{k+m-1}_c({\mathbb G} _{m+1},\delta ) @>{f_c^*}>> \oplus _{2g-2+n=m}H_{2m-k}({\mathcal M} _g^n)_{{\mathcal S} _n} @>{f_*}>> H_{2m-k}({\mathbb G}_{m+1})\cr \end{CD} $$ which commutes up to sign and whose rows are complexes. The maps in the top row are the natural maps and the bottom row is the sequence defined in Section~\ref{subsec:outer}. \end{theorem} The proof is an intelligent application of classical invariant theory. For each of the three Lie algebras one writes down the standard complex. The subcomplex of invariants with respect to the symplectic group is quasi-isomorphic to the full complex. Weyl's invariant theory furnishes a natural basis for this subcomplex. Kontsevich then observes that this makes the subcomplex naturally isomorphic to a cellular chain (or cochain) complex of one of the cell complexes ${\mathbb G} _*$ and ${\mathbb M} _*^*$ whose (co)homology appears in the bottom row. The diagram in this theorem suggests that the sequence of natural transformations $\Lie\to\Ass\to\Com$ is self dual in some sense. This can actually be pinned down by looking at the corresponding operads: Ginzburg and Kapranov \cite{ginz} observed that these operads have ``quadratic relations'' and they proved the self duality of the operad sequence in a Koszul sense. However, our main reason for displaying this diagram is that it pertains to the cohomology of the moduli spaces of curves in two apparently unrelated ways. The first one is evident. The ${\mathcal S} _n$ coinvariants of the homology of ${\mathcal M} _g^n$ features in the middle column, but the righthand column has something to do with the cohomology of a `linearization' of $\Gamma _{\infty}$: we will see that ${\mathfrak g} _{{\rm lie}}$ is intimately related to the Lie algebra of the relative Malcev completions discussed in Section~\ref{sec:malcev}. We explain this in the next subsection after a giving a restatement of Kontsevich's Theorem. In this restatement the Lie algebra cohomology of ${\mathfrak g}_\ast (V)$ is replaced by the relative Lie algebra cohomology of the pair $({\mathfrak g}_\ast (V),{\mathfrak k}(V))$, where ${\mathfrak k}(V)$ is a maximal compact Lie subalgebra of ${\mathfrak g}_\ast (V)_0$, and therefore of ${\mathfrak g}_\ast (V)$ (${\mathfrak k}(V)$ is a unitary Lie algebra of rank $\dim V/2$). As above, the Lie algebra cohomology $H^k({\mathfrak g}_\ast (V),{\mathfrak k}(V))$ depends only on the dimension of $V$ and stabilizes once $\dim V$ is sufficiently large. We denote the inverse limit of these groups by $H^{\bullet}({\mathfrak g}_\ast,{\mathfrak k}_\infty)$. Likewise, we denote the stable cohomology of the pair $(\sp_g,{\mathfrak k}_g)$ by $H^{\bullet}(\sp_\infty,{\mathfrak k}_\infty)$. By a theorem of Borel \cite{borel:triv}, this is naturally isomorphic to the stable cohomology of ${\mathcal A}_g$ and is a polynomial algebra generated by classes $c_1, c_3, c_5, \dots$, where $c_k$ has degree $2k$. Combining Kontsevich's Theorem~\ref{thm:lie} with Borel's computation and an elementary spectral sequence argument, we obtain the following result. (Use the fact that $(\sp_g,{\mathfrak k}_g)$ is both a sub and a quotient of $({\mathfrak g}_\ast(V),{\mathfrak k}(V))$.) \begin{corollary}\label{rel_lie} We have $$ H^k_{{\rm pr}}({\mathfrak g}_\ast,{\mathfrak k}_\infty)_0=H^k_{{\rm pr}}(\sp_\infty,{\mathfrak k}_\infty )\cong \begin{cases} {\mathbb Q} &\text{for }k=2,6,10,\dots\, ;\\ 0 &\text{otherwise}. \end{cases} $$ Furthermore, when $m > 0$, the natural maps $H^{\bullet}({\mathfrak g}_\ast,{\mathfrak k}_\infty)_m \to H^{\bullet}({\mathfrak g}_\ast)_m$ are isomorphisms. \end{corollary} \subsection{Relation with the relative Malcev completion} \label{subsec:relation} We begin with an observation. For a symplectic vector space $V$ and $\ast \in \{{\rm lie},{\rm ass},{\rm com}\}$, denote by ${\mathfrak g} _*^{\flat}(V)$ the subalgebra of ${\mathfrak g} _*(V)$ generated by its summands of weight $0$ and $-1$. Kontsevich's computation shows: \begin{proposition} The graded cohomology groups $H^k ({\mathfrak g} _*^{\flat} (V),{\mathfrak k}(V))_l$ stabilize and the sum of the stable terms is a bigraded bicommutative Hopf algebra $H^{\bullet}({\mathfrak g}_*^{\flat},{\mathfrak k}_\infty)_{{\bullet}}$. In addition, the restriction map $H^k_{{\rm pr}}({\mathfrak g}_*,{\mathfrak k}_\infty)_l\to H^k_{{\rm pr}}({\mathfrak g}_*^{\flat}(V),{\mathfrak k}_{\rm lie}(V))_l$ is an isomorphism when $l\le k$. \end{proposition} The case of interest here is that of lie where ${\mathfrak g} _{\rm lie}^{\flat}(V)_0=\sp (V)$ and ${\mathfrak g} _{\rm lie}^{\flat}(V)_{-1}\cong \wedge ^3V$. Denote by $z_m$ the element of $H^{2m}_{{\rm pr}}({\mathfrak g}_{\rm lie} )_{2m}$ that corresponds, via Theorem~\ref{thm:lie}, to $1\in H_0({\mathbb G} _{m+1})$. The preceding proposition yields: \begin{corollary}\label{cor:weightcontrol} We have a natural isomorphism of bigraded Hopf algebras $$ \sum_{l\le k} H^k({\mathfrak g}_{{\rm lie}}^{\flat},{\mathfrak k}_\infty)_l\cong H^{\bullet}(\sp_\infty,{\mathfrak k}_\infty)[z_1,z_2,\dots ] \cong {\mathbb C}[c_1,c_3,c_5,\dots,z_1,z_2,z_3,\dots] $$ where each $z_i$ and $c_j$ is primitive. \end{corollary} The graded Lie algebra ${\mathfrak g}^\flat_{\rm lie}(V)$ is the semi-direct product of $\sp(V)$ and its elements of positive weight, which we shall denote by ${\mathfrak u}^\flat_{\rm lie}$. Consequently, there are natural inclusions $$ H^{\bullet}({\mathfrak u}^\flat_{\rm lie}(V))^{Sp} \hookrightarrow H^{\bullet}({\mathfrak g}^\flat_{\rm lie}(V),{\mathfrak k}(V)) \text{ and } H^{\bullet}(\sp(V),{\mathfrak k}(V)) \hookrightarrow H^{\bullet}({\mathfrak g}^\flat_{\rm lie}(V),{\mathfrak k}(V)). $$ Together these induce an algebra homomorphism $$ H^{\bullet}(\sp(V),{\mathfrak k}(V))\otimes H^{\bullet}({\mathfrak u}^\flat_{\rm lie}(V))^{Sp} \to H^{\bullet}({\mathfrak g}^\flat_{\rm lie}(V),{\mathfrak k}(V)) $$ which is compatible with stabilization. \begin{proposition}\label{stab_prod} Upon stabilization, these maps induce an isomorphism $$ H^{\bullet}(\sp_\infty,{\mathfrak k}_\infty)\otimes H^{\bullet}({\mathfrak u}^\flat_{\rm lie})^{Sp} \to H^{\bullet}({\mathfrak g}^\flat_{\rm lie},{\mathfrak k}_\infty). $$ \end{proposition} Next we relate the graded Lie algebra ${\mathfrak g}^\flat_{\rm lie}$ to the filtered Lie algebra ${\mathfrak g}_{g,1}$ of the relative Malcev completion ${\mathcal G} _{g,1}$ of $\Gamma_{g,1}\to Sp_g$. Recall from Section~\ref{sec:groups} that $\pi_g^1$ is freely generated by $2g$ generators named $\alpha _{\pm},\dots ,\alpha_{\pm g}$ so that the commutator $\beta:=(\alpha_1,\alpha_{-1})\cdots (\alpha_g,\alpha_{-g})$ represents a simple loop around $x_1$. Using Latin letters for the logarithms of the images of elements of $\pi _g^1$ in its Malcev completion, we find that $$ b\equiv [a_1,a_{-1}]+\cdots +[a_g,a_{-g}]\mod{({\mathfrak p} _g^1)^{(3)}}. $$ So the image of $b$ in $\Gr^2{\mathfrak p}_g^1\cong \wedge^2 V_g$ is the symplectic form $\omega_S$. The obvious homomorphism $\Gamma_{g,1} \to \Aut (\pi_g^1)$ induces a Lie algebra homomorphism \begin{equation}\label{map} {\mathfrak g} _{g,1} \to \Der {\mathfrak p}_g^1. \end{equation} whose image we denote by $\overline{\g} _{g,1}$.\footnote{It is possible that this map is injective so that ${\mathfrak g} _{g,1}\cong\overline{\g} _{g,1}$. Note that it is not surjective --- see \cite{morita:trace} and \cite{nakamura:obstn}, and Section~\ref{subsec:understanding}.} Notice that $\overline{\g}_{g,1}$ is contained in the subalgebra $\Der({\mathfrak p}_{g,1},b)$ consisting of those derivations that kill $b$. Since (\ref{map}) is (Malcev) filtration preserving, it induces Lie algebra homomorphisms $$ \Gr^{\bullet} {\mathfrak g}_{g,1}\to \Gr^{\bullet}\overline{\g}_{g,1}\to \Der^{\bullet}(\Gr{\mathfrak p}_g^1,\omega_S). $$ Notice that the last term is just ${\mathfrak g} _{{\rm lie}}(V_g)$. In view of (\ref{stab_prod}), to construct an algebra homomorphism $$ H^{\bullet}({\mathfrak g}^\flat_{\rm lie},{\mathfrak k}_\infty) \to H^{\bullet}(\Gamma_\infty) $$ it suffices to construct an algebra homomorphism $H^{\bullet}({\mathfrak u}^\flat_{\rm lie})^{Sp_g} \to H^{\bullet}(\Gamma_\infty)$. At this stage we need Hodge theory. Choose a conformal structure on $S_g$. Then, by~(\ref{mhs}), there are natural MHSs on ${\mathfrak g}_{g,1}$ and ${\mathfrak p}_{g,1}$ whose weight filtrations are the Malcev filtrations and such that (\ref{map}) is a MHS morphism. Hence the image $\overline{\g}_{g,1}$ has a natural MHS. Since $\Gr^{\bullet} {\mathfrak g}_{g,1}$ is generated by its summands in degree $0$ and $1$, the same is true for $\Gr^{\bullet}\overline{\g}_{g,1}$. On the other hand, the summands in degree $0$ and $1$ of $\Gr^{\bullet}\overline{\g}_{g,1}$ are equal to the summands of weight $0$ and $-1$ of ${\mathfrak g} _{{\rm lie}}(V_g)$, and so the graded Lie algebra $\Gr^{\bullet}\overline{\g}_{g,1}$ may be identified with ${\mathfrak g}_{\rm lie}^\flat (V_g)_{{\bullet}}$ (except that the indexing of the summands differs by sign). Denote the pronilpotent radical $W_{-1}\overline{\g}_{g,1}$ of $\overline{\g}_{g,1}$ by $\overline{\u}_{g,1}$. We know from Section~\ref{sec:hodgemap} that the homomorphisms \begin{equation}\label{sequence} H^{\bullet} (\overline{\u}_{g,1})^{Sp_g}\to H^{\bullet}({\mathfrak u}_{g,1})^{Sp_g}\to H^{\bullet}(\Gamma_{g,1}) \end{equation} are morphisms of MHS. After weight grading these become bigraded algebra homomorphisms $$ H^{\bullet}({\mathfrak u}_{\rm lie}^\flat (V_g)_{{\bullet}})^{Sp_g}\to H^{\bullet}(\Gr^W_{{\bullet}}{\mathfrak u}_{g,1})^{Sp_g} \to \Gr^W_{\bullet} H^{\bullet}(\Gamma_{g,1}). $$ The sequence (\ref{sequence}) stabilizes with $g$ to a sequence of Hopf algebras in the MHS category. The corresponding weight graded sequence is $$ H^{\bullet}(({\mathfrak u}_{\rm lie}^\flat )_{{\bullet}})^{Sp_g}\to H^{\bullet}(\Gr^W_{{\bullet}}{\mathfrak u}_{\infty,1})^{Sp_g} \to \Gr^W_{\bullet} H^{\bullet}(\Gamma_{\infty}). $$ Each term in this sequence is a Hopf algebra and each map a Hopf algebra homomorphism. But by Pikaart's Purity Theorem we know that the last term is pure of weight $k$ in degree $k$, so that we can replace it by $H^{\bullet}(\Gamma_{\infty})$ and obtain a map $H^{\bullet}({\mathfrak u}^\flat_{\rm lie})^{Sp_g} \to H^{\bullet}(\Gamma_\infty)$. We therefore have a Hopf algebra homomorphism $$ H^{\bullet}({\mathfrak g}^\flat_{\rm lie},{\mathfrak k}_\infty) \cong H^{\bullet}(\sp_\infty,{\mathfrak k}_\infty)\otimes H^{\bullet}({\mathfrak u}^\flat_{\rm lie})^{Sp_g} \to H^{\bullet}(\Gamma_\infty). $$ If we compose the natural restriction map $H^{\bullet}({\mathfrak g}_{\rm lie}, {\mathfrak k}_\infty)\to H^{\bullet}({\mathfrak g}_{\rm lie}^\flat,{\mathfrak k}_\infty)$ with the above maps we get a homomorphism $$ H^{\bullet}({\mathfrak g}_{\rm lie},{\mathfrak k}_\infty)\to H^{\bullet}(\Gamma_{\infty}). $$ Kontsevich asked (at the end of \cite{kontsevich:feynman}) about the meaning of this map.\footnote{Actually, Kontsevich asks this for with $H^{\bullet}({\mathfrak g}_{\rm lie})$ in place of the relative Lie algebra cohomology. However, there does not seem to be a natural homomorphism to $H^{\bullet}(\Gamma_\infty)$ in this case.} This can now be answered by invoking the theorem of Kawazumi and Morita (\ref{kawazumi-morita}), or rather a weaker form, which says that $z_i$ is mapped to a nonzero multiple of $\kappa_i$. This result was obtained with Kawazumi and Morita. \begin{theorem} There is a natural Hopf algebra homomorphism $$ H^{\bullet}({\mathfrak g}_{\rm lie},{\mathfrak k}_\infty)\to H^{\bullet}(\Gamma_{\infty}). $$ The left hand side is a polynomial algebra generated by primitive elements $z_1, z_2,\dots$ and $c_1, c_3, \dots $ where $z_i$ has degree $2i$ and $c_j$ has degree $2j$. The image of this homomorphism is precisely the subalgebra generated by the $\kappa_i$s. The kernel is generated by elements of the form $$ c_{2k+1} - a_k z_{2k+1} - P_{2k+1}, \quad k \in \{0,1,2,\dots \} $$ where $P_{2k+1}$ is a polynomial in the $z_i$ and $c_j$ with no linear terms , and $a_k$ is a non zero rational number. \end{theorem} Here we have used the fact, due to Mumford \cite{mumford}, that the image of $c_{2k+1}$ in $H^{\bullet} ({\mathcal M}_{g,1})$ is a polynomial in the odd $\kappa_i$'s. The theorem indicates that no new stable classes are to be expected from Hodge theory --- that is, a de~Rham version of the Mumford conjecture holds. Recent work of Kawazumi and Morita attempts to explain the kernel of the homomorphism $H^{{\bullet}}(\overline{\g}_{\infty})\to H^{{\bullet}}(\Gamma_{\infty })$ in terms of secondary characteristic classes of surface bundles. The first element of the kernel is the difference $c_1 - 12\, z_1$. Its restriction to the Torelli group can be interpreted as the Casson Invariant (cf.\ \cite{morita:casson}.)
1996-07-17T01:43:24
9607
alg-geom/9607014
en
https://arxiv.org/abs/alg-geom/9607014
[ "alg-geom", "math.AG" ]
alg-geom/9607014
null
Eriko Hironaka
Torsion Points on an Algebraic Subset of an Affine Torus
LaTeX, 25 pages, 4 figures. email: [email protected]
null
null
null
null
Work of Laurent and Sarnak, following a conjecture of Lang, shows that the number of torsion points of order n on an algebraic subset of an affine complex torus is polynomial periodic. In this paper, we find bounds on the degree and period of this number as a function of n. Some examples, including the number of n torsion points on Fermat curves, are computed to illustrate the methods.
[ { "version": "v1", "created": "Tue, 16 Jul 1996 23:33:34 GMT" } ]
2008-02-03T00:00:00
[ [ "Hironaka", "Eriko", "" ] ]
alg-geom
\section{Introduction.} Let $V \subset ({\Bbb C}^*)^r$ be an algebraic subset. Laurent has shown that the torsion points on $V$ lie on a finite union of translates of affine subtori contained in $V$. It follows that the number, $p_V(n)$, of torsion points of order $n$ on $V$ is polynomial periodic. In this paper, we find formulas and bounds for the degree and period of $p_V(n)$ in terms of defining equations for $V$. Counting torsion points on algebraic subsets of an the affine torus is useful for studying abelian representations of a finitely presented group $\Gamma$. For example, the Alexander invariants define a stratification of the character variety for $\Gamma$, which naturally embeds in $({\Bbb C}^*)^r$, and the first Betti number of unbranched coverings can be computed from the torsion points on these strata (cf. \cite{A-S:Betti}, \cite{Hiro:Alex}). In \cite{Laur:Equ}, Laurent gives the following description of the set of torsion points ${\mathrm{Tor}}(V)$ on $V$. \begin{theorem} (Laurent) For any algebraic subset $V$ of $({\Bbb C}^*)^r$ there is a finite set of rational planes $Q_1,..,Q_\ell \subset V$ such that $$ {\mathrm{Tor}}(V) = \bigcup_{i=1}^\ell {\mathrm{Tor}}(Q_i). $$ \end{theorem} Here, a {\it rational plane} $Q \subset ({\Bbb C}^*)^r$ is a subset of the form $\eta P$, where $P \subset ({\Bbb C}^*)^r$ is an {\it affine subtorus}, or connected algebraic subgroup, and $\eta \in ({\Bbb C}^*)^r$ is an element of finite order. Laurent's result extends further than the statement we give here and settles a more general conjecture of Lang (\cite{Lang:Conj}, p.220). It follows from Theorem 1 (cf. \cite{A-S:Betti}) that $p_V(n)$ is a {\it polynomial periodic} function in $n$, that is, there exist periodic functions $a_0(n), \dots, a_d(n)$ such that $$ p_V(n) = a_0(n) + a_1(n)n + \dots + a_d(n)n^d. $$ For, by Theorem 1, the Zariski closure of $V$ has a decomposition into a finite union of rational planes $$ \Zar{{\mathrm{Tor}}(V)} = Q_1 \cup \dots \cup Q_\ell $$ and hence $p_V(n)$ is given by the following formula \begin{eqnarray} p_V(n) = \sum_{k=1}^\ell \quad \sum_{1\leq i_1 < \dots < i_k \leq\ell} (-1)^{\ell-k} p_{Q_{i_1} \cap \dots \cap Q_{i_k}}(n). \end{eqnarray} It is not hard to see that a finite intersection of rational planes is a finite union of disjoint rational planes (see Prop. 3.6 for a more precise description). Furthermore, the number of $n$-torsion points on a rational plane $Q$ is given by \begin{eqnarray*} p_Q(n) &=& \left\{\begin{array}{ll} n^{\dim (Q)} &\quad\mbox{if ${\mathrm{ord}}(Q)\ |\ n$,}\\ 0 &\quad\mbox{otherwise,} \end{array}\right . \end{eqnarray*} where ${\mathrm{ord}}(Q)$ is the least integer $n$ such that $Q^n$ contains the identity element of $({\Bbb C}^*)^r$. We will concentrate on finding the degree and period of $p_V$. The {\it degree} of $p_V$, written $\deg (p_V)$, is the largest $d$ such that $a_d(n)$ is not constantly zero and the {\it period} of $p_V$, written ${\mathrm{per}} (p_V)$, is the least common multiple of the periods of $a_0(n),\dots,a_d(n)$. Thus, for example, if $Q \subset ({\Bbb C}^*)^r$ is a rational plane, then ${\mathrm{ord}} (Q) = {\mathrm{per}} (p_Q)$. As with ordinary polynomials the degree of $p_V$ determines the order of growth of $p_V(n)$: if $\deg(p_V) = d$, then $$ p_V(n) = O(n^d) $$ and for some fixed integer $c$ $$ p_V(n) \asymp n^d $$ for all $n \equiv c\ ({\mathrm{mod}}\ {\mathrm{per}}(p_V))$. If $\Zar{{\mathrm{Tor}}(V)} = Q_1 \cup \dots \cup Q_\ell$, where $Q_1,\dots,Q_\ell$ are rational planes, then (1) implies that \begin{description} \item{(i)} $\deg (p_V) = \max\ \{ \dim(Q_i) : i=1,\dots,\ell \} $, \item{(ii)} ${\mathrm{per}} (p_V)$ divides the least common multiple of ${\mathrm{ord}}(Q)$ where $Q$ ranges among connected components of $$ Q_{i_1} \cap \dots \cap Q_{i_k}, \quad 1 \leq i_1 < \dots < i_k \leq \ell. $$ \end{description} In section 2, Theorem 2, we give bounds for $\deg(p_V)$ and ${\mathrm{per}}(p_V)$, when $V$ is defined over ${\Bbb Q}$. More precise formulas for $\deg(p_V)$ and ${\mathrm{per}}(p_V)$, for general $V$, are obtained later in section 5 after developing notation and theory in sections 3 and 4. The main ingredients of Laurent's proof of Theorem 1 can be stated as follows. Given an algebraic subset $V \subset ({\Bbb C}^*)^r$, one defines a finite set $\Pi$ of mappings, $$ \phi: ({\Bbb C}^*)^r \rightarrow \Gamma_{\phi}, \qquad \phi \in \Pi $$ to algebraic groups $\Gamma_{\phi}$. These mappings have the following properties. \begin{description} \item{(A)} All the fibers of $\phi$ are finite unions of rational planes; and \item{(B)} $$ {\mathrm{Tor}}(V) = \bigcup_{\phi\in \Pi} \phi^{-1}({\cal S}_{\phi}), $$ where ${\cal S}_{\phi} \subset \Gamma_{\phi}$ are finite subsets. \end{description} We will go further in this paper by finding the fibers in (A) and the sets $S_\phi$ in (B) explicitly and relating them to the degree and period of $p_V$. Techniques for finding the fibers in (A) are given in section 3, where we develop some tools for studying rational planes and give properties of monomial mappings. Our approach is to focus on the relationship between algebraic subgroups of $({\Bbb C}^*)^r$ and subgroups of ${\Bbb Z}^r$. We use this correspondence to describe the rational planes in fibers of monomial mappings (see Lemma 3.4 and Cor. 3.5) and, as a consequence, in the intersection of a finite set of rational planes (see Prop. 3.6). In many natural applications, for example, the Alexander strata mentioned above, the algebraic subsets $V$ are defined over ${\Bbb Q}$, so we will concentrate on this case. Then, the sets $S_\phi$ in (B) are related to formal ${\Bbb Q}$-linear combinations of roots of unity. In section 4, we review some of the theory of ${\Bbb Q}$-linear relations among roots of unity using ideas of Schoenberg. The main idea is to view formal ${\Bbb Q}$-linear relations as convex polygons with rational sides and angles. Schoenberg shows in \cite{Sch:Cyc} that all the convex polygons can be obtained from regular $p$-gons, where $p$ ranges over prime numbers. We describe these ideas in section 4 and include proofs of Mann's bound on the order of roots of unity satisying a linear equation of a given length (see Prop. 4.3) and of Schoenberg's result (see Cor. 4.4). The results in sections 3 and 4 are used in section 5 to find the rational planes in an algebraic subset $V \subset ({\Bbb C}^*)^r$ (see Prop. 5.2) and to give formulas for $\deg(p_V)$ and bounds on ${\mathrm{per}}(p_V)$ (see Prop. 5.3 and Theorem 3) in terms of defining equations for $V$. In section 6 we conclude with some illustrative examples. For instance, we find $p_V$ when $V$ is the Fermat curve $$ x^m + y^m = 1, \qquad m \ge 1; $$ and show (see Example 3) that, in this case, the bounds for the degree and period of $p_V$ given in Theorem 2 are attained. \heading{Acknowledgement.} I would like to thank E. Jahangard for useful discussions during the research for this paper. \section{Notation and main result.} In this section we will set up some notation to state our main result: Theorem 2. This theorem gives bounds on the degree and period of $p_V$ rather than exact formulas for them and applies to varieties $V$ defined over ${\Bbb Q}$. A more general result, Theorem 3, is stated and proved in section 5, but the bounds in Theorem 2 are easier to state and compute. \heading{Notation.} Let $\Lambda_r = {\Bbb C}[t_1,\dots,t_r]$ denote the ring of Laurent polynomials. A monomial $t_1^{\lambda_1}\dots t_r^{\lambda_r} \in \Lambda_r$ will be written as $t^\lambda$ where $\lambda = (\lambda_1,\dots,\lambda_r)$. Thus, coordinate functions $t_1,\dots,t_r$ and the usual basis for ${\Bbb Z}^r$ determine a canonical isomorphism between $\Lambda_r$ and ${\Bbb C}[{\Bbb Z}^r]$. A Laurent polynomial $f \in \Lambda_r = {\Bbb C}[t_1^{\pm 1},\dots,t_r^{\pm 1}]$ will be written as $$ f(t) = \sum_{\lambda \in {\cal L}(f)} a_\lambda t^{\lambda}, $$ where ${\cal L}(f) \subset {\Bbb Z}^r$ is a finite subset and $a_\lambda \in {\Bbb C}^*$. For a finite subset ${\cal F} \subset \Lambda_r$, our main theorem will describe the rational planes in the zero set $V({\cal F})$ and bounds for the degree and period of $p_{V({\cal F})}$ in terms of the functions in ${\cal F}$. Two main ingredients will be numbers $N[R({\cal F})]$ and $D({\cal F})$ which can be computed from the number of coefficients of functions in ${\cal F}$ and certain subgroups of ${\Bbb Z}^r$ associated to the exponents of functions in ${\cal F}$. Define $$ R({\cal F}) = \max\ \{|{\cal L}(f)| : f \in {\cal F}\}, $$ where $|{\cal L}(f)|$ denotes the number of elements in ${\cal L}(f)$. For any positive integer $R$, let $N[R]$ denote the product of primes less than or equal to $R$. The number $D({\cal F})$ takes longer to define. For any $f \in \Lambda_r$, let $\Pi_f$ be the set of all partitions ${\cal P}$ of ${\cal L}(f)$ such that each $\nu \in {\cal P}$ has at least two elements and for any finite subset ${\cal F} = \{f_1,\dots,f_\ell\} \in \Lambda_r$, let $$ \Pi_{{\cal F}} = \Pi_{f_1} \times \dots \times \Pi_{f_\ell}. $$ For any partition ${\cal P} \in \Pi_f$, let $\varepsilon({\cal P},f) \subset {\Bbb Z}^r$ be the subgroup generated by $$ \{\lambda-\mu : \ \exists\nu \in {\cal P},\ \lambda,\mu \in \nu\}; $$ for any $\pi = ({\cal P}_1,\dots,{\cal P}_k) \in \Pi_{{\cal F}}$, let $\varepsilon(\pi,{\cal F}) \subset {\Bbb Z}^r$ be the sum $$ \varepsilon({\cal P}_1,f_1) + \dots + \varepsilon({\cal P}_k,f_k); $$ and for any subset ${\cal U} \subset \Pi_{{\cal F}}$, let $\varepsilon({\cal U},{\cal F}) \subset {\Bbb Z}^r$ be the subgroup generated by $$ \bigcup_{\pi \in {\cal U}}\varepsilon(\pi,{\cal F}). $$ For any finite abelian group $G$, let $D(G)$ be the largest order of any element in $G$. Define $$ D({\cal U}, {\cal F}) = D(\overline{\varepsilon({\cal U},{\cal F})}/{\varepsilon({\cal U},{\cal F})}) $$ and $$ D({\cal F}) = \mathop{{\mathrm{lcm}}}\ \{D({\cal U},{\cal F}) : {\cal U} \subset \Pi_{{\cal F}}\}. $$ \smallskip Our main result is the following. \begin{theorem} Let $V = V({\cal F}) \subset ({\Bbb C}^*)^r$ be any algebraic subset defined over ${\Bbb Q}$. \begin{description} \item{(i)} For each maximal rational plane $Q$ in $V$, there is a $\pi \in \Pi_{{\cal F}}$ so that $Q$ is a translate of the affine subtorus of $({\Bbb C}^*)^r$ defined by the set of binomials $$ \{t^\lambda - 1 : \lambda \in \overline{\varepsilon(\pi,{\cal F})} \}; $$ \item{(ii)} $$\deg(p_V) \leq r - \min_{\pi\in \Pi_{{\cal F}}} {\mathrm{rank}}(\overline{\varepsilon(\pi,{\cal F})}); $$ \item{(iii)} $$ {\mathrm{per}}(p_V) \quad | \quad N[R({\cal F})]\ D({\cal F}). $$ \end{description} \end{theorem} This theorem together with finer versions will be proved in section 5. \section{Character groups and rational planes.} The main tool we use in this paper is a natural correspondence between between affine subtori of $({\Bbb C}^*)^r$ and subgroups of ${\Bbb Z}^r$. The relation is explained in section 3.1 and applied in section 3.2 to give properties of rational planes. \subsection{Subgroups of ${\Bbb Z}^r$ and algebraic subgroups of $({\Bbb C}^*)^r$.} For any subgroup $\varepsilon \subset {\Bbb Z}^r$, let $V(\varepsilon)$ be the closed points of ${\mathrm{Spec}} ({\Bbb C}[{\Bbb Z}^r/\varepsilon])$. Then, there is a natural identification of $V(\varepsilon)$ with the group of characters of ${\Bbb Z}^r/\varepsilon$, since any closed point corresponds to a ring epimorphism ${\Bbb C}[{\Bbb Z}^r/\varepsilon] \rightarrow {\Bbb C}$. Define $I_\varepsilon \subset \Lambda_r$ to be the ideal generated by $$ \{ t^\lambda - 1 : \lambda \in \varepsilon\}. $$ Then the kernel of the map ${\Bbb C}[{\Bbb Z}^r] \rightarrow {\Bbb C}[{\Bbb Z}^r/\varepsilon]$ is the image of $I_\varepsilon$ in ${\Bbb C}[{\Bbb Z}^r]$ under the identification $\Lambda_r \cong {\Bbb C}[{\Bbb Z}^r]$. Thus, there is a natural embedding of $V(\varepsilon)$ into $({\Bbb C}^*)^r$ with $V(\varepsilon) = V(I_\varepsilon)$ and $I_\varepsilon = I(V(\varepsilon))$. This embedding preserves group structure as well as the algebraic structure of $V(\varepsilon)$. For any algebraic subset $V \subset ({\Bbb C}^*)^r$, define $$ \varepsilon(V) = \{\lambda \in {\Bbb Z}^r : t^\lambda - 1\ \mbox{vanishes on}\ V\}. $$ \begin{lemma} If $\varepsilon \subset {\Bbb Z}^r$ is any subgroup, then $\varepsilon(V(\varepsilon)) = \varepsilon$. \end{lemma} \heading{Proof.} Observe that for any subgroup $\varepsilon \subset {\Bbb Z}^r$, if $t^\lambda-1 \in I_\varepsilon$, then $\lambda$ must be a sum of elements in $\varepsilon$. This gives the inclusion $\varepsilon(V(\varepsilon)) \subset \varepsilon$. The other inclusion is clear. \qed \begin{lemma} Any subtorus $P$ of $({\Bbb C}^*)^r$ is of the form $V(\varepsilon)$ for some $\varepsilon \subset {\Bbb Z}^r$. \end{lemma} \heading{Proof.} Since $P$ is itself isomorphic to $({\Bbb C}^*)^s$ for some $0 \leq s \leq r$, the embedding $\psi : P \rightarrow ({\Bbb C}^*)^r$ defines a ring epimorphism on coordinate rings $$ \psi^*: {\Bbb C}[{\Bbb Z}^r] \rightarrow {\Bbb C}[{\Bbb Z}^s]. $$ We will show that $\psi^*$ restricted to ${\Bbb Z}^r$ induces an epimorphism ${\Bbb Z}^r \rightarrow {\Bbb Z}^s$. This is the same as saying that the components of $\psi$ are monomials. Since $\psi$ preserves multiplication it also preserves the multiplication of $({\Bbb C}^*)$. Thus, the components of $\psi$ must be homogeneous. Furthermore, since the components of $\psi$ can have no zeros or poles, other than the origin, they must be monomials. Let $\varepsilon \subset {\Bbb Z}^r$ be the kernel of the epimorphism ${\Bbb Z}^r \rightarrow {\Bbb Z}^s$ induced by $\psi$. Then, the kernel of $\psi^*$ is $I_\varepsilon$, so $P = V(\varepsilon)$.\qed. \heading{Remark.} Another way to say the above is that there is a naturally duality between algebraic subgroups of $({\Bbb C}^*)^r$ and quotient groups of ${\Bbb Z}^r$ given by the contravariant functors ${\mathrm{Hom}}_{{\cal A}}(-,{\Bbb C}^*)$ and ${\mathrm{Hom}}(-,{\Bbb C}^*)$, where ${\mathrm{Hom}}_{{\cal A}}(-,-)$ are morphisms of algebraic subgroups of $({\Bbb C}^*)^r$ which preserve both the algebraic and multiplicative structure. For brevity and because they are not necessary for the results of this paper, we omit the details. \subsection{Rational planes and monomial mappings.} Given a subgroup $\varepsilon \subset {\Bbb Z}^r$, let $\overline{\varepsilon}$ be the subgroup of ${\Bbb Z}^r$ defined by $$ \overline{\varepsilon} = \{\lambda \in {\Bbb Z}^r : \exists n \in {\Bbb N},\ n\lambda \in \varepsilon \}. $$ For any subgroup $\varepsilon \subset {\Bbb Z}^r$ and $m \in {\Bbb N}$, let $$ \varepsilon_{m} = \{\lambda \in {\Bbb Z}^r : m\lambda \in \varepsilon \}. $$ \smallskip \begin{lemma} For any subgroup $\varepsilon \in {\Bbb Z}^r$, $V(\overline\varepsilon)$ is an affine subtorus of $({\Bbb C}^*)^r$. Furthermore, there is an injective (non-canonical) endomorphism $$ T : \chargp{\overline{\Epsilon}/\Epsilon} \hookrightarrow ({\Bbb C}^*)^r, $$ from the character group $\chargp{\overline{\Epsilon}/\Epsilon}$ into $({\Bbb C}^*)^r$, such that $V(\varepsilon)$ decomposes as a disjoint union $$ V(\varepsilon) = \bigcup_{\eta \in T(\chargp{\overline{\Epsilon}/\Epsilon})} \eta V(\overline\varepsilon). $$ \end{lemma} \heading{Proof.} The epimorphisms $$ {\Bbb Z}^r \rightarrow {\Bbb Z}^r/\varepsilon \mapright{} {\Bbb Z}^r/{\overline\varepsilon} $$ induce inclusions $$ V(\overline{\varepsilon}) \hookrightarrow V(\varepsilon) \hookrightarrow ({\Bbb C}^*)^r. $$ Since ${\Bbb Z}^r/{\overline{\varepsilon}}$ is a free abelian group, $V(\overline{\varepsilon})$ is an affine subtorus of $({\Bbb C}^*)^r$. The inclusion $\overline{\Epsilon}/\Epsilon \hookrightarrow {\Bbb Z}^r/\varepsilon$ induces a surjective map on character groups $$ \psi : V(\varepsilon) \rightarrow \chargp{\overline{\Epsilon}/\Epsilon} $$ whose identity fiber $F_1$ equals $V(\overline\varepsilon)$. We need to find a splitting for $\psi$. Write ${{\Bbb Z}^r}/\varepsilon$ as the product $$ {{\Bbb Z}^r}/\varepsilon = {\Bbb Z}^s \times G, $$ where $G$ is finite. Then $G \cong \overline{\Epsilon}/\Epsilon$ and ${\Bbb Z}^s \cong {\Bbb Z}^r/\overline{\varepsilon}$. Thus, there is a surjection ${\Bbb Z}^r/\varepsilon \rightarrow \overline{\Epsilon}/\Epsilon$ whose restriction to $\overline{\Epsilon}/\Epsilon$ is the identity. Let $T : \chargp{\overline{\Epsilon}/\Epsilon}\rightarrow V(\varepsilon)$ be the induced endomorphism of character groups. Then $T$ defines a splitting for $\psi$ and we are done. \qed \smallskip \heading{Notation.} A map $$ \psi : \chargp{H} \rightarrow \chargp{G} $$ between character groups which is induced by a homomorphism $\psi^* : G \rightarrow H$ is called a {\it monomial mapping}, since the induced map on coordinate rings is just the linear extension of $\psi^*$ to a mapping ${\Bbb C}[G] \rightarrow {\Bbb C}[H]$. Given a monomial mapping $\psi : ({\Bbb C}^*)^r \rightarrow \chargp{G}$, let $\varepsilon(\psi) \subset {\Bbb Z}^r$ be the image of the induced map $$ \psi^* : G \rightarrow {\Bbb Z}^r. $$ \smallskip \begin{lemma} Let $\psi : ({\Bbb C}^*)^r \rightarrow \chargp{G}$ be a monomial map and set $\varepsilon = \varepsilon(\psi)$. For each $\mu \in {\mathrm{Tor}}({\mathrm{im}}(\psi))$, there is a $\tau \in \overline{\Epsilon}/\Epsilon$ such that, for some $\eta \in ({\Bbb C}^*)^r$ with ${\mathrm{ord}}(\eta) = {\mathrm{ord}}(\mu){\mathrm{ord}}(\tau)$, $$ \psi^{-1}(\mu) = \eta V(\varepsilon). $$ \end{lemma} \heading{Proof.} If $m = {\mathrm{ord}}(\mu)$, then the fiber $\psi^{-1}(\mu)^m$ lies in the identity fiber $V(\varepsilon)$. Furthermore, if $\eta \in \psi^{-1}(\mu)$ has finite order, then $m$ divides ${\mathrm{ord}} (\eta)$. We have $V(\varepsilon)^m = V(\varepsilon_{m})$, so by Lemma 3.3, $$ V(\varepsilon_{m}) = \bigcup_{\gamma \in T(\overline{\varepsilon}/ {\varepsilon_{m}})} \gamma V(\overline{\varepsilon}). $$ Thus, for any $\eta \in \psi^{-1}(\mu)$, $$ \psi^{-1}(\mu)^m = \eta^m V(\varepsilon)^m = \eta^m V(\varepsilon_{m}) $$ and $\psi^{-1}(\mu)^m$ is a translate of $V(\varepsilon_{m})$ in $V(\varepsilon)$. Thus, $\psi^{-1}(\mu)^m$ must contain an element of $T(V(\overline{\Epsilon}/\Epsilon))$. This means we could have chosen $\eta \in \psi^{-1}(\mu)$ such that $\eta^m \in T(V(\overline{\Epsilon}/\Epsilon))$. If $\varepsilon_{m} = \varepsilon$, then $V(\varepsilon)^m = V(\varepsilon)$ and we could have chosen $\eta$ so that $\eta^m = 1$, which would imply ${\mathrm{ord}}(\eta) = m$. Otherwise, set $\tau = \eta^m$. We will show that ${\mathrm{ord}}(\eta) = {\mathrm{ord}}(\tau) m$. We've seen that $m$ divides ${\mathrm{ord}}(\eta)$. If $k = {\mathrm{ord}}(\eta)/m$, then $\tau^k = \eta^{mk} = 1$, so ${\mathrm{ord}}(\tau)$ divides $k$. Since $\eta^{{{\mathrm{ord}}(\tau)}m} = \tau^{{\mathrm{ord}}(\tau)} = 1$, we have ${\mathrm{ord}}(\eta) = {\mathrm{ord}}(\tau)m$.\qed \begin{corollary} Let $\psi : ({\Bbb C}^*)^r \rightarrow \chargp{G}$ be a monomial map and let $\varepsilon = \varepsilon(\psi)$. Then, for any $\mu \in {\mathrm{im}} (\psi)$ and connected component $Q \subset \psi^{-1}(\mu)$, we have \begin{description} \item{(i)} $Q$ is a translate of $V(\overline{\varepsilon})$, \item{(ii)} $\dim(Q) = r - {\mathrm{rank}}(\overline{\varepsilon})$, and \item{(iii)} if $\mu$ has finite order, then $$ {\mathrm{ord}}(\mu)\ | \ {\mathrm{ord}} (Q) \ | \ {\mathrm{ord}}(\mu)D({\overline{\varepsilon}}/\varepsilon), $$ where $D(\overline{\Epsilon}/\Epsilon)$ is the largest order of any element of $\overline{\Epsilon}/\Epsilon$. \end{description} \end{corollary} We will now describe the dimensions and orders of intersections of rational planes. \begin{proposition} Let $\varepsilon_1,\dots,\varepsilon_k \subset {\Bbb Z}^r$ be any subgroups of ${\Bbb Z}^r$ and let $\eta_1,\dots,\eta_k \in {\mathrm{Tor}}(({\Bbb C}^*)^r)$. Let $Q_1,\dots,Q_k \subset ({\Bbb C}^*)^r$ be defined by $$ Q_i = \eta_i V(\varepsilon_i), \qquad i=1,\dots,k. $$ Let $\varepsilon = \varepsilon_1 + \dots + \varepsilon_k$ and $\eta = (\eta_1,\dots,\eta_k)$. Let $$ \begin{array}{rcl} \rho: \varepsilon_1 \times \dots \times \varepsilon_k &\rightarrow& \varepsilon\\ (\lambda_1,\dots,\lambda_k) &\mapsto& \lambda_1+\dots+\lambda_k \end{array} $$ and let $\gamma : \varepsilon_1 \times \dots \times \varepsilon_k \rightarrow ({\Bbb Z}^r )^k$ be the inclusion map. Then $Q_1 \cap \dots \cap Q_k$ is nonempty if and only if $\gamma^*(\eta) \in {\mathrm{im}}(\rho^*)$ and for any connected component $Q \subset Q_1 \cap \dots \cap Q_k$, we have \begin{description} \item{(i)} $Q$ is a translate of $V(\overline{\varepsilon})$, \item{(ii)} $\dim (Q) = r - {\mathrm{rank}} (\overline\varepsilon)$, and \item{(iii)} if $\eta$ has finite order, then $$ {\mathrm{ord}}(\eta)\ |\ {\mathrm{ord}}(Q) \ | \ {\mathrm{ord}}(\eta)D(\overline{\Epsilon}/\Epsilon). $$ \end{description} \end{proposition} \heading{Proof.} Consider the commutative diagram $$ \cd { &{\Bbb Z}^r/{\varepsilon} &\leftarrow &{{\Bbb Z}^r}/{\varepsilon_1} \times\dots\times {{\Bbb Z}^r}/{\varepsilon_k}\cr &\uparrow{} &&\uparrow{}\cr &{\Bbb Z}^r &\leftarrow &{\Bbb Z}^r \times \dots \times {\Bbb Z}^r \cr &\uparrow{}&&\uparrow{}\cr &\varepsilon &\leftarrow &\varepsilon_1 \times \dots \times \varepsilon_r } $$ where the horizontal arrows are given by adding coordinates, the bottom vertical arrows are inclusions and the top vertical arrow are quotient maps. For $i=1,\dots,k$, let $P_i = V(\varepsilon_i)$. Then we have a commutative diagram of induced maps on the associated algebraic sets: $$ \cd { &V(\varepsilon) &\mapright{} &P_1 \times \dots \times P_k \cr &\downarrow{} &&\downarrow{}\cr &({\Bbb C}^*)^r &\mapright{\alpha} &({\Bbb C}^*)^r\times\dots\times ({\Bbb C}^*)^r\cr &\downarrow{\beta} &&\downarrow{\gamma^*}\cr &\chargp{\varepsilon} &\mapright{\rho^*}&\chargp{\varepsilon_1} \times \dots \times \chargp{\varepsilon_k}.} $$ The preimage $\alpha^{-1}(Q_1\times\dots \times Q_k)$ equals the intersection $Q_1 \cap \dots \cap Q_k$. If $Q \subset Q_1 \cap \dots \cap Q_k$ is a connected component, then it is a connected component of a fiber of $\beta$. The intersection is nonempty if and only if $$ \gamma^*(Q_1\times\dots\times Q_k) \cap {\mathrm{im}}(\rho^*) \neq \emptyset. $$ The claim now follows from Cor. 3.5.\qed \smallskip \section{Formal ${\Bbb Q}$-linear relations.} In this section, we describe ${\Bbb Q}$-linear relations between roots of unity and polar rational polygons following the work of Schoenberg \cite{Sch:Cyc} and Mann \cite{Mann:LinRels}. The aim is to give a constructive method for producing all ${\Bbb Q}$-linear relations and give a bound on the orders of roots of unity satisfying a linear equation in terms of the number of coefficients of the equation. Other work in this area can be found in {\cite{Len:Van}} and \cite{C-J:Trig}. For any $n \in N$, let $\zeta_n = \exp(2 \pi \sqrt{-1}/n)$. For $z\in {\Bbb C}^*$, let $\theta(z) = \arg(z)/{2 \pi}$. Thus, $$ z = |z|\exp(2 \pi \sqrt{-1}\ \theta(z)). $$ A {\it polar rational polygon}, or prp, is an oriented polygon in the complex plane with no parallel edges, such that each edge is a vector with rational length and with angle equal to a rational multiple of $2 \pi$. Then prp's can be put into one-to-one correspondence with formal ${\Bbb Q}$-linear relations of the form \begin{eqnarray} \sum_{i=1}^k a_i \epsilon_i = 0. \end{eqnarray} where $a_i \in {\Bbb Q}^*$ and $$ 0 \leq \theta({\mathrm{sign}}(a_1)\epsilon_1) < \dots < \theta({\mathrm{sign}}(a_k)\epsilon_k) < 1. $$ Given a prp $T$ defined by (2), we will call $a_i\epsilon_i$ the sides of $T$, $|a_i|$ the {\it side lengths} and $\epsilon_i$ the {\it side angles}. The {\it order} of $T$, written ${\mathrm{ord}}(T)$, is the least $n$ such that, for some root of unity $\eta$, we have $$ \eta^n \epsilon^n = 1, $$ for all side angles $\epsilon$ of $T$. The length of $T$, written ${\mathrm len} (T)$ is $k$, the number of sides. Any formal ${\Bbb Q}$-linear equation $$ \sum_{j=1}^k b_j \eta_j $$ can be put in the form (2) by the following: \begin{description} \item{(i)} if $\theta(\eta_j)\ ({\mathrm{mod}}\ 1) \ge 1/2$, then replace $a_j$ by $-a_j$ and $\eta_j$ by $\zeta_2 \eta_j$; \item{(ii)} if $\theta(\eta_j) = \theta(\eta_\ell)$, then replace $b_j \eta_j + b_\ell \eta_\ell$ by $(b_j + b_\ell)\eta_j$; \item{(iii)} remove any summand whose coefficient is zero; and \item{(iv)} reorder the summands. \end{description} It is easy to see that to any formal ${\Bbb Q}$-linear equation there is a unique corresponding equation of the form (2), which can be obtained using the above steps, and hence a unique prp. It follows that the set of prp's forms a commutative ${\Bbb Q}[{\mathrm{Tor}}({\Bbb C}^*)]$-algebra, where ${\mathrm{Tor}}({\Bbb C}^*)$ is the set of roots of unity. One just notes that formal ${\Bbb Q}$-linear equations are closed under addition, multiplication and scalar multiplication by elements of ${\mathrm{Tor}}({\Bbb C}^*)$. The algebra operations on a prp will be the same as those for the corresponding formal ${\Bbb Q}$-linear equation. The resulting ${\Bbb Q}$-linear equation may not be of the form (2), but there is a unique prp associated to it which we take as the result of the operation. The algebra operations have the following geometric interpretations. Multiplication by a positive rational $a \in {\Bbb Q}_+$ scales the corresponding polygon by $a$. Multiplication by any root of unity $\eta \in {\mathrm{Tor}}({\Bbb C}^*)$ rotates the polygon by the argument of $\eta$. For example, multiplying the polygon in figure 1 by $\eta = -1$ yields the polygon rotated by 180 degrees shown in figure 2. $$ \epsffile{neg} $$ Summing is like taking the union except that one needs to reorder the sides and get rid of redundancies. Two prp's will be said to be {\it disjoint} if they do not share any sides angles. A prp $T$ is {\it primitive} if there are no disjoint prp's $S$ and $U$ such that $T = S + U$. Geometrically, a prp is primitive if there is no way to rearrange the edges of the polygon to get a union of polygons joined at vertices. While any prp has a decomposition into a sum of disjoint primitives, this decomposition is not necessarily unique. For example, consider the ${\Bbb Q}$-linear equation given by expanding out \begin{eqnarray} (1 + \zeta_3 + \zeta_3^2)(1 + \zeta_5 + \zeta_5^2 + \zeta_5^3 + \zeta_5^4) = 0. \end{eqnarray} The associated prp is given in figure 3. $$ \epsffile{tot} $$ The prp defined by (3) can be decomposed into primitive prp's as in figure 4 and figure 5. $$ \epsffile{decomp} $$ The decomposition in figure 4 comes from writing (3) as $$ \sum_{i=1}^2 \zeta_3^i\ [1 + \zeta_5 + \zeta_5^2 + \zeta_5^3 + \zeta_5^4] = 0 $$ and figure 5 comes from writing (3) as $$ \sum_{i=1}^4 \zeta_5^i\ [1 + \zeta_3 + \zeta_3^2] = 0. $$ \smallskip The sum of nondisjoint prp's could have smaller length than the sum of the total as we see in the next example. Let $A$ be the prp defined by $$ \zeta_6 + \zeta_6^2 + (-1) = 0 $$ Then $\zeta_5\ A + T_5$ is the prp (see figure 6) given by the following ${\Bbb Q}$-linear equation $$ 1 + (\zeta_6 + \zeta_6^5)\zeta_5 + \zeta_5^2 + \zeta_5^3 + \zeta_5^4 = 0 $$ or $$ 1 + \zeta_{30} + \zeta_{30}^{11} + \zeta_5^2 + \zeta_5^3 + \zeta_5^4 = 0. $$ $$ \epsffile{pent} $$ In general we have the following. \begin{lemma} The lengths of sums satisfies the following inequalities $$ \max\ \{{\mathrm len}(T_1),{\mathrm len}(T_2)\} \leq {\mathrm len}(T_1 + T_2) \leq {\mathrm len} (T_1) + {\mathrm len} (T_2), $$ with equality on the right hand side if $T_1$ and $T_2$ are disjoint and equality on the left hand side if the set of side angles of one prp is contained in the other's. \end{lemma} We will make use of the following types of prp's. For any prime $p$, let $\sigma_p(x)$ be the cyclotomic polynomial $$ \sigma_p(x) = 1 + x + \dots + x^{p-1}. $$ Let $T_p$ be the prp defined by $\sigma_p(\zeta_p) = 0$. Let $n = mp$, where $p$ is a prime not dividing $m$. (For example, if $n$ is prime then $m = 1$.) Then the minimal polynomial $\sigma_{n,p}$ for $\zeta_n$ over ${\Bbb Q}[\zeta_m]$ is given by $$ \sigma_{n,p}(x) = \zeta_m^{a(p-1)} + \zeta_m^{a(p-2)}x + \dots + x^{p-1}, $$ where $a$ is the integer $0 \leq a < m$, such that $ap \equiv 1\ ({\mathrm{mod}}\ m)$. Let $T_{n,p}$ be the prp defined by $\sigma_{n,p}(\zeta_n) = 0$. That is, the prp corresponding to the formal ${\Bbb Q}$-linear relation $$ \sum_{i=0}^{p-1} \zeta_n^{ap(p-1-i) + i} = 0. $$ \begin{lemma} Let $n$ be an integer, $p$ a prime dividing $n$, such that $p^2$ doesn't divide $n$. The prp $T_{n,p}$ is a multiple of $T_p$ by some element of ${\Bbb Q}[{\mathrm{Tor}}({\Bbb C}^*)]$. \end{lemma} \heading{Proof.} Let $m = n/p$ and let $a$ be such that $ap \equiv 1 \ ({\mathrm{mod}}\ m)$. Let $r$ be any integer such that $ap = 1 + mr$. Then, since $\zeta_n^m = \zeta_p$, we have \begin{eqnarray*} \zeta_n^{ap(p-1-i) + i} &=& \zeta_n^{mr(p-1-i) + (p-1)}\\ &=& \zeta_p^{r(p-1-i)} \zeta_n^{p-1}. \end{eqnarray*} Thus, $T_{n,p}$ is defined by the formal linear relation $$ \zeta_n^{p-1}\sum_{i=0}^{p-1} \zeta_p^{r(p-1-i)} = 0. $$ Since $p$ doesn't divide $r$, $r(p-1-i)$ ranges in $0,\dots,p-1$ as $i$ ranges in $0,\dots,p-1$. Therefore, $$ T_{n,p} = \zeta_n^{p-1} T_p. $$ \qed Our proof of the following result uses essentially the same ideas as Mann uses in (\cite{Mann:LinRels}, Theorem 1), except that we take more advantage of the structure of prp's as a ${\Bbb Q}[{\mathrm{Tor}}({\Bbb C}^*)]$-algebra. \begin{proposition} Let $T$ be a primitive prp of length $r$ and order $n$. Then we have \begin{description} \item{(i)} $n$ is square free; \item{(ii)} for any prime $p$ dividing $n$, $p \leq r$; \item{(iii)} for any prime $p$ dividing $n$, $T$ is a ${\Bbb Q}[{\mathrm{Tor}}({\Bbb C}^*)]$-linear combination of $T_{n,p}$ and prp's of order $n/p$. \end{description} \end{proposition} \heading{Proof.} Let \begin{eqnarray} \sum_{i=1}^r a_i \epsilon_i = 0 \end{eqnarray} be the ${\Bbb Q}$-linear relation corresponding to $T$ of the form (2). If $T$ has order $n$, then by multiplying $T$ by an element of ${\mathrm{Tor}}({\Bbb C}^*)$, if necessary, we can assume that all the $\epsilon_i$ in (4) have order $n$. We will show that $n$ is square free and for any prime $p$ dividing $n$, $p \leq r$. Let $p$ be a prime dividing $n$ and let $m = n/p$. Since $\zeta_n^p = \zeta_m$, we can write $$ \epsilon_i = \eta_i\zeta_{n}^{\alpha_i}, $$ for $i=1,\dots,r$, where $\eta_i \in {\Bbb Q}[\zeta_{m}]$ and $0 \leq \alpha_i \leq p-1$. Define $$ q_{T,n,p}(x) = \sum_{i=1}^r a_i \eta_i x^{\alpha_i}. $$ This is a polynomial with coefficients in ${\Bbb Q}[\zeta_{m}]$ satisfied by $\zeta_{n}$ and hence the minimal polynomial $\sigma_{n,p}(x)$ for $\zeta_{n}$ over ${\Bbb Q}[\zeta_{m}]$ divides $q_{T,n,p}(x)$. If $p^2$ divides $n$ then $\deg(\sigma_{n,p}) = p$ which is strictly greater than the degree of $q_{T,n,p}(x)$, and hence $q_{T,n,p}(x)$ is identically zero. Since $T$ is primitive, this means that all the $\alpha_i$ are the same, but then we can multiply $T$ by $\zeta_p^{-1}$ to get a prp of order $n/p$ which is a contradiction, since the order of a prp is preserved under multiplication by roots of unity. Thus, $n$ is square free (proving (i)) and $\sigma_{n,p}(x)$ divides $q_{T,n,p}(x)$. Set $$ A_\alpha = \sum_{\alpha_i = \alpha} a_i \eta_i $$ and write $q_{T,n,p}(x)$ as $$ q_{T,n,p}(x) = \sum_{\alpha = 0}^{p-1} A_\alpha x^\alpha. $$ Since $q_{T,n,p}(x)$ is not identically zero and has degree less than or equal to $p-1$, we have \begin{eqnarray} q_{T,n,p}(x) = B \sigma_{n,p}(x) \end{eqnarray} for some invertible $B \in {\Bbb Q}[\zeta_m]$. By (5), evaluating $B^{-1}\ A_\alpha$ as a complex number gives the $\ell$th coefficient of $\sigma_{n,p}(x)$. Thus, the formal ${\Bbb Q}$-linear equation $$ B^{-1}\ A_\alpha - \zeta_m^{a(p-1-\ell)} = 0 $$ defines a prp, which we'll call $T_\alpha$. Then $T$ has the following decomposition into a sum of prp's $$ T = B\ \sum_{\alpha = 0}^{p-1} T_\alpha + B\ T_{n,p}. $$ Since ${\mathrm{ord}}(T_\alpha) = m$, we have proved (iii). By Lemma 4.1, ${\mathrm len} (T) \ge {\mathrm len} (T_{n,p}) = p$, so $p \leq r$ which proves (ii). \qed Proposition 4.3 leads to a geometric method for generating all prps as we see in the following result, originally proved by Schoenberg in \cite{Sch:Cyc}. \begin{corollary} The set of prp's, considered as a ${\Bbb Q}[{\mathrm{Tor}}({\Bbb C}^*)]$-module, is generated by the set of $T_p$, where $p$ ranges over all primes. \end{corollary} \heading{Proof.} By Lemma 4.2, $T_{n,p}$ is a multiple of $T_p$ by an element of ${\Bbb Q}[\zeta_{n/p}]$. The rest follows from Prop. 4.3 and induction on the number of primes dividing $n$. \qed It follows that geometrically one can construct all prp's by starting with regular $p$-gons, for primes $p$, and doing the operations of rotation, stretching and ``adding", where ``adding" means taking the union of sides, grouping like side angles, getting rid of sides of zero length and reordering the sides, as described above. \section{Behavior of torsion points.} We start in this section with a review of the proof of Theorem 1. Then, using the relation between finitely generated groups and rational planes developed in section 3, we study the degree and period of $p_V$. Finally, we give a proof of Theorem 2. \subsection{Proof of Theorem 1.} Let $V = V(f_1,\dots,f_r) \subset ({\Bbb C}^*)^r$ be any algebraic subset. We want to find a finite number of rational planes $Q_1,\dots,Q_\ell$ contained in $V$ so that $$ \Zar{Tor(V)} = \bigcup_{i=1}^\ell Q_i. $$ \heading{Step 1.} We can reduce to the case of hypersurfaces, since $V = V(f_1) \cap \dots \cap V(f_r)$ and, by Prop. 3.6, a finite intersection of rational planes is a finite union of rational planes. \smallskip \heading{Step 2.} Let $V = V(f)$, where $f \in \Lambda_r$ defined by $$ f = \sum_{\lambda \in {\cal L}(f)} a_\lambda t^\lambda, \quad a_\lambda \in {\Bbb C}^*, $$ for ${\cal L}(f) \subset {\Bbb Z}^r$ is a finite subset. We'll show that $$ \Zar{{\mathrm{Tor}}(V)} = \bigcup_{i=1}^k \psi_i^{-1}(S_i), $$ for some finite collection of monomial maps $\psi_i: ({\Bbb C}^*)^r \rightarrow \chargp{G_i}$, some character groups $\chargp{G_i}$, and some finite sets of torsion elements $S_i \subset {\mathrm{Tor}}(\chargp{G_i})$. It then follows from Cor. 3.5 that $\Zar{{\mathrm{Tor}}(V)}$ is a finite union of rational planes. Let ${\cal L} = {\cal L}(f)$ and let $\ell(f)$ be the linear polynomial $$ \ell(f) = \sum_{\lambda \in {\cal L}} a_\lambda x_\lambda $$ where $x_\lambda$ are independent variables, for $\lambda$ ranging in ${\cal L}$. Then $$ V(f) = \alpha_f^{-1}(V(\ell(f))), $$ where $\alpha_f : ({\Bbb C}^*)^r \rightarrow ({\Bbb C}^*)^{{\cal L}}$ is the monomial map induced by \begin{eqnarray*} {\Bbb Z}^{{\cal L}} &\rightarrow& {\Bbb Z}^r\\ e_\lambda &\mapsto& \lambda. \end{eqnarray*} \smallskip \heading{Step 3.} Let $\Pi_f$ be the set of partitions of ${\cal L}$, where for all $\nu \in {\cal P} \in \Pi_f$, $\nu$ has at least 2 elements. Fix ${\cal P} \in \Pi_f$. Let $L_{{\cal P}}$ be the system of linear equations $$ \ell_\nu(f) = \sum_{\lambda \in \nu}a_{\lambda}t_{\lambda}, $$ for each $\nu \in {\cal P}$. This defines a system of linear equations defined on $$ W({\cal P}) = \prod_{\nu \in {\cal P}} ({\Bbb C}^*)^{\nu}. $$ Let ${\cal T}({\cal P},f)$ be the set of elements $(\epsilon_\lambda)_{\lambda \in \nu, \nu \in {\cal P}} \in W({\cal P})$ satisfying the system $L_{{\cal P}}$: $$ \sum_{\lambda \in \nu} a_\lambda \varepsilon_\lambda = 0, \qquad\mathrm{for all}\ \nu \in {\cal P}. $$ Then $$ V(\ell(f)) = \bigcup_{{\cal P} \in \Pi_f} {\cal T}({\cal P},f). $$ For each $\nu \in {\cal P}$, choose $\lambda_\nu \in \nu$ and let $\nu^* = \nu \setminus \{\lambda_\nu\}$. Define $$ W({\cal P})^* = \prod_{\nu \in {\cal P}} ({\Bbb C}^*)^{\nu^*}. $$ Let $$ \beta_\nu : ({\Bbb C}^*)^\nu \rightarrow ({\Bbb C}^*)^{\nu^*} $$ be the map defined by \begin{eqnarray*} {\Bbb Z}^{\nu^*} &\rightarrow& {\Bbb Z}^\nu\\ e_\lambda &\mapsto& e_\lambda - e_{\lambda_\nu}. \end{eqnarray*} Let $$ \beta_{\cal P} : W({\cal P}) \rightarrow W({\cal P})^* $$ be the product of the maps $\beta_{\nu}$. Let ${\cal V}({\cal P},f)$ be the set of solutions $(\varepsilon_\lambda)_{\lambda \in \nu^*, \nu \in {\cal P}} \in W({\cal P})^*$ to the inhomogeneous equations $$ \sum_{\lambda \in \nu^*} (-a_{\lambda}/{a_{\lambda_\nu}}) t_\lambda = 1, \qquad \nu \in {\cal P}. $$ Then ${\cal T}({\cal P},f) = \beta_{{\cal P}}^{-1}({\cal V}({\cal P},f))$, so we have \begin{eqnarray} V(f) = \bigcup_{{\cal P}\in \Pi_f} \alpha_f^{-1} \beta_{{\cal P}}^{-1}({\cal V}({\cal P},f)). \end{eqnarray} \smallskip \heading{Step 4.} Let $$ \psi_{{\cal P},f} : ({\Bbb C}^*)^r \rightarrow W({\cal P})^* $$ be the composition $\psi_{{\cal P},f} = \beta_{{\cal P}}\circ \alpha_f$. Then $\psi_{{\cal P},f}$ is induced by the maps \begin{eqnarray*} \bigoplus_{\nu \in {\cal P}} {\Bbb Z}^{\nu^*} &\rightarrow& {\Bbb Z}^r\\ e_\lambda &\mapsto& \lambda - \lambda_\nu \quad \mbox{if $\lambda \in \nu$.} \end{eqnarray*} Equation (6) now becomes \begin{eqnarray} V(f) = \bigcup_{{\cal P}\in \Pi_f}\psi_{{\cal P},f}^{-1}({\cal V}({\cal P},f)). \end{eqnarray} Since monomial maps take torsion points to torsion points, (7) implies $$ \Zar{{\mathrm{Tor}}(V(f))} = \bigcup_{{\cal P}\in \Pi_f} \psi_{{\cal P},f}^{-1}({\mathrm{Tor}}({\cal V}({\cal P},f))) $$ If ${\cal P}$ and ${\cal P}'$ are two elements in $\Pi_f$, ${\cal P}'$ is a {\it refinement} of ${\cal P}$ if for any $\nu' \in {\cal P}'$, there is a $\nu \in {\cal P}$ such that $\nu' \subset \nu$. A refinement ${\cal P}'$ of ${\cal P}$ is {\it proper} if ${\cal P}'$ is not equal to ${\cal P}$. If ${\cal P}'$ is a refinement of ${\cal P}$, then $$ {\cal V}({\cal P}',f) \subset {\cal V}({\cal P},f). $$ Let ${\cal V}_m({\cal P},f)$ be the set of elements of ${\cal V}({\cal P},f)$ which are not in ${\cal V}({\cal P}',f)$ for any proper refinement ${\cal P}'$ of ${\cal P}$. Let $S({\cal P},f) = {\mathrm{Tor}}({\cal V}_m({\cal P},f))$. Then we have \begin{eqnarray} \Zar{{\mathrm{Tor}}(V(f))} = \bigcup_{{\cal P} \in \Pi_f} \psi_{{\cal P},f}^{-1}(S({\cal P},f)). \end{eqnarray} Thus, the proof of Theorem 1 is completed by the following lemma. \begin{lemma} (Laurent, Sarnak) The set $S({\cal P},f)$ is finite. \end{lemma} This follows from (\cite{Laur:Equ}, Theorem 1) and (\cite{A-S:Betti}, Lemma 3.1). \qed The rational planes inside a given algebraic subset of $({\Bbb C}^*)^r$ can now be described in the following way. \begin{proposition} Let $V\subset ({\Bbb C}^*)^r$ be any algebraic subset defined by Laurent polynomials ${\cal F} = \{f_1,\dots,f_k\} \subset \Lambda_r$. Let $\Pi_{\cal F} =\Pi_{f_1} \times \dots \times \Pi_{f_k}$. Then \begin{eqnarray} \Zar{{\mathrm{Tor}}(V)} = \bigcup_{{\cal P}_1 \times \dots \times {\cal P}_k \in \Pi_{{\cal F}}}\ \bigcap_{i=1}^k \psi_{{\cal P}_i,f_i}^{-1}(S({\cal P}_i,f)). \end{eqnarray} \end{proposition} \heading{Proof.} We have $$ \Zar{{\mathrm{Tor}}(V)} = \bigcap_{i=1}^k \Zar{{\mathrm{Tor}}(V(f_i))} $$ so the statement follows from (8). \qed \smallskip \subsection{Degree and Periodicity.} We will now give some results concerning the degree and period of $p_V$. First let $f \in \Lambda_r$ be a Laurent polynomial. We will begin by studying the hypersurface $V(f) \subset ({\Bbb C}^*)^r$ defined by $f$. For a partition ${\cal P} \in \Pi_f$, define $$ \delta_{{\cal P},f} = \left \{ \begin{array}{ll} 1 &\mbox{if $S({\cal P},f) \neq \emptyset$,}\\ 0 &\mbox{otherwise.} \end{array}\right . $$ For ${\cal P} \in \Pi_f$, let $\varepsilon({\cal P},f) \subset {\Bbb Z}^r$ be the subset generated by $$ \{ \lambda - \mu : \exists \nu \in {\cal P}, \lambda,\mu \in \nu\} $$ Then $\varepsilon({\cal P},f) = \varepsilon(\psi_{{\cal P},f})$, where $\psi_{{\cal P},f}$ is the map defined in the previous section, so by Cor. 3.5, any connected component of a fiber of $\psi_{{\cal P},f}$ is a translate of $V(\overline{\varepsilon({\cal P},f)})$. For any subset ${\cal U} \subset \Pi_f$, let $\varepsilon({\cal U},f) \subset {\Bbb Z}^r$ be the subgroup generated by $$ \bigcup_{{\cal P} \in {\cal U}} \varepsilon({\cal P},f). $$ Then by Prop. 3.6 any rational plane $Q$ in the intersection $$ \bigcap_{{\cal P} \in {\cal U}} F_{{\cal P}} $$ where each $F_{{\cal P}}$ is a fiber of $\psi_{{\cal P},f}$, is a translate of $V(\overline{\varepsilon({\cal U},f)})$. Let $$ D({\cal U},f) = D(\overline{\varepsilon({\cal U},f)}/{\varepsilon({\cal U},f)}) $$ and let $$ D(f) = \mathop{{\mathrm{lcm}}} \ \{D({\cal U},f) : {\cal U} \subset \Pi_f\}. $$ Let $$ M({\cal P},f) = \max \ (\{{\mathrm{ord}}(x) : x \in S({\cal P},f)\} \cup \{1\}) $$ and let $$ M(f) = \mathop{{\mathrm{lcm}}}_{{\cal P} \in \Pi_f}\ M({\cal P},f). $$ \begin{proposition} Let $f \in \Lambda_r$ any Laurent polynomial and $V = V(f)$. Then any rational plane $Q \subset V(f)$ is a translate of $V(\overline{\varepsilon({\cal P},f)})$, for some partition ${\cal P} \in \Pi_f$. Furthermore, $$ \deg(p_{V(f)}) = \max_{{\cal P} \in \Pi_f}\ (r - {\mathrm{rank}}(\overline{\varepsilon({\cal P},f)}) \delta_{{\cal P},f} $$ and $$ {\mathrm{per}}(p_{V(f)}) \quad | \quad M(f)D(f). $$ \end{proposition} \heading{Proof.} Recall that $$ \Zar{{\mathrm{Tor}}(V(f))} = \bigcup_{{\cal P}\in \Pi_f} \psi_{{\cal P},f}^{-1}(S({\cal P},f)). $$ Thus, the degree of $p_{V(f)}$ is the maximum dimension of rational planes in $\psi_{{\cal P},f}^{-1}(S({\cal P},f))$. Since $\varepsilon(\psi_{{\cal P},f}) = \varepsilon({\cal P},f)$, by Cor. 3.5, if $S({\cal P},f)$ is not empty, then for any rational plane $Q \subset \psi_{{\cal P},f}^{-1}(S({\cal P},f))$, we have $$ \dim(Q) = r - {\mathrm{rank}}(\overline{\varepsilon({\cal P},f)}). $$ This gives the formula for $\deg(p_{V(f)})$. The period of $p_V$ depends on the orders of rational planes in intersections of fibers of $\psi_{{\cal P},f}^{-1}$ over $S({\cal P},f)$. Any such intersection is empty if it involves more than one fiber of $\psi_{{\cal P},f}$, for some ${\cal P}$. Let ${\cal U} \subset \Pi_f$ be any subset such that, for any ${\cal P} \in {\cal U}$, $\delta_{{\cal P},f} = 1$. Choose a connected component $Q_{\cal P} \subset \psi_{{\cal P},f}^{-1}(S({\cal P},f))$ for each ${\cal P} \in{\cal U}$. Let $$ M = \mathop{{\mathrm{lcm}}}_{{\cal P} \in {\cal U}} \ {\mathrm{ord}}(\psi_{{\cal P}}(Q_{{\cal P}})). $$ Then, by Prop. 3.6, for any rational plane $Q$ in the intersection $$ \bigcap_{{\cal P} \in {\cal U}}Q_{{\cal P}}, $$ we have $$ {\mathrm{ord}}(Q) \ | \ M\ D({\cal U},f). $$ Taking the least common multiple of both sides, the claim follows. \qed \smallskip While the various groups denoted by $\varepsilon(-)$ can be computed routinely, the sets $S({\cal P},f)$ need to be studied in a case by case manner. When $f$ is defined over ${\Bbb Q}$ the results described in section 4 are aids to computation. In particular, we have the following result of Mann \cite{Mann:LinRels} (cf. Prop. 4.3). \begin{proposition} (Mann) Let $\epsilon = (\epsilon_1,\dots,\epsilon_r)$ be any finite order maximal solution to a linear equation \begin{eqnarray*} \sum_{i=1}^r a_i \epsilon_i = 1, \end{eqnarray*} with $a_i \in {\Bbb Q}$ for all $i=1,\dots,r$. Let $n$ be the order of $\epsilon$ as an element of $({\Bbb C}^*)^r$. Then $n$ is square free and, if $n=p_1\dots p_k$ is a factorization, then $p_i \leq r+1$ for $i=1,\dots,k$. \end{proposition} As a consequence of Prop. 5.4, we have the following bound on $M({\cal P},f)$. For any partition ${\cal P}$, let $$ R({\cal P}) = \max\ \{|\nu | : \nu \in {\cal P}\}. $$ For any positive integer $R$, let $N[R]$ be the product of distinct primes less than or equal to $R$. \begin{corollary} If $f \in \Lambda_r$ is defined over ${\Bbb Q}$ and has $R$ coefficients, then $$ M({\cal P},f) \ | \ N[R({\cal P})] $$ for any partition ${\cal P} \in \Pi_f$. \end{corollary} We will now give bounds on $p_{V({\cal F})}$, where ${\cal F} = \{f_1,\dots,f_k\} \subset \Lambda_r$ is any finite subset. For any $\pi= ({\cal P}_1,\dots,{\cal P}_k) \in \Pi_{{\cal F}}$, let \begin{eqnarray*} \varepsilon(\pi,{\cal F}) &=& \varepsilon({\cal P}_1,f_1) + \dots + \varepsilon({\cal P}_k,f_k),\\ S(\pi,{\cal F}) &=& S({\cal P}_1,f_1) \times \dots \times S({\cal P}_k,f_k),\\ R(\pi) &=& \max\ \{R({\cal P}_i) : i=1,\dots,k\}, \\ M(\pi,{\cal F}) &=& \mathop{{\mathrm{lcm}}} \ \{{\mathrm{ord}}(x) : x \in S(\pi,{\cal F})\}, \ \mbox{and}\\ M({\cal F}) &=& \mathop{{\mathrm{lcm}}}\ \{M(\pi,{\cal F}) : \pi \in \Pi_{{\cal F}}\}. \end{eqnarray*} For any subset ${\cal U} \subset \Pi_{\cal F}$, define $\varepsilon({\cal U},{\cal F}) \subset {\Bbb Z}^r$ to be the subgroup generated by $$ \bigcup_{\pi \in {\cal U}}\varepsilon(\pi,{\cal F}). $$ Let \begin{eqnarray*} D({\cal U}, {\cal F}) &=& D(\overline{\varepsilon({\cal U},{\cal F})}/{\varepsilon({\cal U},{\cal F})})\\ D({\cal F}) &=& \mathop{{\mathrm{lcm}}}\ \{D({\cal U},{\cal F}) : {\cal U} \subset \Pi_{{\cal F}}\}. \end{eqnarray*} Thus, for any $\lambda \in \overline{\varepsilon({\cal U},{\cal F})}/{\varepsilon({\cal U},{\cal F})}$, we have ${\mathrm{ord}}(\lambda)\ | \ D({\cal F})$. Let $\rho$ be defined by \begin{eqnarray*} \rho: \varepsilon({\cal P}_1,f_1) \times \dots \times \varepsilon({\cal P}_k,f_k) &\rightarrow & \varepsilon({\cal P}_1,f_1) + \dots + \varepsilon({\cal P}_k,f_k)\\ (\lambda_1,\dots,\lambda_k)&\mapsto& \lambda_1 + \dots + \lambda_k \end{eqnarray*} and let $$ \gamma: \varepsilon({\cal P}_1,f_1) \times \dots \times \varepsilon({\cal P}_k,f_k) \hookrightarrow Z^r \times \dots \times {\Bbb Z}^r $$ be the inclusion map. Let $$ \delta_{\pi,{\cal F}} = \left\{\begin{array}{ll} 1 &\quad\mbox{if $\gamma^*(S(\pi,{\cal F})) \cap {\mathrm{im}} (\rho^*) \neq \emptyset$,} \\ 0 &\quad\mbox{otherwise.} \end{array}\right . $$ Prop. 5.3 extends to arbitrary algebraic subsets as follows. \begin{theorem} Let ${\cal F} \subset \Lambda_r$ be a finite subset let and $V = V({\cal F})$. Then any rational plane $Q \subset V$ is a translate of $V(\overline{\varepsilon(\pi,{\cal F})})$ for some partition $\pi \in \Pi_{{\cal F}}$. Furthermore, $$ \deg(p_V) = \max_{\pi \in \Pi_{{\cal F}}} \ (r - {\mathrm{rank}}(\overline{\varepsilon(\pi,{\cal F})})) \delta_{{\cal P},{\cal F}} $$ and $$ {\mathrm{per}}(p_V) \ | \ M({\cal F})\ D({\cal F}). $$ \end{theorem} \heading{Proof.} From (9), we know that $\Zar{{\mathrm{Tor}}(V({\cal F}))}$ is the union over $\pi \in \Pi_{{\cal F}}$ of $$ Q(\pi,{\cal F}) = \bigcap_{i=1}^k \psi_{{\cal P}_i,f}^{-1}(S({\cal P}_i,f)). $$ Hence $\deg(p_V)$ is the maximum dimension of any rational plane $Q \subset Q(\pi,{\cal F})$. By Prop. 3.6, $Q(\pi,{\cal F})$ is nonempty if and only if $\delta_{\pi,{\cal F}} = 1$. Also by Prop. 3.6, any rational plane $Q \subset Q(\pi,{\cal F})$ is a translate of $V(\overline{\varepsilon(\pi,{\cal F})})$ and its dimension equals $r-{\mathrm{rank}}(\overline{\varepsilon(\pi,{\cal F})})$. This gives the formula for the degree of $p_V$. The period ${\mathrm{per}}(p_V)$ is the $\mathop{{\mathrm{lcm}}}$ of ${\mathrm{ord}}(Q)$, where $Q$ is a rational plane in the intersection $\bigcap_{\pi \in {\cal U}} Q_{\pi}$, where $Q_{\pi} \subset Q(\pi,{\cal F})$, is some choice of rational planes and $\pi$ ranges in a subset ${\cal U} \subset \Pi_{\cal F}$. Note that $$ M({\cal U},{\cal F}) = \mathop{{\mathrm{lcm}}}\ \{ M(\pi,{\cal F}) : \pi\in {\cal U}\} $$ divides $M({\cal F})$. By Prop. 3.6, we have $$ {\mathrm{ord}}(Q) \ | \ M({\cal U},{\cal F})\ D({\cal U},{\cal F}) $$ and taking the least common multiple of both sides, we get the bound for the period of $p_{V(f)}$. \qed \subsection{Proof of Theorem 2.} Theorem 2 follows easily from Theorem 3 and Corollary 5.5. To prove (i), we need to find find polynomials defining the rational planes in $V$. Since by Theorem 3, any rational plane $Q$ in $V$ is of the form $Q = \eta V(\varepsilon)$, where $\varepsilon = \overline{\varepsilon(\pi,{\cal F})}$ and $\eta$ is a finite order element of $({\Bbb C}^*)^r$. By Lemma 3.1, $V(\varepsilon)$ is defined by $I_\varepsilon$ and (i) follows. The bound for $\deg(p_V)$ in (ii) comes from ignoring the $\delta_{{\cal P},V}$ in Theorem 3. By Corollary 5.5, we have the inequality $$ M({\cal F}) \ | \ N[R({\cal F})], $$ where $R({\cal F})$ is the maximum number of coefficients of a Laurent polynomial $f \in {\cal F}$. This implies the bound for ${\mathrm{per}}(p_V)$ in (iii). \qed \section{Examples.} In this section we consider some algebraic subsets $V \subset ({\Bbb C}^*)^r$ defined by a finite set of Laurent polynomials ${\cal F} \subset \Lambda_r$ and study the torsion points on $V$ in terms of ${\cal F}$. We start with some notation. For any $m \in {\Bbb N}$, let ${\cal C}_m : {\Bbb N} \rightarrow {\Bbb N}$ be the periodic function defined by $$ {\cal C}_m(n) = \left\{\begin{array}{ll}1 &\quad\mbox{if $m$ divides $n$,}\\ 0 &\quad\mbox{otherwise}\end{array}\right . $$ For any root of unity $\epsilon$, let $\theta(\epsilon) \in {\Bbb Q}/{\Bbb Z}$ be such that $$ \epsilon = \exp(2\pi\ \sqrt {-1} \ \theta(\epsilon)). $$ For any positive integer $n$, denote by $\zeta_{n}$ the primitive $n$th root of unity such that $\theta(\zeta_n) = 1/n$. For any $\theta \in ({\Bbb Q}/{\Bbb Z})^r$, let ${\mathrm{ord}}(\theta)$ be the least positive integer $n$, such that $n\theta = 0 ({\mathrm{mod}}\ 1)$. For any torsion point $\epsilon = (\epsilon_1,\dots,\epsilon_r)$ in $({\Bbb C}^*)^r$ define $$ \theta({\epsilon}) = (\theta(\epsilon_1),\dots,\theta(\epsilon_r)) \ \in \ ({\Bbb Q}/{\Bbb Z})^r. $$ Then $$ {\mathrm{ord}}(\epsilon) = {\mathrm{ord}}(\theta(\epsilon)). $$ A binomial equation \begin{eqnarray} t^\lambda - \zeta = 0 \end{eqnarray} where $\lambda=(\lambda_1,\dots,\lambda_r) \in {\Bbb Z}^r$ and $\zeta$ is a root of unity $\zeta = \zeta_{n}^k$, $0 \leq k \leq n-1$, corresponds to a linear equation \begin{eqnarray} \lambda_1 \theta_1 + \dots + \lambda_r \theta_r = k/n\ ({\mathrm{mod}}\ 1), \end{eqnarray} in the sense that $\epsilon$ satisfies (10) if and only if $\theta(\epsilon)$ is a solution to (11). We'll call equation (11) the {\it exponential form} of equation (10). We'll use the following simplification of Theorem 3 in Examples 2 and 3. \begin{proposition} Suppose $\Pi_{{\cal F}}$ contains only one element $\pi$. Let $$ M({\cal F}) = \mathop{{\mathrm{lcm}}} \ \{{\mathrm{ord}}(x) : x \in S(\pi,{\cal F})\} \cup \{1\} $$ and let $$ D({\cal F}) = D(\overline{\varepsilon(\pi,{\cal F})}/{\varepsilon(\pi,{\cal F})}). $$ If $S(\pi,{\cal F}) = 0$, then $p_{V({\cal F})} = 0$. Otherwise, $$ \deg (p_{V({\cal F})}) = r - {\mathrm{rank}}(\overline{\varepsilon(\pi,{\cal F})}). $$ and $$ M({\cal F}) \ | \ {\mathrm{ord}}(p_{V({\cal F})}) \ |\ M({\cal F})\ D({\cal F}). $$ \end{proposition} \heading{Proof.} We have $$ \Zar{{\mathrm{Tor}}(V)} = \psi_{\pi,{\cal F}}^{-1}(S(\pi,{\cal F})), $$ which is a union of disjoint rational planes contained in fibers over the points in $S(\pi,{\cal F})$. Since $\varepsilon(\pi,{\cal F}) = \varepsilon(\psi_{\pi,{\cal F}})$, the rest follows from Cor. 3.5. \qed \heading{Example 1.} We begin with examples where the degree bound in Theorem 2 is attained. Let $r$ be an even number $r = 2k$. Let $$ f = \sum_{i=1}^r (-1)^i t_i \in \Lambda_r $$ and $V = V(f)$. \smallskip We can see immediately that $V(f)$ contains the affine subtorus $$ P = V(\{t_it_{i+1}^{-1} - 1 : i=2j-1, j=1,\dots,k\}). $$ The dimension of $P$ equals the dimension of solutions $\theta=(\theta_1,\dots,\theta_r) \in ({\Bbb Q}/{\Bbb Z})^r$ such that $$ \theta_{2j-1} - \theta_{2j} = 0\ ({\mathrm{mod}}\ 1), \qquad \mbox{for all $j = 1,\dots,k$}. $$ The dimension is clearly $r - k = k$, so $\deg (p_{V(f)}) \ge k$. Now take any partition ${\cal P} \in \Pi_f$. Then $\varepsilon({\cal P},f) = \overline{\varepsilon({\cal P},f)}$ and $$ {\mathrm{rank}}(\varepsilon({\cal P},f)) = \sum_{\nu \in {\cal P}} (|\nu|-1) \ge k. $$ Therefore, $$ \deg (p_{V(f)}) \leq n - k = k $$ and we have equality. \smallskip \heading{Example 2.} Let $V(f) \in ({\Bbb C}^*)^r$ be defined by $$ f = a_1 t^{\lambda_1} + a_2 t^{\lambda_2} + a_3 t^{\lambda_3}, \qquad a_i \in {\Bbb Q}^* $$ Since there are only three coefficients there is only one parition in $\Pi_f$, namely ${\cal P} = \{1,2,3\}$. We have \begin{eqnarray*} &&\varepsilon = \varepsilon({\cal P},f) = \{\lambda_1 - \lambda_3, \lambda_2 - \lambda_3\}\\ && L_{{\cal P},f} : \frac{-a_1}{a_3}x + \frac{-a_1}{a_2}y - 1\\ \end{eqnarray*} The following is a table of all possible values for $a=-a_1/{a_3}$ and $b=-a_2/a_3$ which give nonempty $S({\cal P},f)$. \bigskip \begin{center} \begin{tabular}{|l|l|}\hline $(a,b)$ & $S({\cal P},f)$\\ \hline $(1,1)$&$(\zeta_{6},\zeta_{6}^{-1}),(\zeta_{6}^{-1},\zeta_{6})$\\ $(1,-1)$&$(\zeta_{6},\zeta_{3}),(\zeta_{6}^{-1},\zeta_{3}^{-1})$\\ $(-1,1)$&$(\zeta_{3},\zeta_{6}),(\zeta_{3}^{-1},\zeta_{6}^{-1})$\\ $(-1,-1)$&$(\zeta_{3},\zeta_{3}^{-1}),(\zeta_{3}^{-1},\zeta_{3})$\\ $(1/2,1/2)$&$(1,1)$\\ $(1/2,-1/2)$&$(1,-1)$\\ $(-1/2,1/2)$&$(-1,1)$\\ $(-1/2,-1/2)$&$(-1,-1)$\\ \hline \end{tabular} \end{center} \bigskip Thus, for example, if $f = t_1 + t_2 + t_3$, then $$ p_{V(f)}(n) = 2n \ {\cal C}_3(n). $$ If ${\cal F} \subset \Lambda_r$ is any finite collection of Laurent polynomials with three coefficients, then $\Pi_{{\cal F}}$ contains a single element $\pi$ so we can use Prop. 6.1. Let $V = V({\cal F})$, ${\cal F} = \{f_1,f_2\} \subset \Lambda_r$, where \begin{eqnarray*} f_1 &=& t_1t_3 + t_4 + \alpha\\ f_2 &=& t_1 + t_2 + \beta t_3 \end{eqnarray*} where $\alpha,\beta \in \{ \pm 1\}$. Then $\Pi_{{\cal F}}$ has a single element $\pi = ({\cal P},{\cal P})$, where ${\cal P} = \{1,2,3\}$. Since \begin{eqnarray*} &&\varepsilon = \varepsilon(\pi,{\cal F}) = \varepsilon({\cal P},f_1) + \varepsilon({\cal P},f_2)\\ &&\quad = \{(1,0,1,0),(0,0,0,1), (1,0,-1,0),(0,1,-1,0)\}\\ &&\overline{\varepsilon} = {\Bbb Z}^4,\\ &&D({\cal F}) = D(\overline{\varepsilon}/\varepsilon) = D({\Bbb Z}/{2{\Bbb Z}}) = 2, \end{eqnarray*} by Prop. 6.1, the bounds on the degree and period of $p_V$ are $$ 0 \leq \deg(p_V) \leq r - {\mathrm{rank}}(\overline{\varepsilon})=4 - 4 = 0\\ $$ which implies $\deg(p_V) = 0$ and $$ M({\cal F}) \ | \ {\mathrm{per}}(p_V) \ | \ 2M({\cal F}). $$ For any $\alpha$ and $\beta$, the exponential form of equations defining $$ \Zar{{\mathrm{Tor}}(V)} = \psi_{\pi,{\cal F}}^{-1}(S(\pi,{\cal F})), $$ are the linear equations \begin{eqnarray*} \theta_1 + \theta_3 &=& c\ ({\mathrm{mod}}\ 1)\\ \theta_4 &=& -c\ ({\mathrm{mod}}\ 1)\\ \theta_1 - \theta_3 &=& d\ ({\mathrm{mod}}\ 1)\\ \theta_2 - \theta_3 &=& -d\ ({\mathrm{mod}}\ 1).\\ \end{eqnarray*} where $c$ and $d$ range in $A \times B$, and depend on $\alpha$ and $\beta$. The following table shows the $p_V$ corresponding to the different choices of $\alpha$ and $\beta$. \begin{center} \begin{tabular}{|l|l|l|l|l|l|}\hline $(\alpha,\beta)$ & $A$ & $B$ & $M({\cal F})$ & $p_V$ & ${\mathrm{per}}(p_V)$\\ \hline $(1,1)$ & $\{1/3,2/3\}$ & $\{1/3,2/3\}$ & $3$ & $4\ {\cal C}_3 + 4\ {\cal C}_6$ &$6$\\ $(1,-1)$ & $\{1/3,2/3\}$ & $\{1/6,5/6\}$ & $6$ & $8\ {\cal C}_{12}$ &$12$\\ $(-1,1)$ & $\{1/6,5/6\}$ & $\{1/3,2/3\}$ & $6$ & $8\ {\cal C}_{12}$ &$12$\\ $(-1,-1)$ & $\{1/6,5/6\}$ & $\{1/6,5/6\}$ & $6$ & $8\ {\cal C}_6$& $6$\\ \hline \end{tabular} \end{center} Note that only in the last example where $(\alpha,\beta) = (-1,-1)$ is the period of $p_V$ strictly less than $M({\cal F})\ D({\cal F})$. We will now justify the entries in the column under $p_V$ in the table. If $\alpha = \beta =1$, we have solutions \bigskip \begin{center} \begin{tabular}{|l|l|}\hline $(c,d)$ & solutions $(\theta_1,\theta_2,\theta_3,\theta_4)$\\ \hline $(1/3,1/3)$&$(1/3,2/3,0,2/3), (5/6,1/6,1/2,2/3)$\\ $(1/3,2/3)$&$(0,2/3,1/3,2/3), (1/2,1/6,5/6,2/3)$\\ $(2/3,1/3)$&$(0,1/3,2/3,1/3), (1/2,5/6,1/6,1/3)$\\ $(2/3,2/3)$&$(2/3,1/3,0,1/3), (1/6,5/6,1/2,1/3)$\\ \hline \end{tabular} \end{center} There are four solutions with order 3 and four with order 6, thus, $$ p_V(n) = \left\{\begin{array}{ll} 8 &\mbox{if $6 \ |\ n$,}\\ 4 &\mbox{if $3 \ |\ n$ and $6 \ \not | \ n$}\\ 0 &\mbox{otherwise} \end{array}\right . $$ and hence $p_V(n) = 4 \ {\cal C}_6(n) + 4 \ {\cal C}_3(n)$. \smallskip If $\alpha = 1$ and $\beta = -1$, we have the solutions: \bigskip \begin{center} \begin{tabular}{|l|l|}\hline $(c,d)$& solutions $(\theta_1,\theta_2,\theta_3,\theta_4)$\\ \hline $(1/3,1/6)$&$(1/4,11/12,1/12,2/3), (3/4,5/12,7/12,2/3)$\\ $(1/3,5/6)$&$(1/12,5/12,1/4,2/3), (7/12,11/12,3/4,2/3)$\\ $(2/3,1/6)$&$(11/12,7/12,3/4,1/3), (5/12,1/12,1/4,1/3)$\\ $(2/3,5/6)$&$(3/4,1/12,11/12,1/3), (1/4,7/12,5/12,1/3)$\\ \hline \end{tabular} \end{center} All solutions have order $12$ so we have $p_V(n) = 8 \ {\cal C}_{12}(n)$. \smallskip If $\alpha = -1$ and $\beta = 1$, we have the solutions: \bigskip \begin{center} \begin{tabular}{|l|l|}\hline $(c,d)$& solutions $(\theta_1,\theta_2,\theta_3,\theta_4)$\\ \hline $(1/6,1/3)$&$(1/4,7/12,11/12,5/6), (3/4,1/12,5/12,5/6)$\\ $(1/6,2/3)$&$(5/12,1/12,3/4,5/6), (11/12,7/12,1/4,5/6)$\\ $(5/6,1/3)$&$(7/12,11/12,1/4,1/6), (1/12,5/12,3/4,1/6)$\\ $(5/6,2/3)$&$(3/4,5/12,1/12,1/6), (1/4,11/12,7/12,1/6)$\\ \hline \end{tabular} \end{center} All solutions have order $12$ so $p_V(n) = 8\ {\cal C}_{12}(n)$. If $a = b = -1$, we have solutions \bigskip \begin{center} \begin{tabular}{|l|l|}\hline $(c,d)$ & solutions $(\theta_1,\theta_2,\theta_3,\theta_4)$\\ \hline $(1/6,1/6)$&$(1/6,5/6,0,5/6), (2/3,1/3,1/2,5/6)$\\ $(1/6,5/6)$&$(0,1/3,1/6,5/6), (1/2,5/6,2/3,5/6)$\\ $(5/6,1/6)$&$(0,2/3,5/6,1/6), (1/2,1/6,1/3,1/6)$\\ $(5/6,5/6)$&$(5/6,1/6,0,1/6), (1/3,2/3,1/2,1/6)$\\ \hline \end{tabular} \end{center} All solutions have order 6, so $p_V(n) = 8 \ {\cal C}_6(n)$. \bigskip \heading{Example 3.} We now study the Fermat curve $V(f)$, defined by $$ f(t_1,t_2) = t_1^m + t_2^m -1. $$ Then \begin{eqnarray*} &&\Zar{{\mathrm{Tor}}(V(f))} = \phi_{{\cal P},f}^{-1}(S({\cal P},f))\\ &&S({\cal P},f) = \{\mu_1,\mu_2\} \qquad \mu_1 = (\zeta_{6},\zeta_{6}^{-1}), \quad \mu_2 = (\zeta_{6}^{-1},\zeta_{6}),\\ &&M(f) = 6,\\ &&\varepsilon = \varepsilon({\cal P},f) = \{(m,0),(0,m)\},\\ &&\overline{\varepsilon} = \{(1,0),(0,1)\} = {\Bbb Z}^2,\\ &&\overline{\varepsilon}/\varepsilon = ({\Bbb Z}/{m{\Bbb Z}})^2,\\ &&D(f) = D(({\Bbb Z}/{m{\Bbb Z}})^2) = m\\ \end{eqnarray*} We thus have $$ \deg(p_{V(f)})= 0 $$ and $$ 6= M(f) \ | \ {{\mathrm{per}}}(p_{V(f)})\ |\ M(f)\ D(f)= 6m. $$ For integers $a,b \in {\Bbb Z}$, let $(a,b)$ denote their greatest common divisor. We'll show that $$ p_V(f) = \left\{\begin{array}{ll} 2(m,n)^2&\mbox{if $6 \ |\ \frac{n}{(n,m)}$,}\\ 0 & \mbox{otherwise.} \end{array}\right . $$ and hence $p_{V(f)}$ has period $6m$. \smallskip For instance, if $m = 3$, $p_V(n) = 18\ C_{18}(n)$. \smallskip The exponential linear equations associated to $\phi_{{\cal P},f}$ are \begin{eqnarray} \left\{\begin{array}{rcl} m\theta_1 &=& 1/6\ ({\mathrm{mod}}\ 1)\\ m\theta_2 &=& 5/6\ ({\mathrm{mod}}\ 1) \end{array} \right . \qquad \left\{\begin{array}{rcl} m\theta_1 &=& 5/6\ ({\mathrm{mod}}\ 1)\\ m\theta_2 &=& 1/6\ ({\mathrm{mod}}\ 1) \end{array} \right . \end{eqnarray} Consider the first equation and suppose there is a solution $$ (\theta_1, \theta_2) = (a/n,b/n) \in ({\Bbb Q}/{\Bbb Z})^2. $$ We will show that $6$ divides $n/{(m,n)}$. The equations in (12) imply that $$ ma/n - 1/6 \in {\Bbb Z} $$ so $6n$ divides $6ma - n$ and hence $6$ divides $n$. Setting $n_1 = n/6$, we have $n$ divides $ma - n_1$. Set $m_1 = m/(m,n)$. Then $$ m_1(m,n)a = n_1\ ({\mathrm{mod}}\ n). $$ Since $(m_1,n) = 1$, $m_1$ is invertible modulo $n$ and there is some $m_2 \in {\Bbb Z}$ such that $m_2m_1 = 1 ({\mathrm{mod}}\ n)$. Thus, $$ (m,n)a = m_2n_1\ ({\mathrm{mod}}\ n) $$ and hence $$ (m,n)a = m_2n_1 + nr, $$ for some integer $r$. Since $n$ is relatively prime to $m_1$ and $m_2$, so is $(m,n)$. Thus, $(m,n)$ divides $n_1$ and hence $6$ divides $n/(m,n)$. Conversely, suppose $6$ divides $n/{(m,n)}$. Let $m_1$ be a representative for the multiplicative inverse of $m/{(m,n)}$ modulo $n$. Then $$ a = \frac{n}{6(m,n)}m_1 \qquad b = \frac{-n}{6(m,n)}m_1 $$ is a solution to the first system and $(b,a)$ is a solution to the second. To find the number of solutions of order $n$. We need to look at the homogeneous equations $$ m(r/n) = 0\ ({\mathrm{mod}}\ 1) \qquad m (s/n) = 0\ ({\mathrm{mod}}\ 1). $$ These have solutions $r,s\in {\Bbb Q}/{\Bbb Z}$ given by $$ r = kn/{(m,n)},\quad s = \ell n/{(m,n)} $$ for $k,\ell=0,\dots,(m,n) -1$. Thus, in total, there are $2(m,n)^2$ possible solutions to (12) of order $n$ when $6$ divides $n/{(m,n)}$. \qed \bibliographystyle{math}
1996-07-24T15:36:56
9607
alg-geom/9607026
en
https://arxiv.org/abs/alg-geom/9607026
[ "alg-geom", "hep-th", "math.AG", "math.QA", "q-alg" ]
alg-geom/9607026
Ashok Raina
Indranil Biswas, A.K. Raina
Projective structures on a Riemann surface
Plain LATEX file, to appear in Int. Math. Res. Not
Int.Math.Res.Not. 15 (1996) 753
null
TIFR/TH/96-17
null
For a compact Riemann surface $X$ of any genus $g$, let $L$denote the line bundle $K_{X\times X}\otimes {\cal O}_{X\times X}(2\Delta)$ on $X\times X$, where $K_{X\times X}$ is the canonical bundle of $X\times X$ and $\Delta$ is the diagonal divisor. We show that $L$ has a canonical trivialisation over the nonreduced divisor $2\Delta$. Our main result is that the space of projective structures on $X$ is canonically identified with the space of all trivialisations of $L$ over $3\Delta$ which restrict to the canonical trivialisation of $L$ over $2\Delta$ mentioned above. We give a direct identification of this definition of a projective structure with a definition of Deligne.We also describe briefly the origin of this work in the study of the so-called "Sugawara form" of the energy-momentum tensor in a conformal quantum field theory.
[ { "version": "v1", "created": "Wed, 24 Jul 1996 15:59:17 GMT" } ]
2008-02-03T00:00:00
[ [ "Biswas", "Indranil", "" ], [ "Raina", "A. K.", "" ] ]
alg-geom
\section{Introduction} A {\it projective structure} (also called a {\it projective connection}) on a Riemann surface is an equivalence class of coverings by holomorphic coordinate charts such that the transition functions are all M\"obius transformations. There are several equivalent notions of a projective structure \cite{D}, \cite{G}. For a compact Riemann surface $X$ of any genus $g$, let $L$ denote the line bundle $K_{X\times X}\otimes {{\cal O}}_{X\times X}(2\Delta)$ on $X\times X$, where $K_{X\times X}$ is the canonical bundle of $X\times X$ and $\Delta$ is the diagonal divisor. This line bundle $L$ is trivialisable over a Zariski open neighborhood of $\Delta$ and has a {\it canonical trivialisation} over the nonreduced divisor $2\Delta$. Our main result [Theorem 3.2] is that the space of projective structures on $X$ is canonically identified with the space of all trivialisations of $L$ over $3\Delta$ which restrict to the canonical trivialisation of $L$ over $2\Delta$ mentioned above. In (\cite{D}, page 31, Definition 5.6 bis) Deligne gave another definition of a projective structure (what he calls ``forme infinit\'esimale''). We give a direct identification of this definition with our definition of a projective structure [Theorem 4.2]. In Section 5, which is independent of the rest of the paper, we describe briefly the origin of this work in the study of the so-called ``Sugawara form" of the energy-momentum tensor in a conformal quantum field theory. \section{Trivialisability of the line bundle $L$} Let $X$ be a compact connected Riemann surface, equivalently, a smooth connected projective curve over ${\Bbb C}$, of genus $g$. We denote by $S$ the complex surface $X\times X$, by $\Delta$ the diagonal divisor of $S$, and by $K_S$ the canonical bundle of $S$. Thus $K_S = p^*_1K_X\otimes p^*_2K_X$, where $p_i$ ($i=1,2$) is the projection of $X\times X$ onto the $i$-th factor and $K_X$ is the canonical bundle of $X$. Let $\sigma $ be the involution of $S$ defined by $(x,y) \longmapsto (y,x)$, of which $\Delta$ is the fixed point set. We note that $\sigma $ has a canonical lift ${\tilde \sigma}$ to $L=K_S\otimes {{\cal O}}_S(2\Delta)$; in other words, ${\tilde \sigma}$ is an isomorphism between $L$ and $\sigma ^*(L)$ with ${\tilde \sigma}\circ {\tilde \sigma}$ being the identity isomorphism. The aim of this section is to establish the following theorem: \noindent {\bf Theorem 2.1.}~ {\it The line bundle $L := K_S\otimes {{\cal O}}_S(2\Delta)$ on $S$ is (a), trivialisable on every infinitesimal neighborhood $n\Delta$ of $\Delta$ in $S$ and (b), has a canonical trivialisation on the first infinitesimal neighborhood $2\Delta$, which is the unique trivialisation of $L$ on $2\Delta$ invariant under the action of ${\tilde \sigma}$ and coinciding with the canonical trivialisation of $L$ on $\Delta$.} We denote by $J^d$ the component of the Picard group of $X$ consisting of all line bundles of degree $d$ and by ${{\cal O}}(\Theta )$ the line bundle on $J^{g-1}$ corresponding to the theta divisor, viz. the reduced theta divisor on $J^{g-1}$ defined by the subset $\{\xi \in J^{g-1}\vert H^0(X , \xi) \neq 0\}$ (when $g =0$, it is the zero divisor). Let $\theta $ denote the natural section of the line bundle ${{\cal O}}(\Theta )$ on $J^{g-1}$ given by the constant function $1$; it vanishes precisely on the theta divisor. For $\xi\in J^{g-1}$ let $\xi^*\Theta $ denote the divisor $\{\zeta\otimes \xi^{-1}\mid \zeta\in\Theta \}\subset J^0$. We now recall \cite{NR} that the linear equivalence class of $\xi^*\Theta + (K\otimes\xi^{-1})^*\Theta $ on $J^0$ is independent of $\xi\in J^{g-1}$ and defines {\it canonically} a line bundle on $J^0$, which we denote by ${{\cal O}}(2\Theta _0)$. We require the following property of the line bundle $L$ in the proof of Theorem 2.1: \noindent {\bf Lemma 2.2.}~ {\it Let $\phi~: ~~S~\longrightarrow ~J^0$ be the morphism defined by $(x,y)\longmapsto {\cal O}_X(x-y)$. Then $$ L~~=~~\phi^*{{\cal O}}(2\Theta_0)\leqno{(2.3)} $$} \noindent {\bf Proof.} Clearly we can write $$ L ~~ =~~ {\cal M}_{\alpha }\otimes {\sigma }^*{\cal M}_{\alpha } \leqno{(2.4)} $$ where, for $\alpha \in J^{g-1}$, the line bundle ${\cal M}_{\alpha }$ on $S$ is defined as follows: $$ {\cal M}_{\alpha } ~ := ~ p^*_1(K_X\otimes {\alpha }^{-1}) \otimes p^*_2({\alpha })\otimes {{\cal O}}_S(\Delta) \leqno{(2.5)} $$ As shown in \cite{R1}, ${\cal M}_\alpha $ is isomorphic to $\phi _\alpha ^*{{\cal O}}(\Theta )$, where $$ {\phi }_{\alpha } ~ : ~~ S ~ \longrightarrow ~ J^{g-1} \leqno{(2.6)} $$ is the morphism defined by $(x,y) \longmapsto \alpha \otimes {{\cal O}}_X(x-y)$. Theorem 2.2 of \cite{BR} is, however, preferable, since it gives, in this special case, a {\it natural} isomorphism between ${\phi }^*_{\alpha }{{\cal O}}(\Theta )$ and ${\cal M}_{\alpha }\otimes {\zeta }_{\alpha }$, where ${\zeta }_{\alpha }$ denotes the trivial line bundle on $S$ with fiber ${\Theta }_{\alpha }$, the fiber of ${{\cal O}}(\Theta )$ at the point $\alpha $. (For any $\alpha $ outside the theta divisor, the nonzero vector ${\theta }(\alpha ) \in {\Theta }_{\alpha }$ identifies ${\zeta }_{\alpha }$ with the trivial line bundle.) Using this in (2.4), we see that Lemma 2.2 follows immediately from the definition of ${{\cal O}}(2\Theta _0)$.$\hfill{\Box}$ \noindent {\it Proof of part (a) of Theorem 2.1.}~ Simply observe that the image ${\phi }(\Delta) = 0\in J^0$. In fact, since ${{\cal O}}(2\Theta _0)$ has no base points, $L$ has a global section which is nowhere zero on $\Delta$.$\hfill{\Box}$ \medskip \noindent {\bf Corollary 2.7.}~ {\it Let $\alpha \in J^{g-1}\setminus\Theta $. Then the section $$ {\omega }_{\alpha } ~~ = ~~ {\phi }^*_{\alpha }\theta \otimes ({\phi }_{\alpha }\circ \sigma )^*\theta ~~ \in ~~ H^0(S,L) $$ is 1 at any point of the diagonal $\Delta$. In particular, this section gives a trivialisation of $L$ over some Zariski open neighborhood of $\Delta$. The existence of ${\omega }_{\alpha }$ implies that $$ \dim H^0(S,L)\, \geq \, \dim H^0(S, K_S) +1 \, = \, \dim H^0(X,K_X)^{\otimes 2} +1 \, = \, g^2+1\leqno{(2.8)} $$} \noindent {\bf Proof.}~ Using the natural trivialisation of ${{\cal O}}(\Theta )$ outside the theta divisor given by the section $\theta $ and the above identification of ${\cal M}_{\alpha }$ with the pullback of ${{\cal O}}(\Theta )$, we have a trivialisation of ${\cal M}_{\alpha }$ over some Zariski open neighborhood of $\Delta$. This gives a trivialisation of ${\sigma }^*{\cal M}_{\alpha }$ over some Zariski open neighborhood of $\Delta$. Now the equality (2.4) completes the proof.$\hfill{\Box}$ \medskip \noindent {\bf Notation:}~ For $n\geq 1$, we shall denote the restriction of $L$ to the divisor $n\Delta$ (the $(n-1)$-th order infinitesimal neighborhood of $\Delta$) by $L\mid n\Delta$. \medskip \noindent {\it Proof of part (b) of Theorem 2.1.}~ Now $L\mid \Delta={{\cal O}}_{\Delta}$ and ${{\cal O}}_{\Delta}$ has a one-dimensional space of sections invariant under the action induced by the involution $\sigma $ on $S$. Hence $L$ has a canonical trivialisation on $\Delta$ defined by the section ``1". The situation on $2\Delta$ is more complicated. We know that $\omega _{\alpha }$ in Corollary 2.7, which defines a trivialisation of $L$ on $2\Delta$, is symmetric under $\sigma $. The claim that $L$ has a {\it canonical} trivialisation on $2\Delta$ will then follow from the following lemma: \noindent {\bf Lemma 2.9.}~ {\it The restriction of $L$ to $ 2\Delta$ has a one-dimensional space of sections invariant under the action induced by the involution $\sigma : (x,y)\mapsto (y,x)$ on $S$.} \noindent {\bf Proof.}~ Consider the exact sequence $$ 0~\longrightarrow K_{\Delta}~\longrightarrow ~L\mid 2\Delta~\longrightarrow ~{{\cal O}}_{\Delta}~\longrightarrow 0 $$ where we have made use of the canonical trivialisation of $L$ on $\Delta$. Now note that the global sections form a short exact sequence. Observe that the natural invariant section ``1" of ${{\cal O}}_{\Delta}$ lifts (by averaging over ${\tilde \sigma}$) to an invariant section of $L\mid 2\Delta$, so that the dimension of the space of invariant sections of the latter is at least one. On the other hand, ${\tilde \sigma}$ operates on $H^0(K_{\Delta})$ as {\it -Id}. Indeed, the tangent space at $(x,x)\in\Delta$ is $T_xX \oplus T_xX$, and it is the direct sum of the subspace spanned by $(v_x,v_x)$ with the subspace spanned by $(v_x,-v_x)$, where $v_x$ is a nonzero vector in $T_xX$. The former are invariant under ${\tilde \sigma}$ and belong to the tangent bundle of $\Delta$, while the latter are anti-invariant under ${\tilde \sigma}$ and belong to the normal bundle of $\Delta$. Now, since $K_{\Delta}$ is the conormal bundle to $\Delta$, the involution ${\tilde \sigma}$ operates as {\it -Id} on $H^0(K_{\Delta})$. Thus we conclude that $K_{\Delta}$ has no nonzero section which is invariant under ${\tilde \sigma}$. This proves the lemma and also completes the proof of Theorem 2.1. What is happening is that, under the quotient map $q:S\rightarrow S/\sigma $, the line bundle $L$ descends to ${\tilde L}$ on $S/\sigma $, since ${\tilde \sigma}$ acts trivially on the fibers of $L$ at each point of the fixed point set $\Delta$ of $\sigma $. The trivialisation of $L$ over $\Delta$ induces a trivialisation of ${\tilde L}$ over $\Delta/\sigma $. Since the scheme-theoretic inverse image $q^{-1}(\Delta /\sigma )$ is $2\Delta$ and $q^*{\tilde L}=L$, the trivialisation of ${\tilde L}$ over $\Delta/\sigma $ induces a trivialisation of $L$ over $2\Delta$.$\hfill{\Box}$ It is useful to have an alternative view of the canonical trivialisation of $L$ on $2\Delta$: {\bf Proposition 2.10.}~{\it The canonical trivialisation of $L$ on $2\Delta$ is given by the unique section of $L\mid 2\Delta$ which restricts to the canonical trivialisation on $\Delta$ and lifts to a global section of $L$.} The proof rests on the following lemma, which shows that the inequality (2.8) is actually an equality. \medskip \noindent {\bf Lemma 2.11.}~~~~~ $\dim H^0(S,L) ~ = ~ g^2+1$. \medskip \noindent {\bf Proof.}~ In view of (2.8), we merely have to establish the upper bound. Indeed, tensoring the following exact sequence of sheaves on $S$ $$ 0 ~ \longrightarrow ~ {{\cal O}}_S(-\Delta) ~ \longrightarrow ~{{\cal O}}_S ~ \longrightarrow ~ {{\cal O}}_{\Delta} ~ \longrightarrow ~ 0 \leqno{(2.12)} $$ by $L$ and passing to cohomology, this follows from the observation that $K_S(\Delta)$ has $g^2$ sections. To establish the latter, tensor $(2.12)$ by $K_S(\Delta)$ and pass to cohomology; it then suffices to show that the injection $H^0(S,K_S)\rightarrow H^0(S,K_S(\Delta))$ is an isomorphism. Taking the direct image of this short exact sequence by the projection, $p_1$, to the first factor of $S$, gives the long exact sequence $$ 0 ~ \longrightarrow ~K_X\otimes {\Bbb C}^g ~\stackrel{\iota}{\longrightarrow} ~ p_{1*}(K_S(\Delta))~\longrightarrow ~K_X~ \longrightarrow ~\cdots $$ of which the first three terms are locally free sheaves. The first two terms are rank $g$ vector bundles and hence $\iota$ must be an isomorphism, which completes the proof. $\hfill{\Box}$ \medskip \noindent {\it Proof of Proposition 2.10.}~ From the exact sequence $$ 0~\longrightarrow ~K_S~\longrightarrow ~L~\longrightarrow ~ L\mid 2\Delta ~\longrightarrow~0 $$ and the fact that $L$ has only $g^2 + 1$ sections, we conclude that the space of sections of $L$ has a one-dimensional image in $L\mid 2\Delta$.$\hfill {\Box}$ \section{Projective structures and the line bundle $L$} We will recall the definition of a {\it projective structure on a Riemann surface subordinate to the complex structure}. This is defined (see \cite{G} page 167) to be a holomorphic coordinate covering, $\{U_i,z_i\}_{i \in I}$, of $X$ such that for any pair $i, j \in I$, the holomorphic transition function $f_{i,j}$ (defined by $z_i = f_{i,j}(z_j)$) is a M\"obius transformation, i.e., a function of the form $$ z ~~ \longmapsto ~~ {{az+b}\over {cz+d}} \leqno{(3.1)} $$ where $a,b,c,d \in {\Bbb C}$ with $ad-bc = 1$. The space ${\mbox{\Large $\wp$}}$ of all projective structures on $X$ subordinate to the complex structure is an affine space for the complex vector space $H^0(X, K^2_X)$ (\cite{G} page 172). The main result of this section is the following theorem: \noindent {\bf Theorem 3.2.}~{\it Let ${\cal Q}$ denote the space of all trivialisations of $L\mid 3\Delta$, which, on restriction to $2\Delta$, give the canonical trivialisation of $L\mid 2\Delta$. Then ${\cal Q}$ is an affine space for the vector space $H^0(X, K_X^2)$, which is canonically isomorphic to the affine space $\mbox{\Large $\wp$}$ of projective structures on $X$.} {\bf Proof.} The obvious exact sequence $$ 0 ~ \longrightarrow ~ K^2_X ~ \longrightarrow ~ L\mid 3\Delta ~ \longrightarrow~ L\mid 2\Delta ~ \longrightarrow ~ 0 \leqno{(3.3)} $$ shows that ${\cal Q}$ is an affine space for the vector space $H^0(X,K^2_X)$. We shall now construct a map from ${\mbox{\Large $\wp$}}$ to ${\cal Q}$. Let $M = CP^1\times CP^1$, and consider the {\it trivial} line bundle $L_M ~ := ~ K_M \otimes {{\cal O}}_M(2{\Delta}_M)$ on $M$, where ${\Delta}_M$ is the diagonal on $M$. Let $$ s ~\in ~ H^0(M,L_M) \leqno{(3.4)} $$ be the trivialisation of $L_M$ whose restriction to ${\Delta}_M$ coincides with the canonical trivialisation given by Theorem 2.1(b). The group of all automorphisms of $CP^1$, namely ${\rm Aut}(CP^1)$, acts naturally on $M$ by the diagonal action; this action lifts to $L_M$. The section $s$ in (3.4) is evidently invariant under the induced action of ${\rm Aut}(CP^1)$ on $H^0(M, L_M)$. This invariance property of the section $s$ immediately implies that if we have a projective structure on $X$, the section $s$ induces a trivialisation of $L$ on some analytic open neighborhood of the diagonal $\Delta$. Now, restricting this trivialisation of $L$ to $3\Delta$ we get an element in $\cal Q$. This gives the required map $$ F~: ~ \mbox{\Large $\wp$} ~ \longrightarrow ~ {\cal Q} \leqno{(3.5)} $$ The above construction of the map $F$ has been motivated by \cite{Bi}. The proof of Theorem 3.2 is now completed by the following lemma which describes how the map $F$ relates the affine structures on $\mbox{\Large $\wp$}$ and ${\cal Q}$. \medskip \noindent {\bf Lemma 3.6.}~ {\it For any ${\cal I}\in {\mbox{\Large $\wp$}}$ and $\gamma \in H^0(X, K^2_X)$, the following equality holds. $$ F({\cal I} +{\gamma }) ~~ =~~ F({\cal I} ) ~+ ~{\gamma \over 6} $$} \medskip \noindent {\bf Proof.}~ Let us first recall how the affine $H^0(X, K^2_X)$ structure of ${\mbox{\Large $\wp$}}$ is defined (\cite{G} page 170, Theorem 19). We start by recalling the definition of the {\it Schwarzian derivative}, denoted by $\mbox{${\cal{S}}$}$, which is the differential operator: $$ \mbox{${\cal{S}}$} (f)(z) ~:= ~ {{2f'(z)f'''(z) - 3(f''(z))^2} \over {2(f'(z))^2}} $$ defined over ${\Bbb C}$. Take any ${\cal I} = \{U_i,z_i\}_{i\in I} \in {\mbox{\Large $\wp$}}$ and $\gamma \in H^0(X,K^2_X)$. On each $U_i$ there is a holomorphic function $h_i$ such that $\gamma = h_i dz_i\otimes dz_i$. For $i \in I$, let $w_i$ be a holomorphic function on $z_i(U_i)$ satisfying the equation $$ h_i ~= ~ \mbox{${\cal{S}}$} (w_i)(z_i) \leqno{(3.7)} $$ Another function $w'_i$ satisfies the equation (3.7) if and only if $w'_i (z_i) = \rho \circ w_i(z_i)$, where $\rho $ is a M\"obius transformation. The element ${\cal I} + \gamma \in {\mbox{\Large $\wp$}}$ is given by $\{U_i, w_i\circ z_i\}_{i \in I}$. (Actually we may have to shrink each $U_i$ a bit so that $w_i\circ z_i$ is a coordinate function.) We require an explicit description of the section $s$ defined in (3.4) in terms of local coordinates. Identify $CP^1$ with ${\Bbb C} \cup \{\infty\}$ and let $(z_1,z_2)$ be the natural coordinates on $M$. In these coordinates the section $s$ can be written as: $$ s_z ~~:= ~~{{dz_1\wedge dz_2}\over {(z_1-z_2)^2}} \leqno{(3.8)} $$ Let $ {\cal I} ~ := ~ \{ U_i,z_i\}_{i\in I}$ be a projective structure on $X$, as before. Take a coordinate chart $(U , z)$ in ${\cal I}$. On the open set $U\times U \subset S$ there is a natural coordinate function $(z_1,z_2)$ obtained from $z$. Now $s_z$ in (3.8) gives a trivialisation of $L$ over $U\times U$. Let $(V,y)$ be another coordinate chart in ${\cal I}$ with $y = (az+b)/(cz+d)$ as in (3.1). This implies that the following identity holds: $$ s_z ~:= ~ {{dz_1\wedge dz_2}\over {(z_1-z_2)^2}} ~~ = ~~ {{dy_1\wedge dy_2}\over {(y_1-y_2)^2}} ~ =: ~ s_y $$ where $(y_1,y_2)$ is the coordinate function on $V\times V$. This equality implies that the two local sections of $L$, viz. $s_z$ and $s_y$, coincide on the intersection $(U\cap V) \times (U\cap V) \subset S$. Thus various local trivialisations of $L$ of the form $s_z$ patch together to give a trivialisation of $L$ on some analytic open neighborhood of $\Delta$. In particular, we get a trivialisation of $L\mid n\Delta$ for any $n$. Since the section $s_z$ takes the value $1$ on $U\times U$ and is invariant under the involution ${\tilde \sigma}$, the trivialisation of $L\mid 2\Delta$ obtained this way is the canonical trivialisation. Thus the trivialisation of $L\mid 3\Delta$ is actually an element of ${\cal Q}$. Evidently, the element of $\cal Q$ obtained in this way coincides with $F({\cal I})$, where $F$ is the map defined in (3.5). Let $(U_i,z_i)$, $i\in I$, be a coordinate chart around $x \in X$, with $z_i(x) = 0$. Assume that $$w_i ~= ~ z_i + \sum_{j=2}^{\infty}a_jz_i^j \leqno{(3.9)} $$ is a solution of (3.7) (we may assume that $w_i$ is of this form since we may compose $w_i$ with any M\"obius transformation). Then equation (3.7) gives $$ h(0) ~=~ 6a_3 -6a^2_2 \leqno{(3.10)} $$ Set $y = w_i\circ z_i$, and define $s_y$ as in (3.8). For ${\tilde x} := (x,x) \in S$, using (3.9) we find that $$ s_y({\tilde x} ) ~=~ s_{z_i}({\tilde x} ) + (a_3 -a^2_2)dz_i\otimes dz_i $$ Comparing this with (3.10) we get that $s_y({\tilde x} ) = s_{z_i}({\tilde x} ) + \gamma (x)/6$. This completes the proof of the lemma and also of Theorem 3.2.$\hfill{\Box}$ \medskip \noindent {\bf Remark 3.11.}~ A consequence of Theorem 3.2 is the following alternative definition of the Schwarzian derivative. Let $f$ be a holomorphic function around $z_0 \in {\Bbb C}$ such that $f'(z_0) \neq 0$. Then the function ${\bar f} := (f,f)$ is a biholomorphism defined on some neighborhood, $U$, of $(z_0,z_0) \in {\Bbb C}\times{\Bbb C}$. Consider the section $s$ defined in (3.4). The restriction of $$ {\hat s} ~ := ~ {\bar f}^*s ~ - ~ s $$ to the (nonreduced) divisor $3{\Delta}_U$ is actually a local section of $K^{2}_{{\Bbb C}}$ around $z_0$. From the computation in the proof of Lemma 3.6 it follows that $\hat s$ is actually ${\mbox{${\cal{S}}$}}(f)(dz)^{\otimes 2}/6$. \medskip \noindent {\bf Remark 3.12.} An interesting question, arising naturally from Theorem 3.2, is whether an element of ${\cal Q}$ comes necessarily from a global section of $L$. Thus let $\Lambda \subset H^0(S,L)$ denote the affine subspace consisting of those sections of $L$ which restrict to the canonical trivialisation on $2\Delta$. Then $\Lambda $ is an affine space for the subspace $H^0(S,K_S)=H^0(X,K_X)^{\otimes 2}$. Associating to any $s \in \Lambda $, the corresponding trivialisation of $L$ over $3\Delta$, we get a map from $\Lambda $ to ${\cal Q}$. Then from Theorem 3.2 we have a (holomorphic) map $\lambda $ from $\Lambda $ to $\mbox{\Large $\wp$}$, the space of all projective structures on $X$. Our question is now whether $\lambda $ is surjective. Let $$ R ~: ~ H^0(X, K_X)^{\otimes 2} ~= ~ H^0(S,K_S) ~\longrightarrow ~ H^0(\Delta , K_S{\vert}_{\Delta}) ~ = ~ H^0(X, K^2_X) \leqno{(3.13)} $$ denote the obvious restriction map. From Lemma 3.6 it follows that for any $s \in \Lambda $ and $\beta \in H^0(X,K_X)^{\otimes 2}$, the equality $$ \lambda (s + \beta ) ~ = ~ \lambda (s) + R(6\beta ) ~\in ~ \mbox{\Large $\wp$} \leqno{(3.14)} $$ holds. This equality implies that $\lambda $ is surjective if and only if the homomorphism $R$ in (3.13) is surjective. From M. Noether's theorem (\cite{ACGH}, page 117) we know that if $X$ is non-hyperelliptic then $R$ is surjective. Moreover, for elements $s,t\in \Lambda $ to have the same image under $\lambda $, we must have $s-t\in H^0(S,K_S(-\Delta))$. A similar argument shows that in the non-hyperelliptic case $\lambda $ remains surjective when restricted to the subspace of $\Lambda $ consisting of sections symmetric under the map ${\tilde \sigma}$ induced from $\sigma :S\rightarrow S ~((x,y)\mapsto (y,x))$. Related observations have been made by Tyurin \cite{T}. \medskip \noindent{\bf Remark 3.15.}~ Let $(X_T, {\Gamma}_T) \longrightarrow T$ be a family of Riemann surfaces with theta characteristic. This means that $\Gamma_T$ is a holomorphic line bundle on $X_T$, the total space of the family of Riemann surfaces, and for any $t\in T$, the restriction ${\Gamma}_t$ to the Riemann surface $X_t$ satisfies the condition that ${\Gamma}^{\otimes 2}_t = K_{X_t}$. Consider the line bundle $$ {\cal M}_T ~ := ~ p^*_1(K_{\rm rel}\otimes {\Gamma}^*_T) \otimes p^*_2({\Gamma}_T)\otimes {{\cal O}}(\Delta_T) $$ on the fiber product $X_T\times_TX_T$, where $p_i$ denote the projection to the $i$-th factor, ${\Delta}_T$ is the diagonal divisor in the fiber product, and $K_{\rm rel}$ is the relative canonical bundle on $X_T$. Let ${\cal M}_t$ be the line bundle on $X_t \times X_t$ obtained by setting $\alpha = \Gamma_t$ in the proof of Lemma 2.2. Clearly the restriction of ${\cal M}_T$ to $X_t\times X_t$ is ${\cal M}_t$. The natural isomorphism between ${\phi}^*_{\alpha }{{\cal O}}(\Theta)$ and ${\cal M}_{\alpha }\otimes {\zeta }_{\alpha }$ mentioned in the proof of Lemma 2.2 shows that the restriction of ${\cal M}_T$ to the diagonal $\Delta_T$ is the trivial line bundle. We may extend this trivialisation to some analytic neighborhood of $\Delta_T$. Now using the equality (2.4) for the given family of Riemann surfaces we get a holomorphic family of trivialisations of the restriction of $L$ to some neighborhood of the diagonal. Using Theorem 3.2 this family of trivialisations equips the family $X_T$ with a holomorphic family of projective structures. Given a family of Riemann surfaces, $X_{T'} \longrightarrow T'$, consider the corresponding family of Riemann surfaces with theta characteristic $$ (X_T ,{\Gamma}_T) ~\longrightarrow ~T $$ where $p : T\longrightarrow T'$ is the finite \'etale Galois cover with the fiber of $p$ over $t \in T'$ being the set of all theta characteristics on the corresponding Riemann surface $X_t$. We earlier saw that there is a holomorphic family of projective structures for the family $X_T \longrightarrow T$. For $x \in T$ let ${\mbox{\Large $\wp$}}_x$ denote the projective structure on the Riemann surface over $x$. For any $t \in T'$ consider the projective structure on $X_t$ given by the average $$ {1\over {\# p^{-1}(t)}}\sum_{x \in p^{-1}(t)} {\mbox{\Large $\wp$}}_x $$ which is defined using the affine space structure on the space of all projective structures on $X_t$. Using this construction we conclude that the family of Riemann surfaces, $X_{T'}$, admits a holomorphic family of projective structures. \section{Relation with Deligne's definition} We shall now recall another definition of a projective structure given in \cite{D} (Definition 5.6 bis). The fibers of the natural projection, $\nu$, of the second order infinitesimal neighborhood of the diagonal $\Delta$ (in $S$) onto $\Delta$ are isomorphic to ${\rm Spec}(R)$, where $R$ is the algebra ${{\Bbb C}}[\epsilon ]/{\epsilon}^3$. Let $P$ denote the principal ${\rm Aut}({\rm Spec}(R))$ bundle on $X$ whose fiber over $x \in X$ is the space of all isomorphisms between ${\rm Spec}(R)$ and the fiber of $\nu$ over $x$. On the other hand, ${\rm Aut}({\rm Spec}(R))$ is same as the group of all automorphisms of $CP^1$ that fix the point $0 \in {\Bbb C} \cup \{\infty\} = CP^1$. Let $P_{tg}$ denote the projective bundle on $X$ associated to $P$. Since ${\rm Aut}({\rm Spec}(R))$ fixes a point in $CP^1$, the bundle $P_{tg}$ has a natural section which we shall denote by $\tau$. There is a natural isomorphism between the second order infinitesimal neighborhood of $\Delta$ and the second order neighborhood of the image of $\tau$ (in $P_{tg}$). \medskip \noindent {\bf Definition 4.1} ``Definition 5.6 bis of \cite{D}''. A projective structure on $X$ is an isomorphism between the third order infinitesimal neighborhood of the diagonal $\Delta$ (in $S$) with the third order infinitesimal neighborhood of $\tau$ (in $P_{tg}$) such that the restriction of this isomorphism to the second order infinitesimal neighborhood of $\Delta$ is the canonical isomorphism with the second order infinitesimal neighborhood of the image of $\tau$ mentioned above. \medskip If $X = CP^1$, the projective line, then $P_{tg} = CP^1\times CP^1$. Thus there is a canonical projective structure on $CP^1$ in the sense of (\cite{D}, Definition 5.6 bis) given by the identity map of of the third order neighborhood of the diagonal in $CP^1 \times CP^1$. Let $\cal H$ denote the sheaf on $X$ which to any open set, $U \subset X$, associates the space of all embeddings of the third order infinitesimal neighborhood of the diagonal of $U\times U$ into the restriction of $P_{tg}$ to $U$ which lift the canonical embedding of the second order neighborhood of the diagonal of $U\times U$. Let us recall \cite{D} that a $K^2_X$-{\it torsor} is a holomorphic fiber bundle over $X$ such that its fiber over any $x\in X$ is equipped with a free, transitive holomorphic action of the fiber $K^2_x$. In other words, the result of the action of a local holomorphic section of $K^2_X$ on a local holomorphic section of the torsor is again a local holomorphic section. Proposition 5.8 (page 32) of \cite{D} says that the sheaf $\cal H$ defined above is a $K^2_X$-torsor. We shall denote the restriction of $L$ to $n\Delta$ by $L(n)$. For an analytic open set $U$ of $X$ let ${\Delta}_U$ be the diagonal divisor on $U\times U$. Let $\cal G$ denote the sheaf on $X$ which to any open set $U \subset X$ associates the space of all trivialisations of the restriction of $L(3)$ to $3\Delta_U$ giving the canonical trivialisation on $2\Delta_U$. From (3.3) it follows that the restriction ${\cal G} (U)$ is an affine space for $H^0(U,K^2_U)$, where $K_U$ is the canonical bundle of $U$. In other words, $\cal G$ is a torsor for the sheaf $K^2_X$. Our aim in this section is to prove the following theorem: \medskip \noindent {\bf Theorem 4.2.}~ {\it The two $K^2_X$-torsors on $X$, namely $\cal G$ and $\cal H$, are canonically isomorphic.} \medskip Theorem 4.2 gives a natural identification of the space ${\cal Q}$, the space of global sections of $\cal G$, with the space of global sections of $\cal H$, which is the space of all projective structures on $X$ in the sense of (\cite{D}, Definition 5.6 bis). \medskip \noindent {\it Proof of Theorem 4.2.}~ We shall prove the theorem by constructing a third $K^2_X$-torsor, $\cal T$, on $X$ and identifying both $\cal G$ and $\cal H$ with $\cal T$. A reason for introducing $\cal T$ as the intermediate step is that its construction might be of some interest. For $n\geq 0$, let $J^n(X)$ denote the sheaf of jets of order $n$ on $X$, which is a vector bundle on $X$ of rank $n+1$. Define ${J}^n_0(X)$ to be the kernel of the obvious projection of $J^n(X)$ onto $J^0(X)$. Note that there is a canonical splitting of the inclusion of ${J}^n_0(X)$ into $J^n(X)$ given by the constant functions. Let $\mbox{${\cal{P}}$} (X)$ denote the subset of the total space of ${J}^3_0(X)$ given by the inverse image of $\{K_X -0\}$ (the set of all nonzero vectors in the total space of $K_X$) under the projection of ${J}^3_0(X)$ onto ${J}^1_0(X) = K_X$. The space $\mbox{${\cal{P}}$} (X)$ admits a natural action (by composition of functions) of the group $M(0)$, the isotropy group of $0 \in {\Bbb C}$ for the M\"obius group action on $CP^1$. The action of $M(0)$ on $\mbox{${\cal{P}}$}(X)$ is free, since the only M\"obius transformation of $CP^1$, which acts as the identity map on the second order neighborhood of a point, is actually the identity transformation (\cite{D}, page 29). Let $\cal T$ denote the quotient of $\mbox{${\cal{P}}$} (X)$ by $M(0)$. A projective structure on $X$ gives maps from neighborhoods of points of $X$ into $CP^1$ which differ only by a M\"obius transformation, and hence gives a section of the obvious projection of $\cal T$ onto $X$. An identification between the space of all sections of $\cal T$ and $\mbox{\Large $\wp$}$, the space of projective structures on $X$, is obtained in this way. We shall now give a $K^2_X$-torsor structure on $\cal T$. Let $f \in {\mbox{${\cal{P}}$}}(X)$ be an element over $x \in X$, and let $v \in (K^2_X)_x$ be an element of the fiber of $K^2_X$ over $x$. Let ${\bar f}$ be a function defined around $x$ which represents $f$. Since $d{\bar f}(x) \neq 0$ (by the definition of ${\mbox{${\cal{P}}$}}(X)$), there is a number $\lambda \in {\Bbb C}$ such that $v = \lambda .d{\bar f}(x)\otimes d{\bar f}(x)$. Consider the function $$ {\bar f}_{\lambda }~ := ~ {\bar f} ~+ ~ \lambda . {\bar f}^3 \leqno{(4.3)} $$ defined around $x$. The element in ${\mbox{${\cal{P}}$}}(X)$ over $x$ represented by the function ${\bar f}_{\lambda }$ clearly does not depend upon the choice of the representative $\bar f$ of $f$. An action of $K^2_X$ on ${\mbox{${\cal{P}}$}}(X)$ is obtained by mapping the pair $(v ,f)$ to the element of ${\mbox{${\cal{P}}$}}(X)_x$ represented by ${\bar f}_{\lambda }$. This is a free (but not transitive) action of the abelian group scheme $K^2_X$ over $X$. This action of $K^2_X$ on ${\mbox{${\cal{P}}$}} (X)$ induces a $K^2_X$-torsor structure on the quotient space $\cal T$ of ${\mbox{${\cal{P}}$}}(X)$. Indeed, this is a consequence of the following fact: let $J^n_0(0)$ be the jets of order $n$ of functions vanishing at $0\in {\Bbb C}$; in this notation, the group $M(0)$ acts freely and transitively on the subset of $J^2_0(0)$ consisting of all elements whose image in $J^1_0(0)$ is nonzero. (If $i$ denotes the isomorphism from the space of all sections of $\cal T$ to $\mbox{\Large $\wp$}$, then $i(A +\gamma ) = i(A)+ 6\gamma $ for any $\gamma \in H^0(X,K^2_X)$.) Theorem 4.2 is a consequence of the assertion that both $\cal G$ and $\cal H$ coincide with this $K^2_X$-torsor $\cal T$. We shall first show that $\cal G$ coincides with $\cal T$. Take any $x \in X$, and let $f \in {\cal T}_x$ be an element of the fiber over $x$. Let $z$ be a function defined in a neighborhood, $U$, of $x$, that represents $f$. Since $d{z}(x) \neq 0$, we may assume that $z$ is a biholomorphism onto its image. Let ${\bar z} = (z,z)$ be the biholomorphism defined on $U\times U$. Pull back the section $s$ (defined in (3.4)) to $U\times U$ using this map $\bar z$. Let $\hat f$ denote the local section of $\cal G$ obtained by restricting this section to the second order infinitesimal neighborhood of the diagonal. The evaluation at $x$, namely ${\hat f}(x)$, depends only on $f$ and not on the representing function $z$. Thus we have a map from $\cal T$ to $\cal G$ which is evidently an isomorphism. We want to check that this isomorphism preserves the $K^2_X$-torsor structures of $\cal T$ and $\cal G$. Take an element $v = \lambda (dz)^{\otimes 2} \in K^2_x$, where $\lambda \in {\Bbb C}$. From the definition of the $K^2_X$-torsor structure on $\cal T$ in (4.3) it follows that the local function $z +\lambda z^3$ represents the result of the action of $v$ on $f$. Remark 3.11 says that the two sections of $\cal G$, represented by $z$ and $z+\lambda z^3$ respectively, differ by $\mbox{${\cal{S}}$} (z + \lambda z^3)(0)/6$. Since $$ \mbox{${\cal{S}}$} (z+\lambda z^3)(0) ~ = ~ 6\lambda $$ the preservation of the $K^2_X$-torsor structures of $\cal T$ and $\cal G$ is established. Next we want to show that $\cal H$ coincides with $\cal T$. Take $x$, $f$ and $z$ as above. We noted earlier that $CP^1$ has a canonical projective structure (in the sense of \cite{D}, Definition 5.6 bis) given by the identity map of the third order neighborhood of the diagonal. This projective structure induces a projective structure on $U$ by the biholomorphism $z$. The evaluation of the section (over $U$) of $\cal H$, thus obtained, at the point $x$, does not depend upon the choice of the representative $z$ of $f$. This gives the required $K^2_X$-torsor structure preserving isomorphism between $\cal T$ and $\cal H$. As an alternative proof of Theorem 4.2 we shall give a direct identification between $\cal G$ and $\cal H$ using coordinate charts. Let $(U,z)$ be a coordinate chart around $x \in X$ with $z(x) = 0$. Using (3.8) we get a section of $\cal G$ over $U$. We shall denote this section as $f_z$. Since the only M\"obius transformation of $CP^1$, which acts as the identity map on the second order neighborhood of a point, is actually the identity map, and the group of M\"obius transformations acts transitively on $CP^1$, there is a natural projective structure on $CP^1$ in the sense of (\cite{D}, Definition 5.6 bis). Since the function $z$ identifies $U$ with an open set in $CP^1$, we get a local section of $\cal H$, which we shall denote by $g_z$. By mapping the section $f_z$ to $g_z$ we get an identification of the restriction ${\cal G} (U)$ with ${\cal H} (U)$ which preserves the torsor structures. We shall show that this identification does not depend upon the choice of the coordinate function $z$. Let $(V,w)$ be another coordinate chart around $x$. Thus $$ w ~ = ~ \sum_{i=0}^{\infty}a_i z^i \leqno{(4.4)} $$ with $a_1 \neq 0$. Let $f_w$ (resp. $g_w$) denote the local section of $\cal G$ (resp. $\cal H$) for $(V,w)$. It is a simple calculation using (4.4) to check that $$ f_w(x)\, - \, f_z(x)~= ~ {{a_1a_3 - a^2_2} \over {a^2_1}} dz\otimes dz ~ = ~ g_w(x) \, - \, g_z(x) $$ This completes the proof of the theorem.$\hfill{\Box}$ \section{Genesis in conformal field theory} In this section we explain how the above definition of projective connection in terms of trivialisations of $L$ on $3\Delta$ came out of some investigations on a model quantum field theory on a curve (see \cite{R1}-\cite{R3}), which give it the intuitive picture of a {\it generalized cross ratio} on a compact Riemann surface, in the limit when all of its arguments are made to coalesce. The application of algebraic geometry to quantum field theory in \cite{R1}-\cite{R3} rests on replacing the study of ``quantum fields", which are not geometric objects, by their so-called ``$n$-point functions" which are hypothesised to be so. Thus in \cite{R1} and \cite{R2} we identified the ``$n$-point functions" of the defining ``quantum fields" of the model with meromorphic sections of certain line bundles on the $n$-fold Cartesian product of the curve $X$. In \cite{R3} we showed how the $n$-point functions of the ``current" $j$, which is a ``regularised product" of the defining fields, could be computed by the use of {\it schemes} having {\it nilpotent elements} to give a precise meaning to the coalescing of arguments involved in the definition of the current. A similar regularised product of currents gives the ``energy-momentum tensor" $T$ of the system, a fact usually expressed by saying that $T$ is in ``Sugawara form". This is a feature of many conformal quantum field theories and plays an important role in the theory of the Virasoro algebra \cite{KR}. The heuristic expectation in (conformal) quantum field theory \cite{BPZ} is that its ``one point function" $<T(z)>$ is a {\it projective connection}. Our study of $<T(z)>$ proceeds from the calculation of the two point function of currents $<j(z)j(w)>$ in \cite{R3}. The salient point is the introduction of the remarkable line bundle ${\cal A}:={\cal O}(D_{12}+D_{34}-D_{14}-D_{23})$ on $X^4:=X_1\times X_2\times X_3\times X_4$, the product of 4 copies of X, where $D_{ij}$ denotes the divisor of $X^4$ defined by the diagonal of $X_i$ and $X_j$. It was pointed out in \cite{R3} that the canonical meromorphic section $1_{\cal A}$, associated with the divisor defining ${\cal A}$, is a natural generalisation to an arbitrary compact, connected Riemann surface of the {\it cross ratio} of 4 points in the complex plane. It was shown in \cite{R3} that the calculation of $<j(z)j(w)>$ requires the trivialisability of ${\cal A}$ on the product scheme $Z:=2\Delta_{13}\times 2\Delta_{24}$, where $\Delta_{ij}$ is the diagonal of $X_i\times X_j$ and $2\Delta_{ij}$ denotes its first infinitesimal neighborhood. \noindent {\bf Proposition 5.1}.~ {\it The line bundle ${\cal A}:={\cal O}(D_{12}+D_{34}-D_{14}-D_{23}) $ is trivialisable on $Z~ := 2\Delta_{13}\times 2\Delta_{24}$ and, moreover, if $\rho\in H^0(Z,{\cal A}\mid Z)$ denotes such a trivialisation, then $$ 1_{\cal A}\mid Z~-~ \rho = \omega_B \leqno{(5.2)} $$ where $\omega_B$ denotes a symmetric meromorphic section of $K_{X\times X}$ with double pole on the diagonal, defined by a holomorphic section of $K_{X\times X}(2\Delta)$ which restricts to $1$ on the diagonal.} As pointed out in \cite{R3}, equation (5.2) can be regarded as the precise algebro-geometric formulation of the following well known formula expressing the meromorphic bidifferential $\omega_B$ in terms of the ``prime form" $E(x,y)$ (see Fay\cite{F}, eqn.(28) p.20): $$ \omega_B(x,y)~=~\frac{\partial^2\ln E(x,y)}{\partial x\partial y}\leqno{(5.3)} $$ In this way it was shown in \cite{R3} that the two point function of currents $<j(z)j(w)>$ is a {\it symmetric meromorphic bidifferential with a double pole on the diagonal}. The computation of the ``one point function" $<T(z)>$ from $<j(z)j(w)>$ now requires that the line bundle $K_{X\times X}(2\Delta)$ should be trivialisable on the {\it second} infinitesimal neighborhood $3\Delta$ of $\Delta$ in $X\times X$ (see \cite{R4} for further details), the validity of which follows from results in \cite{R1}. In this way we arrive at our proposed definition of projective connection, having started with the generalised cross ratio and ended with the coalescing of all of its arguments. The fact that a symmetric meromorphic bidifferential gives rise to a projective connection appears to have been first observed in \cite{HS} (see also \cite{F} p.20, following eqn.(28) cited above). The techniques used in these references, however, do not give a {\it characterisation} of a projective structure, as is provided by Theorem 3.2, nor an understanding of when all projective structures arise in this way, as is provided by Remark 3.12. Moreover, their approach cannot be adapted to the study of $<T(z)>$, for which we require an algebro-geometric approach, which will make possible the study of the {\it higher} point functions as well as other related problems. It also appears to be a fact that \cite{HS} and \cite{F} are inaccessible to most geometers and so we hope that the present treatment clarifies some of these results. The present formulation of the concept of projective connection was announced in several conferences and also in \cite{R4}, where the interested reader will find, in addition, a survey for mathematicians of the papers \cite{R1}, \cite{R2} and \cite{R3}. \medskip \noindent {\bf Acknowledgments:} The authors are very grateful to Prof. M. S. Narasimhan for his useful comments. The first named author is thankful to the Institut Fourier and the Acad\'emie des Sciences, Paris, for their hospitality and support. The second named author thanks the International Centre for Theoretical Physics, Trieste, for its hospitality.
1996-07-19T01:44:05
9607
alg-geom/9607019
en
https://arxiv.org/abs/alg-geom/9607019
[ "alg-geom", "math.AG" ]
alg-geom/9607019
Richard Hain
Richard Hain
The Hodge de Rham theory of relative Malcev completion
36 pages. Author supplied dvi available at http://www.math.duke.edu/faculty/hain/
null
null
null
null
The Hodge de Rham theory of relative Malcev completion is developed in this paper. In the special case where one takes the corresponding reductive group to be trivial, one recovers Chen's de Rham theory of the fundamental group and the corresponding Hodge theory due to Morgan and the author. This work is a principal technical tool in the author's work on the mapping class groups.
[ { "version": "v1", "created": "Thu, 18 Jul 1996 23:41:08 GMT" } ]
2008-02-03T00:00:00
[ [ "Hain", "Richard", "" ] ]
alg-geom
\section{Introduction} Suppose that $\pi$ is an abstract group, that $S$ is a reductive algebraic group defined over a field $F$ of characteristic zero, and that $\rho : \pi \to S(F)$ is a homomorphism with Zariski dense image. The completion of $\pi$ relative to $\rho$ is a proalgebraic group ${\mathcal G}$ which is an extension $$ 1 \to {\mathcal U} \to {\mathcal G} \stackrel{p}{\to} S \to 1 $$ where ${\mathcal U}$ is prounipotent, and a homomorphism $\tilde{\rho} : \pi \to {\mathcal G}(F)$ which lifts $\rho$: $$ \begin{CD} \pi @>{\rho}>> S \cr @V{\tilde{\rho}}VV @| \cr {\mathcal G} @>p>> S \end{CD} $$ It is characterized by the following universal mapping property. If $\phi$ is a homomorphism of $\pi$ to a (pro)algebraic group $G$ over $F$ which is an extension $$ 1 \to U \to G \to S \to 1 $$ of $S$ by a unipotent group $U$, and if the the composite $$ \pi \to G \to S $$ is $\rho$, then there is a unique homomorphism ${\mathcal G} \to G$ of $F$-proalgebraic groups which commutes with the projections to $S$ and through which $\phi$ factors. When $S$ is the trivial group, ${\mathcal G}$ is simply the classical Malcev (or unipotent) completion of $\pi$. In this case, with $F={\mathbb R}$ or ${\mathbb C}$, and $\pi$ the fundamental group of a smooth manifold, there is a de~Rham theorem for ${\mathcal O}({\mathcal G})$ which was proved by K.-T.~Chen \cite{chen}. In these notes we generalize Chen's de~Rham Theorem from the unipotent case to the general case. Our approach is based on the notes \cite{deligne:letter} of Deligne where an approach to computing the Lie algebra of the prounipotent radical of ${\mathcal G}$ via Sullivan's minimal models is sketched. Before explaining our result in general, we recall Chen's de~Rham Theorem in the unipotent case. If $M$ is a smooth manifold and $w_1,\dots, w_r$ are smooth 1-forms on $M$, then Chen defined $$ \int_\gamma w_1\dots w_r = \idotsint\limits_{0\le t_1\le \cdots \le t_r \le 1} f_1(t_1)\dots f_r(t_r)\, dt_1 \dots dt_r $$ where $\gamma : [0,1] \to M$ is a piecewise smooth path and $\gamma^\ast w_j = f_j(t)\, dt$. These are viewed as functions on the path space of $M$. An iterated integral is a linear combination of such functions and the constant function. Fix a base point $x\in M$. Set $\pi = \pi_1(M,x)$. Denote the iterated integrals on the space of loops in $M$ based at $x$ by ${\mathcal I}_x$. Denote by $H^0({\mathcal I}_x)$ those elements of ${\mathcal I}_x$ whose value on a loop depends only on its homotopy class. Then Chen's $\pi_1$ de~Rham Theorem asserts that integration induces a Hopf algebra isomorphism $$ {\mathcal O}({\mathcal U}) \cong H^0({\mathcal I}_x) $$ where ${\mathcal U}$ denotes the real points of the unipotent completion of $\pi$ and ${\mathcal O}({\mathcal U})$ its coordinate ring. Another important ingredient of Chen's theorem is that it gives an algebraic description of ${\mathcal I}_x$ and $H^0({\mathcal I}_x)$ in terms of the (reduced) bar construction on the de~Rham complex of $M$ and the augmentation induced by the base point. In this paper we generalize the definition of iterated integrals and prove a more general de~Rham theorem in which the Hopf algebra ${\mathcal O}({\mathcal G})$ of functions on the completion of $\pi_1(M,x)$ relative to a homomorphism $\rho : \pi_1(M,x) \to S$ is isomorphic to a Hopf algebra of ``locally constant iterated integrals,'' defined algebraically in terms of a suitable (2-sided) bar construction on a complex $\Efin^{\bullet}(M,{\mathcal O}(P))$. This complex of forms plays a central role in all our constructions and was introduced by Deligne in his notes \cite{deligne:letter}, the main result of which is that the pronilpotent Lie algebra associated to the 1-minimal model of $\Efin^{\bullet}(M,{\mathcal O}(P))$ is the Lie algebra of the prounipotent radical ${\mathcal U}$ of ${\mathcal G}$. In Section~\ref{groupoid} we define the completion of the fundamental groupoid of a manifold $M$ with respect to the representation $\rho$. This is a category (in fact, a groupoid) whose objects are the points of $X$ and where the Hom sets are proalgebraic varieties; the automorphism of the object $x\in M$ is the completion of $\pi_1(M,x)$ relative to $\rho$. There is a canonical functor of the fundamental groupoid of $M$ to this category. We give a de~Rham description of the coordinate ring of each Hom variety in terms of a suitable 2-sided bar construction on $\Efin^{\bullet}(M,{\mathcal O}(P))$ and of the functor from the fundamental groupoid to its relative completion using iterated integrals. One of the main applications of Chen's $\pi_1$ de~Rham Theorem is to give a direct construction of Morgan's mixed Hodge structure \cite{morgan} on the unipotent completion of the fundamental group of a pointed complex algebraic variety as is explained in \cite{hain:geom}. In this paper we prove that if $X$ is a smooth complex algebraic variety (or the complement of a normal crossings divisor in a compact K\"ahler manifold) and ${\mathbb V} \to X$ is an admissible variation of Hodge structure with polarization $\langle\phantom{x},\phantom{x}\rangle$ whose monodromy representation $$ \rho : \pi_1(X,x) \to S := \Aut(V_x,\langle\phantom{x},\phantom{x}\rangle) $$ has Zariski dense image\footnote{The assumption that the monodromy have Zariski dense monodromy can probably be removed. What one needs to know is that the Zariski closure of the image of $\rho$ is reductive and that its coordinate ring has a natural real Hodge structure --- see Remark~\ref{extended}. This should follow from the work of Simpson and Corlette as each of them has pointed out.}, then the coordinate ring ${\mathcal O}({\mathcal G})$ of the completion of $\pi_1(X,x)$ relative to $\rho$ has a natural mixed Hodge structure. More generally, we show that the coordinate rings of the Hom sets of the relative completion of the fundamental groupoid of $X$ with respect to $\rho$ have canonical mixed Hodge structures. Our principal application of the Hodge theorem for relative completion appears in \cite{hain:torelli} where we use it to prove that the unipotent completion of each Torelli group (genus $\neq 2$) has a canonical mixed Hodge structure given the choice of a smooth projective curve of genus $g$. Another application suggested by Ludmil Kartzarkov, and proved in Section~\ref{hodge_str}, is a generalization of the theorem of Deligne-Griffiths-Morgan-Sullivan (DGMS) on fundamental groups of compact K\"ahler manifolds: If $X$ is a compact K\"ahler manifold and ${\mathbb V}\to X$ is a polarized variation of Hodge structure with Zariski dense monodromy, then the prounipotent radical of the completion of $\pi_1(X,x)$ relative to the monodromy representation has a presentation with only quadratic relations. The theorem of DGMS is recovered by taking ${\mathbb V}$ to be the trivial variation ${\mathbb Q}_X$. In Section~\ref{connection} we show that if $X$ is a smooth variety and ${\mathbb V}$ is an admissible variation of Hodge structure over $X$ with Zariski dense monodromy representation $\rho$, then there is a canonical integrable 1-form $$ \omega \in E^1(X')\comptensor \Gr^W_{\bullet} {\mathfrak u} $$ where $X'$ is the Galois covering of $X$ with Galois group $\im\rho$, and ${\mathfrak u}$ the Lie algebra of the prounipotent radical ${\mathcal U}$ of the completion ${\mathcal G}$ of $\pi_1(X,x)$ with respect to $\rho$. This form is $\im \rho$ invariant under the natural actions of $\im \rho$ on $X'$ and ${\mathfrak u}$. It can be integrated to the canonical representation $$ \tilde{\rho} : \pi_1(X,x) \to S\ltimes {\mathcal U} \cong {\mathcal G}. $$ In the particular case where $X$ is the complement of the discriminant locus in ${\mathbb C}^n$, where $\pi_1(X,x)$ is the braid group$B_n$ and $S$ the symmetric group, this connection is the standard one $$ \omega = \sum_{i<j} d\log(x_i-x_j)\, X_{ij} $$ on $X'$, the complement in ${\mathbb C}^n$ of the hyperplanes $x_i=x_j$. Kohno \cite{kohno} used the $\Sigma_n$ invariant form $\omega$ and finite dimensional representations of $\Gr^W_{\bullet} {\mathfrak u}$ to construct Jones's representations of $B_n$. Our construction is used in \cite{hain:torelli} to construct an analogous ``universal projectively flat connection'' for the mapping class groups in genus $\ge 3$. I am very grateful to Professor Deligne for sharing his notes on the de~Rham theory of relative completion with me and for his interest in this work. I would also like to thank M.~Saito for explaining some of his work to me, and Hiroaki Nakamura for his careful reading the manuscript and his many useful comments. I'd also like to thank Kevin Corlette and Carlos Simpson for freely sharing their ideas on (\ref{extended}). The bulk of this paper was written when I was visiting Paris in spring 1995. I would like to thank the Institute Henri Poincar\'e and the Institute des Hautes \'Etudes Scientifiques for their generous hospitality and support. \section{Conventions} Here, to avoid confusion later on, we make explicit our basic conventions and review some basic constructions that depend, so some extent, on these conventions. Throughout these notes, $X$ will be a connected smooth manifold. By a path in $X$ from $x\in X$ to $y\in Y$, we shall mean a piecewise smooth map $\alpha : [0,1] \to X$ with $\alpha(0)=x$ and $\alpha(1) =y$. The set of all paths in $X$ will be denoted by $PX$. There is a natural projection $PX \to X\times X$; it takes $\alpha$ to its endpoints $(\alpha(0),\alpha(1))$. The fiber of this map over $(x,y)$ will be denoted by $P_{x,y}X$, and the inverse image of $\{x\}\times X$ will be denoted by $P_{x,-}$. The sets $PX$, $P_{x,y}X$, $P_{x,-}X$, each endowed with the compact-open topology, are topological spaces. We shall multiply paths in their natural order, as distinct from the functional order. That is, if $\alpha$ and $\beta$ are two paths in $X$ with $\alpha(1) = \beta(0)$, then the path $\alpha\beta$ is defined and is the path obtained by first traversing $\alpha$, and then $\beta$. Suppose that $(\widetilde{X},\tilde{x}_o) \to (X,x_o)$ is a pointed universal covering of $X$. With our path multiplication convention, $\pi_1(X,x_o)$ acts on the {\em left} of $\widetilde{X}$. One way to see this clearly is to note that there is a natural bijection $$ \coprod_{y\in X}\pi_0(P_{x_o,y}X) \to \widetilde{X}. $$ This bijection is constructed by taking the homotopy class of the path $\alpha$ in $X$ that starts at $x_o$ to the endpoint $\tilde{\alpha}(1)$ of the unique lift $\tilde{\alpha}$ of $\alpha$ to $\widetilde{X}$ that starts at $\tilde{x}_o$. With respect to this identification, the action of $\pi_1(X,x_o)$ is by left multiplication. Another consequence of our path multiplication convention is that $\pi_1(X,x_o)$ naturally acts on the {\em right} of the fiber over $x_o$ of a flat bundle over $X$, as can be seen from an elementary computation. Conversely, if $$ \rho : F\times \pi_1(X,x_o) \to F $$ is a right action of $\pi_1(X,x_o)$ on $F$, then one can define $F\times_\rho \widetilde{X}$ to be the quotient space $F\times \widetilde{X}/\sim$, where the equivalence relation is defined by $$ (f,gx) \sim (fg,x) $$ for all $g \in \pi_1(X,x_o)$. This bundle has a natural flat structure --- namely the one induced by the trivial flat structure on the bundle $F \times \widetilde{X} \to \widetilde{X}$. The composite $$ F \cong F \times \{\tilde{x}_o\} \hookrightarrow F \times \widetilde{X} \to F\times_\rho\widetilde{X} $$ gives a natural identification of the fiber over $x_o$ with $F$. With respect to this identification, the monodromy representation of the flat bundle $F\times_\rho \widetilde{X} \to X$ is $\rho$. Of course, left actions can be converted into right actions by using inverses. Presented with a natural left action of $\pi_1(X,x_o)$ on a space, we will convert it, in this manner, into a right action in order to form the associated flat bundle. The flat bundle over $X$ corresponding to the right $\pi_1(X,x_o)$-module $V$ will be denoted by ${\mathbb V}$. For a flat vector bundle ${\mathbb V}$ over $X$, we shall denote the complex of smooth forms with coefficients in the corresponding $C^\infty$ vector bundle by $E^{\bullet}(X,{\mathbb V})$. This is a complex whose cohomology is naturally isomorphic to $H^{\bullet}(X,{\mathbb V})$. In particular, the $C^\infty$ de~Rham complex of $X$ will be denoted by $E^{\bullet}(X)$. By definition, mixed Hodge structures (MHSs) are usually finite dimensional. When studying MHSs on completions of fundamental groups, one encounters two kinds of infinite dimensional MHSs $$ ((V_{\mathbb R},W_{\bullet}),(V_{\mathbb C},W_{\bullet},F^{\bullet})). $$ In both cases, the weight graded quotients are finite dimensional. In one, the weight filtration is bounded below (i.e. $W_lV=0$, for some $l$) so that each $W_mV$ is finite dimensional. In this case we require that each $W_mV$ with the induced filtrations be a finite dimensional MHS in the usual sense. The other case is dual. Here the weight filtration is bounded above (i.e., $V=W_lV$ for some $l$). In this case, each $V/W_mV$ is finite dimensional. We require that $V$ be complete in the topology defined by the weight filtration (i.e., $V$ is the inverse limit of the $V/W_mV$), that each part of the Hodge filtration be closed in $V$, and that each $V/W_mV$ with the induced filtrations be a finite dimensional MHS in the usual sense. Such mixed Hodge structures form an abelian category, as is easily verified. Finally, if $V^{\bullet}$ is a graded module and $r$ is a integer, $V[r]^{\bullet}$ denotes the graded module with $$ V[r]^n = V^{r+n}. $$ \section{The Coordinate Ring of a Reductive Linear Algebraic Group} \label{coord} Suppose that $S$ is a reductive linear algebraic group over a field $F$ of characteristic zero. The right and left actions of $S$ on itself induce commuting left and right actions of $S$ on its coordinate ring ${\mathcal O}(S)$. If $V$ is a right $S$ module, its dual $V^\ast := \Hom_F(V,F)$ is a left $S$ module via the action $$ (s \cdot \phi) (v) := \phi(v\cdot s), $$ where $s\in S$, $\phi \in \Hom_F(V,F)$ and $v\in V$. The following result generalizes to reductive groups a well known fact about the group ring of a finite group. \begin{proposition}\label{decomp} If $\left(V_\alpha\right)_\alpha$ is a set of representatives of the isomorphism classes of irreducible right $S$-modules, then, as an $(S,S)$ bimodule, ${\mathcal O}(S)$ is canonically isomorphic to $$ \bigoplus_\alpha V_\alpha^\ast\boxtimes V_\alpha. $$ \end{proposition} \begin{proof} This follows from the following facts: \begin{enumerate} \item If $V$ is an $S$ module, then the set of matrix entries of $V$ is the dual $(\End V)^\ast$ of $\End V$. It has commuting right and left $S$ actions. The right action is induced by left multiplication of $S$ on itself by left translation, and the left action by the right action of $S$ on itself. \item As a vector space, $(\End_F V)^\ast$ is naturally isomorphic to $V^\ast\otimes V$. The isomorphism takes $\phi\otimes v\in V^\ast \otimes V$ to the matrix entry $$ \{f:V \to V\} \mapsto \left\{F \stackrel{v}{\to} V \stackrel{f}{\to} V \stackrel{\phi}{\to} F \right\}. $$ It is easily checked that this isomorphism gives an isomorphism $(\End V)^\ast \cong V^\ast \boxtimes V$ of $(S,S)$-bimodules. \item By standard arguments (cf.\ \cite{cartier}), the fact that $S$ is reductive implies that the subspace of ${\mathcal O}(S)$ spanned by the matrix entries of all irreducible linear representations is a subalgebra of ${\mathcal O}(S)$. That is, the image of the linear map $$ \Phi : \sum_\alpha V_\alpha^\ast \boxtimes V_\alpha \to {\mathcal O}(S) $$ is a subalgebra of ${\mathcal O}(S)$. Since $\Phi$ is $S\times S$ equivariant, and since the $V_\alpha^\ast \boxtimes V_\alpha$ are pairwise non-isomorphic irreducible representations of $S\times S$, $\Phi$ is injective. \item Since $S$ is linear, it has a faithful linear representation $V_0$, say and ${\mathcal O}(S)$ is generated by the matrix entries of $V_0$. It follows that $\Phi$ is surjective, and therefore an algebra isomorphism. \end{enumerate} \hfill \end{proof} Recall that if $G$ is an affine algebraic group, then the Lie algebra ${\mathfrak g}$ of $G$ can be recovered from ${\mathcal O}(G)$ as follows: Denote the maximal ideal in ${\mathcal O}(G)$ of functions that vanish at the identity by ${\mathfrak m}$. Then, as a vector space, ${\mathfrak g}$ is isomorphic to the dual ${\mathfrak m}/{\mathfrak m}^2$ of the Zariski tangent space of $G$ at the identity. The bracket is induced by the comultiplication $$ \Delta : {\mathcal O}(G) \to {\mathcal O}(G)\otimes {\mathcal O}(G) $$ as we shall now explain. Evaluation at the identity and inclusion of scalars give linear maps ${\mathcal O}(G) \to k$ and $k\to {\mathcal O}(G)$. There is therefore a canonical isomorphism $$ {\mathcal O}(G) \cong k \oplus {\mathfrak m}. $$ Using this decomposition, we see that the diagonal induces a diagonal map $$ \overline{\Delta} : {\mathfrak m} \to {\mathfrak m}\otimes {\mathfrak m}. $$ Denote the involution $f\otimes g \mapsto g\otimes f$ of ${\mathfrak m} \otimes {\mathfrak m}$ by $\tau$. The map $$ \overline{\Delta} - \tau \circ \overline{\Delta} : {\mathfrak m} \to {\mathfrak m}\otimes {\mathfrak m} $$ induces the map $$ \Delta^c : {\mathfrak m}/{\mathfrak m}^2 \to {\mathfrak m}/{\mathfrak m}^2\otimes {\mathfrak m}/{\mathfrak m}^2 $$ dual to the bracket. \section{A Basic Construction} {}From this point on $S$ will be a linear algebraic group defined over ${\mathbb R}$. We will abuse notation and also denote its group of real points by $S$. We will assume now that we have a representation $$ \rho : \pi_1(X,x_o) \to S $$ whose image is Zariski dense. We will fix a set of representatives $\left(V_\alpha\right)_\alpha$ of the isomorphism classes of rational representations of $S$. Composing $\rho$ with the action of $S$ on itself by {\it right} multiplication, we obtain a right action of $\pi_1(X,x_o)$ on $S$. Denote the corresponding flat bundle by $$ p: P \to X. $$ This is a left principal $S$ bundle whose fiber $p^{-1}(x_o)$ over $x_o$ comes with an identification with $S$ ; the $S$ action and the marking of $p^{-1}(x_o)$ are induced by the obvious left action of $S$ on $S \times \widetilde{X}$ and by the composite $$ S \cong S \times \{x_o\} \hookrightarrow S \times \widetilde{X} \to P. $$ The point $\tilde{x}_o$ of $p^{-1}(x_o)$ corresponding to $1 \in S$ will be used as a basepoint of $P$. Each rational representation of $S$ gives rise to a representation of $\pi_1(X,x_o)$, and therefore to a local system over $X$. We shall call such a local system a {\it rational local system}. The action of $\pi_1(X,x_o)$ on $S$ by right multiplication induces a left action of $S$ on ${\mathcal O}(S)$, the coordinate ring of $S$. Convert this to a right action using inverses: $$ (f\gamma)(s) = f(s\gamma^{-1}), $$ where $f\in {\mathcal O}(S)$, $\gamma\in \pi_1(X,x_o)$, and $s\in S$. Denote the associated flat bundle by $$ {\mathcal O}(P) \to X. $$ This is naturally a {\em right} flat principal $S$ bundle over $X$. It follows from (\ref{decomp}) that it is the direct sum of its rational sub-local systems: \begin{equation}\label{decomp2} {\mathcal O}(P) = \bigoplus_\alpha {\mathbb V}_\alpha^\ast \otimes V_\alpha. \end{equation} In particular, it is the direct limit of its rational sub-local systems. Define $$ \Efin^{\bullet}(X,{\mathcal O}(P)) = \lim_\to E^{\bullet}(X,{\mathbb M}), $$ where ${\mathbb M}$ ranges over the rational sub-local systems of ${\mathcal O}(P)$. Denote the cohomology $$ \lim_\to H^{\bullet}(X,{\mathbb M}) $$ of this complex by $H^{\bullet}(X,{\mathcal O}(P))$. The right action of $S$ on ${\mathcal O}(P)$ induces a right action of $S$ on $$ H^{\bullet}(X,{\mathcal O}(P)). $$ {}From (\ref{decomp2}), it follows that there is a natural isomorphism $$ \Efin^{\bullet}(X,{\mathcal O}(P)) \cong \bigoplus_\alpha E^{\bullet}(X,{\mathbb V}_\alpha^\ast)\otimes V_\alpha $$ of right $S$ modules. The following result is an immediate consequence. \begin{proposition}\label{isom} For each irreducible representation $V$ of $S$, there is a natural isomorphism $$ \left[ H^k(X,{\mathcal O}(P))\otimes V\right]^S \cong H^k(X,{\mathbb V}). \qed $$ \end{proposition} The bundle $P\to X$ is foliated by its locally flat sections. Denote this foliation by ${\mathcal F}$. We view it as a sub-bundle of $TP$, the tangent bundle of $P$. Denote by $E^k(P,{\mathcal F})$ the vector space consisting of $C^\infty$ sections of the dual of the bundle $$ \Lambda^k {\mathcal F} \to P. $$ One can differentiate sections along the leaves to obtain an exterior derivative map $$ d : E^k(P,{\mathcal F}) \to E^{k+1}(P,{\mathcal F}). $$ With this differential, $E^{\bullet}(P,{\mathcal F})$ is a differential graded algebra. Moreover, the left action of $S$ on $P$ induces a natural right action of $S$ on it, and the natural restriction map \begin{equation}\label{res} E^{\bullet}(P) \to E^{\bullet}(P,{\mathcal F}) \end{equation} is an $S$-equivariant homomorphism of differential graded algebras. The base point $\tilde{x}_o\in P$ induces augmentations $$ \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathbb R}\text{ and } E^{\bullet}(P,{\mathcal F}) \to {\mathbb R}. $$ \begin{proposition}\label{dga_homom} There is a natural, augmentation preserving d.g.\ algebra homomorphism $$ \Efin^{\bullet}(X,{\mathcal O}(P)) \to E^{\bullet}(P,{\mathcal F}) $$ which is injective and $S$-equivariant with respect to the natural right $S$ actions.\hfill \qed \end{proposition} \section{Iterated Integrals and Monodromy of Flat Bundles} Consider the category ${\mathcal B}(X,S)$ whose objects are flat vector bundles ${\mathbb V}$ over $X$ that admit a finite filtration $$ {\mathbb V} = {\mathbb V}^0 \supset {\mathbb V}^1 \supset {\mathbb V}^2 \supset \cdots $$ by sub-local systems with the properties: \begin{enumerate} \item the intersection of the ${\mathbb V}^i$ is trivial; \item each graded quotient ${\mathbb V}^i/{\mathbb V}^{i+1}$ is the local system associated with a rational representation of $S$. \end{enumerate} Denote the fiber over the base point $x_o$ by $V_o$. It has a filtration corresponding to the filtration ${\mathbb V}^{\bullet}$ of ${\mathbb V}$: $$ V_o = V_o^0 \supset V_o^1 \supset V_o^2 \supset \cdots $$ The second condition above implies that there are rational representations $\tau_i : S \to \Aut \Gr^i V_o$ such that the representation of $\pi_1(X,x_o)$ on $\Gr^i V_o$ is the composite $$ \pi_1(X,x_o) \stackrel{\rho}{\to} S \stackrel{\tau_i}{\to} \Aut \Gr^i V_o. $$ Let $\tau : S \to \prod \Aut \Gr^i V_o$ be the product of the representations $\tau_i$. Let $$ G = \left\{\phi \in \Aut V_o : \phi\text{ preserves } V_o^{\bullet}\text{ and } \Gr^{\bullet}\phi\in \im \tau \right\}. $$ This is a linear algebraic group which is an extension of $\im \tau$ by the unipotent group $$ U = \left\{\phi \in \Aut V_o : \phi\text{ preserves } V_o^{\bullet}\text{ and acts trivially on } \Gr^{\bullet} V_o \right\} $$ whose Lie algebra we shall denote by ${\mathfrak u}$. We shall denote the monodromy representation at $x_o$ of ${\mathbb V}$ by $$ \tilde{\rho} : \pi_1(X,x_o) \to G. $$ Denote the $C^\infty$ vector bundles associated to the flat bundles ${\mathbb V}$ and ${\mathbb V}^i$ by ${\mathcal V}$ and ${\mathcal V}^i$, respectively. We would like to trivialize ${\mathcal V}$. In order to do this, we pull it back to $P$ along the projection $p:P \to X$. \begin{proposition}\label{triv} There is a trivialization $$ p^\ast {\mathcal V} \stackrel{\cong}{\to} P\times V_o $$ and a splitting of the natural map $G \to \im \tau$ which satisfy \begin{enumerate} \item the corresponding connection form% \footnote{Our convention is that the connection form associated to the trivialized bundle $V\times X \to X$ with connection $\nabla$ is the 1-form $\omega$ on $X$ with values in $\End V$ which is characterized by the property that for all sections $f:X \to V$ $$ \nabla f = df - f\omega \in E^1(X)\otimes V. $$} $\widetilde{\omega}$ satisfies $$ \widetilde{\omega} \in E^1(P)\otimes {\mathfrak u}; $$ \item\label{cond} the isomorphism $V_o \to V_o$, induced by the trivialization of $p^\ast {\mathcal V}$ between the fiber over the points $\tilde{x}_o$ and $s\cdot\tilde{x}_o$ of $p^{-1}(x_o)$, is $\tau(s)^{-1}$. \end{enumerate} \end{proposition} Note that the second condition implies that the isomorphism $V_o \to V_o$, induced by the trivialization of $p^\ast {\mathcal V}$ between the fiber over the points $a\cdot \tilde{x}_o$ and $sa\cdot\tilde{x}_o$ of $p^{-1}(x_o)$, is $\tau(s)^{-1}$. The first step in the proof is the following elementary result. It can be proved by induction on the length of the filtration. It gives the splitting of $G \to \im \tau$. \begin{lemma}\label{splitting} There is an isomorphism $$ {\mathcal V} \cong \bigoplus_{i\ge 0} \Gr^i {\mathcal V} $$ of $C^\infty$ vector bundles that splits the filtration ${\mathcal V}^{\bullet}$. That is, \begin{enumerate} \item the sub-bundle ${\mathcal V}^i$ corresponds to $\oplus_{j\ge i} \Gr^j {\mathcal V}$; \item the isomorphism $$ \Gr^i {\mathcal V} \to {\mathcal V}^i/{\mathcal V}^{i+1} $$ induced by the trivialization is the identity. \hfill \qed \end{enumerate} \end{lemma} \begin{proof}[Proof of (\ref{triv})] Pulling back the splitting given by (\ref{splitting}) of the filtration ${\mathcal V}^i$ to $P$, we obtain a splitting $$ p^\ast{\mathcal V} \cong \bigoplus_i p^\ast \Gr^i {\mathcal V} $$ of $p^\ast {\mathcal V}$. So it suffices to trivialize each $p^\ast \Gr^i {\mathcal V}$. To do this, we first do it on a single leaf ${\mathcal L}$ of $P$. The restriction of the monodromy representation $\tau$ to ${\mathcal L}$ is clearly trivial. Consequently, the restriction of $p^\ast {\mathcal V}$ to ${\mathcal L}$ is trivial as a flat bundle. Observe that if this leaf contains $\tilde{x}_o$, then this trivialization satisfies condition (\ref{cond}) in the statement of (\ref{triv}). Next, change the trivialization of $p^\ast\Gr^i{\mathcal V}$ on $p^{-1}(x_o)$ so that it satisfies condition (\ref{cond}) in the statement of (\ref{triv}). Extend this to a trivialization of $p^{-1}\Gr^i{\mathcal V}$ on all of $P$ by parallel transport along the leaves of $P$. This gives a well defined local trivialization which is a global trivialization by the argument in the previous paragraph. We thus obtain a trivialization of $p^\ast{\mathcal V}$ which is compatible with the filtration ${\mathcal V}^{\bullet}$ and which is flat on each $\Gr^i{\mathcal V}$. It follows that the connection form $\widetilde{\omega}$ associated to this trivialization satisfies $\widetilde{\omega} \in E^1(P)\otimes {\mathfrak u}$. \end{proof} If $S$ is not finite, this connection is not flat as it is not flat in the vertical direction. We can make it flat by restricting it to the leaves of the foliation ${\mathcal F}$ of $P$. Denote the image of $\widetilde{\omega}$ under the restriction homomorphism $$ E^1(P)\otimes {\mathfrak u} \to E^1(P,{\mathcal F}) \otimes {\mathfrak u} $$ by $\omega$. It defines the connection in the leaf direction. This connection is clearly flat, and it follows that $\omega$ is integrable. The following assertion is a consequence of the properties (1) and (2) in the statement of Proposition \ref{triv} and (\ref{dga_homom}). Note that we view $S$ as acting on the left of ${\mathfrak u}$ via the adjoint action --- that is, via the composite $S \to \im \tau \hookrightarrow G \to \Aut {\mathfrak u}$. \begin{proposition} The connection form $\omega$ is integrable and lies in the subspace $\Efin^1(X,{\mathcal O}(P))\otimes {\mathfrak u}$ of $E^1(P,{\mathcal F})\otimes {\mathfrak u}$. Moreover, if $s\in S$, then $s^\ast \omega = Ad(s)\omega$. \qed \end{proposition} \begin{remark}\label{converse} There is a converse to this result. Suppose that ${\mathfrak u}$ is a nilpotent Lie algebra in the category of rational representations of $S$. Then we can form the semi-direct product $G=S\ltimes U$, where $U$ is the corresponding unipotent group. If $V$ is a $G$ module, and if $$ \omega \in \Efin^1(X,{\mathcal O}(P))\otimes {\mathfrak u} $$ satisfies the conditions \begin{enumerate} \item $d\omega + \omega \wedge \omega = 0$; \item $s^\ast \omega = Ad(s)\omega$; \end{enumerate} then we can construct an object of ${\mathcal B}(X,S)$ with fiber $V$ over $x_o$ whose pullback to $P$ has connection form $\omega$ with respect to an appropriate trivialization. \end{remark} We are now ready to express the monodromy representation of ${\mathbb V}$ in terms of iterated integrals of $\omega$. Recall that K.-T.~Chen \cite{chen} defined, for 1-forms $w_i$ on a manifold $M$ taking values in an associative algebra $A$, $$ \int_\gamma w_1 w_2 \dots w_r $$ to be the element $$ \idotsint\limits_{0 \le t_1 \le \dots \le t_r \le 1} f_1(t_1)f_2(t_2) \dots f_r(t_r)\, dt_1dt_2 \dots dt_r $$ of $A$. This is regarded as an $A$-valued function $PM \to A$ on the path space of $M$. An $A$-valued iterated integral is a function $PM \to A$ which is a linear combination of functions of this form together with a constant function. Suppose that $V\times M \to M$ is a trivial bundle with a connection given by the connection form $$ \omega \in E^1(M) \otimes \End(V). $$ In this case we can define the parallel transport map $$ T : PM \to \Aut(V) $$ where $PM$ denotes the space of piecewise smooth paths in $M$. A path goes to the linear transformation of $V$ obtained by parallel transporting the identity along it. Chen \cite{chen} obtained the following expression for $T$ in terms of $\omega$. \begin{proposition}\label{transp} With notation as above, we have $$ T(\gamma) = 1 + \int_\gamma \omega + \int_\gamma \omega\omega + \int_\gamma \omega\omega\omega + \cdots \qed $$ \end{proposition} Note that since ${\mathfrak u}$ is nilpotent, this is a finite sum. Armed with this formula, we can express the monodromy of ${\mathbb V} \to X$ in terms of $\omega \in \Efin^1(X,{\mathcal O}(P))$. Suppose that $\gamma \in P_{x_o,x_o}X$. Denote the unique lift of $\gamma$ to $P$ which is tangent to ${\mathcal F}$ and begins at $\tilde{x}_o\in p^{-1}(x_o)$, by ${\tilde{\gamma}}$. \begin{proposition}\label{monod} The monodromy of ${\mathbb V} \to X$ takes $\gamma \in P_{x_o,x_o}X$ to $$ \tilde{\rho}(\gamma) = \left(1 + \int_{{\tilde{\gamma}}} \omega + \int_{{\tilde{\gamma}}} \omega\omega + \int_{{\tilde{\gamma}}} \omega\omega\omega + \cdots \right)\tau(\rho(\gamma)) \in G. $$ \end{proposition} The proof is a straightforward consequence of Chen's formula (\ref{transp}) and condition (\ref{cond}) of (\ref{triv}). This formula motivates the following generalization of Chen's iterated integrals. \begin{definition}\label{defn} For $\phi\in {\mathcal O}(S)$ and $w_1,\dots,w_r$ elements of $\Efin^1(X,{\mathcal O}(P))$, we define $$ \int \left(w_1 \dots w_r | \phi\right) : P_{x_o,x_o}X \to {\mathbb R} $$ by $$ \int_\gamma \left(w_1 \dots w_r | \phi\right) = \phi(\rho(\gamma))\int_{{\tilde{\gamma}}}w_1\dots w_r. $$ \end{definition} We will call linear combinations of such functions {\it iterated integrals with coefficients in ${\mathcal O}(S)$}. They will be regarded as functions $P_{x_o,x_o}X \to {\mathbb R}$. We will denote the set of them by $I(X,{\mathcal O}(S))_{x_o}$. Such an iterated integral will be said to be {\it locally constant} if it is constant on each connected component of $P_{x_o,x_o}X$. We shall denote the set of locally constant iterated integrals on $P_{x_o,x_o}X$ by $H^0(I(X,{\mathcal O}(S))_{x_o})$. Evidently, each such locally constant iterated iterated integral defines a function $$ \pi_1(X,x_o) \to {\mathbb R}. $$ By taking matrix entries in (\ref{monod}), we obtain the following result. \begin{corollary}\label{matrix_entries} Each matrix entry of the monodromy representation $$ \tilde{\rho}: \pi_1(X,x_o) \to G $$ of an object of ${\mathcal B}(X,S)$ can be expressed as a locally constant iterated integral on $X$ with coefficients in ${\mathcal O}(S)$. \qed \end{corollary} The following results imply that $H^0(I(X,{\mathcal O}(S))_{x_o})$ is a Hopf algebra with coproduct dual to the multiplication of paths, and antipode dual to the involution of $P_{x_o,x_o}X$ that takes each path to its inverse. \begin{proposition}\label{props} Suppose that $\gamma$ and $\mu$ are in $P_{x_o,x_o}X$, that $\phi,\psi\in {\mathcal O}(S)$ and that $w_1,w_2,\dots \in \Efin^1(X,{\mathcal O}(P))$. Then we have: \begin{equation} \int_\gamma \left(w_1 \dots w_p | \phi\right) \int_\gamma \left(w_{p+1} \dots w_{p+q} | \phi\right) = \sum_{\sigma \in Sh(p,q)} \int_\gamma \left( w_{\sigma(1)}\dots w_{\sigma(p+q)}| \phi\psi \right) \end{equation} where $Sh(p,q)$ denotes the set of shuffles of type $(p,q)$; \begin{equation} \int_{\gamma^{-1}} \left(w_1 \dots w_r | \phi\right) = (-1)^r \int_\gamma \left(\rho(\gamma^{-1})^\ast w_r \dots \rho(\gamma^{-1})^\ast w_1| i_S^\ast\phi\right) \end{equation} where $i_S^\ast : {\mathcal O}(S) \to {\mathcal O}(S)$ is the antipode of ${\mathcal O}(S)$; \begin{equation} \int_{\gamma\mu}\left(w_1 \dots w_r | \phi\right) = \sum_{i=0}^r \sum_j \int_\gamma\left( w_1\dots w_i| \phi_j'\right) \int_\mu \left(\rho(\gamma)^\ast w_{i+1}\dots \rho(\gamma)^\ast w_r | \phi_j''\right) \end{equation} where $\Delta_S : {\mathcal O}(S) \to {\mathcal O}(S)\otimes {\mathcal O}(S)$ is the coproduct of ${\mathcal O}(S)$, and $$ \Delta_S \phi = \sum_j \phi_j'\otimes \phi_j''. $$ \end{proposition} \begin{proof} This proof is a straightforward using the definition (\ref{defn}) and basic properties of classical iterated integrals due to Chen \cite{chen}. \end{proof} \begin{corollary} The set of iterated integrals $I(X,{\mathcal O}(S))_{x_o}$ is a commutative Hopf algebra. \end{corollary} \begin{remark} Let $\pi_1(X,x_o) \to {\mathcal G}$ be the completion of $\pi_1(X,x_o)$ relative to $\rho : \pi_1(X,x_o) \to S$. Since the coordinate ring ${\mathcal O}({\mathcal G})$ of ${\mathcal G}$ is the ring of matrix entries of representations of $G$, it follows from (\ref{matrix_entries}) that there is a Hopf algebra inclusion $$ {\mathcal O}({\mathcal G}) \hookrightarrow H^0(I(X,{\mathcal O}(S))_{x_o}). $$ To prove this assertion, it would suffice to show that $$ H^0(I(X,{\mathcal O}(P)))\otimes_{{\mathcal O}(S)}{\mathbb R} $$ is the direct limit of coordinate rings of a directed system of unipotent groups, each with an $S$ action. This is surely true, but we seek a more algebraic de~Rham theorem for ${\mathcal O}({\mathcal G})$ which is more convenient for Hodge theory. \end{remark} \section{Higher Iterated Integrals} As a preliminary step to defining the algebraic analogue of $I(X,{\mathcal O}(S))_{x_o}$, we generalize the the definition of iterated integrals with values in ${\mathcal O}(S)$ to higher dimensional forms. Denote by $E^n(P_{x_o,x_o}X)$ the differential forms of degree $n$ on the loop space $P_{x_o,x_o}X$. One can surely use any reasonable definition of differential forms on $P_{x_o,x_o}X$, but we will use Chen's definition from \cite{chen} where, to specify a differential form on $P_{x_o,x_o}X$, it is enough to specify its pullback along each ``smooth map'' $\alpha : U\to P_{x_o,x_o}X$ from an open subset $U$ of some finite dimensional euclidean space. By a smooth map, we mean a map $\alpha : U\to P_{x_o,x_o}X$ whose ``suspension'' $$ \widehat{\alpha} : [0,1] \times U \to X; \quad (t,u) \mapsto \alpha(u)(t) $$ is continuous and smooth on each $[t_{j-1},t_j]\times U$ for some partition $$ 0=t_0 \le t_1 \le \dots \le t_m = 1 $$ of $[0,1]$. \begin{definition}\label{higherdef} Suppose that $\phi \in {\mathcal O}(S)$, and that $w_j\in \Efin^{n_j}(X,{\mathcal O}(P))$ with each $n_j>0$. Set $n=-r + \sum_j n_j$. Define $$ \int \left(w_1 \dots w_r | \phi\right) \in E^n(P_{x_o,x_o}X) $$ by specifying that for each smooth map $\alpha : U \to P_{x,x}X$, $$ \alpha^\ast\int \left(w_1 \dots w_r | \phi\right) $$ is the element $$ \idotsint\limits_{0\le t_1 \le \cdots \le t_r \le 1} \widehat{w}_1(t_1)\wedge \dots \wedge \widehat{w}_r(t_r)\, dt_1\! dt_2\dots dt_r\, \phi(\rho(\alpha(u))) $$ of $E^n(U)$, where $$ \widehat{w}_j : (\partial/\partial t) \lrcorner \tilde{\alpha}^\ast w_j $$ and $\tilde{\alpha} : [0,1]\times U \to P$ is the smooth map with the property that for each $x\in U$, the map $t\mapsto \tilde{\alpha}(t,x)$ is the unique lift of $t\mapsto \widehat{\alpha}(t,x)$ that begins at $\tilde{x}_o$ and is tangent to ${\mathcal F}$. \end{definition} These iterated integrals form a subspace $I^{\bullet}(X,{\mathcal O}(S))_{x_o}$ of $E^{\bullet}(P_{x_o,x_o}X)$. Chen's arguments \cite{chen} can be adapted easily to show that this is, in fact, a sub d.g.\ Hopf algebra of $E^{\bullet}(P_{x_o,x_o}X)$. In particular, we have: \begin{proposition} The space of locally constant iterated integrals on $X$ with coefficients in ${\mathcal O}(S)$ is $H^0(I^{\bullet}(X,{\mathcal O}(S))_{x_o})$. \qed \end{proposition} \section{The Reduced Bar Construction} In this section we review Chen's definition of the reduced bar construction which he described in \cite{chen:bar}. Suppose that $A^{\bullet}$ is a commutative differential graded algebra (hereafter denoted d.g.a.) and that $M^{\bullet}$ and $N^{\bullet}$ are complexes which are modules over $A^{\bullet}$. That is, the structure maps $$ A^{\bullet} \otimes M^{\bullet} \to M^{\bullet} \text{ and } A^{\bullet}\otimes N^{\bullet} \to N^{\bullet} $$ are chain maps. We shall suppose that $A^{\bullet}$, $M^{\bullet}$ and $N^{\bullet}$ are all positively graded. Denote the subcomplex of $A^{\bullet}$ consisting of elements of positive degree by $A^+$. The {\it (reduced) bar construction} $B(M^{\bullet},A^{\bullet},N^{\bullet})$ is defined as follows. We first describe the underlying graded vector space. It is a quotient of the graded vector space $$ T(M^{\bullet},A^{\bullet},N^{\bullet}) := \bigoplus_s M^{\bullet} \otimes\left(A^+[1]^{\otimes r}\right) \otimes N^{\bullet}. $$ We will use the customary notation $m[a_1|\dots|a_r]n$ for $$ m\otimes a_1\otimes \dots \otimes a_r \otimes n \in T(M^{\bullet},A^{\bullet},N^{\bullet}). $$ To obtain the vector space underlying the bar construction, we mod out by the relations $$ m[dg|a_1|\dots|a_r]n = m[ga_1|\dots|a_r]n - m\cdot g[a_1|\dots|a_r]n; $$ \begin{multline*} m[a_1|\dots|a_i|dg|a_{i+1}|\dots|a_r]n = \hfill \cr m[a_1|\dots|a_i|g\,a_{i+1}|\dots|a_r]n - m[a_1|\dots|a_i\,g|a_{i+1}|\dots|a_r]n \quad 1\le i < s; \end{multline*} $$ m[a_1|\dots|a_r|dg]n = m[a_1|\dots|a_r]g\cdot n - m[a_1|\dots|a_r\,g]n; $$ $$ m[dg]n = 1 \otimes g\cdot n - m\cdot g \otimes 1 $$ Here each $a_i \in A^+$, $g\in A^0$, $m\in M^{\bullet}$, $n\in N^{\bullet}$, and $r$ is a positive integer. Before defining the differential, it is convenient to define an endomorphism $J$ of each graded vector space by $J: v\mapsto (-1)^{\deg v}v$. The differential is defined as $$ d = d_M\otimes 1_T \otimes 1_N + J\otimes d_B \otimes 1 + J_M \otimes J_T \otimes d_N + d_C. $$ Here $T$ denotes the tensor algebra on $A^+[1]$, $d_B$ is defined by \begin{multline*} d_B[a_1|\dots|a_r] = \sum_{1\le i \le r} (-1)^i [Ja_1|\dots|Ja_{i-1}|da_i|a_{i+1}|\dots|a_r] \cr \hfill + \sum_{1 \le i < r} (-1)^{i+1}[Ja_1|\dots|Ja_{i-1}|Ja_i\wedge a_{i+1}|a_{i+2}|\dots|a_r] \end{multline*} and $d_C$ is defined by $$ d_C m[a_1|\dots|a_r]n = (-1)^s Jm[Ja_1|\dots|Ja_{r-1}]a_r \cdot n - Jm\cdot a_1 [a_2|\dots|a_r]n. $$ One can check that these differentials are well defined. If both $M^{\bullet}$ and $N^{\bullet}$ are themselves d.g.a.s over $A^{\bullet}$, then $B(M^{\bullet},A^{\bullet},N^{\bullet})$ is also a differential graded algebra. The product is defined by \begin{multline}\label{prod} m'[a_1|\dots|a_p]n' \otimes m''[a_{p+1}|\dots |a_{p+q}]n'' \mapsto \hfill \cr \hfill \sum_{\sigma\in \Sigma(p,q)} \pm m'\wedge m'' [a_{\sigma(1)}| a_{\sigma(2)}|\dots |a_{\sigma(p+q)}]n'\wedge n''. \end{multline} Here $\Sigma(p,q)$ denotes the set of shuffles of type $(p,q)$. The sign in front of each term on the right hand side is determined by the usual sign conventions that apply when moving a symbol of degree $k$ past one of degree $l$ --- one considers each $a_j$ to be of degree $-1 + \deg a_j$. The reduced bar construction $B(M^{\bullet},A^{\bullet},N^{\bullet})$ has a standard filtration $$ {\mathbb R} = B_0(M^{\bullet},A^{\bullet},N^{\bullet}) \subseteq B_1(M^{\bullet},A^{\bullet},N^{\bullet}) \subseteq B_2(M^{\bullet},A^{\bullet},N^{\bullet}) \subseteq \cdots $$ which is often called the {\it bar filtration}. The subspace $$ B_s(M^{\bullet},A^{\bullet},N^{\bullet}) $$ is defined to be the span of those $m[a_1|\dots|a_r]n$ with $r\le s$. When $A^{\bullet}$ has connected homology (i.e., $H^0(A^{\bullet}) = {\mathbb R}$), the corresponding spectral sequence, which is called the {\it Eilenberg-Moore spectral sequence}, has $E_1$ term $$ E_1^{-s,t} = \left[M^{\bullet}\otimes H^+(A^{\bullet})^{\otimes s}\otimes N^{\bullet}\right]^t. $$ A proof of this can be found in \cite{chen:bar}. The following basic property of the reduced bar construction is a special case of a result proved in \cite{chen:bar}. It is easily proved using the Eilenberg-Moore spectral sequence. Suppose that $\psi : A_1^{\bullet} \to A_2^{\bullet}$ is a d.g.a.\ homomorphism, and that $M^{\bullet}$ is a right $A_2^{\bullet}$ module and $N^{\bullet}$ a right $A_2^{\bullet}$ module. Then $M^{\bullet}$ and $N^{\bullet}$ can be regarded as $A_1^{\bullet}$ modules via $\psi$. We therefore have a chain map \begin{equation}\label{map} B(M^{\bullet},A_1^{\bullet},N^{\bullet}) \to B(M^{\bullet},A_2^{\bullet},N^{\bullet}). \end{equation} \begin{proposition}\label{qism} If $\psi$ is a quasi-isomorphism, then so is (\ref{map}). \qed \end{proposition} \section{The Construction of $\Gdr$} \label{gdr} In this section we construct a proalgebraic group $\Gdr$ which is an extension $$ 1 \to \Udr \to \Gdr \stackrel{p}{\to} S \to 1, $$ where $\Udr$ is prounipotent, and a homomorphism $$ \tilde{\rho} : \pi_1(X,x_o) \to \Gdr $$ whose composition with $p : \Gdr \to S$ is $\rho$. We do this by constructing the coordinate ring of $\Gdr$ using the bar construction. In the two subsequent sections, we will show that $\tilde{\rho} : \pi_1(X,x_o) \to \Gdr$ is the Malcev completion of $\pi_1(X,x_o)$ relative to $\rho$. The fixed choice of a base point $\tilde{x}_o \in p^{-1}(x_o)$ determines augmentations $$ \epsilon_{\tilde{x}_o} : \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathbb R} $$ and $$ \delta_{x_o} : \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathcal O}(S) $$ as we shall now explain. Since these augmentations are compatible with restriction, it suffices to give them in a neighbourhood of $x_o$. Over a contractible neighbourhood $U$ of $x_o$, the local system $P$ is trivial and may therefore by identified with the trivial flat bundle $S\times U \to S$ in such a way that $\tilde{x}_o$ corresponds to $(1,x_o) \in S\times U$. The restriction of an element of $\Efin^k(X,{\mathcal O}(P))$ to $U$ is then of the form $$ \sum_i \phi_i \otimes w_i $$ where $\phi_i \in {\mathcal O}(S)$, and $w_i \in E^k(U)$. Denote the augmentation $E^{\bullet}(U) \to {\mathbb R}$ induced by $x_o$ by $\mu_{x_o}$. Then the augmentations $\delta_{x_o}$ and $\epsilon_{\tilde{x}_o}$ are defined by $$ \delta_{x_o} : \sum_i \phi_i \otimes w_i \mapsto \sum_i \mu_{x_o}(w_i)\, \phi_i $$ and $$ \epsilon_{\tilde{x}_o} : \sum_i \phi_i \otimes w_i \mapsto \sum_i \mu_{x_o}(w_i)\, \phi_i(1). $$ One can regard ${\mathbb R}$ and ${\mathcal O}(S)$ as algebras over $\Efin^{\bullet}(X,{\mathcal O}(P))$ where the actions of $\Efin^{\bullet}(X,{\mathcal O}(P))$ on these is defined using these two augmentations. We can therefore form the bar construction $$ B({\mathbb R},\Efin^{\bullet}(X,{\mathcal O}(P)),{\mathcal O}(S)) $$ which we shall denote by $B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})$. It is a commutative d.g.a.\ when endowed with the product (\ref{prod}). It is, in fact, a d.g.~Hopf algebra, with coproduct defined as follows: \begin{multline*} \Delta : [w_1|\dots|w_r]\phi \mapsto\\ \sum_i [w_1|\dots |w_i] \left(\sum \psi_i^{(k_i)}\dots\psi_r^{(k_r)}\phi'\right) \otimes [w_{i+1}^{(k_i)}|\dots|w_r^{(k_r)}]\left(\sum \phi''\right) \end{multline*} where $$ \Delta_S(\phi) = \sum \phi'\otimes \phi'' $$ is the diagonal of ${\mathcal O}(S)$, and the map $$ \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathcal O}(S) \otimes \Efin^1(X,{\mathcal O}(P)) $$ which gives the $S$ action takes $w_j$ to $$ \sum \psi_j^{(k_j)} \otimes w_j^{(k_j)}. $$ The following proposition is a direct consequence of the definition (\ref{higherdef}) and the basic properties of iterated integrals which may be found in \cite{chen}. \begin{proposition}\label{pro} The map $$ B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)}) \to I^{\bullet}(X,{\mathcal O}(S))_{x_o} $$ defined by $$ [w_1|w_2|\dots |w_r]\phi \mapsto \int\left(w_1 w_2 \dots w_r|\phi\right) $$ is a well defined d.g.~Hopf algebra homomorphism. \qed \end{proposition} \begin{proposition}\label{alg_gp} If $\pi_1(X,x_o)$ is finitely generated, then $$ H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ is the coordinate ring of a linear proalgebraic group which is an extension of $S$ by a prounipotent group. \end{proposition} In the proof, we shall need the following technical result, the proof of which is a straightforward modification of Sullivan's proof of the existence of minimal models (cf.\ \cite{sullivan}.) \begin{proposition}\label{sub} There is a d.g.~subalgebra $A^{\bullet}$ of $\Efin^{\bullet}(X,{\mathcal O}(P))$ with $A^0={\mathbb R}$, which is also an $S$ submodule, with the properties that the inclusion is a quasi-isomorphism. \qed \end{proposition} \begin{proof}[Proof of (\ref{alg_gp})] Choose a d.g.\ subalgebra $A^{\bullet}$ of $\Efin^{\bullet}(X,{\mathcal O}(P))$ as given by (\ref{sub}). It follows from (\ref{qism}) that the natural map $$ H^0(B({\mathbb R},A^{\bullet},{\mathcal O}(S))) \to H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ is an isomorphism. Since $A^0={\mathbb R}$, we have that $$ H^0(B({\mathbb R},A^{\bullet},{\mathcal O}(S))) = H^0(B({\mathbb R},A^{\bullet},{\mathbb R}))\otimes {\mathcal O}(S). $$ It is not difficult to check that ${\mathcal O}(S)$ is a sub Hopf algebra, and that this is a tensor product of algebras, but where the coproduct is twisted by the action of $S$ on $H^0(B({\mathbb R},A^{\bullet},{\mathbb R}))$. So, if we can show that $H^0(B({\mathbb R},A^{\bullet},{\mathbb R}))$ is the limit of the coordinate rings of an inverse system of unipotent groups, each with an $S$ action, then we will have shown that $$ H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ is the coordinate ring of $$ S\ltimes \spec H^0(B({\mathbb R},A^{\bullet},{\mathbb R})) $$ and therefore proved the proposition. From \cite{hain:bar}, we know that there is a canonical splitting (in particular, it is $S$ equivariant) of the projection $$ H^0(B({\mathbb R},A^{\bullet},{\mathbb R})) \to QH^0(B({\mathbb R},A^{\bullet},{\mathbb R})) =: Q $$ onto the indecomposable elements $Q$. This splitting induces an $S$-equivariant algebra isomorphism $$ {\mathbb R}[Q] \to H^0(B({\mathbb R},A^{\bullet},{\mathbb R})). $$ The bar filtration induces a filtration $$ Q_1\subseteq Q_2 \subseteq Q_3 \subseteq \cdots \subseteq Q $$ of the indecomposables such that $Q= \cup Q_r$. Each $Q_r$ is a Lie coalgebra, and the cobracket $\Delta^c$ satisfies $$ \Delta^c : Q_r \to \sum_{i+j = r} Q_i \otimes Q_j $$ and is injective when $r>1$. Since $\pi_1(X,x_o)$ is finitely generated, each of the cohomology groups $H^1(X,{\mathbb V})$ is finite dimensional for each rational local system ${\mathbb V}$ over $X$. It follows from (\ref{isom}) that each isotypical component of $H^1(X,{\mathcal O}(P))$ is finite dimensional. Since $Q_r/Q_{r-1}$ is a subquotient of $$ H^1(X,{\mathcal O}(P))^{\otimes r}, $$ it follows that each $S$-isotypical component of each $Q_r$ is finite dimensional. One can now prove by induction on $r$ using the nilpotence, that as an $S$-module, each $Q_r$ is the direct limit of duals of nilpotent Lie algebras, each of which has an $S$ action. This completes the proof. \end{proof} \begin{definition}\label{def} Define proalgebraic groups $\Gdr$ and $\Udr$ by $$ \Gdr = \spec H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ and $$ \Udr = \spec H^0(B({\mathbb R},\Efin^{\bullet}(X,{\mathcal O}(P)),{\mathbb R})). $$ \end{definition} Evidently, we have an extension $$ 1 \to \Udr \to \Gdr \to S \to 1 $$ of proalgebraic groups, where $\Udr$ is prounipotent. When we want to emphasize the dependence of $\Gdr$ and $\Udr$ on $(X,x)$, we will write them as $\Gdr(X,x)$ and $\Udr(X,x)$, respectively. \begin{proposition}\label{homom} There is a natural homomorphism $\tilde{\rho} : \pi_1(X,x_o) \to \Gdr$ whose composition with $\Gdr \to S$ is $\rho$. \end{proposition} \begin{proof} Define a map from $P_{x_o,x_o}X$ to the linear functionals on $$ B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)}) $$ by $$ \gamma : [w_1|\dots|w_r]\phi \mapsto \int_\gamma\left(w_1\dots w_r|\phi\right). $$ This induces a function $$ \Phi : \pi_1(X,x_o) \to \Hom_{\mathbb R}({\mathcal O}(\Gdr),{\mathbb R}). $$ Define $\tilde{\rho}$ by taking the class of $\gamma$ in $\pi_1(X,x_o)$ to the maximal ideal of $$ H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ consisting of those elements on which $\gamma$ vanishes. (Note that $\gamma$ acts via integration.) That this is a group homomorphism follows from (\ref{props}). \end{proof} \section{Construction of Homomorphisms from $\Gdr$} Suppose that $G$ is a linear algebraic group which can be expressed as an extension $$ 1 \to U \to G \to S \to 1 $$ where $U$ is unipotent. Choose an isomorphism of $G$ with $S\ltimes U$. Denote the Lie algebra of $U$ by ${\mathfrak u}$. \begin{proposition} Each one form $\omega \in \Efin^1(X,{\mathcal O}(P))\otimes {\mathfrak u}$ that satisfies \begin{enumerate} \item $d\omega + \omega\wedge \omega = 0$; \item for all $s \in S$, $s^\ast \omega = Ad(s)\omega$; \end{enumerate} determines a homomorphism $\Gdr \to G$ that commutes with projection to $S$. \end{proposition} \begin{proof} First note that since the exponential map ${\mathfrak u} \to U$ is a polynomial isomorphism, ${\mathcal O}(U)$ is isomorphic to the polynomials ${\mathbb R}[{\mathfrak u}]$ on the vector space ${\mathfrak u}$. Further, there is a natural isomorphism \begin{equation}\label{iso} {\mathcal O}(U) \cong {\mathbb R}[{\mathfrak u}] \to \lim_\to \Hom(U{\mathfrak u}/I^n,{\mathbb R}) \end{equation} which is defined by noting that $U{\mathfrak u}$ is, by the PBW Theorem, the symmetric coalgebra $S^c{\mathfrak u}$ on ${\mathfrak u}$. The isomorphism (\ref{iso}) is an isomorphism of Hopf algebras. Set $$ T = 1 + [\omega] + [\omega|\omega] + [\omega|\omega|\omega] + \cdots $$ which we view as an element of $$ B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)}) \comptensor \widehat{U}{\mathfrak u} $$ of degree zero, where $\widehat{U}{\mathfrak u}$ denotes the completion $$ \lim_\leftarrow U{\mathfrak u}/I^n $$ of $U{\mathfrak u}$ with respect to the powers of its augmentation ideal, and where $\comptensor$ denotes the completed tensor product $$ \lim_\leftarrow B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})\otimes \widehat{U}{\mathfrak u}/I^n $$ The coordinate ring of $G$ is isomorphic to ${\mathcal O}(U)\otimes{\mathcal O}(S)$. Define a linear map $$ \Theta : {\mathcal O}(G) \to B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})^0 $$ by $$ f\otimes \phi \mapsto \langle T,f\rangle \cdot \phi. $$ It is not difficult to check that $\Theta$ is a well defined Hopf algebra homomorphism. This uses the fact that $s^\ast \omega = Ad(s)\omega$. That $\omega$ satisfies the integrability condition $$ d \omega + \omega \wedge \omega = 0 $$ implies that $\im \Theta$ is contained in $H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)}))$. It follows that $\Theta$ induces a Hopf algebra homomorphism $$ {\mathcal O}(G) \to H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,(x_o)})) $$ and therefore a group homomorphism $$ \theta : \Gdr \to G $$ which commutes with the projections to $S$. Finally, it follows from (\ref{converse}), (\ref{transp}) and (\ref{homom}) that the composite $$ \pi_1(X,x_o) \to \Gdr \to G $$ is the homomorphism induced by $\omega$. \end{proof} \begin{corollary}\label{monod_rep} If ${\mathbb V}$ is a local system in ${\mathcal B}(X,S)$, then the monodromy representation $$ \tau : \pi_1(X,x_o) \to \Aut V_o $$ factors through $\tilde{\rho} : \pi_1(X,x_o) \to \Gdr$. \qed \end{corollary} \begin{corollary}\label{factor} If $\tau: \pi_1(X,x_o) \to G$ is a homomorphism into a linear algebraic group which is an extension of $S$ by a unipotent group, and whose composite with the projection to $S$ is $\rho$, then there is a homomorphism $\Gdr \to G$ whose composite with $$ \tilde{\rho} : \pi_1(X,x_o) \to \Gdr $$ is $\tau$. \end{corollary} \begin{proof} Denote the kernel of $G \to S$ by $U$. One can construct a faithful, finite dimensional representation $V$ of $G$ which has a filtration $$ V = V^0 \supset V^1 \supset V^2 \supset \cdots $$ by $G$-submodules whose intersection is zero and where each $V^j/V^{j+1}$ is a trivial $U$-module. The corresponding local system over $X$ lies in ${\mathcal B}(X,S)$. The result now follows from (\ref{monod_rep}). \end{proof} \section{Isomorphism with the Relative Completion} Denote $\pi_1(X,x_o)$ by $\pi$. In Section~\ref{gdr} we constructed a homomorphism $\pi \to \Gdr$. In this section, we prove: \begin{theorem}\label{main} If $\pi$ is finitely generated, then the homomorphism $\pi \to \Gdr$ is the completion of $\pi$ relative to $\rho$. \end{theorem} To prove the theorem, we first fix a completion $\pi \to {\mathcal G}$ of $\pi$ relative to $\rho$. The universal mapping property of the relative completion gives a homomorphism ${\mathcal G} \to \Gdr$ of proalgebraic groups that commutes with the canonical projections to $S$. It follows from (\ref{factor}), that there is a natural homomorphism \begin{equation}\label{univ} \Gdr \to {\mathcal G} \end{equation} that also commutes with the projections to $S$. It follows from the universal mapping property of the relative completion that the composite $$ {\mathcal G} \to \Gdr \to {\mathcal G} $$ is the identity. Denote the prounipotent radical of ${\mathcal G}$ by ${\mathcal U}$. Since $\pi$ is finitely generated, each of the groups $H^1(\pi,V)$ is finite dimensional for each rational representation $V$ of $S$. In view of the following proposition and the assumption that $\pi$ is finitely generated, all we need do to show that the natural homomorphism ${\mathcal G} \to \Gdr$ is an isomorphism is to show that either of the induced maps $$ \Hom_S(H_1({\mathcal U}),V) \to \Hom_S(H_1(\Udr),V) \to \Hom_S(H_1({\mathcal U}),V) $$ is an isomorphism for all $S$-modules $V$. \begin{proposition} Suppose that $G_1$ and $G_2$ are extensions of the reductive group $S$ by unipotent groups $U_1$, $U_2$, respectively: $$ 1 \to U_j \to G_j \to S \to 1. $$ Suppose that $\theta : G_1 \to G_2$ is a split surjective homomorphism of algebraic groups that commutes with the projections to $S$. If either of the induced maps $$ \Hom_S(H_1(U_1),V) \to \Hom_S(H_1(U_2),V) \to \Hom_S(H_1(U_1),V) $$ is an isomorphism for all $S$ modules $V$, then both are, and $\theta$ is an isomorphism. \end{proposition} \begin{proof} The proof reduces to basic fact that a split surjective homomorphism between nilpotent Lie algebras is an isomorphism if and only if it induces an isomorphism on $H_1$. The details are left to the reader. \end{proof} Our first task in the proof of Theorem~\ref{main} is to compute $\Hom_S(H_1({\mathcal U}),V)$. \begin{proposition}\label{h1_comp} For all $S$-modules $V$, there is a canonical isomorphism $$ H^1(\pi,V) \cong \Hom_S(H_1({\mathcal U}),V). $$ \end{proposition} \begin{proof} We introduce an auxiliary group for the proof. Let $$ \Hom_\rho(\pi,S\ltimes V) $$ be the set of group homomorphisms $\pi \to S\ltimes V$ whose composite with the projection $S\ltimes V \to S$ is $\rho$. Then there is a natural bijection between $\Hom_\rho(\pi,S\ltimes V)$ and the set of splittings $\pi \to \pi \ltimes V$ of the projection $\pi \ltimes V \to \pi$: the splitting $\sigma$ corresponds to $\tilde{\rho} : \pi \to S\ltimes V$ if and only if the diagram $$ \begin{CD} \pi @>\sigma>> \pi\ltimes V \cr @| @VV{\rho\ltimes id}V\cr \pi @>\tilde{\rho}>> S\ltimes V \cr \end{CD} $$ commutes. The kernel $V$ acts on both $\Hom_\rho(\pi,S\ltimes V)$ and on the set of splittings, in both cases by inner automorphisms. The action commutes with the bijection. Since $H^1(\pi,V)$ is naturally isomorphic to the set of splittings of $\pi\ltimes V \to \pi$ modulo conjugation by $V$ \cite[p.~106]{maclane}, the bijection induces a natural isomorphism $$ H^1(\pi,V) \cong \Hom_\rho(\pi,S\ltimes V)/\sim. $$ On the other hand, by the universal mapping property of the relative completion, each element of $\Hom_\rho(\pi,S\ltimes V)$ induces a homomorphism ${\mathcal G} \to S\ltimes V$ which commutes with the projections to $S$. Such a homomorphism induces a homomorphism ${\mathcal U} \to V$, and therefore an $S$-equivariant homomorphism $H_1({\mathcal U}) \to V$. Since $V$ is central, this induces a homomorphism $$ \Hom_\rho(\pi,S\ltimes V) \to \Hom_S(H_1({\mathcal U}),V). $$ To complete the proof, we show that this is an isomorphism. Denote the commutator subgroup of ${\mathcal U}$ by ${\mathcal U}'$. Then the quotient ${\mathcal G}/{\mathcal U}'$ is an extension of $S$ by $H_1({\mathcal U})$; the latter being a possibly infinite product of representations of $S$ in which each isotypical factor is finite dimensional. Using the fact that every extension of $S$ by a rational representation in the category of algebraic groups splits and that any two such splittings are conjugate by an element of the kernel, we see that the extension \begin{equation}\label{seqce} 0 \to H_1({\mathcal U}) \to {\mathcal G}/{\mathcal U}' \to S \to 1 \end{equation} is split and that any two splittings are conjugate by an element of $H_1({\mathcal U})$. Choose a splitting of this sequence. This gives an isomorphism $$ {\mathcal G}/{\mathcal U}' \cong S\ltimes H_1({\mathcal U}). $$ An $S$-equivariant homomorphism $H_1({\mathcal U}) \to V$ induces a homomorphism $$ {\mathcal G}/{\mathcal U}' \cong S\ltimes H_1({\mathcal U}) \to S\ltimes V $$ of proalgebraic groups. Composing this with the homomorphism $$ \pi \to {\mathcal G} \to {\mathcal G}/{\mathcal U}', $$ we obtain an element of $\Hom_\rho(\pi,S\ltimes V)/\sim$. Since all splittings of (\ref{seqce}) differ by an inner automorphism by an element of $H_1({\mathcal U})$, we have constructed a well defined map $$ \Hom_S(H_1({\mathcal U}),V) \to \Hom_\rho(\pi,S\ltimes V). $$ This is easily seen to be the inverse of the map constructed above. This completes the proof. \end{proof} The following result completes the proof of Theorem~\ref{main}. \begin{proposition} The map $$ H_1({\mathcal U}) \to H_1(\Udr) $$ induced by (\ref{univ}) is an isomorphism. \end{proposition} \begin{proof} It suffices to show that for all rational representations $V$ of $S$, the map $$ \left[H^1(\Udr)\otimes V\right]^S \to \left[H^1({\mathcal U})\otimes V\right]^S $$ is an isomorphism. Both groups are isomorphic to $H^1(X,{\mathbb V})$. Choose a de~Rham representative $w \in E^1(X,{\mathbb V})$ of a class in $H^1(X,{\mathbb V})$. Let $\delta \in V^\ast\otimes V$ be the element corresponding to the identity $V\to V$. Set $$ \omega := w\otimes \delta \in E^1(X,{\mathbb V})\otimes V^\ast \otimes V. $$ Regard $V$ as an abelian Lie algebra. Then $$ \omega \in \Efin^1(X,{\mathcal O}(P))\otimes V. $$ It is closed, and therefore satisfies the integrability condition $d\omega + \omega\wedge\omega = 0$. Since the identity $V\to V$ is $S$ equivariant, $$ s^\ast \omega = Ad(s) \omega $$ for all $s\in S$. Set $W = V\oplus {\mathbb R}$. Filter this by $$ W = W^0 \supset W^1 \supset W^2 =0 $$ where $W^1 = V$. Then $V\subset \End W$. It follows from (\ref{converse}) that $\omega$ defines a connection on $P\times V$ which is flat along the leaves of the foliation ${\mathcal F}$ and descends to a flat bundle over $X$. The monodromy representation of this bundle is a homomorphism $$ \tau:\pi_1(X,x_o) \to S\ltimes V \subset \Aut W. $$ It follows from the monodromy formula (\ref{transp}) that $\tau$ takes the class of the loop $\gamma$ to $$ \left(\rho(\gamma),\int_{\tilde{\gamma}} w\right) \in S\ltimes V. $$ The result follows. \end{proof} \section{Naturality} Suppose that $\pi_X$ and $\pi_Y$ are two groups, and that $\rho_X : \pi_X \to S_X$ and $\rho_Y : \pi_Y \to S_Y$ are homomorphisms into the $F$-points of reductive algebraic groups, each with Zariski dense image. We have the two corresponding relative completions $$ \tilde{\rho}_X : \pi_X \to {\mathcal G}_X \text{ and } \tilde{\rho}_Y : \pi_Y \to {\mathcal G}_Y. $$ Fix an algebraic group homomorphism $\Psi : S_X \to S_Y$. \begin{proposition}\label{induced} If $\psi : \pi_X \to \pi_Y$ is a homomorphism such that the diagram $$ \begin{CD} \pi_X @>{\rho_X}>> S_X \cr @V{\psi}VV @VV{\Psi}V \cr \pi_Y @>{\rho_Y}>> S_Y \end{CD} $$ commutes, then there is a canonical homomorphism $\widehat{\Psi} : {\mathcal G}_X \to {\mathcal G}_Y$ such that the diagram $$ \begin{CD} \pi_X @>{\tilde{\rho}_X}>> {\mathcal G}_X \cr @V{\psi}VV @VV{\widehat{\Psi}}V \cr \pi_Y @>{\tilde{\rho}_Y}>> {\mathcal G}_Y \end{CD} $$ \end{proposition} \begin{proof} Let $\Psi^\ast {\mathcal G}_Y$ be the pullback of $G_Y$ along $\Psi$: $$ \begin{CD} \Psi^\ast {\mathcal G}_Y @>>> S_X \cr @VVV @VV{\Psi}V \cr {\mathcal G}_Y @>>> S_Y. \end{CD} $$ This group is an extension of $S_X$ by the prounipotent radical of ${\mathcal G}_Y$. The homomorphisms $\pi_X \to S_X$ and $\pi_X \to \pi_Y \to {\mathcal G}_Y$ induce a homomorphism $\pi_X \to \Psi^\ast {\mathcal G}_Y$. By the universal mapping property of $\tilde{\rho}_X : \pi_X \to {\mathcal G}_X$, there is a homomorphism ${\mathcal G}_X \to \Psi^\ast{\mathcal G}_Y$ which extends the homomorphism $\pi_X \to \Psi^\ast{\mathcal G}_Y$. The sought after homomorphism $\widehat{\Psi}$ is the composite ${\mathcal G}_X \to \Psi^\ast{\mathcal G}_Y \to {\mathcal G}_Y$. \end{proof} Next we explain how to realize $\widehat{\Psi}$ using the bar construction. Suppose that $(X,x)$ and $(Y,y)$ are two pointed manifolds. Denote $\pi_1(X,x)$ and $\pi_1(Y,y)$ by $\pi_X$ and $\pi_Y$, respectively. Suppose that $f:(X,x) \to (Y,y)$ is a smooth map which induces the homomorphism $\psi : \pi_X \to \pi_Y$ on fundamental groups. Denote the principal bundles associated to $\rho_X$ and $\rho_Y$ by $P_X \to X$ and $P_Y \to Y$. We have the d.g.a.s $$ \Efin^{\bullet}(X,{\mathcal O}(P_X)) \text{ and } \Efin^{\bullet}(Y,{\mathcal O}(P_Y)). $$ Since the diagram in Proposition~\ref{induced} commutes, $f$ and $\psi$ induce a d.g.a.\ homomorphism $$ (f,\phi)^\ast : \Efin^{\bullet}(Y,{\mathcal O}(P_Y)) \to \Efin^{\bullet}(X,{\mathcal O}(P_X)). $$ This homomorphism respects the augmentations induced by $x \in X$ and $y \in Y$, and therefore induces a d.g.~Hopf algebra homomorphism $$ B(\Efin^{\bullet}(Y,{\mathcal O}(P_Y))_{\tilde{y},(y)}) \to B(\Efin^{\bullet}(X,{\mathcal O}(P_X))_{\tilde{x},(x)}). $$ This induces a homomorphism \begin{equation}\label{induced_homom} \Gdr(X,x) \to \Gdr(Y,y) \end{equation} after taking $H^0$ and then $\spec$. \begin{proposition} Under the canonical identifications of $\Gdr(X,x)$ with ${\mathcal G}(X,x)$ and $\Gdr(Y,y)$ with ${\mathcal G}(Y,y)$, the homomorphism $\widehat{\Psi} : {\mathcal G}(X,x) \to {\mathcal G}(Y,y)$ corresponds to the homomorphism (\ref{induced_homom}). \end{proposition} \begin{proof} If $\gamma$ is a loop in $X$ based at $x$, $w_1,\dots,w_r \in \Efin^{\bullet}(Y,{\mathcal O}(P_Y))$, and $\phi \in {\mathcal O}(S_Y)$, then $$ \int_{f\circ \gamma}\left(w_1 \dots w_r | \phi\right) \int_{\gamma}\left((f,\phi)^\ast w_1 \dots (f,\phi)^\ast w_r | \Psi^\ast\phi\right). $$ It follows that (\ref{induced_homom}) is the homomorphism $\widehat{\Psi}$ induced by $f$ and $\psi$. \end{proof} \section{Relative Completion of the Fundamental Groupoid} In this section we explain how the fundamental groupoid of $X$ can be completed with respect to $\rho : \pi_1(X,x) \to S$ and we give a de~Rham construction of it. In the unipotent case, the de~Rham theorem is implicit in Chen's work \cite{chen}, and is described explicitly in \cite{hain-zucker}. Recall that the fundamental groupoid $\pi(X)$ of a topological space $X$ is the category whose objects are the points of $X$ and whose morphisms from $a \in X$ to $b \in X$ are homotopy classes $\pi(X;a,b)$ of paths $[0,1] \to X$ from $a$ to $b$. We can think of $\pi(X)$ as a torsor over $X\times X$; the fiber over $(a,b)$ being $\pi(X;a,b)$. Observe that there is a canonical isomorphism between the fiber over $(a,a)$ and $\pi_1(X,a)$. The torsor is the one over $X\times X$ corresponding to the representation \begin{equation}\label{action} \pi_1(X\times X,(a,a)) \cong \pi_1(X,a)\times \pi_1(X,a) \to \Aut \pi_1(X,a) \end{equation} where $$ (\gamma, \mu) \to \left\{g \mapsto \gamma^{-1}g \mu\right\}. $$ As in previous sections, $X$ will be a connected smooth manifold and $x_o$ a distinguished base point. Suppose, as before, that $\rho : \pi_1(X,x_o) \to S$ is a Zariski dense homomorphism to a reductive real algebraic group. Denote the completion of $\pi_1(X,x_o)$ relative to $\rho$ by $\pi_1(X,x_o) \to {\mathcal G}$. The representation (\ref{action}) extends to a representation $$ \pi_1(X\times X,(x_o,x_o)) \cong \pi_1(X,x_o)\times \pi_1(X,x_o) \to \Aut {\mathcal G} $$ Denote the corresponding torsor over $X\times X$ by $\boldsymbol{\G}$. This is easily seen to be a torsor of real proalgebraic varieties. Denote the fiber of $\boldsymbol{\G}$ over $(a,b)$ by ${\mathcal G}_{a,b}$. There is a canonical map $$ \pi(X;a,b) \to {\mathcal G}_{a,b} $$ which induces a map of torsors. It follows from standard arguments that, for all $a$, $b$ and $c$ in $X$, there is a morphism of proalgebraic varieties \begin{equation}\label{mult_map} {\mathcal G}_{a,b} \times {\mathcal G}_{b,c} \to {\mathcal G}_{a,c} \end{equation} which is compatible with the multiplication map $$ \pi(X;a,b) \times \pi(X;b,c) \to \pi(X;a,c). $$ An efficient way to summarize the properties of $\boldsymbol{\G}$ and the multiplication maps is to say that they form a category (in fact, a groupoid) whose objects are the elements of $X$ and where $\Hom(a,b)$ is ${\mathcal G}_{a,b}$ with composition defined by (\ref{mult_map}). In addition, the natural map $\pi(X;a,b) \rightsquigarrow {\mathcal G}_{a,b}$ from the fundamental groupoid of $X$ to this category is a functor. We shall call this functor the {\it relative completion of the fundamental groupoid of $X$ with respect to $\rho$}.\footnote{We shall see that the torsor $\boldsymbol{\G}$ is independent of the choice of base point $x_o$, so it may have been better to call $\boldsymbol{\G}$ the completion of the fundamental groupoid relative to the principal bundle $P$.} Our goal is to give a description of it in terms of differential forms. We also have the torsor $\boldsymbol{\cP}$ over $X\times X$ associated to the representation $$ \pi_1(X\times X,x_o) \cong \pi_1(X,x_o)\times \pi_1(X,x_o) \to \Aut S $$ where $$ (\gamma, \mu)\to\left\{g \mapsto\rho(\gamma)^{-1}g\rho(\mu)\right\}. $$ given by $\rho$. Denote the fiber of $\boldsymbol{\cP}$ over $(a,b)$ by ${\mathcal P}_{a,b}$. As above, we have a category whose objects are the points of $X$ and where $\Hom(a,b)$ is ${\mathcal P}_{a,b}$. There is also a functor from the fundamental groupoid of $X$ to this category which is the identity on objects. Denote the restriction of this torsor to $\{a\}\times X$ by $\boldsymbol{\cP}_{a,\underline{\blank}}$. We shall view this as a torsor over $(X,a)$. The image $\id_a$ of the identity in $\pi_1(X,a)$ in ${\mathcal P}_{a,a}$ gives a canonical lift of the base point $a$ of $X$ to $\boldsymbol{\cP}_{a,\underline{\blank}}$. Observe that $\boldsymbol{\cP}_{x_o,\underline{\blank}}$ is the principal $S$ bundle $P$ used in the construction of ${\mathcal G}$. For each $a \in X$, we can from the corresponding local system ${\mathcal O}(P_{a,\underline{\blank}})$ whose fiber over $b\in X$ is the coordinate ring of ${\mathcal P}_{a,b}$. We can form the complex $$ \Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}})) = \varinjlim E^{\bullet}(X,{\mathbb M}) $$ where ${\mathbb M}$ ranges over all finte dimensional sub-local systems of ${\mathcal O}({\mathcal P}_{a,\underline{\blank}})$. This has augmentations $$ \epsilon_{a} : \Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}})) \to {\mathbb R} $$ and $$ \delta_{a,b} : \Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}})) \to {\mathcal O}({\mathcal P}_{a,b}) $$ given by evaluation at $\id_a$ and on the fiber over $b$, respectively. We view ${\mathbb R}$ as a right $\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))$ module via $\epsilon_a$ and ${\mathcal O}({\mathcal P}_{a,b})$ as a left $\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))$ module via $\delta_{a,b}$. We can therefore form the two sided bar construction $$ B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{\id_a,(b)}) := B({\mathbb R},\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}})),{\mathcal O}({\mathcal P}_{a,b})) $$ Define \begin{multline}\label{comult} B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{\id_a,(c)}) \\ \to B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{\id_a,(b)}) \otimes B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{b,\underline{\blank}}))_{\id_b,(c)}) \end{multline} by \begin{multline*} \Delta : [w_1|\dots|w_r]\phi \mapsto \cr \sum_i [w_1|\dots |w_i] \left(\sum \psi_i^{(k_i)}\dots\psi_r^{(k_r)}\phi'\right) \otimes [w_{i+1}^{(k_i)}|\dots|w_r^{(k_r)}]\left(\sum \phi''\right) \end{multline*} where $$ \Delta_S(\phi) = \sum \phi'\otimes \phi'' $$ is the map ${\mathcal O}(P_{a,c})\to {\mathcal O}(P_{a,b})\otimes{\mathcal O}(P_{b,c})$ dual to the multiplication map $P_{a,b}\times P_{b,c}\to P_{a,c}$; and the map $$ \Efin^{\bullet}(X,{\mathcal O}(P_{a,\underline{\blank}})) \to {\mathcal O}(P_{a,b}) \otimes \Efin^1(X,{\mathcal O}(P_{b,\underline{\blank}})) $$ induced by multiplication $P_{a,b} \times \boldsymbol{\cP}_{b,\underline{\blank}} \to \boldsymbol{\cP}_{a,\underline{\blank}}$ takes $w_j$ to $$ \sum \psi_j^{(k_j)} \otimes w_j^{(k_j)}. $$ Definition~\ref{defn} generalizes: \begin{definition} For $\gamma$ a path in $X$ from $a$ to $b$, $\phi\in {\mathcal O}(P_{a,b})$ and $w_1,\dots,w_r$ elements of $\Efin^1(X,{\mathcal O}(P_{a,\underline{\blank}}))$, we define $$ \int_\gamma \left(w_1 \dots w_r | \phi\right) = \phi({\tilde{\gamma}}(1))\int_{{\tilde{\gamma}}}w_1\dots w_r $$ where ${\tilde{\gamma}}$ is the unique lift of $\gamma$ to a horizontal section of $\boldsymbol{\cP}_{a,\underline{\blank}}$ which begins at $\id_a \in P_{a,a}$. \end{definition} There is an analogous extension of the definition of higher iterated integrals (\ref{higherdef}) to this situation. As in that case, one has a d.g.\ algebra homomorphism $$ B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{\id_a,(b)}) \to E^{\bullet}(P_{a,b}X) $$ to the de~Rham complex of $P_{a,b}X$, the space of paths in $X$ from $a$ to $b$. It is defined by $$ [w_1|\dots|w_r]\phi \mapsto \int(w_1\dots w_r|\phi). $$ By taking a homotopy class $\gamma \in \pi(X;a,b)$ to the ideal of functions that vanish on it, we obtain a function $$ \pi(X;a,b) \to \spec H^0(B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{\id_a,(b)})). $$ \begin{theorem}\label{gpoid_dr} This function gives a natural algebra isomorphism $$ {\mathcal O}({\mathcal G}_{a,b}) \cong H^0(B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{a,(b)})). $$ Moreover, the map $$ {\mathcal O}({\mathcal G}_{a,c}) \to {\mathcal O}({\mathcal G}_{a,b})\otimes {\mathcal O}({\mathcal G}_{b,c}) $$ induced by (\ref{mult_map}) corresponds to (\ref{comult}) under this isomorphism. \end{theorem} \begin{proof}[Sketch of Proof] Define $\Gdr_{a,b}$ by $$ \Gdr_{a,b} = \spec H^0(B(\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))_{a,(b)})). $$ The coproduct above induces morphisms $$ \Gdr_{a,b} \times \Gdr_{b,c} \to \Gdr_{a,c}. $$ We therefore have a groupoid whose objects are the points of $X$ and where $\Hom(a,b)$ is $\Gdr_{a,b}$ and a function $$ \pi(X;a,b) \to \Gdr_{a,b}. $$ This map is easily seen to be compatible with path multiplication (use the generalization of the last property of (\ref{props})), and therefore a functor of groupoids. Since $X$ is connected, it suffices to prove that $\Gdr_{a,b}$ is isomorphic to ${\mathcal G}_{a,b}$ for just one pair of points $a,b$ of $X$. But these are isomorphic in the case $a=b=x_o$ by Theorem~\ref{main}. \end{proof} \section{Hodge Theory} \label{hodge_str} Now suppose that $X$ is a smooth complex algebraic variety (or the complement of a normal crossings divisor in a compact K\"ahler manifold) and that ${\mathbb V}$ is an admissible variation of Hodge structure over $X$. Denote the semisimple group associated to the fiber $V_o$ over the base point $x_o\in X$ by $S$. This is the ``orthogonal'' group $$ S = \Aut(V_o,\langle\phantom{x},\phantom{x}\rangle) $$ associated to the polarization $\langle\phantom{x},\phantom{x}\rangle$. It is semi-simple. Suppose that the image of the monodromy representation $$ \rho : \pi_1(X,x_o) \to S $$ is Zariski dense. Denote the completion of $\pi_1(X,x_o)$ relative to $\rho$ by $$ \tilde{\rho} : \pi_1(X,x_o) \to {\mathcal G}(X,x_o). $$ \begin{theorem}\label{hodge} Under these assumptions, the coordinate ring ${\mathcal O}({\mathcal G}(X,x_o))$ of the completion of $\pi_1(X,x_o)$ with respect to $\rho$ has a canonical real mixed Hodge structure with weights $\ge 0$ for which the product, coproduct, antipode and the natural inclusion $$ {\mathcal O}(S) \hookrightarrow {\mathcal O}({\mathcal G}(X,x_o)) $$ are all morphisms of mixed Hodge structure. Moreover the canonical homomorphism ${\mathcal G}(X,x_o) \to S$ induces an isomorphism $\Gr^W_0{\mathcal O}({\mathcal G}(X,x_o)) \cong {\mathcal O}(S)$. \end{theorem} Denote the Lie algebra of $S$ by ${\mathfrak s}$. This has a canonical Hodge structure of weight 0. The following result is an important corollary of the proof of Theorem~% \ref{hodge}. It follows immediately from the theorem and the standard description of the Lie algebra of an affine algebraic group given at the end of Section~\ref{coord}. \begin{corollary}\label{hodge-lie} Under the assumptions of the theorem, the Lie algebra ${\mathfrak g}(X,x_o)$ of ${\mathcal G}(X,x_o)$ has a canonical MHS with weights $\le 0$, and the homomorphism ${\mathfrak g}(X,x_o) \to {\mathfrak s}$ is a morphism of MHS which induces an isomorphism $$ \Gr^W_0 {\mathfrak g} \cong {\mathfrak s}. $$ In particular, there is a canonical MHS with weights $< 0$ on ${\mathfrak u}(X,x_o)$, the Lie algebra of the prounipotent radical of ${\mathcal G}(X,x_o)$. \qed \end{corollary} The principal assertion of Theorem~\ref{hodge} is a special case of the following result when $a=b=c=x_o$. \begin{theorem}\label{groupoid} With $X$, ${\mathbb V}$ and $S$ as above, if $a,\, b\in X$, then the coordinate ring ${\mathcal O}({\mathcal G}_{a,b})$ of the completion of $\pi(X;a,b)$ relative to $\rho$ has a canonical mixed Hodge structure with weight $\ge 0$ and whose multiplication is a morphism of MHS. If $a$, $b$ and $c$ are three points of $X$, then the map $$ {\mathcal O}({\mathcal G}_{a,c}) \to {\mathcal O}({\mathcal G}_{a,b}) \otimes {\mathcal O}({\mathcal G}_{b,c}) $$ induced by path multiplication is a morphism of MHS. Moreover, the mixed Hodge structure on ${\mathcal O}({\mathcal G}_{a,b})$ depends only on the variation ${\mathbb V}$ and not on the choice of the base point $x_o$. \end{theorem} Because of the last assertion, it may be more appropriate to say that {\it ${\mathcal G}_{a,b}$ is the completion of $\pi(X;a,b)$ with respect to the variation ${\mathbb V}$.} The reader is assumed to be familiar with the basic methods for constructing mixed Hodge structures on the cohomology of bar constructions as described in \cite[\S3]{hain:dht}. In the previous section we showed how to express ${\mathcal O}({\mathcal G}_{a,b})$ as the 0th cohomology group of a suitable reduced bar construction. So in order to show that it has a canonical MHS we need only find a suitable augmented, multiplicative mixed Hodge complex $\mathbf{A}^{\bullet}$ which is quasi-isomorphic to $\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))$. To do this, we shall use the work of M.~Saito on Hodge modules. First, some notation: Assume that $X = {\overline{X}} - D$, where ${\overline{X}}$ is a compact K\"ahler manifold and $D$ is a normal crossings divisor. Denote the inclusion $X \hookrightarrow {\overline{X}}$ by $j$. Denote Deligne's canonical extension of ${\mathbb V}\otimes{\mathcal O}_X$ to ${\overline{X}}$ by $\overline{\mathcal{V}}$. Saito proves that there is a Hodge module over ${\overline{X}}$ canonically associated to ${\mathbb V}$, whose complex part is a bifiltered $D$-module $(M,W_{\bullet},F^{\bullet})$, and whose real part is $Rj_\ast{\mathbb V}_{\mathbb R}$ endowed with a suitable weight filtration. There is a canonical inclusion $$ \Omega^{\bullet}_{\overline{X}}({\overline{X}}\log D)\otimes_{\mathcal O} \overline{\mathcal{V}} \hookrightarrow M\otimes_{\mathcal O} \Omega_{\overline{X}}^{\bullet}. $$ Saito defines Hodge and weight filtrations on $\Omega^{\bullet}_{\overline{X}}({\overline{X}}\log D)\otimes_{\mathcal O} \overline{\mathcal{V}}$ by restricting those of $M$. The Hodge filtration is simply the tensor product of those of $\Omega^{\bullet}_{\overline{X}}({\overline{X}}\log D)$ and $\overline{\mathcal{V}}$. The weight filtration is more difficult to describe. \begin{theorem}[Saito \protect{\cite[(3.3)]{saito}}]\label{saito:mhc} The pair \begin{equation}\label{mhc} M^{\bullet}(X,{\mathbb V}) := ((Rj_\ast{\mathbb V}_{\mathbb R}, W_{\bullet}), (\Omega^{\bullet}_{\overline{X}}({\overline{X}}\log D)\otimes_{\mathcal O} \overline{\mathcal{V}}, F^{\bullet}, W_{\bullet})) \end{equation} is a cohomological mixed Hodge complex whose cohomology is canonically isomorphic to $H^{\bullet}(X,{\mathbb V})$. \qed \end{theorem} We can therefore obtain a mixed Hodge complex (MHC) which computes $H^{\bullet}(X,{\mathbb V})$ by taking the standard fine resolution of these sheaves by $C^\infty$ forms. (So the complex part of this will be the $C^\infty$ log complex $E^{\bullet}({\overline{X}}\log D,\overline{\mathcal{V}})$ with suitable Hodge and weight filtrations.) To apply Saito's machinery to the current situation, we will need to know that ${\mathcal O}({\mathcal P}_{x_o,\underline{\blank}})$ is a direct limit of admissible variations over $X$. \begin{lemma}\label{loc_sys} The local system associated with an irreducible representation of $S({\mathbb R})$ underlies an admissible variation of Hodge structure over $X$. These structures are compatible with the decomposition of tensor products. Moreover, these variations are unique up to Tate twist. \end{lemma} \begin{proof} The connected component of the identity of $S({\mathbb R})$ is a real form of $Sp_n({\mathbb C})$ when ${\mathbb V}$ has odd weight, and $SO_n({\mathbb C})$ when ${\mathbb V}$ has even weight. In both cases each irreducible representations of the complex group can be constructed by applying a suitable Schur functor the the fundamental representation and then taking the intersection of the kernels of all contractions with the polarization. (This is Weyl's construction of the irreducible representations; it is explained, for example, in \cite[\S17.3,\S19.2]{fulton-harris}.) Since these operations preserve variations of Hodge structure, it follows that a local system corresponding to an irreducible representation of $S$ underlies a variation of Hodge structure. Since the monodromy representation of ${\mathbb V}$ is Zariski dense, the structure of a polarized variation of Hodge structure on this local system is unique up to Tate twist. (Cf.\ the proof of \cite[(9.1)]{hain:normal}.) \end{proof} This, combined with (\ref{decomp}) yields: \begin{corollary} With our assumptions, ${\mathcal O}({\mathcal P}_{x_o,\underline{\blank}})$ is a direct limit of admissible variations of Hodge structure over $X$ of weight 0, and the multiplication map is a morphism. \qed \end{corollary} \begin{corollary} For each $b\in X$, there is a canonical Hodge structure on ${\mathcal O}(P_{x_o,b})$. \qed \end{corollary} Combining (\ref{loc_sys}) with (\ref{h1_comp}), we obtain: \begin{corollary}\label{varmhs} The local system over $X$ whose fiber over $x\in X$ is $H^1({\mathcal U}(X,x))$ is an admissible variation of mixed Hodge structure whose weights are positive. \end{corollary} Using Saito's machine \cite{saito}, we see that there is a MHC $\mathbf{A}^{\bullet}$ which is quasi-isomorphic to $\Efin^{\bullet}(X,{\mathcal O}({\mathcal P}_{a,\underline{\blank}}))$. The complex part of this MHC is simply the complex of $C^\infty$ forms on ${\overline{X}}$ with logarithmic singularities along $D$ and which have coefficients in the canonical extension $\overline{\O}$ of ${\mathcal O}({\mathcal P}_{a,\underline{\blank}})$ to ${\overline{X}}$. The Hodge filtration is the obvious one inducedd by the Hodge filtration of $\overline{\O}$ and the Hodge filtration of forms on ${\overline{X}}$. The weight filtration is not so easily described, and we refer to Saito's paper for that. We need to know that the multiplication is compatible with the Hodge and weight filtrations. This follows from the next result. \begin{proposition} If ${\mathbb V}_1\otimes{\mathbb V}_2 \to {\mathbb W}$ is a pairing of admissible variations of Hodge structure over $X$, then the multiplication map $$ M^{\bullet}(X,{\mathbb V}_1) \otimes M^{\bullet}(X,{\mathbb V}_2) \to M^{\bullet}(X,{\mathbb W}) $$ is a morphism of cohomological mixed Hodge complexes. \end{proposition} \begin{proof} This follows immediately from the naturality of Saito's construction, its compatibility with exterior tensor products, and the fact that admissible variations of Hodge structure are closed under exterior products --- use restriction to the diagonal. \end{proof} There are two augmentations $$ {\mathbb R} \leftarrow \mathbf{A}^{\bullet} \to {\mathcal O}(P_{a,b}) $$ corresponding to the inclusions $P_{a,b}\hookrightarrow {\mathcal P}_{a,\underline{\blank}}$ and $\id_a \in P_{a,a}$, and these are compatible with all filtrations. It follows from \cite[(3.2.1)]{hain:dht}, (\ref{qism}) and (\ref{gpoid_dr}) that $$ B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(P_{a,b})) $$ is a MHC whose $H^0$ is isomorphic to ${\mathcal O}({\mathcal G}_{a,b})$. Moreover, the multiplication is compatible with the Hodge and weight filtrations. Consequently, $$ {\mathcal O}({\mathcal G}_{a,b}) \cong H^0(B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(P_{a,b}))) $$ has a canonical MHS and its multiplication is a morphism of MHS. Since the MHS on $P_{a,b}$ depends only on ${\mathbb V}$ and not on $x_o$, the same is true of the MHS on ${\mathcal O}({\mathcal G}_{a,b})$. If $a$, $b$, and $c$ are three points of $X$, then it follows directly from the definitions that the map $$ B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(P_{a,c})) \to B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(P_{a,b}))\otimes B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(P_{b,c})) $$ corresponding to path multiplication is a morphism of MHCs. It follows that the induced map $$ {\mathcal O}({\mathcal G}_{a,c}) \to {\mathcal O}({\mathcal G}_{a,b}) \otimes {\mathcal O}({\mathcal G}_{b,c}) $$ is a morphism of MHS. This completes the proof of Theorem~\ref{groupoid}; Theorem~\ref{hodge} follows by taking $a=b=x_o$ except for the assertion that ${\mathcal O}(S) \hookrightarrow {\mathcal O}({\mathcal G})$ is a morphism of MHS. This follows as this is induced by the natural inclusion $$ {\mathcal O}(S) \hookrightarrow B({\mathbb R},\mathbf{A}^{\bullet},{\mathcal O}(S)), $$ that takes $\phi$ to $[\phantom{x}]\phi$. It is a morphism of MHCs. This completes the proofs of Theorems~\ref{hodge} and \ref{groupoid}. \qed We now turn our attention to the variation of the Hodge filtration. Suppose that $X$ and $S$ are as above. Consider the real local system over $X\times X$ whose fiber over $(a,b)$ is ${\mathcal O}({\mathcal G}_{a,b})$. Denote it by $\boldsymbol{\O}$. Next we establish that this underlies a ``pre-variation of MHS.'' Denote by ${\mathcal F}^p\boldsymbol{\O}$the subset of the associated complex local system with fiber $F^p{\mathcal O}({\mathcal G}_{a,b}({\mathbb C}))$ over $(a,b)$. Denote by $W_m\boldsymbol{\O}$ the subset of $\boldsymbol{\O}$ with fiber $W_m{\mathcal O}({\mathcal G}_{a,b})$ over $b$. \begin{theorem}\label{pre_var} The subset $W_m\boldsymbol{\O}$ is a flat sub-bundle of $\boldsymbol{\O}$, and $F^p\boldsymbol{\O}$ is a holomorphic sub-bundle of $\boldsymbol{\O}_{\mathbb C}$ whose corresponding sheaf of sections ${\mathcal F}^p$ satisfies Griffiths transversality: $$ \nabla : {\mathcal F}^p \to {\mathcal F}^{p-1}\otimes \Omega^1_X. $$ \end{theorem} \begin{proof}[Sketch of Proof] We will prove the result for the restriction $\boldsymbol{\O}_a$ of $\boldsymbol{\O}$ to $\{a\}\times X$. The result for the restriction of $\boldsymbol{\O}$ to $X\times \{b\}$ is proved similarly. The general result then follows as the tangent spaces of $\{a\}\times X$ and $X\times \{b\}$ span the tangent space of $X\times X$ at $(a,b)$. First we need a formula for the connection on $\boldsymbol{\O}_a$ at the point $b$ in $X$. Fix a path $\gamma$ in $X$ from $a$ to $b$. Suppose that $\mu : [-\epsilon, \epsilon] \to X$ is a smooth path with $\mu(0) = b$. For $s\in [-\epsilon, \epsilon]$ let $\gamma_s : [0,1] \to X$ be the piecewise smooth path obtained by following $\gamma$ and then $\mu$ from $t=0$ to $t=s$. Suppose that $w_1,\dots, w_r$ are in $\Efin^{\bullet}(X,{\mathcal O}(P_{a,b}))$. Suppose that $U$ is a contractible neighbourhood of $b$ in $X$. With respect to a flat trivialization of the restriction of ${\mathcal P}_{a,\underline{\blank}}$ to $U$, we have $$ w_r|_U = \sum_j w_r^j\otimes \psi_j $$ where $w_r^j\in E^1(U)$ and $\psi_j \in {\mathcal O}(P_{a,b})$. It follows from the analogue of (\ref{props}) in this situation that $$ \frac{d}{ds}\bigg\vert_{s=0} \int_{\gamma_s}(w_1\dots w_r|\phi) = \sum_j \int_\gamma (w_1\dots w_{r-1}|\phi \psi_j) \langle w_r^j,\dot{\mu}(0)\rangle. $$ The restriction of the connection on $\boldsymbol{\O}_a$ to the stalk at $b$ is therefore induced by the map $$ [w_1|\dots |w_r]\phi \mapsto \sum_j [w_1|\dots |w_{r-1}]\phi \psi_j \otimes w_r^j $$ on the bar construction. The flatness of the weight filtration follows immediately from the definition of the weight filtration on the bar construction. Further, if $(z_1,\dots, z_n)$ is a holomorphic coordinate in $X$ centered at $b$, then it follows immediately from the definition of the Hodge and weight filtrations on ${\mathcal O}({\mathcal G}_{a,b}({\mathbb C}))$ and the formula for the connection that $$ \nabla_{\partial/\partial {\overline{z}}_k} : {\mathcal F}^p \to {\mathcal F}^p $$ for each $k$, so that the Hodge filtration varies holomorphically at $b$. Similarly, on the stalk of $F^p$ at $b$ we have $$ \nabla_{\partial/\partial z_k} : {\mathcal F}^p \to {\mathcal F}^{p-1} $$ as each $w_r^j$ contributes at most 1 to the Hodge filtration of ${\mathcal O}({\mathcal G}_{a,b})$. \end{proof} When $X$ is compact, we have: \begin{corollary}\label{good_varn} If $X$ is a compact K\"ahler manifold, then $\boldsymbol{\O}$ is an admissible variation of MHS over $X\times X$. \qed \end{corollary} In order to prove the corresponding result in the non-compact case, it is necessary to study the asymptotic behaviour of $\boldsymbol{\O}$. I plan to consider this in a future paper. We now consider naturality. Suppose that $X$ and ${\mathbb V}$ are as above, and that $Y$ is a smooth variety and that ${\mathbb W}$ is an admissible variations of Hodge structure over $Y$. We will now denote $S$ by $S_X$: $$ S_X = \Aut(V_o,\langle\phantom{x},\phantom{x}\rangle). $$ Set $$ S_Y = \Aut(W_o,\langle\phantom{x},\phantom{x}\rangle) $$ where $W_o$ denotes the fiber of ${\mathbb W}$ over $y_o$. Suppose that the monodromy representation $$ \rho_Y : \pi_1(Y,y_o) \to S_Y $$ has Zariski dense image. Denote the completion of $\pi_1(X,x_o)$ with respect to $\rho_X:\pi_1(X,x_o) \to S_X$ by $\pi_1(X,x_o)\to {\mathcal G}_X$, and the completion of $\pi_1(Y,y_o)$ with respect to $\rho_Y$ by $\pi_1(Y,y_o) \to {\mathcal G}_Y$. Suppose that $f:(Y,y_o) \to (X,x_o)$ is a morphism of varieties, and that we have fixed an inclusion $$ \End {\mathbb V} \hookrightarrow \End f^\ast {\mathbb W} $$ of variations of Hodge structure. This fixes a group homomorphism $$ \Psi : S_X \hookrightarrow S_Y $$ that is compatible with the Hodge theory. By (\ref{induced}), there is a homomorphism $\widehat{\Psi} : {\mathcal G}_X \to {\mathcal G}_Y$ such that the diagram $$ \begin{CD} \pi_1(X,x_o) @>>> {\mathcal G}_X @>>> S_X \cr @V{f_\ast}VV @VV{\widehat{\Psi}}V @VV{\Psi}V \cr \pi_1(Y,y_o) @>>> {\mathcal G}_Y @>>> S_Y \cr \end{CD} $$ commutes. \begin{theorem}\label{hodge_nat} Under these hypotheses, the induced map $$ \widehat{\Psi}^\ast : {\mathcal O}({\mathcal G}_Y) \to {\mathcal O}({\mathcal G}_X) $$ is a morphism of MHS. \end{theorem} \begin{proof} First, choose smooth compactifications ${\overline{X}}$ of $X$ and ${\overline{Y}}$ of $Y$ such that $X={\overline{X}} - D$ and $Y = {\overline{Y}} - E$, where $D$ and $E$ are normal crossings divisors. We may choose these such that $f$ extends to a morphism ${\overline{X}} \to {\overline{Y}}$, which we shall also denote by $f$. Denote by $P_X$ the flat left $S_X$ principal bundle over $X$ associated the the representation of $\pi_1(X,x_o)$ on $S_X$ via $\rho_X$. Denote the analogous principal $S_Y$ principal bundle over $Y$ by $P_Y$. Associated to these we have the local systems ${\mathcal O}(P_X)$ over $X$ and ${\mathcal O}(P_Y)$ over $Y$. The construction above gives multiplicative mixed Hodge complexes $$ \mathbf{A}^{\bullet}(X,{\mathcal O}(P_X)),\quad \mathbf{A}^{\bullet}(X,f^\ast{\mathcal O}(P_Y)), \text{ and } \mathbf{A}^{\bullet}(Y,{\mathcal O}(P_Y)) $$ which compute the canonical mixed Hodge structures on $$ H^{\bullet}(X,{\mathcal O}(P_X)),\quad H^{\bullet}(X,f^\ast{\mathcal O}(P_Y)),\text{ and } H^{\bullet}(X,f^\ast{\mathcal O}(P_Y)) $$ respectively. The map $f$ induces a morphism $$ \mathbf{A}^{\bullet}(Y,{\mathcal O}(P_Y)) \to \mathbf{A}^{\bullet}(X,f^\ast{\mathcal O}(P_Y)) $$ of MHCs, while the inclusion $S_X \hookrightarrow S_Y$ induces a morphism $$ \mathbf{A}^{\bullet}(X,f^\ast{\mathcal O}(P_Y)) \to \mathbf{A}^{\bullet}(X,{\mathcal O}(P_X)) $$ of MHCs. The composition of these corresponds to the induced map $$ \Efin^{\bullet}(Y,{\mathcal O}(P_Y)) \to \Efin^{\bullet}(X,{\mathcal O}(P_X)) $$ induced by $f$ under the canonical quasi-isomorphisms. It follows that the induced map $$ B({\mathbb R},\mathbf{A}^{\bullet}(Y,{\mathcal O}(P_Y)),{\mathcal O}(P_Y)) \to B({\mathbb R},\mathbf{A}^{\bullet}(Y,{\mathcal O}(P_X)),{\mathcal O}(P_X)) $$ is a morphism of MHCs and that the induced the map $$ f^\ast : {\mathcal O}({\mathcal G}_Y) \to {\mathcal O}({\mathcal G}_X) $$ on $H^0$ is the ring homomorphism induced by $f$. The result follows. \end{proof} \begin{remark}\label{extended} Suppose that ${\mathbb V}$ is an admissible variation of Hodge structure over the complement $X$ of a normal crossings divisor in a compact K\"ahler manifold. We will say that the pair $(X,{\mathbb V})$ is {\it neat} if the Zariski closure $S$ of the image of the monodromy map $$ \rho : \pi_1(X,x_o) \to \Aut(V_o,\langle\phantom{x},\phantom{x}\rangle) $$ is semi-simple, and that the canonical MHS on the coordinate ring of $$ \Aut(V_o,\langle\phantom{x},\phantom{x}\rangle) $$ induces one on $S$. For example, every variation where $S$ is finite is neat. I believe that every admissible $(X,{\mathbb V})$ is neat, but have not yet found a proof. The results (\ref{hodge}), (\ref{hodge-lie}), (\ref{groupoid}), (\ref{varmhs}), (\ref{pre_var}), (\ref{good_varn}), (\ref{hodge_nat}) and their proofs are valid with the assumption that $\im \rho$ be Zariski dense in $\Aut(V_o,\langle\phantom{x},\phantom{x}\rangle)$ replaced by the assumption that the pairs $(X,{\mathbb V})$ and $(Y,{\mathbb W})$ be neat. \end{remark} The following is an application suggested by Ludmil Kartzarkov. \begin{theorem} Suppose that $X$ is a compact K\"ahler manifold and that ${\mathbb V}$ is a polarized variation of Hodge structure over $X$ whose monodromy representation $\rho$ has Zariski dense image. Then the prounipotent radical of the completion of $\pi_1(X,x_o)$ relative to $\rho$ has a quadratic presentation. \end{theorem} \begin{proof} It is well known that if $X$ is compact K\"ahler and ${\mathbb V}$ is a polarized variation of Hodge structure over $X$ of weight $m$, then $H^k(X,{\mathbb V})$ has a pure Hodge structure of weight $k+m$. In particular, as the variation ${\mathcal O}(P)$ over $X$ has weight zero, $H^k(X,{\mathcal O}(P))$ is pure of weight weight $k$ for all $k$. Denote the Lie algebra of the prounipotent radical of the relative completion of $\pi_1(X,x_o)$ by ${\mathfrak u}$. It follows from (\ref{def}) and (\ref{main}) that ${\mathfrak u}$ is the Lie algebra canonically associated to the d.g.a.\ $\Efin^{\bullet}(X,{\mathcal O}(P))$ by rational homotopy theory (either via Sullivan's theory of minimal models, or via Chen's theory as the dual of the indecomposables of the bar construction on $\Efin^{\bullet}(X,{\mathcal O}(P))$.) There are canonical maps $$ H^1({\mathfrak u}) \cong H^1(X,{\mathcal O}(P))\text{ and } H^2({\mathfrak u}) \hookrightarrow H^2(X,{\mathcal O}(P)). $$ These are morphisms of MHS \cite[(7.2)]{hain:torelli}. It follows that $H^1({\mathfrak u})$ is pure of weight 1 and $H^2({\mathfrak u})$ is pure of weight 2. The result now follows from \cite[(5.2),(5.7)]{hain:torelli}. \end{proof} \begin{remark} It is not necessarily true that ${\mathfrak u}$ is a quotient of the unipotent completion of $\ker \rho$. A criterion for surjectivity is given in \cite[(4.6)]{hain:comp}. For this reason it may not be easy to apply this result in general situations without artificially restrictive hypotheses. \end{remark} \section{A Canonical Connection} \label{connection} For the time being, let $X$, $x_o$, ${\mathbb V}$, $\rho$, etc.\ be as in the previous section. However, all groups and Lie algebras in this section will be complex, and ${\mathcal G}$, ${\mathcal U}$, ${\mathfrak u}$, etc.\ denote the {\em complex} points of the relative completion of $\pi_1(X,x_o)$, its prounipotent radical, its Lie algebra, etc. Denote the image of $\rho$ by $\Gamma$, and the Galois covering of $X$ with Galois group $\Gamma$ by $X'$. In this section, we show how the Hodge theory of ${\mathcal G}$ gives a canonical (given the choice of $x_o$), $\Gamma$ invariant integrable 1-form $$ \omega \in E^1(X')\comptensor \Gr^W_{\bullet}{\mathfrak u} $$ on $X'$ which can be integrated to the canonical representation $$ \tilde{\rho} : \pi_1(X,x_o) \to S \ltimes {\mathcal U} \cong {\mathcal G}. $$ Here $\comptensor$ denotes the completed tensor product, which is defined below. At the end of the section, we shall explain what this means when $X$ is the complement of the discriminant locus in ${\mathbb C}^n$ and $S$ is the symmetric group $\Sigma_n$. In this case, $X'$ is the complement of the hyperplanes $x_i = x_j$ in ${\mathbb C}^n$, $\pi_1(X,x_o)$ is the classical braid group $B_n$, and the form is $$ \omega = \sum_{i<j} d\log(x_i-x_j). $$ First, we shall define the completed tensor product $\comptensor$. Suppose that ${\mathfrak u}$ is a topological Lie algebra and that $$ {\mathfrak u} = {\mathfrak u}^1 \supseteq {\mathfrak u}^2 \supseteq {\mathfrak u}^3 \supseteq \cdots $$ is a base of neighbourhoods of 0. Suppose that $E$ is a vector space. Define $$ E\comptensor {\mathfrak u} = \lim_{\leftarrow} E\otimes {\mathfrak u}/{\mathfrak u}^m. $$ We can regard a graded Lie algebra ${\mathfrak u} = \oplus_{m < 0} {\mathfrak u}_m$ as a topological Lie algebra where the basic neighbourhoods of 0 are $$ \bigoplus_{l\le m} {\mathfrak u}_l, \quad m < 0. $$ The definition of completed tensor product therefore extends to the case where ${\mathfrak u}$ is graded. Finally, if ${\mathfrak u}$ is a Lie algebra in the category of mixed Hodge structures where ${\mathfrak u} = W_{-1}{\mathfrak u}$ which is complete with respect to the weight filtration, and if $E$ is a complex vector space, then there is a canonical isomorphism $$ E\comptensor {\mathfrak u}_{\mathbb C} \cong E \comptensor \Gr^W_{\bullet} {\mathfrak u}_{\mathbb C} $$ as ${\mathfrak u}_{\mathbb C}$ is canonically isomorphic to $\prod \Gr^W_m{\mathfrak u}_{\mathbb C}$. (Cf.\ \cite[(5.2)]{hain:torelli}.) We view a principal bundles with structure group a proalgebraic group to be the inverse limit of the principal bundles whose structure groups are the finite dimensional quotients of the proalgebraic group. A connection on a principal bundle with proalgebraic structure group is the inverse limit of compatible connections on the corresponding bundles with finite dimensional structure group. \subsection{The unipotent case.} We begin with the unipotent case, $S=1$. Suppose that $X$ is a smooth manifold with distinguished base point $x_o$. Denote the complex form of the unipotent completion of $\pi_1(X,x_o)$ by ${\mathcal G}$ and the complex points of the completion of $\pi(X;x_o,x)$ by ${\mathcal G}_{x_o,x}$. The family $$ \left({\mathcal G}_{x_o,x}\right)_{x\in X} $$ forms a flat principal left ${\mathcal G}$ bundle over $X$ that we shall denote by ${\mathcal G}_{x_o,\underline{\blank}}$. Since the structure group is contractible, (more precisely, an inverse limit of contractible groups), this bundle has a section. Pulling back the canonical connection form, we obtain an integrable connection form $$ \omega \in E^{\bullet}(X)\comptensor {\mathfrak g} $$ where ${\mathfrak g}$ denotes the Lie algebra of ${\mathcal G}$. The monodromy representation of this form is the monodromy of ${\mathcal G}_{x_o,\underline{\blank}}$, which is the canonical homomorphism $$ \pi_1(X,x_o) \to {\mathcal G}. $$ When $X$ is an algebraic manifold, there is a canonical choice of section and therefore a canonical connection form. To see this, note that for each $a \in X$, the weights on ${\mathcal O}({\mathcal G}_{x_o,a})$ are $\ge 0$ and that $$ \Gr^W_0 {\mathcal O}({\mathcal G}_{x_o,a}) \cong {\mathbb C}. $$ Since there is a canonical ring isomorphism $$ {\mathcal O}({\mathcal G}_{x_o,a}) \cong \bigoplus_{l\ge 0} \Gr^W_l {\mathcal O}({\mathcal G}_{x_o,a}) $$ there is a canonical augmentation $$ {\mathcal O}({\mathcal G}_{x_o,a}) \to {\mathbb C} $$ whose kernel is $$ \bigoplus_{l > 0} \Gr^W_l {\mathcal O}({\mathcal G}_{x_o,a}). $$ This determines a canonical point in ${\mathcal G}_{x_o,a}$. Since the family $\left\{{\mathcal O}({\mathcal G}_{x_o,a})\right\}_{a\in X}$ is a variation of MHS over $X$ (see \cite{hain-zucker}), these distinguished points vary smoothly as $a$ varies. They therefore determine a smooth section of ${\mathcal G}_{x_o,\underline{\blank}}$. We therefore have a canonical integrable 1-form $$ \omega \in E^1(X)\comptensor{\mathfrak g} \cong E^1(X)\comptensor\Gr^W_{\bullet}{\mathfrak g}. $$ \subsection{The general case} The first step in doing this in general is to explain the necessary constructions in the $C^\infty$ case. So suppose for the time being that $X$ is a smooth manifold; $\rho$, $S$, ${\mathcal G}$ and $P\to X$ are as before. We also have the torsor $$ {\mathcal G}_{x_o,\underline{\blank}} \to X. $$ It is a flat principal left ${\mathcal G}$ bundle over $X$. There is a map $$ \begin{CD} {\mathcal G}_{x_o,\underline{\blank}} @>\pi>> P \cr @VVV @VVV \cr X @= X \end{CD} $$ of flat bundles. It is compatible with the canonical homomorphism ${\mathcal G} \to S$. Denote by $X'$ the leaf of $P$ containing the distinguished lift $\tilde{x}_o$ to $P$ of $x_o$. The projection $P\to X$ induces a covering map $X' \to X$. It is the Galois covering corresponding to $\ker \rho$. Define ${\mathcal U}_{x_o,\underline{\blank}}$ to be the subset $\pi^{-1}X'$ of ${\mathcal G}_{x_o,\underline{\blank}}$. There is a natural projection ${\mathcal U}_{x_o,\underline{\blank}}\to X'$ indued by $\pi$. Note that the fiber of this over $\tilde{x}_o$ is ${\mathcal U}$, the prounipotent radical of ${\mathcal G}$. Denote the fiber of ${\mathcal U}_{x_o,\underline{\blank}}$ over $a\in X'$ by ${\mathcal U}_{x_o,a}$. Each point $a$ of $P$ determines an augmentation $$ \epsilon_a : \Efin^{\bullet}(X,{\mathcal O}(P)) \to {\mathbb C}. $$ Given two points $a$ and $b$ of $P$, we may form the two sided bar construction \begin{equation}\label{bar} B({\mathbb C},\Efin^{\bullet}(X,{\mathcal O}(P)),{\mathbb C}) \end{equation} where the left hand ${\mathbb C}$ is viewed as a module over $\Efin^{\bullet}(X,{\mathcal O}(P))$ via $\epsilon_a$, and the right hand ${\mathbb C}$ via $\epsilon_b$. We shall denote the d.g.a.\ (ref{bar}) by $B(\Efin^{\bullet}(X,{\mathcal O}(P))_{a,b})$ \begin{proposition} Each ${\mathcal U}_{x_o,a}$ is a proalgebraic variety with coordinate ring $$ {\mathcal O}({\mathcal U}_{x_o,a}) \cong H^0(B(\Efin^{\bullet}(X,{\mathcal O}(P))_{\tilde{x}_o,a})). $$ Moreover, ${\mathcal U}_{x_o,\underline{\blank}}\to X'$ is a principal ${\mathcal U}$ bundle with respect to the natural ${\mathcal U}$ action on ${\mathcal G}_{x_o,\underline{\blank}}$. \end{proposition} Choose a splitting $S\to {\mathcal G}$ of the natural homomorphism ${\mathcal G} \to S$. This induces an isomorphism ${\mathcal G}\cong S\ltimes {\mathcal U}$. The splitting enables us to lift the action of $S$ to ${\mathcal G}_{x_o,\underline{\blank}}$ in such a way that the projection ${\mathcal G}_{x_o,\underline{\blank}}\to P$ is $S$ equivariant. Since $\Gamma$ is a subgroup of $S$, and since it preserves $X'\subset P$, it follows that there is a natural left action of $\Gamma$ on ${\mathcal U}_{x_o,\underline{\blank}}$ and that, with respect to this action, the projection ${\mathcal U}_{x_o,\underline{\blank}}$ is $\Gamma$ equivariant. Denote the pullback of the extension $$ 1 \to {\mathcal U} \to {\mathcal G} \to S \to 1 $$ along $\Gamma \hookrightarrow S$ by ${\mathcal G}_\Gamma$. This is an extension $$ 1 \to {\mathcal U} \to {\mathcal G}_\Gamma \to \Gamma \to 1. $$ The splitting $S\to {\mathcal G}$ induces a splitting $\Gamma \to {\mathcal G}_\Gamma$, and therefore a semi-direct productu decomposition ${\mathcal G}_\Gamma \cong \Gamma \ltimes {\mathcal U}$. The image of the canonical homomorphism $\pi_1(X,x_o) \to {\mathcal G}$ lies in ${\mathcal G}_\Gamma$. The composite ${\mathcal U}_{x_o,\underline{\blank}} \to X' \to X$ is a flat principal left ${\mathcal G}_\Gamma$ bundle over $X$. The associated monodromy representation is the canonical homomorphism $\pi_1(X,x_o) \to {\mathcal G}_\Gamma$. The monodromy therefore induces the canonical homomorphism $$ \pi_1(X,x_o) \to {\mathcal G} \cong S \ltimes {\mathcal U}. $$ Next, we explain that the pullback of this bundle to $X'$ is trivial, and therefore given by an integrable 1-form. \begin{proposition} There is a $\Gamma$ equivariant section of ${\mathcal U}_{x_o,\underline{\blank}}\to X'$. \end{proposition} \begin{proof} The action of $\Gamma$ on $X'$ is free. It follows that the action of $\Gamma$ on ${\mathcal U}_{x_o,\underline{\blank}}$ is also free. Consequently, the square $$ \begin{CD} {\mathcal U}_{x_o,\underline{\blank}} @>>> \Gamma\backslash {\mathcal U}_{x_o,\underline{\blank}} \cr @VVV @VVV \cr X' @>>p> X \end{CD} $$ is a pullback square. Since the fibers of $\Gamma\backslash {\mathcal U}_{x_o,\underline{\blank}} \to X$ are connected, it has a $C^\infty$ section. This section pulls back to a $\Gamma$ invariant section of ${\mathcal U}_{x_o,\underline{\blank}} \to X'$. \end{proof} Let $\Gamma$ act on ${\mathcal U}$ on the left via the adjoint action: $$ Ad(\gamma) : u \mapsto \gamma u \gamma^{-1}. $$ Then $\Gamma$ acts on $X'\times {\mathcal U}$ on the left via the diagonal action. It follows from the previous result that the flat principal bundle ${\mathcal U}_{x_o,\underline{\blank}}\to X'$ has a $\Gamma$ invariant trivialization. We therefore have a connection form $$ \omega \in E^1(X')\comptensor {\mathfrak u}. $$ \begin{proposition} This connection form satisfies $\gamma^\ast \omega = Ad(\gamma)\omega$ for all $\gamma \in \Gamma$. \end{proposition} \begin{proof} Since the bundle is trivial, its (locally defined) sections can be identified with (locally defined) functions $X' \to {\mathcal U}$. Since $\Gamma$ preserves the connection, we see that for each $\gamma\in \Gamma$ the local section $u$ is flat if and only if the local section $(\gamma^{-1})^\ast Ad(\gamma)(u)$ is flat. That is, $Ad(\gamma)(u)$ is flat if and only if $\gamma^\ast u$ is flat. The result now follows from a standard and straight forward computation. \end{proof} Now suppose that $X$ is an algebraic manifold. We have to show that this construction can be made canonical. Note that, given the choice of the base point $x_o$, the only choices made in the construction of $\omega$ were the choice of a splitting of the homomorphism ${\mathcal G} \to S$, and the choice of a $\Gamma$ invariant section of ${\mathcal U}_{x_o,\underline{\blank}} \to X'$. We will now explain how Hodge theory provides canonical choices of both. It follows from (\ref{hodge-lie}) that $\Gr^W_0{\mathfrak g} \cong {\mathfrak s}$. Consequently, there is a canonical splitting of the homomorphism ${\mathfrak g} \to {\mathfrak s}$. This induces a canonical splitting of the homomorphism ${\mathcal G} \to S$, and therefore a canonical action of $\Gamma$ on ${\mathcal U}_{x_o,\underline{\blank}}$ and a canonical identification ${\mathcal G} \cong S \ltimes {\mathcal U}$. It remains to explain why there is a $\Gamma$ equivariant section of ${\mathcal U}_{x_o,\underline{\blank}}$. This is an elaboration of the argument in the unipotent case. For each $b\in X$, Hodge theory provides canonical ring isomorphisms $$ {\mathcal O}(G_{x_o,b}) \cong \bigoplus_{m\ge 0} Gr^W_m {\mathcal O}({\mathcal G}_{x_o,b}) $$ and $$ {\mathcal O}({\mathcal P}_{x_o,b}) \cong \Gr^W_0 {\mathcal O}({\mathcal G}_{x_o,b}). $$ Moreover, it follows from (\ref{pre_var}) that these identifications depend smoothly on $b$. Consequently, there is a smooth section $\sigma$ of the canonical projection ${\mathcal G}_{x_o,\underline{\blank}} \to {\mathcal P}_{x_o,\underline{\blank}}$. Restricting to $X'$, we obtain a canonical smooth section of the projection ${\mathcal U}_{x_o,\underline{\blank}} \to X'$. \begin{proposition} This section is $\Gamma$ equivariant. \end{proposition} \begin{proof} For each $x\in X$ we have the action ${\mathcal G}\times {\mathcal G}_{x_o,x}\to{\mathcal G}_{x_o,x}$. By (\ref{gpoid_dr}) the corresponding map of coordinate rings is a morphism of MHS. By the choice of splitting of ${\mathcal G} \to S$, the action of $S$ given by the splitting preserves the canonical isomorphism $$ {\mathcal O}({\mathcal G}_{x_o,x}) \cong \bigoplus_{l\ge 0}\Gr^W_l {\mathcal O}({\mathcal G}_{x_o,x}). $$ It follows that the section of ${\mathcal G}_{x_o,\underline{\blank}} \to {\mathcal P}_{x_o,\underline{\blank}}$ defined above is equivariant with respect to the left $S$ actions. It follows that the restriction of this section to $X'$ is $\Gamma$ equivariant. \end{proof} \begin{example} In this example, we take $X$ to be the complement in ${\mathbb C}^n$ of the universal discriminant locus. (View ${\mathbb C}^n$ as the space of monic polynomials of degree $n$.) Pick a base point $x_o$. The fundamental group of this space is the classical braid group. Denote the symmetric group on $n$ letters by $\Sigma_n$. There is a natural homomorphism $\rho : B_n \to \Sigma_n$. Denote the corresponding covering of $X$ by $\pi : X' \to X$. Its fundamental group is the pure braid group $P_n$. As is well known, $X'$ is the complement of the hyperplanes $x_i = x_j$ in ${\mathbb C}^n$ where $i\neq j$. The projection takes $(x_1,\dots,x_n)$ to the monic polynomial $\prod(T-x_j)$. The natural left action of $\Sigma_n$ on $X'$ is given by $$ \sigma : (x_1,\dots,x_n) \mapsto (x_{\sigma^{-1}(1)},\dots,x_{\sigma^{-1}(n)}). $$ The local system $\pi_\ast{\mathbb Q}_{X'}$ is an admissible variation of Hodge structure over $X$ of weight 0, rank $n$, and type $(0,0)$. The closure of the image of the monodromy is $\Sigma_n$, a semi-simple group. So we can apply Theorem~\ref{extended} to deduce the existence of a MHS on the relative completion, and the existence of a universal connection. In this case, the canonical connection is well known by the work \cite{kohno} of Kohno. Denote the free Lie algebra over ${\mathbb C}$ generated by the $Y_j$ by ${\mathbb L}(Y_1,\dots,Y_m)$. Denote the unipotent completion of $P_n$ by ${\mathbb P}_n$ and its Lie algebra by ${\mathfrak p}_n$. The associated graded of ${\mathfrak p}_n$ of is the graded Lie algebra $$ {\mathbb L}(X_{ij} : ij\text{ is a two element subset of }\{1,\dots,n\})/R $$ where $R$ is the ideal generated by the quadratic relations \begin{align*}\label{braid_relns} [X_{ij},X_{kl}]&\text{ when $i,j,k$ and $l$ are distinct;}\cr [X_{ij},X_{ik} + X_{jk}]& \text{ when $i,j$ and $k$ are distinct}. \end{align*} The natural (left) action of the symmetric group on it is defined by $$ Ad(\sigma): X_{ij} \mapsto X_{\sigma(ij)}. $$ The canonical invariant form $$ \omega \in E^1(X')\otimes \Gr^W_{\bullet} {\mathfrak p}_n $$ is $$ \omega = \sum_{ij} d\log(x_i - x_j) X_{ij}. $$ It is invariant because $$ \sigma^\ast \omega = \sum_{ij} d\log(x_{\sigma^{-1}(i)} - x_{\sigma^{-1}(j)}) X_{ij} = \sum_{ij} d\log(x_i - x_j) X_{\sigma(ij)} = Ad(\sigma)\omega. $$ We therefore obtain a homomorphism $$ B_n \to \Sigma_n\ltimes {\mathcal P}_n $$ where ${\mathcal P}_n$ denotes the complex form of the Malcev completion of $P_n$. This is the completion of $B_n$ relative to $\rho : B_n \to \Sigma_n$. \end{example}
1997-01-17T19:28:33
9607
alg-geom/9607006
en
https://arxiv.org/abs/alg-geom/9607006
[ "alg-geom", "math.AG" ]
alg-geom/9607006
Mikhail Zaidenberg
G. Dethloff, S. Orevkov, M. Zaidenberg
Plane curves with a big fundamental group of the complement
23 pages LaTeX. A revised version. The unnecessary restriction $d \ge 2g - 1$ of the previous version has been removed, and the main result has taken its final form
Amer. Math. Soc. Transl. (2) 184, 63-84 (1998)
null
Duke preprint DUKE-M-95-00
null
Let $C \s \pr^2$ be an irreducible plane curve whose dual $C^* \s \pr^{2*}$ is an immersed curve which is neither a conic nor a nodal cubic. The main result states that the Poincar\'e group $\pi_1(\pr^2 \se C)$ contains a free group with two generators. If the geometric genus $g$ of $C$ is at least 2, then a subgroup of $G$ can be mapped epimorphically onto the fundamental group of the normalization of $C$, and the result follows. To handle the cases $g=0,1$, we construct universal families of immersed plane curves and their Picard bundles. This allows us to reduce the consideration to the case of Pl\"ucker curves. Such a curve $C$ can be regarded as a plane section of the corresponding discriminant hypersurface (cf. [Zar, DoLib]). Applying Zariski--Lefschetz type arguments we deduce the result from `the bigness' of the $d$-th braid group $B_{d,g}$ of the Riemann surface of $C$.
[ { "version": "v1", "created": "Thu, 4 Jul 1996 12:40:41 GMT" }, { "version": "v2", "created": "Fri, 17 Jan 1997 18:28:10 GMT" } ]
2014-12-01T00:00:00
[ [ "Dethloff", "G.", "" ], [ "Orevkov", "S.", "" ], [ "Zaidenberg", "M.", "" ] ]
alg-geom
\section*{Introduction} \bigskip \noindent The fundamental groups of the plane curve complements are of permanent interest (see e.g. [Di, DoLib, Lib, MoTe, No, O, Zar] and the literature therein). Here we look for the most coarse properties of these groups (cf. e.g. [MoTe]). Namely, we distinguish between {\it big} and {\it small} groups. \bigskip \noindent {\bf 0.1. Definition.} We say that a group $G$ is {\it big} if it contains a non--abelian free subgroup. We call $G$ {\it small} if it is {\it almost solvable}, i.e. it has a solvable subgroup of finite index. \bigskip Recall the Tits alternative [Ti]: {\it any subgroup $G$ of a general linear group $GL(n,\,k)$ over a field $k$ of characteristic zero is either big or small.} This alternative holds true, even in a stronger form, for some classes of discrete groups, such as hyperbolic groups in sense of Gromov and the mapping class groups (see sect.1 below for references). \smallskip In [MoTe] classes of plane Pl\"ucker curves were indicated with infinite almost solvable (i.e. small) non--abelian fundamental groups of the complement. An example is the branching divisor of a generic projection of the Veronese surface $V_3$ of order $3$ onto $ I \!\! P^2$ [MoTe]. The well known Deligne--Fulton Theorem asserts that the complement of a nodal plane curve has abelian fundamental group. Here we show (and this is the main purpose of the paper) that the fundamental group of the complement of the dual of a nodal plane curve is big (with two evident exceptions). More precisely, we have \bigskip \noindent {\bf 0.2. Theorem.} {\it Let $C \subset I \!\! P^2$ be an irreducible immersed curve\footnote{i.e. all the analytic branches at the singular points of $C$ are smooth.} which is neither a line nor a conic nor a nodal cubic. Let $C^* \subset I \!\! P^{2*}$ be the dual curve. Then the group $\pi_1( I \!\! P^{2*} \setminus C^*)$ is big.} \bigskip \noindent Obviously, the statement does not hold for a line, nor for a conic. If $C$ is a nodal cubic then $C^*$ is a three--cuspidal quartic and $\pi_1 ( I \!\! P^{2*} \setminus C^*)$ is the metacyclic group of order $12$ [Zar, p.143--145]. Let $d$ be the degree and $g$ be the geometric genus of $C$. For $g\ge2$ the proof of Theorem 0.2 is easy. Indeed, denote by $\nu: C_{\rm norm} \to C$ a normalization of $C$. Set $R = \{(p,l)\in C_{\rm norm} \times I \!\! P^{2*}\,|\,\nu(p)\in l\}$. It is a smooth surface. Let $\mu_1 : R \to C_{\rm norm}$ and $\mu_2 : R \to I \!\! P^{2*}$ be the canonical projections. Since $C$ is immersed, it is easily seen that $\mu_2$ is ramified exactly over $C^*$. Denote by $R_0$ the part of $R$ over $ I \!\! P^{2*} \setminus C^*$. Then $\mu_1 : R_0 \to C_{\rm norm}$ is a holomorphic surjection with connected fibres. It follows that $(\mu_1)_* : \pi_1(R_0) \to \pi_1(C_{\rm norm})$ is an epimorphism, and hence the group $\pi_1(R_0)$ is big as soon as $g(C_{\rm norm}) \ge 2$ (see e.g. 1.2(a) below). Since $\mu_2 : R_0 \to I \!\! P^{2*} \setminus C^*$ is a finite unramified covering, we have that $(\mu_2)_* (\pi_1(R_0))$ is a finite index subgroup of $\pi_1( I \!\! P^{2*} \setminus C^*)$. Therefore, the group $\pi_1( I \!\! P^{2*} \setminus C^*)$ is also big. Thus, the only non-trivial cases are $g=0$ and $g=1$. However, the proofs of most of the intermediate results needed for these two cases are valid for any $g$, some of them under the additional assumption that $d\ge 2g-1$ (which is automatically fulfilled for $g = 0,\,1$). Therefore, we formulate everything for an arbitrary genus. This provides another proof of Theorem 0.2 for the case $g\ge2$, $d\ge 2g-1$. \smallskip The paper is organized as follows. In Section 1 we provide some (mostly well known) examples of big groups. Besides, by several examples we illustrate a conjectural relation between bigness of the fundamental group and C-hyperbolicity. These include, in particular, the quasi--projective quotients of bounded symmetric domains and the complements of certain reducible plane curves. The proof of Theorem 0.2 is done in Sect. 4. The results in Sect. 2 and 3 (which we believe to be of some independent interest) reduce the proof to the case of a nodal Pl\"ucker curve. In Theorem 2.1 we show that the part $Imm_{d,\,g}$ of the Hilbert scheme of degree $d$ genus $g$ plane curves, which corresponds to the immersed curves, is smooth, and the universal family of curves admits a simultaneous normalization over $Imm_{d,\,g}$ (see [AC, Ha] for related results, especially concerning ($a$) and ($b$) of Theorem 2.1). We show also that the nodal and (for $d \ge 2g-1$) the Pl\"ucker curves form Zariski open subsets of $Imm_{d,\,g}$. \medskip A preliminary version of Theorem 0.2 (under the additional restriction $d\ge 2g-1$) was announced in [DeZa1, Sect. 7] (see also [DeZa2]). After the preprint [DeOrZa] had been distributed the authors have received the preprint\footnote{We are thankful to I. Shimada for sending us this preprint.} [KuShi] where a presentation of the group $\pi_1( I \!\! P^{2*} \setminus C^*)$ for a generic nodal curve $C \subset I \!\! P^2$ of genus $g$ and degree $d \ge 2g+1$ has been computed (see Remark 4.6 below). It is our pleasure to thank D. Akhiezer, E. Artal, H. Flenner, V. Guba, S. Kosarev, V. Lin, V. Sergiescu for their friendly assistance and contributions to the paper. \section{Big groups and C-hyperbolicity} \noindent {\bf 1.1. Generalities on big groups} \smallskip \noindent By a theorem of von Neumann, a big group is non--amenable. The converse is not true, in general; the corresponding examples are due to A. Ol'shanskij, S. I. Adian and M. Gromov (see [OSh]). Note that the group in all these examples is not finitely presented. For a finitely presented group the equivalence of bigness and non--amenability is unknown\footnote {We are grateful to V. Sergiescu and V. Guba for this information.}. Being non--amenable, a big group can not be almost nilpotent or even almost solvable. As follows from the Nielsen--Schreier Theorem, a subgroup of finite index of a big group is big, as well as a normal subgroup with a solvable quotient. Clearly, a group with a big quotient is big. \medskip We remind several classical examples of big groups. First of all, for $g \ge 1$ the Siegel modular group $Sp_{2g}( Z \!\!\! Z)$ is big. In addition, it has no infinite normal solvable subgroup (see (1.3)-(1.4) below). \smallskip Another examples are: the Artin group $B_{d, \,g}$ of the $d$--string braids of a genus $g$ compact Riemann surface $R_g$, and the mapping class group Mod$_{g,\,n}$, i.e. the group of classes of isotopy of orientation preserving diffeomorphisms of a genus $g$ Riemann surface with $n$ punctures (see e.g. [Bi]). Namely, we have the following \bigskip \noindent {\bf 1.2. Lemma.} \noindent {\it ($a$) If $g\ge2$ then $\pi_1(R_g)$ is big. \noindent ($b$) The braid group $B_{d, \,g}\,\,(d \ge 1)$ is big iff $(d,\,g) \neq (1,\,0),\,(2,\,0),\,(3,\,0),\,(1,\,1)$. \noindent ($c$) The mapping class group Mod$_{g,\,n}$ is big iff $g \ge 1$, or $g = 0$ and $n \ge 4$.} \bigskip \noindent {\it Proof.} ($a$) By a theorem of Magnus [CoZi, (2.5.1)], after removing any of the standard generators $a_1,\,b_1,\dots,\,a_g,\,b_g$ of $\pi_1(R_g)$, the subgroup generated by the remaining ones is the free group {\bf F}$_{2g - 1}$. ($b$) By definition, $B_{d, \,g} = \pi_1 (S^dR_g \setminus {\Delta}_{d, \,g})$, where $S^dR_g$ denotes the $d$-th symmetric power of $R_g$ and ${\Delta}_{d, \,g} \subset S^dR_g$ denotes the discriminant hypersurface consisting of the $d$-tuples of points with coincidences. The pure braid group $P_{d,\,g} := \pi_1((R_g)^{\,d} \setminus D_{d,\,g})$, where $D_{d,\,g} \subset (R_g)^{\,d}$ is the union of diagonal hypersurfaces, is the normal subgroup of $B_{d, \,g}$ of index $d$! which corresponds to the Vieta covering $(R_g)^{\,d} \setminus D_{d,\,g} \to S^dR_g \setminus {\Delta}_{d, \,g}$. The fibration $(R_g)^{\,d+1} \setminus D_{d+1,\,g} \to (R_g)^{\,d} \setminus D_{d,\,g}$ with the fibre $R_g \setminus \{d\,\,\,{\rm points}\}$ yields the short exact sequence [Bi, sect. 1.3] $${\bf 1} \to \pi_1(R_g \setminus \{d\,\,\,{\rm points}\}) \to P_{d+1,\,g} \to P_{d,\,g} \to {\bf 1}\,.$$ For $d > 0$ the group $\pi_1(R_g \setminus \{d\,\,\,{\rm points}\})$ is a free group {\bf F}$_k$ with $k = 2g + d - 1$ generators. For $d = 0$ see (a). Hence, under the above restrictions the pure braid group $P_{d,\,g}$, and therefore also the braid group $B_{d, \,g}$, contains a subgroup isomorphic to a non-abelian free group. In the exceptional cases when $(d,\,g) = (1,\,1)$ or $g = 0,\,1 \le d \le 3$ the same exact sequence shows that the corresponding group $B_{d, \,g}$ is not big. This proves ($b$). ($c$) There is a natural surjection $j\,:\,$ Mod$_{g,\,n} \to $ Mod$_g :=$ Mod$_{g,\,0}$, where the kernel Ker$\,j$ is the braid group $B_{n, \,g}$ if $g \ge 2$ and its quotient by the center if $g = 1,\,n\ge 2$ or $g = 0,\, n\ge 3$ [Bi, Theorem 4.3]. Therefore, the group Mod$_{g,\,n}$ is big as soon as the corresponding braid group $B_{n, \,g}$ is so. For $g \ge 1$ the induced representation of Mod$_{g}$ into the first homology group of $R_g$ yields a surjection Mod$_{g}\to$ Sp$_{2g}( Z \!\!\! Z)$ (actually, Mod$_{1} \cong GL(2,\, Z \!\!\! Z)$). This shows that Mod$_{g},\,g \ge 1$, is a big group. For $g = 0$ we have that Mod$_{0,\,3} = B_{3, \,0}/$(center) is a finite group, the groups Mod$_{0,\,0}$ and Mod$_{0,\,1}$ are trivial, whereas Mod$_{0,\,2} = Z \!\!\! Z/2 Z \!\!\! Z$ [Bi, Theorem 4.5]. This completes the proof. \hfill $\Box$ \bigskip \noindent {\it Remark.} In fact, the Tits alternative holds in Mod$_{g},\,g \ge 1$ [Iv, MC] (note that for $g \ge 2$ the latter group is not isomorphic to any arithmetic linear group [Iv]). Furthermore, for $g\ge 2$ any almost solvable subgroup of Mod$_{g}$ is almost abelian [BiLuMC]. \bigskip Let us make certain remarks concerning a conjectural relation of bigness of the fundamental group of a complex space $X$ and its C-hyperbolicity. Recall that $X$ is said to be {\it (almost) C-hyperbolic} if it has an {\it (almost) Carath\'eodory hyperbolic} covering $Y \to X$, i.e. such that the bounded holomorphic functions on $Y$ separate points of $Y$ (up to finite subsets). As follows from Lin's Theorem [Lin, Theorem B], {\it the fundamental group of an almost C-hyperbolic algebraic variety can not be almost nilpotent.} Note that for $g \ge 1$ the complement $ I \!\! P^{2*} \setminus C^*$ is C--hyperbolic, and it is almost C-hyperbolic if $C$ is a generic rational curve of degree $d \ge 5$ [DeZa1, Thm. 1.1]. Thus, by Lin's Theorem, in all these cases the group $\pi_1( I \!\! P^{2*} \setminus C^*)$ is not almost nilpotent. Actually, by Theorem 0.2 above this group is big. This leads us to the following \medskip \noindent {\bf Question.} {\it Let $X$ be an almost C--hyperbolic algebraic variety. Is then necessarily $\pi_1(X)$ a big group?} \medskip By another theorem of Lin [Lin, Thm. B($b$)], $\pi_1(X)$ cannot be an amenable group with a non--trivial center assuming that the universal covering space $\tilde X$ is Carath\'eodory hyperbolic. An easy observation is that the answer is `yes' for dim$\,X=1$. Indeed, an algebraic curve $C$ is C--hyperbolic iff it is hyperbolic, or, in turn, iff its normalization $C_{\rm norm}$ has a non-abelian fundamental group. In the latter case the group $\pi_1(C_{\rm norm})$ is big (see 1.2($a$)). Note, however, that by a result of [LySu], any compact Riemann surface $R$ of genus $g \ge 2$ admits a Galois covering $\tilde R$ with a metabelian (i.e. two-step solvable) Galois group such that $\tilde R$ carries a non--constant bounded holomorphic function. Modifying this result, one may even assume $\tilde R$ being Carath\'eodory hyperbolic [LinZa, Sect. 3]. \medskip More generally, we have the following fact. Its proof given below was communicated to us by D. Akhiezer\footnote{and it is placed here with his kind permission.}. \bigskip \noindent {\bf 1.3. Theorem.} {\it Let $D \subset I \!\!\!\! C^n$ be a bounded symmetric domain, and let $\Gamma \subset {\rm Aut}\,D$ be a discrete subgroup. If the Bergman volume of a fundamental domain of $\Gamma$ is finite, then $\Gamma$ is a big group and it has no infinite solvable normal subgroup.} \bigskip \noindent {\it Proof.} According to a result of A. Borel and J.--L. Koszul [Bo, Kos], a homogeneous domain $D$ is symmetric iff the identity component $G$ of the automorphism group Aut$\,D$ is semisimple. Recall that $G$ has trivial center, and therefore it is a connected linear group [He, Ch. VIII.6]. Being semisimple $G$ is not solvable. Moreover, since $G$ is connected, it is not small. We have $D \cong G/K$, where $K \subset G$ is a maximal compact subgroup [ibid, Ch. VIII. 7]. The automorphism group Aut$\,D$ has finitely many connected components, i.e. [Aut$\,D : G] < \infty$ (indeed, being a compact Lie group the stabilizer Stab$_z \subset$ Aut$\,D$ of a point $z \in D$ has a finite number of connected components, which is the same as those of Aut$\,D$, because the quotient $D \simeq$ Aut$\,D/$Stab$_z$ is connected). Hence, $\Gamma \cap G$ has finite index in $\Gamma$, and the Bergman volume of $(\Gamma \cap G) \setminus D$ is finite, too. Therefore, the invariant volume Vol$\,((\Gamma\cap G) \setminus G)$ is finite, and so $\Gamma\cap G$ is a lattice of $G$. Fix a faithful linear representation $G \hookrightarrow GL(n,\, I \!\!\!\! C)$. Let $G_{ I \!\!\!\! C}$ be the Zariski closure of $G$ in $GL(n,\, I \!\!\!\! C)$. By Borel's Density Theorem (see e.g. [Ra, 5.16]), the conditions "$G$ is semisimple and Vol$\,((\Gamma\cap G) \setminus G) < \infty$" imply that the subgroup $\Gamma\cap G$ is Zariski dense in $G_{ I \!\!\!\! C}$. Hence, if $\Gamma$ is almost solvable, $G_{ I \!\!\!\! C}$ should be also almost solvable, which is not the case. By the Tits alternative, $\Gamma$ must be big. The last assertion follows from a theorem of V. Gorbatsevich [GoShVi, Proposition 3.7]. According to this theorem, the lattice $\Gamma\cap G$ in a connected Lie group $G$ possesses no infinite solvable normal subgroup iff $G$ is reductive and its semisimple part has a finite center. It is easily seen that in our case both conditions are fulfilled. \hfill $\Box$ \bigskip \noindent {\bf 1.4.} {\it Remark.} In fact, it would be enough in the above theorem that $\Gamma$ was a Zariski dense subgroup of a semisimple linear algebraic group $G$ with a finite center, which acts holomorphically in $D$. This may be illustrated by the following example 1.5($a$). \medskip \noindent {\bf 1.5.} {\bf Examples.} \smallskip \noindent ($a$) Let $D = {\cal Z}_g$ be the Siegel upper half--plane and $\Gamma = $Sp$_{2g}( Z \!\!\! Z),\,\,G =$ Sp$_{2g}( I \!\! R),\,g \ge 1$, are resp. the Siegel modular group and the simplectic group. Then $\Gamma \setminus D$ is a coarse moduli space of principally polarized abelian varieties of dimension $g$, which is a quasiprojective variety. Here $\Gamma$ is Zariski dense in $G$. Actually, by a theorem of A. Borel and Harish--Chandra [BoHC, Thm. 7.8], the arithmetic subgroup $G_{ Z \!\!\! Z}$ of a semisimple real algebraic group $G_{ I \!\! R}$ defined over $\bf Q$ is a lattice in $G_{ I \!\! R}$, and so by Borel's Density Theorem, it is Zariski dense in $G_{ I \!\!\!\! C}$. (By the way, these arguments show that $\Gamma = $Sp$_{2g}( Z \!\!\! Z)$ is a big group without infinite normal solvable subgroups.) \medskip \noindent ($b$) Let, furthermore, $D := T_{g,\,n} \subset\s I \!\!\!\! C^{3g-3+n}$ be the Teichm\"uller space of the $n$--punctured genus $g$ marked Riemann surfaces under the Bers realization, where $2-2g-n <0$. By Royden's Theorem, $\Gamma:=$ Aut$\,D$ is the Teichm\"uller modular group, which coincides with the mapping class group Mod$_{g,\,n}$. The quotient $\Gamma \setminus D$ is a coarse moduli space ${\cal M}_{g,\,n}$ of genus $g$ $\,\,n$--punctured Riemann surfaces, which is a quasiprojective variety. By Lemma 1.2 above, except the case when $(g,\,n) = (0,\,3)$ the group Mod$_{g,\,n}$ is big. \medskip \noindent (c) (see e.g. [Sh1, 2]). A smooth projective surface $S$ is called {\it a Kodaira surface} if there is a smooth fibration $\pi\,:\,S \to B$ over a curve $B$, where both $B$ and a generic fibre $F$ of $\pi$ are of genus $\ge 2$ (usually $\pi$ is supposed being a non-trivial deformation of $F$, but we don't need this assumption here). It is well known that the universal covering $\tilde S$ of $S$ can be realized as a bounded pseudo--convex Bergman domain in $ I \!\!\!\! C^2$. Thus, the projective surface $S = \Gamma \setminus D$ is C--hyperbolic; clearly, $\Gamma \simeq \pi_1(S)$ is a big group. More generally, the same is true when both $B$ and $F$ are quasiprojective hyperbolic curves. \bigskip Next we pass to the simplest examples of reducible plane projective curves with a big fundamental group of the complement. \medskip \noindent {\bf 1.6.} {\bf Examples.} \smallskip \noindent ($a$) Let $C \subset I \!\! P^2$ be a finite line arrangement. If these lines are in general position, then by the Deligne--Fulton Theorem, $\pi_1( I \!\! P^2 \setminus C)$ is abelian. Otherwise, this group is big. Indeed, let $C$ has a point $A$ of multiplicity at least $3$. The union $L$ of lines in $C$ passing through $A$ contains at least three members of the associated linear pencil. The linear projection $ I \!\! P^2 \setminus C \to I \!\! P^1 \setminus \{3\,\,{\rm points}\}$ with center at $A$ yields an epimorphism of the fundamental groups. Thus, $\pi_1( I \!\! P^2 \setminus C)$ dominates the free group {\bf F}$_2 = \pi_1( I \!\! P^1 \setminus \{3\,\,{\rm points}\})$, and therefore, it is big. In particular, if $C$ is an arrangement of six lines with four triple points, then $\pi_1( I \!\! P^2 \setminus C)$ is a finite index subgroup of the mapping class group Mod$_{0,\,5}$ (see [DeZa1, 6.1($a$)]; cf. also 1.5($b$) above). \medskip \noindent ($b$) Consider further a configuration $C \subset I \!\! P^2$ of a plane conic together with two of its tangent lines (cf. [DeZa1, 6.1($b$)]). The Zariski--van Kampen method yields a presentation $$G := \pi_1( I \!\! P^2 \setminus C) = \langle\,a,\,b\,|\,abab = baba \,\rangle\,.$$ The following proof of the bigness of $G$ was communicated to us by V. Lin\footnote{We are grateful to V. Lin for a kind permission to place here this proof.}. Remind that the Coxeter group ${\rm {\bf B}}_k$ is the group generated by the orthogonal reflections in $ I \!\! R^k$ with respect to the coordinate planes and the diagonals $x_i - x_j = 0,\,i,j=1,\dots,k$. The corresponding Artin--Brieskorn braid group is the fundamental group $\pi_1(G_k({\rm {\bf B}}_k))$ of the domain $$G_k({\rm {\bf B}}_k) := \{z = (z_1,\dots,z_k) \in I \!\!\!\! C^k\,|\,d_k(z)\cdot z_k \neq 0\}\,, $$ where $d_k(z)$ is the discriminant of the universal polynomial $p_k(t) = p_k(t,\,z):= t^k + z_1t^{k-1} +\dots +z_k$ of degree $k$. Put $G_k := \{z \in I \!\!\!\! C^k\,|\,d_k(z) \neq 0\}$, and let $E^1_k \to G_k$ be the standard $k$--sheeted covering over $G_k$, where $$E^1_k := \{(z,\,\lambda) = (z_1,\dots,z_k,\,\lambda) \in I \!\!\!\! C^{k+1}\,|\, p_k(\lambda,\,z) = 0\}\,.$$ Define a mapping $\,\varphi\,:\,E^1_{k+1} \to G_k({\rm {\bf B}}_k) \times \, I \!\!\!\! C$ as follows: $$\varphi (z_1,\dots,z_{k+1},\,\lambda) = (q_k,\,\lambda) = (\xi_1,\dots,\xi_k,\,\lambda)\,,$$ where $$q_k = q_k(t,\,\xi) = t^k + \xi_1t^{k-1} +\dots+\xi_k:= p_{k+1}(t+\lambda,\,z)/t \in G_k({\rm {\bf B}}_k) \,.$$ Note that $t\,|\,p_{k+1}(t+\lambda,\,z)$, because $p_{k+1}(\lambda,\,z) \equiv 0$ for $(z,\,\lambda) \in E^1_{k+1}$. Since $p_{k+1}(t+\lambda)$ is a polynomial with simple roots, the same is true for $q_k(t)$. Moreover, $q_k(0) \neq 0$; thus, indeed, $q_k \in G_k({\rm {\bf B}}_k)$. It is easily seen that $\varphi$ is a biregular isomorphism. Hence, the isomorphism $$\pi_1(G_k({\rm {\bf B}}_k)) \cong \pi_1(E^1_{k+1}) \hookrightarrow \pi_1(G_{k+1})$$ represents the Artin--Brieskorn braid group $\pi_1(G_k({\rm {\bf B}}_k))$ as a subgroup of finite index (equal to $k+1$) of the standard Artin braid group\footnote{From now on we denote $B_m = \pi_1(G_m)$ the standard Artin braid group with $n$ strings; don't confuse with the Coxeter group ${\rm {\bf B}}_k$.} $B_{k+1} := \pi_1(G_{k+1})$. Therefore, the former group is big as soon as the latter one is so. Both of them are big starting with $k = 2$ (for the Artin group $B_{k+1}$ this can be checked in the same way as it was done in the proof of Lemma 1.2 for the braid groups $B_{k,\,g}$). It remains to note that $ I \!\! P^2 \setminus C \cong G_2({\rm {\bf B}}_2)$, and therefore $G = \pi_1( I \!\! P^2 \setminus C)$ is isomorphic to the braid group $\pi_1(G_2({\rm {\bf B}}_2))$ which is big. \section{Nodal approximation of immersed curves} Due to Theorem 2.1 below, Theorem 0.2 can be reduced to the case where $C$ is a generic nodal Pl\"ucker curve. We also believe that Theorem 2.1 has an independent interest. We use the following notation and terminology. Let $ I \!\! P^N,\,N = N(d) = {d+2 \choose 2} - 1,\,d \ge 1$, be the Hilbert scheme of degree $d$ plane curves. Denote $Imm_{d,\,g}$ the locus of points of $ I \!\! P^N$ which correspond to reduced irreducible immersed curves of geometric genus $g,\,\, 0 \le g \le {d-1 \choose 2}$, and by $Nod_{d,\,g}$ resp. $PlNod_{d,\,g}$ the subset of points of $Imm_{d,\,g}$ which correspond to the nodal resp. to the Pl\"ucker nodal curves. Remind that an irreducible curve $C \subset I \!\! P^2$ is called {\it Pl\"ucker} if the only singular points of $C$ and the dual curve $C^*$ are ordinary nodes and cusps. Let $Pl\ddot uNod_{d,\,g} \subset PlNod_{d,\,g}$ be the subset of curves which have no flex at a node. Denote ${\cal S}_d \to I \!\! P^N$ the universal family of curves over the Hilbert scheme $ I \!\! P^N$, and let ${\cal S}_{d,\,g} \to Imm_{d,\,g}$ be its restriction to $Imm_{d,\,g}$. By {\it a family of curves} we mean a proper morphism $\varphi\,:X \to Y$ of relative dimension one of quasiprojective varieties. If $X,\,Y$ are smooth and $\varphi$ is a submersion, then the family $\varphi$ is called {\it smooth}. We say that $\varphi$ admits {\it a simultaneous normalization} if $Y$ is smooth and there exists a smooth family of curves $\varphi'\,:\,X' \to Y$ and a morphism $f\,:\,X' \to X$ commuting with the projections onto $Y$ such that for every point $y \in Y$ the restriction $f\,|\,X'_y\,:\,X'_y \to X_y$ onto the fibre over $y$ is a normalization map. \bigskip \noindent {\bf 2.1. Theorem.} {\it \noindent a) $Imm_{d,\,g} \subset I \!\! P^N$ is a smooth locally closed subvariety of pure dimension $3d + g - 1$. \smallskip \noindent b) The universal family of curves ${\cal S}_{d,\,g} \to Imm_{d,\,g}$ admits a simultaneous normalization $f\,:\, {\cal M}_{d,\,g} \to {\cal S}_{d,\,g}$. \smallskip \noindent c) $Nod_{d,\,g}$ and, for $n \ge 2g-1,\,\,\,PlNod_{d,\,g}$ are Zariski open subsets of $Imm_{d,\,g}$.} \bigskip \noindent {\it Remark.} The first statement of (c) and the dimension count in ($a$) can be found in [Ha, Sect. 2], while the proofs are quite different. Note that, by Harris [Ha], the variety $Imm_{d,\,g}$ is irreducible; it is non--empty for any $(d,\,g)$ with $0 \le g \le {d - 1 \choose 2}$ [Se, sect.11, p.347; Ha; O]. \medskip In this section we prove ($a$), ($b$) and the first part of ($c$) of Theorem 2.1; the proof of ($c$) is completed in sect. 3. First we study $Imm_{d,\,g}$ locally, in a neighborhood of a given curve $C \in Imm_{d,\,g}$. This needs certain preparation, including a portion of plane curve singularities. \bigskip \noindent {\bf 2.2. The Gorenstein--Rosenlicht invariant, the boundary braid and its algebraic length} \medskip \noindent Recall that the Gorenstein--Rosenlicht invariant ${\delta}_P$ of a singular analytic plane curve germ $(A,\,P)$ can be expressed as ${\delta}_P = {1 \over 2}(\mu + r - 1)$, where $\mu$ is the Milnor number and $r$ is the number of local branches of $A$ at $P$ [Mi, sect. 10]. For a reduced curve $F$ on a smooth surface $W$ we set ${\delta} (F) = \sum_{P \in {\rm Sing} F} {\delta}_P$. If $F$ is a complete irreducible curve, then by the genus formula and the adjunction formula [BPVV, II.11] we have \begin{equation} \pi_a (F) = g(F) + {\delta} (F) = 1/2\, F (K_W + F) + 1\,, \end{equation} where $\pi_a$ resp. $g$ denotes arithmetic resp. geometric genus, $K_W$ is the canonical divisor of $W$, and where for a non--compact surface $W$ we put $F K_W = {\rm deg}\,(K_W\,|\,F)$. \smallskip Let $U \subset I \!\!\!\! C$ be the unit disc, ${\Sigma} = U \times I \!\!\!\! C \subset I \!\!\!\! C^2$ be the solid cylinder ${\Sigma} = \{(u,\,v) \in I \!\!\!\! C^2\,|\,|u| < 1\}$, and $p\,:\, I \!\!\!\! C^2 \to I \!\!\!\! C$ be the first projection. Let $A \subset {\Sigma}$ be an analytic curve extendible transversally through the boundary $\partial {\Sigma}$, so that {\it the link} $\partial A = {\bar A} \cap \partial {\Sigma}$ is smooth. Suppose also that the projection $p\,:\,A \to U$ is proper, i.e. it is a (ramified) covering over the unit disc $U$ of degree, say, $m$. The link $\partial A$ carries a (closed) braid with $m$ strings $b_A \in B_m$ defined uniquely up to conjugation, where $B_m$ is the Artin braid group (see (1.6($b$) above) \footnote{ To define the braid $b_A$, cut the cylinder $\partial U = S^1 \times I \!\!\!\! C$ along its generator $1 \times I \!\!\!\! C$ and then identify it with $[0,\,1] \times I \!\! R^2 \subset I \!\! R^3$. Fix a numbering of the points of the fibre of $\partial A$ over $1 \in \partial U$. Passing once along the circle $S^1 = \partial U$ counterclockwise, we obtain the braid $b_A $.}. Let $\sigma_1,\dots, \sigma_{m-1}$ be the standard generators of $B_m$. For a braid $b = \sigma_{i_1}^{\alpha_1}\dots \sigma_{i_n}^{\alpha_n} \in B_m$ its {\it algebraic length} is defined as $l($b$) := \sum_{k=1}^n \alpha_k$. \bigskip \noindent {\bf 2.3. Lemma.} {\it Let $A \subset {\Sigma}$ as above be a nodal curve with ${\delta}$ nodes. Suppose that all the ramification points of the covering $p\,:\,A \to U$ are simple (i.e. with ramification indices $2$) and no two of them are at the same fibre. If the branching divisor $D \subset U$ consists of ${\delta} + \tau$ points, then $$l(b_A) = 2{\delta} + \tau\,.$$} {\it Proof.} Choose small disjoint discs $\omega_i$ in $U$, $i=1,\dots,{\delta} + \tau$, centered at the points of $D$. Fix a point at the boundary of the disc $\omega_i$ and join it by a path $\gamma_i$ with the point $1 \in \partial U$, where $\gamma_i,\, i=1,\dots,{\delta} + \tau$, are disjoint. The complement $U \setminus \bigcup_{i=1}^{{\delta} + \tau} ({\bar \omega}_i \cup \gamma_i)$ being simply connected, the braid $b_A$ is the product of the local braids $b_{A_i}$ which correspond to the curves $A_i := A \cap p^{-1}(\omega_i)$. It is easily seen that the local braid which corresponds to a node of $A$ is conjugate in the braid group $B_m$ with the square of a generator, and those at an irreducible ramification point is conjugate with a generator. Now the lemma easily follows. \hfill $\Box$ \bigskip With each plane curve singularity $(A,\,{\bar 0}) \subset ( I \!\!\!\! C^2,\,{\bar 0})$ we associate its {\it braid} $b_{A,\,{\bar 0}}$ defined as follows. Fix a generic linear projection $p\,:\,( I \!\!\!\! C^2,\,{\bar 0}) \to ( I \!\!\!\! C,\,{\bar 0})$, so that the direction of $p$ is different from the tangent directions of the branches of $A$ at $\bar 0$, and proceed in the same way as above. \bigskip \noindent {\bf 2.4. Lemma.} {\it Suppose that $(A,\,{\bar 0}) \subset ( I \!\!\!\! C^2,\,{\bar 0})$ is an immersed singularity (i.e. a singular point of a reduced curve having only smooth local branches $A_1,\dots, A_r$) with the Gorenstein--Rosenlicht invariant ${\delta} = {\delta}(A, \,{\bar 0})$. Then \noindent a) $\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,{\delta} = {1 \over 2} \, l(b_{A,\,{\bar 0}})\,.$ \smallskip \noindent b) Let $\tilde A$ be a small nodal deformation of $A$ defined in a fixed small ball $B_{\epsilon}$ centered at the origin. Denote by $r$ resp. $\tilde r$ the number of irreducible components of $A$ resp. of $\tilde A$ in $B_{\epsilon}$. Then ${\delta} ({\tilde A}) \le {\delta} (A)$, and ${\delta} ({\tilde A}) = {\delta} (A)$ iff $r = {\tilde r}$. In the latter case the irreducible components ${\tilde A}_1,\dots, {\tilde A}_r$ of ${\tilde A}$ in $B_{\epsilon}$ approximate the corresponding irreducible components $A_1,\dots, A_r$ of $A \cap B_{\epsilon}$. } \bigskip \noindent {\it Proof.} ($a$) We have ${\delta} = \sum_{1\le k < l \le r} (A_k\,\cdot\,A_l)_{\bar 0}$ [Mil, (10.20)]. Let $A_i' \subset B_{\epsilon}$ be a small generic deformation of the branch $A_i,\,\,i=1,\dots,r$. Set $A' = \bigcup_{i=1}^r A_i'$. Then $A'$ is a nodal curve with $${\delta} = \sum_{1\le k < l \le r} A'_k \cdot A'_l = \sum_{1\le k < l \le r} (A_k\,\cdot\,A_l)_{\bar 0}$$ nodes, and clearly, $b_{A,\,{\bar 0}} = b_{A',\,{\bar 0}}$. Since for all $i=1,\dots,r$ the generic linear projection $p\,:\,A_i \to U_{\epsilon'}$ is non--ramified, the same is true for the branches $A'_i,\,i=1,\dots,r$. Thus, $p\,:\,A' \to U_{\epsilon'}$ is ramified only at ${\delta}$ nodes, and therefore, in the notation of Lemma 2.3, $\tau = \tau (A') =0$. By this lemma, we have ${\delta} = 1/2\, l(b_{A',\,{\bar 0}}) = 1/2\, l(b_{A,\,{\bar 0}})$. This proves ($a$). \smallskip \noindent ($b$) Once again here $b_{A,\,{\bar 0}} = b_{{\tilde A},\,{\bar 0}}$. Due to ($a$) and to Lemma 2.3, we have $$2{\delta}(A) = l(b_{A,\,{\bar 0}}) = l(b_{{\tilde A},\,{\bar 0}}) = 2{\delta} ({\tilde A}) + \tau ({\tilde A})\,,$$ and the inequality of ($b$) follows. The equality holds iff $ \tau ({\tilde A}) =0$, which means that the projection $p \,:\,{\tilde A} \to U_{{\tilde \epsilon}}$ is ramified only at nodes of ${\tilde A}$. Therefore, for any irreducible component ${\tilde A}_i$ of ${\tilde A} \cap B_{\epsilon}$ the composition of the normalization map $({\tilde A}_i)_{\rm norm} \to {\tilde A}_i$ with the projection $p \,:\,{\tilde A}_i \to U_{{\tilde \epsilon}}$ is non--ramified and hence, one--sheeted. It follows that both of these mappings are biholomorphic, so that the irreducible components ${\tilde A}_i$ of ${\tilde A} \cap B_{\epsilon}$ are smooth. The degree of the branched covering $p\,:\,{\tilde A} \to U_{{\tilde \epsilon}}$ being equal to $r,\,\,{\tilde A} \cap B_{\epsilon}$ consists of $r$ smooth irreducible components close to those of $A$. \hfill $\Box$ \bigskip Let $X$ be a smooth projective surface, $C \subset X$ be an irreducible immersed curve with a normalization $\varphi_0\,:\,M_0 \cong C_{\rm norm} \to C$. By [No, (1.8)-(1.12)], there exists a smooth open complex surface $V$ which contains $M_0$ as a closed subvariety, and a holomorphic immersion $\varphi\,:\, V \to X$ that extends $\varphi_0$; it is called {\it a tubular neighborhood of} $\varphi_0$. To obtain $V$ one simply normalizes $C$ together with a tubular neighborhood of $C$ in $X$. \bigskip \noindent {\bf 2.5. Lemma.} {\it Let $C \subset X$, $M_0$ and $V$ be as above, and let $N \to M_0$ be the normal bundle of $M_0$ in $V$. Then \begin{equation} {\rm deg}\, N = M_0^2 = C^2 - 2{\delta} (C)\,.\end{equation} If $X = I \!\! P^2$ and $C \in Imm_{d,\,g}$, then \begin{equation}{\rm deg}\, N = 3d +2(g-1)\,.\end{equation}} \noindent {\it Proof.} By the adjunction formula, we have \begin{equation} 2g-2 = C^2 + CK_X - 2 {\delta} (C) = M_0^2 + M_0 K_V \,.\end{equation} Since $K_V = \varphi^* K_X$, by the projection formula we have $M_0 K_V = CK_X$, and so (2) follows. (3) is a corollary of (2) and the genus formula (1). \hfill $\Box$ \bigskip \noindent {\bf 2.6. Corollary.} {\it a) $N$ is very ample iff $C^2 - 2{\delta} (C) \ge 2g+1$. For $X = I \!\! P^2$ and $C \in Imm_{d,\,g}$ this is always the case, and furthermore, $h^1 (M_0,\,{\cal O}(N)) = 0$ and $h^0 (M_0,\,{\cal O}(N)) = 3d + g - 1$. \smallskip \noindent b) For any pair of points $p_1,\,p_2 \in M_0$ the line bundle $N_{p_1,\,p_2} = N - [p_1] - [p_2]$ on $M_0$ is spanned\footnote{i.e. the linear system $|N_{p_1,\,p_2}|$ has no base point.} if $C^2 - 2{\delta} (C) \ge 2g+2$. In particular, this is so if $X = I \!\! P^2$ and $C \in Imm_{d,\,g}$, where $d \ge 2$.} \bigskip \noindent {\it Proof.} The first statement of (a) and (b) follow from Lemma 2.5 by the well known criteria of ampleness or spannedness of a line bundle over a curve (see e.g. [Hart,IV.3.2] or [Na, 5.1.12]). By the Kodaira Vanishing Theorem, we obtain that $h^1 (M_0,\,{\cal O}(N)) = 0$, and hence, by the Riemann--Roch Formula, we have $h^0 (M_0,\,{\cal O}(N)) = {\rm deg}\, N + 1 - g = 3d + g - 1$. \hfill $\Box$ \bigskip The Kodaira Theorem on embedded deformations [Ko] implies such a \bigskip \noindent {\bf 2.7. Corollary.} {\it There exists a maximal smooth family $\pi_{loc}\,:\,{\cal M}_{loc} \to T_{loc}$ of embedded deformations of the curve $M_0 \cong \pi^{-1} (t_0)$ in $V$ over a smooth base $T_{loc}$ such that the Kodaira--Spencer map $T_{s_0}T_{loc} \to H^0 (M_0,\,{\cal O} (N))$ is an isomorphism. In particular, if $X = I \!\! P^2$ and $C \in Imm_{d,\,g}$, then\footnote{cf. [GH, sect.2.4; Ha].} ${\rm dim}\,T_{loc} = 3d + g - 1$.} \bigskip \noindent {\bf 2.8. Definition.} We say that a curve $C \in Imm_{d,\,g}$ is {\it strongly approximated} by curves $C' \subset Imm_{d,\,g}$ if $C'$ approximate $C$ in the Hausdorff topology, and for any singular point $P$ of $C$ of multiplicity $r(C,\,P)$ and for a fixed small neighborhood $B_{\epsilon, \,P}$ of $P$, the number $r(C',\,P)$ of irreducible components of $C' \cap B_{\epsilon, \,P}$ is equal to $r(C,\,P)$, and the irreducible components of $C' \cap B_{\epsilon, \,P}$ approximate those of $C \cap B_{\epsilon, \,P}$. Or, which is equivalent, if for a given tubular neighborhood $\varphi\,:\,V \to I \!\! P^2$ of a normalization $\varphi_0\,:\,M_0 \to C$, the curves $C'$ have normalizations $\varphi\,|\,M'\,:\,M' \to C'$, where $M' \subset V$ are obtained from $M_0$ by a small deformation. \bigskip We use below the following simple observation: a curve $C \in Nod_{d,\,g}$ is Pl\"ucker iff $C$ has only ordinary flexes, no multitangent line, i.e. a line tangent to $C$ in at least three points, and no bitangent line which is an inflexional tangent. One says that a curve $C \subset I \!\! P^2$ {\it has only ordinary singularities} iff all the local branches of $C'$ at any of its singular point are smooth and pairwise transversal. Denote by $Ord_{d,\,g}$ the set of all such curves of degree $d$ and genus $g$; clearly, $Ord_{d,\,g} \subset Imm_{d,\,g}$. \smallskip The next proposition should be known at least partially; in view of the lack of references, we give its proof. \bigskip \noindent {\bf 2.9. Proposition.} {\it The subspaces $Ord_{d,\,g},\, Nod_{d,\,g},\,PlNod_{d,\,g}$ and $Pl\ddot uNod_{d,\,g}$ are dense in $Imm_{d,\,g}$ in the topology of strong approximation, and hence also in the Hausdorff topology of $ I \!\! P^N$.} \bigskip \noindent {\it Proof.} Fix an arbitrary curve $C \in Imm_{d,\,g}$. We may assume that $d \ge 3$. First we show that $C$ can be strongly approximated by curves $C' \in Ord_{d,\,g}$. For a curve $C' \in Imm_{d,\,g}$ denote by ${\delta}_1(C')$ the number of all non--ordered pairs $(A'_i,\,A'_j)$ of local analytic branches of $C'$ which meet normally at their common center $P \in C'$, so that $(A'_i,\,A'_j)_P = 1$. Clearly, $C' \in Ord_{d,\,g}$ iff ${\delta}(C') = {\delta}_1(C')$. Suppose that $C$ as above has a non--ordinary singular point $P$ of multiplicity $m$. Consider the blow up $\sigma\,:\,X \to I \!\! P^2$ of $ I \!\! P^2$ at $P$, and let ${\hat C} \subset X$ be the proper transform of $C$. It is easily seen that ${\delta}({\hat C}) = {\delta}(C) - {m \choose 2}$. Since ${\hat C}^2 = C^2 - m^2$ we have $${\hat C}^2 - 2{\delta}({\hat C}) = C^2 - 2 {\delta}(C) - m = 3d + 2(g-1) - m \ge 2g +2\,.$$ Let $\varphi \,:\,V \to X$ be a tubular neighborhood of a normalization $\varphi_0\,:\, M_0 \to {\hat C}$ of ${\hat C}$. For a pair $(A_i,\,A_j)$ of local branches of $C$ at $P$ with $(A_i,\,A_j)_P > 1$ let $\hat P \in X$ be the common center of their proper preimages ${\hat A_i},\,{\hat A_j}$ in $X$, and let $P_i,\,P_j \in M_0$ be resp. the centers of the branches $\varphi^{-1}({\hat A_i}),\,\varphi^{-1}({\hat A_j})$ of the curve $M_0 \subset V$. By Lemma 2.5, for the normal bundle $N$ of $M_0$ in $V$ we have $${\rm deg}\,(N - [P_i]) = M_0^2 - 1 = {\hat C}^2 - 2{\delta}({\hat C}) - 1 \ge 2g + 1\,.$$ Therefore, being spanned, the line bundle $N - [P_i]$ possesses a section which does not vanish at $P_j,\,j \neq i$. It follows that $N$ has a section that vanishes at $P_i$, but not at $P_j$. This yields a deformation $M'_0$ of $M_0$ in $V$ which passes through $P_i$, but not through $P_j$. Thus, for the curve $C' := \sigma\varphi (M_0') \in Imm_{d,\,g}$ close enough to $C$ we have ${\delta}_1(C') > {\delta}_1(C)$. By induction on ${\delta}_1(C')$, we get a strong approximation $C' \in Ord_{d,\,g}$ of $C$. Suppose further that $C \in Ord_{d,\,g}$ is not nodal, i.e. it has a point $P$ of multiplicity $m \ge 3$. Applying the same procedure as above to a triple of points $P_i,\,P_j,\,P_k \in M_0$ which lie over $P$, and using the inequality $${\rm deg}\,(N - [P_i] - [P_j]) \ge 2g\,,$$ by the spannedness of the line bundle $N - [P_i] - [P_j]$, we obtain a section of $N$ which vanishes at the points $P_i$ and $P_j$, but not at $P_k$. This leads to a curve $C' \in Ord_{d,\,g}$ which strongly approximates $C$ and is simpler than $C$ in the following sense: $m(C') < m(C)$, where $$m(C) := \sum_{P_i \in {\rm sing}\,(C)} ({\rm mult}\,(P_i) - 1)\,.$$ Induction on $m(C')$ now shows that $C$ can be strongly approximated by curves $C' \in Nod_{d,\,g}$. \smallskip Next we show that a curve $C \in Nod_{d,\,g}$ can be strongly approximated by curves $C' \in Nod_{d,\,g}$ with only ordinary flexes. We proceed by induction on the number $ofl(C')$ of ordinary flexes of $C'$. Since such a flex is a normal intersection point of the Hesse curve of $C'$ with a smooth local branch of $C'$, clearly, the bounded function $ofl(C')$ is lower semi--continuous on $Nod_{d,\,g}$ with respect to the Hausdorff topology. Suppose that $C$ has a non--ordinary flex at a local branch $A$ of $C$ centered at $P \in C$, so that $(A,\,L)_P \ge 4$, where $L$ is the tangent line to $A$ at $P$. In the notation as above, let $\sigma\,:\,X \to I \!\! P^2$ be the composition of three successive blow ups over $P$ with centers at the proper preimages of $A$. Let ${\hat L} \subset X$ be the proper transform of $L$ and $\hat P$ be the center of the proper transform ${\hat A} \subset X$ of $A$. We have $({\hat A},\, {\hat L})_{\hat P} \ge 1$. If $P$ is a smooth point of $C$ then ${\hat C}^2 = C^2 - 3$ and ${\delta}({\hat C}) = {\delta} (C)$. If $P$ is a node of $C$ then ${\hat C}^2 = C^2 - 6$ and ${\delta}({\hat C}) = {\delta} (C) - 1$. In any case, $${\rm deg}\,N = {\hat C}^2 - 2{\delta}({\hat C}) \ge C^2 - 2{\delta}(C) - 4 \ge 2g\,.$$ Therefore, the normal bundle $N$ of $M_0$ in $V$ is spanned, and hence it has a section which does not vanish at the point $\hat P$. The corresponding Kodaira-Spencer deformation yields a curve $M_0'$ on $X$ close enough to $M_0$ which does not pass through $\hat P$. It is easily seen that the projection $C' := \sigma\varphi(M_0') \subset I \!\! P^2$ is a nodal curve with an ordinary flex at $P$ and such that $ofl(C') > ofl(C)$. After a finite number of steps we obtain a strong approximation $C' \in Nod_{d,\,g}$ of $C$ with only ordinary flexes. Suppose further that $C \in Nod_{d,\,g}$ has only ordinary flexes (note that this is an open condition). We will find a strong approximation $C'$ of $C$ without multiple tangents. Denote by $b(C')$ the total number of distinct intersection points with $C$ of all the bitangent lines of $C'$. Clearly, the bounded function $b(C')$ is lower semi--continuous on $Nod_{d,\,g}$. Let $C$ have a multitangent line $L$ which is tangent to $C$ at points $P,\,Q,\,R \in C$ and, perhaps, at some other points. Let $\sigma\,:\,X \to I \!\! P^2$ be the composition of the blow-ups of $ I \!\! P^2$ at the points $P,\,Q$ and $R$, and let $\hat C$ be the proper transform of $C$ at $X$. Note that $d = $ deg$\,C = L\cdot C \ge 6$. As above, the blow up at a smooth point (resp. at a node) of $C$ decreases the difference $C^2 - 2{\delta} (C)$ by $1$ (resp. by $2$). Thus, we have $${\hat C}^2 - 2{\delta} ({\hat C}) \ge C^2 - 2{\delta} (C) - 6 = 2g + 3d - 8 \,.$$ Let $\varphi\,:\,V \to X$ be a tubular neighborhood of a normalization $\varphi_0\,:\,M_0 \to {\hat C}$ of $\hat C$, and let $N$ be the normal bundle of $M_0$ in $V$. The line bundle $N - [{\hat P}] - [{\hat Q}]$ on $M_0$ of degree $\ge 2g + 3d - 10 > 2g + 2$ is spanned (cf. Corollary 2.6). This yields a deformation $C' := \sigma \varphi(M_0') \subset Nod_{d,\,g}$ of $C$ such that $L$ is still tangent to $C'$ at the points $P$ and $Q$, and meets $C'$ normally at $R$, so that $b(C') > b(C)$. Maximizing $b(C')$ we get a strong approximation $ C' \in Nod_{d,\,g}$ of $C$ with only ordinary flexes and without multiple tangents. Suppose now that $C \in Nod_{d,\,g}$ has only ordinary flexes and no multiple tangent line, which is an open condition. To find a strong Pl\"ucker approximation $C'$ of $C$, we will proceed by induction on the total number $inf(C')$ of distinct intersection points of $C'$ with all of its inflexional tangent lines. We have to ensure that no inflexional tangent line of $C'$ is a bitangent line. Let a bitangent line $L$ of $C$ be an inflexional tangent of $C$ at a point $P \in C$ and tangent to $C$ at a point $Q \in C$. Then $d =$ deg$\,C = C\cdot L \ge 5$. Blowing up $ I \!\! P^2$ at $Q$ we get a surface $X = \sigma_Q( I \!\! P^2)$. In the notation as above, we have $${\rm deg}\,(N - 3[{\hat P}]) \ge 2g + 3d - 7 > 2g + 2\,.$$ Therefore, there exists a deformation $ C' = \sigma \varphi(M_0') \in Nod_{d,\,g}$ of $C$ such that $L$ is still an inflexional tangent of $C$ at $P$, but it meets $C$ normally at $Q$. Thus, $inf(C') > inf(C)$. By induction, we obtain a strong approximation $C'$ of $C$ which belongs to $PlNod_{d,\,g}$. Suppose finally that $C \in PlNod_{d,\,g} \setminus Pl\ddot uNod_{d,\,g}$, so that, although all the flexes of $C$ are ordinary, one of them, say $(A,\,P)$, is located at a node of $C$ with the second branch, say, $B$. This time we proceed by induction on the number $sfl(C')$ of flexes of $C'$ which are smooth points. Evidently, $sfl(C')$ is a bounded lower semi--continuous function on $Nod_{d,\,g}$. Performing two successive blow ups, the first one at $P \in C$ and the second one at the center of the proper transform of the branch $A$, we obtain a surface $X$. Denote by ${\hat Q}$ the center of the proper preimage $\hat B$ of the branch $B$ in $X$. We have $${\rm deg}\,N = {\hat C}^2 - 2{\delta}({\hat C}) = C^2 - 2{\delta}(C) - 3 \ge 2g + 1\,,$$ so that the line bundle $N - [{\hat Q}]$ on $V$ is spanned. Hence, we can find a section of $N$ which vanishes at $\hat Q$ and does not vanish at $\hat P$. This yields a small deformation $M_0'$ of $M_0$ on $V$ which passes through $\hat Q$ but not through $\hat P$. The curve $C' := \sigma\varphi(M_0') \subset I \!\! P^2$ is close enough to $C$, still has a node at $P$ which is not any more a flex, while $L$ is still a tangent line of $C'$ at $P$. Note that a small deformation of $C$ yields a small deformation of the Hesse curve $H_C$ of $C$, so that the flexes of $C$ which are the (normal) intersection points of $C$ and $H_C$ are also perturbed a little. Thus, $C'$ has a flex at a smooth point close to $P$. It follows that $sfl(C') > sfl(C)$. In a finite number of steps we obtain a desired strong approximation $C' \in Pl\ddot uNod_{d,\,g}$ of $C$. This completes the proof. \hfill $\Box$ \bigskip The next lemma shows that the strong approximation of immersed curves coincides with the usual one. \bigskip \noindent {\bf 2.10. Lemma.} {\it Let $C \in Imm_{d,\,g}$, and let $\varphi\,:\, V \to I \!\! P^2$ be a tubular neighborhood of its normalization $\varphi_0\,:\,M_0 \to C$. Then any curve $C' \in Imm_{d,\,g}$ close enough to $C$ in the Hausdorff topology of $ I \!\! P^N$ (or, which is the same, coefficientwise) is the image of a unique smooth curve $M \cong C'_{\rm norm} \subset V$ under the holomorphic mapping $\varphi\,:\,V \to I \!\! P^2$.} \bigskip \noindent {\it Proof.} Let $P$ be a singular point of $C$, and let $B_{\epsilon, \,P}$ be a fixed small neighborhood of $P$. Denote by $r(C,\,P)$ the multiplicity of $C$ at $P$, and by $r(C',\,P)$ the number of irreducible components in $B_{\epsilon, \,P}$ of a curve $C'$ close enough to $C$ (cf. Definition 2.8). Once we show that $\,r(C,\,P) = r(C',\,P)$ for any singular point $P$ of $C$, then the irreducible components of $C' \cap B_{\epsilon, \,P}$ approximate those of $C \cap B_{\epsilon, \,P}$, i.e. $C'$ is a strong approximation of $C$, and the statement follows. Actually, it is sufficient to prove the equality $r(C,\,P) = r(C',\,P)$ under the additional assumption that the approximating curve $C'$ is nodal. Indeed, by Proposition 2.9, the curve $C'\subset Imm_{d,\,g}$ can be, in turn, strongly approximated by a curve $C'' \in Nod_{d,\,g}$. Since $C''$ approximates both $C$ and $C'$ in the Hausdorff topology, from the equalities $r(C'',\,P) = r(C,\,P)$ and $r(C'',\,P) = r(C',\,P)$ it follows that $r(C',\,P) = r(C,\,P)$. Assuming further that $C'$ is nodal, by (1) and Lemma 2.4($b$), we obtain $${n-1 \choose 2} - g = {\delta} (C') = \sum_{P\in {\rm Sing}\,C'} {\delta} (C'\cap B_{\epsilon, \,P}) \le \sum_{P\in {\rm Sing}\,C} {\delta}(C,\,P) = {n-1 \choose 2} - g\,.$$ Henceforth, ${\delta} (C'\cap B_{\epsilon, \,P}) = {\delta} (C, \,P)$ for all $P\in {\rm Sing}\,C$. Applying Lemma 2.4($b$) once again, we get that $r(C',\,P) = r(C,\,P)$ for all $P\in {\rm Sing}\,C$, as desired. \hfill $\Box$ \bigskip \noindent {\bf 2.11. Lemma.} {\it ($a$) $Imm_{d,\,g}$ is a locally closed complex analytic submanifold of $ I \!\! P^N$ of dimension $3d + g - 1$. \smallskip \noindent ($b$) The universal family of curves ${\cal S}_{d,\,g} \to Imm_{d,\,g}$ over $Imm_{d,\,g}$ admits a complex analytic simultaneous normalization $f = f_{d,\,g}\,:\,{\cal M}_{d,\,g} \to {\cal S}_{d,\,g}$.} \bigskip \noindent {\it Proof.} Fix a curve $C \in Imm_{d,\,g}$, and consider a tubular neighborhood $\varphi\,:\,V \to I \!\! P^2$ of a normalization $\varphi_0$ of $C$. By Corollary 2.7 and Lemma 2.10, the projection $\varphi$ yields a local analytic chart $U_C$ of dimension $3d + g - 1$ on $Imm_{d,\,g}$ centered at $C$ which covers the whole intersection of $Imm_{d,\,g}$ with a sufficiently small ball in $ I \!\! P^N$ around $C$. This proves ($a$). \smallskip To prove ($b$) denote by ${\cal S}_C$ the restriction of the family ${\cal S}_{d,\,g}$ onto the chart $U_C$. Note that the same projection $\varphi$ yields an analytic simultaneous normalization $f_C\,:\,{\cal M}_C \to {\cal S}_C$ of ${\cal S}_C$. Any two such normalizations $f_C\,:\,{\cal M}_C \to {\cal S}_C$ and $f'_C\,:\,{\cal M}'_C \to {\cal S}_C$ over the same chart $U_C$ which arise from two different tubular neighborhoods $\varphi,\,\varphi'$, can be naturally biholomorphically identified via their projections. Hence, the equivalence class of these normalizations over the same chart $U_C$ in $Imm_{d,\,g}$ can be regarded as an equivalence class of charts on a new complex manifold ${\cal M}_{d,\,g}$ of dimension $3d + g$. Indeed, suppose that two charts $U_C$ and $U_{C'}$ on $Imm_{d,\,g}$ have a non--empty intersection $U_{C,\,C'} := U_C \cap U_{C'}$. Consider a fibrewise bimeromorphic mapping of smooth manifolds $f_{C,\,C'} := f_{C'}^{-1} \circ f_C\,:\, {\cal M}_C\,|\,U_{C,\,C'} \to {\cal M}_{C'}\,|\,U_{C,\,C'}$. It is biholomorphic at the complement of the `multiple point locus' $D_{C,\,C'} := f_C^{-1}($sing$\,S_{C,\,C'})$, where $S_{C,\,C'}:= S_C\,|\, U_{C,\,C'}$, and by Riemann's extension Theorem, it has a holomorphic extension through $D_{C,\,C'}$. Clearly, the projection $f_{d,\,g}\,:\,{\cal M}_{d,\,g} \to {\cal S}_{d,\,g}$ induced by the local mappings $f_C\,:\,{\cal M}_C \to {\cal S}_C$ is a holomorphic simultaneous normalization, which proves ($b$). \hfill $\Box$ \bigskip Next we show that all the above subvarieties of the Hilbert scheme $ I \!\! P^N$ are algebraic. Although the following statement holds in much bigger generality\footnote{We are grateful to H. Flenner who introduced to us this circle of ideas.} (cf. e.g. [BinFl, Theorem 2.2]), it will be enough for us this restricted version which has a rather easy proof. \bigskip \noindent {\bf 2.12. Lemma.} {\it Let $f\,:\,X \to Y$ be a family of curves over an irreducible base $Y$. Then there exists a Zariski open subset $U \subset Y$ such that the restriction $f\,|\,f^{-1}(U)$ of $f$ over $U$ admits a simultaneous normalization.} \bigskip \noindent {\it Proof.} Without loss of generality we may suppose $Y$ being smooth. Let $\nu\,:\,X_{\rm norm} \to X$ be a normalization. Consider the induced family of curves $f' := f \circ \nu$. Since the singular locus $S$ of the normal variety $X_{\rm norm}$ has codimension at least $2$, its image $f'(S) \subset Y$ has codimension at least $1$. Restricting $f$ and $f'$ onto the complement of the Zariski closure $\overline {f'(S)}$ of the constructible subset $f'(S)$ in $Y$, we may suppose $X_{\rm norm}$ being smooth. By the Bertini--Sard Theorem [Hart, III.10.7], $f'$ is an immersion over a Zariski open subset $U \subset Y$. Therefore, each fibre $(f')^{-1}(y),\,y \in U$, is smooth, and the restriction $\nu\,|\,(f')^{-1}(y)$ yields a normalization of the curve $X_y :=f^{-1}(y)$. Thus, we have obtained the desired simultaneous normalization of the original family $f$ over $U$. \hfill $\Box$ \bigskip We use below the following notation. Given a family of curves $f\,:\,X \to Y$, for any $g \ge 0$ denote by $Curv_g(f)$ the subset of points $y \in Y$ such that the fibre $X_y$ over $y$ is a reduced irreducible curve of geometric genus $g$. For the universal family $f_d\,:\,{\cal S}_d \to I \!\! P^N$ of degree $d$ curves in $ I \!\! P^2$, set $Curv_{d,\,g} = Curv_g(f_d)$. We say that an abstract reduced irreducible curve $C$ is {\it of immersed type} if its normalization map $\nu\,:\,C_{\rm norm} \to C$ has a nowhere vanishing differential. Let $Imm_g(f)$ be the subset of points $y \in Curv_g(f)$ which correspond to the curves of immersed type, so that, in particular, $Imm_g(f_d)= Imm_{d,\,g}$. \bigskip \noindent {\bf 2.13. Corollary.} {\it ($a$) Given a family of curves $f\,:\,X \to Y$, the base $Y$ can be represented as a disjoint union of smooth irreducible quasi--projective subvarieties $Y_i \subset Y,\,\,i=1,\dots,n = n(f)$, such that for each $i=1,\dots,n $ the restriction of $f$ onto $Y_i$ admits a simultaneous normalization. \smallskip \noindent ($b$) For any $g \ge 0$ the subsets $Curv_g(f) \subset Y$ and $Imm_g(f) \subset Y$ are constructible. In particular, $Curv_{d,\,g}$ and $Imm_{d,\,g}$ are constructible subsets of the Hilbert scheme $ I \!\! P^N$.} \bigskip \noindent {\it Proof.} ($a$) Assuming for simplicity that $Y$ is irreducible we start with $Y_1 := U$, where $U \subset Y$ is as in Lemma 2.12 above. Next we apply Lemma 2.12 to the restriction of $f$ onto each of the irreducible components of the regular part of the Zariski closed subvariety $Y^{(1)}:=Y \setminus Y_1$ of $Y$. Following this way, in a finite number of steps we obtain the desired partition of $Y$. \hfill $\Box$ \smallskip \noindent ($b$) Since $f\,|\,Y_i$ admits a simultaneous normalization, for any $i=1,\dots,n$ the number and the geometric genera of the irreducible components of a fibre $X_y = f^{-1}(y)$ do not depend on $y \in Y_i$. Thus, $Curv_g(f)$ is a union of some of the $Y_i$, and hence it is constructible. Set $X_i = f^{-1}(Y_i)$ and $f_i = f\,|\,X_i$, where $Y_i \subset Curv_g(f)$ is a stratum of the above stratification. Let $$ \begin{picture}(800,60) \unitlength0.2em \thicklines \put(88,1){$Y_i$} \put(64,23){$X_i'$} \put(108,23){$X_i$} \put(83,24){$\vector(1,0){15}$} \put(105,19){$\vector(-1,-1){12}$} \put(72,19){$\vector(1,-1){12}$} \put(70,10){$p_i$} \put(89,29){$\nu_i$} \end{picture} $$ be a simultaneous normalization. Denote by $T_{Y_i} X_i' = $ Ker$\,dp_i$ the relative tangent bundle of $p_i$; $p_i$ being a smooth family of curves, $T_{Y_i} X_i'$ is a smooth line bundle on $X_i'$. Let $D_i \subset X_i'$ be the locus of points where the restriction $d\nu_i\,|\,T_{Y_i} X_i'$ vanishes. Since $D_i$ is Zariski closed its image $p_i(D_i) \subset Y_i$ is a constructible subset of $Y_i$. Clearly, the complement $Y_i \setminus p_i(D_i)$ coincides with $Imm_g(f) \cap Y_i$. Thus, the latter subset is constructible for all $i= 1,\dots,n$. Hence, $Imm_g(f)$ is constructible, too. \hfill $\Box$ \bigskip \noindent {\bf 2.14.} {\it Starting the proof of Theorem 2.1.} ($a$) directly follows from Lemma 2.11($a$) and Corollary 2.13($b$). From Corollary 2.13($b$) it also follows that the total space ${\cal S}_{d,\,g}$ of the universal family of curves ${\cal S}_{d,\,g} \to Imm_{d,\,g}$ is a quasi--projective variety. The holomorphic mapping $f = f_{d,\,g}\,:\,{\cal M}_{d,\,g} \to {\cal S}_{d,\,g}$ which realizes an analytic simultaneous normalization is finite and proper (see Lemma 2.11($b$)). Therefore, by the Grauert--Remmert Theorem [Ha, B 3.2], ${\cal M}_{d,\,g}$ possesses a structure of a quasi--projective variety, so that $f$ is a finite morphism of quasi--projective varieties. Thus, $f$ yields an algebraic simultaneous normalization of the universal family of curves over $Imm_{d,\,g}$. This proves ($b$). To prove the first part of (c) denote $T = Imm_{d,\,g},\,\,{\cal S}_T = {\cal S}_{d,\,g},\,\,{\cal M}_T = {\cal M}_{d,\,g}$ and $ I \!\! P^2_T = I \!\! P^2 \times T$. There is a natural embedding $\,i\,:\,{\cal S}_T \hookrightarrow I \!\! P^2_T$. Consider the composition $\varphi:= i \circ f\,:\,{\cal M}_T \to I \!\! P^2_T$ and its relative square $\varphi^{(2)} := \varphi^2_T \,:\,{\cal M}_T^2 \to ( I \!\! P^2_T)^2$, where ${\cal M}_T^2 := {\cal M}_T \times_T {\cal M}_T$ and $( I \!\! P^2_T)^2 := I \!\! P^2_T\times_T I \!\! P^2_T$. Let ${\cal D}_T \subset {\cal M}_T^2$ resp. $D_T \subset ( I \!\! P^2_T)^2$ be the diagonals. Clearly, $E:= (\varphi^{(2)})^{-1}(D_T) \setminus {\cal D}_T$ is a closed subvariety of ${\cal M}_T^2$, and the restriction $\pi^{(2)}\,|\,E\,:\,E \to T$ of the projection $\pi^{(2)}\,:\,{\cal M}_T^2 \to T$ has finite fibres. Its fibre over a point $t \in T = Imm_{d,\,g}$ corresponds to the multiple point divisor on the normalization $M_t$ of the immersed curve $S_t \subset I \!\! P^2$. The restriction $\varphi^{(2)}\,|\,E\,:\,E \to D_T$ is a finite morphism. The image ${\tilde E} := \varphi^{(2)}(E)$ is proper over $T$. Moreover, the fibre $\tau^{-1}(t) \subset {\tilde E}$ over a point $t \in T$ under the restriction to ${\tilde E}$ of the projection $\tau\,:\,D_T \to T$ corresponds to the set of singular points of the curve $S_t$. Therefore, it consists of ${\delta} = {d-1 \choose 2} - g$ points iff $S_t$ is a nodal curve. By Proposition 2.9, any irreducible component of $T$ contains points which correspond to nodal curves. Thus, the finite morphism $\tau\,:\,{\tilde E} \to T$ has degree ${\delta}$ over every such component, and so, the complement $Imm_{d,\,g} \setminus Nod_{d,\,g} \subset T = Imm_{d,\,g}$ coincides with the ramification divisor $R_{\tau}$ of $\tau$. Hence, $Nod_{d,\,g} \subset Imm_{d,\,g}$ is, indeed, a Zariski open subvariety. \hfill $\Box$ \bigskip \noindent {\it Remark.} If $S_t$ is a nodal curve with ${\delta} = {d-1 \choose 2} - g$ nodes, then the fibre $p^{-1}(t)$ of the above projection $p := \pi^{(2)}\,|\,E\,:\,E \to T$ consists of $2{\delta}$ points. The latter holds true if $S_t$ has only ordinary singularities. Hence, the subset $Imm_{d,\,g} \setminus Ord_{d,\,g}$ is contained in the ramification divisor $R_p \subset T$ of $p$. \section{Pl\"ucker conditions} It is known [Au] that in general, the subset of the rational Pl\"ucker curves is not Zariski open in the space ${\cal R}_d$ of all the rational plane curves of a given degree $d$, although it always contains a Zariski open subset of ${\cal R}_d$. Nevertheless, we will show that $PlNod_{d,\,g}$ is a Zariski open subset of $Imm_{d,\,g}$ for $d \ge 2g-1$, which proves Theorem 2.1(c). \bigskip \noindent {\bf 3.1. Lemma.} {\it Let $C \subset I \!\! P^2$ be an irreducible nodal curve of degree $d$ with the normalization $M$, and let $g^2_d$ be the linear system on $M$ of all line cuts of $C$. Then $C$ is a Pl\"ucker curve iff $g^2_d$ contains no divisor $D$ of the form $$(i) \,\,\,D = 4p_1 +\dots\,;\,\,\,\,\,\,or\,\,\,\,\,\,(ii)\,\,\,D = 3p_1 + 2p_2 +\dots\,;\,\,\,\,\,\,or\,\,\,\,\,\,(iii) \,\,\,D = 2p_1 + 2p_2 + 2p_3+\dots\,,$$ where $p_i \in M$ are not necessarily distinct.} \bigskip \noindent {\it Proof.} The system $g^2_d$ contains no divisor of type (i) iff all the flexes of $C$ are ordinary, i.e. all the singular branches of $C^*$ are ordinary cusps. Under this condition, at most two of the local branches of $C^*$ meet at a point iff $g^2_d$ does not contain any divisor of type (iii). Furthermore, two branches of $C^*$ meet at a point and one of them is singular iff $g^2_d$ contains a divisor $D$ as in (ii). Since $C$ being nodal has no tacnode, $C^*$ has no one, too. Therefore, $C$ is a Pl\"ucker curve iff $g^2_d$ does not contain any divisor $D$ as in (i)--(iii). \hfill $\Box$ \bigskip Recall the following notion (see e.g. [ACGH]). \medskip \noindent {\bf 3.2. Picard bundles} \smallskip Let $M$ be a smooth projective curve of genus $g$. The $d$--th symmetric power $S^dM$ (which is a smooth manifold) might be regarded as the space of degree $d$ effective divisors on $M$. Let $J_d(M) = $Pic$\,^d(M)$ be the component of the Picard group Pic$\,(M)$ which parametrizes the degree $d$ line bundles on $M$, and let $\phi_d \,:\,S^dM \to J_d(M)$ be the morphism sending a degree $d$ effective divisor on $M$ into its linear equivalence class. Chosing a base point $p_0 \in M$ we may identify $ J_d(M)$ with the Jacobian variety $J_0(M)$ and $\phi_d$ with the $d$-th Abel--Jacobi mapping. By a theorem of Mattuck [Ma] (see also [ACGH, Ch.IV]) for $d \ge 2g -1$ the morphism $\phi_d$ is a submersion and moreover, it defines a projective bundle (i.e. a projectivization of an algebraic vector bundle) with the standard fibre $ I \!\! P^{d-g}$. This bundle is called {\it the $d$-th Picard bundle of $M$}. Given a smooth family $\pi\,:\,{\cal M} \to T$ of complete genus $g$ curves and given $d \ge 2g -1$, there is the associated Picard bundle $\Phi_d \,:\,S^d{\cal M} \to {\cal J}_d({\cal M})$ of relative smooth schemes over $T$. Consider also the associated grassmanian bundle $Grass_{2,\,d-g}({\cal M}) \to {\cal J}_d({\cal M})$ which parametrizes the two--dimensional linear series $g^2_d$ of degree $d$ on the fibres $M_t = \pi^{-1}(t),\,t\in T$. Let $\pi\,:\,{\cal M} \to T,\,\,T = Imm_{d,\,g}$, be the family constructed in 2.14 above. Then for each $t \in T$ there is the linear series $g^2_d = g^2_d(t)$ on $M_t$ of the line cuts of the plane curve $C_t = f(M_t) \subset I \!\! P^2$. This defines a regular section $\sigma\,:\,T \to Grass_{2,\,d-g}({\cal M})$. \bigskip \noindent {\bf 3.3.} {\it Finishing up the proof of Theorem 2.1(c).} Let $\pi\,:\,{\cal M} := {\cal M}_T \to T$ be the family as in 2.14, and let $\Phi_d \,:\,S^d{\cal M} \to {\cal J}_d({\cal M})$ be the associated Picard bundle. Denote ${\cal D}^{(i)}$ resp. ${\cal D}^{(ii)},\,\,{\cal D}^{(iii)}$ the subvariety of $S^d{\cal M}$ which consists of the degree $d$ effective divisors on the fibres $M_t$ of $\pi$ of the form (i) resp. (ii), (iii) of Lemma 3.1. Set ${\cal D} = {\cal D}^{(i)} \cup {\cal D}^{(ii)} \cup {\cal D}^{(iii)}$. Note that ${\cal D}$ is a closed subvariety of $S^d{\cal M}$ of codim$\,_{S^d{\cal M}}{\cal D} \ge 3$ (and moreover, codim$\,_{S^dM_t}{\cal D}_t \ge 3$ for each $t \in T$). Indeed, to be in ${\cal D}_t$ a divisor on $M_t$ must satisfy a system of three independent equations. Let ${\cal Z} \subset Grass_{2,\,d-g}({\cal M}) \times S^d{\cal M}$ be the incidence relation. Its fibre ${\cal Z}_t$ over a point $t \in T$ consists of all pairs $(L,\,v)$, where $L$ is a two--plane in $ I \!\! P^{d-g}_j := \phi_d^{-1}(j),\,\,j \in J_d(M_t)$, and $v \in I \!\! P^{d-g}_j$ is a point of $L$. Let $pr_1\,:\,{\cal Z} \to Grass_{2,\,d-g}({\cal M}),\,\,pr_2\,:\,{\cal Z} \to S^d{\cal M}$ be the canonical projections, and let $\sigma\,:\, T \to Grass_{2,\,d-g}({\cal M})$ be the regular section as in (3.2) above. Put ${\cal Z}_{{\cal D}} := pr_2^{-1}({\cal D}) \subset {\cal Z}$, ${\hat {\cal D}} := pr_1({\cal Z}_{{\cal D}}) \subset Grass_{2,\,d-g}({\cal M})$ and $T' := \sigma^{-1}({\hat {\cal D}}) \subset T$. Since the projection $pr_1$ is proper, ${\hat {\cal D}} \subset Grass_{2,\,d-g}({\cal M})$, and therefore also $T' \subset T$ are closed subvarieties of the corresponding varieties. Clearly, $t \in T'$ iff the linear series $g^2_d(t) = \sigma(t)$ on $M_t$ contains a divisor from ${\cal D}_t$. Recall that $Nod_{d,\,g} = T \setminus R_{\tau}$, where $R_{\tau}$ is the ramification divisor as in (2.14). By Lemma 3.1, we have that $PlNod_{d,\,g} = T \setminus (R_{\tau} \cup T')$. By Proposition 2.9, any irreducible component $I$ of $T = Imm_{d,\,g}$ contains a Pl\"ucker curve. Thus, $T' \cap I$ is a proper subvariety of $I$; in particular, codim$\,_{T}T' \ge 1$. Hence, $PlNod_{d,\,g}$ is, indeed, a Zariski open subset of $T = Imm_{d,\,g}$. This completes the proof of Theorem 2.1. \hfill $\Box$ \bigskip Theorem 2.1 implies \bigskip \noindent {\bf 3.4. Corollary.} {\it Any irreducible plane curve $C^*$ of genus $g$ and degree $n = 2(d + g - 1)$, where $d \ge 2g-1$, whose dual $C$ is an immersed curve, is a specialization of generic maximal cuspidal Pl\"ucker curves $C'^*$ of the same degree and genus\footnote{i.e. $C'^*$ has the maximal number of cusps allowed by Pl\"ucker's formulas.}. Hence, there is an epimorphism $\pi_1( I \!\! P^{2*} \setminus C^*) \to \pi_1( I \!\! P^{2*} \setminus C'^*)$. In particular, the former group is big (resp. non--amenable, non--almost solvable, non--almost nilpotent) if the latter one is so.} \bigskip \noindent {\it Proof.} By the class formula [Na, 1.5.4], the dual of an irreducible immersed plane curve of degree $d$ and genus $g$ has degree $d^* = 2(g + d - 1)$. By Theorem 2.1($a$) and ($b$), there is the diagram $$ \begin{picture}(800,60) \unitlength0.2em \thicklines \put(88,23){${\cal M}_T$} \put(64,1){$ I \!\! P^2_T$} \put(108,1){$ I \!\! P^{2^*}_T$} \put(86,2){$\longleftrightarrow$} \put(87,18){$\vector(-1,-1){12}$} \put(93,18){$\vector(1,-1){12}$} \put(70,14){$f$} \put(110,14){$f^*$} \end{picture} $$ where the morphism $f^*$ yields a simultaneous normalization of the dual family, so that for each $t \in T = Imm_{d,\,g}$ the image $f^*(M_t) =S_t^*$ is the dual curve of the curve $S_t = f(M_t)$ (see e.g. [Na, 1.5.1]). By ($c$), the subset $PlNod_{d,\,g}\subset T$ is Zariski open. The dual $S_t^*$, where $t \in PlNod_{d,\,g}$, is a maximal cuspidal curve of degree $d$ and genus $g$. Vice versa, any such curve is the dual $S_t^*$ of a nodal Pl\"ucker curve $S_t,\,t \in PlNod_{d,\,g}$. This yields the first assertion. The second one follows from a well known theorem of Zariski (see [Zar, p.131, Thm.5] or [Di, 4.3.2]). As for the third one, see (1.1) above. \hfill $\Box$ \section{Proof of Theorem 0.2} \bigskip The following lemma is a particular case of the Varchenko Equisingularity Theorem [Va, Theorem 5.3]. \bigskip \noindent {\bf 4.1. Lemma.} {\it Let $p\,:\,E \to B$ be a surjective morphism, where $E$, $B$ are smooth connected quasi--projective varieties. Then there exist a proper subvariety $A \subset B$ such that the restriction $p\,|\,(E \setminus H)$, where $H = p^{-1}(A)$, determines a smooth locally trivial fibre bundle $p \,:\,E \setminus H \to B \setminus A$.} \bigskip Let ${\Delta}$ be a hypersurface in a complex manifold $E$, $e \in$ reg$\,{\Delta}$ be a smooth point of ${\Delta}$, and ${\omega}$ be a small disc in $E$ centered at $e$ and transversal to ${\Delta}$. By {\it a vanishing loop} of ${\Delta}$ at $e$ we mean a loop ${\delta}$ in $E \setminus {\Delta}$ consisting of a path $\alpha$ which joins a base point $e_0 \in E \setminus {\Delta}$ with a point $e' \in {\omega} \setminus {\Delta}$ and a loop $\beta$ in ${\omega} \setminus {\Delta}$ with the base point $e'$ (i.e. $e$ is in the interior of $\beta$ in $\omega$). The next simple lemma is well known; for the sake of completeness we give its proof. \bigskip \noindent {\bf 4.2. Lemma.} {\it Let, as before, ${\Delta}$ be a hypersurface in a complex manifold $E$, and let ${\gamma}_0,\,{\gamma}_1\,:\,S^1 \to E \setminus {\Delta}$ be two loops with the base point $e_0 \in E \setminus {\Delta}$ joined in $E$ by a smooth homotopy ${\gamma}\,:\, S^1 \times [0,\,1] \to E$ transversal to ${\Delta}$, such that the image $S =$ Im$\,{\gamma}$ meets ${\Delta}$ at the points $e_1,\dots,e_k \in {\rm reg}\,{\Delta}$. Then ${\gamma}_0$ is homotopic in $E \setminus {\Delta}$ to a product ${\gamma}_1{\delta}_{i_1}\dots {\delta}_{i_k}$, where $(i_1,\dots,i_k)$ is a permutation of $(1,\dots,k)$ and ${\delta}_i$ is a vanishing loop of ${\Delta}$ at the point $e_i,\,\,i=1,\dots,k$. } \bigskip \noindent {\it Proof.} Slightly modifying the original homotopy and changing the numeration of the intersection points $e_1,\,\dots,\,e_n \in {\rm reg}\,{\Delta}$ we may assume that $e_i \in \gamma_{t_i} \cap {\Delta},\,\,i=1,\dots,k,$ correspond to different values $0 < t_1 < \dots < t_n < 1$ of the parameter of homotopy $t \in [0, 1]$. If $s_i \in [0,\,1],\,\,0 < s_1 < t_1 < \dots < t_n < s_{n+1} < 1,$ and ${\bar \gamma}_i = \gamma_{s_i}\,:\,S^1 \to E \setminus {\Delta} , \, i=1,\dots,n+1$, then clearly ${\bar \gamma}_{i+1}^{-1}\cdot {\bar \gamma}_i \approx {\delta}_i$, i.e. ${\bar \gamma}_i \approx {\bar \gamma}_{i+1}\cdot{\delta}_i$ in $E \setminus {\Delta}$, where ${\delta}_i$ is a vanishing loop of ${\Delta}$ at the point $e_i$, and ${\bar\gamma}_1 \approx {\gamma}_0,\,\,{\bar\gamma}_{n+1} \approx {\gamma}_1$. Thus, $\gamma_0 \approx {\gamma}_1 {\delta}_n\cdot \dots \cdot {\delta}_1$ in $E \setminus {\Delta}$, and the lemma follows. \hfill $\Box$ \bigskip In the proof of Theorem 0.2 below we use the following proposition. Actually, it follows from Lemma 1.5(C) in [No]. However, we give a proof which is different from that in [No]. \bigskip \noindent {\bf 4.3. Proposition.} {\it Let a morphism $p\,:\,E \to B$ of smooth quasiprojective varieties be a smooth fibration over $B$ with a connected generic fibre $F$ of positive dimension. Let ${\Delta} \subset E$ be a Zariski closed hypersurface which contains no entire fibre of $p$, i.e. $p^{-1}(b) \not\subset {\Delta}$ for each $b \in B$. Then we have the following exact sequence}: $$\pi_1(F \setminus {\Delta}) \stackrel{i_*}{\rightarrow} \pi_1(E \setminus {\Delta}) \stackrel{p_*}{\rightarrow} \pi_1(B) \to {\bf 1}\,.$$ \noindent {\it Proof.} By Lemma 4.1, there exist hypersurfaces $A \subset B$ and $D = H \cup {\Delta} \subset E$, where $H := f^{-1}(A)$, such that $p\,|\,(E \setminus D) \,:\, E \setminus D \to B \setminus A$ is a smooth fibration. In particular, $p\,|\,(E \setminus D)$ induces an epimorphism of the fundamental groups. Since the same is also true for the embedding $i\,:\,B \setminus A \hookrightarrow B$, and since $p_* = i_* \circ (p\,|\,E \setminus D)_*$, the exactness at the third term follows. It remains to prove that the homomorphism $$i_*\,:\,\pi_1(F \setminus {\Delta}) \to {\rm Ker}\,p_* \subset \pi_1(E \setminus {\Delta})$$ is surjective. Fix a generic fibre $F \not\subset D$ and base points $e_0 \in F \setminus D$ and $b_0 = p(e_0) \in B \setminus A$. Let a class $[{\gamma}_0] \in {\rm Ker}\,p_*$ be represented by a loop ${\gamma}_0\,:\,S^1 \to E \setminus {\Delta}$ with the base point $e_0$. We will show that ${\gamma}_0$ is homotopic in $E \setminus {\Delta}$ to a loop ${\gamma}'_0\,:\,S^1 \to F \setminus {\Delta}$ with the same base point. The loop ${\bar {\gamma}}_0 := p \circ {\gamma}_0\,:\,S^1 \to B$ with the base point $b_0 \in B$ is contractible. Let ${\bar {\gamma}} \,:\,S^1 \times [0;\,1] \to B$ be a contraction to the constant loop ${\bar {\gamma}}_1 \equiv b_0$. Since $p\,:\,E \to B$ is a fibration, there exists a covering homotopy ${\gamma}\,:\,S^1 \times [0;\,1] \to E$. Thus, we have ${\bar {\gamma}} = p \circ {\gamma}$ and ${\gamma}_1\,:\,S^1 \to F$. Fix a stratification of $D = {\Delta} \cup H$ which satisfies the Whitney condition A and contains the regular part ${\rm reg}\,D$ of $D$ as an open stratum. By Thom's Transversality Theorem, the homotopy ${\gamma}$ can be chosen being transversal to the strata of this stratification, and therefore such that its image meets the divisor $D$ only in a finite number of its regular points. Let it meet ${\Delta}$ at the points $e_1,\dots,e_k \in$ reg$\,({\Delta} \setminus H)$. We may also assume that the loop ${\gamma}_1\,:\,S^1 \to F$ does not meet $D$; in particular, $[{\gamma}_1] \in \pi_1 (F \setminus {\Delta};\,e_0)$. By Lemma 4.2, ${\gamma}_0$ is homotopic in $E \setminus {\Delta}$ to the product ${\gamma}_1{\delta}_{i_1}\dots {\delta}_{i_k}$, where ${\delta}_{i}$ is a vanishing loop of ${\Delta}$ at the point $e_i,\,\, i=1,\dots,k$. Note that all the transversal discs to ${\Delta}$ in $E$ centered at $e_i$ are homotopic (via the family of such discs). Hence, all the simple positive local vanishing loops of ${\Delta}$ at $e_i$ are freely homotopic in $E \setminus {\Delta}$. Therefore, performing further deformation of the vanishing loops ${\delta}_{i},\,\, i=1,\dots,k$, and taking into account our assumptions that dim$\,F > 0$ and ${\Delta}$ does not contain entirely a fibre of $p$, we may suppose that \smallskip \noindent (i) for each $ \,\, i=1,\dots,k$ the fibre of $p$ through the point $e_i$ is transversal to ${\Delta}$; \smallskip \noindent (ii) the loops ${\delta}_{i},\,\, i=1,\dots,k$, do not meet $H$, and the corresponding local loops $\beta_i,\,\,i=1, \dots, k$, are contained in the fibres of $p$. \smallskip \noindent Since $p^{-1}(A) \subset D$, we have that for each $i = 1,\dots,k$ the projection ${\bar {\delta}}_i := p \circ {\delta}_i$ of the loop ${\delta}_i$ does not meet the hypersurface $A \subset B$. By the construction, the loops ${\bar {\delta}}_i,\,\,i = 1,\dots,k,$ are contractible in $B \setminus A$. Applying the covering homotopy theorem to the smooth fibration $p\,:\,E \setminus D \to B \setminus A$ we may conclude that for each $i = 1,\dots,k,$ the loop ${\delta}_i$ is homotopic in $E \setminus D \subset E \setminus {\Delta}$ to a loop ${\delta}_i'\,:\,S^1 \to F \setminus D \subset F \setminus {\Delta}$. Hence, ${\gamma}_0$ is homotopic in $E \setminus {\Delta}$ to the product ${\gamma}'_0 := {\gamma}_1{\delta}'_{i_1}\dots {\delta}'_{i_k}\,,\,\,{\gamma}'_0\,:\,S^1 \to F \setminus {\Delta}$, and we are done. \hfill $\Box$ \bigskip \noindent {\bf 4.4. Duality, discriminants and the Zariski embedding} \medskip The following construction was used, for instance, in [Zar, pp.307, 326] and in [DoLib, sect.1, 3]. Let $M$ be an irreducible smooth projective variety, and let $L \subset H^0(M,\,{\cal L})$ be a linear system of effective divisors on $M$, where $\cal L$ is a linear bundle on $M$. It defines a rational mapping $\Phi_L\,:\,M \to I \!\! P (L^*)$. If $K \subset L$ is a linear subsystem, then the mapping $\Phi_K\,:\,M \to I \!\! P (K^*)$ is composed of the mapping $\Phi_L$ followed by the linear projection $\pi_{L,\,K}\,:\, I \!\! P (L^*) \to I \!\! P (K^*)$ which is dual to the tautological embedding $\rho_{K,\,L}\,:\, I \!\! P (K) \hookrightarrow I \!\! P (L)$. Set $C_L := \Phi_L (M) \subset I \!\! P (L^*)$ and $C_K := \Phi_K (M) \subset I \!\! P (K^*)$, so that $C_K = \pi_{L,\,K} (C_L)$. The dual variety ${\Delta}_L \subset I \!\! P(L)$ of $C_L \subset I \!\! P (L^*)$ is usually a hypersurface, which is called {\it the discriminant hypersurface of the linear system $L$}. The embedding $\rho_{K,\,L}$ yields the embedding of the discriminants ${\Delta}_K = I \!\! P (K) \cap {\Delta}_L \hookrightarrow {\Delta}_L$. \smallskip In particular, starting with a degree $d$ irreducible plane curve $C \subset I \!\! P^2$ with a normalization $M \to C$, denote by $K = g^2_d$ the linear system on $M$ of line cuts of $C$ and by $L = |g^2_d|$ the corresponding complete linear system. Since $g^2_d$ and therefore, also $L$ are base point free, they define morphisms $\Phi_K\,:\,M \to C \subset I \!\! P^2 = I \!\! P (K^*)$ resp. $\Phi_L\,:\,M \to C_L := \Phi_L(M) \hookrightarrow I \!\! P (L^*)$, and $C \subset I \!\! P^2$ is a projection of the curve $C_L \subset I \!\! P (L^*)$. Set $ I \!\! P^2_C := I \!\! P(K) \hookrightarrow I \!\! P(L)$. The discriminant ${\Delta}_L = C_L^*$ is, indeed, a projective hypersurface, and the dual curve $C^* \subset I \!\! P^2_C$ is an irreducible component of the plane cut ${\Delta}_K = I \!\! P(K) \cap {\Delta}_L$. The other irreducible components of ${\Delta}_K$ are special tangent lines of $C^*$ dual to the cusps of $C$ (by {\it a cusp} we mean here a singular point of a local irreducible analytic branch of $C$). We call these tangent lines {\it artifacts} [DeZa1]. Thus, the plane cut $ I \!\! P(K) \cap {\Delta}_L$ of the discriminant hypersurface ${\Delta}_L$ is irreducible iff $C$ is an immersed curve. The embedding $ I \!\! P^{2*} \cong I \!\! P^2_C \hookrightarrow I \!\! P(L)$ which represents $C^*$ as a plane cut of the discriminant hypersurface ${\Delta}_L$ is called {\it the Zariski embedding} (see [Zar, pp.307, 326; DeZa1]). By definition, the dual variety $C_L^* = {\Delta}_L$ consists of the points $x \in I \!\! P(L)$ such that the dual hyperplane $x^* \subset I \!\! P_L$ cuts out of $C_L$ a non-reduced divisor on the normalization $M$ of $C_L$. If $x \in C_L$ is a cusp, then, clearly, the dual hyperplane $x^*$ is an irreducible component of ${\Delta}_L$. Thus, the discriminant ${\Delta}_L$ is irreducible iff $C_L$ was an immersed curve. In particular, this is the case if $C = \pi_{L,\,K} (C_L)$ is an immersed curve. Vice versa, if $C_L$ is an immersed curve, then the same is true for its generic projection onto the plane. Or, what is the same, if the discriminant ${\Delta}_L$ is irreducible, then its generic plane section is irreducible, too. The projectivization $ I \!\! P(L)$ of the complete linear system $L$ of degree $d$ divisors on $M$ coincides with a fibre of the Abel--Jacobi map $\phi_d\,:\,S^dM \to J_d(M)$ (see (3.2)), so that $ I \!\! P^2_C$ is a plane in this fibre. We still call the morphism $ I \!\! P^2_C \hookrightarrow S^dM$ {\it the Zariski embedding}. The hypersurface ${\Delta}_d \subset S^dM$ which consists of the non-reduced degree $d$ effective divisors on $M$ is also called {\it the discriminant hypersurface}. It is the image of the diagonal hypersurfaces of the direct product $M^d$ via the Vieta map $M^d \to S^dM$. Thus, ${\Delta}_L = I \!\! P(L) \cap {\Delta}_d$, where $ I \!\! P(L)$ has been identified with a fibre $F_j := \phi_d^{-1}(j),\,j \in J_d(M)$, of $\phi_d$. \bigskip \noindent {\bf 4.5.} {\it Proof of Theorem 0.2.} By Corollary 3.4, we may suppose that $C$ is a generic nodal Pl\"ucker curve of degree $d$ and geometric genus $g$, where $d \ge 2g-1$. By Mattuck's Theorem (see (3.2)), the $d$-th Picard bundle $\phi_d\,:\,S^dM \to J_d(M)$, where $M$ is a normalization of $C$, is a projective bundle with a generic fibre $F \cong I \!\! P^{d - g}$. By (4.4), the dual curve $C^*$ can be identified with the plane cut of the discriminant hypersurface ${\Delta}_d \subset S^dM$ by the plane $ I \!\! P^2_C$ via its Zariski embedding $ I \!\! P^2_C \hookrightarrow F_0 := I \!\! P(L) \subset S^dM$, where $L = |g^2_d|$ and $g^2_d$ is the linear system on $M$ of line cuts of $C$. If the group $\pi_1 ( I \!\! P^2_C \setminus \Delta_d)$ is big for a generic plane $ I \!\! P^2_C \subset F_0 \cong I \!\! P^{d - g}$, then by Zariski's Lefschetz type Theorem [Zar, p.279; Di, 4.1.17], it is big for any such plane, so that $ I \!\! P^2_C \subset F_0$ might be assumed being generic. Indeed, a section $S$ of the discriminant hypersurface ${\Delta}_L = F_0 \cap {\Delta}_d$ by a generic plane $ I \!\! P \subset F_0$ is an irreducible curve with the same normalization $M$ and with the dual $S^* \subset I \!\! P^2$ an immersed curve of degree $d$ and genus $g$. Thus, we may start with $C = S^*$ and obtain $C^* = S = I \!\! P^2_C \cap {\Delta}_d$. Note that such a generic linear system $K = g^2_d \subset L$, where $ I \!\! P = I \!\! P(K)$, defines a morphism $M \to I \!\! P^2$ such that its image coincides with $C = S^*$. Since by Theorem 2.1(c), $PlNod_{d,\,g}$ is a Zariski open subset of $Imm_{d,\,g}$, the curve $C=S^*$ obtained in this way is a nodal Pl\"ucker one. Applying the Zariski Lefschetz type Theorem we get an isomorphism $$\pi_1 (F_0 \setminus \Delta_d) \cong \pi_1 ( I \!\! P^2_C \setminus \Delta_d) \cong \pi_1 ( I \!\! P^2 \setminus C^*)\,.$$ By Proposition 4.3, we have the exact sequence $$\pi_1 (F_0 \setminus \Delta_d) \to \pi_1 (S^d M \setminus \Delta_d) \to \pi_1 (J_d(M)) \cong Z \!\!\! Z^{2g} \to {\bf 1}\,.$$ It follows that $ \pi_1 ( I \!\! P^2 \setminus C^*)$ is a big group if $\pi_1 (S^d M \setminus \Delta_d)$ is big (cf. (1.1)). But $\pi_1 (S^d M \setminus \Delta_d)$ is the braid group $B_{d,\,g}$ of $M$ with $d$ strings which is big (see Lemma 1.2($b$)). This completes the proof. \hfill $\Box$ \bigskip \noindent {\it Remark.} A presentation of the group $\pi_1( I \!\! P^2 \setminus C)$ for a generic maximal cuspidal curve $C \subset I \!\! P^2$ of genus 0 or 1 was found by Zariski [Zar, p. 307]; see also [Ka] for $g \le {d-1 \over 2}$, where $d = {\rm deg}\,C^*$. The result of [Ka] is based on the statement in [DoLib] that for $d \ge 2g-1$ the $d$--th Abel--Jacobi mapping $\phi_d\,:\,S^dM \to J(M)$ restricted to the complement of the discriminant hypersurface ${\Delta}_d \subset S^dM$ is a Serre fibration, so that the long exact homotopy sequence is available. But the indication given in [DoLib] does not seem to be sufficient for the proof. Another proof of the exactness of the above sequence of fundamental groups extended to the left by the term $\bf 1$ has been recently obtained in [KuShi]. Once again, this leads to a presentation of the group $\pi_1( I \!\! P^2 \setminus C)$. \bigskip \begin{center} {\LARGE References} \end{center} {\footnotesize \noindent [ACGH] E. Arbarello, M. Cornalba, P.A. Griffiths, J. Harris, {\sl Geometry of algebraic curves. I}, N.Y. e.a.: Springer, 1985 \noindent [AC] E. Arbarello, M. Cornalba. {\sl Su una conjettura di Petri}, Comment. Math. Helv. 56 (1981), 1-38 \noindent [Au] A. B. Aure. {\sl Pl\"ucker conditions on plane rational curves}, Math. Scand. 55 (1984), 47--58, with {\sl Appendix} by S. A. Str\"omme, ibid. 59--61 \noindent [BPVV] W. Barth, C. Peters, A. Van de Ven. {\sl Compact complex surfaces}, N.Y. e.a.: Springer, 1984 \noindent [Be] D. Bennequin. {\sl Entrelacements et \'equations de Pfaff}, Ast\'erisque, 107--108 (1983), 87--161 \noindent [BinFl] J. Bingener, H. Flenner. {\sl On the fibres of analytic mappings}, in: Complex Analysis and Geometry, V. Ancona and A. Silva, eds., N.Y.: Plenum Press, 1993, 45--101 \noindent [Bi] J.S. Birman, {\sl Braids, links, and mapping class groups}, Princeton Univ. Press, Princeton, NJ, 1974 \noindent [Bo] A. Borel. {\sl K\"ahlerian coset spaces of semisimple Lie groups}, Proc. Nat. Acad. Sci., USA, 40, No. 12 (1954), 1147--1154 \noindent [BoHC] A. Borel, Harish--Chandra. {\sl Arithmetic subgroups of algebraic groups}, Ann. of Math. 75 (1962), 485--535 \noindent [CoZi] D. J. Collins, H. Zieschang. {\sl Combinatorial group theory and fundamental groups}, In: Algebra VII, Encyclopaedia of Math. Sci. Vol. 58, Berlin e.a.: Springer, 1993, 3--166 \noindent [DeZa1] G. Dethloff, M. Zaidenberg. {\sl Plane curves with hyperbolic and C--hyperbolic complements}, Ann. Sci. Ecole Norm. Super. Pisa (to appear); Pr\'epublication de l'Institut Fourier de Math\'ematiques, 299, Grenoble 1995, 44p. \noindent [DeZa2] G. Dethloff, M. Zaidenberg. {\sl Examples of plane curves of low degrees with hyperbolic and C--hyperbolic complements.} In: Geometric Complex Analysis, J. Noguchi e.a. eds. World Scientific Publ. Co., Singapore 1996., 176--193 \noindent [DeOrZa] G. Dethloff, S. Orevkov, M. Zaidenberg. {\sl Plane curves with a big fundamental group of the complement.} Pr\'epublication de l'Institut Fourier de Math\'ematiques, 354, Grenoble 1996, 26p.; E-print alg-geom/9607006 \noindent [Di] A. Dimca, {\sl Singularities and topology of hypersurfaces}, Berlin e.a.: Springer, 1992 \noindent [DoLib] I. Dolgachev, A. Libgober. {\sl On the fundamental group of the complement to a discriminant variety}, In: Algebraic Geometry, Lecture Notes in Math. 862, 1--25, N.Y. e.a.: Springer, 1981 \noindent [GoShVi] V. V. Gorbatsevich, O. V. Shvartsman, E. B. Vinberg. {\sl Discrete subgroups of Lie groups.} In: Lie Groups and Lie Algebras II, Encyclopaedia of Math. Sci. Vol. 21, Springer : NY e.a., 1995 \noindent [Ha] J. Harris. {\sl On the Severi problem}, Invent. Math. 84 (1986), 445--461 \noindent [Hart] R. Hartshorn. {\sl Algebraic geometry}, NY e.a.: Springer, 1977 \noindent [He] S. Helgason. {\sl Differential geometry, Lie groups and symmetric spaces}, N.Y. e.a.: Academic Press, 1978 \noindent [Iv] N. V. Ivanov. {\sl Algebraic properties of the Teichm\"uller modular group}, Soviet Math. Dokl. 29 (1984), No.2, 288--291 \noindent [Ka] J. Kaneko. {\sl On the fundamental group of the complement to a maximal cuspidal plane curve}, Mem. Fac. Sci. Kyushu Univ. Ser. A. 39 (1985), 133-146 \noindent [Ko] K. Kodaira. {\sl A theorem of completeness of characteristic systems for analytic families of compact submanifolds of complex manifolds}, Ann. Math. 75 (1962), 146--162 \noindent [Kos] J.-L. Koszul. {\sl Sur la forme hermitienne canonique des espaces homog\`enes complexes}, Can. J. Math. 7 (1955), 562--576 \noindent [KuShi] Vik. S. Kulikov, I. Shimada. {\sl On the fundamental group of complements to the dual hypersurfaces of projective curves.} Preprint Max-Planck-Institute f\"ur Mathematik, MPI 96-32, Bonn, 1996, 1--15 \noindent [Lib] A. Libgober. {\sl Fundamental groups of the complements to plane singular curves}, Proc. Sympos. in Pure Mathem. 46 (1987), 29--45 \noindent [Lin] V. Ja. Lin. {\sl Liouville coverings of complex spaces, and amenable groups}, Math. USSR Sbornik, 60 (1988), 197--216 \noindent [LinZa] V. Ja. Lin, M. Zaidenberg. {\sl Liouville and Carath\'eodory coverings in Riemannian and complex geometry.} Preprint Max-Planck-Institute f\"ur Mathematik, MPI 96-110, Bonn, 1996, 1--20 (see in this volume) \noindent [LySu] T. Lyons, D. Sullivan. {\sl Function theory, random paths, and covering spaces}, J. Diff. Geom. 19 (1984), 299--323 \noindent [Ma] A. Mattuck. {\sl Picard bundles}, Illinois J. Math. 5 (1961), 550--564 \noindent [MC] J. McCarthy. {\sl A "Tits-alternative" for subgroups of surface mapping class groups}, Trans. Amer. Math. Soc. 291 (1985), 583--612 \noindent [Mi] J. Milnor. {\sl Singular points of complex hypersurfaces}, Princeton, New Jersey : Princeton University Press, 1968 \noindent [MoTe] B. Moishezon, M. Teicher. {\sl Fundamental groups of complements of branch curves as solvable groups}, Duke E-print alg--geom/9502015, 1995, 17p. \noindent [Na] M. Namba. {\sl Geometry of projective algebraic curves}, N.Y. a.e.: Marcel Dekker, 1984 \noindent [No] M.V. Nori. {\sl Zariski's conjecture and related problems}, Ann. scient. Ec. Norm. Sup. 16 (1983), 305--344 \noindent [O] M. Oka. {\sl Symmetric plane curves with nodes and cusps}, J. Math. Soc. Japan, 44, No. 3 (1992), 375--414 \noindent [OSh] A. Yu. Ol'shanskij, A. L. Shmel'kin. {\sl Infinite groups}, In: Algebra IV, Enciclopaedia of Math. Sci. 37, Berlin e.a.: Springer, 1993, 3--95 \noindent [Ra] M. S. Raghunathan. {\sl Discrete subgroups of Lie groups}, Berlin e.a.: Springer, 1972 \noindent [Se] F. Severi. {\sl Vorlesungen \"{u}ber algebraische Geometrie}, Leipzig: Teubner, 1921 \noindent [Sh1] G. B. Shabat. {\sl The complex structure of domains covering algebraic surfaces}, Functional Analysis Appl. 11 (1977), 135--142 \noindent [Sh2] G. B. Shabat. {\sl On families of curves covered by bounded symmetric domains}, Serdica Bulg. Math. Public. 11 (1985), 185--188 (in Russian) \noindent [Ti] J. Tits. {\sl Free subgroups in linear groups}, J. Algebra, 20 (1972), 250--270 \noindent [Va] A.N. Varchenko. {\sl Theorems on the topological equisingularity of families of algebraic varieties and families of polynomial mappings}, Math. USSR Izvestija, Vol. 6 (1972), No. 5, 957--1019 \noindent [Zar] O. Zariski. {\sl Collected Papers}. Vol III : {\sl Topology of curves and surfaces, and special topics in the theory of algebraic varieties}, Cambridge, Massachusets e. a.: The MIT Press, 1978 \bigskip \noindent Gerd Dethloff, Mathematisches Institut der Universit\"at G\"ot\-tin\-gen, Bunsenstrasse 3-5, 37073 G\"ot\-tin\-gen, Germany. e-mail: [email protected] \bigskip \noindent Stepan Orevkov, System Research Institute RAN, Moscow, Avtozavodskaja 23, Russia. e-mail: [email protected] \bigskip \noindent Mikhail Zaidenberg, Universit\'{e} Grenoble I, Institut Fourier et Laboratoire de Math\'ematiques associ\'e au CNRS, BP 74, 38402 St. Martin d'H\`{e}res--c\'edex, France. e-mail: [email protected]} \end{document}
1995-01-18T06:20:53
9501
alg-geom/9501009
en
https://arxiv.org/abs/alg-geom/9501009
[ "alg-geom", "math.AG" ]
alg-geom/9501009
Fantechi Barbara
Barbara Fantechi and Rita Pardini
On the Hilbert scheme of curves in higher-dimensional projective space
latex, 12 pages, no figures
null
null
UTM 448
null
In this paper we prove that, for any $n\ge 3$, there exist infinitely many $r\in \N$ and for each of them a smooth, connected curve $C_r$ in $\P^r$ such that $C_r$ lies on exactly $n$ irreducible components of the Hilbert scheme $\hilb(\P^r)$. This is proven by reducing the problem to an analogous statement for the moduli of surfaces of general type.
[ { "version": "v1", "created": "Tue, 17 Jan 1995 15:27:37 GMT" } ]
2015-06-30T00:00:00
[ [ "Fantechi", "Barbara", "" ], [ "Pardini", "Rita", "" ] ]
alg-geom
\section{Introduction} It is well-known that the Hilbert scheme parametrizing subschemes of $\P^r$ can be singular at points corresponding to smooth curves as soon as $r\ge 3$; actually Mumford \cite{Mu} gave an example of an everywhere singular irreducible component. If $r=3$, it has been proven in \cite{EHM} that the open subset of the Hilbert scheme parametrizing smooth curves in $\P^3$ with given genus and degree can have arbitrarily many components when the genus and the degree grow (in fact, they prove that no polynomial estimate on the number of such components holds). Our main result is the following: \smallskip \noindent {\bf Theorem \ref{mainthm2}.}{\em\ Let $n\ge 3$ be an integer. Then there exist infinitely many integers $r$, and for each of them a smooth, irreducible curve $C_r\subset \P^r$ such that $C_r$ lies exactly on $n$ components of the Hilbert scheme of $\P^r$.} \smallskip The idea of the proof is very simple. Firstly, we modify a construction of \cite{FP} to obtain a regular surface $S$ of general type which lies on $n$ components of the moduli space; secondly, we consider a suitable pluricanonical embedding of this surface and intersect its image with a high-degree hypersurface $F$ to construct the curve $C$ we are interested in. Finally, we prove that all embedded deformations of $C$ are induced by embedded deformations of $F$ and $S$. \par\noindent {\em Acknowledgements}. We are grateful to Ciro Ciliberto, who told us about this problem and suggested that we might apply to it the results of \cite{FP}. \section{Notation and preliminaries} All varieties will be assumed smooth and projective over the complex numbers unless the contrary is explicitly stated. A variety $Y$ will be called regular if $H^1(Y,\O_Y)=0$. If ${\cal F}$ is a sheaf on $Y$, let $h^i(Y,{\cal F})=\dim H^i(Y,{\cal F})$. If $t$ is a real number, we denote its integral part by $[t]$. Let $\zeta_3=\exp(2\pi i/3)$. In this paper we will be concerned with abelian covers of a very special type; we collect here the necessary notational set-up. Let $n$ be an integer $\ge 2$, and let $G={\bf Z}_3^n$, $G^*$ its dual; let $e_1,\ldots,e_n$ be the canonical basis of $G$, and $\chi_1,\ldots,\chi_n$ the dual basis of $G^*$ (i.e., $\chi_j(e_i)=1$ if $i\ne j$ and $\chi_j(e_j)=\zeta_3$). Let $e_0=-(e_1+\ldots+e_n)$. Let $I=\{0,\ldots,n\}$, and to each $i\in I$ associate the pair $(H_i,\psi_i)$ where $H_i$ is the cyclic subgroup of $G$ generated by $e_i$, and $\psi_i\in H_i^*$ is the character such that $\psi_i(e_i)=\zeta_3$. Let $Y$ be a smooth projective variety, and $(G,I)$ as above: a $(G,I)$-cover of $Y$ is a normal variety $X$ and a Galois cover $f:X\to Y$ with Galois group $G$ and (nonempty) branch divisors $D_i$ (for $i\in I$) having $(H_i,\psi_i)$ as inertia group and induced character (see \cite{Pa} for details). \begin{lem} To give a smooth $(G,I)$-cover of $Y$ is equivalent to giving line bundles $L$ and $F_j$, for $j=1,\ldots,n$, together with smooth nonempty divisors $D_i\in |M_i|$ {\rm(}where $M_0=L$ and, for $i\ge 1$, $M_i=L-3F_i${\rm)} such that the union of the $D_i$'s has normal crossings. \end{lem} \begin{Pf} {}From \cite{Pa} we know that the cover is determined by its reduced building data, divisors $D_i$ for $i\in I$ and line bundles $L_j$ for $j=1,\ldots,n$ satisfying the relation $3L_j\equiv D_j+2D_0$. Letting $M_i=\O(D_i)$, and putting $M_0=L$, $F_j=L-L_j$, the equations become precisely $M_j=L-3F_j$. \end{Pf} As the natural map $\bigoplus_{i\in I}H_i\to G$ is surjective, the covers we consider will be totally ramified. For $\chi\in G^*$, let as usual $L_\chi^{-1}$ be the corresponding eigensheaf in the direct sum decomposition of $f_*\O_X$; in the above notation, we will have (for $\chi=\chi_1^{\alpha_1}\cdots\chi_n^{\alpha_n}$): \setcounter{equation}{0} \begin{equation}\label{lchi} L_\chi=n_\chi L-\sum_{j=1}^n \alpha_jF_j, \end{equation} where $n_\chi=-[(-\alpha_1-\ldots-\alpha_n)/3]$. In particular note that $n_\chi\ge 1$ when $\chi\ne 1$, and $n_\chi=1$ if and only if $1\le\sum \alpha_j\le 3$. We will write $L_j$ instead of $L_{\chi_j}$. Recall from \cite{Pa}, proof of proposition 4.2 on page 208, that \begin{equation}\label{canonico} 3K_X=\pi^*(3K_Y+2(n+1)L-6\sum F_j). \end{equation} \smallskip We now recall some results from \cite{FP} in a simplified form (fit for our situation). For details and proofs see \cite{FP}, \S 5. \begin{rem}\label{rema}{\rm (1) Let ${\cal Y}\to B$ be a smooth projective morphism (with $B$ a smooth, connected quasiprojective variety) together with an isomorphism between ${\cal Y}_o$ and $Y$ for some $o\in B$, and assume that $Y$ is regular and that the morphism ${\cal Y}\to B$ has a section $\sigma$. Let $L$ be a line bundle on $Y$; assume that $c_1(L)$ is kept fixed by the monodromy action of $\pi_1(B,o)$ on $H^2(Y,{\bf Z})$. Then for each $b\in B$ there is a canonical induced class $c_1(L_b)$ on ${\cal Y}_b$. If, for all $b\in B$, the class $c_1(L_b)$ is of type $(1,1)$, then $L$ can be extended to a line bundle $\L$ over ${\cal Y}$, flat over $B$; this extension is unique if we require that its restriction to $\sigma(B)$ be trivial. This follows by applying the results on p.~20 of \cite{mumford}, and by noting that the relative Picard scheme of ${\cal Y}$ over $B$ is \'etale over $B$ since all fibres are smooth and regular (it is surjective as $c_1(L_b)$ is always of type $(1,1)$); the condition on the monodromy action implies then that the component of the relative Picard scheme containing $[L]$ is in fact isomorphic to $B$. Let $L_b$ be the restriction of $\L$ to ${\cal Y}_b$. \noindent (2) If $h^0({\cal Y}_b,L_b)$ is either constant in $b$, or if it only assumes the values $1$ (for $b\in Z$) and $0$, then there is a (nonunique) quasiprojective variety $W^L\to B$ such that $W^L_b$ is canonically isomorphic to $H^0({\cal Y}_b,L_b)$; $W^L$ is smooth and irreducible in the former case, while in the latter it is the union of one component isomorphic to $B$ and another being the total space of a line bundle over $Z$ (compare with \cite{FP}, theorem 5.8 and remark 5.11). }\end{rem} \begin{assu}\label{ass}{\rm Let $S=\{({i,\chi})\in I\times G^*|\chi_{|H_i}\ne\psi_i^{-1}\}$. Let $X\to Y$ be a smooth $(G,I)$-cover as in lemma 2.1, and ${\cal Y}\to B$ be a smooth projective morphism (with $(B,o)$ a pointed space, and ${\cal Y}_o$ isomorphic to $Y$), such that remark \ref{rema}, (1) applies to ${\cal Y}\to B$, for the line bundle $L$ and for each of the $F_j$'s. Assume moreover that remark \ref{rema}, (2) applies for the line bundles $M_i-L_\chi$ for $({i,\chi})\in S$, yielding varieties $W^{i,\chi}$: let $W$ be the fibred product of the $W^{i,\chi}$ over $B$. Finally, assume that the germ of $B$ at $o$ maps smoothly to the base of the Kuranishi family of $Y$, and that the cohomology groups $H^1(Y,L_\chi^{-1})$ and $H^1(Y,T_Y\otimes L_\chi^{-1})$ vanish for each $\chi\in G^*\setminus 1$. }\end{assu} \begin{thm}\label{fromFP} Assume that assumption {\rm\ref{ass}} holds, and let $w\in W$ be a point over $o\in B$ corresponding to sections $s_{i,\chi}$ such that $s_{i,\chi}=0$ if $\chi\ne 1$, and $s_{i,1}$ defines $D_i$ for $i=0,\ldots,n$. Assume also that $X$ has ample canonical class. One can construct a family of natural deformations of $(G,I)$-covers ${\cal X}\to W$; the induced map from the germ of $w$ in $W$ to the Kuranishi family of $X$ is smooth {\rm(}and, in particular, surjective{\rm)}. Moreover, the flat, projective morphism ${\cal X}\to W$ defines a rational map from $W$ to the moduli of surfaces with ample canonical class, regular at $w$; this map is dominant on each irreducible component of the moduli containing $X$. \end{thm} \begin{Pf} Let $\L$ (resp.~${\cal F}_j$) be the line bundle induced by $L$ (resp.~$F_j$) on $W$; as we {\em define} ${\cal M}_0$ to be $\L$, $\L_j$ to be $\L-{\cal F}_j$ and ${\cal M}_j$ to be $\L-3{\cal F}_j$ for $j=1,\ldots,n$, there are global, canonical isomorphisms $\phi_j:3\L_j\to {\cal M}_j+2{\cal M}_0$. By \cite{FP}, theorem 5.12, the germ of $W$ at $w$ maps smoothly to the base of the Kuranishi family of $X$. If $M$ is an irreducible component of the moduli containing $[X]$, by the previous result the image of $W$ contains an open set in $M$ (in the strong topology), hence it cannot be contained in a closed subset (in the Zariski topology) and is therefore dominant. \end{Pf} \section{Moduli of surfaces of general type} The aim of this section is the proof of theorem \ref{mainthm1}, i.e., the explicit construction of regular surfaces with ample canonical class lying on arbitrarily many components of the moduli. This construction can be carried out in a much more general setting (see remark \ref{+gen}); we consider only the case needed for our applications, since it is easier to describe. For $S$ a smooth projective surface and $\xto.n$ pairwise distinct points of $S$, we let $B\ell(S;\xto.n)$ denote the surface obtained by blowing up $S$ at $\xto.n$. \begin{constr}{\rm Let $S$ be a regular surface, $x_0\in S$, $n$ a positive integer; let $B=B(S,n)$ be the variety parametrizing data $(\xto.n,\yto.n)$ where the $x_i$'s are pairwise distinct points in $S$ (for $i=0,\ldots,n$), the $y_i$'s are pairwise distinct points in $B\ell(S;\xto.n)$, such that $y_i$ is not infinitely near to $x_j$ for $i\ne j\ge 1$ and none of the $y_i$'s lies over $x_0$. $B$ is a smooth quasiprojective variety, which is naturally isomorphic to an open subset of the product of $n$ copies of $S\times S$ blown up along the diagonal. Let ${\cal Y}\to B$ be the smooth projective family such that ${\cal Y}_b$, the fibre of ${\cal Y}$ over the point $b$, is isomorphic to $B\ell(B\ell(S;\xto.n);\yto.n)$ for $b=(\xto.n,\yto.n)$.} \end{constr} Note that the morphism ${\cal Y}\to B$ has a section, given by mapping $b\in B$ to the inverse image of $x_0$ in ${\cal Y}_b$. \begin{lem} Assume that $S$ is rigid. Let $B^0$ be the open set in $B$ where $Aut(S)$ acts freely (the action being the natural one). Then if $b\in B^0$, the natural map from the germ of $B$ in $b$ to the Kuranishi family of ${\cal Y}_b$ is smooth of relative dimension $h^0(S,T_S)$. \end{lem} \begin{Pf} The proof is easy and left to the reader. \end{Pf} \begin{rem}{\rm For any $b\in B$, $b=(\xto.n,\yto.n)$, there is a canonical isomorphism $$N\!S({\cal Y}_b)=N\!S(S)\oplus{\bf Z} e'_1\oplus\ldots\oplus{\bf Z} e_n'\oplus{\bf Z} e''_1\oplus \ldots\oplus{\bf Z} e_n'',$$ where $e_i'$ is the pullback from $B\ell(S;\xto.n)$ of the class of the exceptional divisor over $x_i$, and $e_i''$ is the class of the exceptional divisor over $y_i$. We will consider this isomorphism fixed, and denote this group by $N\!S$. We also let $f_i$ denote $e_i'-e_i''$. Since $S$ is regular, so are all the ${\cal Y}_b$'s and we will not need to distinguish between line bundles and their Chern classes. }\end{rem} \begin{defn}{\rm Let $L\in N\!S$, $G={\bf Z}_3^n$ as in \S 2; for $\chi\in G^*$, let $L_\chi\in N\!S$ be defined by equation (\ref{lchi}), with $F_i=f_i$. Let $B_{L}$ be the open subset of $B$ consisting of the $b$'s such that \begin{enumerate} \item the cohomology groups $H^1({\cal Y}_b,L_\chi^{-1})$, $H^1({\cal Y}_b,T_{{\cal Y}_b}\otimes L_\chi^{-1})$ are zero for each $\chi\in G^*\setminus 1$; \item the line bundles $L$ and $L-3F_j$ are very ample on ${\cal Y}_b$, for $j=1,\ldots,n$; \item the line bundles $L-K_{{\cal Y}_b}$ and $L-3F_j-K_{{\cal Y}_b}$ are ample on ${\cal Y}_b$, for $j=1,\ldots,n$; \item the line bundle $3K_Y+2(n+1)L-6\sum F_j$ is ample on ${\cal Y}_b$. \end{enumerate}} \end{defn} Note that the first condition is needed to ensure that assumption \ref{ass} is satisfied; the second allows one to choose smooth divisors in the linear systems $|L|$ and $|L-3F_j|$ meeting transversally; the third implies that these linear systems have constant dimension when $b$ varies; and the fourth ensures, in view of equation (\ref{canonico}), that the cover so obtained has ample canonical class (recall that the pullback of an ample line bundle via a finite map is again ample). \begin{lem}\label{basic} Let $Y$ be a smooth surface containing $m$ disjoint irreducible curves $C_1,\ldots,C_m$, such that $C_i^2<0$. Then: \begin{enumerate} \item for any choice of nonnegative integers $a_1,\ldots,a_m$, the linear system $|a_1C_1+\ldots+a_mC_m|$ contains only the divisor $a_1C_1+\ldots+a_mC_m$; \item for any choice of nonnegative integers $a_1,\ldots,a_{m-1}$, and for any $b>0$, the linear system $|a_1C_1+\ldots+a_{m-1}C_{m-1}-bC_m|$ is empty. \end{enumerate}\end{lem} \begin{Pf} (1) We prove the theorem by induction on $a_1+\ldots+a_m$, the case where this sum is zero being trivial. Assume without loss of generality that $a_1\ge 1$, and let $C\in |a_1C_1+\ldots+a_mC_m|$; then $C\cdot C_1=a_1C_1^2<0$, hence $C$ must have a common component with $C_1$; therefore $C=C_1+C'$, $C'\in |(a_1-1)C_1+\ldots+a_mC_m|$, and by induction the proof is complete.\par \noindent (2) Assume that there exists $C\in |a_1C_1+\ldots+a_{m-1}C_{m-1}-bC_m|$. Then $C+bC_m\in |a_1C_1+\ldots+a_{m-1}C_{m-1}|$, contradicting (1). \end{Pf} \begin{cor} Let $b\in B$, $b=(\xto.n,\yto.n)$ and let $a_1,\ldots,a_m$ be nonnegative integers. Then the line bundle $a_1f_1+\ldots+a_mf_m$ on ${\cal Y}_b$ is effective if and only if $y_i$ is infinitely near to $x_i$ for every $i$ such that $a_i>0$, and in this case it has only one section. \end{cor} \begin{Pf} If $y_i$ is infinitely near to $x_i$, then $f_i$ is a $(-2)$ curve; otherwise it is the difference of two disjoint $(-1)$ curves. In the former case lemma \ref{basic}, (1) applies and in the latter case \ref{basic}, (2) applies. \end{Pf} \begin{nota}{\rm We will denote by $E$ the closed subset of $B$ consisting of the points $b$ such that $f_i$ is effective on ${\cal Y}_b$ for $i=1,\ldots,n$.} \end{nota} \begin{lem} Let $L\in NS$ be a line bundle and assume that $E\cap B_L\ne\emptyset$. Then assumption {\rm\ref{ass}} holds for the restriction of ${\cal Y}\to B$ to $B_L$; applying theorem {\rm\ref{fromFP}} yields a quasiprojective variety $W$. In this case $W$ is the union of $2^n$ smooth irreducible components $W_A$, indexed by subsets $A\subset \{1,\ldots,n\}$. The dimension of $W_A$ and $W_{A'}$ are equal if $\#A=\#A'$, and in particular one has: $$\dim W_A-\dim W_{\emptyset}=\frac{1}{6}(\#A^3+6\#A^2-\#A).$$ The $W_A$'s have a nonempty intersection. \end{lem} \begin{Pf} The verification that assumption \ref{ass} holds is easy and we leave it to the reader. For $A\subset \{1,\ldots,n\}$, let $E_A=\{b\in B_L|f_i\ \mbox{is effective for}\ i\in A\}$ and let $W_A\subset W$ be defined by $$W_A=\{(b,s_{{i,\chi}})|b\in E_A\ \mbox{and}\ s_{{i,\chi}}=0\ \mbox{for}\ \chi \ne 1\ \mbox{and}\ i\notin A\}.$$ It is easy to check that $W_A$ is smooth over $E_A$ of dimension $1/6(\#A^3+6\#A^2+5\#A)$; on the other hand $E_A$ is smooth of codimension $\#A$ in $B_L$. Finally, $W$ is the union of the $W_A$'s, which are easily seen to be irreducible components. Moreover, the intersection of the $W_A$'s is clearly equal to $W_\emptyset$ intersected with the inverse image of $E\cap B_L$. \end{Pf} \begin{thm}\label{mainthm1} Let $L\in N\!S$ and $b\in E\cap B_L\cap B^0$. Let $f:X\to Y={\cal Y}_b$ be a smooth $(G,I)$-cover with building data $(D_i,L_\chi)$; let $w\in W_b$ be a point corresponding to a choice of equations $s_i\in \O_Y(D_i)$ defining $D_i$, with $s_{{i,\chi}}=0$ for all $\chi\ne 1$. Then the natural map from the germ of $W$ in $w$ to the Kuranishi family of $X$ is smooth. In particular, the Kuranishi family of $X$ is the union of $2^n$ irreducible components, $n+1$ of which have pairwise different dimension. Moreover, the surface $X$ lies on exactly $n+1$ components of the moduli space, having pairwise different dimensions. \end{thm} \begin{Pf} The first statement is a straightforward application of theorem \ref{fromFP}, in view of the previous lemma. To prove the second, note that the map from $W$ to the moduli space factors through the action of the symmetric group on $n$ letters, $\Sigma_n$. The quotient $W/\Sigma_n$ has exactly $n+1$ irreducible components of pairwise different dimensions. By the previous result, each of these components dominates a component of the moduli; the $n+1$ components so obtained must all be distinct, as they have different dimensions. \end{Pf} \begin{rem}\label{suffample} {\rm If $L\in NS$ is sufficiently ample, then the intersection $E\cap B_L\cap B^0$ is nonempty, hence the theorem applies yielding infinitely many surfaces with different invariants. } \end{rem} \begin{cor}\label{HilbX} Given integers $n\ge 2$ and $m\ge 5$, for infinitely many values of $r$ there exists a smooth, regular surface $X$ in $\P^r$ such that $\O_X(1)=mK_X$ and $X$ lies on exactly $n+1$ irreducible components of the Hilbert scheme. \end{cor} \begin{Pf} Let $X$ be a regular surface, with ample $K_X$, lying on exactly $n+1$ irreducible components of the moduli; let $M$ be the union of the irreducible components of the moduli space of surfaces with ample canonical class which contain $[X]$. Let $r=h^0(X,mK_X)-1$; infinitely many such $X$'s (with distinct values of $r$) can be constructed by applying theorem \ref{mainthm1}, in view of remark \ref{suffample}. Fix an $m$-canonical embedding of $X$ in $\P^r$. Every small embedded deformation of $X$ in $\P^r$ is again a smooth surface, $m$-canonically embedded as $X$ is regular. Let $H$ be the union of the irreducible components of the Hilbert scheme of $\P^r$ containing $[X]$, and $H^0$ be the open dense subset of $H$ parametrizing smooth, $m$-canonically embedded surfaces. The natural map $H^0\to M$ is dominant, and each fibre is irreducible of dimension $(r+1)^2-1$; in fact, the fibre over $[X']$ is the set of bases of $H^0(X',mK_{X'})$ modulo the action of the finite group $Aut(X')$ (and modulo the obvious ${\bf C}^*$-action). In particular there is an induced bijection between irreducible components of $M$ and of $H$, which increases the dimension by $(r+1)^2-1$. \end{Pf} \begin{rem}\label{+gen} {\rm The constructions in this section generalize easily to the case where $Y$ is neither regular nor rigid and $G$ is any abelian group. In fact, they also work if the dimension of $Y$ is bigger than $2$ (using a suitable, modified form of lemma \ref{basic}).} \end{rem} \section{The Hilbert scheme of curves in $\P^r$} In this section we apply the results on the Hilbert scheme of surfaces to the Hilbert scheme of curves. We first introduce some notation. If $Z\subset \P^r$ is a subscheme, we will denote by $\hbox{\rm Hilb}(Z)$ the union of the irreducible components of the Hilbert scheme of $\P^r$ containing $[Z]$; we let $H(Z)$ be the germ of $\hbox{\rm Hilb}(Z)$ at $[Z]$. Note that if $F$ is a hypersurface of degree $l$, $\hbox{\rm Hilb}(F)$ is naturally isomorphic to $\P(H^0(\P^r,\O(l)))$. \begin{lem} \label{splitting} Let $X\subset \P^r=\P$ be a smooth surface, $F\subset \P$ be a smooth hypersurface of degree $l$ transversal to $X$, and let $C=X\cap F$. Then there is a natural isomorphism $N_{C/\P}|_C\cong N_{X/\P}|_C\oplus \O_C(l)$. \end{lem} \def\phantom{.}\\{\phantom{.}\\} \begin{Pf} We have a natural diagram: $$ \begin{array}{ccccccccc} &&&& 0 &&&&\\ &&&& \downarrow &&&&\\ \phantom{.}\\ &\phantom{\sum_I^J}&&& N_{C/F} &&&&\\ \phantom{.}\\ &\phantom{\sum_I^J}&&& \downarrow &\searrow&&&\\ \phantom{.}\\ 0&\longrightarrow &N_{C/X}&\longrightarrow&\phantom{\sum_I^J}N_{C/\P}\phantom{\sum_I^J}&\longrightarrow&N_{X/\P}|_{C}&\longrightarrow&0\\ \phantom{.}\\ &\phantom{\sum_I^J}&&\searrow& \downarrow &&&&\\ \phantom{.}\\ &\phantom{\sum_I^J}&&& N_{F/\P}|_{C} &&&&\\ \phantom{.}\\ &&&& \downarrow &&&&\\ &&&& 0 &&&& \end{array}$$ However $N_{F/\P}|_{C}=N_{C/X}=\O_C(l)$, and it is easy to check that the morphism $N_{C/X}\to N_{F/\P}|_{C}$ in the diagram is the identity. This implies that the natural map $N_{C/F}\to N_{X/\P}|_{C}$ is also an isomorphism, hence the claimed splitting. \end{Pf} \begin{prop}\label{smooth} Let $X\subset \P^r=\P$ be a smooth, regular, projectively normal surface. Let $F$ be a smooth hypersurface of degree $l$ in $\P$ cutting $X$ transversally along a curve $C$, and let $U\subset \hbox{\rm Hilb}(X)\times \hbox{\rm Hilb}(F)$ be the open set of pairs $({X'},{F'})$ such that ${X'}$ and ${F'}$ are smooth and transversal and ${X'}$ is projectively normal. If $l>>0$, then for every $({X'},{F'})\in U$ the map $H({X'})\times H({F'})\to H({C'})$ induced by intersection is smooth, where ${C'}={X'}\cap{F'}$. \end{prop} \begin{Pf} The germ of the Hilbert scheme $H(Z)$ represents the functor of embedded deformations of $Z$ in $\P$; when $Z$ is smooth, this functor has tangent (resp.~obstruction) space $H^0(Z,N_{Z/\P})$ (resp.\ $H^1(Z,N_{Z/\P})$). Let $({X'},{F'})\in U$, and ${C'}={X'}\cap {F'}$. The map $H({X'})\times H({F'})\to H({C'})$ induces natural maps on tangent and obstruction spaces; to prove the required smoothness it is enough to prove that the induced maps are surjective on tangent spaces and injective on obstruction spaces. Note that $H^i({C'},N_{{C'}/\P})= H^i({C'},N_{{X'}/\P}|_{{C'}})\oplus H^i({C'},N_{{C'}/\P}|_{{C'}})$ by lemma \ref{splitting}. Via this isomorphism, the maps we are interested in are induced by the long exact sequences associated to:$$ \begin{array}{ccccccccc} 0&\to &N_{{X'}/\P}\otimes{\cal I}_{{C'}\subset {X'}}&\longrightarrow&N_{{X'}/\P}& \longrightarrow N_{{X'}/\P}|_{{C'}}&\to&0\\ \phantom{1}\\ 0&\to &N_{{F'}/\P}\otimes{\cal I}_{{C'}\subset {F'}}&\longrightarrow&N_{{F'}/\P}& \longrightarrow N_{{F'}/\P}|_{{C'}}&\to&0. \end{array}$$ Therefore it is enough to prove that $$H^1({X'},N_{{X'}/\P}\otimes{\cal I}_{{C'}\subset {X'}})=0$$ and that $H^0({F'},N_{{F'}/\P})\to H^0(C,N_{{F'}/\P}|_{{C'}})$ is surjective (remark that $N_{{F'}/\P}=\O_{F'}(l)$, hence $H^1({F'},N_{{F'}/\P})=0$ by Kodaira vanishing). For the claimed surjectivity, note that there is a commutative diagram $$ \begin{array}{ccc} H^0(\P,\O_{\P}(l))&\longrightarrow &H^0({X'},\O_{{X'}}(l))\cr \downarrow&\phantom{{2 choose 3}}&\downarrow\cr H^0({F'},\O_{{F'}}(l))&\longrightarrow &H^0({C'},\O_{{C'}}(l)) \end{array} $$ As ${X'}$ is projectively normal, the upper horizontal arrow is onto, and as ${X'}$ is regular, the right vertical arrow is onto. Hence the lower horizontal arrow is also onto. To prove the vanishing, as ${\cal I}_{{C'}\subset {X'}}=\O_{X'}(-l)$, it is enough to prove that $H^1({X'},N_{{X'}/\P}(-l))=0$ if $l$ is sufficiently large. For any given ${X'}$, this follows immediately from the definition of ampleness; on the other hand it is easy to prove (by a standard semicontinuity argument) that in fact an $l_0$ can be found such that the claimed vanishing holds for all $l\ge l_0$ and for all ${X'}\in \hbox{\rm Hilb}(X)$. \end{Pf} \begin{prop}\label{inj} Let $X\subset \P^r=\P$ be a smooth surface and let $F$ be a smooth hypersurface of degree $l$ meeting $X$ transversally in a smooth curve $C$. Let $U\subset \hbox{\rm Hilb}(X) \times \hbox{\rm Hilb}(F)$ be the open set of pairs $({X'},{F'})$ such that ${X'} \cap {F'}$ is a smooth curve. If $l>>0$, then each fibre of the map $U\to \hbox{\rm Hilb}(C)$ given by $({X'},{F'}) \mapsto {X'}\cap {F'}$ is contained in a fibre of the projection $U\to\hbox{\rm Hilb}(X)$. In other words, each curve contained in the image of $U$ in $\hbox{\rm Hilb}(C)$ lies on exactly one surface in $\hbox{\rm Hilb}(X)$. \end{prop} \begin{Pf} Let ${\cal X}\to \hbox{\rm Hilb}(X)$ be the universal family. Inside the product ${\cal X} \times {\cal X}$, consider the diagonal subvariety ${\cal I}$ consisting of the pairs $(x,x)$. Let $W$ be the locus of $\hbox{\rm Hilb}(X)\times \hbox{\rm Hilb}(X)$ over which the map ${\cal I}\to \hbox{\rm Hilb}(X)\times \hbox{\rm Hilb}(X)$ has one-dimensional fibres. One may choose a stratification $\{W_j\}$ of $W$ such that each of the restricted families ${\cal I}_j\to W_j$ is flat. Thus, one has induced maps from $W_j$ to the Hilbert scheme of one-dimensional subschemes of $\P$. Since the union of the images of the $W_j$'s is contained in a finite number of components of the Hibert scheme, the degree of the curves contained in the intersection of two distinct surfaces of $\hbox{\rm Hilb}(X)$ is bounded by an integer $l_0$. Therefore it is enough to choose $l>l_0$. \end{Pf} \begin{thm}\label{mainthm2} Let $n\ge 3$ be an integer. Then there exist infinitely many integers $r$, and for each of them a smooth, irreducible curve $C_r\subset \P^r$ such that $C_r$ lies exactly on $n$ components of the Hilbert scheme of $\P^r$. \end{thm} \begin{Pf} By corollary \ref{HilbX}, for infinitely many values of $r$ one can construct a regular surface $X$ of general type, embedded in $\P^r$ by a complete $m$-canonical system, such that $X$ lies on exactly $n$ components of the Hilbert scheme of $\P^r$, having pairwise different dimensions. By \cite{And}, $X$ is projectively normal in $\P^r$ if $m>>0$: in fact, by the theorem on page 362 together with the fact that if $K_X$ is ample then $5K_X$ is very ample, it is enough to assume $m\ge 11$. Choose an integer $l$ with $l>>0$, such that propositions \ref{smooth} and \ref{inj} hold for $l$. Let $F$ be a smooth hypersurface of degree $l$ meeting $X$ transversally. Let $U\subset \hbox{\rm Hilb}(X)\times \hbox{\rm Hilb}(F)$ be the locus of pairs $({X'},{F'})$ where both are smooth and meeting transversally, and ${X'}$ is projectively normal. $U$ is the union of $n$ irreducible components of pairwise different dimensions, each of them being the inverse image of an irreducible component of $\hbox{\rm Hilb}(X)$. Let now $C=C_r$ be the intersection of $X$ and $F$, and consider the natural map $U\to \hbox{\rm Hilb}(C)$ given by $({X'},{F'})\mapsto {X'}\cap {F'}$. By proposition \ref{smooth} this morphism is dominant and smooth on its image $V$. By \ref{inj} there is an induced morphism $V\to \hbox{\rm Hilb}(X)$, which is also dominant and smooth on its image. Therefore there is a natural bijection between the irreducible components of $\hbox{\rm Hilb}(X)$ and those of $\hbox{\rm Hilb}(C)$. \end{Pf}
1995-09-15T05:57:21
9501
alg-geom/9501001
en
https://arxiv.org/abs/alg-geom/9501001
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9501001
Misha Verbitsky
Misha Verbitsky
Cohomology of compact hyperkaehler manifolds
87 pages LaTeX 2.09
null
null
null
null
Let M be a compact simply connected hyperk\"ahler (or holomorphically symplectic) manifold, \dim H^2(M)=n. Assume that M is not a product of hyperkaehler manifolds. We prove that the Lie algebra so(n-3,3) acts by automorphisms on the cohomology ring H^*(M). Under this action, the space H^2(M) is isomorphic to the fundamental representation of so(n-3,3). Let A^r be the subring of H^*(M) generated by H^2(M). We construct an action of the Lie algebra so(n-2,4) on the space A, which preserves A^r. The space A^r is an irreducible representation of so(n-2,4). This makes it possible to compute the ring A^r explicitely.
[ { "version": "v1", "created": "Tue, 3 Jan 1995 02:00:13 GMT" }, { "version": "v2", "created": "Sat, 13 May 1995 21:49:21 GMT" } ]
2008-02-03T00:00:00
[ [ "Verbitsky", "Misha", "" ] ]
alg-geom
\section{Introduction.}\label{_introduction_to_so(b_2,...)_Section_} Here we give a brief introduction to results of this paper. The subsequent sections are independent from the introduction. The main object of this paper is theory of compact hyperk\"ahler manifolds. A hyperk\"ahler manifold (see \ref{_hyperkaehler_manifold_Definition_} for more precise wording) is a Riemannian manifold $M$ equipped with three complex structures $I$, $J$ and $K$, such that $I\circ J=-J\circ I=K$ and $M$ is K\"ahler with respect to the complex structures $I$, $J$ and $K$. Let $M$ be a complex manifold which admits a hyperk\"ahler structure. A simple linear-algebraic argument implies that $M$ is equipped with a holomorphic symplectic form. Calabi-Yau theorem shows that, conversely, every compact holomorphically symplectic K\"ahler manifold admits a hyperk\"ahler structure, which is uniquely defined by these data. Further on, we do not discriminate between compact holomorphic symplectic manifolds of K\"ahler type and compact hyperk\"ahler manifolds. Let $(M,I)$ be a compact K\"ahler manifold which admits a holomorphic symplectic form $\Omega$. For simplicity of statements, we assume in this introduction that $\Omega$ is unique up to a constant. Denote by $M$ the $C^\infty$-manifold underlying $(M,I)$. Let $(M,I')$ be another compact holomorphically symplectic manifold which lies in the same deformation class as $(M,I)$. Fixing a diffeomorphism of underlying $C^\infty$-manifolds, we may identify the smooth manifold which underlies $(M,I)$ with that underlying $(M,I')$. Let $X$ be an arbitrary compact K\"ahler manifold, $\dim_{\Bbb C} X=n$. We associate with a K\"ahler structure on $X$ so-called Riemann-Hodge pairing \[ (\cdot,\cdot): \;\; H^2(X,{\Bbb R})\times H^2(X,{\Bbb R})\longrightarrow {\Bbb R} \] which is a map associating a number \begin{equation}\label{_Hodge-Riemann-correct-Equation_} \int_X \omega^{n-2}\wedge \eta_1\wedge\eta_2 - \frac{n-2}{(n-1)^2}\cdot \frac{\int_X \omega^{n-1}\eta_1 \cdot \int_X\omega^{n-1}\eta_2} {\int_X \omega^n} \end{equation} to classes $\eta_1,\eta_2 \in H^2(X,{\Bbb R})$, where $\omega\in H^2(X,{\Bbb R})$ is the K\"ahler class (see also \ref{_Hodge_Riema_general_Claim_}). Consider the positively defined scalar product $(\cdot,\cdot)_{Metr}$ induced by Riemannian metric on the space of harmonic 2-forms, identified by Hodge with $H^2(X,{\Bbb R})$. The pairing \eqref{_Hodge-Riemann-correct-Equation_} is defined in such a way that on primitive $(1,1)$-forms with coefficients in ${\Bbb R}$, $(\cdot,\cdot)$ is equal to $-(\cdot,\cdot)_{Metr}$. Similarly, on the space \[ \bigg({\Bbb R}\omega \oplus H^{2,0}(X)\oplus H^{0,2}(X)\bigg) \cap H^2(X,{\Bbb R}), \] the form $(\cdot,\cdot)$ is equal to $(\cdot,\cdot)_{Metr}$. \hfill Let $(\cdot,\cdot)$, $(\cdot,\cdot)': \; H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})\longrightarrow {\Bbb R}$ be the Hodge-Riemann forms associated with $(M,I)$ and $(M,I')$ respectively. The most surprising result of this paper is following (see \ref{_Hodge_Riemann_independent_Theorem_} for a different wording of the same statement): \hfill \theorem \label{_Riemann_Hodge_unique_in_intro_Theorem_} The forms $(\cdot,\cdot)$ and $(\cdot,\cdot)'$ are {\bf proportional}. \hfill Taking a K\"ahler class $\omega$ such that $Vol(M)= 1$, where $Vol(M):= \int_M \omega^{n}$, we get rid of the ambiguity in the choice of a constant. If $(M,I)$ and $(M,I')$ both satisfy $Vol(M)= 1$, then the Hodge-Riemann forms $(\cdot,\cdot)$ and $(\cdot,\cdot)'$ are {\bf equal}. We call this form {\bf the normalized Hodge-Riemann pairing}, denoted as $(\cdot,\cdot)_{\c H}$. This form is an invariant of a deformation class of complex manifolds. One may regard $(\cdot,\cdot)_{\c H}$ as {\bf topological} invariant. \footnote{In fact, the form $(\cdot,\cdot):\;H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})\longrightarrow {\Bbb R}$ of \eqref{_Hodge-Riemann-correct-Equation_} is a {\it topological} invariant associated with every cohomology class $\omega\in H^2(M,{\Bbb R})$. Our proof immediately implies that for a dense set of $\omega,\omega'\in H^2(M,{\Bbb R})$, \ref{_Riemann_Hodge_unique_in_intro_Theorem_} holds.} We use the form $(\cdot,\cdot)_{\c H}$ to compute the cohomology algebra $H^*(M)$. We explicitely compute the subalgebra of $H^*(M)$ generated by $H^2(M)$. These computations also give the relations between elements generated by $H^2(M)$ and the rest of $H^*(M)$. The main ideas of this computation are due to the observations which involve Torelli theorem and deformation spaces. As follows from results of Bogomolov (\cite{_Bogomolov_}, \cite{_Besse:Einst_Manifo_}, \cite{_Todorov_}), the deformation space of a holomorphically symplectic manifold is a smooth complex manifold, which is a quotient of a Stein space by an arithmetic group. From this description we use only the calculation of dimension of this moduli space. The map of Kodaira and Spencer \[ KS:\; T_{[M]}\c M\longrightarrow H^1(TM) \] is a linear homomorphism from the Zariski tangent space $T_{[M]}\c M$ of a moduli space $\c M$ of deformations of $M$ to the first cohomology of the sheaf of holomorphic vector fields on $M$. It is proven by Kodaira and Spencer \cite{_Kodaira_Spencer_} that this map is an embedding. Results of Bogomolov et al imply that $\c M$ is smooth and $KS$ is an isomorphism. This is often called ``Torelli theorem''. This statement could be translated to the language of period maps. Together with \ref{_Riemann_Hodge_unique_in_intro_Theorem_}, this observation implies a nice description of the period map. By ``period map'' associated with the holomorphic symplectic manifolds (see Section \ref{_perio_and_forge_Section_} for details) we mean the following geometrical object. We define the (coarse, marked) moduli of holomorphic symplectic manifolds as the space of different complex holomorphic symplectic structures on a given $C^\infty$-manifold $M$ up to diffeomorphisms which act trivially on $H^*(M)$. Let $Comp$ be a connected component of this moduli space. For more accurate definition of $Comp$, we refer the reader to \ref{_Comp_Hyp_Definition_}. For all points $I\in Comp$, we denote the corresponding complex manifold by $(M,I)$. {}From the definition of $Comp$ we obtain a canonical identification of cohomology spaces $H^*(M,I)$ for all $I\in Comp$. For simplicity of statements, we deal with the simply connected holomorphically symplectic manifolds $M$ such that $H^{2,0}(M)={\Bbb C}$ (such manifold are called {\bf simple}). The period map (Section \ref{_perio_and_forge_Section_}) in this context is a map $P_c:\; Comp \longrightarrow \Bbb P(H^2(M,{\Bbb C}))$ which relies a line $H^{2,0}(M,I)\subset H^2(M,{\Bbb C})$ to every $I\in Comp$. Classical results of Bogomolov et al imply that $P_c$ is an immersion. Hodge-Riemann relations together with \ref{_Riemann_Hodge_unique_in_intro_Theorem_} let one to describe the image of $P_c$ in following terms. An immediate consequence of Hodge-Riemann relations is that for all lines $x\in Im(P_c)\subset \Bbb P(H^2(M,{\Bbb C}))$, and all vectors $l\in x\subset H^2(M,{\Bbb C})$, we have $(l,l)_{\c H} =0$. Therefore, $P_c$ maps $Comp$ to a quadric $C\subset \Bbb P(H^2(M,{\Bbb C}))$, which is defined by the quadratic form associated with the Riemann-Hodge pairing. Dimension of $Comp$ is computed from Torelli theorem. As one can easily check, it is equal to the dimension of $C$. Since $P_c$ is an immersion, we obtain the following theorem: \hfill \theorem The period map $P_c:\; Comp \longrightarrow C$ is etale. \hfill This is the main observation used in computations of the cohomology algebra $H^*(M)$. \hfill Let $I\in Comp$. Let $ad I\in End(H^*(M))$ be a linear endomorphism which maps a $(p,q)$-form $\eta\in H^{p,q}(M)$ to $(p-q)\sqrt{-1}\: \eta$. Let $\goth M\subset End(H^*(M))$ be a Lie algebra generated by the endomorphisms $ad I$ for all $I\in Comp$. Hodge-Riemann relations imply that the action of $\goth M$ on $V= H^2(M)$ preserves the scalar product $(\cdot,\cdot)_{\c H}$. This defines a Lie algebra homomorphism \begin{equation} \label{_from_Mum-Tate_to_SO(V)_homom_Equation_} \rho:\; \goth M\longrightarrow \goth{so}(V). \end{equation} Using period maps and estimation on dimensions of moduli spaces, we prove that $\rho$ is an {\bf isomorphism} (\ref{_g_0_computed_Theorem_}, \ref{_g_0_is_Mumf_Tate_Theorem_}). This statement is an ingredient in the computation of the algebra $H^*(M)$. The algebraic structure on $H^*(M)$ is studied using the general theory of Lefschetz-Frobenius algebras, introduced in \cite{_Lunts-Loo_}. Lefschetz-Frobenius algebras are associative graded commutative algebras whose properties approximate that of cohomology of compact manifolds which admit K\"ahler structure. We give an exact definition of this term in Section \ref{_Lefshe_Frob_Section_} (\ref{_Lefschetz_Frob_alge_Definition_}). With no loss of generality, reader may think of these algebras as of cohomology algebras. With the Lefschetz-Frobenius algebra $A$ we associate so-called {\bf structure Lie algebra} ${\goth g}(A)\subset End(A)$ which acts on $A$ by linear endomorphism. The action of structure Lie algebra is often sufficient to reconstruct multiplication on $A$. This algebra is defined using the algebraic version of the strong Lefschetz theorem. Let $A= \bigoplus\limits^{2d}_{i=0} A_i$ be a graded commutative associative algebra over a field of characteristic zero. Let $H\in End(A)$ be a linear endomorphism of $A$ such that for all $\eta \in A_i$, $H(\eta)= (i-d) \eta$. This endomorphism is usually considered in Hodge theory. For all $a\in A_2$, denote by $L_a:\;\; A\longrightarrow A$ the linear map which associates with $x\in A$ the element $ax\in A$. Again, this operator is a counterpart of the operator $L$ considered in Hodge theory. The triple $(L_a, H, \Lambda_a)\in End(A)$ is called {\bf a Lefschetz triple} if \[ [ L_a, \Lambda_a] = H,\;\; [ H, L_a ] = 2 L_a, \;\; [ H, \Lambda_a] = -2 \Lambda_a. \] A Lefschetz triple establishes a representation of the Lie algebra $\goth{sl}(2)$ in the space $A$. For cohomology algebras, this representation arises as a part of Lefschetz theory. V. Lunts noticed that in a Lefschetz triple, the endomorphism $\Lambda_a$ is uniquely defined by the element $a\in A_2$ (\ref{_Lefshe_tri_unique_Proposition_}). For arbitrary $a\in A_2$, $a$ is called {\bf of Lefschetz type} if the Lefschetz triple $(L_a, H, \Lambda_a)$ exists. If $A= H^*(X)$ where $X$ is a compact complex manifold of K\"ahler type, then all K\"ahler classes $\omega\in H^2(M)$ are elements of Lefschetz type. On the other hand, not all elements of Lefschetz type are K\"ahler classes. For instance, if $\omega$ is of Lefschetz type, then $-\omega$ is also of Lefschetz type, but $\omega$ and $-\omega$ cannot be K\"ahler classes simultaneously. As one can easily check (see \cite{_Lunts-Loo_}), the set $S\subset A_2$ of all elements of Lefschetz type is Zariski open in $A_2$. Now we can define the structure Lie algebra ${\goth g}(A)$ of $A$. By definition, ${\goth g}(A)\subset End(A)$ is a Lie subalgebra of $End(A)$ generated by $L_a$, $\Lambda_a$, for all elements of Lefschetz type $a\in S$. This Lie algebra is often sufficient to reconstruct the multiplicative structure on $A$. Returning to the hyperk\"ahler manifolds, we consider the structure Lie algebra ${\goth g}(A)$ of the ring $A= H^*(M)$, where $M$ is a compact hyperk\"ahler manifold. It turns out that the structure Lie algebra ${\goth g}(A)$ can be explicitely computed. In particular, \ref{_g(A)_for_hyperkae_Theorem_} gives us the following theorem: \hfill \theorem Let $M$ be a compact holomorphically symplectic manifold. Assume for simplicity\footnote{The structure Lie algebra can be computed without this trivial assumption, but the statement is less starightforward. See the discussion after \ref{_simple_hyperkaehler_mfolds_Definition_} for details.} that $M$ admits a unique up to a constant holomorphic symplectic form. Let $n= \dim(H^2(M))$. Then ${\goth g}(A)= \goth{so}(4, n-2)$. \hfill This computation, which also involves the computation of the Lie algebra $\goth M$ of \eqref{_from_Mum-Tate_to_SO(V)_homom_Equation_}, takes up 4 sections of this paper. However, the main idea of this computation is easy. \hfill Let $M$ be a compact hyperk\"ahler manifold with the complex structures $I$, $J$, $K$. Consider the K\"ahler forms $\omega_I$, $\omega_J$, $\omega_K$ associated with these complex structures. Let $\rho_I:\; \goth{sl}(2)\longrightarrow End(H^*(M))$, $\rho_J:\; \goth{sl}(2)\longrightarrow End(H^*(M))$, $\rho_K:\; \goth{sl}(2)\longrightarrow End(H^*(M))$ be the corresponding Lefschetz homomorphisms. Let $\goth a\subset End(H^*(M))$ be the minimal Lie subalgebra which contains images of $\rho_I$, $\rho_J$, $\rho_K$. The algebra $\goth a$ was computed explicitely in \cite{_so5_on_cohomo_}. \hfill \theorem \label{_so_5_Theorem_} (\cite{_so5_on_cohomo_}) The Lie algebra $\goth a$ is isomorphic to $\goth{so}(4,1)$. \hfill This statement can be regarded as a ``hyperk\"ahler Lefschetz theorem''. Indeed, its proof parallels the proof of Lefschetz theorem. One can check that the cohomology classes $\omega_I$, $\omega_J$, $\omega_K\in H^2(M,{\Bbb R})$ are orthogonal with respect to Riemann-Hodge. Let $Hyp$ be the classifying space of the hyperk\"ahler structures on $M$ (see Section \ref{_moduli_Section_}). Let $P_{hyp}:Hyp\longrightarrow H^2(M)\times H^2(M)\times H^2(M)$ be the map which associates with the hyperk\"ahler structure $\c H= (I, J, K, (\cdot,\cdot))$ the triple $(\omega_I,\omega_J,\omega_K)$. Then the image of $P_{hyp}$ in $H^2(M)\times H^2(M)\times H^2(M)$ satisfies \begin{equation}\label{_image_of_P_hyp_Equation_} \forall (x,y,z)\in im P_{hyp}\;\; \bigg |\;\; \begin{array}{l} (x,y)_{\c H}=(x,z)_{\c H}=(y,z)_{\c H}=0,\\[3mm] (x,x)_{\c H}=(y,y)_{\c H}=(z,z)_{\c H}, \end{array} \end{equation} where $(\cdot,\cdot)_{\c H}$ is the Hodge-Riemann pairing of \eqref{_Hodge-Riemann-correct-Equation_}. Let $D\subset H^2(M)\times H^2(M)\times H^2(M)$ be the set defined by the equations \eqref{_image_of_P_hyp_Equation_}. Using Torelli theorem and Calabi-Yau, we prove the following statement: \hfill \theorem\label{_image_of_P_hyp_Theorem_} The image of $P_{hyp}$ is Zariski dense in $D$. \hfill \ref{_image_of_P_hyp_Theorem_} shows that all algebraic relations which are true for \[ (x,y,z)\in P_{hyp}(Hyp) \] are true for all $(x,y,z)\in D$. Computing the Lie algebra $\goth a$ as in \ref{_so_5_Theorem_}, we obtain a number of relations between $x,y,z\in H^2(M)$ which hold for all $(x,y,z)\in Im(P_{hyp})$. Using the density argument, we obtain that these relations are universally true. This idea leads to the following theorem. \hfill \theorem \label{_structure_alge_for_coho_hyperkahe_Theorem_} Let $A= H^*(M)$ be a cohomology algebra of a compact simply connected holomorphically symplectic manifold $M$ with $H^{2,0}(M)\cong {\Bbb C}$. Let $n = \dim H^2(M)$. Then the structure Lie algebra ${\goth g}(A)$ of $A$ is isomorphic to $\goth{so}(4,n-2)$. \hfill It remains to recover the multiplication on $H^*(M)$ from the structure Lie algebra. This is done as follows. Let $A = \oplus A_i$ be a Lefschetz-Frobenius algebra, ${\goth g} = {\goth g}(A)$ be its structure Lie algebra. Clearly, ${\goth g}$ is graded: ${\goth g}=\bigoplus\limits_i{\goth g}_{2i}$, ${\goth g}_{2i}(A_j)\subset A_{j+2i}$. Let $k$ denote the one-dimensional commutative Lie algebra. In the case of \ref{_structure_alge_for_coho_hyperkahe_Theorem_}, ${\goth g}\cong \goth{so}(4,n-2)$, ${\goth g}_0 \cong \goth{so}(3,n-3)\oplus k$, $\dim {\goth g}_2=\dim{\goth g}_{-2}=n$ and $\dim {\goth g}_{2i}=0$ for $|i|>1$ (see \ref{_calculation_of_g(A)_for_minim_Theorem_}). We say that the Lefschetz-Frobenius algebra $A$ is {\bf of Jordan type} if ${\goth g}_{2i}(A)=0$ for $|i|>1$. For such algebras, the subspaces ${\goth g}_2(A)$, ${\goth g}_{-2}(A)\subset A$ are commutative Lie subalgebras of ${\goth g}(A)$. Let $U_{\goth g}$ be the universal enveloping algebra of ${\goth g}= {\goth g}(A)$, and $U_{{\goth g}_{2}}\subset U_{\goth g}$ be the enveloping algebra of the subalgebra ${\goth g}_2={\goth g}_2(A)\subset {\goth g}$. Consider the space $A$ as $A$-module. Then, for all $v\in A$ we have a map $t_v:\; U_{{\goth g}}\longrightarrow A$ which associates $P(v)$ with $P\in U_{{\goth g}}$. The Lie algebra ${\goth g}_2$ is commutative, and therefore $U_{{\goth g}_2}\cong S^*({\goth g}_2)$. According to \cite{_Lunts-Loo_}, the natural map $A_2\longrightarrow {\goth g}_2$, $a \longrightarrow L_a$, is an isomorphism (see \ref{_g_2_is_A_2_Corollary_} for details). Let $v\in H^0(M)\subset A$ be a unit element of the ring $A$. Consider the restriction $t$ of $t_v:\; U_{\goth g}\longrightarrow A$ to $U_{{\goth g}_2}\subset U_{\goth g}$. Then $t$ is a map from $S^* {\goth g}_2\cong S^* A_2$ to $A$. Clearly, this map coinsides with the map $S^* A_2\longrightarrow A$ defined by the multiplication. This implies that multiplication by elements from $H^2(M)$ can be recovered from the action of the structure Lie algebra ${\goth g}$. Using the calculations of \ref{_structure_alge_for_coho_hyperkahe_Theorem_}, we obtain, in particular, the following theorem (see Section \ref{_cohomolo_compu_Section_}): \hfill \theorem \label{_S^*H^2_is_H^*M_intro-Theorem_} Let $M$ be a compact hyperk\"ahler manifold. Let \[ \bar H^* (M)\subset H^*(M) \] be the subalgebra in $H^*(M)$ generated by $H^2(M)$. Let $\dim_{\Bbb C} M=2n$. Then \[ \bar H^{2i}(M)\cong S^i H^2(M) \] for $i\leq n$, and \[ \bar H^{2i}(M)\cong S^{2n-i} H^2(M) \] for $i\geq n$. \hfill \hfill \centerline{\Large \bf Contents:} \hfill \begin{description} \item [Section \ref{_introduction_to_so(b_2,...)_Section_}:]\hspace{2mm} {Introduction.} \item [Section \ref{hyperk_manif_Section_}:]\hspace{2mm} {Hyperk\"ahler manifolds.} \item [Section \ref{_moduli_Section_}:]\hspace{2mm} {Moduli spaces for hyperk\"ahler and holomorphically symplectic manifolds.} \item [Section \ref{_perio_and_forge_Section_}:]\hspace{2mm} {Periods and forgetful maps.} \item [Section \ref{_Period_and_Hodge_Riemann_Section_}:]\hspace{2mm} {Hodge-Riemann relations for the hyperk\"ahler manifolds and period map.} \item [Section \ref{_Hodge-Rie_independent_Section_}:]\hspace{2mm} {The Hodge-Riemann metric on $H^2(M)$ does not depend on complex structure.} \item [Section \ref{_Q_c_defini_Section_}:]\hspace{2mm} {Period map and the space of 2-dimensional planes in \\ $H^2(M,{\Bbb R})$.} \item [Section \ref{_Lefshe_Frob_Section_}:]\hspace{2mm} {Lefschetz-Frobenius algebras.} \item [Section \ref{_minimal_Fro_Section_}:]\hspace{2mm} {The minimal Frobenius algebras and cohomology of compact K\"ahler surfaces.} \item [Section \ref{_^dA(V)_Section_}:]\hspace{2mm} {Representations of $SO(V,+)$ leading to Frobenius algebras.} \item [Section \ref{_computing_g_for_hyperk_pt-I_Section_}:]\hspace{2mm} {Computing the structure Lie algebra for the cohomology of a hyperk\"ahler manifold, part I.} \item [Section \ref{_compu_g_0_part_1_Section_}:]\hspace{2mm} {Calculation of a zero graded part of the structure Lie algebra of the cohomology of a hyperk\"ahler manifold, part I.} \item [Section \ref{_compu_g_0_part_2_Section_}:]\hspace{2mm} {Calculation of a zero graded part of the structure Lie algebra of the cohomology of a hyperk\"ahler manifold, part II.} \item [Section \ref{_computing_g_for_hyperk_pt-2_Section_}:]\hspace{2mm} {Computing the structure Lie algebra for the cohomology of a hyperk\"ahler manifold, part II.} \item [Section \ref{_cohomolo_compu_Section_}:]\hspace{2mm} {The structure of the cohomology ring for compact hyperkaehler manifolds.} \item [Section \ref{_calcu_dimensi_Section_}:]\hspace{2mm} {Calculations of dimensions.} \end{description} \hfill \hfill \begin{itemize} \item Section \ref{_introduction_to_so(b_2,...)_Section_} tries to supply motivations and heuristics for the further study. In the body of this article, we never refer to Section \ref{_introduction_to_so(b_2,...)_Section_}. Reading Section \ref{_introduction_to_so(b_2,...)_Section_} is not necessarily for understanding of this paper. \item In Section \ref{hyperk_manif_Section_}, we give the definition of a hyperk\"ahler manifold. We explain the geopetry of quaternionic action on $H^*(M)$. This section ends with the statement of Calabi-Yau for compact holomorphically symplectic manifolds. Results and definitions of this section are well known. \item Section \ref{_moduli_Section_} begins with a definition of simple hyperk\"ahler manifolds. This notion stems from the theory of holonomy groups. Let $M$ be a compact hyperk\"ahler manifold, $\dim_{\Bbb R} M=4n$. Then the holonomy group of $M$ is naturally a subgroup of $Sp(n)$. The simple hyperk\"ahler manifold is a hyperk\"ahler manifold such that its restricted holonomy group in exactly $Sp(n)$ and not a proper subgroup of $Sp(n)$. In \ref{_simple_hyperkaehler_mfolds_Definition_} we give a different, but equivalent treatment of this notion (see also \cite{_Beauville_}).Using a formalism by de Rham and Berger, Bogomolov and Beauville proved that a hyperk\"ahler manifold $M$ is simple if and only if there is no finite covering $\tilde M$ of $M$ such that $\tilde M$ can be represented as a product of two (or more) non-trivial hyperk\"ahler manifolds. Therefore, it is usually harmless to assume that a given compact hyperk\"ahler manifold is simple. For simple hyperk\"ahler manifolds, $\dim H^{2,0}(M)=1$. \item After we define simple hyperk\"ahler manifolds, we go for the marked moduli spaces. We define moduli spaces for complex, hyperk\"ahler and holomorphically symplectic structures. We do this in topological, rather than in algebro-geometrical setting. The most obvious reason for this lopsided treatment is that moduli of hyperk\"ahler structures don't have any structures in addition to topology. Results and definitions of this section are well known. \item In Section \ref{_perio_and_forge_Section_}, we define several kinds of period maps. Let $Comp$ be the marked moduli of complex structures on a simple holomorphically symplectic manifold. For all complex structures $I\in Comp$, the space $H^{2,0}(M,I)$ of $(2,0)$-forms is one-dimensional. With every complex structure $I\in Comp$, the period maps $P_c:\; Comp \longrightarrow {\Bbb P}(H^2(M, {\Bbb C}))$ associates a line $H^{2,0}(M,I)\subset H^2(M)$ in ${\Bbb P}(H^2(M, {\Bbb C}))$. With every holomorphic symplectic structure, period map associates a class in $H^2(M,{\Bbb C})$ represented by a holomorphic symplectic form. Finally, with every hyperk\"ahler structure, period map $P_{hyp}$ associates a triple of K\"ahler classes which correspond to the complex structures $I$, $J$, $K$. We define a number of forgetful maps (from hyperk\"ahler moduli to complex moduli etc.) and compare these maps against period mappings. \item Sections \ref{hyperk_manif_Section_} - \ref{_perio_and_forge_Section_} contain no new results. We establish setting for the further study of hyperk\"ahler manifolds. \item Section \ref{_Period_and_Hodge_Riemann_Section_} answers the following query. Hodge-Riemann pairing on cohomology satisfies a certain type of positivity conditions, called {\bf Hodge-Riemann relations}. What happens with these relations on a hyperk\"ahler manifold? It turns that for a Hodge-Riemann form $(\cdot,\cdot)$ defined on $H^2(M)$ and an action of $SU(2)$ on $H^2(M)$ which comes from quaternions, the following conditions are satisfied: \begin{description} \item [{\rm (i)}] Let $H^{inv}\subset H^2(M)$ be the space of $SU(2)$-invariants, and $H^\bot$ be its orthogonal complement. Then the Riemann-Hodge pairing is negatively defined on $H^{inv}$ and positively defined on $H^\bot$. \item [{\rm (ii)}] The space $H^\bot$ is three-dimensional. \item [{\rm (iii)}] The space $H^\bot$ is generated by the K\"ahler forms associated with the complex structures $I$, $J$ and $K$. \end{description} \item Section \ref{_Hodge-Rie_independent_Section_} is the crux of the first part of this paper. We prove that the Hodge-Riemann form \eqref{_Hodge-Riemann-correct-Equation_} on $H^2(M)$ is independent (up to a constant) from the K\"ahler structure. The idea of the proof is the following. \item The group $SO(3)$ acts on the space of the hyperk\"ahler structures, replacing the triple $(I, J, K)$ by another orthogonal triple of imaginary quaternions. As Section \ref{_Period_and_Hodge_Riemann_Section_} shows, this action does not change the Hodge-Riemann form. Let $\c H \in Hyp$, and let $(\omega_I, \omega_J,\omega_K)\in H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})$ be the image of $\c H$ under the action of period map $P_{hyp}$. By definition, the Hodge-Riemann form depends only on $\omega_I$. Therefore, we may replace $\c H$ by $\c H'\in Hyp$ such that $P_{hyp}(\c H')= (\omega_I, \omega_J',\omega_K')$, and such replacement does not change the Hodge-Riemann pairing. We show that by iterating such replacements and action of $SO(3)$, we can connect any two hyperk\"ahler structure $\c H$ and $\c H'$ satisfying $Vol_{\c H} (M)= Vol_{\c H'}(M)$, where by $Vol_{\c H}(M)$ we undertstand the volume of $M$ computed with respect to the Riemannian structure associated with $\c H$. Since these operations don't change the Riemann-Hodge form, this form is equal for all hyperk\"ahler structure $\c H$ of given volume. Section \ref{_Hodge-Rie_independent_Section_} depends on Sections \ref{hyperk_manif_Section_} - \ref{_Period_and_Hodge_Riemann_Section_}. \item Section \ref{_Q_c_defini_Section_} gives a description of the period map \[ P_c:\; Comp\longrightarrow \Bbb B(H^2(M, {\Bbb C}))\] in terms of the manifold $Pl$ of 2-dimensional planes in $H^2(M, {\Bbb R})$. It turns out that there exist an etale mapping $Q_c:\; Comp\longrightarrow Pl$. This is a standard material, covered also in \cite{_Todorov_}. Our exposition adds a twist to \cite{_Todorov_}, because we use the normalized Hodge-Riemann form, which was unknown before. Otherwise, this section depends only on Sections \ref{hyperk_manif_Section_} - \ref{_perio_and_forge_Section_}. \item Sections \ref{_Lefshe_Frob_Section_} - \ref{_^dA(V)_Section_} are completely independed on the preceding sections. \item In Section \ref{_Lefshe_Frob_Section_}, we give a number of algebraic definitions. We give an exposition of the theory of Lefschetz-Frobenius algebras, following \cite{_Lunts-Loo_}. The aim of this section is a purely algebraic version of strong Lefschetz theorem. \item Roughly speaking, Frobenius algebra is a graded algebra for which an algebraic version of Poincare duality holds. A typical example of such algebra is an algebra of cohomology of a compact manifold. Similarly, the Lefschetz-Frobenius algebra is a Frobenius algebra for which the strong Lefschetz theorem holds - typically, an algebra of cohomology of a K\"ahler manifold. In Section \ref{_Lefshe_Frob_Section_} we explain these notions and define a structure Lie algebra ${\goth g}$ of a Lefschetz-Frobenius algebra. Here we also give a definition of Lefschetz-Frobenius algebras of Jordan type. \item Section \ref{_minimal_Fro_Section_} is dedicated to explicit examples of Lefschetz-Frobenius algebras, called {\bf minimal Lefhsetz-Frobenius algebras}. By definition, a minimal Lefschetz-Frobenius algebra is a Lefschetz-Frobenius algebra $A= A_0 \oplus A_2\oplus A_4$. The ``Poincare'' form on $A$ defines a bilinear symmetric pairing on $A_2$. It turns out that this pairing uniquely determines $A$. Conversely, with every linear space equipped with non-degenerate symmetric sclalar product we associate a minimal Lefschetz-Frobenius algebra. We prove that every minimal Lefschetz-Frobenius algebra is of Jordan type, and explicitely compute its structure Lie algebra. For a minimal Lefschetz-Frobenius algebra $A(V)$ associated with a space $V$, we denote the corresponding structure Lie algebra by $\goth{so}(V,+)$. If $V$ is a linear space over ${\Bbb R}$ equipped with a scalar product of signature $(p,q)$, then $\goth{so}(V,+)\cong \goth{so}(p+1,q+1)$. \item In Section \ref{_^dA(V)_Section_} we find all reduced Lefschetz-Frobenius algebras $A= A_0\oplus A_2 \oplus ... \oplus A_n$ with the structure Lie algebra $\goth{so}(V,+)$, where $\dim V\geq 3$. By ``reduced'' we understand the Lefschetz-Frobenius algebras generated by $A_2$. It turns out that for $n$ even, such algebra is unique (we denote it by ${}^{\frac{n}{2}}A(V)$), and for $n$ odd, there is no such algebras. \item Sections \ref{_Lefshe_Frob_Section_} - \ref{_^dA(V)_Section_} are purely algebraic, and Sections \ref{hyperk_manif_Section_} - \ref{_Q_c_defini_Section_} are dealing with geometry. These sections are mutually independent, and Sections \ref{_computing_g_for_hyperk_pt-I_Section_} - \ref{_computing_g_for_hyperk_pt-2_Section_} draw heavily on both parts, geometrical and algebraic. In these sections, we compute the structure Lie algebra of an algebra of cohomology of a simple hyperk\"ahler manifold. The basic result is that this algebra is isomorphic to $\goth{so}(V,+)$, where $V$ is the linear space $H^2(M, {\Bbb R})$ equipped with the normalized Hodge-Riemann pairing. \item In Section \ref{_computing_g_for_hyperk_pt-I_Section_}, we prove that the Lefschetz-Frobenius algebra $A=H^*(M)$ is of Jordan type: ${\goth g}(A)={\goth g}_{-2}(A)\oplus {\goth g}_0(A)\oplus {\goth g}_2(A)$. To prove this we introduce the standard ``density and periods'' argument, which is also used in Sections \ref{_compu_g_0_part_2_Section_} - \ref{_computing_g_for_hyperk_pt-2_Section_}. \item In Section \ref{_compu_g_0_part_1_Section_}, we construct a map ${\goth g}_0(A)\longrightarrow \goth{so}(V)\oplus k$, where $k$ is a one-dimensional commutative Lie algebra. We prove that this map is an isomorphism. We also prove that the Lie subalgebra $\goth M\subset End(V)$ generated by $ad I$ for all $I\in Comp$ is isomorphic to $\goth{so}(V)$. By definition, $ad I:\; H^i(M)\longrightarrow H^i(M)$ is an endomorphism which maps $\eta\in H^{p,q}_I(M)$ to $(p-q)\sqrt{-1}\: \eta$. We use the computations related to the $\goth{so}(5)$-action on $H^*(M)$ (see \cite{_so5_on_cohomo_}). \item In Section \ref{_compu_g_0_part_2_Section_}, we prove that the map $u:\; {\goth g}_0\longrightarrow \goth{so}(V)\oplus k$, constructed in Section \ref{_compu_g_0_part_1_Section_}, is an isomorphism. The proof is computational. \item In Section \ref{_computing_g_for_hyperk_pt-2_Section_}, we use the results of Sections \ref{_computing_g_for_hyperk_pt-I_Section_} - \ref{_compu_g_0_part_2_Section_} to finish the computation of the structure Lie algebra ${\goth g}(A)$. This is done by writing down a linear isomorphism ${\goth g}(A)\longrightarrow \goth{so}(4,n-2)$ explicitely. By computations, we check that this isomorphism is in fact an isomorphism of Lie algebras. \item Section \ref{_cohomolo_compu_Section_} is, again, algebraic. It depends only on Sections \ref{_Lefshe_Frob_Section_} - \ref{_^dA(V)_Section_}. In this section, we explicitely compute the graded commutative algebra ${}^dA(V)$ of Section \ref{_^dA(V)_Section_}. This computation has the following geometrical interpretation. Let $A=H^*(M)$ be the algebra of cohomology of a simple hyperk\"ahler manifold $M$, $\dim_{\Bbb R} M=4d$, and $A^r\subset A$ be its subalgebra generated by $V= H^2(M)$. The main theorem of \ref{_^dA(V)_Section_}, together with an isomorphism ${\goth g}(A)\cong\goth{so}(V,+)$ immediately imply that $A^r\cong {}^dA(V)$. Therefore, by computing ${}^dA(V)$, we compute a big part of the cohomology algebra of $M$. This way, we obtain a proof of \ref{_S^*H^2_is_H^*M_intro-Theorem_}. The computation of ${}^dA(V)$ is based on the classical theory of representations of $\goth{so}(V)$ and their tensor invariants (\cite{_Weyl_}). \item The final section applies the result of \ref{_S^*H^2_is_H^*M_intro-Theorem_} to obtain numerical lower bounds on Betti and Hodge numbers of a hyperk\"ahler manifold. \end{itemize} \section{Hyperk\"ahler manifolds.}\label{hyperk_manif_Section_} \definition \label{_hyperkaehler_manifold_Definition_} (\cite{_Beauville_}, \cite{_Besse:Einst_Manifo_}) A {\bf hyperk\"ahler manifold} is a Riemannian manifold $M$ endowed with three complex structures $I$, $J$ and $K$, such that the following holds. \hspace{5mm} (i) $M$ is K\"ahler with respect to these structures and \hspace{5mm} (ii) $I$, $J$ and $K$, considered as endomorphisms of a real tangent bundle, satisfy the relation $I\circ J=-J\circ I = K$. \hfill This means that the hyperk\"ahler manifold has the natural action of quaternions ${\Bbb H}$ in its real tangent bundle. Therefore its complex dimension is even. Let $\mbox{ad}I$, $\mbox{ad}J$ and $\mbox{ad}K$ be the endomorphisms of the bundles of differential forms over a hyperk\"ahler manifold $M$ which are defined as follows. Define $\mbox{ad}I$. Let this operator act as a complex structure operator $I$ on the bundle of differential 1-forms. We extend it on $i$-forms for arbitrary $i$ using Leibnitz formula: $\mbox{ad}I(\alpha\wedge\beta)=\mbox{ad}I(\alpha)\wedge\beta+ \alpha\wedge \mbox{ad}I(\beta)$. Since Leibnitz formula is true for a commutator in a Lie algebras, one can immediately obtain the following identities, which follow from the same identities in ${\Bbb H}$: \[ [\mbox{ad}I,\mbox{ad}J]=2\mbox{ad}K;\; [\mbox{ad}J,\mbox{ad}K]=2\mbox{ad}I;\; \] \[ [\mbox{ad}K,\mbox{ad}I]=2\mbox{ad}J \] Therefore, the operators $\mbox{ad}I,\mbox{ad}J,\mbox{ad}K$ generate a Lie algebra $\goth g_M\cong \goth{su}(2)$ acting on the bundle of differential forms. We can integrate this Lie algebra action to the action of a Lie group $G_M=SU(2)$. In particular, operators $I$, $J$ and $K$, which act on differential forms by the formula $I(\alpha\wedge\beta)=I(\alpha)\wedge I(\beta)$, belong to this group. \hfill \proposition \label{_there_is_action_of_G_M_Proposition_} There is an action of the Lie group $SU(2)$ and Lie algebra $\goth{su}(2)$ on the bundle of differential forms over a hyperk\"ahler manifold. This action is parallel, and therefore it commutes with Laplace operator. {\bf Proof:} Clear. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill If $M$ is compact, this implies that there is a canonical $SU(2)$-action on $H^i(M,{\Bbb R})$ (see \cite{_so5_on_cohomo_}). \hfill Let $M$ be a hyperk\"ahler manifold with a Riemannian form $\inangles{\cdot,\cdot}$. Let the form $\omega_I := \inangles{I(\cdot),\cdot}$ be the usual K\"ahler form which is closed and parallel (with respect to the connection). Analogously defined forms $\omega_J$ and $\omega_K$ are also closed and parallel. The simple linear algebraic consideration (\cite{_Besse:Einst_Manifo_}) shows that \hfill $\omega_J+\sqrt{-1}\omega_K$ is of type $(2,0)$ and, being closed, this form is also holomorphic. \hfill \definition \label{_canon_holo_symple_form_Definition_ Let $\Omega:= \omega_J+\sqrt{-1}\omega_K$. This form is called {\bf the canonical holomorphic symplectic form of a manifold M}. \hfill Let $M$ be a complex manifold which admits a holomorphic symplectic form $\Omega$. Take the Riemannian metric $(\cdot,\cdot)$ on $M$, and the corresponding Levi-Civitta connection. Assume that $\Omega$ is parallel with respect to the Levi-Civitta connection. Then the metric $(\cdot,\cdot)$ is hyperkaehler% \footnote% {This means that the $(\cdot,\cdot)$ is induced by a hyperk\"ahler structure on $M$.} (\cite{_Besse:Einst_Manifo_}). If some {\bf it compact} K\"ahler manifold $M$ admits non-degenerate holomorphic symplectic form $\Omega$, the Calabi-Yau (\cite{_Yau:Calabi-Yau_}) theorem implies that $M$ is hyperk\"ahler (\ref{_symplectic_=>_hyperkaehler_Proposition_}) This follows from the existence of a K\"ahler metric on $M$ such that $\Omega$ is parallel under the Levi-Civitta connection associated with this metric. \hfill Let $M$ be a hyperk\"ahler manifold with complex structures $I$, $J$ and $K$. For any real numbers $a$, $b$, $c$ such that $a^2+b^2+c^2=1$ the operator $L:=aI+bJ+cK$ is also an almost complex structure: $L^2=-1$. Clearly, $L$ is parallel with respect to a connection. This implies that $L$ is a complex structure, and that $M$ is K\"ahler with respect to $L$. \hfill \definition \label{_induced_structures_Definion_} If $M$ is a hyperk\"ahler manifold, the complex structure $L$ is called {\bf induced by a hyperk\"ahler structure}, if $L=aI+bJ+cK$ for some real numbers $a,b,c\:|\:a^2+b^2+c^2=1$. \hfill \hfill If $M$ is a hyperk\"ahler manifold and $L$ is induced complex structure, we will denote $M$, considered as a complex manifold with respect to $L$, by $(M,L)$ or, sometimes, by $M_L$. \hfill Consider the Lie algebra $\goth{g}_M$ generated by ${ad}L$ for all $L$ induced by a hyperk\"ahler structure on $M$. One can easily see that $\goth{g}_M=\goth{su}(2)$. The Lie algebra $\goth{g}_M$ is called {\bf isotropy algebra} of $M$, and corresponding Lie group $G_M$ is called an {\bf isotropy group} of $M$. By Proposition 1.1, the action of the group is parallel, and therefore it commutes with the action of Laplace operator on differential forms. In particular, this implies that the action of the isotropy group $G_M$ preserves harmonic forms, and therefore this group canonically acts on cohomology of $M$. \hfill \proposition \label{_G_M_invariant_forms_Proposition_} Let $\omega$ be a differential form over a hyperk\"ahler manifold $M$. The form $\omega$ is $G_M$-invariant if and only if it is of Hodge type $(p,p)$ with respect to all induced complex structures on $M$. {\bf Proof:} Assume that $\omega$ is $G_M$-invariant. This implies that all elements of ${\goth g}_M$ act trivially on $\omega$ and, in particular, that $\mbox{ad}L(\omega)=0$ for any induced complex structure $L$. On the other hand, $\mbox{ad}L(\omega)=(p-q)\sqrt{-1}\:$ if $\omega$ is of Hodge type $(p,q)$. Therefore $\omega$ is of Hodge type $(p,p)$ with respect to any induced complex structure $L$. Conversely, assume that $\omega$ is of type $(p,p)$ with respect to all induced $L$. Then $\mbox{ad}L(\omega)=0$ for any induced $L$. By definition, ${\goth g}_M$ is generated by such $\mbox{ad}L$, and therefore ${\goth g}_M$ and $G_M$ act trivially on $\omega$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition \label{_degree_Definition_ Let $M$ be a Kaehler manifold and $\omega\in H^2(M,{\Bbb R})$ be the Kaehler class of $M$. Let $dim_{\Bbb C}(M)=n$. Let $\eta\in H^{2i}(M)$. We define the {\bf degree} of the cohomology class $\eta$ by the formula \[ deg(\eta):= \frac{\int\limits_M \eta\wedge\omega^{n-i}} {{\mbox Vol}(M)}. \] Clearly, if $\eta$ is pure of Hodge type $(p,q)$ and $deg(\eta)\neq 0$, then $p=q$. Let $M$ be a hyperkaehler manifold, and $I$ be an induced complex structure. Then $(M,I)$ is equipped with the canonical Kaehler metric. Consider $(M,I)$ as a Kehler manifold. We define {\bf the degree associated with the induced complex structure $I$} as the linear homomorphism $deg_I:\; H^{2i}(M, {\Bbb R})\longrightarrow {\Bbb R}$ which is equal to degree map \[ deg:\; H^{2i}((M,I), {\Bbb R})\longrightarrow {\Bbb R} \] defined on the cohomology of the Kaehler manifold $(M,I)$. \hfill The following statement follows from a trivial local computation. The more general form of this claim is proven in \cite{Verbitsky:Symplectic_I_}. \hfill \claim \label{_inv_2-forms_have_zero_degree_Claim_} Let $M$ be a hyperkaehler manifold and $\eta\in H^2(M)$ be a $G_M$-invariant hohomology class. Then $deg_I(\eta)=0$ for all induced complex structures $I$. {\bf Proof:} See Theorem 2.1 of \cite{Verbitsky:Symplectic_I_} $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Calabi-Yau theorem provides an elegant way to think of hyperkaehler manifolds in holomorphic terms. Heuristically speaking, compact hyperkaehler manifolds are holomorphic manifolds which admit a holomorphic symplectic form. \hfill \definition \label{_holomorphi_symple_Definition_} The compact complex manifold $M$ is called holomorphically symplectic if there is a holomorphic 2-form $\Omega$ over $M$ such that $\Omega^n=\Omega\wedge\Omega\wedge...$ is a nowhere degenerate section of a canonical class of $M$. There, $2n=dim_{\Bbb C}(M)$. Note that we assumed compactness of $M$.% \footnote{If one wants to define a holomorphic symplectic structure in a situation when $M$ is not compact, one should require also the equation $\nabla'\Omega$ to held. The operator $\nabla':\;\Lambda^{p,0}(M)\longrightarrow\Lambda^{p+1,0}(M)$ is a holomorphic differential defined on differential $(p,0)$-forms (\cite{_Griffiths_Harris_}).} One observes that the holomorphically symplectic manifold has a trivial canonical bundle. A hyperk\"ahler manifold is holomorphically symplectic (see Section \ref{hyperk_manif_Section_}). There is a converse proposition: \hfill \theorem \label{_symplectic_=>_hyperkaehler_Proposition_ (\cite{_Beauville_}, \cite{_Besse:Einst_Manifo_}) Let $M$ be a holomorphically symplectic K\"ahler manifold with the holomorphic symplectic form $\Omega$, a K\"ahler class $[\omega]\in H^{1,1}(M)$ and a complex structure $I$. Assume that \[ deg ([\Omega]\wedge \bar [\Omega]) =2 deg([\omega]\wedge[\omega]) \] Then there exists a {\it unique} hyperk\"ahler structure $(I,J,K,(\cdot,\cdot))$ over $M$ such that the cohomology class of the symplectic form $\omega_I=(\cdot,I\cdot)$ is equal to $[\omega]$ and the canonical symplectic form $\omega_J+\sqrt{-1}\:\omega_K$ is equal to $\Omega$. \hfill \ref{_symplectic_=>_hyperkaehler_Proposition_} immediately follows from the Calabi-Yau theorem (\cite{_Yau:Calabi-Yau_}). $\:\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{Moduli spaces for hyperkaehler and holomorphically symplectic manifolds.} \label{_moduli_Section_} \definition \label{_simple_hyperkaehler_mfolds_Definition_} (\cite{_Beauville_}) The connected simply connected compact hyperkaehler manifold $M$ is called {\bf simple} if $M$ cannot be represented as a Cartesian product of two (non-trivial) hyperkaehler manifolds: \[ M\neq M_1\times M_2, \] where $M_1$, $M_2$ are hyperkaehler manifolds such that $dim\; M_1>0$, $dim\; M_2>0$. \hfill Bogomolov proved that every compact hyperkaehler manifold has a finite covering which is a Cartesian product of a compact torus and simple hyperkaehler manifolds. Even if our results could be easily carried over for all compact hyperkaehler manifolds, we restrict ourselves to the case of non-decomposable manifold to simplify the argument. \hfill Let $M$ be a simple hyperkaehler manifold. According to Bogomolov's theorem (\cite{_Beauville_}), for every induced complex structure $I$, \[dim_{\Bbb C} \bigg(H^{2,0}((M,I))\bigg)=1.\] This means that the space of holomorphic symplectic forms on $(M,I)$ is one-dimensional. {}From now on, we assume that $M$ is a simple compact hyperkaehler manifold, which is not a torus. \hfill The moduli spaces of the hyperkaehler and holomorphically symplectic manifolds were first studied by Bogomolov (\cite{_Bogomolov_}). The studies were continued by Todorov ([Tod]). \hfill Let $M_{C^\infty}$ be $M$ considered as a differential manifold. Let $\mbox{\it Diff}$ be the group of diffeomorphisms of $M$. Recall that the {\bf hyperkaehler structure} on $M_{C^\infty}$ was defined as a quadruple $(I, J, K, (\cdot, \cdot ))$ where \[ I, J, K \in End(TM_{C^\infty}), \;\; I^2=J^2=K^2=-1 \] are operators on the tangent bundle $TM_{C^\infty}$ and $(\cdot, \cdot )$ is a Riemannian form. This quadruple must satisfy certain relations (\ref{_hyperkaehler_manifold_Definition_}). Let $\widetilde{\mbox{\it Hyp}}$ be the set of all hyperkaehler structures on $M_{C^\infty}$. Clearly, the group $\mbox{\it Diff}$ acts on $\widetilde{\mbox{\it Hyp}}$. The set of all non-isomorphic hyperkaehler structures on $M_{C^\infty}$ is in bijective correspondence with the set of orbits of $\mbox{\it Diff}$ on $\widetilde{\mbox{\it Hyp}}$. However, the geometrical properties of $\widetilde{\mbox{\it Hyp}}/\mbox{\it Diff}$ are not satisfactory: as a rule, the natural topology on $\widetilde{\mbox{\it Hyp}}/\mbox{\it Diff}$ is not separable etc. To produce a more geometrical moduli space of the hyperkaehler structures, we will refine the space $\widetilde{\mbox{\it Hyp}}/\mbox{\it Diff}$ in accordance with the general algebro-geometrical formalism of marked coarse moduli spaces. \hfill Let $\underline{\widetilde{Comp}}$ be the set of all integrable complex structures on $M_{C^\infty}$. In other words, $\underline{\widetilde{Comp}}$ is the set of all operators \[ I\in End(TM_{C^\infty}),\;\; I^2=-1 \] such that the almost complex structure defined by $I$ is integrable. The set of all non-isomorphic complex structures on $M_{C^\infty}$ is in one-to-one correspondence with $\underline{\widetilde{Comp}}/\mbox{\it Diff}$. For every $I\in \underline{\widetilde{Comp}}$, we say that $I$ {\bf admits a hyperkaehler structure} when there exist a hyperkaehler structure $(I, J, K, (\cdot,\cdot))$ on $M$. We say that $I$ is {\bf holomorphically symplectic} when the manifold $(M, I)$ admits a non-degenerate holomorphic symplectic form. The set $\underline{\widetilde{Comp}}$ is endowed with a natural topology. \hfill Let $M$ be a compact hyperkaehler manifold. Clearly, each of induced complex structures on $M$ is contained in the same connected component of $\underline{\widetilde{Comp}}$. Denote this component by $\underline{\widetilde{Comp}}^\circ $. \ref{_symplectic_=>_hyperkaehler_Proposition_} immediately implies the following statement: \hfill \corollary \label{_defo_simple_if_it's_Kaeh_Corollary_} Let $I\in \underline{\widetilde{Comp}}^\circ$. Then $(M,I)$ admits the hyperkaehler structure if and only if $(M,I)$ admits a holomorphic symplectic structure and is of Kaehler type% \footnote{We say that a complex manifold $X$ is {\bf of Kaehler type} if $X$ admits a Kaehler metric.}. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Denote the set of all $I\in \underline{\widetilde{Comp}}^\circ$ which admit the hyperkaehler structure by $\widetilde{Comp}^\circ$. \hfill The pair $(I,\Omega)$ is called {\bf the holomorphic symplectic structure on the differential manifold $M_{C^\infty}$} if $I$ is a complex structure on $M_{C^\infty}$ and $\Omega$ is a holomorphic symplectic form over $(M,I)$. Let $\widetilde{Symp}^\circ$ denote the set of all holomorphic symplectic structures $(I,\Omega)$ on $M_{C^\infty}$ such that $I \in {\widetilde{Comp}}^\circ$. Let $\mbox{\it Diff}\,^\circ $ be the set of all $x\in\mbox{\it Diff}$ which act trivially on the cohomology $H^*(M,{\Bbb R})$. Denote by $\widetilde{\mbox{\it Hyp}}^\circ $ the connected component of $\widetilde{\mbox{\it Hyp}}$ which contains the initial hyperkaehler structure on $M$. Let $\mbox{\it Hyp}:= \widetilde{\mbox{\it Hyp}}^\circ /\mbox{\it Diff}\,^\circ$, $Symp:=\widetilde{Symp}^\circ/\mbox{\it Diff}\,^\circ$ and $Comp:= \widetilde{Comp}^\circ /\mbox{\it Diff}\,^\circ $. These spaces are endowed with the natural topology. Their points can be considered as the classes of hyperkaehler (respectively, holomorphically symplectic and complex) structures on $M_{C^\infty}$ up to the action of $\mbox{\it Diff}\,^\circ $. Slightly abusing the language, we will refer to these points as to hyperkaehler (resp., holomorphically symplectic and complex) structures. For each $I\in Comp$, we denote $M$, considered as a complex manifold with the complex structure $I$, by $(M,I)$. It is clear that $(M,I)$ is holomorphically symplectic and admits a hyperkaehler structure for all $I\in Comp$. \hfill \definition \label{_Comp_Hyp_Definition_} The spaces $\mbox{\it Hyp}$, $Symp$, $Comp$ are called {\bf the coarse moduli spaces of deformations of the hyperkaehler (respectively, holomorphically symplectic and complex) structure on the marked compact manifold of hyperkaehler type.} \hfill The word {\bf marked} refers to considering the factorization by $\mbox{\it Diff}\,^\circ $ instead of $\mbox{\it Diff}$. This is roughly equivalent to fixing the basis in the cohomology $H^*(M)$, hence ``marking''. \section {Periods and forgetful maps.}\label{_perio_and_forge_Section_} In assumptions of Section \ref{_moduli_Section_}, let $\mbox{\it Hyp}$, $Symp$, $Comp$ be the moduli spaces of \ref{_Comp_Hyp_Definition_}. We define the {\bf period map} \[ \tilde P_{hyp}: \;\widetilde{\mbox{\it Hyp}}\longrightarrow H^2(M,{\Bbb R})\otimes {\Bbb R}^3 \] as a rule which associates with every hyperkaehler structure $(I,J,K,(\cdot,\cdot))$ on $M_{C^\infty}$ the triple \[ ([\omega_I],\: [\omega_J],\: [\omega_K])\in H^2(M_{C^\infty},{\Bbb R})\times H^2(M_{C^\infty},{\Bbb R}) \times H^2(M_{C^\infty},{\Bbb R}) \] of Kaehler classes corresponding to $I$, $J$ and $K$ respectively. By definition, the group $\mbox{\it Diff}\,^\circ $ acts trivially on $H^2(M)$. Therefore, $\tilde P_{hyp}$ descends to a map \[ P_{hyp}: \;\mbox{\it Hyp}\longrightarrow H^2(M,{\Bbb R})\otimes {\Bbb R}^3 \] Similarly, define the Griffiths' period map \[ P_c:\; Comp\longrightarrow {\Bbb P}^1(H^2(M_{C^\infty},{\Bbb C})) \] as a rule which relies the 1-dimensional complex subspace \[ H^{2,0}((M_{C^\infty},I))\subset H^2(M_{C^\infty},{\Bbb C}) \] to the complex structure $I$ on $M_{C^\infty}$. Using Dolbeault spectral sequence, one can easily see that the subspace $H^{2,0}((M_{C^\infty},I))\subset H^2(M_{C^\infty},{\Bbb C})$ is defined independently on the Kaehler metric. Let $P_s:\; Symp\longrightarrow H^2(M,{\Bbb C})$ map a pair $(I,\Omega)\in Symp$ to the class $[\Omega]\in H^2(M,{\Bbb C})$ which is represented by the closed 2-form $\Omega$. There exist a number of natural ``forgetful maps'' between the spaces $\mbox{\it Hyp}$, $Symp$ and $Comp$. Here we define some of these maps and find how these maps relate to period maps. Let $\c H= (I,J,K, (\cdot,\cdot))\in \mbox{\it Hyp}$ be a hyperkaehler structure. As in \ref{_canon_holo_symple_form_Definition_}, consider the canonical holomorphic symplectic form $\Omega:= \omega_J+\sqrt{-1}\: \omega_K$ associated with $\c H$. Let $\Phi^{hyp}_s:\; \mbox{\it Hyp}\longrightarrow Symp$ map $\c H$ to the pair $(I, \Omega)\in Symp$. Let $\Phi^{hyp}_s:\; \mbox{\it Hyp}\longrightarrow Comp$ map $\c H$ to $I\in Comp$. Let $\Phi^s_c:\; Symp\longrightarrow Comp$ map $\c S=(I, \Omega)\in Symp$ to $I\in Comp$. For $h=(x_1,x_2,x_3)\in H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})\times H^2(M,{\Bbb R})$, let $\pi_i(h)=x_i$. \hfill \claim \label{_forgetting-n-periods_Claim_} Let $\c H\in Hyp$, $\c S\in Symp$. Then (i) $P_s(\Phi^{hyp}_s(\c H))= \pi_2(P_{hyp}(\c H))+ \sqrt{-1}\:\pi_3(P_{hyp}(\c H))$ (ii) The point $P_c(\Phi^{s}_c(\c S))\in {\Bbb P} H^2(M,{\Bbb C})$ corresponds to a line generated by $P_s(\c S)$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill According to the general formalism of Kodaira and Kuranishi, $Comp$ is endowed with a canonical structure of a complex variety. Using the complex structure on $Comp$, we describe the map $P_c^s:\; Symp\longrightarrow Comp$ in terms of algebraic geometry. Let $L$ be a holomorphic vector bundle over a complex variety $X$. Let $Tot(L)$ be the total space of $L$. By definition, $Tot(L)$ is a complex variety which is smoothly fibered over $X$. Every holomorphic section $f\in \Gamma_X(L)$ gives a standard holomorphic map $s_f:\; X\longrightarrow Tot(L)$. Consider the map $s_0:\; X\longrightarrow Tot(L)$ corresponding to the zero section of $L$. This map identifies $X$ with the closed analytic subspace of $Tot(L)$. Let $Tot^*(L):= Tot(L)\backslash X$ be the completion of $Tot(L)$ to $X$. \hfill \proposition \label{_Symp_as_a_total_space_Proposition_} Let $M$ be a hyperkaehler manifold, $Comp$ and $Symp$ be the moduli spaces associated with $M$ as in \ref{_Comp_Hyp_Definition_}. Then there exist a natural holomorphic linear bundle $\tilde \Omega$ on $Comp$ such that the following conditions hold. (i) There exist a natural homeomorphism $i:\; Tot^*(\tilde \Omega)\longrightarrow Symp$. (ii) Let $\pi:\; Tot^*(\tilde \Omega)\longrightarrow Comp$ be the standard projection. Then the diagram \[ \begin{array}{ccccc} Tot^*(\tilde \Omega)&\!\!\!\stackrel {i}\longrightarrow \!\!\!& Symp\\[3mm] \;\;\;\;\;\searrow\!\!{}^\pi\!\!\!\!\! && \!\!\!\swarrow\!\!{}_{\Phi^s_c}\;\;\;\;\; \\[3mm] & \!\!\!\!\!Comp &\\ \end{array} \] is commutative. \hfill The homeomorphism $i$ defines a complex analytic structure on $Symp$. Further on, we consider both $Comp$ and $Symp$ as complex analytic varieties. \hfill {\bf Proof of \ref{_Symp_as_a_total_space_Proposition_}:} Let ${A}:= Comp\times H^2(M, {\Bbb C})$. The holomorphic symplectic structure $(I,\Omega)\in Symp$ is uniquely defined by $I\in Comp$ and the cohomology class $[\Omega]\in H^2(M, {\Bbb C})$. This defines an injection \[ j:\; Symp\hookrightarrow {A}, \;\; j(I,\Omega)= (I, [\Omega]). \] Consider ${A}$ as the total space of a trivial holomorphic bundle ${A}_b$ with the fiber $H^2(M,{\Bbb C})$. We construct $\tilde \Omega$ as a linear subbundle of ${A}_b$, such that its total space coinsides with $j(Symp)$. \hfill Let $U\subset Comp$ be an open set. We say that {\bf there exists a universal fibration over $U$} if there exist a smooth complex analytic fibration $\pi:\;\goth M\longrightarrow U$ such that for all $J\in U$, the fiber $\pi^{-1}(J)$ is isomorphic to $(M,J)$. \hfill \claim \label{_unive_fibra_exi_loca_Claim_} For all $I\in Comp$, there exist an open set $U\subset Comp$, $I\in U$, which admits universal fibration. {\bf Proof:} This is a consequence of Kodaira-Spencer theory (see \cite{_Kodaira_Spencer_}). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $U\subset Comp$ be an open subset which admits a universal fibration $\goth M\stackrel \pi \longrightarrow U$. Let ${\Bbb C}_{\goth M}$ be the constant sheaf over $\goth M$, and $\pi_\bullet$ be the sheaf-theoretic direct image. Let $H^2:=R^2\pi_\bullet {\Bbb C}_{\goth M}$ be the second derived functor of $\pi_\bullet$ applied to ${\Bbb C}_{\goth M}$. Since ${\Bbb C}_{\goth M}$ is a constructible sheaf, and $\pi$ is a proper morphism, the sheaf $H^2$ is also constructible. For every point $I\in U$, the restriction $H^2\restrict{I}$ is isomorphic to $H^2(M,{\Bbb C})$. Hence, $H^2\restrict{I}$ is a locally constant sheaf. This sheaf is equipped with a natural flat connection, known as Gauss-Manin connection. Since $U$ is a subset in the space of {\it marked} deformations of $M$, the monodromy of Gauss-Manin connection is trivial. Hence, the bundle $H^2$ is naturally isomorphic to ${A}_b\restrict{U}$. Let $F^0\subset F^1\subset F^2={A}_b\restrict{U}$ be the variation of Hodge structures associated with $\pi$. By definition, $F^i$ are holomorphic sub-bundles of ${A}_b\restrict{U}$. For every $I\in U$, we have $F^0\cong H^{2,0}((M,I))$. Therefore, $F^0$ is a linear sub-bundle of ${A}_b\restrict{U}$. \hfill \lemma \label{_hol_bundle_from_Hodge_str_and_j_Lemma_} Let $U\subset Comp$ be an open set which admits an universal fibration $\goth M\longrightarrow U$. Let $F^0\subset {A}_b\restrict{U}$ be the holomorphic linear bundle defined as above. Let \[ Symp(U):= \{ (I,\Omega)\in Symp\;\;|\;\; I\in U\}. \] Let $Tot(F^0)\subset U\times H^2(M,{\Bbb R})$ be the total space of $F^0$ considered as a subspace in a total space of ${A}_b\restrict{U}$. Then $Tot^*(F^0)$ coinsides with $j(Symp(U))$ {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill For every open set which admits an universal fibration $\goth M\longrightarrow U$, we defined the holomorphic linear bundle $F^0\subset {A}_b\restrict{U}$. \ref{_hol_bundle_from_Hodge_str_and_j_Lemma_} implies that $F^0\subset {A}_b\restrict{U}$ is independent from the choice of the universal fibration, and that the locally defined sub-bundles $F^0$ can be glued to a globally defined holomorphic linear sub-bundle in ${A}_b$. Denote this linear sub-bundle by $\tilde \Omega$. It is clear that $Tot^*(\tilde \Omega)$ coinsides with $j(Symp)$. Since $j$ is injective, there exist an inverse homomorphism $i:\; Tot^*(\tilde \Omega)\longrightarrow Simp$. The condition (ii) of \ref{_Symp_as_a_total_space_Proposition_} is obvious. It remains to show that the bijective maps $i$ and $j= i^{-1}$ are continuous. This is left to the reader as an exercise. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section[Hodge-Riemann relations for the hyperkaehler manifolds and period map.] {Hodge-Riemann relations for the \\ hyperkaehler manifolds and period map.} \label{_Period_and_Hodge_Riemann_Section_} \subsection{What do we do in this section:} With every hyperkaehler manifold $M$, we associate the action of the group $G_M\cong SU(2)$ on the cohomology of $M$ (see \ref{_there_is_action_of_G_M_Proposition_}). Let $P_{hyp}(M):= (\omega_I,\omega_J,\omega_K)$ be the periods of $M$. We show that the action of $G_M$ may be reconstructed from the periods. This follows from \ref{_Lambda_dual_to_L_Proposition_} and \ref{_g_m_from_L_Lambda_Claim_}. The action of $G_M\cong SU(2)$ on $H^2(M)$ induces a weight decomposition of $H^2(M)$. Using this decomposition, we obtain an interesting version of Hodge-Riemann relations (\ref{_restrictions_of_pairings_to_H^2_Lemma_}). In particular, we obtain that for every hyperkaehler structure $(I,J,K,(\cdot,\cdot))$ on $M$, the Hodge-Riemann pairings associated with the complex structures $I$, $J$ and $K$ are equal (\ref{_pairings_on_H^2_are_equal_Proposition_}). \subsection{Hodge-Riemannian pairing.} \vspace{3mm} \hspace{5.5mm}In this subsection we follow \cite{_Weil_}. \hfill Let $X$ be a compact Kaehler manifold, and $\Lambda^*(X)=\oplus \Lambda^{p,q}(X)$ be a space of differential forms equipped with Hodge decomposition. The Riemannian structure on $X$ equips $\Lambda^*(X)$ with a positively defined Hermitian metric (see \cite{_Weil_} for correct normalization of this metric). Integrating the scalar product of two forms over $X$, we obtain a Hermitian positively defined pairing on the space of global sections of $\Lambda^*(X)$. Let us identify the cohomology space of $X$ with the space of harmonic differential forms. This gives a positively defined Hermitian product on the space of cohomology $H^*(X)$. We denote it by \[ (\cdot,\cdot)_{Her}:\; H^i(X,{\Bbb C})\times H^i(X,{\Bbb C}) \longrightarrow {\Bbb C}. \] Let $I:\; H^i(X,{\Bbb C})\times H^i(X,{\Bbb C}) \longrightarrow {\Bbb C}$ map $(x,y)$ to $(x,\bar y)_{Her}$. Clearly, $I$ is a complex-linear non-degenerate 2-form on $H^i(X,{\Bbb C})$, which is defined over reals. Let $A:\; H^i(X,{\Bbb C})\times H^i(X,{\Bbb C}) \longrightarrow {\Bbb C}$ map $x,y \in H^i(X,{\Bbb C})$ to \[ \int_X x\wedge y\wedge \omega^{n-i}, \] where $n=\dim_{\Bbb C} X$, and $i\leq n$. Let $C:\; H^*(X,{\Bbb C})\longrightarrow H^*(X,{\Bbb C})$ be the Weil operator, which maps a cohomology class $\omega\in H^{p,q}(X)\subset H^{p+q}(X)$ to $\sqrt{-1}\:^{p-q} \omega$. Let $L:\; H^i(X)\longrightarrow H^{i+2}(X)$, $\Lambda:\; H^i(X)\longrightarrow H^{i-2}(X)$ be the Hodge operators, and $P^i(X)\subset H^i(X)$ be the space of primitive cohomology classes: \[ P^i(X) = \{ \alpha \in H^i(X) \;\; |\;\; \Lambda(\alpha) \} =0 \] By Lefshetz theorem, \[ H^i(X) = \oplus L^r P^{i-r}(X). \] Let $p_r:\; H^i(X) \longrightarrow P^{i-r}(X)$ be a map corresponding to this decomposition, such that for all $a\in H^*(M)$, \[ a = \sum_r L^r p_r(a). \] The forms $A$ and $I$ are related by the so-called Hodge-Riemann equation: \begin{equation}\label{_Hodge_Riemann_general_Equation_} (-1)^\frac{(n-i)(n-i-1)}{2} A(a,C b) = \sum_r \mu_r\frac{(n-p+r)!}{r!} I(L^r p_r(a), L^r p_r(b)), \end{equation} where $\mu_r$ are positive real constants which depend only on $r$ and dimension of $X$. Let $\omega\in H^2(X, {\Bbb R})$ be a Kaehler class of $X$. We call the form \begin{equation}\label{_Hodge_Riemann_form_general_Equation_} (-1)^\frac{(n-i)(n-i-1)}{2} A(a,C b): \; H^i(X,{\Bbb C})\times H^i(X,{\Bbb C}) \longrightarrow {\Bbb C} \end{equation} {\bf the Hodge-Riemann pairing associated with a Kaehler class $\omega$} and denote this form by $(\cdot,\cdot)_{\omega}$. \hfill \claim \label{_Hodge_Riema_general_Claim_} (i) The form $(\cdot,\cdot)_{\omega}$ depends only on the Kaehler class $\omega\in H^2(X,{\Bbb R})$ of $X$. In other words, $(\cdot,\cdot)_{\omega}$ would not change if we modify the complex structure or Kaehler metric, provided that the Kaehler class stays the same. (ii) The form $(\cdot,\cdot)_{\omega}$ is defined over reals. (iii) If $X$ is a surface, then the restriction of $(\cdot,\cdot)_{\omega}$ to the primitive cohomology $P^2(X)\subset H^2(X)$ coincides with the intersection form. (iv) Restriction of $(\cdot,\cdot)_{\omega}$ to $H^2(X)$ can be written as follows: \begin{equation} \label{_H_R_to_H^2_formula_Equation_} (\eta_1,\eta_2)_{\omega}=\int_X \omega^{n-2}\wedge \eta_1\wedge\eta_2 - \frac{n-2}{(n-1)^2} \cdot \frac{ \int_X \omega^{n-1}\eta_1 \cdot \int_X\omega^{n-1}\eta_2} {\int_X \omega^n}. \end{equation} \hfill {\bf Proof:} Follows from \eqref{_Hodge_Riemann_general_Equation_} (see also \cite{_Weil_}). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \subsection{Riemann-Hodge relations in hyperkaehler case.} Let $M$ be a compact hyperkaehler manifold and $\c H\in Hyp$ be a hyperkaehler structure on $M$. Let $P_{hyp}(\c H)$ be denoted by $(x_1, x_2, x_3)$, $x_i\in H^2(M,{\Bbb R})$. The hyperkaehler structure $\c H\in Hyp$ defines a Riemannian metric on $M$. This metric establishes a positively defined Hermitian scalar product on the space of ${\Bbb C}$-valued differential forms over $M$. Realizing the cohomology classes as harmonic forms, we obtain a Hermitian pairing $(\cdot,\cdot)_{Her}$ on $H^i(M,{\Bbb C})$. Let $\Lambda_{x_i}\!:\; H^i(M)\longrightarrow H^{i-2}(M)$ be the operator adjoint to $L_{x_i}$ with respect to the pairing $(\cdot,\cdot)_{Her}$. Clearly, $\Lambda_{x_i}$ is the Hodge operator associated with the Kaehler structure on $M$ which is defined by $x_i$ and ${\cal H}$. Let $\inangles{\cdot,\cdot}_{x_i}$ be the Hodge-Riemann form \eqref{_Hodge_Riemann_form_general_Equation_} associated with the Kaehler form $x_i$. Let $L\check{\;}_{x_i}$ be an operator adjoint to $L_{x_i}$ with respect to $\inangles{\cdot,\cdot}_{x_i}$. \hfill \proposition \label{_Lambda_dual_to_L_Proposition_} \ \ $L\check{\;}_{x_j}=\Lambda_{x_j}$ \ \ \ for $j=1,2,3$. {\bf Proof:} Fix a choice of $j\in\{1,2,3\}$. For simplicity, assume that $j=1$. We abbreviate $L_{x_1}$ by $L$, $\Lambda_{x_1}$ by $\Lambda$. Let $I$ be the complex structure induced by ${\cal H}$, such that $x_1\in H^2(M,{\Bbb R})$ is the Kaehler class of $I$. Take the Lefschetz decomposition (\cite{_Griffiths_Harris_}) \[ H^k(M)=\bigoplus L^i P^{k-2i}(M),\;\;\; P^{k-2i}\subset H^{k-2i}(M), \] where the space $P^i(M)$ is a space of all primitive classes: \[ P^i(M)= ker\bigg(\Lambda:\; H^i(M)\longrightarrow H^{i+2}(M)\bigg). \] Let $P^{p,q}(M):= P^{p+q}(M)\cap H^{p,q}(M)$. It is well known that \[ P^{i}(M)=\bigoplus\limits_{p+q=i}P^{p,q}(M). \] (see \cite{_Griffiths_Harris_}, \cite{_Weil_}). Hodge-Riemann relations (\eqref{_Hodge_Riemann_general_Equation_}; see also \cite{_Weil_}) describe $(\zeta_1,\zeta_2)_{Her}$ in terms of $\inangles{\zeta_1,\bar\zeta_2}_{x_1}$ and Lefschetz decomposition. Let $n=dim_{\Bbb C}(M)$. When \[ \zeta_1\in L^i P^{p,q}(M), \;\; \zeta_2\in L^{i'} P^{p',q'}(M) \;\;\;\mbox{\it and} \;\;\;(i,p,q)\neq (i',p',q'), \] both scalar products vanish: \begin{equation}\label{_Hodge_Riemann_relations_vanishing_Equation} (\zeta_1,\zeta_2)_{Her}= \inangles{\zeta_1,\bar\zeta_2}_{x_1}=0. \end{equation} \hfill When $\zeta_1\in L^i P^{p,q}(M)$ and $\zeta_2\in L^i P^{p,q}(M)$, we have \begin{equation}\label{_Hodge_Riemann_relations_Equation_} (\zeta_1,\zeta_2)_{Her}= \sqrt{-1}\:^{p-q}(-1)^{\frac{(n-p-q)(n-p-q-1)}{2}} \inangles{\zeta_1,\bar\zeta_2}_{x_1}. \end{equation} \hfill The operator $L\check{\;}_{x_1}$ is adjoint to $L_{x_1}$ with respect to $\inangles{\cdot,\cdot}_{x_1}$ and $\Lambda_{x_1}$ is adjoint to $L_{x_1}$ with respect to $(\cdot,\cdot)_{Her}$. Let $\zeta\in L^i P^{p,q}(M)$. By definition, $t=L\check{\;}_{x_1}(\zeta)$ is the element of $H^{2i-2+p+q}(M)$ such that $\forall \xi \in H^{2i-2+p+q}(M)$ we have \begin{equation} \label{_L_galochka_definition_Equation_} \inangles{t,\xi}_{x_1} = \inangles{\zeta,L_{x_1}(\xi)}_{x_1}. \end{equation} Using \eqref{_Hodge_Riemann_relations_vanishing_Equation}, we see that $t\in L^{i-1} P^{p,q}(M)$. For $t\in L^{i-1} P^{p,q}(M)$, \eqref{_Hodge_Riemann_relations_vanishing_Equation} shows that if \eqref{_L_galochka_definition_Equation_} holds for all $\xi\in L^{i-1} P^{p,q}(M)$, this equation holds for all $\xi \in H^{2i-2+p+q}(M)$. On the other hand, Hodge-Riemann relations \eqref{_Hodge_Riemann_relations_Equation_} imply that for $\xi,t\in L^{i-1} P^{p,q}(M)$, \[ (t,\bar \xi)_{Her}= \sqrt{-1}\:^{p-q}(-1)^{\frac{(n-p-q)(n-p-q-1)}{2}} \inangles{t,\xi}_{x_1} = \] \[ = \sqrt{-1}\:^{p-q}(-1)^{\frac{(n-p-q)(n-p-q-1)}{2}} \inangles{\zeta,L_{x_1}(\xi)}_{x_1}= (\zeta,L_{x_1}\bar\xi)_{Her}. \] Therefore $L\check{\;}_{x_1}$ is adjoint to $L_{x_1}$ with respect to $(\cdot,\cdot)_{Her}$. \ref{_Lambda_dual_to_L_Proposition_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill In Section \ref{hyperk_manif_Section_}, we defined the action of $G_M\cong SU(2)$ on the cohomology of a hyperkaehler manifold. For every hyperkaehler structure ${\cal H}\in \mbox{\it Hyp}$, there is an action of $SU(2)$ on $H^*(M_{C^\infty})$ which is determined by ${\cal H}$. We proceed to describe this $SU(2)$-action in terms of the triple \[ (x_1,x_2,x_3)=P_{hyp}({\cal H})\in H^2(M)\oplus H^2(M)\oplus H^2(M). \] \ref{_Lambda_dual_to_L_Proposition_} expresses the Hodge operators $\Lambda_{x_i}$ via the multiplicative structure on $H^*(M_{C^\infty})$. Let $\goth a_{\cal H}\subset End(H^*(M,{\Bbb R})$ be the Lie algebra generated by $L_{x_i}$, $\Lambda_{x_i}$, $i=1,2,3$. According to \cite{_so5_on_cohomo_}, $\goth a_{\cal H}\cong \goth{so}(4,1)$. \hfill Let ${\goth g}_{\cal H}\subset \goth a_{\cal H}$ be the subalgebra of $\goth a_{\cal H}$ consisting of all elements which respect the grading on $H^*(M)$ induced by the degree: \[ {\goth g}_{\cal H}:= \{ x\in \goth a_{\cal H}\;\; | \;\; x(H^i(M))\subset H^i(M), \;\;i=0,1, ... \; 2n.\} \] \hfill \claim \label{_g_m_from_L_Lambda_Claim_} The Lie algebra ${\goth g}_{\cal H}$ is isomorphic to $\goth{su}(2)$. Its action coincides with that of ${\goth g}_M$ defined in Section \ref{hyperk_manif_Section_}. {\bf Proof:} This is Theorem 2 of \cite{_so5_on_cohomo_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill The forms $(\cdot,\cdot)_{x_i}$, $i=1,2,3$ depend only on the value of $x_i\in H^2(M,{\Bbb R})$. Therefore, \ref{_Lambda_dual_to_L_Proposition_} has the following interesting consequence: \hfill \corollary \label{_so(5)_inde_from_H_Corollary_} % Let $\c H\in Hyp$ be a hyperkaehler structure on $M$, and ${\goth g}_{\cal H}\cong \goth{su}(2)$, $\goth a_{\c H}\cong \goth{so}(4,1)$ be the corresponding Lie algebras defined as above. Then the action of ${\goth g}_{\cal H}$, $\goth a_{\c H}$ on $H^*(M,{\Bbb R})$ depends only on hyperkaehler periods of $\c H$. In other words, if $\c H_1$, $\c H_2$ are hyperkaehler structures such that \[P_{hyp}(\c H_1)= P_{hyp}(\c H_2),\] then action of $\goth a_{\c H_1}$, ${\goth g}_{\c H_1}$ on $H^*(M)$ coinsides with action of $\goth a_{\c H_2}$, ${\goth g}_{\c H_2}$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \proposition \label{_pairings_on_H^2_are_equal_Proposition_} In assumptions of \ref{_Lambda_dual_to_L_Proposition_}, let $\inangles{\cdot,\cdot}_i$ be the restriction of the pairing $\inangles{\cdot,\cdot}_{x_i}$ to $H^2(M)$. Then $\inangles{\cdot,\cdot}_1 = \inangles{\cdot,\cdot}_2=\inangles{\cdot,\cdot}_3$. {\bf Proof:} Let $(\cdot,\cdot)$ be the restriction of $(\cdot,\cdot)_{\cal H}$ to $H^2(M)$. Let $V$ be the subspace of $H^2(M)$ spanned by $(x_1,x_2,x_3)$. Earlier, we defined the action of the Lie algebra ${\goth g}_{\cal H}\cong\goth{su}(2)$ on $H^2(M)$. Let $H_{inv}$ be the space of all ${\goth g}_{\cal H}$-invariant elements in $H^2(M)$. According to \cite{_so5_on_cohomo_}, the action of ${\goth g}_{\cal H}$ on $H^*(M)$ induces the Hodge decomposition. Namely, for every induced complex structure $I$ there exist a Cartan subalgebra $\goth h\in {\goth g}_{\cal H}$ such that the weight decomposition on $H^*(M)$ induced by $\goth h$ coincides with the Hodge decomposition \[ H^i(M)=\bigoplus\limits_{p+q=i}H^{p,q}(M). \] The space $H^{2,0}(M)$ is one-dimensional for every induced complex structure. Using the theory of representations of $\goth{sl}(2)$, one can check that this implies that $H^2(M)/H_{inv}$ is a simple 3-dimensional representation of ${\goth g}_{\cal H}$. Another trivial calculation shows that $V$ is a ${\goth g}_{\cal H}$-invariant subspace of $H^2(M)$, and ${\goth g}_{\cal H}$ acts on $V$ non-trivially. Therefore $H^2(M)=H_{inv}\oplus V$. \ref{_pairings_on_H^2_are_equal_Proposition_} is implied by the following lemma. \hfill \lemma \label{_restrictions_of_pairings_to_H^2_Lemma_} Consider the restrictions of $\inangles{\cdot,\cdot}_i$ and $(\cdot,\cdot)$ to $V$ and $H_{inv}$. Then \begin{equation} \label{_restriction_to_V_Equation_} \inangles{x,\bar y} = (x,\bar y) \;\;\;\mbox{for} \;\;x,y\in V, \end{equation} \begin{equation} \label{_restriction_to_H_inv_Equation_} \inangles{x,\bar y}_i = -(x,\bar y)\;\;\;\mbox{for}\;\; x,y\in H_{inv}, \;\: \mbox{\it\ and finally,} \end{equation} \begin{equation} \label{_restriction_to_H_and_V_Equation_} \inangles{x,\bar y}_i = (x,\bar y)=0\;\;\;\mbox{for} \;\;x\in V,\; y\in H_{inv}. \end{equation} \hfill {\bf Proof:} Let $I$ be an induced complex structure. Then $V=H^{2,0}\oplus L_I (H^{0,0})\oplus H^{0,2}(M)$. where the Hodge decomposition is taken with respect to $I$ and $L_I$ is the Hodge operator of exterrior multiplication by the Kaehler class of $I$. Then \eqref{_restriction_to_V_Equation_} immediately follows from Hodge-Riemann relations \eqref{_Hodge_Riemann_relations_Equation_}. By \ref{_inv_2-forms_have_zero_degree_Claim_} all elements of $H_{inv}$ are primitive. On the other hand, $H_{inv}\subset H^{1,1}$ by \ref{_G_M_invariant_forms_Proposition_}. Therefore \eqref{_restriction_to_H_inv_Equation_} follows from \eqref{_Hodge_Riemann_relations_Equation_}, and \eqref{_restriction_to_H_and_V_Equation_} follows from \eqref{_Hodge_Riemann_relations_vanishing_Equation}. \ref{_restrictions_of_pairings_to_H^2_Lemma_} and consequently \ref{_pairings_on_H^2_are_equal_Proposition_} are proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \corollary \label{_indu_comple_same_HR_Corollary_} Let $\c H= (I, J, K, (\cdot,\cdot))$ be a hyperkaehler structure on $M$, and $L=a I + bJ + cK$ be an induced complex structure, $a^2+b^2+c^2=1$. Let $\omega_1\in H^2(M,{\Bbb R})$ be the Kaehler class associated with the Kaehler manifold $(M, I)$, and $\omega\in H^2(M,{\Bbb R})$ be the Kaehler class associated with the Kaehler manifold $(M, L)$. Let $(\cdot,\cdot)_\omega$, $(\cdot,\cdot)_{\omega_1}:\; H^2(M)\times H^2(M)\longrightarrow {\Bbb C}$ be the Hodge-Riemann forms associated with $\omega$, $\omega_1$. Then $(\cdot,\cdot)_{\omega_1}=(\cdot,\cdot)_\omega$ {\bf Proof:} Follows from \ref{_restrictions_of_pairings_to_H^2_Lemma_} $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{The Hodge-Riemann metric on $H^2(M)$ does not depend on complex structure.} \label{_Hodge-Rie_independent_Section_} Let $M_{C^\infty}$ be a compact manifold which admits a hyperkaehler structure. Let $Hyp$, $Symp$, $Comp$ be the moduli spaces constructed in Section \ref{_moduli_Section_}. \hfill \definition \label{_Hodge_Riemann_asso_w_hyperkeahler_Definition_} Let $M$ be a hyperkaehler manifold, $\c H =(I, J, K, (\cdot, \cdot))$ be its hyperkaehler structure and $\omega_1$, $\omega_2$, $\omega_3\in H^2(M, {\Bbb R})$ be Kaehler classes associated with induced complex structures $I$, $J$, $K$. Consider the Riemann-Hodge pairing $(\cdot,\cdot)_{\omega_i}:\; H^2(M,{\Bbb R})\times H^2(M, {\Bbb R})\longrightarrow {\Bbb R}$, $i=1,2,3$ defined as in \eqref{_Hodge_Riemann_form_general_Equation_} (see also \eqref{_H_R_to_H^2_formula_Equation_}). According to \ref{_pairings_on_H^2_are_equal_Proposition_}, \[ (\cdot,\cdot)_{\omega_1}= (\cdot,\cdot)_{\omega_2} = (\cdot,\cdot)_{\omega_3}. \] Let \[ \inangles{x,y}_{\c H} := (x,y)_{(vol(M))^{-1/n}\cdot x_i} \] where the volume $vol(M)= \int_M x_i^n$ is volume calculated with respect to the Riemannian metric $(\cdot,\cdot)$, and $n=\frac{\dim_{\Bbb R}(M)}{2}$. This pairing is called {\bf the normalized Hodge-Riemann pairing associated with the hyperkaehler structure $\c H$}. According to \eqref{_H_R_to_H^2_formula_Equation_}, the normalized Hodge-Riemann pairing $\inangles{\cdot,\cdot}_{\c H}$ can be expressed by \begin{equation} \label{_normalized_HR_Equation_} (\eta_1,\eta_2)_{\omega}= \lambda^{n-2}\int_X \omega^{n-2}\wedge \eta_1\wedge\eta_2 - \frac{n-2}{(n-1)^2} \lambda^{2n-2}\int_X \omega^{n-1}\eta_1 \cdot \int_X\omega^{n-1}\eta_2 \end{equation} where $\lambda = (vol(M))^{-1/n}$. \hfill The main result of this section is the following theorem: \hfill \theorem \label{_Hodge_Riemann_independent_Theorem_} Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$ be hyperkaehler structures on $M_{C^\infty}$. Then $\inangles{\cdot, \cdot}_{\c H_1} = \inangles{\cdot, \cdot}_{\c H_2}$ In other words, the normalized Hodge-Riemann pairing associated with the point $\c H\in \mbox{\it Hyp}$ does not depend on the choice of $\c H$ in $Hyp$. \hfill {\bf Proof:} The space $\mbox{\it Hyp}$ is endowed with the homogenous action of the group $SO(3)$ as follows. Let $(I, J, K, (\cdot, \cdot))\in \mbox{\it Hyp}$ be a hyperkaehler structure. Consider the complex structures $I$, $J$, $K$ as endomorphisms of the tangent bundle $TM$. We express this by $I, J, K\in \Gamma_M(End(TM))$. Consider the three-dimensional subspace $V\subset \Gamma_M(End (TM))$ generated by $I$, $J$, $K$. By definition of a hyperkaehler structure, $V$ is a three-dimensional vector space equipped with a canonical isomorphism with the space of anti-self-adjoint quaternions. The space $\c V$ of anti-self-adjoint quaternions is a Lie subalgebra of the quaternion algebra. The space of sections $\Gamma_M(End (TM))$ has a canonical algebra structure. By definition, $V$ is a Lie subalgebra of $\Gamma_M(End(TM))$, and the Lie algebra structure on $V$ coincides with that on $\c V\subset \Bbb H$. The Lie algebra $\c V\subset \Bbb H$ is isomorphic to $\goth{so}(3)$. Consider the adjoint action of $SO(3)$ on $V\cong \c V\cong \goth {so}(3)$. Let $A\in SO(3)$. By definition of adjoint action, \begin{equation} \label{_A_of_quate_quate_Equation_} \begin{array}{l} A(I)^2=A(J)^2=A(K)^2, \mbox{ \ and \ } \\ A(I)\circ A(J)=A(K)=-A(J)\circ A(I) \end{array} \end{equation} The operators $I$, $J$, $K$ are parallel with respect to the Levi-Civita connection. The operators $A(I)$, $A(J)$, $A(K)$ are linear combinations of $I$, $J$, $K$ with constant coefficients. Hence, these operators are also parallel. They are orthogonal by trivial reasons. \hfill \claim \label{_Newla_Niere_for_para_Claim_} Let $X$ be a Riemannian manifold equipped with Levi-Civita connection. Let $\c I$ be an orthogonal almost complex structure which is parallel with respect to the connection. Then $\c I$ is an integrable (i. e., defines a complex structure). Moreover, the Riemannian metric on $X$ is Kaehler. {\bf Proof:} This follows from Newlander-Nierenberg theorem. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_Newla_Niere_for_para_Claim_} implies that $A(I)$, $A(J)$, $A(K)$ are operators of complex structure. Now, \eqref{_A_of_quate_quate_Equation_} implies that $(A(I), A(J), A(K), (\cdot, \cdot))$ is a hyperkaehler structure. We obtain an action of $SO(3)$ on $Hyp$. \hfill \definition \label{_action_SO(3)_on_Hyp_Definition_} This action of $SO(3)$ on $\mbox{\it Hyp}$ is called {\bf a standard action of $SO(3)$ on the space of hyperkaehler structures}. Two hyperkaehler structures are called {\bf equivalent} if one can be obtained from another by the standard action of $SO(3)$. \hfill It is easy to describe the action of $SO(3)$ on $\mbox{\it Hyp}$ in terms of the period map: \hfill \claim\label{_action_SO(3)_on_Hyp_via_periods_Lemma_} Let $\c H \in \mbox{\it Hyp}$ and $A\in SO(3)$. Consider $P_{hyp}(\c H)$ and $P_{hyp}(A(\c H))$ as elements of the space \[ W:= H^2(M,{\Bbb R})\otimes{\Bbb R}^3 \cong H^2(M,{\Bbb R})\oplus H^2(M,{\Bbb R})\oplus H^2(M,{\Bbb R}). \] Then $P_{hyp}(A(\c H))$ is obtained from $P_{hyp}(\c H)$ by applying $Id\otimes A$ to $P_{hyp}(\c H)\in H^2(M,{\Bbb R})\otimes{\Bbb R}^3$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \lemma \label{_Hodge_Riemann_independe_for_equiva_hyperkae_Lemma_} In assumptions of \ref{_Hodge_Riemann_independent_Theorem_}, Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$ be equivalent hyperkaehler structures on $M_{C^\infty}$. Then \[ \inangles{\cdot, \cdot}_{\c H_1} = \inangles{\cdot, \cdot}_{\c H_2} \] {\bf Proof:} Follows from \ref{_indu_comple_same_HR_Corollary_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition \label{_admissible_substi_Definition_} Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$. We say that $\c H_2$ {\bf is obtained from $\c H_1$ by an admissible substitution} if either of the following two conditions hold: (i) There exists $\lambda\in {\Bbb R}$ such that $P_3(\c H_1)= \lambda P_3(\c H_2)$. (ii) $\c H_1$ is equivalent to $\c H_2$. We say that $\c H_1$ and $\c H_2$ are {\bf well connected} if $\c H_2$ can be obtained from $\c H_1$ by a sequence of admissible substitutions. Obviously, this relation is an equivalence relation. \hfill \lemma \label{_Hodge_Riemann_independe_for_well_connected_Lemma_} In assumptions of \ref{_Hodge_Riemann_independent_Theorem_}, Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$ be the hyperkaehler structures on $M_{C^\infty}$. Assume that $\c H_1$ and $\c H_2$ are well connected. Then \[ \inangles{\cdot, \cdot}_{\c H_1} = \inangles{\cdot, \cdot}_{\c H_2}. \] {\bf Proof:} It is sufficient to prove \ref{_Hodge_Riemann_independe_for_well_connected_Lemma_} assuming that $\c H_2$ is obtained from $\c H_1$ by admissible substitution. In other words, we may assume that one of conditions (i) and (ii) of \ref{_admissible_substi_Definition_} holds. When (i) holds, \ref{_Hodge_Riemann_independe_for_well_connected_Lemma_} follows from \ref{_Hodge_Riemann_independe_for_equiva_hyperkae_Lemma_}. When (ii) holds, \ref{_Hodge_Riemann_independe_for_well_connected_Lemma_} is a direct consequence of \eqref{_normalized_HR_Equation_} (see also \ref{_Hodge_Riema_general_Claim_} (iv)). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill We obtain that \ref{_Hodge_Riemann_independent_Theorem_} is a consequence of \ref{_Hodge_Riemann_independe_for_well_connected_Lemma_} and the following statement: \hfill \proposition \label{_hyperk_are_well_connected_Proposition_} Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$ be the hyperkaehler structures on $M_{C^\infty}$. Then $\c H_1$ and $\c H_2$ are well connected. {\bf Proof:} \hfill \lemma \label{_hyp.st._w/the_same_I_w/conne_Lemma_} Let $\c H=(I, J, K, (\cdot,\cdot))$ and $\c H'=(I', J', K', (\cdot,\cdot)')$ be two hyperkaehler structures with $I=I'$. Then $\c H$ and $\c H'$ are well connected. {\bf Proof:} Since $I=I'$, we have \[ P_c(\Phi^{hyp}_c(\c H))=P_c(\Phi^{hyp}_c(\c H')). \] By \ref{_forgetting-n-periods_Claim_}, the spaces spanned by $\inangles{P_2(\c H), P_3(\c H)}$ and $\inangles{P_2(\c H'), P_3(\c H')}$ coinside. Denote $\inangles{P_2(\c H), P_3(\c H)}$ by $W$. Let $U$ be the group of linear automorphisms of $W$ which preserve the Hodge-Riemann pairing $(\cdot,\cdot)_{\c H}$ associated with $\c H$. Using the basis $W=\inangles{P_2(\c H), P_3(\c H)}$, we may identify $U$ with $U(1)$. Let $u\in U\cong U(1)$ be represented by the matrix \[ u = \bigg(\begin{array}{rr} \cos(\alpha) & \sin(\alpha) \\ -\sin(\alpha) & \cos(\alpha) \end{array} \bigg). \] Let $u(J)= \cos(\alpha) J+ \sin(\alpha) K$ and $u(K)= \cos(\alpha) K- \sin(\alpha) J$. Checking the definition of the hyperkaehler structure, one obtains that $u(\c H):= (I, u(J), u(K), (\cdot, \cdot))$ is a hyperkaehler structure which is equivalent to $\c H$. By \ref{_action_SO(3)_on_Hyp_via_periods_Lemma_}, $P_2(u(\c H))= u(P_2(\c H))$ and $P_3(u(\c H))= u(P_3(\c H))$. Choosing a suitable $u$, we can make $P_3(u(\c H))$ proportional to $P_3(\c H')$. For such $u$, $u(\c H)$ is well connected to $\c H'$. Since $u(\c H)$ is equivalent to $\c H$, we obtain that $\c H$ is well connected to $\c H'$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \lemma\label{_hyper_w/same_ind_comp_str_well_connect_Corollary_ Let $\c H_1$, $\c H_2\in \mbox{\it Hyp}$ be the hyperkaehler structures, and $I\in Comp$ be the complex structure. Assume that $I$ is induced by $\c H_1$ and $\c H_2$. Then $\c H_1$ is well connected with $\c H_2$. {\bf Proof:} Clearly, \ref{_hyper_w/same_ind_comp_str_well_connect_Corollary_} follows from \ref{_hyp.st._w/the_same_I_w/conne_Lemma_} and the following statement: \hfill \sublemma \label{_induced_compl_str_turn_to_I_Sublemma_} Let $(I, J, K, (\cdot,\cdot))=\c H\in \mbox{\it Hyp}$, $I'\in Comp$ be a complex structure which is induced by $\c H$. Then $\c H$ is equivalent to a hyperkaehler structure $\c H' = (I', J', K', (\cdot,\cdot)')$ for some $J', K', (\cdot,\cdot')$. {\bf Proof:} Consider the action of $SO(3)$ on the space $V:=\inangles{I,J,K}\subset \Gamma(End(TM))$ (see \ref{_action_SO(3)_on_Hyp_Definition_}). Clearly, $I'$, considered as a section of $\Gamma(End(TM))$, belongs to $V$. Take a matrix $A\in SO(3)$ which maps $I\in V$ to $I'$. Then $\c H':=A(\c H)$ satisfies conditions of \ref{_induced_compl_str_turn_to_I_Sublemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition \label{_well_conne_comple_str_Definition_ Let $I_1$, $I_2\in Comp$. The complex structures $I_1$, $I_2$ are called {\bf well connected} if there exist well connected hyperkaehler structures $\c H_1$, $\c H_2$ such that $\c H_1$ induces $I_1$ and $\c H_2$ induces $I_2$. By \ref{_hyper_w/same_ind_comp_str_well_connect_Corollary_}, this is an equivalence relation. \hfill Let $\omega\in H^2(M,{\Bbb R})$. Let $Comp^\omega$ be the set of all $I\in Comp$ such that $\omega$ belongs to the Kaehler cone of $I$. \hfill \claim \label{_Comp^omega_well_conne_Claim_} Let $\omega\in H^2(M,{\Bbb R})$, $I, I'\in Comp^\omega$. Then the complex structures $I$ and $I'$ are well connected. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $\mbox{\it Kah}$ be the set of all $\omega\in H^2(M,{\Bbb R})$ such that $\omega$ is a Kaehler class for some complex structure $I\in Comp$. \hfill \definition \label{_well_conne_kah_str_Definition_ Let $\omega$, $\omega'\in \mbox{\it Kah}$. The classes $\omega$ and $\omega'$ are called {\bf well connected} if there exist well connected hyperkaehler structures $\c H$, $\c H'$ such that $P_1(\c H)= \omega$ and $P_1(\c H')= \omega'$. \hfill \lemma \label{_well_conne_Kah_classe_indu_w/c_hype_Lemma_} Let $\omega$, $\omega'$ be two well connected classes from $\mbox{\it Kah}$. Let $\c H$, $\c H'$ be the hyperkaehler structures such that $P_1(\c H)= \omega$ and $P_1(\c H')= \omega'$. Then $\c H$, $\c H'$ are well connected. {\bf Proof:} Consider the spaces $Comp^\omega$ and $Comp^{\omega'}$. Since $\omega$ is well connected to $\omega'$, there exist well connected hyperkaehler structures \[ \c F=\bigg(A, B, C, (\cdot,\cdot)\bigg),\;\; \; \c F'=\bigg(A', B', C', (\cdot,\cdot)'\bigg) \] such that $P_1(\c F)= \omega$ and $P_1(\c F')= \omega'$. Therefore, $A\in Comp^\omega$ is well connected to $A'\in Comp^{\omega'}$. Let $\c H = (I, J, K, (\cdot,\cdot))$, $\c H'=(I', J', K', (\cdot,\cdot)')$. By definition, $I\in Comp^\omega$ and $I'\in Comp^{\omega'}$. By \ref{_Comp^omega_well_conne_Claim_}, $I$ is well connected to $A$ and $I'$ is well connected to $A'$. Therefore, $I$ is well connected to $I'$. By definition of well connected complex structures, there exist well connected hyperkaehler structures $\c G$ and $\c G'$ such that $\c G$ induces $I$ and $\c G'$ induces $I'$ By \ref{_hyper_w/same_ind_comp_str_well_connect_Corollary_}, $\c G$ is well connected to $\c H$ and $\c G'$ is well connected to $\c H'$. Since the relation of being well connected is transitive, $\c H$ is well connected to $\c H'$. \ref{_well_conne_Kah_classe_indu_w/c_hype_Lemma_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill To finish the proof of \ref{_hyperk_are_well_connected_Proposition_}, it is sufficient to show that for all $x,y \in \mbox{\it Kah}$, the cohomology classes $x$ and $y$ are well connected. By \ref{_symplectic_=>_hyperkaehler_Proposition_}, $\mbox{\it Kah}=P_1(Hyp)$. Since $\mbox{\it Hyp}$ is connected, $\mbox{\it Kah}$ is also connected. Therefore \ref{_hyperk_are_well_connected_Proposition_} is implied by the following lemma: \hfill \lemma \label{_C(omega)_open_in_Kah_Lemma_} Let $\omega\in \mbox{\it Kah}$, $C(\omega)$ be the set of all classes $\omega'\in \mbox{\it Kah}$ which are well connected to $\omega$. Then $C(\omega)$ is open in $\mbox{\it Kah}$. {\bf Proof:} Since $\mbox{\it Kah}\subset H^2(M,{\Bbb R})$, it is sufficient to show that $C(\omega)$ is open in $H^2(M,{\Bbb R})$. Since the relation of being well connected is transitive, it is sufficient to show that $C(\omega)\subset \mbox{\it Kah}$ contains an open neighbourhood of $\omega$ for all $\omega\in Kah$. Let $I\in Comp^\omega$. Let $\c H$ be a hyperkaehler structure associated with $I$ and $\omega$ as in \ref{_symplectic_=>_hyperkaehler_Proposition_}. Let $K(\c H)\subset Kah$ be the set of all $\eta\in H^2(M,{\Bbb R})$ such that the following condition holds. The hyperkaehler structure $\c H$ induces a complex structure $L$ such that $\eta\in K(L)$. As usually, $K(L)$ is the Kaehler cone of $L$. As the following lemma implies, $K(\c H)\subset C(\omega)$. \hfill \sublemma \label{_K(H)_well_conne_to_omega_Sublemma_} Let $\omega'\in K(\c H)$. Then $\omega'$ is well connected to $\omega$. {\bf Proof:} Let $\c H\in Hyp$, $\omega=P_1(\c H)$. Let $L$ be an induced complex structure such that $\omega'\in K(L)$. By \ref{_induced_compl_str_turn_to_I_Sublemma_}, there exist a hyperkaehler structure $\c H'=(L, J, K, (\cdot,\cdot))$ which is equivalent to $\c H$. Then, $\c H$ is well connected to $\c H'$. Let $\c H''$ be the hyperkaehler structure associated with $L$ and $\omega'$ as in \ref{_symplectic_=>_hyperkaehler_Proposition_}. By \ref{_hyper_w/same_ind_comp_str_well_connect_Corollary_}, $\c H'$ and $\c H''$ are well connected. By definition, $P_1(\c H'')=\omega'$. Since $\c H$ is well connected to $\c H''$, $\omega$ is well connected to $\omega'$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $\c H \in Hyp$ and $L$ be an induced complex structure. As usually, we denote the intersection $H^{1,1}(M,L)\cap H^2(M,{\Bbb R})$ by $H^{1,1}_L(M,{\Bbb R})$. Let $\omega\in H^{1,1}_L(M,{\Bbb R})$. The hyperkaehler structure $\c H$ induces a Riemannian metric on $M$. Let $\tilde \omega\in \Lambda^2(M,{\Bbb R})$ be the harmonic form which represents the cohomology class $\omega$. Hodge theory implies that $\tilde\omega$ is a form of Hodge type (1,1) with respect to the complex structure $L$. Under these assumptions, we introduce the following definition. \hfill \definition \label{_positive_classes_Definition_} We say that the cohomology class $\omega$ is {\bf positive} with respect to $(\c H,L)$ if the corresponding harmonic (1,1)-form $\tilde \omega$ is everywhere positively defined. In other words, $\omega\in H^{1,1}_L(M,{\Bbb R})$ is {\bf positive} if the symmetric form \[ S_p: \; T_pM\times T_pM \longrightarrow {\Bbb R},\;\; S_p(x,y):= \tilde \omega(x, L(y)) \] is positively defined in every point of $p\in M$. We denote by $K_{\c H}(L)$ the set of all $\omega\in H^{1,1}_L(M,{\Bbb R})$ such that $\omega$ is positive with respect to $(\c H,L)$. \hfill \claim \label{_positive_form_is_Kaehler_Claim_} In assumptions of \ref{_positive_classes_Definition_}, let $\omega\in H^{1,1}_L(M,{\Bbb R})$ be the two-form which is positive with respect to $(\c H,L)$. Then $\omega$ is a Kaehler class: $\omega\in K(L)$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $\c H\in Hyp$. Denote by $K_c(\c H)$ the set of all $\omega\in H^2(M,{\Bbb R})$ such that there exists an induced complex structure $L$ and $\omega\in K_{\c H}(L)$. By \ref{_positive_form_is_Kaehler_Claim_}, $K_c(\c H)\subset K(\c H)$. Let $\omega\in C(\omega)$. This means that $\omega=P_1(\c H)$ for some $\c H\in Hyp$. Since $K_c(\c H)\subset K(\c H)\subset C(\omega)$, to prove that $C(\omega)$ is open in $H^2(M,{\Bbb R})$ it is sufficient to show that $K_c(\c H)$ contains an open neighbourhood of $\omega$ (we use here the transitiveness of well-connectedness). Therefor, \ref{_C(omega)_open_in_Kah_Lemma_} is a consequence of the following statement: \hfill \proposition \label{_K(H)_open_in_H^2(M,R)_Sublemma_} Let $M$ be a hyperkaehler manifold with the hyperkaehler structure $\c H$. Then the set $K_c(\c H)\subset H^2(M,{\Bbb R})$ contains an open neighbourhood of $P_1(\c H)$. {\bf Proof:} Consider the action of the group of unit quaternions $G_M\cong SU(2)$ defined as in Section \ref{hyperk_manif_Section_}. The action of $G_M$ is defined on the tangent bundle $T(M)$. We naturally extend this action to the tensor powers of $T(M)$, including $End(T(M))\cong T(M)\otimes T^*(M)$. Consider the set $R$ of induced complex structures as subset of the space of sections $\Gamma_M(End(TM))$. An easy local computation shows that $G_M$ acts transitively on $R\cong S^2$ (see also \ref{_induced_compl_str_turn_to_I_Sublemma_}). Let $L$ be an induced complex structure, $\omega\in K_{\c H}(L)$. Let $g\in G_M$, $L':=g(L)$. Consider the Kaehler form $\omega$ as the section of $\Lambda^2(TM)\subset T^*(M)\otimes T^*(M)$. Obviously, the 2-form \[ \inbfpare{\cdot,\cdot}:= g(\omega)(L'(\cdot),\cdot)= \omega(g \circ g^{-1}\circ L\circ g (\cdot),g(\cdot)) = \omega (L(g(\cdot)),g(\cdot)) \] is symmetric and positively defined. To show that the Riemannian form $\inbfpare{\cdot,\cdot}$ is Kaehler, we have to prove that the form \[ \inbfpare{L'(\cdot),\cdot} = - g(\omega)(\cdot,\cdot) \] is symplectic. Since $G_M$ commutes with Laplacian, it maps harmonic forms to harmonic ones. Hence, $g(\omega)$ is a symplectic form. Therefore $\inbfpare{\cdot,\cdot}$ is a Kaehler metric. This implies that $g(\omega)\in K(L')$. We proved the following statement: \hfill \claim \label{_G_M_acts_transi_on_K(H)_Claim_} Let $M$ be a hyperkaehler manifold with the hyperkaehler structure $\c H$. Consider the action of the group of unit quaternions $G_M\cong SU(2)$ on $H^2(M,{\Bbb R})$ (see \ref{_there_is_action_of_G_M_Proposition_}). Let $g\in G_M$, $L, L'\in R$, $L'=g(L)$. Then $g:\; H^2(M, {\Bbb R})\longrightarrow H^2(M,{\Bbb R})$ induces an isomorphism from $K_{\c H}(L)\subset H^2(M,{\Bbb R})$ to $K_{\c H}(L')\subset H^2(M,{\Bbb R})$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill The following statement is clear: \hfill \claim \label{_K_c(L)_is_open_Claim_} In assumptions of \ref{_positive_classes_Definition_}, the set $K_{\c H}(L)$ is open in $H^{1,1}_L(M,{\Bbb R})$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_G_M_acts_transi_on_K(H)_Claim_} establishes a smooth map $\delta:\; G_M\times K_{\c H}(I)\longrightarrow K_c(\c H)$. Let $b_2:=dim(H^2(M))$. By \ref{_K_c(L)_is_open_Claim_}, $dim_{\Bbb R} K_{\c H}(I)= dim_{\Bbb C} (H^{11}((M, I))$. Since $M$ is a simple hyperkaehler manifold, $dim_{\Bbb C} (H^{11}((M, I)) = b_2-2$. According to Section \ref{hyperk_manif_Section_}, $R\cong S^2$. Therefore, $dim_{\Bbb R}(R\times K_{\c H}(I))= b_2$. By definition, $G_M$ is identified with the group of unit quaternions, and $R$ is identified with the set $x\in G_M \ \ | \ \ x^2=-1$. This identification defines a canonical embedding $R\hookrightarrow G_M$. Let $\phi:\; R\times K_{\c H}(I)\longrightarrow K_c(\c H)$ be the restriction of $\delta$ to $R\times K_{\c H}(I)\subset G_M\times K_{\c H}(I)$. The dimension of $R\times K_{\c H}(I)$ is equal to $b_2$. Therefore, to prove that $K_c(\c H)$ is open in $H^2(M,{\Bbb R})$ it is sufficient to prove the following: \hfill \sublemma \label{_K(H)cong_S^2_times_K(I)_Sublemma_} The map $\phi:\; R\times K_{\c H}(I)\longrightarrow K_c(\c H)$ is a diffeomorphism. {\bf Proof:} As in \ref{_restrictions_of_pairings_to_H^2_Lemma_}, consider the decomposition $H^2(M,{\Bbb R})= H_{inv}\oplus V$. As we have established previously, $V$ is generated by $P_i(\c H)$, $i=1,2,3$ and $(H_{inv},V)_{\c H}=0$. For $x\in H^2(M, {\Bbb R})$, let $\pi_i(x)$ be the orthogonal projection of $x$ to $H_{inv}$ and $\pi_v(x)$ be the orthogonal projection of $x$ to $V$. The bilinear form $(\cdot,\cdot)_{\c H}$ is $G_M$-invariant by \ref{_Hodge_Riemann_independe_for_equiva_hyperkae_Lemma_}. Therefore, $\pi_i(g(x))=g(\pi_i(x))$. For every induced complex structure $L$, $\c H$ defines a Kaehler structure on the complex manifold $(M, L)$. Hence, for every induced complex structure $L$, the hyperkaehler structure $\c H$ defines a Kaehler form $\omega_L$ and a degree map $deg_L:\; H^{2i}(M,{\Bbb R})\longrightarrow {\Bbb R}$. According to \ref{_inv_2-forms_have_zero_degree_Claim_}, for all $x\in H_{inv}$ and all induced complex structures $L$, $deg_L(x)=0$. Therefore for all $x\in K_c(\c H)$, we have $\pi_v(x)\neq 0$. Let $y\in K_c(\c H)$. Let $l(y)$ be the line in the three-dimensional space $V$ generated by $\pi_v(y)$. The space $V$ is generated by the set of induced complex structures, which constitute a unit sphere in $V$. Hence, the space of lines in $V$ is canonically identified with the set of complex structures up to a sign. Let $R^{\pm}\cong {\Bbb R} P^2$ be the quotient of $R$ by $\pm 1$. Let $\theta: \; K_c(\c H)\longrightarrow R^{\pm}$ map $y$ to the point of $R^{\pm}$ which corresponds to $l(y)$. Denote the induced complex structures which correspond to $\theta(y)$ by $L_1$, $L_2$, where $L_2=-L_1$. Denote the Hodge decomposition associated with an arbitrary complex structure $L\in Comp$ by $H^{pq}_L$. According to \ref{_G_M_invariant_forms_Proposition_}, $x\in H^{pp}_L$ if and only if $L(x)=x$. Obviously, $L(y)=\pi_i(y)+ L(\pi_v(y))$. Realizing $L$ and $\pi_v(y)$ as quaternions in a usual way, we may check that $L(\pi_v(y))= L\pi_v(y)L^{-1}$. Since the centralizator of all elements in $SU(2)$ is one-dimensional, $L\in l$ whenever $L(y)=y$. Therefore for $y\in H^{11}_L$, we have $L=\pm L_1$. Since $\pi_v$ is an orthogonal projection, \begin{equation} \label{_pi_v(y)_Equation_} \pi_v(y)=\frac{deg_L(y)}{deg_L(\omega_L)}\omega_L, \end{equation} where $\omega_L$ is the Kaehler form of $(M,L)$ considered as an element of $V$, where \[ deg_L(y):= \int_M \omega_L^{n-1}\wedge y. \] By definition, $deg_{L_1}(x)=-deg_{L_2}(x)$. On the other hand, for $x\in K(L)$, we have $deg_L(x)>0$. Therefore for all $y\in K_c(\c H)$, there exist only one induced complex structure $L$ such that $y\in K(L)$. This implies that $\phi$ is a monomorphism. We need to construct the inverse of $\phi$ and prove that it is smooth. Let $\Theta^{\pm}: K_c(\c H) \longrightarrow R^{\pm}\times (H_{inv}\oplus {\Bbb R})$ map $y\in K_c(\c H)$ to the pair \[ \bigg(\theta (y),\; \pi_i(y)\oplus |deg_{\pi_i(y)}(y)|\bigg) \] where $deg_{\pi_i(y)}$ is well defined up to a sign. According to \eqref{_pi_v(y)_Equation_}, up to a sign, one can reconstruct $\pi_v(y)$ by $\Theta^{\pm}(y)$. Therefore, $\Theta^{\pm}$ is a double covering. Let $\rho:\;R^{\pm}\times (H_{inv}\oplus {\Bbb R})\longrightarrow H^2(M,{\Bbb R})$ map $(s,h+t)\in R^{\pm}\times(H_{inv}\oplus {\Bbb R})$ to \[ \frac{t}{deg_I(\omega_I)}\omega_I +h. \] The $I$-degree of $\omega\in K_{\c H}(I)$ is positive. Therefore \eqref{_pi_v(y)_Equation_} implies that the map \[ \phi\circ \Theta^{\pm} \circ \rho:\; R\times K_c(\c H)\longrightarrow R^{\pm}\times K_c(\c H) \] acts as identity on $K_c(\c H)$ and acts as a double covering on $R$. Since $\Theta^{\pm}$ is a double covering, $\phi$ is an open embedding. \ref{_K(H)cong_S^2_times_K(I)_Sublemma_} and \ref{_C(omega)_open_in_Kah_Lemma_} is proven. The proof of \ref{_hyperk_are_well_connected_Proposition_} and consequently \ref{_Hodge_Riemann_independent_Theorem_} is finished. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{Period map and the space of 2-dimensional planes in $H^2(M,\protect {\Bbb R})$.} \label{_Q_c_defini_Section_} There is an alternative way of looking at Griffiths period map $P_c:\; Comp\longrightarrow {\Bbb P}(H^2(M,{\Bbb C}))$. This enhanced version of period map is a map from $Comp$ to an open subset in Grassmanian of all 2-dimensional planes in $H^2(M, {\Bbb R})$. To define this map, we remind the reader certain well-known results from linear algebra. Let $V_{\Bbb R}$ be an ${\Bbb R}$-linear space endowed with the non-degenerate symmetric bilinear form $(\cdot,\cdot)_{\Bbb R}$. Let $V_{\Bbb C}:= V_{\Bbb R}\otimes{\Bbb C}$ be the complexification of $V_{\Bbb R}$, and $(\cdot,\cdot)_{\Bbb C}$ be the ${\Bbb C}$-linear form on $V_{\Bbb C}$ obtained as a complexification of $(\cdot,\cdot)_{\Bbb R}$. In applications, $V_{\Bbb R}= H^2(M,{\Bbb R})$, $V_{\Bbb C}= H^2(M,{\Bbb C})$, and $(\cdot,\cdot)_{{\Bbb R}}$ is the normalized Hodge-Riemann pairing $(\cdot,\cdot)_{\c H}$. \hfill Consider the projectivization ${\Bbb P} V_{\Bbb C}$ as a space of lines in $V_{\Bbb C}$. For all $x\in V_{\Bbb C}$, let $\bar x$ denote the complex conjugate to $x$. Let \[ C:= \{ t\in V_{\Bbb C} \;\; |\;\; \forall x\in t,\, (x,x)_{\Bbb C}=0, (x,\bar x)_{\Bbb C}> 0 \}. \] Let $\tilde Pl$ be the space of all oriented 2-dimensional linear subspaces in $V_{\Bbb R}$. Let $Pl\subset \tilde Pl$ be the set of all $L\in \tilde Pl$ such that the restriction of $(\cdot,\cdot)_{\Bbb R}$ to the 2-dimensional space $L\subset V_R$ is positively defined. Clearly, $Pl$ is open in $\tilde Pl$. For $t\in {\Bbb P} V_{\Bbb C}$, take $x\in t$, $x\neq 0$. Let $i_x(t)\subset V_{\Bbb R}$ be the linear span of $Re(x)$, $Im(x)\in V_{\Bbb R}$. If $i_x(t)$ is two-dimensional, we consider $i_x(t)$ as the oriented space with the orientation defined by the basis $(Re(x), Im(x))$. \hfill \proposition \label{_from_PV_C_to_Pl_linear-alg_Proposition_} Let $t\in C\subset {\Bbb P} V_{\Bbb C}$. Then the space $i_x(t)$ is 2-dimensional and independent on the choice of $x\in t$. Let $i:\; C\longrightarrow \tilde Pl$ map $t\in C$ to $i_x(t)$. The image of $i:\; C\longrightarrow \tilde Pl$ coinsides with $Pl\subset \tilde Pl$. Established this way map $i:\; C\longrightarrow Pl$ is bijective. {\bf Proof:} Let $t\in C$, $x\in t\subset V_{\Bbb C}$. Let $y=Re(x), \;z= Im(x)$. Then \[ (x,x)_{\Bbb C}= (y,y)_{\Bbb R}-(z,z)_{\Bbb R} + 2\sqrt{-1}\: (y,z)_{\Bbb R}= 0. \] Therefore $(y,y)_{\Bbb R}=(z,z)_{\Bbb R}$ and $(y,z)_{\Bbb R}=0$. On the other hand, \[ (x,\bar x)_{\Bbb C} = (y,y)_{\Bbb R}+ (z,z)_{\Bbb R}> 0. \] We obtain that \begin{equation} \label{_y^2=z^2>0_Equation_} (y,y)_{\Bbb R} = (z,z)_{\Bbb R} > 0 \;\; \mbox{and}\;\; (y,z)_{\Bbb R}=0 \end{equation} This implies that the vectors $y$ and $z$ are linearly independent. For all $c=\lambda e^{\sqrt{-1}\: \alpha}\in {\Bbb C}$, where $\lambda,\alpha\in {\Bbb R}$, we have \[ Re(cx)=\lambda \cos (\alpha) y +\lambda \sin(\alpha) z, \; Im(cx)=-\lambda \sin (\alpha) y +\lambda \cos(\alpha) z. \] Therefore, $i_x(t)=i_{cx}(t)$. This implies that the map $i:\; C\longrightarrow \tilde Pl$ is well defined. According to \eqref{_y^2=z^2>0_Equation_}, $i(C)\subset Pl$. \hfill Let us construct the inverse map $j:\; Pl\longrightarrow C$. For $L\in Pl$, take an oriented orthonormal basis $(y,z)$ in the Euclidean space $L$. For another orthonormal basis $(y',z')$ in $L$, we have \begin{equation} \label{_rotation_on_y,z_Equation_} \begin{array}{rrrrr} y'& =& \cos (\alpha) y &+&\sin(\alpha) z \\[3mm] z'& = & -\sin (\alpha) y &+&\cos(\alpha) z \end{array} \end{equation} for some $\alpha\in {\Bbb R}$. Let $j(L)\in {\Bbb P}V_{\Bbb C}$ be the line generated by $y+ \sqrt{-1}\: z\in V_{\Bbb C}$. The equations \eqref{_rotation_on_y,z_Equation_} imply that $j(L)$ is independent of the choice of the oriented orthonormal basis $(y,z)$. Since $(y,z)$ is orthonormal basis, \eqref{_y^2=z^2>0_Equation_} holds. This equation immediately implies that $j(L)\in C\subset {\Bbb P}V_{\Bbb C}$. Finally, it is clear from defintions that $j\circ i= Id$ and $i\circ j= Id$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Returning to the hyperkaehler manifolds, consider the space \[ \goth C\subset {\Bbb P}(H^2(M,{\Bbb C})) \] consisting of lines $l\in H^2(M,{\Bbb C})$ such that for all $x\in l$, $(x,x)_{\c H}=0$ and $(x,\bar x)_{\c H}>0$. Here, as elsewhere, \[ (\cdot,\cdot)_{\c H}:\; H^2(M,{\Bbb C})\times H^2(M,{\Bbb C})\longrightarrow {\Bbb C} \] is a complexification of the normalized Hodge-Riemann pairing. Hodge-Riemann relations imply that $P_c(Comp)\subset \goth C$. \ref{_from_PV_C_to_Pl_linear-alg_Proposition_} establishes a diffeomorphism $\goth i:\; \goth C\longrightarrow Pl$, where $Pl$ is the space of 2-dimensional subspaces $L\subset H^2(M,{\Bbb R})$ such that $(\cdot,\cdot)_{\c H}$ is positively defined on $L$. Let $Q_c:\; Comp\longrightarrow Pl$ be the composition of $P_c$ and $\goth i$. As results of Bogomolov and Todorov imply (see \cite{_Bogomolov_}, \cite{_Beauville_}, \cite{_Todorov_}), the map $Q_c$ is an immersion. Since $\dim Comp = \dim \goth C= \dim H^2(M) -2$, this map is etale. It can be described in more straightforward terms as follows. Let $I\in Comp$. Let $\tilde \Omega$ be a holomorphic symplectic form over $(M, I)$. Let $\Omega\in H^2(M, {\Bbb C})$ be the cohomology class represented by the closed differential form $\tilde \Omega$. Let $\omega_2:= Re(\Omega)$, $\omega_3:= Im(\Omega)$. Then one can define $Q_c(I)$ as the linear span of $\omega_2$, $\omega_3$. \section {Lefschetz-Frobenius algebras.} \label{_Lefshe_Frob_Section_} In this section, we give a number of preliminary definitions, which eventually lead to a calculation of the cohomology of a compact hyperkaehler manifold. Some of these definitions are due to V. Lunts (see \cite{_Lunts-Loo_}). Further on, by ``algebra'' we understand an associative algebra with unit. \hfill \definition Let $A$ be an algebra over a field $k$ and $(\cdot,\cdot)$ be a $k$-valued bilinear form on $A$. The form $(\cdot,\cdot)$ is called {\bf invariant} if for all $a,b,c\in A$, $(ab,c)=(a,bc)$. A $k$-algebra equipped with an invariant non-degenerate bilinear form is called {\bf Frobenius algebra}. \hfill \definition Let $A=\oplus A_i, i=0,...,d$ be a graded supercommutative algebra over a field $k$ of characteristic zero. Assume that $A$ is equipped with an invariant bilinear form, such that for all $a\in A_n, b\in A_m$, the following holds: \begin{equation} \label{_graded_scalar_pro_Equation_} \begin{array}{l} (a,b) = (-1)^{nm}(b,a), \mbox{\ and} \ \ \ \ \ \ \ \ \ \ \ \ \\[2mm] (a,b)=0 \mbox{\ for\ } n+m\neq d. \ \ \ \ \ \ \ \ \ \ \ \ \\[2mm] \end{array} \end{equation} Then $A$ is called {\bf a graded Frobenius algebra of degree $d$}. The prototypical example of graded Frobenius algebras is the cohomology algebra of a compact manifold. \hfill Further on, we consider only the graded Frobenius algebras. For brevity, we sometimes omit the word ``graded''. Let $A=\oplus A_i, i=0,...,d$ be a graded Frobenius algebra of degree $d$ over the field $k$. Let $End_k(A)$ be the space of all $k$-linear endomorphisms of $A$. Let $H\in End_k(A)$ be the endomorphism which maps $a\in A_i$ to $(2d-i)\cdot a$. This endomorphism is introduced by Hodge in his study of harmonic forms and Lefschetz isomorphism. One can check that $H$ is a derivative of an algebra $A$: \[ H(ab)= H(a) b + a H(b). \] For all $A\in A$, let $L_a:\; A\longrightarrow A$ map $b\in A$ to $ab$. \hfill \definition \label{_Lefschetz_triple_Definition_} Let $A=\oplus A_i$ be a graded Frobenius algebra, $a\in A_2$. Let $L_a$, $H\in End_k(A)$ be as above. The triple of endomorphisms $L_a$, $H$, $\Lambda_a\in End_k(A)$ is called {\bf a Lefshets triple} if \[ [H, L_a] = 2L_a, [H,\Lambda_a] = -2 \Lambda_a, [L_a,\Lambda_a] =H. \] Clearly, Lefschetz triples correspond to some representations of the Lie algebra $\goth{sl}(2)$ in $End_k(A)$. Lefschetz theorem (\cite{_Griffiths_Harris_}) gives examples of Lefschetz triples for $A=H^*(M)$ and $M$ is a Kaehler manifold. \hfill \proposition \label{_Lefshe_tri_unique_Proposition_} Let $A$ be a graded Frobenius algebra, $a\in A_2$. Let $(L_a, H, \Lambda_a)$, $(L_a, H, \Lambda'_a)$ be two Lefschetz triples. Then $\Lambda_a=\Lambda'_a$. In other words, $\Lambda_a$ is uniquely determined by $a$. {\bf Proof:} (V. Lunts) Consider the representations $\rho$, $\rho'$ of the Lie algebra $\goth{sl}(2)$ associated with these triples. Take a basis $(x,y,h)$ in $\goth{sl}(2)$, \[ [h,x]=2x, [h,y] =-2y, [x,y]=h, \] such that $\rho(x) =\rho'(x) = L_a$, $\rho(h)= \rho'(h)= H$, $\rho(y)=\Lambda_a$, $\rho'(y)=\Lambda'_a$. Let $T:= \Lambda_a-\Lambda'_a$. Consider the adjoint action of $\goth{sl}(2)$ on the space $End(A)$ obtained from $\rho$: \[ ad (\rho):\; \goth{sl}(2)\longrightarrow End(End(A)) \] Clearly, $ad (\rho)(x)(T)=[L_a,\Lambda_a]-[L_a,\Lambda'_a]=0$. Therefore, $T$ is a highest vector is an $\goth{sl}(2)$-submodule of $End(A)$ generated by $T$ and $ad (\rho)$. On the other hand, $ad (\rho)(h)(T)=-2T$, and therefore, the weight of $T$ is $-2$. This is impossible because $ad (\rho)$ is a finite dimensional representation of $\goth {sl}(2)$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition Let $A=\oplus A_i$ be a graded Frobenius algebra, $a\in A_2$. Then $a$ is called {\bf of Lefschetz type} if a Lefschetz triple $(L_a, H, \Lambda_a)$ exists. \hfill \lemma \label{_a_Lefshe_if_a^i_iso_Lemma_} Let $A=A_0\oplus A_1\oplus ... A_{2d}$ be a graded space. Let $L\in End(A)$ be an endomorphism of grading 2: $L:\; A_i\longrightarrow A_{i+2}$. Let $H$ act on $A_i$ as the multiplication by $d-i$, $i=0,1, ... , 2d$. Then the following conditions are equivalent: \hfill (i) There exist an endomoprhism $\Lambda\in End(A)$ of grading -2 such that the relations \[ [H,\Lambda] =-2 \Lambda, [H,L] =-2 L, [L,\Lambda] = H \] hold\footnote{The first two of these relations hold trivially because of the grading.}. \hfill (ii) For all $i=0,1,...,d-1$, the map \[ L^{d-i}:\; A^i\longrightarrow A^2d-i, \] is an isomorphism. \hfill {\bf Proof:} Clear (see \cite{_Lunts-Loo_} for details). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill When $A$ is a cohomology algebra of a compact Kaehler manifold $M$, all Kaehler classes are obviously of Lefschetz type. On the other hand, a class of Lefschetz type is not necessarily a Kaehler class. For example, for a Kaehler class $\omega$, the class $-\omega$ is of Lefschetz type, but $-\omega$ cannot be a Kaehler class by trivial reasons. \hfill \definition \label{_Lefschetz_Frob_alge_Definition_} Let $A$ be a graded Frobenius algebra. Let $S\subset A_2$ be the set of all elements of Lefschetz type. The algebra $A$ is called {\bf a Lefschetz-Frobenius algebra} if the following conditions hold: (i) The space $A_0$ is one-dimensional over $k$. (ii) The set $S$ is Zariski dense in $A_2$. \hfill {\bf Example}: Let $M$ be a compact Kaehler manifold. Then the algebras $H^*(M)$ and $\oplus H^{p,p}(M)$ are Lefschetz-Frobenius, as \ref{_Lefshe_Frob_if_a_Lefshe_ele_exists_Proposition_} implies. \hfill By \ref{_a_Lefshe_if_a^i_iso_Lemma_}, the set $S$ of all elements of Lefschetz type is given by an open condition. Therefore $S$ is open in $A_2$. Therefore, $A$ is a Frobenius-Lefschetz algebra if and only if $S$ is non-empty in $A_2$. We obtained the following statement: \hfill \proposition \label{_Lefshe_Frob_if_a_Lefshe_ele_exists_Proposition_} Let $A$ be a graded Frobenius algebra. Assume that $A_2$ contains at least one element of Lefschetz type. Then $A$ is Lefschetz-Frobenius. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill For every Lefschetz triple $T$, we define a Lie algebra homomorphism \[ \rho_T:\; \goth{sl}(2) \longrightarrow End_k(A) \] in an obvious way. For a Lefschetz-Frobenius algebra $A$, let ${\goth g}(A)$ be the Lie subalgebra of $End_k(A)$ generated by the images of $\rho_T$ for all Lefschetz triples $T$. This algebra is our main object of study. The algebra ${\goth g}(A)$ is graded: ${\goth g}(A)= \oplus{\goth g}_{2i}(A)$, $g(A_n)\subset A_{n+2i}$ for all $g\in {\goth g}_{2i}(A)$. This algebra is called {\bf the structure Lie algebra of $A$}. \hfill \definition \label{_Lefshe_Fro_Definition_} Let $A$ be a Lefschetz-Frobenius algebra. Assume that ${\goth g}_{2i}(A)=0$ for $i\neq -1,0,1$: \[ {\goth g}(A)={\goth g}_{-2}(A)\oplus {\goth g}_0(A)\oplus {\goth g}_2(A). \] Then $A$ is called {\bf a Lefschetz-Frobenius algebra of Jordan type}. Such $A$ are closely related with Jordan algebras (\cite{_Springer_}). \hfill If $A$ is generated by $A_2$ and $A_0\cong k$, $A$ is called {\bf reduced}. The subalgebra $A^r\subset A$ generated by $A_2$ and $A_0$ is called {\bf reduction of $A$}. We use the following result. \hfill \proposition \label{_Lunts_about_FLJ_Proposition_} (\cite{_Lunts-Loo_}) Let $A$ be a Frobenius-Lefschetz algebra, ${\goth g}=\oplus {\goth g}_{2i}$ be its structure Lie algebra. Then the following conditions are equivalent: (i) ${\goth g}_2$ is spanned by $L_a$ for all Lefschetz elements $a$, (ii) $A$ is of Jordan type, (iii) $[\Lambda_a,\Lambda_b]=0$ for all Lefschetz elements $a,b\in A_2$. {\bf Proof:} See Proposition 2.6 and Claim 2.6.1 of \cite{_Lunts-Loo_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_Lunts_about_FLJ_Proposition_} immediately implies the following statement: \hfill \corollary \label{_g_2_is_A_2_Corollary_} Let $A=\oplus A_i$ be a Lefschetz-Frobenius algebra of Jordan type, ${\goth g}={\goth g}_{-2}\oplus {\goth g}_0\oplus {\goth g}_2$ be its sturtcure Lie algebra. Then ${\goth g}_2=A_2$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill In the future, we often assume that $A$ is reduced. In this case, the multiplicative structure on $A$ can be recovered from the ${\goth g}(A)$-action. This is done as follows. Since $[{\goth g}_2(A),{\goth g}_2(A)]\subset {\goth g}_4(A) =0$, the space ${\goth g}_2(A)\subset {\goth g}(A)$ is a commutative subalgebra of ${\goth g}(A)$. Consider the corresponding embedding of enveloping algebras: \[ U_{{\goth g}_2(A)}\cong S^*({\goth g}_2(A))\hookrightarrow U_{{\goth g}(A)}. \] Let ${\Bbb I}\in A_0$ be the unit. The representation ${\goth g}(A)\longrightarrow End(A)$ induces the canonical map \[ U_{{\goth g}(A)} \stackrel{\tilde p}\longrightarrow A, \; P\longrightarrow P({\Bbb I}), \] where $P\in U_{{\goth g}(A)}$ is an ``polynomial'' over ${\goth g}(A)$. Consider the restriction of $\tilde p$ to $U_{{\goth g}_2(A)}\subset U_{{\goth g}(A)}$: \begin{equation}\label{_p_from_U_g_2_to_A_Equation_} p:\; U_{{\goth g}_2(A)} \longrightarrow A. \end{equation} Clearly, for all $a\in A_2$, $a$ of Lefschetz type, $L_a\in {\goth g}_2(A)$. Since the set of elements of Lefschetz type is Zariski dense in $A_2$, we have $L_a\in {\goth g}(A)\subset End(A)$ for all $a\in A_2$. One can easily check that the corresponding map $i:\; A_2\longrightarrow {\goth g}_2(A)$ is an isomorphism (see \ref{_Lunts_about_FLJ_Proposition_}). Therefore, $S^*({\goth g}_2(A))\cong S^*(A_2)$. Applying the isomorphism $U_{{\goth g}_2(A)}\cong S^*({\goth g}_2(A))\cong S^*(A_2)$ to the map \eqref{_p_from_U_g_2_to_A_Equation_}, we obtain the map \[ p':\; S^* A_2\longrightarrow A. \] \hfill \claim \label{_p_is_multiplication_Claim_} The map $p'$ coinsides with the map induced by multiplication in $A$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_p_is_multiplication_Claim_} implies that the kernel of the map $p:\; U_{{\goth g}_2(A)} \longrightarrow A$ is an ideal in $U_{{\goth g}_2(A)}$. This leads to a more general construction. \hfill \definition \label{_multi_associ_with_representa_Definition_} Let ${\goth g}={\goth g}_{-2}\oplus {\goth g}_0\oplus {\goth g}_2$ be a graded Lie algebra. Let $V$ be a representation of ${\goth g}$ and $v\in V$ be a vector. Assume that by applying ${\goth g}_2$ to $v$ repeatedly we obtain the whole space $V$ (i. e., the vector $v$ generates $V$ as a representation of ${\goth g}_2$). Assume that ${\goth g}_{-2}(v)=0$, and that for all $g\in {\goth g}_0$, $g(v)$ is proportional to $v$. Let $p:\; U_{{\goth g}_2}\longrightarrow V$ be the map which associates with the polynomial $P\in U_{{\goth g}_2}$ the vector $P(v)\in V$. Clearly, $\ker(p)$ is a left ideal in $U_{{\goth g}_2}$. Since ${\goth g}_2$ is commutative, this ideal is two-sided. Therefore, $V\cong \bigg(U_{{\goth g}_2}/ \ker (p)\bigg)$ is equipped with a structure of commutative algebra. We denote this algebra by $V_{{\goth g}, v}$. \hfill \claim \label{_g_structure_defines_algebr_Claim_ Let $A$ be a Lefschetz-Frobenius algebra of Jordan type. Assume that $A$ is reduced (generated by $A_2$). Consider $A$ as a representation of ${\goth g}(A)\cong {\goth g}_{-2}(A)\oplus {\goth g}_0(A)\oplus {\goth g}_2(A)$. Take the unity vector ${\Bbb I}\in A_0$. Then the algebra $A_{{\goth g}(A),{\Bbb I}}$ coinsides with $A$. {\bf Proof:} Follows from definitions. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \hfill {\large\bf Appendix. Reduction and the structure Lie algebra.} \hfill Let $A$ be a Lefschetz-Frobenius algebra, and $A^r$ be its reduction. Assume that the restriction $(\cdot,\cdot)_r$ of $(\cdot,\cdot)$ to $A^r$ is non-degenerate. Then $(\cdot,\cdot)_r$ establishes a structure of Frobenius algebra on $A^r$. We are going to show that $A^r$ is Lefschetz-Frobenius, and relate ${\goth g}(A)$ to ${\goth g}(A^r)$. \hfill \proposition \label {_A^r_Lef-Frob,_pres_by_g(A)_Proposition_} Let $A$ be a Lefschetz-Frobenius algebra. Assume that the restriction of $(\cdot,\cdot)$ to $A^r$ is non-degenerate. Then $A^r$ is also Lefschetz-Frobenius. Moreover, the action of ${\goth g}(A)$ on $A$ preserves the subspace $A^r\subset A$. {\bf Proof:} Let $A^r_\bot$ be the orthogonal complement to $A^r$ in $A$. Since $(\cdot,\cdot)\restrict{A^r}$ is nondegenerate, $A=A^r\oplus A^r_\bot$. \hfill \lemma \label{_A^r^bot_preserved_by_mult_by_A^r_Lemma_} Let $a\in A^r$, $b\in A^r_\bot$. Then $ab\in A^r_\bot$. {\bf Proof:} It is sufficient to show that for all $c\in A^r$, $(ab,c)=0$. Since $(\cdot,\cdot)$ is invariant, for all $c\in A$ we have $(ab,c)=(b, ac)$. Since $ac\in A^r$ and $b\in A^r_\bot$, $(ab,c)=0$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill {}From \ref{_A^r^bot_preserved_by_mult_by_A^r_Lemma_}, we obtain that the operators $L_a$, $a\in A_2$ preserve the decomposition $A=A^r\oplus A^r_\bot$. By \ref{_a_Lefshe_if_a^i_iso_Lemma_}, the map $L_a^{d-i}:\; A_i\longrightarrow A_{2d-i}$ is an isomorphism. Therefore, the restriction of $L_a^{d-i}$ to the $i$-th grading component $(A^r)_i$ of $A^r$ is an embedding to $(A^r)_{2d-i}$. Since $A^r$ is Frobenius, $dim_k(A^r)_{2d-i}=dim_k(A^r)_i$. Therefore, the restriction of $L_a^{d-i}$ to $(A^r)_i$ is an isomporphism. Applying \ref{_a_Lefshe_if_a^i_iso_Lemma_} again, we obtain that for all Lefschetz-type elements $a\in A_2$, these elements are of Lefschetz type in $A^r$. Therefore, $A^r$ is a Lefschetz-Frobenius algebra. It remains to prove that $A^r$ is preserved by ${\goth g}(A)$. Clearly, the generators $L_a$ and $H$ of ${\goth g}(A)$ preserve $A^r$. Therefore, to show that ${\goth g}(A)$ preserves $A^r$ it is sufficient to prove that $\Lambda_a$ preserve $A^r$ for all Lefschetz-type elements $a\in A_2$. Let $L=L_a$. Let $L_r$, $H_r$ be the restrictions of $L$, $H$ to $A^r$ and $L_{\bot}$ be the restrictions of $L$, $H$ to $A^r_\bot$. By \ref{_a_Lefshe_if_a^i_iso_Lemma_}, there exists an endomorphism $\Lambda_\bot:\; A^r_\bot\longrightarrow A^r_\bot$ of grading $-2$ such that $[L_\bot,\Lambda_\bot]= H_\bot$. Let $\Lambda_r:\; A^r\longrightarrow A^r$ be the endomorphism of $A^r$ such that $(L_r, H_r, \Lambda_r)$ is a Lefschetz triple. Let $\Lambda_r+\Lambda_\bot$ be an endomorphism of $A$ such that for all $a=b+c$, $b\in A^r, c\in A_\bot^r$, \[ \Lambda_r+\Lambda_\bot(a)= \Lambda_r(b)+\Lambda_\bot(c). \] Checking relations, we obtain that $(L, H, \Lambda_r+\Lambda_\bot)$ is a Lefschetz triple. By \ref{_Lefshe_tri_unique_Proposition_}, $\Lambda_a=\Lambda_r+\Lambda_\bot$. This implies that $\Lambda_a$ preserves $A^r$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_A^r_Lef-Frob,_pres_by_g(A)_Proposition_} immediately implies the following useful statement: \hfill \corollary Let $A$ be a Lefschetz-Frobenius algebra such that its reduction $A^r$ is also Frobenius. Then $A^r$ is Lefschetz-Frobenius, and there exists a natural Lie algebra epimorphism ${\goth g}(A)\longrightarrow {\goth g}(A^r)$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section {The minimal Frobenius algebras and cohomology of compact Kaehler surfaces.} \label{_minimal_Fro_Section_} In this section we concentrate on the simplest case of Frobenius algebras related to Lefschetz theory. Namely, we analyze the graded Frobenius algebras $A= A_0\oplus A_2\oplus A_4$, where $dim_k A_0=1$. Such algebras are called minimal. These algebras are naturally related to the complex surfaces. \hfill \definition Let $A= \oplus A_i$ be a graded Frobenius algebra. Assume that $A= A_0\oplus A_2\oplus A_4$, $dim_k A_0=dim_k A_4=1$. Then $A$ is called {\bf a minimal graded Frobenius algebra}. \hfill \proposition \label{_minimal_is_Lefschetz_Proposition_} Let $A= A_0\oplus A_2\oplus A_4$ be a minimal graded Frobenius algebra. Then $A$ is Lefschetz-Frobenius. {\bf Proof:} Let $(\cdot,\cdot):\; A\times A\longrightarrow k$ denote the invariant scalar product on $A$. The restriction of $(\cdot,\cdot)$ to $A_2$ is a non-degenerate bilinear symmetric form (it is non-degenerate because of grading conditions \eqref{_graded_scalar_pro_Equation_}). The following statement immediately implies \ref{_minimal_is_Lefschetz_Proposition_}: \hfill \lemma \label{_el-t_with_non_zero_square_Lefschetz_Lemma_} Let $A= A_0\oplus A_2\oplus A_4$ be a minimal graded Frobenius algebra. Let $a\in A_2$ be a vector such that $(a,a)\neq 0$. Then $a$ is a Lefschetz element. {\bf Proof:} Let $a^\bot$ be the orthogonal complement of $a$ in $A_2$: \[ a^\bot:= \{b\in A_2\; |\; (a,b)=0 \}. \] Let ${\Bbb I}\in A_0$ be the unit in $A$. For all $b\in a^\bot$, $(ab,{\Bbb I})=(a,b)=0$. Since $A_4$ is one-dimensional and its generator has non-zero scalar product with ${\Bbb I}$, we have \begin{equation} \label{_ab=0_for_all_b_in_a^bot_Equation_} \forall b\in a^\bot,\; \;\; ab=0. \end{equation} Let $k_a\subset A_2$ be the one-dimensional space generated by $a$. Let $A_a:= A_0\oplus ka \oplus A_4$. Clearly, $A_a$ is a subalgebra of $A$. By \eqref{_ab=0_for_all_b_in_a^bot_Equation_}, the operator $L_a$ vanishes on $a^\bot$. Since $H(A_2) =0$, the operator $H$ also vanishes on $a^\bot\subset A_2$. Therefore it is sufficient to show that $a$ is a Lefschetz element in the algebra $A_a$. Since $A_a\cong k[x]/(x^3=0)\cong H^*(\Bbb P^2, k)$, this follows from Lefschetz theory. \ref{_el-t_with_non_zero_square_Lefschetz_Lemma_} and consequently, \ref{_minimal_is_Lefschetz_Proposition_}, is proven. We also obtained the following result: \hfill \corollary \label{_Lambda_vanish_Corollary_} Let $(L_a, H, \Lambda_a)$ be the Lefschetz triple on $A$, where $A$ is a minimal graded Frobenius algebra. Then $(L_a, H, \Lambda_a)$ all vanish on $a^\bot$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill There is an easy way to construct the minimal graded Frobenius algebras using spaces with non-degenerate symmetric bilinear forms. Namely, let $V$ be a linear space over $k$, equipped witha bilinear form $(\cdot,\cdot)_V$. Consider the linear space \[ A(V):= k{{\Bbb I}}\oplus V\oplus k\Omega, \] where $k{\Bbb I}$ and $k\Omega$ are one-dimensional spaces generated, respectively, by ${\Bbb I}$ and $\Omega$. We introduce a graded Frobenius algebra structure on $A(V)$ in the following way. The grading of $V$ is 2, the grading of ${\Bbb I}$ is 0, the grading of $\Omega$ is 4. The product on $A(V)$ is defined as follows: \hfill (i) ${\Bbb I}$ is a unit. (ii) for $v_1, v_2\in A_2(V)\cong V$, $v_1 v_2= (v_1,v_2)_V \Omega$. \hfill It remains to establish the invariant bilinear symmetric form $(\cdot,\cdot)$ on $A(V)$. \hfill (iii) On $A_2(V)\cong V$, $(\cdot,\cdot)$ is equal to $(\cdot,\cdot)_V$. (iv) The product of ${\Bbb I}$ and $\Omega$ is 1. \hfill Together with \eqref{_graded_scalar_pro_Equation_}, relations (iii) and (iv) define the form $(\cdot,\cdot)$ is a unique way. One can trivially check that this construction results in a graded Frobenius algebra. Exactly this algebra appears as the even cohomology of the compact Kaehler surface $M$, where $V=H^2(M)$ and $(\cdot,\cdot)_V$ is the intersection form. In fact, every minimal graded Frobenius algebra can be obtained this way (\ref{_all_minim_alge_are_associ_Claim_}). \hfill \definition \label{_associ_gra_Fro_Definition_ The graded Frobenius algebra $A(V)$ is called {\bf the minimal graded Frobenius algebra associated with $V$, $(\cdot,\cdot)_V$}. \hfill \claim\label{_all_minim_alge_are_associ_Claim_} Let $A= A_0\oplus A_2\oplus A_4$ be the minimal graded Frobenius algebra. Denote the restriction of the invariant scalar product to $A_2$ by $(\cdot,\cdot):\; A_2\times A_2\longrightarrow k$. Then $A$ is is isomorphic to the minimal graded Frobenius algebra associated with $A_2$, $(\cdot,\cdot)$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \proposition\label{_minimal_alge_are_Jordan_Proposition_ Let $A= A_0\oplus A_2\oplus A_4$ be a minimal graded Frobenius algebra. Then $A$ is of Jordan type. {\bf Proof:} By \ref{_Lunts_about_FLJ_Proposition_}, it is sufficient to show that for every two elements $a_1, a_2\in A_2$ of Lefschetz type, $[ \Lambda_{a_1},\Lambda_{a_2}]=0$. Denote the generators of $A_0$, $A_4$, by ${\Bbb I}$, $\Omega$, as in \ref{_associ_gra_Fro_Definition_}. The endomorphism $[ \Lambda_{a_1},\Lambda_{a_2}]$ has a grading $-4$. Therefore it is a map from $A_4$ to $A_0$. To prove that $[ \Lambda_{a_1},\Lambda_{a_2}]=0$ it is sufficient to show that \[ [ \Lambda_{a_1},\Lambda_{a_2}]\:\bigg(\Omega\bigg)=0 \] \ref{_Lambda_vanish_Corollary_} implies that $\Lambda_{a_i}(\Omega)$ is proportional to $a_i$. An easy calculation in $\goth{sl}(2)$ implies that $\Lambda_{a_i}(\Omega)=-a_i$. Similarly, $\Lambda_{a_i}(a_j)=(a_i,a_j) {\Bbb I}$. Therefore \[ [ \Lambda_{a_1},\Lambda_{a_2}]\:\bigg(\Omega\bigg) = (a_1,a_2)\cdot {\Bbb I} - (a_2,a_1) \cdot{\Bbb I} =0 \] $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill We proceed to compute the Lie algebra ${\goth g}(A)$ associated with the minimal Frobenius algebra $A=A(V)$, where $V$ is a linear space equipped with a scalar product. Denote by $\goth{so}(V)$ the Lie algebra of skew-symmetric endomorphisms of $V$. Let $\goth H$ be the 2-dimensional space over $k$ with the hyperbolic scalar product. In other words, $\goth H$ has a basis $x, y$ such that $(x,y)=1$, $(x,x)=0$, $(y,y)=0$. By $\goth{so}(V)\oplus k$ we understand a direct sum of $\goth{so}(V)$ and a trivial Lie algebra of dimension 1. \hfill \theorem\label{_calculation_of_g(A)_for_minim_Theorem_} Let $V$ be a $k$-linear space equipped with a non-degenerate scalar product. Let $A=A_0\oplus A_2\oplus A_4$ be the minimal graded Frobenius algebra $A(V)$ constructed by $V$. Take the graded Lie algebra ${\goth g}(A)={\goth g}_{-2}(A)\oplus {\goth g}_0(A)\oplus {\goth g}_2(A)$ (\ref{_Lefshe_Fro_Definition_}). Then \hfill (i) ${\goth g}_0(A)\cong \goth{so}(V)\oplus k$, (ii) ${\goth g}_2(A)\cong {\goth g}_{-2}(A)\cong V$, (ii) ${\goth g}(A)\cong \goth{so}(V\oplus \goth H)$. \hfill {\bf Proof:} Denote the invariant bilinear form on $A(V)$ by $(\cdot,\cdot)$. Let $(\cdot,\cdot)'$ another bilinear symmetric form, defined by \hfill $(a,b)'=(a,b)$ if $a,b\in A_2$, $(a,b)'=-(a,b)$ if $a,b\in A_0\oplus A_4$, $(a,b)'=0$ if $a\in A_2$ and $b\in A_0\oplus A_4$. \hfill Let $A'$ be $A$ equipped with the scalar product $(\cdot,\cdot)'$. Obviously, \[ A'\cong V\oplus \goth H. \] {\bf Step 1:} We are going to show that ${\goth g}(A)\subset \goth{so}(A')$. By trivial reasons, $L_a$ and $H$ belong to $\goth{so}(A')$ for all $a\in A_2$. Let $a\in A_2$ be an element of the Lefschetz type. To prove that $\Lambda_a\in \goth{so}(A')$, we consider the decomposition $A' = a^\bot \oplus A_a$ (see the proof of \ref{_el-t_with_non_zero_square_Lefschetz_Lemma_}). By \ref{_Lambda_vanish_Corollary_}, $\Lambda_a$ acts trivially on $a^\bot$. Since the decomposition $A' = a^\bot \oplus A_a$ is orthogonal, it is sufficient to prove that the restriction of $\Lambda_a$ to $A_a$ is skew-symmetric. Three-dimensional representation $\rho:\; \goth{sl}(2)\longrightarrow A_a$ obtained from the Lefschetz triple $(L_a, \Lambda_a, H)$ is naturally isomorphic to the adjoint representation of $\goth{sl}(2)$. Using this isomorphism, we obtain that $(\cdot,\cdot)$ is the Killing form of the Lie algebra ${\goth g}(A_a)\cong \goth{sl}(2)$. Therefore, ${\goth g}(A_a)\subset End(A_a)$ consists of skew-symmetric matrices. This finishes Step 1. \hfill {\bf Step 2:} We prove \ref{_calculation_of_g(A)_for_minim_Theorem_} (i). Since ${\goth g}_0(A)$ is the grade-preserving part of ${\goth g}(A)$, we have a homomorphism \[ {\goth g}_0(A)\stackrel {\mu}\longrightarrow \goth{so}(A_2) \cong \goth{so}(V) \] which maps an endomorphism $h\in End(A)$ to its restriction $h\restrict{A_2}\in End(A_2)$. The kernel of this homomorphism is the space of all $h\in g_0(A)$ such that $h$ vanishes on $A_2$. Therefore, $\ker\mu\in \goth{so}(A_0\oplus A_4)$. The algebra $\goth{so}(A_0\oplus A_4)\cong \goth{so}(\goth H)\cong \goth{so}(1,1)$ is commutative and one-dimensional. Combining $\mu$ and the embedding \[ \ker\mu\stackrel i\hookrightarrow \goth{so}(A_0\oplus A_4)\cong k, \] we obtain an embedding \[ {\goth g}_0(A)\stackrel m\hookrightarrow \goth{so}(V) \oplus k. \] It remains to show that $m$ is a surjection. Consider the Hodge endomorphism $H\in {\goth g}_0(A)\subset End(A)$ introduced a few sentences before \ref{_Lefschetz_triple_Definition_}. By obvious reasons, $H\in \ker\mu$. The map $i:\; \ker \mu \longrightarrow \goth{so}(A_0\oplus A_4)$ is surjective because $i(H)$ is non-zero, and $\goth{so}(A_0\oplus A_4)$ is one-dimensional. Therefore, to prove that $m$ is surjective, it is sufficient to show that $\mu:\;{\goth g}_0(A)\longrightarrow \goth{so}(A_2)$ is surjective. Let \begin{equation} \label{_condi_on_a,b_Equation_} a,b\in A_2,\;\; (a,a)\neq, \;(b,b)\neq 0,\;(a,b)=0. \end{equation} Let $\inangles{a,b}\subset V$ be the plane generated by $a$ and $b$, and $\inangles{a,b}^\bot\subset V$ be its orthogonal completion. Let $T_{ab}\subset \goth{so}(A_2)$ be the set of all skew-symmetric endomorphisms which vanish on $\inangles{a,b}^\bot$. Since $\goth{so}(2)$ is one-dimensional, for given $a,b\in V$, all elements of $T_{ab}$ are proportional. We notice that the union of all $T_{ab}$ generates the lie algebra $\goth{so}(A_2)$. Therefore, to prove that $\mu$ is surjective it is sufficient to show that $T_{ab}\subset \mu({\goth g}_0(A))$ for all $a,b$ satisfying conditions \eqref{_condi_on_a,b_Equation_}. \ref{_Lambda_vanish_Corollary_} implies that $[L_a,\Lambda_b]\in T_{ab}$. It is easy to check that this element is non-zero. The proof of \ref{_calculation_of_g(A)_for_minim_Theorem_} (i) is finished. \ref{_calculation_of_g(A)_for_minim_Theorem_} (ii) immediately follows from \ref{_Lunts_about_FLJ_Proposition_}. \ref{_calculation_of_g(A)_for_minim_Theorem_} (iii) follows fom the inclusion ${\goth g}(A)\subset \goth{so}(A')$ and comparing dimensions, where dimension of ${\goth g}(A)$ is computed via \ref{_calculation_of_g(A)_for_minim_Theorem_} (i) and \ref{_calculation_of_g(A)_for_minim_Theorem_} (ii). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill We denote the graded Lie algebra ${\goth g}(A)$ constructed by $V$ as in \ref{_calculation_of_g(A)_for_minim_Theorem_} by $\goth{so}(V,+)$. Clearly, over ${\Bbb R}$, when the symmetric form on $V$ has a signature $(a,b)$% \footnote{Of course, this means that $\goth{so}(V)=\goth{so}(a,b)$}% , \[ \goth{so}(V,+)\cong \goth{so}(a+1,b+1). \] \section[Representations of $SO(V,+)$ leading to Frobenius algebras.] {Representations of $SO(V,+)$ leading to \\ Frobenius algebras.} \label{_^dA(V)_Section_} In this section, we describe all reduced Lefschetz-Frobenius algebras $A= A_0\oplus A_2 \oplus ... \oplus A_{2d}$ with ${\goth g}(A)\cong \goth{so}(V,+)$. It turns out that such algebras are uniquely defined by the number $d$, which is {\it even}, whenever $dim V>2$. Let $V$ be a linear space supplied with a non-degenerate symmetric bilinear form $(\cdot,\cdot)$. Let $A=A(V)$ be the minimal Frobenius algebra constructed in Section \ref{_minimal_Fro_Section_}. Take the tensor product \[ A^{\otimes d}:= \underbrace{A\otimes_k A\otimes_k ... A}_{d\mbox{\ \ times}}. \] There is a natural action of $\goth{so}(V,+)$ on $A^{\otimes d}$. Let ${}^{(d)}A$ be the irreducible $\goth{so}(V,+)$-module generated by ${}^d{\Bbb I}:= {\Bbb I} \otimes {\Bbb I}\otimes ... {\Bbb I}$, where ${\Bbb I}\in A$ is the unit. The space ${}^{(d)}A$ is not necessarily a subalgebra in $A^{\otimes d}$. We introduce a new algebra structure on ${}^{(d)}A$ which does not necessarily come from the algebra structure on $A^{\otimes d}$, but instead comes from $\goth{so}(V,+)$-action as in \ref{_multi_associ_with_representa_Definition_}. Denote the graded Lie algebra $\goth{so}(V,+)$ by ${\goth g}= {\goth g}_{-2}\oplus {\goth g}_0\oplus {\goth g}_2$. The algebra ${\goth g}_0$ has an additional decomposition: ${\goth g}_0= \goth{so}(V)\oplus kH$ (\ref{_calculation_of_g(A)_for_minim_Theorem_}). Clearly, all elements of ${\goth g}_{-2}$ vanish on ${}^d{\Bbb I}$, $\goth{so}(V)\subset {\goth g}_0$ vanish on ${}^d{\Bbb I}$ and $H$ acts on ${}^d{\Bbb I}\in {}^{(d)}A$ as multiplication by $-2d$. Therefore we may apply \ref{_multi_associ_with_representa_Definition_} to the representation ${}^{(d)}A$ and the vector ${}^d{\Bbb I}$. Denote the resulting algebra ${}^{(d)}A_{{\goth g}, {\Bbb I}}$ by ${}^dA(V)$. \hfill \theorem \label{_all_alg_with_so_are_^dA_Theorem_} Let $\tilde A=A_0\oplus A_2\oplus ... \oplus A_{2n}$ be a Lefschetz-Frobenius algebra of Jordan type. Assume that the graded Lie algebra ${\goth g}(A)$ is isomorphic to $\goth{so}(V,+)$ and $dim_k V> 2$. Let $A$ be the subalgebra of $\tilde A$ generated by $A_0$ and $A_2$ (also known as {\bf reduction} of $\tilde A$). Then (i) $n$ is even (ii) $A$ is isomorphic to ${}^{n/2} A(V)$ as a graded algebra% \footnote{The invariant Frobenius pairing is unique up to a scalar, which is easy to see.}% {}. \hfill {\bf Remark:} In particlural, this theorem implies that the reduction $A$ of $\tilde A$ is Frobenius, which is not immediately clear in general case. \hfill {\bf Proof:} Let ${\goth g}={\goth g}_{-2}\oplus {\goth g}_0\oplus {\goth g}_2$ be the graded Lie algebra $\goth{so}(V,+)$. Denote the unit in $A$ by ${\Bbb I}$. By definition, $A_0$ is ${\goth g}_0$-invariant one-dimensional space. We have shown that ${\goth g}_0= \goth{so}(V)\oplus kH$. Since $\goth{so}(V)$ is a simple Lie algebra, $\goth{so}(V)({\Bbb I})=0$. Therefore, to prove \ref{_all_alg_with_so_are_^dA_Theorem_} (i), it is sufficient to prove the following lemma. \hfill \lemma \label{_repres_so(V,+)_even-weight_Lemma_} Let $\rho:\;{\goth g}\longrightarrow End(M)$ be a simple representation of ${\goth g}\cong \goth{so}(V,+)$, where $dim_k(V)>2$. Let ${\Bbb I}\in M$ be a vector such that ${\goth g}_{-2}({\Bbb I})=0$, $\goth{so}(V)({\Bbb I})=0$,% \footnote{As elsewhere, we use the decomposition ${\goth g}_0\cong \goth{so}(V)\oplus kH$ provided by \ref{_calculation_of_g(A)_for_minim_Theorem_}.} and $H({\Bbb I})=-2n\cdot{\Bbb I}$. Then $n$ is even. {\bf Proof:} {\bf Step 1:} We show that $k\neq 1$. Assume the contrary. Consider the decomposition $M\cong M_{-1}\oplus M_1$, given by the weights of $H$. Since ${\Bbb I}$ is a highest weight vector for some root system in ${\goth g}$ (see \ref{_Cartan_exists_in_so(V,+)_Claim_}), the corresponding weight space is one-dimensional. This implies that $dim_k(M_{-1})=1$. There exists an automorphism of $\goth{so}(V,+)$ which maps $H$ to $-H$ (an easy check). Therefore, $dim_k(M_{-1})=dim_k(M_1)=1$. In other words, the representation $M$ is two-dimensional. Consider $M$ as a representation of $\goth{so}(V)\subset \goth{so}(V,+)$. Since ${\Bbb I}$ is invariant with respect to $\goth{so}(V)$, the space $M$ is decomposed into a sum of two one-dimensional $\goth{so}(V)$-invariant subspaces. For $dim_k(V)>2$, the Lie algebra $\goth{so}(V)$ is simple. Therefore its one-dimensional representations are trivial. We obtain that $\goth{so}(V)(M)=0$. Since ${\goth g}=\goth{so}(V,+)$ is a simple Lie algebra, the homomorphism $\rho:\; {\goth g}\longrightarrow End(M)$ cannot have proper non-zero kernel. Therefore, $\rho({\goth g})=0$. {\bf Step 2:} We use the following statement, which is easy to check. \hfill \claim \label{_Cartan_exists_in_so(V,+)_Claim_} Let $\goth{so}(V)\oplus kH \stackrel i \hookrightarrow \goth{so}(V,+)$ be the embedding provided by \ref{_calculation_of_g(A)_for_minim_Theorem_} (i). Then there exists a Cartan subalgebra $\goth h'\subset \goth{so}(V)$ such that $i(\goth h'\oplus kH)$ is a Cartan subalgebra in $\goth{so}(V,+)$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Take a Cartan subalgebra $\goth h:= i(\goth h'\oplus kH)\subset {\goth g}$ provided by \ref{_Cartan_exists_in_so(V,+)_Claim_}. Clearly, the linear map \[ -H\check{\;}:= -(H,\cdot), \;\;-H\check{\;}:\; \goth h\longrightarrow k\] is a root. Taking a root system $\alpha_1, ... ,\alpha_m$ in $\goth h'\subset \goth{so}(V)$, we obtain that $\alpha_1, ... ,\alpha_m, -H\check{\;}$ is a root system in ${\goth g}$. In this root system, ${\Bbb I}\in M$ is a highest weight vector of the representation $M$. It is known that the set\footnote{Here, $\goth h\check{\;}$ means $\goth h$ dual} $W\subset \goth h\check{\;}$ of possible weights of the highest weight vector coinsides with the intersection of a weight lattice $L$ in $\goth h\check{\;}$ and a Weyl chamber. In particular, $W\subset\goth h\check{\;}$ is an abelian semi-group with group structure induced from $\goth h\check{\;}$. The weight of ${\Bbb I}$ corresponding to the root system $\alpha_1, ... ,\alpha_m, -H\check{\;}$ is $(0,0 ,..., 2n)$. Let $W_0\subset W$ be the set of all $(0,0 ,..., 2n)\subset W\subset \goth h\check{\;}$ which correspond to representations $M$ satisfying conditions of \ref{_repres_so(V,+)_even-weight_Lemma_}. Clearly, \[ W_0= \bigg \{ (x_1, .... , x_m)\in W \;\;|\;\; x_1=...=x_{m-1}=0\bigg \} \] By Step 1, $n\neq 1$. Since Weyl chamber is invariant with respect to homotheties, the semigroup $W_0$ is isomorphic to ${\Bbb Z}_{\geq 0}$. Therefore, $n$ is never odd. \ref{_repres_so(V,+)_even-weight_Lemma_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill To prove \ref{_all_alg_with_so_are_^dA_Theorem_}, we notice that the simple representation of a reductive Lie algebra is uniquely determined by its highest weight. The weights of representations of $\goth{so}(V,+)$ corresponding to Lefschetz-Frobenius algebras are computed a few lines above. In particular, we obtained that for some root system in ${\goth g}$, the highest weight of $A$ is $(0,0,0... , 2n)$. Simple representations with a given highest weight are isomorphic. Therefore, as a representation of ${\goth g}$, $A\cong {}^dA(V)$. By \ref{_g_structure_defines_algebr_Claim_}, the action of ${\goth g}$ detrermines multiplication in $A$ uniquely. This finishes the proof of \ref{_all_alg_with_so_are_^dA_Theorem_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section[Computing the structure Lie algebra for the cohomology of a hyperkaehler manifold, part \ I.]{Computing the structure Lie algebra for the cohomology of a hyperkaehler manifold, \\part \ I.} \label{_computing_g_for_hyperk_pt-I_Section_} Let $M$ be a compact Kaehler manifold. According to \ref{_Lefshe_Frob_if_a_Lefshe_ele_exists_Proposition_}, the ring $A:=H^*(M)$ is Lefschetz-Frobenius. The aim of this section is to compute ${\goth g}(A)$ in the case when $M$ is a simple compact hyperkaehler manifold. The answer is hinted at by the following statement, which is proven in \cite{_so5_on_cohomo_}: \hfill \proposition \label{_subalg_in_g_genera_by_three_Kae_forms_Proposition_ Let $\c H\in Hyp$ be a hyperkaehler structure on $M$. Consider the Kaehler classes \[ \omega_I= P_1(\c H),\;\; \omega_J=P_2(\c H),\;\; \omega_K= P_3(\c H), \;\;\omega_I, \omega_J,\omega_K\in H^2(M,{\Bbb R}). \] Cohomology classes $\omega_I, \omega_J,\omega_K$ are Lefschetz elements in the graded Frobenius algebra $A:= H^*(M,{\Bbb R})$. Consider the graded subalgebra ${\goth g}(\c H)$ in ${\goth g}(A)$ generated by $L_{\omega_I},L_{\omega_J},L_{\omega_K}$, $\Lambda_{\omega_I}, \Lambda_{\omega_J},\Lambda_{\omega_K}$ and $H$. Then ${\goth g}(\c H)$ is isomorphic to ${\goth g}({\Bbb R}^3)$, where ${\goth g}({\Bbb R}^3)\cong \goth{so}(4,1)$ is the structure Lie algebra of the minimal graded Lefschetz-Frobenius algebra corresponding to the linear space ${\Bbb R}^3$ with positively defined scalar product. In particular, the graded algebra ${\goth g}(\c H)$ is independent from $\c H$ and $M$. {\bf Proof:} This statement is proven in \cite{_so5_on_cohomo_}. It is based on the commutation relations in ${\goth g}(\c H)$ given as follows. Denote $P_i(\c H)$ by $\omega_i$, $i=1,2,3$. Denote $L_{\omega_i}$ by $L_i$, and $\Lambda_{\omega_i}$ by $\Lambda_i$. Let $K_{ij}:= [L_i,\Lambda_j]$, $i\neq j$. Then the following relations are true: \begin{equation} \label{_so5_relations_Equation_} \begin{array}{l} {}[ L_i, L_j] = [\Lambda_i,\Lambda_j] =0;\;\; \\[2mm] [L_i,\Lambda_i]= H;\;\; [H, L_i] = 2 L_i; \;\; [H, \Lambda_i]=-2\Lambda_i \\[2mm]{} K_{ij}=-K_{ji}, [K_{ij}, K_{jk}]=2 K_{ik}, [K_{ij}, H] =0 \\[2mm]{} [K_{ij} L_j]=2 L_i;\;\; [K_{ij} \Lambda_j] = 2 \Lambda_i \\[2mm]{} [K_{ij}, L_k] = [K_{ij}, \Lambda_k] =0\;\; (k\neq i,j)\\[2mm] \end{array} \end{equation} $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $M$ be a simple hyperkaehler manifold, $A=H^*(M,{\Bbb R})$ be its cohomology ring, equipped with the invariant pairing provided by the Poincare duality. We consider $A=\oplus A_i=\oplus H^i(M,{\Bbb R})$ as a graded Frobenius algebra over ${\Bbb R}$. In Section \ref{_Hodge-Rie_independent_Section_}, we defined the normalized Hodge-Riemann pairing $(\cdot,\cdot)_{\c H}$ on $H^2(M,{\Bbb R})$. Let $V$ be a linear space $H^2(M)$ equipped with this pairing. In Section \ref{_minimal_Fro_Section_}, we defined a graded Frobenius algebra ${\goth g}(V)$, also denoted by $\goth{so}(V,+)$. By definition, over ${\Bbb R}$ the algebra $\goth{so}(V,+)$ is isomorphic to $\goth{so}(m+1,n+1)$, where $(m,n)$ is the signature of $V$. \hfill \theorem\label{_g(A)_for_hyperkae_Theorem_} In this notation, ${\goth g}(A)$ is isomorphic to ${\goth g}(V)$. \hfill \ref{_g(A)_for_hyperkae_Theorem_} is the main result of this paper. It is proven in the subsequent sections. The present section is dedicated to proving that the Lefschetz-Frobenius algebra $A$ is of Jordan type. This is a crucial step in proof of \ref{_g(A)_for_hyperkae_Theorem_}. \hfill {\bf Remarks:} For a hyperkaehler manifold with $dim \; H^2(M) =3$ (minimal dimension which is not obviously impossible), \ref{_g(A)_for_hyperkae_Theorem_} is equivalent to \ref{_subalg_in_g_genera_by_three_Kae_forms_Proposition_}. For a hyperkaehler manifold with $dim\; H^2(M) =n$, the Riemann-Hodge metric on $H^2(M,{\Bbb R})$ has a signature $(3,n-3)$. This means that ${\goth g}(V)$ is isomorphic to $\goth{so}(4,n-2)$. \hfill \proposition \label{_H^*_hyp_Jordan_type_Proposition_} Let $A$ be the Lefschetz-Frobenius algebra of cohomology of a simple compact hyperkaehler manifold. Then $A$ is of Jordan type. {\bf Proof:} According to \ref{_Lunts_about_FLJ_Proposition_}, it is sufficient to show that for all $a,b \in A_2$, $a$, $b$ of Lefschetz type, $[\Lambda_a,\Lambda_b]=0$. Let $a,b\in A_2$, $\c H\in Hyp$, $\c H = (I,J,K, (\cdot,\cdot))$. We write $a\bullet_{\c H}b$ if there exist complex structures $I_1, I_2$ which are induced by $\c H$ such that $a, b$ are Kaehler classes corresponding to $I_1$, $I_2$ and the metric $(\cdot,\cdot)$.\footnote{the metric $(\cdot,\cdot)$ is Kaehler with respect to the complex structures $I_1$, $I_2$, as the definition of induced complex structures implies.} Clearly, if $a\bullet_{\c H} b$, then $a$ and $b$ are of Lefschetz type. \hfill \lemma \label{_a_bullet_H_b=>_Lambdas_commute_Lemma_} Let $\alpha,\beta\in A_2$, $\c H\in Hyp$, $\alpha\bullet_{\c H}\beta$. Then $[\Lambda_\alpha,\Lambda_\beta]=0$. {\bf Proof:} Let ${\goth g}(\c H)$ be the graded Lie algebra defined in \ref{_subalg_in_g_genera_by_three_Kae_forms_Proposition_}. Since ${\goth g}(\c H)={\goth g}({\Bbb R}^3)$, the $-4$-th component of ${\goth g}(\c H)$ vanishes: ${\goth g}_{-4}(\c H)=0$. Therefore for all $\lambda,\mu\in {\goth g}_{-2}(\c H)$, we have $[\lambda,\mu]=0$. Therefore, to prove \ref{_a_bullet_H_b=>_Lambdas_commute_Lemma_} it is sufficient to show that $\Lambda_\alpha,\Lambda_\beta\in {\goth g}_{-2}(\c H)$. This is implied by the following lemma. \hfill \lemma\label{_Kaehle_cla_indu_by_H_in_g(H)_Lemma_} Let $\c H\in Hyp$, $\c H=(I,J,K,(\cdot,\cdot))$ and $I'$ be a complex structure induced by $\c H$. Let $\omega\in A_2$ be the Kaehler class corresponding to $I'$ and the metric $(\cdot,\cdot)$. Then $\Lambda_\omega \in {\goth g}_{-2}(\c H)$. {\bf Proof:} Let $\omega_i:=P_i(\c H)$, $i=1,2,3$. By definition of induced complex structures, $I'= a I+b J+ c K$, where $a,b,c\in {\Bbb R}$, $a^2+b^2+c^2=1$. Since $\omega(x,y)= (x, I'y)$, we have $\omega=a\omega_1+b\omega_2+c\omega_3$. Let \[ \Lambda:= a\Lambda_{\omega_1}+ b\Lambda_{\omega_2} + c\Lambda_{\omega_3}. \] Using \eqref{_so5_relations_Equation_}, we see that $[L_\omega,\Lambda]=H$. The other relations defining Lefschetz triples checked automatically, we obtain that $(L_\omega, H, \Lambda)$ is a Lefschetz triple. Therefore, $\Lambda=\Lambda_\omega\in{\goth g}_2(\c H)$. This proves \ref{_a_bullet_H_b=>_Lambdas_commute_Lemma_} and \ref{_Kaehle_cla_indu_by_H_in_g(H)_Lemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $a,b\in A_2$. We write $a\bullet b$ if there exist $\c H\in Hyp$, $\lambda\in {\Bbb R}$, $\lambda\neq 0$ such that $a\bullet_{\c H}\lambda b$. Clearly, $\Lambda_{\lambda b} = \lambda^{-1}\Lambda_b$. Therefore, \ref{_a_bullet_H_b=>_Lambdas_commute_Lemma_} implies the following statement. \hfill \claim \label{_a_bullet_b_Lambdas_commute_Claim_} Let $a,b\in A_2$, $a\bullet b$. Then $[\Lambda_a,\Lambda_b]=0$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill The relation $a\bullet b$ is prominent further on in this paper. We could have given an alternative definition of $a\bullet b$ as follows. \hfill \lemma \label{_a_bullet_b_if_hyperk_exists_Lemma_} Let $a,b \in A_2$. The following conditions are equivalent: (i) $a \bullet b$ (ii) There exists $\c H\in Hyp$ such that $a$ and $b$ can be expressed as linear combinations of $P_i(\c H)$, $i=1,2,3$. {\bf Proof:} Clear from \ref{_action_SO(3)_on_Hyp_via_periods_Lemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill {\bf Remark:} Let $a\bullet b$, $a,b\neq 0$. Then both $a$ and $b$ are elements of Lefschetz type. \hfill Let $S\in A_2$ be a set of all elements of Lefschetz type. By \ref{_a_Lefshe_if_a^i_iso_Lemma_}, $S$ is Zariski open in $A_2$. Let $\nu:\; S\longrightarrow End(A)$ map $a\in S$ to $\Lambda_a\in End(A)$. An easy linear-algebraic check implies that $\nu$ is an algebraic map. Therefore the map $\eta:\; S\times S\longrightarrow End(A)$, $a,b \longrightarrow [\Lambda_a,\Lambda_b]$ is also an algebraic map. To prove \ref{_H^*_hyp_Jordan_type_Proposition_} it is sufficient to show that $\eta$ is identically zero, as shown by \ref{_Lunts_about_FLJ_Proposition_}. By \ref{_a_bullet_b_Lambdas_commute_Claim_}, for all $a,b\in A_2$, such that $a\bullet b$, we have $\eta(a,b)=0$. Therefore, \ref{_H^*_hyp_Jordan_type_Proposition_} is implied by the following statement: \hfill \lemma \label{_a_bullet_b_dense_in_A_2_x_A_2_Lemma_} Let $D\subset A_2\times A_2$ be the set of all pairs $(a,b)\in A_2\times A_2$ such that $a\bullet b$. Then $D$ is Zariski dense in $A_2$. {\bf Proof:} We use the following trivial statement: \hfill \sublemma \label{_a_bullet_b_linear_span_Sublemma_} Let $(a,b)$ and $(a',b')$ be two pairs in $A_2$ such that the linear span of $(a,b)$ is equal to the linear span of $(a',b')$. Then \[ a\bullet b\;\;\Leftrightarrow \;\; a'\bullet b'. \] {\bf Proof:} See \ref{_a_bullet_b_if_hyperk_exists_Lemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Denote the Grassmanian of all 2-dimensional planes in $H^2(M,{\Bbb R})$ by $Gr$. Let $Gr^\bullet$ be the space of planes generated by $a, b$ with $a\bullet b$. To prove \ref{_a_bullet_b_dense_in_A_2_x_A_2_Lemma_} it is sufficient to show that $Gr^\bullet$ is Zariski dense in $Gr$. \hfill In Section \ref{_Q_c_defini_Section_}, we defined the period map $Q_c:\; Comp\longrightarrow G_r$. \hfill \claim \label{_D_in_Q_c(Comp)_Claim_} The space $Gr^\bullet$ coincides with $Q_c(Comp)$. {\bf Proof:} Clear from definitions. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Since the map $Q_c$ is etale (see the end of Section \ref{_Q_c_defini_Section_}), its image contains an open set in $Gr$. Therefore $Q_c(Comp)$ is Zariski dense in $Gr$. \ref{_a_bullet_b_dense_in_A_2_x_A_2_Lemma_}, and consequently, \ref{_H^*_hyp_Jordan_type_Proposition_}, is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{Calculation of a zero graded part of the structure Lie algebra of the cohomology of a hyperkaehler manifold, part I.} \label{_compu_g_0_part_1_Section_} In this section, we make steps related to computation of the grade zero part of the Lie algebra ${\goth g}(A)={\goth g}_{-2}(A)\oplus {\goth g}_0(A)\oplus {\goth g}_2(A)$. As in the previous section, $V$ denotes the space $H^2(M,{\Bbb R})$, considered as a linear space with the scalar product $(\cdot, \cdot)_{\c H}$, and $A$ is the Frobenius algebra $H^*(M,{\Bbb R})$. We construct an epimorphism ${\goth g}_0(A)\longrightarrow \goth{so}(V)\oplus kH$, where $kH$ is one-dimensional Lie algebra. \hfill With every $I\in Comp$, we associate the semisimple endomorphism $ad I$ of $A$, defined as follows (see also Section \ref{hyperk_manif_Section_}). Consider the Hodge decomposition $H^i(M,{\Bbb C})=\oplus_{p+q=i}H^{p,q}(M)$. Let $ad^cI:\; H^i(M,{\Bbb C}) \longrightarrow H^i(M,{\Bbb C})$ multiply $(p,q)$-forms by the scalar $(p-q)\sqrt{-1}\:$. One can check that $ad^c I$ commutes with the standard real structure on $H^i(M,{\Bbb C})$. Therefore, $ad^c I$ is a complexification of a (uniquely defined) endomorphism of $H^i(M,{\Bbb R})$. Denote this endomorphism by $ad I$. This definition coinsides with one given in Section \ref{hyperk_manif_Section_}. Consider the action of $ad I$ on $V=H^2(M, {\Bbb R})$. Using Hodge-Riemann relations, we immediately obtain that $ad I\restrict{V}\in \goth{so}(V)$. Let $\c M_2\subset End(V)$ be the Lie algebra generated by endomorphisms $ad I\restrict{V}$, for all $I\in Comp$. Clearly, $\c M_2\subset \goth{so}(V)$. One can show that $\c M_2$ is a Mumford-Tate group of $(M, I)$ for generic $I\in Comp$, although we never use this observation. Let $v:\; B\longrightarrow B$ be an endomorphism of a linear space $B$. We call the endomorphism \[ v^{\circ{}}\in End(B), \;\;v^{\circ{}}:= v-\frac{1}{\dim B} Tr(B) Id_B \] {\bf the traceless part of $v$}. Clearly, the map $Tl:\; \goth{gl}(V)\longrightarrow \goth{sl}(V)$, $Tl(v)= v^{\circ{}}$ is a Lie algebra homomorphism. For all $g\in {\goth g}_0(A)$, consider the restriction $g\restrict{A_2}:\; A_2\longrightarrow A_2$. Let $g^\circ\in End(A_2)$ be the traceless part of $g\restrict{A_2}$. This defines a Lie algebra homomorphism\footnote{$V$ is $A_2$ is $H^2(M,{\Bbb R})$} $t:\; {\goth g}_0(A)\longrightarrow \goth{sl}(V)$, $t(g)= g^\circ$. Consider the one-dimensional Lie algebra in $End(A)$, generated by the Hodge endomorphism $H\in End(A)$. Denote this algebra by $k H$. Let $s:\; {\goth g}_0(A)\longrightarrow k H$ map $g\in {\goth g}_0(A)$ to \[ -\frac{Tr(g\restrict{A_0})}{\frac{1}{2}\dim_{\Bbb R} M} H\in kH. \] The map $s$ is defined in such a way that $s(H)= H$. Clearly, $s$ is also a Lie algebra homomorphism. \hfill \proposition \label{_str_of_g_0_Proposition_} The following statements are true: \hfill (i) $t({\goth g}_0(A))\subset \c M_2$ (ii) The inclusion $\c M_2\subset \goth{so}(V)$ is an equality: $\c M_2= \goth{so}(V)$. (iii) The map $t\oplus s:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)\oplus k H$ is an epimorphism. \hfill {\bf Proof:} We use the following simple lemma. \hfill \lemma \label{_g_0_gener_by_[L,Lambda]_Lemma_} (see also \cite{_Lunts-Loo_}) Let $\c A$ be a Lefschetz-Frobenius algebra of Jordan type, ${\goth g}(\c A)={\goth g}_{-2}(\c A)\oplus{\goth g}_0(\c A)\oplus{\goth g}_2(\c A)$ be its structure Lie algebra. Then ${\goth g}_0(\c A)$ is a linear span of the elements $[L_a,\Lambda_b]$, where $a,b$ are Lefschetz elements in $\c A_2$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $S\subset A_2$ be the set of all elements of Lefschetz type. Let $\nu:\; S\times S \longrightarrow {\goth g}_0(A)$ be the map $a,b\longrightarrow [L_a,\Lambda_b]$. \ref{_g_0_gener_by_[L,Lambda]_Lemma_} implies that ${\goth g}_0(A)$ is a linear span of the set $\nu(S\times S)$. Therefore, to prove \ref{_str_of_g_0_Proposition_} (i) it is sufficient to show that for all $a,b\in S$, $t(\nu(a,b))\in \c M_2$. As we have seen, $S$ is a Zariski open subset in $A_2$. Consider $S$ as an algebraic manifold with the algebraic structure induced from $A$. With respect to this algebraic structure, both $t$ and $\nu$ are algebraic maps. Therefore it is sufficient to show that for a Zariski dense subset $D\subset S\times S$, $t(\nu(D))\subset \c M_2$. By \ref{_a_bullet_b_dense_in_A_2_x_A_2_Lemma_}, the set of all $a,b\in S$, $a\bullet b$ is Zariski dense in $S$. Therefore, the inclusion $t(\nu(S\times S))\subset \c M_2$ is implied by the following statement: \hfill \lemma \label{_[L_a,Lambda_b]=I_for_(a,b)_in_Q_c(I)_Lemma_} Let $a,b \in S$, $a\bullet b$, where $a$ is not proportional to $b$. Let $\c L$ be a plane in $H^2(M,{\Bbb R})$ spanned by $a$ and $b$. According to \ref{_D_in_Q_c(Comp)_Claim_}, there exist $\c I\in Comp$ such that $Q_c(I)= \c L$. Let \[ \xi_1:= \bigg([L_a,\Lambda_b]\bigg|_{{}_{H^2(M,{\Bbb R})}}\bigg)^\circ \in End(H^2(M, {\Bbb R}) \] be the traceless part of the restriction of the linear operator $[L_a,\Lambda_b]\in End(A)$ to $H^2(M, {\Bbb R})=A_2\subset A$. Let $\xi_2$ be the restriction of $ad I$ to $H^2(M, {\Bbb R})$. Then $\xi_1$ is proportional to $\xi_2$. {\bf Proof:} The space $A_2\cong H^2(M, {\Bbb R})$ is equipped with the normalized Hodge-Riemann pairing $(\cdot,\cdot)_{\c H}$. Let $b,x$ be an orthogonal basis of $\c L$. Clearly, $x=\lambda a +\mu b$, where $\lambda,\mu\in {\Bbb R}$. This implies that \begin{equation} \label{_adding_L_in_[L,Lambda]_Equation_} [L_x,\Lambda_b] = \lambda H +\mu [L_a,\Lambda_b]. \end{equation} Since the endomorphism $\lambda H\restrict{A_2}\in End(A_2)$ is proportional to identity, the traceless parts \[ \bigg([L_a,\Lambda_b]\bigg)^\circ, \;\; \bigg([L_x,\Lambda_b]\bigg)^\circ \] are proportional. Therefore, proving \ref{_[L_a,Lambda_b]=I_for_(a,b)_in_Q_c(I)_Lemma_}, we may assume that \[ (a,b)_{\c H}=0, \;\; (a,a)_{\c H}= (b,b)_{\c H}>0. \] Let $\c H\in Hyp$ be a hyperkaehler structure such that $\Phi^{hyp}_c(\c H)= \c I$. The relation $\Phi^{hyp}_c(\c H)= \c I$ means that $\c H= (\c I, J, K, (\cdot,\cdot))$ for some $J, K, (\cdot,\cdot)$. By definition of $Q_c$, the linear span of $P_2(\c H)$, $P_3(\c H)$ coinsides with $\c L$. Using \ref{_action_SO(3)_on_Hyp_via_periods_Lemma_}, we can easily find a hyperkaehler structure $\c H'\in Hyp$, $\c H'$ equivalent to $\c H$, such that $P_2(\c H)=\lambda a$, $P_3(\c H)=\lambda b$ for some $\lambda \in {\Bbb R}$. Now, \ref{_[L_a,Lambda_b]=I_for_(a,b)_in_Q_c(I)_Lemma_} is a consequence of the following simple statement: \hfill \claim \label{_[L_2,Lambda_3]=adI_Claim_} Let $\c H\in Hyp$, $\c H= (I, J, K, (\cdot,\cdot))$, $a=P_2(\c H)$, $b=P_3(\c H)$. Then the following endomorphisms of $A$ coinside: \[ [L_a,\Lambda_b]=ad I \] {\bf Proof:} See \cite{_so5_on_cohomo_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill This finishes the proof of \ref{_str_of_g_0_Proposition_} (i). To prove \ref{_str_of_g_0_Proposition_} (ii), we recall the following linear-algebraic construction. Let $V$ be a linear space equipped with non-degenerate symmetric bilinear form $(\cdot,\cdot)$, and $\c L\subset V$ be a 2-dimensional plane in $V$, such that the restriction of $(\cdot,\cdot)$ to $\c L$ is non-degenerate. Let $\c L^\bot$ be the orthogonal complement of $V$ to $\c L$. Let $T_{\c L}$ be the set of all non-trivial skew-symmetric endomorphisms of $V$ which vanish on $\c L^\bot$. As we have seen previously, all elements of $T_{\c L}$ are proportional. Let $Gr^\circ$ be the space of all 2-dimensional planes $\c L\subset V$ such that the restriction $(\cdot,\cdot)_{\c H}\restrict{\c L}$ is non-degenerate. \hfill \claim \label{_T_L_generate_SO_Claim_} The linear span of the union \[ \bigcup\limits_{\c L\in Gr^\circ} T_{\c L} \] coinsides with $\goth{so}(V)$. {\bf Proof:} Clear. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill By \ref{_T_L_generate_SO_Claim_}, to prove that $\c M_2=\goth{so}(V)$ it is sufficient to show that for a Zariski dense set $\c E\subset Gr^\circ$, we have \[ \forall \c L\in \c E, \;\; T_{\c L}\subset \c M_2. \] For $\c E$, we take the set $Gr^\bullet$ of \ref{_D_in_Q_c(Comp)_Claim_}. Since $Gr^\bullet = Q_c(Comp)$ is Zariski dense in $Gr$, it is sufficient to show that for all $\c L\in Q_c(Comp)$, we have $T_{\c L}\subset \c M_2$. This is implied by \ref{_[L_2,Lambda_3]=adI_Claim_} and the following statement. \hfill \claim \label{_T_Q(I)_is_[L,Lambda]_Claim_} In assumptions of \ref{_[L_2,Lambda_3]=adI_Claim_}, the following two sets coinside: \[ \bigg\{\lambda [L_a,\Lambda_b],\;\lambda\in{\Bbb R}\backslash 0\bigg\} = T_{Q_c(I)}. \] {\bf Proof:} Since all elements of $T_{Q_c(I)}$ are proportional and $[L_a,\Lambda_b]\neq 0$, it is sufficient to show that $[L_a,\Lambda_b]\in T_{Q_c(I)}$. Let $\c L\subset V$ be the linear span of $a$ and $b$. Obviously from definition, $Q_c(I)=\c L$. Let $\c L^\bot$ be the orthogonal complement to $\c L$ in $V$. We need to show that the restriction of $[L_a,\Lambda_b]$ to $V=H^2(M, {\Bbb R})$ vanishes on $\c L^\bot$. Consider the Hodge decomposition on $H^2(M)$ associated with the complex structure $I$. By definition, \[ \c L = \bigg( H^{2,0}_I(M)\oplus H^{0,2}_I(M)\bigg) \cap H^2(M, {\Bbb R}). \] Hodge-Riemann relations imply that \[ \c L^\bot= H^{1,1}_I(M)\cap H^2(M, {\Bbb R}) \] (see Section \ref{_Period_and_Hodge_Riemann_Section_} for details). By defintion of $ad I$, we have $ad I(H^{1,1}_I(M))=0$. Since $[L_a,\Lambda_b] = ad I$, we obtain that $[L_a,\Lambda_b]\in T_{Q_c(I)}$. This proves \ref{_T_Q(I)_is_[L,Lambda]_Claim_}, and consequently, proves \ref{_str_of_g_0_Proposition_} (ii). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Using the fact that \[ \{ a,b\in A_2\times A_2 \;\; | \; \; a\bullet b \} \] is Zariski dense in $A_2\times A_2$, we can derive from \ref{_T_Q(I)_is_[L,Lambda]_Claim_} the following corollary: \hfill \corollary \label{_[L,Lambda]_in_T_a,b_Corollary_} Let $a,b\in A_2$ be elements of Lefschetz type. Let $\c L\subset A_2$ be a plane generated by $a$ and $b$ and $\c L^\bot$ be its orthogonal complement in $A_2=V$. Let $[L_a,\Lambda_b]\bigg|_{{}_{V}}^\circ$ be a traceless part of $[L_a,\Lambda_b]\bigg|_{{}_{V}}\in End(V)$. Then $[L_a,\Lambda_b]\bigg|_{{}_{V}}^\circ$ vanish on $\c L^\bot$. {\bf Proof:} Using the argument with Zariski dense sets, we see that it is sufficient to check \ref{_[L,Lambda]_in_T_a,b_Corollary_} for $a,b$ such that $a\bullet b$, and $(b,b)_{\c H}\neq 0$. When $(a,b)_{\c H} =0$, \ref{_[L,Lambda]_in_T_a,b_Corollary_} is a direct consequence of \ref{_T_Q(I)_is_[L,Lambda]_Claim_}. If $(a,b)_{\c H} \neq 0$, take \[ x:= a - \frac{(a,b)_{\c H}}{(b,b)_{\c H}} b. \] Clearly, $(x,b)_{\c H} =0$. The traceless part of $[L_b, \Lambda_b]\bigg|_{{}_{V}}= H\restrict{V}$ is zero. Therefore, \[ [L_a,\Lambda_b]\bigg|_{{}_{V}}^\circ = [L_x,\Lambda_b]\bigg|_{{}_{V}}^\circ \] This reduces \ref{_[L,Lambda]_in_T_a,b_Corollary_} to the case $(a,b)_{\c H} =0$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill To prove \ref{_str_of_g_0_Proposition_} (iii), we notice that from the proof of \ref{_str_of_g_0_Proposition_} (ii) it follows also that $t:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)$ is an epimorphism. Therefore, it is sufficient to find an element $e\in {\goth g}_0(A)$ such that $t(e)=0$, $s(e)\neq 0$. This element is a Hodge endomorphism $H\in {\goth g}_0(A)\subset End(A)$. We proved \ref{_str_of_g_0_Proposition_} (iii). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{Calculation of a zero graded part of the structure Lie algebra of the cohomology of a hyperkaehler manifold, part II.} \label{_compu_g_0_part_2_Section_} We work in assumptions of the previous section. \hfill \theorem \label{_g_0_computed_Theorem_} In assumptions of \ref{_str_of_g_0_Proposition_}, consider the homomorphism \[ u:\; {\goth g}_0(A)\longrightarrow \goth{so}(V) \oplus k H, \] $u=t\oplus s$. Then $u$ is an isomorphism. {\bf Proof:} We use the following technical result: \hfill \proposition \label{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_} Let $A$ be a Lefschetz-Frobenius algebra of cohomology of a compact simple hyperkaehler manifold. Let $a, b\in A_2$ be elements of Lefschetz type. Then \[ {{(b,b)_{\c H}}} [L_a, \Lambda_b] - {{(a,a)_{\c H}}} [\Lambda_a,L_b] =2(a,b)_{\c H} H, \] where $(\cdot,\cdot)_{\c H}$ is a normalized Hodge-Riemann pairing. {\bf Proof:} First of all, we prove the following lemma. \hfill \lemma \label{_Lambda_additive_for_a_bullet_b_Lemma_ In assumptions of \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_}, let $a,b\in A_2$, $a\bullet b$. Then $a+b$ is also of Lefschetz type, and \[ {(a+b,a+b)_{\c H}}\Lambda_{a+b} = {(a,a)_{\c H}}\Lambda_a + {(b,b)_{\c H}}\Lambda_b. \] {\bf Proof:} Let $\c H\in Hyp$ be a hyperkaehler structure, such that there exist $x_i, y_i\in {\Bbb R}$, $i=1,2,3$, such that \[ a =\sum x_i\omega_i, \;\; b=\sum y_i\omega_i, \] where $\omega_i=P_i(\c H)$, $i=1,2,3$. Such $\c H$ exists by definition of the relation $a \bullet b$. Clearly, \[ (\omega_1,\omega_1)_{\c H}= (\omega_2,\omega_2)_{\c H} = (\omega_3,\omega_3)_{\c H}. \] Let $c:= {(\omega_1,\omega_1)_{\c H}}$. We are going to show that \begin{equation}\label{_Lambda_a_as_lin_combi_Lambda_omega_Equation_} \Lambda_a = c\frac{\sum x_i\Lambda_{\omega_i}}{{(a,a)_{\c H}}.} \end{equation} First of all, we notice that for all $a =\sum x_i\omega_i, \;\; a\neq 0,$ the triple \[ L_a,\;\; H,\;\; \Lambda := \frac{\sum x_i\Lambda_{\omega_i}}{{\sum x_i^2}} \] is a Lefschetz triple. This can be shown by an easy calculation which uses \eqref{_so5_relations_Equation_}. On the other hand, $\omega_i$ are orthogonal with respect to $(\cdot,\cdot)_{\c H}$, and therefore $(a,a)_{\c H}= c\sum x_i^2$. Therefore, \[ c\frac{\sum x_i\omega_i}{{(a,a)_{\c H}}} = \frac{\sum x_i\omega_i}{{\sum x_i^2}} \] Now, \eqref{_Lambda_a_as_lin_combi_Lambda_omega_Equation_} immediately implies \ref{_Lambda_additive_for_a_bullet_b_Lemma_}, as an easy calculation shows: \[ {(a+b,a+b)_{\c H}}\Lambda_{a+b} =^* c\sum (x_i+y_i)\Lambda_{\omega_i} = \] \[ = c\sum x_i\Lambda_{\omega_i}+c\sum y_i\Lambda_{\omega_i} =^* {(a,a)_{\c H}}\Lambda_{a} +{(b,b)_{\c H}}\Lambda_{b}, \] where $=^*$ marks an application of \eqref{_Lambda_a_as_lin_combi_Lambda_omega_Equation_}. This proves \ref{_Lambda_additive_for_a_bullet_b_Lemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Since \[ \{ (a,b)\in A_a \;\; |\;\; a\bullet b \} \] is Zariski dense in $A_2$, \ref{_Lambda_additive_for_a_bullet_b_Lemma_} has the following corollary: \hfill \corollary \label{_Lambda_additive_Corollary_} Let $a,b\in A_2$, be the elements of Lefschetz type, such that $a+b$ is also of Lefschetz type. Then \[ {(a+b,a+b)_{\c H}}\Lambda_{a+b} = {(a,a)_{\c H}}\Lambda_a + {(b,b)_{\c H}}\Lambda_b. \] $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_Lambda_additive_Corollary_} implies the following interesting result. The set $S\subset A_2$ of Lefschetz elements is preserved by homotheties. Therefore we may speak of homogeneous functions from $S$ to some linear space. Consider the function $\Lambda:\; S\longrightarrow {\goth g}_{-2}$, $x\longrightarrow \Lambda_2$. Clearly, this map is homogeneous of degree $-1$. Therefore, the map $\tilde r:\;S\longrightarrow {\goth g}_{-2}$, $x\longrightarrow (x,x)_{\c H} \Lambda_x$ is homogeneous of degree 1. By \ref{_Lambda_additive_Corollary_}, $\tilde r$ is linear on $S$. Therefore, $\tilde r$ is a restriction of a linear map $r:\; A_2 \longrightarrow {\goth g}_{-2}$ \hfill \corollary \label{_g_-2_is_quotie_of_A_2_Corollary_} The map $r:\; A_2\longrightarrow {\goth g}_{-2}$ is a surjection of linear spaces. {\bf Proof:} For all Lefschetz-Frobenius algebras $C$, ${\goth g}_{-2}(C)$ is spanned by $\Lambda_x$ for $x$ of Lefschetz type (\cite{_Lunts-Loo_}). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill To prove \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_}, we need only the formula \eqref{_Lambda_a_as_lin_combi_Lambda_omega_Equation_}. Since \[ \{ (a,b)\in A_a \;\; |\;\; a\bullet b \} \] is Zariski dense in $A_2$, it is sufficient to prove \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_}, for $a,b$ such that $a\bullet b$. Consider notation introduced in the proof of \ref{_Lambda_additive_for_a_bullet_b_Lemma_}. The formula \eqref{_Lambda_a_as_lin_combi_Lambda_omega_Equation_} together with relations \eqref{_so5_relations_Equation_} implies that \begin{equation}\label{_[Lambda_a,L_b]_via_x_y_Equation_} [ \Lambda_a, L_b ] = \bigg[ \frac{\sum x_i \Lambda_{\omega_i}}{\sum x_i^2}, \sum y_i L_{\omega_i} \bigg] = \frac{1}{\sum x_i^2} \bigg(-\sum x_i y_i H + \sum\limits_{i\neq j} x_i y_j [\Lambda_{\omega_i}, L_{\omega_j}]\bigg), \end{equation} and \begin{equation}\label{_[L_a,Lambda_b]_via_x_y_Equation_} [ L_a, \Lambda_b ] = -\bigg[ \frac{\sum y_i \Lambda_{\omega_i}}{\sum y_i^2}, \sum x_i L_{\omega_i} \bigg] = \frac{1}{\sum y_i^2} \bigg(\sum x_i y_i H - \sum\limits_{i\neq j} x_i y_j [L_{\omega_i}, \Lambda_{\omega_j}]\bigg). \end{equation} Let $c\in {\Bbb R}$ be the constant defined in the proof of \ref{_Lambda_additive_for_a_bullet_b_Lemma_}. We have \[ (a,a)_{\c H}= c \sum x_i^2,\;\; (b,b)_{\c H}= c \sum y_i^2,\;\; (a,b)_{\c H} = c \sum x_i y_i. \] Making the corresponding substitutions in \eqref{_[Lambda_a,L_b]_via_x_y_Equation_}, \eqref{_[L_a,Lambda_b]_via_x_y_Equation_}, we obtain \begin{equation}\label{_[Lambda_a,L_b]_via_()_H_and_x_i_Equation_} [ \Lambda_a, L_b ] = \frac{1}{(a,a)_{\c H}} \bigg(-(a,b)_{\c H} H + \frac{1}{c} \sum\limits_{i\neq j} x_i y_j [\Lambda_{\omega_i}, L_{\omega_j}]\bigg) \end{equation} and \begin{equation}\label{_[L_a,Lambda_b]_via_()_H_and_x_i_Equation_} [ L_a, \Lambda_b ] = \frac{1}{(b,b)_{\c H}} \bigg((a,b)_{\c H} H + \frac{1}{c} \sum\limits_{i\neq j} x_i y_j [L_{\omega_i}, \Lambda_{\omega_j}]\bigg). \end{equation} A linear combination of these equations yields \[ {(b,b)_{\c H}}[L_a, \Lambda_b]-{(a,a)_{\c H}}[\Lambda_a, L_b] = (a,b)_{\c H} H + \] \[ +\frac{1}{c} \sum\limits_{i\neq j} x_i y_j \bigg([\Lambda_{\omega_i}, L_{\omega_j}] - [L_{\omega_i}, \Lambda_{\omega_j}] \bigg) \] To prove \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_} it remains to show that \[ \sum\limits_{i\neq j} x_i y_j \bigg([\Lambda_{\omega_i}, L_{\omega_j}] - [L_{\omega_i}, \Lambda_{\omega_j}] \bigg) =0, \] which follows from the relation \[ [\Lambda_{\omega_i}, L_{\omega_j}] = [L_{\omega_i}, \Lambda_{\omega_j}], \;\;i \neq j \] from \eqref{_so5_relations_Equation_}. \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Let $s:\; {\goth g}_0\longrightarrow kH$ be a Lie algebra homomorphism of \ref{_str_of_g_0_Proposition_}. Either of equations \eqref{_[L_a,Lambda_b]_via_()_H_and_x_i_Equation_} and \eqref{_[Lambda_a,L_b]_via_()_H_and_x_i_Equation_} implies the following useful corollary: \hfill \corollary \label{_s_of_[L_a,Lambda_b]_Corollary_} Let $a,b \in S$. Then \[ s([L_a,\Lambda_b]) = \frac{(a,b)_{\c H}}{(b,b)_{\c H}} H. \] $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Now we can easily prove \ref{_g_0_computed_Theorem_}. Let $n=\dim V$. Clearly, $\dim \goth{so}(V)=\frac{n(n-1)}{2}$. Since $u:\; {\goth g}_0\longrightarrow \goth{so}(V)\oplus k$ is an epimorphism by \ref{_str_of_g_0_Proposition_}, it is sufficient to show that $\dim {\goth g}_0\leq \frac{n(n-1)}{2}+1$. The element $H$ belongs to the center of ${\goth g}_0$. Let $\bar {\goth g}_0:= {\goth g}_0/k H$ be the quotient Lie algebra, and $q:\; {\goth g}_0\longrightarrow \bar{\goth g}_0$ be the quotient map. Let $S\subset A_2$ be the set of all elements of Lefschetz type. By \ref{_g_0_gener_by_[L,Lambda]_Lemma_}, the space $\bar {\goth g}_0$ is spanned by all vectors $q([L_a,\Lambda_b])$, where $a,b\in S$. Denote the map $S\times S\longrightarrow \bar{\goth g}_0$, $a,b \longrightarrow q([L_a,\Lambda_b])$ by $\tilde\nu:\; S\times S \longrightarrow \bar{\goth g}_0$. Let $\nu:\; S\times S\longrightarrow \bar{\goth g}_0$, \[ \nu(a,b):= \frac{q([L_a,\Lambda_b])}{(b,b)_{\c H}}. \] Consider the Zariski open set $S\subset A_2$ as a space with an associative commutative group structure, which is defined by rational maps. In other words, $S$ is equipped with an addition, which is defined not everywhere, but in a Zariski open subset of $S$. This addition is induced from $A_2\supset S$, which is a linear space. By \ref{_Lambda_additive_Corollary_}, the map $\nu$ is bilinear with respect to this operation. Consider $\nu$ as a rational map from $A_2\times A_2$ to $\bar {\goth g}_0$. This rational map is also bilinear. An easy check shows that a linear rational map of linear spaces is defined everywhere. Hence, $\nu$ can be uniquely lifted to a bilinear map \[ \nu:\; A_2\times A_2 \longrightarrow \bar{\goth g}_0, \] such that the square \[ \begin{array}{ccc} S\times S & \hookrightarrow & A_2\times A_2 \\[3mm] \bigg\downarrow \nu &&\bigg\downarrow \nu \\[3mm] \bar{\goth g}_0 &\stackrel{Id}{\longrightarrow} &\bar{\goth g}_0 \end{array} \] is commutative. By \ref{_[L_a,Lambda_b]-[Lambda_a,L_b]_Proposition_}, the bilinear map $\nu:\; A_2\times A_2\longrightarrow \bar{\goth g}_0$ is skew-symmetric. Let $\eta$ be the corresponding linear map from the exterrior square of $A_2$ to $\bar {\goth g}_0$: \[ \eta:\; \bigwedge^2 A_2\longrightarrow \bar{\goth g}_0. \] Clearly, $\nu(A_2\times A_2) = \eta(\bigwedge^2 A_2)$. As \ref{_g_0_gener_by_[L,Lambda]_Lemma_} implies, $\bar{\goth g}_0$ is generated by the image of $\nu(A_2\times A_2)$. Hence, $\eta$ is an epimorphism. Therefore, $\dim \bar{\goth g}_0 \leq \dim \bigwedge^2 A_2 = \frac{n(n-1)}{2}$. We obtained an upper bound on $\dim{\goth g}_0$: \[ \dim {\goth g}_0\leq \frac{n(n-1)}{2} + 1. \] Since the Lie algebra $\goth{so}(V)\oplus k$ has the same dimension, $\frac{n(n-1)}{2}+1$ and the map $u:\; {\goth g}_0\longrightarrow \goth{so}(V)\oplus k$ is an epimorphism as we have seen previously, the map $u$ is an isomorphism. \ref{_g_0_computed_Theorem_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Consider the map $r:\; A_2\longrightarrow {\goth g}_{-2}$ constructed in \ref{_g_-2_is_quotie_of_A_2_Corollary_}. Let $r':\; A_2\longrightarrow {\goth g}_2$ be a standard isomorphism. Then $\nu$ can be defined as a composition of $r'\times r:\; A_2\times A_2 \longrightarrow {\goth g}_2\times {\goth g}_{-2}$ and a commutator map $[,]:\; {\goth g}_2\times {\goth g}_{-2}\longrightarrow {\goth g}_0$. The map $\nu$ was described in terms of the standard bilinear map $A_2\times A_2\longrightarrow \Lambda^2A_2$. This description easily implies that for all $y\in A_2$ there exists $x\in A_2$ such that $\nu(x,y)\neq 0$. Since $\nu=r'\times r\circ [,]$, the map $r$ has no kernel. It is epimorphic by \ref{_g_-2_is_quotie_of_A_2_Corollary_}. We obtained the following statement: \hfill \corollary \label{_g_-2_is_A_2_Corollary_} Consider the linear map $r:\; A_2\longrightarrow {\goth g}_{-2}$ constructed in \ref{_g_-2_is_quotie_of_A_2_Corollary_}. Then $r$ is an isomorphism of linear spaces. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \ref{_g_0_computed_Theorem_} gives a nice insight on the Hodge structures on $H^*(M)$ corresponding to various complex structures $I\in Comp$. As elsewhere, for each $I\in Comp$ we define an endomorphism $ad I\in End(A)$, $ad I(\omega)= (p-q)\sqrt{-1}\: \omega$ for all $\omega \in H^{p,q}(M)$. Let $\goth M\subset End(A)$ be a Lie algebra generated by $ad I$ for all $I\in Comp$. \hfill \lemma \label{_M_acts_by_deriva_Lemma_} The Lie algebra $\goth M$ acts on $A$ by derivations: for all $m\in \goth M$, $a,b \in A$, we have \[ m(ab) = m(a) b+ am(b). \] {\bf Proof:} For all $I\in Comp$, the operator $ad I$ is a derivation, as a calculation shows. A commutator of derivations is a derivation by obvious reasons. $\;\; \hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \theorem\label{_g_0_is_Mumf_Tate_Theorem_} The following Lie subalgebras of $End(A)$ coinside: \[ {\goth g}_0(A)\cong \goth M\oplus kH. \] {\bf Proof:} The inclusion $\goth M\subset {\goth g}_0(A)$ is implied by \ref{_[L_2,Lambda_3]=adI_Claim_}. The inclusion ${\goth g}_0(A)\subset \goth M\oplus kH$ is proven as follows. We have seen that ${\goth g}_0(A)$ is linearly spanned by $H$ and endomorphisms $[L_a,\Lambda_b]$, where $a\bullet b$, $(a,b)_{\c H}=0$, $(a,a)_{\c H}=(b,b)_{\c H}\neq 0$. Let $\c H'$ be a hyperkaehler structure such that $a\bullet_{\c H'} b$. Applying \ref{_action_SO(3)_on_Hyp_via_periods_Lemma_}, we obtain a hyperkaehler structure $\c H$, $\c H$ equivalent to $\c H'$, such that $a=P_2(\c H)$ and $b=P_3(\c H)$. Then, \ref{_[L_2,Lambda_3]=adI_Claim_} implies that $[L_a,\Lambda_b]\in \goth M$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill By trivial reasons, $H$ acts on $A$ as a derivation. Therefore, \ref{_g_0_is_Mumf_Tate_Theorem_} together with \ref{_M_acts_by_deriva_Lemma_} implies the following: \hfill \corollary \label{_g_0_derivatives_Corollary_} The Lie algebra ${\goth g}_0\subset End(A)$ acts on $A$ by derivations. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section[Computing the structure Lie algebra for the cohomology of a hyperkaehler manifold, part \ II.] {Computing the structure Lie algebra for the cohomology of a hyperkaehler manifold, \\part \ II.} \label{_computing_g_for_hyperk_pt-2_Section_} In this section, we prove the isomorphism ${\goth g}\cong \goth{so}(V, +)$. This is done as follows. In previous sections, we have computed dimensions of the components of ${\goth g}\cong {\goth g}_{-2}\oplus {\goth g}_0\oplus{\goth g}_2$. Let $n:= \dim V$. \ref{_g_-2_is_A_2_Corollary_} implies that $\dim {\goth g}_2=n$, \ref{_g_-2_is_A_2_Corollary_} implies that $\dim{\goth g}_{-2}=n$, and \ref{_g_0_computed_Theorem_} implies that $\dim{\goth g}_0 = \frac{n(n-1)}{2}+1$. Therefore, \[ \dim{\goth g}= \frac{n(n-1)}{2}+1 +2n = \frac{(n+2)(n+1)}{2}. \] A trivial computation yields \[ \dim \goth{so}(V,+)= \frac{(n+2)(n+1)}{2}. \] We see that dimensions of ${\goth g}$ and $\goth{so}(V, +)$ are equal; it is easy to see that ${\goth g}$ is isomorphic to $\goth{so}(V, +)$ as a graded linear space. We construct an isomorphism of graded linear spaces $U:\; {\goth g} \longrightarrow \goth{so}(V, +)$ and prove that it commutes with the Lie algebra operation. \hfill Let $B= B_0\oplus B_2\oplus B_4$ be the minimal graded Frobenius algebra associated with $V$, $(\cdot,\cdot)_{\c H}$. By definition, ${\goth g}(B)= \goth{so}(V,+)$. We are going to construct an isomorphism of linear spaces $U=U_{-2}\oplus U_0\oplus U_2$, \[ U:\; {\goth g}(A)\longrightarrow {\goth g}(B), \;\; U_{2i}:\; {\goth g}_{2i}(A)\longrightarrow {\goth g}_{2i}(B). \] We have canonical isomorphisms \[ u_B:\; {\goth g}_0(B)\longrightarrow \goth{so}(V)\oplus kH,\;\; u_A:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)\oplus kH. \] These isomorphisms yield a natural Lie algebra isomorphism $U_0:\; {\goth g}_0(A)\longrightarrow {\goth g}_0(B)$. The homomorphism $U_2:\; {\goth g}_2(A)\longrightarrow {\goth g}_2(B)$ is provided by the natural isomorphism ${\goth g}_2(A)\cong A_2$ (\ref{_g_2_is_A_2_Corollary_}), which exists for every Lefschetz-Frobenius algebra of Jordan type. To construct the isomorphism $U_{-2}:\; {\goth g}_{-2}(A)\longrightarrow {\goth g}_{-2}(B)$, we use \ref{_g_-2_is_A_2_Corollary_}. According to this statement, ${\goth g}_{-2}(A)$ is naturally isomorphic to $V$. The natural isomorphism ${\goth g}_{-2}(B)\cong V$ is constructed in \ref{_calculation_of_g(A)_for_minim_Theorem_}. Composing these isomorphisms, we obtain $U_{-2}$. Now, the isomorphism ${\goth g}(A)\cong \goth{so}(V, +)$ is implied by the following proposition: \hfill \proposition \label{_U_is_Lie_homomo_Proposition_} The map $U:\; {\goth g}(A)\longrightarrow {\goth g}(B)$ is a homomorphism of Lie algebras. {\bf Proof:} By our construction, the restriction of $U$ to ${\goth g}_0(A)$ is a homomorphism of Lie algebras. Therefore, it suffices to check that \begin{equation}\label{_U_commu_with_commutator_Equation_} U([X,Y])=[U(X),U(Y)] \end{equation} in the following cases: (i) $X\in {\goth g}_2$, $Y\in {\goth g}_{-2}$ (ii) $X\in {\goth g}_0$, $Y\in {\goth g}_2$ (iii) $X\in {\goth g}_0$, $Y\in {\goth g}_{-2}$. \hfill We start the proof of \eqref{_U_commu_with_commutator_Equation_} with (i). We represent $u_A:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)\oplus k H$ as a sum $u_A=t_A\oplus s_A$, where $t_A:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)$ and $s_A:\; {\goth g}_0(A)\longrightarrow k H$ are components of $u_A$. Consider the similar decomposition of $u_B:\; {\goth g}_0(B)\longrightarrow \goth{so}(V)\oplus k H$, $u_B=t_B\oplus s_B$. Let $S_A\subset A_2=V$ be the set of all elements of Lefschetz type in $A$, and $S_B\subset B_2=V$ be the set of all elements of Lefschetz type in $B$. Let $S:= S_A\cap S_B$. For $x\in S_A$ ($S_B)$, we denote by $L^A_x$, $\Lambda^A_x$ ($L^B_x$, $\Lambda^B_x$) the corresponding elements in ${\goth g}(A)$ (${\goth g}(B)$). Clearly, $S$ is Zariski open in $V$. Therefore, to prove \eqref{_U_commu_with_commutator_Equation_} in case (i) it is sufficient to show that for all $x,y \in S$, \begin{equation}\label{_U_[L,Lambda]=[U(L),U(Lambda)]_Equation_} U([L^A_x, \Lambda^A_y]) = [ U(L^A_x), U(\Lambda^A_y) ]. \end{equation} Checking the definition of $U$, one can easily see that $U(L_x^A)= L_x^B$ and $U(\Lambda_x^A)= \Lambda_x^B$. Therefore, \eqref{_U_[L,Lambda]=[U(L),U(Lambda)]_Equation_} is equivalent to \begin{equation}\label{_U_[L,Lambda]=[L^B,Lambda^B]_Equation_} U([L^A_x, \Lambda^A_y]) = [L^B_x, \Lambda^B_y ]. \end{equation} By definition of $U$, \eqref{_U_[L,Lambda]=[L^B,Lambda^B]_Equation_} is equivalent to \[ u_A([L^A_x, \Lambda^A_y]) = u_B([L^B_x, \Lambda^B_y ]). \] Using the decomposition of $u_A$, $u_B$, we obtain that this equation is implied by the following two relations: \begin{equation}\label{_t_commu_with_commutato_Equation_} t_A([L^A_x, \Lambda^A_y]) = t_B([L^B_x, \Lambda^B_y ]), \end{equation} \begin{equation}\label{_t_commu_with_commutato_another_Equation_} s_A([L^A_x, \Lambda^A_y]) = s_B([L^B_x, \Lambda^B_y ]). \end{equation} The second of these relations is implied by the equation \[ s_A([L^A_x, \Lambda^A_y]) = \frac{(x,y)_{\c H}}{(y,y)_{\c H}} H \] (\ref{_s_of_[L_a,Lambda_b]_Corollary_}). We proceed to prove \eqref{_t_commu_with_commutato_Equation_}. Consider the action of the operators $t_A([L^A_x, \Lambda^A_y]), t_B([L^B_x, \Lambda^B_y ]) \in \goth{so}(V)$ on $V$. Let $\c L$ be the two-dimensional plane generated by $x,y\in V$. Let $\c L^\bot$ be its orthogonal complement. By \ref{_[L,Lambda]_in_T_a,b_Corollary_}, $t_A([L^A_x, \Lambda^A_y])$ acts as zero on $\c L^\bot$. By \ref{_Lambda_vanish_Corollary_}, $t_B([L^B_x, \Lambda^B_y])$ also vanish on $\c L^\bot$. The space of skew-symmetric endomorphisms of $V$ which vanish on $\c L^\bot$ is one-dimensional. Hence, the operators $t_A([L^A_x, \Lambda^A_y])$ and $t_B([L^B_x, \Lambda^B_y])$ are proportional. To prove that they are equal we have to compute only the coefficient of proportionality. \hfill Denote the result of application of $\xi\in \goth{so}(V)$ to $x\in V$ by $\xi x$. To prove \eqref{_t_commu_with_commutato_Equation_} it is sufficient to show that \begin{equation}\label{_L_x_Lambda_y_to_y_Equation_} t_A([L_x^A,\Lambda_y^A])y=t_B([L_x^B,\Lambda_y^B])y\neq 0. \end{equation} In the case when $x,y$ are collinear, \[ t_A([L_x^A,\Lambda_y^A])=t_B([L_x^B,\Lambda_y^B])=0. \] (see \eqref{_t(L_y_Lambda_y)=0_Equation_}). Therefore, in this case \eqref{_L_x_Lambda_y_to_y_Equation_} is not true. However, \eqref{_t_commu_with_commutato_Equation_} is vacuously true in this case. Therefore, to prove \eqref{_t_commu_with_commutato_Equation_} it is sufficient to prove \eqref{_L_x_Lambda_y_to_y_Equation_} assuming that $x$ and $y$ are not collinear. \hfill Let $x,y\in V$ be two vectors which are not collinear. We prove \eqref{_L_x_Lambda_y_to_y_Equation_} as follows. Denote the Hodge operators in $A$, $B$ by $H^A$, $H^B$ respectively. By definition of the Lefschetz triple, \[ [L_y^A,\Lambda_y^A])= H^A, \;\;[L_y^B,\Lambda_y^B] = H^B. \] This implies that \begin{equation}\label{_t(L_y_Lambda_y)=0_Equation_} t_A([L_y^A,\Lambda_y^A])=t_B([L_y^B,\Lambda_y^B])=0 \end{equation} Let $\lambda\in R$. Since the expressions $t_A([L_x^A,\Lambda_y^A])$, $t_B([L_x^B,\Lambda_y^B])$ are bilinear by $x$, we have \[ t_A([L_{x+\lambda y}^A,\Lambda_y^A])=t_A([L_x^A,\Lambda_y^A]) \] and \[ t_B([L_{x+\lambda y}^B,\Lambda_y^B])=t_B([L_x^B,\Lambda_y^B]). \] Therefore, \eqref{_L_x_Lambda_y_to_y_Equation_} is equivalent to \[ t_A([L_{x+\lambda y}^A,\Lambda_y^A])y= t_B([L_{x+\lambda y}^B,\Lambda_y^B])y\neq 0. \] By \ref{_el-t_with_non_zero_square_Lefschetz_Lemma_}, $z\in S_B$ if and only if $(z,z)_{\c H}\neq 0$. Since $S\subset S_B$, the number $(y,y)_{\c H}$ is non-zero. Take $\lambda=\frac{(x,y)_{\c H}}{(y,y)_{\c H}}$. Then $(x+\lambda y,y)_{\c H}=0$. Replacing $x$ by $x+\lambda y$, we see that it is sufficient to prove \eqref{_L_x_Lambda_y_to_y_Equation_} in the case when $(x,y)_{\c H}=0$. Let $\mu\in {\Bbb R}$, $\mu\neq 0$. If we replace $x$ by a vector $\mu x$, both sides of \eqref{_L_x_Lambda_y_to_y_Equation_} are multiplied by the number $\mu$. Choosing the appropriate coefficient $\mu$, we may assume that $(x,x)_{\c H}=(y,y)_{\c H}>0$. We obtained that we may prove \eqref{_L_x_Lambda_y_to_y_Equation_} under the following set of assumptions: \[ (x,x)_{\c H}=(y,y)_{\c H}>0, \;\; (x,y)_{\c H}=0. \] This is implied by the following lemma. \hfill \lemma \label{_[L_x,Lambda_y]_to_y_is_-2x_Lemma_} Let $x,y\in S$, $(x,x)_{\c H}=(y,y)_{\c H}>0$, $(x,y)_{\c H}=0$. Then \hfill (i) $t_A([L_x^A,\Lambda_y^A])y =-2x$, and (ii) $t_B([L_x^B,\Lambda_y^B])y =-2x$. \hfill {\bf Proof:} (i) Let \[ T:= \bigg \{ (x,y)\in S \;\; | \;\; (x,x)_{\c H}=(y,y)_{\c H},\;\; (x,y)_{\c H}=0 \bigg\}. \] Let \[ T^\bullet:= \bigg \{ (x,y)\in S \;\; | \;\; (x,x)_{\c H}=(y,y)_{\c H}, \;\; (x,y)_{\c H}=0, \;\;x \bullet y \bigg\}. \] A standard argument with periods and comparing dimensions implies that $T^\bullet$ is Zariski dense in $T$. Therefore, we may prove (i) assuming that $x\bullet y$. Let $\tilde {\c H}$ be a hyperkaehler structure such that $x\bullet_{\tilde{\c H}} y$. Using \ref{_action_SO(3)_on_Hyp_via_periods_Lemma_}, we replace $\tilde {\c H}$ by an equivalent hyperkaehler structure $\c H=(I,J, K, (\cdot,\cdot))$ such that $P_2(\c H)=x$, $P_3(\c H) =y$. In this case \[ [ L^A_x, L^A_y] = ad I \] (\ref{_[L_2,Lambda_3]=adI_Claim_}). Let $\Omega:= x+ \sqrt{-1}\: y$. By definition of $ad I$, $adI (\Omega) = 2\sqrt{-1}\: \Omega$. This immediately implies \ref{_[L_x,Lambda_y]_to_y_is_-2x_Lemma_} (i). \hfill {\bf Proof of \ref{_[L_x,Lambda_y]_to_y_is_-2x_Lemma_} (ii):} By definition, for all $a, b\in B_2$, \begin{equation}\label{_ab_in_B_Equation_} ab= (a,b)_{\c H}\Omega_B \end{equation} for a fixed vector $\Omega_B\in B_4$. Therefore, $L_x^B y=0$. We obtain that \begin{equation}\label{_commut_applied_to_y__in_B_Equation_} [ L_x^B,\Lambda_y^B]y = L_x^B\Lambda_y^B y. \end{equation} Let ${\Bbb I}_B\in B_0$ be the unit in $B$. Then $\Lambda_y^B y= \Lambda_y^B L_y^B {\Bbb I}_B$. Since $[L_y^B,\Lambda_y^B] =H^B$ and $\Lambda_y^B {\Bbb I}_B=0$, we have \[ \Lambda_y^B L_y^B {\Bbb I}_B = - H^B (\Bbb I) = -2 {\Bbb I}_B. \] Using \eqref{_ab_in_B_Equation_}, we can easily check that \[ [L_x^B,\Lambda_y^B]{\Bbb I}_B=0. \] Therefore, \[ s_B([L_x^B,\Lambda_y^B])=0. \] This implies that the action of $[ L_x^B,\Lambda_y^B]$ on $B_2$ coinsides with the action of $t_B([ L_x^B,\Lambda_y^B])$ on $V$. Hence, \eqref{_commut_applied_to_y__in_B_Equation_} implies \ref{_[L_x,Lambda_y]_to_y_is_-2x_Lemma_} (ii). This finishes the proof of \eqref{_U_commu_with_commutator_Equation_}, case (i). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Relation \eqref{_U_commu_with_commutator_Equation_}, case (ii) is implied by the following explicit description of the commutators between ${\goth g}_0(A)$ and ${\goth g}_2(A)$, ${\goth g}_{-2}(A)$, which is valid for many Lefschetz-Frobenius algebras of Jordan type. \hfill \proposition \label{_commu_between_g_0_and_g_+-2_Proposition_} Let $C=\oplus C_i$ be a Lefschetz-Frobenius algebra of Jordan type. Let $\xi\in {\goth g}_0(C)$. Assume that ${\goth g}_0(C)$ acts on $C$ by derivations. For any $x\in C_2$, denote by $\xi(x)$ the image of $x$ under an action of $\xi:\; C_2\longrightarrow C_2$. Then $[\xi, L_x]= L_{\xi(x)}$ for all $x\in A_2$, $\xi\in{\goth g}_0$. {\bf Proof:} By definition of a derivation, $\xi(xa)= \xi(x) a +x\xi(a)$. By definition, $\xi L_x(a)= \xi(xa)$ and $L_x \xi (a)=x\xi(a)$. Substracting one from another, we obtain $[\xi, L_x](a)= L_{\xi(x)}(a)$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill According to \ref{_g_0_derivatives_Corollary_}, ${\goth g}_0(A)$ acts on $A$ by derivations. To show that ${\goth g}_0(B)$ acts on $B$ by derivations, we notice the following. Let $C$ be an associative algebra over a field and ${\goth g}$ be a Lie algebra, ${\goth g}\subset End(C)$. Consider the corresponding Lie group $G\subset End(C)$. Then ${\goth g}$ acts on $C$ by derivations if and only if $G$ acts on $C$ by algebra automorphisms. Now, ${\goth g}_0(B)\cong \goth{so}(V)\oplus kH$. The algebra $kH$ acts on $B$ by derivations for obvious reasons. On the other hand, the Lie group $SO(V)$ acts on $B$ by automorphisms, as follows from definition of $B$. Therefore, \ref{_commu_between_g_0_and_g_+-2_Proposition_} can be applied to $A$ and $B$. Let $g\in \goth{so}(V)\oplus kH$. Let $g^A$, $g^B$ be the elements of ${\goth g}_0(A)$, ${\goth g}_0(B)$ which correspond to $g$. Then, \eqref{_U_commu_with_commutator_Equation_}, case (ii) is equivalent to \[ U^A([g^A, L_x^A]) = U^B([g^B, L_x^B]),\;\; \forall x\in V. \] By \ref{_commu_between_g_0_and_g_+-2_Proposition_}, this is equivalent to \[ g^A(x)=g^B(x), \] which is clear from definitions. We proved \eqref{_U_commu_with_commutator_Equation_}, case (ii). $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill It remains to prove \eqref{_U_commu_with_commutator_Equation_} in case (iii). Consider the action of ${\goth g}_0(A)$, ${\goth g}_0(B)$ on ${\goth g}_{-2}(A)$, ${\goth g}_{-2}(B)$ by commutators. The action of ${\goth g}_0(B)$ on ${\goth g}_{-2}(B)$ consides with that on $B_2\cong V$ (we use the standard isomorphism ${\goth g}_{-2}(B)\cong V$ which is apparent from the explicit description of ${\goth g}(B)$). This means that $k H\subset {\goth g}_0(B)$ acts on ${\goth g}_{-2}(B)$ trivially, and $\goth{so}(V)\subset {\goth g}_0(B)$ acts on ${\goth g}_{-2}(B)\cong V$ in a standard fashion. Denote this action by $\rho_1:\; \goth{so}(V)\longrightarrow End(V)$. Similarly, $kH\subset {\goth g}_0(A)$ acts trivially on ${\goth g}_{-2}(A)$. It remains to compare the action of $\goth{so}(V)\subset {\goth g}_0(A)$ on ${\goth g}_{-2}(A)$ to $\rho_1$. Consider the isomorphism $r:\; A_2\longrightarrow {\goth g}_{-2}(A)$ constructed in \ref{_g_-2_is_A_2_Corollary_}. The action of ${\goth g}_0(A)$ on ${\goth g}_{-2}(A)$ defines an action of $\goth{so}(V)\subset {\goth g}_0(A)$ on $A_2$. Using the isomorphism $A_2\cong V$, we write this action as a homomorphism $\rho_2:\;\goth{so}(V)\longrightarrow End(V)$. In this notation, the equation \eqref{_U_commu_with_commutator_Equation_}, case (iii) is equivalent to the following statement: \hfill \lemma \label{_rho_1_is_rho_2_Lemma_} The representations $\rho_1$, $\rho_2$ coinside. {\bf Proof:} Let $I\in Comp$. Consider the endomorphism $ad I\in {\goth g}_0(A)$. We identify ${\goth g}_0(A)$ and ${\goth g}_0(B)$ using the isomorphism $U_0$. Let $t:\; {\goth g}_0(A)\longrightarrow \goth{so}(V)$ be a projection on a summand. Let $C\subset \goth{so}(V)$ be the union of $t(ad I)$ for all $I\in Comp$. As we have seen, $C$ is Zariski dense in $\goth{so}(V)$. Therefore it is sufficient to show that $\rho_1$, $\rho_2$ coinside on $C$. Let $x\in A_2$. By definition, $\rho_1(t(ad I))(x)= ad I(x)$, and $\rho_2(t(ad I))x = r^{-1}[ad I, r(x)]$, where $r:\; V\longrightarrow {\goth g}_{-2}(A)$ is a map of \ref{_g_-2_is_A_2_Corollary_}. To prove \ref{_rho_1_is_rho_2_Lemma_} we have to show the following: \begin{equation} \label{_r_commu_w_so(V)_Equation_} r(ad I(x)) = [ad I, r(x)]. \end{equation} Both sides of \eqref{_r_commu_w_so(V)_Equation_} are linear by $x$. Therefore it is sufficient to check \eqref{_r_commu_w_so(V)_Equation_} in two cases: \hfill (i) $x\in H^{1,1}_I(M)$ (ii) $x\in H^{2,0}_I(M)\oplus H^{0,2}_I(M)$, \hfill \hspace{-6mm}where $H^{p,q}_I(M)$ is Hodge decomposition associated with the complex structure $I\in Comp$. In case (i), $ad I(x)=0$, so \eqref{_r_commu_w_so(V)_Equation_} is equivalent to \begin{equation} \label{_adI_commu_w_r(x)_for_x_in_H^1,1_Equation_} [ad I, r(x)] =0. \end{equation} Since elements of Lefschetz type are Zariski dense in $A_2$, it is sufficient to prove \eqref{_adI_commu_w_r(x)_for_x_in_H^1,1_Equation_} assuming that $x$ is of Lefschetz type. In this case, $r(x)=(x,x)_{\c H}\Lambda_x$. Therefore, \eqref{_adI_commu_w_r(x)_for_x_in_H^1,1_Equation_} follows from the equation $[ad I, \Lambda_x] =0$. Since $x\in H^{1,1}_I(M)$, the operator $L_x$ preserves weights of the Hodge decomposition. An easy linear algebraic check insures that in this case $\Lambda_x$ also preserves Hodge weights. Therefore, by definition of $ad I$, we have $[ad I, \Lambda_x] =0$. We proved \eqref{_r_commu_w_so(V)_Equation_}, case (i). \hfill Consider a non-zero holomorphic symplectic form $\tilde \Omega$ over the complex manifold $(M,I)$. Let $\Omega\in H^2(M)$ be cohomology class represented by $\Omega$. Then $Im(\Omega)$, $Re(\Omega)$ constitute basis in two-dimensional space $H^{2,0}_I(M)\oplus H^{0,2}_I(M)$. Let $\c H = (I, J, K, (\cdot,\cdot))$ be a hyperkaehler structure such that $P_2(\c H)= Re(\Omega)$, $P_3(\c H)= Im(\Omega)$. Such $\c H$ exists by Calabi-Yau theorem. Since \ref{_r_commu_w_so(V)_Equation_} is linear by $x$, we may check \ref{_r_commu_w_so(V)_Equation_} case (ii) only for $x_2=P_2(\c H)$, $x_3=P_3(\c H)$. Clearly from definitions, \[ ad I(x_2) = 2 x_3, \; ad I(x_3) = -2 x_2. \] Let $c=(x_2,x_2)_{\c H}=(x_3,x_3)_{\c H}$. By definition, $r(x_i)=c\Lambda_{x_i}$, $i=2,3$. Therefore, \eqref{_r_commu_w_so(V)_Equation_} case (ii) is equivalent to the following pair of equations: \[ 4c \Lambda_{2x_3} = [ad I, c\Lambda_{x_2}], \] and \[ -4 c\Lambda_{2x_2} = [ad I, c\Lambda_{x_3}] \] Since $\Lambda_{2a}=1/2\Lambda_a$, these two equations can be rewritten as \begin{equation}\label{_adI_on_Lambdas_Equation_} 2\Lambda_{x_3} = [ad I, \Lambda_{x_2}], \; -2\Lambda_{x_2} = [ad I, \Lambda_{x_3}] \end{equation} \ref{_[L_2,Lambda_3]=adI_Claim_} implies that in notation of \eqref{_so5_relations_Equation_}, $ad I = K_{23}$. Therefore \eqref {_adI_on_Lambdas_Equation_} is a consequence of \eqref{_so5_relations_Equation_}. This proves \ref{_rho_1_is_rho_2_Lemma_}, and consequently, \eqref{_U_commu_with_commutator_Equation_} case (iii). Proof of \ref{_U_is_Lie_homomo_Proposition_} is finished. This also finishes the proof of \ref{_g(A)_for_hyperkae_Theorem_}, which spanned four sections of this paper. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{The structure of the cohomology ring for compact hyperkaehler manifolds.} \label{_cohomolo_compu_Section_} Let $M$ be a compact simple hyperkaehler manifold and $A=H^*(M)$ be its cohomology ring. Let $\c V=H^2(M,{\Bbb R})$ considered as a linear space with non-degenerate symmetric pairing $(\cdot,\cdot)_{\c H}$. Applying \ref{_g(A)_for_hyperkae_Theorem_} and \ref{_all_alg_with_so_are_^dA_Theorem_} to $A$, we immediately obtain the following statement. \hfill \theorem\label{_cohomo_of_hyperk_are_^dA(V)_Theorem_} Let $A^r\subset H^*(M)$ be the subalgebra of $A=H^*(M)$ generated by $A_0$, $A_2$. Then $A^r\cong {}^dA(\c V)$, where ${}^dA(\c V)$ is a Frobenius algebra considered in \ref{_all_alg_with_so_are_^dA_Theorem_} $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill We see that $A^r\subset H^*(M)$ can be expressed purely in linear-algebraic terms. In this section, we make this description as explicit as possible. Fix a linear space $V$ with a non-degenerate bilinear symmetric form $B:\; S^2V\longrightarrow k$, where $k$ is a ground field. We express ${}^dA(V)$ in terms of $V$ and $B$ as follows. Let \[ C = C_0\oplus C_2\oplus ... \oplus C_{4d} \] be a graded linear space, with \[ C_{2i} = S^iV, \;\;i\leq 2d, \] \[ C_{2i} = S^{2i-d}V, \;\;i\geq 2d. \] We describe a multiplicative structure on $C$ using some classical results of linear algebra. Let $V$ be a linear space equipped with non-degenerate bilinear symmetric product $B:\; V\otimes V\longrightarrow k$. Consider the Lie group $SO(V)$ associated with $V$ and $B$. This group naturally acts on the symmetric powers $S^nV$ for all $n$. The representation $SO(V)\longrightarrow End(S^nV)$ is not irreducible. Its irreducible decomposition is a classical result of linear algebra. We describe this decomposition explicitely, and define $SO(V)$-invariant multiplicative structure on $C$ in terms of this decomposition. Let $\goth V:= \{ x_1,...., x_n\}$ be a basis in $V$. We represent the vectors from $S^nV$ by the polynomials \[ \sum \alpha_{i_1,...,i_m} x_{i_1},..., x_{i_m}, \] where $\alpha_{i_1,...,i_m}\in k$ and $x_{i_j}\in \goth V$, $k$ is a ground field. Consider an $SO(V)$-invariant vector $r\in S^2 V$ represented by the polynomial \[ r:= \sum_{i,j} B(x_i,x_j) x_i x_j. \] Let $L_r:\; S^nV\longrightarrow S^{n+2}V$ be the map multiplying the polynomial $P$ by $r$. Since the product in $S^*V$ commutes with the $SO(V)$-action and $r$ is $SO(V)$-invariant, the map $L_r$ is a homomorphism of $SO(V)$-representations. The scalar product $B$ on $V$ can be extended to an $SO(V)$-invariant scalar product $(\cdot,\cdot)_{V^{\otimes_n}}$ on the space of $n$-tensors $\otimes^n V$ by the law \[ (x_1\otimes x_2\otimes ... \otimes x_n, y_1\otimes y_2\otimes ... \otimes y_n)_{V^{\otimes_n}} = B(x_1,y_1) B(x_2,y_2) ... B(x_n,y_n). \] The space $S^nV\subset \otimes ^n V$ is $SO(V)$-invariant. Using Schur's lemma, one can see that the restriction of $(\cdot,\cdot)_{V^{\otimes_n}}$ to $S^nV$ is non-degenerate. Denote this scalar product by $(\cdot,\cdot)_{S^n V}$. For all $n>1$, the map $L_r:\; S^{n-2}V\longrightarrow S^n V$ is an embedding. Let $R^n V\subset S^n V$ be the orthogonal complement to the image of $L_r\; S^{n-2}V\longrightarrow S^n V$. Using the embedding $L_r:\; S^{i-2}V\longrightarrow S^i V$ for different $i$,we obtain a decomposition \begin{equation}\label{_decompo_of_S^nV_Equation_} S^n V\cong R^n V\oplus R^{n-2} V \oplus ... \oplus R^{n\; \mbox{\tiny mod}\; 2} V, \end{equation} where $R^0V= k$ and $R^1 V = V$. \proposition \label{_R^nV_is_irredu_Proposition_} For all $n\in \Bbb N$, the $SO(V)$-representation $R^n V$ is irreducible. {\bf Proof:} See \cite{_Weyl_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill We obtain that \eqref{_decompo_of_S^nV_Equation_} is an irreducible decomposition of an $SO(V)$-representation $S^n V$. By definition, the representation \[ S^nV / R^n V \oplus R^{n-2}V \oplus... \oplus R^{n-2i +2}V \] is canonically isomorphic to $S^{n-2i}V$. Denote the corresponding quotient map by $B^n_{n-2i}:\; S^n V\longrightarrow S^{n-2i}V$. Using the maps $B^n_{n-2i}$ we define the multiplicative structure on $C$ as follows. Let $S^* V= \oplus S^n V$ be the algebra of symmetric tensors over $V$. Let $\phi:\; S^* V\longrightarrow C$, $\phi=\oplus \phi_i$, where $\phi_i:\; S^iV \longrightarrow C_{2i}$ is the following map. For $2i\leq 2d$, $C_{2i}\cong S^i V$. For such $i$, $\phi_i$ is defined as identity map. For $2d<2i\leq 4d$, we have $C_{2i}= S^{2d-i}V$. Let $\phi_i: S^i V\longrightarrow C_{2i}$ be equal to $B^i_{2d-i}:\; S^i V\longrightarrow S^{2d-i}V$. For $2i>4d$, $C_{2i}=0$ and we take $\phi_i=0$. Clearly, the map $\phi$ is onto. \hfill \lemma \label{_ker_phi_ideal_in_S^*V_Lemma_} Let $I$ be a kernel of $\phi:\; S^*V\longrightarrow C$. Then $I$ is an ideal in $S^*V$. {\bf Proof:} Let $x\in S^l V$, $x\in I$, and $y\in S^m V$. We have to show that $xy\in I$. This relation is vacuously true except in case when $d<l<l+m <2d$. Let $\Lambda_r:\; S^n V\longrightarrow S^{n-2} V$ be equal to $B^n_{n-2}$ for all $n = 2,3,...$. Clearly, $B^n_{n-2i}= \Lambda^i_r$, where $\Lambda^i_r$ is $\Lambda_r$ to the power of $i$. Therefore $x\in I$ is equivalent to $\Lambda^{l-d}_r(x)=0$, and $xy\in I$ is equivalent to $\Lambda^{l+m-d}_r(xy)=0$. Therefore, \ref{_ker_phi_ideal_in_S^*V_Lemma_} is a special case of the following statement. \hfill \lemma \label{_Lambda_r_to_multi_Lemma_} Let $l,m,n$ be positive integer numbers, $x\in S^l V$, $y\in S^m V$. Assume that $\Lambda^{n}_r(x) =0$. Then $\Lambda^{n+m}_r(xy) =0$. {\bf Proof:} Consider an element of $S^n V$ as polynomial function on $V$ considered as an affine space. Let $\Delta:\; S^n V\longrightarrow S^{n-2}V$ be the Laplace operator associated with the metric structure $B$ on $V$. By definition, \[ \Delta(P) = \sum B(x_i,x_j) \frac{\partial^2}{\partial x_i \partial x_j} P, \] where $x_1,..., x_n$ is a basis in $V$. This definition is independent of the choice of basis in $V$. The operator $\Delta$ commutes with an action of $SO(V)$. Checking that $\Delta(S^n V)$ contains $y^{n-2}$ for all $y\in V$, we obtain that the map $\Delta$ is onto. This imples that $\ker(\Delta^i:\; S^n V \longrightarrow S^{n-2i} V)$ coinsides with $R^n V\oplus R^{n-2} V\oplus ... \oplus R^{n-2i+2}V\subset S^n V$. Therefore, $\ker(\Delta^i)=\ker(\Lambda_r^i)= \ker(B^n_{n-2i})$. We obtain that \ref{_Lambda_r_to_multi_Lemma_} is equivalent to the following statement: \hfill \lemma \label{_laplace_multi_Lemma_} Let $x\in S^l V$, $y\in S^m V$. Assume that $\Delta^i x =0$. Then $\Delta^{i+m}(xy)=0$. {\bf Proof:} We prove \ref{_laplace_multi_Lemma_} using induction by $l$, $m$. We denote by $\bf L(l_0, m_0)$ the statement of \ref{_laplace_multi_Lemma_} applied to $l=l_0$, $m=m_0$. Clearly, \begin{equation}\label{_Laplace_of_produ_Equation_} \Delta(xy) = \Delta(x) y + x\Delta(y) + 2\sum \frac{\partial x}{\partial x_i} \frac{\partial y}{\partial x_j} B(x_i,x_j). \end{equation} By $\bf L(l-1, m)$, we have \[ \Delta^{i+m -1}(\Delta(x) y) =0. \] By $\bf L(l, m-2)$, \[ \Delta^{i+m -1}(\Delta(x) y) =0. \] Laplacian commutes with partial derivatives, and therefore $\Delta^{i}(\frac{\partial x}{\partial x_i})=0$. Hence, by the virtue of $\bf L(l, m-1)$, \[ \Delta^{i+m -1} \bigg(\sum \frac{\partial x}{\partial x_i} \frac{\partial y}{\partial x_j} B(x_i,x_j)\bigg) =0 \] Therefore, $\Delta^{i+m -1}$ applied to the right hand side of \eqref{_Laplace_of_produ_Equation_} is zero. This finishes the proof of \ref{_laplace_multi_Lemma_}. We proved \ref{_Lambda_r_to_multi_Lemma_} and \ref{_ker_phi_ideal_in_S^*V_Lemma_}. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill By definition, $C= S^*V / I$. Since $I$ is an ideal in $S^* V$, the space $C$ inherits a canonical ring structure. Denote this algebra by ${}^dC(V)$. The following theorem characterizes the $H^2(M)$-generated subring of cohomology ring of a simple compact hyperkaehler manifold in terms of $C$. \hfill \theorem \label{_^dA(V)_is_C_Theorem_} Let $V$ be a linear space equipped with bilinear symmetric pairing $B_V$. Then the algebra ${}^dA(V)$ is naturally isomorphic to ${}^dC(V)$. {\bf Proof:} We prove \ref{_^dA(V)_is_C_Theorem_} as follows. We consider ${}^dC(V)$ and ${}^dA(V)$ as graded linear spaces with an action of the group $SO(V)$. These spaces are isomorphic as graded $SO(V)$-representations. We notice that ${}^dA(V)$ is by definition a quotient of $SO(V)$ by the $SO(V)$-invariant ideal $J$. We show that there is a unique graded $SO(V)$-invartiant ideal $J$ in $S^*V$ such that $S^*V/J$ is isomorphic as a graded $SO(V)$-representation to ${}^dC(V)\cong S^*(V)/I$. This implies that $I$ coinsides with $J$, which proves \ref{_^dA(V)_is_C_Theorem_}. In Section \ref{_minimal_Fro_Section_} we considered the graded Lie algebra $\goth{so}(V,+)$. By definition, $\goth{so}(V,+)$ is a Lie algebra of skew-symmetric endomorphisms of the space $V\oplus \goth H$. Denote $V\oplus \goth H$ by $W$. The minimal Frobenius algebra $A(V)\cong {}^1A(V)$ is isomorphic to $W$ as $\goth{so}(W)$-representation. Therefore ${}^dA(V)$ is an irreducible $\goth{so}(W)$-subrepresentation of $S^d W$ generated by ${\Bbb I}\otimes{\Bbb I}\otimes...\otimes{\Bbb I}$. Consider the action of the group $SO(W)$ on ${}^dA(V)$ which corresponds to this Lie algebra action. We immediately obtain the following: \hfill \claim \label{_^dA(V)_is_R^d(W)_Claim_} Let $V$ be a linear space equipped with a non-degenerate bilinear symmetric form, and ${}^dA(V)$ be a Frobenius algebra defined in Section \ref{_^dA(V)_Section_}. Let $\goth H$ be the two-dimensional space with the hyperbolic metric, and $W:= V\oplus \goth H$. As shown above, there is a natural action of $SO(W)$ on the space ${}^dA(V)$. Earlier in this section, we defined the $SO(W)$-representation $R^d W$. Then ${}^dA(V)$ is isomorphic to $R^d W$ as a representation of $SO(W)$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill Consider a natural embedding $SO(V)\subset SO(W)$ which corresponds to the decomposition $W=V\oplus \goth H$. Consider $R^d W$ as a graded space, with the grading inherited from ${}^dA(V)$. We proceed to demonstrate that, as a graded $SO(V)$-representation, $R^d W\cong {}^dC(V)$. This is done as follows. Let $r_V\in S^2 V$, $r_W\in S^2 W$ be $SO(V)$-invariant (respectively, $SO(W)$-invariant) polynomials of degree 2defined earlier in this section. Denote the scalar product in $W$ by $B_W(\cdot,\cdot)$. Earlier we denoted the scalar product in $V$ by $B_V(\cdot,\cdot)$. Let $x_1, ..., x_n$ be a basis in $V$. Consider the vectors ${\Bbb I}$, $\Omega\in W$ which were introduced when we gave definition of $A(V)\cong W$. By definition, all vectors $x_i$ are orthogonal to ${\Bbb I}$ and $\Omega$, and $B_W({\Bbb I}, \Omega)=1$, while $B_W({\Bbb I}, {\Bbb I})=0$ and $B_W(\Omega, \Omega)=0$. Clearly, the vectors ${\Bbb I}, \Omega, x_1,...,x_n$ form a basis in $W$. By definition, \[ r_W = {\Bbb I}\Omega - r_V. \] Consider a linear map $\gamma:\; S^d W\longrightarrow S^d W$ which maps a monomial $P = {\Bbb I}^i \Omega^j T$ to \[ \gamma(P) := \begin{array}{l} r_V^i \Omega^{j-i}, \;\; j\geq i \\[5mm] {\Bbb I}^{i-j} r_V^j, \;\; j< i, \end{array} \] where $T= x_1^{\alpha_1} x_2^{\alpha_2}... x_n^{\alpha_n}$, $\sum \alpha_i= d-i-j$. Clearly, $\ker \gamma = r_W S^{d-2}W$. Therefore, the image of $\gamma$ is naturally isomorphic as a linear space to $R^dW$. Consider a multiplicative grading on $S^d W$ defined as follows: $gr(\Bbb I)=0$, $gr(x_i)=2$, $i=1,2,..., n$, $gr(\Omega)=4$. Clearly, this grading induces the standard one on the space ${}^dA(V)\cong R^dW\subset S^d V$. By definition, the map $\gamma:\; S^d W\longrightarrow S^d W$ preserves this grading. Since $\Bbb I$, $r_V$ and $\Omega$ are $SO(V)$-invariant, the map $\gamma$ commutes with an action of $SO(V)$ on $S^d W$. Therefore, $\gamma(S^d W)$ is isomorphic to $R^d W$ as a graded representation of $SO(V)$. On the other hand, $\gamma(S^d W)$ is isomorphic to ${}^d C(V)$ (again, as a graded representation of $SO(V)$ as the following argument shows. For $2i\leq 2d$, the grading-$2i$ subspace $\bigg( \gamma(S^d W)\bigg)_{2d}\subset \gamma(S^d W)$ is a linear span of monomials \[ {\Bbb I}^{d-i} x_1^{\alpha_1} x_2^{\alpha_2} ... x_n^{\alpha_n}, \;\; \sum \alpha_i = i. \] Similarly, for $2i>2d$, the space $\bigg( \gamma(S^d W)\bigg)_{2d}$ is a linear span of monomials \[ \Omega^{i-d} x_1^{\alpha_1} x_2^{\alpha_2} ... x_n^{\alpha_n}, \;\; \sum \alpha_i = 2d-i. \] Therefore, the grade $2i$ part of $R^dW\cong {}^d(W)$ is $S^iV$ for $i<d$ and is $S^{2d-i}V$ for $i>d$. We proved the following statement: \hfill \lemma \label{_C_iso_to_A(V)_as_SO(V)_repre_Lemma_} The spaces ${}^d C(V)$ and ${}^dA(V)$ are isomorphic as graded representations of $SO(V)$. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill By construction, ${}^d C(V)$ and ${}^dA(V)$ are quotient algebras of $S^*V$. The canonical epimorphisms $S^*V \longrightarrow {}^d C(V)$ and $S^*V \longrightarrow {}^d A(V)$ are $SO(V)$-invariant. Therefore, \ref{_^dA(V)_is_C_Theorem_} is a consequence of \ref{_C_iso_to_A(V)_as_SO(V)_repre_Lemma_} and the following lemma. \hfill \lemma \label{_quotie_alge_of_S^*V_iso_as_repre_iso_as_alge_Lemma_} Let $V$ be a linear space equipped with a scalar product. Let $D$, $E$ be quotients of $S^*V$ by graded ideals $I$, $J\subset S^*V$. Consider $D$, $E$ as graded algebras, with the grading inherited from $S^*V$. Assume that $I$, $J$ are $SO(V)$-invariant subspaces in $S^*V$ and $D$ is isomorphic to $E$ as a graded representation of $SO(V)$. Then $D$ is isomorphic to $E$ as an algebra. {\bf Proof:} Consider the irreducible decomposition of $S^i V$ given by \eqref{_decompo_of_S^nV_Equation_}. The summands of this decomposition are pairwise non-isomorphic. Therefore, by Schur's lemma every $SO(V)$-invariant subspace of $S^nV$ is a direct sum of several components of the decomposition \eqref{_decompo_of_S^nV_Equation_}. Let $I_n$, $J_n\subset S^n V$ be the $n$-th grade components of $I$, $J$. Since $I_n$, $J_n$ are $SO(V)$-invariant, these subspaces are direct sum of several components of \eqref{_decompo_of_S^nV_Equation_}. The quotient spaces $D_n =S^n V/I_n$, $E_n =S^n V/J_n$ are isomorphic as $SO(V)$-representations. These spaces can be identified with the direct sums of those components of the decomposition \eqref{_decompo_of_S^nV_Equation_} which don't appear in the decomposition of $I_n$, $J_n$. Since $E_n$ is isomorphic to $D_n$, the spaces $I_n$ and $J_n$ are isomorphic (as representations of $SO(V)$). Since the components of \eqref{_decompo_of_S^nV_Equation_} are pairwise non-isomorphic, the spaces $I_n$ and $J_n$ coinside. This proves \ref{_quotie_alge_of_S^*V_iso_as_repre_iso_as_alge_Lemma_}. \ref{_^dA(V)_is_C_Theorem_} is proven. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \section{Calculations of dimensions.} \label{_calcu_dimensi_Section_} We obtain easy numerical lower bounds on the dimensions \[ \dim H^{p,q}(M)\] of Hodge components of cohomology spaces of compact hyperkaehler manifolds. We have computed the part of cohomology generated by $H^2(M)$. The dimension of $\dim H^{p,q}(M)$ cannot be lower than the dimension of the space $\bar H^{p,q}(M)$ of all $(p,q)$ cohomology classes which are generated by $H^2(M)$. In this section, we compute dimensions of $\bar H^{p,q}$ for all $p$, $q$. Let $\bar H^*(M)= \oplus_{p,q}\bar H^{p,q}\subset H^*(M)$ be the subring of $H^*(M)$ generated by $H^2(M)$. Clearly, \[ \dim H^{p,q}(M) \geq \dim \bar H^{p,q}(M). \] By $p(n,m)$, we denote dimension of the space of homogeneous polynomials of degree $m$ of $n$ variables. This number is known from combinatorics as partition number. It is given by the following formal serie, which was discovered by Euler: \[ \sum p(n,m) s^n t^m = \prod\limits_{i=1}^\infty \bigg(\frac{1}{1-t^is}\bigg) \] Consider the ring $S_2$ of polynomials of $n+1$ variables, where $n$ variables are assigned degree 1 and one variable is assigned degree 2. Let $p_2(n,m)$ be the space of homogeneous polynomials of degree $m$ in $S_2$. Clearly, \[ p_2(n,m) = \sum_{i=0}^{i=[\frac{m}{2}]} p(n,m-i). \] \hfill \theorem \label{_dimens_of_bar_H_^pq_in_terms_of_p_Theorem_} Let $M$ be a simple compact hyperkaehler manifold, $\dim_{\Bbb R} M =4d$, $b_2(M)=n$ (we denote by $b_2$ the second Betti number). Then \hfill (i) $\dim \bar H^*(M) = p(n,d) - p(n, d-2)$. \hfill (ii) $\dim \bar H^{2i}(M)= \begin{array}{l} p(n,i),\;\; i\leq d\\[2mm] p(n,2d-i),\;\; i\geq d \end{array}$ \hfill (iii) $\dim\bar H^{p,q}(M)=0$ for $p+q$ odd. \hfill (iv) $\dim \bar H^{p,q}(M) = \dim \bar H^{p,2d-q}(M) =$ \\ \centerline{$=\dim \bar H^{2d-p,q}(M)=\dim \bar H^{2d-p,2d-q}(M)$.} \hfill (v) For $p+q\leq 2d$, $p\leq q$, \[ \dim \bar H^{p,q}(M) = p_2(n-2,p) \] \hfill {\bf Proof:} \hfill (i) Follows from \ref{_^dA(V)_is_R^d(W)_Claim_}. \hfill (ii) \ref{_^dA(V)_is_C_Theorem_} \hfill (iii) Clear \hfill (iv) See \cite{_so5_on_cohomo_} \hfill (v) Let $V= H^2(M)$. \ref{_^dA(V)_is_C_Theorem_} implies that $\bar H^{2m}(M) \cong S^m V$. Clearly, the Hodge decomposition on $\bar H^{2m}(M)= S^mV$ is induced from that on $V= H^{2,0}(M)\oplus H^{1,1}(M)\oplus H^{0,2}(M)$. The spaces $H^{2,0}(M)$ and $H^{0,2}(M)$ are one-dimensional. Let $z\in H^{2,0}(M)$, $\bar z \in H^{0,2}(M)$ be the non-zero vectors, and $z,\bar z, x_1,...,x_{n-2}$ be the basis in $H^2(M)\cong V$. Then the space $\bar H^{p,q}(M)$, $p+q\leq 2d$, $p\leq q$, is a linear span of the monomials \[ T_{a,b,\alpha_1,...\alpha_{n-2}} = z^a \bar z^b x_1^{\alpha_1} x_2^{\alpha_2}... x_{n-2}^{\alpha_{n-2}} \] where $b-a=q-p$, $\sum \alpha_i = p+q-(b+a)$. Let $\Theta=z \bar z\in S^2 V$. Then \[ T_{a,b,\alpha_1,...\alpha_{n-2}} = \Theta^{a} \bar z^{b-a} x_1^{\alpha_1} x_2^{\alpha_2}... x_{n-2}^{\alpha_{n-2}}, \] where $a+b + \sum \alpha_i = p+q$. Since $b-a= q-p$, \[ T_{a,b,\alpha_1,...\alpha_{n-2}} = \Theta^{a} \bar z^{p-q} x_1^{\alpha_1} x_2^{\alpha_2}... x_{n-2}^{\alpha_{n-2}}, \] where $2a + (q-p) +\sum \alpha_i = p+q$. Translating $q-p$ to the right hand side, we obtain that $T_{a,b,\alpha_1,...\alpha_{n-2}}$ is numbered by the different combinations of $a, \alpha_1,...\alpha_{n-2}$, which satisfy the condition $2a +\sum \alpha_i = 2p$. This number is by definition $ p_2(n-2,p)$. $\;\;\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill {\bf Acknowledgements:} I am very grateful to my advisor David Kazhdan for warm support and encouragement. This paper owes its existence to F. A. Bogomolov and A. Todorov, whose studies of hyperkaehler manifolds made this work possible. The last parts of this paper were inspired by joint work with Valery Lunts. I am extremely grateful to Pierre Deligne, who was most kind and helpful. I owe Deligne several important corrections in the final version of this manuscript. Thanks due to B. A. Dubrovin, P. Etingof, R. Bezrukavnikov, L. Positsel'sky, A. Todorov, A. Polishchuk, T. Pantev and M. Bershadsky for stimulating discussions. Prof. Y.-T. Siu and Prof. J. Bernstein kindly answered the questions vital for the development of this work. I am also grateful to MIT math department for allowing me the use of their computing facilities. \hfill
1996-03-05T06:15:37
9501
alg-geom/9501003
en
https://arxiv.org/abs/alg-geom/9501003
[ "alg-geom", "math.AG" ]
alg-geom/9501003
Martin Pikaart
Martin Pikaart and Johan de Jong
Moduli of curves with non-abelian level structure
25 pages, latex, only hand-drawn figures.
null
null
null
null
Following Deligne and Mumford we construct a coarse moduli space of smooth curves with non-abelian level structure, involving higher order commutators. We prove that its Deligne-Mumford compactification is smooth over an open part of Spec(${\msy Z}$).
[ { "version": "v1", "created": "Mon, 9 Jan 1995 12:02:53 GMT" } ]
2015-06-30T00:00:00
[ [ "Pikaart", "Martin", "" ], [ "de Jong", "Johan", "" ] ]
alg-geom
\section{Introduction} Deligne and Mumford introduced the moduli stack ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ parametrizing smooth genus $g$ curves with Teichm{\"u}ller structure of level $G$, a finite group. For example, if $G=\{e\}$, resp.\ $G\cong (\msy Z/n\msy Z)^{2g}$, this moduli stack is just the moduli stack of smooth curves ${\cal M}_g} \def\Mg{M_g$, resp.\ of smooth curves with abelian level structure (sometimes denoted ${\cal M}_g} \def\Mg{M_g[n]$). They also defined $\overline{{\cal M}_g}} \def\bMg{\overline{M_g}$, the moduli stack of stable curves and proved it is proper over ${\rm Spec}(\msy Z)$. Taking the normalization of $\overline{{\cal M}_g}} \def\bMg{\overline{M_g}\times {\rm Spec}(\msy Z[1/\# G])$ in the function field of ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ defines $\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}$, proper over ${\rm Spec}(\msy Z[1/\# G])$. Let $\Pi$ denote the standard fundamental group of a compact Riemann surface of genus $g$. The nth powers together with the kth order commutators generate a normal subgroup $\Pi^{(k),n}$. (We regard $[a,b]$ as a commutator of order 2.) Let $G$ be the quotient $\Pi/\Pi^{(k+1),n}$. We show that if $k\geq 1$, $n\geq 3$ the coarse moduli scheme $\bGMg$ for $\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}$ exists, and we actually have $\bGMg\cong\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}$. The following theorem is our main result. \medskip \noindent{\bf Theorem\enspace \ref{glad}} {\sl Suppose $k\in \{1,2,3\}$ and $n\geq 3$. The structural morphism $\bGMg\to {\rm Spec}(\msy Z[1/n])$ is smooth if and only if \begin{itemize} \item{$k=1$} and $g=2$, \item{$k=2$} and $n$ is odd, \item{$k=3$} and $n$ is odd or $n$ is divisible by $4$. \end{itemize} Furthermore, if $k \geq 4$, $n \geq 3$ and $n$ relatively prime to $6$, then $\bGMg\to {\rm Spec}(\msy Z[1/n])$ is smooth.} \medskip Looijenga constructed a smooth and compact cover $\overline{M_g[_2^n]}^{an},\ n\geq 3$ of $\bMgan$ using Prym level structures. He proves that these coverings are universally ramified along the boundary of ${\cal M}^{an}_g} \def\Mgan{M_g^{an}$, see \cite{Looijenga}. This holds also for our coverings as can be seen from \ref{mono}. These coverings can be applied to the construction of the Chow rings of the stacks $\overline{{\cal M}_g}} \def\bMg{\overline{M_g}[n]$: for example it is clear that the specialization maps are ring homomorphisms. The paper is organized as follows. Section 2 deals with the definition of the moduli problem. The formulation is in terms of the relative fundamental group of a family of curves. Its final proposition states that $\bGMg\to {\rm Spec}(\msy Z[1/n])$ is smooth if and only if the associated compact analytic space $\overline{{}_GM_g}^{an}$ is is a manifold. Section 3 contains a precise description of the monodromy along the boundary of $\overline{{}_GM_g}^{an}$ for general groups $G$ and the statement of the main result. Sections 4 and 5 are of a topological nature. We describe the monodromy on the relative fundamental group of the universal curve along the boundary in terms of Dehn twists. We have to describe the situation in some detail in order to understand how this monodromy acts on the finite quotients $\Pi/\Pi^{(k),n}$. These considerations prove the main theorem, using Section 6, which contains the necessary computations for a free group on three generators. The authors were stimulated by the article \cite{Looijenga}. We thank Prof.~Looijenga for numerous discussions explaining his and other results. We thank Prof.~Oort, who remarked that it should be possible to do everything algebraically and drew our attention to the article \cite{Oda}. \subsection{Notations and conventions}\label{notation} \begin{enumerate} \item Throughout the paper $g$ is a fixed natural number at least 2. \item $G$ is a finite group. \item $\msy L$ is a set of primes, $\msy L$ contains the primes dividing $\# G$. \item The stack of stable curves of genus $g$ is denoted $\overline{{\cal M}_g}} \def\bMg{\overline{M_g}$, the open substack of smooth curves ${\cal M}_g} \def\Mg{M_g$. Stacks are denoted by script letters. For definitions and results concerning stacks we refer to \cite{DM}. \item Suppose $\Gamma$ is a (pro-finite) group. A characteristic subgroup of $\Gamma$ is a normal subgroup fixed by any automorphism of $\Gamma$. A characteristic quotient is one whose kernel is a characteristic subgroup. If $\Gamma$ is profinite and topologically finitely generated then it is the direct limit of its finite characteristic quotients. \item Let $\Gamma$ denote a (finitely generated) group. For $a,b\in \Gamma$ we put $[a,b]=a^{-1}b^{-1}ab$, so that $ab=ba[a,b]$. We define the lower central series $\Gamma^{(k)}$ of $\Gamma$ by $\Gamma^{(1)}=\Gamma$ and $\Gamma^{(k+1)}=[\Gamma^{(k)},\Gamma]$. The subgroup of $\Gamma$ generated by nth-powers is denoted $\Gamma^n$. We write $\Gamma^{(k),n}$ to indicate the subgroup generated by $\Gamma^{(k)}$ and nth-powers, $\Gamma^{(k),n}=\Gamma^{(k)}\cdot \Gamma^n$. Any group homomorphism maps commutators to commutators and nth-powers to nth-powers, hence preserves these subgroups. In particular, the subgroups $\Gamma^{(k),n}$ are characteristic subgroups of $\Gamma$. \item Let $\Pi=\Pi_g$ denote the standard fundamental group $\Pi=\pi_1(S)$ of a compact Riemann surface $S$ of genus $g$. \end{enumerate} \section{Definition of the moduli problem} In this section we recall the definition of the moduli problem of Teichm\"uller level structures, see \cite[ Section 5]{DM}. Furthermore, we prove that the Deligne-Mumford compactification of the associated stack is smooth if and only if the corresponding analytic orbifold is smooth. \subsection{The relative fundamental group} In this section we define the relative fundamental group for a proper smooth morphism $f: X\to S$ with connected fibres and endowed with a section $s: S\to X$. In order to motivate the definition in the algebraic case (and since we need it also) we first do the analytic case. \subsubsection{The analytic case} Here $f:X\to S$ is a proper smooth morphism of analytic spaces with connected fibres. In addition we are given a section $s:S\to X$ of $f$. We define a locally constant sheaf of groups $\pi_1(X/S,s)$ over $S$ such that for all points $p\in S$ we have an isomorphism of groups $$\pi_1(X/S,s)_p\cong \pi_1(X_p,s(p)),$$ of the fibre of the sheaf at $p$ with the topological fundamental group of the fibre $X_p$ of $f$ at $p$ with base point $s(p)$. To construct $\pi_1(X/S,s)$ we choose for any point $p$ of $S$ a connected open neighbourhood $U_p\subset S$ and a topological isomorphism $$\phi_p: f^{-1}(U_p)\cong X_p\times U_p.$$ Such can be found compatible with $f$ and the projection to $U_p$, inducing the identity on $X_p$ and such that $\phi_p\circ s$ equals $q\mapsto (s(p), q)$. Over $U_p$ we take $\pi_1(X/S,s)$ constant with fibre $\pi_1(X_p,s(p))$. To glue these we note that given two points $p_1, p_2\in S$ there is for any $q\in U_{p_1}\cap U_{p_2}$ an identification $$X_{p_1}=X_{p_1}\times \{q\} \mapright{\phi_{p_1}^{-1}} X_q \mapright{\phi_{p_2}} X_{p_2}\times \{q\}=X_{p_2}.$$ This identification is compatible with base points $s(p_i)$ and depends continuously on $q\in U_{p_1}\cap U_{p_2}$. This means that the induced isomorphism $$ \pif{p_1}\cong \pif{p_2}$$ is constant on the connected components of $U_{p_1}\cap U_{p_2}$. Hence we get the desired gluing. We leave to the reader the trivial verification that these gluings satisfy the desired cocycle condition on $U_{p_1}\cap U_{p_2}\cap U_{p_3}$. If we choose other $\phi_p$, say $\phi'_p$, then for $q\in U_p$ the map $X_q\to X_p\times \{q\}\to X_q$, using first $\phi_p^{-1}$ then $\phi'_p$, is homotopic to the identity. Hence the resulting sheaves $\pi_1(X/S,s)$ are canonically isomorphic. A similar argument deals with the shrinking of the neighbourhoods $U_p$. \begin{proposition}The construction given above defines a locally constant sheaf of groups $\pi_1(X/S,s)$ over $S$, characterized by the following properties: \begin{enumerate} \item For any point $p\in S$ there is given an isomorphism $\pi_1(X/S,s)_p\cong \pi_1(X_p,s(p))$. \item The monodromy action of $\pi_1(S,p)$ on the fibre of the locally constant sheaf $$\rho : \pi_1(S,p)\to {\rm Aut}\big(\pi_1(X/S,s)_p\big) $$ agrees, via the isomorphism of 1), with the action of $\pi_1(S,p)$ on $\pi_1(X_p,s(p))$ deduced from the split exact sequence $$1\longrightarrow \pi_1(X_p,s(p)) \longrightarrow \pi_1(X, s(p)) \lower3pt\hbox{$\longrightarrow$}\llap{\raise3pt\hbox{$\mapleft{s_\ast}$}} \pi_1(S,p)\longrightarrow 1.$$ \item The construction of $\pi_1(X/S,s)$ commutes with arbitrary base change $S'\to S$. \end{enumerate}\end{proposition} \begin{proof}Suppose $\gamma\in \pi_1(S,p)$ and $\alpha\in \pi_1(X_p,s(p))$. It is well known and easy to prove that $s_\ast(\gamma)\alpha s_\ast(\gamma^{-1})$ is equal to the horizontal transportation of the loop $\alpha$ over $\gamma$. Clearly this describes the monodromy representation for the sheaf $\pi_1(X/S,s)$. The proof of the other assertions is left to the reader.\end{proof} \begin{remark}\label{innerconjugation} {\rm Suppose $s':S\to X$ is a second section of $f$. In general $\pi_1(X/S,s)$ is not isomorphic to $\pi_1(X/S,s')$. However, locally on $S$, say over $U\subset S$, we can choose a homotopy $H$ between $s$ and $s'$. This will induce an identification $$ i_H: \pi_1(X/S,s)|_U\cong \pi_1(X/S,s')|_U.$$ This is unique up to an {\it inner automorphism} of $\pi_1(X/S,s)$. Indeed, if $H'$ is another such homotopy, then combining $H$ and $H'$ gives a familly of loops in $X$ over $U$, with base points $s(u)$, i.e., a section of $\pi_1(X/S,s)$ over $U$. The map $(i_{H'})^{-1}\circ i_H$ is equal to conjugation with this section.}\end{remark} \subsubsection{The algebraic case} Here we consider a proper smooth morphism of schemes $f:X\to S$ with connected geometric fibres. As before we have a section $s:S\to X$ of the morphism $f$. Further, we assume given a set of primes $\msy L} \def\cC{{\cal C}$ such that all residue characteristics of $S$ are {\it not} in $\msy L} \def\cC{{\cal C}$. We recall some general notations concerning algebraic fundamental groups. We refer to \cite{SGA1} and \cite{Murre} for more details. If $Y$ is a scheme, then ${\rm \acute Et}(Y)$ denotes the category of finite \'etale coverings $Y'\to Y$. If ${\bar p}$ is a geometric point of $Y$ then we denote by $F_{Y,{\bar p}}$ the fundamental functor (or fibre functor) $ F_{Y,{\bar p}} : {\rm \acute Et}(Y)\longrightarrow Set $ which associates to $Y'$ the set of geometric points ${\bar p}\to Y'$ lying over ${\bar p}\to Y$. By definition we have $\pi_1(Y,{\bar p})={\rm Aut}(F_{Y,{\bar p}})$; this is the fundamental group of $Y$ with base point ${\bar p}$. To compare the fundamental groups with base points ${\bar p}$, resp.\ ${\bar q}$ we use a path from ${\bar p}$ to ${\bar q}$, i.e., an isomorphism of fibre functors $ \alpha: F_{Y,{\bar p}}\longrightarrow F_{Y,{\bar q}}.$ Obviously, $\alpha$ gives an isomorphism $\alpha_\ast : \pi_1(Y,{\bar p})\longrightarrow \pi_1(Y,{\bar q}).$ We note that it is independent of the choice of $\alpha$ up to conjugation. A morphism of schemes $h: Y\to Z$, defines a functor $h^\ast : {\rm \acute Et}(Z)\to {\rm \acute Et}(Y)$, which satisfies $F_{Y,{\bar p}}=F_{Z,h({\bar p})}\circ h^\ast$. Therefore we get $h_\ast$ on loops and on paths. A slight modification of the above gives $\pip Y{\bar p}$, the algebraic fundamental group classifying Galois coverings of degree in $\msy L} \def\cC{{\cal C}$. Formally it can be defined as $$\pip Y{\bar p} = \lim_{{\longleftarrow}} G, $$ where the limit is taken over all surjections $\pi_1(Y,{\bar p})\to G$ onto finite groups $G$ whose orders have only prime factors from $\msy L} \def\cC{{\cal C}$. In the sequel we will use the following results from \cite{SGA1}: The sequence $$\pi_1(X_{\bar p},s({\bar p}))\longrightarrow \pi_1(X,s({\bar p})) \longrightarrow \pi_1(S,{\bar p})\longrightarrow 1 $$ is exact. If we take the pushout of this sequence with the surjection $\pif{\bar p}\to\pip {X_{\bar p}}{\bar p}$ then the sequence also becomes left exact $$ 1\longrightarrow \pip {X_{\bar p}}{\bar p} \longrightarrow\pi'_1(X,s({\bar p})) \lower3pt\hbox{$\longrightarrow$}\llap{\raise3pt\hbox{$\mapleft{s_\ast}$}} \pi_1(S,{\bar p})\longrightarrow 1.\eqno{(*)}$$ See \cite[Expos\'e XII 4.3, 4.4]{SGA1}. As before the section $s$ defines a splitting $s_\ast : \pi_1(S,{\bar p})\to \pi_1(X,s({\bar p}))$. \begin{proposition}(\cite[Expos\'e XII 4.5]{SGA1})\label{piprel} There is a pro-object in the category of locally constant sheaves of groups on $S_{\acute et}$, denoted $\pi_1^{\L}(X/S,s)$, determined up to unique isomorphism by the following properties: \begin{enumerate} \item For any geometric point ${\bar p}$ of $S$ there is given an isomorphism $$\pi_1^{\L}(X/S,s)_{\bar p}\longrightarrow \pip {X_{\bar p}}{s({\bar p})}.$$ \item The monodromy presentation $$\rho : \pi_1(S,{\bar p})\longrightarrow {\rm Aut}\big(\pi_1^{\L}(X/S,s)_{\bar p}\big)$$ equals, via 1), the action of $\pi_1(S,{\bar p})$ on $\pip {X_{\bar p}}{s({\bar p})}$ deduced from {\rm (*)}. \item The construction of $\pi_1^{\L}(X/S,s)$ commutes with arbitrary base change $S'\to S$.\end{enumerate}\end{proposition} \begin{proof}Of course we may assume that $S$ is connected. Take a geometric point ${\bar p}$ of $S$. First we note that, since $\pi_1^{\L}(X_\p, s(\p))$ is topologically finitely generated, it is the direct limit of its characteristic finite quotients: $$\pi_1^{\L}(X_\p, s(\p))=\lim_{{\scriptstyle \longleftarrow}\atop {\scriptstyle \omega}} G_\omega$$ The action of $\pi_1(S,{\bar p})$ on $\pi_1^{\L}(X_\p, s(\p))$ deduced from (*) gives an action $\rho_\omega$ on each $G_\omega$. This defines a finite locally constant \'etale sheaf ${\cal F}_\omega$ on $S_{\acute et}$ whose fibre in ${\bar p}$ is given by $G_\omega$ and monodromy action equal to $\rho_\omega$. We put $$ \pi_1^{\L}(X/S,s)=\lim_{{\scriptstyle \longleftarrow}\atop {\scriptstyle \omega}} {\cal F}_\omega.$$ This immediately gives 1) and 2) for our chosen point ${\bar p}$. Part 3) is also clear if there exists a lift of ${\bar p}$ to a geometric point of $S'$, see \cite{SGA1}. Thus it suffices to prove 1) and 2) for a second geometric point ${\bar q}$ of $S$. Note that by definition $$\pi_1(X,s({\bar p}))={\rm Aut}\big(F_{S,{\bar p}}\circ s^\ast\big).$$ Hence if we choose an isomorphism $\alpha : F_{S,{\bar p}}\to F_{S,{\bar q}}$ then we get a commutative diagram $$\matrix{\pi_1(X, s({\bar p}))& \lower3pt\hbox{$\longrightarrow$}\llap{\raise3pt\hbox{$\mapleft{s_\ast}$}}& \pi_1(S,{\bar p})\cr \mapdown{\alpha_\ast}&&\mapdown{\alpha_\ast}\cr \pi_1(X, s({\bar q}))& \lower3pt\hbox{$\longrightarrow$}\llap{\raise3pt\hbox{$\mapleft{s_\ast}$}}& \pi_1(S,{\bar q}).\cr}$$ By (*) this induces an isomorphism $$ \alpha^{\msy L} \def\cC{{\cal C}} : \pi_1^{\L}(X_\p, s(\p)) \longrightarrow \pi_1^{\msy L} \def\cC{{\cal C}}(X_{\bar q}, s({\bar q})).$$ This is an isomorphism compatible with the actions of $\pi_1(S,{\bar p})$ and $\pi_1(S,{\bar q})$, compared via $\alpha_\ast$. The fibre of $\pi_1^{\L}(X/S,s)$ at ${\bar q}$ is by definition $$F_{S,{\bar q}}\big(\pi_1^{\L}(X/S,s)\big)\mapright{\alpha^{-1}} F_{S,{\bar p}}\big(\pi_1^{\L}(X/S,s)\big)= \pi_1^{\L}(X_\p, s(\p)).$$ If we use $\alpha^{\msy L} \def\cC{{\cal C}}$ to identify this with $\pi_1^{\msy L} \def\cC{{\cal C}}(X_{\bar q}, s({\bar q}))$ then we see by the above that the monodromy action on this exactly corresponds to the action deduced from (*) (with ${\bar q}$ in stead of ${\bar p}$). We leave to the reader the verification that another choice of $\alpha$ gives the same identification $$\pi_1^{\L}(X/S,s)_{\bar q}\longrightarrow \pi_1^{\msy L} \def\cC{{\cal C}}(X_{\bar q}, s({\bar q})).$$\end{proof} Suppose $s': S\to X$ is a second section of $f$. Take a geometric point ${\bar p}$ of $S$. We write $i_{\bar p}$ for the morphism $X_{\bar p}\to X$. We say that a path $\beta$ on $X$ connecting $s({\bar p})$ to $s'({\bar p})$ lies in $X_{\bar p}$ if there exists a path $\tilde \beta$ in $X_{\bar p}$ such that $i_{{\bar p},\ast}(\tilde \beta)=\beta$. We remark that any path $\beta$ connecting $s({\bar p})$ to $s'({\bar p})$ lies in $X_{\bar p}$ if and only if $f_\ast(\beta)=1$ (in $\pi_1(S,{\bar p})$). This is easily seen using the first exact sequence above. Let us take such a $\beta$ lying in $X_{\bar p}$. It gives rise to a commutative diagram $$\matrix{1&\longrightarrow& \pi_1^{\L}(X_\p, s(\p))&\longrightarrow&\pi'_1(X,s({\bar p}))&\longrightarrow&\pi_1(S,{\bar p})& \longrightarrow&1\cr &&\mapdown{\cong}&&\mapdown{\beta_\ast}&&\mapdown{{\rm id}}&&\cr 1&\longrightarrow&\pi_1^{\msy L} \def\cC{{\cal C}}(X_{\bar p}, s'({\bar p})) &\longrightarrow&\pi'_1(X,s'({\bar p}))&\longrightarrow&\pi_1(S,{\bar p})& \longrightarrow&1\cr}$$ The isomorphism $\pi_1^{\L}(X_\p, s(\p))\cong \pi_1^{\msy L} \def\cC{{\cal C}}(X_{\bar p}, s'({\bar p}))$ determined in this way is unique up to inner conjugation. We constructed $\pi_1^{\L}(X/S,s)$, resp.\ ${\pi_1^{\msy L} \def\cC{{\cal C}}(X/S,s')}$ as the limit of sheaves ${\cal F}_\omega$, resp.\ ${\cal F}'_\omega$ corresponding to characteristic quotients $\pi_1^{\L}(X_\p, s(\p))\to G_\omega$, resp.\ ${\pi_1^{\msy L} \def\cC{{\cal C}}(X/S,s')}\to G_\omega$. Note that the isomorphisms $G_\omega\to G'_\omega$ induced from the above are also unique up to inner conjugation. Let $S_\omega\to S$ be the finite \'etale covering of $S$ trivializing the action of $\pi_1(S,{\bar p})$ on both $G_\omega$ and $G'_\omega$. We can use the above to get an isomorphism of (constant) sheaves of groups $$ {\cal F}_\omega|_{S_\omega}\longrightarrow {\cal F}'_\omega|_{S_\omega}.$$ We claim this isomorphism is unique up to inner conjugation. This is clear if we only change $\beta$, but what happens if we change ${\bar p}$ to ${\bar q}$? Take $\alpha : F_{S,{\bar p}}\to F_{S,{\bar q}}$ as in the proof of Proposition \ref{piprel}. What we have to check is that $s'_\ast(\alpha)\circ\beta\circ s_\ast(\alpha^{-1})$ is a path connecting $s({\bar q})$ to $s'({\bar q})$ lying in $X_{\bar q}$. But this is clear since $f_\ast\big(s'_\ast(\alpha)\circ\beta\circ s_\ast(\alpha^{-1})\big) =\alpha\circ \alpha^{-1}=1.$ \begin{corollary}\label{invariance}The construction above defines locally in the \'etale topology on $S$ identifications of the finite quotients of the sheaves $\pi_1^{\L}(X/S,s)$ and ${\pi_1^{\msy L} \def\cC{{\cal C}}(X/S,s')}$. These identifications are unique up to inner conjugation and agree via 1) of Proposition \ref{piprel} with the usual identifications of $\pip {X_{\bar p}}{s({\bar p})}$ and $\pip {X_{\bar p}}{s'({\bar p})}$.\end{corollary} \subsection{Exterior homomorphisms} In this section we define the sheaf of exterior homomorphisms of the relative fundamental group of $X$ over $S$ into a fixed finite group $G$. As in \cite[5.5]{DM} this sheaf will be denoted ${\cal H}om^{ext}(\pi_1(X/S),G)$ and will be a finite locally constant sheaf of sets on $S$ (or $S_{\acute et}$). We note that $\pi_1(X/S)$ has not been defined. \subsubsection{The analytic case} Here $f:X\to S$ is a proper smooth morphism of analytic spaces with connected fibres. If $f$ has a section $s$ then we can look at the locally constant sheaf $$ {\cal F}={\cal H}om(\pi_1(X/S,s), G)$$ on $S$. It has finite fibres since $\pi_1(X_p,s(p))$ is finitely generated for all $p$ in $S$. There is a natural action of the sheaf $\pi_1(X/S,s)$ on the sheaf ${\cal F}$ given by conjugation. We define the sheaf of exterior homomorphisms as the quotient of ${\cal F}$ by this action $$ {\cal H}om^{ext}(\pi_1(X/S),G):={\cal F}\big/\pi_1(X/S,s) .$$ It is clear from Remark \ref{innerconjugation} that the right hand side does not depend on the chosen section $s$. In general, we choose an open covering $S=\bigcup U_i$ such that $X\to S$ has a section over each $U_i$. The sheaf ${\cal H}om^{ext}(\pi_1(X/S),G)$ is then defined by gluing the sheaves constructed above. The fibres are described by the formula: $${\cal H}om^{ext}(\pi_1(X/S),G)_p={\rm Hom}(\pi_1(X_p,q),G)\big/ \pi_1(X_p,q), $$ where $q$ is any point of the fibre $X_p$. Note that the monodromy action of $\gamma\in \pi_1(S,p)$ on this is given by horizontal transport of loops in $\pi_1(X_p,q)$. \subsubsection{The algebraic case} Here we assume that $S$ is a scheme over ${\rm Spec}(\msy Z[1/\#G])$. The morphism $f:X\to S$ is still assumed proper smooth with connected geometric fibres. The construction of ${\cal H}om^{ext}(\pi_1(X/S),G)$ in this case is exactly the same as for the analytic case. We use $\pi_1^{\L}(X/S,s)$, where $\msy L} \def\cC{{\cal C}$ is the set of primes dividing $\# G$, and we use that sections of $f$ exist locally in the \'etale topology on $S$. We use also Corollary \ref{invariance}. The geometric fibres of the resulting sheaf can be described as follows: $${\cal H}om^{ext}(\pi_1(X/S),G)_{\bar p}={\rm Hom}(\pi^{\msy L} \def\cC{{\cal C}}_1(X_{\bar p},{\bar q}),G)\big/ \pi^{\msy L} \def\cC{{\cal C}}_1(X_{\bar p},{\bar q})= {\rm Hom}(\pi_1(X_p,q),G)\big/ \pi_1(X_p,q), $$ the last equality holds in view of our definition of $\msy L} \def\cC{{\cal C}$. This justifies dropping $\msy L} \def\cC{{\cal C}$ from the notations. \subsubsection{Comparison} We note that if $S$ is a scheme of finite type over $\msy C$ there is a canonical homomorphism $\pi_1(X^{an}/S^{an},s)\to \pi_1^{\L}(X/S,s)^{an}$, identifying the relevant finite quotient sheaves. Clearly this gives rise to an identification of sheaves of exterior homomorphisms. \subsection{Teichm\"uller level structures} In this section the morphism $f:X\to S$ will be a familly of smooth projective curves of genus $g$. The abstract finite group $G$ will be fixed. In both the analytic case and the algebraic case we make the following definition. \begin{definition}{\rm \cite[5.6]{DM}} A Teichm\"uller structure $\alpha$ of level $G$ on $X\to S$ is a surjective exterior homomorphism $$\alpha\in \Gamma(S, {\cal H}om^{ext}(\pi_1(X/S),G)).$$ Thus locally on $S$ (resp.\ $S_{\acute et}$) $\alpha$ corresponds to a surjective homomorphism $\pi_1(X/S,s)\to G$.\end{definition} We want to consider the moduli spaces parametrizing smooth stable curves of genus $g$ with a Teichm\"uller structure of level $G$. However, as usual, it is more convenient to work with stacks. Thus let ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ denote the stack whose category of sections over the scheme $S$ (lying over ${\rm Spec}(\msy Z[1/\#G])$) is the category of smooth stable curves $X\to S$ of genus $g$ endowed with a Teichm\"uller structure of level $G$, see \cite[Section 5]{DM}. The construction of ${\cal H}om^{ext}(\pi_1(X/S),G)$ shows that the stack ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ is representable finite \'etale over ${\cal M}_g} \def\Mg{M_g[1/\#G]$. Thus ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ is a separated algebraic stack, smooth over ${\rm Spec}(\msy Z[1/\#G])$. In a similar way we define the analytic stack ${}_G{\cal M}^{an}_g} \def\GMgan{{}_GM_g^{an}$ classifying complex curves with a Teichm\"uller structure of level $G$. By comparing the stacks ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ and ${}_G{\cal M}^{an}_g} \def\GMgan{{}_GM_g^{an}$ with the stack ${\cal M}_g} \def\Mg{M_g$ and using the known result for $M_g$ one derives easily the following result. \begin{theorem}\label{existence}With notations as above. \begin{enumerate} \item A coarse moduli scheme ${}_GM_g$ for ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ exists. It is separated of finite type over ${\rm Spec}(\msy Z[1/\#G])$. \item A coarse analytic moduli space ${}_GM_g^{an}$ for ${}_G{\cal M}^{an}_g} \def\GMgan{{}_GM_g^{an}$ exists. It is canonically isomorphic to the analytic space associated to the complex variety ${}_GM_g\otimes \msy C$. \end{enumerate}\end{theorem} Suppose we have a surjection $G\to G'$. There is a natural forgetful morphism ${}_G{\cal M}_g} \def\GMg{{}_GM_g\to {}_{G'}{\cal M}_g} \def\Mg{M_g$. This morphism is representable finite \'etale. Hence, if ${}_{G'}{\cal M}_g} \def\Mg{M_g$ is a scheme (i.e., isomorphic to ${}_{G'}M_g$), then so is ${}_G{\cal M}_g} \def\GMg{{}_GM_g$. In this case the morphism ${}_{G'}M_g\to \GMg$ is finite \'etale (perhaps of degree $0$). Finally, we note that ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ might be empty, this being the case if $G$ is not isomorphic to a quotient of the fundamental group of a Riemann surface of genus $g$. \subsubsection{Abelian level structures} In this section we treat the algebraic case. Let $f:X\to S$ be a smooth stable curve of genus $g$ over $S$. Fix a natural number $m$. An abelian structure structure of level $m$ on $X$ over $S$ is defined as an isomorphism of \'etale sheaves over $S$ $$ (\msy Z/m\msy Z)^{2g}_S\longrightarrow R^1f_\ast(\msy Z/m\msy Z).$$ Such a level structure can only exist if the base scheme $S$ lies over ${\rm Spec}(\msy Z[1/m])$; let us assume this is the case. Note that the sheaf ${\cal H}om^{ext}(\pi_1(X/S),\msy Z/m\msy Z)$ is isomorphic to the subsheaf of primitive elements in $R^1f_\ast(\msy Z/m\msy Z)$. Thus we see that the moduli stack of curves with an abelian level $m$ structure is isomorphic to the stack ${}_G{\cal M}_g} \def\GMg{{}_GM_g$ with $G=(\msy Z/m\msy Z)^{2g}$. In particular, if $m\geq 3$, then $\GMg$ is a fine moduli scheme smooth over ${\rm Spec}(\msy Z[1/m])$. The following is deduced from the above. \begin{proposition} If the finite group $G$ allows a surjection onto $(\msy Z/m\msy Z)^{2g}$ for some $m\geq 3$ then the coarse moduli scheme ${}_GM_g$ is a fine moduli scheme.\end{proposition} \subsubsection{Compactifications} In order to get compact moduli spaces we just take the normalization with respect to the Deligne-Mumford compactification. In this subsection the finite group $G$ will be fixed, of order $n=\#G$. Consider the Deligne-Mumford compactification ${\cal M}_g} \def\Mg{M_g\subset \overline{{\cal M}_g}} \def\bMg{\overline{M_g}$. We define $\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}$ as the normalization of $\bMg[1/n]$ with respect to ${}_G{\cal M}_g} \def\GMg{{}_GM_g$. Similarly we define $\bcGMgan$ as the normalization of $\overline{{\cal M}_g}^{an}} \def\bMgan{\overline{M_g}^{an}$ with respect to ${}_G{\cal M}^{an}_g} \def\GMgan{{}_GM_g^{an}$. As per convention we denote $\bGMg$ (resp.\ $\overline{{}_GM_g}^{an}$) the associated coarse moduli scheme (resp.\ analytic moduli space). The morphism $$\bGMg\longrightarrow {\rm Spec}(\msy Z[1/n])$$ is proper, since $\bMg[1/n]$ is proper over ${\rm Spec}(\msy Z[1/n])$. To see whether this morphism is smooth we have the following criterion. \begin{proposition}\label{criterium} For finite groups $G$ which allow a surjection $G\to (\msy Z/m\msy Z)^{2g}$ for some $m\geq 3$ the following statements are equivalent: \begin{enumerate} \item The morphism $\bGMg\longrightarrow {\rm Spec}(\msy Z[1/n])$ is smooth. \item The analytic space $\overline{{}_GM_g}^{an}$ is a (nonsingular) complex manifold. \end{enumerate}\end{proposition} \begin{proof}We use that $\overline{{}_GM_g}^{an}$ is isomorphic to the analytic space associated to the variety $\bGMg\otimes \msy C$. Thus it suffices to show that $\bGMg\otimes \msy C$ nonsingular implies $\bGMg\otimes \bar {\msy F}_p$ nonsingular, where $p>0$ and $p$ does not divide $n$. The argument will be based on the fact that the morphism $\varphi : \bGMg\to \bMg[1/n]$ is tamely ramified along the boundary. To see this we need a description of the complete local rings of $\bGMg$ in points on the boundary. Suppose that $C$ is a stable curve of genus $g$ over an algebraically closed field $k$ of characteristic $p$. The singular points of $C$ are $P_1,\ldots, P_\ell$. Let $\cC\to {\rm Spf}\big(W(k)[[t_1,\ldots,t_{3g-3}]]\big)$ be the universal deformation of $\cC$. We choose the parameters $t_i$ such that $t_i=0$ defines the locus where $P_i$ survives as a singular point, for $i=1,\ldots, \ell$. Put $A=W(k)[[t_1,\ldots,t_{3g-3}]]$. Since $\cC$ is (uniquely) algebraizable we have a morphism ${\rm Spec}(A)\to \overline{{\cal M}_g}} \def\bMg{\overline{M_g}$ and $${\rm Spec}\big(A[1/t_1\ldots t_\ell]\big)\longrightarrow {\cal M}_g} \def\Mg{M_g.$$ We consider the fibre product $${\rm Spec}\big(A[1/t_1\ldots t_\ell]\big)\times_{{\cal M}_g} \def\Mg{M_g} {}_G{\cal M}_g} \def\GMg{{}_GM_g .$$ This is finite \'etale over $A[1/t_1\ldots t_\ell]$ hence affine. The normalization of this over ${\rm Spec}(A)$ is ${\rm Spec}(B)$; here $B$ is a product of complete local rings finite over $A$, ramified only over $t_1\ldots t_\ell=0$. Thus we have $${\rm Spec}\big(B[1/t_1\ldots t_\ell]\big)= {\rm Spec}\big(A[1/t_1\ldots t_\ell]\big)\times_{{\cal M}_g} \def\Mg{M_g} {}_G{\cal M}_g} \def\GMg{{}_GM_g .$$ We claim that the morphism ${\rm Spec}(B)\to \bGMg$ identifies complete local rings at the points of $\bGMg$ lying over $[C]\in \bMg(k)$. By general theory we know that the completion of $\bMg$ at $[C]$ is ${\rm Spf}(A)/{\rm Aut}(C)$. A formal argument gives that the completion of $\bGMg$ along $\varphi^{-1}([C])$ is isomorphic to ${\rm Spf}(B)/{\rm Aut}(C)$. Thus it suffices to show that the action of ${\rm Aut}(C)$ on $\varphi^{-1}([C])$ is free. By comparing levels, it suffices to show this for $G=(\msy Z/m\msy Z)^{2g}$; this is the content of \cite[Proposition 3.5]{De}. See also references in remark below. We use Abhyankar's lemma which asserts that any normal local domain $A'$, finite generically \'etale over $A$ and ramified only along $t_1\ldots t_\ell=0$ is contained in $A[t_i^{1/n_i}]$ for some $n_1,\ldots,n_\ell$ relatively prime to $p$. In addition, it is easily seen that $A'$ is formally smooth over $W(k)$ if and only if $A'$ is actually equal to $A[t_i^{1/n_i}]$ for some $n_1,\ldots,n_\ell$. Let $C_K$ denote the lift of $C$ to $K=Q\big(W(k)\big)$ given by setting $t_1=\ldots=t_{3g-3}=0$. The homomorphism $A\to K[[t_1,\ldots,t_{3g-3}]]$ is such that $$B\otimes_A K[[t_1,\ldots,t_{3g-3}]]$$ describes the complete local rings of $\bGMg$ along $\varphi^{-1}([C_K])$. The result follows by comparing $B$ to this ring.\end{proof} \begin{remark} {\rm The arguments above actually show that in the situation of the proposition the stacks $\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}$ are schemes, i.e., $\overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}\cong \bGMg$. Thus we get a stable curve over $\bGMg$ from the morphism $\bGMg\cong \overline{{}_G{\cal M}_g}} \def\bGMg{\overline{{}_GM_g}\to \overline{{\cal M}_g}} \def\bMg{\overline{M_g}$; this also follows from the existence of such a stable curve in the case of abelian level structures, see \cite[Thm. 10.9]{Po}, \cite[page 12]{GO} and \cite[Bemerkung 1]{Mo}.}\end{remark} \section{Monodromy along the boundary} In this section we study the moduli spaces $\bGMg$ along the boundary. To do this it suffices to understand the monodromy along the boundary on the relative fundamental group of the universal curve over ${\cal M}^{an}_g} \def\Mgan{M_g^{an}$. \subsection{The results}\label{results} Let us formulate the main result. Let $\Pi$ denote the standard fundamental group of a compact Riemann surface of genus $g$. We fix natural numbers $k,n$ with $k\geq 1$ and $n\geq 3$. We will consider moduli of curves of genus $g$ with a Teichm\"uller structure of level $G$, where $$G=\Pi\big/ \Pi^{(k+1),n}.$$ (For notations see \ref{notation}.) By a result of Labute \cite{Labute}, the quotients $\Pi^{(k)}/\Pi^{(k+1)}$ are finitely generated free abelian groups. Thus $G$ has a filtration whose successive quotients are finite abelian groups of exponent $n$. Any prime dividing $\# G$ also divides $n$. Further, there is a surjection $G\to (\msy Z/n\msy Z)^{2g}$. By Section 2 we get a moduli scheme $$\bGMg\longrightarrow {\rm Spec}(\msy Z[1/n])$$ whose interior $\GMg$ classifies smooth genus $g$ curves $C$ with a surjection $\pi_1(C)\to G$ given up to inner automorphisms. \begin{theorem}\label{glad} Suppose $k\in \{1,2,3\}$ and $n\geq 3$. The structural morphism $\bGMg\to {\rm Spec}(\msy Z[1/n])$ is smooth if and only if \begin{itemize} \item{$k=1$} and $g=2$, \item{$k=2$} and $n$ is odd, \item{$k=3$} and $n$ is odd or $n$ is divisible by $4$. \end{itemize} Furthermore, if $k \geq 4$, $n \geq 3$ and $n$ relatively prime to $6$, then $\bGMg\to {\rm Spec}(\msy Z[1/n])$ is smooth. \end{theorem} \begin{remark}{\rm The abelian case of this theorem, i.e., the case $k=1$, has been proven by Mostafa \cite{Mo} and Van Geemen-Oort \cite{GO}. A smooth and compact cover $\overline{M_g[_2^n]}^{an} \rightarrow \bMgan$ with $n$ even and $n\geq 6$ has been constructed by Looijenga using Prym level structures, see \cite{Looijenga}. This has been extended to higher level structures $\overline{ M_g[_q^n]}^{an}$ by one of us (Pikaart). Using an analog of Proposition \ref{criterium} this result may be extended to an open set of ${\rm Spec}(\msy Z)$, at least if $q\geq 3$. We note that $\bGMg\to\bMg[1/n]$ is a Galois cover\label{galois} with group ${\rm Aut}(G)/{\rm Inn}(G)$ if $g\geq 3$ (if $g=2$ one has to divide out an additional $\{\pm 1\}$). This follows since for any surjection $\Pi\to G$ the kernel is equal to $\Pi^{(k),n}$. However, $\bGMg\otimes\msy C$ need not be connected, see \cite[5.13]{DM} for a description of the set of connected components.}\end{remark} To prove the theorem, it suffices to consider $\overline{{}_GM_g}^{an}$, see Proposition \ref{criterium}. We already know that $\GMgan$ is smooth. As in the proof of Proposition \ref{criterium} we describe analytic neighbourhoods of points in the boundary. Let $C$ be a complex stable curve of genus $g$ with singular points $\{ P_1,\ldots,P_\ell\}$. Let $\Gamma=\Gamma(C)$ be its dual graph; an edge for each point $P_j$, a vertex for an irreducible component of $C$. Let $ \pi: (\cC,C) \rightarrow (B,0)$ be a local universal deformation of $C$, where $B\subset \msy C^{3g-s}$ is a polydisc neighbourhood of $0$. The coordinates $z_i$ are chosen such that $z_j=0$, $1\leq j\leq \ell$ parametrizes curves where the singular point $P_j$ subsists. The discriminant locus $\Delta\subset B$ of $\pi$ is thus given by $z_1\ldots z_\ell=0$. Put $U=B\setminus \Delta$, let $x\in U$ and choose $y\in \cC_x=\pi^{-1}(x)$. The fundamental group of $U$ is an abelian group, freely generated by simple loops around the divisors $z_j=0$, thus naturally isomorphic to the free abelian group on the edges of $\Gamma$, i.e., $\pi_1(U,x) \cong \bigoplus_{e \in {\rm Edges}( \Gamma )} {\msy Z} e$. The map $\cC|_U\to U$ is a locally trivial fibration, hence we have the exact sequence $$1 \longrightarrow \pi_1(\cC_x,y) \longrightarrow \pi_1(\cC|_U,y) \longrightarrow \pi_1(U,x) \longrightarrow 1.$$ (Use that $\pi_2(U)=(0)$.) This short exact sequence provides us with the monodromy representation $$\rho: \pi_1(U,x) \longrightarrow {\rm Out}\big(\pi_1(\cC_x,y)\big).$$ The points $P_j$ determine non-trivial distinct isotopy classes of circles on $\cC_x$, which have pairwise disjoint representatives $c_j$. The fundamental group of $U$ is also naturally isomorphic to the free abelian group on these circles, $\pi_1(U,x) \cong \bigoplus_{i=1}^l {\msy Z} c_i$. Under this identification we have that $$\rho(c_i)=D_{c_i},$$ where $D_{c_i}$ is the exterior automorphism of $\pi_1(C_x)$ given by a Dehn twist (also written $D_{c_j}$) around the circle $c_i$ (see \cite{Dehn},\cite{Lamotke}). We will describe of a neighbourhood of a point in $\overline{{}_GM_g}^{an}$ lying above $[C]$. Let $Z$ be the fibre product $$Z=U\times_{\Mgan}\GMgan.$$ The normalization of $B$ in the function field of $Z$ is denoted $\bar Z$. Note that $Z\to U$ is a finite topological covering space given by the set $$S=\hbox{Hom-surj}\big(\pi_1(\cC_x,y),G\big)\big/\pi_1(\cC_x,y)$$ with $\pi_1(U,x)$-action defined via $\rho$. As in the proof of Proposition \ref{criterium} there is an action of ${\rm Aut}(C)$ on $\bar Z$ and $\bar Z/{\rm Aut}(C)$ defines a neighbourhood of $\varphi^{-1}([C])\subset \overline{{}_GM_g}^{an}$. As in that proof we get that $\overline{{}_GM_g}^{an}$ is smooth along $\varphi^{-1}([C])$ if and only if $\bar Z$ is smooth. (Here we use again that ${\rm Aut}(C)$ acts freely on $\varphi^{-1}([C])$.) Finally there is the following criterion: $\bar Z$ is smooth if and only if for all $s\in S$ we have $${\rm Stab}(s)=\bigoplus_{e \in {\rm Edges}(\Gamma)}n_e {\msy Z} e $$ for certain $n_e\in \msy Z$. Notice that if $m \in \msy Z $, then $mF^0 +m F^1$ ($F^i$ as below) is of this form, but $2mF^0 +m F^1$ is not. We remark that the arguments above go through for arbitrary finite groups $G$, with a surjection onto $(\msy Z/n\msy Z)^{2g}$. In order to describe the stabilizers in our more special situation we introduce the following notation: \begin{eqnarray*} \Akn &= &{\rm Ker}\left( {\rm Aut}(\Pi) \to {\rm Aut}(\Pi/\Pi^{(k+1),n})\right),\\ \Ikn &= &{\rm Ker}\left( {\rm Inn}(\Pi) \to {\rm Inn}(\Pi/\Pi^{(k+1),n})\right),\\ \Okn &= &{\rm Ker}\left( {\rm Out}(\Pi) \to {\rm Out}(\Pi/\Pi^{(k+1),n})\right). \end{eqnarray*} Oda et al.\ consider a variant with $n=0$. By choosing an isomorphism $\pi_1(\cC_x,y)\cong\Pi$ we may view $\rho$ as a map into ${\rm Out}(\Pi)$. It is clear that ${\rm Stab}(s)=\rho^{-1}(\Okn)$ for any $s\in S$ (use Remark \ref{galois}). Therefore, Theorem \ref{glad} follows from Theorem \ref{mono} below. We will describe a decreasing filtration $F^i$ on $\bigoplus_{e \in {\rm Edges}( \Gamma)} {\msy Z} e$. An edge $e$ such that $\Gamma \backslash e$ is disconnected is called a {\em bridge}. A bridge $b$ is said to {\em bound a genus one curve} if one of the two components of $C_x \backslash \{ $ the circle corresponding to $b \}$ has genus one. A pair of distinct edges $\{e,f\}$ is called a {\em cut pair} if neither $e$ nor $f$ is a bridge and $\Gamma \backslash \{e,f \}$ is disconnected. A subset $E$ of the edges of $\Gamma$ is called a {\em maximal cut system} if $E$ contains at least one cut pair, any two elements of $E$ form a cut pair and no element of $E$ forms a cut pair with an element outside $E$. Let $B$ be the set of bridges of $\Gamma$ and let $B_1$ be the subset of $B$ consisting of bridges which bound genus one curves. Let $\{E_i\}_{i \in I}$ denote the maximal cut systems.Set $$D_i:=Ker(\bigoplus_{e \in E_i} \msy Z e \stackrel{deg}{\rightarrow } \msy Z).$$ We define a decreasing filtration $F^i$ on $\bigoplus_{e \in {\rm Edges}(\Gamma)}\msy Z e$, as follows: \begin{eqnarray*} F^0&= & \bigoplus\nolimits_{e \in {\rm Edges}(\Gamma)}{\msy Z} e,\\ F^1&= & \bigoplus\nolimits_{i\in I} D_i \bigoplus \left(\bigoplus\nolimits_{b \in B} {\msy Z}b\right), \\ F^2&= & \bigoplus\nolimits_{b \in B} {\msy Z}b ,\\ F^3&= & (0).\end{eqnarray*} Furthermore, set $F^2_1:=\bigoplus_{b \in B_1} {\msy Z}b$. This refines the filtration into $F^0\supset F^1\supset F^2\supset F^2_1\supset F^3=(0)$. Here is the main result of this article. \begin{theorem}\label{mono} Notations as above. For $n,l \in \msy Z$, define $n_l:=n/ {\rm gcd}(l,n)$. \[ \begin{array}{cll} 1. & \mbox{If $k=1$, then} & \rho^{-1}(\Okn ) =nF^0 +F^1 \\ 2. & \mbox{If $k=2$, then} & \rho^{-1}(\Okn )=nF^0 +n_2F^1+F^2 \\ 3. & \mbox{If $k=3$ and $2 ||n$, then} & \rho^{-1}(\Okn )=nF^0 + \frac{1}{2}nF^1+n_2F^2 +n_6F^2_1 \\ & \mbox{If $k=3$ and $n$ is odd or $4 | n$, then} & \rho^{-1}(\Okn ) =nF^0 + nF^1+n_2F^2 +n_6F^2_1\\ 4. & \mbox{If $k \geq 4$ and $(n,6)=1$, then} & \rho^{-1}(\Okn ) =nF^0 \end{array} \] \end{theorem} \begin{remark} {\rm The case $k=1$ has been proven by Brylinski (\cite{Brylinski}). The case $k \geq 4$ follows from 3, the easy inclusions of Section \ref{zeer easy} and the inclusions $\Okn \subset O^{(l),n}$ if $k \geq l$. If we take $n=0$ then Theorem \ref{mono} reduces to the non-pointed case of \cite[Main Theorem]{Oda}.} \end{remark} \section{Description of Dehn twists and easy inclusions} \label{zeer easy} In this section we will prove the inclusions ``$\supset$'' from Theorem \ref{mono}. Let $\Gamma$ be the graph of a stable curve and $(S,\{ c_i \})$ the smooth model with a set of circles as described in Subsection \ref{results}. We will describe the Dehn twist associated to a bridge, cut pair or circle separatedly. \subsection{Bridges}\label{bridges} Let $b$ be a bridge on $S$, let $g$ be the genus of $S$. Modulo a homeomorphism the situation looks as follows: \hfill \vspace{4 cm} $$\hbox{\sl Fig.~1}$$ Cutting $S$ along $b$ yields the decomposition $S= S_1 \cup S_2$. Let $g_i$ be the genus of $S_i$. Choose a base point $p$ in $S_1$ and standard generators $\alpha_{\pm i},$ $1 \leq i \leq g$, for $\pi :=\pi_1(S,p)$ such that $\alpha_{\pm i}$ are in $S_1$ if $i \leq g_1$ and $\alpha_{\pm i}$ for $i\geq g_1+1$ hits $b$ exactly twice. We set $v=[\alpha_1,\alpha_{-1}] \cdots [\alpha_{g_1}, \alpha_{-g_1}]$, it is freely homotopic to $b$ for a suitable orientation of $b$. We list the effect of the Dehn twist $D_b$ on the standard generators: \[ \begin{array}{ll} i \in \{\pm 1,\ldots, \pm g_1\} & {D}_{b}: \alpha_i \mapsto \alpha_i ,\\ i \in \{ \pm ( g_1 +1),\ldots,\pm g \} & {D}_{b}: \alpha_i \mapsto v^{-1} \alpha_i v \end{array} \] This gives for the mth-powers: \[ \begin{array}{ll} i \in \{\pm 1,\ldots, \pm g_1\} & {D}^m_{b}: \alpha_i \mapsto \alpha_i ,\\ i \in \{ \pm ( g_1 +1),\ldots,\pm g \} & {D}^m_{b}: \alpha_i \mapsto v^{-m} \alpha_i v^m \end{array} \] We see that $D_b(\alpha_{\pm i})\alpha_{\pm i}^{-1} \in \pi^{(3)}$ for all $i$. This proves that $\rho(b) \in O^{(2)}$ and thus by linearity $\rho(F^2) \subset O^{(2)}$. Suppose $b$ bounds a genus one curve, say $S_2$ has genus one. In this case $v = [\alpha_{-g},\alpha_g]$ and we only have to consider $$D_b^m(\alpha_{\pm g})\alpha_{\pm g}^{-1}= [[\alpha_g,\alpha_{-g}]^m,\alpha_{\pm g}]\equiv [[\alpha_g,\alpha_{-g}],\alpha_{\pm g}]^m \bmod \pi^{(4)}.$$ To prove this element lies in $\pi^{(4),n}$ we define $f:F(x,y,z) \ra \pi$ by $x \mapsto \alpha_g$, $y \mapsto \alpha_{-g}$ and $z \mapsto 1$. Then $f([[x,y],x])=[[\alpha_g,\alpha_{-g}],\alpha_{ g}]$ and $f([[x,y],y])=[[\alpha_g,\alpha_{-g}],\alpha_{ -g}]$. {}From Lemma \ref{berekening} we see that $n_6|m$ implies that the element above lies in $\pi^{(4),n}$. Thus $\rho(n_6F^2_1) \subset O^{(3),n}$. Suppose $b$ does not bound a genus one curve. In that case we have \begin{eqnarray*} D_b^m(\alpha_{\pm i})\alpha_{\pm i}^{-1}&=& [v^m,\alpha_{\pm i}^{-1}]\ \equiv\ [v,\alpha_{\pm i}]^{-m}\\ &\equiv & [[\alpha_{1},\alpha_{-1}],\alpha_{\pm i}]^{-m}\cdots [[\alpha_{g_1},\alpha_{-g_1}],\alpha_{\pm i}]^{-m} \end{eqnarray*} for $i>g_1$ in $\pi/\pi^{(4)}$. Define $g_j:F(x,y,z) \ra \pi$ by $x \mapsto \alpha_j,~y \mapsto \alpha_{-j}$ and $z \mapsto \alpha_{\pm i}$. Then $g_j([[x,y],z]^m)= [[\alpha_j,\alpha_{-j}],\alpha_{\pm i}]^m$. {}From Lemma \ref{berekening} we see that $n_2 |m$ implies $\rho(mb) \in O^{(3),n}$, and thus $\rho(n_2F^2) \subset O^{(3),n}$. \subsection{Edges which are not bridges}\label{not bridges} Let $c$ be a circle on $S$ which is not a bridge. Modulo a homeomorphism the situation looks as follows: \hfill \vspace{4 cm} $$\hbox{\sl Fig.~2}$$ We choose a point $p$ in $S$ and standard generators $\alpha_{\pm i}$ for $\pi =\pi_1(S,p)$ such that $\alpha_{-g}$ is the only one intersecting $c$ and $\alpha_g$ is freely homotopic to $c$. The action of the Dehn twist becomes: $D_c(\alpha_i)=\alpha_i$ if $i \neq -g$ and $D_c(\alpha_{-g}) =\alpha_g \alpha_{-g}$. Thus $D_c^m(\alpha_i)\alpha_i^{-1}=1$ or $\alpha_g^m$. Evidently, if $n|m$ then $\rho(mc) \in O^{(k),n}$ for all $k$. Together with the results of \ref{bridges} this gives $\rho(nF^0)\subset O^{(k),n}$ for all $k$. \subsection{Cut systems}\label{cut systems} Let $e_1,e_2$ be a cut pair on $S$. Modulo a homeomorphism the situation is as follows: \hfill \vspace{4 cm} $$\hbox{\sl Fig.~3}$$ Cutting $S$ along $e_1$ and $e_2$ yields the decomposition $S=S_1 \cup S_2$. Let $g_1$ be the genus of $S_1$. Choose a base point $p$ in $S_1$. We take standard generators $\alpha_{\pm i}$ for $\pi_1(S,p)$, such that for $i \in \{ \pm 1,\ldots, \pm g_1, g_1+1\} $, $ \alpha_i $ is in $S_1$, for $i \in \{ \pm (g_1+2),\ldots, \pm g \} $, $ \alpha_i $ enters $S_2$ via $e_1$ and leaves $S_2$ also via $e_1$ and $\alpha_{-(g_1+1)}$ enters $S_2$ via $e_1$ and leaves $S_2$ via $e_2$. Furthermore, we want $\alpha_{g_1+1}$ to be freely homotopic to $e_2$, for a suitable orientation of $e_2$. It follows that after suitable orientation of $e_1$ the loop $[\alpha_1,\alpha_{-1}] \cdots [\alpha_{g_1},\alpha_{-g_1}]\alpha_{g_1+1}^{-1}$ is freely homotopic to ${e_1}$. Let us write $v=[\alpha_1,\alpha_{-1}] \cdots [\alpha_{g_1},\alpha_{-g_1}]$. We list the effect of the Dehn twists ${D}_{e_i}$ on these generators: \[ \begin{array}{ll} \mbox{$i \in \{ \pm 1,\ldots, \pm g_1, g_1+1\} $}& \left\{ \begin{array}{l} {D}_{e_1}: \alpha_i \mapsto \alpha_i \\ {D}_{e_2}: \alpha_i \mapsto \alpha_i \end{array} \right. \\[12pt] \mbox{$i = -(g_1+1)$} & \left\{ \begin{array}{l} {D}_{e_1}: \alpha_i \mapsto \alpha_i v \alpha_{g_1+1}^{-1}\\ {D}_{e_2}: \alpha_i \mapsto \alpha_{g_1+1}^{-1}\alpha_i \end{array} \right. \\[12pt] \mbox{$i \in \{ \pm (g_1+2),\ldots, \pm g \}$} & \left\{ \begin{array}{l} {D}_{e_1}: \alpha_i \mapsto \alpha_{g_1+1} v^{-1} \alpha_i v \alpha_{g_1+1}^{-1}\\ {D}_{e_2}: \alpha_i \mapsto \alpha_i \end{array} \right. \end{array} \] Thus we get the following formulae for $D_{e_2}D_{e_1}^{-1}$: \[ \begin{array}{ll} |i| \leq g_1 & D_{e_2}D_{e_1}^{-1}(\alpha_i)=\alpha_i, \\[2pt] i=g_1+1 & D_{e_2}D_{e_1}^{-1}(\alpha_i)=\alpha_i, \\[2pt] i = -(g_1+1)& D_{e_2}D_{e_1}^{-1}(\alpha_i) =\alpha_{g_1+1}^{-1}\alpha_i\alpha_{g_1+1}v^{-1}\alpha_i^{-1}\alpha_i =[\alpha_{g_1+1},\alpha_i^{-1}]v^{-1}[v^{-1},\alpha_i^{-1}]\alpha_i,\\[2pt] |i| \geq g_1+2 & D_{e_2}D_{e_1}^{-1}(\alpha_i) =v\alpha_{g_1+1}^{-1}\alpha_i\alpha_{g_1+1}v^{-1}\alpha_i^{-1}\alpha_i =v[\alpha_{g_1+1},\alpha_i^{-1}]v^{-1}[v^{-1},\alpha^{-1}]\alpha_i \end{array} \] and for the mth powers: \[ \begin{array}{ll} i \in \{ \pm 1,\ldots, \pm g_1, g_1+1\}& (D_{e_2}D_{e_1}^{-1})^m(\alpha_i)=\alpha_i,\\[2pt] i = -(g_1+1)& (D_{e_2}D_{e_1}^{-1})^m(\alpha_i) =\alpha_{g_1+1}^{-m}\alpha_i (\alpha_{g_1+1}v^{-1})^m,\\[2pt] i \in \{ \pm (g_1+2),\ldots, \pm g \}& (D_{e_2}D_{e_1}^{-1})^m(\alpha_i) =(v\alpha_{g_1+1}^{-1})^m \alpha_i (\alpha_{g_1+1}v^{-1})^m. \end{array} \] This proves that for a cut pair $\{e_1,e_2\}$ we have $\rho(e_2-e_1) \in O^{(1)}$. It follows from this and \ref{bridges} that $\rho(F^1) \subset O^{(1)}$, because $F^1$ is generated by elements of the form $e_2-e_1$ for a cut pair $\{e_1,e_2\}$ and the elements $b,~b \in B$. To finish the proof of the inclusions ``$\supset$'' we have to show that $\rho(me_2-me_1)\in O^{(2),n}$ if $n_2|m$ and $\rho(me_2-me_1)\in O^{(3),n}$ if $n|m$, or $\frac{1}{2}n|m$ in case $2||m$. We have to show that the divisibility conditions imply $(D_{e_2}D_{e_1}^{-1})^m(\alpha_i)\alpha_i^{-1}\in \pi^{(3),n}$, respectively $\pi^{(4),n}$. We will make computations in $\pi$ modulo $\pi^{(4)}$. In the case that $i\in\{\pm1,\ldots,\pm g_1, g_1+1\}$ there is nothing to prove. We have $(\alpha_{g_1+1}v^{-1})^m\equiv \alpha_{g_1+1}^mv^{-m}[v^{-1},\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} $ mod $ \pi^{(4)}$, as one can prove by induction. {}From equality 5 of Subsection \ref{group structure} we have $$\alpha_{g_1+1}^{-m}\alpha_i\alpha_{g_1+1}^m \alpha_i^{-1}\equiv [\alpha_{g_1+1},\alpha_i^{-1}]^m [[\alpha_{g_1+1},\alpha_i^{-1}],\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)}.$$ Thus for $i=-(g_1+1)$ we get \begin{eqnarray*} &&(D_{e_2}D_{e_1}^{-1})^m(\alpha_i)\alpha_i^{-1}\\ &\equiv&\alpha_{g_1+1}^{-m} \alpha_i (\alpha_{g_1+1}v^{-1})^m \alpha_i^{-1}\\ &\equiv&\alpha_{g_1+1}^{-m} \alpha_i \alpha_{g_1+1}^mv^{-m} [v^{-1},\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} \alpha_i^{-1}\\ &\equiv&{[\alpha_{g_1+1},\alpha_i^{-1}]}^{m} [[\alpha_{g_1+1},\alpha_i^{-1}],\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} v^{-m} [v, \alpha_i]^m [v,\alpha_{g_1+1}]^{-\frac{1}{2}m(m-1)}\\ &\equiv& {[\alpha_{g_1+1},\alpha_i^{-1}]}^{m} [[\alpha_{g_1+1},\alpha_i^{-1}],\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} [\alpha_1,\alpha_{-1}]^m \cdots [\alpha_{g_1},\alpha_{-g_1}]^m [[\alpha_1,\alpha_{-1}],\alpha_i]^m \cdots \\ && {[[\alpha_{g_1},\alpha_{-g_1}],\alpha_i]}^m {[[\alpha_1,\alpha_{-1}],\alpha_{g_1+1}]}^{-\frac{1}{2}m(m-1)} \cdots [[\alpha_{g_1},\alpha_{-g_1}],\alpha_{g_1+1}]^{-\frac{1}{2}m(m-1)}. \end{eqnarray*} In this product there are, apart from factors in $\pi^{(3)}$, a number of terms of the form $[x,y]^m$, $x,y\in \pi$. By the divisibility conditions all these are in $\pi^{(4),n}$, since this is true in the case of a free group $\langle x, y, z\rangle$ by Lemma \ref{berekening}. Thus it is clear that the whole product lies in $\pi^{(3),n}$ for $n_2|m$. The other terms in the product are of the form $[[x,y],z]^{\frac{1}{2}m(m-1)}$, where $x,y,z\in \pi$. It follows from Lemma \ref{berekening}, by mapping $G=\langle x, y, z\rangle$ into $\pi$ that these terms are in $\pi^{(4),n}$ as soon as $n_2| \frac{1}{2}m(m-1)$. This is equivalent to the condition $n|m$ or $\frac{1}{2}n|m$ in case $2||n$. For $i \geq g_1+2$ we have: \begin{eqnarray*} (D_{e_2}D_{e_1}^{-1})^m(\alpha_i)\alpha_i^{-1}&=& (v\alpha_{g_1+1}^{-1})^m \alpha_i (\alpha_{g_1+1}v^{-1})^m \alpha_i^{-1} \\ &\equiv& v^m \alpha_{g_1+1}^{-m} [\alpha_{g_1+1}^{-1},v]^{\frac{1}{2}m(m-1)}\alpha_i \alpha_{g_1+1}^mv^{-m} [v^{-1},\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} \alpha_i^{-1}\\ &\equiv & {[\alpha_{g_1+1},\alpha_i^{-1}]}^m [[\alpha_{g_1+1},\alpha_i^{-1}],\alpha_{g_1+1}]^{\frac{1}{2}m(m-1)} \end{eqnarray*} The same argument works to show that our divisibility conditions imply this product lies in either $\pi^{(3),n}$ or $\pi^{(4),n}$. \section{Completion of proof} In this section we will use the following argument. Suppose we are given a morphism $f:S \ra S'$ of compact connected oriented surfaces and a circle $c$ on $S$. Assume that $f$ maps a tubular neighbourhood of $c$ isomorphically into $S'$, denote by $c'$ the image of $c$. Then we have: $f_\ast \circ D_c =D_{c'} \circ f_\ast$. Let $\Gamma$ be the graph of a stable curve and $(S,\{ c_i \})$ the smooth model with a set of circles as described in Subsection \ref{results}. Let $H$ be the set of circles which are not bridges and are not involved in any cut system. Let $B$ be the set of bridges of $\Gamma$ and let $\{ E_i|i=1,\ldots,s \}$ be the set of maximal cut systems. Choose a numbering $E_i= \{ e_{i,0},\ldots,e_{i,f_i} \}$ for each $i$. Remark that ${\rm Edges}(\Gamma)=H\cup \bigcup E_i\cup B$. \subsection{The case $k=1$} We want to prove the inclusion ``$\subset$'' for $k=1$. Suppose $\sigma \in \oplus {\msy Z}e$ is a counterexample which involves a minimal number of edges. This means that $\rho(\sigma)$ lies in $O^{(1),n}$ but not all coefficients of $\sigma$ are divisible by $n$; the number of nonzero coefficients is minimal. We write $\sigma =\sum n_c c$. We may subtract elements of $nF_0+F_1$ from $\sigma$: we already know that $nF_0+F_1$ maps into $O^{(1),n}$. Thus, by minimality, we may suppose that: (\romannumeral1) none of the nonzero coefficients $n_c$ is divisible by $n$, (\romannumeral2) none of the circles $c$ is a bridge, and (\romannumeral3) of each cutsystem $E_i$ at most one element occurs in $\sigma$. Modulo a homeomorphism the situation looks as follows: \hfill \vspace{4 cm} $$\hbox{\sl Fig.~4}$$ Take a point $p$ near $c$. It is clear that we can find a loop $\alpha$ which intersects $c$ exactly once and none of the other edges involved in $\sigma$. But now we have a contradiction, because by Section \ref{not bridges} we know: $\rho(\sigma)(\alpha)\alpha^{-1}= c^{n_c} \notin \pi^n$ if $n$ does not divide $n_c$. Thus we have proven $\rho^{-1}(O^{(1),n})\subset nF^0+F^1$, equality follows from sections \ref{not bridges} and \ref{cut systems}. \subsection{The case $k=2$}\label{k is 2} We want to prove the inclusion ``$\subset$'' for $k=2$. Let $\sigma=\sum_{b \in B}m_b b +\sum_{i,j}m_{i,j}(e_{i,j}-e_{i,0}) +\sum_i m_i e_{i,0} + \sum_{c \in H} n_c c$ be such that $\rho(\sigma ) \in O^{(2),n}$. By the above we know that $n|m_i$ and $n|n_c, c\in H$. We have to show that $n_2|m_{i,j}$ for all possible $i,j$. Suppose this does not hold and suppose furthermore that $\sigma$ is a minimal counterexample with respect to the number of edges involved. Arguing as above we may assume that $\sigma =\sum_{i,j}m_{i,j}(e_{i,j}-e_{i,0})$, where no (nonzero) $m_{i,j}$ is divisible by $n_2$. \begin{proposition} \label{nice cut system} There is a maximal cut system which is such that if we cut $S$ along every element of the cut system, one connected component contains all other cut systems. \end{proposition} \begin{proof} (Cf.\ \cite{Oda}, Lemma 5.3.) Take any cut system $E$. If it does not have the required property, that means that there is more than one connected component of $S \setminus E$, say $S_{E,1}, \dots , S_{E,l}$, which contain maximal cut systems. Suppose $S_{E,j}$ contains $a_j$ maximal cut systems and suppose $a_1= \mbox{ min}_j \{ a_j \}$. Take a maximal cut system $F$ in $S_{E,1}$. Clearly there is a connected component of $S\setminus F$ which contains at least $\sum_{j=2}^la_j+1$ components, namely the one containing $E$. This component contains more maximal cut systems then the one we started with. Continuing in this way, we arrive at our result. \end{proof} Let $E_1=\{e_{1,0}, \dots, e_{1,f_1} \}$ be a maximal cut system such that one of the two components bounded by $e_{1,0}$ and $e_{1,f_1}$, contains all other cut systems involved in $\sigma$. We suppose these edges are numbered cyclically: one of the connected components of $S\setminus \{e_{1,i},e_{1,i+1}\}$ contains all other $e_{1,j}$ (see figure 6). If $E_1$ contains more than two edges, we proceed as follows. Replace the component of $S \setminus \{ e_{1,1},~e_{1,f_1} \}$ which does not contain $e_{1,0}$ by a cylinder to get an oriented surface $S'$. There is a continuous map $f: S\to S'$ such that for all elements $e_{1,j}$, except $e_{1,0}$, we have that $f(e_{1,j})$ is homotopic to $f(e_{1,1})$. As explained above, we get that $f_\ast \circ \rho(\sigma)=\rho(f_\ast\sigma)\circ f_\ast$. Thus $\rho(\sigma)\in O^{(2),n}$ implies $\rho(f_\ast \sigma)\in O^{(2),n}$. It is also clear that $f_\ast(\sigma)= m_{1,1}(f(e_{1,1})-f(e_{1,0}))+ \sum_{i>1} m_{i,j}(f(e_{i,j})-f(e_{i,0}))$. We are reduced to the case that $E_1=\{e_{1,0},e_{1,1}\}$ consists of two elements. The situation now is as in \ref{cut systems} (figure 3) where all the cut systems $E_i$, $i>1$, lie in the component $S_2$. The generator $\alpha_{-(g_1+1)}$ may be chosen such that it does not intersect the circles $e_{i,j}$, $i>1$. Thus we get from the formulae of Section \ref{cut systems} that \begin{eqnarray*} \rho(\sigma)(\alpha_{-(g_1+1)})\cdot \alpha_{-(g_1+1)}^{-1}& =&(D_{e_{1,1}}D_{e_{1,0}}^{-1})^{m_{1,1}}(\alpha_{-(g_1+1)}) \cdot \alpha_{-(g_1+1)}^{-1}\\ &=&\alpha_{g_1+1}^{-m_{1,1}}\alpha_{-(g_1+1)}(\alpha_{g_1+1}v^{-1})^{m_{1,1}} \alpha_{-(g_1+1)}^{-1}.\end{eqnarray*} We define $f:\pi\to G=\langle x,y,z\rangle$ by $\alpha_1, \alpha_{-g} \mapsto x$ , $\alpha_{-1}, \alpha_g \mapsto y$, $\alpha_{g_1+1} \mapsto z$, other generators $ \mapsto 1$. (The defining relation for $\pi$ is indeed mapped to 1.) The expression above is mapped to $z^{-m_{1,1}}\cdot(z\cdot [y,x])^{m_{1,1}}$. Modulo $G^{(3)}$ this equals $[y,x]^{m_{1,1}}$, hence $\rho(\sigma)\in O^{(3),n}$ implies $n_2|m_{1,1}$ (see Corollary \ref{free comp}). In this way we prove the desired contradiction; the inclusion ``$\subset$'' for $k=2$ follows. \subsection{The case $k=3$} We want to prove the inclusion ``$\subset$'' for $k=3$. Let $\sigma=\sum_{b \in B}m_b b +\sum_{i,j}m_{i,j}(e_{i,j}-e_{i,0}) +\sum_i m_i e_{i,0} + \sum_{c \in H} n_c c$ be such that $\rho(\sigma) \in O^{(3),n}$. We have to show that $n_2|m_{i,j}$ or $n|m_{i,j}$, for all $i,j$, depending on $n$ being exactly divisible by $2$ or not; and $n_6$ respectively $n_2$ divides $m_b$ depending on $b \in B_1$ or not. Suppose this does not hold and suppose furthermore that $\sigma$ is a minimal counterexample with respect to the number of edges involved. Thus $\sigma =\sum_{b \in B}m_b b +\sum_{i,j}m_{i,j}(e_{i,j}-e_{i,0})$ and none of the nonzero $m_b$, $m_{i,j}$ satisfy the divisibility conditions. Case 1. For some $b$ involved in $\sigma$ one of the connected components, say $S'$, of $S\setminus \{b\}$ contains no other edges involved in $\sigma$. The situation looks as follows. \hfill \vspace{4.8 cm} $$\hbox{\sl Fig.~5}$$ We take a basepoint $p$ in $S'$ and standard generators $\alpha_{\pm i}$ which are loops in $S'$ for $i \leq g(S')$ and which are not loops in $S'$ for $i > g(S')$. We may choose $\alpha_{g(S')+1}$ such that it does not intersect any circles involved in $\sigma$ but $b$. To see this let $S''$ denote the connected component of $S\setminus \{c\in \sigma, c\not = b\}$ containing $S'$. If $g(S'')>g(S')$, then there is a `hole' between $b$ and the boundary of $S''$, which we can take to be hole number $g(S')+1$ and take $\alpha_{g(S')+1}$ accordingly. This is the case for example if $S''=S$ (i.e., if $\sigma = m_b b$) or if $S''$ is bounded by another bridge $b'$. If $g(S'')=g(S')$, then $S''$ is bounded by a cut pair $\{e_1, e_2\}$ (belonging to some maximal cut system $E_i$), the part $S''\setminus S'$ looks like a pair of pants. In this case we can choose $\alpha_{g(S')+1}$ to go around a pants leg of $S''\setminus S'$, for example freely homotopic to $e_2$ as in Section \ref{cut systems}. It is now clear that $\rho(\sigma)(\alpha_{g(S')+1})= D_b^{m_b}(\alpha_{g(S')+1})$ and by Section \ref{bridges} we get $$ \xi:= \rho(\sigma)(\alpha_{g(S')+1})\cdot \alpha_{g(S')+1}^{-1}= v^{-m_b}\alpha_{g(S')+1}v^{m_b}\alpha_{g(S')+1}^{-1}$$ with $v=[\alpha_1,\alpha_{-1}]\cdot\ldots\cdot [\alpha_{g(S')},\alpha_{-g(S')}]$. Thus $\rho(\sigma)\in O^{(3),n}$ implies $\xi\in \pi^{(4),n}$. If $b$ bounds a curve of genus $1$, say $g(S')=1$ (the case $g-g(S')=1$ is similar), then $v=[\alpha_1,\alpha_{-1}]$. We define $f :\pi \to G=\langle x,y,z\rangle$ by $\alpha_1,\alpha_{-2}\mapsto x$, $\alpha_{-1},\alpha_{2}\mapsto y$ and other generators $\mapsto 1$. We see that $f(\xi)= [x,y]^{-m_b}y[x,y]^{m_b}y^{-1}\equiv [[x,y],y]^{-m_b}$ modulo $G^{(4)}$. Thus $\xi\in \pi^{(4),n}$ implies $n_6 | m_b$ by Lemma \ref{berekening}. If $b$ does not bound a curve of genus 1 (i.e., $g(S')\geq 2$ and $g-g(S')\geq 2$) we define $g:\pi \to G=\langle x,y,z\rangle$ by $\alpha_1, \alpha_{-g} \mapsto x$, $\alpha_{-1}, \alpha_g \mapsto y$ , $\alpha_{g(S')+1} \mapsto z^{-1}$ and other generators $\mapsto 1$. We get $g(\xi)=[x,y]^{-m_b}z^{-1}[x,y]^{m_b}z\equiv [[x,y],z]^{m_b}$. Application of Lemma \ref{berekening} gives $n_2|m_b$ as desired. Case 2: case 1 does not occur. This means that every connected component of $S$ cut out by one bridge involved in $\sigma$ contains a maximal cut system. Choose a maximal cut system as in Proposition \ref{nice cut system}. Let $E_1=\{e_{1,0}, \dots, e_{1,f_1} \}$ be this cut system. Suppose the edges are numbered cyclically and suppose the connected component cut out by $e_{1,0}$ and $e_{1,1}$ which does not contain any edge of $E_1$ contains all other maximal cut systems. Call this component $S'$. By the choice of $E_1$ and the assumption, we get that $S \setminus S'$ does not contain any bridges involved in $\sigma$. The situation thus looks as shown below: \vspace{6.5 cm} $$\hbox{\sl Fig.~6}$$ Arguing as in the proof for $k=2$, we arrive at a curve with a cut pair $e_{1,0}$ and $e_{1,1}$ which is a maximal cut system with the property that one of the two components cut out by the cut pair contains all other edges involved in the counter example $\sigma$. Considering $f : \pi \to G$ as in Section \ref{k is 2}, we get, as in case 1, that the element $z^{-m_{1,1}}\cdot (z\cdot [y,x])^{m_{1,1}}$ lies in $G^{(4),n}$. Note that $$z^{-m_{1,1}}\cdot (z\cdot [y,x])^{m_{1,1}}\equiv [y,x]^{m_{1,1}}[[y,x],z]^{\frac{1}{2}m_{1,1}(m_{1,1}-1)}.$$ Thus we get our divisibility condition on $m_{1,1}$ by Lemma \ref{berekening}. \section{Computations for a free group on three generators} \label{group structure} \label{berekeningen} \begin{lemmatje} Let $G$ be a free group on three generators, $G=\langle x,y,z\rangle$. Then $$ G/G^{(4)} \cong \{(i_1,\ldots,i_{14})\in {\msy Z}^{14} \}$$ where multiplication on the right-hand side is given as follows: $$(i_1,i_2,i_3,i_4,i_5,i_6,i_7,i_8,i_9,i_{10},i_{11},i_{12},i_{13},i_{14}) \cdot $$ $$(j_1,j_2,j_3,j_4,j_5,j_6,j_7,j_8,j_9,j_{10},j_{11},j_{12},j_{13},j_{14})=$$ \[ \left( \begin{array}{c} i_1 +j_1 \\ i_2 +j_2 \\ i_3 +j_3 \\ i_4+j_4 -i_2 j_1 \\ i_5+j_5-i_3 j_2 \\ i_6+j_6-i_3j_1 \\ i_7+j_7 +i_4j_1- \frac{1}{2}(j_1-1)j_1i_2 \\ i_8+j_8 +i_4j_2- \frac{1}{2}(i_2-1)i_2j_1 -i_2j_1j_2 \\ i_9+j_9 +i_4j_3+i_6j_2 -i_2i_3j_1-i_2j_1j_3-i_3j_1j_2 \\ i_{10}+j_{10}+i_5j_1+i_6j_2 -i_3j_1j_2 \\ i_{11}+j_{11}+i_5j_2 -\frac{1}{2}(j_2-1)j_2i_3 \\ i_{12}+j_{12}+i_5j_3-\frac{1}{2}(i_3-1)i_3j_2-i_3j_2j_3 \\ i_{13}+j_{13}+i_6j_1-\frac{1}{2}(j_1-1)j_1i_3 \\ i_{14}+j_{14}+i_6j_3-\frac{1}{2}(i_3-1)i_3j_1-i_3j_1j_3 \end{array} \right) \] \end{lemmatje} \begin{proof} We have the exact sequences $$1 \rightarrow \Gk/\Gkeen \rightarrow G/ \Gkeen \rightarrow G/ \Gk \rightarrow 1 ,$$ where $\Gk/\Gkeen$ is a finitely generated free abelian group (\cite[Th.5.12]{Magnus}) of rank $N_i:=\frac{1}{i} \sum_{d|i} \mu(d) 3^{i/d}$ (\cite[Th.5.11]{Magnus}), where $\mu$ is the M\"obius-function. Explicitly, we have, with $r:=[x,y]$, $s:=[y,z]$ and $t:=[x,z]$: \[ \begin{array}{l} G^{(1)}/G^{(2)} = \langle x,y,z\rangle_{ab} ~,\\ G^{(2)}/G^{(3)} = \langle r,s,t\rangle_{ab}~, \\ G^{(3)}/G^{(4)} = \langle [r,x],[r,y],[r,z],[s,x],[s,y],[s,z],[t,x],[t,z]\rangle_{ab}~, \end{array} \] by the Jacobi-relation $[t,y] \equiv [r,z][s,x]$ mod $G^{(4)}$. Now recall that if $a \in \Gk$ and $b \in G^{(l)}$ then $ab=ba[a,b]$ with $[a,b] \in G^{(k+l)}$, so that \begin{eqnarray} ab\equiv ba \mbox{ mod } G^{(k+l)}. \end{eqnarray} Furthermore we have the following identities: (\cite[Th.5.1]{Magnus}) \begin{eqnarray} \ [a,b] & = & [b,a]^{-1},\\ \ [a,bc] & = & [a,c]~[a,b]~[[a,b],c], \\ \ [ab,c] & = & [a,c]~[[a,c],b]~[b,c]. \end{eqnarray} It is clear now that any element $a$ of $G/G^{(4)}$ can be written in the form $a=x^{i_{1}}\cdots [t,z]^{i_{14}}$ in a unique manner. Note that the last eleven factors commute by (1). We define a map $\phi:G/G^{(4)} \rightarrow \{(i_1,\ldots,i_{14}) \in {\msy Z}^{14} \}$ by $$x^{i_1}y^{i_2}z^{i_3}r^{i_4}s^{i_5}t^{i_6}[r,x]^{i_7}[r,y]^{i_8}[r,z]^{i_9} [s,x]^{i_{10}}[s,y]^{i_{11}}[s,z]^{i_{12}}[t,x]^{i_{13}}[t,z]^{i_{14}} \mapsto (i_1,\ldots,i_{14}) .$$ Before we start the computation, notice that modulo $G^{(4)}$ we have \begin{eqnarray} [a^i,b^j]\equiv [a,b]^{ij}[[a,b],a]^{\frac{1}{2}ij(i-1)}[[a,b],b]^{\frac{1}{2}ij(j-1)}, \end{eqnarray} as one proves easily by induction. We set $(n,m):=\frac{1}{2}(n-1)nm$. The following identities hold in the group $G/G^{(4)}$: \begin{eqnarray*} &&x^{i_1}y^{i_2}z^{i_3}x^{j_1}y^{j_2}z^{j_3}\\ &=& x^{i_1+j_1}x^{-j_1}y^{i_2}z^{i_3}x^{j_1}z^{-i_3}y^{-i_2}y^{i_2}z^{i_3} y^{j_2}z^{j_3}\\ &=& x^{i_1+j_1}[x^{j_1},z^{-i_3}y^{-i_2}]y^{i_2+j_2}y^{-j_2}z^{i_3} y^{j_2}z^{-i_3}z^{i_3+j_3}\\ &=& x^{i_1+j_1}[x^{j_1},y^{-i_2}][x^{j_1},z^{-i_3}][[x^{j_1},z^{-i_3}],y^{-i_2}] y^{i_2+j_2}[y^{j_2},z^{-i_3}]z^{i_3+j_3}\\ &=& x^{i_1+j_1}r^{-i_2j_1}[r,x]^{(j_1,-i_2)}[r,y]^{(-i_2,j_1)}t^{-i_3j_1} [t,x]^{(j_1,-i_3)}[t,z]^{(-i_3,j_1)} \\ &&\ \ \ y^{i_2+j_2}s^{-i_3j_2}[s,y]^{(j_2,-i_3)} [s,z]^{(-i_3,j_2)}z^{i_3+j_3}[t,y]^{i_2i_3j_1}\\ &=& x^{i_1+j_1}y^{i_2+j_2}z^{i_3+j_3}r^{-i_2j_1}s^{-i_3j_2}t^{-i_3j_1} [r,y]^{-i_2j_1(i_2+j_2)}[t,y]^{-i_3j_1(i_2+j_2)}[r,z]^{-i_2j_1(i_3+j_3)} \\ &&\ \ \ [t,z]^{-i_3j_1(i_3+j_3)}[s,z]^{-i_3j_2(i_3+j_3)} [r,x]^{(j_1,-i_2)}[r,y]^{(-i_2,j_1)}[t,x]^{(j_1,-i_3)}[t,z]^{(-i_3,j_1)} \\ &&\ \ \ [s,y]^{(j_2,-i_3)}[s,z]^{(-i_3,j_2)}[r,z]^{j_1i_2i_3}[s,x]^{j_1i_2i_3}\\ &=&x^{i_1+j_1}y^{i_2+j_2}z^{i_3+j_3}r^{-i_2j_1}s^{-i_3j_2}t^{-i_3j_1} [r,x]^{(j_1,-i_2)}[r,y]^{-(i_2,j_1)-i_2j_1j_2}[r,z]^{-i_2j_1j_3-i_3j_1j_2 -i_2i_3j_1} \\ &&\ \ \ [s,x]^{-i_3j_1j_2}[s,y]^{(j_2,-i_3)}[s,z]^{-(i_3,j_2)-i_3j_2j_3} [t,x]^{(j_1,-i_3)}[t,z]^{-(i_3,j_1)-i_3j_1j_3}. \end{eqnarray*} Also \begin{eqnarray*} r^{i_4}s^{i_5}t^{i_6}x^{j_1}y^{j_2}z^{j_3}r^{j_4}s^{j_5}t^{j_6}&=& x^{j_1}y^{j_2}z^{j_3}r^{i_4}s^{i_5}t^{i_6}[r^{i_4}s^{i_5}t^{i_6}, x^{j_1}y^{j_2}z^{j_3}]r^{j_4}s^{j_5}t^{j_6}\\ &=& x^{j_1}y^{j_2}z^{j_3}r^{i_4+j_4}s^{i_5+j_5}t^{i_6+j_6}[r,x]^{i_4j_1} [r,y]^{i_4j_2}[r,z]^{i_4j_3+i_6j_2}\\ &&\ \ \ {[s,x]}^{i_5j_1+i_6j_2}[s,y]^{i_5j_2} [s,z]^{i_5j_3}[t,x]^{i_6j_1}[t,z]^{i_6j_3}. \end{eqnarray*} Combining these proves the lemma. \end{proof} \begin{lemmatje} $(i_1,\cdots,i_{14})^n=$ \[ \left( \begin{array}{c} ni_1 \\ ni_2 \\ ni_3 \\ni_4-\frac{1}{2}(n-1)ni_2i_1\\ ni_5-\frac{1}{2}(n-1)ni_3i_2\\ ni_6-\frac{1}{2}(n-1)ni_3i_1\\ ni_7+\frac{1}{2}(n-1)ni_4i_1-\frac{1}{12}(n-1)n(2n-1)i_1^2i_2 +\frac{1}{4}(n-1)ni_1i_2 \\ ni_8 +\frac{1}{2}(n-1)ni_4i_2-\frac{1}{6}(n-1)n(2n-1)i_1i_2^2 +\frac{1}{4}(n-1)ni_1i_2-\frac{1}{4}(n-1)ni_1i_2^2\\ ni_9 +\frac{1}{2}(n-1)ni_4i_3 +\frac{1}{2}(n-1)ni_6i_2 -\frac{1}{3}(n-1)n(2n-1)i_3i_1i_2 -\frac{1}{2}(n-1)ni_1i_2i_3\\ ni_{10} +\frac{1}{2}(n-1)ni_5i_1 +\frac{1}{2}(n-1)ni_6i_2 -\frac{1}{6}(n-1)n(2n-1)i_3i_1i_2 \\ ni_{11}+\frac{1}{2}(n-1)ni_5i_2-\frac{1}{12}(n-1)n(2n-1)i_2^2i_3 +\frac{1}{4}(n-1)ni_2i_3 \\ ni_{12} +\frac{1}{2}(n-1)ni_5i_3-\frac{1}{6}(n-1)n(2n-1)i_2i_3^2 +\frac{1}{4}(n-1)ni_3i_2-\frac{1}{4}(n-1)ni_2i_3^2\\ ni_{13}+\frac{1}{2}(n-1)ni_6i_1-\frac{1}{12}(n-1)n(2n-1)i_1^2i_3 +\frac{1}{4}(n-1)ni_1i_3 \\ ni_{14} +\frac{1}{2}(n-1)ni_6i_3-\frac{1}{6}(n-1)n(2n-1)i_1i_3^2 +\frac{1}{4}(n-1)ni_3i_1-\frac{1}{4}(n-1)ni_1i_3^2\\ \end{array} \right) \] \end{lemmatje} \begin{proof} By induction on $n$. \end{proof} \begin{lemmatje} \label{berekening} Notations as above. \[ \begin{array}{rcl} G^{(4),n}/G^{(4)} & \cong & \{(i_1,\ldots,i_{14})| i_1,i_2,i_3 \in n{\msy Z},~ i_4,i_5,i_6 \in n_2{\msy Z}, \\ & & i_7,\ldots,i_{14} \in n_6{\msy Z},~i_9 + i_{10} \in n_2 {\msy Z} \}. \end{array} \] (Recall that $n_i:= n / {\rm gcd} (n,i).) $\end{lemmatje} \begin{proof} The inclusion ``$\subset$'' follows at once from Lemma 4.2 and the following two observations: $$-\frac{1}{12}(n-1)n(2n-1)i_1^2i_2+\frac{1}{4}(n-1)ni_1i_2=$$ $$\frac{1}{12}(n-1)ni_1i_2[-(2n-1)i_1+3]$$ is an element of $n_6 \msy Z$ (either $i_1$ is even or the two terms between brackets have the same parity). The same holds for $$-\frac{1}{6}(n-1)n(2n-1)i_1i_2^2 +\frac{1}{4}(n-1)ni_1i_2-\frac{1}{4}(n-1)ni_1i_2^2=$$ $$\frac{1}{12}(n-1)ni_1i_2[-(4n+1)i_2+3].$$ The other inclusion follows from a direct computation: \[ \begin{array}{rl} a:= & (1,0,\ldots,0)^n(-1,1,0,\ldots,0)^n(0,-1,0,\ldots,0)^n \\ =& (0,0,0,\frac{1}{2}(n-1)n,0,0,-\frac{1}{6}(n-1)n(n+1),-\frac{1}{6}(n-1)n(n+1), 0,\ldots,0),\\ b:= & (-1,0,\ldots,0)^n (1,1,0\ldots,0)^n (0,-1,\ldots,0)^n\\ =& (0,0,0,-\frac{1}{2}(n-1)n,0,0,-\frac{1}{6}(n-2)(n-1)n, \frac{1}{6}(n-1)n(n+1),0,\ldots,0),\\ c:= & (1,0\ldots,0)^n(-1,-1,0,\ldots,0)^n(0,1,0,\ldots.,0)^n \\ =& (0,0,0,-\frac{1}{2}(n-1)n,0,0,\frac{1}{6}(n-1)n(n+1),-\frac{1}{6}(n-2)(n-1)n, 0,\ldots,0). \end{array} \] So we have: $$ab=(0,0,0,0,0,0,-\frac{1}{6}(n-1)n(2n-1),0,\ldots,0),$$ $$ac=(0,0,0,0,0,0,0,-\frac{1}{6}(n-1)n(2n-1),0,\ldots,0).$$ The same computation holds for the eleventh till the fourteenth coefficient, so it is possible to get $n_6$ as one of the coefficients $i_7,i_8,i_{11},..,i_{14}$ and all the other coefficients zero. Thus, looking again at $a$, we see that it is possible to get $n_2$ as the coefficient $i_4$ and all the other coefficients zero. By symmetry the same holds for the coefficients $i_5,i_6$.\\ Furthermore, if we put $$d:= (0,0,-1,0,\ldots,0)^n(0,0,1,1,0,\ldots,0)^n(0,0,0,-1,0,\ldots.,)^n,$$ $$e:= (-1,0,\ldots,0)^n(1,0,0,0,1,0,\ldots,0)^n(0,0,0,0,-1,0,\ldots.,)^n,$$ then we have: \[ \begin{array}{l} d=(0,0,0,0,0,0,0,0,\frac{1}{2}(n-1)n,0,\ldots,0), \\ e=(0,0,0,0,0,0,0,0,0,\frac{1}{2}(n-1)n,0,\ldots,0). \end{array} \] Thus, we can get $n_2$ in the ninth and tenth coefficient, independently. The rest now follows from what we have proven already and the computation of $$(-1,0,\ldots,0)^n(0,-1,0,\ldots,0)^n(0,0,-1,0,\ldots,0)^n(1,1,1,0,\ldots,0)^n.$$ \end{proof} \begin{cortje} \label{free comp} Notations as above. \hfill \break 1. $(G^{(2)} \cap G^{(3),n})/G^{(3)} \cong n_2G^{(2)}/G^{(3)}.$ \hfill\break 2. $(i_7,\ldots,i_{14}) \in (G^{(3)} \cap G^{(4),n})/G^{(4)} $ if and only if $i_7,\ldots,i_{14} \in n_6 {\msy Z},~i_9 +i_{10} \in n_2 {\msy Z}.$ \end{cortje}
1996-03-05T06:16:17
9501
alg-geom/9501007
en
https://arxiv.org/abs/alg-geom/9501007
[ "alg-geom", "math.AG" ]
alg-geom/9501007
Gerd Dethloff
Gerd Dethloff and Mikhail Zaidenberg
Plane Curves with Hyperbolic and C-hyperbolic Complements
Final version, published in 1996, subsuming version 1 of this preprint and another work by the same authors on "Examples of Plane Curves with Hyperbolic and C-hyperbolic Complements"
Ann. Scuola Norm. Sup. Pisa 23, 749-778 (1996)
null
null
http://arxiv.org/licenses/nonexclusive-distrib/1.0/
The general problem which initiated this work is: What are the quasiprojective varieties which can be uniformized by means of bounded domains in $\cz^n$ ? Such a variety should be, in particular, C--hyperbolic, i.e. it should have a Carath\'{e}odory hyperbolic covering. We study here the plane projective curves whose complements are C--hyperbolic. For instance, we show that most of the curves whose duals are nodal or, more generally, immersed curves, belong to this class. We also give explicit examples of irreducible such curves of any even degree d greater or equal 6.
[ { "version": "v1", "created": "Tue, 17 Jan 1995 11:49:50 GMT" }, { "version": "v2", "created": "Sat, 29 Nov 2014 14:18:27 GMT" } ]
2014-12-02T00:00:00
[ [ "Dethloff", "Gerd", "" ], [ "Zaidenberg", "Mikhail", "" ] ]
alg-geom
\section{Introduction} \noindent {\bf 1.1.} A complex space $X$ is called {\it C--hyperbolic} if it has a (non--ramified) covering $\tilde{X}$ which is {\it Carath\'{e}odory hyperbolic}, i.e. the points of $\tilde{X}$ can be separated by bounded holomorphic functions [Ko1, pp. 129--130] (see also [LiZa, 1.3]). In this paper we study C--hyperbolicity of the complements of plane projective curves. In particular, we are interesting in what the minimal degree of a plane curve with C--hyperbolic complement is. It is well known that any C--hyperbolic space is Kobayashi hyperbolic [Ko1, p. 130]. The opposite property to Carath\'{e}odory hyperbolicity is {\it liouvilleness}. A complex space $X$ is called {\it Liouville} if it has no non-constant bounded holomorphic functions. For example, any quasi--projective variety $X$ is Liouville, and by the theorem of V. Lin its liouvilleness is preserved by passing to a nilpotent covering over $X$, i.e. a Galois covering with nilpotent group of deck transformations [Li, Theorem B] (see also [LiZa], Theorem 3 at p.119). Thus, if $X$ is a quasi--projective variety whose Poincar\'{e} group $\pi_1(X)$ is (almost) nilpotent, then any covering over $X$ is Liouville and therefore $X$ can not be C--hyperbolic. In particular, this is the case for $X= I \!\! P^2 \setminus C$, where $C$ is a (not necessarily irreducible) nodal curve, i.e. a plane curve with normal crossing singularities only. Indeed, in this case by the Deligne-Fulton theorem [Del, Fu] the fundamental group $\pi_1 (X)$ is abelian, and thus by Lin's Theorem any covering over $X= I \!\! P^2 \setminus C$ is a Liouville one. The fundamental group $\pi_1 ( I \!\! P^2 \setminus C)$ for an irreducible plane curve $C$ of degree $d$, which is not necessarily nodal, is known to be abelian in a number of other cases, and hence to be isomorphic to $ Z \!\!\! Z / d Z \!\!\! Z$ (see e.g. the survey article [Lib] and the references therein). For instance, this is so for any rational or elliptic Pl\"ucker curve except those of even degree with the maximal number of cusps, and therefore also for the curves that can be specialized to such ones [Zar, pp. 267, 327-330] (cf. [DL], [Kan]). This is true as well for any irreducible curve of degree $d \le 4$ with the only exception of the three--cuspidal quartic; in the latter case $\pi_1 ( I \!\! P^2 \setminus C)$ is a finite non-abelian metacyclic group of order $12$ [Zar, pp. 135, 145], and so it is almost abelian. Therefore, in all these cases any covering over $ I \!\! P^2 \setminus C$ is a Liouville one. \\ [1ex] \noindent {\bf 1.2.} At the same time, the complement of a nodal plane curve can be Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. The well known example is the complement of five lines in $ I \!\! P^2$ in general position [Gr3; KiKo, Corollary 3 in section 4]; for further examples of reducible curves see e.g. [DSW1,2] and the literature therein. There exist even the irreducible smooth quintics with these properties [Za3]. Moreover, Y.-T. Siu and S.-K. Yeung [SY] have announced recently a proof of a long standing conjecture that generic (in Zariski sense) smooth curve in $ I \!\! P^2$ of degree $d$ large enough ($d \ge 1,200,000$) has the complement which is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$ (while all its coverings are Liouville). This shows that for the complement of a curve in $ I \!\! P^2$ the property to be C--hyperbolic is much stronger than those of Kobayashi hyperbolicity, and it can occure only for the curves with singularities worse than the ordinary double points.\\[1ex] \noindent {\bf 1.3.} However, plane curves with C--hyperbolic complements do exist. The simplest example is a reducible quintic $C_5$ with the ordinary triple points as singularities at worst. Namely, $C_5$ is the union of five lines which is given in homogeneous coordinates $(x_0:x_1:x_2)$ in $ I \!\! P^2$ by the equation $$x_0 x_1 x_2 (x_0 -x_1 )(x_0 -x_2) =0\,\,$$ \begin{center} \begin{picture}(300,90) \thicklines \put(30,85){\line(0,-1){85}} \put(20,85){\line(1,-1){85}} \put(25,85){\line(1,-2){45}} \put(20,10){\line(1,0){95}} \put(20,47){\line(2,-1){90}} \put(50,-15){$C_5$} \put(200,85){\line(0,-1){85}} \put(190,85){\line(1,-1){85}} \put(195,85){\line(1,-2){45}} \put(190,10){\line(1,0){95}} \put(190,47){\line(2,-1){90}} \put(190,0){\line(1,1){55}} \put(218,-15){$C_6$} \end{picture} \end{center} \begin{center} Figure 1 \end{center} \noindent Indeed, $ I \!\! P^2 \setminus C_5$ is biholomorphic to $( I \!\!\!\! C^{**})^2$, where $ I \!\!\!\! C^{**} = I \!\! P^1 \setminus \{3\,{\rm points}\}$, and thus its universal covering is the bidisk $\Delta^2$ (hereafter $\Delta=\{z\in I \!\!\!\! C\,|\,|z| < 1\}$ denotes the unit disc). Slightly modifying the previous example, consider further the reducible sextic $C_6 \subset I \!\! P^2$ which is the line arrangement given by the equation $$x_0 x_1 x_2 (x_0 -x_1 )(x_0 -x_2)(x_1 - x_2) =0\,\,.$$ \noindent It is known [Kal] that the universal covering of the complement $M_2 := I \!\! P^2 \setminus C_6$ is biholomorphic to the Teichm\"uller space $T_{0,\, 5}$ of the Riemann sphere with five punctures. Furthermore, via the Bers embedding $T_{0, \,5} \hookrightarrow I \!\!\!\! C^2$ it is biholomorphic to a bounded Bergman domain of holomorphy in $ I \!\!\!\! C^2$, which is contractible and Kobayashi complete hyperbolic. The automorphism group of $T_{0,\,5}$ is discrete and isomorphic to the mapping class group, or modular group, ${\rm Mod}(0,\,5)$ [Ro]. Note that $5$ is the minimal degree of a plane curve whose complement is C--hyperbolic. Indeed, the complement of a quartic curve is not even Kobayashi hyperbolic [Gr2, section 6]. While there do exist irreducible plane sextics with C--hyperbolic complements (see Proposition 4.5 below), there does not exist such an irreducible plane quintic (see (4.8)-(4.10)), and so the minimal degree of an irreducible plane curve with C--hyperbolic complement is $6$ (Proposition 4.10). \\[1ex] \noindent {\bf 1.4.} To obtain examples of irreducible plane curves whose complements are C--hyperbolic we can apply the method that was used by M. Green [Gr1] (see also [CaGr, GP]) for constructing curves with hyperbolic complements. In this paper we study systematically the class of curves with C--hyperbolic complements which can be obtained by this method. Let us describe briefly its main idea. Let $S^n X$ denote the n--th symmetric power of a variety $X$ and $R_n \subset S^n X$ be its discriminant variety, i.e. the ramification locus of the branched Galois covering $s_n \,:\, X^n \to S^n X$. For a plane curve $C \subset I \!\! P^2$ there is a natural embedding $\rho_C \,:\, I \!\! P^2 \hookrightarrow S^n C^*_{norm}$, where $C^* \subset I \!\! P^{2*}$ is the dual curve, $n = {\rm deg}\, C^*$ and $C^*_{norm}$ is the normalization of $C^*$. It may happen that this gives an embedding of the complement $ I \!\! P^2 \setminus C$ into the n--th configuration space $S^n C^* \setminus R_n$, and that either this configuration space, or some subspace of it containing the image $\rho_C ( I \!\! P^2 \setminus C )$ has nice hyperbolic properties. Here we give an example. Suppose that the dual curve $C^* \subset I \!\! P^{2*}$ is smooth and of degree $n \ge 4$. For $z = (a:b:c) \in I \!\! P^2$ denote by $\rho_C (z)$ the non--ordered set of $n$ points of intersection $l_z \cap C^*$, where $l_z$ is the dual line $ax_0^* + bx_1^* + cx_2^* = 0$ in $ I \!\! P^{2*}$; here an intersection point of multiplicity $m$ is repeated $m$ times. In this way we obtain a morphism $\rho_C : I \!\! P^2 \to S^nC^*$ into $n$--th symmetric power of $C^*$, which is a smooth variety (see e.g. [Zar, p.253] or [Na, (5.2.15)]). The ramified covering $s_n : {C^*}^n \to S^nC^*$ has the ramification divisors $D_n := \bigcup\limits_{1\le i< j\le n} D_{ij} \subset {C^*}^n$ resp. $R_n = s_n (D_n) \subset S^nC^*$, where $D_{ij} := \{ x = (x_1,\dots,x_n) \in {C^*}^n\,|\,x_i=x_j\}$ is a diagonal hypersurface. Following Zariski [Zar, p.266] we call $R_n$ {\it the discriminant hypersurface}. Since $C^*$ is smooth, the preimage $\rho_C^{-1} (R_n)$ coincides with $C$, and so we have the commutative diagram \begin{center} \begin{picture}(1000,60) \thicklines \put(100,5){$ I \!\! P^2 \setminus C = X$} \put(155,45){$Y$} \put(157,38){\vector(0,-1){20}} \put(130,25){${\tilde s}_n$} \put(200,58){${\tilde \rho}_C$} \put(200,20){$\rho_C$} \put(200,45){$\hookrightarrow$} \put(200,5){$\hookrightarrow$} \put(245,5){$S^n C^* \setminus R_n \,\,\,\,$} \put(250,45){${C^*}^n \setminus D_n$} \put(275,38){\vector(0,-1){20}} \put(255,25){$s_n$} \put(400,25){(1)} \put(320,45){$\hookrightarrow$} \put(350,45){${C^*}^n$} \end{picture} \end{center} \noindent where ${\tilde s}_n \,:\,Y \to X$ is the induced covering. The genus $g(C^*) \ge 3$, therefore ${C^*}^n$ has the polydisc $\Delta ^n$ as the universal covering. Passing to the induced covering $Z \to Y$ we can extend (1) to the diagram \begin{center} \begin{picture}(1000,80) \thicklines \put(155,75){$Z$} \put(158,69){\vector(0,-1){18}} \put(155,40){$Y$} \put(158,36){\vector(0,-1){18}} \put(100,5){$ I \!\! P^2 \setminus C = X$} \put(133,23){${\tilde s}_n$} \put(200,75){$\hookrightarrow$} \put(200,40){$\hookrightarrow$} \put(200,5){$\hookrightarrow$} \put(200,55){${\tilde \rho}_C$} \put(200,20){$\rho_C$} \put(253,75){$\Delta^n$} \put(258,69){\vector(0,-1){18}} \put(252,40){${C^*}^n$} \put(258,36){\vector(0,-1){18}} \put(247,5){$S^n C^*$} \put(264,25){$s_n$} \put(400,40){(2)} \end{picture} \end{center} \noindent Being a submanifold of the polydisc, $Z$ is Carath\'eodory hyperbolic, and so $X$ is C--hyperbolic. Therefore, we have proved the following assertion.\\[1ex] \noindent {\bf 1.5. Proposition.} {\it Let $C \subset I \!\! P^2$ be an irreducible curve whose dual curve $C^*$ is smooth and of degree at least $4$. Then $ I \!\! P^2 \setminus C$ is C--hyperbolic.}\\[1ex] \noindent {\bf 1.6.} Note that, furthermore, $ I \!\! P^2 \setminus C$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. The latter is also true under the weaker assumptions that (a) the geometric genus $G$ of $C$ is at least two; (b) each tangent line to $C^*$ intersects with $C^*$ in at least two points, and (c) the following inequality is fulfilled: $2n < d$, where $d = {\rm deg}\,C$ and $n = {\rm deg}\,C^*$, or, what is equivalent, $$\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\sum \limits_{i=1}^l (m^*_i - 1) < 2g-2 \,\,\,,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,(3)$$ where $m^*_1,\dots,m^*_l$ are the multiplicities of the singular branches of $C^*$ [Gr1, CaGr, GP]. Moreover, under these assumptions a stronger conclusion is valid. Namely, there exists a continuous hermitian metric on $ I \!\! P^2 \setminus C$ whose holomorphic sectional curvature is bounded from above by a negative constant and which dominates some positive multiple of the Fubini-Study metric on $ I \!\! P^2$ [GP]. \\ [1ex] \noindent {\bf 1.7.} It is clear that a subspace of a C--hyperbolic space is also C--hyperbolic. In particular, if $D \subset I \!\! P^n$ is a hypersurface whose complement $ I \!\! P^n \setminus D$ is C--hyperbolic, then any plane section of $D$ is a plane curve with C--hyperbolic complement. In this way, considering curve complements, one might at least obtain necessary conditions for $ I \!\! P^n \setminus D$ to be C--hyperbolic (cf. [Za2]).\\[1ex] \noindent {\bf 1.8.} {\it Contents of the paper.} The main results are summarized at Theorem 7.12 at the very end of the paper. Besides this Introduction, the paper contains six sections. Sections 2 and 3 are preliminary. The first of them deals with some necessary facts from hyperbolic analysis, while in the second one certain generalities on plane curves are given. In section 4 we prove C--hyperbolicity of the complements of irreducible curves of genus at least $1$, whose duals are immersed curves (for instance, nodal curves) (see Theorem 4.1). We give examples of such curves of any even degree $d \ge 6$. Furthermore, we study the general case when the dual curve $C^*$ may have cusps. Then, under the morphism $\rho_C \,:\, I \!\! P^2 \to S^n C^*_{norm}$, the discriminant divisor $R_n \subset S^n C^*_{norm}$, besides the curve $C$ itself, cuts out a line configuration $L_C \subset I \!\! P^2$ which consists of the dual lines of cusps of $C^* \subset I \!\! P^{2*}$; they are the inflexional tangents and some cuspidal tangents of $C$. We call $L_C$ {\it the artifacts of $C$} (see (3.3)). In Theorem 4.1 we prove C--hyperbolicity of $ I \!\! P^2 \setminus (C \cup L_C )$, where $C$ is an irreducible curve of genus $g \ge 1$. The case of rational curves is studied in sections 5 - 7. In section 5 we give necessary preliminaries. If $C$ is rational, then $S^n C^*_{norm} \cong I \!\! P^n \,,\,\,\,\rho_C \,:\, I \!\! P^2 \to I \!\! P^n \cong S^n C^*_{norm}$ is a linear embedding and the discriminant hypersurface $R_n \subset I \!\! P^n$ is the projective hypersurface defined by the usual discriminant of the universal polynomial of degree $n$. Therefore, we have the following assertion.\\[1ex] \noindent {\bf 1.9. Lemma} (cf. {\rm [Zar, p.266])}. {\it Any rational curve $C \subset I \!\! P^2$ whose dual $C^*$ is of degree $n$, together with its artifacts $L_C$ is a plane section of the discriminant hypersurface $R_n \subset I \!\! P^n$.} \\[1ex] \noindent We call $\rho_C$ {\it the Zariski embedding} \footnote{In an unexplicit way it is contained already in [Ve, Ch. IV, p.208]}. Using this lemma as well as a duality between the Zariski embedding and a projection of the rational normal curve (see 5.4-5.5), we establish an analog of Theorem 4.1 for a rational curve whose dual has at least one cusp. This is done in Theorem 6.5, where also all exceptions are listed. A classification of the orbits of the natural $ I \!\!\!\! C^*$--actions is an important ingredient of the proof. We give several concrete examples. In section 7 we deal with the rational curves whose duals are nodal Pl\"ucker curves, i.e. with the maximal cuspidal rational curves. For such a curve of degree $d \ge 8$ we prove that its complement is almost C--hyperbolic (Corollary 7.10; see 2.4 below for the terminology). The proof is based on passing to the moduli space of the punctured Riemann sphere and on a study of the orbits of the natural representation of the group $ I \!\! P GL (2;\, I \!\!\!\! C )$ on the projectivized space of binary forms. \\[1ex] The second of the authors had fruitful discussions on the content of section 7 with D. Akhiezer, M. Brion, Sh. Kaliman and H. Kraft; its his pleasure to thank all of them. He also is gratefull to the SFB-170 `Geometry and Analysis' at G\"ottingen University for its hospitality and excellent working conditions. \\[2ex] \section {Preliminaries in hyperbolic complex analysis} \noindent {\bf 2.1.} {\it Lin's Theorem.} Here we recall some definitions and facts from [Li]. A complex space $X$ is called {\it ultra--Liouville} if any bounded plurisubharmonic function on $X$ is constant. For instance, any quasi--projective variety is ultra--Liouville. By Lin's Theorem [Li, Theorems B and 3.5] any almost nilpotent (or even almost $\omega$--nilpotent) Galois covering of an ultra--Liouville complex space $X$ is Liouville. A covering is called {\it almost nilpotent} (resp. {\it almost $\omega$--nilpotent}) if its group of deck transformations is so. Recall that a group $G$ is {\it almost nilpotent} if it has a nilpotent subgroup of a finite index ($G$ is {\it almost $\omega$--nilpotent} if the union of the members of its upper central series is a subgroup of $G$ of finite index; for a finitely generated group $G$ the almost $\omega$--nilpotency is equivalent to the almost nilpotency). \\ [1ex] \noindent {\bf 2.2.} {\it Super--liouvilleness}. Let us say that a complex space $X$ is {\it super-Liouville} if any covering over $X$ is Liouville. Super-liouvilleness is a property which in a sense is opposite to C-hyperbolicity. It is clear that $X$ is super-Liouville iff the universal covering $U_X$ of $X$ is Liouville. By Lin's theorem any ultra-Liouville complex space $X$ which has almost $\omega$--nilpotent fundamental group $\pi_1(X)$ is a super-Liouville one. In particular, a smooth quasi-projective curve $C$ is super-Liouville iff the group $\pi_1(C)$ is abelian, i.e. iff $C$ is non-hyperbolic. Note that if any two points of $X$ can be connected by a finite chain of Liouville subspaces (which are assumed to be connected but not necessarily closed), then $X$ itself is Liouville. More generally, we have the following lemma. \\[1ex] \noindent {\bf 2.3. Lemma.} {\it Let $X$ be a complex space (with countable topology) such that any two points of $X$ can be connected by a finite chain of super-Liouville subspaces of $X$. Then $X$ is super-Liouville. In particular, if $X$ is a quasi-projective variety such that each pair of points of $X$ can be connected by a finite chain of non-hyperbolic curves, then $X$ is super-Liouville.} \\[1ex] \noindent {\it Proof.} Let $\, \pi : U_X \rightarrow X \,$ be the universal covering. Suppose that there exists a non-constant bounded holomorphic function $f$ on $U_X$. Let ${\cal F}$ be the collection of all super-Liouville subspaces of $X$, and let $\tilde{{\cal F}}$ be the collection of subspaces consisting of all connected components of preimages $\, \pi^{-1}(A)$, where $\, A \in {\cal F}$. We define an equivalence relation on $U_X$ as follows: \noindent {\it Two points in $U_X$ are equivalent iff they can be connected by a finite chain of members of $\tilde{{\cal F}}$.} \noindent By the condition of the lemma it is easily seen that the union of the equivalence classes of the points of a given fibre of $\pi$ coincides with the whole space $U_X$. Since $\pi_1(X)$ is an at most countable group, the fibre of $\pi$ is at most countable, too, and therefore there exists an at most countable set of equivalence classes. If $M$ is any of them, then clearly $f|M \equiv const$. Therefore, $f$ takes at most countable set of values, which is impossible. The second statement is an easy corollary of the first one. \qed \noindent {\bf 2.4.} {\it Weak C--hyperbolicity}. We say that a complex space $X$ is {\it almost} resp. {\it weakly Carath\'eodory hyperbolic} if for any point $p \in X$ there exist only finitely many resp. countably many points $q \in X$ which cannot be separated from $p$ by bounded holomorphic functions (i.e. if the equivalence relation defined on $X$ by the functions from the algebra $H^{\infty} (X)$ is finite resp. at most countable). It will be called {\it almost} resp. {\it weakly C--hyperbolic} if $X$ has a covering $Y \to X$, where $Y$ is almost resp. weakly Carath\'eodory hyperbolic. These notions are meaningful due to the following reasons. It is unknown whether the universal covering space $U_X$ of a C--hyperbolic complex space $X$ is Carath\'eodory hyperbolic, or more generally, whether there is a Carath\'eodory hyperbolic covering $Y \to X$ which can be defined in a functorial way. In contrary, one can make the following observation.\\ \noindent {\it A complex space $X$ is weakly C--hyperbolic iff the universal covering space $U_X$ is weakly Carath\'eodory hyperbolic.}\\ \noindent Hereafter we assume $X$ to be reduced and with countable topology. In particular, the universal covering $U_X$ of a C--hyperbolic space $X$ is weakly Carath\'eodory hyperbolic. One may consider on $X$ the pseudo--distance which is the quotient of the Carath\'eodory pseudo--distance $c_{U_X }$ on $U_X$ (resp. the quotient of the inner Carath\'eodory pseudo--distance $c'_{U_X }$ resp. of the differential Carath\'eodory--Reiffen pseudo--distance $C_{U_X }$; see [Re]). All three of these quotient pseudo--distances are contracted by holomorphic mappings. Furthermore, the deck transformations on $U_X$ being isometries, the quotient of $C_{U_X }$ on $X$ is locally isometric to $C_{U_X }$ itself, and thus it is non-degenerate iff $C_{U_X }$ is so (for a weakly C--hyperbolic space $X$ it is at least non--trivial). The proof of the following lemma is easy and can be omitted.\\ [1ex] \noindent {\bf 2.5. Lemma.} {\it Let $f: Y \to X$ be a holomorphic mapping of complex spaces. If f is injective (resp. f has finite resp. at most countable fibres) and $X$ is C--hyperbolic (resp. almost resp. weakly C--hyperbolic), then so is $Y$.} \\[1ex] \noindent {\bf 2.6.} {\it Brody hyperbolicity}. Recall that a complex space $X$ is {\it Brody hyperbolic} if it contains no entire curve, i.e. if every holomorphic mapping $ I \!\!\!\! C \to X$ is constant. Note that sometimes by Brody hyperbolicity one means absence of {\it Brody entire curves} in $X$, i.e. entire curves whose derivatives are uniformly bounded with respect to a fixed hermitian metric on $X$ (see e.g. [Za3]). Usually this is enough in applications. But in this paper we do not need such a precision. It is clear that any weakly C--hyperbolic complex space is Brody hyperbolic. \\ [1ex] \noindent {\bf 2.7.} {\it Kobayashi hyperbolicity}. For a curve $C \subset I \!\! P^2$ denote by ${\rm sing}\, C$ the set of all singular points of $C$. Put ${\rm reg}\, C = C \setminus {\rm sing}\, C$. The next statement follows from Theorem 2.5 in [Za1]. \\[1ex] \noindent {\bf Proposition.} {\it Let the Riemann surface ${\rm reg}\, C$ be hyperbolic and $ I \!\! P^2 \setminus C$ be Brody hyperbolic (the latter is true, in particular, if $ I \!\! P^2 \setminus C$ is weakly C--hyperbolic). Then $ I \!\! P^2 \setminus C$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$.}\\ [1ex] Note that in Example 1.3 in the Introduction the first condition fails while the second one is fulfilled. It is easily seen that in this example $ I \!\! P^2 \setminus C$ is not hyperbolically embedded into $ I \!\! P^2$. In fact, the condition ` ${\rm reg}\, C$ is hyperbolic' is necessary for $ I \!\! P^2 \setminus C$ being hyperbolically embedded into $ I \!\! P^2$ [Za1, Corollary 1.3].\\[1ex] \noindent {\bf 2.8.} {\it Relative hyperbolicities}. Let $X$ be a complex space resp. a quasi--projective variety and $Z \subset X$ be a closed analytic subset resp. a closed algebraic subvariety. We say that $X$ is {\it Brody hyperbolic modulo $Z$} if any (non-constant) entire curve $\, I \!\!\!\! C \to X$ is contained in $Z$. For instance, this is the case if $X$ is Kobayashi hyperbolic modulo $Z$ (see [KiKo]). (We mention that in [Za4] the above property of relative Brody hyperbolicity was called {\it strong algebraic degeneracy}.) We will say that $X$ is {\it C--hyperbolic modulo $Z$} ({resp. \it almost resp. weakly C--hyperbolic modulo $Z$}) if there is a covering $\pi \,:\,Y \to X$ such that for each point $p \in Y$ any other point $q \in Y \setminus \pi^{-1} (Z)$ (resp. any other point $q \in Y \setminus \pi^{-1} (Z)$ besides only finitely many resp. besides only countably many such points) can be separated from $p$ by bounded holomorphic functions. By the monodromy theorem weak C--hyperbolicity of $X$ modulo $Z$ implies Brody hyperbolicity of $X$ modulo $Z$. The next lemma, which is a generalization of Lemma 2.5, easily follows from the definitions. \\[1ex] \noindent {\bf 2.9. Lemma.} {\it Let $f: Y \to X$ be a holomorphic mapping of complex spaces and let $Z$ be a closed complex subspace of $X$. If $f\,|\,(Y \setminus f^{-1} (Z) )$ is injective (resp. has finite resp. at most countable fibres) and $X$ is C--hyperbolic (resp. almost resp. weakly C--hyperbolic) modulo $Z$, then $Y$ is C--hyperbolic (resp. almost resp. weakly C--hyperbolic) modulo $f^{-1} (Z).$} \\ [2ex] \section {Background on plane algebraic curves} \noindent {\bf 3.1.} {\it Classical singularities. Immersed curves}. We say that a curve $C$ in $ I \!\! P^2$ has {\it classical singularities} if its singular points are nodes and ordinary cusps. It is called {\it a Pl\"ucker curve} if both $C$ and the dual curve $C^*$ have classical singularities only and no flex at a node. If the normalization mapping $\nu : C^*_{norm} \to C \hookrightarrow I \!\! P^2$ is an immersion, or, which is equivalent, if all irreducible local analytic branches of $C$ are smooth, then we say that $C$ is {\it an immersed curve}. An immersed curve $C$ is called {\it a curve with tidy} or {\it ordinary singularities}, or simply {\it a tidy curve}, if at each point $p \in C$ the local irreducible branches of $C$ have pairwise distinct tangents. By M. Noether's Theorem [Co] any plane curve can be transformed into a tidy curve by means of Cremona transformations.\\ \noindent {\bf 3.2.} {\it Cusps and flexes}. In the sequel by {\it a cusp} we mean an irreducible plane curve singularity. In particular, an irreducible local analytic branch $A$ of a plane curve $C$ with centrum $p \in C$ is a cusp iff it is singular. The tangent line to a cusp $A$ at $p$ is called {\it a cuspidal tangent of $C$}. Recall that {\it the multiplicity sequence} of a plane analytic germ $A$ at $p_0 \in A$ is the sequence of multiplicities of $A$ at $p_0$ and its infinitely nearby points. Following [Na, (1.5)] $A$ is called {\it a simple cusp} if its first Puiseaux pair is $(m, m+1)$, where $m = {\rm mult}_p A \ge 2$, or, what is the same, if the multiplicity sequence of $A$ is $(m,\, 1, \,1, \dots)$. By Lemma 1.5.7 in [Na] a cusp $A$ is a simple cusp iff the corresponding branch $A^*$ of the dual curve $C^*$ is smooth. In this case $A^*$ is a flex of order $m-1$ (see the definition below), and vice versa, if $A^*$ is a flex of order $m-1$, then $A$ is a simple cusp of multiplicity $m$. A simple c sp $A$ of multiplicity $2$ is called {\it an ordinary cusp} if $C$ is locally irreducible at $p$. A smooth irreducible local branch $A$ of the curve $C$ at a point $p \in C$ is called {\it a flex of order} $k$ if $i(A, T_p A; p) = k+2 \ge 3$. The tangent line $T_p A$ to a flex $A$ at $p$ is called {\it an inflexional tangent}. {\it An ordinary flex} is a flex of order $1$ at a smooth point of $C$. Thus, the dual $C^*$ is an immersed curve iff $C$ has no flex and all its cusps are simple. \\[1ex] \noindent {\bf 3.3.} {\it Artifacts}. If $C^*$ has cusps, denote by $L_C$ the union of their dual lines in $ I \!\! P^2$. Clearly, $L_C$ consists of the inflexional tangents of $C$ and the cuspidal tangents at those cusps of $C$ which are not simple. Due to some analogy in tomography, we call $L_C$ {\it the artifacts} of $C$. Note that the dual curve $C^*$ of $C$ is an immersed curve iff $L_C = \emptyset$. Such a curve $C$ may have complicated reducible singularities, which correspond to multiple tangents of $C^*$; for instance, it may have tacnodes, etc. \\[1ex] \noindent {\bf 3.4.} {\it The Class Formula}. Let $C \subset I \!\! P^2$ be an irrducible curve of degree $d\ge 2$, of geometric genus $g$ and of class $c$. Then $c=d^* ={\rm deg}\, C^*$ is defined by the Class Formula (see [Na, (1.5.4)]) $$\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,c=d^* = 2(g+d-1) - \sum\limits_{p \in {\rm sing}\, C} (m_p - r_p) =\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \,\,\,\,\,\,\,\,\,\,\,\,$$ $$\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \,\,\,\,\,\,\,\,\,\,\,\, = 2(g+d-1) - \sum\limits_{A = (A, p),\,p \in {\rm sing}\, C} (m_A - 1) \,\,\,,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, \,\,\,\,\,\,\,\,\,\,\,\,(4)$$ where $m_p = {\rm mult}_p C ,\,\,\,r_p$ is the number of irreducible analytic branches of $C$ at $p$, $A = (A, p)$ is a local analytic branch of $C$ at $p$ and $m_A$ is its multiplicity at $p$. In particular, $C$ is an immersed curve iff $d^* = 2(g+d-1)$ (indeed, this is the case iff the last sum in (4) vanishes). Furthermore, for an immersed curve $C$ one has $\,\,\,d^* \ge 2d+2 \ge 10\,\,\,$ if $\,\,\,g\ge 2 ,\,\,\,$ $d^* = 2d \ge 6\,\,\,$ if $g=1 ,\,\,\,$ and $d^* = 2d-2 \ge 2\,\,\,$ if $g=0$. For reader's convenience we recall here also the usual Pl\"ucker formulas: $$g = 1/2 (d - 1)(d - 2) - \delta - \kappa = 1/2 (d^* - 1)(d^* - 2) - b - f$$ $$d^* = d(d - 1) - 2\delta - 3\kappa\,\,\,\,\,{\rm and}\,\,\,\,\,d = d^* (d^* - 1) - 2b - 3f$$ for a Pl\"ucker curve $C$ with $\delta$ nodes, $\kappa$ cusps, $b$ bitangent lines and $f$ flexes. \\[2ex] \noindent {\bf 3.5.} {\it The $n$--th Abel--Jacobi map}. Let $M$ be a compact Riemann surface of genus $g$, and let $j\,:\,M \to J(M)$ be a fixed Abel--Jacobi embedding of $M$ into its Jacobian variety $J(M) \cong {\rm Pic^0}\, (M)$. The $n$--th symmetric power $S^n M$ may be identified with the space of effective divisors of degree $n$ on $M$. Let $\phi_n \,:\, S^n M \to J(M)$ be the $n$--th Abel--Jacobi map $D = p_1 + \dots + p_n \longmapsto \phi_n (D) := j(p_1 ) + \dots + j(p_n )$, so that $j = \phi_1$. We recall the following well known facts (see e.g. [GH, 2.2], [Zar, pp.352--353] or [Na, (5.2), (5.3)] and references therein): \\ \noindent i) $\phi_n$ is holomorphic; \noindent ii) (Abel's Theorem) $\phi_n^{-1} (\phi_n (D)) = |D| = I \!\! P H^0 (M, O([D])) \cong I \!\! P^{{\rm dim}\,|D|}$ is the complete linear system of $D$, where $D \in S^n M$ is an effective divisor of degree $n$ on $M$; \noindent iii) the natural injection $|D| \hookrightarrow S^n M$ is a holomorphic embedding, i.e. $|D| = \phi_n^{-1} (\phi_n (D))$ is a smooth subvariety in $S^n M$; \noindent iv) if $n \le g$, then $\phi_n \,:\,S^n M \to J(M)$ is generically one--to--one; in particular, the image $\phi_{g-1} (S^{g-1} M) \subset J(M)$ is a translation of the theta divisor $\Theta$ on $J(M)$; \noindent v) (Jacobi Inversion) if $n \ge g$, then $\phi_n \,:\,S^n M \to J(M)$ is surjective. For $n > 2g - 2$ it is an algebraic projective bundle (see [Ma]); in particular, if $g = 1$, then it is a $ I \!\! P^{n-1}$--bundle over $J(M) \cong M$. \\[1ex] \noindent {\bf 3.6.} {\it The Zariski embedding}. Let $C \subset I \!\! P^2$ be an irreducible curve of degree $d \ge 2$ and let $\nu : C^*_{norm} \to C^*$ be the normalization of the dual curve. Following Zariski [Zar, p.307, p.326] and M. Green [Gr1] (see also [DL]), as in (1.4) we consider the mapping $\rho_C \,: \, I \!\! P^2 \to S^n C^*_{norm}$ of $ I \!\! P^2$ into the $n$-th symmetric power of $C^*_{norm}$, where $n = {\rm deg}\, C^*$. We put $\rho_C (z) = \nu^* (l_z) \subset S^n C^*_{norm}$, where $z \in I \!\! P^2$ and $l_z \subset I \!\! P^{2*}$ is the dual line. Clearly, $ \rho_C: I \!\! P^2 \to S^n C^*_{norm}$ is holomorphic. We still denote by $D_n$ the union of the diagonal divisors in $(C^*_{norm})^n$ and by $R_n = s_n (D_n)$ the discriminant divisor, i.e. the ramification locus of the branched covering $s_n: (C^*_{norm})^n \to S^n C^*_{norm}$ (cf. (1.4)). It is clear that $\rho_C$ is a holomorphic embedding, which we call in the sequel {\it the Zariski embedding}. More precisely, it is composed of two embeddings $i_1 \,:\, I \!\! P^2 \hookrightarrow I \!\! P^{h(C)}$ and $i_2 \,:\, I \!\! P^{h(C)} \hookrightarrow S^n C^*_{norm}$ which are described below. Denote by $H$ a divisor of degree $n$ on $M:= C^*_{norm}$, which is the trace of a line cut of $C^*$ in $ I \!\! P^{2*}$. The two dimensional linear system $g^2_n$ of all such line cuts is naturally identified with the original projective plane $ I \!\! P^2 = ( I \!\! P^{2*})^*$. Let $h(C) := {\rm dim}\,|H|$; then $i_1 \,:\, I \!\! P^2 = g^2_n \hookrightarrow I \!\! P^{h(C)} = |H| $ is defined to be the canonical linear embedding of $g^2_n $ into the complete linear system $|H|$. (Let us mention that $g^2_n$ itself might be complete; for instance, this is the case when $C^*$ is a nodal curve with $\delta$ nodes, where $\delta < {n(n-2) \over 4}$ for $n$ even or $\delta < {(n-1)^2 \over 4}$ for $n$ odd; see [Na, p.115].) The Abel embedding $i_2 \,:\, I \!\! P^{h(C)} = |H| \hookrightarrow S^n C^*_{norm} = S^n M$ identifies $|H|$ with the fibre $ \phi_n^{-1} (\phi_n (H))$ of the $n$--th Abel--Jacobi map $\phi_n \,:\,S^n M \to J(M)\,$ (see (3.5)). What we really need in section 4 is that the restriction $\rho_C\,|\, ( I \!\! P^2 \setminus C)$ is injective, which can also be shown as follows. We have to show that any projective line $l$ in $ I \!\! P^{2*}$ which is not tangent to $C^*$ meets $C^*$ in at least two distinct points, and so it is uniquely defined by its image $\rho (l) = l \cap C^*$. Suppose that there exists a line $l_0 \subset I \!\! P^{2*}$ which has only one point $p_0$ in common with $C^*$ and which is not tangent to $C^*$ at this point. Let $l_1$ be the tangent to a local analytic branch of $C^*$ at $p_0$. Then we have $$i(l_1, C^*; p_0) > i(l_0, C^*; p_0) = n = l_1 \cdot C^* \,\,\,,$$ which is impossible. In the same elementary way it can be shown that $\rho_C$ is a holomorphic injection. \\[1ex] \noindent {\bf 3.7. Lemma.} {\it The preimage $\rho_C^{-1} (R_n) \subset I \!\! P^2$ is the union of $C$ with its artifacts $L_C$. In particular, $\rho_C^{-1} (R_n) = C$ iff the dual curve $C^*$ is an immersed curve. }\\ \noindent {\it Proof.} Note that a point $z \in I \!\! P^2 \setminus C$ is contained in $\rho_C^{-1} (R_n)$ iff its dual line $l_z$ passes through a cusp of $C^*$. The lines in $L_C$ are just the dual lines to the cusps of $C^*$. \qed \section {C-hyperbolicity of complements of plane curves of genus $g \ge 1$} \noindent In this section we keep all the notation from 3.6. The main result here is the following theorem. \\[1ex] \noindent {\bf 4.1. Theorem.} {\it Let $C \subset I \!\! P^2$ be an irreducible curve of genus $g \ge 1$ and $L_C$ be its artifacts. Then \noindent a) $ I \!\! P^2 \setminus (C\cup L_C)$ is C--hyperbolic. \noindent b) If the dual curve $C^* \subset I \!\! P^{2*}$ is an immersed curve, then $ I \!\! P^2 \setminus C$ is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$.} \\ [1ex] \noindent {\it Proof.} a) Consider first the case when $g\ge 2$. In this case $S^n C^*_{norm}\setminus R_n$ is C--hyperbolic (cf. (1.4)). Indeed, its covering $(C^*_{norm})^n \setminus D_n$ is a domain in $(C^*_{norm})^n$. Since the universal covering of $(C^*_{norm})^n$ is the polydisc $\Delta^n$, it is C--hyperbolic. Therefore, $ (C^*_{norm})^n \setminus D_n$, and hence also $S^n C^*_{norm} \setminus R_n$ are C--hyperbolic, too. By Lemma 3.7 the image $\rho_C ( I \!\! P^2 \setminus (C \cup L_C )) \subset S^n C^*_{norm}$ does not meet the discriminant variety $R_n$ and by 3.6 $\rho_C\,|\, ( I \!\! P^2 \setminus (C \cup L_C )) \,:\, I \!\! P^2 \setminus (C \cup L_C ) \to S^n C^*_{norm} \setminus R_n$ is injective. Therefore, by Lemma 2.5 $ I \!\! P^2 \setminus (C \cup L_C )$ is C--hyperbolic. Next we consider the case when $C$ is an elliptic curve. Denote $E = C^*_{norm}$. Note that both $E^n \setminus D_n$ and $S^nE \setminus R_n$ are not C--hyperbolic or even hyperbolic, and so we can not apply the same arguments as above. Represent $E$ as $E = J(E) = I \!\!\!\! C / \Lambda_{\omega}$, where $\Lambda_{\omega}$ is the lattice generated by $1$ and $\omega \in I \!\!\!\! C_{+}$ (here $ I \!\!\!\! C_{+} := \{z \in I \!\!\!\! C\,|\,{\rm Im}z > 0\}$). By Abel's Theorem we may assume this identification of $E$ with its jacobian $J(E)$ being chosen in such a way that the image $\rho_C ( I \!\! P^2)$ is contained in the hypersurface $s_n (H_0) = \phi_n^{-1} ({\bar 0}) \cong I \!\! P^{n-1} \subset S^n E$, where $$H_0 := \{z=(z_1,\dots, z_n) \in E^n\,|\,\sum\limits_{i=1}^n z_i = 0\}$$ is an abelian subvariety in $E^n$ (see (3.5)). The universal covering $\tilde{H}_0$ of $H_0$ can be identified with the hyperplane $\sum\limits_{i=1}^n x_i = 0$ in $ I \!\!\!\! C^n = \tilde{E}^n$. Consider the countable families $\tilde{D}_{ij}$ of parallel affine hyperplanes in $ I \!\!\!\! C^n$ given by the equations $x_i - x_j \in \Lambda_{\omega}\,,\,\,\,i,j = 1,\dots,n,\,\,i<j$.\\[1ex] \noindent {\it Claim. The domain $\tilde{H}_0 \setminus \bigcup\limits_{i=1}^{n-1} \tilde{D}_{i,i+1}$ is biholomorphic to $( I \!\!\!\! C \setminus \Lambda_{\omega})^{n-1}$.} \\ [1ex] Indeed, put $y_k := (x_k - x_{k+1})\,|\,\tilde{H}_0\,,\,\,i = 1,\dots,n-1$. It is easily seen that $(y_1,\dots,y_{n-1})\,:\, \tilde{H}_0 \to I \!\!\!\! C^{n-1}$ is a linear isomorphism whose restriction yields a biholomorphism as in the claim. The universal covering of $( I \!\!\!\! C \setminus \Lambda_{\omega})^{n-1}$ is the polydisc $\Delta^n$, and so $( I \!\!\!\! C \setminus \Lambda_{\omega})^{n-1}$ is C--hyperbolic. Put $\tilde{D}_n:= \bigcup\limits_{i,j=1,\dots,n} \tilde{D}_{ij}$. The open subset $\tilde{H}_0 \setminus \tilde{D}_n$ of $\tilde{H}_0 \setminus \bigcup\limits_{i=1}^{n-1} \tilde{D}_{i,i+1} \cong ( I \!\!\!\! C \setminus \Lambda_{\omega})^{n-1} $ is also C--hyperbolic. Denote by $p$ the universal covering map $ I \!\!\!\! C^n \to ( I \!\!\!\! C /\Lambda_{\omega})^n$. The restriction $$p\,|\,\tilde{H}_0 \setminus \tilde{D}_n\,:\,\tilde{H}_0 \setminus \tilde{D}_n \to H_0 \setminus D_n \subset E^n \setminus D_n$$ is also a covering mapping. Therefore, $H_0 \setminus D_n$ is C--hyperbolic, and so $s_n (H_0) \setminus R_n$ is C--hyperbolic, too. Since $\rho_C\,|\, ( I \!\! P^2 \setminus (C \cup L_C )) \,:\, I \!\! P^2 \setminus (C \cup L_C ) \to s_n (H_0) \setminus R_n$ is an injective holomorphic mapping, by Lemma 2.5 $ I \!\! P^2 \setminus (C \cup L_C )$ is C--hyperbolic. b) Assume further that $C^*$ is an immersed curve. Then $C$ can not be smooth. Indeed, being smooth $C$ would have flexes at the points of intersection of $C$ with its Hesse curve (see [Wa]), and hence $C^*$ would have cusps. Thus, ${\rm reg}\,C$ is hyperbolic, and by that what has been proven above $ I \!\! P^2 \setminus C$ is C--hyperbolic. Therefore, by Proposition 2.7 it is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. This completes the proof. \qed \noindent {\bf 4.2. Remark.} The complement to a ${\underline {\rm rational}}$ curve whose dual is an immersed curve is not necessarily C--hyperbolic; it even may be not Brody or Kobayashi hyperbolic. An example is a plane quartic $C$ with three cusps. Such a quartic $C$ is projectively equivalent to the curve given by the equation $$4x_1^2 (x_1 -2x_0 )(x_1 + x_2 ) - (2x_0 x_2 - x_1^2 )^2 =0$$ (see [Na, (2.2.5)]). Its dual curve $C^*$ is a nodal cubic with equation $$x_0 x_1^2 + x_1^3 - x_0 x_2^2 =0\,\,\,.$$ Thus, $g(C) = 0$ and $C^*$ is an immersed curve. The complement $ I \!\! P^2 \setminus C$ is not Kobayashi hyperbolic, because its Kobayashi pseudo--distance vanishes on any of three cuspidal tangent lines of $C$, on any of three lines passing through two cusps of $C$ each one and on the only bitangent line of $C$. Indeed, each of these seven lines meets $C$ in only two points; but $k_{ I \!\!\!\! C^*} =0 $, where $ I \!\!\!\! C^* = I \!\! P^1 \setminus \{2 \,{\rm points}\}$. Therefore, $ I \!\! P^2 \setminus C$ is not C--hyperbolic. Note that $\pi_1 ( I \!\! P^2 \setminus C)$ is a finite group of order $12$ [Zar, p.145], and thus $ I \!\! P^2 \setminus C$ is super--Liouville (see (2.2)). Note, furthermore, that by an analogous reason the complement of any quartic curve $C \subset I \!\! P^2$ is neither Kobayashi hyperbolic nor Brody hyperbolic [Gr2]. The fundamental group $\pi_1 ( I \!\! P^2 \setminus C)$ for an irreducible quartic $C$ being almost abelian (see (1.1)), by Lin's Theorem $ I \!\! P^2 \setminus C$ is super--Liouville. \\[1ex] Next we give some examples, or at least computations of numerical characters of plane curves which satisfy the assumptions of Theorem 4.1. First we consider examples of curves of genus $g \ge 2$ with the dual an immersed curve. \\[1ex] \noindent {\bf 4.3. Example.} Let $C \subset I \!\! P^2$ be an irreducible curve whose dual $C^*$ is a nodal curve of degree $n=d^* \ge 3$ with $\delta$ nodes. Assume that the genus $g(C) = g(C^*) = {(n-1)(n-2)\over 2} - \delta$ is at least $2$. Such a curve does exist for any given $\delta$ with $0\le\delta \le {(n-1)(n-2)\over 2} -2$ (see [Se, \S 11, p.347]; [O, Proposition 6.7]). By the Class Formula (4) $C$ has degree $d=n(n-1) -2\delta $, which can be any even integer from the interval $[2(n+1), n(n-1)]$. The least possible value of $n$ resp. $d$ is $n=4$ resp. $d=10$, which corresponds to the case when $C^*$ is a nodal quartic with one node ($\delta =1$) (see e.g. [Na, p.130]). To be a Pl\"ucker curve (see (3.1)) such $C$ should be a curve of degree $10$ with $16$ nodes and $18$ ordinary cusps (cf. [Zar, p.176]). If $C$ is a curve of genus $g\ge 2$ whose dual is a nodal curve, then by Theorem 4.1 the complement $ I \!\! P^2 \setminus C$ is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. This yields examples of such curves $C$ of any even degree $d\ge 10$. \\[1ex] There are similar examples with elliptic curves.\\ \noindent {\bf 4.4. Example.} If the dual $C^*$ of $C$ is an immersed elliptic curve, then by the Class Formula (4) $d= {\rm deg}\,C =2n \ge 6$, where $n = {\rm deg}\,C^* \ge 3$ (see 3.4). Thus, the least possible value of the degree $d$ of such a curve $C$ is $d=6$. Let $C$ be a sextic in $ I \!\! P^2$ with $9$ cusps. Then $C$ is an elliptic Pl\"ucker curve whose dual $C^*$ is a smooth cubic; vice versa, the dual curve to a smooth cubic is a sextic with $9$ ordinary cusps (see e.g. [Wa]). By Theorem 4.1 the complement of such a curve is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. The same is true if $C$ is the dual curve to a nodal quartic $C^*$ with two nodes; here $d= {\rm deg}\,C =8$ (see e.g. [Na, p.133]). To be Pl\"ucker such a curve $C$ must have $8$ nodes and $12$ ordinary cusps. \\[1ex] These examples lead to the following conclusion.\\[1ex] \noindent {\bf 4.5. Proposition.} {\it For any even $d\ge6$ there are irreducible plane curves of degree $d$ and of genus $g\ge 1$ whose duals are nodal curves, and so which satisfy the assumptions of Theorem 4.1, b).} \\[1ex] Next we pass to examples to part a) of Theorem 4.1. \\ \noindent {\bf 4.6. Examples.} Let $C^* \subset I \!\! P^{2*}$ be an irreducible curve of genus $g\ge 1$ with classical singularities. If $C^*$ has $\delta$ nodes and $\kappa$ cusps, then the dual curve $C \subset I \!\! P^2$ has $\kappa$ ordinary flexes as the only flexes, and so $L_C$ is the union of inflexional tangents of $C$. By the Class Formula (4) we have $d = {\rm deg}\,C = 2(n+g-1)- \kappa$. Since all $\kappa$ inflexional tangents of $C$ are distinct, it follows that ${\rm deg}\,(C\cup L_C) = 2(n+g-1) \ge 2n \ge 6$.. Assume that $\kappa>0$ to exclude the case considered in Examples 4.3 and 4.4 above, when $C^*$ was an immersed curve. Since $g\ge 1$, the case when $C$ is a singular cubic has also been excluded. Thus, we have $n\ge 4$, and hence ${\rm deg}\,(C\cup L_C) \ge 8$. The simplest example is a quartic $C^*$ with an ordinary cusp and a node as the only singularities; such a quartic does exist (see [Na, p.133]). The dual curve $C$ is an elliptic septic with the only inflexional tangent line $l=L_C$. To be a Pl\"ucker curve, $C$ must have $4$ nodes and $10$ ordinary cusps. Another example is a quartic $C^*$ with two ordinary cusps as the only singular points; it also does exist [Na, p.133]. Here $C$ is an elliptic sextic and $L_C$ is the union of two inflexional tangents of $C$. To be a Pl\"ucker curve, $C$ should have one node and $8$ ordinary cusps. In all these examples the assumptions of Theorem 4.1 are fulfilled. \\[1ex] \noindent {\bf 4.7. Remark.} Of course, it may happen that the complement of the artifacts $ I \!\! P^2 \setminus L_C$ is itself C--hyperbolic. For instance, this is so if $L_C$ contains an arrangement of five lines with two triple points on one of them, which is projectively equivalent to those $C_5$ of Example 1.3 in the Introduction. But this is not the case if $L_C$ consists only of few lines like in the examples 4.6, or if it consists of lines in general position (cf. (1.2)). For instance, if $C = F_d$ is the Fermat curve of degree $n \ge 3$ in $ I \!\! P^2$, then it is easily checked that the inflexional tangents are in general position (note that here all the flexes are hyperflexes of high order). We suppose that for a generic smooth plane curve $C$ of degree $d \ge 4$ its inflexional tangents are in general position, and therefore the complement $ I \!\! P^2 \setminus L_C$ is super--Liouville. \\[1ex] Next we consider the problem of existence of an irreducible quintic with C--hyperbolic complement. \\ \noindent {\bf 4.8.} We have already noted that there is no plane quartic $C$ with Kobayashi hyperbolic complement $ I \!\! P^2 \setminus C$ [Gr2]. The obstructions are lines in $ I \!\! P^2$ which intersect $C$ in at most two points; e.g. cuspidal tangents, inflexional tangents, bitangents, etc. Recall that if $C \subset I \!\! P^2$ is a nodal quintic, then $ I \!\! P^2 \setminus C$ is super--Liouville (see (1.1) and (2.2)). Although there are irreducible quintics whose complements are Kobayashi hyperbolic [Za3], there is no one with C--hyperbolic complement (cf. (1.3) and (6.1) for examples of reducible quintics with C--hyperbolic complements). It is shown in the next lemma that there is no one among the non-Pl\"ucker quintics; as for the Pl\"ucker ones, see Proposition 4.10 below. \\[1ex] \noindent {\bf 4.9. Lemma.} {\it Let $C \subset I \!\! P^2$ be an irreducible quintic. Suppose that $C$ is not a Pl\"ucker curve. Then $ I \!\! P^2 \setminus C$ is not Brody hyperbolic. Moreover, there exists a line $l_0 \subset I \!\! P^2$ which intersects with $C$ in at most two points.} \\[1ex] \noindent {\it Proof.} Assume that $C$ has a singular point $p_0$ which is not a classical one. Let $l_0$ be the tangent line to an irreducible local analytic branch of $C$ at $p_0$. If ${\rm mult}_{p_0}\,C \ge 3$, then $i(C,\,l_0 ;\,p_0 ) \ge 4$, and so the line $l_0$ has at most one more intersection point with $C$. If ${\rm mult}_{p_0}\,C = 2$, then either \noindent 1) $C$ has two smooth branches at $p_0$ with the same tangent $l_0$ (e.g. $p_0 \in C$ is {\it a tacnode}), \noindent or \noindent 2) $C$ is locally irreducible in $p_0$ and has the multiplicity sequence $(2,\,2,\,\dots)$ at $p_0$ (see (3.2)). \noindent In both cases we still have $i(C,\,l_0 ;\,p_0 ) \ge 4$, and the same conclusion as before holds. It holds also in the case when $l_0$ is the inflexional tangent to $C$ at a point where $C$ has a flex of order at least $2$ (see (3.2)). Therefore, from now on we may suppose that $C$ has only classical singularities and ordinary flexes. Let $q_0$ be a singular point of $C^*$ which is not classical. It can not be locally irreducible, since $C$ has only ordinary flexes. If one of the irreducible local branches of $C^*$ at $q_0$ is singular, then the dual line $l_0$ of $q_0$ is a multiple tangent line to $C$ which is an inflexional tangent at some flex of $C$. Therefore, by the Bezout Theorem $l_0$ is a bitangent line with intersection indices $2$ and $3$. The remaining case to consider is the case when $C^*$ has only smooth local branches at $q_0$. If two of them, say, $A^*_0$ and $A^*_1$, are tangent to each other, then by duality the corresponding local branches $A_0$ and $A_1$ of $C$ should have common center and moreover, they should be tangent to each other, too. But this is impossible since $C$ is assumed to have only classical singularities. Thus, we are left with the case that $q_0$ is a tidy singularity consisting of at least three disinct irreducible local branches of $C^*$. But then the dual line $l_0$ of $q_0$ is tangent in at least three different points of $C$. Since $C$ is of degree five, by Bezout's Theorem this is also impossible. This completes the proof. \qed From this lemma, Proposition 4.5 and the computations done by A. Degtjarjov [Deg 1, 2] we obtain the following statement. \\[1ex] \noindent {\bf 4.10. Proposition.} {\it The minimal degree of an irreducible plane curve with C--hyperbolic complement is 6}. \\ \noindent {\it Proof}. Indeed, it is shown in [Deg 1, 2] that two irreducible plane quintics with the same type of singular points are isotopic in $ I \!\! P^2$, and there are only two types of them such that the fundamental groups of the complement are not abelian.. In both cases these quintics have non-classical singularities. It follows that for an irreducible Pl\"ucker quintic $C \subset I \!\! P^2$ the complement $ I \!\! P^2 \setminus C$ has cyclic fundamental group, and therefore it is super-Liouville. \qed \section {Rational plane curves and duality} Here we precise the construction of (3.6) in the case of a rational curve. \\ \noindent {\bf 5.1.} {\it The Vieta covering}. The symmetric power $S^n I \!\! P^1$ can be identified with $ I \!\! P^n$ in such a way that the canonical projection $s_n \,:\, ( I \!\! P^1 )^n \to S^n I \!\! P^1$ becomes {\it the Vieta ramified covering}, which is given by $$((u_1 : v_1),\dots,(u_n : v_n)) \longmapsto $$ $$\longmapsto (\prod\limits_{i=1}^n v_i)\,(1 : \sigma_1 (u_1 /v_1 ,\dots, u_n /v_n )\,:\,\dots \,:\,\sigma_n (u_1 /v_1 ,\dots, u_n /v_n ))\,\,\,,$$ where $\sigma_i (x_1 ,\dots, x_n)\,,\,\,i=1,\dots,n$, are elementary symmetric polynomials. This is a Galois covering with the Galois group being the n-th symmetric group $S_n$. In the case when $z_i := u_i /v_i \in I \!\!\!\! C ,\,\,i=1,\dots,n$, we have $s_n (z_1 ,\dots, z_n) = (a_0 : \dots : a_n)$, where the equation $a_0 z^n + \dots +a_n =0$ has the roots $z_1,\dots, z_n$ (see [Zar, p.252] or [Na, (5.2.18)]). In general, $z_i \in I \!\! P^1 ,\,i= 1,\dots, n,\,$ are the roots of the binary form $\, \sum\limits_{i=0}^n a_i u^{n - i} v^i$ of degree $n$. \\ \noindent {\bf 5.2.} {\it Plane cuts of the discriminant hypersurface}. If $C \subset I \!\! P^2$ is a rational curve of degree $d > 1$, then $C^*_{norm} \cong I \!\! P^1$, and so the Zariski embedding is a morphism $\rho_C \,:\, I \!\! P^2 \to I \!\! P^n \cong S^n I \!\! P^1$, where $n = {\rm deg}\,C^*$. The normalization map $ \nu \,:\, I \!\! P^1 \to C^* \subset I \!\! P^2$ can be given as $\nu = (g_0 : g_1 : g_2)\,$, where $g_i (z_0 , z_1) = \sum\limits_{j=0}^n b^{(i)}_j z_0^{n-j} z_1^j \,\,,\,\,i=0,1,2, $ are homogeneous polynomials of degree $n$ without common factor. If $x = (x_0 : x_1 : x_2) \in I \!\! P^2$ and $l_x \subset I \!\! P^{2*}$ is the dual line, then $\rho_C (x) = \nu^* (l_x) \in S^n I \!\! P^1 = I \!\! P^n$ is defined by the equation $\sum\limits_{i=0}^2 x_i g_i (z_0 : z_1 ) =0$. Thus, $\rho_C (x) = (a_0 (x) : \dots : a_n (x) )$, where $a_j (x) = \sum\limits_{i=0}^2 x_i b^{(i)}_j$. Therefore, in the case of a plane rational curve $C$ the Zariski embedding $\rho_C \,:\, I \!\! P^2 \to I \!\! P^n$ is the linear embedding given by the $3 \times (n+1)$--matrix $B_C := (b^{(i)}_j )$, $i=0,1,2,\,j=0,\dots,n$. In what follows we denote by $ I \!\! P^2_C$ the image $\rho_C ( I \!\! P^2)$, which is a plane in $ I \!\! P^n$. By Lemma 3.7 the curve $C$ is an irreducible component of the plane cut of the discriminant hypersurface $R_n \subset I \!\! P^n$, which has degree $2n-2$, by the plane $\rho_C ( I \!\! P^2)$; the other irreducible components come from the artifacts $L_C$ of $C$. This yields Lemma 1.9 in the Introduction: $$ I \!\! P_C^2 \cap R_n = C \cup L_C \,\,.$$ The embedding $C^* \hookrightarrow I \!\! P^{2*}$ composed with the normalization $\nu\,:\, I \!\! P^1 \to C^*$ is uniquely, up to projective equivalence, defined by the corresponding linear series $g^2_n$ on $ I \!\! P^1$, and the embedding $\rho_C$ is uniquely defined by $C$ up to a choice of normalization of $C^*$. Thus, $ I \!\! P^2_C$ is uniquely defined by $g^2_n$ up to the action on $ I \!\! P^n$ of the group $ I \!\! P {\rm GL} (2,\, I \!\!\!\! C ) \times I \!\! P {\rm GL} (3,\, I \!\!\!\! C )$ via its natural representation in $ I \!\! P {\rm GL} (n+1,\, I \!\!\!\! C )$, where the second factor leaves $ I \!\! P^2_C$ invariant. \\ \noindent {\bf 5.3.} {\it The rational normal curve.} The dual map ${\rho_C}^* \,:\, I \!\! P^{n*} \to I \!\! P^{2*}$, given by the transposed $(n+1) \times 3$--matrix $\,^t B_C = (b^{(j)}_i )$, $i=0,\dots,n,\, j=0,1,2$, defines a linear projection with the center $N_C := {\rm Ker\,} \,^t B_C \subset I \!\! P^{n*}$ of dimension $n - 3$. The curve $C^*$ is the image under this projection of the rational normal curve $C_n^* = (z_0^n : z_0^{n-1} z_1 : \dots : z_1^n ) \subset I \!\! P^{n*}$ (cf. [Ve, p.208]), i.e. $$\rho_C^* (C_n^*) = C^* \,\,..$$ Furthermore, $C_n^*$ is the image of $ I \!\! P^1 \cong C^*_{norm}$ under the embedding $i \,:\, I \!\! P^1 \hookrightarrow I \!\! P^{n*}$ defined by the complete linear system $|H| = |n(\infty)| \cong I \!\! P^n$. Therefore, $\nu = {\rho_C}^* \circ i \,:\, I \!\! P^1 \to C^* \subset I \!\! P^{2*}$ is the normalization map. \\ \noindent {\bf 5.4.} {\it The duality picture}. It is easily seen that the rational normal curve $C_n^* \subset I \!\! P^{n*}$ and the discriminant hypersurface $R_n \subset I \!\! P^{n}$ are dual to each other. This yields the following duality picture: \begin{center} \begin{picture}(500,70) \put(150,55){$( I \!\! P^2 ,\,C \cup L_C)\,\,\,\,\,\,\, \hookrightarrow \,\,\,\,\,\,\, ( I \!\! P^n ,\,R_n )$} \put(155,5){$( I \!\! P^{2*} ,\,C^* ) \,\,\,\,\,\,\,\,\,\, \longleftarrow \,\,\,\,\,\,\, ( I \!\! P^{n*} ,\,C_n^* )$} \put(175,30){$\updownarrow$} \put(291,30){$\updownarrow$} \put(238,20){$\rho^*_C$} \put(240,68){$\rho_C$} \end{picture} \end{center} \noindent To describe this duality in more details, fix a point $q=(z_0^n:z_0^{n-1}z_1:...:z_1^n) \in C_n^* \subset I \!\! P^{n*}$, and let $$F_q C^*_n = \{ T^0_q C_n^* \subset T^1_q C_n^* \subset \dots \subset T^{n - 1}_q C_n^* \subset I \!\! P^{n*} \}$$ be the flag of the osculating subspaces to $C_n^*$ at $q$, where ${\rm dim}\,T^k_q C_n^* = k,\,\,T^0_q C_n^* = \{q\}$ and $T^1_q C_n^* = T_q C_n^*$ is the tangent line to $C_n^*$ at $q$ (see [Na, p.110]). For instance, for $q = q_0 = (1 : 0 : \dots : 0) \in C_n^*$ we have $T^k_q C_n^* = \{x_{k + 1} = \dots = x_n = 0 \} \subset I \!\! P^{n*}$. The dual curve $C_n \subset I \!\! P^n$ of $C_n^*$ is in turn projectively equivalent to a rational normal curve; namely, $$C_n = \{ p \in I \!\! P^n \,|\, p = (z_1^n : -nz_0 z_1^{n - 1} : \dots : (-1)^k { n \choose k } z_0^k z_1^{n - k} : \dots : (-1)^n z_0^n )\}$$ Furthermore, the dual flag $F_q^{\perp} = \{ I \!\! P^n \supset H^{n - 1}_q \supset \dots \supset H^0_q \}$, where $H^{n - k}_q := (T^{k - 1}_q C_n^*)^{\perp}$, is nothing else but the flag of the osculating subspaces $F_p C_n = \{T^{k - 1}_p C_n \}_{k = 1}^n$ of the dual rational normal curve $C_n \subset I \!\! P^n$. An easy way to see this is to look at the flags at the dual points $q_0 = (1 : 0 : \dots : 0 ) \in C_n^*$ and $p_0 = (0 : \dots : 0 : 1 ) \in C_n$, where all the flags consist of coordinate subspaces, and then to use the ${\rm Aut}\, I \!\! P^1$-homogeneity (cf. 7.2 - 7.4 below). The points of the osculating subspace $H^k_q = T^k_p C_n$ correspond to the binary forms of degree $n$ for which $(z_0 : z_1) \in I \!\! P^1$ is a root of multiplicity at least $n - k$. In particular, $H_q^{n - 2} = (T_q C^*_n )^{\perp}$ consists of the binary forms which have $(z_0 : z_1)$ as a multiple root. Therefore, the discriminant hypersurface $R_n$ is the union of these linear subspaces $H_q^{n - 2} \cong I \!\! P^{n - 2}$ for all $q \in C_n^*$, and thus it is the dual hypersurface of the rational normal curve $C_n^*$, i.e. each of its points corresponds to a hyperplane in $ I \!\! P^{n*}$ which contains a tangent line of $C_n^*$. At the same time, $R_n$ is the developable hypersurface of the $(n-2)$--osculating subspaces $H_q^{n - 2} = T_p^{n - 2} C_n$ of the dual rational normal curve $C_n \subset R_n$; here $T_p^{n - 2} C_n \cap C_n = \{p\}$. If $z_0 \neq 0$, then the subspace $H_q^{n - 2}$ in $ I \!\! P^n$ can be given as the image of the linear embedding $$ I \!\! P^{n-2} \ni (c_0:...:c_{n-2}) \longmapsto (a_0:a_1:a_2:...:a_n) =$$ $$= (\sum_{k=2}^n c_{k-2} (k-1) z_0^{n-k} z_1^k: -\sum_{k=2}^n c_{k-2} k z_0^{n-k+1} z_1^{k-1} : c_0 z_0^n :...: c_{n-2} z_0^n) \in (T_qC_n^*)^* \subset I \!\! P^n \,$$ (and symmetrically for $z_0 = 0$). Consider the decomposition $D_{ij} = d_{ij} \times ( I \!\! P^1 )^{n-2}$ of the diagonal hyperplane $D_{ij} \subset D_n$, where $d_{ij} \equiv I \!\! P^1$ is the diagonal line in $ I \!\! P^1_i \times I \!\! P^1_j$, as the trivial fibre bundle $D_{ij} \to I \!\! P^1$ with the fibre $( I \!\! P^1 )^{n-2}$. The subspaces $H_q^{n - 2} \subset R_n$ are just the images of the fibres under the Vieta map; moreover, the restriction of the Vieta map $s_n\, :\,( I \!\! P^1)^n \to I \!\! P^n$ to a fibre yields the Vieta map $s_{n-2}\, :\,( I \!\! P^1)^{n-2} \to I \!\! P^{n-2}$. The dual rational normal curve $C_n \subset I \!\! P^n$ is the image $s_n (d_n )$ of the diagonal line $d_n := \bigcap_{i,\,j} D_{ij} = \{z_1 = \dots = z_n \} \subset ( I \!\! P^1 )^n$. \\[1ex] \noindent {\bf 5.5.} {\it Artifacts as linear sections and the hyperplanes dual to the cusps}. By duality we have $N_C={\rm Ker}\, {\rho_C}^* = ({\rm Im}\,\rho_C )^{\perp}$, i.e. $N_C = ( I \!\! P^2_C )^{\perp}$. Therefore, $$ I \!\! P^2_C = N_C^{\perp} = \bigcap\limits_{x^* \in N_C } {\rm Ker}\,x^* = \{ x \in I \!\! P^n \,|\, <x, x^* > = 0 \,\,{\rm for\,\, all\,\,}x^* \in N_C \} \,\,\,.$$ A point $q$ on the rational normal curve $C_n^* \subset I \!\! P^{n*}$ corresponds to a cusp of $C^*$ under the projection ${\rho_C}^*$ iff the center $N_C$ of the projection meets the tangent developable $TC_n^*$, which is a smooth ruled surface in $ I \!\! P^{n*}$, in some point $x_{q^*}$ of the tangent line $T_q C_n^*$ (see [Pi]). In this case it meets $T_q C_n^*$ at the only point $x_q^*$, because otherwise $N_C$ would contain $T_q C_n^*$ and thus also the point $q$, which is impossible since ${\deg}\,C^* = {\deg}\,C_n^* = n$. Let $C^*$ have a cusp $B$ at the point $ q_0 = {\rho_C}^* (q)$, which corresponds to a local branch of $C_n^*$ at the point $q \in C_n^*$ under the normalizing projection ${\rho_C}^* \,:\,C_n^* \to C^*$. Define $L_{B, q_0} := {\rm Ker}\,x_q^* \subset I \!\! P^n$ to be the dual hyperplane of the point $x_q^* \in N_C \cap T_q C_n^*$. Since $x_q^* \in N_C$, this hyperplane $L_{B, q_0}$ contains the image $ I \!\! P^2_C = \rho_C ( I \!\! P^2 )$. This yields a correspondence between the cusps of $C^*$ and certain hyperplanes in $ I \!\! P^n$ containing the plane $ I \!\! P^2_C$. From the definition it follows that $L_{B,q_0}$ contains also the dual linear space $H_q^{n - 2} = (T_q C_n^*)^{\perp} \subset R_n$ of dimension $n - 2$. Since the plane $ I \!\! P^2_C$ is not contained in $R_n$, we have $L_{B,q_0} = {\rm span}\, ( I \!\! P^2_C ,\,H_q^{n - 2} )$. It is easily seen that the intersection $ I \!\! P^2_C \cap H_q^{n - 2}$ coincides with the tangent line $l_{q_0} \subset L_C$ of $C$, which is dual to the cusp $q_0$ of $C^*$. Thus, the artifacts $L_C$ of $C$ are the sections of $ I \!\! P^2_C$ by those osculating linear subspaces $H_q^{n - 2} \subset R_n$ for which $q$ is a cusp of $C^*$; any other subspace $H_q^{n - 2}$ meets the plane $ I \!\! P^2_C$ in one point of $C$ only. \\[1ex] \noindent {\bf 5.6. Lemma.} {\it Let $C \subset I \!\! P^2$ be a rational curve whose dual curve $C^* \subset I \!\! P^{2*}$ has degree $n$. Let $B$ be a cusp of $C^*$ with center $q_0 \in C^*$, and let $L_{B, q_0} \subset I \!\! P^n$ be the corresponding hyperplane which contains the plane $ I \!\! P^2_C = \rho_C ( I \!\! P^2 )$ (see (5.5)). Then under a suitable choice of a normalization of the dual curve $C^*$ we have $L_{B, q_0} = {\bar A}_1$, where $${\bar A}_1 := \{(a_0 : \dots : a_n) \in I \!\! P^n \,|\,a_1 = 0 \}\,\,.$$ \noindent The preimage ${\bar H}_0 := s_n^{-1} ({\bar A}_1) \subset ( I \!\! P^1)^n$ is the closure of the linear hyperplane in $ I \!\!\!\! C^n$ $${H}_0 := \{z = (z_1,\dots, z_n ) \in I \!\!\!\! C^n \,|\, \sum\limits_{i=1}^n z_i =0 \} \,\,\,.$$} \noindent {\it Proof.} Up to a choice of coordinates in $ I \!\! P^{2*}$, which does not affect the statement, we may assume that $C^*$ has a cusp $B$ at the point $q_0 = (0:0:1)$. Let $\infty = (1:0) \in I \!\! P^1$, and let $\nu \,:\, I \!\! P^1 \cong C^*_{norm} \to C^* \hookrightarrow I \!\! P^2$ be composition of an isomorphism $ I \!\! P^1 \cong C^*_{norm}$ with the normalization map. This isomorphism may be chosen in such a way that the cusp $B$ corresponds to the local branch of $ I \!\! P^1$ at $\infty$, and so $\nu (\infty) = q_0$. Here as above $\nu = (g_0 : g_1 : g_2)\,$ is given by a triple of homogeneous polynomials $g_i (z_0 , z_1) = \sum\limits_{j=0}^n b^{(i)}_j z_0^{n-j} z_1^j \,\,,\,\,i=0,1,2, $ of degree $n$. Since $\nu (\infty) = q_0$ we have ${\rm deg}_{z_0}\,g_0 < n\,,\,{\rm deg}_{z_0}\,g_1 < n\,,\,{\rm deg}_{z_0}\,g_2 = n$, i.e. $b^{(0)}_0 =b^{(1)}_0 =0\,,\,b^{(2)}_0 \neq 0$. Performing Tschirnhausen transformation $$ I \!\! P^1 \ni (z_0 \,:\,z_1) \longmapsto (z_0 - {b_1^{(2)} \over nb_0^{(2)}} z_1 \,: \, z_1) \in I \!\! P^1$$ we may assume, furthermore, that $b_1^{(2)} =0$.\\ \noindent {\it Claim 1. Under the above choice of parametrization the image $ I \!\! P^2_C = \rho_C ( I \!\! P^2 )$ is contained in the hyperplane ${\bar A}_1 := \{(a_0 :\dots : a_n) \in I \!\! P^n\,|\,a_1 = 0\}$. } \\ Indeed, since $C^*$ has a cusp at $q_0$ we have $(g_0 / g_2)'_{z_1} = (g_1 / g_2)'_{z_1} = 0$ at the point $(1:0) \in I \!\! P^1$, i.e. $(g_0)'_{z_1} = (g_1)'_{z_1} = 0$ when $z_1 = 0$. This means that ${\rm deg}_{z_0}\,g_0 < n-1\,,\,{\rm deg}_{z_0}\,g_1 < n-1$, i.e. $b^{(0)}_1 =b^{(1)}_1 =0$. And also $b_1^{(2)} =0$, as it has been achieved above by making use of Tschirnhausen transformation. Since $b_1^{(i)} =0\,,\,i=0,1,2$, we have $a_1 (x) \equiv 0$. Therefore, $\rho_C (x) \in {\bar A}_1$ for any $x \in I \!\! P^2$, which proves the claim. \qed \noindent {\it Claim 2. The dual space $H_q^{n - 2}$ to $T_qC_n^*$ is contained in ${\bar A}_1$.}\\ Indeed, since $\nu ( \infty) = q_0$ and $\nu = \rho_C^* \circ i$ with $i: I \!\! P^1 \rightarrow C_n^* \subset I \!\! P^{n*}$ we get $q=(1:0:...:0)$. Thus, by (5.4) the subspace $H_q^{n - 2} = (T_qC_n^*)^{\perp}$ is given by the equations $\{a_0=a_1=0 \} $, and hence it is contained in ${\bar A}_1$. \qed By (5.5) we have $L_{B,q_0} = {\rm span}\, ( I \!\! P^2_C ,\,H_q^{n - 2} )$. Therefore, from these two claims we get $L_{B,q_0} = {\bar A}_1$. \\ To conclude the proof of the lemma it is enough to note that if $a_0 \neq 0$ and $a_1 = 0$, then the sum of the roots $z_1 + \dots + z_n$ of the equation $a_0 z^n + a_1 z^{n-1} + \dots + a_n = 0$ is identically zero. Thus, $${\bar H}_0 = s_n^{-1} ({\bar A}_1) = \{(\,(u_1 : v_1),\dots,(u_n : v_n)\,) \in ( I \!\! P^1)^n \,|\, \sum\limits_{i=1}^n u_i / v_i = 0 \}\,,$$ which is the closure of the linear hyperplane $H_0 \subset I \!\!\!\! C^n$ as in the lemma. \qed \noindent {\bf 5.7.} {\it Monomial and quasi--monomial rational plane curves}. A rational curve $C \subset I \!\! P^2$, which can be normalized (up to permutation) as follows: $(x_0 (t) : x_1 (t) : x_2 (t)) = (at^k : bt^m : g(t))$, where $a,\,b \in I \!\!\!\! C^* , \,k,\,m \in Z \!\!\! Z_{\ge 0}$ and $g \in I \!\!\!\! C [t]$, will be called {\it a quasi--monomial curve}. If here $g(t) = ct^l ,\, c\in I \!\!\!\! C^* ,\,l \in Z \!\!\! Z_{\ge 0}$, then $C$ is {\it a monomial curve}; in this case we may assume that ${\rm min} \,\{k,\,l,\,m \} =m =0$ and gcd$(k, \,l) =1\,,\,l > k$. Note that a linear pencil of monomial curves $C_{\mu} = \{\alpha x_0^l + \beta x_1^{l-k} x_2^k =0\}$, where $\mu = \alpha /\beta \in I \!\! P^1$, is self--dual, i.e. the dual curve of a monomial one is again monomial and belongs to the same pencil. In contrast, the dual curve to a quasi--monomial one is not necessarily projectively equivalent to a quasi--monomial curve (recall that two plane curves $C,\,C'$ are {\it projectively equivalent} if $C' = \alpha ( C)$ for some $\alpha \in I \!\! P {\rm GL} (3;\, I \!\!\!\! C ) \cong {\rm Aut}\, I \!\! P^2$). The simplest example is the nodal cubic $C = \{(x_0 : x_1 : x_2 ) = (t : t^3 : t^2 -1 )\}$. Indeed, its dual curve is a quartic with three cusps (cf. Remark 4.2); but a quasi--monomial curve may have at most two cusps. \\ The statement of the next lemma is easy to check, so the proof is omitted. \\[1ex] \noindent {\bf 5.8. Lemma.} {\it A quasi--monomial curve $C = (t^k : t^m : g(t))$, where $k < m$ and $g(t) = \sum\limits_{j=0}^n b_j t^{n-j}$ is a polynomial of degree $n \ge 3$, has no cusp iff it is one of the following curves:} $$(t : t^{n \pm 1} : g(t)),\,\,b_n \neq 0$$ $$(t : t^n : g(t)),\,\,b_1 \neq 0, b_n \neq 0$$ $$(1 : t^{n \pm 1} : g(t)),\,\,b_{n-1} \neq 0$$ $$(1 : t^n : g(t)),\,\,b_1 \neq 0 \,\,{\rm and}\,\,b_{n-1} \neq 0 \,\,.$$ {\it In particular, a monomial curve $C = (t^k : t^l : 1)$, where $k < l $ and gcd$(k,\,l) = 1$, has no cusp iff it is a smooth conic $C = (t : t^2 : 1)$.} \\[1ex] \noindent {\bf 5.9.} {\it Parametrized rational plane curves}. Note that, while the action of the projective group $PGL(3, I \!\!\!\! C)$ on $ I \!\! P^2$ does not affect the image $ I \!\! P^2_C = \rho_C ( I \!\! P^2 ) \subset I \!\! P^n = S^n I \!\! P^1$, the choice of the normalization $ I \!\! P^1 \to C^*$, defined up to the action of the group $PGL(2, I \!\!\!\! C ) = {\rm Aut} I \!\! P^1$, usually does (cf. (5.2)). This is why in the next lemma we have to fix the normalization of a rational plane curve $C$. This automatically fixes a normalization of its dual curve $C^*$, and vice versa. Indeed, recall that if $C = (g_0 : g_1 : g_2 )$, where $g_i \in I \!\!\!\! C [t] ,\,\,i=0,1,2$, is a parametrized rational plane curve, then the dual curve $C^*$ has, up to cancelling of the common factors, the parametrization $C^* = (M_{12} : M_{02} : M_{01} )$, where $M_{ij}$ are the $2 \times 2$--minors of the matrix $$ \left( \begin{array}{ccc} g_0 & g_1 & g_2 \\ g'_0 & g'_1 & g'_2 \end{array} \right)$$ \noindent Furthermore, the equation of $C$ can be written as $\frac{1}{x_2^d } {\rm Res} \,(x_0 g_2 - x_2 g_0 , x_1 g_2 - x_2 g_1 ) = 0$, where $d = {\rm deg}\,C$ and ${\rm Res}$ means resultant (see e.g. [Au, 3.2]). We will use the following terminology. By {\it a parametrized rational plane curve} we will mean a rational curve $C$ in $ I \!\! P^2$ with a fixed normalization $ I \!\! P^1 \to C$ of it. {\it A parametrized monomial} resp. {\it a parametrized quasi--monomial plane curve} is a parametrized rational plane curve such that all resp. two of its coordinate functions are monomials. Clearly, projective equivalence between parametrized curves is a stronger relation than just projective equivalence between underlying projective curves themselves. \\[1ex] \noindent {\bf 5.10. Lemma.} {\it A parametrized rational plane curve $C^* \subset I \!\! P^{2 *}$ of degree $n$ is projectively equivalent to a parametrized monomial resp. quasi--monomial curve iff $ I \!\! P^2_C \subset I \!\! P^n$ is a coordinate plane resp. containes a coordinate axis. This axis is unique iff $C^*$ is projectively equivalent to a parametrized quasi--monomial curve, but not to a monomial one.} \\[1ex] \noindent {\it Proof.} Let $\nu \,:\,t \longmapsto (at^k : bt^m : g(t))$, where $a,\,b \in I \!\!\!\! C^* ,\,\,g \in I \!\!\!\! C [t]$ and $t=z_0 / z_1 \in I \!\! P^1$, define a parametrized quasi--monomial curve $C^* \subset I \!\! P^{2 *}$ of degree $n$. Denote $e_k = (0 : \dots : 0 : 1_k : 0 : \dots : 0) \in I \!\! P^n$. Then $\rho_C$ is given by the matrix $B_C = (b^{(0)},\,b^{(1)},\,b^{(2)}) = (ae_{n-k} ,\, be_{n-m} ,\,b^{(2)})$, and therefore $ I \!\! P^2_C = \rho_C ( I \!\! P^2 ) = {\rm span}\,(b^{(0)},\,b^{(1)},\,b^{(2)})$ contains the coordinate axis $l_{n-k,\,n-m}$, where $l_{i,j} := {\rm span}\,(e_i ,\,e_j ) \subset I \!\! P^n$. If $C^*$ is a paramatrized monomial curve, i.e. if $g(t) = ct^r$, where $c \in I \!\!\!\! C^*$, then clearly $ I \!\! P^2_C$ is the coordinate plane $ I \!\! P_{n-k,\,n-m,\,n-r} := {\rm span}\,(e_{n-k},\,e_{n-m},\,e_{n-r})$. Since the projective equivalence of parametrized plane curves does not affect the $ I \!\! P^2_C$, this yields the first statement of the lemma in one direction. Vice versa, suppose that $ I \!\! P^2_C$ coincides with the coordinate plane $ I \!\! P_{n-k,\,n-m,\,n-r}$. Performing a suitable linear coordinate change in $ I \!\! P^{2 *}$ we may assume that $b^{(0)} = e_{n-k} , \,b^{(1)} = e_{n-m} ,\,b^{(2)} = e_{n-r}$, i.e. that $\nu (t) = (t^k : t^m : t^r)$. Therefore, in this case the parametrized curve $C^*$ is projectively equivalent to a monomial curve. Suppose now that $ I \!\! P^2_C$ contains the coordinate axis $l_{n-k ,\,n-m}$. Performing as above a suitable linear coordinate change in $ I \!\! P^{2 *}$ we may assume that $b^{(0)} = e_{n-k} , \,b^{(1)} = e_{n-m}$, and so $\nu (t) = (t^k : t^m : g(t))$. In this case $C^*$ is projectively equivalent to a parametrized quasi--monomial curve. This proves the first assertion of the lemma. Let $C^* = (at^{n-k} : bt^{n-m} : g(t))$ be a parametrized quasi--monomial curve which is not projectively equivalent to a monomial one. Then as above $ I \!\! P^2_C \supset l_{k,\,m}$, and this is the only coordinate axis contained in $ I \!\! P^2_C$. Indeed, if $l_{i,\,j} \subset I \!\! P^2_C$, where $\{i,\,j\} \neq \{k,\,m\}$, then $ I \!\! P^2_C$ would contain at least three distinct vertices $e_{\alpha}$, where $\alpha \in \{i,\,j,\,k,\,m\}$, and so $ I \!\! P^2_C$ would be a coordinate plane, what has been excluded by our assumption. The opposite statement is evidently true. This concludes the proof. \qed \noindent {\bf 5.11. Remarks.} {\it a}. Let $C^* = (at^k : bt^m : ct^r)$, where $a,\, b,\, c \in I \!\!\!\! C^*$, be a parametrized monomial curve of degree $n$. To be a normalization, this parametrization should be irreducible, i.e. up to permutation there should be $0 = k < m < r = n$, where ${\rm gcd}\,(m,\,n) = 1$. Thus, $ I \!\! P^2_C = I \!\! P_{0,\,n-m,\,n}$ is a rather special coordinate plane. \\[1ex] {\it b}. Let $C^*$ be the parametrized quasi--monomial curve $C_{k, m, g} := (at^k : bt^m : g(t))$, which is not equivalent to a monomial one. Then the only coordinate axis contained in $ I \!\! P^2_C$ is the axis $l_{n-k,\,n-m} := {\rm span}\,(e_{n-k},\,e_{n-m}) = \rho_C (l_2 )$, where $l_2 := \{x_2 = 0 \} \subset I \!\! P^2$. Furthermore, if $C^*$ is obtained from such a curve by a permutation of the coordinates, then still the only coordinate axis contained in $ I \!\! P^2_C$ is $l_{n-k,\,n-m}$. \\[1ex] \noindent {\bf 5.12.} {\it An equivariant meaning of the Vieta map}. Consider the following $\, I \!\!\!\! C^*$--actions on $( I \!\! P^1)^n$ resp. on $ I \!\! P^n = S^n I \!\! P^1$: $${\tilde{G}} \,:\, I \!\!\!\! C^* \times ( I \!\! P^1)^n \ni (\lambda,\,((u_1 : v_1),\dots, (u_n : v_n))) \longmapsto ((\lambda u_1 : v_1), \dots, (\lambda u_n : v_n)) \in ( I \!\! P^1)^n$$ \noindent resp. $$G \,:\, I \!\!\!\! C^* \times I \!\! P^n \ni (\lambda,\, (a_0 : a_1 : \dots : a_n )) \longmapsto (a_0 : \lambda a_1 : \lambda^2 a_2 : \dots : \lambda^n a_n ) \in I \!\! P^n$$ \noindent Note that the Vieta map $s_n : ( I \!\! P^1)^n \to I \!\! P^n$ (see (5.1)) is equivariant with respect to these $\, I \!\!\!\! C^*$--actions and its branching divisors $D_n$ resp. $R_n$ are invariant under ${\tilde{G}}$ resp. $G$. Identifying $ I \!\!\!\! C$ with $ I \!\! P^1 \setminus \{(1:0)\}$, we fix an embedding $ I \!\!\!\! C^n \hookrightarrow ( I \!\! P^1)^n$; denote its image by $ I \!\!\!\! C_z^n$. Both this Zariski open part of $( I \!\! P^1)^n$ and its complementary divisor are ${\tilde{G}}$--invariant. In turn, the hyperplane $ I \!\! P_0^{n-1} := \{a_0 = 0 \}$ in $ I \!\! P^n$, as well as any other coordinate linear subspace of $ I \!\! P^n$, and its complement $ I \!\!\!\! C_a^n := I \!\! P^n \setminus I \!\! P_0^{n-1}$ are $G$--invariant. \\[1ex] The next lemma is a usefull supplement to Lemma 5.10. \\ \noindent {\bf 5.13. Lemma.} {\it A parametrized rational plane curve $C^* \subset I \!\! P^{2*}$ is projectively equivalent to a parametrized quasi--monomial curve iff $ I \!\! P^2_C \subset I \!\! P^n$ contains a one--dimensional $G$--orbit. This orbit is unique iff $C^*$ is projectively equivalent to a parametrized quasi--monomial curve, but not to a monomial one.} \\[1ex] \noindent {\it Proof.} Let $\lambda \longmapsto (a_0 : \lambda a_1 :\dots :\lambda^n a_n )$, where $\lambda \in I \!\!\!\! C^*$, be a parametrization of the $G$--orbit $O_p$ throuh the point $p = (a_0 : \dots : a_n ) \in I \!\! P^n$. Since the non-zero coordinates here are linearly independent as functions of $\lambda$, the orbit $O_p \subset I \!\! P^n$ is contained in a projective plane iff all but at most three of coordinates of $p$ vanish. If $p$ has exactly three non--zero coordinates, then the only plane that containes ${\bar O}_p$ is a coordinate plane. If only two of the coordinates of $p$ are not zero, then the closure ${\bar O}_p$ is a coordinate axis. Since we consider a one--dimensional orbit, the case of one non--zero coordinate is excluded. Now the lemma easily follows from Lemma 5.10. \qed \section{C--hyperbolicity of complements of rational curves in presence of artifacts} Before proving an analog of Theorem 4.1 for the case of a rational curve (see Theorem 6.5 below), let us consider simple examples which illustrate some ideas used in the proof. In (1.3) we gave an example of a quintic $C_5 \subset I \!\! P^2$ (union of five lines) whose complement is C--hyperbolic. Here is another one.\\[1ex] \noindent {\bf 6.1. Example.} Let $C \subset I \!\! P^2$ be a smooth conic and $L = l_1 \cup l_2 \cup l_3$ be the union of three distinct tangents of $C$. \begin{center} \begin{picture}(500, 90) \thicklines \put(212,85){\line(1,-2){40}} \put(222,85){\line(-1,-2){40}} \put(174,16){\line(1,0){92}} \put(217,34){\circle{37}} \end{picture} \\ Figure 2 \end{center} \noindent {\it Claim. a) $X_1 := I \!\! P^2 \setminus (C \cup l_1)$ is super--Liouville and its Kobayashi pseudo--distance $k_{X_1}$ is identically zero. \\ \noindent b) Put $X_2 = I \!\! P^2 \setminus (C \cup l_1 \cup l_2 )$. Let $(C_{\alpha}), \,\alpha \in I \!\! P^1 $, be the linear pencil of conics generated by $C$ and $l_1 + l_2$, where $C = C_{(1 : 1)}$ and $l_1 + l_2$ = $C_{(1 : 0)}$. Then the image of any entire curve $ I \!\!\!\! C \to X_2$ is contained in one of the conics $C_{\alpha}$, and $k_{X_2} (p,\,q) = 0$ iff $p,\,q \in C_{\alpha}$ for some $\alpha \in I \!\! P^1$. Furthermore, $X_2$ is neither C--hyperbolic, nor super--Liouville (see (2.2)). \\ \noindent c) $X = X_3 := I \!\! P^2 \setminus (C \cup L)$ is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$.} \\ \noindent {\it Proof.} a) is easily checked by applying, for instance, Lemma 2.3. An alternative way is to note that $X_1$ is isomorphic to the product $ I \!\!\!\! C \times I \!\!\!\! C^*$. \qed \noindent b) Consider the affine chart $ I \!\!\!\! C^2 \cong I \!\! P^2 \setminus l_2$ in $ I \!\! P^2$. We have $X_2 \cong I \!\!\!\! C^2 \setminus \Gamma$, where the affine curve $\Gamma := (C \cup l_1) \setminus l_2$ can be given in appropriate coordinates by the equation $y(x^2 - y) = 0$. Let the double covering $\pi \,:\, I \!\!\!\! C^2 \to I \!\!\!\! C^2$ branched over the axis $l_1$ be given as $(x, y) = \pi (x, z) := (x, z^2)$. It yields the non--ramified double covering $Y \to X_2$, where $Y := I \!\!\!\! C^2 \setminus \pi^{-1} (\Gamma )$. Here $ \pi^{-1} (\Gamma )$ is union of three affine lines $m_0 = \{z = 0 \},\,\,m_1 = \{x = z \},\,\,m_{-1} = \{x = -z \}$, which are level sets of the rational function $\phi (x,\,z) := z / x$. It defines a holomorphic mapping $\phi \,|\,Y \,:\, Y \to I \!\! P^1 \setminus \{0,\,1,\,-1 \}$. Therefore, for any entire curve $f\,:\, I \!\!\!\! C \to X_2$ its covering curve ${\tilde f} \,:\, I \!\!\!\! C \to Y$ has the image contained in an affine line $l_{\beta_0}$ from the linear pencil $l_{\beta} := \{x = \beta z\} ,\,\,\beta \in I \!\!\!\! C$. Thus, the image $f ( I \!\!\!\! C )$ is contained the conic $C_{\alpha_0}$ from the linear pencil $C_{\alpha} = \{x^2 = \alpha y \}$, where $\alpha = \beta^2$. This proves the first assertion in b). The second one easily follows from the inequality $k_Y \ge \phi^* k_{ I \!\! P^1 \setminus \{3 \,{\rm points}\}}$ and the equality $k_{X_2} = \pi_* k_Y$. Finally, since the tautological line bundle $\phi \,:\, I \!\!\!\! C^2 \setminus \{{\bar 0}\} \to I \!\! P^1$ is trivial over $ I \!\! P^1 \setminus \{{\rm a\,\,point}\}$, there is an isomorphism $Y \cong I \!\!\!\! C^* \times ( I \!\! P^1 \setminus \{3 \,{\rm points}\})$.. Therefore, the universal covering $U_Y \cong U_{X_2}$ of $Y$ resp. of $X_2$ is biholomorphic to $ I \!\!\!\! C \times \Delta$. Hence, $X_2$ is neither C--hyperbolic, nor super--Liouville. \qed \noindent c) We can treat the dual curve of $C \cup L$ as the dual conic $C^* \subset I \!\! P^{2*}$ with three distinguished points $q_1 ,\, q_2 , \,q_3$ on it, whose dual lines are, respectively, $l_1 , \,l_2 ,\, l_3$. Choose an isomorphism $C^* \cong I \!\! P^1$ in such a way that $q_1 , \,q_2 ,\, q_3 \in C^*$ correspond, respectively, to the points $(0 : 1),\, (1 : 0),\, (1 : 1) \in I \!\! P^1$. The Vieta map $s_2 \,:\,( I \!\! P^1 )^2 \to I \!\! P^2 = S^2 I \!\! P^1$ is given by the formula $$s_2 \,:\,((u_1 : v_1),\, (u_2 : v_2)) \longmapsto (v_1 v_2 : -(u_1 v_2 + u_2 v_1 ) : u_1 u_2 )\,\,.$$ To the distinguished points $(0 : 1),\, (1 : 0),\, (1 : 1) \in I \!\! P^1$ there correspond six generators of the quadric $ I \!\! P^1 \times I \!\! P^1$, three vertical ones and three horizontal ones. Their images under the Vieta map $s_2$ is the union $L_0$ of three lines $x_0 =0 ,\, x_2 =0, \,x_0 + x_1 + x_2 =0\,$ in $ I \!\! P^2 = S^2 I \!\! P^1$, which are tangent to the conic $C_0 := s_2 ({\bar D}_2 ) = \{x_1^2 - 4x_0 x_2 =0 \} \subset I \!\! P^2$, where ${\bar D}_2 ={\bar D}_{1,2} $ is the diagonal of $( I \!\! P^1 )^2$. Thus, we have the commutative diagram: \\ \begin{picture}(270,150) \unitlength0.2em \thicklines \put(45,65){$Y\hookrightarrow {\bar Y}$} \put(66,67){{\vector(1,0){40}}} \put(112,65){$( I \!\! P^1)^2$} \put(40,49){${\tilde s}_2$} \put(63,57){${\tilde\rho}_C$} \put(50,62){{\vector(1,-1){10}}} \put(47,60){\vector(0,-1){23}} \put(120,60){\vector(0,-1){23}} \put(60,44){$( I \!\!\!\! C^{**} )^2 \setminus D_2 \hookrightarrow ( I \!\!\!\! C^{**} )^2$} \put(105,52){{\vector(1,1){10}}} \put(76,41){\vector(0,-1){23}} \put(8,29){${ I \!\! P^2} \setminus (C\cup L) = X \hookrightarrow I \!\! P^2$} \put(102,29){$ I \!\! P^2 = S^2 I \!\! P^1 \hbox{\hskip 1cm} $} \put(67,30){\vector(1,0){30}} \put(45,20){$\rho_C$} \put(48,27){{\vector(1,-1){10}}} \put(60,11){${ I \!\! P^2} \setminus (C_0 \cup L_0)$} \put(92,18){{\vector(1,1){10}}} \put(160,35){(5)} \put(122,49){$s_2$} \end{picture} \noindent where ${\tilde{s_2}} \,:\,Y \to X$ is the induced covering, $\, I \!\!\!\! C^{**} = I \!\! P^1 \setminus \{3 \,{\rm points}\}$ and the horizontal arrows are isomorphisms. It follows that $Y \cong ( I \!\!\!\! C^{**} )^2 \setminus D_2 \subset ( I \!\!\!\! C^{**} )^2$ is C--hyperbolic, and therefore, $X$ is C--hyperbolic, too. It is easily seen that ${\rm reg}\,(C\cup L) = (C\cup L) \setminus {\rm sing}\,(C\cup L)$ is hyperbolic. Therefore, by Proposition 2.7 $X = I \!\! P^2 \setminus (C\cup L)$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. \qed Here is one more example of a curve with properties as in Claim c) above. \\[1ex] \noindent {\bf 6.2. Example.} Let the things be as in the previous example. Performing the Cremona transformation $\sigma$ of $ I \!\! P^2$ with center at the points of intersections of the lines $l_1 , l_2 , l_3$, we obtain a 3-cuspidal quartic $C' := \sigma (C)$ together with three new lines $m_1 , m_2 , m_3$, passing each one through a pair of cusps of $C'$ (they are images of the exceptional curves of the blow-ups by $\sigma$ at the above three points; see Fig. 3). \begin{center} \begin{picture}(500, 90) \thicklines \put(212,85){\line(1,-2){40}} \put(222,85){\line(-1,-2){40}} \put(174,15){\line(1,0){92}} \end{picture} \\ Figure 3 \end{center} $$ $$ \noindent Put $L' = m_1 \cup m_2 \cup m_3$ and $X' = I \!\! P^2 \setminus (C' \cup L' )$. Since $\sigma\,|\,X\,:\,X \to X'$ is an isomorphism and $X$ is C--hyperbolic, we have that $X'$ is also C--hyperbolic. The same reasoning as above ensures that $X'$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. \\[1ex] The next lemma will be used in the proof of Theorem 6.5. From now on `bar' over a letter will denote a projective object, in contrast with the affine ones. \\[1ex] \noindent {\bf 6.3. Lemma.} {\it Let ${\bar H}_0$ be the hyperplane in $ I \!\! P^{n-1}$ given by the equation $\sum\limits_{i=1}^n x_i = 0$, and let ${\bar D}_{n-1} = \bigcup\limits_{1\le i<j\le n} {\bar D}_{ij}$ be the union of the diagonal hyperplanes, where ${\bar D}_{ij} \subset I \!\! P^{n-1}$ is given by the equation $x_i - x_j =0$. Then ${\bar H}_0 \setminus {\bar D}_{n-1}$ is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into ${\bar H}_0 \cong I \!\! P^{n-2}$.}\\ [1ex] \noindent {\it Proof.} Put $y_i = x_1 - x_{i+1}\,\,,\,\,i=1,\dots,n-1$. Then $z_i = y_i / y_{n-1}\,\,,\,\,i=1,\dots,n-2 $, are coordinates in the affine chart ${\bar H}_0 \setminus {\bar D}_{1, n} \cong I \!\!\!\! C^{n-2}$. In these coordinates ${\bar D}_{1, i+1} \cap {\bar H}_0$ resp. ${\bar D}_{i+1, n} \cap {\bar H}_0$ is given by the equation $z_i = 0$ resp. $z_i =1\,\,,\,\,i=1,\dots,n-2 $. Thus, ${\bar H}_0 \setminus {\bar D}_{n-1} \hookrightarrow ( I \!\!\!\! C^{**} )^{n-2}$, where $ I \!\!\!\! C^{**} := I \!\! P^1 \setminus \{3\,\,points\}$. By Lemma 2.5 it follows that ${\bar H}_0 \setminus {\bar D}_{n-1}$ is C--hyperbolic. To prove Kobayashi complete hyperbolicity and hyperbolic embededdness we may use the following criterion [Za1, Theorem 3.4] :\\ \noindent {\it The complement of a finite set of hyperplanes $L_1,\dots,L_N$ in $ I \!\! P^n$ is hyperbolically embedded into $ I \!\! P^n$ iff (*) for any two distinct points $p, q$ in $ I \!\! P^n$ there is a hyperplane $L_i\,\,,\,\,i \in \{1,\dots,N\},$ which does not contain any of them. }\\ Note that the complement of a hypersurface is locally complete hyperbolic [KiKo, Proposition 1], and therefore its hyperbolic embededdness implies the complete hyperbolicity (see [Ki] or [KiKo, Theorem 4]). Therefore, it is enough to check that the union of hyperplanes ${\bar H}_0 \cap {\bar D}_{n-1}$ in ${\bar H}_0 \cong I \!\! P^{n-1}$ satisfies the above condition (*). Supposing the contrary we would have that there exists a pair of points $p, q \in {\bar H}_0 \,\,,\,\,p \neq q\,,$ such that each of the diagonal hyperplanes ${\bar D}_{ij}$ contains at least one of these points. Put $p= (x'_1 :\dots :x'_n )$ and $q= (x''_1 :\dots :x''_n )$. Since $(\bigcap\limits_{i, j} {\bar D}_{ij} ) \cap {\bar H}_0 = \emptyset$, we may assume that up to permutation $x'_1 =\dots =x'_k$ and $x''_{k+1} =\dots =x''_n$, where $2\le k \le n-1$, and moreover, that $x'_l \neq x'_i$ for each $i\le k < l \le n$. The latter means that $p \notin {\bar D}_{il}$ for such $i\,,\,l$. Therefore, we must have $q \in {\bar D}_{il}$ for $i\le k < l$. In particular, $q\in {\bar D}_{i, k+1} \,\,,\,\,i=1,\dots,k$, and so $x''_1=\dots =x''_k = x''_{k+1} =\dots =x''_n$, which is impossible, since $q \in {\bar H}_0$. \qed \noindent {\bf 6.4. Remark.} If $n = 4$, so that ${\bar H}_0 \cong I \!\! P^2$, it is easily seen that ${\bar D}_3 \cap {\bar H}_0$ is a complete quadruple in $ I \!\! P^2$, i.e. the union of six lines defined by four points in general position. \\[1ex] Now we are ready to extend Theorem 4.1, under certain additional restrictions, to the case of a rational curve. \\[1ex] \noindent {\bf 6.5. Theorem.} {\it Let $C \subset I \!\! P^2$ be a rational curve whose dual curve $C^*$ has at least one cusp, so that $C$ has the artifacts $L_C \neq \emptyset$. Let $X := I \!\! P^2 \setminus (C \cup L_C)$. Then the following statements hold. \\ a) If the dual curve $C^*$ is not projectively equivalent to a quasi--monomial one, then $X$ is almost C--hyperbolic.\\ b) $X$ is still almost C--hyperbolic if $C^*$ is projectively equivalent to a quasi--monomial curve $C_{k,\,m,\,g} :=\{(t^k : t^m : g(t))\}$ of degree $n$, but not to a monomial one, except the cases when, up to a choice of normalization, $C_{k,\,m,\,g}$ is one of the curves $\{(1 : t^n : g(t))\}$ or $\{(t : t^n : g(t))\}$, where $g \in I \!\!\!\! C [t]$ and ${\rm deg}\,g \le n-2$. In the latter cases $X$ is almost C--hyperbolic modulo the line $l_2 := \{x_2 = 0 \} \subset I \!\! P^2$ in the coordinates where $\,C^* = C_{k,\,m,\,g}$.\\ c) Let $C=C_{\mu_0}$ be a monomial curve \footnote {The case when $C^*$ is projectively equivalent to a monomial curve is easily deduced to this one.} from the linear pencil $C_{\mu} = \{\alpha x_0^n + \beta x_1^k x_2^{n-k} =0\}$, where $\mu = (\alpha : \beta ) \in I \!\! P^1$. Then $k_X (p,\,q) = 0$ iff $p,\,q \in C_{\mu}$ for some $\mu \in I \!\! P^1 \setminus \{\mu_0 \}$. In particular, any entire curve $ I \!\!\!\! C \to X$ is contained in one of the curves of the linear pencil $(C_{\mu})$.} \\[1ex] \noindent {\it Proof.} The proof will be done in several steps. We will start with the main construction used in the proof. \\ \noindent {\it Basic construction.} Fix a cusp $q_0$ of $C^*$, and let $q_0^* \subset I \!\! P^2$ be the dual line of $q_0$. Clearly, $q_0^* \subset L_C$. Choosing an appropriate isomorphism $ I \!\! P^1 \cong C^*_{norm}$ and coordinates in $ I \!\! P^2$ as in the proof of Lemma 5.6, by this lemma we may assume that $\nu (\infty ) = q_0 = (0 : 0 : 1) \in I \!\! P^{2*}$, $q_0^* = l_2 = \{x_2 = 0 \} \subset I \!\! P^2$ and $ I \!\! P^2_C = \rho_C ( I \!\! P^2 ) \subset {\bar A}_1 \subset I \!\! P^n = S^n I \!\! P^1$, where $n = {\rm deg}\,C^*$ and ${\bar A}_1 = \{(a_0 :\dots : a_n) \in I \!\! P^n\,|\,a_1 = 0\} $. Let $ I \!\!\!\! C^n_z \subset ( I \!\! P_1)^n$ and $ I \!\!\!\! C^n_{(a)} \subset I \!\! P^n$ be as in (5.12). Then, as it is easily seen, $\rho_C (X) \subset \rho_C ( I \!\! P^2 \setminus l_2 ) \subset s_n ( I \!\!\!\! C^n_{(z)} ) \cong I \!\!\!\! C^n_{(a)} \subset I \!\! P^n$, where $s_n \,:\, I \!\!\!\! C^n_{(z)} \to I \!\!\!\! C^n_{(a)}$ is the restriction of the Vieta map (see (5.1)). By (5.12) this affine Vieta map yields the non--ramified covering $s_n\,:\,{H}_0 \setminus D_n \to A_1 \setminus R_n$, where as in Lemma 5.6 above ${H}_0 = \{z = (z_1,\dots, z_n ) \in I \!\!\!\! C^n \,|\, \sum\limits_{i=1}^n z_i =0 \}$, $D_n$ is the union of the affine diagonal hyperplanes $D_{ij} = \{z \in I \!\!\!\! C^n \,|\, z_i = z_j \}\,\,,\,1\le i < j \le n$, $A_1 := \{a = (a_1,\dots, a_n ) \in I \!\!\!\! C^n_{(a)} \,|\,a_1 = 0 \} \cong I \!\!\!\! C^{n-1} $ and $R_n \subset I \!\!\!\! C^n_{(a)}$ is the affine discriminant hypersurface. The Zariski map gives the linear embedding $\rho_C \,|\,X \,:\,X \to A_1 \setminus R_n$. Let $ {\tilde{s}}_n \,:\,Y \to X$ be the non--ramified covering induced by the Vieta covering via this embedding. Denote by $\pi$ the canonical projection $ I \!\!\!\! C^n_{(z)} \setminus \{{\bar 0}\} \to I \!\! P^{n-1}$. Put ${\bar H}_0 := \pi ({H}_0 ) \cong I \!\! P^{n-2} \subset I \!\! P^{n-1}$ and ${\bar D}_{ij} := \pi (D_{ij} )\,,\,{\bar D}_{n-1} := \pi (D_n ) = \bigcup\limits_{1\le i < j \le n} {\bar D}_{ij}$. By Lemma 6.3 ${\bar H}_0 \setminus {\bar D}_{n-1}$ is C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into ${\bar H}_0 \cong I \!\! P^{n-2}$. Thus, we have the following commutative diagram: \\ \begin{picture}(250,80) \unitlength0.2em \thicklines \put(-4,5) {${ I \!\! P^2} \setminus (C\cup L_C) = X$} \put(35,25){$Y$} \put(37,22){\vector(0,-1){11}} \put(27,16){${\tilde s}_n$} \put(44,27){{\vector(1,0){14}}} \put(47,30){${\tilde\rho}_C$} \put(44,6){\vector(1,0){12}} \put(47,10){$\rho_C$} \put(62,5){$A_1 \setminus R_n$} \put(62,25){${H}_0 \setminus D_n$} \put(70,22){\vector(0,-1){11}} \put(75,16){$s_n$} \put(170,15){(6)} \put(82,27){{\vector(1,0){14}}} \put(87,30){$\pi$} \put(100,25){${\bar H}_0 \setminus {\bar D}_{n-1} \hookrightarrow I \!\! P^{n-2}$} \end{picture} \noindent where ${\tilde\rho}_C$ is an injective holomorphic mapping. Note that here the Vieta map $s_n$ is equivariant with respect to the $\, I \!\!\!\! C^*$--actions ${\tilde{G}}$ on $ {H}_0 \setminus D_n$ and $G$ on $A_1 \setminus R_n$, respectively, and all the fibres of the projection $\pi$ are one--dimensional ${\tilde{G}}$--orbits (see(5.12)). \\ \noindent {\it Proof of a)}. Under the assumption of a) $C^*$ is not projectively equivalent to a quasi--monomial curve. Then we have the following assertion.\\ \noindent {\it Claim. The mapping $\pi \circ {\tilde \rho}_C \,:\,Y \to {\bar H}_0 \setminus {\bar D}_{n-1}$ has finite fibres.}\\ \noindent Indeed, since the fibres of $\pi$ are $\tilde G$--orbits, it is enough to show that any $\tilde G$--orbit in $H_0 \subset I \!\!\!\! C^n_{(z)}$ has a finite intersection with ${\tilde \rho}_C (Y)$. Or, what is equivalent, that any $G$--orbit in $A_1 \subset I \!\!\!\! C^n_{(a)}$ has a finite intersection with $\rho_C (X) \subset I \!\! P^2_C$. We have shown in Lemma 5.13 above that if the latter fails, i.e. if $ I \!\! P^2_C$ contains a one-dimensional $G$--orbit, then $C^*$ (paramatrized as above) is projectively equivalent to a (paramatrized) quasi--monomial curve, which is assumed not to be the case. This yields the claim. Since ${\bar H}_0 \setminus {\bar D}_{n-1}$ is C--hyperbolic, by Lemma 2.5 this implies that $X$ is almost C--hyperbolic. \qed \noindent {\it Proof of b)}. We still fix a parametrization of $C^*$ as in the basic construction above, and so we fix the $ I \!\! P^2_C$ in $ I \!\! P^n$. If $ I \!\! P^2_C$ does not contain any coordinate line, we can finish up the proof like in a) and conclude that $X$ is almost C--hyperbolic. So, assume further that $ I \!\! P^2_C$ does contain a coordinate line. By Lemma 5.10 this means that $C^*$ as a parametrized curve is projectively equivalent to a quasi--monomial curve. Since by our assumption it is not equivalent to a monomial one, the plane $ I \!\! P^2_C$ is not a coordinate one. After an appropriate change of coordinates in $ I \!\! P^2$ which does not affect $ I \!\! P^2_C$ we may assume that $C^* = C_{f,\,g,\,h} := (f : g : h)$, where $f,\,g,\,h \in I \!\!\!\! C[t]$ and two of them are the monomials $t^k ,\,t^m$. We have that $l_{n-k,\,n-m} \subset I \!\! P^2_C$ is the only coordinate axis contained in $ I \!\! P^2_C$ (see Lemma 5.10 and Remark 5.11, b)). By Lemma 5.13 it is the closure of the only one-dimensional $G$--orbit $O_p$ contained in $ I \!\! P^2_C$. Now we have to distinguish between two cases: \\ \noindent i) $\rho_C^{-1} (l_{n-k,\,n-m}) \subset L_C$ {\hskip 0.5in} and {\hskip 0.5in} ii) $\rho_C^{-1} (l_{n-k,\,n-m}) \not\subset L_C$. \\ \noindent In case i) we have, as in the Claim above, that $\pi \circ {\tilde \rho}_C \,:\,Y \to {\bar H}_0 \setminus {\bar D}_{n-1}$ has finite fibres, and therefore $X$ is almost C--hyperbolic. In case ii) we have $O_p \subset \rho_C (X)$; the preimage ${\tilde s}_n^{-1} (O_p)$ is the union of $n!$ distinct $\tilde G$--orbits, which are $\pi$--fibres, and all the others $\pi$--fibres in $Y$ are finite. Thus, by Lemma 2.9 it follows that $Y$ is almost C--hyperbolic modulo ${\tilde s}_n^{-1} (O_p)$, and hence $X$ is almost C--hyperbolic modulo $O_p$. Next we show that ii) corresponds exactly to the two exceptional cases mentioned in b), which proves b). By the assumption of the theorem $C^*$ has a cusp, and we suppose as above this cusp being at the point $q_0 = (0 : 0 : 1)$ and corresponding to the value $t = \infty$. This means that ${\rm deg}\,f \le n-2,\, {\rm deg}\,g \le n-2$ and ${\rm deg}\,h = n$ (see the proof of Lemma 5.6). Thus, the dual line $l_2 = q_0^* \subset I \!\! P^2$ belongs to $L_C$. If $f = t^k$ and $g = t^m$ are monomials, then $\rho_C^{-1} (l_{n-k,\,n-m}) = l_2$ and we have case i). Therefore, up to the transposition of $f$ and $g$ we may suppose further that $f = t^k$ and $h = t^m$ are monomials, while $g(t)$ is not. In that case $k \le n-2, \,{\rm deg}\,g \le n-2 ,\,m = n$ and $\rho_C^{-1} (l_{n-k,\,n-m}) = l_1 := \{x_1 = 0 \} \subset I \!\! P^2$. The dual point $q_1 = (0 : 1 : 0) = l_1^* \in I \!\! P^{2*}$ is a cusp of $C^*$ iff $k \ge 2$. Hence, ii) occurs iff here $k \le 1$, i.e. iff $C^*$ was projectively equivalent to one of the curves $(1 : t^n : g(t))$ or $(t : t^n : g(t))$, where ${\rm deg}\,g \le n-2$. If $C^*$ is one of these curves, then $X = I \!\! P^2 \setminus (C \cup L_C)$ is C--hyperbolic modulo $l_2 = \rho_C^{-1} (l_{n-k,\,0})$. \qed \noindent {\it Proof of c)}. Let $C = C_{\mu_0}$ be a monomial curve from the linear pencil $C_{\mu} = \{\alpha x_0^n + \beta x_1^k x_2^{n - k} =0\}$, where $\mu = (\alpha : \beta ) \in I \!\! P^1$. The pencil $(C_{\mu})$ is self--dual, i.e. $C_{\mu}^* = C_{\mu^*}$, where $\mu^*$ depends on $\mu$ (see 5.7, 5.9), and so without loss of generality we may assume that $C^* = C_{\mu}^* = C_{(1 : -1)}$. Thus, $C^*$ has the parametrizations $C^* = (\tau^k : \tau^n : 1) = (t^{n - k} : 1 : t^n )$, where $\tau = t^{-1}$. Since $C^*$ has a cusp, we have $n = {\rm deg}\, C^* \ge 3$ and ${\rm max}\,(k, n-k) \ge 2$. By permuting coordinates, if necessary, we may assume that $k \ge 2$. In this case the second parametrization, which we denote by $\nu$, fits in with the basic construction, i.e. $\nu (\infty ) = q_0 = (0 : 0 : 1)$ is a cusp of $C^*$ and $b^{(i)}_1 = 0,\,i = 0,\,1,\,2$. The parametrization $\nu$ being fixed as above, the Zariski embedding $\rho_C$ is given by the matrix $B_C = (e_k ,\,e_n ,\,e_0 )$. Therefore, $\rho_C \,:\, I \!\! P^2 \to I \!\! P^2_C = I \!\! P_{0, k, n} \subset {\bar A}_1 \subset I \!\! P^n$ is coordinatewise (cf. Remark 5.11, a): $$\rho_C (x_0 : x_1 : x_2 ) = (a_0 : \dots : a_n ) = (x_2 : 0 : \dots : \underbrace{x_0}_k : 0 : \dots : 0 : x_1)$$ The $ I \!\!\!\! C^*$--action $G$ on $ I \!\! P^n$ induces the $ I \!\!\!\! C^*$--action $G'$ on $ I \!\! P^2$, where $$G' \,:\,(\lambda, \,(x_0 : x_1 : x_2 )) \longmapsto (\lambda^k x_0 : \lambda^n x_1 : x_2 ) = (x_0 /\lambda^{n - k} : x_1 : x_2 /\lambda^n )$$ It is easily seen that the closure of a one-dimensional $G'$--orbit is an irreducible component of a member of the linear pencil $(C_{\mu})$. In what follows we identify $X$ resp. $Y$ with its image under $\rho_C$ resp. $\tilde \rho_C$. Let $f\,:\, I \!\!\!\! C \to X$ be an entire curve and ${\tilde f}\,:\, I \!\!\!\! C \to Y$ be its covering curve. From Lemma 6.3 it follows that the map $\pi \circ {\tilde \rho}_C \circ {\tilde f}$ is constant. This means that ${\tilde f}( I \!\!\!\! C)$ is contained in an orbit of $\tilde G$, and so $f( I \!\!\!\! C)$ is contained in a $G$--orbit, which in turn is contained in one of the curves $C_{\mu}$, as it is stated in c). Furthermore, ${\bar H}_0 \setminus {\bar D}_{n-1}$ being Kobayashi hyperbolic, the $k_Y$--distance between any two distinct $\tilde G$--orbits in $Y$ is positive. Over each $G$--orbit $O$ in $X$ there is $n! \,\,\tilde G$--orbits in $Y$, and each of them is maped by ${\tilde s}_n$ isomorphically onto $O$. Therefore, the $k_X$--distance between two different $G$--orbits in $X$, which is equal to the $k_Y$--distance between their preimages in $Y$, is positive, too. This proves c). Now the proof of Theorem 6.5 is complete. \qed \noindent {\bf 6.6. Remark.} In general, b) is not true for a plane curve whose dual is a quasi-monomial curve without cusps. Indeed, if $C$ is a three--cuspidal plane quartic, then $C^*$ is a nodal cubic, which is projectively equivalent to a quasi--monomial curve $t \longmapsto (t : t^3 : t^2 - 1 )$, where the node corresponds to $t = \pm 1$. The Kobayashi pseudo--distance of $ I \!\! P^2 \setminus C$ is degenerate on at least seven lines (see Remark 4.2), and thus $ I \!\! P^2 \setminus C$ is not C--hyperbolic modulo a line. \\[1ex] The next examples illustrate Theorem 6.5. \\[1ex] \noindent {\bf 6.7. Example.} Let $C \subset I \!\! P^2$ be the cuspidal cubic $4x_0^3 - 27x_1^2 x_2 =0$. Its dual curve $C^* \subset I \!\! P^{2*}$ is the cuspidal cubic with the equation $y_0^3 + y_1^2 y_2 =0$. The cusp of $C^*$ at the point $q_0 = (0 :0 :1)$ corresponds to the only flex of $C$ at the point $p_0 = (0 : 1 : 0)$, with the inflexional tangent $l_2 = \{x_2 = 0 \} \subset I \!\! P^2$, so that $L_C = l_2$. Consider the curve $C \cup l_2$. Its complement $X:= I \!\! P^2 \setminus (C \cup l_2)$ is neither C--hyperbolic nor Kobayashi hyperbolic. Indeed, $C$ is a member of the linear pencil of cubics $C_{\mu} = \{\alpha x_0^3 - \beta x_1^2 x_2 = 0 \}$, where $\mu = (\alpha : \beta ) \in I \!\! P^1$ (here $C = C_{\mu_0}$, where $\mu_0 = (4 : 27)$). This pencil is generated by its only non--reduced members $C_{(1:0)} = 3 l_0$ and $C_{(0 : 1)} = 2 l_1 + l_2$, where $l_i = \{x_i =0 \}\,,\,i=0,1,2$. The Kobayashi pseudo-distance $k_X$ is identically zero along any of the cubics $C_{\mu}\,,\mu \neq \mu_0$, because $C_{\mu} \cap X = C_{\mu} \setminus (C \cup l_2 ) \cong I \!\!\!\! C^*$ and $k_{ I \!\!\!\! C^*} \equiv 0$. Nevertheless, by Theorem 6.5, c) any entire curve $ I \!\!\!\! C \to X = I \!\! P^2 \setminus (C \cup l_2)$ is contained in one of the cubics $C_{\mu}$, where $\mu \in I \!\! P^1 \setminus \{\mu_0\}$. Moreover, $k_X (p,\,q) = 0 $ iff $p,\,q \in C_{\mu}$ for some $\mu \in I \!\! P^1 \setminus \{\mu_0 \}$. \\[1ex] \noindent {\bf 6.8. Example.} Let $C \subset I \!\! P^2$ be the nodal cubic $x_1^2 x_2 = x_0^3 + x_0^2 x_2$, and let $l_1 ,\,l_2 ,\,l_3$ be the three inflexional tangents of $C$. They correspond to the cusps of the dual curve $C^* \subset I \!\! P^{2*}$, which is the 3-cuspidal quartic $(2y_1 y_2 + y_0^2 )^2 = 4y_0^2 (y_0 - 2y_2 )(y_0 +y_2)$ (see Remark 4.2). Thus, $L_C = l_1 \cup l_2 \cup l_3 $. By Theorem 6.5, a) we have that $X := I \!\! P^2 \setminus (C \cup L_C) $ is almost C--hyperbolic. Hence, it is also Brody hyperbolic (see (2.6)). By Bezout Theorem three cusps of $C^*$ are not at the same line in $ I \!\! P^{2*}$. Therefore, their dual lines, which are inflexional tangents $l_1 ,\,l_2 ,\,l_3$ of $C$, are not passing through the same point. From this it easily follows that ${\rm reg}\,(C \cup l_1 \cup l_2 \cup l_3 )$ is hyperbolic. Thus, by Proposition 2.7 $X$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. \\[1ex] \noindent {\bf 6.9. Example.} Let $C \subset I \!\! P^2$ be the rational quintic $t \longmapsto (2t^5 - t^2 : -(4t^3 + 1) : 2t)$ with a cusp at the only singular point $(1 : 0 : 0)$. The dual curve $C^* \subset I \!\! P^{2*}$ is the quasi--monomial quartic $t \longmapsto (1 : t^2 : t^4 + t)$ given by the equation $(y_0 y_2 - y_1^2 )^2 = y_0^3 y_1$. It has the only singular point $q_0 = (0 : 0 : 1)$, which is a ramphoid cusp, i.e. it has the multiplicity sequence $(2,\,2,\,2,\,1,\,\dots)$ and $\delta = \mu /2 = 3$, where $\mu$ is the Milnor number. Any rational quartic with a ramphoid cusp is projectively equivalent to $C^*$ (see [Na, 2.2.5(a)]). The artifacts $L_C$ consist of the only cuspidal tangent line $l_2 = \{x_2 = 0 \}$ of $C$. By Theorem 6.5, b) the complement $ I \!\! P^2 \setminus (C \cup l_2)$ is almost C--hyperbolic. Note that $\Gamma := C \setminus l_2$ is a smooth rational affine curve in $ I \!\!\!\! C^2 \cong I \!\! P^2 \setminus l_2$, which is isomorphic to $ I \!\!\!\! C^* := I \!\!\!\! C \setminus \{0\}$. Thus, $X := I \!\!\!\! C^2 \setminus \Gamma$ is almost C--hyperbolic. \\[1ex] \noindent {\bf 6.10. Example.} Let $C' \subset I \!\! P^2$ be the rational quartic $t \longmapsto (t^3 (2t + 1) : -t(4t + 3) : -2)$. It has two singular points, a double cusp at the point $(0 : 0 : 1)$ (i.e. a cusp with the multiplicity sequence $(2, \,2,\,1,\,\dots)$ and $\delta = 2$) and another one, which is an ordinary cusp. The dual curve $C'^* \subset I \!\! P^{2*}$ is the quasi--monomial quartic $t \longmapsto (1 : t^2 : t^4 + t^3)$ given by the equation $(y_0 y_2 - y_1^2 )^2 = y_0 y_1^3$. It has the same type of singularities as $C'$, namely a double cusp at the point $q_0 = (0 : 0 : 1)$ and an ordinary cusp at the point $(1 : 0 : 0)$. Therefore, $L_{C'} = l_0 \cup l_2$, where $l_0 = \{x_0 = 0 \}$ and $l_2 = \{x_2 = 0 \}$. By Theorem 6.5, b) the complement $X:= I \!\! P^2 \setminus (C' \cup L_{C'})$ is almost C--hyperbolic. \\[2ex] \section{C--hyperbolicity of complements of maximal cuspidal rational curves} In Corollary 7.10 below we show that the complement of a maximal cuspidal rational curve of degree $d \ge 8$ in $ I \!\! P^2$ is almost C--hyperbolic. In a sense, this completes the study on C--hyperbolicity of $ I \!\! P^2 \setminus (C \cup L_C)$. The deep reason of this fact, which actually does not appear in the proof, is that the Teichm\"uller space $T_{0,n}$ of the Riemann sphere with $n$ punctures is a bounded domain in $ I \!\!\!\! C^n$ (cf. [Kal]).\\ Let us start with necessary preliminaries. \\[2ex] \noindent {\bf 7.1.} {\it Maximal cuspidal rational curves as generic plane sections of the discriminant.} Let $C \subset I \!\! P^2$ be a rational curve of degree $d > 1$. By the Class Formula (4) its dual curve $C^* \subset I \!\! P^{2*}$ is an immersed curve (or, equivalently, $L_C = \emptyset$) iff $d = 2(n - 1)$, where $n = {\rm deg}\,C^*$ (cf. 3.4). If in addition $C$ is a Pl\"ucker curve, then it has the maximal possible number of ordinary cusps, which is equal to $3(n - 2)$, and besides this it has also $2(n - 2)(n - 3)$ nodes. Such a curve $C$ is called {\it a maximal cuspidal rational curve} [Zar, p. 267]. Note that the dual $C^*$ of such a curve $C$ is a rational nodal curve of degree $n$ in $ I \!\! P^{2*}$. In particular, a generic maximal cuspidal rational curve $C$ naturally appears via the Zariski embedding $\rho_C \,:\, I \!\! P^2 \to I \!\! P^2_C \hookrightarrow I \!\! P^n$ as a generic plane section of the discriminant hypersurface $R_n \subset I \!\! P^n$ (see 3.6-3.7, 5.4). \\[2ex] \noindent {\bf 7.2.} {\it The moduli space of the $n$--punctured sphere as an orbit space}. Note, first of all, that the Vieta map $s_n\,:\,( I \!\! P^1)^n \to S^n I \!\! P^1 = I \!\! P^n$ is equivariant with respect to the natural actions of the group $ I \!\! P GL(2,\, I \!\!\!\! C) = {\rm Aut}\, I \!\! P^1$ on $ ( I \!\! P^1)^n$ and on $ I \!\! P^n$, respectively. The branching divisors $D_n$ (the union of the diagonals) resp. $R_n$ (the discriminant divisor), as well as their complements are invariant under the corresponding actions. It is easily seen that for $n \ge 3$ the orbit space of the $ I \!\! P GL(2,\, I \!\!\!\! C)$--action on $ I \!\! P^n \setminus R_n$ is naturally isomorphic to the moduli space $M_{0,\,n}$ of the Riemann sphere with $n$ punctures. Denote by $\tilde M_{0,\,n}$ the quotient $(( I \!\! P^1)^n \setminus D_n)\, / I \!\! P GL(2,\, I \!\!\!\! C)$. We have the following commutative diagram of equivariant morphisms \begin{picture}(200,80) \unitlength0.2em \thicklines \put(54,25){$( I \!\! P^1)^n \setminus D_n$} \put(110,25){$\tilde M_{0,\,n}$} \put(87,27){$\vector(1,0){15}$} \put(60,5){$ I \!\! P^n \setminus R_n$} \put(87,6){$\vector(1,0){15}$} \put(110,5){$M_{0,\,n}$} \put(69,22){$\vector(0,-1){11}$} \put(114,22){$\vector(0,-1){11}$} \put(90,10){$\pi_n$} \put(90,30){$\tilde \pi_n$} \put(60,16){$s_n$} \put(170,15){(7)} \end{picture} \noindent {\bf 7.3.} {\it Description of $\tilde M_{0,\,n}$}. The cross--ratios $\sigma_i (z) = (z_1,\,z_2;\,z_3 ,\,z_i)$, where $z = (z_1 ,\dots, z_n ) \in ( I \!\! P^1)^n$ and $4 \le i \le n$, define a morphism $$\sigma^{(n)} = (\sigma_4 ,\,\dots, \,\sigma_n )\,:\,( I \!\! P^1)^n \setminus D_n \to ( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3}$$ (here as before $\, I \!\!\!\! C^{**} = I \!\! P^1 \setminus \{0,\,1,\,\infty\}$). By the invariance of cross--ratio $\sigma^{(n)}$ is constant along the orbits of the action of $ I \!\! P GL(2,\, I \!\!\!\! C)$ on $( I \!\! P^1 )^n \setminus D_n$. Therefore, it factorizes through a mapping of the orbit space $\tilde M_{0,\,n} \to ( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3}$. On the other hand, for each point $z \in ( I \!\! P^1)^n \setminus D_n$ its $ I \!\! P GL(2,\, I \!\!\!\! C)$--orbit $O_z$ contains the unique point $z'$ of the form $z' = (0,\,1,\,\infty,\,z'_4 ,\,\dots,\,z'_n )$. This defines a regular section $\tilde M_{0,\,n} \to ( I \!\! P^1)^n \setminus D_n$, and its image coincides with the image of the biregular embedding $$( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3} \ni u = (u_4 ,\,\dots ,\,u_n ) \longmapsto (0,\,1,\,\infty,\,u_4 ,\,\dots,\,u_n ) \in ( I \!\! P^1)^n \setminus D_n \,\,.$$ This shows that the above mapping $\tilde M_{0,\,n} \to ( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3}$ is an isomorphism. \\[2ex] \noindent {\bf 7.4.} {\it $ I \!\! P GL(2,\, I \!\!\!\! C)$--orbits.} Here as before we treat $ I \!\! P^n$ as the projectivized space of the binary forms of degree $n$ in $u$ and $v$. For instance, $e_k = (0:\dots :0:1_k :0:\dots :0) \in I \!\! P^n$ corresponds to the forms $cu^{n - k}v^k$, where $c \in I \!\!\!\! C^*$. Denote by $O_q$ the $ I \!\! P GL(2,\, I \!\!\!\! C)$--orbit of a point $q \in I \!\! P^n$. Clearly, $O_{e_i} = O_{e_{n - i}},\,i=0,\dots, n$; $O_{e_0}$ is the only one--dimensional orbit and, at the same time, the only closed orbit; $O_{e_i},\,i=1,\dots,[n/2]$, are the only two-dimensional orbits, and any other orbit has dimension $3$. Note that $O_{e_0} = C_n$ is the dual rational normal curve, and $S:= O_{e_0} \cup O_{e_1}$ is its developable tangent surface (see 5.4). If $O_q$ is an orbit of dimension $3$, then its closure $\bar {O_q}$ is the union of the orbits $O_q, O_{e_0}$ and those of the orbits $O_{e_i}, i=1,\dots, n-1$, for which the form $q$ has a root of multiplicity $i$ [AlFa, Proposition 2.1]. Furthermore, for any point $q \in I \!\! P^n \setminus R_n$, i.e. for any binary form $q$ without multiple roots, its orbit $O_q$ is closed in $ I \!\! P^n \setminus R_n$ and $\bar {O_q} = O_q \cup S$, where $S = \bar {O_q} \cap R_n$. Therefore, any Zariski closed subvariety $Z$ of $ I \!\! P^n$ such that ${\rm dim}\,(O_q \cap Z) > 0$ must meet the surface $S$. These observations yield the following lemma. \footnote {We are gratefull to H. Kraft who pointed out to us an approach which is based on the notion of the associated cone of an orbit [Kr] (here we have used a simplified version of it), and to M. Brion for mentioning to us of the paper [AlFa].} \\[2ex] \noindent {\bf 7.5. Lemma.} {\it If a linear subspace $L$ in $ I \!\! P^n$ does not meet the surface $S = {\bar O_{e_1}} \subset R_n$, then it has at most finite intersection with any of the orbits $O_q$, where $q \in I \!\! P^n \setminus R_n$. In particular, this is so for a generic linear subspace $L$ in $ I \!\! P^n$ of codimension at least $3$.} \\[2ex] \noindent {\bf 7.6. Remark.} Fix $k$ distinct points $z_1 ,\dots, z_k \in I \!\! P^1$, where $3 \le k \le n$. Let $g_0$ be a binary form of degree $k$ with the roots $z_1 ,\dots, z_k$. Consider the projectivized linear subspace $L_0 \subset I \!\! P^n$ of codimension $k$ consisting of the binary forms of degree $n$ divisible by $g_0$. It is easily seen that $L_0 \cap S = \emptyset$. This gives a concrete example of such a subspace. \\ The next tautological lemma is used below in the proof of Theorem 7.9.\\[2ex] \noindent {\bf 7.7. Lemma.} {\it Let $C \subset I \!\! P^2$ be a rational curve. Put $n = {\rm deg}\,C^*$, and let as before $ I \!\! P^2_C = \rho_C ( I \!\! P^2 ) \hookrightarrow I \!\! P^n$ be the image under the Zariski embedding. \\ \noindent a) The plane $ I \!\! P^2_C$ meets the surface $S = {\bar O_{e_1}}$ iff there exists a local irreducible analytic branch $(A^*, p^* )$ of the dual curve $C^*$ such that $\,i(T_{p^*} A^* , \,A^* ;\,p^*) \ge n - 1$. \\ \noindent b) Furthermore, if $C^*$ has a cusp $(A^*, p^* )$ of multiplicity $n - 1$, then $\rho_C (l_{p^*}) \subset I \!\! P^2_C \cap S$, where $l_{p^*} \subset L_C \subset I \!\! P^2$ is the dual line of the point $p^* \in I \!\! P^{2*}$. \\ \noindent c) If the dual curve $C^*$ has only ordinary cusps and flexes and $n = {\rm deg}\,C^* \ge 5$, then $ I \!\! P^2_C \cap S = \emptyset$. } \\[2ex] \noindent{\it Proof.} a) By the definition of the Zariski embedding $q \in I \!\! P^2_C \cap S$ iff, after passing to the normalization $\nu :\, I \!\! P^1 \to C^*$ and identifying $ I \!\! P^2$ with its image $ I \!\! P^2_C$ under the Zariski embedding $\rho_C$, the dual line $l_q \subset I \!\! P^2$ cuts out on $C^*$ a divisor of the form $(n - 1)a + b$, where $a, b \in I \!\! P^1$. Then $p^* := \nu (a) \in C^*$ is the center of a local branch $A^*$ of $C^*$ which satisfies the condition in a). The converse is evidently true. \qed \noindent b) For any point $q \in l_{p^*}$ its dual line $l_q \subset I \!\! P^{2*}$ passes through $p^*$, and hence by the above consideration we have $\rho_C (q) \in I \!\! P^2_C \cap S$. \qed \noindent c) By the condition we have that $i(T_{p^*} A^* , \,A^* ;\,p^*) \le 3 < n - 1$ for any local analytic branch $(A^*, p^* )$ of $C^*$. Now the result follows from a). \qed \noindent {\bf 7.8. Lemma.} {\it Let $C^* \subset I \!\! P^{2*}$ be a rational curve of degree $n$. Then the complement $X = I \!\! P^2 \setminus (C \cup L_C)$ is almost C--hyperbolic, whenever $ I \!\! P^2_C \cap S = \emptyset$.} \\[1ex] \noindent{\it Proof.} Consider the following commutative diagram of morphisms: \\ \begin{picture}(250,80) \unitlength0.2em \thicklines \put(-5,5) {${ I \!\! P^2} \setminus (C \cup L_C) = X$} \put(35,25){$Y$} \put(36,22){\vector(0,-1){11}} \put(27,16){${\tilde s}_n$} \put(44,27){{\vector(1,0){14}}} \put(49,30){${\tilde\rho}_C$} \put(44,6){\vector(1,0){14}} \put(49,10){$\rho_C$} \put(66,5){$ I \!\! P^n \setminus R_n$} \put(62,25){$( I \!\! P^1)^n \setminus D_n$} \put(76,22){\vector(0,-1){11}} \put(128,22){\vector(0,-1){11}} \put(124,5){$M_{0,\,n}$} \put(67,16){$s_n$} \put(170,15){(8)} \put(92,27){{\vector(1,0){14}}} \put(94,6){{\vector(1,0){20}}} \put(95,30){${\tilde \pi}_n$} \put(100,10){$\pi_n$} \put(110,25){$ ( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3} \hookrightarrow ( I \!\!\!\! C^{**})^{n - 3}$} \end{picture} \noindent where ${\tilde s}_n \,:\,Y \to X$ is the induced covering (cf. (7) and 7.2--7.3 above). From Lemma 7.5 it follows that the mapping $\pi_n \circ \rho _C \,:\,X \to M_{0,\,n}$ has finite fibres. Hence, the same is valid for the mapping ${\tilde \pi}_n \circ {\tilde\rho}_C \,:\, Y \to ( I \!\!\!\! C^{**})^{n - 3} \setminus D_{n - 3}$. By Lemma 2.5 $Y$, and thus also $X$, are almost C--hyperbolic. \qed {}From this lemma and Lemma 7.7 we have the following theorem, which is a useful supplement to Theorem 6.5. \\[1ex] \noindent {\bf 7.9. Theorem.} {\it Let $C^* \subset I \!\! P^{2*}$ be a rational curve of degree $n$ such that $\,i(T_{p^*} A^* , \,A^* ;\,p^*) \le n - 2$ for any local analytic branch $(A^* , \,p^* )$ of $C^*$. Let $C = (C^*)^* \subset I \!\! P^2$ be the dual curve. Then the complement $X = I \!\! P^2 \setminus (C \cup L_C)$ is almost C--hyperbolic. In particular, this is so if $n \ge 5$ and $C^*$ has only ordinary cusps and flexes.} \\[2ex] The next corollary is an addition to Theorem 4.1, b). \\ \noindent {\bf 7.10. Corollary.} {\it Let $C \subset I \!\! P^2$ be a maximal cuspidal rational curve of degree $d = 2(n - 1) \ge 8$. Then $X = I \!\! P^2 \setminus C$ is almost C--hyperbolic, Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. In particular, this is the case if the dual curve $C^*$ is a generic rational nodal curve of degree $n \ge 5$ in $ I \!\! P^{2*}$. } \\ \noindent {\it Proof.} The first statement immediately follows from Theorem 7.9, while the second one follows from Proposition 2.7. Indeed, under our assumptions we have $n \ge 5$, and therefore the curve $C$ has at least $9$ cusps. Hence ${\rm reg}\,C$ is a hyperbolic curve. The last statement is evident. \qed The next example shows that our method is available not for all rational curves whose dual curves are nodal. \\ \noindent {\bf 7.11. Example.} Let $C^* = (p(t) : q(t) : 1)$ be a parametrized plane rational curve, where $p, q \in I \!\!\!\! C [t]$ are generic polynomials of degrees $n$ and $n - 1$, respectively. Then $C^*$ is a nodal curve of degree $n$ which is the projective closure of an affine plane polynomial curve with one place at infinity, at the point $(1 : 0 : 0)$, and this is a smooth point of $C^*$. Thus, the line $l_2 = \{x_2 = 0 \}$ is an inflexional tangent of order $n - 2$ of $C^*$, and so by Lemma 7.7, a) $ I \!\! P^2_C \cap S \neq \emptyset$. Therefore, we can not apply in this case the same approach as above. \\[2ex] At last, we can summarize the main results of the paper (cf. Theorems 4.1, 6.5 and 7.9). \\[1ex] \noindent {\bf 7.12. Theorem.} {\it Let $C \subset I \!\! P^2$ be an irreducible curve of genus $g$. Put $n = {\rm deg}\,C^*$ and $X = I \!\! P^2 \setminus (C \cup L_C )$. \\ \noindent a) If $g \ge 1$, then $X$ is C--hyperbolic. If $g = 0$, then $X$ is almost C--hyperbolic if at least one of the following conditions is fulfilled: \\ \noindent i) $\,i(T_{p^*} A^* , \,A^* ;\,p^*) \le n - 2$ for any local analytic branch $(A^* , \,p^* )$ of $C^*$; \\ \noindent ii) $C^*$ has a cusp and it is not projectively equivalent to one of the curves $(1 : g(t) : t^n ),\,\,(t : g(t) : t^n )$, where $g \in I \!\!\!\! C [t],\, {\rm deg}\,g \le n - 2$. \footnote{The monomial curves correspond here to $g(t) = t^k,\,k \le n - 2$. Note that the curves $(1 : t^{n - 1} : t^n )$ and $(1 : t : t^n )$, being considered as non--parametrized ones, are projectively equivalent, and therefore all monomial curves have been excluded.} \\ \noindent b) Let, furthermore, $C^*$ be an immersed curve. If $g \ge 1$, then $ I \!\! P^2 \setminus C$ is C--hyperbolic. If $g = 0$ and i) is fulfilled, then $ I \!\! P^2 \setminus C$ is almost C--hyperbolic; in particular, this is so if $C^*$ is a generic rational nodal curve of degree $n \ge 5$. In both cases $ I \!\! P^2 \setminus C$ is Kobayashi complete hyperbolic and hyperbolically embedded into $ I \!\! P^2$. } \\[2ex] \newpage \begin{center} {\LARGE References} \end{center} \noindent [AlFa] P. Aluffi, C. Faber. {\sl Linear orbits of $d$--tuples of points in $ I \!\! P^1$}, J. reine angew. Math. 445 (1993), 205--220 \\ \noindent [Au] A. B. Aure. {\sl Pl\"ucker conditions on plane rational curves}, Math. Scand. 55 (1984), 47--58, with {\sl Appendix} by S. A. Str\"omme, ibid. 59--61 \\ \noindent [CaGr] J. A. Carlson, M. Green. {\sl Holomorphic curves in the plane}, Duke Math. J. 43 (1976), 1--9 \\ \noindent [Co] J. L. Coolidge. {\sl A Treatise on Algebraic Plane Curves}, N.Y.: Dover, 1959 \\ \noindent [Deg 1] A. I. Degtjarjov. {\sl Isotopy classification of complex plane projective curves of degree 5}, Preprint LOMI P--3-87, Leningrad, 1987, 1--17 \\ \noindent [Deg 2] A. I. Degtjarjov. {\sl Topology of plane projective algebraic curves}, PhD Thesis, Leningrad State University, 1987 (in Russian) \\ \noindent [Del] P. Deligne. {\sl Le groupe fondamental du complement d'une courbe plane n'ayant que des points doubles ordinaires est ab\'elien}, Sem. Bourbaki, 1979/1980, Lect. Notes in Math. vol. 842, Springer-Verlag (1981), 1--25 \\ \noindent [DSW1] G. Dethloff, G. Schumacher, P.-M. Wong. {\sl Hyperbolicity of the complements of plane algebraic curves}, preprint, Math. G\"otting. 31 (1992), 38 p. (to appear in Amer. J. Math.) \\ \noindent [DSW2] G. Dethloff, G. Schumacher, P.-M. Wong. {\sl On the hyperbolicity of the complement of curves in algebraic surfaces: The three component case}, preprint, Essen e.a., 1993, 25 p. (to appear in Duke Math. J.) \\ \noindent [DL] I. Dolgachev, A. Libgober. {\sl On the fundamenthal group of the complement to a discriminant variety}, In: Algebraic Geometry, Lecture Notes in Math. 862, 1--25, N.Y. e.a.: Springer, 1981 \\ \noindent [Fu] W. Fulton. {\sl On the fundamental group of the complement to a node curve}, Ann. of Math. (2) 111 (1980), 407--409 \\ \noindent [GP] H. Grauert, U. Peternell. {\sl Hyperbolicity of the complement of plane curves}, Manuscr. Math. 50 (1985), 429--441 \\ \noindent [Gr1] M. Green. {\sl The complement of the dual of a plane curve and some new hyperbolic manifolds}, in: 'Value Distribution Theory`, Kujala and Vitter, eds., N.Y.: Marcel Dekker, 1974, 119--131 \\ \noindent [Gr2] M. Green. {\sl Some examples and counterexamples in value distribution theory}. Compos. Math. 30 (1975), 317-322 \\ \noindent [Gr3] M. Green. {\sl The hyperbolicity of the complement of $2n+1$ hyperplanes in general position in $ I \!\! P_n$ and related results}, Proc. Amer. Math. Soc. 66 (1977), 109--113 \\ \noindent [GK] G. M. Greuel, U. Karras. {\sl Families of varieties with prescribed singularities}, Compositio Math. 69 (1989), 83--110 \\ \noindent [GH] Ph. Griffiths, J. Harris. {\sl Principles of Algebraic Geometry}. N.Y. e.a.: J. Wiley and Sons Inc., 1978 \\ \noindent [Kal] Sh. I. Kaliman. {\sl The holomorphic universal covers of spaces of polynomials without multiple roots}, Selecta Mathem. form. Sovietica, 12 (1993) No. 4, 395--405 \\ \noindent [Kan] J. Kaneko. {\sl On the fundamental group of the complement to a maximal cuspidal plane curve}, Mem. Fac. Sci. Kyushu Univ. Ser. A. 39 (1985), 133--146 \\ \noindent [Ki] P. Kiernan. {\sl Hyperbolically imbedded spaces and the big Picard theorem}, Math. Ann. 204 (1973), 203--209 \\ \noindent [KiKo] P. Kiernan, Sh. Kobayashi. {\sl Holomorphic mappings into projective space with lacunary hyperplanes}, Nagoya Math. J. 50 (1973), 199-216 \\ \noindent [Ko1] Sh. Kobayashi. {\sl Hyperbolic manifolds and holomorphic mappings}. N.Y. a.e.: Marcel Dekker, 1970 \\ \noindent [Ko2] Sh. Kobayashi, {\sl Intrinsic distances, measures and geometric function theory}, Bull. Amer. Math. Soc 82 (1976), 357-416 \\ \noindent [Kr] H. Kraft. {\sl Geometrische Methoden in der Invariantentheorie}. Braunschweig/Wiesbaden: Vieweg und Sohn, 1985 \\ \noindent [Lib] A. Libgober. {\sl Fundamental groups of the complements to plane singular curves}, Proc. Sympos. in Pure Mathem. 46 (1987), 29--45 \\ \noindent [Li] V. Ja. Lin. {\sl Liouville coverings of complex spaces, and amenable groups}, Math. USSR Sbornik, 60 (1988), 197--216\\ \noindent [LiZa] V. Ya. Lin, M.G. Zaidenberg, {\sl Finiteness theorems for holomorphic mappings}, Encyclopaedia of Math. Sci. 9 (1986), 127-194 (in Russian). English transl. in Encyclopaedia of Math. Sci. Vol.9. Several Complex Variables III. N.Y. e.a.: Springer Verlag, 1989, 113-172 \\ \noindent [Ma] A. Mattuck, {\sl Picard bundles}, Illinois J. Math. 5 (1961), 550--564 \\ \noindent [Na] M. Namba. {\sl Geometry of projective algebraic curves}, N.Y. a.e.: Marcel Dekker, 1984 \\ \noindent [Pi] R. Piene. {\sl Cuspidal projections of space curves}, Math. Ann. 256 (1981), 95--119 \\ \noindent [O] M. Oka. {\sl Symmetric plane curves with nodes and cusps}, J. Math. Soc. Japan, 44, No. 3 (1992), 375--414 \\ \noindent [Re] H. J. Reiffen. {\sl Die Carath\'eodorysche Distanz und ihre zugeh\"orige Differentialmetrik}, Math. Ann. 161 (1965), 315--324 \\ \noindent [Ro] H. Royden. {\it Automorphisms and isometries of Teichm\"uller space}, In: Advances in the Theory of Riemann Surfaces, 1969 Stony Brook Conf. Ann. of Math. St. 66, Princeton, N.J.: Princeton Univ. Press, 1971, 369--383 \\ \noindent [Se] F. Severi. {\sl Vorlesungen \"uber algebraische Geometrie}, Leipzig: Teubner, 1921 \\ \noindent [SY] Y.--T. Siu, S.--K. Yeung. {\sl Hyperbolicity of the complement of a generic smooth curve of high degree in the complex projective plane}, preprint, 1994, 56 p. \\ \noindent [Ve] G. Veronese. {\sl Behandlung der projectivischen Verh\"altnisse der R\"aume von verschiedenen Dimensionen durch das Princip des Projicirens und Schneidens}, Math. Ann. 19 (1882), 193--234 \\ \noindent [Wa] R. J. Walker. {\sl Algebraic curves}. Princeton Meth. Series 13, Princeton, N.J.: Princeton University Press, 1950 \\ \noindent [Za1] M. Zaidenberg. {\sl On hyperbolic embedding of complements of divisors and the limiting behavior of the Kobayashi-Royden metric}, Math. USSR Sbornik 55 (1986), 55--70 \\ \noindent [Za2] M. Zaidenberg. {\sl The complement of a generic hypersurface of degree $2n$ in $ I \!\!\!\! C I \!\! P^n$ is not hyperbolic}. Siberian Math. J. 28 (1987), 425--432 \\ \noindent [Za3] M. Zaidenberg. {\sl Stability of hyperbolic imbeddedness and construction of examples}, Math. USSR Sbornik 63 (1989), 351--361 \\ \noindent [Za4] M. Zaidenberg. {\sl Hyperbolicity in projective spaces}, Proc. Conf. on Hyperbolic and Diophantine Analysis, RIMS, Kyoto, Oct. 26--30, 1992. Tokyo, TIT, 1992, 136--156 \\ \noindent [Zar] O. Zariski. {\sl Collected Papers}. Vol III : {\sl Topology of curves and surfaces, and special topics in the theory of algebraic varieties}. Cambridge, Massachusets e. a.: The MIT Press, 1978 \\ \noindent Gerd Dethloff\\ Mathematisches Institut der Universit\"at G\"ottingen\\ Bunsenstra\3e 3-5\\ 3400 G\"ottingen\\ Germany\\ \vspace{0.8cm}e-mail: [email protected]\\ \noindent Mikhail Zaidenberg\\ Universit\'{e} Grenoble I \\ Institut Fourier des Math\'ematiques\\ BP 74\\ 38402 St. Martin d'H\`{e}res--c\'edex\\ France\\ \vspace{0.8cm}e-mail: [email protected] \end{document}
1995-01-23T06:20:11
9501
alg-geom/9501010
en
https://arxiv.org/abs/alg-geom/9501010
[ "alg-geom", "math.AG" ]
alg-geom/9501010
E. Looijenga
Eduard Looijenga
On the tautological ring of $\M _g$
6 pages, amstex 2.1
null
10.1007/BF01884306
null
null
We prove among other things that any product of tautological classes of $\M_g$ of degree $d$ (in the Chow ring of $\M _g$ with rational coefficients) vanishes for $d>g-2$ and is proportional to the class of the hyperelliptic locus in degree $g-2$.
[ { "version": "v1", "created": "Fri, 20 Jan 1995 09:07:01 GMT" } ]
2015-06-30T00:00:00
[ [ "Looijenga", "Eduard", "" ] ]
alg-geom
\section{\global\advance\headnumber by1\global\labelnumber=0{{\the\headnumber}.\ }} \define\label{(\global\advance\labelnumber by1 \the\headnumber .\the\labelnumber )\enspace} \NoBlackBoxes \topmatter \title On the tautological ring of $\M _g$ \endtitle \rightheadtext{Tautological ring} \author Eduard Looijenga \endauthor \address Faculteit Wiskunde en Informatica, Universiteit Utrecht, P.O. Box 80.010, 3508 TA Utrecht, The Netherlands\endaddress \email looijeng\@math.ruu.nl\endemail \abstract We prove that any product of tautological classes of $\M _g$ of degree $d$ (in the Chow ring of $\M _g$ with rational coefficients) vanishes for $d>g-2$ and is proportional to the class of the hyperelliptic locus in degree $g-2$. \endabstract \endtopmatter \document \head \section Results \endhead Fix an integer $g\ge 2$ and denote by $\CC _g^n$ the moduli space of tuples $(C,x_1,\dots ,x_n)$, where $C$ is a smooth connected projective curve of genus $g$ and $x_1,\dots ,x_n$ are (not necessarily distinct) points of $C$; we also write $\M _g$ for $\CC _g^0$. Forgetting the $i$th point defines a morphism $\CC _g^n\to\CC _g^{n-1}$ whose relatively dualizing sheaf is denoted by $\omega _i$ ($i=1,\dots ,n)$. We write $K_i$ for the first Chern class of $\omega _i$, considered as an element of the Chow group $A^1(\CC _g^n)$ (with rational coefficients); for $n=1$ we also write $K$. Our main result is: \proclaim{\label Theorem} Any monomial of degree $d$ in the classes $K_1,\dots ,K_n$ is a linear combination of the classes of the irreducible components of the locus parametrizing tuples $(C,x_1,\dots ,x_n)$ admitting a morphism $C\to\Pone$ of degree $\le g+n$ such that the fiber over $0$ (resp.\ $\infty$) has at most $g+n-d-1$ points (resp.\ is a singleton) and $\{ x_1,\dots ,x_n\}$ is contained in one of these two fibers. (Hence such a class is zero when $d>g+n-2$.) All monomials of degree $g+n-2$ are proportional to the class of the locus parametrizing tuples $(C,x_1,\dots ,x_n)$ with $C$ hyperelliptic and $x_1=\cdots =x_n$ a Weierstra\ss\ point. \endproclaim The direct image of $K^{d+1}$ in $A^d(\M _g)$ is the Mumford--Morita--Miller {\it tautological class} $\kappa _d$. Mumford showed in his fundamental paper \cite{4} that the subring of $A^{\bullet}(\M _g)$ generated by these classes (the {\it tautological ring} of $\M _g$) is already generated by $\kappa _1,\dots ,\kappa _{g-2}$. On the basis of many calculations Carel Faber has made the intriguing conjecture that this ring has the formal properties of the even-dimensional cohomology ring of a projective manifold of dimension $g-2$, i.e., satisfies Poincar\'e duality and a Lefschetz decomposition. We offer the following support for this conjecture: \proclaim{\label Theorem} Any product of tautological classes that has degree $d$ is a linear combination of the classes of the irreducible components of the locus parametrizing curves $C$ admitting a morphism $C\to\Pone$ of degree $\le 2g-2$ totally ramified over $\infty$ and with at most $g-1-d$ points over $0$ (hence is zero when $d>g-2$). All such classes of degree $g-2$ are proportional to the class of the hyperelliptic locus. \endproclaim A finer analysis of our proof may well yield that $\kappa _1^{g-2}$ is a nonzero multiple of the hyperelliptic class, but it is not known whether the latter is actually nonzero. The proof of the theorems uses the flag of subvarieties of $\M _g$ introduced by Arbarello \cite{1}, a variant of which was exploited by Diaz \cite{2} to prove that $\M _g$ has no complete subvarieties of dimension $>g-2$. Our simple key result \refer{2.4} serves as a substitute for Diaz's lemma $2$ in \cite{2} and can be used in that paper to eliminate the use of compactifications of Hurwitz schemes (see \refer{2.8}). The proof of the second assertion of each theorem involves an application of the Fourier transform for abelian varieties, due to Mukai and Beauville. \smallskip In this paper we only consider Chow groups with respect to rational equivalence, tensorized with $\Q$, and graded by codimension, notation: $A^{\bullet}$. If $X$ is a variety that is smooth, or more generally, that admits a smooth Galois covering, then there is an intersection product $A^k(X)\otimes A^l(X)\to A^{k+l}(X)$. \smallskip I thank Johan de Jong for drawing my attention to the paper by Deninger--Murre \cite{2} and for comments on a first draft. \head \section Proofs \endhead \label Let $C$ be a smooth projective curve of genus $g$ and let $D_0$ and $D_{\infty}$ be positive divisors on $C$ that are linearly equivalent, but whose supports are disjoint. Then there is a finite morphism $\pi :C\to\Pone$ such that $\pi ^*(i)=D_i$ ($i=0,\infty $). If $p\in C$ occurs in $D_i$ with multiplicity $m_p>0$, then $\pi$ determines an isomorphism of $\C\cong T^*_i\Pone$ onto $T^*_pC ^{\otimes m_p}$. However, $\pi$ is not unique for it is defined up to natural action of $\C ^{\times}$ on $\Pone$. That ambiguity can be eliminated as follows. Let $R$ denote the part of the ramification divisor of $\pi$ that lies over $\Pone -\{ 0,\infty \}$. If $c$ denotes the number of points of $\supp (D_0+D_{\infty})$, then the Riemann-Hurwitz formula implies that the degree $r$ of $R$ is equal to $2g-2+c$. If $\pi _*(R)= \sum _i n_i(z_i)$, then $\pi$ can be normalized in such a way that $\prod _i z_i ^{n_i}=1$. This normalization is unique up to multiplication by an $r$th root of unity. So for $p$ and $m_p$ as above, and $\pi$ normalized, the corresponding generator of $T^*_pC ^{\otimes m_p}$ raised to the $r$th power gives a {\it canonical} generator of $T^*_pC ^{\otimes m_pr}$. This argument works just as well in families and so we obtain: \proclaim{\label Proposition} Let $f:\CC \to S$ be a projective family of smooth genus $g$ curves with reduced base. Let $D_0$ and $D_{\infty}$ be positive relative divisors on $\CC$ whose supports are disjoint and are \'etale over $S$. Suppose that their difference is linearly equivalent to the pull-back of a divisor on $S$. Then for every section $x:S\to \CC$ of $f$ with image in the support of $D_0+D_{\infty}$, a suitable positive tensor power of $x^*\omega _{\CC /S}$ is trivial. \endproclaim We shall use the following simple fact: \proclaim{\label Lemma} Let $L_1,\dots ,L_d$ be line bundles on a variety $V$ and let $V=V^0\supset V^1\supset\cdots\supset V^d$ be a chain of closed subvarieties such that $L_k$ is trivial on $V^{k-1}-V^k$. Then $c_1(L_1)\cdots c_1(L_d)$ has support in $V^d$. \endproclaim The key result we need is: \proclaim{\label Lemma} Let $d$ be a positive integer and let $\{(C_t,x_t,P_t)\} _{t\in\Delta}$ be an analytic family of triples consisting of a smooth connected projective curve $C_t$, a point $x_t\in C_t$, and a pencil $P_t$ on $C_t$ containing $d(x_t)$. Assume that for $t\not=0$, $P_t$ has no base points and let $R_t$ be the part of the ramification divisor on $C_t-x_t$ of the associated morphism $C_t\to P_t$. If $R_0$ is the limit of $R_t$ for $t\to 0$, then the multiplicity of $x_0$ in $R_0$ is also the multiplicity of $x_0$ as a fixed point of $P_0$. \endproclaim \demo{Proof} Represent the family by a smooth analytic morphism $t:\CC\to\Delta$ with section $x:\Delta \to\CC$. Extend $t$ to a chart $(z,t)$ at $x_0$ such that $z=0$ is the image of $x$ at $x_0$. In terms of these coordinates generators of $P_t$ can be represented by $z^d$ and a holomorphic function $A(z,t)=\sum _{i\not= d} a_i(t)z^i$ which is divisible neither by $t$ nor by $z$. The first index $k$ for which $a_k(0)\not= 0$ is $<d$ and is equal to the multiplicity of $x_0$ as fixed point of $P_0$. In the domain of the chart, the divisor $R_t$ is given by locus where the $z$-derivatives of $A$ and $z^d$ are proportional, i.e., by the divisor of $\sum _{i \not= d} (i-d)a_i(t)z^i$ ($t\not= 0$). This expression is not divisible by $z$ or $t$ so that $R_0$ is given by $\sum _{i\not= d} (i-d)a_i(0)z^i$. So $x_0$ occurs with multiplicity $k$ in $R_0$. \enddemo An immediate consequence is an amplification of a result due to Arbarello \cite{1} and Diaz \cite{2}: \proclaim{\label Corollary} Suppose that in the situation of \refer{2.4} there exists an analytic section $\{ D_t\in P_t\} _{t\in\Delta}$ such that for $t\not= 0$, $\supp (D_t)$ is disjoint with $x_t$ and has $d-r$ points, whereas $D_0=d(x_0)$. Then $P_0$ can be written as $r(x_0) +P'$. \endproclaim \medskip \label If $d$ is positive integer, then we have moduli space $P(d)$ of triples $(C,x,P)$ with $C$ a smooth projective curve of genus $g$, $x\in C$ and $P$ a pencil on $C$ containing $(d)x$. The existence of this is clear if $d>2g-2$, for then this is just a bundle of projective spaces of dimension $d-g-1$ over $\CC _g$; the remaining cases $d\le 2g-2$ follows from this by simply viewing $P(d)$ as the locus in $P(2g-1)$ parametrizing triples $(C,x,P)$ for which $x$ is a fixed point in $P$ of multiplicity $2g-1-d$. This implies that we also have defined a moduli space $Z$ of tuples $(C,x_1,\dots ,x_n,x,D,P)$ with $C$ a smooth projective curve of genus $g$, $x_1,\dots ,x_n,x\in C$, $P$ a pencil on $C$ containing $(n+g)x$, $D$ a degenerate member of $P$ and $\{ x_1,\cdots ,x_n\}\subset \supp (D)$. Notice that $D$ and $x$ determine $P$ unless $D=(n+g)(x)$. The forgetful morphism $f:Z\to\CC ^n_g$ is clearly proper. The tuples for which $\supp (D)$ has at most $g+n-1-k$ points outside $x$ define a closed subvariety $Z^k$ of $Z$. It is clear that $Z^{n+g-1}$ can be identified with the set of tuples $(C,x,\dots ,x,x,(n+g)x,P)$ with $P$ a pencil through $(n+g)(x)$. \proclaim{\label Lemma} For $k<g+n-1$, $Z^k-Z^{k+1}$ is Zariski-open in an affine variety of pure dimension $3g-3+n-k$ and $f^*K_i|Z^k-Z^{k+1}=0$ ($i=1,\dots ,n$). \endproclaim \demo{Proof} Let $k<g+n-1$ and let $W$ be a connected component of $Z^k-Z^{k+1}$. If $(C,x_1,\dots ,x_n,x,D,P)$ represents an element of $W$, then write $D=m(x)+D'$ with $x\notin\supp (D')$ so that $\supp (D')$ has exactly $n+g-k$ points. There is a finite morphism $\pi :C\to\Pone$ with $\pi ^*(0)=D'$ and $\pi ^*(\infty )=(g+n-m)(x)$. The part of the ramification divisor $R$ of $\pi$ over $\Pone -\{ 0,\infty \}$ has by Riemann-Hurwitz degree $2g-2+(g+n-k)=3g-2+n-k$. The multiplicity $m$, the multiplicity of $x_i$ in $D$, and the stratum of the diagonal stratification of $\CC ^{n+1}_g$ containing $(C,x_1,\dots ,x_n,x)$ only depend on $W$. So assigning to $(C,x_1,\dots ,x_n,x,P)$ the $\C ^{\times}$-orbit of $\pi _*R$ defines a flat, quasi-finite morphism from $W$ to the quotient of a $(3g-3+n-k)$-dimensional torus by an action of the symmetric group. So $W$ is pure of dimension $3g-3+n-k$. Proposition \refer{2.2} implies that $f^*K_i|W$ is trivial. \enddemo \demo{Proof of the first clause of \refer{1.1}} Let $X^k$ be the union of irreducible components of $Z^k$ that are distinct from $Z^{n+g-1}$. (It can be shown that $Z^{n+g-1}$ is actually an irreducible component of $Z$ and so $X^0\not= Z$.) The restriction $f: X^0\to\CC ^n_g$ is clearly proper. It is also surjective, because for given $(C,x_1,\dots ,x_n)$, the morphism $$ (y,y_1,\dots ,y_{g-1})\in C^g\mapsto [-(n+g)y + 2(x_1)+\sum _{i=2}^n(x_i)+\sum _{j=1}^{g-1}(y_j)]\in J(C) $$ is onto. Observe that $X^{n+g-1}=\emptyset$. We claim that $f(X^k\cap Z^{n+g-1})\subset f(X^{k+1})$. For if $(C,x,\dots ,x,x,(n+g)x,P)$ represents an element of $X^k\cap Z^{n+g-1}$, then by \refer{2.5}, $P$ will be of the form $(k+1)x +P'$ with $P'$ a pencil of degree $n+g-k-1$. So $P$ has a member $\not= (n+g)(x)$ with at most $n+g-k-2$ points. It follows that the pre-image $U^k$ of $f(X^k)-f(X^{k+1})$ in $X^k$ is contained in $Z^k-Z^{k+1}$. In particular, $f^*K_i|U^k=0$ for $i=1,\dots ,n$. Since $f:U ^k\to f(X^k)-f(X^{k+1})$ is proper and onto, we also have $K_i|f(X^k)-f(X^{k+1})$ =0. So by \refer{2.3}, a monomial of degree $k$ in $K_1,\dots ,K_n$ is a linear combination of irreducible components of $f(X^k)$ of codimension $k$. One easily checks that these components are as described in the theorem. \enddemo \label Since $f(X_k)-f(X_{k+1})$ admits a finite covering that is Zariski-open in an affine variety, it cannot contain a complete curve. From this we recover Diaz's theorem which asserts that $\CC _g^n$ does not contain a complete subvariety of dimension $>g+n-2$. In order to complete the proof of \refer{1.1} we need two more results, one algebraic, one topological. \proclaim{\label Lemma} Let $f:\A \to S$ be a family of abelian varieties of dimension $g$ and let $d$ be a positive integer. Then the class of the locus $\A \la d\ra $ of points of order $d$ is a positive multiple of the class of the zero section in $A^g(\A )$. (The coefficient is the number of elements in $(\Z /d)^{2g}$ of order $d$.) \endproclaim \demo{Proof} We use the Fourier transform for abelian varieties introduced by Mukai, developed by Beauville and extended to abelian schemes by Deninger--Murre \cite{2}. Mukai's transform gives an (in general inhomogeneous) isomorphism $\FF : A(\A )\to A(\hat\A )$, where $\hat\A\to S $ is the dual family. We shall compare the images of the two classes in $A(\hat\A )$ under $\FF$. Let $k$ be an positive integer relative prime to $d$. Multiplication by $k$ in $A$ maps $\A \la d\ra $ isomorphically onto itself. So the class of $\A \la d\ra $ in $A^g(\A )$ is fixed under $k_*$. Lemma \refer{2.18} of \cite{2} implies that then $\FF ([\A \la d\ra ])\in A^0(\hat\A )$. Since the projection induces an isomorphism $A^0(S )\to A^0(\hat\A )$, the lemma follows. \enddemo \proclaim{\label Lemma} Let $\pi :C\to\Pone$ be a covering of degree $d$ by a smooth connected curve that is totally ramified over $0$ and $\infty$ such that the part $D$ of the discriminant in $\Pone -\{ 0,\infty\}$ is reduced. Then there exists a disk neighborhood $B$ of $\supp (D)$ in $\Pone -\{0,\infty\}$ such that for $p\in\partial B$, the monodromy group of $\pi$ over $(B-\supp (D),p)$ is a single transposition $(a',a'')$. Moreover, if $\sigma$ is the monodromy of a simple loop in $\Pone -\inw (B)$ around $0$ based at $p$, then $a''=\sigma ^r(a')$ for some divisor $r$ of $d$ and $\pi$ factorizes through the covering $z\in\Pone\to z^r\in\Pone$. \endproclaim \demo{Proof} We choose a base point $p\in\Pone$ outside the discriminant and we put $F:=\pi ^{-1}(p)$. By a {\it simple arc} we shall mean an embedded interval connecting $p$ with a point of the discriminant that does not meet the discriminant along the way. A simple arc $\alpha$ determines up to isotopy (relative $p$ and the discriminant) a simple loop based at $p$ around a point of the discriminant and hence a monodromy transformation $\tau _{\alpha}\in\Aut (F)$. A collection of simple arcs that do not meet outside $p$ shall be called an {\it arc system}. Notice that the directions of departure of the members of such a collection determine a cyclic ordering (our preference is clockwise) of these. We begin by fixing a simple arc $\omega$ connecting $p$ with $0$. We write $\sigma$ for $\tau _{\omega}$; this is a $d$-cycle in $\Aut (F)$. Any transposition $\tau$ in $\Aut (F)$ can be written $(a,\sigma ^l(a))$ for some $l\in\{ 0,1,\dots {1\over 2}d\}$; this means that $\sigma\tau$ is the product of two disjoint cycles of length $l$ and $d-l$. Let us call $l$ the {\it mesh} of $\tau$. Let $\alpha _1$ be an simple arc to a point of $\supp (D)$ that forms with $\omega$ an arc system and is such that $\tau :=\tau _{\alpha _1}$ has minimal mesh $r$. Write $\sigma\tau =\sigma '\sigma ''$ with $\sigma '$ and $\sigma ''$ disjoint cycles of length $r$ resp.\ $d-r$ and denote by $F'$ and $F''$ the corresponding parts of $F$. Notice that $\tau _{\alpha _1}$ interchanges some $a'\in F'$ with some $a''\in F''$. Let $\beta$ be another simple arc to a point of $\supp (D)$ such that $(\omega ,\alpha _1 ,\beta )$ is a clockwise oriented arc system. Then $\tau _{\beta}$ cannot commute with $\sigma ''$: if it did, then it would interchange two points of $F'$ and would therefore have a mesh $<r$. It may happen that $\tau _{\beta}$ commutes with $\sigma '$. But not every choice for $\beta$ can be like this, for then $\sigma '$ would commute with the monodromy around $\infty$ and this is impossible as the latter is a $d$-cycle. So for some $\beta$, $\tau _{\beta}$ interchanges some $b'\in F'$ with some $b''\in F''$. If we modify $\beta$ by letting it first wind $k$ times around the union of $\omega$ and $\alpha _1$, then its monodromy gets conjugated by $(\sigma '\sigma '')^{\pm k}$. In this way we can arrange that $b''=a''$. If $b'\not= a'$, then a straightforward verification shows that $\tau _{\beta}$ would have a smaller mesh than $r$. So $b'=a'$ and hence $\tau _{\beta}=\tau$. This argument proves more: the fact that for every integer $k$ the mesh of the $(\sigma '\sigma '')^k$-conjugate of $\tau _{\beta}$ is $\ge r$ implies that $r$ divides $d$. We put $\alpha _2:=\beta$. We now prove with induction on $l$ that for $l\le \deg (D)$ there is an arc system $(\alpha _1,\alpha _2,\dots ,\alpha _l)$ in clockwise cyclic order such that $\tau _{\alpha _i}=\tau$ for $i=1,\dots ,l$. The lemma then follows: we already showed that $r$ divides $d$, and it is easy to see that the asserted factorization exists. So suppose we found such an arc system $(\alpha _1,\alpha _2,\dots ,\alpha _l)$ for some $l\ge 2$. First assume $l$ even. Then the monodromy around the union of these arcs is equal to $\sigma$ and so the above argument yields simple arcs $\beta _1,\beta _2$ such that $\tau _{\beta _1}=\tau _{\beta _2}$ and $(\omega ,\alpha _1,\dots ,\alpha _l,\beta _1,\beta _2)$ is an arc system in clockwise order. Since $\tau _{\beta _i}$ does not commute with $\sigma ''$, we can modify $\beta _1$ and $\beta _2$ by letting both go round the union of $(\omega ,\alpha _1,\dots ,\alpha _l)$ the same number of times first, to ensure that $\tau _{\beta _1}=\tau _{\beta _2}$ moves $a''$. If $\tau _{\beta _i}$ does not commute with $\sigma '$, then the argument above shows that in fact $\tau _{\beta _i}=\tau$ and so we managed to take two induction steps. If $\tau _{\beta _i}$ does commute with $\sigma '$, then let $\beta '_i$ be obtained from $\beta _i$ by going round $\alpha _l$ first. Then $(\omega ,\alpha _1,\dots ,\alpha _{l-1},\beta '_1,\beta ' _2,\alpha _l)$ is in clockwise order and $\tau _{\beta '_1}=\tau _{\beta '_2}$ interchanges an element of $F'$ with an element of $F''$. Next modify the $\beta '_1$ and $\beta '_2$ by letting them first encircle $(\omega ,\alpha _1,\dots ,\alpha _{l-1})$ the same number of times as to arrange that $\tau _{\beta '_i}$ moves $a''$ (this might cause them to meet $\alpha _l$ in a point $\not= p$). Then $\tau _{\beta '_i}=\tau$ and hence we have made the induction step. It remains to do the induction step for $l$ odd. That is handled in the same way as the case $l=1$. \enddemo \demo{Proof of the second clause of \refer{1.1}} Notice that $X^{n+g-2}$ parametrizes the triples $(C,x,y)$, where $C$ is smooth of genus $g$, $x,y\in C$ are distinct and $d(x)\equiv d(y)$ for some $d\in \{ 2,\dots ,n+g\}$. By our previous discussion this defines a closed subvariety $Y$ of $\CC ^2_g$ of pure codimension $g$. The assertion that is to be proved will follow if we show that the classes of the irreducible components of $Y$ are proportional in $A^g(\CC ^2_g)$. Our first business is therefore to describe these irreducible components. For $d\ge 2$, let $Y_d\subset \CC ^2_g$ be the locus parametrizing triples $(C,x,y)$ for which $(x)-(y)$ has order $d$ in $J(C)$. For such $(C,x,y)$ we have a morphism $\pi :C\to\Pone$ of degree $d$ such that $\pi ^*(0)=d(x)$, $\pi ^*(\infty )=d(y)$ and $\pi$ does not factor through a cover $z\in\Pone\mapsto z^r\in\Pone$ for some $r>1$. The previous lemma shows that all such covers are of the same topological type. This implies that $Y_d$ is irreducible. So every irreducible component of $Y$ is equal to some $Y_d$. Let $\JJ_g\to\M _g$ be the universal Jacobian and let $q:\CC _g^2\to\JJ _g$ be the Abel-Jacobi map $(C,x,y)\mapsto (x)-(y)\in J(C)$. Then $Y_d=q^{-1}\JJ _g\la d\ra $. Since $Y_d$ has the correct codimension $g$ in $\CC _g^2$, it follows that $[Y_d]$ is a positive multiple of $q^*[\JJ _g\la d\ra ]$. According to \refer{2.9}, $[\JJ _g\la d\ra ]$ is a positive multiple of the class of the zero section in $A^g(\JJ _g )$ and so the proof is complete. \enddemo \demo{Proof of \refer{1.2}} First observe that the direct image of $K_1^{1+d_1}\cdots K_n^{1+d_n}$ under the forgetful morphism $\CC _g^n\to\M _g$ equals $\kappa _{d_1}\cdots\kappa _{d_n}$. Now the direct image of the class of an irreducible component of $X^k$ of codimension $k$ under $\CC _g^n\to\M _g$ is zero unless the image has the correct codimension $k-n$. In particular, a nonzero image requires $k\ge n$. It follows that any product in the tautological classes of degree $d$ can be represented by a linear combination of the irreducible components of the locus in $\M _g$ that parametrizes the curves $C$ that admit a covering $\pi :C\to \Pone$ of degree $\le 2g-2$ totally ramified over $\infty$ and with at most $g-1-d$ points over $0$. The rest follows immediately from \refer{1.1}. \enddemo \Refs \ref\no 1 \paper Weierstrass points and moduli of curves \by E.\ Arbarello \jour Compositio Math. \vol 29 \yr 1974 \pages 325--342 \endref \ref\no 2 \paper Motivic decomposition of abelian schemes and the Fourier transform \by C.\ Deninger \&\ J.\ Murre \jour J. reine angew. Math. \vol 422 \yr 1991 \pages 201--219 \endref \ref\no 3 \paper A bound on the dimensions of complete subvarieties of $\M _g$ \by S.\ Diaz \jour Duke Math.\ J. \vol 51 \yr 1984 \pages 405--408 \endref \ref\no 4 \by D. Mumford \paper Towards an enumerative geometry of the moduli space of curves \inbook Arithmetic and Geometry.~{\rm II} \eds M. Artin and J. Tate \publ Birkha\"user Verlag \publaddr Boston--Basel--Berlin \pages 271--328 \yr 1983 \endref \endRefs \enddocument
1995-10-19T05:20:10
9411
alg-geom/9411006
en
https://arxiv.org/abs/alg-geom/9411006
[ "alg-geom", "math.AG" ]
alg-geom/9411006
Fumiharu Kato
Fumiharu Kato
Logarithmic Embeddings and Logarithmic Semistable Reductions
LaTeX
null
null
null
null
In this paper, we give a criterion for the existence of logarithmic embeddings -- which was first introduced by Steenbrink -- for general normal crossing varieties. Using this criterion, we also give a new proof of the theorem of Kawamata--Namikawa which states a criterion for the existence of the log structures of semistable type.
[ { "version": "v1", "created": "Fri, 11 Nov 1994 05:04:56 GMT" }, { "version": "v2", "created": "Wed, 18 Oct 1995 16:34:51 GMT" } ]
2008-02-03T00:00:00
[ [ "Kato", "Fumiharu", "" ] ]
alg-geom
\section{Introduction} Let $X$ be a connected, geometrically reduced algebraic scheme over a field $k$. Then $X$ is said to be a {\it normal crossing variety} of dimension $n-1$ if there exists an isomorphism of $k$-algebras $$ \widehat{\O}_{X,x}\stackrel{\sim}{\longrightarrow} k(x)[[T_1,\ldots,T_n]]/(T_1\cdots T_{l_x}) $$ for each closed point $x\in X$, where $\widehat{\O}_{X,x}$ denotes the completion of the local ring $\O_{X,x}$ along its maximal ideal (Definition \ref{ncvdef}). Normal crossing varieties usually appear in contexts of algebraic geometry via degenerations and normal crossing divisors. In the first case, they appear as a specialization of a family of smooth varieties. Normal crossing varieties are usually considered and expected to be limits of smooth varieties, and --- as is well--known --- they are important to the theory of moduli. As for the second situation, a {\it normal crossing divisor} is a divisor of a smooth variety which itself is a normal crossing variety. Normal crossing divisors play important roles in various fields of algebraic geometry. For example, a pair of smooth variety and its normal crossing divisor is usually called a log variety. Considering log varieties instead of smooth varieties --- or usually admitting some mild singularities --- alone, some algebro geometric theories ({\it e.g.}, minimal model theory, etc.) are well generalized. Relating with a normal crossing variety $X$, there are two problems, {\it smoothings} and {\it embeddings}, in light of degenerations and normal crossing divisors, respectively. The {\it smoothing problem} is a problem to find a Cartesian diagram $$ \begin{array}{ccc} X&\longrightarrow&\hbox{\maxid X}\\ \vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\\ 0&\longrightarrow&\Delta\rlap{,} \end{array} $$ for a normal crossing variety $X$ (in this situation, we should assume that $X$ is proper over $k$), where $\Delta$ is a one-dimensional regular scheme, $\hbox{\maxid X}$ is a regular scheme proper flat and generically smooth over $\Delta$, and $0$ is a closed point of $\Delta$ whose residue field is $k$. We usually take, as the base scheme $\Delta$, the spectrum of a discrete valuation ring, {\it e.g.}, the ring of formal power series over $k$ or --- in case $k$ is perfect --- the ring of Witt vectors over $k$. In the complex analytic situation, Friedmann \cite{Fri1} studied the smoothing problem generally, and solve it for degenerated K3 surfaces. Recently, Kawamata--Namikawa \cite{K-N1} approached this problem by introducing a new method; the {\it logarithmic} method. The Cartesian diagram as above with $\Delta$ a spectrum of an Artinian local ring $A$ is called an {\it infinitesimal smoothing}, if it is \'{e}tale locally isomorphic to the diagram $$ \begin{array}{ccc} \mathop{\mbox{\rm Spec}}\nolimits k[Z_1,\ldots,Z_n]/(Z_1\cdots Z_l)&\longrightarrow& \mathop{\mbox{\rm Spec}}\nolimits A[Z_1,\ldots,Z_n]/(Z_1\cdots Z_l-\pi)\\ \vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\\ \mathop{\mbox{\rm Spec}}\nolimits k&\longrightarrow&\mathop{\mbox{\rm Spec}}\nolimits A\rlap{,} \end{array} $$ where $\pi$ is an element of the maximal ideal of $A$ and $\pi\neq 0$. The central problem to find such an infinitesimal smoothing is to compute the obstruction class of $X$ to have such a diagram and to show vanishing or non--vanishing of it. The {\it embedding problem} is a problem to find a closed embedding $X\hookrightarrow V$ over $k$ of $X$ as a normal crossing divisor, where $V$ is a smooth variety over $k$. If $X$ is smoothable in the above sense with $\Delta$ a smooth algebraic variety over $\mathop{\mbox{\rm Spec}}\nolimits k$, the smoothing family $X\hookrightarrow\hbox{\maxid X}$ gives an embedding of this sense. If $X$ is smooth, this problem becomes trivial, since we can take as $V$ the product of $X$ and, for example, ${\mbox{\bf P}}^1$. But for a general normal crossing variety, this problem seems far from satisfactory solutions. Similarly to the smoothing problem, we can consider this problem in the infinitesimal sense. In this paper, we consider the above problems in a {\it logarithmic} sense. We consider logarithmic generalizations of smoothings and embeddings of normal crossing varieties according to Kajiwara \cite{Kaj1}, Kawamata--Namikawa \cite{K-N1} and Steenbrink \cite{Ste1}, and we solve their existence problems. These generalizations are done in terms of logarithmic geometry of Fontaine, Illusie and Kazuya Kato. {\it Logarithmic geometry} --- or {\it log geometry} --- was first founded by Fontaine and Illusie based on their idea of, so--called, {\it log structures}; afterwards, it was established as a generally organized theory and applied to various fields of algebraic and arithmetic geometry by Fontaine, Illusie and Kazuya Kato (cf. \cite{Kat1}, \cite{Kat4}). In various kinds of geometries including algebraic geometry, we usually consider local ringed spaces, {\it i.e.}, the pairs of topological spaces --- possibly in the sense of Grothendieck topologies --- and sheaves of local rings over them. The basic idea of Fontaine and Illusie is that, instead of local ringed spaces alone, they consider local ringed spaces equipped with some additional structure --- which they call the logarithmic structures --- written in terms of sheaves of commutative and unitary monoids (see \cite{Kat1} for the precise definition). In algebro geometric situations, these log structures usually represent ``something'' of the underlying local ringed spaces, {\it e.g.}, divisors or the structure of torus embeddings, etc. Through these foundations, they suggested to generalize the ``classical'' geometries by considering ``log objects'' --- such as {\it log schemes} --- which are the pairs of local ringed spaces and log structures on them. In the present paper, we recall and generalize the {\it logarithmic embedding} (Definition \ref{logembdef}) introduced by Steenbrink \cite{Ste1}. A logarithmic embedding --- which is regarded as a logarithmic generalization of a log variety --- is a certain log scheme $(X,{\cal M}_X)$ with $X$ a normal crossing variety. Then we prove the following theorem which gives a criterion for the existence of logarithmic embeddings: \vspace{3mm}\noindent {\bf Theorem}\ ({\it Theorem \ref{mainthm}})\ {\it For a normal crossing variety $X$, a logarithmic embedding of $X$ exists if and only if there exists a line bundle ${\cal L}$ on $X$ such that ${\cal L}\otimes_{\O_X}\O_D\stackrel{\sim}{\rightarrow} {\cal T}^1_X$, where $D$ is the singular locus of $X$.} \vspace{3mm} Here, ${\cal T}^1_X$ is an invertible $\O_D$-module, called the {\it infinitesimal normal bundle} (cf. \cite{Fri1}), which is naturally isomorphic to ${\cal E}xt^1_{\O_X}(\Omega^1_X,\O_X)$; we recall the construction of it in \S 3. A normal crossing variety $X$ is said to be {\it $d$-semistable} if ${\cal T}^1_X$ is a trivial bundle on $D$ (cf. \cite{Fri1}). By the above theorem, any $d$-semistable normal crossing variety $X$ has a logarithmic embedding. As for the smoothing problem, we recall and generalize the concept, the {\it logarithmic semistable reduction} (Definition \ref{logsemidef}) introduced by Kajiwara \cite{Kaj1} (in one dimensional case) and Kawamata--Namikawa \cite{K-N1} (by a different but essentially the same method). Using the above theorem, we get a criterion for the existence of logarithmic semistable reductions, which was first proved by Kawamata--Namikawa \cite{K-N1} in the complex analytic situation, as follows: \vspace{3mm}\noindent {\bf Theorem}\ ({\it Theorem \ref{mainthm2}}) {\rm (cf. \cite{K-N1})}\ {\it For a normal crossing variety $X$, the log structure of semistable type on $X$ exists if and only if $X$ is $d$-semistable.} \vspace{3mm} The composition of this paper is as follows. In \S 2, we study the geometry of normal crossing varieties in general. In particular, we define good \'{e}tale local charts on normal crossing varieties, and prove the existence of them. In \S 3, we recall the basic construction of the tangent complex of a normal crossing variety, and introduce the invertible sheaf ${\cal T}^1_X$ on $D$. We introduce the logarithmic embedding in \S 4. This section also contains the proof of our main theorem. The logarithmic semistable reduction is studied in \S 5. The author thanks T. Fujisawa for useful communications. The author is also grateful to Professors K. Ueno, S. Usui and T. Yusa for their helpful comments. {\sc Conventions}:\ All sheaves are considered with respect to \'{e}tale topology. By a monoid, we mean --- as usual in the contexts of log geometry --- a set with a commutative and associative binary operation and the neutral element. For such a monoid $M$, we denote by $\gp{M}$ the Grothendieck group of $M$. We denote by $\mbox{\bf N}$ the monoid of non--negative integers. \section{Normal crossing varieties} Throughout this paper, we always work over a fixed base field $k$. As usual, an algebraic $k$-scheme is, by definition, a seperated scheme of finite type over $k$. Let $X$ be an algebraic $k$-scheme and $x\in X$ a point. We denote the residue field at $x\in X$ by $k(x)$. \begin{dfn}\label{ncvdef} Let $X$ be a connected and geometrically reduced algebraic $k$-scheme. Then $X$ is said to be a {\it normal crossing variety} over $k$ of dimension $n-1$ if the following condition is satisfied: For any closed point $x\in X$, there exists an isomorphism \begin{equation}\label{ncvdefloc} \widehat{\O}_{X,x}\stackrel{\sim}{\longrightarrow} k(x)[[T_1,\ldots,T_n]]/(T_1\cdots T_{l_x}) \end{equation} of $k$-algebras, where $l_x$ is an integer $(1\leq l_x\leq n)$ depending on $x$. Here, we denote by $\widehat{\O}_{X,x}$ the completion of the local ring $\O_{X,x}$ by its maximal ideal. \end{dfn} The integer $l_x$ is called the {\it multiplicity} at $x\in X$. We sometimes denote it by $l^X_{x}$ if we want to emphasize the scheme $X$. The Zariski closure of the set of closed points whose multiplicity is greater than 1 is the singular locus of $X$, which we denote by $D$. A standard example of normal crossing varieties is an affine scheme \begin{equation}\label{ncvstandard} \mathop{\mbox{\rm Spec}}\nolimits k[T_1,\ldots,T_n]/(T_1\cdots T_l)\ \ (1\leq l\leq n). \end{equation} This scheme consists of $l$ irreducible components which intersect transversally along the singular locus \begin{equation} \mathop{\mbox{\rm Spec}}\nolimits k[T_1,\ldots,T_n]/(T_1\cdots\widehat{T_j}\cdots T_l\ :\ 1\leq j\leq l). \end{equation} Each irreducible component is isomorphic to the affine $(n-1)$-space over $k$. In general, a normal crossing variety $X$ is said to be {\it simple} if each irreducible component of $X$ is smooth over $k$. For example, a smooth $k$-variety is a simple normal crossing variety. Let $V$ be a smooth $k$-variety of dimension $n$. A reduced divisor $X$ on $V$ is called a {\it normal crossing divisor} if $X$ itself is a normal crossing variety of dimension $n-1$. In this case, the closed embedding $X\hookrightarrow V$ is called a {\it NCD embedding} of $X$. For example, the affine normal crossing variety (\ref{ncvstandard}) is a normal crossing divisor in the affine $n$-space over $k$. The proof of the following proposition is straightforward and is left to the reader. \begin{pro}\label{ncvetale} Let $Y$ be a connected scheme \'{e}tale over a connected algebraic $k$-scheme $X$. If $X$ is a normal crossing variety, then so is $Y$. The converse is also true if the \'{e}tale morphism $Y\rightarrow X$ is surjective. \end{pro} It is clear that an \'{e}tale morphism leaves invariant the multiplicity at every closed point, {\it i.e.}, if $\varphi\colon Y\rightarrow X$ is an \'{e}tale morphism of normal crossing varieties and $y\in Y$ is a closed point, then we have $l^Y_y=l^X_{\varphi(y)}$. In the following paragraphs of this section, we shall study the local nature of normal crossing varieties for the later purpose. In the subsequent sections, we need to take a good \'{e}tale neighborhood around every closed point. We require that these \'{e}tale neighborhoods have good coordinate systems which serve for several explicit calculations. To clarify the notion of ``good'' \'{e}tale neighborhoods, we define them as follows: \begin{dfn}\label{ncvchart}{\rm Let $X$ be a normal crossing variety and $x\in X$ a closed point. Let $\varphi\colon U\rightarrow X$ be an \'{e}tale morphism with $U$ a simple normal crossing variety and $z_1,\ldots,z_{l_x}\in\Gamma(U,\O_U)$, where $l_x$ is the multiplicity at $x$. Then $(\varphi\colon U\rightarrow X;\ z_1,\ldots,z_{l_x})$ is said to be a {\it local chart} around $x$ if the following conditions are satisfied: \begin{description} \item[{\rm (a)}] There exists a unique point $y\in U$ such that $\varphi(y)=x$. \item[{\rm (b)}] There exists a closed immersion $\iota\colon U\hookrightarrow V$, where $V$ is an affine smooth $k$-scheme. \item[{\rm (c)}] There exist $Z_1,\ldots,Z_n\in\Gamma(V,\O_V)$ which form a regular parameter system at $\iota(y)\in V$ such that $z_i=\iota^*Z_i$ for $1\leq i\leq l_x$, and $U$ is defined as a closed subset in $V$ by the ideal $(Z_1\cdots Z_{l_x})$. \item[{\rm (d)}] each ideal $(z_i)$ is prime and the irreducible components of $U$ are precisely the closed subsets of $U$ corresponding to the ideals $(z_1),\ldots,(z_{l_x})$. \end{description} } \end{dfn} Note that $\iota\colon U\hookrightarrow V$ is, due to (c), a NCD embedding. Moreover, due to (d), all the irreducible components intersect and contain the point $y$. The following theorem assures the existence of local chart around every closed point of normal crossing variety $X$. We prove this theorem later in this section. \begin{thm}\label{chartexist} Let $X$ be a normal crossing variety and $x\in X$ a closed point. Then there exists a local chart $(\varphi\colon U\rightarrow X;\ z_1,\ldots,z_{l_x})$ around $x$. \end{thm} Since any \'{e}tale open set of $X$ is again a normal crossing variety, we have the following: \begin{cor}\label{specialcov} Let $X$ be a normal crossing variety. Then the set of all local charts forms an open basis with respect to the \'{e}tale topology on $X$. \end{cor} \begin{rem}\label{ncvemb}{\rm Theorem \ref{chartexist} implies that any normal crossing variety is realized as a simple normal crossing divisor on some smooth $k$-variety \'{e}tale locally. But a normal crossing variety, in general, cannot be a normal crossing divisor globally on a smooth $k$-variety. In the next section, we will see a necessary condition for a normal crossing variety to be a normal crossing divisor (Proposition \ref{suffcondemb}). } \end{rem} For the proof of Theorem \ref{chartexist}, we need one lemma: \begin{lem}\label{htzero} Let $\hbox{\maxid q}$ be a height zero prime ideal in $K[T_1,\ldots,T_n]/(T_1\cdots T_l)$ $(1\leq l\leq n)$, where $K$ is a field. Then $\hbox{\maxid q}=(T_j)$ for some $j$ $(1\leq j\leq l)$. \end{lem} {\sc Proof.}\hspace{2mm} By Krull's principal ideal theorem, any non--zero element in $\hbox{\maxid q}$ is a zero factor. Hence any element in $\hbox{\maxid q}$ is a multiple of $T_j$'s $(1\leq j\leq l)$. Since $\hbox{\maxid q}$ is a prime ideal, $\hbox{\maxid q}$ must contain $T_j$ for some $j$ $(1\leq j\leq l)$, {\it i.e.}, $(T_j)\subseteq\hbox{\maxid q}$. But since the height of $\hbox{\maxid q}$ is zero and $(T_j)$ is a prime ideal, we have $\hbox{\maxid q}=(T_j)$. $\Box$ \vspace{3mm} {\sc Proof of Theorem \ref{chartexist}.}\hspace{2mm} The complete local ring $\widehat{\O}_{X,x}$ is isomorphic to the complete local ring $k(x)[[T_1,\ldots,T_n]]/(T_1\cdots T_{l_x})$ which is a completion of the local ring $(k(x)[T_1,\ldots,T_n]/(T_1\cdots T_{l_x}))_{0}$. Then due to \cite[Corollary (2.6)]{Art1}, there exist a scheme $U$ and \'{e}tale morphisms $\varphi\colon U\rightarrow X$ and $\phi\colon U\rightarrow\mathop{\mbox{\rm Spec}}\nolimits k(x)[T_1,\ldots,T_n]/(T_1\cdots T_{l_x})$ such that $\varphi(y)=x$ and $\phi(y)=0$ for some $y\in U$. We fix this closed point $y\in U$. Since $\varphi$ is \'{e}tale, we may assume --- replacing $U$ by its Zariski open subset if necessary --- that $y$ is the only point which is mapped to $x$ by $\varphi$. Obviously we may assume that $U$ is connected and affine. We can remove all the irreducible components which do not contain $y$. Then we may assume that all the irreducible components of $U$ contain $y$. We set $U=\mathop{\mbox{\rm Spec}}\nolimits A$ and $B\colon =k(x)[T_1,\ldots,T_n]/(T_1\cdots T_{l_x})$. Since $U$ is \'{e}tale over a reduced $k$-scheme $\mathop{\mbox{\rm Spec}}\nolimits B$, the $k$-algebra $A$ is reduced. Take a minimal prime factorization \begin{equation}\label{prmfac} (0)=\hbox{\maxid p}_1\cap\cdots\cap\hbox{\maxid p}_{l}. \end{equation} of the ideal $(0)=\sqrt{(0)}$. Since each $\hbox{\maxid p}_i$ is minimal in the set of all prime ideals, the height of each $\hbox{\maxid p}_i$ is zero. Obviously the prime decomposition (\ref{prmfac}) precisely corresponds to the decomposition of $U$ into irreducible components. Set $\hbox{\maxid q}_i\colon =\phi(\hbox{\maxid p}_i)$ which is a prime ideal of height zero in $B$ for $1\leq i\leq l$. Due to Lemma \ref{htzero}, we have $\hbox{\maxid q}_i=(T_{j_i})$ for some $j_i$ $(1\leq j_i\leq l_x)$, {\it i.e.}, any generic point of a irreducible component of $U$ is mapped by $\phi$ to a generic point of a irreducible component of $\mathop{\mbox{\rm Spec}}\nolimits B$. Let us suppose that the map $i\mapsto j_i$ is not injective, {\it i.e.}, there exist $i$ and $j$ $(i\neq j)$ such that $\hbox{\maxid q}_i=\hbox{\maxid q}_j$. Consider the Cartesian diagram $$ \begin{array}{ccc} \overline{\{\hbox{\maxid q}_i\}}\times_{\mathop{\mbox{\rm Spec}}\nolimits B}U&\lhook\joinrel\longrightarrow&U\\ \llap{$\phi_i$}\vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\rlap{$\phi$}\\ \overline{\{\hbox{\maxid q}_i\}}&\lhook\joinrel\longrightarrow&\mathop{\mbox{\rm Spec}}\nolimits B\rlap{,} \end{array} $$ where the horizontal arrows are closed immersions and the vertical ones are \'{e}tale. The scheme $U_i\colon =\overline{\{\hbox{\maxid q}_i\}}\times_{\mathop{\mbox{\rm Spec}}\nolimits B}U$ is also a normal crossing variety. Since $\overline{\{\hbox{\maxid p}_i\}}\cap\overline{\{\hbox{\maxid p}_j\}}$ is a closed subscheme (which contains $y$) of $U_i$, the multiplicity $l^{U_i}_y$ at $y$ in $U_i$ is greater than 1. But since the irreducible component $\overline{\{\hbox{\maxid q}_i\}}$ is smooth, we have $l_{\phi_i(y)}=1$. This is a contradiction since $l^{U_i}_y=l_{\phi_i(y)}$. Thus, the map $i\mapsto j_i$ is injective, {\it i.e.}, there is at most one component over each component of $\mathop{\mbox{\rm Spec}}\nolimits B$. Moreover, in this case, we have $U_i=\overline{\{\hbox{\maxid p}_i\}}$. Then the irreducible component $U_i$ is \'{e}tale over a smooth scheme $\overline{\{\hbox{\maxid q}_i\}}$, and hence the normal crossing variety $U$ is simple. Moreover, the prime ideal $\hbox{\maxid p}_i$ is a principal ideal $(z_i)$, where $z_i\colon =\phi^*T_{j_i}$ for $1\leq i\leq l$, since $\overline{\{\hbox{\maxid p}_i\}}=\overline{\{\hbox{\maxid q}_i\}}\times_{\mathop{\mbox{\rm Spec}}\nolimits B}U$ implies that $\hbox{\maxid p}_i=\hbox{\maxid q}_i\otimes_BA$. Since the map $i\mapsto j_i$ is injective, we have $l\leq l_x$. Note that the multiplicity $l_y$ at y in $U$ equals to $l_x$. Since the simple normal crossing variety $U$ consists of $l$ irreducible components, we have $l_y=l_x\leq l$. Hence we have $l_x=l$. The scheme $\mathop{\mbox{\rm Spec}}\nolimits B$ is a normal crossing divisor in the $n$ dimensional affine space over $k(x)$. Hence, due to \cite[Expos\'{e} 1. Proposition 8.1]{Gro1}, any point in $U$ has a Zariski open neighborhood which is embedded in a smooth $k(x)$-variety as a normal crossing divisor. This implies that, replacing $U$ by its Zariski open neighborhood of $y$, we may assume that $U$ can be embedded in an affine smooth $k(x)$-scheme $V=\mathop{\mbox{\rm Spec}}\nolimits R$ of dimension $n$ as a normal crossing divisor. Let $\iota\colon U\hookrightarrow V$ be the closed immersion. Finally, consider the Cartesian diagram $$ \begin{array}{ccc} U&\stackrel{\iota}{\lhook\joinrel\longrightarrow}&V\\ \llap{$\phi$}\vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\rlap{$\Phi$}\\ \mathop{\mbox{\rm Spec}}\nolimits B&\lhook\joinrel\longrightarrow&\mathop{\mbox{\rm Spec}}\nolimits k[T_1,\ldots,T_n]\rlap{,} \end{array} $$ where $\Phi$ is an \'{e}tale morphism. Set $Z_i\colon =\Phi^*T_i\in\Gamma(V,\O_V)$ for $1\leq i\leq n$. Then $Z_1,\ldots,Z_n$ form a regular parameter system at $\iota(y)\in V$. We also have $z_i=\iota^* Z_i$ for $1\leq i\leq l_x$. It is clear that the closed subscheme $U$ in $V$ is defined by an ideal $(Z_1\cdots Z_{l_x})$. Then the proof of the theorem is completed. $\Box$ The following lemma will be needed in the later arguments. \begin{lem}\label{inj-zero} Let $(\varphi'\colon U'\rightarrow X;z'_1,\ldots,z'_{l'})$ be a local chart on $X$ around some closed point and $(\psi\colon U\rightarrow U';z_1,\ldots,z_{l})$ a local chart on $U'$ around some closed point. Then $\psi\colon U\rightarrow U'$ is injective in codimension zero, {\it i.e.}, it maps the generic points of irreducible components on $U$ injectively to those of $U'$. \end{lem} {\sc Proof.}\hspace{2mm} Let $\eta\in U'$ be a codimension zero point. Since $U'$ is simple, $\overline{\{\eta\}}$ is regular and so is $\overline{\{\eta\}}\times_{U'}U$ whenever it is not empty. Then each connected component of $\overline{\{\eta\}}\times_{U'}U$ is irreducible and its generic point is of codimension zero. Hence each connected component of $\overline{\{\eta\}}\times_{U'}U$ is an irreducible component of $U$. Since any two of irreducible components of $U$ intersect, $\overline{\{\eta\}}\times_{U'}U$ itself is an irreducible component of $U$. Hence, if $\xi\in U$ is a codimension zero point such that $\psi(\xi)=\eta$, we have $\overline{\{\xi\}}=\overline{\{\eta\}}\times_{U'}U$. In particular, there exists at most one such $\xi$. $\Box$ For a normal crossing variety $X$, the {\it normalization} $\nu\colon\widetilde{X}\rightarrow X$ of $X$ is defined as usual: The scheme $\widetilde{X}$ is defined by the disjont union of the normalizations of irreducible components of $X$ and $\nu\colon\widetilde{X} \rightarrow X$ is the natural morphism. The normalization $\widetilde{X}$ is a smooth $k$-scheme due to Theorem \ref{chartexist} and the following lemma. \begin{lem}\label{normalization} Let $U\rightarrow Z$ be a \'{e}tale morphism of $k$-varieties. Let $\widetilde{U}\rightarrow U$ and $\widetilde{Z}\rightarrow Z$ be normalizations of $U$ and $Z$, respectively. Then there exists a natural isomorphism $\widetilde{U}\stackrel{\sim}{\rightarrow}U\times_{Z}\widetilde{Z}$. In particular, the natural morphism $\widetilde{U}\rightarrow\widetilde{Z}$ is \'{e}tale. \end{lem} {\sc Proof.}\hspace{2mm} Since $U\times_{Z}\widetilde{Z}\rightarrow\widetilde{Z}$ is \'{e}tale and $\widetilde{Z}$ is normal, $U\times_{Z}\widetilde{Z}$ is a normal variety. Hence there exists a unique morphism $\phi\colon U\times_{Z}\widetilde{Z}\rightarrow \widetilde{U}$ which factors the morphism $U\times_{Z}\widetilde{Z}\rightarrow U$. Moreover $\phi$ also factors the morphism $U\times_{Z}\widetilde{Z} \rightarrow\widetilde{Z}$ since the last morphism is the unique morphism determined by the morphism $U\times_{Z}\widetilde{Z}\rightarrow Z$. Hence the natural morphism $\varphi\colon \widetilde{U}\rightarrow U\times_{Z}\widetilde{Z}$ is the inverse morphism of $\phi$. $\Box$ For a local chart $(\varphi\colon U\rightarrow X; z_1,\ldots,z_l)$, the normalization of $U$ is given by the disjoint union of all irreducible components and the natural morphism, {\it i.e.}, $$ \nu_U\colon\widetilde{U}=\coprod^l_{i=1}U_i\longrightarrow U, $$ where $U_i$ is the irreducible component of $U$ corresponding to the ideal $(z_i)$. Set $\overline{D}\colon =D\times_X\widetilde{X}$, which is a divisor of $\widetilde{X}$. \begin{lem} $\overline{D}$ is a normal crossing divisor of $\widetilde{X}$. \end{lem} {\sc Proof.}\hspace{2mm} Let $(\varphi\colon U\rightarrow X; z_1,\ldots,z_l)$ be a local chart on $X$. Then $D_U\colon =D\times_{X}U$ is nothing but the singular locus of $U$ and is \'{e}tale over $D$. Consider the normalization $\nu_U\colon \widetilde{U}\rightarrow U$ as above. Set $$ \overline{D_U}\colon =D_U\times_{U}\widetilde{U}. $$ Clearly, $\overline{D_U}$ is a normal crossing divisor of $\widetilde{U}$ defined by an ideal $(z_1\cdots\widehat{z_i}\cdots z_l)$ on $U_i$. There exists a natural morphism $\overline{D_U}\rightarrow\overline{D}$. Since one can easily see that there exists a natural isomorphism $$ \overline{D_U}\stackrel{\sim}{\longrightarrow} \overline{D}\times_{\widetilde{X}}\widetilde{U} $$ and the morphism $\widetilde{U}\rightarrow\widetilde{X}$ is \'{e}tale due to Lemma \ref{normalization}, the morphism $\overline{D_U}\rightarrow \overline{D}$ is \'{e}tale. Then, considering all the local charts on $X$, $\overline{D}$ is a normal crossing divisor on $\widetilde{X}$ due to Proposition \ref{ncvetale}. $\Box$ \section{Tangent complex on a normal crossing variety} In this section, we recall the tangent complex and the infinitesimal normal bundle ${\cal T}^1_X$ of a normal crossing variety $X$ which will play important roles in the subsequent sections. Let $X$ be a normal crossing variety over a field $k$. For a local chart $(\varphi\colon U=\mathop{\mbox{\rm Spec}}\nolimits A\rightarrow X; z_1,\ldots,z_l)$ of $X$ around some closed point, we use the folowing notation in this and subsequent sections: Let $V=\mathop{\mbox{\rm Spec}}\nolimits R$ and $Z_1,\ldots,Z_l$ be as in Definition \ref{ncvchart}. Set $I_j\colon=(Z_j)$ and $J_j\colon=(Z_1\cdots\widehat{Z_j}\cdots Z_l)$ for $1\leq j\leq l$. (If $l=1$, we set $J_1=R$ for the convention.) Then $A=R/I$ where $I\colon=I_1\cdots I_l$. Moreover, the ideal $I_j/I\subset A$ is generated by $z_j=(Z_j\,\mbox{\rm mod}\,I)$ and is prime of height zero. Set $J\colon=J_1+\cdots+J_l$. Then the singular locus $D_U\colon=D\times_XU$ of $U$ is the closed subscheme defined by $J$. We set $Q\colon=R/J$. Note that, for $1\leq j\leq l$, $I_j/II_j$ is a free $A$-module of rank one and is generated by $\zeta_j\colon=(Z_j\,\mbox{\rm mod}\,II_j)$. There exists a natural isomorphism $I_j/II_j\otimes_AQ\stackrel{\sim} {\rightarrow}I_j/JI_j$ of $Q$-modules which maps $\zeta_j\otimes 1$ to $\xi_j\colon=(Z_j\,\mbox{\rm mod}\,JI_j)$. Moreover, there exists a natural isomorphism \begin{equation}\label{conormal} I/I^2\stackrel{\sim}{\rightarrow} I_1/II_1\otimes_A\cdots\otimes_AI_l/II_l \end{equation} of $A$-modules, and hence, the $A$-module $I/I^2$ is free of rank one and is generated by $\zeta_1\otimes\cdots\otimes\zeta_l$. We denote by $\pi_j$ the natural projection $I_j/II_j\rightarrow I_j/I\subset A$. The cotangent complex of the morphism $k\rightarrow A$ is given by $$ L^{\cdot}:0\longrightarrow R\otimes_{R}A\stackrel{\delta}{\longrightarrow} \Omega^1_{R/k}\otimes_{R}A\longrightarrow 0, $$ where $\delta$ is defined by $R\rightarrow F\cdot R \stackrel{d}{\rightarrow}\Omega^1_{R/k}$ with $F\colon=Z_1\cdots Z_l$ (cf. \cite{L-S1}). Then the tangent complex of $U$ is the complex $$ \mathop{\mbox{\rm Hom}}\nolimits_A(L^{\cdot},A):0\longrightarrow\Theta_{R/k}\otimes_{R}A \stackrel{\delta^*}{\longrightarrow}\mathop{\mbox{\rm Hom}}\nolimits_A(R\otimes_{R}A,A) \longrightarrow 0, $$ where $\Theta_{R/k}\colon =\mathop{\mbox{\rm Hom}}\nolimits_R(\Omega^1_{R/k},R)$. We define \begin{equation}\label{tangentdef} T^1_A=\mathop{\mbox{\rm Hom}}\nolimits_A(R\otimes_{R}A,A)/\delta^*(\Theta_{R/k}\otimes_{R}A). \end{equation} \begin{lem}\label{tangentloc} We have the natural isomorphism \begin{equation}\label{tangentloc1} T^1_A\stackrel{\sim}{\rightarrow}\mathop{\mbox{\rm Hom}}\nolimits_A(I/I^2,A)\otimes_{A}Q. \end{equation} \end{lem} {\sc Proof.}\hspace{2mm} Consider the exact sequence $$ 0\longrightarrow I/I^2\longrightarrow\Omega^1_{R/k}\otimes_{R}A \longrightarrow\Omega^1_{A/k}\longrightarrow 0. $$ By definition, we have $T^1_A=\mathop{\mbox{\rm Coker}}\nolimits(\mathop{\mbox{\rm Hom}}\nolimits_A(\Omega^1_{R/k}\otimes_{R}A,A)\rightarrow \mathop{\mbox{\rm Hom}}\nolimits_A(I/I^2,A))$. Then one can show --- by direct calculations --- that $\mathop{\mbox{\rm Hom}}\nolimits_A(I/I^2,A)\rightarrow T^1_A$ is nothing but the ``tensoring'' morphism $\otimes_AQ$. Moreover, we have $T^1_A\stackrel{\sim}{\rightarrow}\mathop{\mbox{\rm Ext}}\nolimits^1_A(\Omega^1_{A/k},A)$. $\Box$ Considering all the local charts $U$ on $X$, these modules $T^1_A$ glue to an invertible $\O_D$-module on $X$, which is denoted by ${\cal T}^1_X$; this is well--known (cf. \cite{L-S1}) but, for the later purpose, we prove it in the following. Suppose we have two local charts $(\varphi\colon U\rightarrow X; z_1,\ldots,z_{l})$ and $(\varphi'\colon U'\rightarrow X; z'_1,\ldots,z'_{l'})$ and an \'{e}tale morphism $\psi\colon U\rightarrow U'$ such that $\varphi=\varphi'\circ\psi$. (Because we are interested in the singular locus, we shall assume $l>1$ and $l'>1$.) For these local charts, we use all the notation as above. (For $U'$, we denote them by $A'$, $I'$, $J'$, $\zeta'_j$, etc.) Let $f\colon A'\rightarrow A$ be the ring homomorphism corresponding to $\psi$. We shall show that the morphism $\psi$ induces naturally an isomorphism $T^1_{A'}\otimes_{Q'}Q\stackrel{\sim}{\rightarrow}T^1_A$ of $Q$-modules. Let $U_j$ (resp. $U'_j$) be the irreducible component of $U$ (resp. $U'$) corresponding to $I_j/I$ (resp. $I'_j/I'$) for $1\leq j\leq l$ (resp. $1\leq j\leq l'$). Since $\psi$ is \'{e}tale and injective in codimension zero (Lemma \ref{inj-zero}), we may assume that the generic point of $U_j$ is mapped to that of $U'_j$ by $\psi$ for $1\leq j\leq l$. In particular, we have $l\leq l'$. Then one sees easily that $U\times_{U'}U'_j\cong U_j$ for $1\leq j\leq l$. This implies that $A/(I_j/I)\cong (A'/(I'_j/I')\otimes_{A'}A) (\cong A/((I'_j/I')\otimes_{A'}A))$, and hence, \begin{equation}\label{ideals} I_j/I=(I'_j/I')\otimes_{A'}A,\ (1\leq j\leq l) \end{equation} as ideals in $A$. For $1\leq j\leq l$, we can set $f(z'_j)=u_jz_j$ for some $u_j\in A$. Here, each $u_j$ is determined up to modulo $J_j/I$. Due to (\ref{ideals}), $u_jz_j$ generates the ideal $I_j/I$, and hence, $u_j$ is a unit in $A/(J_j/I)$ (and, of course, in $A/(J/I)$). (Note that $u_j$ is not necessarily a unit in $A$, since $A$ is not an integral domain for $l>1$.) Then there exists an isomorphism (naturally induced by $f$) of $Q$-modules \begin{equation}\label{def-tau} \tau_j\colon I'_j/I'I'_j\otimes_{A'}Q\stackrel{\sim}{\longrightarrow} I_j/II_j\otimes_AQ \end{equation} by $\xi'_j\mapsto(u_j\,\mbox{\rm mod}\,J/I)\xi_j$. The natural projection $\pi'_j\colon I'_i/I'I'_i\rightarrow I'_i/I' \subset A'$ $(1\leq i\leq l')$ and $f$ induce an $A$-module morphism \begin{equation}\label{def-proj} \widetilde{\rho}_i\colon I'_i/I'I'_i\otimes_{A'}A\longrightarrow A. \end{equation} For $1\leq j\leq l$, $\widetilde{\rho}_j$ maps $I'_i/I'I'_i\otimes_{A'}A$ surjectively onto $I_j/I$, and for $i>l$, $\widetilde{\rho}_i$ is an isomorphism; because, for $i>l$, one sees that $\widetilde{\rho}_i(\zeta'_i\otimes 1)=f(z'_i)$ is an invertible element of $A$ as folows: Since $\psi$ is injective in codimension zero, the point $I'_i/I'$ does not belong to $\psi(U)$; hence $\psi$ maps $U=\mathop{\mbox{\rm Spec}}\nolimits A$ to $\mathop{\mbox{\rm Spec}}\nolimits A'_{(I'_i/I')}$, and this implies the image of elements in $I'_i/I'$ under $f$ is invertible. Set $\rho_i\colon=\widetilde{\rho}_i\otimes_AQ$. Then these isomorphisms induce \begin{equation}\label{def-tau2} \tau\colon=\tau_1\otimes_Q\cdots\otimes_Q\tau_l\otimes_Q \rho_{l+1}\otimes_Q\cdots\otimes_Q\rho_{l'}\colon I'/I'^2\otimes_{A'}Q\stackrel{\sim}{\rightarrow}I/I^2\otimes_AQ. \end{equation} The $Q$-dual of $\tau$ is the desired isomorphism (cf. Lemma \ref{tangentloc}). One can easily check that this isomorphism $\tau$ does not depends on parameters $z'_j$, $z_j$; it is cannonically induced by $f\colon A'\rightarrow A$. Hence, for any sequence of \'{e}tale morphisms of local charts $U\stackrel{\psi}{\rightarrow}U'\stackrel{\psi'}{\rightarrow}U''$, we obviously have $\tau''=\tau\circ(\tau'\otimes_{Q'}Q)$, where $\tau\colon I'/I'^2\otimes_{A'}Q\stackrel{\sim}{\rightarrow}I/I^2\otimes_AQ$, $\tau'\colon I''/I''^2\otimes_{A''}Q'\stackrel{\sim}{\rightarrow} I'/I'^2\otimes_{A'}Q'$ and $\tau''\colon I''/I''^2\otimes_{A''}Q\stackrel{\sim}{\rightarrow} I/I^2\otimes_AQ$ are the isomorphisms defined as above with respect to $\psi$, $\psi'$ and $\psi'\circ\psi$, respectively. Then one sees easily that there exists a unique $\O_D$-module whose restriction to each $U$ is the $\O_{D_U}$-module corresponding to $T^1_A$; and it is nothing but our desired $\O_D$-module ${\cal T}^1_X$. Note that there exists a natural isomorphism ${\cal T}^1_X\stackrel{\sim}{\rightarrow}{\cal E}xt^1_{\O_X}(\Omega^1_{X/k},\O_X)$. Suppose $X$ has a global NCD embedding $X\hookrightarrow V$. Then by Lemma \ref{tangentloc}, the restriction of the normal bundle ${\cal N}_{X|V}$ to the singular locus $D$ is isomorphic to ${\cal T}^1_X$. Hence we have the following: \begin{pro}\label{suffcondemb} If a normal crossing variety $X$ over $k$ is embedded into a smooth $k$-variety as a normal crossing divisor, then there exists a line bundle $\L_X$ on $X$ such that $\L_X\otimes_{\O_X}\O_D\stackrel{\sim}{\rightarrow}{\cal T}^1_X$. \end{pro} Let $\hbox{\maxid X}\rightarrow\Delta$ be a semistable reduction of schemes, {\it i.e.}, a flat and generically smooth morphism between regular schemes with $\Delta$ one-dimensional and every closed fiber is a normal crossing variety. Suppose $X\rightarrow \mathop{\mbox{\rm Spec}}\nolimits k$ is isomorphic to a closed fiber of this family. Then one sees that the normal bundle ${\cal N}_{X|V}$ is trivial on $X$, and so is ${\cal T}^1_X$. \begin{dfn}{\rm (cf. \cite{Fri1})}\ \label{dsemistable}{\rm A normal crossing variety $X$ is said to be {\it $d$-semistable} if ${\cal T}^1_X$ is the trivial line bundle on $D$.} \end{dfn} Due to the above observation, we have the following: \begin{pro}{\rm (cf. \cite{Fri1})}\label{suffcondsmoothing} The $d$-semistablilty is a necessary condition for the existence of global smoothings of $X$. \end{pro} \section{Logarithmic embeddings} In this section, we define the logarithmic embedding of a normal crossing varieties (cf. \cite{Ste1}). This concept is defined in terms of log geometry of Fontaine, Illusie, and Kazuya Kato (cf. \cite{Kat1}). Let $X$ be a normal crossing variety over a field $k$. Suppose that $X$ has a NCD embedding $\iota\colon X\hookrightarrow V$. We denote the open immersion $V\setminus X\hookrightarrow V$ by $j$. We define a log structure on $X$ by $$ \iota^*(\O_V\bigcap j_*{\cal O}^\times_{V\setminus X})\longrightarrow\O_X, $$ where $\iota^*$ denotes the pull--back of log structures (cf. \cite[(1.4)]{Kat1}). We call this the log structure associated to the NCD embedding $\iota\colon X\hookrightarrow V$. For a general normal crossing variety $X$, we cannot define the log structure of this type on $X$, because $X$ may not have a NCD embedding. But, as we have seen in Remark \ref{ncvemb}, $X$ has \'{e}tale locally a NCD embedding. Then we can consider the log strcuture of this type for a general $X$ defined as follows: \begin{dfn}{\rm (cf. \cite{Ste1})}\ \label{logembdef}{\rm A log structure ${\cal M}_X\rightarrow\O_X$ is said to be of {\it embedding type}, if the following condition is satisfied: There exists an \'{e}tale covering $\{\varphi_{\lambda}\colon U_{\lambda}\rightarrow X\}_{\lambda\in\Lambda}$ by local charts --- with the NCD embeddings $\iota_{\lambda}\colon U_{\lambda}\hookrightarrow V_{\lambda}$ as in Definition \ref{ncvchart} --- such that, for each $\lambda\in\Lambda$, the restriction $$ {\cal M}_{U_{\lambda}}\colon =\varphi^*_{\lambda}{\cal M}_X\longrightarrow\O_{U_{\lambda}} $$ is isomorphic to the log structure associated to the NCD embedding $\iota_{\lambda}$. If ${\cal M}_X\rightarrow\O_X$ is a log structure of embedding type of $X$, we call the log scheme $(X,{\cal M}_X)$ the {\it logarithmic embedding}. } \end{dfn} Let $(X,{\cal M}_X)$ be a logarithmic embedding. We can explicitly write this log structure ${\cal M}_X$ \'{e}tale locally. Let $\nu\colon \widetilde{X}\rightarrow X$ be a normalization of $X$. Take a local chart $\varphi\colon U\rightarrow X$ with parameters $z_1,\ldots,z_l$ such that ${\cal M}_U\colon =\varphi^*{\cal M}_X\rightarrow\O_U$ is the log structure associated to the NCD embedding $\iota\colon U\hookrightarrow V$. Let $U=\bigcup^l_{i=1}U_i$ be the decomposition into irreduclbile components, where $U_i$ is the irreducible component corresponding to the ideal $(z_i)$. The normalization $\widetilde{U}=\coprod^{l}_{i=1}U_i\rightarrow U$ is denoted by $\nu_U$. Note that, due to Lemma \ref{normalization}, we have $U\times_X\widetilde{X} \cong\widetilde{U}$. Define a homomorphism of monoids \begin{equation}\label{logembchart} \alpha\colon (\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\rightarrow\O_U \end{equation} by $\alpha(e_{U_i})=z_i$ for $i=1,\ldots,l$, where $(e_{U_i})$ is the standard base of $(\nu_U)_*\mbox{\bf N}_{\widetilde{U}}=\bigoplus^l_{i=1}\mbox{\bf N}_{U_i}$. Then $\alpha$ induces a log structure \begin{equation}\label{logembloc} {\cal O}^\times_U\bigoplus(\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\rightarrow\O_U. \end{equation} \begin{pro}\label{logembloc-imp} The log structure ${\cal M}_U\rightarrow\O_U$ is isomorphic to (\ref{logembloc}). \end{pro} {\sc Proof.}\hspace{2mm} Let $Z_1,\ldots,Z_l\in\Gamma(V,\O_V)$ be as in Definition \ref{ncvchart}. By definition of the log structure associated to the embedding $\iota\colon U\hookrightarrow V$, these $Z_1,\ldots,Z_l$ are sections of the sheaf ${\cal M}_U$. Define a morphism $$ \psi\colon (\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\longrightarrow{\cal M}_U $$ by $\psi(e_{U_i})\colon =Z_i$ for $1\leq i\leq l$. Let $$ \widetilde{\psi}\colon (\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\rightarrow {\cal M}_U/{\cal O}^\times_U $$ be the composition of $\psi$ followed by the natural projection ${\cal M}_U\rightarrow{\cal M}_U/{\cal O}^\times_U$. It is easy to see that $\widetilde{\psi}$ is injective. Since sections of $\O_V\cap j_*{\cal O}^\times_{V\setminus U}$ are precisely those of $\O_V$ which may take zeros along $\iota(U)$, these are written in the form $uZ^{a_1}_1\cdots Z^{a_l}_l$ where $u\in{\cal O}^\times_V$ and $a_1,\ldots,a_l \in\mbox{\bf N}$. This implies that the morphism $\widetilde{\psi}$ is an isomorphism. Then, consider the exact sequence of sheaves of monoids $$ 1\rightarrow{\cal O}^\times_U\rightarrow{\cal M}_U\rightarrow{\cal M}_U/{\cal O}^\times_U\rightarrow 1, $$ where the second arrow is injective. This exact sequence splits since $\widetilde{\psi}$ is an isomorphism and $\psi$ defines a cross section ${\cal M}_U/{\cal O}^\times_U\rightarrow{\cal M}_U$. By this, we can easily obtain the desired result. $\Box$ Thus, a log structure of embedding type is determined by the morphism $\alpha\colon (\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\rightarrow\O_U$ such that $\alpha(e_{U_i})$ is a local defining function of the component $U_i$ for each $i=1,\ldots,l$. Let ${\alpha}'$ be another such homomorphism. Then --- replacing $U$ by sufficiently small Zariski open subset --- we can take $u_i\in\Gamma(U,{\cal O}^\times_X)$ such that ${\alpha}'(e_{U_i})=u_i\alpha(e_{U_i})$ for each $i$. Then the isomorphism of log structures of embedding type determined by $\alpha$ and ${\alpha}'$ is described by the following commutative diagram \begin{equation}\label{logemb-iso} \begin{array}{ccccc} {\cal O}^\times_U\oplus(\nu_U)_*\mbox{\bf N}_{\widetilde{U}}&&\stackrel{\phi}{\longrightarrow}&& {\cal O}^\times_U\oplus(\nu_U)_*\mbox{\bf N}_{\widetilde{U}}\\ &\llap{$\mbox{\rm by}\ {\alpha}'$}\searrow&&\swarrow\rlap{$\mbox{\rm by}\ \alpha$}\\ &&\O_X\rlap{,} \end{array} \end{equation} where $\phi$ is defined by $\phi(1,e_{U_i})=(u_i,e_{U_i})$ for each $i=1,\ldots,l$. In particular, the log structure of embedding type exists \'{e}tale locally, and is unique up to isomorphisms. \begin{cor} For any logarithmic embedding $(X,{\cal M}_X)$, we have an exact sequence of abelian sheaves \begin{equation}\label{abelian} 1\longrightarrow{\cal O}^\times_X\longrightarrow\gp{{\cal M}_X}\longrightarrow \nu_*\mbox{\bf Z}_{\widetilde{X}}\longrightarrow 0. \end{equation} \end{cor} {\sc Proof.}\hspace{2mm} Due to the local expression (\ref{logembloc}). $\Box$ In the rest of this section, we prove the following theorem, which is the main theorem of this paper. \begin{thm}\label{mainthm} For a normal crossing variety $X$, the logarithmic embedding of $X$ exists if and only if there exists a line bundle ${\cal L}$ on $X$ such that ${\cal L}\otimes_{\O_X}\O_D\stackrel{\sim}{\rightarrow} {\cal T}^1_X$. \end{thm} For the proof of this theorem, we shall prove some lemmas as follows. Let $(\varphi\colon U=\mathop{\mbox{\rm Spec}}\nolimits A\rightarrow X; z_1,\ldots,z_l)$ be a local chart on $X$. Let the NCD embedding $U\hookrightarrow V=\mathop{\mbox{\rm Spec}}\nolimits R$ and the ideals $I_j,I, J_j,J$ of $R$ be as in the previous section. \begin{lem}\label{lem-1} The natural morphism $$ \bigoplus^l_{j=1}J_j/I\longrightarrow J/I $$ of $A$-modules, induced by $J_j\hookrightarrow J$, is an isomorphism. \end{lem} {\sc Proof.}\hspace{2mm} The surjectivity is clear. We are going to show the injectivity. Take $a_jZ_1\cdots\widehat{Z_j}\cdots Z_l\in J_j$ --- where $Z_1,\ldots,Z_l$ are as in the previous section --- for $1\leq j\leq l$ such that $$ \sum^l_{j=1}a_jZ_1\cdots\widehat{Z_j}\cdots Z_l=b\cdot Z_1\cdots Z_l, $$ where $a_j,b\in R$. Since $R$ is an integral domain, $a_j$ is divisible by $Z_j$, and hence, we have $a_jZ_1\cdots\widehat{Z_j}\cdots Z_l\equiv 0\ (\mbox{\rm mod}\,I)$. $\Box$ Let $\pi_j\colon I_j/II_j\rightarrow I_j/I$ and $q_j\colon I_j/I \rightarrow I_j/JI_j(\cong I_j/II_j\otimes_AQ \ \mbox{\rm where}\ Q=R/J)$ be the natural projections and set $p_j\colon=q_j\circ\pi_j$. Let $q\colon I/I^2\rightarrow I/JI(\cong I/I^2\otimes_AQ)$ be the natural projection. \begin{lem}\label{lem-2} Let $M_1,\ldots,M_l$ be free $A$-modules of rank one and set $M\colon=M_1\otimes_A\cdots\otimes_AM_l$. Suppose we are given an $A$-module isomorphism $\widetilde{g}\colon M\stackrel{\sim}{\rightarrow}I/I^2$ and $A$-module homomorphisms $g_j\colon M_j\rightarrow I_j/I$, for $1\leq j\leq l$, such that, \begin{enumerate} \item for each $j$, there exists a free generator $\delta_j$ of $M_j$ such that $g_j(\delta_j)=z_j$, \item $(q_1\circ g_1)\otimes_Q\cdots\otimes_Q(q_l\circ g_l)=q\circ\widetilde{g}$. \end{enumerate} Then there exists a unique collection $\{\widetilde{g}_j\colon M_j \stackrel{\sim}{\rightarrow}I_j/II_j\}^{l}_{j=1}$ of $A$-isomorphisms such that $\pi_j\circ\widetilde{g}_j=g_j$ for each $j$ and $\widetilde{g}_1\otimes_A\cdots\otimes_A\widetilde{g}_l=\widetilde{g}$. \end{lem} {\sc Proof.}\hspace{2mm} We fix the free generators $\delta_j$ of $M_j$ as above. Then $M$ is generated by $\delta_1\otimes\cdots\otimes\delta_l$. Set $\widetilde{g}(\delta_1\otimes\cdots\otimes\delta_l)= v\zeta_1\otimes\cdots\otimes\zeta_l$ where $v\in A^{\times}$. By the second condition, we have $v\equiv 1\ (\mbox{\rm mod}\,J/I)$, {\it i.e.}, $$ v=1+\sum^l_{j=1}a_jz_1\cdots\widehat{z_j}\cdots z_l $$ for $a_j\in A$. We set $u_j=1+a_jz_1\cdots\widehat{z_j}\cdots z_l$ and define $\widetilde{g}_j$ by $\widetilde{g}_j(\delta_j)\colon=u_j\zeta_j$ for $1\leq j\leq l$. Then, since $v=u_1\cdots u_l$, each $u_j$ is a unit in $A$ and $\widetilde{g}_j$ is an isomorphism. Moreover, we have $\widetilde{g}_1\otimes\cdots\otimes\widetilde{g}_l=\widetilde{g}$ as desired. The uniqueness follows from Lemma \ref{lem-1}. $\Box$ \vspace{3mm} {\sc Proof of Theorem \ref{mainthm}}. We first prove the ``if'' part. This part is divided into four steps. {\sc Step 1}: Here, we shall describe the log structure of embedding type by another \'{e}tale local expression. Let $(\varphi\colon U=\mathop{\mbox{\rm Spec}}\nolimits A\rightarrow X; z_1,\ldots,z_l)$ be a local chart. For $m=(m_1,\ldots,m_l)\in\mbox{\bf N}^l$, define an $A$-module $P_m$ by $$ P_m\colon=(I_1/II_1)^{\otimes m_1}\otimes_A\cdots\otimes_A (I_l/II_l)^{\otimes m_l}. $$ Each $P_m$ is a free $A$-module of rank one and $P_{(1,\ldots,1)}\cong I/I^2$. The natural projections $\pi_j$ induce a natural $A$-homomorphism $$ \sigma_m\colon P_m\longrightarrow A. $$ Define a monoid $$ M\colon =\left\{ \begin{array}{c|l} (m,a)&m\in\mbox{\bf N}^l,\\ &a:\mbox{a generator of $P_m$} \end{array} \right\}, $$ and a homomorphism $M\rightarrow A$ of monoids by $(m,a)\mapsto\sigma_m(a)$. Then the associated log structure $\alpha_U\colon{\cal M}_U\rightarrow\O_U$ of the pre--log structure $M\rightarrow A$ is that of embedding type on $U$. {\sc Step 2}: Now, we assume that we are given a line bundle $\L$ on $X$ satisfying $\L\otimes_{\O_X}\O_D\cong({\cal T}^1_X)^{\vee}$. Suppose we have two local charts $(\varphi\colon U\rightarrow X; z_1,\ldots,z_{l})$ and $(\varphi'\colon U'\rightarrow X; z'_1,\ldots,z'_{l'})$ and an \'{e}tale morphism $\psi\colon U\rightarrow U'$ such that $\varphi=\varphi'\circ\psi$. For these local charts, we use the notation as in the previous section; such as $U=\mathop{\mbox{\rm Spec}}\nolimits A\hookrightarrow V=\mathop{\mbox{\rm Spec}}\nolimits R$, $U'=\mathop{\mbox{\rm Spec}}\nolimits A'\hookrightarrow V'=\mathop{\mbox{\rm Spec}}\nolimits R'$, $f\colon A'\rightarrow A$, $I$, $I'$, etc. As in the previous section, we may assume $(I'_j/I')\otimes_{A'}A=I_j/I$ as ideals in $A$ for $1\leq j\leq l$, and set $f(z'_j)=u_jz_j$ (each $u_j$ is determined up to modulo $J_j/I$). To give the line bundle $\L$ as above is equivalent to give a compatible system of isomorphisms $$ \widetilde{\tau}\colon I'/I'^2\otimes_{A'}A\stackrel{\sim}{\longrightarrow}I/I^2, $$ for all such $U\rightarrow U'$, with $\widetilde{\tau}\otimes_AQ=\tau$, where $\tau$ is defined as in (\ref{def-tau2}). Then we shall show that $\widetilde{\tau}$ induces canonically an isomorphism of log structures $\psi^*{\cal M}_{U'}\stackrel{\sim}{\rightarrow}{\cal M}_U$, and prove that these isomorphisms form so a compatible system that the log structures ${\cal M}_U$ glue to a log structure of embedding type on $X$. Moreover --- since local charts form an \'{e}tale open basis (Corollary \ref{specialcov}) --- we can pass through this procedure replacing $U$ by its Zariski open subset if necessary. In particular, we may assume that each $u_j$ as above is a unit in $A$, because $(u_j\,\mbox{\rm mod}\,J/I)$ is a unit in $A/(J/I)$ (in case $l>1$). Fix a locally constant section $w\in\H^0(D,{\cal O}^\times_D)$. (Actually, we can take $w$ as any global section in $\H^0(D,{\cal O}^\times_D)$ but, if we do so, the following argument have to be modified slightly.) {\sc Step 3}: (i) If $l=l'=1$, {\it i.e.}, $I_1=I$ and $I'_1=I'$, then we set $\widetilde{\tau}_1\colon I'_1/I'I'_1\otimes_{A'}A\stackrel{\sim}{\rightarrow} I_1/II_1$ by $\widetilde{\tau}_1\colon=\widetilde{\tau}$. (ii) If $l=1$ and $l'>1$, we define $\widetilde{\tau}_1\colon I'_1/I'I'_1\otimes_{A'}A\stackrel{\sim}{\rightarrow} I_1/II_1$ as follows: Suppose $\widetilde{\tau}$ maps $\zeta'_1\otimes\cdots\otimes\zeta'_{l'}\otimes 1$ to $v\zeta_1$, where $v\in A^{\times}$. Let $\widetilde{\rho}_j\colon I'_i/I'I'_i\otimes_{A'}A\rightarrow A$ be as (\ref{def-proj}), for $1\leq i\leq l'$. Suppose, moreover, each $\widetilde{\rho}_i$, for $i>1$, maps $\zeta'_i\otimes 1$ to $v_i\in A^{\times}$. Then, define $\widetilde{\tau}_1$ by $\widetilde{\tau}_1(\zeta'_1\otimes 1) \colon=w_Uvv^{-1}_2\cdots v^{-1}_{l'}\zeta_1$, where $w_U$ is a non--zero scalar which coincides with $w$ restricted to $D_U$. (iii) Suppose $l>1$ and $l'>1$. We claim that, under the conditions \begin{equation}\label{cond1} \pi_j\circ\widetilde{\tau}_j=\widetilde{\rho}_j,\ (1\leq j\leq l) \end{equation} and \begin{equation}\label{cond2} \widetilde{\tau}_1\otimes_A\cdots\otimes_A\widetilde{\tau}_l\otimes_A \widetilde{\rho}_{l+1}\otimes_A\cdots\otimes_A\widetilde{\rho}_{l'}= \widetilde{\tau}, \end{equation} the $A$-isomorphisms $$ \widetilde{\tau}_j\colon I'_j/I'I'_j\otimes_{A'}A\stackrel{\sim} {\longrightarrow}I_j/II_j $$ exist uniquely for $1\leq j\leq l$. Set $M_j\colon=I'_j/I'I'_j\otimes_{A'}A$ and $g_j\colon=\widetilde{\rho}_j$ for $1\leq j\leq l$. Define $\widetilde{g}$ by $\widetilde{g}\otimes_A\widetilde{\rho}_{l+1}\otimes_A \cdots\otimes_A\widetilde{\rho}_{l'}=\widetilde{\tau}$ (this is possible since $\widetilde{\rho}_i(\zeta'_i\otimes 1)$ is a unit element in $A$ for $i>l$), which is obviously an isomorphism. Then --- since we assumed each $u_j$ to be a unit in $A$ --- $M_j\colon= I'_j/I'_jI'\otimes_{A'}A$, $g$, and $g_j$ satisfy the conditions in Lemma \ref{lem-2}. Hence our claim follows from this lemma. Note that, in any cases, we have the following commutative diagram: \begin{equation}\label{com-mor} \begin{array}{ccc} I'_j/I'I'_j&\longrightarrow&I_j/II_j\\ \llap{$\pi'_j$}\vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\rlap{$\pi_j$}\\ A'&\underrel{\longrightarrow}{f}&A\rlap{,} \end{array} \end{equation} for $1\leq j\leq l$; this follows from (\ref{cond1}) in case $l, l'>1$, and is quite obvious in the other cases. {\sc Step 4}: These morphisms $\widetilde{\tau}_j$ induce the morphisms $$ \gamma_{m'}\colon P'_{m'}\longrightarrow P_{m}, $$ where $m=(m_1,\ldots,m_l)$ for $m'=(m_1,\ldots,m_{l'})\in\mbox{\bf N}^{l'}$. Then these $\gamma_{m'}$ induce naturally a morphism of monoids $M'\rightarrow M$ compatible with $M'\rightarrow A'$, $M\rightarrow A$ and $f$. By the construction of these morphisms, the induced morphism of sheaves of monoids $\gamma\colon \psi^*{\cal M}_{U'}\stackrel{\sim}{\rightarrow}{\cal M}_U$ is an isomorphism. By the commutative diagram (\ref{com-mor}), this isomorphism commutes the following diagram: $$ \begin{array}{ccc} \psi^*{\cal M}_{U'}&\stackrel{\sim}{\longrightarrow}&{\cal M}_{U}\\ \llap{$\psi^*\alpha_{U'}$}\vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\rlap{$\alpha_U$}\\ \O_U&=&\O_U\rlap{;} \end{array} $$ hence $\gamma$ is an isomorphism of log structures. Our construction of the isomorphism $\gamma$ is canonical in the following sense: Suppose we are given a sequence of \'{e}tale morphisms $U\stackrel{\psi}{\rightarrow}U'\stackrel{\psi'}{\rightarrow}U''$ of local charts (with $U$ and $U'$ sufficiently small), we have $\gamma''=\gamma\circ(\psi^*\gamma')$, where $\gamma\colon \psi^*{\cal M}_{U'}\stackrel{\sim}{\rightarrow}{\cal M}_U$, $\gamma'\colon \psi'^*{\cal M}_{U''}\stackrel{\sim}{\rightarrow}{\cal M}_{U'}$ and $\gamma''\colon \psi^*\psi'^*{\cal M}_{U''}\stackrel{\sim}{\rightarrow}{\cal M}_U$ are the isomorphisms of log structures defined as above corresponding to $\psi$, $\psi'$ and $\psi'\circ\psi$, respectively. This follows from the naturality of $\pi_j$ and $\widetilde{\rho}_j$, and the compatibility of $\widetilde{\tau}$'s. Then there exists a unique log structure ${\cal M}_X$ on $X$ which is of embedding type. Hence the ``if'' part is now proved. Conversely, suppose we are given a log structure ${\cal M}_X$ of embedding type. Then we have an exact sequence (\ref{abelian}) of abelian sheaves. Considering the cohomology exact sequence, we obtain a morphism $$ \delta\colon \H^0(X,\nu_*\mbox{\bf Z}_{\widetilde{X}})\longrightarrow\H^1(X,{\cal O}^\times_X) (\cong\mathop{\mbox{\rm Pic}}\nolimits X). $$ In $\H^0(X,\nu_*\mbox{\bf Z}_{\widetilde{X}})$, we consider the element $\hbox{\maxid{d}}$ which is defined by the image of $1\in\mbox{\bf Z}_X$ under the diagonal morphism $\mbox{\bf Z}_X\rightarrow \nu_*\mbox{\bf Z}_{\widetilde{X}}$. Then $\delta(\hbox{\maxid{d}})$ defines a line bundle $\L=\L_{{\cal M}_X}$ on $X$. We shall show that this line bundle satisfies $\L\otimes_{\O_X}\O_D \stackrel{\sim}{\rightarrow}({\cal T}^1_X)^{\vee}$. The line bundle $\L$ is constructed as follows: the inverse image of $\hbox{\maxid{d}}$ under $\gp{{\cal M}_X}\rightarrow\nu_*\mbox{\bf Z}_{\widetilde{X}}$ defines a principally homogeneous space over ${\cal O}^\times_X$ and hence defines a line bundle, which is nothing but $\L$. Let $U=\mathop{\mbox{\rm Spec}}\nolimits A$ be a local chart as above. Then the inverse image of $\hbox{\maxid{d}}$ restricted to $U$ gives a generator of an $A$-module $I/I^2$ which is --- due to Lemma \ref{tangentloc} --- a local lifting of ${\cal T}^1_X$ restricted to $U$. Hence $\L$ satisfies the desired condition. $\Box$ \begin{rem}\label{important}{\rm 1. As we have seen above, the log structure of embedding type exists locally and is unique up to isomorphisms. The sheaf of germs of automorphisms of such a log structure is naturally isomorphic to ${\cal K}$, where ${\cal K}$ is defined by the exact sequence \begin{equation}\label{important2} 1\longrightarrow{\cal K}\longrightarrow{\cal O}^\times_X\longrightarrow {\cal O}^\times_D\longrightarrow 1. \end{equation} This can be shown by the following steps: (i) any automorphism over a sufficiently small local chart $U$ is given by $ \phi$ in the diagram (\ref{logemb-iso}) with $\alpha=\alpha'$; (ii) $\phi$ is determined by $\{u_i\}$ with $u_i\in\Gamma(U,{\cal O}^\times_X)$ such that $z_i=u_i\cdot z_i$ for each $i$; (iii) hence such $u_i$'s are written in the form of $u_i= 1+a_i\cdot z_1\cdots\widehat{z_i}\cdots z_l$; (iv) due to Lemma \ref{lem-1}, to give a system $\{u_i\}$ is equivarent to give $u=u_1\cdots u_l$ which is a section of ${\cal K}$. Hence the obstruction for the existence of log structures of embedding type lies in $\H^2(X,{\cal K})$. The proof of Theorem \ref{mainthm} shows that this class coincides with the obstruction class for a lifting of $(T^1_X)^{\vee}$ on $X$, {\it i.e.}, the image of $(T^1_X)^{\vee}$ under $\H^1(D,{\cal O}^\times_D)\rightarrow\H^2(X,{\cal K})$. 2. One sees easily --- by the proof of Theorem \ref{mainthm} --- that there exists a natural surjective map \begin{equation}\label{important1} \left\{ \begin{array}{ll} \mbox{isom. class of log structures}\\ \mbox{of embedding type on $X$} \end{array} \right\} {\longrightarrow} \left\{ \begin{array}{c|c} \L\in\mathop{\mbox{\rm Pic}}\nolimits X&\L\otimes_{\O_X}\O_D\stackrel{\sim}{\rightarrow}({\cal T}^1_X)^{\vee} \end{array} \right\} \end{equation} by ${\cal M}\mapsto\L_{{\cal M}}$, where $\L_{{\cal M}}$ is defined as in the proof of Theorem \ref{mainthm}. If ${\cal M}_X$ is associated to a global NCD embedding $X\hookrightarrow V$, then $\L_{{\cal M}_X}$ is nothing but the conormal bundle of $X$ in $V$. The set of isomorphism classes of log structures of embedding type on $X$, is a principally homogeneous space over $\H^1(X,{\cal K})$. Then one sees easily that the map (\ref{important1}) is equivariant to $\H^1(X,{\cal K})\rightarrow \mathop{\mbox{\rm Ker}}\nolimits(\H^1(X,{\cal O}^\times_X)\rightarrow\H^1(D,{\cal O}^\times_D))$ induced by the cohomology exact sequence of (\ref{important2}). In particular, if $X$ is proper and $D$ is connected, the map (\ref{important1}) is a bijection since $\H^1(X,{\cal K})\stackrel{\sim}{\rightarrow}\mathop{\mbox{\rm Ker}}\nolimits(\H^1(X,{\cal O}^\times_X)\rightarrow \H^1(D,{\cal O}^\times_D))$; in this case, the logarithmic embeddings are determined by their ``normal bundles.'' 3. By the exact sequence (\ref{abelian}), a log structure of embedding type ${\cal M}_X$ on $X$ defines an extension class in $\mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\nu_*\mbox{\bf Z}_{\widetilde{X}},{\cal O}^\times_X)$. Under the morphism $\mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\nu_*\mbox{\bf Z}_{\widetilde{X}},{\cal O}^\times_X) \rightarrow\mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X)$, induced by the diagonal morphism $\mbox{\bf Z}_X\rightarrow\nu_*\mbox{\bf Z}_{\widetilde{X}}$, and the natural identification $\mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X)\stackrel{\sim}{\rightarrow}\mathop{\mbox{\rm Pic}}\nolimits X$, this class is mapped to the class corresponding to the line bundle $\L_{{\cal M}_X}$ defined as above. (The proof is straightforward and left to the reader.) } \end{rem} \section{Logarithmic semistable reductions} \begin{dfn}{\rm (cf. \cite{Kaj1}, \cite{K-N1})}\label{logsemidef}\ {\rm A log strcuture of embedding type ${\cal M}_X\rightarrow\O_X$ is said to be of {\it semistable type}, if there exists a homomorphism $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ of abelian sheaves on $X$ such that the diagram $$ \begin{array}{ccccc} \gp{{\cal M}_X}&&\longrightarrow&&\nu_*\mbox{\bf Z}_{\widetilde{X}}\\ &\nwarrow&&\nearrow\rlap{$\hbox{\maxid{d}}$}\\ &&\mbox{\bf Z}_X \end{array} $$ commutes, where $\hbox{\maxid{d}}\colon\mbox{\bf Z}_X{\rightarrow}\nu_*\mbox{\bf Z}_{\widetilde{X}}$ is the diagonal homomorphism, and $\gp{{\cal M}_X}\rightarrow\nu_*\mbox{\bf Z}_{\widetilde{X}}$ is the projection in (\ref{abelian}). } \end{dfn} If ${\cal M}_X$ is a log strcuture of semistable type, the homomorphism $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ induces the homomorphism $\mbox{\bf N}_X\rightarrow{\cal M}_X$ of monoids by the following Cartesian diagram: $$ \begin{array}{ccc} \mbox{\bf N}_X&\longrightarrow&{\cal M}_X\\ \vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\\ \mbox{\bf Z}_X&\longrightarrow&\gp{{\cal M}_X}\rlap{;} \end{array} $$ this follows easily from the local expression (\ref{logembloc}). This morphism defines a morphism of log schemes $$ (X,{\cal M}_X)\longrightarrow(\mathop{\mbox{\rm Spec}}\nolimits k,\mbox{\bf N}) $$ Here, $(\mathop{\mbox{\rm Spec}}\nolimits k,\mbox{\bf N})$ is the {\it standard point} defined by $\mbox{\bf N}\rightarrow k$ which maps $m\in\mbox{\bf N}$ to $0^m$. We call this morphism of log schemes the {\it logarithmic semistable reduction}. Logarithmic semistable reductions are {\it log smooth} in the sense of \cite{Kat1}. \begin{rem}\label{genuine-semistable}{\rm Let $f\colon\hbox{\maxid X}\rightarrow\Delta$ be a semistable reduction of schemes; {\it i.e.}, a proper flat generically smooth morphism $f$ with $\hbox{\maxid X}$ a regular scheme and $\Delta$ a one-dimensional regular local scheme, with the closed fiber $X\rightarrow 0=\mathop{\mbox{\rm Spec}}\nolimits k$ a normal crossing variety. Then this morphism induces canonically a logarithmic semistable reduction $(X,{\cal M}_X)\rightarrow(\mathop{\mbox{\rm Spec}}\nolimits k,\mbox{\bf N})$ on the closed fiber as follows: We define a log structure ${\cal M}_{\hbox{\maxid X}}\rightarrow\O_{\hbox{\maxid X}}$ by $$ {\cal M}_{\hbox{\maxid X}}\colon=\O_{\hbox{\maxid X}}\bigcap j_*{\cal O}^\times_{\hbox{\maxid X}\setminus X} \lhook\joinrel\longrightarrow\O_{\hbox{\maxid X}} $$ where $j\colon \hbox{\maxid X}\setminus X\hookrightarrow\hbox{\maxid X}$ is an open immersion. Take a local parameter $t\in\O_{\Delta}$ around $0=\mathop{\mbox{\rm Spec}}\nolimits k$. Then $f^{-1}(t)$ belongs to ${\cal M}_{\hbox{\maxid X}}$. We define a homomorphism of monoids $\mbox{\bf N}_{\hbox{\maxid X}}\rightarrow{\cal M}_{\hbox{\maxid X}}$ by $1\mapsto f^{-1}(t)$. Then this homomorphism extends to a morphism of log schemes \begin{equation}\label{france} (\hbox{\maxid X},{\cal M}_{\hbox{\maxid X}})\longrightarrow(\Delta,0), \end{equation} where the log structure on $\Delta$ is the associated log structure of $$ \mbox{\bf N}\longrightarrow\O_{\Delta}\quad \mbox{\rm by}\quad m\mapsto t^m. $$ Taking the pull--back of (\ref{france}) to the closed fiber, we get a logarithmic semistable reduction. Note that the monoid morphism $\mbox{\bf N}_{\hbox{\maxid X}}\rightarrow{\cal M}_{\hbox{\maxid X}}$ induces $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ which satisfies the condition in Definition \ref{logsemidef}. Hence, such a morphism $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ for a general log structure of semistable type can be regarded as a ``parametrization.'' } \end{rem} \begin{rem}{\rm The logarithmic semistable reduction induced by a semistable reduction family, as in Remark \ref{genuine-semistable}, is regarded as the ``closed fiber'' of the morphism (\ref{france}) of log schemes. Then, conversely, one can consider the theory of deformations which deal with liftings of the logarithmic semistable reductions. This is nothing but the {\it logarithmic deformation} of Kawamata--Namikawa \cite{K-N1}, and also a part of the {\it log smooth deformation} developed in \cite{Kat2}. } \end{rem} Using Theorem \ref{mainthm} --- which is proved in the previous section --- we get a new proof of the theorem of Kawamata--Namikawa as follows: \begin{thm}\label{mainthm2}{\rm (cf. \cite{K-N1})}\ For a normal crossing variety $X$, the log structure of semistable type on $X$ exists if and only if $X$ is $d$-semistable. \end{thm} To prove the theorem, we need the following lemma: \begin{lem}\label{l1} Let $(X,{\cal M}_X)$ be a logarithmic embedding. Consider the exact sequence (\ref{abelian}) of abelian sheaves and the induced morphism $$ \mathop{\mbox{\rm Hom}}\nolimits_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,\nu_*\mbox{\bf Z}_{\widetilde{X}})\stackrel{\delta}{\longrightarrow} \mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X). $$ Let $\hbox{\maxid{d}}\in\mathop{\mbox{\rm Hom}}\nolimits_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,\nu_*\mbox{\bf Z}_{\widetilde{X}})$ be the diagonal morphism. Then, under the natural identification $\mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X) \stackrel{\sim}{\rightarrow}\mathop{\mbox{\rm Pic}}\nolimits X$, we have $$ \delta(\hbox{\maxid{d}})=[\L_{{\cal M}_X}], $$ where $\L_{{\cal M}_X}$ is the line bundle defined in the previous section. \end{lem} {\sc Proof.}\hspace{2mm} This lemma follows from the commutative diagram $$ \begin{array}{ccc} \mathop{\mbox{\rm Hom}}\nolimits_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,\nu_*\mbox{\bf Z}_{\widetilde{X}})&\stackrel{\delta}{\longrightarrow}& \mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X)\\ \llap{$\cong$}\vphantom{\bigg|}\Big\downarrow&&\vphantom{\bigg|}\Big\downarrow\rlap{$\cong$}\\ \H^0(X,\nu_*\mbox{\bf Z}_{\widetilde{X}})&\longrightarrow&\H^1(X,{\cal O}^\times_X) \end{array} $$ where the vertical morphisms are natural isomorphisms and the definition of the line bundle $\L_{{\cal M}_X}$. $\Box$ \vspace{3mm} {\sc Proof of Theorem \ref{mainthm2}.}\hspace{2mm} Suppose ${\cal M}_X$ is a log structure of semistable type. Consider the exact sequence \begin{equation}\label{extension} \mathop{\mbox{\rm Hom}}\nolimits_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,\gp{{\cal M}_X})\stackrel{\pi}{\longrightarrow} \mathop{\mbox{\rm Hom}}\nolimits_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,\nu_*\mbox{\bf Z}_{\widetilde{X}})\stackrel{\delta}{\longrightarrow} \mathop{\mbox{\rm Ext}}\nolimits^1_{\mbox{\bf Z}_X}(\mbox{\bf Z}_X,{\cal O}^\times_X) \end{equation} induced by (\ref{abelian}). The ``parametrization'' morphism $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ is mapped to $\hbox{\maxid{d}}$ by $\pi$. This implies that the line bundle $\L_{{\cal M}_X}$ is trivial. Then so is $(T^1_X)^{\vee}$ because $\L_{{\cal M}_X}\otimes_{\O_X}\O_D$ is isomorphic to $(T^1_X)^{\vee}$. Conversely, if $X$ is $d$-semistable, there exists at least one log structure of embedding type on $X$ due to Theorem \ref{mainthm}. Since $\L_{{\cal M}_X}\otimes_{\O_X}\O_D$ is trivial, we can take the log structure ${\cal M}_X$ of embedding type such that the corresponding line bundle $\L_{{\cal M}_X}$ is trivial (due to the natural surjection (\ref{important1})). Since the obstruction for the existence of a morphism $\mbox{\bf Z}_X\rightarrow\gp{{\cal M}_X}$ which is mapped to $\hbox{\maxid{d}}$, is nothing but the class $[\L_{{\cal M}_X}]$, we deduce that ${\cal M}_X$ is of semistable type. $\Box$ As is shown in the above proof, the log structure of semistable type on $X$ is --- considering the natural surjection (\ref{important1}) --- the log structure of embedding type which is mapped to the trivial bundle on $X$. Hence we have the following: \begin{cor} Let $X$ be a proper, $d$-semistable normal crossing variety, and assume that the singular locus $D$ is connected. Then, the log structure of semistable type on $X$ exists uniquely. \end{cor} \begin{exa}{\rm Let $X\colon =X_0\cup\cdots\cup X_N$ be a chain of surfaces defined as follows: Each $X_i$ is the Hirzebruch surface of degree $a_i\leq 0$. The surfaces $X_{i-1}$ and $X_i$ are connected by identifying the section $s'_{i-1}$ on $X_{i-1}$ and the one $s_i$ on $X_i$, where $(s'_{i-1})^2=a_{i-1}$ and $(s_i)^2= -a_i$, for $1\leq i\leq N$. Then $X$ has a log structure of embedding type if and only if $a_i|(a_{i-1}+a_{i+1})$ for $1\leq i\leq N-1$, while $X$ has a log structure of semistable type if and only if $a_0=a_1=\cdots =a_N$. } \end{exa} \begin{small}
2006-03-24T21:30:51
9411
alg-geom/9411012
en
https://arxiv.org/abs/alg-geom/9411012
[ "alg-geom", "math.AG" ]
alg-geom/9411012
null
David B. Jaffe
On sextic surfaces having only nodes (preliminary report)
14 pages, AMS-LaTeX
null
null
null
null
Let S be a surface in CP^3, having only nodes as singularities. Let pi: S~ --> S be a minimal resolution of singularities. A set N of nodes on S is EVEN if there exists a divisor Q on S~ such that 2Q ~ pi^{-1}(N). Suppose that S has degree 6. It is known (Basset) that S cannot have 67 or more nodes. It is also known (Barth) that S can have 65 nodes. It is not known if S can have 66 nodes. Likewise, it is not known exactly what sizes can occur for an even set of nodes on S. We show that an nonempty even set of nodes on S must have size 24, 32, 40, 56, or 64. We do not know if the sizes 56 and 64 can occur. We show that if S has 66 nodes, then it must have an even set of 64 nodes, and it cannot have an even set of 56 nodes. THUS IF ONE COULD RULE OUT THE CASE OF A 64 NODE EVEN SET, IT WOULD FOLLOW THAT S CANNOT HAVE 66 NODES. The existence or nonexistence of large even node sets is related to the following vanishing problem. Let S be a normal surface of degree s in CP^3. Let D be a Weil divisor on S such that D is Q-rationally equivalent to rH, for some r \in \Q. Under what circumstances do we have H^1(O_S(D)) = 0? For instance, this holds when r < 0. For s=4 and r=0, H^1 can be nonzero. For s=6 and r=0, if a 56 or 64 node even set exists, then H^1 can be nonzero. The vanishing of H^1 is also related to linear normality, quadric normality, etc. of set-theoretic complete intersections in P^3. Hard copy is available from the author. E-mail to [email protected].
[ { "version": "v1", "created": "Thu, 17 Nov 1994 21:44:35 GMT" } ]
2008-02-03T00:00:00
[ [ "Jaffe", "David B.", "" ] ]
alg-geom
\section{#1}} \def\abs#1{{\vert{#1}\vert}} \def\floor#1{\lfloor#1\rfloor} \def\makeaddress{ \vskip 0.15in \par\noindent {\footnotesize Department of Mathematics and Statistics, University of Nebraska} \par\noindent {\footnotesize Lincoln, NE 68588-0323, USA\ \ (jaffe{\kern0.5pt}@{\kern0.5pt}cpthree.unl.edu)}} \def \def\arabic{footnote}}\setcounter{footnote}{0}{\fnsymbol{footnote}{ \def\arabic{footnote}}\setcounter{footnote}{0}{\fnsymbol{footnote}} \par\noindent David B. Jaffe\protect\footnote{Partially supported by the National Science Foundation.} \makeaddress\def\arabic{footnote}}\setcounter{footnote}{0}{\arabic{footnote}}\setcounter{footnote}{0}} \newenvironment{proof}{\trivlist \item[\hskip \labelsep{\sc Proof.\kern1pt}]}{\endtrivlist \newenvironment{alphalist}{\begin{list}{(\alph{alphactr})}{\usecounter{alphactr}}}{\end{list} \def\underline{\underline} \def\overline{\overline} \hfuzz 3pt \documentclass[12pt]{article}\usepackage{amssymb} \newtheorem{theorem}{Theorem}[section] \setlength{\parindent}{9mm} \setcounter{tocdepth}{3} \newtheorem{fact}[theorem]{Fact \newtheorem{proposition}[theorem]{Proposition \newtheorem{lemma}[theorem]{Lemma \newtheorem{sublemma}[theorem]{Sublemma \newtheorem{conjecture}[theorem]{Conjecture \newtheorem{cor}[theorem]{Corollary \newtheorem{corollary}[theorem]{Corollary \newtheorem{theoremdefinition}[theorem]{Theorem/Definition \newtheorem{propdefinition}[theorem]{Proposition/Definition \newtheorem{problemdefinition}[theorem]{Problem/Definition \newtheorem{prop}[theorem]{Proposition \newtheorem{claim}[theorem]{Claim \def\ifundefined#1{\expandafter\ifx\csname#1\endcsname\relax} \begin{document} \vskip 0.15in \def2 \left[ 1 + {s-1 \choose 3} \right]{2 \left[ 1 + {s-1 \choose 3} \right]} \begin{center} \bf\Large On sextic surfaces having only nodes\\ (preliminary report) \end{center} \vskip 0.15in \def\arabic{footnote}}\setcounter{footnote}{0}{\fnsymbol{footnote} \vspace{0.25in} \block{Introduction} Over ${\Bbb C}\kern1pt$, let $S \subset \xmode{\Bbb P\kern1pt}^3$ be a surface having only nodes as singularities. Let \mp[[ \pi || {\tilde{S}} || S ]] be a minimal resolution of singularities. A set $N$ of nodes on $S$ is {\it even\/} if there exists a divisor $Q$ on ${\tilde{S}}$ such that $2Q \sim \pi^{-1}(N)$. Suppose that $S$ has degree $6$. It is known (Basset) that $S$ cannot have $67$ or more nodes. It is also known (Barth) that $S$ can have $65$ nodes. It is not known if $S$ can have $66$ nodes. Likewise, it is not known exactly what sizes can occur for an even set of nodes on $S$. We show that an nonempty even set of nodes on $S$ must have size $24$, $32$, $40$, $56$, or $64$. We do not know if the sizes $56$ and $64$ can occur. We show that if $S$ has $66$ nodes, then it must have an even set of $64$ nodes% ,\footnote{J.\ Wahl informed me that he has independently obtained these results, but I have not seen his methods.} and it cannot have an even set of $56$ nodes. {\bf\it Thus if one could rule out the case of a 64 node even set, it would follow that S cannot have 66 nodes.} The existence or nonexistence of large even node sets is related to the following vanishing problem. Let $S \subset \xmode{\Bbb P\kern1pt}^3$ be a normal surface of degree $s$. Let $D$ be a Weil divisor on $S$ such that $D \sim_{\Bbb Q}\kern1pt rH$, for some $r \in {\Bbb Q}\kern1pt$. Under what circumstances do we have $H^1({\cal O}_S(D)) = 0$? For instance, this holds when $r < 0$. For $s=4$ and $r=0$, $H^1$ can be nonzero. For $s=6$ and $r=0$, if a $56$ or $64$ node even set exists, then $H^1$ can be nonzero. The vanishing of $H^1$ is also connected to linear normality, quadric normality, etc.\ of set-theoretic complete intersections\ in $\xmode{\Bbb P\kern1pt}^3$. See (\Lcitemark 5\Rcitemark \ 11.8--11.11). There is no further discussion of these issues in this paper. \block{Basic tools and notation} Since some of the calculations in this report apply as well to surfaces of arbitrary degree, we will for a while allow $S$ to be any surface of degree $s$ in $\xmode{\Bbb P\kern1pt}^3$ having only nodes as singularities. Many ideas used here are taken from Beauville\Lspace \Lcitemark 3\Rcitemark \Rspace{}, who analyzes quintic surfaces having only nodes. The following result is well-known: \begin{lemma}\label{torsion-free} $\mathop{\operatoratfont Pic}\nolimits({\tilde{S}})$ is torsion-free. \end{lemma} \begin{proof} Let ${\cal{L}}$ be a line bundle on ${\tilde{S}}$ such that ${\cal{L}}^{\o* n} \cong {\cal O}_{\tilde{S}}$ for some $n \in \xmode{\Bbb N}$. Let $S_0$ be the disjoint union of the spectra of the local rings of the singular points of $S$, and let ${\tilde{S}}_0 = \pi^{-1}(S_0)$. Since ${\cal{L}}$ restricted to each exceptional curve is trivial, it follows (see\Lspace \Lcitemark 7\Rcitemark \Rspace{}\ \S10, p.\ 157) that ${\cal{L}}|_{{\tilde{S}}_0}$ is trivial. Therefore there exist effective divisors $D_1, D_2$ on ${\tilde{S}}$ which do not meet the exceptional curves and are such that ${\cal{L}} \cong {\cal O}_{\tilde{S}}(D_1-D_2)$. Hence ${\cal{L}}$ is the pullback of a line bundle from $S$. But $\mathop{\operatoratfont Pic}\nolimits(S)$ is torsion-free and $\mathop{\operatoratfont Pic}\nolimits(\pi)$ is injective, so ${\cal{L}} \cong {\cal O}_{\tilde{S}}$. {\hfill$\square$} \end{proof} The even sets of nodes on $S$ comprise the codewords of a binary linear code $C$. That is, if $C$ is the set of all even sets of nodes, then $C$ may be regarded as a sub-vector-space of ${\Bbb F}\kern1pt_2^n$, where $n$ is the number of nodes on $S$. We use without much explanation some standard tools from coding theory to investigate $C$; the standard reference is \Lcitemark 6\Rcitemark \Rspace{}. For $w \in C$, let $\abs{w}$ denote the size of the node set, which in coding theory would be called the {\it weight\/} of $w$. It is the number of $1$'s which appear in the expression for $w$ as an element of ${\Bbb F}\kern1pt_2^n$. For $n \in \xmode{\Bbb N}$, let $C_n$ denote the number of words of weight $n$ in $C$. Let $C^\perp$ denote the dual code to $C$. If $J$ is a set of positive integers, by a $[n,k,J]$ code we mean a code with lives in ${\Bbb F}\kern1pt_2^n$, has dimension $k$, and has nonzero weights in the set $J$. We also underscore weights to indicate that they must occur. Thus, for example, a $[66,13,\setof{24,32,40,\underline{64}}]$ code is a code which lives in ${\Bbb F}\kern1pt_2^{66}$, has dimension $13$, has nonzero weights in the set \setof{24,32,40,64}, and has a word of weight $64$. For $w \in C$, we let $Q_w$ denote a divisor on ${\tilde{S}}$ with the property that $2Q_w$ is rationally equivalent to the sum of the exceptional curves corresponding to the points in $w$. By \pref{torsion-free}, $Q_w$ is determined up to rational equivalence. For $w \in C - \setof{0}$, let \mp[[ \tau_w || X_w || {\tilde{S}} ]] be the double cover branched along the union of the exceptional curves corresponding to the singular points in $w$, and determined by $Q_w$. (See\Lspace \Lcitemark 2\Rcitemark \Rspace{}\ I\ \S17 for background on this.) Since anything in $\mathop{\operatoratfont Ker}\nolimits[\mathop{\operatoratfont Pic}\nolimits(\tau_w)]$ must be $2$-torsion (\Lcitemark 2\Rcitemark \ I\ 16.2), it follows from \pref{torsion-free} that $\mathop{\operatoratfont Pic}\nolimits(\tau_w)$ is injective. For any divisor $D$ on ${\tilde{S}}$, we can push $D$ down to a Weil divisor $\overline{D}$ on $S$; the exceptional components of $D$ (if any) are ignored. \block{Results on linear systems and curves} If ${\cal{L}}$ is a globally generated line bundle on a smooth projective variety $T$, we consider the complete linear system corresponding to ${\cal{L}}$, and let $\varphi_{\cal{L}}$ denote the induced morphism from $T$ to its image, which is a projective variety. The following result is no doubt well-known; we include it for lack of a reference. \begin{lemma}\label{birational} Let \mp[[ f || W || T ]] be a surjective morphism of smooth projective varieties of dimension $\geq 2$, which is generically two-to-one, and such that $\mathop{\operatoratfont Pic}\nolimits(f)$ is injective. Let ${\cal{L}}$ be a globally generated line bundle on $T$. Assume that $h^0(f^*{\cal{L}}) > h^0({\cal{L}})$. Assume that $\varphi_{\cal{L}}$ is birational. Then $\varphi_{f^*{\cal{L}}}$ is birational. \end{lemma} \begin{proof} Let $D$ be an effective divisor in the complete linear system associated to ${\cal{L}}$. Let $Q$ be a general member of $\abs{f^*D}$. By Bertini's theorem, $Q$ is smooth and connected. Since $\dim \abs{f^*D} > \dim \abs{D}$, we see that $Q \notin f^*\abs{D}$. Let $f_*(Q)$ denote $f(Q)$, thought of as a reduced effective divisor. The divisor $f^*f_*(Q) - Q$ is effective. If $f^*f_*(Q) = Q$, then $f^*({\cal{L}}) \cong f^*{\cal O}_T(f_* Q)$, so ${\cal{L}} \cong {\cal O}_T(f_*Q)$ [since $\mathop{\operatoratfont Pic}\nolimits(f)$ is injective], and hence $Q \in f^*\abs{D}$: contradiction. Hence $f^*f_*(Q) - Q \not= 0$. Since $Q$ is general it follows that $Q \subsetneq f^{-1}(f(Q))$ as sets. Therefore if $x \in Q$ is a general point, $\setof{x} \subsetneq f^{-1}(f(x))$. Since $Q$ passes through $x$ and there is exactly one other point $y \in f^{-1}(f(x))$, which $Q$ does not pass through, we see that the restriction of $\varphi_{f^*{\cal{L}}}$ to $f^{-1}(f(x))$ is injective. Since $x$ ranges over general points on general members of $\abs{f^*D}$, in fact $x$ ranges over a nonempty open subset of $W$. Hence there is a nonempty open subset of $W$ on which $\varphi_{f^*{\cal{L}}}$ is injective. {\hfill$\square$} \end{proof} We state Castelnuovo's bound for the genus of a curve, in a form which follows immediately from (\Lcitemark 1\Rcitemark \ p.\ 116). \begin{theorem}\label{castelnuovo} Let $M$ be a smooth curve of genus $g$, that admits a birational mapping onto a nondegenerate curve of degree $d$ in $\xmode{\Bbb P\kern1pt}^r$, $r \geq 2$. Let $x = \floor{{d-1 \over r-1}}$, $y = {d-1 \over r-1} - x$. Then: $${g \over r-1} \leq {x \choose 2} + xy.$$ \end{theorem} \block{Results valid for any degree {\it s}} First we make an elementary calculation, and then we give a very useful proposition, which is probably known in some sense. \begin{lemma}\label{chi} Let $w \in C - \setof{0}$, and write $X = X_w$ and $Q = Q_w$ for simplicity. Let $p = \abs{w}$. Let $n \in \xmode{\Bbb Z}$. Then $$\chi({\cal O}_X(n))\ =\ n(n+4-s)s + 2 \left[ 1 + {s-1 \choose 3} \right] - p/4.$$ \end{lemma} \begin{proof} \begin{eqnarray*} \chi({\cal O}_X(n)) & = & \chi({\cal O}_{\tilde{S}}(n)) + \chi({\cal O}_{\tilde{S}}(-Q)(n))\\ & = & {1\over2}(nH) \cdot (nH-K_S) + {1\over2}(-Q + nH) \cdot (-Q + nH - K_S) + 2\chi({\cal O}_{\tilde{S}})\\ & = & (nH)\cdot(n+4-s)H + 2\chi({\cal O}_{\tilde{S}}) + {1\over2} Q^2\\ & = & n(n+4-s)s + 2 \left[ 1 + {s-1 \choose 3} \right] + {1\over8}(\svec E1p)^2.\kern1.15cm\square \end{eqnarray*} \end{proof} Let $w \in C - \setof{0}$, and simplify notation by writing $Q = Q_w$, $X = X_w$, $\tau = \tau_w$. Let $E$ be the union of the exceptional curves corresponding to the singular points in $w$. For any $n \in \xmode{\Bbb Z}$, consider the exact sequence \sesdot{{\cal O}_{\tilde{S}}(nH-Q)}{{\cal O}_{\tilde{S}}(nH+Q)}{{\cal O}_E(nH+Q)% }Since the restriction of $nH+Q$ to any component of $E$ has degree $-1$, $h^j({\cal O}_E(nH+Q)) = 0$ for all $j$. Hence $h^k({\cal O}_{\tilde{S}}(nH-Q)) = h^k({\cal O}_{\tilde{S}}(nH+Q))$ for all $k$. This fact will be used a number of times. \begin{prop}\label{mmmmm} Assume that $s \geq 5$. Assume that $h^0({\cal O}_X(1)) \geq 5$. Then $h^0({\cal O}_X(1)) = 5$ and $\abs{w} \leq s(s-1)$. \end{prop} \begin{proof} Let $p = \abs{w}$. Since $h^0({\cal O}_{{\tilde{S}}}(1)) = h^0({\cal O}_S(1)) = 4$, it follows by \pref{birational} that $\varphi_{{\cal O}_X(1)}$ is birational. Let $H$ be a general member of the complete linear system on ${\tilde{S}}$ corresponding to ${\cal O}_{{\tilde{S}}}(1)$. Let $M = \tau^{-1}(H)$. By Bertini's theorem, $M$ is a smooth connected curve. Let ${\cal O}_M(1)$ denote the pullback of ${\cal O}_X(1)$ to $M$. Since $\varphi_{{\cal O}_X(1)}$ is birational, $\varphi_{{\cal O}_M(1)}$ is birational. Let $r = h^0({\cal O}_M(1)) - 1$. The map \mapx[[ M || H ]] is an \'etale double cover. Since $H$ has genus $(s-1)(s-2)/2$, $M$ has genus $(s-1)(s-2)-1$. Since the image of $H$ in $\xmode{\Bbb P\kern1pt}^3$ has degree $s$, it follows that $\varphi_{{\cal O}_M(1)}(M)$ is a nondegenerate curve of degree $2s$ in $\xmode{\Bbb P\kern1pt}^r$. We have $r \geq 3$; we will show that in fact $r = 3$. Apply \pref{castelnuovo}. We assume that $s \geq 8$, leaving the cases $s = 5,6,7$ to the reader. We have $x = \floor{{2s-1 \over r-1}}$. Since $y \leq (r-2)/(r-1)$, we have: \begin{eqnarray*} {g \over r-1} & = & {(s-1)(s-2)-1 \over r-1}\ \leq\ {x(x-1) \over 2} + \left( {r-2 \over r-1} \right) x\\ & \leq & {(2s-1)(2s-r) \over 2(r-1)^2} + {(2s-1)(r-2) \over (r-1)^2}, \end{eqnarray*} which implies that $$2(r-1)(s^2 - 3s + 1)\ \leq\ (2s-1)(2s+r-4).$% $In particular, if $r \geq 4$, then the above inequality holds even when $r = 4$ is formally substituted in, and it follows that $s^2 - 8s + 3 \leq 0$. Hence $s < 8$: contradiction. Hence $r = 3$. {}From the exact sequence \ses{{\cal O}_X}{{\cal O}_X(1)}{{\cal O}_M(1)% }we get $h^0({\cal O}_X(1)) \leq 5$, so in fact $h^0({\cal O}_X(1)) = 5$. Since $h^0({\cal O}_X(1)) = h^0({\cal O}_{{\tilde{S}}}(H)) + h^0({\cal O}_{{\tilde{S}}}(H-Q))$, we get $h^0({\cal O}_{{\tilde{S}}}(H-Q)) = 1$. Let $I \in \abs{H-Q}$. Since $I \cdot H = s$, $\pi(\red{I})$ is a curve (possibly reducible) of degree $\leq s$ on $S$, which passes through at least $p$ of its nodes. It follows that $p \leq s(s-1)$. {\hfill$\square$} \end{proof} \block{Words of weight < 24 in {\it C}} {\bf From now on we will assume that {\it s} = 6.} We will want to get as much information as possible about the code $C$. First, an argument of Reid (\Lcitemark 4\Rcitemark \ 2.11) shows that all the weights of words in $C$ are divisible by $8$. Suppose that $w \in C$ is a nonzero word of weight $< 24$. Then $\abs{w} \leq 16$. Let $X = X_w$. By \pref{chi}, $\chi({\cal O}_X(1)) \geq 12$. By Serre duality (roughly), $h^0({\cal O}_X(1)) = h^2({\cal O}_X(1))$, so $h^0({\cal O}_X(1)) \geq 6$, contradicting \pref{mmmmm}. Hence $C$ has no nonzero words of weight $< 24$. \block{Properties of words of weight 24 in {\it C}}\label{weight-24-section} Let $w \in C_{24}$, and let $X = X_w$. By \pref{chi}, $\chi({\cal O}_X(1)) = 10$, and since $h^0({\cal O}_X(1)) = h^2({\cal O}_X(1))$ (as above), we have $h^0({\cal O}_X(1)) \geq 5$. By \pref{mmmmm}, $h^0({\cal O}_X(1)) = 5$. Hence $h^0({\cal O}_{\tilde{S}}(H-Q_w)) = 1$. In other words $\abs{H-Q_w}$ has a unique element, which we henceforth denote by $D_w$. We will show now that $\overline{D_w}$ is reduced. Since $\overline{D_w}$ passes through $\geq 24$ singular points of $S$, $\deg(\red{(\overline{D_w})}) \geq 5$. Suppose that $\overline{D_w}$ is nonreduced. Then $\deg(\red{(\overline{D_w})}) = 5$ so $\overline{D_w}$ contains a double line. Let $M$ be the remainder of $\overline{D_w}$, sans the doubled line. Since $\overline{D_w}$ must pass with {\it odd multiplicity\/} through exactly $24$ singular points of $S$, we conclude that $M$ must pass with odd multiplicity through exactly $24$ singular points of $S$. Since $\deg(M) = 4$, this is impossible. Hence $\overline{D_w}$ is reduced. Now there is a ``quadric surface'' $V_w \subset \xmode{\Bbb P\kern1pt}^3$ such that $2\overline{D_w} = V_w \cap S$ as Weil divisors on $S$. There are four possibilities for $V_w$. The worst case, that $V_w$ might be a doubled plane, can be ruled out, since then $\overline{D_w} = \red{(V_w)} \cap S$ as Weil divisors on $S$, so $\overline{D_w}$ is Cartier on $S$, which yields a contradiction. Unfortunately, we do not know if $V_w$ can be the union of two distinct planes, a possibility which substantially complicates the proofs. For starters, we must analyze this case. We could rule it out if we could show that a cubic curve on $S$ cannot pass through $15$ singular points of $S$. \begin{prop}\label{two-planes} Let $w \in C_{24}$. Suppose that $V_w$ is the union of two distinct planes $H_1$, $H_2$. For each $i$, let $\overline{D_i} = \red{(H_i \cap S)}$. Then: \begin{alphalist} \item for each $i$, $\overline{D_i}$ passes through exactly $15$ singular points of $S$, and it passes with multiplicity one through each of these; \item of the $15$ singular points, exactly $12$ are in $w$, and the remaining $3$ lie on the line $H_1 \cap H_2$. Both $\overline{D_1}$ and $\overline{D_2}$ pass with multiplicity one through these $3$ nodes. \end{alphalist} \end{prop} \begin{proof} Let $D_i \subset {\tilde{S}}$ be the strict transform of $\overline{D_i}$, for each $i$. Let $L = H_1 \cap H_2$. Because $2|\deg(S)$ and $\overline{D_w}$ is reduced, we cannot have $L \subset S$. Hence $D_1$ meets $D_2$ properly. Note that $\deg(D_i) = 3$. Let $Q_i = {1\over2} \sum_E (D_i \cdot E) E$, as $E$ ranges over the exceptional curves of ${\tilde{S}}$. Then $Q$ is a ${\Bbb Q}\kern1pt$-divisor, but (caution) it is not a divisor. We have $D_i \sim_{\Bbb Q}\kern1pt {1\over2} H - Q_i$. Suppose that $\overline{D_1}$ has only nodes as singularities. Then $\chi({\cal O}_{D_1})$ is equal to the number (say $n$) of such singular points which are resolved by $\pi$. Then $D_1^2 = -2n-6$, from which it follows that $Q_1$ has the form $${1\over2}\left(\svec E1{15} + \sum_{i=1}^n 2E_i'\right),$% $where $\vec E1{15}, E_1',\ldots,E_n'$ are distinct exceptional curves. But $\overline{D_1}$ has degree $3$, and so can pass through at most $3(\deg(S)-1) = 15$ singular points of $S$. Hence $n = 0$, so (a) holds, for $D_1$. In the three cases where $\overline{D_1}$ has a singularity other than a node, arguments similar to those of the last paragraph still yield the conclusion that (a) holds, for $D_1$. Of course, the results of the last two paragraphs apply equally to $\overline{D_2}$. Since $\overline{D_w}$ cannot pass with multiplicity one through a node outside $w$, it follows that the number of nodes of $w$ (say $r$) which $\overline{D_i}$ passes through is independent of $i$. Since $\overline{D_1}$ and $\overline{D_2}$ must together pass with multiplicity one through all $24$ of the singular points in $w$, we must have $r = 12$. Statement (b) follows. {\hfill$\square$} \end{proof} \begin{lemma}\label{line-through-four} Let $L \subset S$ be a line. Let $w \in C$. Then $L$ can pass through at most $4$ singular points of $S$ which are in $w$. \end{lemma} \begin{proof} Let ${\tilde{L}} \subset {\tilde{S}}$ be the strict transform of $L$. Certainly $L$ can pass through at most $5$ singular points of $S$. But ${\tilde{L}} \cdot Q_w \in \xmode{\Bbb Z}$, so the lemma follows. {\hfill$\square$} \end{proof} \begin{prop}\label{empty-intersection} Let $w_1, w_2 \in C$, and assume that $\abs{w_1} = \abs{w_2} = \abs{w_1+w_2} = 24$. Then either (a) $D_{w_1} \cap D_{w_2} = \varnothing$, or else (b) we can write $V_{w_1} = H_1 \cup H_2$ and $V_{w_2} = H_2 \cup H_3$, where $H_1, H_2, H_3$ are planes. \end{prop} \begin{proof} First note that $(H-Q_{w_1}) \cdot (H-Q_{w_2}) = 6 - {1\over2}(12) = 0$, so for case (a) it suffices to show that $D_{w_1}$ and $D_{w_2}$ share no component. Let $M$ be the shared part of $D_{w_1}$ and $D_{w_2}$. We may assume that $V_{w_1}$ meets $V_{w_2}$ properly, since otherwise conclusion (b) holds. Then $\deg(V_{w_1} \cap V_{w_2}) = 4$. We have ${\overline{M}} \subset V_{w_1} \cap V_{w_2}$, and since $V_{w_i}$ is tangent to $S$ along ${\overline{M}}$ for each $i$, in fact $V_{w_1}$ is tangent to $V_{w_2}$ along ${\overline{M}}$. Hence $2{\overline{M}} \subset V_{w_1} \cap V_{w_2}$. Hence $\deg(M) \leq 2$. If $\deg(M) = 2$, then $V_{w_1} \cap V_{w_2} = 2{\overline{M}}$. Hence ${\overline{M}}$ passes through all $12$ singular points in $w_1 \cap w_2$. But $\deg({\overline{M}}) = 2$, so ${\overline{M}}$ can pass through at most $2[\deg(S)-1] = 10$ singular points of $S$: contradiction. Hence $\deg(M) \leq 1$. We have $$0\ \leq\ (H-Q_{w_1}-M) \cdot (H-Q_{w_2}-M) = M(Q_{w_1}+Q_{w_2}-2H)+M^2. \eqno(*)$% $We will complete the proof by showing that if $M \not= 0$, then the right hand side\ of $(*)$ is negative. Write $M = R+E$, where $R$, $E$ are effective, $E$ is exceptional, and $R$ has no exceptional components. Expanding out the right hand side\ of $(*)$ yields $$(R+E)\cdot(Q_{w_1}+Q_{w_2}) - 2\deg(R) + R^2 + 2(E \cdot R) + E^2.$% $If $R=0$, then $E=0$, or else this quantity is negative. Hence $\deg(R) = 1$. By the adjunction formula, $R^2 = -4$. Therefore to get a contradiction, it is enough to show that $$(R+E) \cdot (Q_{w_1} + Q_{w_2}) + 2(E \cdot R) + E^2 < 6.$% $By \pref{line-through-four}, $R \cdot (Q_{w_1} + Q_{w_2}) \leq 4$, so it is enough to show that $$E \cdot (Q_{w_1} + Q_{w_2}) + 2(E \cdot R) + E^2 < 2.$% $Thus it is sufficient to show that if $F$ is an exceptional curve, then for all $n \in \xmode{\Bbb N}$, $$nF \cdot (Q_{w_1} +Q_{w_2}) + 2(nF \cdot R) + (nF)^2 \leq 0.$% $Clearly it is enough to do the case $n=1$, so we need $$F \cdot(Q_{w_1} + Q_{w_2}) + 2(F \cdot R) \leq 2.$% $Since $F \cdot R \leq 1$, this is clear. {\hfill$\square$} \end{proof} \block{Words of weight 48 in {\it C}} We will show that $C$ has no words of weight $48$. So let $v \in C_{48}$, heading for a contradiction. Write $Q$ for $Q_v$, $X$ for $X_v$. The first step is to show that $h^0({\cal O}_{\tilde{S}}(2H-Q)) = 1$. Let $N \subset {\Bbb F}\kern1pt_2^{48}$ be $\setof{w \in C: w \subset v}$. We have $h^0({\cal O}_{\tilde{S}}(H-Q)) = 0$, since otherwise there is a curve of degree $6$ on $S$ which passes through $\geq 48$ singular points of $S$. Hence $h^0({\cal O}_X(1)) = 4$. By \pref{chi} we have $\chi({\cal O}_X) = 10$, so $h^2({\cal O}_X) - h^1({\cal O}_X) = 9$. But $h^2({\cal O}_X) \geq h^2({\cal O}_{\tilde{S}}) = 10$, so $h^1({\cal O}_X) \geq 1$. Hence $\dim_{{\Bbb F}\kern1pt_2}(\mathop{\operatoratfont Pic}\nolimits(X)_2) \geq 2$. Hence by (\Lcitemark 3\Rcitemark \ Lemma 2), $\dim(N) \geq 3$. Since we know that $w \in N-\setof{0}\ \Longrightarrow\ \abs{w} \geq 24$, we in fact have $\setof{\abs{w}: w \in N} = \setof{0,24,48}$. Choose $w_1, w_2 \in N$ with $\abs{w_1} = \abs{w_2} = 24$ and $w_1 \cap w_2 = \varnothing$. By the first paragraph of \S\ref{weight-24-section}, $h^0({\cal O}_{\tilde{S}}(H-Q_{w_i})) \not= 0$. Since $v = w_1 + w_2$, $h^0({\cal O}_{\tilde{S}}(2H-Q)) \not= 0$. By Serre duality, $h^2({\cal O}_{\tilde{S}}(Q)) \not= 0$. Hence $h^2({\cal O}_{\tilde{S}}(-Q)) \not= 0$. Then $h^2({\cal O}_X) = h^2({\cal O}_{\tilde{S}}) + h^2({\cal O}_{\tilde{S}}(-Q)) \geq 11$, so $h^1({\cal O}_X) \geq 2$. Hence $\dim(N) \geq 5$. Suppose that $h^0({\cal O}_{\tilde{S}}(2H-Q)) > 1$. Then (arguing as above) $h^1({\cal O}_X) \geq 3$, so $\dim(N) \geq 7$. Hence there exists a $[48,7,\setof{24,48}]$ code, which contradicts the Griesmer bound for codes. (One also gets a contradiction by the linear programming method.) Hence $h^0({\cal O}_{\tilde{S}}(2H-Q)) = 1$. Let $w \in N_{24}$. Since $h^0({\cal O}_{\tilde{S}}(2H-Q)) = 1$, it follows that $D_w + D_{v+w}$ is independent of $w$. Since $\dim(N) \geq 3$, we can find $w_1, w_2 \in N$ with $\abs{w_1} = \abs{w_2} = \abs{w_1+w_2} = 24$. We will show that $V_{w_1}$ does not meet $V_{w_2}$ properly. Suppose otherwise. We have $D_{w_1} + D_{v+w_1} = D_{w_2} + D_{v+w_2}$. By \pref{empty-intersection}, $D_{w_1} \cap D_{w_2} = \varnothing$. Hence $D_{w_1} \subset D_{v+w_2}$, so $\overline{D_{w_1}} = \overline{D_{v+w_2}}$, so $w_1 = v+w_2$: contradiction. Hence $V_{w_1}$ does not meet $V_{w_2}$ properly, i.e.\ each consists of two planes, with one in common. The same holds for the pair $(V_{w_1}, V_{v+w_2})$, the pair $(V_{v+w_1}, V_{w_2})$, and the pair $(V_{v+w_1}, V_{v+w_2})$. We consider now the configuration $\cal C$ of planes which occur in $V_{w_1}$, $V_{w_2}$, $V_{v+w_1}$, and $V_{v+w_2}$. In total, there are four planes. First suppose that no three of these planes share a common line. Then we may visualize the configuration by means of a tetrahedron $T$. (The faces correspond to planes.) For any pair of faces, the intersection of the corresponding two planes with $S$ gives rise to a word $w$ of weight $24$, and so \pref{two-planes} applies. Therefore every edge of $T$ has exactly $3$ singular points of $S$ on it, and every face of $T$ has exactly $15$ singular points of $S$ on it. Here is the generic picture, showing only the singular points which appear on the edges: \vspace{0.1in} \par\noindent{\bf[There was a postscript picture of a tetrahedron here, with three little spheres along each edge.]} \vspace{0.1in} \par\noindent The picture is generic because some of the edge singular points could be at the vertices. Restricting to the generic case for the moment, let us compute the total number of singular points of $S$ which lie on $T$. Since each face contains $15$, the interior of each face has $6$. Hence $$\abs{\mathop{\operatoratfont Sing}\nolimits(S) \cap T}\ =\ (4 \cdot 6) + (6 \cdot 3)\ =\ 42.$% $But $T$ must go through at least $\abs{v} = 48$ singular points of $S$: contradiction. In the nongeneric cases, one similarly gets a contradiction. (In the least generic case, where all $4$ vertices of $T$ are in $\mathop{\operatoratfont Sing}\nolimits(S)$, one gets $\abs{\mathop{\operatoratfont Sing}\nolimits(S) \cap T} = 46$, which still gives a contradiction.) Hence at least $3$ of the planes in $\cal C$ must share a common line. Suppose now that the four planes in ${\cal{C}}$ do not all share a common line. Taking a cross section by a suitable plane, we may represent each plane by a line segment and each intersection of two planes by a vertex. (The single point where all planes meet is not shown.) In the generic case, we can label each vertex in the picture with a $3$. \widepost{85}{25}{ newpath 0 -15 moveto -50 50 lineto stroke newpath 0 -15 moveto 50 50 lineto stroke newpath 0 -15 moveto 0 50 lineto stroke newpath -50 50 moveto 50 50 lineto stroke 0 -15 PointAt -50 50 PointAt 50 50 PointAt 0 50 PointAt {6 270 (3) AnglePrint} 0 -15 xyput {6 90 (3) AnglePrint} 0 50 xyput {6 135 (3) AnglePrint} -50 50 xyput {6 45 (3) AnglePrint} 50 50 xyput} \par\noindent One finds that the four planes together pass through at most $45$ singular points of $S$: contradiction. In the nongeneric case, one finds that ${\cal{C}}$ passes through $47$ points of $\mathop{\operatoratfont Sing}\nolimits(S)$, still shy of $48$. Now we know that the four planes in ${\cal{C}}$ must contain a common line. It follows that there exists a line $L$ such that for any $w \in N_{24}$, $L$ is contained in both planes of $V_w$. Let ${\cal{C}}^+$ be the configuration of all planes which appear in $V_w$ for some $w \in N_{24}$. Since $\dim(N) \geq 5$, $\abs{N_{24}} \geq 30$. Since ${8 \choose 2} < 30$, ${\cal{C}}^+$ must have at least $9$ planes in it. Hence ${\cal{C}}^+$ must pass through at least $(9 \cdot 12) + 3 = 111$ singular points of $S$, which is absurd. We conclude that $C$ has no word of weight $48$. \block{Apply coding theory} Suppose now that $S$ has $66$ nodes. Since $H^2(S,\xmode{\Bbb Z}) \cong \xmode{\Bbb Z}^{106}$, it follows\Lspace \Lcitemark 3\Rcitemark \Rspace{} that $\dim(C) \geq 66 - 106/2 = 13$. We know that the nonzero weights appearing in $C$ are all in the set \setof{24,32,40,56,64}. Now we show that weight $64$ must occur and that weight $56$ cannot occur. \begin{theorem}\label{code-doesnt-exist} There is no $[66,13,\setof{24,32,40,56}]$ code. \end{theorem} \begin{proof} Let $R = \setof{24,32,40,56}$ and $T = \setof{4,8,12,16,20}$. By the linear programming method (LP), there is no $[21,10,T]$ code. (See the appendix for some {\sc Maple} code to do calculations like this.) Therefore by taking the residual code with respect to\ a codeword of weight $40$ (hereafter referred to as res!40), one sees that there is no $[61,11,\setof{24,32,\underline{40},56}]$ code. Let $C$ be a $[58,11,\setof{24,32,56}]$ code. By LP, $C_{56} = 1$. By LP, $C^\perp_1 = 2$. By LP, $C^\perp_2 = 6.5$: contradiction. Hence there is no such code $C$. Now let $C$ be a $[59,11,\setof{24,32,56}]$ code. Then $C^\perp_1 = 0$. Then LP yields a contradiction, so there is no such code $C$. Similarly, one sees that there is no $[60,11,\setof{24,32,56}]$ code, and finally in the same way that $(*)$ there is no $[61,11,\setof{24,32,56}]$ code. By the preceeding paragraph, we conclude $(\dag)$ that there is no $[61,11,R]$ code. By LP, there is no $[22,11,T]$ code. Therefore by res!40, there is no $[62,12,\setof{24,32,\underline{40},56}]$ code. By LP, there is no $[62,12,\setof{24,32,56}]$ code $C$ with $C^\perp_1 = 0$, and therefore by $(*)$ of the preceeding paragraph, there is no $[62,12,\setof{24,32,56}]$ code at all. Hence there is no $[62,12,R]$ code. {}From $(\dag)$ we see that a $[63,13,R]$ code $C$ must have $C^\perp_2 = C^\perp_3 = 0$. By LP, one concludes that there is no $[63,13,R]$ code. {}From the preceeding three paragraphs we see that a $[64,13,R]$ code $C$ must have $C^\perp_1 = C^\perp_2 = C^\perp_3 = 0$. By LP, one concludes that a $[64,13,R]$ code $C$ must have $C_{56} = 2.5$: contradiction. Hence there is no such code $C$. Let $C$ be a $[65,13,R]$ code. Then $C^\perp_1 = 0$. By LP, $C^\perp_2 \geq 5$. Consider two distinct words of weight two in $C^\perp$. If they are disjoint [resp.\ not disjoint], then there is a $[61,11,R]$ code [resp.\ $[62,12,R]$ code], in either case contradicting earlier results. Hence there is no $[65,13,R]$ code. Let $C$ be a $[66,13,R]$ code. (This will lead to a contradiction.) Then $C^\perp_1 = 0$. Form a graph $G$ whose vertices correspond to the coordinate positions covered by the words of weight $2$ in $C^\perp$, and such that the edges of $G$ correspond to the words of weight $2$ in $C^\perp$. Then $G$ is a disjoint union of complete graphs. By LP, $C^\perp_2 \geq 7$, so $G$ has at least $7$ edges. Now $G$ has no $K_4$ component, because then one would have a $[62,12,R]$ code. Similar arguments show that $G$ cannot have a $K_n$ (for $n \geq 5$), it cannot have $2$ $K_3$'s, and it cannot have a $K_3$ and a $K_2$. Therefore, since $G$ has at least $7$ edges, it must consist entirely of $K_2$'s. Otherwise said, the words of weight $2$ in $C^\perp$ are pairwise disjoint. If $C$ has a word $v$ of weight $56$, it occupies all but $10$ bits of $C$, and so at most $5$ of the words of weight $2$ in $C^\perp$ are disjoint from $v$. Thus we can choose some $w \in C^\perp_2$ with the property that if such a word $v$ exists, then $w \subset v$. Let $D$ be the subcode of $C$ consisting of words disjoint from $w$. Then $D$ is a $[64,12,\setof{24,32,40}]$ code and $D^\perp$ has at least $7-1=6$ words of weight $2$. Adjoin the unique word of weight $64$ to $D$, yielding a $[64,13,\setof{24,32,40,64}]$ code $E$ with $E^\perp_2 \not= 0$. Hence there exists a $[62,12,\setof{24,32,40}]$ code: contradiction. {\hfill$\square$} \end{proof} We conclude that $C$ must be a $[66,13,\setof{24,32,40,\underline{64}}]$ code. Unfortunately, such a code exists. \vspace{0.1in} \par\noindent{\bf Appendix} \vspace{0.1in} Here we give some {\sc Maple} code which can be used to carry out the linear programming calculations which appear in the proof of \pref{code-doesnt-exist}. The first four lines can be adjusted to check for particular types of codes. As written, the program will analyze $[64,13,\setof{24,32,40,56}]$ codes $C$ with $C^\perp_1 = C^\perp_2 = C^\perp_3 = 0$. If you run it, you will find that such a code must have $C_{56} = 2.5$ (which is absurd), as claimed in the proof of \pref{code-doesnt-exist}. \vspace{0.2in} \begin{verbatim} n := 64; k := 13; w := {24,32,40,56}; extraconstraints := [mu1=0, mu2=0, mu3=0]; K := (x,i) -> sum( (-1)^j * binomial(i,j) * binomial(n-i,x-j), j=0..x ): for m from 0 to n/2 do mu.m := (K(m,0) + convert(map(i -> 'a.i' * K(m,i), w), `+`)) / 2^k: od: objective := 1 + convert(map(z -> a.z, w), `+`); mus := {seq( mu.m >= 0, m=0..n/2 ), objective = 2^k, op(extraconstraints)}: for i from 1 to 2 do M := simplex[minimize](mu.i,mus,NONNEGATIVE); `minimum value of mu`.i, evalf(subs(M,mu.i)); M := simplex[maximize](mu.i,mus,NONNEGATIVE); `maximum value of mu`.i, evalf(subs(M,mu.i)); od; for q in w do M := simplex[minimize](a.q,mus,NONNEGATIVE); q, `min`, evalf(subs(M,a.q)); M := simplex[maximize](a.q,mus,NONNEGATIVE); q, `max`, evalf(subs(M,a.q)); od; quit(): \end{verbatim} \message {DOCUMENT TEXT} \section*{References} \addcontentsline{toc}{section}{References} \ \par\noindent\vspace*{-0.25in} \hfuzz 5pt \message{REFERENCE LIST} \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{4}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{1}% \def\Atest{ }\def\Astr{Arbarello\Revcomma E\Initper % \Acomma M\Initper Cornalba% \Acomma P\Initper \Initgap A\Initper Griffiths% \Aandd J\Initper Harris}% \def\Ttest{ }\def\Tstr{Geometry of Algebraic Curves, Vol.\ I}% \def\Itest{ }\def\Istr{Spring\-er-Ver\-lag}% \def\Ctest{ }\def\Cstr{New York}% \def\Dtest{ }\def\Dstr{1985}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{3}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{2}% \def\Ven{Van de Ven}{}% \def\Atest{ }\def\Astr{Barth\Revcomma W\Initper % \Acomma C\Initper Peters% \Aandd A\Initper \Ven}% \def\Ttest{ }\def\Tstr{Compact Complex Surfaces}% \def\Itest{ }\def\Istr{Spring\-er-Ver\-lag}% \def\Ctest{ }\def\Cstr{New York}% \def\Dtest{ }\def\Dstr{1984}% \def\Qtest{ }\def\Qstr{access via "barth peters ven"}% \def\Xtest{ }\def\Xstr{local-triv: I 10.1 (theorem of Grauert-Fischer)}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{1}\def\Ecnt{1}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{3}% \def\Atest{ }\def\Astr{Beauville\Revcomma A\Initper }% \def\Ttest{ }\def\Tstr{Sur le nombre maximum de points doubles d'une surface in $\xmode{\Bbb P\kern1pt}^3$ $(\mu(5) = 31)$}% \def\Btest{ }\def\Bstr{Algebraic Geometry, Angers, 1979}% \def\Etest{ }\def\Estr{A\Initper Beauville}% \def\Itest{ }\def\Istr{Sijthoff \& Noordhoff}% \def\Dtest{ }\def\Dstr{1980}% \def\Ptest{ }\def\Pcnt{ }\def\Pstr{207--215}% \def\Qtest{ }\def\Qstr{access via "beauville double points"}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{1}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{4}% \def\Atest{ }\def\Astr{Catanese\Revcomma F\Initper }% \def\Ttest{ }\def\Tstr{Babbage's conjecture, contact of surfaces, symmetric determinantal varieties and applications}% \def\Jtest{ }\def\Jstr{Invent. Math.}% \def\Vtest{ }\def\Vstr{63}% \def\Dtest{ }\def\Dstr{1981}% \def\Ptest{ }\def\Pcnt{ }\def\Pstr{433--465}% \def\Qtest{ }\def\Qstr{access via "catanese contact"}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{1}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{5}% \def\Atest{ }\def\Astr{Jaffe\Revcomma D\Initper \Initgap B\Initper }% \def\Ttest{ }\def\Tstr{Applications of iterated curve blowup to set theoretic complete intersections in $\xmode{\Bbb P\kern1pt}^3$}% \def\Rtest{ }\def\Rstr{preprint}% \def\Qtest{ }\def\Qstr{access via "jaffe rational singularities" or preferably "jaffe applications intersections"}% \def\Htest{ }\def\Hstr{7}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{2}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{6}% \def\Atest{ }\def\Astr{MacWilliams\Revcomma F\Initper \Initgap J\Initper % \Aand N\Initper \Initgap J\Initper \Initgap A\Initper Sloane}% \def\Ttest{ }\def\Tstr{The Theory of Error-Correcting Codes}% \def\Itest{ }\def\Istr{North-Holland}% \def\Ctest{ }\def\Cstr{Amsterdam}% \def\Dtest{ }\def\Dstr{1977}% \def\Qtest{ }\def\Qstr{access via "macwilliams sloane"}% \Refformat\egroup% \bgroup\Resetstrings% \def\Loccittest{}\def\Abbtest{ }\def\Capssmallcapstest{}\def\Edabbtest{ }\def\Edcapsmallcapstest{}\def\Underlinetest{ }% \def\NoArev{1}\def\NoErev{0}\def\Acnt{1}\def\Ecnt{0}\def\acnt{0}\def\ecnt{0}% \def\Ftest{ }\def\Fstr{7}% \def\Atest{ }\def\Astr{Pinkham\Revcomma H\Initper }% \def\Ttest{ }\def\Tstr{Singularites rationnelles de surfaces}% \def\Btest{ }\def\Bstr{S\'eminaire sur les Singularit\'es des Surfaces}% \def\Stest{ }\def\Sstr{Lecture Notes in Mathematics}% \def\Itest{ }\def\Istr{Spring\-er-Ver\-lag}% \def\Ctest{ }\def\Cstr{New York}% \def\Vtest{ }\def\Vstr{777}% \def\Dtest{ }\def\Dstr{1980}% \def\Ptest{ }\def\Pcnt{ }\def\Pstr{147--178}% \def\Qtest{ }\def\Qstr{access via "pinkham rational singularities"}% \Refformat\egroup% \end{document}
1996-03-31T05:41:55
9411
alg-geom/9411010
en
https://arxiv.org/abs/alg-geom/9411010
[ "alg-geom", "math.AG" ]
alg-geom/9411010
Miles Reid
Yukari Ito and Miles Reid
The McKay correspondence for finite subgroups of SL(3,\C)
AMSTeX, amsppt and optional epsf.tex , This paper will appear in Higher Dimensional Complex Varieties (Trento, Jun 1994), ed. M. Andreatta, De Gruyter, Mar 1996. It has been circulated as a Univ. of Tokyo, Dept. of Math Sciences preprint, UTMS 94--66, 19 pp
null
null
null
null
This is the final draft, containing very minor proof-reading corrections. Let G in SL(n,\C) be a finite subgroup and \fie: Y -> X = \C^n/G any resolution of singularities of the quotient space. We prove that crepant exceptional prime divisors of Y correspond one-to-one with ``junior'' conjugacy classes of G. When n = 2 this is a version of the McKay correspondence (with irreducible representations of G replaced by conjugacy classes). In the case n = 3, a resolution with K_Y = 0 is known to exist by work of Roan and others; we prove the existence of a basis of H^*(Y, \Q) by algebraic cycles in one-to-one correspondence with conjugacy classes of G. Our treatment leaves lots of open problems.
[ { "version": "v1", "created": "Wed, 16 Nov 1994 14:00:39 GMT" }, { "version": "v2", "created": "Wed, 10 Jan 1996 09:13:15 GMT" } ]
2008-02-03T00:00:00
[ [ "Ito", "Yukari", "" ], [ "Reid", "Miles", "" ] ]
alg-geom
1994-11-16T06:20:22
9411
alg-geom/9411009
en
https://arxiv.org/abs/alg-geom/9411009
[ "alg-geom", "math.AG" ]
alg-geom/9411009
V. Batyrev
Victor V. Batyrev and Yuri Tschinkel
Rational Points of Bounded Height on Compactifications of Anisotropic Tori
45 pages, Latex
null
null
null
null
We investigate the analytic properties of the zeta-function associated with heights on equivariant compactifications of anisotropic tori over number fields. This allows to verify conjectures about the distribution of rational points of bounded height.
[ { "version": "v1", "created": "Tue, 15 Nov 1994 17:59:52 GMT" } ]
2008-02-03T00:00:00
[ [ "Batyrev", "Victor V.", "" ], [ "Tschinkel", "Yuri", "" ] ]
alg-geom
\section{Toric varieties over arbitrary fields} \subsection{Algebraic tori} Let $K$ be an arbitrary field, $\overline{K}$ the algebraic closure of $K$, ${\bf G}_m(\overline{K}) = \overline{K}^*$ the multiplicative group of $\overline{K}$. Let $X$ be an arbitrary algebraic variety over $\overline{K}$. Let $E/K$ be a finite extension such that $X$ is defined over $E$. To stress this fact we sometimes will denote $X$ also by $X_E$. The set of $E$-rational points of $X_E$ will be denoted by $X_E(E)$. \begin{dfn} {\rm A linear algebraic group $T$ over $K$ is called a {\em $d$-dimen\-sio\-nal algebraic torus} if its base extension $T_{\overline{K}} = T \times_{{\rm Spec}(K)} {\rm Spec }(\overline{K})$ is isomorphic to $({\bf G}_m(\overline{K}))^d$. } \label{opr.tori} \end{dfn} We notice that an isomorphism between ${T}$ and $({\bf G}_m(\overline{K}))^d$ is always defined over a finite Galois extension $E$ of $K$. \begin{dfn} {\rm Let $T$ be an algebraic torus over $K$. A finite Galois extension $E$ of $K$ such that $T_E = T \times_{{\rm Spec}(K)} {\rm Spec}(E)$ is isomorphic to $({\bf G}_m(E))^d$ is called a {\em splitting field } of $T$.} \end{dfn} \begin{dfn} {\rm We denote by $\hat{T} = {\rm Hom}\,( T, \overline{K}^*)$ the group of regular $\overline{K}$-rational characters of $T$. For any subfield $E \subset \overline{K}$ containing $K$, we denote by $\hat{T}_E$ the group of characters of $T$ defined over $E$. } \end{dfn} There is well-known correspondence between Galois representations by integral matrices and algebraic tori \cite{ono1,vosk}: \begin{theo} Let $G = {\rm Gal }(E/K)$ be the Galois group of the splitting field $E$ of a $d$-dimen\-sional torus $T$. Then ${\hat{T}}$ is a free abelian group of rank $d$ with a structure of $G$-module defined by the natural representation \medskip \[ \rho \; :\; G \rightarrow {\rm Aut}(\hat{T}) \cong {\rm GL}(d, {\bf Z}).\] Every d-dimensional integral representation of $G$ defines a $d$-dimen\-sional algebraic torus over $K$ which splits over $E$. One obtains a one-to-one correspondence between $d$-dimensional algebraic tori over $K$ with the splitting field $E$ up to isomorphism, and $d$-dimensional integral representations of $G$ up to equivalence. \label{represent} \end{theo} \begin{rem} {\rm The group $\hat{T}_K$ is a sublattice in $\hat{T} \cong {\bf Z}^d$ consisting of all $G$-invariant elements.} \label{prop.char} \end{rem} \begin{dfn} {\rm An algebraic torus $T$ over $K$ is called {\em anisotropic} if $\hat{T}_K$ has rank zero.} \end{dfn} \begin{exam} {\rm Let $f(z) \in K\lbrack z \rbrack$ be a separable polynomial of degree $d$. Consider the $d$-dimensional $K$-algebra \[ A(f) = K \lbrack z \rbrack / (f(z)). \] Then the multiplicative group $A^*(f)$ is a $d$-dimensional algebraic torus over $K$. This torus has the following properties: (i) The rank of the group of characters of $A^*(f)$ is equal to the number of irreducible components of ${\rm Spec}(A(f))$. (ii) If $f(z)$ splits in linear factors over some finite Galois extension $E$ of $K$, then $A(f) \otimes_K E \cong E^n$, and $A^*(f)\otimes_K E \cong (E^*)^n$. Thus, $E$ is a splitting field of $A^*(f)$. (iii) Since the classes of $1, z, \ldots , z^{d-1}$ in $A(f)$ give rise to a $K$-basis of the $d$-dimensional algebra $A(f)$, we can consider $A^*(f)$ as a commutative subgroup in ${\rm GL}(d,K)$. Thus, the determinant of the matrix defines a regular $K$-character \[{\cal N} \; : \; A^*(f) \rightarrow K^*. \] We denote by $A^*_1(f)$ the $(d-1)$-dimensional algebraic torus which is the kernel of ${\cal N}$. (iv) The multiplicative group $K^*$ is a subgroup of $A^*(f)$ and the restriction of ${\cal N}$ to $K^*$ sends $x \in K^*$ to $x^d \in K^*$. The factor-group $A^*(f) / K^*$ is a $(d-1)$-dimensional torus which is isogeneous to $A^*_1(f)$. } \label{exam.tori} \end{exam} \begin{exam} {\rm Let $K'$ be a finite separable extension of $K$. By primitive element theorem, $K' \cong A(f)$ for some irreducible polynomial $f(z) \in K \lbrack z \rbrack $. Thus, we come to a particular case of the previous example. In this case, ${\cal N}$ is the norm $N_{K'/K}$, the algebraic torus $A^*(f)$ is usually denoted by \[ R_{K'/K} ({\bf G}_m) , \] and the torus $A_1^*(f)$ is usually denoted by \[ R_{K'/K}^1 ({\bf G}_m) . \] Since ${\rm Spec}(K')$ is irreducible, $R_{K'/K}^1 ({\bf G}_m)$ and $R_{K'/K} ({\bf G}_m)/ K^*$ are examples of anisotropic tori.} \label{exam.field} \end{exam} \subsection{Compactifications of split tori} We recall standard facts about toric varieties over algebraically closed fields \cite{danilov,demasur,oda}. Let $M$ be a free abelian group of rank $d$ and $N = {\rm Hom}(M, {\bf Z})$ the dual abelian group. \begin{dfn} {\rm A finite set $\Sigma$ consisting of convex rational polyhedral cones in $N_{\bf R} = N \otimes {\bf R}$ is called a {\em complete regular $d$-dimensional fan} if the following conditions are satisfied: (i) every cone $\sigma \in \Sigma$ contains $0 \in N_{\bf R}$; (ii) every face $\sigma'$ of a cone $\sigma \in \Sigma$ belongs to $\Sigma$; (iii) the intersection of any two cones in $\Sigma$ is a face of both cones; (iv) $N_{\bf R}$ is the union of cones from $\Sigma$; (v) every cone $\sigma \in \Sigma$ is generated by a part of a ${\bf Z}$-basis of $N$.\\ We denote by $\Sigma(i)$ the set of all $i$-dimensional cones in $\Sigma$. For each cone $\sigma \in \Sigma$ we denote by $N_{{\sigma}, \bf R}$ the minimal linear subspace containing $\sigma$. } \label{def.fan} \end{dfn} \noindent Every complete regular $d$-dimensional fan defines a smooth equivariant compactification ${\bf P}_{\Sigma}$ of the split $d$-dimensional algebraic torus $T$. The variety ${\bf P}_{\Sigma}$ has the following two geometric properties: \begin{prop} The toric variety ${\bf P}_{\Sigma}$ is the union of split algebraic tori $T_{\sigma}$ $($${\rm dim}\, T_{\sigma} = d - {\rm dim}\, \sigma$$)$: \[ {\bf P}_{\Sigma} = \bigcup_{ \sigma \in \Sigma } T_{\sigma}. \] For each $k$-dimensional cone $\sigma \in \Sigma{(k)}$, $T_{\sigma}$ is the kernel of a homomorphism $T \rightarrow ({\bf G}_m(\overline{K})^k)$ defined by a ${\bf Z}$-basis of the sublattice $N \cap N_{{\sigma},{\bf R}} \subset N$. \end{prop} \noindent Let $\check{\sigma}$ denote the cone in $M_{\bf R}$ which is dual to $\sigma$. \begin{prop} The toric variety ${\bf P}_{\Sigma}$ has a $T$-invariant open covering by affine subsets $U_{\sigma}$: \[ {\bf P}_{\Sigma} = \bigcup_{ \sigma \in \Sigma} U_{\sigma} \] where $U_{\sigma} = {\rm Spec}(\overline{K}) \lbrack M \cap \check{\sigma} \rbrack$. \end{prop} \begin{dfn} {\rm A continuous function $\varphi\; : \; N_{\bf R} \rightarrow {\bf R}$ is called {\em $\Sigma$-piecewise linear} if the restriction $\varphi_{\sigma}$ of $\varphi$ to every cone $\sigma \in \Sigma$ is a linear function. It is called {\em integral} if $\varphi(N) \subset {\bf Z}$.} \end{dfn} \begin{dfn} {\rm For any integral $\Sigma$-piecewise linear function $\varphi\; : \; N_{\bf R} \rightarrow {\bf R}$ and any cone $\sigma \in \Sigma(d)$, we denote by $m_{\sigma, \varphi}$ the restriction of $\varphi$ to $\sigma$ considered as an element in $M$. We put $m_{\sigma', \varphi} = m_{\sigma, \varphi}$ if $\sigma'$ is a face of a $d$-dimensional cone $\sigma \in \Sigma$. } \end{dfn} \begin{dfn} {\rm For any integral $\Sigma$-piecewise linear function $\varphi\; : \; N_{\bf R} \rightarrow {\bf R}$, we define the invertible sheaf $L(\varphi)$ as the subsheaf of the constant sheaf of rational functions on ${\bf P}_{\Sigma}$ generated over $U_{\sigma}$ by the element $- m_{\sigma,\varphi}$ considered as a character of $T \subset {\bf P}_{\Sigma}$. } \end{dfn} \begin{rem} {\rm The $T$-action on the sheaf of rational functions restricts to the subsheaf $ L(\varphi)$ so that we can consider $ L(\varphi)$ as a $T$-linearized line bundle over ${\bf P}_{\Sigma}$. } \end{rem} Denote by $e_1, \ldots, e_n$ the primitive integral generators of all $1$-dimensional cones in $\Sigma$. Let $T_{i}$ $(i =1, \ldots, n)$ be the $(d-1)$-dimensional torus orbit corresponding to the cone ${\bf R}_{\geq 0}e_i \in \Sigma$ and $D_i$ the Zariski closure of $T_i$ in ${\bf P}_{\Sigma}$. Define ${\bf D}({\Sigma}) \cong {\bf Z}^n$ as the free abelian group of $T$-invariant Weil divisors on ${\bf P}_{\Sigma}$ with the basis $D_1, \ldots, D_n$. \begin{prop} The correspondence $\varphi \rightarrow L(\varphi)$ gives rise to an isomorphism between the group of $T$-linearized line bundles on ${\bf P}_{\Sigma}$ and the group $PL(\Sigma)$ of all $\Sigma$-piecewise linear integral functions on $N_{\bf R}$. There is the canonical isomorphism \[ PL(\Sigma) \cong {\bf D}(\Sigma), \;\; \varphi \mapsto (\varphi(e_1), \ldots, \varphi(e_n)). \] The Picard group ${\rm Pic}({\bf P}_{\Sigma})$ is isomorphic to $PL(\Sigma)/M$ where elements of $M$ are considered as globally linear integral functions on $N_{\bf R}$, so that we have the exact sequence \begin{equation} 0 \rightarrow M \rightarrow {\bf D}(\Sigma) \rightarrow {\rm Pic}({\bf P}_{\Sigma}) \rightarrow 0 \end{equation} \end{prop} \begin{dfn} {\rm Let $\Lambda_{\rm eff}(\Sigma)$ be the cone in ${\rm Pic}({\bf P}_{ \Sigma })$ generated by classes of effective divisors on ${\bf P}_{\Sigma}$. Denote by ${\Lambda}_{\rm eff}^*(\Sigma)$ the dual to $\Lambda_{\rm eff}(\Sigma)$ cone.} \end{dfn} \begin{prop} $\Lambda_{\rm eff}(\Sigma)$ is generated by the classes $\lbrack D_1 \rbrack, \ldots, \lbrack D_n \rbrack$. \label{generators} \end{prop} {\em Proof.} Any divisor $D$ on ${\bf P}_{\sigma}$ is linearly equivalent to an integral linear combination of $D_1, \ldots, D_n$. Assume that $D = a_1 D_1 + \cdots + a_n D_n$ is effective. Then there exists a rational function $f$ on ${\bf P}_{\Sigma}$ having no poles and zeros on $T$ such that \begin{equation} (f) + D \geq 0. \label{effective} \end{equation} We can assume that $f$ is character of $T$ defined by an element $m_f \in M$. Then the condition (\ref{effective}) is equivalent to \begin{equation} b_i = \langle m_f, e_i \rangle + a_i \geq 0,\; i =1, \ldots, n \end{equation} Then $D' = b_1 D_1 + \cdots + b_n D_n$ is linearly equivalent to $D$. So every effective class $\lbrack D \rbrack $ is a non-negative integral linear combination of $\lbrack D_1 \rbrack, \ldots, \lbrack D_n \rbrack$. \hfill $\Box$ \begin{prop} Let $\varphi_{\Sigma}$ be the $\Sigma$-piecewise linear integral function such that $\varphi(e_1) = \cdots = \varphi(e_n)= 1$. Then $L(\varphi_{\Sigma})$ is isomorphic to the $T$-linearized anticanonical line bundle on ${\bf P}_{\Sigma}$. \end{prop} \begin{exam} {\sl Projective spaces}. {\rm Consider a $d$-dimensional fan $\Sigma$ whose $1$-dimensional cones are generated by $d+1$ elements $e_1, \ldots , e_{d}, e_{d+1} = -(e_1 + \cdots + e_{d})$, where $\{ e_1, \ldots , e_{d} \}$ is a ${\bf Z}$-basis of ${d}$-dimensional lattice $N$, and $k$-dimensional cones in $\Sigma$ are generated by all possible $k$-element subsets in $\{ e_1,\ldots, e_{d+1} \}$. Then the corresponding compactification ${\bf P}_{\Sigma}$ of the $d$-dimensional split torus is ${\bf P}^d$.} \label{proj.space} \end{exam} \begin{rem} {\rm It is easy to see that the combinatorial construction of toric varieties ${\bf P}_{\Sigma}$ immediatelly extends to arbitrary fields $E$; i.e., using a rational complete polyhedral fan $\Sigma$, one can define the toric variety ${\bf P}_{\Sigma,E}$ as the equivariant compactification of the split torus $({\bf G}_m(E))^d$. } \end{rem} \subsection{Compactifications of nonsplit tori} Let $T$ be a $d$-dimensional algebraic torus over $K$ with a splitting field $E$ and $G = {\rm Gal}\, (E/K)$. Denote by $M$ the lattice $\hat{T}$ and put $N ={\rm Hom}\, (M, {\bf Z})$. Let $\rho^*$ be the integral representation of $G$ in ${\rm GL}(N)$ which is dual to $\rho$. In order to construct a projective compactification of $T$ over $K$, we need a complete fan $\Sigma$ of cones having an additional combinatorial structure: an {\em action of the Galois group } $G$ \cite{vosk}: \begin{dfn} {\rm A complete fan $\Sigma \subset N_{\bf R}$ is called {\em $G$-invariant} if for any $g \in G$ and for any $\sigma \in \Sigma$, one has $\rho^*(g) (\sigma) \in \Sigma$. } \label{opr.invar} \end{dfn} \begin{theo} Let $\Sigma$ be a complete regular $G$-invariant fan in $N_{\bf R}$. Then there exists a complete algebraic variety ${\bf P}_{\Sigma,K}$ over $K$ such that its base extension ${\bf P}_{\Sigma,K} \otimes_{{\rm Spec}\, K} {\rm Spec}\, E$ is isomorphic to the toric variety ${\bf P}_{\Sigma,E}$ defined over $E$ by $\Sigma$. Let $\Sigma^G$ be the subset of all $G$-invariant cones $\sigma \in \Sigma$. Then \[ {\bf P}_{\Sigma}(K) = \bigcup_{\sigma \in \Sigma^G} T_{\sigma}(K), \] where $T_{\sigma}$ is the $(d - {\rm dim}\, \sigma)$-dimensional algebraic torus over $K$ corresponding to the restriction of the integral $G$-representation in ${\rm GL}(M)$ to the sublattice $(\hat{\sigma} \cap - \hat{\sigma}) \cap M \subset M$. \label{decompos} \end{theo} Taking $G$-invariant elements in the short exact sequence \[ 0 \rightarrow M \rightarrow {\bf D}(\Sigma) \rightarrow {\rm Pic}({\bf P}_{\Sigma,E}) \rightarrow 0 \] we obtain the exact sequence \begin{equation} 0 \rightarrow M^G \rightarrow {\bf D}(\Sigma)^G \rightarrow {\rm Pic}({\bf P}_{\Sigma,E})^G \rightarrow H^1(G,M) \rightarrow 0 \label{short3} \end{equation} \begin{prop} The group ${\rm Pic}({\bf P}_{\Sigma,E})^G$ is canonically isomorphic to the Picard group ${\rm Pic}({\bf P}_{\Sigma,K})$. Moreover $H^1(G, M)$ is the Picard group of $T$. \end{prop} \begin{coro} The correspondence $\varphi \rightarrow L(\varphi)$ induces an isomorphism between the group of $T$-linearized invertible sheaves on ${\bf P}_{\Sigma,K}$ and the group $PL(\Sigma)^G$ of all $\Sigma$-piecewise linear integral $G$-invariant functions on $N_{\bf R}$. An invertible sheaf $ L$ on ${\bf P}_{\Sigma,K}$ admits a $T$-linearization if and only if the restriction of $L$ on $T$ is trivial. In particular, some tensor power of $ L$ always admits a $T$-linearization. \end{coro} \begin{coro} Let $\Lambda_{\rm eff}(\Sigma,K)$ be the cone of effective divisors of ${\bf P}_{\Sigma,K}$. Then $\Lambda_{\rm eff}(\Sigma,K)$ consists of $G$-invariant elements in $\Lambda_{\rm eff}(\Sigma)$. \end{coro} \begin{coro}{\rm \cite{colliot1}} Let ${\bf P}_{\Sigma,K}$ be a compactification of an anisotropic torus $T$. Then all $K$-rational points of ${\bf P}_{\Sigma,K}$ are contained in $T$ itself. \label{point.aniso} \end{coro} {\em Proof.} By \ref{decompos}, it is sufficient to prove that for an anisotropic torus $T$ defined by some Galois representation of $G$ in ${\rm GL}(M)$, there is no $G$-invariant cone $\sigma$ of positive dimension in $\Sigma$. Assume that a $k$-dimensional cone $\sigma$ with the generators $\{ e_{i_1}, \dots , e_{i_k} \}$ is $G$-invariant. Then $ e_{i_1} + \cdots + e_{i_k}$ is a nonzero $G$-invariant integral vector in the interior of $\sigma$. Hence the sublattice $N^G$ of $G$-invariant elements in $N$ has positive rank. Thus $M^G \cong \hat{T}_K$ also has positive rank. Contradiction. \hfill $\Box$ \begin{prop} Let ${\bf P}_{\Sigma,K}$ be a a compactification of an anisotropic torus $T$. Then the cone of effective divisors ${\Lambda}_{\rm eff}({\Sigma,K})$ is simplicial. The rank of the Picard group ${\rm Pic}({\bf P}_{\Sigma,K})$ equals to the number of $G$-orbits in $\Sigma(1)$. \label{simp} \end{prop} {\em Proof.} Let $A_1({\bf P}_{\Sigma})$ be the group of $1$-cycles on ${\bf P}_{\Sigma,E}$ modulo numerical equivalence. We identify $A_1({\bf P}_{\Sigma})$ with the dual to ${\rm Pic}({\bf P}_{\Sigma})$ group. Consider the dual cone ${\Lambda}^*_{\rm eff}({\Sigma,K})$. Since ${\Lambda}^*_{\rm eff}({\Sigma,K}) = {\Lambda}^*_{\rm eff}({\Sigma})^G$, by \ref{generators}, ${\Lambda}^*_{\rm eff}({\Sigma,K})$ consists of non-negative $G$-invariant ${\bf R}$-linear relations among primitive generators of $\Sigma(1)$. Let \[ \Sigma(1) = \Sigma_1(1) \cup \ldots \cup \Sigma_l(1) \] be the decomposition of $\Sigma(1)$ into a union of $G$-orbits. Then every $G$-invariant linear relation among the primitive generators $e_1 , \ldots , e_n$ of the $1$-dimensional cones has the form \[ \sum_{1 \leq i \leq l } \lambda_i \left( \sum_{ \sigma_j \in \Sigma_i(1) } e_j \right) = 0\;\;\;(\sigma_j = {\bf R}_{\geq 0} e_j). \] For every $i$ ($1 \leq i \leq l$), the sum \[ \sum_{ \sigma_j \in \Sigma_i(1) } e_j \] is a $G$-invariant element of the lattice $N$. Since $T$ is anisotropic, $N^G = 0$ and all sums $\sum_{ \sigma_j \in \Sigma_i(1) } e_j $ must be equal to zero. These integral relations give rise to a ${\bf Z}$-basis $r_1, \dots , r_l$ of the group of integral linear relations among $e_1 , \ldots , e_n$. Thus $A_1 ({\bf P}_{\Sigma})^G_{\bf R}$ is isomorphic to ${\bf Z}^l$ and the cone $\Lambda_{\rm eff}^*({\Sigma,K})$ consists of nonnegative linear combinations of $r_1, \dots , r_l$. So the cone $\Lambda_{\rm eff}({\Sigma,K})$ is also an $l$-dimensional simplicial cone in ${\rm Pic}({\bf P}_{\Sigma,K}) \otimes {\bf R}$. \hfill $\Box$ Below we consider several examples of compactifications of anisotropic tori. \begin{exam} {\rm Consider a $d$-dimensional fan $\Sigma$ as in \ref{proj.space}. It has a natural action of the symmetric group $S_{d+1}$. Let $E$ be a Galois extension of $K$ such that the Galois group ${\rm Gal}\, (E/K)$ is a subgroup of $S_{d+1}$ (for instance, $E$ is a simple algebraic extension defined by an $K$-irreducible polynomial $f$). Then the action of $G$ on $\Sigma$ defines a ${d}$-dimensional toric variety ${\bf P}_{\Sigma,K}$ which over $E$ is isomorphic to $d$-dimensional projective space; i.e. ${\bf P}_{\Sigma,K}$ is a Severi-Brauer variety. In particular, if $E = K(f)$, then ${\bf P}_{\Sigma,K}$ is a compactification of the ${d}$-dimensional anisotropic torus $R_{E/K}({\bf G}_m)/K^*$. Since ${\bf P}_{\Sigma,K}$ contains infintely many $K$-rational points, ${\bf P}_{\Sigma,K}$ is in fact isomorphic to ${\bf P}^{d}$ over $K$. } \label{exam.pn} \end{exam} \begin{exam} {\rm A complete fan $\Sigma$ is called {\em centrally symmetric} if it is invariant under the map $-Id$ of $N_{\bf R}$. Let $\Sigma$ be a centrally symmetric $4$-dimensional fan and let $E$ be an exension of $K$ of degree 2. The $d$-dimensional torus $T$ corresponding to the integral representation of ${\rm Gal}(E/K) \cong {\bf Z}/2{\bf Z}$ by $Id$ and $-Id$ is isomorphic to the anisotropic torus $(R^1_{E/K})^d$. The ${\bf Z}/2{\bf Z}$-invariant fan $\Sigma$ defines the compactification ${\bf P}_{\Sigma,K}$ of $(R^1_{E/K})^d$. } \label{two.aniso} \end{exam} \begin{exam} {\rm Let $K'$ be a cubic extension of a number field $K$. We construct a smooth compactification of the $2$-dimensional anisotropic $K$-torus $R^1_{K'/K}({\bf G}_m)$ as follows. Let $Y$ be the cubic surface in ${\bf P}^3$ defined by the equation \[ N_{K'/K}(z_1,z_2,z_3) = z_0^3 \] where $ N_{K'/K}(z_1,z_2,z_3)$ is the homogeneous cubic norm-form. Over the algebraic closure $\overline{K}$ it is isomorphic to the singular cubic surface $z_1 z_2 z_3 = z_0^3$. The $3$ quadratic singular points $p_1, p_2, p_3 \in Y_{\overline{K}}$ are defined over a splitting field $E$ of $R^1_{K'/K}({\bf G}_m)$ and the Galois group $G = {\rm Gal}\, K'/K$ acts on $\{ p_1, p_2, p_3 \}$ by permutations. There exists a minimal simultaneous resolution $\psi\; : Y' \rightarrow Y$ of singularities which is defined over $K$. By contraction $\psi' : Y' \rightarrow X$ of the proper pull-back of three $(-1)$ curves which are preimages of lines passing through the singular points we obtain a Del Pezzo surface $X$ of anticanonical degree $6$ which is a smooth compactification of the anisotropic torus $R^1_{K'/K}({\bf G}_m)$. } \end{exam} Let $k$ be a finite field of characteristic $p$ containing $q = p^n$ elements. Any finite extension $k'$ of $k$ is a cyclic Galois extension and the group $G = {\rm Gal }(k'/k)$ is generated by the Frobenius automorphism $\phi \; : \; z \rightarrow z^q$. By ~\ref{represent}, any $d$-dimensional algebraic torus $T$ over $k$ splitting over $k'$ is uniquely defined by the conjugacy class in ${\rm GL} (d, {\bf Z})$ of the integral matrix \[ \Phi = \rho ( \phi ) .\] \noindent The characteristic polynomial of the matrix $\Phi$ gives the following formula obtained by T. Ono \cite{ono1} for the number of $k$-rational points in $T$ : \begin{theo} Let $T$ be a $d$-dimensional algebraic torus defined over a finite field $k$. In the above notations, one has the following formula for the number of $k$-rational points of $T$: \[ {\rm Card} \lbrack T(k) \rbrack = (-1)^d{\rm det} ( \Phi - q\cdot Id ). \] \label{fin.tori} \end{theo} \begin{prop} Let ${\bf P}_{\Sigma}$ be a toric variety over a finite field $k$ defined by a $\Phi$-invariant fan $\Sigma \subset N_{\bf R}$. For any $\Phi$-invariant cone $\sigma \in \Sigma^G$, let $M_{{\bf R}, \sigma} = \check{\sigma} \cap (- \check{\sigma})$ be the maximal linear subspace in the dual cone $\check{\sigma} \subset M_{\bf R}$. Let $\Phi_{\sigma}$ be the restriction of $\Phi$ on $M_{{\bf R},\sigma}$. Then \[ {\rm Card}\lbrack {\bf P}_{\Sigma}(k)\rbrack = \sum_{\sigma \in \Sigma^G} (-1)^{{\rm dim}\,\sigma} {\rm det}(\Phi_{\sigma} - q \cdot Id). \] \label{point.var} \end{prop} \medskip {\em Proof.} By \ref{decompos}, \[ {\bf P}_{\Sigma} (k) = \bigcup_{\sigma \in \Sigma^G} T_{\sigma}(k). \] Observe that $k'$ is a splitting field for every algebraic torus $T_{\sigma,k}$ defined by the $\rho$-action of $\Phi_{\sigma}$. Now the statement follows from \ref{fin.tori}. \hfill $\Box$ \subsection{Algebraic tori over local and global fields} First we fix our notations. Let ${\rm Val}(K)$ be the set of all valuations of a global field $K$. For any $v \in {\rm Val}(K)$, we denote by $K_v$ the completion of $K$ with respect to $v$. Let $v$ be a non-archimedian absolute valuation of a number field $K$ and $E$ a finite Galois extension of $K$. Let ${\cal V}$ be an extension of $v$ to $E$, $E_{\cal V}$ the completion of $E$ with respect to ${\cal V}$. Then \[ {\rm Gal}(E_{\cal V}/ K_v ) \cong G_v \subset G, \] where $G_{v}$ is the decomposition subgroup of $G$ and $ K_v \otimes_K E \cong \prod_{{\cal V} \mid v} E_{\cal V}. $ Let $T$ be an algebraic torus over $K$ with the splitting field $E$. Denote by $T_{K_v}=T\otimes K_v$. \begin{dfn} {\rm We denote the group of characters $\hat{T}_{K_v} = M^{G_v}$ by $M_v$ and the dual group ${\rm Hom}(\hat{T}_{K_v} , {\bf Z}) = N^{G_v}$ by $N_v$. } \end{dfn} \noindent Let ${(K_v \otimes_K E)}^*$ and $E^*_{\cal V}$ be the multiplicative groups of $K_v \otimes_K E$ and $E_{\cal V}$ respectively. One has \[ T_{K_v} = {\rm Hom}_G (\hat{T}, {(K_v \otimes_K E )}^* ) = {\rm Hom}_{G_v} (M, E_{\cal V}^* ). \] Denote by ${\cal O}_{\cal V}$ the maximal compact subgroup in $E^*_{\cal V}$. There is a short exact sequence \[ 1 \rightarrow {\cal O}_{\cal V} \rightarrow E_{\cal V}^* \rightarrow {\bf Z} \rightarrow 1, \;\; b \rightarrow {\rm ord} \mid b \mid_{\cal V}. \] \noindent Denote by $T({\cal O}_v)$ the maximal compact subgroup in $T(K_v)$. Applying the functor ${\rm Hom}_{G_v}(M_v, * )$ to the short exact sequence above, we obtain the short exact sequence \[ 1 \rightarrow N_v \otimes {\cal O}_{\cal V} \rightarrow N_v \otimes E_{\cal V}^* \rightarrow N_v \rightarrow 1 \] which induces an injective homomorphism \[ \pi_v\; : \; T(K_v) / T({\cal O}_v) \hookrightarrow N_v = N^{G_v}. \] \begin{prop} {\rm \cite{drax1}} The homomorphism $\pi_v$ has finite cokernel. Moreover, $\pi_v$ is an isomorphism if $E$ is unramified in $v$. \label{pi-image} \end{prop} \begin{dfn} {\rm Let $S$ be a finite subset of ${\rm Val}(K)$ containing all archimedian and ramified non-archimedian valuations of $K$. We denote by $S_{\infty}$ the set of all archimedian valuations of $K$ and put $S_0 = S \setminus S_{\infty}$. } \end{dfn} \medskip Now we assume that $v$ is an archimedian absolute valuation, i.e., $K_v$ is ${\bf R}$ or ${\bf C}$. It is known that any torus over ${\bf R}$ is isomorphic to the product of some copies of ${\bf C}^*$, ${\bf R}^*$, or $S^1 = \{ z \in {\bf C} \mid z \overline{z} =1 \}$. The quotient $T(K_v) / T({\cal O}_v)$ is isomorphic to the ${\bf R}$-linear space $N_v \otimes {\bf R}$. The homomorphism $T(K_v) \rightarrow T(K_v) / T({\cal O}_v)$ is simply the logarithmic mapping onto the Lie algebra of $T(K_v)$. Hence, we obtain: \begin{prop} For any archimedian absolute valuation $v$, the quotient $T(K_v) / T({\cal O}_v)$ can be canonically identified with the real Lie algebra of $T(K_v)$ embedded in the $d$-dimensional ${\bf R}$-subspace $N_{\bf R}$. \end{prop} \medskip \begin{dfn} {\rm Denote by $T({\bf A}_K)$ the adele group of $T$, i.e., the restricted topological product \[ {\prod_{v \in {\rm Val}(K)}} T(K_v) \] consisting of all elements ${\bf t} = \{ t_v \} \in \prod_{v \in {\rm Val}(K)}T(K_v)$ such that $t_v \in T({\cal O}_v)$ for almost all $v \in {\rm Val}(K)$. Let \[T^1({\bf A}_K) = \{ {\bf t} \in T({\bf A}_K) \mid \prod_{v \in {\rm Val}(K)} \mid m(t_v) \mid_v = 1, \; {\rm for \; all}\; m \in \hat{T}_K \subset M \}. \] We put also \[ {\bf K}_T = \prod_{v \in {\rm Val}(K)} T({\cal O}_v), \] } \end{dfn} \begin{prop} {\rm \cite{ono1}} The groups $T({\bf A}_K)$, $T^1({\bf A}_K)$, $T(K)$, ${\bf K}_T$ have the following properties which are generalizations of the corresponding properties of the adelization of ${\bf G}_m(K)$: {\rm (i)} $T({\bf A}_K)/T^1({\bf A}_K) \cong {\bf R}^k$, where $k$ is the rank of $\hat{T}_K$; {\rm (ii)} $T^1({\bf A}_K)/T(K)$ is compact; {\rm (iii)} $T^1({\bf A}_K)/ {\bf K}_T \cdot T(K) $ is isomorphic to the direct product of a finite group ${\bf cl}(T_K)$ $($this is an analog of the idele-classes group ${Cl}(K)$$)$ and a connected compact abelian topological group which dimension equals the rank $r'$ of the group of ${\cal O}_K$-units in $T(K)$ $($this rank equals $r_1 + r_2 -1$ for ${\bf G}_m$$)$; {\rm (iv)} $W(T) = {\bf K}_T \cap T(K)$ is a finite group of all torsion elements in $T(K)$ $($this is the analog of the group of roots of unity in ${\bf G}_m(K)$$)$. \label{subgroups} \end{prop} The following theorem of A. Weil plays a fundamental role in the definition of adelic measures on algebraic varieties. \begin{theo} {\rm \cite{peyre,weil1} } Let $X$ be an $d$-dimensional smooth algebraic variety over a global field $K$. Denote by ${\cal K}$ the canonical sheaf on $X$ with a family of local metrics $\|\cdot \|_v $. Then these local metrics uniquely define natural $v$-adic measures $\omega_{{\cal K},v}$ on $X(K_v)$. Let $U \subset X$ be a Zariski open subset of $X$. Then for almost all $v \in {\rm Val}(K)$ one has \[ \int_{U({\cal O}_v)} \omega_{{\cal K},v} = \frac{{\rm Card} \lbrack U(k_v) \rbrack}{q^d_v}, \] where $k_v$ is the residue field of $K_v$ and $q_v = {\rm Card} \lbrack k_v \rbrack$. \end{theo} \begin{rem} {\rm We notice that the structure sheaf ${\cal O}_X$ of any algebraic variety $X$ has a natural metrization defined by $v$-adic valuations of the field $K$. If $X = {\cal G}$ is an algebraic group, then there exists a natural way to define a metrization of the canonical sheaf ${\cal K}$ on ${\cal G}$ by choosing a ${\cal G}$-invariant algebraic differential $d$-form $\Omega$. Such a form defines an isomorphism of ${\cal K}$ with the structure sheaf ${\cal O}_{\cal G}$. We denote the corresponding local measure on ${\cal G}(K_v)$ by $\omega_{\Omega,v}$.} \label{l.measures} \end{rem} Let $T$ be a $d$-dimensional torus over $K$ with a splitting field $E$. Take a $T$-invariant differential $d$-form $\Omega$ on $T$ (it is unique up to a constant from $K$). According to A. Weil (\ref{l.measures}), we obtain a family of local measures $\omega_{\Omega,v}$ on $T(K_v)$. \begin{dfn} {\rm \cite{ono1} Let \[ L_S(s, T;E/K) = \prod_{v \not\in S} L_v(s, T ;E/K) \] be the Artin $L$-function corresponding to the representation \[ \rho \; :\; G= {\rm Gal}(E/K) \rightarrow {\rm GL}(M) \] and a finite set $S \subset {\rm Val}(K)$ containing all archimedian valuations and all non-archimedian valuations of $K$ which are ramified in $E$. By definition, $L_v(s,T;E/K) \equiv 1$ if $v \in S$. The numbers \[ c_v = L_v(1, T; E/K) = \frac{1}{{\rm det}(Id - q^{-1}_v \Phi_v)}, \; v \not\in S \] are called {\em canonical correcting factors} for measures $\omega_{\Omega,v}$ ($\Phi_v$ is the $\rho$-image of a local Frobenius element in $G$). } \end{dfn} By \ref{fin.tori}, one has \[ c_v^{-1} = \int_{T({\cal O}_v)} \omega_{\Omega,v} = \frac{{\rm Card} \lbrack T(k_v)\rbrack}{q^d_v}, \; \; v \not\in S. \] Let $d\mu_v = c_v \omega_{\Omega,v}$. We put $c_v =1$ for $v \in S$. Since \[ \int_{T({\cal O}_v)} d\mu_v = 1 \] for $v \not\in S$, the $\{ c_v \}$ defines the canonical measure \[ \omega_{\Omega,S} = \prod_{v \in {\rm Val}(K)} d\mu_v \] on the adele group $T({\bf A}_K)$. By the product formula, $\omega_{\Omega,S}$ does not depend on the choice of $\Omega$. Let $dx$ be the standard Lebesgue measure on $T({\bf A}_K)/T^1({\bf A}_K) \cong {\bf R}^k = M_{\bf R}^G$. There exists a unique Haar measure $\omega^1_{\Omega,S}$ on $T^1({\bf A}_K)$ such that $\omega^1_{\Omega,S}dx = \omega_{\Omega,S}$. \begin{dfn} {\rm The {\em Tamagawa number of } $T_K$ is defined as \[ \tau(T_K) = \frac{b_S(T_K)}{l_S(T_K)} \] where \[ b_S(T_K) = \int_{T^1({\bf A}_K)/T(K)} \omega^1_{\Omega,S} , \] \[ l_S(T_K) = \lim_{s \rightarrow 1} (s-1)^k L_S(s, T; E/K). \]} \label{tamagawa1} \end{dfn} \begin{rem} {\rm Although the numbers $b_S(T_K)$ and $l_S(T_K)$ do depend on the choice of the finite subset $S \subset {\rm Val}(K)$, the Tamagawa number $\tau(T_K)$ does not depend on $S$.} \end{rem} \begin{theo} {\rm \cite{ono1,ono2} } The Tamagawa number $\tau(T)$ of $T$ does not depend on the choice of a splitting field $E$. It satisfies the following properties: {\rm (i)} $\tau({\bf G}_m(K)) =1;$ {\rm (ii)} $\tau(T \times T') = \tau(T) \cdot \tau(T') $ where $T'$ and $T$ are tori over $K$; {\rm (iii)} $\tau_K (R_{K'/K}(T)) = \tau_{K'}(T)$ for any torus $T$ over $K'/K$. Moreover, $\tau(T)$ is the ratio of two positive integers \[ h(T_K) = {\rm Card} \lbrack H^1(G, M) \rbrack \] and $i(T_K) = {\rm Card}\lbrack {\rm III}(T) \rbrack$ where \[ {\rm III}(T) = {\rm Ker}\, \lbrack H^1(G, T(K)) \rightarrow \prod_{v} H^1(G_v, T(K_v)) \rbrack; \] in particular, $\tau(T_K)$ is a rational number. \label{tamagawa2} \end{theo} \begin{dfn} {\rm Let $\overline{T(K)}$ be the closure of $T(K)$ in $\prod_vT(K_v) $ in the direct product topology. Define the {\em obstruction group to weak approximation} as \[ A(T)= \prod_v T(K_v)/\overline{T(K)}. \] } \end{dfn} \begin{theo} {\rm \cite{sansuc}} Let ${\bf P}_{ \Sigma }$ be a complete smooth toric variety over $K$. There is an exact sequence: \[ 0\rightarrow A(T) \rightarrow Hom (H^1(G,{\rm Pic}({\bf P}_{ \Sigma ,E})),{\bf Q }/{\bf Z })\rightarrow {\rm III}(T)\rightarrow 0. \] The group $H^1(G,{\rm Pic}({\bf P}_{ \Sigma ,E}))$ is canonically isomorphic to ${\rm Br}({\bf P}_{\Sigma,K})/{\rm Br}(K)$, where ${\rm Br}({\bf P}_{\Sigma,K}) = H^2_{\rm et}({\bf P}_{\Sigma,K}, {\bf G}_m)$. \label{weak} \end{theo} \begin{coro} Denote by $\beta({\bf P}_{\Sigma})$ the cardinality of $H^1(G,{\rm Pic}({\bf P}_{ \Sigma ,E}))$. Then \[ {\rm Card} \lbrack A(T) \rbrack = \frac{\beta({\bf P}_{\Sigma})}{i(T_K)}. \] \label{weak1} \end{coro} \section{Heights and their Fourier transforms} \subsection{Complexified local Weil functions and heights} A theory of heights on an algebraic variety $X$ defined over a number field $K$ is the unique functorial homomorphism from ${\rm Pic}(X) $ to equivalence classes of functions $X(K)\rightarrow {\bf R}_{\ge 0}$ which on metrized line bundles ${\cal L}$ is given by the formula $$ H_{{\cal L}}(x) = \prod_v \| f(x)\|^{-1}_v $$ where $f$ is a rational section of $L$ not vanishing in $x\in X(K)$. Two functions are equivalent if they differ by a bounded on $X(K)$ function. For our purposes it will be convenient to extend these notions to the complexified Picard group ${\rm Pic}(X)\otimes {\bf C}$. Let ${\bf P}_{\Sigma}$ be a compact toric variety over a global field $K$. We define a canonical compact covering of ${\bf P}_{\Sigma}(K_v)$ by compact subsets ${\bf C}_{\sigma,v} \subset U_{\sigma}(K_v)$. For this purpose we identify lattice elements $m \in M$ with characters of $T$ and define the compact subset ${\bf C}_{\sigma,v} \subset U_{\sigma}(K_v)$ as follows \[ {\bf C}_{\sigma,v} = \{ x_v \in U_{\sigma}(K_v) \mid \| m(x_v) \|_v \leq 1\; {\rm for}\; {\rm all}\; m \in M^{G_v} \cap \check{\sigma} \}. \] \begin{prop} The compact subsets ${\bf C}_{\sigma,v}$ $(\sigma \in \Sigma)$ form a compact covering of ${\bf P}_{\Sigma}(K_v)$ such that for any two cones $\sigma, \sigma' \in \Sigma$ one has \[ {\bf C}_{\sigma,v} \cap {\bf C}_{\sigma',v} = {\bf C}_{\sigma \cap \sigma',v}. \] \end{prop} {\em Proof.} The last property of the compact subsets ${\bf C}_{\sigma,v}$ follows immediatelly from their definition. Since the $T(K_v)$-orbit of maximal dimension is dense in ${\bf P}_{\Sigma}(K)$, it is sufficient to prove that the compacts ${\bf C}_{\sigma,v}$ cover $T(K_v)$. Let $x_v \in T(K_v)$. Denote by $\overline{x_v}$ the image of $x_v$ in $T(K_v) / T({\cal O}_v) \subset N_{\bf R}$. By completeness of the fan $\Sigma$, the point $-\overline{x_v}$ is contained in some cone $\sigma \in \Sigma$. Hence $x_v \in {\bf C}_{\sigma,v}$. \hfill $\Box$ \bigskip Now we define canonical metrizations of $T(K_v)$-linearized line bundles on ${\bf P}_{\Sigma}(K_v)$. \medskip Let $ L(\varphi)$ be a line bundle on ${\bf P}_{\Sigma}( K_v)$ corresponding to a $\Sigma$-piecewise linear integral $G_v$-invariant function $\varphi$ on $N_{\bf R}$. \begin{prop} Let $f$ be a rational section of $ L(\varphi)$. We define the $v$-norm of $f$ at a point $x_v \in {\bf P}_{\Sigma}(K_v)$ as \[ \| f(x_v) \|_v = \mid \frac{f(x_v)}{m_{\sigma,\varphi}(x_v)} \mid_v \] where $\sigma$ is a cone in $\Sigma$ such that $x_v \in {\bf C}_{\sigma,v}$ and $m_{\sigma,\varphi} \in M$ is the restriction of $\varphi$ on $\sigma$. Then this $v$-norm defines a $T({\cal O}_v)$-invariant $v$-adic metric on $ L(\varphi)$. \end{prop} {\em Proof.} The statement follows from the fact that \[ \mid m_{\sigma,\varphi}(x_v) \mid_v = \mid m_{\sigma',\varphi}(x_v) \mid_v \] if $x_v \in {\bf C}_{\sigma,v} \cap {\bf C}_{\sigma',v}$. \hfill $\Box$ A family of local metrics on all $T$-linearized line bundles on ${\bf P}_{\Sigma}$ corresponding to $\Sigma$-piecewise linear $G$-invariant functions $\varphi \in PL(\Sigma)^G$ uniquely determines a family of local Weil functions on $({\bf P}_{\Sigma})$ corresponding to $T$-invariant divisors \[ D_{\varphi} = \varphi(e_1)D_1 + \cdots + \varphi(e_n)D_n. \] We extend these local Weil functions to the group of $T$-invariant Cartier divisors with {\em complex coefficients} as follows. \begin{dfn} {\rm A $T$-invariant ${\bf C}$-Cartier divisor is a formal linear combination $D_s = s_1 D_1 + \cdots + s_n D_n$, with $s=(s_1,\ldots,s_n)\in {\bf C}^n$ or equivalently a complex piecewise linear function $\varphi$ in $PL(\Sigma)^G_{\bf C}$ having the property $\varphi(e_i) = s_i$ $(i = 1, \ldots ,n)$. } \end{dfn} \begin{dfn} {\rm Let $\varphi \in PL(\Sigma)^G_{\bf C}$. For any point $x_v \in T(K_v) \subset {\bf P}_{\Sigma}(K_v)$, denote by $\overline{x_v}$ the image of $x_v$ in $N_v$ (resp. $N_v\otimes {\bf R}$ for archimedian valuations), where $N_v$ is considered as a canonical lattice in the real space $N_{\bf R}$. Define the {\em complexified local Weil function} $H_{ \Sigma ,v}(x_v, \varphi)$ by the formula \[H_{ \Sigma ,v}(x_v,\varphi) = e^{\varphi(\overline{x_v})\log q_v }\] where $q_v$ is the cardinality of the residue field $k_v$ of $K_v$ if $v$ is non-archimedian and $\log q_v = 1$ if $v$ is archimedian. } \end{dfn} \begin{prop} The complexified local Weil function $H_{ \Sigma ,v}(x_v, \varphi)$ satisfies the following properties: {\rm (i)} If $s_i = \varphi(e_i) \in {\bf Z}^n$ $( i =1, \ldots, n)$, then $H_{ \Sigma ,v}(x_v,\varphi)$ is a classical local Weil function $H_{L(\varphi),v}(x_v)$ corresponding to a $T$-invariant Cartier divisor \[ D_s = s_1 D_1 + \cdots + s_n D_n \] on ${\bf P}_{\Sigma}$. {\rm (ii)} $H_{ \Sigma ,v}(x_v,\varphi)$ is $T({\cal O}_v)$-invariant. {\rm (iii)} $H_{ \Sigma ,v}(x_v, \varphi + \varphi') = H_{ \Sigma ,v}(x_v,\varphi)H_{ \Sigma ,v}(x_v,\varphi')$. \label{local.f} \end{prop} \begin{dfn} {\rm Let $\varphi \in PL(\Sigma)^G_{\bf C}$. We define the {\em complexified height function on ${\bf P}_{\Sigma,K}$} by \[ H_{\Sigma}(x,\varphi) = \prod_{v \in {\rm Val}(K)} H_{ \Sigma ,v}(x,\varphi). \]} \end{dfn} \begin{rem} {\rm Although all local factors $H_{ \Sigma ,v}(x,\varphi)$ of $H_{\Sigma}(x,\varphi)$ are functions on $PL(\Sigma)_{\bf C}^G$, by the product formula, the global complex height function $H_{\Sigma}(x,\varphi)$ depends only on the class of $\varphi \in PL(\Sigma)_{\bf C}^G$ modulo complex global linear $G$-invariant functions on $N_{\bf C}$, i.e., $H_{\Sigma}(x,\varphi)$ depends only on the class of $\varphi$ in ${\rm Pic}({\bf P}_{\Sigma,K})\otimes {\bf C}$. } \end{rem} \begin{dfn} {\rm We define the zeta-function of the complex height-function $H_{\Sigma}(x, \varphi)$ as \[ Z_{\Sigma}(\varphi) = \sum_{x \in T(K)} H_{\Sigma}(x,-\varphi). \]} \end{dfn} \begin{rem} {\rm One can see that the series $Z_{\Sigma}(\varphi)$ converges absolutely and uniformly in the domain ${\rm Re}(\varphi(e_j)) \gg 0$ for all $j$. Since $H_{\Sigma}(x,\varphi)$ is the product of the local complex Weil functions $H_{ \Sigma ,v}(x,\varphi)$ and $H_{ \Sigma ,v}(x,\varphi) = 1$ for almost all $v$ ($x \in T(K)$), we can immediately extend $H_{\Sigma}(x,\varphi)$ to a function on the adelic group $T({\bf A}_K)$.} \end{rem} \subsection{Fourier transforms of non-archimedian heights} \begin{dfn} {\rm Let $\Sigma$ be a complete regular fan of cones in $N$ whose 1-dimensional cones are generated by $e_1, \ldots, e_n$. We establish a one-to-one correspondence between $e_1, \ldots, e_n$ and $n$ independent variables $z_1, \ldots, z_n$. The {\em Stanley-Reisner ring} $R(\Sigma)$ is defined as the factor of the polynomial ring $A[z]= {\bf C} [z_1, \ldots, z_n ]$ by the ideal $I(\Sigma)$ generated by all monomials $z_{i_1} \ldots z_{i_k}$ such that $e_{i_1}, \ldots, e_{i_k}$ are not generators of a $k$-dimensional cone in $\Sigma$.} \label{opr.stenley} \end{dfn} \begin{prop} There is a natural identification between the the elements of the lattice $N$ and the monomial {\bf C}-basis of the ring $R(\Sigma)$. \label{prop.basis} \end{prop} {\em Proof.} Every integral point $x \in N$ belongs to the interior of a unique cone $\sigma \in \Sigma$. Let $e_{i_1}, \ldots, e_{i_k}$ be an integral basis of $\sigma$. Then there exist positive integers $a_1, \ldots, a_k$ such that \[ \overline{x} = a_1e_{i_1} + \cdots a_k e_{i_k}. \] Therefore, $\overline{x}$ defines the monomial $m(\overline{x}) = z_{i_1}^{a_1} \cdots z_{i_k}^{a_k}$. By definition, $m(\overline{x}) \notin I(\Sigma)$. It is clear that $I(\Sigma)$ has a monomial {\bf C}-basis. Hence, we have constructed a mapping $\overline{x} \rightarrow m(\overline{x})$ from $N$ to the monomial basis of $R(\Sigma)$. It is easy to see that this mapping is bijective. \hfill $\Box$ \bigskip Now choose a valuation $v \not\in S$. Then we obtain a cyclic subgroup $G_v = \langle \Phi_v \rangle \subset G$ generated by a lattice automorphism $\Phi_v \,:\, N \rightarrow N$ representing the local Frobenius element at place $v$. Then $\Sigma(1)$ splits into a disjoint union of $G_v$-orbits \[ \Sigma(1) = \Sigma_1(1) \cup \cdots \cup \Sigma_l(1). \] Let $d_j$ be the length of the $G_v$-orbit $ \Sigma _j(1)$. One has \[ \sum_{ i =1}^{l} d_j = n. \] \begin{dfn} {\rm Define the ${\bf Z}^l{\geq 0}$-grading of the polynomial ring $A[z] = {\bf C} \lbrack z_1, \ldots, z_n \rbrack$ and the Stanley-Reisner ring $R(\Sigma)$ by the decomposition of the set of variables $\{z_1, \ldots, z_n\}$ into the disjoint union of $l$ sets $Z_1 \cup \cdots \cup Z_l$ which is induced by the decomposition of $ \Sigma (1)$ into $G_v$-orbits. The standard ${\bf Z}_{\geq 0}$-grading of the polynomial ring $A[z] = {\bf C} \lbrack z_1, \ldots, z_n \rbrack$ and the Stanley-Reisner ring $R(\Sigma)$ will be called the {\em total grading}. } \end{dfn} \begin{dfn} {\rm We define the power series $P(\Sigma, \Phi_v; t_1, \ldots, t_l)$ by the formula \[ P(\Sigma,\Phi_v; t_1, \ldots, t_l) = \sum_{(i_1,\ldots, i_l) \in {\bf Z}_{{\geq 0}}^l} ({\rm Tr}\, \Phi_v^{i_1, \ldots, i_l}) t_1^{i_1} \cdots t_l^{i_l}, \] where $\Phi_v^{i_1, \ldots, i_l}$ is the linear operator induced by $\Phi_v$ on the homogeneous $(i_1, \ldots, i_l)$-component of $R(\Sigma)$. } \end{dfn} \begin{prop} One has \[ P(\Sigma,\Phi_v; t_1, \ldots, t_l) = \frac{Q_{ \Sigma }(t_1^{d_1}, \ldots, t_l^{d_l})} {(1- t_1^{d_1}) \cdots (1 - t_l^{d_l}) } \] where $Q_{ \Sigma }(t_1^{d_1}, \ldots, t_l^{d_l})$ is a polynomial in $t_1^{d_1}, \ldots, t_l^{d_l}$ having the total degree $n$ such its all nonconstant monomials have the total degree at least $2$. \label{p-function} \end{prop} {\em Proof.} Since ${\rm dim} A \lbrack z \rbrack - {\rm dim} R(\Sigma) = n-d$, there exists the minimal ${\bf Z}^l_{\geq 0}$-graded free resolution \[ 0 \rightarrow F^{n-d} \rightarrow \cdots \rightarrow F^1 \rightarrow F^0 = A \lbrack z \rbrack \rightarrow R(\Sigma) \rightarrow 0 \] of the Stanley-Reisner ring $R(\Sigma)$ considered as a module over the polynomial ring $A \lbrack z \rbrack$. Let $\overline{\Phi}_v^{\,i_1, \ldots, i_l}$ be the linear operator on the homogeneous $(i_1, \ldots, i_l)$-component of $A\lbrack z \rbrack$ induced by the action of $\Phi_v$ on $z_1, \ldots, z_n$. Then \[ \sum_{(i_1,\ldots, i_l) \in {\bf Z}_{{\geq 0}}^l} ({\rm Tr}\, \overline{\Phi}_v^{\,i_1, \ldots, i_l}) t_1^{i_1} \cdots t_r^{i_l} = \frac{1}{\prod_{j = 1}^l (1 - t_j^{d_j})}. \] We notice that ${\rm Tr}\, \overline{\Phi}_v^{\,i_1, \ldots, i_l}$ and ${\rm Tr}\, {\Phi}_v^{\,i_1, \ldots, i_l}$ can be nonzero only if the length $d_k$ of the $G_v$-orbit $ \Sigma _k(1)$ divides $i_k$ $( k =1, \ldots, l)$. Therefore the polynomial $Q(t_1^{d_1}, \ldots, t_l^{d_l})$ is defined by ranks and ${\bf Z}_{\geq 0}^l$-degrees of generators of the free $A\lbrack z \rbrack$-modules $F^i$. Notice that every monomial in $I(\Sigma)$ has the total degree at least $2$, because every element $e_i \in \{ e_1, \ldots, e_n \}$ generates a $1$-dimensional cone of $\Sigma$. So all generators of $F^i$ $(i \geq 1)$ have the total degree at least $2$. Therefore, the polynomial $Q_{ \Sigma }$ has only monomials of the total degree at least $2$. Since $R(\Sigma)$ is a Gorenstein ring, $F^{n-d}$ is a free $A\lbrack z \rbrack$-module of rank $1$ with a generator of the degree $(d_1, \ldots, d_l)$. Therefore, $Q$ has the total degree $n = d_1 + \cdots + d_l$. \hfill $\Box$ \bigskip Let $\chi$ be a topological character of $T({\bf A}_K)$ such that its $v$-component $\chi_v\, : \, T(K_v) \rightarrow S^1 \subset {\bf C}^*$ is trivial on $T({\cal O}_v)$. For each $ j \in \{ 1, \ldots, l\}$, we denote by $n_j$ the sum of $d_j$ generators of all $1$-dimensional cones of the $G_v$-orbit $ \Sigma _j(1)$. Then $n_j$ is a $G_v$-invariant element of $N$. By \ref{pi-image}, $n_j$ represents an element of $T(K_v)$ modulo $T({\cal O}_v)$. Therefore, $\chi_v(n_j)$ is well defined. \begin{prop} Let $v \not\in S$. Denote by $q_v$ the cardinality of the finite residue field $k_v$ of $K_v$. Then for any local topological character $\chi_v$ of $T(K_v)$, one has \[ \hat{H}_{\Sigma,v} (\chi, -\varphi) = \int_{T(K_v)} H_{ \Sigma ,v}(x_v,-\varphi) \chi_v(x_v) d\mu_v = \] \[ \frac{Q_{ \Sigma }\left( \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}}, \ldots, \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_l)}} \right)} {(1- \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}} ) \cdots (1 - \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_l)}} ) } \] if $\chi_v$ is trivial on $T({\cal O}_v)$, and \[ \int_{T(K_v)} H_{ \Sigma ,v}(x_v,-\varphi) \chi_v(x_v) d\mu_v = 0 \] otherwise. \label{integral.1} \end{prop} {\em Proof.} Since the local Haar measure $\mu_v$ is $T({\cal O}_v)$-invariant, one has \[ \int_{T(K_v)} H_{ \Sigma ,v}(x_v, -\varphi) \chi_v(x_v) d \mu_v = \] \[ = \sum_{\overline{x}_v \in T(K_v)/T({\cal O}_v)} H_{ \Sigma ,v}(\overline{x}_v, -\varphi) \chi_v(\overline{x}_v) \int_{T({\cal O}_v)} \chi_v d\mu_v \] where $\overline{x}_v$ denotes the image of $x_v$ in $T(K_v)/T({\cal O}_v) = N_v$. Notice that $\int_{T({\cal O}_v)} \chi_v d\mu_v = 0$ if $\chi_v$ has nontrivial restriction on $T({\cal O}_v)$. By \ref{prop.basis}, there exists a natural identification between $G_v$-invariant elements of $N$ and $G_v$-invariant monomials in $R(\Sigma)$. Since $\Phi_v$ acts by permutations on monomials in the homogeneous $(i_1, \ldots, i_l)$-component of $R(\Sigma)$, the number of $G_v$-invariant monomials in $R^{i_1, \ldots, i_l}(\Sigma)$ equals ${\rm Tr}\,\Phi_v^{i_1, \ldots, i_l}$. Take a $G_v$-invariant element $\overline{x}_v \in N$ such that $m(\overline{x}_v) \in R^{i_1, \ldots, i_r}(\Sigma)$. Put $i_k = d_k b_k$ $( k =1, \ldots, l)$. Then \[ \varphi(\overline{x}_v) = b_1\varphi(n_1) + \cdots + b_l \varphi(n_l) \] and \[ \chi_v(\overline{x}_v) = \chi_{v}^{b_1}(n_1) \cdots \chi_{v}^{b_l}(n_l). \] This implies the claimed formula. \hfill $\Box$ \bigskip Let $A^*({\bf P}_{\Sigma}) = \bigoplus_{i =0}^d A^i({\bf P}_{\Sigma})$ be the Chow ring of ${\bf P}_{\Sigma,E_{\cal V}}$. The groups $A^i({\bf P}_{\Sigma})$ have natural $G_v$-action. Denote by $\Phi_v(i)$ the operator on $A^i({\bf P}_{\Sigma})$ induced by $\Phi_v$. \begin{prop} Denote by ${\bf 1}_v$ the trivial topological character of $T(K_v)$. Then the restriction of \[ \int_{T(K_v)} H_{ \Sigma ,v}(x_v,-\varphi) {\bf 1}_v(x_v) d \mu_v \] to the line $s_1 = \cdots = s_r = s$ is equal to \[ L_v( s,T; E/K) \left( \sum_{k =0}^d \frac{{\rm Tr}\, \Phi_v(i)}{q_v^{ks}} \right). \] \label{loc-integ} \end{prop} {\em Proof.} The Chow ring is the quotient of $R(\Sigma)$ by a regular sequence \cite{danilov}. This gives the ${\bf Z}_{\geq 0}$-graded (by the total degree) Koszul resolution having a $\Phi_v$-action: \[ 0 \rightarrow \Lambda^d M \otimes R(\Sigma) \rightarrow \cdots \rightarrow \Lambda^1 M \otimes R(\Sigma) \rightarrow R(\Sigma) \rightarrow A^*({\bf P}_{\Sigma}) \rightarrow 0. \] We apply the trace operator to the $k$-homogeneous component of the Koszul complex, then we multiply the result by $1/q_v^{ks}$ and take the sum over $k \geq 0$. By \ref{integral.1}, we have \[ L_v^{-1}(s,T;E/K) \cdot \int_{T(K_v)} H_{ \Sigma ,v}(x_v,-\varphi) {\bf 1}_v(x_v) d \mu_v = \sum_{k =0}^d \frac{{\rm Tr}\, \Phi_v(i)}{q_v^{ks}} \] because \[ \sum_{k =0}^d \frac{(-1)^k}{q_v^{ks}} {\rm Tr}( \Lambda^k {\Phi_v}) = {\rm det}(Id - q_v^{-s} \Phi_v) = L_v^{-1}(s,T;E/K). \] \hfill $\Box$ \subsection{Fourier transforms of archimedian heights} \begin{prop} Let $\chi_v(y) = e^{-2\pi i \langle x,y \rangle}$ be a topological character of $T(K_v)$ which is trivial on $T({\cal O}_v)$. Then the Fourier transform $\hat{H}_{ \Sigma ,v}(\chi_v,-\varphi)$ of a local archimedian Weil function $H_{ \Sigma ,v} (x,-\varphi)$ is a rational function in $s_j = \varphi(e_j)$ for ${\rm Re}(s_j) > 0$. \end{prop} {\em Proof.} First we consider the case $K_v = {\bf C}$. Then $T(K_v)/T({\cal O}_v) = N_{\bf R}$ and \[ \frac{\hat{H}_{ \Sigma ,v}(\chi_v(y),-\varphi)}{\int_{T({\cal O}_v)} d\mu_v} = \int_{N_{\bf R}} e^{-\varphi(x) - 2\pi i \langle x,y \rangle} dx = \sum_{\sigma \in \Sigma(d)} \int_{\sigma} e^{-\varphi(x) - 2\pi i \langle x,y \rangle} dx \] where $dx=d\overline{x}_v$ is the standard measure on $N_v\otimes {\bf R }\simeq {\bf R }^d $. On the other hand, \[ \int_{\sigma} e^{-\varphi(x) - 2\pi i \langle x,y \rangle} dx = \frac{1}{\prod_{e_j \in \sigma} (s_j + 2\pi i \langle x,y \rangle)}. \] If $K_v = {\bf R}$, then the following simple statements allows to repeat the arguments: \begin{lem} Let $\Sigma \subset N_{\bf R}$ be a complete regular $G$-invariant fan of cones. Denote by $\Sigma^G \subset N_{\bf R}^G$ the fan consisting of $\overline{\sigma} = \sigma \cap N_{\bf R}^G$, $\sigma \in \Sigma$. Then $\Sigma^G$ is again a complete regular fan. \end{lem} \hfill $\Box$ \medskip \newline The proof of the following proposition was suggested to us by W. Hoffmann. \begin{prop} Let ${\bf K} \subset {\bf C}^r$ be a compact such that ${\rm Re}( s_j) > \delta$ for all $(s_1, \ldots, s_r) \in {\bf K}$. Then there exists a constant $c({\bf K},\Sigma)$ such that \[ \mid \hat{H}_{ \Sigma ,v} (y,-\varphi) \mid \leq c({\bf K}, \Sigma ) \sum_{\sigma \in \Sigma(d)} \frac{1}{\prod_{e_k \in \sigma} (1 + \mid \langle y, e_k \rangle \mid)^{1 + 1/d}}. \] \label{l-estimation} \end{prop} {\em Proof.} Let $f_1, \ldots, f_d$ be a basis of $M$. Put $x_i = \langle x, f_i \rangle$. We denote by $y_1, \ldots, y_d$ the coordinates of $y$ in the basis $f_1, \ldots, f_d$. Let $\varphi_i(x) = \frac{\partial}{\partial x_i}\varphi (x)$. $\varphi_i(x)$ has a constant value $\varphi_{i,\sigma}$ in the interior of a cone $\sigma \in \Sigma(d)$. \[ \hat{H}_{ \Sigma ,v} (y,-\varphi) = \int_{N_{\bf R}} e^{-\varphi(x)-2\pi i<y,x>}dx = \frac{1}{2\pi iy_j} \int_{N_{\bf R}} \frac{\partial}{\partial x_j} (e^{-\varphi(x)}) e^{-2\pi i<y,x>}dx \] \[ = - \frac{1}{2\pi iy_j} \int_{N_{\bf R}} \varphi_j(x) e^{-\varphi(x)-2\pi i<y,x>}dx \] \[ = \frac{i}{2\pi y_j} \sum_{\sigma \in\Sigma(d)} \frac{\varphi_{j,\sigma}}{\prod_{e_k \in \sigma} (s_k+2\pi i<y,e_k>)} \] Notice that $M_{\bf R}$ is covered by $d$ domains: \[ V_j = \{ y = \sum_i y_i f_i \in M_{\bf R} \mid\; \; \mid y_j \mid = \max_i \mid y_i \mid \}. \] Let $\| y \|^2 = \sum_{i} y_i^2$. Then $\| y \| \leq \sqrt{d} \mid y_j \mid$ for $ y \in V_j$. Then \[ | \hat{H}_{ \Sigma ,v} (y,-\varphi) |\leq \frac{\sqrt{d}}{\|y\|}\sum_{\sigma \in \Sigma (d)}\frac{1}{ \prod_{e_k\in \sigma } |s_k+2\pi i<y,e_k>|} \] for $y\in V_j$. Furthermore, we obtain \[ | \hat{H}_{ \Sigma ,v} (y,-\varphi) |\leq \frac{C'(\delta)}{1 + \|y\|} \sum_{\sigma \in \Sigma (d)}\frac{1}{ \prod_{e_k\in \sigma } (1 + \mid <y,e_k>\mid } \] using the following obvious statement: \begin{lem} Assume that ${\rm Re}(s) > \delta > 0$. Then there exists a positive constant $C(\delta)$ such that for all $t$ one has $ C(\delta) (\mid s + 2\pi i t \mid) \geq 1 + \mid t \mid$. \end{lem} Since $\mid \langle y, e_k \rangle \mid \leq \| y \| \|e_k \|$, it follows that there exist constants $c_{\sigma}$ such that \[ c_{\sigma} (1+|<y,e_k>|)^d \geq \prod_{e_k\in \sigma }(1+|<y,e_k>|). \] Finally, we obtain \[ \mid \hat{H}_{ \Sigma ,v} (y,-\varphi) \mid \leq c_j(\delta, \Sigma) \cdot \sum_{\sigma \in \Sigma(d)} \frac{1}{\prod_{e_k \in \sigma} (1 + \mid \langle y, e_k \rangle \mid)^{1 + 1/d}} \] for all $y \in V_j$. It remains to put $c({\bf K}, \Sigma ) = \max_j c_j(\delta, \Sigma)$. \hfill $\Box$ \begin{coro} Let $g \, : \, M_{\bf C} \rightarrow {\bf C}$ be a continious function such that \[ \mid g(iy) \mid \leq \|y\|^{\varepsilon}, \; \varepsilon < 1, \;\; y \in M_{\bf R}. \] Then \[ \sum_{y \in O} g(iy) \hat{H}_{ \Sigma ,v}(y, -\varphi) \] is absolutely and uniformly convergent on ${\bf K}$ for any function $g(iy)$ for any lattice $O \subset M_{\bf R}$. \label{lconver} \end{coro} \section{Characteristic functions of convex cones} Let $V$ be an $r$-dimensional real vector space, $V_{\bf C}$ its complex scalar extension, $\Lambda \subset V$ a convex $r$-dimensional cone such that $\Lambda \cap - \Lambda = 0 \in V$. Denote by $ \Lambda ^{\circ}$ the interior of $ \Lambda $, ${ \Lambda }_{\bf C}^{\circ} = { \Lambda }^{\circ} + iV$ the complex tube domain over ${ \Lambda }^{\circ}$, by $V^*$ the dual space, by ${ \Lambda }^* \subset V^*$ the dual to ${ \Lambda }$ cone and by $dy$ a Haar measure on $V^*$. \begin{dfn} {\rm The {\em characteristic function of} ${ \Lambda }$ is defined as the integral \[ {\cal X}_{ \Lambda }(dy,u) = \int_{{ \Lambda }^*} e^{- \langle u, y \rangle} dy, \] where $u \in { \Lambda }_{\bf C}$. } \end{dfn} \begin{rem} {\rm Characteristic functions of convex cones have been investigated in the theory of homogeneous cones by M. K\"ocher, O.S. Rothaus, and E.B. Vinberg \cite{koecher,vinberg,rothaus}.} \end{rem} \begin{rem} {\rm We will be interested in characteristic functions of convex cones ${ \Lambda }$ in real spaces $V$ which have natural lattices $L \subset V$ of the maximal rank $r$. Let $L^*$ be the dual lattice in $V^*$, then we can normalize the Haar measure $dy$ on $V^*$ so that the volume of the fundamental domain $V^*/L^*$ equals $1$. In this case the corresponding characteristic function will be denoted simply by ${\cal X}_{ \Lambda }(u)$. } \end{rem} \begin{prop} {\rm \cite{vinberg} } Let $u \in { \Lambda }^{\circ} \subset V$ be an interior point of ${\Lambda}$. Denote by ${ \Lambda }^*_u(t)$ the convex $(r-1)$-dimensional compact \[ \{ y \in { \Lambda }^* \mid \langle u, y \rangle = t \} \] We define the $(r-1)$-dimensional measure $dy_t'$ on ${ \Lambda }^*_u(t)$ in such a way that for any function $f\; : \; V \rightarrow {\bf R}$ with compact support one has \[ \int_{{V}^*} f(y) dy = \int_{-\infty}^{+\infty}dt \left(\int_{\langle u,y \rangle=t} f(y)dy_t'\right). \] Then \[ {\cal X}_{ \Lambda }(u) = (r-1)!\int_{{ \Lambda }^*_u(1)} dy_1'. \] \end{prop} The characteristic function ${\cal X}_{ \Lambda }(u)$ has the following properties \cite{rothaus,vinberg}: \begin{prop} {\rm (i)} If ${\cal A}$ is any invertible linear operator on ${V}$, then \[ {\cal X}_{ \Lambda } ({\cal A}u) = \frac{{\cal X}_{ \Lambda }(u)} {{\rm det}{\cal A}}; \] {\rm (ii)} If ${ \Lambda }^{\circ} = {\bf R}^r_{\geq 0}$, $L = {\bf Z}^r \subset {\bf R}^r$, then \[ {\cal X}_{ \Lambda }(u) = (u_1 \cdots u_r)^{-1}, \;{\rm for } \;{\rm Re}(u_i) > 0 ; \] {\rm (iii)} If $z \in { \Lambda }^{\circ}$, then \[ \lim_{z \rightarrow \partial { \Lambda }} {\cal X}_{ \Lambda }(z) = \infty; \] {\rm (iv)} ${\cal X}_{ \Lambda }(u) \neq 0$ for all $u \in { \Lambda }_{\bf C}^{\circ}$. \label{zeta.cone} \end{prop} \begin{prop} If ${ \Lambda }$ is an $r$-dimensional finitely generated polyhedral cone, then ${\cal X}_{ \Lambda }(u)$ is a rational function of degree $-r$. In particular, ${\cal X}_{ \Lambda }(u)$ has a meromorphic extension to the whole complex space $V_{\bf C}$. \end{prop} {\em Proof.} It follows from Proposition \ref{zeta.cone}(i) that ${\cal X}_{ \Lambda }(\lambda u) = {\lambda}^{-r} {\cal X}_{ \Lambda }(u)$. Hence ${\cal X}_{ \Lambda }(u)$ has degree $-r$. In order to calculate ${\cal X}_{ \Lambda }(u)$, we subdivide the dual cone ${ \Lambda }^*$ into a union of simplicial subcones \[ { \Lambda }^* = \bigcup_{j} { \Lambda }^*_j .\] Then ${ \Lambda }$ is the intersection \[ { \Lambda } = \bigcap_j { \Lambda }_j. \] For ${\rm Re}(u) \in \bigcap_j { \Lambda }_{j}^{\circ}$, one has \[ {\cal X}_{ \Lambda }(u) = \sum_j {\cal X}_{{ \Lambda }_j}(u). \] By Proposition \ref{zeta.cone}(i),(ii), every function ${\cal X}_{{ \Lambda }_j}(u)$ is rational. \hfill $\Box$ \begin{dfn} {\rm Let $X$ be a smooth proper algebraic variety. Denote by $ \Lambda _{\rm eff} \subset {\rm Pic}(X)_{\bf R}$ the cone generated by classes of effective divisors on $X$. Assume that the anticanonical class $ \lbrack {\cal K}^{-1} \rbrack \in {\rm Pic}(X)_{\bf R}$ is contained in the interior of $ \Lambda _{\rm eff}$. We define the constant $\alpha(X)$ by \[ \alpha(X) = {\cal X}_{ \Lambda _{\rm eff}}( \lbrack {\cal K}^{-1} \rbrack). \] } \end{dfn} \begin{coro} If ${ \Lambda }_{\rm eff}$ is a finitely generated polyhedral cone, then $\alpha(X)$ is a rational number. \end{coro} \begin{exam} {\rm Let ${\bf P}_{\Sigma,K}$ be a smooth compactification of an anisotropic torus $T_K$. By \ref{simp}, $ \Lambda _{\rm eff} \subset {\rm Pic}({\bf P}_{\Sigma,K}) \otimes {\bf R}$ is a simplicial cone. Using \ref{zeta.cone} and the exact sequence \[ 0 \rightarrow PL(\Sigma)^G \rightarrow {\rm Pic}({\bf P}_{\Sigma,K}) \rightarrow H^1(G,M) \rightarrow 0 \] we obtain \[ {\cal X}_{ \Lambda _{\rm eff}}(u) = \frac{1}{h(T_K) u_1 \cdots u_r}, \] where $u = \varphi$, $\varphi(e_j) = u_j$ $(j =1, \ldots l)$. In particular, \[ \alpha({\bf P}_{ \Sigma }) = \frac{1}{h(T_K)}. \] } \end{exam} \begin{exam} {\rm Consider an example of a non-simplicial cone of Mori $ \Lambda _{\rm eff}$ in $V = {\rm Pic}(X)_{\bf R}$ where $X$ is a Del Pezzo surface of anticanonical degree 6. The cone ${ \Lambda }$ has 6 generators corresponding to exceptional curves of the first kind on $X$. We can construct $X$ as the blow up of 3 points $p_1, p_2, p_3$ in general position on ${\bf P}^2$. The exceptional curves are $C_1, C_2, C_3, C_{12}, C_{13}, C_{23}$, where $C_{ij}$ is the proper pullback of the line joining $p_i$ and $p_j$. If $u = u_1 [C_1] + u_2 [C_2] + u_3 [C_3] + u_{12}[C_{12}] + u_{13}[C_{13}] + u_{23} [C_{23}] \in \Lambda _{\rm eff}^{\circ}$, then \[ {\cal X}_{ \Lambda _{\rm eff}}(u) = \frac{ u_1 + u_2 + u_3 + u_{12} + u_{13} + u_{23} } {(u_1 + u_{23}) (u_2 + u_{13})(u_3 + u_{12})(u_1 + u_2 + u_3 ) (u_{12} + u_{13} + u_{23})} \] and \[ \alpha(X) = 1/12. \]} \end{exam} \begin{prop} Assume that ${ \Lambda }$ is a finitely generated polyhedral cone and ${ \Lambda } \cap -{ \Lambda } = 0$. Let $p_0$ and $p_1$ be two points in ${E}$ such that $p_1 \not\in { \Lambda }$ and $p_0 \in { \Lambda }^{\circ}$. Let $t_0$ be a positive real number such that $t_0p_0 + p_1 \in \partial { \Lambda }$. We define a meromorphic function in one complex variable $t$ as \[ Z(p_0,p_1, t) = {\cal X}_{ \Lambda }(tp_0 + p_1 ). \] Let $k$ be the codimension of the minimal face of ${ \Lambda }$ containing $t_0p_0 + p_1$. Then the rational function $Z(p_0,p_1, t)$ is analytic for ${\rm Re}(t) > t_0$ and it has a pole of order $k$ at $t = t_0$. \end{prop} {\em Proof.} As in the proof of the previous statement, we can subdivide the dual cone ${ \Lambda }^*$ into simplicial subcones ${ \Lambda }^*_j$ such that $t_0p_0 + p_1 \in \partial { \Lambda }_1$ and $t_0p_0 + p_1 \not\in \partial { \Lambda }_j$ $(j >1)$. It suffices now to apply Proposition \ref{zeta.cone}(i),(ii) to ${ \Lambda }_1$. \hfill $\Box$ \begin{coro} Assume that ${ \Lambda }$ is only locally polyhedral at the point $t_0p_0 + p_1$ and $k$ is the codimension of a minimal polyhedral face of ${ \Lambda }$ containing $t_0p_0 + p_1$. Then {\rm (i)} $Z(p_0,p_1, t)$ is an analytical function for ${\rm Re}(t) > t_0$. {\rm (ii)} $Z(p_0,p_1, t)$ has meromorphic continuation to some neigbourhood of $t_0$. {\rm (iii)} $Z(p_0,p_1, t)$ has a pole of order $k$ at $t = t_0$. \end{coro} \section{Distribution of rational points} \subsection{The method of Draxl} Let $\Sigma$ be a $G$-invariant regular fan, $\Sigma(1) = \Sigma_1(1) \cup \cdots \cup \Sigma_r(1)$ be the decomposition of $\Sigma(1)$ into $G$-orbits. We choose a representative $\sigma_j$ in each $\Sigma_j(1)$ $( j =1, \ldots, r)$. Let $e_j$ be the primitive integral generator of $\sigma_j$, $G_j \subset G$ be the stabilizer of $e_j$. Denote by $k_j$ the length of $G$-orbit of $e_j$, and by $K_j \subset E$ the subfield of $G_j$-fixed elements. Then $k_j = \lbrack K_j : K \rbrack$ $(j =1, \ldots, r)$. Consider the $n$-dimensional torus \[ T' := \prod_{j =1}^r R_{K_j/K}({\bf G}_m). \] Notice that the group ${\bf D}(\Sigma)$ can be identified with the $G$-module $\hat{T}'_K$. The homomorphism of $G$-modules $ M \rightarrow {\bf D}(\Sigma)$ induces the homomorphism $T' \rightarrow T$ and a map $$ \gamma:\, \prod_{j=1}^r {\bf G}_{m}({\bf A}_{K_j})/ {\bf G}_{m}(K_j) \rightarrow T({\bf A}_{K})/T(K) $$ We get a map of characters $$ \gamma^*:\, (T({\bf A}_{K})/T(K))^*\rightarrow \prod_{j =1}^r ({\bf G}_{m}({\bf A}_{K_j})/{\rm G}_m(K_j))^*. $$ \begin{rem}{\rm The kernel of $\gamma^*$ is dual to the obstruction group to weak approximation $A(T) $ defined above. } \label{obstr} \end{rem} Let \[ \chi \;: \; T({\bf A}_K) \rightarrow S^1 \subset {\bf C}^* \] be a topological character which is trivial on $T(K)$. Then $\chi \circ \gamma$ defines Hecke characters of the idele groups \[ \chi_j \; :\; {\bf G}_m({\bf A}_{K_j}) \rightarrow S^1 \subset {\bf C}^*. \] If $\chi$ is trivial on ${\bf K}_T$, then all characters $\chi_j$ $(j =1, \ldots, r)$ are trivial on the maximal compact subgroups in ${\bf G}_m({\bf A}_{K_j})$. We denote by $L_{K_j}(s,\chi_j)$ the Hecke $L$-function corresponding to the character $\chi_j$. The following statement is well-known: \begin{theo} The function $L_{K_j}(s,\chi_j)$ is holomorphic in the whole plane unless $\chi_j$ is trivial. In the later case, $L_{K_j}(s,\chi_j)$ is holomorphic for ${\rm Re}(s) >1$ and has a meromorphic extension to the complex plane with a pole of order $1$ at $s = 1$. \end{theo} We come to the main statement which describes the analytical properties of the Fourier transform of height functions. \begin{theo} Define affine complex coordinates $\{s_1, \ldots, s_r \}$ on the vector space $PL(\Sigma)_{\bf C}^G$ by $s_j = \varphi(e_j)$ $( j =1, \ldots, r)$. Then the Fourier transform $\hat{H}_{ \Sigma }(\chi, -\varphi)$ of the complex height function $H_{ \Sigma }(x,-\varphi)$ is always an analytic function for ${\rm Re}(s_j) > 1$ $(1 \leq j \leq l)$, and \[ \hat{H}_{ \Sigma }(\chi,-\varphi) \prod_{i=1}^r L^{-1}_{K_j}(s_j,\chi_j) \] has an analytic extension to the domain ${\rm Re}(s_j) > 1/2$ $(1 \leq j \leq r)$. \label{dmethod} \end{theo} {\em Proof.} The idea of the proof is essentialy due to Draxl \cite{drax1}. We have the Euler product \[ \hat{H}_{ \Sigma }(\chi,-\varphi) = \prod_{v \in {\rm Val}(K)} \hat{H}_{ \Sigma ,v}(\chi_v,-\varphi) \] In order to prove the above properties of $\hat{H}_{ \Sigma }(\chi, -\varphi)$, it is sufficient to investigate the product \[ \hat{H}_{ \Sigma ,S}(\chi,-\varphi) = \prod_{v \not\in S} \hat{H}_{ \Sigma ,v}(\chi_v,-\varphi). \] Choose a valuation $v \not\in S$. Then we obtain a cyclic subgroup $G_v = \langle \Phi_v \rangle \subset G$ generated by a lattice automorphism $\Phi_v \,:\, N \rightarrow N$ representing the local Frobenius element at place $v$. Let $l$ be the number of $G_v$-orbits in $ \Sigma (1)$. By \ref{integral.1}, \[ \hat{H}_{ \Sigma ,S}(\chi, -\varphi) = \prod_{v \not\in S} P\left( \Sigma, \Phi_v; \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}}, \ldots, \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_l)}} \right). \] By \ref{p-function}, we have \[ \hat{H}_{ \Sigma ,v}(\chi, -\varphi) = \frac{Q_{ \Sigma }\left( \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}}, \ldots, \frac{\chi_{v}(n_l)} {{q_v}^{\varphi(n_l)}} \right)} {(1- \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}} ) \cdots (1 - \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_l)}} )} . \] Moreover, \[ \prod_{v \not\in S} Q_{ \Sigma }\left( \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}}, \ldots, \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_r)}} \right) \] is an absolutely convergent Euler product for ${\rm Re}(s_j) > 1/2$ $( j =1, \ldots, l)$. It remains to show the relation between \[ \left( 1 - \frac{\chi_{v}(n_1)}{{q_v}^{\varphi(n_1)}} \right)^{-1} \cdots \left( 1 - \frac{\chi_{v}(n_l)}{{q_v}^{\varphi(n_l)}} \right)^{-1} \] and local factors of the product of the Hecke $L$-functions \[ \prod_{j =1}^r L_{K_j}(s_j, \chi_j). \] For this purpose, we compare two decompositions of $ \Sigma (1)$ into the disjoint union of $G_v$-orbits and $G$-orbits. Notice that for every $j \in \{ 1, \ldots, r \}$, the $G$-orbit $\Sigma^j(1)$ decomposes into a disjoint union of $G_v$-orbits \[ \Sigma_j(1) = \Sigma_{j1}(1) \cup \cdots \cup \Sigma_{j l_j}(1). \] Let $d_{ji}$ be the length of the $G_v$-orbit $ \Sigma ^j_i(1)$; i.e., we put $\{ d_{ji} \} = \{ d_1, \ldots, d_l \}$. One has \[ \sum_{ i =1}^{l_j} d_{ji} = k_j. \] and \[ l = \sum_{i =1}^r l_j. \] On the other hand, $l_j$ is the number of different valuations ${\cal V}_{j1}, \ldots, {\cal V}_{jl_j} \in {\rm Val}(K_j)$ over of $v \in {\rm Val}(K)$. Let $k_v$ be the residue field of $v \in {\rm Val}(K)$, $k_{{\cal V}_{ji}}$ the residue field of ${\cal V}_{ji} \in {\rm Val}(K_j)$. Then \[ d_{ji} = \lbrack k_{{\cal V}_{ji}} : k_v \rbrack. \] We put also $\{ n_1, \ldots, n_l \} = \{ n_{ji} \}$, where $n_{ji}$ denotes the sum of $d_{ji}$ generators of all $1$-dimensional cones of the $G_v$-orbit $ \Sigma _{ji}(1)$. Therefore, $\chi_v(n_{ji})$ is the ${\cal V}_{ji}$-adic component of the Hecke character $\chi_j$. Hence \[ \prod_{i =1}^{l_j} \left( 1 - \frac{ \chi_{v}(n_{ji})}{{q_v}^{\varphi(n_{ji})}} \right)^{-1} \] equals the product of the local factors \[ \prod_{{\cal V}_{ji}} \left( 1 - \frac{\chi_{{\cal V}_{ji}}}{q_{{\cal V}_{ji}}^{\varphi(n_{ji})}} \right)^{-1} \] of the Hecke $L$-function $L_{K_j}(s_j, \chi_j)$. \hfill $\Box$ \subsection{The meromorphic extension of $Z_{\Sigma}(\varphi)$} \begin{theo} For any $\varepsilon > 0$ there exists a $\delta >0$ and a constant $c({\varepsilon})$ such that \[ \mid L_K(s,\chi) \mid \leq c(\varepsilon)(\mbox{\rm Im}(s))^{\varepsilon}\; \; \mbox{\rm for}\; u= \mbox{\rm Re}(s) > 1 - \delta \] for every Hecke $L$-function $L_K(s,\chi)$ with a nontrivial nonramified character $\chi$. \label{estim} \end{theo} {\em Proof.} We use the following standard statement based on the Phragm\'en-Lindel\"of principle: \begin{lem} {\rm ( \cite{titchmarsh}, p.181)} Let $f(s) $ be a single valued analytic function in the strip $u_1 \le {\rm Re}(s)\le u_2$ satisfying the conditions: {\rm (i)} $|f(u + it )| < A_0 \exp ( e^{C|t|})$ for some real constants $A_0 >0$ and $ 0 < C < \pi/(u_2 - u_1)$; {\rm (ii)} $| f(u_1 + it) | \leq A_1 |t|^{a_1}$, $| f(u_2 + it) | \leq A_2 |t|^{a_2}$ for some constants $a_1$, $a_2$. Then for all $u_1 \leq u \leq u_2$, we have the estimate \[ | f(u + it) | \leq A_3 |t|^{a(u)} \] where \[ a(u) =a_1 \frac{u_2-u}{u_2-u_1} + a_2 {\frac{u-u_1}{u_2-u_1}}.\] \label{FL} \end{lem} Choose a sufficiently small $\delta_1$. Then $| L_K(s,\chi) |$ is bounded by $A_2(\delta_1)= \zeta_K(1+\delta_1)$ for ${\rm Re}(s) = 1 + \delta_1$. Consider the functional equation $L_K(s,\chi) = C(s)L_K(1-s,\overline{\chi})$. Since $\chi$ (as well as $\overline{\chi}$) is unramified, the function $C(s)$ depends only on the field $K$. Using standard estimates for $\Gamma$-factors in $C(s)$, we obtain $| L_K(s, \chi) | < A_1 ({\rm Im}(s))^{a_1}$ for ${\rm Re}(s) = - \delta_1$ and some sufficiently large explicit constants $A_1(\delta_1), a_1(\delta_1)$. We apply Lemma \ref{FL} to the $L$-function $L_K(s,\chi)$ where $u_1 = - \delta_1$, $u_2 = 1 + \delta_1$, and $a_2 = 0$. Then for $1 - \delta < {\rm Re}(s) < 1 + \delta_1$, one has \[ | L_K(s,\chi) | \leq A_3(\delta,\delta_1)({\rm Im} (s))^{a_1\frac{\delta + \delta_1}{1 + \delta + \delta_1}}. \] It is possible to choose $\delta$ and $\delta_1$ in such a way that \[ a_1(\delta_1) \frac{\delta + \delta_1}{1 + 2 \delta_1} < \varepsilon. \] \hfill $\Box$ \begin{theo} Let $s_j = \varphi(e_j)$ $( j =1, \ldots, r)$. Then the height zeta function $Z_{\Sigma}(\varphi)$ is holomorphic for ${\rm Re}(s_j) > 1$. There exists an analytic continuation of $Z_{\Sigma}(\varphi)$ to the domain ${\rm Re}(s_j) > 1- \delta$ such that the only singularities of $Z_{\Sigma}(\varphi)$ in this domain are poles of order $ \leq 1$ along the hyperplanes $s_j = 1$ $(j =1, \ldots, r)$. \label{extension.m} \end{theo} {\em Proof.} By the Poisson formula, \[ Z_{\Sigma}(\varphi) = \frac{1}{{\rm vol}(T^1({\bf A}_K)/T(K)) } \sum_{\chi \in (T({\bf A}_K)/T(K))^* } \hat{H}_{ \Sigma }(\chi, -\varphi). \] Since $H_{\Sigma}(x, -\varphi)$ is ${\bf K}_T$-invariant, we can assume that in the above formula $\chi$ runs over the elements of the group ${\cal P}$ consisting of characters of $T({\bf A}_K)$ which are trivial on ${\bf K}_T \cdot T(K)$. Let $J$ be a subset of $I = \{ 1, \ldots, r \}$. Denote by ${\cal P}_J$ the subset of ${\cal P}$ consisting of all characters $\chi \in {\cal P}$ such that the corresponding Hecke character $\chi_j$ is trivial if and only if $j \in J$. Then \[ Z_{\Sigma}(\varphi) = \sum_{J \subset I} Z_{\Sigma,J}(\varphi) \] where \[ Z_{\Sigma,J}(\varphi) = \frac{1}{{\rm vol}(T^1({\bf A}_K)/T(K)) } \sum_{\chi \in {\cal P}_J} \hat{H}_{ \Sigma }(\chi, -\varphi). \] Consider the logarithmic space \[ N_{{\bf R}, \infty} = \prod_{ v \in {\rm Val}_{\infty}(K)} T(K_v)/T({\cal O}_v) \] containing the full sublattice $T({\cal O}_K)/W(T)$ of ${\cal O}_K$-integral points of $T(K)$ modulo torsion. Let $\chi_{\infty}$ be the restriction of a character $\chi \in {\cal P}$ to $N_{{\bf R}, \infty}$. Then $\chi_{\infty}(x) = e^{2\pi i < x,y_{\chi} >}$ where $y_{\chi}$ is an element of the dual logarithmic space \[ M_{{\bf R}, \infty} = \prod_{ v \in {\rm Val}_{\infty}(K)} {\rm Hom}\,( T(K_v)/T({\cal O}_v), {\bf R}). \] Moreover, $y_{\chi}$ belongs to the dual lattice $(T({\cal O}_K)/W(T))^* \subset M_{{\bf R}, \infty}$. Let $u_1, \ldots, u_{r'}$ be a basis of the lattice $T({\cal O}_K)/W(T) \subset N_{{\bf R}, \infty}$, $f_1, \ldots, f_{r'}$ the dual basis of the dual lattice $(T({\cal O}_K)/W(T))^* \subset M_{{\bf R}, \infty}$. We extend \[ e^{2\pi i < x, f_1>} , \ldots, e^{2 \pi i<x, f_{r'}>} \] to some adelic characters $\eta_1, \ldots, \eta_{r'} \in {\cal P}$. Using \ref{subgroups} and the basis $\eta_1, \ldots, \eta_{r'}$, we can extend \[ \chi_{\infty} = \prod_{k = 1}^{r'} e^{2\pi i a_k < x,f_k >}, \;\; a_k \in {\bf Z} \] to a character \[ \tilde{\chi} = \prod_{k =1}^{r'} \eta_i^{a_k} \in {\cal P} \] such that $\tilde{\chi}_{\infty} = \chi_{\infty}$ and $\chi \cdot \tilde{\chi}^{-1}$ is a character of the finite group ${\bf cl}(T)$. We fix a character of $\chi_c$ of ${\bf cl}(T)$. Denote by ${\cal P}_{J,\chi_c}$ the set of all characters $\chi \in {\cal P}_J$ such that $\chi \cdot \tilde{\chi}^{-1} = \chi_c$. Then a character $\chi \in {\cal P}_{J,\chi_c}$ is uniquely defined by its archimedian component $\chi_{\infty}$. By \ref{dmethod}, \[ Q_{ \Sigma } (\chi,-\varphi) = \prod_{ v \not\in S_{\infty}} \hat{H}_{ \Sigma ,v}(\chi_v, \varphi) \prod_{i =1}^r L_{K_j}^{-1}(s_j, \chi_j) \] is absolutely convergent Euler product for ${\rm Re}(s_j) > 1 - \delta > 1/2$. By \ref{estim}, \[ \prod_{j \not\in J} L_{K_j}(s_j, \chi_j) < C(\varepsilon) \prod_{j \not\in J} {\rm Im}\, (s_j)^{\varepsilon} \] ${\rm Re}(s_j) > 1 - \delta$. We apply \ref{l-estimation} to the archimedian Fourier transform \[ \hat{H}_{ \Sigma ,\infty}(\chi, \varphi) = \prod_{ v \in S_{\infty}} \hat{H}_{ \Sigma ,v}(\chi_v, \varphi). \] Then, by \ref{lconver}, \[ \sum_{\chi \in {\cal P}_{J,\chi_c}} \hat{H}(\chi, -\varphi) \prod_{j \in J} \zeta_{K_j}(s_j)^{-1} \] is absolutely convergent for ${\rm Re}(s_j) > 1 - \delta$. Therefore, we have obtained that \[ Z_{ \Sigma ,J} (\varphi) \prod_{j \in J} \zeta_{K_j}(s_j)^{-1} \] is a holomorphic function for ${\rm Re}(s_j) > 1 - \delta$ and for any $J \subset I$. It remains to notice, that in the considered domain $\prod_{j \in J} \zeta_{K_j}(s_j)$ has only poles of order $1$ along hyperplanes $s_j = 1$. \hfill $\Box$ \subsection{Rational points of bounded height} Recall the standard tauberian statement: \begin{theo}\cite{delange} Let $X$ be a countable set, $F\, : \, X \rightarrow {\bf R}_{>0}$ a real valued function. Assume that \[ Z_F(s) = \sum_{x \in X} F(x)^{-s} \] is absolutely convergent for ${\rm Re}(s) > a> 0$ and has a representation $$ Z_F(s)=(s-a)^{-r}g(s) + h(s) $$ with $g(s)$ and $h(s)$ holomorphic for $Re(s)\ge a$, $g(a)\neq 0$, $r\in N$. Then for any $B>0$ there exists only a finite number $N(F,B)$ of elements $x \in X$ such that $F(x) \leq B$. Moreover, \[ N(F,B) = \frac{g(a)}{a(b-1)!} B^a (\log B)^{r-1}(1+o(1)). \] \label{tauberian} \end{theo} Let ${\cal L} = {\cal L}(\varphi_0)$ be a metrized invertible sheaf over a smooth compactification ${\bf P}_{\Sigma}$ of an anisotropic torus $T_K$ defined by a $G$-invariant fan $\Sigma$. We denote by $Z_{\Sigma,{\cal L}}(s) = Z_{\Sigma}(s \varphi_0)$ the restriction of $Z_{ \Sigma }(\varphi)$ to the line $s\lbrack {\cal L} \rbrack \subset {\rm Pic}({\bf P}_{ \Sigma })_{\bf R}$. Let $a({\cal L})$ be the abscissa of convergence of $Z_{\Sigma,{\cal L}}(s)$ and $b({\cal L})$ the order of the pole of $Z_{\Sigma,{\cal L}}(s)$ at $s = a({\cal L})$. By \ref{extension.m}, \[ a({\cal L}) \leq \min_{j = 1}^r \frac{1}{\varphi(e_j)}. \] By \ref{tauberian}, we obtain: \begin{theo} Assume that $\varphi(e_j) > 0$ for all $j =1, \ldots, r$; i.e., the class $\lbrack {\cal L} \rbrack$ is contained in the interior of the cone of effective divisors $\Lambda_{\rm eff}(\Sigma)$. Then there exists only finite number $N({\bf P}_{\Sigma}, {\cal L},B)$ of $K$-rational points $x \in T(K)$ having the ${\cal L}$-height $H_{\cal L}(x) \leq B$. Moreover, \[ N({\bf P}_{\Sigma}, {\cal L},B) =B^{a({\cal L})}\cdot (\log B)^{b({\cal L})-1}(1+o(1)) \;\; , B \rightarrow \infty. \] \end{theo} The following statement implies Batyrev-Manin conjectures about the distribution of rational points of bounded ${\cal L}$-height for smooth compactifications of anisotropic tori: \begin{theo} The number $a({\cal L})$ equals \[ a({\cal L})={\rm inf} \, \{\lambda \mid \lambda \lbrack {\cal L}\rbrack + \lbrack {\cal K} \rbrack \in \Lambda_{\rm eff}(\Sigma)\}; \] i.e., \[ a({\cal L}) = \min_{j = 1,r} \frac{1}{\varphi(e_j)}. \] Moreover, $b({\cal L})$ equals the codimension of the minimal face of $\Lambda_{\rm eff}(\Sigma)$ containing $a({\cal L})\lbrack {\cal L}\rbrack + \lbrack {\cal K} \rbrack$. \end{theo} {\em Proof. } By \cite{ono1}, we can choose the finite set $S$ such that the natural homomorphism \[ \pi_S \, : \, T(K) \rightarrow \prod_{v \not\in S} T(K_v)/T({\cal O}_v) = \prod_{v \not\in S} N_v \] is surjective. Denote by $T({\cal O}_S)$ the kernel of $\pi_S$ consisting of all $S$-units in $T(K)$. The group $T({\cal O}_S)/W(T)$ has the natural embedding in the finite-dimensional space \[ N_{S,{\bf R}} = \prod_{v \in S} T(K_v)/T({\cal O}_v) \otimes {\bf R} \] as a full sublattice. Let $\Delta$ be the fundamental domain of $T({\cal O}_S)/W(T)$ in $N_{S,{\bf R}}$. For any $x \in T(K)$, denote by $\overline{x}_S$ the image of $x$ in $N_{S,{\bf R}}$. Define $\phi(x)$ to be the element of $T({\cal O}_S)$ such that $\overline{x}_S - \phi(x) \in \Delta$. Thus, we have obtained the mapping \[ \phi_S \, : \, T(K) \rightarrow T({\cal O}_S). \] Define the new height function $\tilde{H}_{ \Sigma }(x,\varphi)$ on $T(K)$ by \[ \tilde{H}_{ \Sigma }(x, \varphi) = H_{ \Sigma }(\varphi, \phi_S(x)) \prod_{v \not\in S} H_{ \Sigma ,v}(x_v, \varphi). \] \noindent Notice the following easy statement: \begin{lem} Choose a compact subset ${\bf K} \subset {\bf C}^r$ such that ${\rm Re}(s_j) > \delta$ $( j =1, \ldots, r)$ for $\varphi \in {\bf K}$. Then there exist positive constants $C_1$, $C_2$ such that \[ 0 < C_1 < \frac{\tilde{H}_{ \Sigma }(x, \varphi)}{H_{ \Sigma }(x, \varphi)} < C_2, \; \mbox{\rm for}\; \varphi \in {\bf K}, \; x \in T(K). \] \label{compare} \end{lem} Define $\tilde{Z}_{ \Sigma }(\varphi)$ by \[ \tilde{Z}_{ \Sigma }(\varphi) = \sum_{x \in T(K)} \tilde{H}_{ \Sigma }(x, -\varphi). \] Then $\tilde{Z}_{ \Sigma }(\varphi)$ splits into the product \[ \tilde{Z}_{ \Sigma }(\varphi) = \prod_{v \not\in S} \left( \sum_{z \in N_v} H_{ \Sigma ,v}(z,-\varphi) \right) \cdot \left( \sum_{ u \in T({\cal O}) } H_{ \Sigma }(u, -\varphi) \right). \] By \cite{drax1}, the Euler product \[ \prod_{j =1}^r \zeta_{K_j}(s_j) \prod_{v \not\in S} \left( \sum_{z \in N_v} H_{ \Sigma ,v}(z,-\varphi) \right) \] is a holomorphic function without zeros for ${\rm Re}(s_j) > 1/2$. On the other hand, \[ \sum_{ u \in T({\cal O}) } H_{ \Sigma }(u, -\varphi) \] is an absolutely convergent series nonvanishing for ${\rm Re}(s_j) > 0$. Therefore, $\tilde{Z}_{ \Sigma }(\varphi)$ has a meromorphic extension to the domain ${\rm Re}(s_j) > 1/2$ where it has poles of order $1$ along the hyperplanes $s_j = 1$. By \ref{compare} and \ref{tauberian}, $\tilde{Z}_{ \Sigma }(\varphi)$ and $Z_{ \Sigma }(\varphi)$ must have the same poles in the domain ${\rm Re}(s_j) > 1 - \delta$. Therefore, $Z_{ \Sigma }(\varphi)$ has poles of order $1$ along the hyperplanes $s_j = 1$. By taking the restriction of $Z_{ \Sigma }(\varphi)$ to the line $\varphi = s \varphi_0$, we obtain the statement. \hfill $\Box$ \subsection{The residue at $s_j = 1$} Recall the definition of the Tamagawa number of Fano varieties \cite{peyre}. This definition immediately extends to arbitrary algebraic varieties $X$ with a metrized canonical sheaf ${\cal K}$. Let $x_1, \ldots, x_d$ be local analytic coordinates on $X$. They define a homeomorphism $f\,: \, U \rightarrow {K_v}^d$ in $v$-adic topology between an open subset $U \subset X$ and $f(U) \subset {K_v}^d$. Let $dx_1 \cdots dx_d$ be the Haar measure on $K_v^d$ normalized by the condition \[ \int_{{\cal O}_v^d} dx_1 \cdots dx_d = \frac{1}{(\sqrt{\delta_v})^d} \] where $\delta_v$ is the absolute different of $K_v$. Denote by $dx_1 \wedge \cdots \wedge dx_d$ the standard differential form on $K_v^{d}$. Then $g = f^*(dx_1 \wedge \cdots \wedge dx_d)$ is a local analytic section of the metrized canonical sheaf ${\cal K}$. We define the local measure on $U$ by \[ \omega_{{\cal K},v} =f^*( \| g(f^{-1}(x)) \|_v dx_1 \cdots dx_d) . \] The adelic Tamagawa measure $\omega_{{\cal K},S}$ is defined by \[ \omega_{{\cal K},S} = \prod_{v \in {\rm Val}(K)} \lambda_v^{-1} \omega_{{\cal K},v} \] where $\lambda_v = L_v(1, {\rm Pic}(X_{\overline{K}});\overline{K}/K))$ if $v \not\in S$, $\lambda_v = 1$ if $v \in S$. \begin{dfn} {\rm \cite{peyre} Let $\overline{X(K)}$ be the closure of $X(K) \subset X({\bf A}_K)$ in the direct product topology. Then the {\em Tamagawa number} of $X$ is defined by \[ \tau_{\cal K}(X) = \lim_{s \rightarrow 1} (s-1)^r L_S(s, {\rm Pic}(X_{\overline{K}});\overline{K}/K)) \cdot \int_{\overline{X(K)}} \omega_{{\cal K},S}. \]} \end{dfn} \begin{prop} Let ${\cal K} = {\cal L}(-\varphi_{\Sigma})$ be the metrized canonical sheaf on a toric variety ${\bf P}_{\Sigma}$. Then the restriction of the $v$-adic measure $\omega_{{\cal K},v}$ to $T(K_v) \subset {\bf P}_{\Sigma}(K_v)$ coincides with the measure \[ H_{ \Sigma ,v}(x, -\varphi_{\Sigma}) \omega_{\Omega,v}. \] \label{restriction} \end{prop} {\em Proof.} The rational differential $d$-form $\Omega$ is a rational section of ${\cal K}$. By definition of the $v$-adic metric on ${\cal L}(-\varphi_{\Sigma})$, $H_{ \Sigma ,v}(x, -\varphi_{\Sigma})$ equals the norm $\| \Omega \|_v$ of the $T$-invariant section $\Omega$. This implies the statement. \hfill $\Box$ \begin{prop} One has \[ \int_{\overline{T(K)}} \omega_{{\cal K},S} = \int_{\overline{{\bf P}_{ \Sigma }(K)}} \omega_{{\cal K},S}. \] \end{prop} {\em Proof.} Since $\overline{{\bf P}_{ \Sigma }(K)} \setminus \overline{T(K)}$ is a subset of ${\bf P}_T({\bf A}_K) \setminus T({\bf A}_K)$, it is sufficient to prove that \[ \int_{T(K_v)} \omega_{{\cal K},v} = \int_{{\bf P}_{ \Sigma }(K_v)} \omega_{{\cal K},v}. \] Since the measure $\omega_{{\cal K},v}$ is $T({\cal O}_v)$-invariant and the stabilizer in $T({\cal O}_v)$ of any point $x \in {\bf P}_{ \Sigma }(K_v) \setminus T(K_v)$ is uncountable, the $\omega_{{\cal K},v}$-volume of ${\bf P}_{ \Sigma }(K_v) \setminus T(K_v)$ is zero. \label{2int} \hfill $\Box$. \begin{theo} Let $\Theta(\Sigma,K)$ be the the residue of the zeta-function $Z_{\Sigma}(\varphi)$ at $s_1 = \cdots = s_r = 1$. Then \[ \Theta(\Sigma,K) = \alpha({\bf P}_{\Sigma})\beta({\bf P}_{\Sigma}) \tau_{\cal K}({\bf P}_{\Sigma}). \] \end{theo} {\em Proof.} By the Poisson formula, \[ Z_{\Sigma}(\varphi) = \frac{1}{{\rm vol}(T^1({\bf A}_K)/T(K)) } \sum_{\chi \in (T({\bf A}_K)/T(K))^* } \hat{H}_{ \Sigma }(\chi, -\varphi). \] By \ref{extension.m}, the residue of $Z_{\Sigma}(\varphi)$ at $s_1 = \cdots = s_r = 1$ appears from $Z_{ \Sigma ,I}(\varphi)$ containing only the terms $\hat{H}_{ \Sigma }(\chi, -\varphi)$ such that $\chi_1, \ldots , \chi_r$ are trivial characters of ${\bf G}_m({\bf A}_{K_j})/{\bf G}_m(K_j)$ $(j = 1, \ldots, l)$; i.e., $\chi$ is a character of the finite group $A(T)$. We apply again the Poisson formula to the finite sum \[ Z_{ \Sigma ,I}(\varphi) = \frac{1}{{\rm vol}(T^1({\bf A}_K)/T(K)) } \sum_{\chi \in (A(T))^* } \hat{H}_{ \Sigma }(\chi, -\varphi). \] Using \ref{weak}, \ref{weak1}, \ref{obstr}, we have \[ Z_{ \Sigma ,I}(\varphi) = \frac{\beta({\bf P}_{\Sigma})}{i(T_K)\cdot{\rm vol}(T^1({\bf A}_K)/T(K)) } \int_{\overline{T(K)}} H_{ \Sigma }(x,-\varphi) \omega^1_{\Omega,S}. \] Notice that $\omega^1_{\Omega,S} = \omega_{\Omega,S}$ for anisotropic tori. Now we assume that $\varphi(e_1) = \ldots = \varphi(e_r) = s$. Our purpose is to compute the constant \[ \Theta(\Sigma,K) = \lim_{ s \rightarrow 1} (s-1)^r Z_{ \Sigma ,I}(s\varphi_{\Sigma}). \] By \ref{tamagawa1}, \ref{tamagawa2}, \[ \Theta(\Sigma,K) = \frac{\beta({\bf P}_{\Sigma})}{h(T_K)} L_S^{-1}(1, T; E/K) \lim_{ s \rightarrow 1} (s-1)^r \int_{\overline{T(K)}} H_{ \Sigma }(x,-s\varphi_{\Sigma}) \omega_{\Omega,S}. \] Notice that $\overline{T(K)}$ contains $T(K_v)$ for $v \not\in S$. Denote by $\overline{T(K)}_S$ the image of $\overline{T(K)}$ in $\prod_{v \in S} T(K_v)$. By \ref{loc-integ}, we have \[ \int_{\overline{T(K)}} H_{ \Sigma ,v}(x,-s\varphi_{\Sigma}) \omega_{\Omega,S} = \] \[ = \prod_{v \not\in S} \int_{T(K_v)} H_{ \Sigma ,v}(x,-s\varphi_{\Sigma}) d\mu_v \cdot \int_{\overline{T(K)}_S} \prod_{v \in S} H_{ \Sigma }(x,-s\varphi_{\Sigma}) \omega_{\Omega,v} = \] \[ = L_S(s, T;E/K) \cdot \prod_{v \not\in S} \left( \sum_{k =0}^d \frac{{\rm Tr}\, \Phi_v(i)}{q_v^{ks}} \right) \cdot \int_{\overline{T(K)}_S} \prod_{v \in S} H_{ \Sigma ,v}(x,-s\varphi_{\Sigma}) \omega_{\Omega,v}. \] \[ L_S^{-1}(1, T; E/K) \omega_{\Omega,S} = \prod_{v \in {\rm Val}(K)} \omega_{\Omega,v}. \] By \ref{p-function} and \ref{integral.1}, \[ L_S^{-1}(s, {\rm Pic}({\bf P}_{\Sigma,E}); E/K) \prod_{v \not\in S} \left( \sum_{k =0}^d \frac{{\rm Tr}\, \Phi_v(i)}{q_v^{ks}} \right) \] has no singularity at $s =1$. Moreover, by \ref{restriction}, \[ \prod_{v \not\in S} L_S^{-1}(1, {\rm Pic}({\bf P}_{\Sigma,E});E/K) \left( \sum_{k =0}^d \frac{{\rm Tr}\, \Phi_v(i)}{q_v^{k}} \right) = \prod_{v \not\in S} \int_{T(K_v)} \lambda_v^{-1} \omega_{{\cal K},v}. \] Therefore \[ \Theta(\Sigma,K) = \frac{\beta({\bf P}_{\Sigma})}{h(T_K)} \lim_{ s \rightarrow 1} (s-1)^r L_S(s, {\rm Pic}({\bf P}_{\Sigma,E}); E/K) \int_{\overline{T(K)}} \omega_{{\cal K},S}. \] It remains to apply \ref{2int}. \hfill $\Box$ Using \ref{tauberian}, we obtain: \begin{coro} Let $T$ be an anisotropic torus and ${\bf P}_{\Sigma}$ its smooth compactification $($notice that we do not need to assume that ${\bf P}_{\Sigma}$ is a Fano variety$)$. Let $r$ be the rank of ${\rm Pic}({\bf P}_{\Sigma,K})$. Then the number $N( {\bf P}_{\Sigma},{\cal K}^{-1}, B)$ of $K$-rational points $x \in T(K)$ having the anticanonical height $H_{{\cal K}^{-1}}(x) \leq B$ has the asymptotic \[ N({\bf P}_{\Sigma},{\cal K}^{-1}, B) = \frac{\Theta(\Sigma,K)}{(r-1)!} \cdot B (\log B)^{r-1}(1+o(1)), \hskip 0,3cm B\rightarrow \infty.\] \end{coro}
1996-01-22T01:48:53
9411
alg-geom/9411013
en
https://arxiv.org/abs/alg-geom/9411013
[ "alg-geom", "math.AG" ]
alg-geom/9411013
Jean-Bruno Erismann
A. Belkasri, M. Hamade
A New Approach to Characterise all the Transitive Orientations for an Undirected Graph
20 pages, LaTeX, macro epsf.sty
null
null
CPT-94/P.3069
null
A new approach to find all the transitive orientations for a comparability graph (finite or infinite) is presented. This approach is based on the link between the notion of ``strong'' partitive set and the forcing theory (notions of simplices and multiplices). A mathematical algorithm is given for the case of a comparability graph which has only non limit sub-graphs.
[ { "version": "v1", "created": "Fri, 18 Nov 1994 08:16:05 GMT" } ]
2008-02-03T00:00:00
[ [ "Belkasri", "A.", "" ], [ "Hamade", "M.", "" ] ]
alg-geom
\section{Introduction} The problem of the transitive orientation for a comparability graph was studied by Golumbic using the forcing theory [1]. The problem was solved for {\it finite} comparability graphs and an algorithm was given which gives one transitive orientation for a {\it finite} comparability graph. The purpose of this paper is to study the transitive orientation for the case of {\it infinite} comparability graphs. The results for the finite case could not be extended to the infinite case because of the finite type like of the approach used in Ref.[1]. We were then obliged to consider a new approach, but remaining within the forcing theory. This was possible by introducing the notion of a {\it ``strong'' partitive set}. It happens that this idea permits to solve the problem of the transitive orientations by inducing a lot of characteristics of the forcing theory, from one hand and the undirected graphs, in general, from the other hand. The paper is organized as follows: the section 2 is devoted to the definitions and notations used throughout the paper and also to recall the main results of the forcing theory which still valid for infinite graphs. In section 3, we establish narrow links between the notion of a {\it ``strong'' partitive set} and {\it simplices} (and {\it multiplices}) which are the principal touls of the forcing theory. In section 4, we used the results of section 3 to face the problem of transitive orientation of comparability graphs. We proved a theorem which is in fact a mathematical algorithm which gives all the transitive orientations for a comparability graph which has all its sub-graphs non limit. \section{Preliminairies} This section is devoted to the definitions and notations which will be used throughout this article. We also recall some results about the partitive sets and the implication classes. These results can be found with more details in Ref.[1]. We consider here any kind of graphs; finite or infinite. In what follows we denote by $G=(V,E)$ any graph, where $V$ is the set of vertices, and by $E(\subseteq V^2)$ the set of edges. Directed edge will be denoted by $(a,b)$ \newline (for $a,b\in V$) and an undirected one is denoted by $ab, ab=\left\{(a,b),(b,a)\right\}$. We say that a graph $G=(V,E)$ is {\it empty} if $E=\phi$. For any $X\subseteq V$, $G(X)=(X,E(X))$ will denote the sub-graph of $G$ induced on $X$; where $$ E(X)=\left\{(a,b)\in E\quad ;\quad \{a,b\}\subseteq X\right\} $$ we define also the set of vertices $\widetilde{A}$ spanned by a set of edges $A$ as $$ \widetilde{A} =\left\{a\in V;\ \hbox{there exists}\ b\in V \hbox{so that}\ (a,b)\in A \ \hbox{or}\ (b,a)\in A\right\} $$ \subsection{Partitive set and ``strong'' partitive set} Let $\cong$ be a binary relation acting on $V^2$ defined by $$(a,b)\cong (c,d)\Leftrightarrow\left(\{(a,b),(c,d)\}\cap E=\phi\ \hbox{or}\ \{(a,b),(c,d)\}\subseteq E\right) $$ This means that the edges $(a,b)$ and $(c,d)$ are both belonging to $E$ or are both out of $E$. Let $X\subseteq V$ be a sub-set of $V$. $X$ is a partitive set of $G$ (or $V$) if: $$ \hbox{for every}\ \{a,b\}\subseteq X\ \hbox{and every}\ c\in V-X $$ $$ \hbox{we have}\ (a,c)\cong (b,c)\ \hbox{and}\ (c,a)\cong (c,b) $$ It means that the elements of $X$ are related to any external element in the same manear. The notion of the partitive set is the analogue of the notion of interval in an ordered set. We will denote by $I(G)$ the class of partitive sets of the graph $G$. A partitive set is trivial if it is a singleton or equal to $V$. By $I^{\star}(G)$ we will denote the class of non-trivial partitive sets. \noindent A graph is indecomposable if all its partitive sets are trivial; otherwise, it is decomposable. \noindent A partitive set $X\in I(G)$ is called a ``strong'' partitive set of $G$(or $V$) if for every partitive set $Y\in I(G)$ so that $Y\cap X\not =\phi$, we have either $X\subseteq Y$ or $Y\subseteq X$. It means that the eventuality that $X$ and $Y$ shear only a common part is excluded. \noindent We will denote by $I_F(G)$ the class of ``strong'' partitive sets of $G$ and by $I^{\star}_F(G)$ the class of the non-trivial ``strong'' partitive sets. We say that $G$ is {\it limit} if $I_F(G)$ do not contain any element different from $V$ which is maximal for the inclusion$(\subseteq)$; otherwise it is {\it non limit}. \newpage \noindent {\it Isomorphism}: Two graphs $G=(V,E)$ and $G'=(V',E')$ are said to be isomorphic if: \begin{enumerate} \item[(i)] {there is a bijection $f:V\to V'$;} \item[(ii)] {$f$ preserves the edges, i.e., for every $\{a,b\}\subseteq V:$} \end{enumerate} $$ (a,b)\in E\Leftrightarrow (f(a),f(b))\in E'\ . $$ \noindent {\it Quotient graph}: Let $G=(V,E)$ be a graph and $P$ a partition of $G$ made of partitive sets $(P\subseteq I(G))$. We define the {\it quotient graph} of $G$ by $P$, denoted by $G/P$, as the isomorphic graph to $G(f(P))$, where $f$ is a choice function from $P$ to $V$, i.e., $X\in P\Rightarrow f(X)\in X$. \begin{proposition} \label{prop2.1} Let $P$ be a partition of partitive sets (respectively of ``strong'' partitive sets) of $G=(V,E)$ and $X$ a sub-set of $P$. We have then\\ $X\in I(G/P)$ (respectively $X\in I_F(G/P)$) if and only if: $\cup X\in I(G)$ \noindent $(\hbox{respectively}\ \cup X\in I_F(G))\ ,$ \noindent where $\cup X$ means the union of the vertices constituting the partitive sets (respectively the ``strong'' partitive sets) of $X$. \end{proposition} \subsection{Implication classes and simplices} {\it Comparability graph} {\it (or transitively orientable graph)}: Let $G=(V,E)$ be an undirected graph. $G$ is a comparability graph if there exists an orientation of the edges of $G$ so it constitutes a partial order on $V$. Comparability graphs are also known as transitively orientable graphs or partially orderable graphs. \noindent{\it Implication classes}: Let us define the binary relation $\Gamma$ on the edges of an undirected graph $G=(V,E)$ as follows: $$ (a,b)\Gamma(a',b')\Leftrightarrow\left\{\begin{array}{llll} \hbox{either}& a=a'&\hbox{and}&bb'\not\in E\\ \hbox{or}&b=b'&\hbox{and}&aa'\not\in E.\\ \end{array}\right. $$ We say that $(a,b)$ directly forces $(a',b')$ whenever $(a,b)\Gamma (a',b')$. In graphical representation $ab$ and $a'b'$ will have only one common vertex and the orientation is so that arrows are both pointed to the extremities or are both pointed to the common vertex (see Figure 2.1). Notice that $\Gamma$ is not transitive. In Figure~2.1 $(a,d)$ forces directly $(a,b)$ and $(a,c)$. The edges $(b,c)$ and $(c,b)$ are not forced by any other edge. \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 2.1} \end{center} Let $\Gamma^{\star}$ be the transitive closure of $\Gamma$. It is easy to verify that $\Gamma^{\star}$ is an equivalence relation on $E$ and the equivalence classes related to $\Gamma^{\star}$ are what one call the {\it implication classes} of $G$. In what follows we will see that it is useful to define what one call color classes of $G$ (or shortly colors of $G$). If $A$ is an implication class, the color class associated to $A$ and denoted by $\widehat A$, is the union of $A$ and $A^{-1}$ \\ $(\widehat A=A\cup A^{-1}\subseteq E)$; where $A^{-1}=\{(a,b)\in E$ with $(b,a) \in A\}$. \begin{theorem}\label{th2.2} ({\it Golumbic Ref.[1]}) Let $G=(V,E)$ be a comparability graph and $A$ an implication class of $G$. If $O=(V,E')$ is a partial order associated to $G$, we have necessarily either $E'\cap\widehat A=A$ or $E'\cap \widehat A=A^{-1}$ and, in either case, $A\cap A^{-1}=\phi$. \end{theorem} \begin{lemma}\label{lem2.3} {\it (The Triangle Lemma).} Let $A,B$ and $C$ be implication classes of an undirected graph $G=(V,E)$ with $A \not = B$ and $A\not = C^{-1}$ and having edges $(a,b)\in C$, $(a,c)\in B$ and $(b,c)\in A$, we have then \begin{enumerate} \item[(i)] $(b',c')\in A\Rightarrow \left((a,b')\in C\right.$ and $\left.(a,c')\in B\right)$ ; \item[(ii)] $\left((b',c')\in A\right.$ and $\left.(a',b')\in C\right)\Rightarrow (a',c')\in B$ ; \item[(iii)] $a\not\in\widetilde A$. \end{enumerate} \end{lemma} The following results are consequences of the triangle lemma. \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 2.2 } {\it Illustration of the triangle lemma.} \end{center} \begin{theorem}\label{th2.4} ({\it Golumbic Ref.[1]}) Let $A$ be an implication class of an undirected graph $G=(V,E)$. Exactly one of the following alternatives holds: \begin{enumerate} \item[(i)] $A=A^{-1}=\widehat A$ and $\widehat A$ is not transitively orientable ; \item[(ii)] $A\cap A^{-1}=\phi$ and then $A$ and $A^{-1}$ are two transitive orientations of $\widehat A$. \end{enumerate} \end{theorem} \begin{proposition}\label{prop2.5} Let $X$ be a partitive set $(X\in I(G))$ and $\widehat A$ a color class of an undirected graph $G=(V,E)$ so that $E(X)\cap\widehat A\not= \phi$ we have then $\widehat A\subseteq E(X)$. \end{proposition} \begin{proposition}\label{prop2.6} If $\widehat A$ is a color class of $G=(V,E)$ then $\widetilde A$ is a partitive set of $G$ $(\widetilde A\in I(G))$. \end{proposition} \noindent {\it Simplex} : Let $G=(V,E)$ be an undirected graph. A $K_{r+1}$ complete sub-graph of $G$, $S=(V_S,E_S)$ on $r+1$ vertices is called a {\it simplex of rank} $r$ if each undirected edge $ab$ of $E_S$ is contained in a different color class of $G$. A simplex is {\it maximal} if it is not properly contained in any larger simplex. The {\it multiplex} $M$ generated by a simplex $S$ of rank $r$ is defined to be the part of $E$ constituted of all edges which their color classes are present in the simplex $S$ ($M(S)={{\cup}_{\widehat A\cap E_S \neq\phi}}\widehat A$). $M$ is said also a multiplex of rank $r$. The multiplex $M$ is said to be maximal if $S$ is maximal. We will denote by $\widehat M$ the collection of color classes present in the multiplex $M$. \section{Connection between multiplices and ``strong'' partitive sets} In this section we make connection between the notion of ``strong'' partitive set and multiplices. In our knowledge this was never made before. It happens that this connection allows us to recover results of Golumbic [1] for finite graphs and generalize them to the infinite graphs. The results presented in this section will be used in the following section to state a theorem on the decomposability of undirected graphs. In what follows $G$ will denote an undirected graph unless other mention is pointed out. \begin{proposition}\label{prop3.1} If $\widehat A$ and $\widehat B$ are two color classes of $G=(V,E)$ we have then $(\widehat A=\widehat B)\Leftrightarrow (\widetilde A=\widetilde B)$. \end{proposition} {\it Proof.} It is obvious that $\widehat A=\widehat B \Rightarrow \widetilde A=\widetilde B$. Let us suppose that $\widetilde A= \widetilde B$ and $\widehat A\not=\widehat B$, we have then for every $x\in\widetilde A$ there exists $a\in\widetilde A$ and $b\in\widetilde B$ so that $ax\in \widehat A$ and $bx\in\widehat B$. Since $\widehat A\not=\widehat B$ we have necessarily $ab\in E$. Let $C$ be the color class which contains $ab$. We have then two alternatives: \begin{enumerate} \item[(i)] $\widehat C\not=\widehat A\not= \widehat B\Rightarrow b\not\in\widetilde A$ (Triangle Lemma) which is absurd since\\ $b\in \widetilde B=\widetilde A$; \item[(ii)] $\widehat C=\widehat A\not= \widehat B\Rightarrow a\not\in\widetilde B$ (Triangle Lemma) which is also absurd, since $a\in \widetilde A=\widetilde B$. \end{enumerate} \begin{proposition}\label{prop3.2} Let $\widehat A$ be a color class of $G=(V,E)$ and $X$ a partitive set of $G$ so that $X\subset \widetilde A$, there exists then $a\in\widetilde A-X$ so that for every $x\in X$, $ax\in\widehat A$. \end{proposition} {\it Proof.} Let $x\in X\subset\widetilde A$, there exists $a\in\widetilde A$ so that $ax \in \widehat A$. Necessarily $a\in \widetilde A-X$ (otherwise $(ax\in\widehat A\cap E(X))\Rightarrow\widehat A\subseteq E(X)$~[P.\rf{prop2.5}) then for every $x\in X$ we have $xa\in E$, since $X\in I(G)$. Let us assume that there exists $y\in X$ and a color $ \widehat B\not=\widehat A$ so that $ay\in \widehat B$, then $xy\in E$. Let $\widehat C$ be the color which contains $xy$. We have $\widehat C\not=\widehat A$ and $\widehat C\not=\widehat B$ (otherwise: $(\widehat C=\widehat A\Rightarrow \widetilde A\subseteq X)$ and $(\widehat C=\widehat B\Rightarrow \widetilde B\subseteq X)$ which contradicts the fact that $a\in(\widetilde A\cap\widetilde B)-X$). Then using the triangle lemma we will have $y\not\in\widetilde A$ which is absurd. Finally we have for every $x\in X,ax\in \widehat A$. \begin{theorem}\label{th3.3} Let $\widehat A$ and $\widehat B$ be two color classes of $G=(V,E)$ so that $\widetilde A-\widetilde B\not=\phi$ and $\widetilde B-\widetilde A\not=\phi$. We have then that $X=\widetilde A\cap \widetilde B$ is a ``strong'' partitive set of $G$. \end{theorem} {\it Proof.} Since $\widetilde A$ and $\widetilde B$ are partitive sets[P.2.6] we have that $X=\widetilde A\cap \widetilde B$ is a partitive set. If $X=\phi$ or $X$ is a singleton $(|X|=1)$ then $X$ is a ``strong'' partitive set. Let us suppose now that $|X|>1$. Let $Y$ be a partitive set $(Y\in I(G))$ so that $X\cap Y\not=\phi$ and $Y-X\not=\phi$ and let $z\in X\cap Y$ and $y\in Y-X$. We have to show that $X\subset Y$. Applying~[P.\rf{prop3.2} we have: $X\subset\widetilde A\Rightarrow$ there exists $a\in\widetilde A-X$ so that for every $x\in X,\ ax\in\widehat A$; $X\subset\widetilde B\Rightarrow$ there exists $b\in\widetilde B-X$ so that for every $x\in X,\ bx\in\widehat B$. Thus $za\in\widehat A$ and $zb\in \widehat B$. If $a\in Y$ then $az\in E(Y)\cap\widehat A\not=\phi$, thus $X\subseteq\widetilde A\subseteq Y$~[P.\rf{prop2.5}. If $b\in Y$ then $bz\in E(Y)\cap\widehat B\not=\phi$, thus $X\subseteq\widetilde B\subseteq Y$~[P.\rf{prop2.5}. Let us now suppose that $\{a,b\}\cap Y=\phi$, we have then: \begin{enumerate} \item[(i)] if $y\not\in\widetilde A$ then: $(\{y,z\}\subseteq Y\in I(G)$ and $za\in E)\Rightarrow ya\in E$; \\ $(\widetilde A\in I(G)$ and $ay\in E)\Rightarrow$ for every $x\in X\subset \widetilde A, yx\in E$. \\ Let $\widehat K$ be the color class of $ya$ and $\widehat C$ that one of $yz$. We have then: $(y\in(\widetilde K\cap\widetilde C)-\widetilde A$, $a\in\widetilde K-Y$ and $yz\in\widehat C\cap E(Y)\not=\phi)$ implies that $(\widetilde K\not=\widetilde A\not= \widetilde C)\Leftrightarrow (\widehat K\not=\widehat A\not=\widehat C)$, where at the last step we have applied~[P.\rf{prop3.1}. \\ Let $v\in X$ (with $v\not= z$) and $\widehat D$ be the color of $yv$. We have then: \\ $\widehat K\not=\widehat A\not=\widehat D$ (since $\widehat K\not=\widehat A,yv\in\widehat D\cap E(Y\cup X)\not=\phi$ and\\ $a\in (\widetilde A\cap\widetilde K)-(Y\cup X)$). The two tricolor triangles $(a,y,z;\widehat K,\widehat A,\widehat C)$ and $(a,y,v;\widehat K,\widehat A,\widehat D)$ have two common colors, thus $\widehat D=\widehat C$, thus for every $x\in X, yx\in\widehat C$ which implies that $X\subset\widetilde C\subseteq Y$ ($\widetilde C\subseteq Y$ because $yz\in \widehat C\cap E(Y)\neq\phi$[P.2.5]) (see Figure~3.1). \item[(ii)] If $y\in \widetilde A$ then $(y\not\in X=\widetilde A\cap\widetilde B$ and $y\in\widetilde A)\Rightarrow y\not\in\widetilde B$. By replacing $a$ by $b$ and $\widehat A$ by $\widehat B$ up here in (i), we get $X\subset \widetilde C \subseteq Y$. \end{enumerate} \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 3.1} {\em We have : $\widehat D=\widehat C$[Triangle Lemma].} \end{center} \smallskip \begin{theorem}\label{th3.4} Let $X$ and $Y$ be two ``strong'' partitive sets of $G=(V,E)$ so that $X\cap Y=\phi$. $X$ and $Y$ can be related to each other by only one color at most. \end{theorem} \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 3.2} {\em If $X$ and $Y$ are ``strong'' partitive sets we have necessarily $\widehat A = \widehat B$}. \end{center} \medskip {\it Proof.} Let $\widehat A$ and $\widehat B$ be two colors connecting $X$ and $Y$. We have then $X\subset \widetilde A\cap\widetilde B$ and $Y\subset \widetilde A\cap \widetilde B$ (applying [P.2.6] and the definition of the ``strong'' partitive set). Let us suppose that $\widehat A\not=\widehat B$ and define the following sets $$I_A=\left\{Z\in I_F(G)\ \hbox{with}\ Z\subset\widetilde A\right\}\ \hbox{and}\ I_B=\left\{Z\in I_F(G)\ \hbox{with}\ Z\subset\widetilde B\right\}$$ we have $\cup I_A=\widetilde A$ and $\cup I_B=\widetilde B$ since: $\cup I_A\subseteq \widetilde A$ and if $x\in \widetilde A$, then $\{x\}\in I_F(G)\cap I_A$, thus $x\in \cup I_A$ and then $\widetilde A\subseteq \cup I_A$. The same situation holds for $I_B$. Since $\{X,Y\}\subseteq I_A\cap I_B$ we have $\widetilde A\cap \widetilde B\not=\phi$, $\widehat A\cap E(\widetilde A\cap \widetilde B)\not=\phi$ and $\widehat B\cap E(\widetilde A\cap \widetilde B)\not= \phi$ we have then only two alternatives: \begin{enumerate} \item[(i)] either $\widetilde A=\widetilde B\Rightarrow\widehat A=\widehat B$~[P.\rf{prop3.1} \item[(ii)] or $\widetilde A\not=\widetilde B\Rightarrow\left(\widetilde A-(\widetilde A\cap \widetilde B)\neq\phi \right.$ or $\left.\widetilde B -(\widetilde A\cap \widetilde B)\neq\phi\right)$ which is absurd since it contradicts with $\widehat A\cap E(\widetilde A\cap\widetilde B)\not= \phi$ and $\widehat B\cap E(\widetilde A\cap\widetilde B)\not=\phi$~[P.\rf{prop2.5}. \end{enumerate} \begin{corollary}\label{cor3.5} Let $P$ be a partition of ``strong'' partitive sets of $G=(V,E)$ and $f$ a choice mapping from $P$ to $V$, i.e., $f:X\in P\to f(X)\in X\subseteq V$. We have then that the isomorphism from $G/P$ to $G(f(P))$ conserves the color classes. \end{corollary} \begin{corollary}\label{cor3.6} Let $X$ be a ``strong'' partitive set of $G=(V,E)$ with $X\not= V$ and let $u\in V-X$. We have then: \begin{enumerate} \item[(i)] either for every $x\in X, ux\not\in E$; \item[(ii)] or there exists a color $\widehat A$ of $G$ so that for every $x\in X, ux\in \widehat A$. \end{enumerate} \end{corollary} {\it Proof.} This is true because for every $u\in V$ the singleton $\{u\}$ is a ``strong'' partitive set of $G$. \begin{theorem}\label{th3.7} Let $X$ be a ``strong'' partitive set of $G=(V,E)$ and $M(S)$ a multiplex of $G$ generated by $S=(V_S,E_S)$. We have then the following implication: $M\cap E(X)\not=\phi\Rightarrow M\subseteq E(X)$. \end{theorem} {\it Proof.} If $E_S\subseteq E(X)$ then for every color $\widehat A\subseteq M$ we have $\widehat A\cap E(X)\not =\phi$ and thus for every $\widehat A\subseteq M$, $\widehat A\subseteq E(X)$ which implies that $M\subseteq E(X)$. In the other hand if $E_S\cap E(X)=\phi$ then $M\cap E(X)=\phi$. Let us assume now that $E_S-E(X)\not=\phi$ and $E_S\cap E(X)\not=\phi$, then $V_S-X\not=\phi$. Let $u\in V_S-X$ and $ab\in E_S\cap E(X)$, then $\{a,b\}\subseteq X\cap V_S$ and $u$ is related to $a$ and $b$ by two different colors (applying the definition of a simplex). This is absurd because it contradicts with the corollary~[\rf{cor3.6}. This result is the analogue for multiplices and "strong" partitive sets of [P.2.5] which deals with colors and partitive sets. \begin{lemma}\label{lem3.8} Let $G=(V,E)$ be an undirected graph with the number of vertices greater than 2 $(|V|>2)$. If $G$ is decomposable and has a color $\widehat A$ so that $\widetilde A= V$, then $G$ has a non-trivial maximal ``strong'' partitive set. \end{lemma} {\it Proof.} Let $X$ be a non-trivial partitive set of $G$ $(X\in I^{\star}(G))$, thus $X\not=\widetilde A$. Let us define on $I(G)$ the following binary relation $R$ : \newline $XRY\Leftrightarrow ((X=Y)$ or $(X\cap Y\not=\phi, X-Y\not=\phi$ and $Y-X\not=\phi))$. And let $R^{\star}$ be the transitive closure of $R$. It is easy to verify that $R^{\star}$ is an equivalence relation on $I(G)$. Let $X^{\star}$ be the equivalence class of $X$ modulo $R^{\star}$. By the definition itself of $R^{\star}$, $\cup X^{\star}$ is a ``strong'' partitive set of $G$. In the other hand $X\not=\widetilde A\Rightarrow$ for every $Y\in X^{\star}$, $E(Y)\cap\widehat A=\phi$ (otherwise $\widetilde A=Y$ and for every $Z\in X^{\star}$, $Z\subset Y$ which contradicts with the definition of $R^{\star}$). We have then $\cup X^{\star}\subset V$. There exists then $a\in V-\cup X^{\star}$ so that for every $x\in \cup X^{\star}$, $ax\in\widehat A$[P.3.2]. But $\{ a\}\in I^{\star}_F(G)$ and if $Y\in I^{\star}_F(G)$ so that $\{ a\}\subset Y$ then $Y\subseteq V-\cup X^{\star}$ ( otherwise $\widehat A\cap E(Y)\neq\phi\Rightarrow\widetilde A =V\subseteq Y $[P.2.5]), consequently $a$ is contained in a maximal "strong" partitive set different from $V$. Thus $\cup X^{\star}$ is contained in a non trivial maximal "strong" partitive set. \begin{theorem}\label{th3.9} Let $M (S)$ be a multiplex of $G = (V,E)$ generated by the simplex $S = (V_S,E_S)$. $M (S)$ is maximal if and only if the set of vertices $\widetilde M$ spanned by $M$ is a ``strong'' partitive set. \end{theorem} {\it Proof.} Let us suppose that $M (S)$ is maximal, then $S$ is maximal. $\widetilde M$ is a partitive set of $G$. Let $Y$ be a partitive set of $G$ so that $Y \cap \widetilde M \not= \phi$ and $Y - \widetilde M \not= \phi$. Assume that $\widetilde M - Y \not= \phi$ and let us show that it is absurd. Let $y \in Y - \widetilde M$. $G (\widetilde M)$ is connected and then there exists $u \in Y \cap \widetilde M$ and $v \in \widetilde M - Y$ so that $u v \in M$. The following statements hold: \arraycolsep2.5pt $$ \begin{array}{ll} ( Y \in I (G),\ u \in Y\ \mbox{and}\ v \in \widetilde M - Y ) & \Rightarrow\ \mbox{For any} x\in Y, xv\in E\Rightarrow\ y v \in E\ ; \\[3mm] ( {\widetilde M} \in I (G),\ v \in \widetilde M\ \mbox{and}\ y \in Y - \widetilde M ) & \Rightarrow\ \mbox{for any}\ x \in \widetilde M,\ x y \in E \end{array} $$(see Figure 3.3). \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 3.3} {\em For every $x\in {\widetilde M},$ $xy\in E$}. \end{center} \medskip The colors connecting $y$ to the summits of $S$ can not be all different, otherwise, $M$ will not be maximal. Let us suppose that there exists $\{ a,b \} \in S$ and a color $\widehat A$ so that $\{ ya, yb \} \subseteq \widehat A$. Since $y \in \widetilde A - \widetilde M$ and $\widetilde M \in I (G)$, then we have $\widehat A \cap M = \phi$[P.2.5]. If rank$(M)=1$ then there exists a color $\widehat K$ so that $\widehat K = M$. But $(ab\in \widehat K \cap E (\widetilde A) \not= \phi ) \Rightarrow \widetilde K \subseteq \widetilde A$ : $\{ uv,ab\}\subseteq\widehat K \Rightarrow \{ yu,yv\}\in\widehat A$[Triangle Lemma], then we have $(y u \in \widehat A \cap E (Y)) \Rightarrow \widetilde M = \widetilde K \subseteq \widetilde A \subseteq Y$[P.2.5] ( see Figure 3.4) which is in contradiction with our proposition. \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 3.4} {\em $yu\in\widehat A\cap E(Y)\neq\phi$ and $ab\in\widehat K\cap E(\widetilde A)\neq \phi$}. \end{center} \medskip Let us assume now that rank$(M) \ge 2$. Let $c \in V_S - \{ a,b \},\ \widehat B$ the color of $y c,\ \widehat C$ the color of $a c$ and $\widehat D$ the color of $b c$. If $\widehat A \not= \widehat B$ then $S$ contains two tricolor triangles: $(y,a,c~;\ a y \in \widehat A,\ c y \in \widehat B,\ a c \in \widehat C)$ and $(y,b,c~;\ b y \in \widehat A,\ c y \in \widehat B,\ b c \in \widehat D)$ having two common colors, then $\widehat C = \widehat D$~[Triangle lemma] which is absurd because $S$ is a simplex, then $\widehat A = \widehat B$ (see Figure 3.5). \begin{center} \vglue 0,2cm \vglue 0,4cm {\bf Figure 3.5} {\em $\widehat B\neq\widehat A\Rightarrow \widehat C=\widehat D$}. \end{center} \medskip Thus for every $x \in V_S,\ y x \in \widehat A$ and by consequence $V_S \subset \widetilde A$ which implies that for every color $\widehat H \subset M$ we have $ \widehat H \cap E (\widetilde A) \not= \phi$ then $\widetilde M \subset \widetilde A$. But $(y u \in \widehat A \cap E(Y) \not= \phi) \Rightarrow \widetilde A \subseteq Y$ hence $\widetilde M \subset \widetilde A \subseteq Y$, which also contradicts the first assumption $(\widetilde M - Y \not= \phi)$. Finally we have showed that $\widetilde M \subset Y$ which means that $\widetilde M$ is a ``strong'' partitive set. Let us now prove the converse. Let us assume that $\widetilde M$ is a ``strong'' partitive set of $G$ and let us show that $M$ is a maximal. If the summits of $\widetilde M$ are related to $y \in V - \widetilde M$, then they are related by the same color which achieves the proof. \begin{corollary}\label{cor3.10} Let $M (S)$ be a multiplex of $G = (V,E)$ generated by the simplex $S=(V_S,E_S)$ and let $a \in V - \widetilde M$. The simplex $S$ is extensible to a simplex $S'=(V_{S'},E_{S'})$ (with $S$ a sub-graph of $S'$) so that $V_{S'} = V_S \cup \{ a \}$ if and only if there exists $\{ b,c \} \subseteq \widetilde M$ and two colors $\widehat A$ and $\widehat B$ of $E - M$ so that $a b \in \widehat A$ and $a c \in \widehat B$. \end{corollary} {\it Proof.} Use [C.3.6] and [T.3.9]. \section{Transitive orientations of an undirected graph} \noindent In this section, using the results of the previous section we prove the existence of a partition of maximal multiplices for the set of edges of an undirected graph. Therefore, the transitive orientations of a comparability graph turn up to the transitive orientations of their multiplices. These orientations are independent to each other. A theorem of decomposability for a non limit undirected graph is proved. \begin{lemma}\label{lem4.1} Let $G=(V,E)$ be any graph. Let $X$ and $Y$ be two partitive sets of $G$ so that $X \subseteq Y$. The following statements hold: \begin{itemize} \item[(i)] $X \in I_F (G) \Rightarrow X \in I_F (G (Y))\ ;$ \item[(ii)] $Y \in I_F (G) \Rightarrow (X \in I_F (G) \Leftrightarrow X \in I_F (G (Y)).$ \end{itemize} \end{lemma} {\it Proof.} Let $X$ be a ``strong'' partitive set of $G$ $(X \in I_F (G))$. We have for every $Z \in I (G (Y)),\ Z \in I (G)$. Hence $X \in I_F (G (Y))$. Let $X \in I_F (G (Y)),\ Y \in I_F (G)$ and $Z \in I (G)$ so that $Z \cap X \not= \phi$ and $Z - X \not= \phi$. Then $Z \cap X \not= \phi \Rightarrow Z \cap Y \not= \phi$. But $Y \in I_F (G)$, thus either $Z \subseteq Y$ or $Y \subseteq Z$. But $(Y \subseteq Z \Rightarrow X \subseteq Z)$ and $(X \subseteq Y \Rightarrow Z \in I (G (Y)) \Rightarrow X \subseteq Z)$. \begin{lemma}\label{lem4.2} Let $G=(V,E)$ be any graph. Let $F$ and $F'$ be two partitions of $G$ constituted of maximal ``strong'' partitive sets. We have then $F = F'$. \end{lemma} {\it Proof.} Let us assume that $F \not= F'$. Since $F$ and $F'$ are patitions of $V$, if $X \in F$ then there exists $X' \in F'$ so that $X \cap X' \not= \phi$. Thus either $X \subseteq X'$ or $X' \subseteq X$. But $X$ and $X'$ are maximal ``strong'' partitive sets, thus $X = X'$. Hence $F = F'$. \begin{proposition} \label{prop4.3} Let $M (S)$ a multiplex of $G = (V,E)$ generated by the simplex $S$ with rank$(M) \ge 2$. We have then for every $x \in \widetilde M$ there exists two colors $\{ \widehat A,\widehat B \} \subseteq \widehat M$ with $\widetilde A - \widetilde B \not= \phi$ and $\widetilde B - \widetilde A \not= \phi$ so that $x \in \widetilde A \cap \widetilde B$. \end{proposition} {\it Proof.} Let $x \in M$, then there exists $\widehat A \subset M$ and $y \in \widetilde A$ so that $x y \in \widehat A$. Moreover ($\widehat A \subset M$ and rank $(M) \ge 2$) $\Rightarrow S$ contains one tricolor triangle: $(a, b, c\ ;\ \widehat A, \widehat B, \widehat C)$ so that $b c \in \widehat A,\ a c\in \widehat B$ and $a b \in \widehat C$. Thus $a \not\in \widetilde A,\ b \not\in \widetilde B$ and $c \not\in \widetilde C$~[Triangle lemma]. Hence $\widetilde A - \widetilde B \not= \phi$ and $\widetilde B - \widetilde A \not= \phi$. In the other hand: $x y \in \widehat A \Rightarrow ((a x \in \widehat B\ \hbox{and}\ a y \in \widehat C)\ \hbox{or}\ (a x \in \widehat C\ \hbox{and}\ a y \in \widehat B))$~[Triangle lemma]. If one suppose $a x \in \widehat B$ then $x \in \widetilde A \cap \widetilde B$. \begin{theorem}\label{th4.4} Let $M (S)$ be a multiplex of $G = (V,E)$ generated by the simplex $S=(V_S,E_S)$. We have then the following statements: \begin{itemize} \item[(i)] $G (\widetilde M)$ has a partition of maximal ``strong'' partitive sets $F_M \not= \{ \widetilde M \}$~; \item[(ii)] If rank$(M) = 1$, we have either $G (\widetilde M) / F_M$ is isomorphic to $S$ or $G (\widetilde M) / F_M$ is indecomposable and isomorphic to a sub-graph $G'=(V',E')$ of $G (\widetilde M)$ so that $E' \subseteq M$~; \item[(iii)] If rank $(M) \ge 2$, then $G (\widetilde M) / F_M$ is isomorphic to $S$. \end{itemize} \end{theorem} {\it Proof.} 1) If rank$(M) = 1$ then $M$ contains only one color $\widehat A = M$. Moreover if $G (\widetilde M)$ is indecomposable then $F_M = \{ \{ x \}\ ;\ x \in \widetilde M \}$. Hence $G (\widetilde M) / F_M$ is isomorphic to $G (\widetilde M)$ and thus $G (\widetilde M) / F_M$ is indecomposable. Let us assume now that $G (\widetilde M)$ is decomposable. Thus $G (\widetilde M)$ contains a non-trivial partitive set $X$. After~[L.\rf{lem4.2} $F_M$ exists. If $| F_M | = 2$ then $G (\widetilde M) / F_M$ is complete and isomorphic to $S$. Let us suppose that $| F_M | \ge 3$. If $G (\widetilde M) / F_M$ has a non-trivial partitive set $X$, $G (\widetilde M) / F_M$ has a non-trivial maximal ``strong'' partitive set $Y$~[L.\rf{lem4.2} and [C.3.5]. Hence $\cup Y \in I_F^{\star} (G (M))$[P.21] which is absurd because $F_M$ is already maximal. Thus $G (\widetilde M) / F_M$ is indecomposable. \vskip 0.8cm 2) Let us now assume that rank$(M) \ge 2$. Using the proposition~[\rf{prop4.3} one gets that for every $x \in \widetilde M$ there exists two colors $\{ \widehat A, \widehat B \} \subseteq \widehat M$ with $\widetilde A - \widetilde B \not= \phi$ and $\widetilde B - \widetilde A \not= \phi$ so that $x \in \widetilde A \cap \widetilde B$. Applying~[T.\rf{th3.3} we have that the intersection $\widetilde A \cap \widetilde B \not= \widetilde M$ is a ``strong'' partitive set of $G$. Thus $\widetilde A \cap \widetilde B$ is a ``strong'' partitive set of $G (\widetilde M)$ itself since $\widetilde M \in I (G)$~[L.\rf{lem4.1}. Let $F_M$ be the set of the intersections two by two of colors of $M$. Then for every $a \in V_S$ there exists $X \in F_M$ so that $a \in X$. Let $\{ a, b \} \subset V_S$ so that $a \in X \in F_M$ and $b \in Y \in F_M$. Let us assume that there exists a ``strong'' partitive set $Z$ of $G (\widetilde M)$ so that $X \cup Y \subseteq Z$. $S$ will contain a tricolor triangle $(a, b, c\ ;\ \widehat A, \widehat B, \widehat C)$ with $b c \in \widehat A$, $a c \in \widehat B$ and $a b \in \widehat C$. This is absurd since $\{ c \}$ and $Z$ are two ``strong'' partitive sets of $G (\widetilde M)$ and can be related at most by only one color[T.3.4]. Hence $F_M$ is a maximal partition of "strong" partitive sets and separates the summits of $S$. Finally we get that $G (\widetilde M) / F_M$ is isomorphic to $S$. \begin{corollary}\label{cor4.5} The only multiplices $M (S)$ which might be not transitively orientable are those of rank $= 1$ and so that $G (\widetilde M) / F_M$ is non isomorphic to $S$. \end{corollary} {\it Proof.} Because complete graphs are orientable. \begin{corollary}\label{cor4.6} An undirected graph $G = (V, E)$ can have at most one multiplex which spanned all its summits. \end{corollary} {\it Proof.} Let $M (S)$ and $M (S')$ be two multiplices of $G$ so that $\widetilde M = V$ and $\widetilde M' = V$, we have then $G (\widetilde M) = G (\widetilde M') = G$. Let $F$ and $F'$ be two partitions of maximal ``strong'' partitive sets of $G$ related to $M$ and $M'$. Thus $G / F$ is isomorphic to $S$ and $G / F'$ is isomorphic to $S'$. Moreover the two isomorphismes conserve the colors~[C.\rf{cor3.5}. But after~[L.\rf{lem4.2} we have $F = F'$. Hence $S = S'$ and $M = M'$. \begin{corollary}\label{cor4.7} Let $G = (V,E)$ be an undirected graph and $M (S),\ M' (S')$ two maximal multiplices of $G$. We have then $M \cap M' \not= \phi \Rightarrow M = M'$. \end{corollary} {\it Proof.} Since $M$ and $M'$ are ``strong'' partitive sets we have \\ $M \cap M' \not= \phi \Rightarrow \widetilde M\cap \widetilde {M'}\neq \phi \Rightarrow ( \widetilde M \subseteq \widetilde M'\ \hbox{or}\ \widetilde M' \subseteq \widetilde M )$.\newline Let us assume that $\widetilde M \subseteq \widetilde M'$ and let $F, F'$ be two partitions of maximal ``strong'' partitive sets related respectively to $G (\widetilde M)$ and $G (\widetilde M')$. If there exists $X \in F'$ so that $M \cap E (X) \not= \phi$ then $M \subseteq E (X)$~[T.\rf{th3.7} and $M \cap M' = \phi$ ( since $G (M') / F'$ is isomorphic to $S'$ and $S$ would be a sub-graph of $G (X)$. Hence $E_S\cap E_{S'}= \phi$).\newline Finaly we have that for every $X \in F',\ E (X) \cap M = \phi$. Thus $S$ is isomorphic to a sub-graph of $S'$. But $S$ is maximal. Hence $S = S'$ and $M = M'$. \begin{corollary}\label{cor4.8} Let $G = (V,E)$ be an undirected graph with $E \not= \phi$. $E$ has then a partition of maximal multiplices. \end{corollary} The theorem 4.4 tell us that a multiplex has the same number of transitive orientations as the simplex which generated this multiplex. The simplex itself has a number of transitive orientations equal to the number of the possible permutations of its summits. Moreover [C.\rf{cor4.5} asserts that the only multiplices $M(S)$ which might be not transitively orientable are those with rank 1 so that $G(\widetilde M)/F_M$ is not isomorphic to $S$. Thus using [C.\rf{cor4.8} the problem of transitive orientation for a comparability graph come down to the transitive orientation of its multiplices. But the following problem is rised: if we orientate in any way and at certain step a given multiplex, will this orientation influence or not the orientations of the other multiplices at the following steps ? The response is not and we will prove this statement using a theorem [T.\rf{th4.11} which is known and for which we propose a new proof outcoming {}from the forcing theory.\newline Before announcing [T.\rf{th4.11}, we will announce a theorem [T.\rf{th4.9} which, in fact, is a mathematical algorithm permitting to find all the transitive orientations for a comparability graph which has only non limit sub-graphs, {\em e.g.}, case of finite graphs. \begin{lemma}\label{lem4.x} Let $G=(V,E)$ be a connected undirected graph. Then $G$ can not contain a multiplex $M$ so that both $\widetilde M\neq V$ and $\widetilde M$ is maximal for the inclusion among the $\widetilde N$, where $N$ is any multiplex of $G$ \end{lemma} {\it Proof.} Let us assume that such a multiplex $M$ exists and show that it is absurd. Since $\widetilde M$ is maximal for the inclusion it implies that $M$ is maximal. After[T.\rf{th3.9}, $\widetilde M$ is a "strong" partitive set. Since $\widetilde M\neq V$ and $G$ is connected, we have :\newline there exists $x\in V-\widetilde M$ and $y\in \widetilde M$ so that $xy\in E$. Let $\widehat A$ the color containing $xy$. Then $\widehat A\not\subset M$ and $x\in (\widetilde A \cap \widetilde M )\neq\phi$. But $\widetilde M$ is a "strong" partitive set of $G$ and $\widetilde A$ is a partitive set of $G$, thus $\widetilde M\subset \widetilde A$, which is absurd because $\widehat A$ is a multiplex of rank 1. \begin{theorem}\label{th4.9} Let $G = (V,E)$ be an undirected graph having a partition of maximal ``strong'' partitive sets $F_G \not= \{ V \}$. We have then that $G / F_G$ satisfies one of the following exclusive assertions: \begin{itemize} \item[(i)] $G / F_G$ is empty. \item[(ii)] $G / F_G$ is indecomposable and there exists a maximal multiplex $M_G$ of $G$ with rank$(M_G) = 1,\ \widetilde M_G = V$ and $G / F_G$ is isomorphic to a sub-graph $G'=(V',E')$ of $G$ so that $E' \subseteq M_G$. \item[(iii)] $G / F_G$ is complete and isomorphic to a maximal simplex $S$ generating a maximal multiplex $M_G$ so that $\widetilde M_G = V$. \end{itemize} \end{theorem} {\it Proof.} If $G$ is non connected, $F_G$ is the class of the connected components and $G / F_G$ is empty. In the other hand, it is obvious that if $G / F_G $ is empty then $G$ is non connected. Hence $G / F_G$ is empty if and only if $G$ is non connected. Let us assume now that $G$ is connected. Since $G$ is non limit, it implies that $G$ has maximal multiplex $M$ so that $\widetilde M$ is maximal for the inclusion. Thus using the [L.\rf{lem4.x} we have $\widetilde M =V$. Hence $G=G(\widetilde M)$ and using [T.\rf{th4.4} we get the result. \begin{corollary}\label{cor4.10} Let $G = (V,E)$ be a connected and undirected graph. $G$ has then a partition of maximal ``strong'' partitive sets $F_G \not= \{ V \}$ if and only if $G$ has a multiplex $M_G$ so that $\widetilde M_G = V$. \end{corollary} {\it Proof.} Applying~[T.\rf{th4.9} we have: $F_G$ exists $\Rightarrow M_G$ exists. In the other hand applying~[T.\rf{th4.4}, we have the other implication: $M_G$ exists $\Rightarrow F_G$ exists. \begin{theorem}\label{th4.11} Let $O = (V,E)$ be a partial order and $G = (V, E)$ be its comparability graph. $O$ and $G$ have then the same ``strong'' partitive sets $(I_F (G) = I_F (O))$. \end{theorem} Before giving the proof of this theorem, we present some preleminary results which will be used for the proof. \begin{lemma}\label{lem4.12} Let $O = (V,E')$ be a partial order and $G = (V, E)$ its comparability graph. Then every partitive set of $O$ is a partitive set of $G$\\ $ (I (O) \subseteq I (G))$. \end{lemma} {\it Proof.} Let $Y \in I (O)$. If $Y$ is a trivial partitive set, we have $Y \in I (G)$. Let us suppose that $Y$ is non trivial, then if $a \in V - Y$ we have one of the following statements: \begin{itemize} \item for every $y \in Y,\ ((a,y) \in E')\Rightarrow a y \in E$. \item or for every $y \in Y,\ ((y,a) \in E')\Rightarrow a y \in E$. \item or for every $y \in Y,\ (\{ (a,y), (y,a) \} \cap E' = \phi) \Rightarrow a y \not\in E$. \end{itemize} Therefore $Y \in I (G)$. \begin{proposition}\label{prop4.13} Let $O = (V,E')$ be a partial order and $G = (V, E)$ its comparability graph. Then every ``strong'' partitive set of $G$ is a ``strong'' partitive set of $O$ $ (I_F (G) \subseteq I_F (O))$. \end{proposition} {\it Proof.} Let $X$ be a ``strong'' partitive set, {\it i.e.,} $X \in I_F (G)$. If $X$ is trivial then $X \in I_F (O)$. Let us suppose that $X$ is non trivial. We have then one of the following statements: \begin{itemize} \item for every $x \in X,\ x a \not\in E$, then for every $\{ x, y \} \subseteq X,\ (x, a) \cong (y,a)$ (cf.~\S2.1). \item or there exists a color $\widehat A$ of $G$ so that for every $x \in X,\ x a \in \widehat A$, then for every $\{ x, y \} \subseteq X,\ (x, a) \cong (y, a)$ (since $G$ is a comparability graph [T.\rf{th2.2}). \\ Therefore $X \in I (O)$. Let $Y \in I (O)$ so that $Y \cap X \not= \phi$ and $Y - X \not= \phi$. Then using the previous lemma~[\rf{lem4.12} we have $Y \in I (G)$ and then $X \subset Y$. Thus $X \in I_F (O)$. \end{itemize} \begin{proposition}\label{prop4.14} If $X$ is a maximal ``strong'' partitive set of $G$ then $X$ is a maximal ``strong'' partitive set of $O$. \end{proposition} {\it Proof.} If $X \in I_F (G)$ is maximal, it implies that $G$ is non limit. Thus $G$ has a partition ${ F}$ of maximal ``strong'' partitive sets and $X \in { F} \subseteq I_F (O)$. After~[T.\rf{th4.9} we have either $G / { F}$ is empty and then $X$ is a connecting class. Thus $X$ is maximal in $I_F (O)$ ; or $G / { F}$ is complete and then $O / { F}$ is a chain, therefore $G / { F}$ do not has a non trivial ``strong'' partitive sets, thus $X$ is maximal in $I_F (O)$ ; or $G / { F}$ is indecomposable and the partitive sets of $G / { F}$ are trivial, therefore, after~[L.\rf{lem4.12} the partitive sets of $O / { F}$ are trivial, thus $X$ is maximal in $I_F (O)$. \begin{proposition}\label{prop4.15} Let $X$ be a non trivial ``strong'' partitive set of $O$. There exists then a non trivial ``strong'' partitive set $Y$ of $G$ so that $X \subseteq Y$. \end{proposition} {\it Proof.} If $G$ is non limit then it has a partition ${ F}$ of maximal ``strong'' partitive sets which is also a partition of maximal ``strong'' partitive sets of $O$[P.4.15]. Thus there exists $Y \in { F}$ so that $X \subseteq Y$. If $G$ is limit then since $X \in I (O) \Rightarrow X \in I (G)$~[L.\rf{lem4.12}, therefore $Y$ exists. \vskip 0.8cm \noindent In what follows, we present the proof of the theorem~[\rf{th4.11}. {\it Proof.} If $X \in I_F (G)$ then after~[P.\rf{prop4.14} $X \in I_F (O)$. Let $X \in I_F^{\star} (O)$. Then after~[P.\rf{prop4.15} there exists $Y \in I_F^{\star} (G)$ so that $X \subseteq Y$. Let us suppose that $X \not\in I_F (G)$. Thus $X \subset Y$.\newline $Y \in I_F^{\star} (G) \Rightarrow Y \in I_F^{\star} (O)$. Therefore after~[L.\rf{lem4.1} $X \in I_F^{\star} (O (Y))$. Thus there exists $Y_1 \in I_F^{\star} (G (Y))$ so that $X \subset Y_1$. So we have constructed a strictly decreasing suite of elements of $I_F^{\star}(G) \subseteq I_F^{\star} (O)$ (after~[P.\rf{prop4.13}).\newline The intersection $\cap Y_i$ of this suite is a ``strong'' partitive set of $O$ and it is the smallest element of $I_F^{\star} (O)$ which contains $X$. Therefore $X = \cap Y_i$. But $\cap Y_i$ is also a ``strong'' partitive set of $G$. Thus $X \in I_F^{\star} (G)$ and then $I_F (O) = I_F (G)$. \section*{Acknowledgments} One of the authors, M.~H., would like to thank Prof. P. ILLE for helpful and stimulating discussions and also Profs. C. Rauzy and G. Fardoux for their encouragements.
1994-11-22T06:20:13
9411
alg-geom/9411014
en
https://arxiv.org/abs/alg-geom/9411014
[ "alg-geom", "math.AG" ]
alg-geom/9411014
Peter Magyar
Peter Magyar
A Borel-Weil Theorem for Schur Modules
35pp, LaTeX
null
null
null
null
We present a generalization of the classical Schur modules of $GL(N)$ exhibiting the same interplay among algebra, geometry, and combinatorics. A generalized Young diagram $D$ is an arbitrary finite subset of $\NN \times \NN$. For each $D$, we define the Schur module $S_D$ of $GL(N)$. We introduce a projective variety $\FF_D$ and a line bundle $\LL_D$, and describe the Schur module in terms of sections of $\LL_D$. For diagrams with the ``northeast'' property, $$(i_1,j_1),\ (i_2, j_2) \in D \to (\min(i_1,i_2),\max(j_1,j_2)) \in D ,$$ which includes the skew diagrams, we resolve the singularities of $\FD$ and show analogs of Bott's and Kempf's vanishing theorems. Finally, we apply the Atiyah-Bott Fixed Point Theorem to establish a Weyl-type character formula of the form: $$ {\Char}_{S_D}(x) = \sum_t {x^{\wt(t)} \over \prod_{i,j} (1-x_i x_j^{-1})^{d_{ij}(t)}} \ ,$$ where $t$ runs over certain standard tableaux of $D$. Our results are valid over fields of arbitrary characteristic.
[ { "version": "v1", "created": "Mon, 21 Nov 1994 22:44:45 GMT" } ]
2015-06-30T00:00:00
[ [ "Magyar", "Peter", "" ] ]
alg-geom
\section{$GL(N)$ modules} \label{Schur modules and Weyl modules} \subsection{Schur modules} Given a finite set $T$, we will also use the symbol $T$ to denote the order $|T|$ when appropriate. Thus $GL(T) \stackrel{\rm def}{=} GL(|T|)$, etc. Let $\Sigma_T$ be the symmetric group permuting the elements of $T$. For any left $G$-space $X$, $\Sigma_T$ acts on the right, and $G$ acts on the left, of the cartesian product $X^T$ by: $$ g(x_{t_1}, x_{t_2}, \ldots ) \pi = (g x_{\pi t_1}, g x_{\pi t_2}, \ldots ) . $$ A {\em diagram} is a finite subset of ${\bf N} \times {\bf N}$. Its elements $(i,j) \in D$ are called {\em squares}. We shall often think of $D$ as a sequence $(C_1,C_2,\ldots,C_r)$ of columns $C_j \subset {\bf N}$. The Young diagram corresponding to $ \lambda = ( \lambda _1 \geq \lambda _2 \geq \cdots \geq \lambda _N \geq 0)$ is the set $\{(i,j) \mid 1\leq j\leq N,\ 1\leq i\leq \lambda _j \}$. For any diagram $D$, we let $$ \mathop{\rm Col}(D) = \{\pi \in \Sigma_D \mid \pi(i,j) = (i',j) \ \exists i'\} $$ be the group permuting the squares of $D$ within each column, and we define $\mathop{\rm Row}(D)$ similarly for rows. Let $F$ be a field. We shall always write $G = GL(N,F)$, $B = $ the subgroup of upper triangular matrices, $H =$ the subgroup of diagonal matrices, and $V = F^N$ the defining representation. Now let $F$ have characteristic zero. Define the idempotents ${\alpha}_D$, $\beta_D$ in the group algebra $F[\Sigma_D]$ by $$ {\alpha_D} = {1 \over |\mathop{\rm Row} D|} \sum_{\pi \in \mathop{\rm Row} D} \pi, \ \ \ \ {\beta_D} = {1 \over |\mathop{\rm Col} D|} \sum_{\pi \in \mathop{\rm Col} D} \mathop{\rm sgn}(\pi) \pi , $$ where $\mathop{\rm sgn}(\pi)$ is the sign of the permutation. Define the {\em Schur module} $$ S_D \stackrel{\rm def}{=} V^{\otimes D} {\alpha_D} {\beta_D} \subset V^{\otimes D}, $$ a representation of $G$. Note that we get an isomorphic Schur module if we change the diagram by permuting the rows or the columns (i.e., for some permutation $\pi: {\bf N} \rightarrow {\bf N}$, changing $D = \{(i,j)\}$ to $D' = \{(\pi(i), j) \mid (i,j) \in D\}$, and similarly for columns). \subsection{Weyl modules} Let $W = V^*$, the dual of the defining representation of $G=GL(N,F)$, where $F$ is an infinite field. Given a diagram $D$, define the alternating product {\em with respect to the columns} $$ {\bigwedge} ^D W = \{ f:V^D \rightarrow F \mid f \mbox{ multilinear, and } f(v \pi) = \mathop{\rm sgn}(\pi) f(v) \ \forall \pi \in \mathop{\rm Col}(D) \}, $$ where {\em multilinear} means $f(v_1,\ldots,v_d)$ is $F$-linear in each of the $d = |D|$ variables. Consider the multidiagonal {\em with respect to the rows} $$ \Delta^D V = \Delta^{R_1} V \times \Delta^{R_2} V \times \cdots \subset V^{R_1} \times V^{R_2} \times \cdots = V^D , $$ where $R_1, R_2, \ldots $ are the rows of $D$. Now define the {\em Weyl module} $$ W_D \stackrel{\rm def}{=} {\bigwedge} ^D W \mid_{ \Delta^D V}, $$ where $\mid_{ \Delta^D V}$ denotes restriction of functions from $V^D$ to $ \Delta^D V$. Since $ \Delta^D V$ is stable under the diagonal action of $G$, $W_D$ is naturally a $G$-module. \vspace{1em} \noindent {\bf Remark.} For $F$ a finite field, we make the following modification. Consider $W = W(F) \hookrightarrow W(\bar{F})$, where $\bar{F}$ is the algebraic closure. That is, identify $$ W = \{ f:\bar{F}^N \rightarrow \bar{F} \mid f \mbox{ is $\bar{F}$-linear, and } f(F^N) \subset F \}. $$ Then define $$ W_D \stackrel{\rm def}{=} {\bigwedge} ^D W \mid_{ \Delta^D V(\bar{F})}. $$ This keeps the restriction map from killing nonzero tensors which happen to vanish on the finite set $ \Delta^D V(F)$. With this definition, $W_D$ clearly has the base change property $W_D(L) = W_D(F) \otimes_F L$ for any extension of fields $F \subset L$. Now consider $W_D({\bf Z})$. This is a free ${\bf Z}$-module, since it is a submodule of the ${\bf Z}$-valued functions on $ \Delta^D V$. Suppose $D$ satisfies a direction condition. Then our vanishing results of Proposition \ref{config split} (a), along with the appropriate universal coefficient theorems, can be used to show that for any field $F$, $$ W_D(F) = W_D({\bf Z}) \otimes_{{\bf Z}} F . $$ \begin{prop} If $F $ has characteristic zero, then $W_D \cong S_D^*$ as $G$-modules. \end{prop} \noindent {\bf Proof.} $S_D$ is the image of the composite mapping $$ V^{\otimes D} {\alpha_D} \hookrightarrow V^{\otimes D} \stackrel{{\beta_D}}{\rightarrow} V^{\otimes D} {\beta_D} . $$ For $W = V^*$, write $$ W^{\otimes D} = \{f: V^D \rightarrow F \mid f \mbox{ multilinear}\}, $$ $$ \mathop{\rm Sym}\mbox{} ^D W = \{f: V^D \rightarrow F \mid f \mbox{ multilinear, and } f(v\pi) = f(v) \ \forall \pi \in \mathop{\rm Row}(D)\}. $$ Now, representations of $F[\Sigma_D]$ are completely reducible, so $S_D^*$ is the image of $$ W^{\otimes D} {\beta_D} \hookrightarrow W^{\otimes D} \stackrel{{\alpha_D}}{\rightarrow} W^{\otimes D} {\alpha_D} , $$ and $ W^{\otimes D} {\beta_D} \cong \bigwedge^D W$,\ $W^{\otimes D} {\alpha_D} \cong \mathop{\rm Sym}\mbox{}^D W$. Now, let $$ {\mathop{\rm Poly}\mbox{}}^D W = \{f: V^l \rightarrow F \mid f \mbox{ homog poly of multidegree } (R_1, \ldots, R_l) \}, $$ where $l$ is the number of rows of $D$. Then we have a $G$-equivariant map $$ \mathop{\rm rest}\mbox{}_{\Delta} : \mathop{\rm Sym}\mbox{} ^D W \rightarrow \mathop{\rm Poly}\mbox{} ^D W $$ restricting functions from $V^D$ to the row-multidiagonal $ \Delta^D V \cong V^l$. It is well known that $\mathop{\rm rest}\mbox{}_{\Delta}$ is an isomorphism: the symmetric part of a tensor algebra is isomorphic to a polynomial algebra. Thus we have the commutative diagram \begin{eqnarray*} {\bigwedge} ^D W & \hookrightarrow W^{\otimes D} \stackrel{{\alpha_D}}{\rightarrow} & \mathop{\rm Sym}\mbox{} ^D W \\ || & || & \downarrow \mathop{\rm rest}\mbox{}_{\Delta} \\ {\bigwedge} ^D W & \hookrightarrow W^{\otimes D} \stackrel{{\alpha_D}}{\rightarrow} & \mathop{\rm Poly}\mbox{} ^D W . \end{eqnarray*} Now, the image in the top row is $S_D^*$, the image in the bottom row is $W_D$, and all the vertical maps are isomorphisms, so we have $\mathop{\rm rest}\mbox{}_{\Delta} : S_D^* \tilde{\rightarrow} W_D$ an isomorphism. $\bullet$ If $D = \lambda $ a Young diagram, then $W_D$ is isomorphic to Carter and Lusztig's dual Weyl module for $G = GL(N,F)$. This will follow from Proposition \ref{Young diagram} in the following section. \section{Configuration varieties} \label{configuration varieties} {\bf N.B.} Although our constructions remain valid over ${\bf Z}$, for simplicity we will assume for the remainder of this paper that {\em $F$ is an algebraically closed field}. \subsection{Definitions and examples} \label{definitions and examples} Given a finite set C (a column), and $V=F^N$, consider $ V^C \cong M_{N\times C}(F)$, the $N \times |C|$ matrices, with a right multiplication of $GL(C)$. Let $$ \mathop{\rm St}(C) = \{X \in V^C \mid \mathop{\rm rank} X = |C| \}, $$ the Stiefel manifold, and $$ \mathop{\rm Gr}(C) = \mathop{\rm St}(C)/GL(C), $$ the Grassmannian. Also, let $$ {\cal L}_C = \mathop{\rm St}(C) \stackrel{GL(C)}{\times} {\det} ^{-1} \rightarrow \mathop{\rm Gr}(C) $$ be the Plucker determinant bundle, whose sections are regular functions $f: \mathop{\rm St}(C) \rightarrow F$ with $f(XA) = \det(A) f(X) \ \forall A \in GL(C)$. In fact, such global sections can be extended to polynomial functions $f: V^C \rightarrow F$. For a diagram $D$ with columns $C_1, C_2,\ldots$, we let $$ \mathop{\rm St}(D) = \mathop{\rm St}(C_1) \times \mathop{\rm St}(C_2) \times \cdots, \mbox{ } \mathop{\rm Gr}(D) = \mathop{\rm Gr}(C_1) \times \mathop{\rm Gr}(C_2) \times \cdots, \mbox{ } \LL_D = {\cal L}_{C_1} \Box\hspace{-0.76em}\times {\cal L}_{C_2} \Box\hspace{-0.76em}\times \cdots. $$ Recall that $ \Delta^D V \subset V^D$ is the {\em row} multidiagonal (as opposed to the column constructions above). Let $$ \FF_D ^o \stackrel{\rm def}{=} \mathop{\rm Im}\left[ \Delta^D V \cap \mathop{\rm St}(D) \rightarrow \mathop{\rm Gr}(D)\right], $$ and define the {\em configuration variety of $D$} by $$ \FF_D = \overline{ \FF_D ^o} \subset \mathop{\rm Gr}(D), $$ the Zariski closure of $ \FF_D ^o$ in $\mathop{\rm Gr}(D)$. We denote the restriction of $ \LL_D $ from $\mathop{\rm Gr}(D)$ to $ \FF_D $ by the same symbol $ \LL_D $. Some properties follow immediately from the definitions. For instance, $ \FF_D $ is an irreducible variety. Just as for Schur modules and Weyl modules, changing the diagram by permuting the rows or the columns gives an isomorphic configuration variety and line bundle. If we add a column $C$ to $D$ which already appears in $D$, we get an isomorphic configuration variety, but the line bundle is twisted to have higher degree. Since $ \LL_D $ gives the Plucker embedding on $\mathop{\rm Gr}(D)$, it is very ample on $ \FF_D $. \vspace{1em} \noindent {\bf Examples. } Set $N=4$. Consider the diagrams $$ D_1 = \begin{array}{ccc} \Box & & \\ \Box & \Box & \\ & \Box & \end{array} \ \ \ D_2 = \begin{array}{ccc} \Box & & \\ \Box & \Box & \Box \\ & & \Box \end{array} \ \ \ D_3 = \begin{array}{ccc} \Box & \Box & \\ & \Box & \\ & \Box & \Box \end{array} $$ Identifying $\mathop{\rm Gr}(k,F^N)$ with $\mathop{\rm Gr}(k-1,{\bf P}^{N-1}_F)$, we may consider the $ \FF_D $'s as varieties of configurations in ${\bf P}^3$:\\ (1) $ { \cal F } _{D_1}$ is the variety of pairs $(l,l')$, where $l,l'$ are intersecting lines in ${\bf P}^3$. It is singular at the locus where the two lines coincide.\\ (2) $ { \cal F } _{D_2}$ is the variety of triples $(l, p, l')$ of two lines and a point which lies on both of them. The variety is smooth: indeed, it is a fiber bundle over the partial flag variety of a line containing a point. There is an obvious map $ { \cal F } _{D_2} \rightarrow { \cal F } _{D_1}$, which is birational, and is in fact a small resolution of singularities. (C.f. Proposition \ref{smooth}.)\\ (3) $ { \cal F } _{D_3}$ is the variety of planes with two marked points (which may coincide). $ { \cal F } _{D_3}^o$ is the locus where the marked points are distinct. The variety is smooth as in the previous example. \\ $$ D_4 = \begin{array}{ccc} \Box & \Box & \\ \Box & & \Box \\ & \Box & \Box \end{array} \ \ \ D_5 = \begin{array}{ccc} \Box & \Box & \Box \\ \Box & & \\ & \Box & \\ & & \Box \end{array} D_6 = \begin{array}{cccc} \Box & & & \\ & \Box & & \\ & & \Box & \\ & & & \Box \end{array} \ \ \ $$ (4) $ { \cal F } _{D_4}$ is the variety of triples of coplanar lines.\\ (5) $ { \cal F } _{D_5}$ is the variety of triples of lines with a common point. This is the projective dual of the previous variety, since the diagrams are complementary within a $4 \times 3$ rectangle (up to permutation of rows and columns). (See Theorem \ref{complement thm}.) The variety of triples of lines which intersect pairwise cannot be described by a single diagram, but consists of $ { \cal F } _{D_4} \cup { \cal F } _{D_5}$. (See Section \ref{intersection varieties}.)\\ (6) $ { \cal F } _{D_6} \cong ({\bf P}^{3})^4$ contains the $GL(N)$-invariant subvariety where all four points in ${\bf P}^3$ are colinear. Since the cross-ratio is an invariant of four points on a line, this subvariety contains infinitely many $GL(N)$ orbits. $$ D_7 = \begin{array}{cccc} \Box & \Box & \Box & \\ & \Box & \Box & \Box \\ & & \Box & \\ & & & \Box \end{array} \ \ \ D_8 = \begin{array}{ccccc} \Box & \Box & \Box & & \\ & \Box & \Box & \Box & \Box \\ & & \Box & & \\ & & & \Box & \end{array} $$ (7) $ { \cal F } _{D_7} \cong G \stackrel{B}{\times} X_{ \lambda }$, the $G$-orbit version of the Schubert variety $X_{ \lambda } \subset \mathop{\rm Gr}(2,4)$ associated to the partition $ \lambda = (1,2)$. This is the smallest example of a singular Schubert variety. \\ (8) $ { \cal F } _{D_8}$ is a smooth variety which maps birationally to $ { \cal F } _{D_7}$ by forgetting the point associated to the last column. In fact, this is essentially the same resolution as (1) and (2) above. Such resolutions of singularities can be given for arbitrary Schubert varieties of $G = GL(N)$, and generalize Zelevinsky's resolutions in~\cite{Z}. C.f. Section \ref{resolution}. $\bullet$ \begin{thm} \label{weyl mod is image} If $F$ is an algebraically closed field, then $$ W_D \cong \mathop{\rm Im}\left[\mathop{\rm rest}\mbox{}_{\Delta}: H^0(\mathop{\rm Gr}(D), \LL_D ) \rightarrow H^0( \FF_D , \LL_D )\right], $$ where $\mathop{\rm rest}\mbox{}_{\Delta}$ is the restriction map. \end{thm} \noindent {\bf Proof.} Note that for $GL(D) = GL(C_1) \times GL(C_2) \times \cdots$, $$ H^0(\mathop{\rm Gr}(D), \LL_D ) = \{ f:V^D \rightarrow F \mid f(XA) = \det(A) \, f(X) \ \forall A \in GL(D)\}, $$ and recall $$ {\bigwedge} ^D W = \{ f:V^D \rightarrow F \mid f \mbox{ multilinear, and } f(v \pi) = \mathop{\rm sgn}(\pi) \ f(v) \ \forall \pi \in \mathop{\rm Col}(D) \}. $$ But in fact these sets are equal, because a multilinear, anti-symmetric function $g:V^C\rightarrow F$ always satisfies $g(XA) = \det(A) \ g(X) \, \forall A \in GL(C)$. Now $W_D$ and $H^0( \FF_D , \LL_D )$ are gotten by restricting functions in these identical sets to $ \Delta^D V$, so we are done. $\bullet$ \subsection{Diagrams with at most $N$ rows} We say $D$ has $\leq N$ rows if $(i,j) \in D \Rightarrow 1\leq i\leq N$. \begin{prop} \label{bigorbit} If $D$ has $\leq N$ rows, then $ \FF_D $ has an open dense $GL(N)$-orbit $ \FF^{\mbox{\rm \tiny gen}} _D$. \end{prop} \noindent {\bf Proof.} Let $D$ have columns $C_1, C_2,\ldots$. Consider a sequence of vectors $X =(v_1,\ldots,v_n) \in V^N$. For $C = \{i_1, i_2, \ldots \} \subset \{1,\ldots,N\}$, define $X(C) \stackrel{\rm def}{=} {\mathop{\rm Span}} _F(v_{i_1},v_{i_2},\ldots ) \in \mathop{\rm Gr}(C)$ (for $X$ sufficiently general). Consider an element $g \in GL(N)$ as a sequence of column vectors $g = (v_1,\ldots,v_n)$. Then $$ g(C) = g\cdot {\mathop{\rm Span}} _F(e_{i_1},e_{i_2},\ldots ) = g\cdot (I(C_1), I(C_2), \ldots) , $$ where $e_i$ denotes the $i$-th coordinate vector and $I$ the identity matrix. Now define the map $$ \begin{array}{rccc} \Psi: & V^N & \rightarrow & \Delta^D V \subset V^D \\ & (v_1,\ldots,v_N) & \mapsto & (v_i)_{(i,j)\in D}, \end{array} $$ where $(u_{ij})_{(i,j)\in D} $ denotes an element of $V^D$. Then the composite $$ V^N \stackrel{\Psi}{\rightarrow} \Delta^D V \rightarrow \FF_D ^o $$ is an onto map taking $g \mapsto (g(C_1), g(C_2),\ldots) = g\cdot (I(C_1), I(C_2), \ldots)$. Since $GL(N)$ is dense in $V^N$, its image is dense in $ \FF_D ^o$, and hence the composite image $ \FF^{\mbox{\rm \tiny gen}} _D \stackrel{\rm def}{=} G \cdot (I(C_1),I(C_2),\ldots)$ is a dense $G$-orbit in $ \FF_D $. $\bullet$ \begin{prop} \label{Young diagram} If $D$ is the Young diagram associated to a dominant weight $ \lambda $ of $GL(N)$, then: \\ (a) $ \FF_D \cong G/P$, a quotient of the flag variety $ { \cal F } = G/B$. \\ (b) The Borel-Weil line bundle $ {\cal L} _{ \lambda } \stackrel{\rm def}{=} G \stackrel{B}{\times} ( \lambda ^{-1}) \rightarrow { \cal F } $ is the pullback of $ \LL_D $ under the projection $ { \cal F } \rightarrow \FF_D $. \\ (c) $ \mathop{\rm rest}\mbox{}_{\Delta} : H^0(\mathop{\rm Gr}(D), \LL_D ) \rightarrow H^0( \FF_D , \LL_D )$ is surjective, and $W_D \cong H^0( \FF_D , \LL_D )$. \\ \end{prop} \noindent {\bf Proof.} (a) Let $\mu = (\mu_1 \geq \mu_2 \geq \cdots) = \lambda ^t $, the transposed diagram, and let $P = \{(x_{ij}) \in GL(N) \mid x_{ij} = 0 \mbox{ if } \exists k , \ i > \mu_k \geq j > \mu_{k+1} \}$, a parabolic subgroup of $G$. Then $G/P$ is the space of partial flags $V = F^N \supset V_1 \supset V_2 \supset \cdots $ consisting of subspaces $V_j$ with $dim(V_j) = \mu_j$. Clearly $G/P \cong \FF_D $. \\ (b) Let $\Psi: G \rightarrow \Delta^D V \cap \mathop{\rm St}(D) $ be the map in the proof of the previous proposition. Then the map \begin{eqnarray*} G \stackrel{P}{\times} F_{ \lambda ^{-1}} & \rightarrow & \LL_D = (\mathop{\rm St}(D) \stackrel{GL(D)}{\times} {\det} _D^{-1}) \mid_{ \FF_D } \\ (g, \alpha) & \mapsto & (\Psi(g), \alpha) \end{eqnarray*} is a $G$-equivariant bundle isomorphism. Then (b) follows by standard arguments. \\ (c) The surjectivity is a special case of Proposition \ref{config split} in Section \ref{cohomology}. (See also~\cite{In}.) The other statement then follows by Prop \ref{weyl mod is image}. $\bullet$ \subsection{Complementary diagrams} \label{complementary diagrams} \begin{thm} \label{complement thm} Suppose the rectangular diagram Rect $= \{1,\ldots,N\} \times \{1,\ldots,r\}$ is the disjoint union of two diagrams $D$, $D'$. Let $W_D$, $W_{D'}$ be the corresponding Weyl modules for $G =GL(N,F)$. Then: \\ (a) there is an $F$-linear bijection $\tau:W_D \rightarrow W_{D'} $ such that $\tau(g w) = \det_{N\times N}^r(g')\ g'\ \tau(w)$, where $g'$ is the inverse transpose in $GL(N)$ of the matrix $g$; \\ (b) the characters obey the relation $ {\mathop{\rm char}} \, {W_{D'}}(h) = \det_{N\times N}^r(h)\ {\mathop{\rm char}} \, {W_D} (h^{-1}) ,$ for diagonal matrices $h\in G$ ; \\ (c) if $F$ has characteristic zero, then as $G$-modules $$ W_{D'} \cong {\det}\mbox{}^{-r} \otimes W_{D}^* \ \ \ \mbox{ and }\ \ \ S_{D'} \cong {\det}\mbox{}^r \otimes S_{D}^* . $$ \end{thm} \noindent {\bf Proof.} (a) Given $C \subset \{1,\ldots,N\} $ (a column set), we considered above the Plucker line bundle $$ \mathop{\rm St}(C) \stackrel{GL(C)}{\times} \det\mbox{} ^{-1} \rightarrow \mathop{\rm Gr}(D) . $$ We may equally well write this as $$ GL(N) \stackrel{P_C}{\times} \det\mbox{} _C^{-1} \rightarrow \mathop{\rm Gr}(D) , $$ where $P_C \stackrel{\rm def}{=} \{ (x_{ij}) \in GL(N) \mid x_{ij} = 0 \mbox{ if } i \not\in C, \ j \in C\}$ is a maximal parabolic subgroup of $GL(N)$ (not necessarily containing $B$), and $\det_C : P_C \rightarrow F$ is the multiplicative character $\det_C(x_{ij})_{N\times N} \stackrel{\rm def}{=} \det_{C \times C}(x_{ij})_{i,j \in C}$. Hence, if $C_1, C_2, \ldots, C_r$ are the columns of $D$, we may write $$ Gr(D) \cong G ^r /P_D, $$ and the bundle $$ \LL_D \cong G ^r \stackrel{P_D}{\times} \det\mbox{} _D^{-1}, $$ where $P_D \stackrel{\rm def}{=} P_{C_1} \times \cdots \times P_{C_r}$ and $\det_D(X_1,\ldots,X_r) \stackrel{\rm def}{=} \det_{C_1}(X_1)\times \cdots \times \det_{C_r}(X_r)$. Under this identification, $$ \FF_D \cong \mbox{ closure} \mathop{\rm Im}\left[\, \Delta G \hookrightarrow G^r \rightarrow \mathop{\rm Gr}(D) \, \right] $$ (c.f. Proposition \ref{bigorbit}). Now let $\tau: G^r \rightarrow G^r$, $\tau(g_1,\ldots,g_r) = (g_1', \ldots,g_r')$, where $g' = \ ^t g^{-1}$, the inverse transpose of a matrix $g\in G$. Then $\tau(P_D) = P_{D'}$, and $\tau$ induces a map $$ \tau: \mathop{\rm Gr}(D) \rightarrow \mathop{\rm Gr}(D'), $$ as well as a map of line bundles $$ \begin{array}{rccc} \tau:& \LL_D & \rightarrow & {\cal L}_{D'} \\ & \| & & \| \\ & G^r \stackrel{P_D}{\times} \det\mbox{}_D^{-1} & & G^r \stackrel{P_{D'}}{\times} \det\mbox{}_{D'}^{-1} \\ &(g_1, \ldots, g_r, \alpha) & \mapsto & (g_1', \ldots, g_r', \det(g_1',\ldots,g_r') \alpha) . \end{array} $$ This map is not $G$-equivariant. Rather, if we have a section of $ \LL_D $, $f:G^r \rightarrow F$ (with $f(gp) = \det_D(p) f(g)$ for $p \in P_D$), then for $g_0 \in G$, we have $\tau(g_0 f ) = g_0' \det(g_0')^r \tau(f)$ (a section of ${\cal L}_{D'}$). Since $W_D$ is the restriction of such functions $f$ to $\Delta G \subset G^r$, and $\tau(\Delta G) \subset \Delta G$, we have an induced map $$ \tau: W_D \rightarrow W_{D'} $$ (an isomorphism of $F$ vector spaces), satisfying $\tau(g_0 w) = g_0' \det(g_0')^r \tau(w)$ for $g_0 \in G$, $w\in W_D$. This is the map required in (a), and now (b), (c) follow trivially. $\bullet$ \section{Resolution of singularities} \label{resolution} In this section, we define the class of northwest direction diagrams, which includes (up to a permutation of rows and columns) the skew, inversion, Rothe, and column-convex diagrams. We construct an explicit resolution of singularities of the associated configuration varieties by means of ``blowup diagrams''. We also find defining equations for these varieties. One should note that the resolutions constructed are not necessarily geometric blowups, and can sometimes be small resolutions, as in Example 8 above. We shall, as usual, think of a diagram $D$ either as a subset of ${\bf N} \times {\bf N}$, or as a list $(C_1, C_2, \ldots, C_r)$ of columns $C_j \subset {\bf N}$. We shall examine only configuration varieties, as opposed to line bundles on them, so we shall assume that the columns are without multiplicity: $C_j \neq C_{j'}$ for $j \neq j'$. \subsection{Northwest and lexicographic diagrams} A diagram $D$ is {\em northwest} if it possesses the following property: $$ (i_1,j_1),\ (i_2, j_2) \in D \Rightarrow (\min(i_1,i_2),\min(j_1,j_2)) \in D. $$ Given two subsets $C = \{i_1 < i_2 < \ldots < i_l \}, \ \ C' = \{i'_1 < i'_2 < \ldots i'_{l'}\} \subset {\bf N}$, we say $C$ is {\em lexicographically less than } $C'$ \ ($C \stackrel{\rm lex}{<} C'$) if $$ l < l\, ' \mbox{ and } i_1 = i'_1,\ \ldots ,\ i_l = i'_{l'}, $$ $$ \mbox{or } \exists\, m : \ i_1 = i'_1,\ \ldots ,\ i_{m-1} = i'_{m-1},\ i_m < i'_m. $$ In the first case, we say $C$ is an {\em initial subset} of $C'$ \ ($C \stackrel{\rm init}{\subset} C'$). A diagram $D = (C_1, C_2, \ldots)$ is {\em lexicographic} if $C_1 \stackrel{\rm lex}{<} C_2 \stackrel{\rm lex}{<} \cdots$. Note that any diagram can be made lexicographic by rearranging the order of columns. \begin{lem} If $D$ is northwest, then the lexicographic rearrangement of $D$ is also northwest. \end{lem} \noindent {\bf Proof.} (a) I claim that if $j < j'$, then either $C_j \stackrel{\rm lex}{<} C_{j'}$, or $C_{j} \stackrel{\rm init}{\supset} C_{j'}$. Let $C_j = \{i_1 < i_2 < \ldots \}$, $ C_{j'} = \{i'_1 < i'_2 < \ldots \}$. We have assumed $C_j \neq C_{j'}$. Thus $C_j \stackrel{\rm lex}{<} C_{j'}$ or $C_j \stackrel{\rm lex}{>} C_{j'}$. In the second case, $C_j \stackrel{\rm init}{\supset} C_{j'}$ or there is an $r$ such that $i_1 = i'_1, \ldots i_{r-1} = i'_{r-1}, i_r < i'_r$. By the northwest property, this last case would mean $i'_r \in C_j$, with $i_{r-1} = i'_{r-1} < i'_r < i_r$. But this contradicts the definition of $C_j$. Thus the only possibilities are those of the claim. \\ (b) It follows immediately from (a) that if $C_1 \stackrel{\rm lex}{<} C_2 \stackrel{\rm lex}{<} \cdots \stackrel{\rm lex}{<} C_{s-1} \stackrel{\rm lex}{>} C_s$, then there is a $t < s$ with $C_{t-1} \stackrel{\rm lex}{<} C_s$, $C_s \stackrel{\rm init}{\subset} C_t$, $C_s \stackrel{\rm init}{\subset} C_{t+1}$, \ldots, $C_s \stackrel{\rm init}{\subset} C_{s-1}$. \\ (c) From (b), we see that to rearrange the columns lexicographically requires only the following operation: we start with $C_1, C_2, \ldots$, and when we encounter the first column $C_s$ which violates lexicographic order, we move it as far left as possible, passing over those columns $C_i$ with $C_s \stackrel{\rm init}{\subset} C_i$. This operation does not destroy the northwest property, as we can easily check on boxes from each pair of columns in the new diagram. By repeating this operation, we get the lexicographic rearrangement, which is thus northwest. \subsection{Blowup diagrams} The combinatorial lemmas of this section will be used to establish geometric properties of configuration varieties. Given a northwest diagram $D$ and two of its columns $C, C' \subset {\bf N}$, the {\em intersection blowup diagram} $ \widehat{D} _{C,C'}$ is the diagram with the same columns as $D$ except that the new column $C \cap C'$ is inserted in the proper lexicographic position (provided $C \cap C' \neq C, C'$). \begin{lem} \label{intlem} Suppose $D$ is lexicographic and northwest, and $C \stackrel{\rm lex}{<} C'$ are two of its columns. Then: (a) $C \cap C' \stackrel{\rm init}{\subset} C'$, and (b) if $C \subset C'$, then $C \stackrel{\rm init}{\subset} C'$. \end{lem} \noindent {\bf Proof.} (a) If $i \in C_j \cap C_{j'}$ and $i > i' \in C_{j'}$, then $i' \in C_j$ by the northwest property. Similarly for (b). $\bullet$ \begin{lem} If $D$ is lexicographic and northwest, then $ \widehat{D} _{C,C'}$ is also lexicographic and northwest. \end{lem} \noindent {\bf Proof.} If $C = C_j, C' = C_{j'}$ with $j < j'$, and we insert the column $C \cap C' \stackrel{\rm init}{\subset} C'$ immediately before $C'$, then we easily check that the resulting diagram is again northwest. Hence $ \widehat{D} _{C,C'}$, which is the lexicographic rearrangement of this, is also northwest by a previous lemma. $\bullet$ Consider the columns $C_1, C_2, \ldots \subset {\bf N}$ of a northwest diagram $D$, and take the smallest collection $\{ \widehat{C} _1 \stackrel{\rm lex}{<} \widehat{C} _2 \stackrel{\rm lex}{<} \cdots \}$ of subsets of ${\bf N}$ which contains the $C_i$ and is closed under taking intersections. Then we define a new diagram $ \widehat{D} = ( \widehat{C} _1, \widehat{C} _2, \ldots )$ which we call the {\em maximal intersection blowup diagram} of $D$. Clearly $ \widehat{D} \, ^{\widehat{}} = \widehat{D} $. Repeated application of the above lemma shows that if $D$ is lexicographic and northwest, then so is $ \widehat{D} $. \vspace{1em} \noindent {\bf Examples.} For one of the (non-northwest) diagrams considered previously, we have: $$ D_4 = \begin{array}{ccc} \Box & \Box & \\ \Box & & \Box \\ & \Box & \Box \end{array} \ \ \ \ \ \ \widehat{D} _4 = \begin{array}{cccccc} \Box & \Box & \Box & & & \\ & \Box & & \Box & \Box & \\ & & \Box & & \Box & \Box \end{array} $$ For the diagrams $D_7$ and $D_8$ in the previous examples, $D_8 = \widehat{D} _7$. $\bullet$ Consider the columns $C \subset {\bf N}$ of a diagram $D$ as a partially ordered set under $\subset$, ordinary inclusion. Given two distinct columns $C$, $C'$, we say $C'$ {\em minimally covers} $C$ (or simply $C'$ {\em covers} $C$) if $C \subset C'$ and there is no column of $D$ strictly included between $C$ and $C'$. \begin{lem} \label{maxmin} Let $D$ be a lexicographic northwest diagram, and $C_L$ be the last column of $D$. Then: \\ (a) there is a column $C_l \neq C_L$ such that $$ ( \bigcup_{C \neq C_L} C) \cap C_L = C_l \cap C_L ; $$ (b) if $ \widehat{D} = D$, then $C_L$ covers at most one other column $C_l$ and is covered by at most one other column $C_u$. \end{lem} \noindent {\bf Proof.} (a) Now, by Lemma \ref{intlem}, $C \cap C_L \stackrel{\rm init}{\subset} C_L$ for any column $C$. Hence the sets $C \cap C_L$ for $C \neq C_L$ are linearly ordered under inclusion, and there is a largest one $C_l \cap C_L$. Thus $$ ( \bigcup_{C \neq C_L} C) \cap C_L = \bigcup_{C \neq C_L} (C \cap C_L) = C_l \cap C_L . $$ (b) By Lemma \ref{intlem}, the columns with $C \subset C_L$ satisfy $C \stackrel{\rm init}{\subset} C_L$ and are linearly ordered, so there is at most one maximal $C_u$. Now suppose $C_u, C_u' \stackrel{\rm lex}{<} C_L$ are columns of $D$ both covering $C_L$. Then again by Lemma \ref{intlem}, we have $C_u \cap C_u' \stackrel{\rm lex}{\leq} C_u$ or $ \stackrel{\rm lex}{\leq} C_u'$, so that $C_u \cap C_u' \neq C_L$. But $C_u \cap C_u'$ is between $C_L$ and $C_u$, and between $C_L$ and $C_u'$. Hence $C_u = C_u \cap C_u' = C_u'$. $\bullet$ \subsection{Blowup varieties} Let $D = (C_1, C_2, \ldots)$ be a lexicographic northwest diagram, and $ \widehat{D} = ( \widehat{C} _1, \widehat{C} _2, \ldots)$ be its maximal intersection blowup. Recall that $ \widehat{D} $ is obtained by adding certain columns to $D$, so there is a natural projection map \begin{eqnarray*} {\mathop{\rm pr}} : \ \mathop{\rm Gr}( \widehat{D} ) & \rightarrow & \mathop{\rm Gr}(D) \\ (V_{ \widehat{C} })_{ \widehat{C} \in \widehat{D} } & \mapsto & (V_{C})_{C \in D}, \end{eqnarray*} obtained by forgetting some of the linear subspaces $V_{ \widehat{C} } \in \mathop{\rm Gr}( \widehat{C} )$. \begin{prop} \label{birational} If $D$ has $\leq N$ rows, then $$ {\mathop{\rm pr}} : \ \mathop{\rm Gr}( \widehat{D} ) \rightarrow \mathop{\rm Gr}(D) $$ induces a birational map of algebraic varieties $$ {\mathop{\rm pr}} : \ { \cal F } _{ \widehat{D} } \rightarrow \FF_D . $$ \end{prop} \noindent {\bf Proof.} Consider the dense open sets $ \FF^{\mbox{\rm \tiny gen}} _{ \widehat{D} } \subset { \cal F } _{ \widehat{D} }$ and $ \FF^{\mbox{\rm \tiny gen}} _D \subset \FF_D $ of Proposition \ref{bigorbit}, consisting of subspaces in general position. If we consider an element $g \in GL(N)$ as a sequence of column vectors $g = (v_1,\ldots,v_n)$, and $C = \{i_1, i_2, \ldots \} \subset \{1,\ldots,N\}$, recall that we define $g(C) = {\mathop{\rm Span}} _F(v_{i_1},v_{i_2},\ldots ) \in \mathop{\rm Gr}(C)$. By definition, any element of $ \FF^{\mbox{\rm \tiny gen}} _{ \widehat{D} }$ can be written as $(g( \widehat{C} _1), g( \widehat{C} _2), \ldots ) \in \mathop{\rm Gr}(D)$ for some $g \in GL(N)$. Now, any column of $ \widehat{D} $ can be written as an intersection of columns of $D$: $ \widehat{C} = C_{j_1} \cap C_{j_2} \cap \cdots$. Then we have $g( \widehat{C} ) = g(C_{j_1}) \cap g(C_{j_2}) \cap \cdots$, so the projection map \begin{eqnarray*} {\mathop{\rm pr}} : \ \FF^{\mbox{\rm \tiny gen}} _{ \widehat{D} } & \rightarrow & \FF^{\mbox{\rm \tiny gen}} _D \\ (g( \widehat{C} ))_{ \widehat{C} \in \widehat{D} } & \mapsto & (g(C))_{C \in D} \end{eqnarray*} can be inverted: $$ \begin{array}{rrcl} {\mathop{\rm pr}} ^{-1} : & \FF^{\mbox{\rm \tiny gen}} _D & \rightarrow & \FF^{\mbox{\rm \tiny gen}} _{ \widehat{D} } \\ & (g(C))_{C \in D} & \mapsto & (g( \widehat{C} ) = g(C_{j_1}) \cap g(C_{j_2}) \cap \cdots)_{ \widehat{C} \in \widehat{D} }. \end{array} $$ Hence the map is birational on the configuration varieties as claimed. $\bullet$ \subsection{Intersection varieties} \label{intersection varieties} Now, given a diagram $D$, define the {\em intersection variety} $ \II_D $ of $D$ by: $$ \II_D = \{ (V_C)_{C \in D} \in \mathop{\rm Gr}(D) \mid \forall C,C',\ldots \in D, \ \ \dim( V_C \cap V_{C'} \cap \cdots ) \geq |C \cap C' \cap \cdots|\}. $$ Clearly $ \II_D $ is a projective subvariety of $\mathop{\rm Gr}(D)$, and $ \FF_D \subset \II_D $. If $ \widehat{D} = D$ (up to rearrangement of column order), then the intersection conditions reduce to inclusions: $$ \II_D = \{ (V_C)_{C \in D} \in \mathop{\rm Gr}(D) \mid C \subset C' \Rightarrow V_C \subset V_{C'} \}. $$ \vspace{1em} \noindent {\bf Example.} For the diagram $D_4$ of Section \ref{definitions and examples}, and $N=4$, $ { \cal I } _{D_4}$ has two irreducible components, $ { \cal F } _{D_4}$ and $ { \cal F } _{D_5}$. That is, as before, if we have three lines in ${\bf P}^3$ with non-empty pairwise intersections, then either they are coplanar, or they all intersect in a point. $\bullet$ \begin{lem} \label{union conditions} Let $D$ be a northwest diagram, and $ \II_D $ its intersection variety. Then any configuration $(V_C)_{C \in D} \in \II_D $ satisfies $$ \dim(V_C + V_{C'} + \cdots) \leq |C \cup C' \cup \cdots| $$ for any columns $C, C',\ldots$ of $D$. \end{lem} \noindent {\bf Proof.} Without loss of generality, assume $D$ is lexicographic. We use induction on the number of columns in $D$. Now any list $C, C', \ldots$ of columns of $D$ also constitutes a lexicographic northwest diagram, so to carry through the induction we need only prove the statement for {\em all} the columns $C_1, C_2, \ldots, C_L$ of $D$. Now, by Lemma \ref{maxmin}, there is a column $C_l \neq C_L$ such that $( \cup_{C\neq C_L} C) \cap C_L = C_l \cap C_L$. Then we have \begin{eqnarray*} \dim(\ ( \sum_{C\neq C_L} V_C ) \cap V_{C_L}\ ) & \geq & \dim( \sum_{C \neq C_L} (V_C \cap V_{C_L})\ ) \\ & \geq & \dim(\, V_{C_l} \cap V_{C_L} ) \\ & \geq & |\, C_l \cap C_L| \ \ \ \mbox{ since } (V_C) \in \II_D \\ & = & | \ ( \bigcup_{C\neq C_L} C\ ) \cap C_L \ | . \end{eqnarray*} Thus we may write \begin{eqnarray*} \dim( \sum_{C\in D} V_C ) & = & \dim( \sum_{C \neq C_L} V_C \, ) + \dim(V_{C_L}) -\dim(\, ( \sum_{C\neq C_L} V_C \, ) \cap V_{C_L} \ ) \\ & \leq & |\! \bigcup_{C\neq C_L} C \, | + |\, C_L| - |\, ( \bigcup_{C\neq C_L} C \ ) \cap C_L | \ \ \ \mbox{ by induction} \\ & = & |\bigcup_{C\in D} C \ | \ \ \ \bullet \end{eqnarray*} \begin{lem} If $D$ is a northwest diagram with $\leq N$ rows and $ \widehat{D} = D$ (up to rearrangement of column order), then $ \FF_D $ is an irreducible component of $ \II_D $. \end{lem} \noindent {\bf Proof.} Recall that $ \FF_D $ is always irreducible. Thus it suffices to show that $ \FF^{\mbox{\rm \tiny gen}} _D$ is an open subset of $ \II_D $. Consider the set $ \II_D ^{\mbox{\rm \tiny gen}}$ of configurations $(V_C)_{C \in D}$ satisfying, for every list $C, C',\ldots$ of columns in $D$, $$ \dim (V_C + V_{C'} + \cdots) = |C \cup C' \cup \cdots | $$ and $$ \dim (V_C \cap V_{C'} \cup \cdots) = |C \cap C' \cap \cdots | . $$ This is an open subset of $ \II_D $ by the previous lemma. I claim that $ \FF^{\mbox{\rm \tiny gen}} _D = \II_D ^{\mbox{\rm \tiny gen}}$. To see this equality, let $(V_C)_{C \in D} \in \II_D $ satisfy the above rank conditions, and we will find a basis $g = (v_1,\ldots,v_N)$ of $V=F^N$ such that $V_C = g(C)$ for all $C$. (C.f. the proof of Proposition \ref{bigorbit}.) As before, we consider the columns as a poset under ordinary inclusion. We begin by choosing mutually independent bases for those $V_C$ where $C$ is a minimal element of the poset. This is possible because $ \dim {\mathop{\rm Span}} (V_C \mid C \mbox{ minimal}) = \sum_{C \ \mbox{\tiny minml}} |C|. $ Now we consider the $V_C$ where $C$ covers a minimal column. We start with the basis vectors already chosen, and add enough vectors, all mutually independent, to span each space. Again, the dimension conditions ensure there will be no conflict in choosing independent vectors, since the $V_C$ can have no intersections with each other except those due to the intersections of columns. The condition $ \widehat{D} = D$ ensures that all these intersections are (previously considered) columns. We continue in this way for the higher layers of the poset. We will not run out of independent basis vectors because all the columns of $D$ are contained in $\{1,\ldots,N\}$. $\bullet$ \subsection{Smoothness and equations defining varieties} \begin{prop} \label{smooth} Let $D$ be a northwest diagram with $\leq N$ rows and $ \widehat{D} = D$ (up to rearrangement of column order). Then $ \FF_D = \II_D $, and $ \FF_D $ is a smooth variety. \end{prop} \noindent {\bf Proof.} (a) Let $C_L$ be the last column of $D$, and let $D'$ be $D$ without the last column. By lemma \ref{maxmin}, $C_L$ is covered by at most one other column $C_u$, and covers at most one other column $C_l$. If these columns do not exist, take $C_l = \emptyset$, $C_u = \{1,\ldots,N\}$. \\ (b) Now I claim that there is a fiber bundle $$ \begin{array}{ccc} \mathop{\rm Gr}(C_l, C_L, C_u) & \rightarrow & Z \\ & & \downarrow \\ & & \mathop{\rm Gr}(D') \ \, \end{array} $$ where $\mathop{\rm Gr}(C_l, C_L, C_u)$ denotes the Grassmannian of $|C_L|$-dimensional linear spaces which contain a fixed $|C_l|$-dimensional space and are contained in a fixed $|C_u|$-dimensional space; and $$ Z = \{ (\, (V_{C'})_{C'}, V_L ) \in \mathop{\rm Gr}(D') \times \mathop{\rm Gr}(C_L) \mid V_{C_l} \subset V_L \subset V_{C_u} \}. $$ This is clear. See also~\cite{BD1}.\\ (c) Note that $ \II_D = ( { \cal I } _{D'} \times \mathop{\rm Gr}(C_L) ) \cap Z$. This is because of the uniqueness of $C_l$ and $C_u$. Thus the above fiber bundle restricts to \begin{eqnarray*} \mathop{\rm Gr}(C_l, C_L, C_u) & \rightarrow & \II_D \\ & & \downarrow \\ & & { \cal I } _{D'} \ \ , \end{eqnarray*} which is thus also a fiber bundle. \\ (d) Now apply the above construction repeatedly, dropping columns of $D$ from the end. Finally we obtain $ \II_D $ as an iterated fiber bundle whose fibers at each step are smooth and connected (in fact they are Grassmannians). In particular, $ \II_D $ is smooth and connected. \\ (e) Since $ \II_D $ is a smooth, connected, projective algebraic variety, it must be irreducible. But by a previous lemma, $ \FF_D $ is an irreducible component of $ \II_D $. Therefore $ \FF_D = \II_D $, a smooth variety. $\bullet$ \begin{prop} \label{conn fibers} Let $D$ be a northwest diagram with $\leq N$ rows. Then $ \FF_D = \II_D $, and the birational projection map $ { \cal F } _{ \widehat{D} } \rightarrow \FF_D $ has connected fibers. \end{prop} \vspace{1em} \noindent {\bf Proof.} (a) I claim the following: if $ \widehat{C} $ is a column of $ \widehat{D} $ such that for all $C \in \widehat{D} $ with $C \stackrel{\neq}{\subset} \widehat{C} $ we have $C \in D$, then the projection map $ { \cal I } _{D \cup \widehat{C} } \rightarrow \II_D $ is onto, with connected fibers. Suppose $(V_C)_{C\in D}$ is a configuration in $ \II_D $. Let $$ V_u = \bigcap_{C \in D \atop C \supset \widehat{C} } V_C \ \ \ \ \mbox{ and } \ \ \ \ V_l = \sum_{C \in D \atop C \subset \widehat{C} } V_C. $$ Then $\dim(V_u) \geq | \widehat{C} |$ since $(V_C) \in \II_D $, and $\dim(V_l) \leq | \widehat{C} |$ by Lemma \ref{union conditions}. Clearly $V_l \subset V_u$. Now choose an arbitrary $V_{ \widehat{C} }$ between $V_l$ and $V_u$ with $\dim(V_{ \widehat{C} }) = | \widehat{C} |$. Then for any list of columns $C, C', \ldots \in D$, we have either: \\ (i) $ \widehat{C} \cap C \cap C'\cdots = \widehat{C} $, and $$ V_{ \widehat{C} \cap C \cap C'\cdots} \ = \ V_{ \widehat{C} } \ = \ V_{ \widehat{C} } \cap V_u \ \subset \ V_{ \widehat{C} } \cap V_C \cap V_{C'} \cap \cdots ; $$ or (ii) $ \widehat{C} \cap C \cap C'\cdots \stackrel{\neq}{\subset} \widehat{C} $, so that $ \widehat{C} \cap C \cap C'\cdots \in D$ by hypothesis, and $$ V_{ \widehat{C} \cap C \cap C'\cdots} \ \subset \ V_l \cap V_C \cap V_{C'} \cdots \ \subset \ V_{ \widehat{C} } \cap V_C \cap V_{C'} \cap \cdots . $$ In either case $(V_C)_{C \in D \cup \widehat{C} } \in { \cal I } _{D \cup \widehat{C} }$. Thus $ { \cal I } _{D \cup \widehat{C} } \rightarrow \II_D $ is onto, and the fibers are the Grassmannians $\mathop{\rm Gr}(V_l, |C|, V_u)$. \\ (b) We now see that $ { \cal I } _{ \widehat{D} } \rightarrow \II_D $ is onto (with connected fibers) by repeated application of (a), starting with $ \widehat{C} $ minimal in the poset of columns of $ \widehat{D} $ and proceeding upward. \\ (c) By the previous proposition, the projection map takes $ { \cal I } _{ \widehat{D} } = { \cal F } _{ \widehat{D} } \rightarrow \FF_D $. But $ { \cal I } _{ \widehat{D} } \rightarrow \II_D $ is onto, so $ \FF_D = \II_D $, and we are done. $ \bullet $ The above proposition shows that for northwest diagrams, $ \FF_D $ is defined by the rank conditions of $ \II_D $. In general, we state the \begin{conj} For an arbitrary diagram $D$, $ \FF_D $ is the set of configurations satisfying \begin{eqnarray*} \dim(V_C + V_{C'} + \cdots) & \leq & |\ C \cup C' \cup \cdots| \\ \dim( V_C \cap V_{C'} \cap \cdots ) & \geq & |\ C \cap C' \cap \cdots| \end{eqnarray*} for every list $C,C',\ldots $ of columns of $D$. Equivalently(?), the variety defined by these equations is irreducible. \end{conj} \section{Cohomology of line bundles} \label{cohomology} Using the technique of Frobenius splitting, we show certain surjectivity and vanishing theorems for line bundles on configuration varieties. In particular, we show that for any northwest diagram $D$, $ \FF_D $ is normal, and projectively normal with respect to $ \LL_D $ (so that global sections of $L_D$ on $ \FF_D $ extend to $\mathop{\rm Gr}(D)$); and $ \FF_D $ has rational singularities. The material of section \ref{frobenius} was shown to me by Wilberd van der Kallen. \subsection{Frobenius splittings of flag varieties} \label{frobenius} The technique of Frobenius splitting, introduced by V.B. Mehta, S. Ramanan, and A. Ramanathan {}~\cite{MR},~\cite{RR},~\cite{R1},~\cite{R2}, is a method for proving certain surjectivity and vanishing results. Given two algebraic varieties $Y \subset X$ defined over an algebraically closed field $F$ of characteristic $p > 0$, with $Y$ a closed subvariety of $X$, we say that the pair $Y \subset X$ is {\em compatibly Frobenius split} if: \\ (i) the $p^{th}$ power map $F: {\cal O} _X \rightarrow F_* {\cal O} _X$ has a splitting, i.e. an $ {\cal O} _X$-module morphism $\phi: F_* {\cal O} _X \rightarrow {\cal O} _X$ such that $\phi F$ is the identity; and \\ (ii) we have $\phi(F_* I) = I$, where $I$ is the ideal sheaf of $Y$. Mehta and Ramanathan prove the following \begin{thm} \label{vanishing} Let $X$ be a projective variety, $Y$ a closed subvariety, and $L$ an ample line bundle on $X$. If $Y \subset X$ is compatibly split, then $H^i(Y,L) = 0$ for all $i >0$, and the restriction map $H^0(X,L) \rightarrow H^0(Y,L)$ is surjective. Furthermore, if $Y$ and $X$ are defined and projective over ${\bf Z}$ (and hence over any field), and they are compatibly split over any field of positive characteristic, then the above vanishing and surjectivity statements also hold for all fields of characteristic zero. $\bullet$ \end{thm} Our aim is to show that, for $D$ a northwest diagram, $ \FF_D \subset \mathop{\rm Gr}(D)$ is compatibly split. The above theorem and Theorem \ref{weyl mod is image} will then imply that $S_D^* \cong H^0( \FF_D , \LL_D ) = \sum_i (-1)^i H^i( \FF_D , \LL_D )$, the Euler characteristic of $ \LL_D $. We will also need the following result of Mehta and V. Srinivas~\cite{MS}: \begin{prop} \label{normality} Let $Y$ be a projective variety which is Frobenius split, and suppose there exists a smooth irreducible projective variety $Z$ which is mapped onto $Y$ by an algebraic map with connected fibers. Then $Y$ is normal. Furthermore, if $Y$ is defined over ${\bf Z}$, and is normal over any field of positive characteristic, then $Y$ is also normal over all fields of characteristic zero. $\bullet$ \end{prop} \begin{prop} \label{pushforward} Let $f: Z\rightarrow X$ be a separable morphism with connected fibers, where $X$ and $Z$ are projective varieties and $X$ smooth. If $Y \subset Z$ is compatibly split, then so is $f(Y) \subset X$. $\bullet$ \end{prop} We will show our varieties are split by using the above proposition to push forward a known splitting due to Ramanathan~\cite{R2} and O. Mathieu~\cite{M}. For an integer $n$ and $n$ permutations $w, w', \ldots$, define $$ X_n = \underbrace{G \stackrel{B}{\times} G \stackrel{B}{\times} \cdots \stackrel{B}{\times} G}_{n \ \mbox{\rm \tiny factors} }/B, $$ and the twisted multiple Schubert variety $$ Y_{w, w' \ldots} = \overline{B w B} \stackrel{B}{\times} \overline{B w' B} \stackrel{B}{\times} \cdots \subset X_n $$ Note that we have an isomorphism $$ \begin{array}{ccc} X_n & \rightarrow & (G/B)^n \\ (g, g', g'', \ldots) & \mapsto & (g, g g', g g' g'',\ldots). \end{array} $$ \begin{prop}{(Ramanathan-Mathieu)} \label{Ram-Math} Let $G$ be a reductive algebraic group over a field of positive characteristic with Weyl group $W$ and Borel subgroup $B$, and let $w_0, w_1, \ldots w_r \in W$. Then $Y_{w_0, w_1, \ldots} \subset X_{r+1}$ is compatibly split. $\bullet$ \end{prop} Now, for Weyl group elements $u_1, \ldots, u_r$, define a variety $ { \cal F } _{u_1,\ldots,u_r} \subset (G/B)^r$ by $$ { \cal F } _{u_1,\ldots,u_r} = \overline{G \cdot ( u_1 B,\ldots, u_r B)}. $$ \begin{prop}{(van der Kallen)} \label{splitting} Let $w_1, \ldots, w_r$ be Weyl group elements, and define $u_1 = w_1, u_2 = w_1 w_2, \ldots u_r = w_1 \cdots w_r$. Suppose $w_1, \ldots w_r$ satisfy $\ell(w_1 w_2 \cdots w_r) = \ell(w_1) + \ell(w_2) + \cdots + \ell(w_r)$, or equivalently $\ell(u_j) = \ell(u_{j-1}) + \ell(u_{j-1}^{-1} u_j)$ for all $j$. Then the pair $ { \cal F } _{u_1,\ldots,u_r} \subset (G/B)^r$ is compatibly split. \end{prop} \noindent {\bf Proof.} Define $$ \begin{array}{rccc} f: & X_{r+1} & \rightarrow & (G/B)^r \\ & (g_0, g_1, \ldots, g_r) & \mapsto & (g_0 g_1,\, g_0 g_1 g_2,\, \ldots,\, g_0 g_1 \ldots g_r). \end{array} $$ We will examine the image under this map of $$ Y \stackrel{\rm def}{=} Y_{w_0, w_1, \ldots w_r} = G \stackrel{B}{\times} \overline{B w_1 B} \stackrel{B}{\times} \cdots \stackrel{B}{\times} \overline{B w_r B} \subset X_{r+1} , $$ where $w_0$ is the longest element of the Weyl group. It is well known that, under the given hypotheses, we have $(B w_1 B) \cdots (B w_r B) = B w_1 \cdots w_r B$, and that the multiplication map $$ B w_1 B \stackrel{B}{\times} \cdots \stackrel{B}{\times} B w_r B \rightarrow B w_1 \cdots w_r B $$ is bijective. Thus any element $(g, b_1 w_1 b_1',\ldots, b_r w_r b_r' B) $ (for $b_i, b_i' \in B$) can be written as $(g , b w_1, w_2, \ldots, w_r B) = (g b, w_1, w_2, \ldots, w_r B)$ for some $b\in B$, and $$ f(G \stackrel{B}{\times} B w_1 B \stackrel{B}{\times} \cdots \stackrel{B}{\times} B w_n B) = f(G \stackrel{B}{\times} w_1 \stackrel{B}{\times} \ldots \stackrel{B}{\times} w_r B) = G (u_1 B, \ldots , u_r B). $$ Hence $f(Y) = { \cal F } _{u_1,\ldots,u_r}$, since our varieties are projective. Now, $f$ is a separable map with connected fibers between smooth projective varieties, so the compatible splitting of the previous proposition pushes forward by Proposition \ref{pushforward}. $\bullet$ We will need the following lemmas to show that our configuration varieties have rational singularities. \begin{lem}{(Kempf~\cite{K})} \label{Kempf} Suppose $f : Z \rightarrow X$ is a separable morphism with generically connected fibers between projective algebraic varieties $Z$ and $X$, with $X$ normal. Let $L$ be an ample line bundle on $X$, and suppose that $H^i(Z, f^* L^{\otimes n}) = 0$ for all $i > 0$ and all $n >> 0$. Then $R^i f_* {\cal O} _Z = 0 $ for all $i>0$. $\bullet$ \end{lem} Resuming the notation of Prop \ref{Ram-Math}, let $w_1, \ldots w_n$ be arbitrary Weyl group elements, and let $ \lambda _1, \ldots \lambda _n$ be arbitrary weights of $G$. Let $X_n$ be as before, and define the line bundle $ {\cal L} _{ \lambda _1, \ldots \lambda _n}$ on $X_n$ and on $Y_{w_1, \ldots, w_n} \subset X_n$ as the quotient of $G^n \times F$ by the $B^n$-action $$ (b_1, \ldots, b_n)\cdot (g_1, g_2, \ldots, g_n, a) \stackrel{\rm def}{=} (g_1 b_1, b_1^{-1} g_2 b_2, \ldots , b_{n-1}^{-1} g_n b_n,\, \lambda _1(b_1) \cdots \lambda _n(b_n) a) . $$ Note that under the identification $X_n \cong (G/B)^n$, $ {\cal L} _{ \lambda _1, \ldots, \lambda _n}$ is isomorphic to the Borel-Weil line bundle $G^n \stackrel{B^n}{\times} ( \lambda _1^{-1},\ldots \lambda _n^{-1})$. \begin{lem} \label{nonample} Assume $ \lambda _1, \ldots \lambda _n$ are dominant weights (possibly on the wall of the Weyl chamber). Then $H^i(Y_{w_1, \ldots, w_n},\, {\cal L} _{ \lambda _1, \ldots, \lambda _n}) = 0$ for all $i > 0$. \end{lem} \noindent {\bf Proof} (van der Kallen). Note that $ {\cal L} _{ \lambda _1,\ldots, \lambda _n}$ is effective, but not necessarily ample, so we cannot deduce the conclusion directly from Theorem \ref{vanishing}. Recall the following facts from $B$-module theory {}~\cite{P},~\cite{vdK}: \\ (a) An excellent filtration of a $B$-module is one whose quotients are isomorphic to Demazure modules $H^0(\overline{BwB}, {\cal L} _{ \lambda })$, for Weyl group elements $w$ and dominant weights $ \lambda $. \\ (b) If $M$ has an excellent filtration, and $ {\cal E} (M) \stackrel{\rm def}{=} G \stackrel{B}{\times} M$ is the corresponding vector bundle on $G/B$, then $H^i(G/B, {\cal E} (M)) = 0 $ for all $i >0$, and $H^0(G/B, {\cal E} (M))$ has an excellent filtration. \\ (c) Polo's Theorem: If $M$ has an excellent filtration, then so does $( \lambda ^{-1}) \otimes M$ for any dominant weight $ \lambda $. Now consider the fiber bundle $$ \begin{array}{ccc} Y_{w_2,\ldots, w_n} & \rightarrow & Y_{w_1, w_2, \ldots, w_n} \\ & & \downarrow \\ & & \overline{Bw_1B} \end{array} $$ which leads to the spectral sequence $$ H^i(\ \overline{Bw_1B}, \ {\cal E} ( \, ( \lambda _1^{-1}) \otimes H^j(Y_{w_2,\ldots , w_n}, {\cal L} _{ \lambda _2,\ldots, \lambda _n})\, )\ ) \Rightarrow H^{i+j}(Y_{w_1, w_2, \ldots, w_n}, {\cal L} _{ \lambda _1, \lambda _2,\ldots, \lambda _n}) . $$ By induction, assume that $H^j(Y_{w_2, \ldots, w_n}, {\cal L} _{ \lambda _2,\ldots, \lambda _n}) = 0$ for $j>0$, and that $H^0(Y_{w_2, \ldots, w_n}, {\cal L} _{ \lambda _2,\ldots, \lambda _n})$ has an excellent filtration. Then applying (b) and (c), we find $$ H^i(Y_{w_1, w_2, \ldots, w_n}, {\cal L} _{ \lambda _1, \lambda _2,\ldots, \lambda _n}) = H^i(\ \overline{Bw_1B}, \, {\cal E} ( \, ( \lambda _1^{-1}) \otimes H^0(Y_{w_2,\ldots , w_n}, {\cal L} _{ \lambda _2,\ldots, \lambda _n})\, ) \ ) = 0 $$ for $i>0$, and that $H^0(Y_{w_1, w_2, \ldots, w_n}, {\cal L} _{ \lambda _1, \lambda _2,\ldots, \lambda _n})$ has an excellent filtration. $\bullet$ \subsection{Frobenius splitting of Grassmannians} We would now like to push forward the Frobenius splittings found above for flag varieties to get splittings of configuration varieties. For this we need a combinatorial prerequisite. Given a diagram $D = (C_1, C_2,\ldots, C_r)$ with $\leq N$ rows, consider a sequence of permutations (Weyl group elements) $u_1, u_2, \ldots \in \Sigma_N$ such that, for all $j$: \\ ($\alpha$) $\ell(u_j) = \ell(u_{j-1})+\ell(u_{j-1}^{-1} u_{j}) $, and \\ ($\beta$) $u_j(\ \{1,2,\ldots,|C_j| \ \}) = C_j$. The first condition says that the sequence is increasing in the weak order on the Weyl group. In the next section, we will give an algorithm which produces such a sequence for any northwest diagram, so that the following theorem will apply: \begin{prop} \label{config split} If $D$ a diagram which admits a sequence of permutations $u_1, u_2, \ldots$ satisfying ($\alpha$) and ($\beta$) above, then the pair $ \FF_D \subset \mathop{\rm Gr}(D)$ is compatibly split for any field $F$ of positive characteristic. \mbox \\ Hence over an algebraically closed field $F$ of arbitrary characteristic, \\ (a) the cohomology groups $H^i( \FF_D , \LL_D ) = 0$ for $i > 0$; \\ (b) the restriction map $\mathop{\rm rest}\mbox{}_{\Delta} : H^0(\mathop{\rm Gr}(D), \LL_D ) \rightarrow H^0( \FF_D , \LL_D )$ is surjective;\\ (c) $ \FF_D $ is a normal variety. \end{prop} \noindent {\bf Proof.} By ($\beta$), the maximal parabolic subgroups $P_C = \{ (x_{ij}) \in GL(N) \mid x_{ij} = 0 \mbox{ if } i \not\in C , \ j \in C \}$ satisfy $u_i B u_i^{-1} \subset P_{C_i}$. Write $$ \mathop{\rm Gr}(D) = \mathop{\rm Gr}(C_1) \times \cdots \times \mathop{\rm Gr}(C_r) \cong G/P_{C_1} \times \cdots \times G/P_{C_r}, $$ and consider the $G$-equivariant projection $$ \begin{array}{rccc} \phi: & (G/B)^r & \rightarrow & \mathop{\rm Gr}(D) \\ & (g_1 B,\ldots,g_r B) & \mapsto & (g_1 u_1^{-1} P_{C_1},\ldots,g_r u_r^{-1} P_{C_r}) \end{array} $$ Then we have $\phi(u_1 B, \ldots, u_r B) = (I \, P_{C_1},\ldots, I \, P_{C_r})$ and $\phi( { \cal F } _{u_1,\ldots,u_r}) = \FF_D $. Since $\phi$ is a map with connected fibers between smooth projective varieties, we can push forward the compatible splitting for $ { \cal F } _{u_1,\ldots,u_r} \subset (G/B)^r$ found in the previous section. Applying Theorem \ref{vanishing} and Propositions \ref{normality} and \ref{conn fibers}, we have the assertions of the theorem. $\bullet$ Note that (b) and (c) of the Proposition are equivalent to the projective normality of $ \FF_D $ with respect to $ \LL_D $. \begin{conj} For any diagram $D$, and any Weyl group elements $u_1, \ldots u_r$, the pairs $ \FF_D \subset \mathop{\rm Gr}(D)$ and $ { \cal F } _{u_1, \ldots u_r} \subset (G/B)^r$ are compatibly split. \end{conj} In order to prove the character formula in the last section of this paper, we will need stronger relations between the singular configuration varieties and their desingularizations. In particular, we will show that our varieties have rational singularities. \begin{lem} Let $X$, $Y$ be algebraic varities with an action of an algebraic group $G$, and $f: X\rightarrow Y$ an equivariant morphism. Assume that $X$ has an open dense $G$-orbit $G\cdot x_0$, and take $y_0 = f(x_0)$, $G_0 = \mathop{\rm Stab}_G y_0$. Then $f^{-1}(y_0) = \overline{G_0 \cdot x_0}$. In particular, if $G_0$ is connected, then $f^{-1}(y_0)$ is connected and irreducible. \end{lem} \noindent {\bf Proof.} For $F = {\bf C}$, this is trivial. Take $x_1 \in f^{-1}(y_0)$, and consider a path $x(t) \in X$ such that $x(0) = x_1$ and $x(t) \in G\cdot x_0$ for small $t>0$. Then the path $f(x(t))$ lies in $G\cdot y_0$ for small $t \geq 0$, and we can lift it to a path $g(t) \in G$ such that $g(0) = \mathop{\rm id}$ and $f(x(t)) = g(t)\cdot y_0$ for small $t \geq 0$. Then $\tilde{x}(t) \stackrel{\rm def}{=} g(t)^{-1}\cdot x(t)$ satisfies $\tilde{x}(0) = x_1$, $\tilde{x}(t) \in G_0\cdot x_0$ for small $t > 0$. For general $F$, T. Springer has given the following clever argument. Assume without loss of generality that $X$ is irreducible and $G\cdot y_0$ is open dense in $Y$. Since an algebraic map is generically flat, and $G\cdot y_0$ is open, all the irreducible components $C$ of $f^{-1}(y_0)$ have the same dimension $\dim C = \dim X - \dim Y$. Let $Z = \overline{G\cdot C}$ be the closure of one of these components. Now, the restriction $f: Z \rightarrow Y$ also satisfies our hypotheses, with $C \subset Z$ again a component of the fiber of the restricted $f$, so we again have $\dim C = \dim Z - \dim Y$, and $\dim Z = \dim X$. Thus $G\cdot C$ is an open subset of $X$, since $X$ is irreducible. Now consider the open set $G\cdot C \cap G\cdot x_0 \subset X$. Choose a point $z$ in this set which does not lie in any other component $C'$ of our original $f^{-1}(y_0)$. For any other component $C'$, choose a similar point $z'$. But we have $g \cdot z \in C$, $g'\cdot z' = g_0 g\cdot z \in C'$ for some $g, g', g_0 \in G$, and in fact $g_0 \in G_0$. Thus $C' = g_0 \cdot C$, and $G_0$ permutes the components transitively. Hence, $G_0 \cdot x_0 $ has at least as many irreducible components as the whole $f^{-1}(y_0)$, and the lemma follows. $\bullet$ \begin{prop}{(Inamdar-van der Kallen)} \label{rational sing} Suppose $D_1$, $D_2$ are diagrams admitting sequences of permutations with ($\alpha$) and ($\beta$) as above, such that $D_2$ is obtained by removing some of the columns of $D_1$. Denote $ { \cal F } _1 = { \cal F } _{D_1}$, $ { \cal F } _2 = { \cal F } _{D_2}$, $ {\cal L} _2 = {\cal L} _{D_2}$, and consider the projection $ {\mathop{\rm pr}} : { \cal F } _1 \rightarrow { \cal F } _2$.\\ Then: \\ (a) $H^0( { \cal F } _1, {\mathop{\rm pr}} ^* {\cal L} _2 ) = H^0( { \cal F } _2, {\cal L} _2 )$, and this $G$-module has a good filtration (one whose quotients are isomorphic to $H^0(G/B, {\cal L} _{ \lambda })$ for dominant weights $ \lambda $). \\ (b) $H^i( { \cal F } _1, {\mathop{\rm pr}} ^* {\cal L} _2 ) = H^i( { \cal F } _2, {\cal L} _2 ) = 0$ for all $i > 0$. \\ (c) $R^i {\mathop{\rm pr}} _* {\cal O} _{ { \cal F } _2} = 0$ for all $i >0$. \\ (d) If $F$ has characteristic zero, then $ \FF_D $ has regular singularities for any northwest diagram $D$. \end{prop} \noindent {\bf Proof.} (i) Consider a sequence of permutations $w_1, w_2, \ldots w_r$ (where $r$ is the number of columns in $D_1$) such that $u_1 = w_1, u_2 = w_1 w_2, \ldots $ satisfies ($\alpha$) and ($\beta$), and let $Y = Y_{w_0, w_1, \ldots w_r}$, (where $w_0$ is the longest permutation). Then we have a commutative diagram of surjective morphisms $$ \begin{array}{ccc} Y & \stackrel{\Phi_1}{\rightarrow} & { \cal F } _1 \\ & \stackrel{\Phi_2}{\searrow} & \downarrow \mbox{\rm \small pr} \\ & & { \cal F } _2 \end{array} $$ where $\Phi_j = \phi \, \circ f$, where $\phi$ and $f$ are the maps defined in the proofs of Propositions \ref{splitting} and \ref{config split} in the cases $D = D_j$. All of these spaces have dense $G$-orbits. Furthermore, the stabilizer of a general point in $ \FF_D $ is an intersection of parabolic subgroups and is connected. Thus, by the above lemma, the fibers of $\Phi_1$ are generically connected. \\ (ii) Now (i) and Lemma \ref{nonample} insure that the hypotheses of Kempf's lemma (Proposition \ref{Kempf}) are satisfied. Thus $R^i(\Phi_1)_* {\cal O} _Y = 0$ for $i>0$, and by the Leray spectral sequence we have, for all $i \geq 0$, $$ H^i(Y, \Phi_1^* {\mathop{\rm pr}} ^* {\cal L} _2) = H^i( { \cal F } _1, (\Phi_1)_* (\Phi_1)^* {\mathop{\rm pr}} ^* {\cal L} _2) . $$ (iii) Furthermore, $ { \cal F } _1$ is normal by the previous Proposition, and $\Phi_1$ is separable with connected fibers, so \begin{eqnarray*} (\Phi_1)_* (\Phi_1)^* {\mathop{\rm pr}} ^* {\cal L} _2 & \cong & [ (\Phi_1)_* (\Phi_1)^* {\cal O} _{ { \cal F } _1} ] \otimes {\mathop{\rm pr}} ^* {\cal L} _2 \\ & \cong & {\mathop{\rm pr}} ^* {\cal L} _2. \end{eqnarray*} Thus $H^i(Y, \Phi_1^* {\mathop{\rm pr}} ^* {\cal L} _2) = H^i( { \cal F } _1, {\mathop{\rm pr}} ^* {\cal L} _2)$ for all $i \geq 0$. \\ (iv) An exactly similar argument shows that $H^i(Y, \Phi_2^* {\cal L} _2) = H^i( { \cal F } _2, {\cal L} _2)$ for all $i \geq 0$. But $\Phi_2^* = \Phi_1^* {\mathop{\rm pr}} ^*$, so for all $i$, $$ H^i( { \cal F } _1, {\mathop{\rm pr}} ^* {\cal L} _2) = H^i(Y, \Phi_2^* {\cal L} _2) = H^i( { \cal F } _2, {\cal L} _2) . $$ But we saw in Lemma \ref{nonample} that $H^i(Y, \Phi_2^* {\cal L} _2)$ vanishes for $i>0$, so (b) of the present Proposition follows. \\ (iv) We also saw in the proof of Lemma \ref{nonample} that $H^0(Y, \Phi_2^* {\cal L} _2)$ has an excellent filtration as a $B$-module. But this is equivalent to it having a good filtration as a $G$-module, so (a) follows. \\ (v) Now consider the spectral sequence $$ R^i {\mathop{\rm pr}} _* R^j(\Phi_1)_* {\cal O} _Y \Rightarrow R^{i+j} (\Phi_2)_* {\cal O} _Y. $$ For $i>0$, we have $ R^i(\Phi_1)_* {\cal O} _Y = 0$ and $R^i(\Phi_2)_* {\cal O} _Y =0 $ by (ii) above. Because of this and the normality of $ { \cal F } _1$, we have for all $i>0$, \begin{eqnarray*} 0 & = & R^i(\Phi_2)_* {\cal O} _Y \\ & = & R^i {\mathop{\rm pr}} _* (\Phi_1)_* {\cal O} _Y \\ & = & R^i {\mathop{\rm pr}} _* {\cal O} _{F_1} . \end{eqnarray*} this shows (c). \\ (vi) Now take $D_2 = D$ an arbitrary northwest diagram, and $D_1 = \widehat{D} $ its maximal blowup. Then $ {\mathop{\rm pr}} $ is a resolution of singularities by Proposition \ref{smooth}. Assume, as we will show in the next section, that $D_1$ admits a sequence of permutations as required. Then (c) holds, and this is precisely the definition of rational singularities in characteristic zero, so we have (d). $\bullet$ \subsection{Monotone sequences of permutations} Let $D = (C_1, C_2,\ldots, C_r)$ be a northwest diagram with $\leq N$ rows. In this section, we will construct by a recursive algorithm a sequence of permutations $u_1, u_2, \ldots \in \Sigma_N$ satisfying the conditions of the previous section: for all $j$, \\ ($\alpha$) $\ell(u_j) = \ell(u_{j-1})+\ell(u_{j-1}^{-1} u_{j}) $, and \\ ($\beta$) $u_j(\ \{1,2,\ldots,|C_j| \ \}) = C_j$. For each column $C$ of $D$, define the integer $$ {\mathop{\rm gap}} _ N(C) = \left\{ \begin{array}{l} \max \{ i\mid i\not\in C,\ \exists i'\in C : i < i' \} \mbox{, if this set is } \neq \emptyset \\ N \mbox{, if the above set is empty.} \end{array} \right. $$ Since $D$ is northwest, there is an integer $J_N \geq 1$ such that $$ N = {\mathop{\rm gap}} _ N(C_1) = \cdots = {\mathop{\rm gap}} _ N(C_{J_N-1}) > {\mathop{\rm gap}} _ N(C_{J_N}) = \cdots = {\mathop{\rm gap}} _ N(C_r). $$ Now define the {\em derived diagram} $D'$ of $D$ as follows. Given a column $C$ of $D$, there is a corresponding column $C'$ of $D'$: $$ C' = \{i\mid i \in C, \ i < {\mathop{\rm gap}} _ N(C) \} \cup \{ i-1 \mid i\in C,\ i > {\mathop{\rm gap}} _ N(C) \}. $$ That is, we take $C$ and push all squares below the $ {\mathop{\rm gap}} _ N(C)$-th row upward by one place. \begin{lem} If $D$ is northwest with $\leq N$ rows, then $D'$ is northwest with $\leq N-1$ rows. \end{lem} \noindent {\bf Proof.} The only doubtful case in checking the northwest property is that of two squares $(i_1,j_1)$ and $(i_2,j_2)$ in $D'$ with $j_1 < J_N \leq j_2$ and $i_1 > i_2$. Since $j_1 < J_N$, we have $C_{j_1} = \{1,2,\ldots,i_1-1,i_1,\ldots\}$, so that $i_2 \in C_{j_1}$ and $i_2 \in C_{j_1}'$. Hence $(i_2,j_1) \in D'$ as required. $\bullet$ Now, consider the following elements of $\Sigma_N$: $$ \kappa_n^{(N)}(i) = \left\{ \begin{array}{ll} i & \mbox{ if } i<n \\ i+1 & \mbox{ if } n\leq i <N \\ n & \mbox{ if } i = N \end{array} \right. $$ Then $\kappa_1^{(N)},\ldots,\kappa_N^{(N)}$ are minimum length coset representatives of the quotient $\Sigma_N/ \Sigma_{N-1}$, and for any permutation $\pi \in \Sigma_{N-1}$, we have $\ell(\kappa_n \pi) = \ell(\kappa_n) + \ell(\pi)$. Now, starting with $D$, a northwest diagram with $\leq N$ rows, we can define a sequence of derived diagrams $D = D^{(N)}, D^{(N-1)}, \ldots, D^{(1)}$, where $D^{(i)} = (D^{(i+1)})'$ is a northwest diagram with $\leq i$ rows. Let the columns of $D^{(i)}$ be $C_1^{(i)},\ldots,C_r^{(i)}$, and define $ {\mathop{\rm gap}} (i,j) = {\mathop{\rm gap}} _ i(C_j^{(i)})$. For each $i$, we have $$ i = {\mathop{\rm gap}} (i,1) = \cdots = {\mathop{\rm gap}} (i,J_i-1) > {\mathop{\rm gap}} (i,J_i) = \cdots = {\mathop{\rm gap}} (i,r). $$ Then either $\kappa_{ {\mathop{\rm gap}} (i,j)}^{(i)}(\{1, 2, \ldots,i-1\}) \supset C_j^{(i)}$, or $C_j^{(i)} = \{1, 2,\ldots, i\}$. Notice that $J_N \leq J_{N-1} \leq \cdots$. Finally, for each column $j = 1,\ldots, r$, define $$ u_j = \kappa_{ {\mathop{\rm gap}} (N,j)}^{(N)}\, \kappa_{ {\mathop{\rm gap}} (N-1,j)}^{(N-1)} \cdots \kappa_{ {\mathop{\rm gap}} (1,j)}^{(1)}. $$ This is a reduced decomposition of $u_j$, in the sense that $\ell(u_j)$ is the sum of the lengths of the factors. Since $\kappa_i^{(i)} = \mathop{\rm id}$, and $J_N \leq J_{N-1} \leq \cdots$, each $u_j$ is an initial string of $u_{j+1}$. Thus the $u_j$ have the desired monotonicity property ($\alpha$). It only remains to show property ($\beta$): \begin{lem} For each column $C_j$ of $D$, $u_j(\{1,2,\ldots,|C_j|\}) = C_j$. \end{lem} \noindent {\bf Proof.} For each $i$, we have a $u_j^{(i)}$ associated to $D^{(i)}$, with $u_j^{(i+1)} = \kappa_{ {\mathop{\rm gap}} (i+1,j)}^{(i+1)}\, u_j^{(i)}$. For a given $i<N$, assume that $u_j^{(i)}(\{1,2,\ldots, |C_j^{(i)}|\}) \subset C_j^{(i)}$. Then I claim the same is true for $i+1$. This is clear, because $C_j^{(i+1)}$ is $C_j^{(i)}$ with some of its squares pushed down, and $\kappa_{ {\mathop{\rm gap}} (i+1,j)}^{(i+1)}$ pushes down these squares to the proper positions. In the case that $C_j^{(i+1)} = \{1,2,\ldots,l\}$, for some $l$, we have $u_j^{(i+1)} = u_j^{(i)} = \mathop{\rm id}$, and the claim is again true. The lemma now follows by induction on $i$. $\bullet$ \section{A Weyl character formula} The results of the last two sections allow us to apply the Atiyah-Bott Fixed Point Theorem to compute the characters of the Schur modules for northwest diagrams. To apply this theorem, we must examine the points of $ \FF_D $ fixed under the action of $H$, the group of diagonal matrices. We must also understand the action of $H$ on the tangent spaces at the fixed points. \subsection{Fixed points and tangent spaces} \label{fixed points} The following formula is due to Atiyah and Bott~\cite{AB} in the complex analytic case, and was extended to the algebraic case by Nielsen~\cite{N},~\cite{Iv}. \begin{thm} \label{AB thm} Let $F$ be an algebraically closed field, and suppose the torus $H = (F^{\times})^N$ acts on a smooth projective variety $X$ with isolated fixed points, and acts equivariantly on a line bundle $L \rightarrow X$. Then the character of $H$ acting on the cohomology groups of $L$ is given by: $$ \sum_i (-1)^i \mathop{\rm tr}(h \mid H^i(X,L)) = \sum_{p \ \mbox{\tiny fixed}} {\mathop{\rm tr}(h \mid L|_p ) \over \det( \mathop{\rm id} -\, h \mid T^*_p X )}, $$ where $p$ runs over the fixed points of $H$, $L|_p$ denotes the fiber of $L$ above $p$, and $T^*_p X$ is the cotangent space. $\bullet$ \end{thm} We will apply the formula for $X = \FF_D $ a smooth configuration variety, where $D = (C_1, C_2, \ldots)$ is a lexicographic northwest diagram with $\leq N$ rows and $ \widehat{D} = D$. \vspace{1em} \noindent {\bf Fixed points.} Assume for now that the columns are all distinct. Let $H = \{h=\mathop{\rm diag}(x_1,\ldots,x_N) \in GL(N) \}$ act on $\mathop{\rm Gr}(D)$ and $ \FF_D $ by the restriction of the $GL(N)$ action. Then by Proposition \ref{smooth}, we have $ \FF_D = \II_D = \{(V_C)_{C \in D} \in \mathop{\rm Gr}(D) \mid C \subset C' \Rightarrow V_C \subset V_{C'} \}$, a smooth variety. A point in $ \FF_D \subset \mathop{\rm Gr}(D) = \mathop{\rm Gr}(C_1) \times \mathop{\rm Gr}(C_2) \times \cdots$ is fixed by $H$ if and only if each component is fixed. Now, the fixed points of $H$ in $\mathop{\rm Gr}(l, F^N)$ are the coordinate planes $E_{k_1,\ldots,k_l} = {\mathop{\rm Span}} (e_{k_1}, \ldots, e_{k_l})$, where the $e_k$ are coordinate vectors in $F^N$ (c.f.~\cite{H}). For instance, the fixed points in ${\bf P}^{N-1}$ are the $N$ coordinate lines $F e_k$. We may describe the fixed points in $\mathop{\rm Gr}(C)$ as $E_S = {\mathop{\rm Span}} (e_k \mid k\in S)$, where $S \subset \{1,\ldots,N\}$ is any set with $|S| = |C|$. Hence the fixed points in $ \FF_D $ are as follows: Take a function $t$ which assigns to any column $C$ a set $t(C)\subset \{1,\ldots,N\}$ with $|\, t(C)| = |C|$, and $C \subset C' \Rightarrow t(C) \subset t(C')$. We will call such a $t$ a {\em standard column tabloid} for $D$. Then the fixed point corresponding to $t$ is $E_t = (\, E_{t(C)})_{C\in D}$. \vspace{1em} \noindent {\bf Tangent spaces at fixed points.} We may naturally identify the tangent space $T_{V_0} Gr(l, F^N) = \mathop{\rm Hom}_F(V_0, F^N/V_0)$. If $V_0$ is a fixed point (that is, a space stable under $H$), then $h \in H$ acts on a tangent vector $\phi \in Hom_F(V_0,F^N/V_0)$ by $(h\cdot \phi)(v) = h (\phi( h^{-1} v))$. For $(V_C)_{C\in D} \in \mathop{\rm Gr}(D)$, we have $T_{(V_C)} \mathop{\rm Gr}(D) = \bigoplus_{C\in D} Hom(V_C, F^N/V_C)$. Furthermore, if $(V_C)_{C\in D} \in \FF_D $, then $$ \begin{array}{cl} T_{(V_C)} \FF_D = \{ \phi = (\phi_C)_{C\in D} & \! \! \in \bigoplus_{C\in D} Hom(V_C, F^N/V_C) \mid \\ & C \subset C' \Rightarrow \phi_{C'}|_{V_C} \equiv \phi_C \mbox{ mod } V_{C'} \} \end{array} $$ (that is, the values of $\phi_{C}$ and $\phi_{C'}$ on $V_C$ agree up to translation by elements of $V_{C'}$). See~\cite{H}. For a fixed point $E_t = (E_{t(C)})$, we will find a basis for $T_{E_t}$ consisting of eigenvectors of $H$. Now, the eigenvectors in $T_{E_t} \mathop{\rm Gr}(D) = \bigoplus_{C\in D} Hom(V_C, F^N/V_C)$ are precisely $\phi^{ijC_0}= (\phi^{ijC_0}_C)_{C \in D}$, where $i,j\leq N$, $C_0$ is a fixed column of $D$, and $\phi^{ijC_0}_C(e_l) \stackrel{\rm def}{=} \delta_{C_0, C} \delta_{il} e_j$ ($\delta$ being the Kronecker delta). The eigenvalue is $$ h \cdot \phi^{ijC_0} = \mathop{\rm diag}(x_1, \ldots x_N) \cdot \phi^{ijC_0} = x_i^{-1} x_j \ \phi^{ijC_0}. $$ To obtain eigenvectors of $T_{E_t} \FF_D $, we must impose the compatibility conditions. An eigenvector $\phi$ with eigenvalue $x_i^{-1} x_j$ must be a linear combination $\phi = \sum_{C \in D} a_C \phi^{ijC}$ with $a_C \in F$. By the compatibility, we have that $$ C\subset C',\, i\in t(C),\, j\not\in t(C')\ \ \Rightarrow \ \ a_C = a_{C'} . $$ We wish to find the number $d_{ij}$ of linearly independent solutions of this condition for $a_C$. Given a poset with a relation $\subset$, define its {\em connected components} as the equivalence classes generated by the elementary relations $x \sim y$ for $x \subset y$. Now for a given $i,j$ consider the poset whose elements are those columns $C$ of $D$ such that $i\in t(C)$, $j \not\in t(C)$, with the relation of ordinary inclusion. Then $d_{ij}$ is the number of components of this poset. Note that the eigenvectors for all the eigenvalues span the tangent space. Thus $$ \det(\, \mathop{\rm id} - h \ | \ T^*_{E_t}) = \prod_{i\neq j} (1 - x_i x_j^{-1})^{d_{ij}(t)}. $$ \vspace{1em} \noindent {\bf Bundle fibers above fixed points.} Finally, let us examine the line bundles $L$ on $ \FF_D $ obtained by giving each column $C$ a multiplicity $m(C) \geq 0$. If $m(C) > 0$ for all columns $C$ of $D$, then $L \cong {\cal L} _{D'}$ for the diagram $D'$ with the same columns as $D$, each repeated $m(C)$ times. If some of the $m(C) = 0$, then $L$ is the {\em pullback} of $ {\cal L} _{D'}$ for the diagram $D'$ with the same columns as $D$, each taken $m(C)$ times, where 0 times means deleting the column. In the second case, $L$ is effective, but not ample. It follows easily from the definition that $$ \mathop{\rm tr}( h \mid L|_{E_t} ) = x_1^{- {\mathop{\rm wt}} _1(t)} \cdots x_N^{- {\mathop{\rm wt}} _N(t)}, $$ where $$ {\mathop{\rm wt}} _ i(t) = \sum_{C \atop i \in t(C)} m(C) . $$ Hence we obtain: $$ \sum_i (-1)^i \mathop{\rm tr}(h \mid H^i( \FF_D , \LL_D )) = \sum_{t} { \prod_i x_i^{- {\mathop{\rm wt}} _ i(t)} \over \prod_{i\neq j} (1-x_i x_j^{-1})^{d_{ij}(t)} }, $$ where $t$ runs over the standard column tabloids of $D$. \subsection{The character formula} \label{character formula} We summarize in combinatorial language the implications of the previous section. We think of a diagram $D$ as a list of columns $C_1, C_2, \ldots \subset {\bf N}$, possibly with repeated columns. Given a diagram $D$, the {\em blowup diagram} $ \widehat{D} $ is the diagram whose columns consist of all the columns of $D$ and all possible intersections of these columns. We will call the columns which we add to $D$ to get $ \widehat{D} $ the {\em phantom columns}. We may define a {\em standard column tabloid} for the diagram $ \widehat{D} $\, with respect to $GL(N)$, to be a filling (i.e. labeling) of the squares of $ \widehat{D} $ by integers in $\{1,\ldots,N\}$, such that:\\ (i) the integers in each column are strictly increasing, and\\ (ii) if there is an inclusion $C \subset C'$ between two columns, then all the numbers in the filling of $C$ also appear in the filling of $C'$. Given a tabloid $t$ for $ \widehat{D} $, define integers $ {\mathop{\rm wt}} _ i(t)$ to be the number of times $i$ appears in the filling, but {\em not counting i's which appear in the phantom columns}. Also define integers $d_{ij}(t)$ to be the number of connected components of the following graph: the vertices are columns $C$ of $ \widehat{D} $ such that $i$ appears in the filling of $C$, but $j$ does not; the edges are $(C,C')$ such that $C \subset C'$ or $C' \subset C$. (An empty graph has zero components.) Recall that a diagram $D$ is {\em northwest} if $ i \in C_j,\ i' \in C_{j'} \ \Rightarrow \min(i,i') \in C_{\min(j,j')} $. The following theorem applies without change to {\em northeast} diagrams and any other diagrams obtainable from northwest ones by rearranging the order of the rows and the order of the columns. Also, we can combine it with Theorem \ref{complement thm} to compute the character for the complement of a northwest diagram in an $N \times r$ rectangle. Denote a diagonal matrix by $h = \mathop{\rm diag}(x_1,\ldots,x_N).$ \begin{thm} Suppose $D$ is a northwest diagram with $\leq N$ rows, and $F$ an algebraically closed field. Then:\\ (a) The character of the Weyl module $W_D$ (for $GL(N,F)$) is given by $$ {\mathop{\rm char}} _ {W_D}(h) = \sum_{t} { \prod_i x_i^{- {\mathop{\rm wt}} _ i(t)} \over \prod_{i\neq j} (1-x_i x_j^{-1})^{d_{ij}(t)} }, $$ where $t$ runs over the standard tabloids for $ \widehat{D} $.\\ (b) For $F$ of characteristic zero, the character of the Schur module $S_D$ (for $GL(N,F)$) is given by $$ {\mathop{\rm char}} _ {S_D}(h) = \sum_{t} { \prod_i x_i^{ {\mathop{\rm wt}} _ i(t)} \over \prod_{i\neq j} (1-x_i^{-1} x_j)^{d_{ij}(t)} }, $$ where $t$ runs over the standard tabloids for $ \widehat{D} $. \end{thm} \vspace{1em} \noindent {\bf Example.} Consider the following diagram and some of its standard tabloids for $N = 3$: $$ D = \begin{array}{ccc} \Box & \Box & \\ & \Box & \Box \end{array} \ \ \ \ t_1 = \begin{array}{ccc} 1 & 1 & \\ & 2 & 2 \end{array} \ \ \ \ t_2 = \begin{array}{ccc} 1 & 1 & \\ & 2 & 1 \end{array} \ \ \ \ t_3 = \begin{array}{ccc} 3 & 2 & \\ & 3 & 3 \end{array} $$ The tabloid $t_1$ has $d_{12} = d_{13} = d_{21} = d_{23} = 1$, $d_{31} = d_{32} = 0$, and $t_2$ has $d_{12} = 2$, $d_{13} = d_{23} = 1$, $d_{21} = d_{31} = d_{32} = 0$. The other standard tabloids can be obtained from $t_1$ and $t_2$ by applying a permutation of $\{1,2,3\}$ to the entries, and rearranging the entries in the middle column to make them increasing. For instance, $t_3 = \pi t_2$, where $\pi$ is the transposition $(13)$. Note that the standard tabloids are standard tableaux in the usual sense: they are fillings with the columns strictly increasing, and the rows non-increasing. This is true in general when $ \widehat{D} $ is a skew diagram with no repeated columns, though not all the standard tableaux are obtained in this way. Applying our formula we find that $ {\mathop{\rm char}} \, S_D = s_{(3,1,0)} + s_{(2,2,0)}$, where $s_{( \lambda _1, \lambda _2, \lambda _3)}$ is a classical Schur function, the character of an irreducible Schur module. Since $D$ is a skew diagram, we could have obtained this result using the Littlewood-Richardson Rule. It should be possible to prove this rule using the present methods. $\bullet$ \vspace{1em} \noindent {\bf Proof of the Theorem.} (i) Consider the map $ {\mathop{\rm pr}} : { \cal F } _{ \widehat{D} } \rightarrow \FF_D $, and the pullback line bundle $ {\mathop{\rm pr}} ^* \LL_D $. This is the bundle on $ { \cal F } _{ \widehat{D} }$ corresponding to giving the phantom columns $C$ of $ \widehat{D} $ multiplicity $m_C = 0$. Let RHS denote the right hand side of our formula in (a). Then by the analysis of Section \ref{fixed points}, RHS is equal to the right hand side of the Atiyah-Bott formula (Theorem \ref{AB thm}) for $X = { \cal F } _{ \widehat{D} }$, $L = {\mathop{\rm pr}} ^* \LL_D $. Thus $$ \mbox{RHS} = {\mathop{\rm char}} \sum_i (-1)^i H^i( { \cal F } _{ \widehat{D} }, {\mathop{\rm pr}} ^* \LL_D ) . $$ (ii) By Proposition \ref{rational sing}, we have $ H^i( { \cal F } _{ \widehat{D} }, {\mathop{\rm pr}} ^* \LL_D ) = 0 $ for $i>0$, and $ H^0( { \cal F } _{ \widehat{D} }, {\mathop{\rm pr}} ^* \LL_D ) = H^0( \FF_D , \LL_D )$. Thus RHS = $ {\mathop{\rm char}} H^0( \FF_D , \LL_D )$. \\ (iii) By Proposition \ref{config split}, the restriction of global sections of $ \LL_D $ from $\mathop{\rm Gr}(D)$ to $ \FF_D $ is surjective, and we have \begin{eqnarray*} \mbox{RHS} & = & {\mathop{\rm char}} H^0( \FF_D , \LL_D ) \\ & = & {\mathop{\rm char}} \mathop{\rm Im}\left( \mathop{\rm rest}\mbox{} : H^0(\mathop{\rm Gr}(D), \LL_D ) \rightarrow H^0( \FF_D , \LL_D ) \right) \\ & = & {\mathop{\rm char}} W_D. \end{eqnarray*} The last equality holds by Proposition \ref{weyl mod is image}, and we have proved (a). Then (b) follows because $S_D = (W_D)^*$. $\bullet$ \subsection{Betti numbers} In this section, we compute the betti numbers of the smooth configuration varieties of Section \ref{fixed points}. \begin{prop}{(Bialynicki-Birula~\cite{B})} Let $X$ be a smooth projective variety over an algebraically closed field $F$, acted on by the one-dimensional torus $F^{\times}$ with isolated fixed points. Then there is a decomposition $$ X = \coprod_{p \ \mbox{\rm \tiny fixed}} X_p , $$ where the $X_p$ are disjoint, locally closed, $H$-invariant subvarieties, each isomorphic to an affine space $X_p \cong \mbox{\rm \bf A}^{d^+\! (p)}$. The dimensions $d^+(p)$ are given as follows. Let the tangent space $T_p X \cong \bigoplus_{n \in {\bf Z}} a_n(p) F_n$, where $a_n(p) \in {\bf N}$ and $F_n$ is the one-dimensional representation of $F^{\times}$ for which the group element $t \in F^{\times}$ acts as the scalar $t^n$. Then $$ d^+\!(p) = \sum_{n > 0} a_n(p) . $$ $\bullet$ \end{prop} Over ${\bf C}$, the above proposition does not quite give a CW decomposition for $X$, since the boundaries of the cells need not lie in cells of lower dimension. Nevertheless, $\dim_{{\bf R}} \partial X_p \leq \dim_{{\bf R}} X_p - 2$, and this is enough to fix the betti numbers $\beta_i = \dim_{{\bf R}} H^{i}(X, {\bf R})$:\ \ $\beta_{2i} = \#\{p \mid d^+(p) = i\}$, and $\beta_{2 i + 1} = 0$. Now, in our case consider the spaces $X = \FF_D $ of Section \ref{fixed points}, acted on by the $N$-dimensional torus $H$. Consider the embedding $$ \begin{array}{cccc} \check{\rho} : & F^{\times} & \rightarrow & H \\ & t & \mapsto & \mathop{\rm diag}(t^{N-1}, t^{N-2}, \ldots, t, 1), \end{array} $$ corresponding to the coweight $\check{\rho} = $ half sum of the positive coroots. Then $\check{\rho}(F^{\times})$ has the same (isolated) fixed points as $H$, since none of the eigenvectors of $H$ on $T_p X$ is fixed by $\check{\rho}(F^{\times})$. (I.e., $(\alpha, \check{\rho}) \neq 0$ for any root $\alpha$.) Also, a given eigenvector of weight $x_i x_j^{-1}$ is of positive $\check{\rho}(F^{\times})$ weight exactly when $i < j$. Thus, for a fixed point (standard tabloid) $t$ of $D$, define $$ d^+\! (t) = \sum_{i < j} d_{ij}(t) . $$ We then have the \begin{prop} Suppose $F = {\bf C}$, and $D$ is a northwest diagram with $\leq N$ rows and $ \widehat{D} = D$. Then the betti numbers $$ \beta_{2i} = \#\{t \mid d^+\! (t) = i\}, \ \ \ \beta_{2i+1} = 0 , $$ and the Poincare polynomial $$ P(x, \FF_D ) \stackrel{\rm def}{=} \sum_i \beta_i x^i = \sum_t x^{2 d^+\! (t)} , $$ where $t$ runs over the standard tabloids of $D$. $\bullet$ \end{prop} In fact, our proof shows the above propostion for a broader class of spaces. Suppose $D = (C_1, C_2, \ldots)$ is an arbitrary diagram such that the variety $$ \mbox{\rm Inc}_D \stackrel{\rm def}{=} \{ (V_C)_{C \in D} \in \mathop{\rm Gr}(D) \mid C \subset C' \Rightarrow V_C \subset V_{C'} \} $$ is smooth. Then the proposition holds with $ \FF_D $ replaced by $\mbox{\rm Inc}_D$.
1995-07-20T18:16:12
9411
alg-geom/9411005
en
https://arxiv.org/abs/alg-geom/9411005
[ "alg-geom", "math.AG" ]
alg-geom/9411005
Stein Arild Stromme
G. Ellingsrud and S. A. Str{\o}mme
Bott's formula and enumerative geometry
22 pages, amslatex 1.1 The paper is a considerably expanded version of our previous eprint alg-geom/9409006 which had the title "Counting twisted cubics on general complete intersections"
null
null
null
null
We outline a strategy for computing intersection numbers on smooth varieties with torus actions using a residue formula of Bott. As an example, Gromov-Witten numbers of twisted cubic and elliptic quartic curves on some general complete intersection in projective space are computed. The results are consistent with predictions made from mirror symmetry computations. We also compute degrees of some loci in the linear system of plane curves of degrees less than 10, like those corresponding to sums of powers of linear forms, and curves carrying inscribed polygons.
[ { "version": "v1", "created": "Thu, 10 Nov 1994 13:15:01 GMT" } ]
2008-02-03T00:00:00
[ [ "Ellingsrud", "G.", "" ], [ "Strømme", "S. A.", "" ] ]
alg-geom
\section{Introduction} One way to approach enumerative problems is to find a suitable complete parameter space for the objects that one wants to count, and express the locus of objects satisfying given conditions as a certain zero-cycle on the parameter space. For this method to yield an explicit numerical answer, one needs in particular to be able to evaluate the degree of a given zerodimensional cycle class. This is possible in principle whenever the numerical intersection ring (cycles modulo numerical equivalence) of the parameter space is known, say in terms of generators and relations. Many parameter spaces carry natural actions of algebraic tori, in particular those coming from projective enumerative problems. In 1967, Bott \cite{Bott-1} gave a residue formula that allows one to express the degree of certain zero-cycles on a smooth complete variety with an action of an algebraic torus in terms of local contributions supported on the components of the fixpoint set. These components tend to have much simpler structure than the whole space; indeed, in many interesting cases, including all the examples of the present paper, the fixpoints are actually isolated. We show in this note how Bott's formula can be effectively used to attack some enumerative problems, even in cases where the rational cohomology ring structure of the parameter space is not known. Our first set of applications is the computation of the numbers of twisted cubic curves (theorems \ref{main} and \ref{main2}) and elliptic quartic curves (\thmref{main3}) contained in a general complete intersection and satisfying suitable Schubert conditions. The parameter spaces in question are suitable components of the Hilbert scheme parameterizing these curves. These components are smooth, by the work of Piene and Schlessinger \cite{Pien-Schl} in the case of cubics, and Avritzer and Vainsencher \cite{Avri-Vain} in the case of elliptic quartics. The second set of applications is based on the Hilbert scheme of zero-dimensional subschemes of ${\bold P}^2$, which again is smooth by Fogarty's work \cite{Foga-1}. These applications deal with the degree of the variety of sums of powers of linear forms in three variables (\thmref{main4}) and Darboux curves (\thmref{main5}). \begin{ack} Part of this work was done at the Max-Planck-Institut f\"ur Mathematik in Bonn during the authors' stay there in the spring of 1993. We would like to thank the MPI for this possibility. We would also like to express our thanks to S. Katz, J. Le Potier, D. Morrison, and A. Tyurin for raising some of the problems treated here and for many stimulating discussions. \end{ack} \subsection{Main results} The first theorem deals with the number of twisted cubics on a general Calabi-Yau threefold which is a complete intersection in some projective space. There are exactly five types of such threefolds: the quintic in ${\bold P}^4$, the complete intersections $(3,3)$ and $(2,4)$ in ${\bold P}^5$, the complete intersection $(2,2,3)$ in ${\bold P}^6$ and finally $(2,2,2,2)$ in ${\bold P}^7$. \begin{thm}\label{main} For the general complete intersection Calabi-Yau threefolds, the numbers of twisted cubic curves they contain are given by the following table: \smallskip \begin{center} \begin{tabular} {|l|c|c|c|c|c|}\hline Type of complete intersection\vrule height1em width0cm depth0.4em & $5$ & $4,2$ & $3,3$ & $3,2,2$ &$2,2,2,2$\\ \hline Number of twisted cubics \vrule height1em width0cm depth0.4em &$317206375$ &$15655168$ &$6424326$ &$1611504$ &$416256$\\ \hline \end{tabular} \end{center} \end{thm} In the case of a general quintic in ${\bold P}^4$, the number of rational curves of any degree was predicted by Candelas et al.\ in \cite{Cand-Gree-Ossa-Park}, and the cubic case was verified by the authors in \cite{Elli-Stro-3}. In \cite{Libg-Teit} Libgober and Teitelbaum predicted the corresponding numbers for the other Calabi-Yau complete intersections. Our results are all in correspondence with their predictions. Greene, Morrison, and Plesser \cite{Gree-Morr-Ples} have also predicted certain numbers of rational curves on higher dimensional Calabi Yau hypersurfaces. Katz \cite{Katz-2} has verified these numbers for lines and conics for hypersurfaces of dimension up to 10. The methods of the present paper have allowed us to verify the following numbers. All but the last one, $N_3^{1,1,1,1}(8)$, have been confirmed by D.~Morrison (privat communication) to be consistent with \cite{Gree-Morr-Ples}. \begin{thm}\label{main2} For a general hypersurface $W$ of degree $n+1$ in ${\bold P}^n$ $(n\le8)$ and for a partition $\lambda=(\lambda_1\ge\dots\ge\lambda_m>0)$ of $n-4$, the number $N_3^{\lambda}(n)$ of twisted cubics on $W$ which meet $m$ general linear subspaces of codimensions $\lambda_1+1,\dots,\lambda_m+1$ respectively is given as follows: \medskip \begin{center} \begin{tabular}{|l|l|l||l|l|l|}\hline \vrule height1.2em width0cm depth0.6em$n$ & $\lambda$ & $N_3^{\lambda}(n)$& \vrule height1em width0cm depth0.4em$n$ & $\lambda$ & $N_3^{\lambda}(n)$ \\ \hline \vrule height1em width0cm depth0.4em$4$&{} &$317206375$ & \vrule height1em width0cm depth0.4em$7$&$1,1,1$ & $12197109744970010814464$ \\ \hline \vrule height1em width0cm depth0.4em$5$& $1$ & $6255156277440$ & $8$& \vrule height1em width0cm depth0.4em$4$ & $897560654227562339370036$ \\ \hline \vrule height1em width0cm depth0.4em$6$& $2$ & $30528671745480104$ &$8$& \vrule height1em width0cm depth0.4em$3,1$ & $17873898563070361396216980$ \\ \hline \vrule height1em width0cm depth0.4em$6$& $1,1$ & $222548537108926490$ & \vrule height1em width0cm depth0.4em$8$&$2,2$ & $33815935806268253433549768$ \\ \hline \vrule height1em width0cm depth0.4em$7$&$3$ & $154090254047541417984$ & \vrule height1em width0cm depth0.4em$8$&$2,1,1$ & $174633921378662035929052320$ \\ \hline \vrule height1em width0cm depth0.4em$7$&$2,1$ & $2000750410187341381632$ & \vrule height1em width0cm depth0.4em$8$&$1,1,1,1$& $957208127608222375829677128$ \\ \hline \end{tabular} \end{center} \end{thm} The number $N_3^{1,1,1,1}(8)$ is not related to mirror symmetry as far as we know; Greene et.al.~ get numbers only for partitions with at most 3 parts. Our methods also yield other numbers not predicted (so far!) by physics methods: for example, there are 1345851984605831119032336 twisted cubics contained in a general nonic hypersurface in ${\bold P}^7$ (not a Calabi-Yau manifold). A similar method can be used to compute the number of elliptic quartic curves on general Calabi-Yau complete intersections. Here are the results for some hypersurfaces, which we state without proof: \begin{thm}\label{main3} The number of quartic curves of arithmetic genus 1 on a general hypersurface of degree $n+1$ in ${\bold P}^n$ are for $4\le n\le13$ given by the following table. These curves are all smooth. \smallskip \begin{center} \begin{tabular}{|l|l|}\hline $n$\vrule height1.2em width0cm depth0.6em & Smooth elliptic quartics on a general hypersurface of degree $n+1$ in ${\bold P}^n$\\ \hline $4$&\vrule height1em width0cm depth0.4em $3718024750$\\ \hline $5$&\vrule height1em width0cm depth0.4em $387176346729900$\\ \hline $6$&\vrule height1em width0cm depth0.4em $81545482364153841075$\\ \hline $7$&\vrule height1em width0cm depth0.4em $26070644171652863075560960$\\ \hline $8$&\vrule height1em width0cm depth0.4em $12578051423036414381787519707655$\\ \hline $9$&\vrule height1em width0cm depth0.4em $8760858604226734657834823089352310000$\\ \hline $10$&\vrule height1em width0cm depth0.4em $8562156492484448592316222733927180351143552$\\ \hline $11$&\vrule height1em width0cm depth0.4em $11447911791501360069250820471811603020708611018752$\\ \hline $12$&\vrule height1em width0cm depth0.4em $20498612221082029813903827233942127541022477928303274152$\\ \hline $13$&\vrule height1em width0cm depth0.4em $48249485834889092561505032612701767175955799366431126942036480$\\ \hline \end{tabular} \end{center} \end{thm} This computation uses the description given in \cite{Avri-Vain} of the irreducible component of the Hilbert scheme of ${\bold P}^3$ parameterizing smooth elliptic quartics. This Hilbert scheme component can be constructed from the Grassmannian of pencils of quadrics by two explicit blowups with smooth centers, and one may identify the fixpoints for the natural action of a torus in a manner analogous to what we carry out for twisted cubics in this paper. For another related construction, see \cite{Meur-1}, which treats curves in a weighted projective space. The number of elliptic quartics on the general quintic threefold was predicted by Bershadsky et.al.\ \cite{Bers-Ceco-Oogu-Vafa}. Their number, 3721431625, includes singular quartics of geometric genus 1. These are all plane binodal quartics, and their number is $1185*2875=3406875$ by \cite{Vain-1}. Thus the count of \cite{Bers-Ceco-Oogu-Vafa} is compatible with the number above. Recently, Kontsevich \cite{Kont-1} has developed a technique for computing numbers of rational curves of {\em any\/} degree, using the stack of stable maps rather than the Hilbert scheme as a parameter space. He also uses Bott's formula, but things get more complicated than in the present paper because of the presence of non-isolated fixpoints in the stack of stable maps. The next theorem deals with plane curves of degree $n$ whose equation can be expressed as a sum of $r$ powers of linear forms. Let $PS(r,n)$ be the corresponding subvariety of ${\bold P}^{n(n+3)/2}$. Then $PS(r,n)$ is the $r$-th secant variety of the $n$-th Veronese imbedding of ${\bold P}^2$. Let $p(r,n)$ be the number of ways a form corresponding to a general element of $PS(r,n)$ can be written as a sum of $r$ $n$-th powers if this number is finite, and 0 otherwise. The last case occurs if and only if $\dim(PS(r,n))$ is less than the expected $3r-1$. We don't know of an example where $p(r,n)>1$ if $PS(r,n)$ is a proper subvariety. If $p(r,n)=1$, then $p(r,n')=1$ for all $n'\ge n$. It is easy to see that $p(2,n)=1$ for $n\ge3$. \begin{thm} \label{main4} Assume that $n\ge r-1$ and $2\le r\le8$. Then $p(r,n)$ times the degree of $PS(r,n)$ is $s_r(n)$, where {\allowdisplaybreaks \begin{align} 2\,s_2(n)= &\,n^4-10\,n^2+15\,n-6, \notag \\ 3!\,s_3(n)= &\,n^6-30\,n^4+45\,n^3+206\,n^2-576\,n+384,\notag \\ 4!\,s_4(n)= &\,n^8-60\,n^6+90\,n^5+1160\,n^4-3204\,n^3-5349\,n^2+26586 \,n-23760,\notag\\ 5!\,s_5(n)= &\,n^{10}-100\,n^8+150\,n^7+3680\,n^6-10260\,n^5- 52985\,n^4+\notag\\* &\,224130\,n^3+127344\,n^2-1500480\,n + 1664640,\notag\\ 6!\,s_6(n)= &\,n^{12}-150\,n^{10}+225\,n^9+8890\,n^8-25020\,n^7-244995\,n^6 +1013490\,n^5+\notag \\*&\,2681974\,n^4-17302635\,n^3+1583400\,n^2+101094660\,n -134190000,\notag\\ 7!\,s_7(n)=&\,n^{14}-210\,n^{12}+315\,n^{11}+ 18214\,n^{10}-51660\,n^9-802935\,n^8+\notag\\* &\,3318210\,n^7+17619994\,n^6-102712365\,n^5 -136396680\,n^4+\notag\\* &\,1498337820\,n^3-872582544\,n^2-7941265920\,n +12360418560,\notag\\ 8!\,s_8(n)=&\,n^{16}-280\,n^{14}+420\,n^{13}+ 33376\,n^{12}-95256\,n^{11} -2134846\,n^{10}+\notag\\* &\,8858220\,n^9+75709144\,n^8- 427552020\,n^7-1332406600\,n^6+\notag\\* &\,11132416680\,n^5+5108998089\,n^4 -145109970684\,n^3+\notag\\* &\,144763373916\,n^2 +713178632880\,n-1286736675840.\notag \end{align} } \end{thm} For example, $s_5(4)=0$; this corresponds to the classical but non-obvious fact that not all ternary quartics are sums of five fourth powers. (Those who are are called Clebsch quartics; they form a hypersurface of degree 36). Note in particular that $s_3(3)=4$. It is classically known that $PS(3,3)$ is indeed a hypersurface of degree 4, its equation is the so-called $S$-invariant \cite{Salm}. It follows that $p(3,3)=1$, and hence that $p(3,n)=1$ for $n\ge3$. Only the first few of these polynomials are reducible: $s_r(r-1)=0$ for $r\le5$, but the higher $s_r$ in the table are irreducible over ${\bold Q}$. Note that the formulas of the theorem are not valid unless $n\ge r-1$. For example, a general quintic is uniquely expressable as a sum of seven fifth powers (cfr.~the references in \cite{Muka-1}), while $s_7(5)$ is negative. The final application quite similar. A {\em Darboux curve\/} is a plane curve of degree $n$ circumscribing a complete $(n+1)$-gon (this terminology extends the one used in \cite{Bart-1}). This means that there are distinct lines $L_0,\dots,L_n$ such that $C$ contains all intersection points $L_i\cap L_j$ for $i<j$. Equivalently, there are linear forms $\ell_0,\dots,\ell_n$ such that the curve is the divisor of zeroes of the rational section $\sum_{i=0}^n \ell_i^{-1}$ of $\OP2(-1)$. Let $D(n)$ be the closure in ${\bold P}^{n(n+3)/2}$ of the locus of Darboux curves. Let $p(n)$ be the number of inscribed $(n+1)$-gons in a general Darboux curve, if finite, and 0 otherwise. \begin{thm} \label{main5} For $n=5,6,7,8,9$, the product of $p(n)$ and the degree of the Darboux locus $D(n)$ is $2540, 583020, 99951390, 16059395240, 2598958192572$, respectively. \end{thm} We have no guess as to what $p(n)$ is; it might well be 1 for $n\ge5$. It is always positive for $n\ge5$ by an argument of Barth's \cite{Bart-1}. For $n\le4$ it is 0. For $n\le3$, all curves are Darboux. For $n=4$, Darboux curves are L\"uroth quartics, and form a degree 54 hypersurface \cite{Morl,LePo,Tyur-1}. \section{Bott's formula}\label{C} Let $X$ be a smooth complete variety of dimension $n$, and assume that there is given an algebraic action of the multiplicative group ${\bold C}^*$ on $X$ such that the fixpoint set $F$ is finite. Let ${\cal E}$ be an equivariant vector bundle of rank $r$ over $X$, and let $p(c_1,\dots,c_r)$ be a weighted homogenous polynomial of degree $n$ with rational coefficients, where the variable $c_i$ has degree $i$. Bott's original formula \cite{Bott-1} expressed the degree of the zero-cycle $p(c_1({\cal E}),\dots,c_r({\cal E}))\in H^{2n}(X,{\bold Q})$ purely in terms of data given by the representations induced by ${\cal E}$ and the tangent bundle $T_X$ in the fixpoints of the action. Later, Atiyah and Bott \cite{Atiy-Bott-1} gave a more general formula, in the language of equivariant cohomology. Its usefulness in our context is mainly that it allows the input of Chern classes of several equivariant bundles at once. Without going into the theory of equivariant cohomology, we will give here an interpretation of the formula which is essentially contained in the work of Carrell and Lieberman \cite{Carr-Lieb-1,Carr-Lieb-2}. To explain this, first note that the ${\bold C}^*$ action on $X$ induces, by differentiation, a global vector field $\xi\in H^0(X,T_X)$, and furthermore, the fixpoint set $F$ is exactly the zero locus of $\xi$. Hence the Koszul complex on the map $\xi^{\vee}\:\Omega_X \to {\cal O}_X$ is a locally free resolution of ${\cal O}_F$. For $i\ge0$, denote by $B_i$ the cokernel of the Koszul map $\Omega_X^{i+1}\to \Omega_X^{i}$. It is well known that $H^j(X,\Omega_X^i)$ vanishes for $i\ne j$, see e.g. \cite{Carr-Lieb-1}. Hence there are natural exact sequences for all $i$: $$ 0 \to H^i(X,\Omega_X^i) @>p_i>> H^i(X,B_i) @>r_i>> H^{i+1}(X,B_{i+1}) \to0. $$ In particular, there are natural maps $q_i=r_{i-1}\circ\dots\circ r_0\:H^0(F,{\cal O}_F)\to H^i(X,B_i)$. \begin{defn} Let $f\: F \to {\bold C}$ be a function and $c\in H^i(X,\Omega_X^i)$ a non-zero cohomology class. We say that $f$ {\em represents\/} $c$ if $q_i(f)=p_i(c)$. \end{defn} For each $i\ge-1$, put $A_i=\ker q_{i+1}$. Then $$ 0=A_{-1}\subseteq{\bold C}=A_0\subseteq A_1\subseteq A_2\subseteq \dots \subseteq A_n = H^0(F,{\cal O}_F) $$ is a filtration by sub-vector spaces of the ring of complex-valued functions on $F$. The filtration has the property that $A_i A_j \subseteq A_{i+j}$, and the associated graded ring $\operatornamewithlimits{\oplus} A_{i}/A_{i-1}$ is naturally isomorphic to the cohomology ring $H^*(X,{\bold C})\simeq \operatornamewithlimits{\oplus} H^i(X,\Omega_X^i)$. (In \cite{Carr-Lieb-2}, the filtration is constructed as coming from one of the spectral sequences associated to hypercohomology of the Koszul complex above.) An interesting aspect of this is that cohomology classes can be represented as functions on the fixpoint set. The representation is unique up to addition of functions coming from cohomology classes of lower degree (i.e., lower codimension). Since the algebra of functions on a finite set is rather straightforward, this gives an efficient way to evaluate zero-cycles, provided that 1) we know how to describe a function representing a given class, and 2) we have an explicit formula for the composite linear map $$ \epsilon_X\: H^0({\cal O}_F) @>q_n>> H^{n}(X,\Omega_X^n) @>\text{res}_X>\simeq> {\bold C}. $$ These issues are addressed in the theorem below. Let ${\cal E}$ be an equivariant vector bundle of rank $r$ on $X$. In each fixpoint $x\in F$ the fiber of ${\cal E}$ splits as a direct sum of one-dimensional representations of ${\bold C}^*$; let $\tau_1({\cal E},x),\dots,\tau_r({\cal E},x)$ denote the corresponding weights, and for all integers $k\ge0$, let $\sigma_k({\cal E},x)\in{\bold Z}$ be the $k$-th elementary symmetric function in the $\tau_i({\cal E},x)$. \begin{thm}\label{bott1} Let the notation and terminology be as above. Then \begin{enumerate} \item The $k$-th Chern class $c_k({\cal E})\in H^k(X,\Omega_X^k)$ of ${\cal E}$ can be represented by the function $x\mapsto \sigma_k({\cal E},x)$. \item For a function $f\: F \to {\bold C}$, we have $\epsilon_X(f) = \displaystyle{\sum_{x\in F} \frac{f(x)}{\sigma_n(T_X,x)}}.$ \end{enumerate} \end{thm} \begin{pf} See \cite[equation~3.8]{Atiy-Bott-1}, and \cite{Carr-Lieb-2}. \end{pf} Note that the function $\sigma_k({\cal E},-)$ depends on the choice of a ${\bold C}^*$-linearisation of the bundle ${\cal E}$, whereas the Chern class $c_k({\cal E})$ it represents does not. \section{Twisted cubics} \label{A} Let $\operatorname{Hilb}^{3t+1}_{{\bold P}^n}$ be the Hilbert scheme parameterizing subschemes of ${\bold P}^n$ ($n\ge3$) with Hilbert polynomial $3t+1$, and let $H_n$ denote the irreducible component of $\operatorname{Hilb}^{3t+1}_{{\bold P}^n}$ containing the twisted cubic curves. Recall from \cite{Pien-Schl} that $H_3$ is smooth and projective of dimension $12$. Any curve corresponding to a point of $H_n$ spans a unique 3-space, hence $H_n$ admits a fibration \begin{equation} \label{fibration} \Phi\: H_n\to G(3,n) \end{equation} over the Grassmannian of 3-planes in ${\bold P}^n$, with fiber $H_3$. It follows that $H_n$ is smooth and projective of dimension $4n$. There is a universal subscheme ${\cal C}\subset H_n\times {\bold P}^n$. For a closed point $x\in H_n$, we denote by $C_x$ the corresponding cubic curve, i.e., the fiber of ${\cal C}$ over $x$. Also, let ${\cal I}_x\subseteq\OP{n}$ be its ideal sheaf. By the classification of the curves of $H_n$ (see below), it is easy to see that \begin{equation} \label{H1} H^1({\bold P}^n,{\cal I}_x(d))=H^1(C_x,{\cal O}_{C_x}(d))=0\quad \text{for all $d\ge1$ and for all $x\in H_n$.} \end{equation} For a subscheme $W\subseteq {\bold P}^n$, denote by $H_W\subseteq H_n$ the closed subscheme parameterizing twisted cubics contained in $W$. There is a natural scheme structure on $H_W$ as the intersection of $H_n$ with the Hilbert scheme of $W$. If $C_x\subseteq W$ is a Cohen-Macaulay twisted cubic, then locally at $x\in H_n$, the scheme $H_W$ is simply the Hilbert scheme of $W$. Our goal is to compute the cycle class of $H_W$ in $A^*(H_n)$ in the case that $W$ is a general complete intersection in ${\bold P}^n$. In particular, we want its cardinality if it is finite, and its Gromov-Witten invariants (see below) if it has positive dimension. For each integer $d$ we define a sheaf ${\cal E}_{d}$ on $H_n$ by \begin{equation} \label{defE} {\cal E}_{d}=p_{1*}({\cal O}_{\cal C}\* p_{2}^* \OP{n}(d)), \end{equation} where $p_1$ and $p_2$ are the two projections of $H_n\times {\bold P}^n$. If $d\ge1$, then the vanishing of the first cohomology groups \eqref{H1} implies by standard base change theory \cite{AG} that ${\cal E}_{d}$ is locally free of rank $3d+1$, and moreover that there are surjections $\rho\:H^0(\OP{n}(d))_{H_n}\to{\cal E}_{d}$ of vector bundles on $H_n$. In particular, for all $x\in H_n$, there is a natural isomorphism \begin{equation} {\cal E}_{d}(x) @>\simeq>> H^0(C_x,{\cal O}_{C_x}(d)). \label{basechange} \end{equation} A homogenous form $F\in H^0(\OP{n}(d))$ induces a global section $\rho(F)$ of ${\cal E}_{d}$ over $H_n$, and the evaluation of this section at a point $x$ corresponds under the identification (\ref{basechange}) to the restriction of $F$ to the curve $C_x$. Hence the zero locus of $\rho(F)$ corresponds to the set of curves $C_x$ contained in the hypersurface $V(F)$. More generally, in the case of an intersection $V(F_1,\dots,F_p)$ in ${\bold P}^n$ of $p$ hypersurfaces, the section $(\rho(F_1),\dots,\rho(F_p))$ of ${\cal E}={\cal E}_{d_1}\oplus\dots\oplus{\cal E}_{d_p}$ vanishes exactly on the points corresponding to twisted cubics contained in $V(F_1,\dots,F_p)$. \begin{prop} \label{propA} Let $W\subseteq{\bold P}^n$ be the complete intersection of $p$ general hypersurfaces in ${\bold P}^n$ of degrees $d_1,\ldots,d_p$ respectively. Assume that $\sum_i(3d_i+1)=4n$. Then the number of twisted cubic curves contained in $W$ is finite and equals $$\int_{H_n}c_{4n}({\cal E}_{d_1}\oplus\cdots\oplus{\cal E}_{d_p}).$$ These cubics are all smooth. \end{prop} \begin{pf} By the considerations above, the bundle ${\cal E}={\cal E}_{d_1}\oplus\cdots\oplus{\cal E}_{d_p}$ is a quotient bundle of the trivial bundle $\operatornamewithlimits{\oplus} H^0(\OP{n}(d_i))_{H_n}$. Hence Kleiman's Bertini theorem \cite{Klei-1} implies that the zero scheme of the section $(\rho(F_1),\ldots,\rho(F_p))$ is nonsingular and of codimension $\operatorname{rank}({\cal E})$. Since $\operatorname{rank}({\cal E})=\dim(H_n)$, the number of points is finite and given by the top Chern class. \end{pf} \subsection{Gromov-Witten numbers}\label{B} More generally, assume that $W$ is as in \propref{propA}, except that we only assume an inequality $\sum_i(3d_i+1)\le4n$ instead of the equality. The top Chern class of ${\cal E}$ still represents the locus $H_n(W)$ of twisted cubics contained in $W$, although there are infinitely many of them if the inequality is strict. One may assign finite numbers to this family by imposing Schubert conditions. For this purpose say that a {\em Schubert condition\/} on a curve is the condition that it intersect a given linear subspace of ${\bold P}^n$. If the subspace has codimension $c+1$, then the corresponding Schubert condition is of codimension $c$ (corresponding to the class $\gamma_c$ below). \begin{defn} Let $W$ be a general complete intersection in ${\bold P}^n$ of $p$ hypersurfaces of degrees $d_1,\dots,d_p$ respectively. Let $\lambda=(\lambda_1 \ge \lambda_2 \ge \dots \ge \lambda_m > 0)$ be a partition of $4n - \sum_{i=1}^p(3d_i+1)$, and let $P_1,\dots,P_m$ be general linear subspaces such that $\operatorname{codim} P_i = \lambda_i+1$. The number of twisted cubics on $W$ meeting all the $P_i$ is called the $\lambda$-th {\em Gromov-Witten number} of the family of twisted cubic curves on $W$, and is denoted by $N_3^\lambda(W)$. \end{defn} \begin{rem} This is a slight variation on the definition used in \cite{Katz-2}, and differs from that by a factor of 3 (resp.~9) for partitions with 2 (resp.~1) parts. In \cite{Katz-2} only partitions of length at most three are considered, as these numbers are the ones that have been predicted by mirror symmetry computations (when $W$ is Calabi-Yau). We have used the term Gromov-Witten number rather than Gromov-Witten invariant, as the latter term is now being used in a more sophisticated sense \cite{Kont-Mani-1}. \end{rem} Let $h$ denote the hyperplane class of ${\bold P}^n$ as well as its pullback to $H_n\times{\bold P}^n$, and let $[{\cal C}]\in A^*(H_n\times{\bold P}^n)$ be the cycle class of the universal curve ${\cal C}$. If $P\subseteq{\bold P}^n$ is a linear subspace of codimension $c+1\ge2$, then ${\cal C}\cap H_n\times P$ projects birationally to its image under the first projection, which is the locus of curves meeting $P$. Hence the class of the locus of curves meeting $P$ is ${p_1}_* (h^{c+1}[{\cal C}])$. For simplicity, we give this class a special notation: \begin{notation} For a natural number $c$, let $\gamma_c = {p_1}_* (h^{c+1}[{\cal C}]) \in A^c(H_n)$. \end{notation} \begin{prop} \label{propB} Let $W\subseteq{\bold P}^n$ be the complete intersection of $p$ general hypersurfaces in ${\bold P}^n$ of degrees $d_1,\ldots,d_p$ respectively. Assume that $\sum_i(3d_i+1)\le 4n$, and let $\lambda=(\lambda_1 \ge \lambda_2 \ge \dots \ge \lambda_m > 0)$ be a partition of $4n - \sum_{i=1}^p(3d_i+1)$. Then $$ N_3^\lambda(W)= \int_{H_n} c_{{\text{top}}}({\cal E}_{d_1}\oplus\cdots\oplus{\cal E}_{d_p})\cdot \prod_{i=1}^m\gamma_{\lambda_i}. $$ Furthermore, if $P_1,\dots,P_m$ are general linear subspaces such that $\operatorname{codim} P_i = \lambda_i+1$, then the $N_3^\lambda(W)$ twisted cubics on $W$ which meets each $P_i$ are all smooth. \end{prop} \begin{pf} Similar to the proof of \propref{propA}. \end{pf} For later use, we want to express the classes $\gamma_i$ in terms of Chern classes of the bundles ${\cal E}_d$. \begin{prop} \label{gammaformler} Let $a_i=c_i({\cal E}_1)$, $b_i=c_i({\cal E}_2)$, $c_i=c_i({\cal E}_3)$, and $d_i=c_i({\cal E}_4)$. Then we have the following formulas for the $\gamma_c$: {\allowdisplaybreaks \begin{align} \gamma_0 =&\, 3\notag \\ \gamma_1 =&\, 5a_1-14b_1+13c_1-4d_1\notag\\ \gamma_2 =&\, 3a_1^2-9a_1b_1+9a_1c_1-3a_1d_1-3b_1^2+9b_1c_1\notag\\* & -3b_1d_1-6c_1^2+3c_1d_1+a_2-3b_2+3c_2-d_2\notag\\ \gamma_3 =&\, 3a_1^3-9a_1^2b_1+9a_1^2c_1-3a_1^2d_1-3a_1b_1^2+9a_1b_1c_1\notag\\* & -3a_1b_1d_1-6a_1c_1^2+3a_1c_1d_1-4a_1a_2-3a_1b_2+3a_1c_2-a_1d_2\notag\\* &+ 14a_2b_1-13a_2c_1+4a_2d_1+3a_3\notag\\ \gamma_4 =&\, 3a_1^4-9a_1^3b_1+9a_1^3c_1-3a_1^3d_1-3a_1^2b_1^2\notag\\* & +9a_1^2b_1c_1-3a_1^2b_1d_1-6a_1^2c_1^2+3a_1^2c_1d_1-7a_1^2a_2\notag\\* & -3a_1^2b_2+3a_1^2c_2-a_1^2d_2+23a_1a_2b_1-22a_1a_2c_1+7a_1a_2d_1\notag\\* & +3a_2b_1^2-9a_2b_1c_1+3a_2b_1d_1+6a_2c_1^2-3a_2c_1d_1+8a_1a_3\notag\\* & -a_2^2+3a_2b_2-3a_2c_2+a_2d_2-14a_3b_1+13a_3c_1-4a_3d_1-3a_4\notag \end{align} } \end{prop} \begin{pf} Let $\pi\: B={\bold P}({\cal E}_1) \to H_n$. The natural surjection $\rho\:H^0(\OP{n}(1))_{H_n}\to{\cal E}_{1}$ induces a closed imbedding $B\subseteq H_n\times{\bold P}^n$ over $H_n$. Over a closed point $x$ of $H_n$, the fiber of $B$ is just the ${\bold P}^3$ spanned by $C_x$. So the universal curve ${\cal C}$ is actually a codimension 2 subscheme of $B$. It follows by the projection formula that $$ \gamma_c = {\pi}_* (\tau^{c+1}[{\cal C}]_B) \in A^c(H_n), $$ where $[{\cal C}]_B$ denotes the class of ${\cal C}$ in $A^2(B)$, and $\tau\in A^1(B)$ is the first Chern class of the tautological quotient linebundle on $B$. The formulas of the proposition now follow by straightforward computation (for example using \cite{schubert}) from the next lemma. \end{pf} \begin{lem} The class of ${\cal C}$ in $B$ is \begin{equation*} \begin{aligned} [{\cal C}]_B=&\,3\tau^2+(-4d_1+2a_1-14b_1+13c_1)\tau\\& +3c_1d_1+4a_2-3b_2+3c_2-d_2-2a_1^2\\& +5a_1b_1-4a_1c_1+a_1d_1-3b_1^2+9b_1c_1-3b_1d_1-6c_1^2 \end{aligned} \end{equation*} \end{lem} \begin{pf} Let $i\: {\cal C}\to B$ be the inclusion. Then $[{\cal C}]_B$ equals the degree 2 part of the Chern character of the ${\cal O}_B$-module $i_*{\cal O}_{\cal C}(\ell\tau)$, for any integer $\ell$. For $\ell=4$, there is a canonical Beilinson type resolution of $i_*{\cal O}_{\cal C}(4\tau)$: \begin{equation} 0\to\pi^*{\cal E}_1\*\Omega^3_{B/H_n}(3\tau)\to \pi^*{\cal E}_2\*\Omega^2_{B/H_n}(2\tau)\to \pi^*{\cal E}_3\*\Omega^1_{B/H_n}(\tau)\to \pi^*{\cal E}_4 \label{beilinson} \end{equation} Using this it is a straightforward exercise (again using \cite{schubert}) to compute the Chern character of $i_*{\cal O}_{\cal C}$ in terms of $\tau$ and the Chern classes of the ${\cal E}_n$. \end{pf} \subsection{Coarse classification of twisted cubics}\label{D} We divide the curves $C_x$ for $x\in H_3$ into two groups, according to whether they are Cohen-Macaulay or not. A locally Cohen-Macaulay twisted cubic curve $C_x$ is also arithmetically Cohen-Macaulay, and its ideal is given by the vanishing of the $2\times2$ minors of a $3\x2$ matrix $\alpha$ with linear coefficients. There is a resolution of ${\cal O}_{C_x}$: \begin{equation} K^{\bullet}:\quad 0 @>>> F\*{\cal O}_{{\bold P}^3}(-3) @>\alpha>> E\*{\cal O}_{{\bold P}^3}(-2) @>\wedge^2\alpha^t>> \OP3, \end{equation} where $F$ and $E$ are vector spaces of dimensions 2 and 3 respectively. Intrinsically, \begin{align} E=&H^0({\bold P}^3,{\cal I}_{x}(2))\subseteq H^0(\OP3(2))\label{EE}\\ F=&\operatorname{Ker}(E\* H^0(\OP3(1)) @>\text{mult}>> H^0(\OP3(3))).\label{FF} \end{align} \begin{lem} \label{CMlem} Let $C_x$ be Cohen-Macaulay, and let $E$ and $F$ be as above. Then there is a functorial exact sequence: \begin{equation} 0 @>>> {\bold C} \to \operatorname{End}(F)\operatornamewithlimits{\oplus}\operatorname{End}(E) \to \operatorname{Hom}(F,E)\* H^0(\OP3(1)) \to T_{H_3}(x) \to 0 \end{equation} \end{lem} \begin{pf} Recall the canonical identification $T_{H_3}(x)=\operatorname{Hom}_{{\bold P}^3}({\cal I}_{x},{\cal O}_{C_x})$. The sequence now follows from consideration of the total complex associated to the double complex $\operatorname{Hom}_{{\bold P}^3}(K^{\bullet},K^{\bullet})$. \end{pf} Next let us consider the curves $C_x$ for $x\in H_3$ which are not Cohen-Macaulay. By \cite{Pien-Schl}, these are projectively equivalent to a curve with ideal generated by the net of quadrics $x_0(x_0,x_1,x_2)$ plus a cubic form $q$, which can be taken to be of the form $q=Ax_1^2+Bx_1x_2+Cx_2^2$, with $A$, $B$, and $C$ linear forms in ${\bold C}[x_1,x_2,x_3]$. If we furthermore impose the conditions that $B$ is a scalar multiple of $x_3$, then the cubic $q$ is unique up to scalar. (See \cite{Elli-Stro-3}). Let $Y\subseteq H_3$ be the locus of non-Cohen-Macaulay curves, and denote by $I$ the 5-dimensional incidence correspondence $\{(p,H)\in {\bold P}^3\times{{\bold P}^3}^* \mid p\in H \}$. By the above, the quadratic part of ${\cal I}_x$ for $x\in Y$ gives rise to a point of $I$. This gives a morphism \begin{equation} g\: Y\to I, \label{def:g} \end{equation} and again from the above it is clear that this makes $Y$ a ${\bold P}^6$-bundle over $I$. Hence $Y$ is a divisor on $H_3$, and it is clear how to compute the tangent spaces $T_Y(x)$. To get hold of $T_{H_3}(x)$, we need to identify the normal direction of $Y$ in $H_3$. For this, let $C_x$ be the curve above, and consider the family $C_t$ of Cohen-Macaulay curves given for $t\ne0$ by the matrix \begin{equation} \alpha_t =\begin{pmatrix} 0 & -x_0\\ x_0 & 0\\ -x_1 & x_2 \end{pmatrix} + t\begin{pmatrix} C & B\\ 0 & A\\ 0 & 0 \end{pmatrix} \end{equation} Then \begin{equation} \det\left(\begin{array}{c|c} & x_1\\ \alpha_t & -x_2\\ & 0 \end{array}\right) = t(Ax_1^2+Bx_1x_2+Cx_2^2) = tq, \end{equation} which implies that $\lim_{t\to0}C_t=C_x$. The tangent vector $\xi\in T_{H_3}(x)=\operatorname{Hom}({\cal I}_x,{\cal O}_{C_x})$ corresponding to this one-parameter family has this effect on the quadratic equations: \begin{equation} \label{xi} \xi(x_0^2)=Bx_0,\quad \xi(x_0x_1)=-Bx_1-Cx_2, \quad \xi(x_0x_2)=Ax_1. \end{equation} In particular, $\xi\ne0$. (This argument actually shows that the blowup of the space of determinantal nets of quadrics along the locus of degenerate nets maps isomorphically onto $H_3$, cfr.~\cite{Elli-Pien-Stro}). \subsection{The torus action}\label{E} Consider the natural action of $\operatorname{GL}(n+1)$ on ${\bold P}^n$. It induces an action on $H_n$ and on the bundles ${\cal E}_d$ for $d\ge1$. Let $T\subseteq\operatorname{GL}(n+1)$ be a maximal torus, and let $(x_0,\ldots,x_n)$ be homogeneous coordinates on ${\bold P}^n$ in which the action of $T$ is diagonal. A point $x\in H_n$ is fixed by $T$ if and only if the corresponding curve $C_x$ is invariant under $T$, i.e., $t(C_x)=C_x$ for any $t\in T$. This is easily seen to be the case if and only if the graded ideal of $C_x$ is generated by {\em monomials\/} in the $x_i$. In particular, the fixpoints are isolated. We will identify all the fixpoints $x\in H_n$, and for each of them we will compute the representation on the tangent space $T_{H_n}(x)$. The tangent space of the Hilbert scheme is $\operatorname{Hom}({\cal I}_x,{\cal O}_x)$, but special care must be taken at the points where $H_n$ meets another component of the Hilbert scheme. At these points, $T_{H_n}(x)$ is a proper subspace of $\operatorname{Hom}({\cal I}_x,{\cal O}_x)$. By the choice of the coordinates $(x_0,\ldots,x_n)$, there are characters $\lambda_i$ on $T$ such that for any $t\in T$ we have $t.x_i=\lambda^{}_i(t)x_i$. The characters $\lambda^{}_i$ generate the representation ring of $T$, i.e., if $W$ is any finite dimensional representation of $T$ we may, by a slight abuse of notation, write $W=\sum a_{p_0,\dots,p_n}\lambda_0^{p_0}\lambda_1^{p_1} \cdots\lambda_n^{p_n}$, where the $p_i$ and $a_{p_0,\dots,p_n}$ are integers. Recall \eqref{fibration} the morphism $\Phi:H_n\to G(3,n)$ which maps a point $x\in H_n$ to the 3-space spanned by the corresponding curve $C_x$. This morphism clearly is $\operatorname{GL}(n+1)$ equivariant, and its fibers are all isomorphic to $H_3$. If $C_x$ is invariant under $T$, then so is its linear span. Hence up to a permutation of the variables, we may assume that it is given by the equations $x_4=\dots=x_n=0$, so that $x_0,\dots,x_3$ are coordinates on the ${\bold P}^3\subseteq{\bold P}^n$ corresponding to $\Phi(x)$. The torus $T$ acts on ${\bold P}^3$ through the four-dimensional quotient torus $T_3$ of $T$ with character group spanned by $\lambda_0,\dots,\lambda_3$. The tangent space of $H_n$ at a fixpoint $x$ decomposes as a direct sum \begin{equation} \label{dsum} T_{H_n}(x)=T_{H_3}(x)\oplus T_{G(3,n)}(\Phi(x)), \end{equation} and it is well known that \begin{equation}\label{tangrass} T_{G(3,n)}(\Phi(x))= \operatorname{Hom}(H^0({\cal I}_{{\bold P}^3/{\bold P}^n}(1)), H^0(\OP3(1)))= \sum_{j=0}^3\sum_{i=4}^n \lambda_j^{}\lambda_i^{-1}. \end{equation} Hence we need to study the tangent space of $H_3=\Phi^{-1}\Phi(x)$. \begin{prop} Any fixpoint of $T_3$ in $H_3$ is projectively equivalent to one of the following, where the first four are Cohen-Macaulay and the last four are not: \begin{alignat}{2} (1)\quad& (x_0x_1,x_1x_2,x_2x_3) &\qquad \qquad (5)\quad&(x_0^2,x_0x_1,x_0x_2,x_1x_2x_3) \notag\\ (2)\quad&(x_0x_1,x_1x_2,x_0x_2) &\qquad \qquad (6)\quad&(x_0^2,x_0x_1,x_0x_2,x_1^{}x_2^2)\notag\\ (3)\quad&(x_0x_1,x_2^2,x_0x_2) &\qquad \qquad (7)\quad&(x_0^2,x_0x_1,x_0x_2,x_2^2x_3^{})\notag\\ (4)\quad&(x_0^2,x_0x_1,x_1^2) &\qquad \qquad (8)\quad&(x_0^2,x_0x_1,x_0x_2,x_2^3)\notag \end{alignat} \end{prop} \begin{pf} The action of $T_3$ on ${\bold P}^3$ has the four coordinate points as its fixpoints, and the only one-dimensional orbits are the six lines of the coordinate tetrahedron. Hence any curve invariant under $T_3$ must be supported on these lines. If $C_x\in H_3$ is Cohen-Macaulay and $T_3$-fixed, it is connected, has no embedded points and is not plane. Hence there are only four possibilities: (1) the union of three distinct coordinate lines, two of which are disjoint, (2) the union of three concurrent coordinate lines, (3) a coordinate line doubled in a coordinate plane plus a second line intersecting the first but not contained in the plane, and finally (4) the full first-order neighborhood of a coordinate line. If $x\in Y^{T_3}$, then by the description of the curves in $Y$ we may assume that the quadratic part of the ideal is $(x_0^2,x_0x_1,x_0x_2)$, meaning that $C_x$ is a cubic plane curve in the plane $x_0=0$ which is singular in $P=(0,0,0,1)$ plus an embedded point supported at $P$ but not contained in the plane. For the cubic we have these possibilities: (5) the three coordinate lines, (6) one double coordinate line through $P$ plus another simple line passing through $P$, (7) one double coordinate line through $P$ plus another simple line not passing through $P$, and (8) one coordinate line through $P$ tripled in the plane. \end{pf} \begin{rem} There are several fixpoints of each isomorphism class, in fact it is easy to verify by permuting the variables that in a given ${\bold P}^3$ the numbers of fixpoints of the types 1 through 8 are 12, 4, 24, 6, 12, 24, 24, 24, respectively. This is consistent with the fact that the (even) betti numbers of $H_3$ are 1, 2, 6, 10, 16, 19, 22, 19, 16, 10, 6, 2, 1, so that the Euler characteristic of $H_3$ is 130, see \cite{Elli-Pien-Stro}. \end{rem} \begin{prop}\label{cm-akk} Let $x$ be one of the fixpoints 1--4. Then the representation on the tangent space $T_{H_3}(x)$ is given by \begin{equation*} T_{H_3}(x) = \operatorname{Hom}(F,E)\*({\lambda}_0+{\lambda}_1+{\lambda}_2+{\lambda}_3)-\operatorname{End}(E)-\operatorname{End}(F)+1, \end{equation*} where the representations $E$ and $F$ are given in the following table: \smallskip \begin{center} \begin{tabular} {|c|c|c|c|} \hline Type\vrule height1.2em width0cm depth0.6em & ${\cal I}_x$ & $E$ & $F$ \\ \hline (1)\vrule height1em width0cm depth0.4em & $(x_0x_1,x_1x_2,x_2x_3)$ & $\lambda^{}_0\lambda^{}_1+\lambda^{}_1\lambda^{}_2 +\lambda^{}_2\lambda^{}_3$ & $\lambda^{}_0\lambda^{}_1\lambda^{}_2+\lambda^{}_1 \lambda^{}_2\lambda^{}_3$ \\ \hline (2)\vrule height1em width0cm depth0.4em & $(x_0x_1,x_1x_2,x_0x_2)$ & $\lambda^{}_0\lambda^{}_1+\lambda^{}_1\lambda^{}_2 +\lambda^{}_0\lambda^{}_2$ & $2\lambda^{}_0\lambda^{}_1\lambda^{}_2$ \\ \hline (3)\vrule height1em width0cm depth0.4em & $(x_0x_1,x_2^2,x_0x_2)$ & $\lambda^{}_0\lambda^{}_1+\lambda_2^2+\lambda^{}_0 \lambda^{}_2$ & $\lambda^{}_0\lambda^{}_1\lambda^{}_2+\lambda^{}_0 \lambda^{2}_2$ \\ \hline (4)\vrule height1em width0cm depth0.4em & $(x_0^2,x_0x_1,x_1^2)$ & $\lambda^{2}_0+\lambda^{}_0\lambda^{}_1+\lambda^{2}_1$ & $\lambda^{}_0\lambda^{2}_1+\lambda^{2}_0\lambda^{}_1$ \\ \hline \end{tabular} \end{center} \end{prop} \begin{pf} Follows from \lemref{CMlem}, and the fact that $E$ and $F$ are equivariantly given by \eqref{EE} and \eqref{FF}. \end{pf} \begin{prop} \label{pr:akk} Let $x$ be one of the fixpoints 5--8. Let $\mu$ be the character of the minimal cubic generator, i.e., ${\lambda}_1{\lambda}_2{\lambda}_3$, ${\lambda}_1{\lambda}_2^2$, ${\lambda}_2^2{\lambda}_3$, and ${\lambda}_2^3$, respectively, and let \begin{align*} A=&{\lambda}_0^{-1}({\lambda}_1+{\lambda}_2+{\lambda}_3)+{\lambda}_3({\lambda}_1^{-1}+{\lambda}_2^{-1})\\ B=&{\lambda}_1^3+{\lambda}_1^2{\lambda}_2+{\lambda}_1^2{\lambda}_3+{\lambda}_1{\lambda}_2^2+ {\lambda}_1{\lambda}_2{\lambda}_3+{\lambda}_2^3+{\lambda}_2^2{\lambda}_3 \end{align*} Then the tangent space of $H_3$ at $x$ is given by \begin{equation*} T_{H_3}(x) = A + \mu^{-1}(B-\mu) + ({\lambda}_0{\lambda}_1{\lambda}_2)^{-1}\mu \end{equation*} \end{prop} \begin{pf} Let $\beta=g(x) \in I$ be as in \secref{E}. In fact, all types 5--8 lie over the same fixpoint $\beta$. The first term, $A$ in the sum above, is easily seen to be the representation on the tangent space $T_I(\beta)$. Now $g\:Y\to I$ is a projective bundle, and the fiber $g^{-1}(\beta)$ is the projective space associated to the vector space of cubic forms in $(x_1,x_2)^2{\bold C}[x_1,x_2,x_3]$. The representation on this vector space is $B$, and the second term of the formula of the proposition is the representation on $T_{g^{-1}\beta}(x)$. Thus the first two terms make up $T_Y(x)$. The last term, $({\lambda}_0{\lambda}_1{\lambda}_2)^{-1}\mu$, is the character on $N_{Y/H_3}(x)$. This can be seen from equations (\ref{xi}): by checking each case, one verifies that the normal vector $\xi$ is semi-invariant with character is $({\lambda}_0{\lambda}_1{\lambda}_2)^{-1}$ times the character of the cubic form $q$. \end{pf} \subsection{The computation}\label{F} Let us briefly describe the actual computation, carried out using ``Maple'' \cite{Maple}, of the numbers in the introduction. $H_n$ has a natural torus action with isolated fixpoints. By what we have done in the last section, we can construct a list of all the fixpoints of $H_n$; there are $130\binom{n+1}4 $ of these. For each of them we compute the corresponding tangent space representation, by \eqref{dsum} and propositions \ref{cm-akk} and \ref{pr:akk}. A consequence of the fact that all fixpoints are isolated is that none of the tangent spaces contain the trivial one-dimensional representation. Choose a one-parameter subgroup $\psi\:{\bold C}^* \to T$ of the torus $T$, such that all the induced weights of the tangent space at each fixpoint are non-zero. This is possible since we only need to avoid a finite number of hyperplanes in the lattice of one-parameter subgroups of $T$. For example, we may choose $\psi$ in such a way that the weights of the homogeneous coordinates $x_0,\dots,x_n$ are $1,w,w^2,\dots,w^n$ for a sufficiently large integer $w$. In our computations (for $n\le8$) we used instead weights taken from the sequence 4, 11, 17, 32, 55, 95, 160, 267, 441, but any choice that will not produce a division by zero will do. Since all the tangent weights of the ${\bold C}^*$ action on $H_n$ so obtained are non-trivial, it follows that this action has the same fixpoints as the action of $T$, hence a finite number. By \propref{propB}, we need to evaluate the class $$\delta=c_{{\text{top}}}({\cal E}_{d_1}\oplus\cdots\oplus{\cal E}_{d_p})\cdot \prod_{i=1}^m\gamma_{\lambda_i}\in A^{4n}(H_n).$$ Note that the isomorphism \eqref{basechange} is equivariant. Clearly $H^0({\cal O}_{C_x}(d))$ is spanned by all monomials of degree $d$ not divisible by any monomial generator of $I_x$. Thus we know all the representations ${\cal E}_d(x)$ for all fixpoints $x\in H_n$. By \propref{gammaformler}, $\delta$ is a polynomial $p(\dots,c_k({\cal E}_d),\dots)$ in the Chern classes of the equivariant vector bundles ${\cal E}_d$. To find a function $f$ on the fixpoint set which represents $\delta$, simply replace each occurance $c_k({\cal E}_d)$ by the localized equivariant Chern class $\sigma_k({\cal E}_d,-)$, i.e., put $f=p(\dots,\sigma_k({\cal E}_d,-),\dots)$. Then the class is evaluated by the formula in \thmref{bott1} (2). \section{The Hilbert scheme of points in the plane}\label{hilbkap} Let $V$ be a three-dimensional vector space over ${\bold C}$ and let ${\bold P}(V)$ be the associated projective plane of rank-1 quotients of $V$. Denote by $H_r=\operatorname{Hilb}^r_{{\bold P}(V)}$ the Hilbert scheme parameterizing length-$r$ subschemes of ${\bold P}(V)$. There is a universal subscheme ${\cal Z}\subseteq H_r\times{\bold P}(V)$. We will use similar notational conventions as in \secref{A}: for example, if $x\in{H_r}$, the corresponding subscheme of ${\bold P}^2$ is denoted $Z_x$, its ideal sheaf ${\cal I}_x$ etc. As in \eqref{defE}, let for any integer $n$ \begin{equation} \label{def2E} {\cal E}_{n}=p_{1*}({\cal O}_{\cal Z}\* p_{2}^* {\cal O}_{{\bold P}(V)}(n)), \end{equation} where $p_1$ and $p_2$ are the two projections of $H_r\times{\bold P}(V)$. Since ${\cal Z}$ is finite over ${H_r}$, all basechange maps \begin{equation} {\cal E}_{n}(x) @>\simeq>> H^0(Z_x,{\cal O}_{Z_x}(n)). \label{basechange2} \end{equation} are isomorphisms. In particular, ${\cal E}_n$ is a rank-$r$ vector bundle on ${H_r}$. Denote by ${\cal L}$ the linebundle \begin{equation}\label{defL} {\cal L} = \displaystyle{\operatornamewithlimits{\wedge}}^r{\cal E}_0\*\displaystyle{\operatornamewithlimits{\wedge}}^r{\cal E}_{-1}^{\vee}. \end{equation} Then ${\cal L}$ corresponds to the divisor on $H_r$ corresponding to subschemes $Z$ meeting a given line. We are going to compute integrals of the form \begin{equation}\label{generalintegral} \int_{{H_r}} s_{2r}({\cal E}_n\*{\cal L}^{\*m}) \end{equation} for small values of $r$. Afterwards we will give interpretations of some of these numbers in terms of degrees of power sum and Darboux loci in the system ${\bold P}(S_nV)$ of plane curves of degree $n$ in the dual projective plane ${\bold P}(V^{\vee})$. As usual, we start by identifing all the fixpoints and tangent space representations for a suitable torus action. This has been carried out in more detail in \cite{Elli-Stro-1}, the following simpler presentation is sufficient for the present purpose. As in \secref{E}, let $T\subseteq \operatorname{GL}(V)$ be a maximal torus and let $x_0,x_1,x_2$ be a basis of $V$ diagonalizing $T$ under the natural linear action. The eigenvalue of $x_i$ is a character ${\lambda}_i$ of $T$. We identify characters with one-dimensional representations, hence the representation ring of $T$ with the ring of Laurent polynomials in ${\lambda}_0,{\lambda}_1,{\lambda}_2$. For example, the natural representation on the vector space $V^{\vee}$ can be written ${\lambda}_0^{-1}+{\lambda}_1^{-1}+{\lambda}_2^{-1}$. Fixpoints of ${H_r}$ can be described in terms of partitions, i.e., integer sequences $b=\{b_r\}_{r\ge0}$ weakly decreasing to zero. Let $|b|=\sum_{r\ge0}b_r$. The {\em diagram\/} of a partition $b$ is the set $D(b)=\{(r,s)\in{\bold Z}_{\ge0}^2\mid s< b_r\}$ of cardinality $|b|$. A {\em tripartition\/} is a triple $B=(b^{(0)},b^{(1)},b^{(2)})$ of partitions; put $|B|=\sum_i |b^{(i)}|$; the number being partitioned. The {\em $n$-th diagram\/} $D_n(B)$ of a tripartition $B=(b^{(0)},b^{(1)},b^{(2)})$ is defined for $n\ge|B|$ as follows: Letting the index $i$ be counted modulo 3, we put $$ D_n^i(B)=\{ (n_0,n_1,n_2)\in{\bold Z}^3 \mid n_0+n_1+n_2=n \text{ and } (n_{i+1},n_{i+2})\in D(b^{(i)})\} $$ and $D_n(B) = D_n^0(B)\cup D_n^1(B)\cup D_n^2(B)$. Intuitively, the diagram $D_n(B)$ lives in an equilateral triangle with corners $(n,0,0)$, $(0,n,0)$, and $(0,0,n)$, and originating from the $i$-th corner there is a (slanted) copy of $D(b^{(i)})$. When $n\ge|B|$, these don't overlap. As $n$ grows, the shape and size of the three parts of $D_n(B)$ stay the same, whereas the area separating them grows. We may also define $D_n(B)$ for integers $n<r$ by the same formula as above, but where the union is taken in the sense of multisets, i.e., some elements might have multiplicities 2 or even 3. For $n<r$ the diagram $D_n(B)$ may also stick out of the triangle referred to above. A fixpoint $x\in{H_r}$ corresponds to a length-$r$ subscheme $Z_x\subseteq{\bold P}(V)$ defined by a monomial ideal. Fix an integer $n\ge r$ and consider the set $$ D_n(Z_x)=\{(n_0,n_1,n_2)\in {\bold Z}_{\ge0}^3 \mid \sum_i n_i=n \text{ and } \prod x_i^{n_i} \notin H^0({\bold P}(V),{\cal I}_x(n))\} $$ This set is the $n$-th diagram of a tripartition $B$ of $r$, the three constituent partitions corresponding to the parts of $Z_x$ supported in the three fixpoints of ${\bold P}(V)$. Conversely, starting with a tripartition of $r$, we may obviously construct a monomial ideal of colength $r$ in such a way that we get an inverse of the construction above. Hence there is a natural bijection between $H_r^T$ and the set of tripartitions of $r$. In terms of representations, it follows from the above that for a fixpoint $x$ corresponding to the tripartition $B=(b^{(0)},b^{(1)},b^{(2)})$, we have \begin{equation}\label{reprEn} {\cal E}_n(x)=H^0({\cal O}_{Z_x}(n)) = \sum_{(n_0,n_1,n_2)\in D_n(B)} \prod {\lambda}_i^{n_i}. \end{equation} For $n<r$, the summation index needs to be interpreted as running through the multiset $D_n(B)$. The representation on ${\cal L}$ in the same fixpoint is \begin{equation}\label{reprL} {\cal L}(x) = \prod {\lambda}_i^{|b^{(i)}|}. \end{equation} For $n\ge r$, we also have the following formula for $I_n:=H^0({\bold P}(V),{\cal I}_x(n))$ in the representation ring: $$ I_n =S_nV - H^0({\cal O}_{Z_x}). $$ To compute the tangent space representation, we use a trick that is often useful even in higher dimensions: functorial free resolutions. The tangent space of ${H_r}$ in $x$ is canonically isomorphic to $\operatorname{Ext}^1({\cal I}_x,{\cal I}_x)$. Fix an integer $n\ge r+2$. Then there is a canonical resolution of locally free ${\cal O}_{{\bold P}(V)}$-modules $$ K_{\bullet}:\quad 0\to K_2 \to K_1 \to K_0 $$ of ${\cal I}_x(n)$, where $K_p=\Omega_{{\bold P}(V)}^p(p)\*I_{n-p}$. As in \eqref{beilinson}, this is a special case of Beilinson's spectral sequence. $T$ acts on $K_\bullet$. Let $S^\bullet$ be the total complex associated to the double complex $\operatorname{Hom}_{{\bold P}^2}(K_\bullet,K_\bullet)$. Then the $i$-th cohomology group of $S^\bullet$ is $\operatorname{Ext}^i({\cal I}_x(n),{\cal I}_x(n))=\operatorname{Ext}^i({\cal I}_x,{\cal I}_x)$ for $i=0,1,2$ \cite[Lemma 2.2]{Elli-Stro-4}. For $i=0$ this is ${\bold C}$ (with trivial action) and for $i=2$ it is zero. Using the canonical identifications $\operatorname{Hom}(\Omega^p(p),\Omega^q(q)) = \displaystyle{\operatornamewithlimits{\wedge}}^{p-q}V^{\vee}$, we end up with the following formula for the tangent space representation in terms of the data of the tripartition $B$: {\allowdisplaybreaks \begin{align}\label{reprT} T_{{H_r}}(x) = \,& 1-(\sum_{i=0}^2(-1)^i\operatorname{Ext}^i({\cal I}_x,{\cal I}_x))\\ = \,&1 -(S^0-S^1+S^2) \notag\\ =\,&1-(\operatorname{Hom}(I_n,I_{n})+\operatorname{Hom}(I_{n-1},I_{n-1})+\operatorname{Hom}(I_{n-2},I_{n-2}))\notag\\* &+ ({\lambda}_0^{-1}+{\lambda}_1^{-1}+{\lambda}_2^{-1}) (\operatorname{Hom}(I_{n-1},I_n)+\operatorname{Hom}(I_{n-2},I_{n-1}))\notag\\* &-({\lambda}_0^{-1}{\lambda}_1^{-1}+{\lambda}_1^{-1}{\lambda}_2^{-1}+{\lambda}_2^{-1}{\lambda}_0^{-1})\operatorname{Hom}(I_{n-2},I_n)\notag \end{align} } Here are the computational results on the Hilbert scheme which will be used in the following applications: \begin{prop} \label{computation1} Let ${\cal E}_n$ be as in \eqref{def2E}. For $2\le r\le 8$, we have $$ \int_{{H_r}}s_{2r} ({\cal E}_n) = s_r(n), $$ where $s_r(n)$ are as in \thmref{main4}. \end{prop} \begin{prop} \label{computation2} Let ${\cal E}_{-1}$ and ${\cal L}$ be as in \eqref{def2E} and \eqref{defL}. For $r=2,3,4,5,6,7,8,9,10$, we have $$ \int_{{H_r}}s_{2r}( {\cal E}_{-1}\*{\cal L}) = 0,0,0,0,2540, 583020, 99951390, 16059395240, 2598958192572. $$ \end{prop} \begin{pf} For both propositions, apply Bott's formula. The one-parameter subgroup of $T$ such that the ${\lambda}_i$ have weights 0,1,19 will work. The contribution at each fixpoint is given by \eqref{reprEn}, \eqref{reprL}, and \eqref{reprT}. Generate all fixpoints and perform the summation using e.g. \cite{Maple}. \end{pf} \begin{rem} There is no difficulty in principle to evaluate \eqref{generalintegral} directly with symbolic values of both $n$ and $m$, for given values of $r$. For example, for $r=3$, the result is $(n^{6}+24\,n^{5}m+252\,n^{4}m^{2}+1344\,n^{3}m^{3}+3780\, n^{2}m^{4}+5040\,nm^{5}+ 2520\,m^{6}-30\,n^{4}-432\,n^{3}m-2520\,n^{2}m ^{2}-6048\,nm^{3}-5040\,m^{4}+45\,n^{3}+ 504\,n^{2}m+2268\,nm^{2}+3024 \,m^{3}+206\,n^{2}+1200\,nm+1512\,m^{2}-576\,n- 1728\,m+384)/6$. However, with given computer resources, one gets further the fewer variables one needs. On a midrange workstation, we could do this integral up to $r=5$. \end{rem} \begin{rem} Tyurin and Tikhomirov \cite{Tyur-Tikh} and Le Potier have shown that \propref{computation2} implies that the Donaldson polynomial $q_{17} ({\bold P}^2) = 2540 $. It may also be deduced from the proposition that $q_{21}({\bold P}^2) = 233208$, see \cite{Tyur-Tikh} or our forthcoming joint paper with J. Le Potier. \end{rem} \subsection{Power sum varieties of plane curves} Closed points of ${\bold P}(S_nV)$ correspond naturally to curves of degree $n$ in the dual projective plane ${\bold P}(V^{\vee})$. In particular, points of ${\bold P}(V)$ correspond to lines in ${\bold P}(V^{\vee})$, so $H_r=\operatorname{Hilb}^r_{{\bold P}(V)}$ is a compactification of the set of unordered $r$-tuples of linear forms modulo scalars. Let $r$ and $n$ be given integers. Let $U(r,n)$ be the set of pairs $(\{L_1,\dots,L_r\},C)$ where the $L_i\subseteq{\bold P}(V^{\vee})$ are lines in general position, and $C$ is a curve with equation of the form $\sum_{i=1}^r a_i\ell_i^n \in S_nV$, where $\ell_i\in V^{\vee}$ is an equation of $L_i$. Then the power sum variety $PS(r,n)$ is the closure of the image of $U(r,n)$ in ${\bold P}(S_nV)$ under the projection onto the last factor. To compute the degree of the image times the degree $p(r,n)$ of the map $U(r,n)\to PS(r,n)$, we need to find a workable compactification of $U(r,n)$. Recall from \eqref{def2E} the rank-$r$ vector bundle ${\cal E}_n$ on ${H_r}$. It comes naturally with a morphism $S_nV_{H_r}\to {\cal E}_n$, which is surjective if $n\ge r-1$. Now consider the projective bundle over $H_r$: \begin{equation} {\bold P}({\cal E}_n) \subseteq {\bold P}(S_nV) \times {H_r}. \end{equation} It is easy to verify that ${\bold P}({\cal E}_n)$ contains $U(r,n)$ as an open subset. It follows that $p(r,n)$ times the degree of $PS(r,n)$ is given by the self-intersection of the pullback to ${\bold P}({\cal E}_n)$ of ${\cal O}_{{\bold P}(S_nV)}(1)$. This is $\int_{{\bold P}({\cal E}_n)} c_1({\cal O}_{{\bold P}({\cal E}_n)}(1))^{3r-1}$, and pushing it forward to the Hilbert scheme, we get almost by definition, $\int_{{H_r}} s_{2r}({\cal E}_n)$ \cite[Ch.~ 3]{IT}. Together with \propref{computation1}, this proves \thmref{main4}. \subsection{Darboux curves} These curves are also defined in terms of linear forms, and we may take $H_{n+1}$ as a parameter space for the variety of complete $(n+1)$-gons. For a length-$(n+1)$ subscheme $Z\subseteq{\bold P}(V)$, put $E=H^0({\cal O}_{Z}(-1))=H^1({\cal I}_Z(-1))$ and $F=H^1({\cal I}_Z)$. The multiplication map $ V\* E \to F $ gives rise to a bundle map over the dual plane ${\bold P}(V^{\vee})$: $$ m\: E_{{\bold P}(V^{\vee})}(-1) \to F_{{\bold P}(V^{\vee})}. $$ If $Z$ consists of $n+1$ points in general position, the degeneration locus $D(Z)$ corresponds to the set of bisecant lines to $Z$, i.e., the singular locus of the associated $(n+1)$-gon. The Eagon-Northcott resolution of $D(Z)$ gives the following short exact sequence: $$ 0 \to F_{{\bold P}(V^{\vee})}^{\vee}(-1) \to E_{{\bold P}(V^{\vee})}^{\vee} \to L\*_{\bold C}{\cal I}_{D(Z)}(n) \to 0, $$ showing that there is a natural surjection $$ S_nV^{\vee} \to H^0({\bold P}(V^{\vee}),{\cal I}_{D(Z)}(n))^{\vee} \simeq H^0(Z,{\cal O}_Z(-1))\*L. $$ Here $L$ is the onedimensional vector space $\det(F)\*\det(E)^{-1}$. Globalizing this construction over $H_{n+1}$ gives a natural map $$ p\: S_nV_{H_{n+1}} \to {\cal E}_{-1}\*{\cal L} $$ such that the closure of the image of the induced rational map ${\bold P}({\cal E}_{-1}\*{\cal L})\to{\bold P}(S_nV)$ is the Darboux locus $D(n)$. By the lemma below, this rational map is actually a morphism. Thus we may argue as in the power sum case and find that $p(n)$ times the degree of $D(n)$ is $\int_{{\bold P}({\cal E}_{-1}\*{\cal L})} c_1({\cal O}(1))^{3n+2}= \int_{H_{n+1}} s_{2n+2}({\cal E}_{-1}\*{\cal L})$. This together with \propref{computation2} implies \thmref{main5}. \begin{lem} The bundle map $p\:S_nV_{H_{n+1}} \to {\cal E}_{-1}\*{\cal L}$ over $H_{n+1}$ is surjective. \end{lem} \begin{pf} Assume the contrary. Since the support of the cokernel is closed and $\operatorname{GL}(V)$-invariant, there exists a subscheme $Z$ in $\operatorname{Supp}\operatorname{Coker}(p)$ which is supported in one point. Without loss of generality we may assume that $Z$ is supported in the point $x_1=x_2=0$. Let $E$ and $F$ be as above, and let $K\subseteq E$ be a subspace of codimension 1. For a linear form $\ell\in V$, let $m_{\ell}\:E\to F$ be multiplication by $\ell$. The assumption that $p$ is not surjective means that $K$ can be chosen such that the determinant of the restriction of $m_{\ell}$ to $K$ is 0 for all $\ell\in V$. Multiplication by $x_0$ induces an isomorphism $E=H^0({\cal O}_Z(-1)) \simeq H^0({\cal O}_Z) \simeq {\bold C}[x,y]/I_Z$, where $x=x_1/x_0$ and $y=x_2/x_0$. Under these identifications, if $\ell=1-ax-by$ is the image of a general linear form, the kernel of $m_{\ell}$ is generated by $1/\ell$. Consider the set $S$ consisting of all such elements $\ell^{-1}$, with $a,b\in{\bold C}$. The series expansion $\ell^{-1}=1+(ax+by)+(ax+by)^2+\cdots$ shows that $S$ generates ${\bold C}[x,y]/(x,y)^m$ as a vector space for all $m\ge0$. Indeed, a hyperplane $W_m\subseteq{\bold C}[x,y]/(x,y)^m$ containing the image of $S$ would, by induction on $m$, dominate ${\bold C}[x,y]/(x,y)^{m-1}$. Hence $W_m$ could not contain $(x,y)^{m-1}/(x,y)^m$. But the image of $S$ in $(x,y)^{m-1}/(x,y)^m$ is the cone over a rational normal curve of degree $m-1$, hence spans this space. Since $(x,y)^m\subseteq I_Z$ for $m$ large, it follows that $S$ generates ${\cal O}_Z$ and hence $E$ as a vector space. Now for an $\ell$ such that $\ell^{-1}\notin K$, the restriction of $m_\ell$ to $K$ will be an isomorphism. This leads to the desired contradiction. \end{pf} \section{Discussion} How general is the strategy of using Bott's formula in enumerative geometry, as outlined in these examples? The first necessary condition is probably a torus action, although Bott's formula is valid in a more general situation: a vector field with zeros and vector bundles acted on by the vector field. It seems to us, though, that the cases where one stands a chance of analysing the local behaviour of bundles near all zeroes of such a field are those where the both the vector field and its action on the vector bundles are ``natural'' in some sense. If there are parameter spaces with natural flows on them, not necessarily coming from torus actions, presumably Bott's formula could be useful. It is not necessary that the fixpoints be isolated in order for the method to give results. Kontsevich's work \cite{Kont-1} is a significant example of this. Another natural candidate for Bott's formula is the moduli spaces of semistable torsionfree sheaves on ${\bold P}^2$. These admit torus actions, but not all fixpoints are isolated. One can still control the structure of the fixpoint components, however. This may hopefully be used for computing Donaldson polynomials of the projective plane, at least in some cases. A more serious obstacle to the use of Bott's formula is the presence of singularities in the parameter space. For example, all components of the Hilbert scheme of ${\bold P}^n$ admit torus actions with isolated fixpoints, but they are almost all singular. The main non-trivial exceptions are actually the ones treated in the present paper. On the other hand, singularities present inherent problems for most enumerative approaches, especially if a natural resolution of singularities is hard to find. For all examples in this paper, one needs a computer to actually perform the tedious computations. We mentioned already that the number of fixpoints in the case of twisted cubics in ${\bold P}^n$ is $130\binom{n+1}4$. For the Hilbert scheme of length-8 subschemes of ${\bold P}^2$, the corresponding number is 810. From the point of view of computer efficiency, there are some advantages to the use of Bott's formula in contrast to trying to work symbolically with generators and relations in cohomology, as for example in \cite{Elli-Stro-3} or \cite{schubert}. First of all, the method works even if we don't know all relations, as is the case for the Hilbert scheme of the plane, for example. But the main advantage is perhaps that Bott's formula is not excessively hungry for computer resources. It is often straightforward to make a loop over all the fixpoints. The computation for each fixpoint is fairly simple, and the result to remember is just a rational number. This means that the computer memory needed does not grow much with the number of fixpoints, although of course the number of CPU cycles does. For example, most of the numbers of elliptic quartics were computed on a modest notebook computer, running for several days. Finally, Bott's formula has a nice error-detecting feature, which is an important practical consideration: If your computer program actually produces an integer rather than just a rational number, chances are good that the program is correct!
1996-03-05T06:13:25
9411
alg-geom/9411004
en
https://arxiv.org/abs/alg-geom/9411004
[ "alg-geom", "hep-th", "math.AG", "math.QA", "q-alg" ]
alg-geom/9411004
Ezra Getzler
Ezra Getzler
Operads and moduli spaces of genus $0$ Riemann surfaces
26 pages, latex2e with amslatex 1.2beta (available from e-math.ams.com by anonymous ftp), no figures
null
null
null
null
We study a pair of dual operads which arise in the study of moduli spaces of pointed genus 0 curves (this duality is similar to that between commutative and Lie algebras). These operads are both quadratic, and even Koszul, and arise in the theory of quantum cohomology.
[ { "version": "v1", "created": "Wed, 9 Nov 1994 18:51:02 GMT" } ]
2008-02-03T00:00:00
[ [ "Getzler", "Ezra", "" ] ]
alg-geom
\section{$\SS$-modules and operads} In this section, we recall the basic definitions of the theory of operads. For more details, see \cite{n-algebras} and \cite{GK}. \subsection{$\SS$-modules} An $\SS$-module is a collection of chain complexes (all chain complexes in this paper are over the field $\mathbb{C}$, and have finite dimensional total homology) $$ \{ \v(n) \mid n\ge0 \} , $$ together with an action of $\SS_n$ on $\v(n)$. This definition generalizes Joyal's notion of a linear species \cite{Joyal}, which is an ungraded $\SS$-module. A chain complex $V$ may be thought of as an $\SS$-module by setting $$ \v(n) = \begin{cases} V , & n=1 , \\ 0 , & n\ne1 . \end{cases} $$ \subsection{Schur functors} Given an $\SS$-module $\v$ and a finite set $S$, we define $$ \v(S) = \Bigl( \bigoplus\begin{Sb}\text{bijections}\\\begin{CD}f:\{1,\dots,n\}@>>>S \end{CD}\end{Sb} \v(n) \Bigr)_{\SS_n} . $$ It is clear that if $S=\{1,\dots,n\}$, then $\v(S)$ is naturally identified with $\v(n)$. To an $\SS$-module $\v$ is associated an endofunctor of the category of chain complexes, called the Schur functor of $\v$, by the formula $$ V \mapsto \mathsf{S}(\v,V) = \bigoplus_{n=0}^\infty \v\o_{\SS_n} V^{\o n} ; $$ here $V^{\o n}$ is the graded $n$th tensor power of $V$. Introduce a monoidal structure on the category of $\SS$-modules, with tensor product $$ (\v\circ\mathcal{W})(n) = \bigoplus_{k=0}^\infty \Bigl( \v(k) \o \bigoplus_{\begin{CD}f:\{1,\dots,n\}@>>>\{1,\dots,k\} \end{CD}} \bigotimes_{i=1}^k \mathcal{W}(f^{-1}(i)) \Bigr)_{\SS_k} , $$ and unit the $\SS$-module ${1\!\!1}$: $$ {1\!\!1}(n) = \begin{cases} \mathbb{C} , & n=1 , \\ 0 , & n\ne1 . \end{cases}$$ The peculiar formula for $\v\circ\mathcal{W}$ is justified by $$ \mathsf{S}(\v,\mathsf{S}(\mathcal{W},V)) = \mathsf{S}(\v\circ\mathcal{W},V) . $$ Note also that $\mathsf{S}({1\!\!1},V)=V$. For more on this formalism, see Chapter 1 of \cite{n-algebras}. \subsection{Operads} An operad is a monoid in the category of $\SS$-modules, that is, an $\SS$-module $\a$ with product $\begin{CD}\rho:\a\circ\a@>>>\a \end{CD}$ and unit $\begin{CD}\eta:{1\!\!1}@>>>\a \end{CD}$ satisfying the axioms of associativity and unit \cite{Maclane}. We denote the image under the product $\rho$ of $$ a\o b_1\o\dots\o b_k \in \a(k)\o\a(n_1)\o\dots\o\a(n_k) $$ by $a(b_1,\dots,b_k)$ and the unit by $1\in\a(1)$. An operad structure on an $\SS$-module $\a$ such that $\a(n)=0$ for $n\ne1$ is the same thing as an associative algebra structure on $\a(1)$. Whereas an element of an algebra has only one ``input'' and one ``output,'' an element of an operad has multiple inputs and one output. \subsection{The endomorphism operad of a chain complex} If $V$ is a chain complex, its endomophism operad is the $\SS$-module $$ \mathcal{E}_V(n) = \operatorname{Hom}(V^{\o n},V) . $$ This is an operad, whose product is given by composition: if $a\in\operatorname{Hom}(V^{\o k},V)$ and $b_i\in\operatorname{Hom}(V^{\o n_i},V)$, then $$ a(b_1,\dots,b_k) = a \* (b_1\o\dots\o b_k) , $$ where we think of $b_1\o\dots\o b_k$ as an element of $\operatorname{Hom}(V^{\o(n_1+\dots+n_k)},V^{\o k})$. \subsection{Suspension of operads} If $V$ is a chain complex, denote by $\Sigma V$ the chain complex such that $(\Sigma V)_i=V_{i-1}$, with differential $-\delta$. If $\v$ is an $\SS$-module, denote by $\Lambda\v$ the $\SS$-module $$ (\Lambda\v)(n) = \Sigma^{1-n} \operatorname{sgn}_n \o \v(n) , $$ where $\operatorname{sgn}_n$ is the sign character of $\SS_n$. There is a natural isomorphism $$ \mathsf{S}(\Lambda\v,V) \cong \Sigma\mathsf{S}(\v,\Sigma^{-1}V) . $$ It follows that if $\a$ is an operad, then so is $\Lambda\a$, and if $A$ is an $\a$-algebra, then $\Sigma A$ is a $\Lambda\a$-algebra. \subsection{Algebras over an operad} An algebra over an operad $\a$ is a chain complex $A$, together with a morphism of operads $\begin{CD}\a@>>>\mathcal{E}_A \end{CD}$. Thus, if $A$ is an algebra over an operad $\a$ and $\rho\in\a(n)$, there is a product $a_1\o\dots a_n\mapsto\rho(a_1,\dots,a_n)$ from $A^{\o n}$ to $A$. These products are equivariant, under the actions of $\SS_n$ on $\a(n)$ and $A^{\o n}$, associative with respect to the product of $\a$, and $1(a)=a$, where $1\in\a(1)$ is the unit of $\a$. \subsection{Construction of operads} Given an algebraic structure defined by a set of multilinear operations together with a set of multilinear relations among them, one may construct the operad $\a$ having this presentation, in such a way that an $\a$-algebra is the same thing as an instance of the original algebraic structure. Denote by $V\mapsto T(V)$ the free algebra generated by the chain complex $V$ with respect to this algebraic structure. To define $\a(n)$, we form the free algebra $T(x_1,\dots,x_n)$ generated by the free vector space $\mathbb{C}^n$ of rank $n$. The torus $(\mathbb{C}^\times)^n$ acts on this chain complex; let $\a(n)$ be the $\SS_n$-submodule on which it acts by the character $(z_1,\dots,z_n)\mapsto z_1\dots z_n$. The group $\SS_n$ acts on $T(x_1,\dots,x_n)$, and thus on $\a(n)$, by permutation of the letters $x_i$. It is not difficult to see that the $\SS$-module thus constructed is an operad: the unit is the word $x_1\in\a(1)\subset T(x_1)$, while the product is defined by substitution. Let us give some examples of this construction. In the case of commutative algebras, we call the resulting operad ${\operatorname{\mathcal{C}\mathit{om}}}^+$; the $\SS_n$-module ${\operatorname{\mathcal{C}\mathit{om}}}^+(n)$ is spanned by the word $x_1\dots x_n$ the free commutative algebra generated by letters $\{x_1,\dots,x_n\}$, and carries the trivial action of $\SS_n$. (Here, we are dealing with non-unital commutative algebras, so that ${\operatorname{\mathcal{C}\mathit{om}}}^+(0)=0$.) Let ${\operatorname{\mathcal{A}\mathit{ss}}}^+$ be the operad associated to associative algebras; the $\SS_n$-module ${\operatorname{\mathcal{A}\mathit{ss}}}^+(n)$ is spanned by the words $$ \{ x_{\sigma(1)}\dots x_{\sigma(n)} \mid \sigma\in\SS_n \} , $$ and carries the regular representation of $\SS_n$. (Again, we set ${\operatorname{\mathcal{A}\mathit{ss}}}^+(0)=0$.) Finally, the operad associated to Lie algebra is denoted ${\operatorname{\mathcal{L}\mathit{ie}}}^+$. The underlying $\SS$-module may be studied by means of the Poincar\'e-Birkhoff-Witt theorem, which implies that $$ {\operatorname{\mathcal{A}\mathit{ss}}}^+ \cong {\operatorname{\mathcal{C}\mathit{om}}}^+ \circ {\operatorname{\mathcal{L}\mathit{ie}}}^+ . $$ In Proposition \ref{Lie}, we will show how this leads to a formula for the character of ${\operatorname{\mathcal{L}\mathit{ie}}}^+(n)$, due to Klyachko \cite{Klyachko}. \subsection{The configuration spaces $\mathbb{C}^n_0$ and the braid operad $\b$} Let $\mathbb{C}^n_0$ be the configuration space of $n$ labelled points in $\mathbb{C}$. Define the $\SS$-module $\mathcal{B}\mathit{raid}$ by $$ \mathcal{B}\mathit{raid}(n) = \begin{cases} H_{\bullet}(\mathbb{C}^n_0) , & n>0 , \\ 0 , & n=0 . \end{cases}$$ We now construct a natural operad structure on $\mathcal{B}\mathit{raid}$. (We called this the braid operad in \cite{n-algebras}.) In the definition of operads, one can replace the category of chain complexes by the category of topological spaces, and the tensor product by Cartesian product, obtaining the notion of a topological operad. If $\mathcal{O}$ is a topological operad and $H_{\bullet}(-)$ is a generalized homology theory with products, $H_{\bullet}(\mathcal{O})$ is an operad in the category of graded vector spaces; this gives a useful method of constructing operads. Boardman and Vogt have constructed a topological operad called the little discs operad \cite{BoardmanVogt}. Let $D$ be the closed unit disc in $\mathbb{C}$, and let $\mathcal{O}(n)$ be the topological space $$\textstyle \mathcal{O}(n) = \bigl\{ \binom{z_1\dots z_n}{r_1\dots r_n} \in \binom{D^n}{\mathbb{R}^n_+} \big| \text{ the discs $r_iD+z_i$ are disjoint subsets of $D$} \bigr\} . $$ The symmetric group $\SS_n$ acts on $\mathbb{CP}(n)$ by permuting the discs: $$\textstyle \sigma\binom{z_1\dots z_n}{r_1\dots r_n} = \binom{z_{\sigma(1)}\dots z_{\sigma(n)}}{r_{\sigma(1)}\dots r_{\sigma(n)}} . $$ The product in this operad is defined by gluing of disks: if $a=\binom{z_1\dots z_k}{r_1\dots r_k}$ and $b_i=\binom{y_{i,1}\dots y_{i,n_i}}{s_{i,1}\dots s_{i,n_i}}$, then $$\textstyle a(b_1,\dots,b_k) = \begin{pmatrix} r_1y_{1,1}+z_1&\dots&r_1y_{1,n_1}+z_1&\dots& r_ky_{k,1}+z_k&\dots&r_ky_{k,n_k}+z_k \\ r_1s_{1,1}&\dots&r_1s_{1,n_1}&\dots&r_ks_{k,1}&\dots&r_ks_{k,n_k} \end{pmatrix} . $$ The map $\begin{CD}\mathcal{O}(n)@>>>\mathbb{C}^n_0 \end{CD}$ defined by $\binom{z_1\dots z_n}{r_1\dots r_n}\mapsto(z_1,\dots,z_n)$ is a homotopy equivalence, and thus the homology operad $H_{\bullet}(\mathcal{O})$ of this topological operad has $\mathcal{B}\mathit{raid}$ as its underlying $\SS$-module. The operad $\mathcal{B}\mathit{raid}$ has the following presentation (see \cite{Cohen:thesis} and \cite{n-algebras}): it is generated by two operations, a commutative product of degree $0$ and a Lie bracket of degree $1$, satisfying the Poisson relation: $$ [a,bc] = [a,b]c + (-1)^{(|a|+1)|b|} b[a,c] . $$ It follows that the $\SS$-module $\mathcal{B}\mathit{raid}$ is isomorphic to ${\operatorname{\mathcal{C}\mathit{om}}}^+\circ\Lambda^{-1}{\operatorname{\mathcal{L}\mathit{ie}}}^+$. In particular, we obtain another realization of the $\SS_n$-module ${\operatorname{\mathcal{L}\mathit{ie}}}^+(n)$: $$ {\operatorname{\mathcal{L}\mathit{ie}}}^+(n) \cong \operatorname{sgn}_n \o H_{n-1}(\mathbb{C}^n_0) . $$ \section{Cyclic operads} All of the operads which we discuss in this article, notably the hypercommutative and gravity operads, are cyclic operads \cite{cyclic} --- that is, there is a notion of invariant inner product on algebras over these operads. In this section, we recall the basics of the theory of cyclic operads. \subsection{Cyclic $\SS$-modules} A cyclic $\SS$-module is an $\SS$-module $\v$ together with an action of $\SS_{n+1}$ on $\v(n)$ extending the action of $\SS_n$. The name derives from the fact that an action of $\SS_{n+1}$ is determined by compatible actions of $\SS_n$ and the cyclic group $C_{n+1}\subset\SS_{n+1}$ generated by the cycle $(01\dots n)$. We denote the action of $(01\dots n)$ on $\v(n)$ by $v\mapsto v^*$, motivated by the fact that a cyclic $\SS$-module structure on the $\SS$-module associated to a chain complex $V$ is just an involution $v\mapsto v^*$. It is convenient to write $\v\(n\)$ for $\v(n+1)$. If $\v$ is a cyclic $\SS$-module, denote by $\Lambda\v$ the cyclic $\SS$-module $$ (\Lambda\v)\(n\) = \Sigma^{2-n} \operatorname{sgn}_n \o \v\(n\) . $$ Thus, $\Lambda$ applied to the $\SS$-module underlying $\v$ is isomorphic to the $\SS$-module underlying $\Lambda\v$. A stable cyclic $\SS$-modules is a cyclic $\SS$-module $\a$ which satisfies the condition that $\a\(n\)=0$ for $n<3$. (The word ``stable'' comes from the theory of algebraic curves.) \subsection{Cyclic operads} A cyclic operad $\a$ is a cyclic $\SS$-module $\a$ with an operad structure, such that $1^*=1$ and for all $a\in\a(k)$ and $b\in\a(l)$, $$ a(1,\dots,1,b)^* = (-1)^{|a|\,|b|} b^*(a^*,1,\dots,1) . $$ Note that if $\v$ is a cyclic $\SS$-module associated to a chain complex $V$ with involution, then a cyclic operad structure on $\v$ is the same thing as a $\ast$-algebra structure on $V$. \subsection{Stable cyclic operads} A cyclic operad $\a$ whose underlying $\SS$-module is stable is called a stable cyclic operad. Because of this condition, $\a\(2\)=0$, so we can no longer think of $\a$ as having a unit. This requires the introduction of non-unital operads, following Markl. Observe that all of the products $a(b_1,\dots,b_k)$ of a (unital) operad may be obtained by iterating the products $$ a\circ_i b = a(\underset{\text{$i-1$ times}}{\underbrace{1,\dots,1}},b, \underset{\text{$k-i$ times}}{\underbrace{1,\dots,1}}) . $$ The axioms for non-unital operads may be found in \cite{modular}: they are of two types, equivariance and associativity. Finally, a non-unital operad is cyclic if $(a\circ_kb)^*=(-1)^{|a|\,|b|}b^*\circ_1a^*$. \subsection{Invariant inner products and cyclic algebras} If $V$ is a chain complex with inner product $\<-,-\>$ (which we suppose to be non-degenerate), the endomorphism operad $\mathcal{E}_V$ is a cyclic operad, such that if $a\in\a(n)$ and $v_i\in V$, $0\le i\le n$, $$ \< v_0 , a(v_1,\dots,v_n) \> = \< v_n , a^*(v_0,\dots,v_{n-1}) \> . $$ A cyclic algebra over a cyclic operad is a chain complex $A$ with inner product, together with a morphism of cyclic operads $\begin{CD}\a@>>>\mathcal{E}_A \end{CD}$. If $A$ is an algebra over the operad underlying a cyclic operad $\a$, and $\<-,-\>$ is an inner product on $A$, we say that the inner product is invariant. Note that if $\a$ is a cyclic operad, then the operad $\Lambda\a$ is not cyclic, but rather anticyclic: $(a\circ_kb)^*=-(-1)^{|a|\,|b|}b^*\circ_1a^*$. This reflects the fact that if $V$ is a chain complex and $\<-,-\>$ is an inner product on $V$, then there is induced on $\Sigma V$ an antisymmetric non-degenerate bilinear form $(-1)^{|v|}\<\Sigma v,\Sigma w\>$. However, $\Lambda^2\a$ is again a cyclic operad, and if $A$ is a cyclic $\a$-algebra, then $\Sigma^2A$ is a cyclic $\Lambda^2\a$-algebra. \subsection{Examples of cyclic operads} Since associative, commutative and Lie algebras all have well-known notions of invariant inner product, it is not surprising that the corresponding operads are cyclic. For the operads $\a^+$, where $\a\in\{{\operatorname{\mathcal{A}\mathit{ss}}},{\operatorname{\mathcal{C}\mathit{om}}},{\operatorname{\mathcal{L}\mathit{ie}}}\}$, define the $\SS$-module $\a$ by setting $\a(n)=\a^+(n)$ for $n>1$ and $\a(n)=0$ for $n\le1$. The cyclic structure of ${\operatorname{\mathcal{C}\mathit{om}}}$ is simple to describe, since an inner product on a commutative algebra is invariant if and only if \begin{equation} \label{invariant} \<a,bc\> = \<ab,c\> . \end{equation} It follows from this formula that $a=a^*$ for all $a\in{\operatorname{\mathcal{C}\mathit{om}}}\(n\)$. In a similar way, the action of $\SS_n$ on ${\operatorname{\mathcal{A}\mathit{ss}}}\(n\)$ is determined by the condition that an inner product on an associative algebra is invariant if and only if \eqref{invariant} holds. It turns out that the $\SS_n$-module ${\operatorname{\mathcal{A}\mathit{ss}}}\(n\)$ is the induced representation $\operatorname{Ind}_{C_n}^{\SS_n}{1\!\!1}$, where $C_n$ is the subgroup of $\SS_n$ generated by $\tau_n$, and ${1\!\!1}$ is its trivial representation. We may think of ${\operatorname{\mathcal{A}\mathit{ss}}}\(n\)$ as being the $\SS_n$-module spanned by symbols $$ \<x_{\sigma(1)},x_{\sigma(2)}\dots x_{\sigma(n)}\> , \quad \sigma\in\SS_n , $$ representing the inner product of $x_{\sigma(2)}\dots x_{\sigma(n)}$ with $x_{\sigma(1)}$, subject to the relations $$ \<x_{\sigma(1)},x_{\sigma(2)}\dots x_{\sigma(n)}\> \sim \<x_{\sigma(n)},x_{\sigma(1)}\dots x_{\sigma(n-1)}\> . $$ The cyclic structure on the Lie operad is associated with the usual notion of an invariant inner product (or Killing form) on a Lie algebra, satisfying $$ \<[a,b],c\> = \<a,[b,c]\> . $$ The character of the $\SS_{n+1}$-module ${\operatorname{\mathcal{L}\mathit{ie}}}(n)$ is calculated in \cite{modular}. We will obtain a realization of this representation as a homology group in the next section. It is proved in \cite{cyclic} that the braid operad $\mathcal{B}\mathit{raid}$ is not a cyclic operad. \section{The moduli spaces $\overline{\mathcal{M}}_{0,n}$} In this section, we study the combinatorial structure of the compactified moduli spaces $\overline{\mathcal{M}}_{0,n}$. We then define the gravity and hypercommutative operads, and introduce the fundamental exact sequences relating them, which are obtained by considering the spectral sequence associated by Deligne to the stratified space $\overline{\mathcal{M}}_{0,n}$. \subsection{Graphs and trees} The strata of the compactification $\overline{\mathcal{M}}_{0,n}$ are labelled by trees with $n$ legs, and we recall some definitions from the theory of trees in this paragraph (see also \cite{GK} and \cite{modular}). A graph $G=(F,\pi,\tau)$ is a finite set $F=\operatorname{Flag}(G)$, the set of flags of the graph, together with a partition $\pi$ and an involution $\tau$ of $F$. (By a partition, we mean a decomposition of $F$ into disjoint subsets, possibly empty, called its blocks.) The vertices $\operatorname{Vert}(G)$ of the graph $G$ are the blocks of $\pi$, the edges $\operatorname{Edge}(G)$ are the $2$-cycles of $\tau$, while the legs $\operatorname{Leg}(G)$ are the fixed points of $\tau$. To a graph $G$ is associated a cell complex $G$ whose cells have dimension $0$ and $1$, and whose ends correspond to the legs $\operatorname{Leg}(G)$. A graph is called a tree if this complex is simply connected. We will have no further use for non-simply connected graphs in this paper; however, much of the theory we describe has an analogue for general graphs \cite{modular}. The legs $\operatorname{Leg}(v)$ of a vertex $v\in\operatorname{Vert}(G)$ are the flags in the corresponding equivalence class, while the valence $|v|$ of a vertex is the cardinality of $\operatorname{Leg}(v)$. If $S$ is a finite set, let $\mathcal{T}\(S\)$ be the set of isomorphism classes of trees $T$ whose external edges are labelled by the elements of $S$ and such that each vertex has valence at least three. Note that $\mathcal{T}\(S\)$ is finite. The set of trees is graded by the number of edges: $$ \mathcal{T}\(S\) = \bigcup_{i=0}^{|S|-3} \mathcal{T}_i\(S\) . $$ In particular, $\mathcal{T}_0\(S\)$ has a single element, the tree with one vertex whose set of flags equals $S$. Denote by $\det(S)$ the determinant line $\Lambda^{\text{max}}\mathbb{C}^S$, which is a representation of $\operatorname{Aut}(S)$. (For example, $\det(\{1,\dots,n\})$ is just the sign representation $\operatorname{sgn}_n$.) If $T$ is a tree, let $\det(T)$ be the determinant line $\det(\operatorname{Vert}(T))$ of the set of vertices of $T$. There are natural isomorphisms \begin{equation} \label{det} \det(T) \cong \det(\operatorname{Edge}(T)) \cong \det(\operatorname{Leg}(T)) \o \bigotimes_{v\in\operatorname{Vert}(T)} \det(\operatorname{Leg}(v)) . \end{equation} \subsection{Stable curves} A stable curve with $n$ marked points is a projective curve $\Sigma$ whose only singularities are double points, together with an embedding of $\{1,\dots,n\}$ in the set of smooth points of $\Sigma$, such that there are no continuous automorphisms of $\Sigma$ fixing the marked points and double points. Knudsen has proved that the moduli space $\overline{\mathcal{M}}_{g,n}$ of stable curves of arithmetic genus $g$ with $n$ marked points is a compact orbifold, obtained by adjoining to $\mathcal{M}_{g,n}$ a divisor with normal crossings \cite{Knudsen}. We will only be interested in the genus zero case, in which case $\overline{\mathcal{M}}_{0,n}$ is actually a projective variety. Another reference for the genus $0$ case is \cite{Kapranov}. A stable curve $\Sigma$ determines a graph $\Gamma_\Sigma$, called the dual graph of $\Sigma$: the vertices of $\Gamma_\Sigma$ are the components of the subvariety of smooth points of $\Sigma$, the edges are the double points, and $\operatorname{Leg}(\Gamma_\Sigma)$ is the set of marked points. In particular, if $\Sigma$ has arithmetic genus $0$ and $n$ marked points, then $\Gamma_\Sigma\in\mathcal{T}\(n\)$. The moduli space $\overline{\mathcal{M}}_{0,n}$ is a stratified space: it has one stratum $\mathcal{M}\(T\)$ for each tree $T\in\mathcal{T}\(n\)$, consisting of all curves $\Sigma\in\overline{\mathcal{M}}_{0,n}$ such that $\Gamma_\Sigma=T$. The stratum $\mathcal{M}\(T\)$ is isomorphic to the product $$ \prod_{v\in\operatorname{Vert}(T)} \mathcal{M}_{0,|v|} , $$ and has codimension equal to the number of edges of $T$. For example $\overline{\mathcal{M}}_{0,4}$ has the following four strata: $$ \setlength{\unitlength}{0.0075in}% \begin{picture}(800,130)(25,0) \put(100,5){\begin{picture}(80,102)(80,650) \put( 80,740){\line( 1,-1){ 80}} \put(160,740){\line(-1,-1){ 80}} \put( 75,745){0} \put(155,745){2} \put( 75,640){3} \put(156,640){1} \end{picture}} \put(250,0){\begin{picture}(125,107)(55,605) \put( 60,700){\line( 1,-1){ 40}} \put(100,660){\line(-1,-1){ 40}} \put(100,660){\line( 1, 0){ 40}} \put(140,660){\line( 1, 1){ 40}} \put(140,660){\line( 1,-1){ 40}} \put( 55,705){0} \put(175,705){3} \put( 55,600){1} \put(175,600){2} \end{picture}} \put(425,0){\begin{picture}(125,107)(55,605) \put( 60,700){\line( 1,-1){ 40}} \put(100,660){\line(-1,-1){ 40}} \put(100,660){\line( 1, 0){ 40}} \put(140,660){\line( 1, 1){ 40}} \put(140,660){\line( 1,-1){ 40}} \put( 55,705){0} \put(175,705){1} \put( 55,600){2} \put(175,600){3} \end{picture}} \put(600,0){\begin{picture}(125,107)(55,605) \put( 60,700){\line( 1,-1){ 40}} \put(100,660){\line(-1,-1){ 40}} \put(100,660){\line( 1, 0){ 40}} \put(140,660){\line( 1, 1){ 40}} \put(140,660){\line( 1,-1){ 40}} \put( 55,705){0} \put(175,705){1} \put( 55,600){3} \put(175,600){2} \end{picture}} \end{picture}$$ If $T\in\mathcal{T}_k(n)$, denote by $\overline{\mathcal{M}}\(T\)$ the closure of the stratum $\mathcal{M}\(T\)$ of $\overline{\mathcal{M}}_{0,n}$, and by $[\overline{\mathcal{M}}\(T\)]$ the corresponding cycle in $H_{2(n-k-3)}(\overline{\mathcal{M}}_{0,n})$. The following theorem is due to Keel \cite{Keel}. \begin{theorem} \label{Keel:1} The cycles $[\overline{\mathcal{M}}\(T\)]$, $T\in\mathcal{T}\(n\)$ span $H_{\bullet}(\overline{\mathcal{M}}_{0,n})$. \end{theorem} In Section \ref{Keel:proof}, we give a proof of this theorem which differs from Keel's, and uses mixed Hodge theory. \subsection{The gravity operad} Let $\mathcal{G}\mathit{rav}$ be the stable cyclic $\SS$-module $$ \mathcal{G}\mathit{rav}\(n\) = \begin{cases} \Sigma^{3-n}\operatorname{sgn}_n\o H_{\bullet}(\mathcal{M}_{0,n}) , & n\ge3 , \\ 0 , & n<3 . \end{cases}$$ Note that $\mathcal{G}\mathit{rav}\(n\)$ is concentrated in degrees $3-n\le i\le 0$. There is a natural cyclic operad structure on $\mathcal{G}\mathit{rav}$. To define the product $$\begin{CD} \circ_i : \mathcal{G}\mathit{rav}\(m+1\)\o\mathcal{G}\mathit{rav}\(n+1\) @>>> \mathcal{G}\mathit{rav}\(m+n\) , \end{CD}$$ consider the embedding $j$ of $\mathcal{M}_{0,\{0,\dots,m\}}\times\mathcal{M}_{0,\{0',\dots,n'\}}$ as a stratum of $\overline{\mathcal{M}}_{0,\{0,\dots,\hat{\imath},\dots,m,1',\dots,n'\}}$, corresponding to the joining of the point labelled $i$ in the curve $\Sigma_1\in\mathcal{M}_{0,\{0,\dots,m\}}$ to the point labelled $0'$ in the curve $\Sigma_2\in\mathcal{M}_{0,\{0',\dots,n'\}}$. Consider the Poincar\'e residue map associated to this embedding \cite{Deligne}: $$\begin{CD} \operatorname{Res} : H^{\bullet}(\mathcal{M}_{0,\{0,\dots,\hat{\imath},\dots,m,1',\dots,n'\}}) @>>> H^{\bullet}(\mathcal{M}_{0,\{0,\dots,m\}}\times\mathcal{M}_{0,\{0',\dots,n'\}}) \end{CD}$$ Suitably suspending the adjoint of this map, we obtain the product $\circ_i$ of $\mathcal{G}\mathit{rav}$. This construction makes it quite obvious that $\mathcal{G}\mathit{rav}$ satisfies the equivariance and associativity axioms of a cyclic operad. Denote by $[x_1,\dots,x_n]$ the element of $\mathcal{G}\mathit{rav}(n)$ of degree $2-n$ corresponding to the standard basis vector of $H_0(\mathcal{M}_{0,n+1})$. This operation is graded antisymmetric, since $\SS_{n+1}$ acts on it by the sign representation, and it is proved in \cite{weil} that this sequence of operations generates $\mathcal{G}\mathit{rav}$, and that all relations are generated by the quadratic relations of \eqref{gravity}. Note that in that paper, we work with the operad $\Lambda^{-1}\mathcal{G}\mathit{rav}$, and the generators are all in degree $1$: the relationship between these two sets of generators at the level of algebras is $$ \{x_1,\dots,x_n\} = (-1)^{(n-1)|x_1|+(n-2)|x_2|+\dots+|x_{n-1}|} \Sigma^{-1} [\Sigma x_1,\dots,\Sigma x_n] . $$ In that paper, the operad structure is constructed in a different, though equivalent, way, using $\mathbb{C}^\times$-equivariant homology. \subsection{$\mathcal{G}\mathit{rav}$ as a mixed Hodge operad} Operads (and, more specifically, stable cyclic operads) may be defined in any symmetric monoidal category with colimits. Up to this point, we have concentrated on the examples of operads in the categories of chain complexes (differential graded operads) and topological spaces (topological operads). However, the category of mixed Hodge complexes \cite{BZ} is a symmetric monoidal category with colimits, with graded tensor product as the monoidal structure, and operads in this category are called mixed Hodge operads. In fact, the mixed Hodge operads which most concern us are pure and have vanishing differential. The gravity operad is an example of a mixed Hodge operad. This carries a unique mixed Hodge structure compatible with the Poincar\'e residue maps which define the products in $\mathcal{G}\mathit{rav}$: $$ \mathcal{G}\mathit{rav}\(n\) = \Sigma^{3-n} \operatorname{sgn}_n \o H_{\bullet}(\mathcal{M}_{0,n},\mathbb{C}(n-3)) . $$ Here, $\mathbb{C}(n-3)$ is the Tate Hodge structure, which is a line with Hodge numbers $(n-3,n-3)$. \subsection{The hypercommutative operad} Let $\mathcal{H}\mathit{ycom}$ be the stable cyclic $\SS$-module $$ \mathcal{H}\mathit{ycom}\(n\) = \begin{cases} H_{\bullet}(\overline{\mathcal{M}}_{0,n}) , & n\ge3 , \\ 0 , & n<3 . \end{cases}$$ The cyclic $\SS$-space $\overline{\mathcal{M}}\(n\)=\overline{\mathcal{M}}_{0,n}$ is a topological cyclic operad: the product is given by gluing stable curves together at marked points. It follows that $\mathcal{H}\mathit{ycom}$ is a cyclic operad. Kontsevich and Manin found \cite{KM} that algebras over $\mathcal{H}\mathit{ycom}$ are just hypercommutative algebras in the sense of \eqref{hypercommutative}, where the operation $(x_1,\dots,x_n)\in\mathcal{H}\mathit{ycom}\(n+1\)$ corresponds to the fundamental class $[\overline{\mathcal{M}}_{0,n+1}]\in H_{2(n-2)}(\overline{\mathcal{M}}_{0,n+1})$. The fact that the operations $(x_1,\dots,x_n)$ generate $\mathcal{H}\mathit{ycom}$ is an elegant restatement of Theorem \ref{Keel:1}. In Proposition \ref{orthogonal}, we will give a new proof (obtained jointly with Kontsevich) of the relations between the generators of $\mathcal{H}\mathit{ycom}$, which relies on the duality between hypercommutative algebras and gravity algebras, together with our explicit presentation \eqref{gravity} of the gravity operad $\mathcal{G}\mathit{rav}$. \subsection{A spectral sequence of Deligne} Let $M$ be a smooth projective variety of complex dimension $n$, and let $\{D_1,\dots,D_N\}$ be a sequence of smooth divisors with normal crossings; we denote their union by $D$. The sheaf of logarithmic differential forms $\mathcal{E}^{\bullet}_M(\log D)$ on $M$ is generated over the sheaf of differential forms $\mathcal{E}^{\bullet}_M$ by symbols $d(\log f)$, where $f$ is a section of $\mathcal{O}(D)$, subject to the relations $$ d(\log fg) = d(\log f) + d(\log g) \quad\text{and}\quad f\*d(\log f) = df. $$ The sheaf $\mathcal{E}^{\bullet}_M$ has a differential $d$, characterized by $d(d(\log f))=0$, and we have the fundamental isomorphism $$ H^{\bullet}(U) \cong H^{\bullet}(M,\mathcal{E}^{\bullet}_M(\log D)) , $$ where $U=M\setminus D$. The sheaf $\mathcal{E}^{\bullet}_M(\log D)$ is filtered by subsheaves $$ W_k\mathcal{E}^{\bullet}_M(\log D) = \operatorname{span}_\mathcal{O}\{ d(\log f_1)\.\dots\.d(\log f_i) \mid i\le k \} . $$ Let $j^k:D^k\hookrightarrow M$ be the embedding of the closed subvariety $$ D^k = \coprod_{i_1<\dots<i_k} D_{i_1} \cap \dots \cap D_{i_k} , $$ and let $\epsilon_k$ be the locally constant line bundle over $D^k$, which over the component $D_{i_1}\cap\ldots\cap D_{i_k}$ equals the determinant line $\det(\{i_1,\dots,i_k\})$. There is a canonical quasi-isomorphism $$ \operatorname{gr}^W_k \mathcal{E}^{\bullet}_M(\log D) \simeq \Sigma^{-k}j^k_*\mathcal{E}^{\bullet}_{D^k}\o\epsilon_k . $$ The associated spectral sequence $$ E_1^{-p,q} = H^{2p+q}(D^p,\epsilon_p) \Rightarrow E_\infty^{-p,q} = \operatorname{gr}^W_pH^{-p+q}(U) $$ carries a Hodge filtration $F$, induced by the Hodge filtration of $\mathcal{E}^{\bullet}_M(\log D)$, and by the principal of two types, $E_2=E_\infty$ (\cite{Deligne}, Section 3.2). The weight filtration induced on $H^n(U)$ by $W$ defines, up to translation, its mixed Hodge structure: $\operatorname{gr}^W_pH^{-p+q}(M)$ carries a pure Hodge structure of weight $q$. If a finite group $\Gamma$ acts on $M$, preserving $U\subset M$, this spectral sequence carries an action of $\Gamma$ compatible with its action on $H^{\bullet}(U)$. \subsection{The cohomology ring of $\mathcal{M}_{0,n}$} We now describe the cohomology ring of the moduli space $\mathcal{M}_{0,n}$. Our main tool is Arnold's description of the cohomology ring of the configuration space $\mathbb{C}^n_0$ \cite{Arnold}. If $1\le j\ne k\le n$, let $\omega_{jk}$ be the logarithmic differential form on the configuration space $\mathbb{C}^n_0$ given by the formula $$ \omega_{jk} = \frac{d\log(z_j-z_k)}{2\pi i} . $$ Note that the cohomology class of $\omega_{jk}$ is integral. \begin{proposition} \label{Arnold} The cohomology ring $H^{\bullet}(\mathbb{C}^n_0,\mathbb{Z})$ is the graded commutative ring with generators $[\omega_{jk}]$, and relations $\omega_{jk}=\omega_{kj}$ and $\omega_{ij}\omega_{jk}+\omega_{jk}\omega_{ki}+\omega_{ki}\omega_{ij}=0$. The symmetric group $\SS_n$ acts on $H^{\bullet}(\mathbb{C}^n_0,\mathbb{Z})$ through its action on the generators $\sigma\*\omega_{ij}=\omega_{\sigma(i)\sigma(j)}$. \end{proposition} \begin{pf} The Serre spectral sequence for the fibration $$\begin{CD} \mathbb{C}\setminus\{z_1,\dots,z_n\} @>>> \mathbb{C}^{n+1}_0 @>>> \mathbb{C}^n_0 \end{CD}$$ defined by projecting $(z_1,\dots,z_{n+1})$ to $(z_1,\dots,z_n)$ collapses at $E_2$, and the monodromy of $\pi_1(\mathbb{C}^n_0)$ on $H^{\bullet}(\mathbb{C}\setminus\{z_1,\dots,z_n\})$ is trivial. The proof now proceeds by induction on $n$. \end{pf} \begin{corollary} \label{equivariant} The cohomology ring $H^{\bullet}(\mathcal{M}_{0,n+1},\mathbb{C})$ may be identified with the kernel of the differential $\iota$ on $H^{\bullet}(\mathbb{C}^n_0,\mathbb{C})$ whose action on the generators is $\iota\omega_{jk}=1$. \end{corollary} \begin{pf} The isotropy group of the point $\infty\in\mathbb{CP}^1$ under the action of $\operatorname{PSL}(2,\mathbb{C})$ is $$\textstyle \operatorname{Aff}(\mathbb{C}) = \Bigl\{ \left( \begin{smallmatrix} a & b \\ 0 & a^{-1} \end{smallmatrix} \right) \mathop{\Big|} a\in\mathbb{C}^\times , b\in\mathbb{C} \Bigr\} \subset \operatorname{PSL}(2,\mathbb{C}) . $$ Since $\operatorname{PSL}(2,\mathbb{C})$ acts transitively on $\mathbb{CP}^1$, we see that $\mathcal{M}_{0,n+1}\cong\mathbb{C}^n_0/\operatorname{Aff}(\mathbb{C})$. But the group $\operatorname{Aff}(\mathbb{C})$ is homotopy equivalent to the circle group $$\textstyle \Bigl\{ \left( \begin{smallmatrix} a & 0 \\ 0 & a^{-1} \end{smallmatrix} \right) \mathop{\Big|} |a|=1 \Bigr\} , $$ giving a homotopy equivalence $\mathbb{C}^n_0\simeq\mathcal{M}_{0,n+1}\times S^1$. This allows us to identify the cohomology of $\mathcal{M}_{0,n+1}$ with the $S^1$-equivariant cohomology of $\mathbb{C}^n_0$. The infinitesimal generator of the circle action on $\mathbb{C}^n_0$ is the vector field $$ T = 2\pi i \sum_{k=1}^n \bigl( z_k \partial_k - \bar{z}_k \bar\partial_k \bigr) , $$ whose contraction with a generator $\omega_{jk}$ is $\omega_{jk}(T)=1$. \end{pf} The above result leads to yet another realization of the $\SS_n$-module structure on ${\operatorname{\mathcal{L}\mathit{ie}}}\(n\)\cong{\operatorname{\mathcal{L}\mathit{ie}}}(n-1)$: $$ {\operatorname{\mathcal{L}\mathit{ie}}}\(n\) \cong \operatorname{sgn}_n \o H_{n-3}(\mathcal{M}_{0,n}) . $$ \subsection{Application to $\mathcal{M}_{0,n}\subset\overline{\mathcal{M}}_{0,n}$} \label{Deligne} We now apply Deligne's spectral sequence with $U=\mathcal{M}_{0,n}$ and $M=\overline{\mathcal{M}}_{0,n}$. (This spectral sequence is also discussed in \cite{GK}, Section 3.4.5.) Denote the closure of the open stratum $\mathcal{M}\(T\)$ by $\overline{\mathcal{M}}\(T\)$. Then the divisors are the closed strata $\mathcal{M}\(T\)$, $T\in\mathcal{T}_1\(n\)$, while $D^p$ is the union of the closed strata $\overline{\mathcal{M}}\(T\)$, $T\in\mathcal{T}_p\(n\)$. The restriction of $\epsilon_p$ to $\overline{\mathcal{M}}\(T\)$ equals $\det(\operatorname{Edge}(T))$; as this is naturally isomorphic to $\det(T)$, we see that $$ E_1^{-p,q} \cong \bigoplus_{T\in\mathcal{T}_p\(n\)} H^{-2p+q}(\overline{\mathcal{M}}\(T\),\det(T)) . $$ The differential $\begin{CD}d_1:E_1^{-p,q}@>>>E_1^{-p+1,q} \end{CD}$ is easy to describe: it is the composition $$\begin{CD} \displaystyle \bigoplus_{T\in\mathcal{T}_p\(n\)} H^{-2p+q}(\overline{\mathcal{M}}\(T\),\det(T)) @. \displaystyle \bigoplus_{T\in\mathcal{T}_{p-1}\(n\)} H^{-2p+2+q}(\overline{\mathcal{M}}\(T\),\det(T)) \\ @| @| \\ \displaystyle \bigoplus_{T\in\mathcal{T}_p\(n\)} H_{2(n-3)-q}(\overline{\mathcal{M}}\(T\),\det(T)) @>>> \displaystyle \bigoplus_{T\in\mathcal{T}_{p-1}\(n\)} H_{2(n-3)-q}(\overline{\mathcal{M}}\(T\),\det(T)) \end{CD}$$ where the vertical isomorphisms are induced by Poincar\'e duality, and the bottom arrow is the map induced on the homology groups by the inclusion of the $p$-codimensional closed strata of $\overline{\mathcal{M}}_{0,n}$ into the $(p-1)$-codimensional closed strata. The key to unlocking this spectral sequence is the following lemma, which shows that $$ E_2^{-k,2k} \cong H^k(\mathcal{M}_{0,n}) , $$ while $E_2^{pq}=0$ if $2p+q\ne0$. \begin{lemma} \label{mixed} The mixed Hodge structure of $H^k(\mathcal{M}_{0,n})$, $n\ge3$, is pure of weight $2k$. In degree $i$, $\mathcal{G}\mathit{rav}\(n\)$ equals the $\SS_n$-module $\operatorname{sgn}_n\o H_{i+n-3}(\mathcal{M}_{0,n})$, with pure Hodge structure of weight $-2i$. \end{lemma} \begin{pf} By Proposition \ref{Arnold}, the cohomology ring of $\mathbb{C}^n_0$ is generated by the logarithmic differential forms $\omega_{ij}$; it follows that the mixed Hodge structure of $H^k(\mathbb{C}^n_0)$ is pure of weight $2k$. By Corollary \ref{equivariant}, there is an injection of the cohomology of $\mathcal{M}_{0,n+1}$ into the cohomology of $\mathbb{C}^n_0$, induced by the quotient map $\begin{CD}\mathbb{C}^n_0@>>>\mathcal{M}_{0,n+1} \end{CD}$, and the result follows. \end{pf} Thus, for $q$ even, the $E_1$-term of the spectral sequence gives rise to a resolution of the graded vector space $H^{\bullet}(\mathcal{M}_{0,n})$ \begin{equation} \label{bar:even} \begin{CD} 0 @>>> H^p(\mathcal{M}_{0,n}) @>>> \bigoplus_{T\in\mathcal{T}_p\(n\)} H^0(\overline{\mathcal{M}}\(T\),\det(T)) @>>> \bigoplus_{T\in\mathcal{T}_{p-1}\(n\)} H^2(\overline{\mathcal{M}}\(T\),\det(T)) @>>> \dots \end{CD} \end{equation} For $q$ odd, we obtain the exact sequence $$\begin{CD} \dots @>>> \bigoplus_{T\in\mathcal{T}_1\(n\)} H^{q-2}(\overline{\mathcal{M}}\(T\),\det(T)) @>>> H^q(\overline{\mathcal{M}}_{0,n}) = \bigoplus_{T\in\mathcal{T}_0\(n\)} H^q(\overline{\mathcal{M}}\(T\),\det(T)) @>>> 0 . \end{CD}$$ which shows, by induction on odd $q$, that the cohomology groups $H^q(\overline{\mathcal{M}}_{0,n})$ vanish if $q$ is odd. \subsection{Proof of Theorem \ref{Keel:1}} \label{Keel:proof} It is clear that the result holds for $n=3$, since $\overline{\mathcal{M}}_{0,3}$ is a point. We now argue by induction on $n$. As a consequence of the surjectivity of the differential in the exact sequence \eqref{bar:even} $$\begin{CD} \bigoplus_{T\in\mathcal{T}_1\(n\)} H^{2p-2}(\overline{\mathcal{M}}\(T\),\det(T)) @>>> H^{2p}(\overline{\mathcal{M}}_{0,n}) = \bigoplus_{T\in\mathcal{T}_0\(n\)} H^{2p}(\overline{\mathcal{M}}\(T\),\det(T)) \end{CD}$$ for $p\le n-2$, we see that all homology classes of $\overline{\mathcal{M}}_{0,n}$ except the fundamental class are supported on the closures of strata of codimension $1$. Such a closed stratum is isomorphic to $\overline{\mathcal{M}}_{0,i}\times\overline{\mathcal{M}}_{0,j}$ where $i+j=n+2$ and $i,j\ge3$, allowing us to apply the induction. }\end{pf \subsection{The dimension of $H^2(\overline{\mathcal{M}}_{0,n})$} Another consequence of \eqref{bar:even} is a simple formula for the dimension $\dim H^2(\overline{\mathcal{M}}_{0,n})$ of the Picard variety of $\overline{\mathcal{M}}_{0,n}$: \begin{eqnarray*} \dim H^2(\overline{\mathcal{M}}_{0,n}) &=& \sum_{k=3}^{n-1} \binom{n-1}{k} = 2^{n-1} - \frac{n^2-n+2}{2} \\ &=& \binom{n}{n-4} + \binom{n}{n-6} + \dots \end{eqnarray*} \begin{pf} When $p=1$, the short exact sequence \eqref{bar:even} becomes $$\begin{CD} 0 @>>> H^1(\mathcal{M}_{0,n}) @>>> \bigoplus_{T\in\mathcal{T}_1\(n\)} H^0(\overline{\mathcal{M}}\(T\),\det(T)) @>>> H^2(\overline{\mathcal{M}}_{0,n}) @>>> 0 . \end{CD}$$ By Corollary \ref{equivariant}, the dimension of $H^1(\mathcal{M}_{0,n})$ equals the coefficient of $-t$ in $(1-2t)\dots(1-(n-2)t)$, or $$ 2 + \dots + (n-1) = \binom{n-1}{2} - 1 . $$ (See \eqref{poincare-moduli} for more details of this calculation: we will actually show that $H^1(\mathcal{M}_{0,n})$ is isomorphic to the irreducible $\SS_n$-module $V_{n-2,2}$.) Each tree in $\mathcal{T}_1\(n\)$ contributes a copy of $\mathbb{C}$ to $\bigoplus_{T\in\mathcal{T}_1\(n\)} H_{2(n-4)}(\overline{\mathcal{M}}\(T\))$. Let $T$ be such a tree and consider the set $S$ of the external edges attached to one vertex of $T$. We see that trees with two vertices correspond to subsets $S\subset\{1,\dots,n\}$ where $2\le|S|\le n-2$, where we identify the trees corresponding to the subsets $S$ and $S^c$, the complement of $S$. Thus, $$ |\mathcal{T}_1\(n\)| = \frac12 \sum_{k=2}^{n-2} \binom{n}{k} = \frac12 (2^n-2n-2) = 2^{n-1} - n - 1 . $$ The result follows easily. \end{pf} This formula may be compared to the dimension of $H^2(\overline{\mathcal{M}}_{g,n})$, $g>2$, which follows from the work of Arbarello and Cornalba \cite{AC}: $$ \dim H^2(\overline{\mathcal{M}}_{g,n}) = 2^{n-1} (g+1) + n + 1 . $$ We see that this formula is correct for $g=0$ up to a polynomial error. The above dimension formula may be refined, using the realization of $\overline{\mathcal{M}}_{0,n}$ as an iterated blowup, to show that as an $\SS_n$-module, $H^2(\overline{\mathcal{M}}_{0,n})$ is the direct sum of the suitable exterior powers of the permutation representation $\mathbb{C}^n$ of $\SS_n$. Note that there are the same number of $n$-linear relations \eqref{gravity} among the brackets $[x_1,\dots,x_k]$ generating the gravity operad as $\dim H^2(\overline{\mathcal{M}}_{0,n+1})$; as we will see, this is no coincidence. \section{Koszul operads} In this section, we prove our main theorem, the duality of the hypercommutative and gravity operads. To do this, we must generalize Ginzburg and Kapranov's theory of Koszul operads \cite{GK} so that it applies to operads which are not necessarily generated by bilinear operations. First, we recall their cobar construction for operads, an analogue of Hochschild's bar construction for associative algebras. The dual of an operad is only defined up to homotopy, and is represented by the cobar operad. However, there is a class of operads, the Koszul operads, for which there is a particularly nice dual, whose generators are in one-to-one correspondence with those of the original operad. A Koszul operad is quadratic (the relations among its generators are bilinear), as is its dual, and the relations in the dual operad may be characterized as the orthogonal complement of those of the original operad. \subsection{Free operads and trees} We now recall from \cite{cyclic} the structure of the free cyclic operad $\mathbb{T}_+\v$ generated by a cyclic $\SS$-module $\v$. There is an analogous construction for operads, for which we refer to \cite{n-algebras}. From now on, all cyclic $\SS$-modules which we discuss will be stable. If $\v$ is a (stable) cyclic $\SS$-module, let $\mathbb{T}_+\v$ be the (stable) cyclic $\SS$-module defined by $$ \mathbb{T}_+\v\(n\) = \bigoplus_{T\in\mathcal{T}\(n\)} \v\(T\) , $$ where $\v\(T\)=\bigotimes_{v\in\operatorname{Vert}(T)} \v\(\operatorname{Leg}(v)\)$. Note that $\mathbb{T}_+\v$ is graded by subspaces $$ \mathbb{T}_i\v\(n\) = \bigoplus_{T\in\mathcal{T}_i\(n\)} \v\(T\) . $$ Then $\mathbb{T}_+$ is an endofunctor in the category of (stable) cyclic $\SS$-modules. There is a natural structure of a triple on the functor $\mathbb{T}_+$: \begin{enumerate} \item since $\mathbb{T}_+\mathbb{T}_+$ is a sum over trees, each vertex of which is itself a tree, the product of the triple is a natural transformation from $\mathbb{T}_+\mathbb{T}_+$ to $\mathbb{T}_+$ obtained by gluing the trees at the vertices into the larger tree; \item the unit of the triple is the natural transformation from the identity functor to $\mathbb{T}_+$ induced by the inclusion $\mathcal{T}_0\(n\)\subset\mathcal{T}\(n\)$. \end{enumerate} The following theorem is a melding of results from \cite{cyclic} and \cite{modular}. \begin{theorem} A (non-unital, stable) cyclic operad is the same thing as a $\mathbb{T}_+$-algebra in the category of (stable) cyclic $\SS$-modules. \end{theorem} \subsection{The cobar construction for operads} The cobar construction $\mathsf{B}$, introduced by Ginzburg and Kapranov \cite{GK}, is a contravariant functor on the category of operads. We study here the slight variant of this functor which acts on the category of (non-unital, stable) cyclic operads. The dual $V^*$ of a chain complex $V$ is defined as follows: $\begin{CD}V^*_i=(V_{-i})^*$, and $\delta^*:V^*_i@>>>V^*_{i-1} \end{CD}$ is the adjoint of $\begin{CD}\delta:V_{-i+1}@>>>V_{-i}\end{CD}$. If $\v$ is a stable cyclic $\SS$-module, denote by $\v^\vee$ the stable cyclic $\SS$-module $$ \v^\vee\(n\) = \Sigma^{n-3}\operatorname{sgn}_n\o\v\(n\)^* . $$ This functor is an involution on the category of stable cyclic $\SS$-modules, that is, $(\v^\vee)^\vee$ is naturally isomorphic to $\v$. The cobar operad $\mathsf{B}\a$ of a (non-unital, stable) cyclic operad is obtained by perturbing the differential of the free cyclic operad $\mathbb{T}_+\a^\vee$ by a differential $\partial$ which reflects the operad structure of $\a$, and is defined as follows. If $T\in\mathcal{T}\(n\)$, and $e$ is an edge of $T$, denote by $T/e$ the tree in which $e$ is contracted to a point: thus, $T/e$ has one fewer vertices, and one fewer edges, than $T$. There is a natural map of degree $0$ $$\begin{CD} \partial_{T/e} : \a\(T\) @>>> \a\(T/e\) \end{CD}$$ induced by composition in the operad $\a$ along the edge $e$. This induces a map $$\begin{CD} \partial^\vee_{T/e} : \a^\vee\(T/e\) @>>> \a^\vee\(T\) \end{CD}$$ of degree $-1$. We now define the differential $\partial$ to be the operator whose matrix element from $\a^\vee\(\tilde{T}\)\subset\mathsf{B}\a$ to $\a^\vee\(T\)\subset\mathsf{B}\a$ is the sum of the operators $\partial^\vee_{T/e}$ over internal edges $e$ such that $T/e$ is isomorphic to $\tilde{T}$. Paying careful attention to the signs coming from the suspensions, one shows that the differential $\partial$ satisfies the formulas $\partial^2=\delta\partial+\partial\delta=0$, and hence that $\delta+\partial$ is a differential on $\mathbb{T}_+\a^\vee$. It is also not hard to show that $\partial$ is compatible with the cyclic operad structure of $\mathbb{T}_+\a^\vee$, so that $\mathsf{B}\a=(\mathbb{T}_+\a^\vee,\delta+\partial)$ is an operad. The properties of the resulting functor are summarized by the following theorem. \begin{theorem} \label{cobar} \textup{(1)} The cobar construction is a homotopy functor, that is, if $\begin{CD}f:\a@>>>\b \end{CD}$ is a homotopy equivalence, then so is $\begin{CD} \mathsf{B} f:\mathsf{B}\a@>>>\mathsf{B}\b\end{CD}$. \noindent \textup{(2)} There is a natural transformation from $\mathsf{B}\BB$ to the identity functor, and the resulting map $\begin{CD}\mathsf{B}\BB\a@>>>\a \end{CD}$ is a homotopy equivalence for all $\a$. \end{theorem} \begin{pf} The homotopy invariance of $\mathsf{B}$ is easy to see by a double complex argument. The natural map from $\mathsf{B}\BB\a$ to $\a$, which is projection onto the summand $\a\cong(\a^\vee)^\vee\subset\mathbb{T}_+(\mathbb{T}_+\a^\vee)^\vee$, is shown to be a homotopy equivalence of operads in Theorem 3.2.16 of \cite{GK}. \end{pf} If $\a$ and $\b$ are operads and $\begin{CD}\Phi:\mathsf{B}\a@>>>\b \end{CD}$ is a morphism of operads, there is a bar construction $\mathcal{B}_\Phi$ on $\a$-algebras, defined for an $\a$-algebra $A$ by twisting the differential on the chain complex $$ A \oplus \Sigma^{-1} \bigoplus_{n=3}^\infty \operatorname{Hom}_{\SS_n}(\b(n),(\Sigma A)^{\o n}) $$ in such a way as to reflect the $\a$-algebra structure of $A$. (See \cite{n-algebras} for details.) When $\b=\mathsf{B}\a$ and $\Phi$ is the identity map, we denote the resulting functor $\mathcal{B}$. Let $QA$ be the complex of indecomposables, obtained by taking the cokernel of the map $$\begin{CD} \bigoplus_{n=3}^\infty \rho_n : \bigoplus_{n=3}^\infty \a(n)\o_{\SS_n} A^{\o n} @>>> A . \end{CD}$$ The following theorem is proved in Chapter 3 of \cite{n-algebras}. \begin{theorem} \label{Bar} \textup{(1)} There is a natural transformation of functors $\begin{CD} \mathcal{B}@>>>Q\end{CD}$, such that if $A$ is a free algebra, the morphism $\begin{CD}\mathcal{B} A@>>>QA \end{CD}$ is a homotopy equivalence. \noindent \textup{(2)} The functors $\mathcal{B}_\Phi$ are homotopy functors: if $\begin{CD}f:A@>>>B \end{CD}$ is a homotopy equivalence, then so is $\begin{CD}\mathcal{B}_\Phi f:\mathcal{B}_\Phi A@>>>\mathcal{B}_\Phi B\end{CD}$. \end{theorem} If $\begin{CD}\Phi:\mathsf{B}\a@>>>\b \end{CD}$ is a homotopy equivalence of operads, the natural morphism $\begin{CD}\mathcal{B} A@>>>\mathcal{B}_\Phi A \end{CD}$ is a homotopy equivalence for all $\a$-algebras $A$. Thus, the functor $\mathcal{B}_\Phi$ is a left derived functor $\L Q$ of the indecomposable functor $Q$, that is, a homotopy functor homotopy equivalent to $Q$ on free $\a$-algebras. It is proved in \cite{GK} that there are natural homotopy equivalences of operads $$\begin{CD} \mathsf{B}{\operatorname{\mathcal{A}\mathit{ss}}} @>>> {\operatorname{\mathcal{A}\mathit{ss}}} \quad \mathsf{B}{\operatorname{\mathcal{C}\mathit{om}}} @>>> {\operatorname{\mathcal{L}\mathit{ie}}} \quad \mathsf{B}{\operatorname{\mathcal{L}\mathit{ie}}} @>>> {\operatorname{\mathcal{C}\mathit{om}}} . \end{CD}$$ The bar construction associated to the first of these homotopy equivalences is, up to a shift in degree, the Hochschild bar construction on associative algebras, while the bar constructions associated to the other two homotopy equivalences are the functors $\L Q$ on commutative and Lie algebras discussed in the introduction, equal, up to a shift in degree, to the Harrison and Chevalley-Eilenberg complexes respectively. Thus, the duality result Theorem \ref{Harrison} is seen to be a special case of Theorem \ref{Bar}. In Theorem \ref{Kontsevich}, we will prove that $\mathcal{H}\mathit{ycom}$-algebras are the same thing as hypercommutative algebras in the sense of \eqref{hypercommutative}. Thus the duality between hypercommutative and gravity algebras announced in the introduction follows from Theorem \ref{Bar} combined with the following generalization of Theorem 4.25 of \cite{GK}. \begin{theorem} \label{duality} There is a natural homotopy equivalence of operads $\begin{CD} \mathsf{B}\mathcal{H}\mathit{ycom}@>>>\mathcal{G}\mathit{rav}\end{CD}$. \end{theorem} \begin{pf} Let $\v$ be the $\SS$-module obtained by summing the short exact sequences \eqref{bar:even} (minus the terms $H^p(\mathcal{M}_{0,n})$), placing the summand $$ \bigoplus_{T\in\mathcal{T}_p\(n\)} H^q(\overline{\mathcal{M}}\(T\),\det(T)) $$ of $\v\(n\)$ in degree $2(n-3)-p-q$. Using Poincar\'e duality, we see that $$ \v\(n\) \cong \bigoplus_{p=0}^{n-3} \bigoplus_{T\in\mathcal{T}_p\(n\)} \Sigma^p \det(T) \o \mathcal{H}\mathit{ycom}\(T\) . $$ Furthermore, there is a natural homotopy equivalence $\begin{CD}\Sigma^{2(n-3)}H^{\bullet}(\overline{\mathcal{M}}_{0,n})@>>>\v\(n\) \end{CD}$, which induces a homotopy equivalence $\begin{CD}\v^\vee@>>>\mathcal{G}\mathit{rav}\end{CD}$. By the isomorphism \eqref{det}, the $\SS$-module $\v^\vee$ may be rewritten as \begin{eqnarray*} \v^\vee\(n\) &\cong& \bigoplus_{p=0}^{n-3} \Sigma^{n-3-p} \bigoplus_{T\in\mathcal{T}_p\(n\)} \bigotimes_{v\in\operatorname{Vert}(T)} \det(\operatorname{Leg}(v)) \o \mathcal{H}\mathit{ycom}^*\(\operatorname{Leg}(v)\) \\ &\cong& \bigoplus_{p=0}^{n-3} \Sigma^{n-3-p} \bigoplus_{T\in\mathcal{T}_p\(n\)} \bigotimes_{v\in\operatorname{Vert}(T)} \Sigma^{3-|v|} \mathcal{H}\mathit{ycom}^\vee\(\operatorname{Leg}(v)\) \\ &\cong& \bigoplus_{p=0}^{n-3} \Sigma^{n-3-p+\sum_{v\in\operatorname{Vert}(T)}(3-|v|)} \bigoplus_{T\in\mathcal{T}_p\(n\)} \mathcal{H}\mathit{ycom}^\vee\(T\) . \end{eqnarray*} If $T\in\mathcal{T}_p\(n\)$, we have $$ \sum_{v\in\operatorname{Vert}(T)} (3-|v|) = 3 - n + p , $$ showing that $$ \v^\vee\(n\) \cong \bigoplus_{p=0}^{n-3} \bigoplus_{T\in\mathcal{T}_p\(n\)} \mathcal{H}\mathit{ycom}^\vee\(T\) \cong \mathsf{B}\mathcal{H}\mathit{ycom}\(n\) . $$ A little diagram chasing shows that the differentials of the $\SS$-modules $\v^\vee$ and $\mathsf{B}\mathcal{H}\mathit{ycom}$ are the same, and that the resulting homotopy equivalence between $\mathcal{G}\mathit{rav}$ and $\mathsf{B}\mathcal{H}\mathit{ycom}$ is compatible with the operad structures. \end{pf} \subsection{The cobar construction for mixed Hodge operads} The free operad functor $\v\mapsto\mathbb{T}\v$ and the functor $\begin{CD} \v@>>>\v^\vee\end{CD}$ have analogues in the category of mixed Hodge $\SS$-modules, defined in precisely the same way as in the category of $\SS$-modules. (We recall that the dual $\v^*$ in the category of mixed Hodge complexes reverses the weight filtration, sending complexes of weight $k$ to complexes of weight $-k$.) This allows us to extend the cobar construction to the category of mixed Hodge operads, by the same definition as in the category of dg operads. If we follow through the proof of Theorem \ref{duality} paying attention to the mixed Hodge structures, we see that the homotopy equivalence $\begin{CD}\mathsf{B}\mathcal{H}\mathit{ycom}@>>>\mathcal{G}\mathit{rav} \end{CD}$ is compatible with the Hodge structures of $\mathcal{G}\mathit{rav}$ and $\mathcal{H}\mathit{ycom}$, where $\mathcal{H}\mathit{ycom}$ carries the natural (pure) Hodge structure coming from its realization as the cohomology of the smooth K\"ahler manifold $\overline{\mathcal{M}}_{0,n}$. This observation will be essential in our calculation of the defining relations of the operad $\mathcal{H}\mathit{ycom}$. \subsection{Quadratic operads} We now generalize Ginzburg and Kapranov's notion of a Koszul operad to operads whose generators are not necessarily bilinear operations. Once more, we restrict attention to stable cyclic operads. An ideal $\b\subset\a$ of a cyclic operad is a cyclic $\SS$-submodule such that for all operations $\circ_i$, $a\circ_ib$ is in $\b$ if either $a$ or $b$ is. The intersection of two ideals is obviously an ideal. An ideal is generated by a cyclic $\SS$-submodule $\r\subset\b$ if $\b$ is the intersection of all ideals of $\a$ containing $\r$. Let $\a$ be an operad, generated by a cyclic $\SS$-submodule $\v$. The pair $(\a,\v)$ is a cyclic quadratic operad if the ideal $\begin{CD} \ker(\mathbb{T}\v@>>>\a)\end{CD}$ in the free cyclic operad $\mathbb{T}_+\v$ is generated by $$\begin{CD} \ker \Bigl( \mathbb{T}_1\v=\bigoplus_{T\in\mathcal{T}_1} \v\(T\) @>>> \a \Bigr) . \end{CD}$$ The word quadratic is used here because $\v\(T\)$ is quadratic in $\v$ if $T$ has one internal edge and hence two vertices. Thus, $\r$ is itself quadratic in $\v$. The cyclic operads $\a={\operatorname{\mathcal{A}\mathit{ss}}}$, ${\operatorname{\mathcal{C}\mathit{om}}}$ and ${\operatorname{\mathcal{L}\mathit{ie}}}$ are all quadratic, with generating cyclic submodule $\v$, where $$ \v\(n\) = \begin{cases} \a\(3\) , & n=3, \\ 0 , & n\ne3 . \end{cases} $$ For example, ${\operatorname{\mathcal{L}\mathit{ie}}}\(3\)$ is one-dimensional, spanned by $[a_1,a_2]$, and the cyclic $\SS$-module of relations $\r$ is given by the formula $$ \r\(n\) = \begin{cases} \span\{ [a_1,[a_2,a_3]] , [a_2,[a_3,a_1]] \} , & n=4 , \\ 0 , & n\ne4 . \end{cases} $$ \subsection{The naive dual of a quadratic operad} If $\a$ is a cyclic quadratic operad, the naive dual $\a^!$ of $\a$ is the cokernel of the composition $$\begin{CD} \psi : \mathsf{B}\a @>\partial>> \mathsf{B}\a @>>> \mathbb{T}_+\v^\vee , \end{CD}$$ where the second arrow is the surjection of cyclic $\SS$-modules $\begin{CD}\mathsf{B}\a@>>>\mathbb{T}_+\v^\vee \end{CD}$ induced by the inclusion of cyclic $\SS$-modules $\v\subset\a$. \begin{definition} If $\a$ is a cyclic quadratic operad, there is a natural morphism of operads $\begin{CD}\Phi:\mathsf{B}\a@>>>\a^!\end{CD}$, induced by the surjection $\begin{CD}\mathsf{B}\a@>>>\mathbb{T}_+\v^\vee \end{CD}$. The operad $\a$ is Koszul if the surjection of operads $\begin{CD}\mathsf{B}\a@>>>\a^! \end{CD}$ is a homotopy equivalence, or equivalently, if $\begin{CD}\mathsf{B}\a^!@>>>\a\end{CD}$ is. \end{definition} \begin{proposition} \label{orthogonal} If $\a$ is a cyclic quadratic operad with generators $\v$ and relations $\r$, let $\r^\perp$ be the kernel of the natural map from $\mathbb{T}_1\v^\vee\cong(\mathbb{T}_1\v^\vee)^\vee$ to $\r^\vee$. Then $\a^!$ is a cyclic quadratic operad with generators $\v^\vee$ and relations $\r^\perp$. \end{proposition} \begin{pf} It suffices to show that the image of $\mathsf{B}\a$ in $\mathbb{T}_+\v^\vee$ under the above composition is the ideal generated by $\r^\perp$. Denote by $\a_k\subset\a$ the image of $\mathbb{T}_k\v$ in $\a$ under the quotient map $\begin{CD}\mathbb{T}_+\v@>>>\a \end{CD}$. Thus, $\a_0=\v$ and $\a_1=\mathbb{T}_1\v/\r$. Observe that $\a_1^\vee\cong\r^\perp$. If $T\in\mathcal{T}\(n\)$, the summand $\a^\vee\(T\)$ of $\mathsf{B}\a\(n\)$ may be thought of as the vector space spanned by decorations of the tree $T$, in which each vertex of $T$ is assigned an element of $\a^\vee$ of appropriate valence. The map $\begin{CD}\psi:\mathsf{B}\a@>>>\mathbb{T}_+\v^\vee \end{CD}$ vanishes on such a decorated tree unless the vertex decorations lie in $\v$ at all but one vertex $v$, which is decorated by $a\in\a_1^\vee$. The map $\psi$ applied to this decorated tree produces a new decorated tree in which the vertex $v$ is replaced by the tree underlying $a$ (and thus having one additional edge). Thus, the image of $\psi$ is the ideal generated by $\r^\perp$. \end{pf} \begin{corollary} A cyclic quadratic operad $\a$ is Koszul if and only if $\a^!$ is, and $(\a^!)^!\cong\a$. \end{corollary} As examples of naive duals, we have ${\operatorname{\mathcal{A}\mathit{ss}}}^!\cong{\operatorname{\mathcal{A}\mathit{ss}}}$, ${\operatorname{\mathcal{C}\mathit{om}}}^!\cong{\operatorname{\mathcal{L}\mathit{ie}}}$ and ${\operatorname{\mathcal{L}\mathit{ie}}}^!\cong{\operatorname{\mathcal{C}\mathit{om}}}$. It is proved in \cite{GK} that the operads ${\operatorname{\mathcal{A}\mathit{ss}}}$, ${\operatorname{\mathcal{C}\mathit{om}}}$ and ${\operatorname{\mathcal{L}\mathit{ie}}}$ are Koszul. A non-cyclic example of a Koszul operad is $\mathcal{B}\mathit{raid}$ \cite{n-algebras}, which satisfies $\mathcal{B}\mathit{raid}^!\cong\Lambda^{-1}\mathcal{B}\mathit{raid}$. The proof of the following theorem occupies the remainder of this section. This theorem is joint work of the author and M. Kontsevich. \begin{theorem} \label{Kontsevich} Let $\v\subset\mathcal{H}\mathit{ycom}$ be the cyclic $\SS$-submodule spanned by the fundamental classes $$ [\overline{\mathcal{M}}_{0,n}] \in H_{2(n-3)}(\overline{\mathcal{M}}_{0,n}) \subset \mathcal{H}\mathit{ycom}\(n\) . $$ The operad $\mathcal{H}\mathit{ycom}$ is Koszul, with generators $\v$, and $\mathcal{H}\mathit{ycom}^!\cong\mathcal{G}\mathit{rav}$. \end{theorem} \begin{pf} This theorem is proved using the duality between the mixed Hodge operads $\mathcal{G}\mathit{rav}$ and $\mathcal{H}\mathit{ycom}$, and the fact that the operad $\mathcal{G}\mathit{rav}$ is quadratic. If $\a$ is a mixed Hodge operad, the natural homotopy equivalence $\begin{CD}\mathsf{B}\BB\a@>>>\a \end{CD}$ of Theorem \ref{cobar} is a morphism of mixed Hodge operads. It follows that we have a diagram in the category of mixed Hodge operads $$\begin{CD} \mathsf{B}\BB\mathcal{H}\mathit{ycom} @>>> \mathsf{B}\mathcal{G}\mathit{rav} \\ @VVV @. \\ \mathcal{H}\mathit{ycom} @. \end{CD}$$ in which both arrows are homotopy equivalences. The homology of the weight $-2p$ summand of the complex $\mathsf{B}\mathcal{G}\mathit{rav}\(n\)$, which by this argument is isomorphic to $H_{2p}(\overline{\mathcal{M}}_{0,n})$, must be concentrated in degree $2p$; from this, we see that this subcomplex is exact except at the last term, giving a long exact sequence \begin{equation} \label{bar:Hodge}\begin{CD} \dots @>>> \bigoplus_{T\in\mathcal{T}_{n-2-p}\(n\)} \mathcal{G}\mathit{rav}^\vee\(T\)_{2p+1} @>>> \bigoplus_{T\in\mathcal{T}_{n-3-p}\(n\)} \mathcal{G}\mathit{rav}^\vee\(T\)_{2p} @>>> \mathcal{H}\mathit{ycom}\(n\)_{2p} @>>> 0 . \end{CD}\end{equation} Let $\v^\vee\subset\mathcal{G}\mathit{rav}$ be the cyclic $\SS$-module spanned by the generators of $\mathcal{G}\mathit{rav}$; for each $n\ge3$, there is one generator, of degree $3-n$ and weight $2(n-3)$, in $\mathcal{G}\mathit{rav}\(n\)$. Thus $\v\(n\)$ is spanned by an element of degree $2(n-3)$ and weight $2(3-n)$. Taking $p=n-3$ in the long exact sequence \eqref{bar:Hodge}, we see that $\v\(n\)$ may be identified with $H_{2(n-3)}(\overline{\mathcal{M}}_{0,n})$. Let $\r^\vee\subset\mathcal{G}\mathit{rav}$ be the cyclic $\SS$-module spanned by the elements of degree $4-n$ in $\mathcal{G}\mathit{rav}$. An $\SS_n$-module, $\r^\vee\(n\)\cong H_1\(\mathcal{M}_{0,n}\)\o\operatorname{sgn}_n$; it is concentrated in degree $4-n$, has weight $2(n-4)$, and dimension $\binom{n-1}{2}-1$. The case $p=n-4$ of the long exact sequence \eqref{bar:Hodge} is the short exact sequence $$\begin{CD} 0 @>>> \r\(n\) @>>> \bigoplus_{T\in\mathcal{T}_1\(n\)} \v\(T\) @>>> H_{2(n-4)}(\overline{\mathcal{M}}_{0,n}) @>>> 0 , \end{CD}$$ showing that the $\SS$-module $\r$ is a subset of the set of relations for $\mathcal{H}\mathit{ycom}$. Furthermore, there are no further relations, as may be seen from \eqref{bar:Hodge}: for general $p\le n-3$, there is an exact sequence \begin{multline*}\begin{CD} \dots @>>> \bigoplus_{T\in\mathcal{T}_{n-2-p}\(n\)} \bigoplus_{v\in\operatorname{Vert}(T)} \r\(\operatorname{Leg}(v)\) \o \bigotimes_{w\in\operatorname{Vert}(T)\setminus\{v\}} \v\(\operatorname{Leg}(w)\) @>>> \\ \bigoplus_{T\in\mathcal{T}_{n-3-p}\(n\)} \v\(T\) @>>> H_{2p}\(n\)_{2p} @>>> 0 . \end{CD}\end{multline*} This shows that the operad $\mathcal{H}\mathit{ycom}$ is quadratic, with generators $\v$ and relations $\r$. It remains to identify the $\SS$-module $\r$ with the set of relations \eqref{hypercommutative} which hold in a hypercommutative algebra. This is done in two parts: we first show that these relations are are in the orthogonal complement of the relations which define a gravity algebra, and thus form a subset of $\r$, and then show that they form a subspace of $\r\(n\)$ of dimension at least $\binom{n-1}{2}-1$. Since $\r\(n\)$ itself has dimension $\binom{n-1}{2}-1$, this completes the proof. It is simple to check that the relations \eqref{hypercommutative} are orthogonal to those which hold in a gravity algebra. Consider the relation $$ G_0 = \pm [[a_i,a_j],a_1,\dots,\widehat{a_i},\dots,\widehat{a_j},\dots,a_k] . $$ The inner product of this relation with the relation $$ H = \sum_{S_1\coprod S_2=\{1,\dots,n\}} \pm ((a,b,x_{S_1}),c,x_{S_2}) - \sum_{S_1\coprod S_2=\{1,\dots,n\}} \pm (a,(b,c,x_{S_1}),x_{S_2}) $$ vanishes, since only the terms with $S_1=\emptyset$ can contribute: they each contribute a term $1$, but with opposite sign. Turning now to the relation $$ G_\ell = \pm [[a_i,a_j],a_1,\dots,\widehat{a_i},\dots,\widehat{a_j},\dots,a_k] - [[a_1,\dots,a_k],b_1,\dots,b_\ell] , \quad \ell>0 , $$ we see that there are three cases to consider: \begin{enumerate} \item none of the letters $a,b,c$ lie in the set $\{b_1,\dots,b_\ell\}$, in which case the inner product of relation $H$ with the above relation again vanishes, for the same reason as when $\ell=0$; \item one of the letters, say $c$, lies in the set $\{b_1,\dots,b_\ell\}$, in which case the only terms having a non-zero inner product with $H$ are $[[a,b],c,x_S]$ and $[[a_1,\dots,a_k],b_1,\dots,b_\ell]$, whose contributions, each equal to $1$, cancel; \item two or three of the letters $a,b,c$ lie in the set $\{b_1,\dots,b_\ell\}$, in which case the inner product of the above relation with each term of $H$ vanishes. \end{enumerate} Finally, we check that the space of relations among $n$ letters in a hypercommutative algebra has dimension at least $\binom{n}{2}-1$. Consider the projection of these relations into the space $\a(n)$ of all quadratic words in the generators of the hypercommutative operad of the form $$ ((x_i,x_j),x_1,\dots,\widehat{x_i},\dots,\widehat{x_j},\dots,x_n) . $$ (Note that this subspace of $\mathbb{T}\v^\vee\(n+1\)$ is not $\SS_{n+1}$-invariant, but only $\SS_n$-invariant.) The dimension of $\a(n)$ is $\binom{n}{2}$, and the relations \eqref{hypercommutative} project in $\a(n)$ to relations $$ ((a,b),c,x_S) = ((a,c),b,x_S) . $$ Clearly, the quotient of $\a(n)$ by these relations is one-dimensional. This completes the proof that \eqref{hypercommutative} are all the relations in the operad $\mathcal{H}\mathit{ycom}$. \end{pf} \section{The equivariant Poincar\'e polynomials of $\mathcal{M}_{0,n}$ and $\overline{\mathcal{M}}_{0,n}$} In this section, we use the results of \cite{modular} to calculate the character of the $\SS_n$-modules $H_i(\mathcal{M}_{0,n})$ and $H_i(\overline{\mathcal{M}}_{0,n})$. By and large, the results of this section are independent of the rest of this paper. \subsection{Symmetric functions} Let $\Lambda$ be the ring of symmetric functions: this is the limit $$ \Lambda = \varprojlim \mathbb{Z}\[x_1,\dots,x_k\]^{\SS_k} . $$ Then $\Lambda$ is the ring $\mathbb{Z}\[h_1,h_2,\dots\]$ of power series in the complete symmetric functions $$ h_n(x_i) = \sum_{i_1\le\dots\le i_n} x_{i_1}\dots x_{i_n} , $$ and $\Lambda_\mathbb{Q}=\Lambda\o\mathbb{Q}$ is a power series ring $\mathbb{Q}\[p_1,p_2,\dots\]$ in the power sums $$ p_n(x_i) = \sum_i x_i^n . $$ If $\sigma\in\SS_n$ has cycles of length $\lambda_1\ge\dots\ge\lambda_\ell$, its cycle index $\psi(\sigma)$ is the monomial $p_{\lambda_1}\dots p_{\lambda_\ell}$. If $V$ is an $\SS_n$-module, its characteristic is the symmetric function $$ \operatorname{ch}_n(V) = \frac{1}{n!} \sum_{\sigma\in\SS_n} \operatorname{Tr}_V(\sigma) \psi(\sigma) . $$ It may be shown that $\operatorname{ch}_n(V)\in\Lambda\subset\Lambda_\mathbb{Q}$, and that the characteristics of the irreducible representations of the symmetric groups $\SS_n$, $n\ge0$, form a basis of $\Lambda$ over $\mathbb{Z}$, called the Schur functions \cite{Macdonald}. For example, $h_n$ is the characteristic of the trivial representation of $\SS_n$. Define the Poincar\'e characteristic of an $\SS$-module to be $$ \operatorname{ch}_t(\v) = \sum_{n=0}^\infty \sum_{i=0}^\infty (-t)^i \, \operatorname{ch}_n(\v_i(n)) \in \Lambda\(t\) . $$ Setting $t=1$, we obtain the (Euler-Frobenius) characteristic $\operatorname{ch}(\v)$. For example, $$ \operatorname{ch}({\operatorname{\mathcal{C}\mathit{om}}}^+) = \sum_{n=1}^\infty h_n = \exp\Bigl( \sum_{n=1}^\infty \frac{p_n}{n} \Bigr) - 1 , \quad\text{and}\quad \operatorname{ch}({\operatorname{\mathcal{A}\mathit{ss}}}^+) = \sum_{n=1}^\infty p_1^n = \frac{p_1}{1-p_1} . $$ \subsection{Plethysm} Consider the ring $\Lambda\(t\)$ of power series in a variable $t$ with coefficients in $\Lambda$. There is an associative product on $\Lambda\(t\)$, called plethysm and denoted $f\circ g$, characterized by the formulas \begin{enumerate} \item $(f_1+f_2)\circ g=f_1\circ g+f_2\circ g$; \item $(f_1f_2)\circ g=(f_1\circ g)(f_2\circ g)$; \item if $f=f(t,p_1,p_2,\dots)$, then $p_n\circ f=f(t^n,p_n,p_{2n},\dots)$, and $t\circ f=t$. \end{enumerate} The following formula generalizes its analogue for ungraded $\SS$-modules, proved in \cite{Macdonald}: $$ \operatorname{ch}_t(\v\circ\mathcal{W}) = \operatorname{ch}_t(\v)\circ\operatorname{ch}_t(\mathcal{W}) . $$ The operation $$ \operatorname{Exp}(f) = \sum_{n=0}^\infty h_n\circ f $$ plays the role for symmetric functions that exponentiation does for power series. The inverse of $\operatorname{Exp}$ is the operation $$ \operatorname{Log}(f) = \sum_{n=1}^\infty \frac{\mu(n)}{n} \log(p_n\circ f) , $$ where $\mu(n)$ is the M\"obius function. Using this formula and the Poincar\'e-Birkhoff-Witt theorem ${\operatorname{\mathcal{A}\mathit{ss}}}^+={\operatorname{\mathcal{C}\mathit{om}}}^+\circ{\operatorname{\mathcal{L}\mathit{ie}}}^+$, we may calculate $\operatorname{ch}({\operatorname{\mathcal{L}\mathit{ie}}}^+)$. \begin{proposition} \label{Lie} $\displaystyle\operatorname{ch}({\operatorname{\mathcal{L}\mathit{ie}}}^+) = - \sum_{n=1}^\infty \frac{\mu(n)}{n} \log(1-p_n)$ \end{proposition} \begin{pf} We know from the Poincar\'e-Birkhoff-Witt theorem that $\operatorname{ch}({\operatorname{\mathcal{A}\mathit{ss}}}^+)=\operatorname{Exp}(\operatorname{ch}({\operatorname{\mathcal{L}\mathit{ie}}}^+))$; it follows that $$ \operatorname{ch}({\operatorname{\mathcal{L}\mathit{ie}}}^+) = \operatorname{Log}(1+\operatorname{ch}({\operatorname{\mathcal{A}\mathit{ss}}}^+)) = \sum_{n=1}^\infty \frac{\mu(n)}{n} \log(1+p_n\circ\operatorname{ch}({\operatorname{\mathcal{A}\mathit{ss}}}^+)) . $$ Since $$ 1+ p_n \circ \operatorname{ch}({\operatorname{\mathcal{A}\mathit{ss}}}^+) = 1 + \frac{p_n}{1-p_n} = \frac{1}{1-p_n} , $$ the result follows. \end{pf} It follows from this formula that $$ \operatorname{ch}_n({\operatorname{\mathcal{L}\mathit{ie}}}^+(n)) = \frac{1}{n} \sum_{d|n} \mu(d) p_d^{n/d} . $$ This is the characteristic of the induced representation $\operatorname{Ind}_{C_n}^{\SS_n}\chi$, where $\chi$ is a primitive character of the cyclic group. We now turn to calculating the characteristic of the braid operad $\mathcal{B}\mathit{raid}$. First, we need a lemma. \begin{lemma} $\operatorname{ch}_t(\Lambda\v)=-t\operatorname{ch}_t(\v)(-t^{-1}p_1,-t^{-2}p_2,-t^{-3}p_3,\dots)$ \end{lemma} \begin{pf} Tensoring with $\operatorname{sgn}_n$ has the effect of replacing $p_n$ by $(-1)^{n-1}p_n$. Applying $\Sigma^{-n}$ to $\v(n)$ then has the effect of replacing $p_n$ by $(-t)^{-n}p_n$. \end{pf} \begin{proposition} \label{configuration} For each $n\ge1$, let $$ P_n(t) = \frac{1}{n} \sum_{d|n} \frac{\mu(n/d)}{t^d} . $$ Then $$ \operatorname{ch}_t(\mathcal{B}\mathit{raid}) = \prod_{n=1}^\infty \Bigl( 1 + t^np_n \Bigr)^{P_n(t)} - 1 . $$ \end{proposition} \begin{pf} The $\SS$-module $\Lambda^{-1}{\operatorname{\mathcal{L}\mathit{ie}}}^+$ has Poincar\'e characteristic $$ \operatorname{ch}_t(\Lambda^{-1}{\operatorname{\mathcal{L}\mathit{ie}}}^+) = \frac{1}{t} \sum_{n=1}^\infty \frac{\mu(n)}{n} \log(1+t^np_n) . $$ It follows that \begin{eqnarray*} \operatorname{ch}_t(\b) &=& \operatorname{Exp} \circ \operatorname{ch}_t(\Lambda^{-1}{\operatorname{\mathcal{L}\mathit{ie}}}^+) - 1 = \exp\Bigl( \sum_{k=1}^\infty \frac{p_k}{k} \Bigr) \circ \Bigl( t^{-1} \sum_{n=1}^\infty \frac{\mu(n)}{n} \log(1+t^np_n) \Bigr) - 1 \\ &=& \exp\Bigl( \sum_{k=1}^\infty \sum_{n=1}^\infty t^{-k} \frac{\mu(n)}{kn} \log(1+t^{kn}p_{kn}) \Bigr) - 1 \\ &=& \prod_{n=1}^\infty ( 1 + t^np_n )^{P_n(t)} - 1 . }\end{pf \end{eqnarray*} \def}\end{pf{}\end{pf} In particular, setting $p_1=x$ and $p_n=0$, $n>1$, we see that the Poincar\'e polynomial of the space $\mathbb{C}^n_0$ is \begin{eqnarray} \label{poincare-config} \sum_{i=0}^{n-1} (-t)^n \dim H^i(\mathbb{C}^n_0) &=& \text{coefficient of $x^n/n!$ in $(1+tx)^{t^{-1}} - 1$} \\ &=& \binom{t^{-1}}{n} t^n = \prod_{i=1}^{n-1} (1-it) . \end{eqnarray} \subsection{The characteristic of a cyclic $\SS$-module} If $\v$ is a stable cyclic $\SS$-module, we define $\operatorname{Ch}_t(\v)$ in a similar way to $\operatorname{ch}_t(\v)$: $$ \operatorname{Ch}_t(\v) = \sum_{n=3}^\infty \sum_{i=0}^\infty (-t)^i \, \operatorname{ch}_n(\v_i\(n\)) \in \Lambda\(t\) . $$ If $\operatorname{ch}_t(\v)$ denotes the Poincar\'e characteristic of the $\SS$-module underlying $\v$, then we have the formula $$ \operatorname{ch}_t(\v) = \frac{\partial\operatorname{Ch}_t(\v)}{\partial p_1} . $$ We now calculate $\operatorname{Ch}_t(\mathbf{m})$, where $\mathbf{m}$ is the cyclic $\SS$-module $$ \mathbf{m}\(n\) = \begin{cases} H_{\bullet}(\mathcal{M}_{0,n}) , & n\ge3 , \\ 0 , & n<3 . \end{cases}$$ Note that the Poincar\'e polynomial of $\mathcal{M}_{0,n}$ is much easier to calculate than the Poincar\'e characteristic: it is obtained by dividing the Poincar\'e polynomial \eqref{poincare-config} of $\mathbb{C}^{n-1}_0$ by $1-t$: \begin{equation} \label{poincare-moduli} \sum_{i=0}^{n-2} (-t)^n \dim H^i(\mathcal{M}_{0,n}) = \prod_{i=2}^{n-2} (1-it). \end{equation} \begin{theorem} \label{moduli:characteristic} $$ \operatorname{Ch}_t(\mathbf{m}) = \frac{1}{1-t^2} \Bigl( (1+tp_1) \prod_{n=1}^\infty ( 1 + t^np_n )^{P_n(t)} - 1 - (1+t)h_1 - (h_2 + te_2) \Bigr) $$ \end{theorem} \begin{pf} {}From the $\SS_n$-equivariant homotopy equivalence $\mathbb{C}^n_0\simeq\mathcal{M}_{0,n+1}\times S^1$, which holds for $n\ge2$, we see that \begin{equation} \label{differential} \frac{\partial\operatorname{Ch}_t(\mathbf{m})}{\partial p_1} = \operatorname{ch}_t(\mathbf{m}) = \frac{t\bigl(\operatorname{ch}_t(\mathcal{B}\mathit{raid})-p_1\bigr)}{t-1} . \end{equation} The Serre spectral sequence for the $\SS_n$-equivariant fibration $$\begin{CD} \mathbb{C}\setminus\{1,\dots,n\} @>>> \mathcal{M}_{0,n+1} @>>> \mathcal{M}_{0,n} \end{CD}$$ collapses at $E^2$, so $H_{\bullet}(\mathcal{M}_{0,n+1}) \cong H_{\bullet}(\mathbb{C}\setminus\{1,\dots,n\}) \o H_{\bullet}(\mathcal{M}_{0,n})$. Furthermore, this isomorphism is $\SS_n$-equivariant, where $\SS_n$ acts on $H_{\bullet}(\mathbb{C}\setminus\{1,\dots,n\})$ by the monodromy of the Gauss-Manin connection. Now, $H_0(\mathbb{C}\setminus\{1,\dots,n\})$ is the trivial $\SS_n$-module, while $H_1(\mathbb{C}\setminus\{1,\dots,n\})$ is the irreducible representation $V_{n-1,1}$, which is the kernel of the natural map $\begin{CD} \mathbb{C}^n@>>>\mathbb{C}\end{CD}$ obtained by sending $(x_1,\dots,x_n)$ to $x_1+\dots+x_n$. If $\sigma\in\SS_n$ is a transitive permutation, $\operatorname{Tr}(\sigma|V_{n-1,1})=-1$; this shows that \begin{equation} \label{initial} \operatorname{Ch}_t(\mathbf{m})\big|_{p_1=0} = \frac{\operatorname{ch}_t(\mathcal{B}\mathit{raid})}{1+t} \Big|_{p_1=0} . \end{equation} The theorem now follows on solving the differential equation \eqref{differential} with initial condition \eqref{initial}. \end{pf} The first few terms of $\operatorname{Ch}_t(\mathbf{m})$ are as follows: $$\begin{tabular}{|c|l|} $n$ & $\operatorname{Ch}_t(\mathcal{M}_{0,n})$ \\ \hline $3$ & $s_3$ \\ $4$ & $s_4-ts_{2^2}$ \\ $5$ & $s_5-ts_{32}+t^2(s_{31^2})$ \\ $6$ & $s_6-ts_{42}+t^2(s_{41^2}+s_{321})-t^3(s_{41^2}+s_{3^2}+s_{2^21^2})$ \\ \end{tabular}$$ The pattern emerging here, that $H_1(\mathcal{M}_{0,n})\cong V_{n-2,2}$, is easily verified in general using our formula for $\operatorname{Ch}_t(\mathbf{m})$. We have seen that there is a natural identification between $\operatorname{sgn}_n\o H^1(\mathcal{M}_{0,n})$ and the space of relations \eqref{hypercommutative} among $n-1$ letters in the hypercommutative operad: thus, we see that this space of relations is the irreducible $\SS_n$-module $V_{2^21^{n-4}}$. Applying l'H\^opital's rule to Theorem \ref{moduli:characteristic}, we see that the Euler-Frobenius characteristic of $\mathbf{m}$ is given by the formula $$ \operatorname{Ch}(\mathbf{m}) = \lim_{\begin{CD}t@>>>1\end{CD}} \operatorname{Ch}_t(\mathbf{m}) = \frac12 (1+p_1)^2 \sum_{n=1}^\infty \frac{\varphi(n)}{n} \log(1+p_n) - \frac{1}{4} (2p_1 + 3p_1^2 + p_2) . $$ Finally, it follow easily from the formula for $\operatorname{Ch}_t(\mathbf{m})$ that $$ \operatorname{Ch}_t(\mathcal{G}\mathit{rav}) = - \frac{t^3}{1-t^2} \Bigl( (1-p_1) \prod_{n=1}^\infty (1-p_n)^{P_n(t)} - 1 + (1+t^{-1})h_1 - (t^{-1}h_2 + t^{-2}e_2) \Bigr) $$ \subsection{The Poincar\'e characteristic of $\mathbb{T}_+\v$} In \cite{modular}, a formula for the Poincar\'e characteristic of $\mathbb{T}_+\v$ in terms of the Poincar\'e characteristic of $\v$ is derived. If $F=e_2-\operatorname{Ch}_t(\v)$, define the Legendre transform $G=\mathcal{L} F$ of $F$ in the sense of symmetric functions by the formula \begin{equation} \label{Legendre} F\circ \frac{\partial(\mathcal{L} G)}{\partial p_1} + G = p_1 \frac{\partial G}{\partial p_1} . \end{equation} Then $G=h_2+\operatorname{Ch}_t(\mathbb{T}_+\v)$. Note that \eqref{Legendre} implies that $(\partial F/\partial p_1)\circ(\partial G/p p_1)=p_1$, from which it is straightforward to calculate $\partial G/\partial p_1$. Substituting $\partial G/\partial p_1$ into both sides of \eqref{Legendre}, we obtain an explicit formula for $G$. As an application of \eqref{Legendre}, we now calculate the Poincar\'e characteristics of the varieties $\overline{\mathcal{M}}_{0,n}$. We use a slight extension of \eqref{Legendre}, in which $\v=\v_0\oplus\v_1$ has an internal $\mathbb{Z}/2$-grading, and $\operatorname{Ch}_t(\v)=\operatorname{Ch}_t(\v_0)-\operatorname{Ch}_t(\v_1)$. Let $\v$ and $\mathcal{W}$ be the stable cyclic $\SS$-modules \begin{eqnarray*} \v_i\(n\) &=& \begin{cases} 0 , & i=0 , \\ \mathcal{H}\mathit{ycom}\(n\)\o\operatorname{sgn}_n , & i=1 ; \end{cases} \\ \mathcal{W}_{ij}\(n\) &=& \begin{cases} \operatorname{sgn}_n\o H^p(\mathcal{M}_{0,n}) , & \text{$i=2(n-p-3)$ and $j\equiv p+1\pmod{2}$,} \\ 0 , & \text{otherwise.} \end{cases} \end{eqnarray*} Thus, \begin{eqnarray*} \operatorname{Ch}_t(\v) & =& - \operatorname{Ch}_t(\mathcal{H}\mathit{ycom})\big|_{p_n\to(-1)^{n-1}p_n} , \\ \operatorname{Ch}_t(\mathcal{W}) &=& \frac{- t^{-6}\operatorname{Ch}_t(\mathbf{m})\Big|_{t\to-t^{-2}} {p_n\to(-1)^{n-1}t^{2n}p_n}} . \end{eqnarray*} Using \eqref{det}, we may rewrite \eqref{bar:even} in the form $$\begin{CD} 0 @>>> H^p(\mathcal{M}_{0,n})\o\operatorname{sgn}_n @>>> \bigoplus_{T\in\mathcal{T}_p\(n\)} \v\(T\)_{2(n-p-3)} @>>> \bigoplus_{T\in\mathcal{T}_{p-1}\(n\)} \v\(T\)_{2(n-p-3)} @>>> \dots \end{CD}$$ which shows that $\operatorname{Ch}_t(\mathcal{W})=\operatorname{Ch}_t(\mathbb{T}_+\v)$, and hence that $G=h_2+\operatorname{Ch}_t(\mathcal{W})$ is the Legendre transform of $F=e_2-\operatorname{Ch}_t(\v)$. In this way, we have proved the following proposition. \begin{theorem} \label{F-G} The symmetric function $$ F = e_2 + \operatorname{Ch}_t(\mathcal{H}\mathit{ycom})\big|_{p_n\to(-1)^{n-1}p_n} $$ is the Legendre transform of the symmetric function $$ G = h_2 - t^{-6} \operatorname{Ch}_t(\mathbf{m})\Big|_{\frac{t\to t^2}{ p_n\to(-1)^{n-1}t^{2n}p_n}} . $$ \end{theorem} Here are some sample calculations of $\operatorname{Ch}_t(\overline{\mathcal{M}}_{0,n})$ for small $n$: $$\begin{tabular}{|c|l|} $n$ & $\operatorname{Ch}_t(\overline{\mathcal{M}}_{0,n})$ \\ \hline $3$ & $s_3$ \\ $4$ & $(1+t^2)s_4$ \\ $5$ & $(1+t^4)s_5+t^2(s_5+s_{41})$ \\ $6$ & $(1+t^6)s_6+(t^2+t^4)(2s_6+s_{51}+s_{42})$ \\ \end{tabular}$$ These formulas simplify if we are only interested in the dimensions of the vector spaces $H_i(\overline{\mathcal{M}}_{0,n})$. We have the formula \begin{eqnarray*} g(x,t) = G'\Big|_{\frac{p_1\to x}{ p_n\to0,n>1}} &=& x - \sum_{n=2}^\infty \frac{x^n}{n!} \sum_{i=0}^{n-2} (-1)^i t^{2(n-i-2)} \dim H_i(\mathcal{M}_{0,n+1}) \\ &=& x - \frac{(1+x)^{t^2}-(1+t^2x)}{t^2(t^2-1)} . \end{eqnarray*} It is a corollary of Theorem \ref{F-G} that $$ f(x,t) = F'\Big|_{\frac{p_1\to x}{ p_n\to0,n>1}} = x + \sum_{n=2}^\infty \frac{x^n}{n!} \sum_{i=0}^{n-2} t^{2i} \dim H_{2i}(\overline{\mathcal{M}}_{0,n+1}) , $$ is the inverse of $g$, in the sense that $f(g(x,t),t)=x$. This is a reformulation of Fulton and MacPherson's calculation of the Poincar\'e polynomial of $\overline{\mathcal{M}}_{0,n}$. Note that their proof also makes use of mixed Hodge theory, in the form of the ``fake Poincar\'e polynomial.'' Our result Theorem \ref{F-G} is an equivariant version of their calculation. \makeatletter\renewcommand{\@biblabel}[1]{\hfill[#1]}\makeatother
1997-01-22T09:51:28
9610
alg-geom/9610002
en
https://arxiv.org/abs/alg-geom/9610002
[ "alg-geom", "math.AG" ]
alg-geom/9610002
Yves Laszlo
Yves Laszlo
Linearization of group stack actions and the Picard group of the moduli of $\SL_r/\mu_s$-bundles on a curve
13 pages, PlainTex
null
null
null
null
We first study the descent theory of line bundles under a morphism which is tors or under a group stack and then use this technical result to determine the exact structure of $\Pic(\M_G)$ where $G=\SL_r/\mu_s$ (we include a minor modification to explain the genus 0 case).
[ { "version": "v1", "created": "Thu, 3 Oct 1996 07:38:31 GMT" }, { "version": "v2", "created": "Wed, 22 Jan 1997 08:50:18 GMT" } ]
2008-02-03T00:00:00
[ [ "Laszlo", "Yves", "" ] ]
alg-geom
\section{...} et une sous-section avec \font\tengoth=eufm10 at 10 pt \font\sevengoth=eufm10 at 7 pt \font\fivegoth=eufm10 at 5 pt \newfam\gothfam \textfont\gothfam=\tengoth \scriptfont\gothfam=\sevengoth \scriptscriptfont\gothfam=\fivegoth \def\fam\gothfam\tengoth{\fam\gothfam\tengoth} \newcount\subsecno \newcount\secno \newcount\prmno \newif\ifnotfound \newif\iffound \def\namedef#1{\expandafter\def\csname #1\endcsname} \def\nameuse#1{\csname #1\endcsname} \def\typeout#1{\immediate\write16{#1}} \long\def\ifundefined#1#2#3{\expandafter\ifx\csname #1\endcsname\relax#2\else#3\fi} \def\hwrite#1#2{{\let\the=0\edef\next{\write#1{#2}}\next}} \toksdef\ta=0 \toksdef\tb=2 \long\def\leftappenditem#1\to#2{\ta={\\{#1}}\tb=\expandafter{#2}% \edef#2{\the\ta\the\tb}} \long\def\rightappenditem#1\to#2{\ta={\\{#1}}\tb=\expandafter{#2}% \edef#2{\the\tb\the\ta}} \def\lop#1\to#2{\expandafter\lopoff#1\lopoff#1#2} \long\def\lopoff\\#1#2\lopoff#3#4{\def#4{#1}\def#3{#2}} \def\ismember#1\of#2{\foundfalse{\let\given=#1% \def\\##1{\def\next{##1}% \ifx\next\given{\global\foundtrue}\fi}#2}} \def\section#1{\medbreak \global\defth}\medskip{section} \global\advance\secno by1\global\prmno=0\global\subsecno=0 {\bf \par\hskip 0cm\relax\number\secno. {#1}$.-$} } \def\global\def\currenvir{subsection{\global\defth}\medskip{subsection} \global\advance\prmno by1\global\subsecno=0 \par\hskip 0cm\relax{\bf \number\secno.\number\prmno. }} \def\global\def\currenvir{rem{\global\defth}\medskip{rem} \global\advance\prmno by1\global\subsecno=0 \medskip\par\hskip 0cm\relax{\bf Remark \number\secno.\number\prmno$.-$ }} \def\global\def\currenvir{ex{\global\defth}\medskip{ex} \global\advance\prmno by1\global\subsecno=0 \medskip\par\hskip 0cm\relax{\bf Example \number\secno.\number\prmno$.-$ }} \def\global\def\currenvir{formule{\global\defth}\medskip{formule} \ifnum\prmno=0\global\advance\prmno by1 {\number\secno.\number\prmno}\else \global\advance\subsecno by1 {\number\secno.\number\prmno.\number\subsecno}\fi} \def\proclaim#1{\global\advance\prmno by 1\global\subsecno=0 {\bf #1 \the\secno.\the\prmno$.-$ }} \long\def\th#1 \enonce#2\endth{% \medbreak\proclaim{#1}{\it #2}\global\defth}\medskip{th}\medskip} \def\isinlabellist#1\of#2{\notfoundtrue% {\def\given{#1}% \def\\##1{\def\next{##1}% \lop\next\to\za\lop\next\to\zb% \ifx\za\given{\zb\global\notfoundfalse}\fi}#2}% \ifnotfound{\immediate\write16% {Warning - [Page \the\pageno] {#1} No reference found}}% \fi}% \def\ref#1{\ifx\empty{\immediate\write16 {Warning - No references found at all.}} \else{\isinlabellist{#1}\of}\fi} \def\newlabel#1#2{\rightappenditem{\\{#1}\\{#2}}\to} \def{} \def\label#1{% \def\given{th}% \ifx\giventh}\medskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \def\given{section}% \ifx\giventh}\medskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno}}}\f \def\given{formule}% \ifx\giventh}\medskip% { \ifnum\prmno=1\hwrite\lbl { \string\newlabel{#1}{\number\secno.\number\prmno} } \else\hwrite\lbl { \string\newlabel{#1}{\number\secno.\number\prmno.\number\subsecno} }\fi} \fi% \def\given{subsection}% \ifx\giventh}\medskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \def\given{rem}% \ifx\giventh}\medskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \def\given{ex}% \ifx\giventh}\medskip% {\hwrite\lbl{\string\newlabel{#1}{\number\secno.\number\prmno}}}\fi% \def\given{subsubsection}% \ifx\giventh}\medskip% {\hwrite\lbl{\string% \newlabel{#1}{\number\secno.\number\subsecno.\number\subsubsecno}}}\fi \ignorespaces} \newwrite\lbl \def\openout\lbl=\jobname.lbl{\openout\lbl=\jobname.lbl} \def\closeout\lbl{\closeout\lbl} \newread\testfile \def\lookatfile#1{\openin\testfile=\jobname.#1 \ifeof\testfile{\immediate\openout\nameuse{#1}\jobname.#1 \write\nameuse{#1}{} \immediate\closeout\nameuse{#1}}\fi% \immediate\closein\testfile}% \def\lookatfile{lbl{\lookatfile{lbl} \input\jobname.lbl \openout\lbl=\jobname.lbl} \let\bye\closeall\bye \def\closeall\bye{\closeout\lbl\bye} \mathcode`A="7041 \mathcode`B="7042 \mathcode`C="7043 \mathcode`D="7044 \mathcode`E="7045 \mathcode`F="7046 \mathcode`G="7047 \mathcode`H="7048 \mathcode`I="7049 \mathcode`J="704A \mathcode`K="704B \mathcode`L="704C \mathcode`M="704D \mathcode`N="704E \mathcode`O="704F \mathcode`P="7050 \mathcode`Q="7051 \mathcode`R="7052 \mathcode`S="7053 \mathcode`T="7054 \mathcode`U="7055 \mathcode`V="7056 \mathcode`W="7057 \mathcode`X="7058 \mathcode`Y="7059 \mathcode`Z="705A \def\spacedmath#1{\def\packedmath##1${\bgroup\mathsurround =0pt##1\egroup$} \mathsurround#1 \everymath={\packedmath}\everydisplay={\mathsurround=0pt}} \def\nospacedmath{\mathsurround=0pt \everymath={}\everydisplay={} } \spacedmath{2pt} \def\qfl#1{\normalbaselines{\baselineskip=0pt \lineskip=10truept\lineskiplimit=1truept}\nospacedmath\smash {\mathop{\hbox to 6truemm{\rightarrowfill}} \limits^{\scriptstyle#1}}} \def\phfl#1#2{\normalbaselines{\baselineskip=0pt \lineskip=10truept\lineskiplimit=1truept}\nospacedmath\smash {\mathop{\hbox to 8truemm{\rightarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\diagram#1{\def\normalbaselines{\baselineskip=0truept \lineskip=10truept\lineskiplimit=1truept} \matrix{#1}} \def\vfl#1#2{\llap{$\scriptstyle#1$}\left\downarrow\vbox to 6truemm{}\right.\rlap{$\scriptstyle#2$}} \def\note#1#2{\footnote{\parindent 0.4cm$^#1$}{\vtop{\eightpoint\baselineskip12pt\hsize15.5truecm\noindent #2}} \parindent 0cm} \def\mathop{\rm Ker}\nolimits{\mathop{\rm Ker}\nolimits} \def\mathop{\rm Pic}\nolimits{\mathop{\rm Pic}\nolimits} \def\pc#1{\tenrm#1\sevenrm} \def\up#1{\raise 1ex\hbox{\smallf@nt#1}} \def\kern-1.5pt -{\kern-1.5pt -} \def\cqfd{\kern 2truemm\unskip\penalty 500\vrule height 4pt depth 0pt width 4pt\medbreak} \def\vrule height 4pt depth 0pt width 4pt{\vrule height 5pt depth 0pt width 5pt} \def\raise .4ex\hbox{,}{\raise .4ex\hbox{,}} \def\decale#1{\smallbreak\hskip 28pt\llap{#1}\kern 5pt} \defn\up{o}\kern 2pt{n\up{o}\kern 2pt} \def\par\hskip 0cm\relax{\par\hskip 0cm\relax} \def\par\hskip 0.5cm\relax{\par\hskip 0.5cm\relax} \parindent=0cm \def\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}{\mathrel{\hbox{\vrule height 3pt depth -2pt width 6pt}}} \def\kern 1pt{\scriptstyle\circ}\kern 1pt{\kern 1pt{\scriptstyle\circ}\kern 1pt} \baselineskip15pt \overfullrule=0pt \def\boxit#1#2{ \setbox1=\hbox{\kern#1{#2}\kern#1} \dimen1=\ht1 \advance\dimen1 by #1 \dimen2=\dp1 \advance\dimen2 by #1 \setbox1=\hbox{\vrule height\dimen1 depth\dimen2\box1\vrule} \setbox1=\vbox{\hrule\box1\hrule} \advance\dimen1 by .4pt \ht1\dimen1 \advance\dimen2 by .4pt \dp1\dimen2 \box1\relax} \def\boxit{0pt}{$\times$}{\boxit{0pt}{$\times$}} \def\rlap{\raise 1pt\hbox{$\scriptscriptstyle /$}}\alpha{\rlap{\raise 1pt\hbox{$\scriptscriptstyle /$}}\alpha} \def\mathop{\bf Spin}\nolimits{\mathop{\bf Spin}\nolimits} \def\mathop{\bf Sp}\nolimits{\mathop{\bf Sp}\nolimits} \def\mathop{\bf GL}\nolimits{\mathop{\bf GL}\nolimits} \def\mathop{\bf SL}\nolimits{\mathop{\bf SL}\nolimits} \def\mathop{\bf SO}\nolimits{\mathop{\bf SO}\nolimits} \let\bk\backslash \let\lra\longrightarrow \let{\rightarrow}\rightarrow \def\it{\it} \def\hbox{\rm Id}{\hbox{\rm Id}} \def\buildrel\sim\over{\rightarrow}{\buildrel\sim\over{\rightarrow}} \def\mathop{\rm Cl}\nolimits{\mathop{\rm Cl}\nolimits} \def\mathop{\rm Aut}\nolimits{\mathop{\rm Aut}\nolimits} \def\mathop{\rm Stab}\nolimits{\mathop{\rm Stab}\nolimits} \def\mathop{\rm codim}\nolimits{\mathop{\rm codim}\nolimits} \def{\hbox{\rm\'et}}{{\hbox{\rm\'et}}} \def{\bf} Z}\def\Q{{\bf} Q}\def\C{{\bf} C}\def\F{{\bf} F{{\bf} Z}\def\Q{{\bf} Q}\def\C{{\bf} C}\def\F{{\bf} F} \def{\rm Spec}{\mathop{\rm spec}\nolimits} \def\mathop{\oalign{lim\cr\hidewidth$\longrightarrow$\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longrightarrow$\hidewidth\cr}}} \def\mathop{\oalign{lim\cr\hidewidth$\longleftarrow$\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longleftarrow$\hidewidth\cr}}} \def\fhd#1#2{\nospacedmath\smash{\mathop{\hbox to 8mm{\rightarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\fhg#1#2{\nospacedmath\smash{\mathop{\hbox to 8mm{\leftarrowfill}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\fvb#1#2{\nospacedmath\llap{$\scriptstyle #1$}\left\downarrow \vbox to 6mm{}\right.\rlap{$\scriptstyle#2$}} \def\fvh#1#2{\nospacedmath\llap{$\scriptstyle #1$}\left\uparrow \vbox to 6mm{}\right.\rlap{$\scriptstyle#2$}} \def\FFhd#1#2{\nospacedmath\smash{\mathop{\hbox to 8mm{\hfil$\Longrightarrow$\hfil}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\FFhg#1#2{\nospacedmath\smash{\mathop{\hbox to 8mm{$\Longleftarrow$}} \limits^{\scriptstyle#1}_{\scriptstyle#2}}} \def\FFvb#1#2{\nospacedmath\llap{$\scriptstyle #1$}\left\Downarrow \vbox to 6mm{}\right.\rlap{$\scriptstyle#2$}} \def\FFvh#1#2{\nospacedmath\llap{$\scriptstyle #1$}\left\Uparrow \vbox to 6mm{}\right.\rlap{$\scriptstyle#2$}} \def\fvb{}{}{\fvb{}{}} \def\fvh{}{}{\fvh{}{}} \def\fhd{}{}{\fhd{}{}} \def\fvb{}{}{\fvb{}{}} \def\rightarrow \kern -3mm\rightarrow {\rightarrow \kern -3mm\rightarrow } \def\kern 1pt{\scriptstyle\circ}\kern 1pt{\kern 1pt{\scriptstyle\circ}\kern 1pt} \def{\rightarrow}{{\rightarrow}} \def\vrule height 4pt depth 0pt width 4pt{\vrule height 4pt depth 0pt width 4pt} \def\sevenbf{\sevenbf} \def\\{{\setminus}} \def\mathop{\rm pf}\nolimits{\mathop{\rm pf}\nolimits} \def\mathop{\rm Aut}\nolimits{\mathop{\rm Aut}\nolimits} \def\mathop{{\cal A}ut}\nolimits{\mathop{{\cal A}ut}\nolimits} \def\mathop{\rm Hom}\nolimits{\mathop{\rm Hom}\nolimits} \def\mathop{{\cal H}om}\nolimits{\mathop{{\cal H}om}\nolimits} \def\mathop{{E}xt}\nolimits{\mathop{{E}xt}\nolimits} \def\mathop{{\cal E}nd}\nolimits{\mathop{{\cal E}nd}\nolimits} \def\mathop{\rm End}\nolimits{\mathop{\rm End}\nolimits} \def\mathop{\rm Ext}\nolimits{\mathop{\rm Ext}\nolimits} \def\mathop{{\rm Im}\nolimits}{\mathop{{\rm Im}\nolimits}} \def{\bf P}{{\bf P}} \def{\cal O}{{\cal O}} \def{(}}\def\br{{)}{{(}}\def\br{{)}} \let\bk\backslash \let\lra\longrightarrow \let{\rightarrow}\rightarrow \def\mathop{\rm dim}\nolimits{\mathop{\rm dim}\nolimits} \def\mathop{\rm Stab}\nolimits{\mathop{\rm Stab}\nolimits} \def\mathop{\rm codim}\nolimits{\mathop{\rm codim}\nolimits} \def\mathop{\oalign{lim\cr\hidewidth$\longrightarrow$\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longrightarrow$\hidewidth\cr}}} \def\mathop{\oalign{lim\cr\hidewidth$\longleftarrow$\hidewidth\cr}}{\mathop{\oalign{lim\cr\hidewidth$\longleftarrow$\hidewidth\cr}}} \def{\partial}{{\partial}} \def\mathop{\bf S}\nolimits{\mathop{\bf S}\nolimits} \def{\bf}{{\bf}} \def{\rm Spec}{{\rm Spec}} \lookatfile{lbl \null \centerline{{\bf Linearization of group stack actions and the Picard group}} \centerline{{\bf of the moduli of $\mathop{\bf SL}\nolimits_r/\mu_s$-bundles on a curve}} \medskip \centerline{Yves {\pc LASZLO} \footnote{\parindent 0.5cm($\dagger$)}{\sevenrm Partially supported by the European HCM Project ``Algebraic Geometry in Europe" (AGE).}} \def\a{{\cal A}}\def\b{{\cal B}}\def\c{{\cal C}}\def\tp{{\tilde p}}\def\X{{\cal X}} \def{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}} \def{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}{{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def{\bf P}{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}} \def{\cal U}}\def\d{{\cal D}}\def\F{{\cal F}{{\cal U}}\def\d{{\cal D}}\def\F{{\cal F}} \let\scp=\scriptstyle {\bf Introduction} \medskip Let $G$ be a complex semi-simple group and $\tilde G\rightarrow \kern -3mm\rightarrow G$ the universal covering. Let $\M_G$ (resp. $\M_{\tilde G}$) be the moduli stack of $G$-bundles over a curve $X$ of degree $1\in\pi_1(G)$ (resp. of ${\tilde G}$-bundles. In [B-L-S], we have studied the link between the groups $\mathop{\rm Pic}\nolimits(\M_G)$ and $\mathop{\rm Pic}\nolimits(\M_{\tilde G}$), the later being well understood thanks to [L-S]. In particular, it has been possible to give a complete description in the case where $G={\bf PSL}_r$ but not in the case $\mathop{\bf SL}\nolimits_r/\mu_s,\ s\mid r$, although we were able to give partial results. The reason was that we did not have at our disposal the technical background to study the morphism $\M_{\tilde G}{\rightarrow}\M_G$. It turns to be out that it is a torsor under some group stack, not far from a Galois \'etale cover in the usual schematic picture. Now, the descent theory of Grothendieck has been adapted to the set-up of fpqc morphisms of stacks in [L-M] and gives the theorem \ref{theo} in the particular case of a morphism which is torsor under a group stack. We then used this technical result to determine the exact structure of $\mathop{\rm Pic}\nolimits(\M_G)$ where $G=\mathop{\bf SL}\nolimits_r/\mu_s$ (theorem \ref{theo-pic}). \medskip I would like to thank L. Breen to have taught me both the notion of torsor and linearization of a vector bundle in the set-up of group-stack action and for his comments on a preliminary version of this paper. \medskip {\bf Notation} \medskip Throughout this paper, all the stacks will be implicitely assumed to be algebraic over a fixed base scheme and the morphisms locally of finite type. We fix once for all a projective, smooth, connected genus $g$ curve $X$ over an algebraically closed field $\k$ and a closed point $x$ of $X$. For simplicity, we assume $g>0$ (see remarks \ref{rema} and \ref{rem} for the case of ${\bf P}^1$). The Picard stack parametrizing families of line bundles of degree $0$ on $X$ will be denoted by ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ and the jacobian variety of $X$ by $JX$. If $G$ is an algebraic group over $\k$, the quotient stack ${\rm Spec}(\k)/G$ (where $G$ acts trivialy on ${\rm Spec}(\k))$ whose category over a $\k$-scheme $S$ is the category of $G$-torsors (or $G$-bundles) over $S$ will be denoted by $BG$. If $n$ is an integer and $A={\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X), JX$ or $BG_m$ we denote by $n_A$ the $n^{\rm\scp th}$-power morphism $a\longmapsto a^n$. We denote by ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_n$ (resp. $J_n$) the $0$-fiber $A\times_A{\rm Spec}(\k)$ of $n_A$ when $A={\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ (resp. $A=JX$), wh \section{Generalities} Following [Br], for any diagram $$ A\fhd{h}{}B\matrix{\fhd{g}{}\cr\Uparrow\rlap{$\scp\lambda$}\cr \scp f\cr\fhd{}{}}C\fhd{l}{}D$$ of $2$-categories, we'll denote by $l*\lambda:\ l\kern 1pt{\scriptstyle\circ}\kern 1pt f\Rightarrow l\kern 1pt{\scriptstyle\circ}\kern 1pt g$ (resp. $\lambda*h:\ f\kern 1pt{\scriptstyle\circ}\kern 1pt h\Rightarrow g\kern 1pt{\scriptstyle\circ}\kern 1pt h$) the $2$-morphism deduced from $\lambda$. \global\def\currenvir{subsection For the convenience of the reader, let us prove a simple formal lemma which will be usefull in the section \ref{simplicial}. Let $\a,\b,\c$ be three $2$-categories, a $2$-commutative diagram $$\matrix{ &&\c\cr &\llap{${}^{\delta_0}$}\nearrow&\fvh{}{d_0}\cr \a&\fhd{f}{}&\b\cr &\llap{${}_{\delta_1}$}\searrow&\fvb{}{d_1}\cr &&\c\cr}\leqno{(\global\def\currenvir{formule)}$$\label{strict-categorie} and a $2$-morphism $\mu:\ \delta_0\Rightarrow \delta_1$. \th Lemma \enonce Assume that $f$ is an equivalence. There exists a unique $2$-morphism $$\mu*f^{-1}:\ d_0\Rightarrow d_1$$ such that $(\mu*f^{-1})*f=\mu$. \endth\label{lemme-categorie} {\it Proof}: let $\epsilon_k, k=0,1$ the $2$-morphism $d_k\kern 1pt{\scriptstyle\circ}\kern 1pt f\Rightarrow \delta_k$. Let $b$ be an object of $\b$. Pick an object $a$ of $\a$ and an isomorphism $\alpha:\ f(a)\buildrel\sim\over{\rightarrow} b$. Let $\varphi_\alpha :\ d_0(b)\buildrel\sim\over{\rightarrow} d_1(b)$ be the unique isomorphism making the diagram $$\matrix{ \delta_0(a)&\fhd{\epsilon_0(a)}{}&d_0\kern 1pt{\scriptstyle\circ}\kern 1pt f(a)&\fhd{d_0(\alpha )}{}&d_0(b)\cr \fvb{\mu_a}{}&&&&\fvb{}{\varphi_\alpha }\cr \delta_1(a)&\fhd{\epsilon_1(a)}{}&d_1\kern 1pt{\scriptstyle\circ}\kern 1pt f(a)&\fhd{d_1(\alpha )}{}&d_1(b)\cr }$$ commutative. We have to show that $\varphi_\alpha $ does not depend on $\alpha $ but only on $b$. Let $\alpha ':\ f(a')\buildrel\sim\over{\rightarrow} b$ be another isomorphism. There exists a unique isomorphism $\iota:\ a'\buildrel\sim\over{\rightarrow} a$ such that $\alpha\kern 1pt{\scriptstyle\circ}\kern 1pt f(\iota)=\alpha'$. The one has the equality $\varphi_{\alpha'}=d_1(\alpha )\kern 1pt{\scriptstyle\circ}\kern 1pt\Phi\kern 1pt{\scriptstyle\circ}\kern 1pt d_0(\alpha)^{-1}$ where $$\Phi=[d_1\kern 1pt{\scriptstyle\circ}\kern 1pt f(\iota)]\kern 1pt{\scriptstyle\circ}\kern 1pt\epsilon_1(a')\kern 1pt{\scriptstyle\circ}\kern 1pt \mu_{a'}\kern 1pt{\scriptstyle\circ}\kern 1pt\epsilon_0(a')^{-1}\kern 1pt{\scriptstyle\circ}\kern 1pt[d_0\kern 1pt{\scriptstyle\circ}\kern 1pt f(\iota)]^{-1}.$$ The functoriality of $\epsilon_i$ and $\mu$ ensures that one has the equalities $$d_k\kern 1pt{\scriptstyle\circ}\kern 1pt f(\iota)\kern 1pt{\scriptstyle\circ}\kern 1pt\epsilon_k(a')=\epsilon_k(a)\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_k(\iota)$$ and $$\mu_{a}=\delta_1(\iota)\kern 1pt{\scriptstyle\circ}\kern 1pt\mu_{a'}\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_0(\iota)^{-1}.$$ This shows that $$\Phi=\epsilon_1(a)\kern 1pt{\scriptstyle\circ}\kern 1pt\mu_a\kern 1pt{\scriptstyle\circ}\kern 1pt\epsilon_0(a)^{-1}$$ which proves the equality $\varphi_\alpha =\varphi_{\alpha '}$. We can therefore define $\mu_b$ to be the isomorphism $\varphi_\alpha $ for one isomorphism $\alpha:\ f(a)\buildrel\sim\over{\rightarrow} b$. One checks that the construction is functorial in $b$ and the lemma follows.\cqfd \section{Linearizations of line bundles on stacks} Let us first recall following [Br] the notion of torsor in the stack context. \global\def\currenvir{subsection Let $f:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow}\E$ be a faithfully flat morphism of stacks. Let us assume that an algebraic $gr$-stack $\C$ acts on $f$ (the product of $\C$ is denoted by $\m$ and the unit object by $1$). Following [Br], this means that there exists a 1-morphism of $\E$-stacks $m:\ \C\times{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ and a $2$-morphism $\mu:\ m\kern 1pt{\scriptstyle\circ}\kern 1pt (\m\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})\Rightarrow m\kern 1pt{\scriptstyle\circ}\kern 1pt (\hbox{\rm Id}_\C\times m)$ such that the obvious associativity condition (see the diagram (6.1.3) of [Br]) is satisfied and such that there exists a $2$-morphism $\epsilon:\ m\kern 1pt{\scriptstyle\circ}\kern 1pt(1\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})\Rightarrow \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ which is compatible to $\mu$ in the obvious sense (see (6.1.4) of [Br]). \global\def\currenvir{rem To say that $m$ is a morphism of $\E$-stacks means that the diagram $$\matrix{\C\times{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}&\fhd{m}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\cr\searrow&&\swarrow\cr&\E\cr}$$ is $2$-commutative. In other words, if we denote for simplicity the image of a pair of objects $m(g,x)$ by $g.x$, this means that there exists a functorial isomorphism $\iota_{g,x}:\ f(g.x){\rightarrow} f(x)$.\label{Y-morphism} \global\def\currenvir{subsection\label{equi} Suppose that $\C$ acts on another such $f':{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}'{\rightarrow}\E$. A morphism $p:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}'{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ will be said equivariant if there exists a $2$-morphism $$q:\ m\kern 1pt{\scriptstyle\circ}\kern 1pt(\hbox{\rm Id}\times p)\Rightarrow p\kern 1pt{\scriptstyle\circ}\kern 1pt m'$$ which is compatible to $\mu$ (as in [Br] (6.1.6)) and $\epsilon$ (which is implicit in [Br]) in the obvious sense. \th Definition \enonce With the above notations, we say that $f$ (or ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$) is a $\C$-torsor over $\E$ if the morphism $pr_2\times m:\ \C\times {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\times_\E{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ is an isomorphism (of stacks) and the geometrical fibers of $f$ are not empty. \endth \global\def\currenvir{rem In down to earth terms, this means that if $\iota:\ f(x){\rightarrow} f(x')$ is an isomorphism in $\E$ ($x,x'$ being objects of ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$), there exist an object $g$ of $\C$ and a unique isomorphism $(x,g.x)\fhd{\sim}{}(x,x')$ which induces $\iota$ thanks to $\iota_{g,x}$ (cf. \ref{Y-morphism}). \global\def\currenvir{ex If $\M_X(G_m)$ is the Picard stack of $X$, the morphism $\M_X(G_m){\rightarrow}\M_X(G_m)$ of multiplication by $n\in{\bf} Z}\def\Q{{\bf} Q}\def\C{{\bf} C}\def\F{{\bf} F$ is a torsor under $B\mu_n\times J_n(X)$ (cf. (\ref{torseur})). \global\def\currenvir{subsection Let a $\L$ be a line bundle on ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$. By definition, the data $\L$ is equivalent to te data of a morphism $l:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow} BG_m$ (see [L-M],prop. 6.15). If $\L,\L'$ are $2$ line bundles on $\X$ defined by $l,l'$, we will view an isomorphism $\L\buildrel\sim\over{\rightarrow}\L'$ as a $2$-morphism $l\Rightarrow l'$.\label{convention-fibre} \th Definition \enonce A $\C$-linearization of $\L$ is a $2$-morphism $\lambda:\ l\kern 1pt{\scriptstyle\circ}\kern 1pt m\Rightarrow l\kern 1pt{\scriptstyle\circ}\kern 1pt pr_2$ such that the two diagrams of $2$-morphisms $$\matrix{ l\kern 1pt{\scriptstyle\circ}\kern 1pt m\kern 1pt{\scriptstyle\circ}\kern 1pt (\m\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})&\buildrel{l*\mu}\over{\Longrightarrow}&l\kern 1pt{\scriptstyle\circ}\kern 1pt m\kern 1pt{\scriptstyle\circ}\kern 1pt (\hbox{\rm Id}_\C\times m)\cr \hfill\Big\Downarrow{\scriptstyle\lambda*(\m\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})}&&\hfill\Big\Downarrow{\scriptstyle\lambda*(\hbox{\rm Id}_\C\times m)}\cr l\kern 1pt{\scriptstyle\circ}\kern 1pt pr_2\kern 1pt{\scriptstyle\circ}\kern 1pt(\m\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})=l\kern 1pt{\scriptstyle\circ}\kern 1pt pr_2\kern 1pt{\scriptstyle\circ}\kern 1pt pr_{23}&\buildrel \lambda*pr_{23}\over{\Longleftarrow}&l\kern 1pt{\scriptstyle\circ}\kern 1pt pr_2\kern 1pt{\scriptstyle\circ}\kern 1pt(\hbox{\rm Id}_\C\times m)=l\kern 1pt{\scriptstyle\circ}\kern 1pt m\kern 1pt{\scriptstyle\circ}\kern 1pt pr_{23}\cr }\leqno(\global\def\currenvir{formule)$$\label{lineA} and $$ \matrix{ l\kern 1pt{\scriptstyle\circ}\kern 1pt m\kern 1pt{\scriptstyle\circ}\kern 1pt(1\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})&\buildrel l*\epsilon\over{\Longrightarrow}&l\cr \FFvb{\lambda*(1\times \hbox{\rm Id}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})}{}&&\parallel\cr l&=&l\cr }\leqno(\global\def\currenvir{formule)$$\label{lineB}(stritly) commutes. \endth \global\def\currenvir{rem\label{rem-line} In $g_1,g_2$ are objects of $\C$ and $d$ of ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$, the commutativity of the diagram (\ref{lineA}) means that the diagram $$\matrix{ \L_{(g_1.g_2)x}&\fhd{\sim}{}&\L_{g_1(g_2.x)}\cr \fvb{}{\wr}&&\fvb{}{\wr}\cr \L_x&\fhg{\sim}{}&\L_{g_2.x}\cr}$$ is commutative and the commutativity of (\ref{lineB}) that the two isomorphisms $\L_{1.x}\simeq \L_x$ defined by the linearization $\lambda$ and $\epsilon$ respectively are the same. \section{An example} Let me recall that a closed point $x$ of $X$ has been fixed. Let $S$ be a $\k$-scheme. The $S$-points of the jacobian variety of $X$ are by definition isomorphism classes of line bundles on $X_S$ together whith a trivialization along $\{x\}\times S$ (such a pair will be called a rigidified line bundle). For the covenience of the reader, let me state this well known lemma which can be founf in SGA4, exp. XVIII, (1.5.4) \th Lemma \enonce The Picard stack ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ is canonically isomorphic (as a $\k$-group stack) to $JX\times BG_m$. \endth\label{torseur} {\it Proof}: let $f:\ {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X){\rightarrow} JX\times BG_m$ be the morphim which associates -to the line bundle $L$ on $X_S$ the pair $L\otimes L_{\mid \{x\}\times S}^{-1},L_{\mid \{x\}\times S}$ (thought as an object of $JX\times BG_m$ over $S$); -to an isomorphism $L\buildrel\sim\over{\rightarrow} L'$ on $X_S$ its restriction to $\{x\}\times S$. Let $f':\ JX\times BG_m{\rightarrow}{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ be the morphism which associates -to the pair $(L,V)$ where $L$ is a rigidified bundle on $X_S$ and $V$ a line bundle on $S$ (thought as an object of $JX\times BG_m$ over $S$), the line bundle $L\otimes_{X_R}V$; -to an isomorphism $(l,v):\ (L,V)\buildrel\sim\over{\rightarrow} (L',V')$ the tensor product $l\otimes_{X_S} v$. The morphisms $f$ and $f'$ are (quasi)-inverse each other and are morphisms of $\k$-stacks.\cqfd We will identify from now ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ and $JX\times BG_m$. Let $\L$ (resp. ${\bf P}$ and $\T$) be the universal bundle on $X\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ (resp. on $X\times JX$ and $BG_m$) and let $\Theta=(\det R\Gamma{\bf P})^{-1}$ be the theta line bundle on $JX$. The isomorphism $\L\buildrel\sim\over{\rightarrow}{\bf P}\otimes\T$ yields an isomorphism $$\det R\Gamma\L^n(m.x)\buildrel\sim\over{\rightarrow}\Theta^{-n^2}\otimes\T^{(m+1-g)}.\leqno({\global\def\currenvir{formule})$$\label{iso-det} \section{Descent of $\C$-line bundles} \label{simplicial} The object of this section is to prove the following statement \th Theorem \enonce Let $f:{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow}\E$ a $\C$-torsor as above. Let $\mathop{\rm Pic}\nolimits^\C({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})$ be the group of isomorphism classes of $\C$-linearized line bundles on ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$. Then, the pull-back morphism $f^*:\ \mathop{\rm Pic}\nolimits(\E)\buildrel\sim\over{\rightarrow}\mathop{\rm Pic}\nolimits^\C({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})$ is an isomorphism. \endth\label{theo} The descent theory of Grothendieck has been adapted in the case of algebraic $1$-stacks in [L-M], essentially in the proposition (6.23).Let ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_\bullet{\rightarrow}\E$ be the (augmented) simplicial complex of stacks coskeleton of $f$ (as defined in [De] (5.1.4) for instance). By proposition (6.23) of [L-M], one just has to construct a cartesian ${\cal O}_{D_\bullet}$-module $\L_\bullet$ such that $L_0$ is the ${\cal O}_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$-module $\L$ to prove the theorem. The $n$-th piece ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n$ is inductively defined by ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_0={\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M},\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n={\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\times_\E{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n-1}$ for $n>0$. Let $p_n:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ be the projection on the first factor. It is the simplicial morphism associated to the map $$\tp_n:\ \left\{\matrix{\Delta_0&{\rightarrow}&\Delta_n\cr 0&\longmapsto&0\cr}\right.$$ We define $\L_n$ by the morphism $$l_n:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n\fhd{p_n}{}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\fhd{l}{}BG_m\leqno{(\global\def\currenvir{formule)}.$$ \global\def\currenvir{subsection\label{rel-iso} Let $\delta_i$ (resp. $s_j$) be the face (resp. degeneracy) operators (see [De] (5.1.1) for instance) (by abuse of notation, we use the same notation for $\delta_j,s_j$ and their image by ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_\bullet)$. The category $(\Delta)$ is generated by the face and degeneracy operators with the following relations (see for instance the proposition VII.5.2 page 174 of [McL]) $$\matrix{\delta_i\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_j&=&\delta_{j+1}\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_i&\kern .5cm&i\leq j\cr}\leqno{(\global\def\currenvir{formule)}$$\label{A} $$\matrix{s_j\kern 1pt{\scriptstyle\circ}\kern 1pt s_i&=&s_i\kern 1pt{\scriptstyle\circ}\kern 1pt s_{j+1}&\kern .5cm&i\leq j\cr}\leqno{(\global\def\currenvir{formule)}$$\label{B} $$\left\{\matrix{ s_j\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_i&=&\delta_i\kern 1pt{\scriptstyle\circ}\kern 1pt s_{j-1}&\kern .5cm&i<j\cr &=&1&\kern .5cm&i=j,i=j+1\cr &=&\delta_{i-1}\kern 1pt{\scriptstyle\circ}\kern 1pt s_j&\kern .5cm&i>j+1.\cr } \right.\leqno{(\global\def\currenvir{formule)}$$\label{C} Therefore, the data of a cartesian ${\cal O}_{{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_\bullet}$-module $\L_\bullet$ is equivalent to the data of isomorphisms $\alpha _j:\ \delta_j^*\L_n\buildrel\sim\over{\rightarrow}\L_{n+1},\ j=0,\ldots,n+1$ and $\beta _j:\ s_j^*\L_{n+1}\buildrel\sim\over{\rightarrow}\L_{n},\ j=0,\ldots,n$ (where $n$ is a non negative integer) which are compatible with the relations \ref{A}, \ref{B} and \ref{C}. Let $n$ be a non negative integer. \global\def\currenvir{subsection We have first to define for $j=0,\ldots,n+1$ an isomorphism $\alpha _j:\ \delta_j^*\L_n\buildrel\sim\over{\rightarrow}\L_{n+1}$. The line bundle $\delta_j^*\L_n$ is defined by the morphism $l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_j:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+1}{\rightarrow} BG_m$ and $\tp_n\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_j$ is associated to the map $$\left\{\matrix{\Delta_0&{\rightarrow}&\Delta_{n+1}\cr 0&\longmapsto&\delta_j(0)\cr}\right.$$ If $j\not=0$, one has therefore $\tp_n\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_j=\tp_{n+1}$ and $\delta_j^*\L_n=L_{n+1}$. We define $\alpha _j$ by the identity in this case. Suppose now that $j=0$. Let $\pi_n:\ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_1$ be the projection on the $2$ first factors (associated to the canonical inclusion $\Delta_1\hookrightarrow\Delta_n$. The commutativity of the $2$ diagrams $$\matrix{ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+1}&\fhd{\delta_0}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n}\cr\fvb{}{\pi_{n+1}}&&\fvb{p_{n}}{}\cr{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_1&\fhd{\delta_0}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\cr} \ {\rm and}\ \matrix{ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+1}&\fhd{p_{n+1}}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\cr\fvb{}{\pi_{n+1}}&&\fvh{\delta_1}{}{}\cr{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_1&=&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_1\cr}$$ allows to reduce the problem to the construction of an isomorphism $$\delta_0^*\L\buildrel\sim\over{\rightarrow} \delta_1^*\L\ {\rm where}\ \delta_i,{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_1{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\ i=0,1$$ are the face morphisms or, what is amounts to the same, to the construction of a $2$-morphism $\nu:\ l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_0\Rightarrow l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_1$ (the morphism $\alpha _j$ will be $\alpha _j=\nu*\pi_{n+1}$). Now the diagram $$\matrix{ &&BG_m\cr &\llap{${}^{l\kern 1pt{\scriptstyle\circ}\kern 1pt m}$}\nearrow&\fvh{}{l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_0}\cr \C\times\X&\fhd{pr_2\times m}{}&\X\times_\E\X\cr &\llap{${}_{l\kern 1pt{\scriptstyle\circ}\kern 1pt pr_2}$}\searrow&\fvb{}{l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_1}\cr &&BG_m\cr}\leqno{(\global\def\currenvir{formule)}$$ is strict commutative and $pr_2\times m$ is an equivalence by the definition of a torsor. By the lemma \ref{lemme-categorie}, the $2$-morphism $\lambda$ induces a canonical $2$-morphism $\lambda*(pr_2\times m)^{-1}:\ l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_0\Rightarrow l\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_1$ which is the required $2$-morphism $\nu$. \global\def\currenvir{subsection We have then to define for $j=0,\ldots,n$ an isomorphism $\beta _j:\ s_j^*\L_{n+1}\buildrel\sim\over{\rightarrow}\L_{n}$. The line bundle $s_j^*\L$ is defined by the morphism $l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt s_j$ and $p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt s_j$ is associated to the canonical inclusion $\Delta_0\hookrightarrow{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n$ which means $p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt s_j=p_n$. Therefore, $s_j^*\L_{n+1}=\L_n$ and we define $\beta_j$ to be the identity. \global\def\currenvir{subsection We have to show that the data $\L_\bullet,\alpha _j,\beta_j,\ j\geq 0$ defines a line bundle on the simplicial stack ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_\bullet$ as explained in (\ref{rel-iso}). Notice that the fact that the definition of the $\beta_j$'s is compatible with the relations \ref{B} is tautological ($\beta_j$ is the identity on the relevant $\L_n$). \global\def\currenvir{subsection Relation \ref{A}: in terms of $l$, this relation means the following. We have the 2 stricltly commutative diagrams $$\alpha _i\kern 1pt{\scriptstyle\circ}\kern 1pt(\delta_i*\alpha _j):\ l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_j\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_i\FFhd{\delta_i*\alpha _j}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_i\FFhd{\alpha _i}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+2}$$ diagrams $$\matrix{ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+2}&\fhd{\delta_i}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+1}&\fhd{\delta_j}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n\cr &\llap{$\scp p_{n+2}$}\searrow&\fvb{}{\scp p_{n+1}}&\swarrow\rlap{$\scp p_n$}{}\cr &&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}&\fhd{l}{}BG_m\cr }\ {\rm and}\ \matrix{ {\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+2}&\fhd{\delta_{j+1}}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_{n+1}&\fhd{\delta_i}{}&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}_n\cr &\llap{$\scp p_{n+2}$}\searrow&\fvb{}{\scp p_{n+1}}&\swarrow\rlap{$\scp p_n$}{}\cr &&{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}&\fhd{l}{}BG_m\cr }$$ inducing the two $2$-morphisms $$ \alpha _i\kern 1pt{\scriptstyle\circ}\kern 1pt(\alpha _j*\delta_i):\ l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_j\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_i\FFhd{\alpha _j*\delta_i}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_i\FFhd{\alpha _i}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+2}$$ and $$ \alpha _{j+1}\kern 1pt{\scriptstyle\circ}\kern 1pt(\alpha _i*\delta_{j+1}):\ l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_i\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_{j+1}\FFhd{\alpha _i*\delta_{j+1}}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+1}\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_{j+1}\FFhd{\alpha _{j+1}}{}l\kern 1pt{\scriptstyle\circ}\kern 1pt p_{n+2}.$$ The relation \ref{A} means exactly the equality $$\alpha _i\kern 1pt{\scriptstyle\circ}\kern 1pt(\alpha _j*\delta_i)=\alpha _{j+1}\kern 1pt{\scriptstyle\circ}\kern 1pt(\alpha _i*\delta_{j+1}), \ i\leq j.\leqno(\ref{A}')$$ If $j=0$, the relation \ref{A}' is just by definition of $\alpha _j$ the condition \ref{lineA} (see remark \ref{rem-line}). If $j>0$, both the 2 isomorphisms $\alpha_j$ and $\alpha_{j+1}$ are the relevant identity and the relation \ref{A}' is tautological. \global\def\currenvir{subsection Relation \ref{C}: the only non tautological relation in (\ref{C}) is the associated to the equality $s_0\kern 1pt{\scriptstyle\circ}\kern 1pt \delta_0=1$ in $(\Delta)$ which means as before that $\alpha _0*\delta_0$ is the identity functor of $l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n=l\kern 1pt{\scriptstyle\circ}\kern 1pt p_n\kern 1pt{\scriptstyle\circ}\kern 1pt\delta_0\kern 1pt{\scriptstyle\circ}\kern 1pt s_0$. But, this is exactly the meaning of the relation \ref{lineB} (see remark \ref{rem-line}). \section{Application to the Picard groups of some moduli spaces} Let us chose 3 integers $r,s,d$ such that $$r\geq 2\ {\rm and}\ s\mid r\mid ds.$$ If $G$ is the group $\mathop{\bf SL}\nolimits_r/\mu_s$ we denote as in [B-L-S] by $\M_G(d)$ the moduli stack of $G$-bundles on $X$ and by $\M_{\mathop{\bf SL}\nolimits_r}(d)$ the moduli stack of rank $r$ vector bundles and determinant ${\cal O}(d.x)$. If $r=s$ (i.e. $G={\bf PSL}_r$), the natural morphism of algebraic stacks $$\pi:\ \M_{\mathop{\bf SL}\nolimits_r}(d){\rightarrow}\M_G(d)$$ is a ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_r$-torsor (see the corollary of proposition 2 of [Gr] for instance). Let me explain how to deal with the general case. \global\def\currenvir{subsection Let $E$ be a a rank $r$ vector bundle on $X_S$ endowed with an isomorphism $\tau;\ D^{r/s}\buildrel\sim\over{\rightarrow}\det(E)$ where $D$ is some line bundle. Let me define the $\mathop{\bf SL}\nolimits_r/\mu_s$-bundle $\pi(E)$ associated to $E$ (more precisely to the pair $(E,\tau)$). \th Definition \enonce An $s$-trivialization of $E$ on the \'etale neighborhood $T{\rightarrow} X_S$ is a triple $(M,\alpha ,\sigma)$ where $\alpha:\ D\buildrel\sim\over{\rightarrow} M^s$ is an isomorphism ($M$ is a line bundle on $T$); $\sigma:\ M^{\oplus r}\buildrel\sim\over{\rightarrow} E_T$ is an isomorphism; $\det(\sigma)\kern 1pt{\scriptstyle\circ}\kern 1pt \alpha^{r/s}:\ D^{r/s}\buildrel\sim\over{\rightarrow}\det(E)$ is equal to $\tau$. Two $s$-trivializations $(M,\alpha ,\sigma)$ and $(M',\alpha' ,\sigma')$ of $E$ will be said equivalent if there exists an isomorphism $\iota: M\buildrel\sim\over{\rightarrow} M'$ such that $\iota^s\kern 1pt{\scriptstyle\circ}\kern 1pt\alpha =\alpha'$. \endth The principal homogeneous space $$T\longmapsto \{\hbox{equivalence classes of $s$-trivializations of } E_T\}$$ defines the $\mathop{\bf SL}\nolimits_r/\mu_s$-bundle $\pi(E)$\footnote{${}^\dagger$}{\sevenrm We see here a $\scp G$-bundle as a formal homogeneous space under $\scp G$.}. Now, the construction is obviously functorial and therefore defines the morphsim $\pi:\ \M_{\mathop{\bf SL}\nolimits_r}(d){\rightarrow} \M_G(d)$ (observe that an object $E$ of $\M_{\mathop{\bf SL}\nolimits_r}(d)$ has determinant ${\cal O}({ds\over r}.x)^{r/s}$). Let $L$ be a line bundle and $(M,\alpha,\tau)$ an $s$-trivialization of $E_T$. Then, $(M\otimes L,\alpha\otimes \hbox{\rm Id}_{L^s},\sigma\otimes \hbox{\rm Id}_L)$ is an $s$-trivialization of $E\otimes L$ (which has determinant ($D\otimes L^s)^{r/s}$). This shows that there exists a canonical functorial isomorphism $$\pi(E)\buildrel\sim\over{\rightarrow}\pi(E\otimes L)\leqno{(\global\def\currenvir{formule)}.$$\label{iso-pro} In particular, $\pi$ is ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s$-equivariant. \th Lemma \enonce The natural morphism of algebraic stacks $\pi:\ \M_{\mathop{\bf SL}\nolimits_r}(d){\rightarrow}\M_G(d)$ is a ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s$-torsor. \endth \label{pi-tors} {\it Proof}: let $E,E'$ be two rank $r$ vector bundles on $X_S$ (with determinant equal to ${\cal O}(d.x)$) and let $\iota:\ \p(E)\buildrel\sim\over{\rightarrow}\p(E)'$ an isomorphism. As in the proof of the lemma 13.4 of [B-L-S], we have the exact sequence of sets $$1{\rightarrow}\mu_s{\rightarrow} {\rm Isom}(E,E'){\rightarrow}{\rm Isom}(\p(E),\p(E)')\fhd{\pi_{E,E'}}{} H^1_{\rm \acute et}(X_S,\mu_s).$$ Let $L$ be a $\mu_s$-torsor such that $\pi_{E,E'}(\iota)=[L]$. Then, $\p(E\otimes L)$ is canonically equal to $\p(E)$ and $\pi_{E\otimes L,E'}=0$ and $\iota$ is induced by an isomorphism $E\otimes L\buildrel\sim\over{\rightarrow} E'$ well defined up to multiplication by $\mu_s$. The lemma follows.\cqfd \global\def\currenvir{subsection Let ${\cal U}}\def\d{{\cal D}}\def\F{{\cal F}$ be the universal bundle on $X\times \M_{\mathop{\bf SL}\nolimits_r}(d)$. We would like to know which power of the determinant bundle $\d=(\det R\Gamma{\cal U}}\def\d{{\cal D}}\def\F{{\cal F})^{-1}$ on $\M_{\mathop{\bf SL}\nolimits_r}(d)$ descends to $\M_G(d)$. As in I.3 of [B-L-S], the rank $r$ bundle $\F=\L^{\oplus(r-1)}\oplus\L^{1-r}(d.x)$ on $X\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ has determinant ${\cal O}(d.x)$ and therefore defines a morphism $$f:\ {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)=JX\times BG_m{\rightarrow}\M_{\mathop{\bf SL}\nolimits_r}(d)$$ which is ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s$-equivariant. The vector bundle $\F'={\cal O}^{\oplus (r-1)}\oplus\L^{-r/s}(d.x)$ on $X\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ has determinant $[\L^{-1}({ds\over r}.x)]^{r/s}$. The $G$-bundle $\pi(\F')$ on $X\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ defines a morphism $f':\ {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}{\rightarrow}\M_G(d)$. The relation $\L\otimes (\hbox{\rm Id}_X\times s_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H})^**(\F')=\F$ and (\ref{iso-pro}) gives an isomorphism $\pi(\F)=(\hbox{\rm Id}_X\times s_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H})^*\pi(\F')$ which means that the diagram $$\matrix{ {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)&\fhd{f}{}&\M_{\mathop{\bf SL}\nolimits_r}(d)\cr \fvb{}{s_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}}&&\fvb{}{\pi}\cr {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)&\fhd{f'}{}&\M_G(d)\cr}\leqno{(\global\def\currenvir{formule)}$$\label{dia-rest} is $2$-commutative. Exactly as in I.3 of [B-L-S], let me prove the \th Lemma \enonce The line bundle $f^*\d^k$ on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ descends through $s_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$ if and only if $k$ multiple of $s/(s,r/s)$. \endth\label{lem-rest} {\it Proof}: let $\chi=r(g-1)-d$ be the opposite of the Euler characteristic of ($\k$-)points of $\M_{\mathop{\bf SL}\nolimits_r}(d)$. By (\ref{iso-det}), one has an isomorphism $f^*\d^k\buildrel\sim\over{\rightarrow}\Theta^{kr(r-1)}\otimes\T^{k\chi}.$ The theory of Mumford groups says that $\Theta^{kr(r-1)}$ descends through $s_J$ if and only if $k$ is a multiple of $s/(s,r/s)$. The line bundle $\T^{k\chi}$ on $BG_m$ descends through $s_{BG_m}$ if and only if $k\chi$ is a multiple of $s$. The lemma follows from the above isomorphism and from the observation that the condition $s\mid r\mid ds$ forces $s\chi$ to be a multiple of $s$.\cqfd \global\def\currenvir{rem\label{rema} If $g=0$, the jacobian $J$ is a point and the condition on $\Theta$ is empty. The only condition is in this case being $k\chi$ multiple of $s$. Let me recall that $\d$ is the determinant bundle on $\M_{\mathop{\bf SL}\nolimits_r}(d)$ and $G=\mathop{\bf SL}\nolimits_r/\mu_s$. \th Theorem \enonce Assume that the characteristic of $\k$ is $0$. The integers $k$ such that $\d^k$ descends to $\M_G(d)$ are the multiple of $s/(s,r/s)$. \endth\label{theo-pic} {Proof}: by lemma \ref{lem-rest} and diagram (\ref{dia-rest}), we just have to proving that $\d^k$ efectively descends where $k=s/(s,r/s)$. By theorem \ref{theo} and lemma \ref{pi-tors}, this means exactly that $\d^k$ has a ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s$-linearization. We know by lemma \ref{lem-rest} that the pull-back $f^*\d^k$ has such a linearization. \th Lemma \enonce The pull-back morphism $\mathop{\rm Pic}\nolimits({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times\M_{SL_r}(d)){\rightarrow}\mathop{\rm Pic}\nolimits({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X))$ is injective. \endth {\it Proof}: by lemma \ref{torseur}, one is reduced to prove that the natural morphism $$\mathop{\rm Pic}\nolimits(B\mu_s\times\M_{SL_r}(d)){\rightarrow}\mathop{\rm Pic}\nolimits(B\mu_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X))$$ is injective. Let ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$ be any stack. The canonical morphism ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}{\rightarrow}{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\times B\mu_s$ is a $\mu_s$-torsor (with the trivial action of $\mu_s$ on ${\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}$). By theorem \ref{theo}, one has the equality $$\mathop{\rm Pic}\nolimits({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\times B\mu_s)=\mathop{\rm Pic}\nolimits^{\mu_s}({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}).$$ Assume further that $H^0({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M},{\cal O})=\k$. The later group is then canonically isomorphic to $$\mathop{\rm Pic}\nolimits({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})\times\mathop{\rm Hom}\nolimits(\mu_s,G_m)=\mathop{\rm Pic}\nolimits({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})\times\mathop{\rm Pic}\nolimits(B\mu_s).$$ All in all, we get a functorial isomorphism $$\iota_{\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}:\ \mathop{\rm Pic}\nolimits({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M}\times B\mu_s)\buildrel\sim\over{\rightarrow}\mathop{\rm Pic}\nolimits({\cal X}}\def\C{{\cal G}}\def\L{{\cal L}}\def\E{{\cal Y}}\def\m{{m_\C}}\def\M{{\cal M})\times\mathop{\rm Pic}\nolimits(B\mu_s).\leqno(\global\def\currenvir{formule)\label{pic-pro}$$ By [L-S], the Picard group of $\M_{\mathop{\bf SL}\nolimits_r}(d)$ is the free abelian group ${\bf} Z}\def\Q{{\bf} Q}\def\C{{\bf} C}\def\F{{\bf} F.\d$ and the formula (\ref{iso-det}) proves that $$f^*:\ \mathop{\rm Pic}\nolimits(\M_{\mathop{\bf SL}\nolimits_r}(d)){\rightarrow}\mathop{\rm Pic}\nolimits({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X))$$ is an injection. The lemma folows from the commutative diagram $$\diagram{ \mathop{\rm Pic}\nolimits(\M_{\mathop{\bf SL}\nolimits_r}(d))\times\mathop{\rm Pic}\nolimits(B\mu_s)&\hookrightarrow&\mathop{\rm Pic}\nolimits({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X))\times\mathop{\rm Pic}\nolimits(B\mu_s)\cr \fvb{\iota_\M}{\wr}&&\fvb{\iota_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}}{\wr}\cr \mathop{\rm Pic}\nolimits(\M_{\mathop{\bf SL}\nolimits_r}(d)\times B\mu_s)&{\rightarrow}&\mathop{\rm Pic}\nolimits({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)\times B\mu_s)\cr }$$\cqfd Let $\H$ (resp. $\H_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$) be the line bundle on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times\M_{\mathop{\bf SL}\nolimits_r}(d)$ (resp. ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ $$\H=\mathop{{\cal H}om}\nolimits(m_\M^*\d^k,pr_2^*\d^k)\ {\rm resp.}\ \H_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}=\mathop{{\cal H}om}\nolimits(m_\M^*f^*\d^k,pr_2^*f^*\d^k).$$ Let us chose a ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s$-linearization $\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$ of $f^*\d^k$. It defines a trivialization of the line bundle $\H_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$. The equivariance of $f$ implies (cf. \ref{equi}) that there exists a (compatible) $2$-morphism $$q:\ m_\M\kern 1pt{\scriptstyle\circ}\kern 1pt(\hbox{\rm Id}\times f)\Rightarrow f\kern 1pt{\scriptstyle\circ}\kern 1pt m_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$$ making the diagram $$ \matrix{ {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)&\fhd{m_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}}{}&{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)\cr \fvb{{\rm Id}\times f}{}&&\fvb{f}{}\cr {\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times\M_{SL_r}(d)&\fhd{m_\M}{}&\M_{SL_r}(d)\cr }$$ $2$-commutative. The $2$-morphism $q$ defines an isomorphism from the pull-back $m_\M^*\d^k$ on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ to $m_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}^*(f^*\d^k)$. The pull-back of $pr_2^*\d^k$ on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ is tautologically isomorphic to $pr_2^*(f^*\d^k)$. The preceding isomorphisms induce an isomorphism $$(\hbox{\rm Id}\times f)^*\H\buildrel\sim\over{\rightarrow}\H_J.$$ The later line bundle being trivial, so is $(\hbox{\rm Id}\times f)^*\H$. The lemma above proves therefore that $\H$ istself is {\it trivial} . Each ($\k$-)point $j$ of $J_s$ defines a morphism $\M_{SL_r}(d){\rightarrow}{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times\M_{SL_r}(d)$ (resp. ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X){\rightarrow}{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$); let me denote by $\H_j$ (resp. $f^*\H_j$) the pull-back of $\H$ (resp. $(\hbox{\rm Id}\times f)^*\H)$ by this morphism. The pull-back morphism $$H^0({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times\M_{\mathop{\bf SL}\nolimits_r}(d),\H){\rightarrow} H^0({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X),(\hbox{\rm Id}\times f)^*\H)$$ can be identified to the direct sum $$\oplus_{j\in{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s(\k)}H^0(\M_{\mathop{\bf SL}\nolimits_r}(d),\H_j){\rightarrow} H^0({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X),f^*\H_j).$$ Because $$H^0(\M_{\mathop{\bf SL}\nolimits_r}(d),{\cal O})=H^0({\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X),{\cal O})=\k\leqno{(\global\def\currenvir{formule)},\label{cons}$$ this morphism is a direct sum of non-zero morphisms of vector spaces of dimension $1$ and therefore an isomorphism. In particular, a linearization $\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$ of $f^*\d^k$ defines canonicaly an isomophism $$\lambda_\M:\ m_\M^*\d^k\buildrel\sim\over{\rightarrow} pr_2^*\d^k$$ such that $(\hbox{\rm Id}\times f)^*\lambda_\M=\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$. Explicitely, $\lambda_\M$ is characterized as follows: let $x$ be an object of $\M_{\mathop{\bf SL}\nolimits_r}(d)$ over a connected scheme $S$ and $g$ an object of ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s(S)={\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s(\k)$. The preceding dicussion means that the functorial isomorphisms $$\lambda_\M(g,x):\ \d^k_{g.x}\buildrel\sim\over{\rightarrow}\d^k_x$$ are determined when $x$ lies in the essential image of $f$. In this case, let us chose an isomorphism $f(x')\buildrel\sim\over{\rightarrow} x$ (inducing an isomorphism $g.f(x')\buildrel\sim\over{\rightarrow} g.x$). Then, the diagram of isomorphisms of line bundles on $S$ $$\matrix{ L'_{x'}=L_{f(x')}&{\rightarrow}&L_x\cr \fvh{\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(g,x')}{}&&&\llap{$\scp\lambda_\M(g,x)$}\nwarrow&\cr L'_{g.x'}=L_{f(g.x')}&\fhd{q_{g,x'}}{}&L_{g.f(x')}&{\rightarrow}&L_{g.x}\cr }$$ is commutative (where $L=\d^k$ and $L'=f^*\d^k$). Now, the pull-back of $\lambda_\M$ on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}_s\times{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$ satisfies conditions \ref{lineA} and \ref{lineB}. Using (\ref{cons}) and the equivariance of of $f$ as above, this shows that $\lambda_\M$ is a linearization. For instance, keeping the notation above, let us check the condition \ref{lineB}. We have to check that the isomorphism $\iota$ of $L$ induced by $\epsilon$ is the identity. As above, it is enough to check that on ${\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(X)$. By definition, with a slight abuse of notations, the diagrams $$\matrix{ L'_{x'}=L_{f(x')}&{\rightarrow}&L_x\cr \fvh{\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(1,x')}{}&&&\llap{$\scp\lambda_\M(1,x)$}\nwarrow&\cr L'_{1.x'}=L_{f(1.x')}&\fhd{q_{1,x'}}{}&L_{1.f(x')}&{\rightarrow}&L_{1.x}\cr }\kern .5cm {\rm and}\kern .5cm \matrix{ L_x&\fhd{\iota}{}&L_x\cr &\llap{$\scp\lambda_\M(1,x)$}\nwarrow&\fvh{\epsilon(x)}{}\cr &&L_{1.x}\cr }$$ Because $\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}$ is a linearization, condition \ref{lineB} gives the commutative diagram $$\matrix{ L'_{x'}&=&L'_{x'}\cr &\llap{$\scp\lambda_{\cal J}}\def\p{{\pi}}\def\k{{\bf k}}\def\P{{\cal P}}\def\T{{\cal T}}\def\H{{\cal H}(1,x')$}\nwarrow&\fvh{}{\epsilon'(x')}\cr &&L_{1.x'}\cr }$$ showing that the equality $\iota=\hbox{\rm Id}$ remains to prove the commutativity of the diagram $$\matrix{ L_{f(1.x')}&\fhd{\epsilon'}{}&L_{f(x')}\cr \fvb{q_{1,x'}}{}&&\parallel\cr L_{1.f(x')}&\fhd{\epsilon}{}&L_{f(x')}\cr } $$ But this follows from the commutativity of the diagram $$\matrix{ f(1.x')&\fhd{\epsilon'}{}&f(x')\cr \fvb{q_{1,x'}}{}&&\parallel\cr 1.f(x')&\fhd{\epsilon}{}&f(x')\cr } $$ which is by definition the fact that $q$ is compatible to $\epsilon$ as required in (\ref{equi}). One would check condition \ref{lineA} in an analogous way.\cqfd \global\def\currenvir{rem\label{rem}In the case $g=0$, the condition is an in remark \ref{rema}. \global\def\currenvir{rem This linearization can be certainly also deduced from a careful analysis of the first section of [Fa], but the method above seems simpler. \bigskip \centerline{\bf References} \medskip [B-L-S] A.Beauville, Y. Laszlo, C. Sorger, {\it The Picard group of the moduli of $G$-bundles on a curve}, preprint alg-geom/9608002. \medskip [Br] L. Breen, {\it Bitorseurs et cohomologie non ab\'elienne }, in The Grothendieck Festschrift I, Progr. Math. {\bf 86}, Birkh\"auser (1990), 401-476. \medskip [De] P. Deligne, {\it Th\'eorie de Hodhe III}, Publ. Math. I.H.E.S. {\bf 44} (1974), 5-78. \medskip [Fa] G. Faltings, {\it Stable $G$-bundles and Projective Connections}, JAG {\bf 2} (1993), 507-568. \medskip [Gr] A. Grothendieck, {\it G\'eom\'etrie formelle et g\'eom\'etrie alg\'ebrique}, FGA, S\'em. Bourbaki {\bf 182} (1958/59), 1-25. \medskip [L-M] G. Laumon, L. Moret-Bailly, {\it Champs alg\'ebriques}, preprint Universit\'e Paris-Sud (1992). \medskip [L-S] Y. Laszlo, C. Sorger, {\it The line bundles on the moduli of parabolic $G$-bundles over curves and their sections}, preprint alg-geom/9507002. \bigskip [McL] S. Mac Lane, {\it Categories for the working mathematician}, GTM {\bf 5}, Springer-Verlag (1971). \bigskip \hfill\hbox to 5cm{\hfill Y. {\pc LASZLO}\hfill} \smallskip \hfill\hbox to 5cm{\hfill DMI - \'Ecole Normale Sup\'erieure\hfill} \smallskip \hfill\hbox to 5cm{\hfill ( URA 762 du CNRS )\hfill} \smallskip \hfill\hbox to 5cm{\hfill 45 rue d'Ulm\hfill} \smallskip \hfill\hbox to 5cm{\hfill F-75230 {\pc PARIS} Cedex 05\hfill} \smallskip \closeall\bye
1996-10-09T20:46:51
9610
alg-geom/9610008
en
https://arxiv.org/abs/alg-geom/9610008
[ "alg-geom", "math.AG" ]
alg-geom/9610008
Jim Bryan
Jim Bryan and Marc Sanders
The Rank Stable Topology of Instantons on $\cpbar$
LaTeX2e
null
null
null
null
Let $\M_{k}^{n}$ be the moduli space of based (anti-self-dual) instantons on $\cpbar$ of charge $k$ and rank $n$. There is a natural inclusion of rank $n$ instantons into rank $n+1$. We show that the direct limit space is homotopy equivalent to $BU(k)\times BU(k)$. The moduli spaces also have the following algebro-geometric interpretation: Let $\linf$ be a line in the complex projective plane and consider the blow-up at a point away from $\linf$. $\M _{k}^{n}$ can be described as the moduli space of rank $n$ holomorphic bundles on the blownup projective plane with $c_{1}=0$ and $c_{2}=k$ and with a fixed holomorphic trivialization on $\linf$.
[ { "version": "v1", "created": "Wed, 9 Oct 1996 18:41:28 GMT" } ]
2008-02-03T00:00:00
[ [ "Bryan", "Jim", "" ], [ "Sanders", "Marc", "" ] ]
alg-geom
\section{Introduction} In his 1989 paper \cite{Ta}, Taubes studied the stable topology of the based instanton moduli spaces. He showed that if ${\mathcal{M}}_{k}^{n}(X)$ denotes the moduli space of based $SU(n)$-instantons of charge $k$ on $X$, then there is a map ${\mathcal{M}} _{k}^{n}(X)\to {\mathcal{M}} _{k+1}^{n}(X)$ and, in the direct limit topology, ${\mathcal{M}} _{\infty }^{n}(X)$ has the homotopy type of $\operatorname{Map}_{0}(X,BSU(n))$. There is also a map ${\mathcal{M}} _{k}^{n}(X)\hookrightarrow {\mathcal{M}} _{k}^{n+1}(X)$ given by the direct sum of a connection with the trivial connection on a trivial line bundle and one can consider the direct limit ${\mathcal{M}} _{k}^{\infty }(X)$. For the case of $X=S^{4}$ with the round metric, it was shown by Kirwan and also by Sanders (\cite{Kir},\cite{Sa}) that the direct limit has the homotopy type of $BU(k)$. In this note we consider the case of $X=\cpbar $ where $\cpbar $ denotes the complex projective plane with the Fubini-Study metric and the opposite orientation of the one induced by the complex structure. Our result is: \begin{theorem}\label{thm: main result} ${\mathcal{M}} _{k}^{\infty }(\cpbar )$ has the homotopy type of $BU(k)\times BU(k)$. \end{theorem} The main tool in the proof of the theorem is a construction of the moduli spaces ${\mathcal{M}} _{k} ^{n}(\cpbar )$ due to King \cite{Ki}. In general, Buchdahl \cite{Bu} has shown that, for appropriate metrics on the $N$-fold connected sum $\# _{N}\cpbar $, the moduli spaces ${\mathcal{M}} _{k}^{n}(\# _{N}\cpbar )$ are diffeomorphic to certain spaces of equivalence classes of holomorphic bundles on $\cnums \P^{2}$ blown-up at $N$ points. The universal $U(k)\times U(k) $ bundle that appears giving the homotopy equivalence of theorem \ref{thm: main result} can be constructed as higher direct image bundles (see section \ref{section:proof}). \begin{rem}\label{rem:motivation} The cofibration $S^2\to \cpbar \to S^4$ gives rise to the fibration of mapping spaces $\Omega^4BSU(n)\to {\rm Map}_{\ast}(\cpbar , BSU(n))\to \Omega^2BSU(n)$ which for K-theoretic reasons is a trivial fibration in the limit over $n$. The total space of this fibration is homotopy equivalent to the space of based gauge equivalence classes of all connections on $\cpbar$. Thus, from Taubes' result, ${\mathcal{M}}_k^\infty $ must have the property that taking the limit over $k$ gives $BU\times BU$. For $S^4$, similar remarks imply that $\lim_{k\to \infty} {\mathcal{M}} _{k} ^{\infty}(S^4)\simeq BU$ and the inclusion of ${\mathcal{M}} _k ^{\infty} (S^4)$ into this limit has been shown to be (up to homotopy) the natural inclusion $BU(k)\hookrightarrow BU$ (\cite {Sa}). Theorem 1.1 and these results for $S^4$ suggest a general conjecture which is supported by the fact that the higher direct image bundle giving our homotopy equivalence generalizes in an appropriate way. \end{rem} \begin{conj}\label{conj:N fold connected sum} For appropriate metrics on $\# _{N}\cpbar $, ${\mathcal{M}} _{k}^{\infty }(\# _{N}\cpbar )$ has the homotopy type of a product $BU(k)\times \cdots \times BU(k)$ with $N+1$ factors. \end{conj} \begin{rem}\label{rem:Bott periodicity} Combining theorem \ref{thm: main result} with Taubes' stabilization result leads to an alternate proof of Bott periodicity for the unitary group. There is a natural map from instantons on $S^4$ to those on $\cpbar$ (given by pull-back) which in the limit as $n\to \infty$ is homotopy equivalent to the diagonal $BU(k) \to BU(k)\times BU(k)$. Taking the limit as $k\to\infty$ and applying Taubes' result, the diagonal map appears as the inclusion of fibers in the fibration $BU\simeq\Omega_k^4BU\to BU\times BU\to \Omega^2(BSU)$ (see remark 1.1). However, the map $BU\rightarrow BU \times BU$ has homotopy fiber $U\times U \bigg/ U\: \simeq U$, and therefore, given the above fibration we must have $U\simeq\Omega^3BSU\simeq\Omega^2 SU\simeq\Omega^2 U$. Tian (\cite {Ti}) noticed that results for limits of instantons on $S^4$ (see \cite {Kir} or \cite {Sa}) already imply the four-fold periodicity $\Omega^4BU\simeq Z\times BU$. The ${\cpbar}$ case thus gives the finer two-fold periodicity $\Omega^2BU\simeq Z\times BU$, as one might expect due to the nontrivial $S^2\hookrightarrow {\cpbar}$. In a future paper we will study limits of $Sp(n)$ and $SO(n)$ instantons on $\cpbar$ and their relationships to those on $S^4$. As an amusing corollary, we will be able to rederive many of the Bott periodicity relationships among $Sp$, $U$, $SO$, and their homogeneous spaces. \end{rem} \section{The construction of ${\mathcal{M}} _{k}^{n}(\cpbar )$} Let $x_{0}\in \cpbar $ be the base point. Since $\cpbar -\{x_{0} \}$ is conformally equivalent to $\til{\cnums} ^{2}$, the complex plane blown-up at the origin, ${\mathcal{M}} _{k}^{n}(\cpbar )$ can be regarded as instantons on $\til{\cnums} ^{2}$ based ``at infinity''. Buchdahl \cite{Bu} proved an analogue in this non-compact setting of Donaldson's theorem relating instantons to holomorphic bundles: Let $\til{\cnums}^{2}_{N}$ be the complex plane blown-up at $N$ points with a K\"ahler metric. Then $\til{\cnums}^{2}_{N}$ has a ``conformal compactification'' to $\#_{N}\cpbar$ and a ``complex compactification'' to $\til{\cnums \P}^{2}_{N}$ (the projective plane blown-up at $N$ points). We have added a point $x_{0}$ in the former case and a complex projective line $\ell_{\infty }$ in the latter. Define $\msalg{k}{n}(\til{\cnums \P}^{2}_{N})$ to be the moduli space consisting of pairs $({\mathcal{E}},\tau )$ where ${\mathcal{E}} $ is a rank $n$ holomorphic bundle on $\til{\cnums \P} ^{2}_{N}$ with $c_{1}({\mathcal{E}} )=0$, $c_{2}({\mathcal{E}} )=k$, and where $\tau :{\mathcal{E}} |_{\ell_{\infty }}\to \cnums ^{n}\otimes \O _{\ell_{\infty }}$ is a holomorphic trivialization of ${\mathcal{E}} $ on $\ell_{\infty }$. There is a natural map $\Phi:{\mathcal{M}}_{k}^{n}(\# _{N}\cpbar )\longrightarrow\msalg{k}{n}(\til{\cnums \P}^{2}_{N})$ defined as follows. Let $p:\til{\cnums \P}^{2}_{N}\to \# _{N}\cpbar $ be the map that collapses $\ell_{\infty }\mapsto x_{0}$. If $[A]\in\ms{k}{n}$ then the $\bar\partial$ operator that defines the holomorphic bundle $\mathcal{V}=\Phi(A)$ is taken to be $(d_{p^{*}(A)})^{(0,1)}$, the anti-holomorphic part of the covariant derivative defined by the pullback of the connection. The anti-self-duality of $A$ implies that the curvature of $p^{*}(A)$ is a $(1,1)$-form and so $\bar\partial ^{2}=0$. Buchdahl's theorem is then \begin{theorem} \label{thm:buchdahl} The map $\Phi:{\mathcal{M}}_{k}^{n}(\# _{N}\cpbar )\longrightarrow\msalg{k}{n}(\til{\cnums \P}^{2}_{N})$ is a diffeomorphism. \end{theorem} The case $N=1$ was first proved by King \cite{Ki}. We now restrict ourselves to that case and simply write $\ms{k}{n}$ for ${\mathcal{M}}_{k}^{n}(\cpbar )$ and $\msalg{k}{n}(\til{\cnums \P}^{2})$. King constructed $\ms{k}{n}$ explicitly in terms of linear algebra data. We recall his construction. Consider configurations of linear maps: \begin{picture}(100,75)(-135,0) \put(20,50){$W_0$} \put(56,15){\vector(-2,3){21}} \put(55,5){$V_{\infty}$} \put(36,52){\vector(1,0){47}} \put(83,56){\vector(-1,0){47}} \put(84,50){$W_1$} \put(86,45){\vector(-2,-3){19}} \put(35,27){$b$} \put(56,45){$x$} \put(48,60){$a_1,a_2$} \put(81,27){$c$} \end{picture} \noindent where $W_0$, $W_1$ and $V_{\infty }$ are complex vector spaces of dimensions $k$, $k$, and $n$ respectively. A configuration $(a_1,a_2,b,c,x)$ is called {\em integrable} if it satisfies the equation $$a_1xa_2-a_2xa_1+bc=0.$$ A configuration $(a_1,a_2,b,c,x)$ is {\em non-degenerate} if it satisfies the following conditions: $$\forall \hspace{4pt} (\lambda_1,\lambda_2),(\mu_1,\mu_2)\in\cnums^2 \text{ such that }\lambda_1\mu_1 + \lambda_2\mu_2=0\text{ and }(\mu_1,\mu_2)\neq(0,0),$$ \begin{eqnarray*} \exists\hspace{-6pt}\raisebox{1.5pt}{/} v\in W_1\text{ such that } &&\left\{ \begin{array}{lr} xa_1v=\lambda_1v & (\mu_1 a_1 + \mu_2 a_2)v=0 \\ xa_2v=\lambda_2v & cv=0 \end{array} \right. \\ \text{and }\exists\hspace{-6pt}\raisebox{1.5pt}{/} w\in W_0^\star\text{ such that } &&\left\{ \begin{array}{lr} x^*a_1^*w=\lambda_1w & (\mu_1 a^*_1 + \mu_2 a^*_2)w=0 \\ x^*a_2^*w=\lambda_2w & b^*w=0 \end{array} \right. \end{eqnarray*} Let $A_{k}^{n}$ be the space of all integrable non-degenerate configurations. $G=Gl(W_{0})\times Gl(W_{1})$ acts canonically on $A_k^n$. The action is explicitly given by $$ (g_{0},g_{1})\cdot (a_{1},a_{2},b,c,x)= (g_{0}a_{1}g_{1}^{-1},g_{0}a_{2}g_{1}^{-1}, g_{0}b,cg_{1}^{-1},g_{1}xg_{0}^{-1}) $$ \begin{theorem}\label{thm: monad description} The moduli space $\ms{k}{n}$ is isomorphic to $A_{k}^{n}/G$. \end{theorem} \proof King uses such configurations to determine monads that in turn determine holomorphic bundles. Configurations in the same $G$ orbit determine the same bundle. For the sake of brevity we refer the reader to \cite{Ki} or \cite{Br} for details. The construction identifies the vector spaces $W_0$ and $W_{1}$ canonically as $H^1(\mathcal{E}(-\ell_{\infty }))$ and $H^{1}(\mathcal{E}(-\ell_{\infty }+E))$ respectively, where $E\subset \til{\cnums \P}^{2}$ is the exceptional divisor. The vector space $V_{\infty } $ is identified with the fiber over $\ell_{\infty }$. \section{Proof of theorem \ref{thm: main result} }\label{section:proof} We prove the theorem in two steps: We first show that the space of monad data $A_{k}^{n}$ forms a principal $G=Gl(k)\times Gl(k)$ bundle over ${\mathcal{M}} _{k}^{n}$. We then show that the induced $G$-equivariant inclusion $A_k^n\hookrightarrow A_k^{n+2k}$ is null-homotopic so that we can conclude that $A_{k}^{\infty }$ is contractible. \begin{lem}\label{lem:freeness of G action} $G$ acts freely on the space of monad data $A_{k}^{n}$. \end{lem} \begin{proof} This is essentially proved in \cite{Ki} where it is implicitly shown that the non-degeneracy conditions are precisely the conditions that guarantee freeness. We point out that this also follows more conceptually from the existence of a universal family $\Bbb{E}\to \ms{k}{n}\times \til{\cnums \P} ^{2}$ and the cohomological interpretation of $W_{0}$ and $W_{1}$: First, the existence of a universal family can be shown via the gauge theoretic construction: Let $V$ be a smooth hermitian vector bundle on $\til{\cnums \P}^{2}$ with $c_{1}(V)=0$ and $c_{2}(V)=k$. Let $\mathcal{A}^{1,1}_{0}$ denote unitary connections on $V$ with curvature of pure type $(1,1)$ and that restrict to the trivial connection on $\ell_{\infty } $ and let $\mathcal{G}_{0}^{\cnums }$ denote the complex gauge transformations of $V$ that are the identity restricted to $\ell_{\infty } $. Then $\ms{k}{n}=\mathcal{A}^{1,1}_{0}/\mathcal{G}^{\cnums }_{0}$. The quotient $$(\mathcal{A}^{1,1}_{0}\times V )/\mathcal{G}^{\cnums }_{0}\to {\mathcal{M}} _{k}^{n}\times \til{\cnums \P}^{2}$$ will form a universal bundle if the moduli space is smooth and no ${\mathcal{E}} \in {\mathcal{M}} _{k}^{n}$ has non-trivial automorphisms ({\em c.f.} \cite{Fr-Mo} Chapt. IV): \begin{lem}\label{lem:E has no finite automorphisms, moduli space is smooth} ${\mathcal{M}} _{k}^{n}$ is smooth and any ${\mathcal{E}} \in {\mathcal{M}} _{k}^{n}$ has no non-trivial automorphisms preserving $\tau :{\mathcal{E}} |_{\ell_{\infty }} \to \cnums ^{n}\otimes \mathcal{O}_{\ell_{\infty } }$. \end{lem} By Serre duality $H^{2}({\mathcal{E}} \otimes {\mathcal{E}}^{*} )=H^{0}({\mathcal{E}} \otimes {\mathcal{E}} ^{*}\otimes K)^{*}$. Since ${\mathcal{E}} \otimes {\mathcal{E}} ^{*}$ is trivial on $\ell_{\infty } $, it is also trivial on nearby lines. Any section of ${\mathcal{E}} \otimes {\mathcal{E}}^{*} \otimes K$ restricts to a section of $\cnums ^{n^{2}}\otimes \mathcal{O}_{\ell_{\infty } }(-3)$ and so must vanish on $\ell_{\infty } $. Likewise, it must vanish on nearby lines and so it is $0$ on an open set and must be identically $0$. Thus $H^{2}({\mathcal{E}}\otimes {\mathcal{E}} ^{*})=0$ and smoothness follows once we show there are no automorphisms. Suppose that there exists an automorphism $\phi \in H^{0}({\mathcal{E}} \otimes {\mathcal{E}} ^{*})$ such that $\phi \neq {{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l $ and $\phi $ preserves $\tau $ so that $\phi |_{\ell_{\infty } }={{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l |_{\ell_{\infty } }$. Then $\phi -{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l $ is a non-zero section of ${\mathcal{E}} \otimes {\mathcal{E}}^{*} $ vanishing on $\ell_{\infty } $. We then get an injection $0\to \O(\ell_{\infty } )\to {\mathcal{E}} \otimes {\mathcal{E}} ^{*}$. Restricting this sequence to $\ell_{\infty } $ we get an injection $0\to \O_{\ell_{\infty } }(1)\to \O_{\ell_{\infty } }\otimes \cnums ^{n^{2}}$ which is a contradiction. Let $\pi :\ms{k}{n}\times \til{\cnums\P }^{2}\to \ms{k}{n}$. The higher direct image sheaves $R^{1}\pi _{*}(\Bbb{E}(-\ell_{\infty }))$ and $R^{1}\pi _{*}(\Bbb{E}(-\ell_{\infty }+E))$ are locally free and rank $k$. This follows from the index theorem and the vanishing of the $H^{0} $ and $H^{2}$ cohomology of ${\mathcal{E}} (-\ell_{\infty } )$ and ${\mathcal{E}} (-\ell_{\infty } +E)$. The $H^{0}$ vanishing follows by again considering the restriction of a section of the bundles to lines nearby to $\ell_{\infty } $. Using Serre duality and the same argument, one gets the vanishing for $H^{2}$. Thus the vector spaces $W_{0}$ and $W_{1}$ are the fibers of the vector bundles $R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }))$ and $R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }+E))$. The $G$-orbit of a configuration giving a bundle ${\mathcal{E}}$ can be identified with the group of isomorphisms $g_0:H^1({\mathcal{E}}(-\ell_{\infty }))\to \cnums^k$ and $g_1:H^{1}({\mathcal{E}}(-\ell_{\infty }+E))\to \cnums^k$. Thus $A_{k}^{n}$ is realized precisely as the total space of the principal $Gl(k)\times Gl(k)$ bundle associated to $R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }))\oplus R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }+E))$. \end{proof} Recall that the map $\ms{k}{n}\hookrightarrow \ms{k}{n+1}$ is defined by the direct sum with the trivial connection: $[A]\mapsto [A\oplus \theta]$. In terms of holomorphic bundles this is ${\mathcal{E}} \mapsto {\mathcal{E}} \oplus \mathcal{O}$. Tracing through the monad construction, it is easy to see that the inclusion induces the $G$-equivariant map $A_k^n\hookrightarrow A_k^{n+1}$ given by $(a_{1},a_{2},x,b,c)\mapsto (a_{1},a_{2},x,b',c')$ where $b^{\prime}$ is $b$ with an extra first column of zeroes and $c^{\prime}$ is $c$ with an extra first row of zeroes. Define $A_k^{\infty}$ to be the direct limit $\lim_{n\to\infty}A_k^n$ so that there is a homeomorphism between ${\mathcal{M}} _k^{\infty}$ and $A_k^{\infty}/G$ \begin{lem} $A_k^{\infty}$ is a contractible space. \end{lem} \begin{proof} Since the $A_k^{n}$'s are algebraic varieties and the maps $A_k^n\to A_k^{n+1}$ are algebraic, they admit triangulations compatible with the maps. Thus $A_k^\infty$ inherits the structure of a CW-complex and so it is sufficient to show that all of its homotopy groups are zero. To this end we prove that for any $k$ and $l$ there is an $r>l$ such that the natural inclusion from $A_k^n\hookrightarrow A_k^r$ is homotopically trivial. Consider the homotopy $H_t:A_k^n\to A_k^{2k+n}$ defined as follows: \[H_t((a_1,a_2,x,b,c))= ( (1-t)a_1,\; (1-t)a_2,\; (1-t)x,\; b_t,\; c_t)\] where \[c_t= \begin{pmatrix}tI_k\\ 0_{k,k}\\ (1-t)c \end{pmatrix} \;\;{\rm ,}\;\; b_t=(0_{k,k},\; tI_k,\; (1-t)^2b),\] $I_k$ is the $k\times k$ identity matrix and $0_{k,k}$ is the $k\times k$ zero matrix. To see that $H_t(v)\in A_k^{n+2k}$ for any $v\in A_k^n$, we check that the integrability and non-degeneracy conditions are satisfied for all $0\leq t\leq 1$. Integrability holds because $b_tc_t=(1-t)^3bc$. Non-degeneracy is satisfied for all $t\not=0$ because there is a full rank $k\times k$ block, $tI_k$, in both $c_t$ and $b_t$. Furthermore, $H_0$ is just the inclusion $A_k^n\hookrightarrow A_k^{n+2k}$, so non-degeneracy also holds when $t=0$. Finally, note that $H_1$ is a constant map. \end{proof} These lemmas show that $A_{k}^{\infty }$ is a contractible space acted on freely by $G=Gl(k)\times Gl(k)$ and $A_{k}^{\infty }/G=\ms{k}{\infty }$. Thus $\ms{k}{\infty }$ is homotopic to $BG$ which in turn has the homotopy type of $BU(k)\times BU(k)$. We end by remarking that the proof shows that the universal $U(k)\times U(k)$ bundle is the bundle that restricts to any of the finite ${\mathcal{M}}_k^n$'s as $R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }))\oplus R^{1}\pi_{*}(\Bbb{E}(-\ell_{\infty }+E))$.
1996-12-16T11:00:28
9610
alg-geom/9610019
en
https://arxiv.org/abs/alg-geom/9610019
[ "alg-geom", "math.AG" ]
alg-geom/9610019
Alexander G. Kuznetsov
Alexander Kuznetsov
The Laumon's resolution of Drinfeld's compactification is small
Correction of misprints, alternative definition of Drinfeld's compactification included AMSLaTeX v 1.1
null
null
null
null
Let $C$ be a smooth projective curve of genus 0. Let $\FF$ be the variety of complete flags in an $n$-dimensional vector space $V$. Given an $(n-1)$-tuple $\alpha$ of positive integers one can consider the space $\MM\alpha$ of algebraic maps of degree $\alpha$ from $C$ to $\FF$. This space has drawn much attention recently in connection with Quantum Cohomology. The space $\MM\alpha$ is smooth but not compact. The problem of compactification of $\MM\alpha$ proved very important. One compactification $\MML\alpha$ (the space of {\em quasiflags}), was constructed in \cite{L}. However, historically the first and most economical compactification $\MMD\alpha$ (the space of {\em quasimaps}) was constructed by Drinfeld (early 80-s, unpublished). The latter compactification is singular, while the former one is smooth. Drinfeld has conjectured that the natural map $\pi:\MML\alpha\to\MMD\alpha$ is a small resolution of singularities. In the present note we prove this conjecture. As a byproduct, we compute the stalks of $IC$ sheaf on $\MMD\alpha$ and, moreover, the Hodge structure in these stalks. Namely, the Hodge structure is a pure Tate one, and the generating function for the $IC$ stalks is just the Lusztig's $q$-analogue of Kostant's partition function (see \cite{Lu}).
[ { "version": "v1", "created": "Sun, 27 Oct 1996 10:37:49 GMT" }, { "version": "v2", "created": "Mon, 28 Oct 1996 05:33:31 GMT" }, { "version": "v3", "created": "Mon, 16 Dec 1996 09:46:05 GMT" } ]
2008-02-03T00:00:00
[ [ "Kuznetsov", "Alexander", "" ] ]
alg-geom
\section{The space of maps into flag variety} \subsection{Notations} Let $G$ be a complex semisimple simply-connected Lie group, $H\subset B$ its Cartan and Borel subgroups, $N$ the unipotent radical of $B$, $Y$ the lattice of coroots of $G$ (with respect to $H$), $l$ the rank of $Y$, $I=\{i_1,i_2,\dots i_l\}$ the set of simple coroots, $R^+$ the set of positive coroots, $X$ the lattice of weights, $X^+$ the cone of dominant weights, $\Omega=\{\omega_1,\omega_2\dots\omega_l\}$ the set of fundamental weights ($\langle\omega_k,i_l\rangle=\delta_{kl}$), ${\cal B}=G/B$ the flag variety and $C$ a smooth projective curve of genus ~$0$. Recall that there are canonical isomorphisms $$ \oper{H}\nolimits_2({\cal B},{\Bbb Z})\cong Y\qquad\oper{H}\nolimits^2({\cal B},{\Bbb Z})\cong X. $$ For $\lambda\in X$ let ${\mbox{\bf L}}_\lambda$ denote the corresponding $G$-equivariant line bundle on ${\cal B}$. The map $\varphi:C\to{\cal B}$ has degree $\alpha\in{\Bbb N}[I]\subset Y$ if the following equivalent conditions hold: \begin{enumerate} \item $\varphi_*([C])=\alpha$; \item for any $\lambda\in X\quad\text{we have}\quad \deg(\varphi^*{\mbox{\bf L}}_\lambda)=\langle\lambda,\alpha\rangle$. \end{enumerate} We denote by $\MM\alpha$ the space of algebraic maps from $C$ to ${\cal B}$ of degree $\alpha$. It is known that $\MM\alpha$ is smooth variety and $\dim\MM\alpha=2|\alpha|+\dim{\cal B}$. In this paper we compare two natural compactifications of the space $\MM\alpha$, which we presently describe. \subsection{Drinfeld's compactification} The Pl\"ucker embedding of the flag variety ${\cal B}$ gives rise to the following interpretation of $\MM\alpha$. For any irreducible representation $V_\lambda$ ($\lambda\in X^+$) of $G$ we consider the trivial vector bundle ${\cal V}_\lambda=V_\lambda\otimes{\cal O}_C$ over $C$. For any $G$-morphism $\psi:\ V_\lambda\otimes V_\mu\longrightarrow V_\nu$ we denote by the same letter the induced morphism $\psi:\ {\cal V}_\lambda\otimes {\cal V}_\mu \longrightarrow {\cal V}_\nu$. Then $\MM\alpha$ is the space of collections of line subbundles ${\cal L}_\lambda\subset{\cal V}_\lambda,\ \lambda\in X^+$ such that: a) $\deg{\cal L}_\lambda=-\langle\lambda,\alpha\rangle$; b) For any nonzero $G$-morphism $\psi:\ V_\lambda\otimes V_\mu\longrightarrow V_\nu$ such that $\nu=\lambda+\mu$ we have $\psi({\cal L}_\lambda\otimes{\cal L}_\mu)= {\cal L}_\nu$; c) For any $G$-morphism $\psi:\ V_\lambda\otimes V_\mu\longrightarrow V_\nu$ such that $\nu<\lambda+\mu$ we have $\psi({\cal L}_\lambda\otimes{\cal L}_\mu)=0$. \begin{remark}\label{rem1} Certainly, the property b) guarantees that in order to specify such a collection it suffices to give ${\cal L}_{\omega_k}$ for the set $\Omega$ of fundamental weights. \end{remark} If we replace the curve $C$ by a point, we get the Pl\"ucker description of the flag variety ${\cal B}$ as the space of collections of lines $L_\lambda\subset V_\lambda$ satisfying conditions of type (b) and (c) (thus ${\cal B}$ is embedded into $\prod\limits_{\lambda\in X^+}{\Bbb P}(V_\lambda)$). Here, a Borel subgroup ${\mbox{\bf B}}$ in ${\cal B}$ corresponds to a system of lines $(L_\lambda,\lambda\in X^+)$ if lines are the fixed points of the unipotent radical of ${\mbox{\bf B}}$, $L_\lambda=(V_\lambda)^{\mbox{\bf N}}$, or equivalently, if ${\mbox{\bf N}}$ is the common stabilizer for all lines ${\mbox{\bf N}}=\bigcap\limits_{\lambda\in X^+}G_{L_\lambda}$. The following definition in case $G=SL_2$ appeared in \cite{Drinfeld}. \begin{defn}[V.Drinfeld]\label{MMD} The space $\MMD\alpha$ of quasimaps of degree $\alpha$ from $C$ to ${\cal B}$ is the space of collections of invertible subsheaves ${\cal L}_\lambda\subset{\cal V}_\lambda,\ \lambda\in X^+$ such that: a) $\deg{\cal L}_\lambda=-\langle\lambda,\alpha\rangle$; b) For any nonzero $G$-morphism $\psi:\ V_\lambda\otimes V_\mu\longrightarrow V_\nu$ such that $\nu=\lambda+\mu$ we have $\psi({\cal L}_\lambda\otimes{\cal L}_\mu)= {\cal L}_\nu$; c) For any $G$-morphism $\psi:\ V_\lambda\otimes V_\mu\longrightarrow V_\nu$ such that $\nu<\lambda+\mu$ we have $\psi({\cal L}_\lambda\otimes{\cal L}_\mu)=0$. \end{defn} \begin{remark} Here is another version of the Definition, also due to V.Drinfeld. The principal affine space ${\cal A}=G/N$ is an $H$-torsor over ${\cal B}$. We consider its affine closure ${\frak A}$, that is, the spectrum of the ring of functions on ${\cal A}$. The action of $H$ extends to ${\frak A}$ but it is not free anymore. Consider the quotient stack $\tilde{\cal B}={\frak A}/H$. The flag variety ${\cal B}$ is an open substack in $\tilde{\cal B}$. A map $\tilde\phi:C\to\tilde{\cal B}$ is nothing else than an $H$-torsor $\Phi$ over $C$ along with an $H$-equivariant morphism $f:\Phi\to{\frak A}$. The degree of this map is defined as follows. Let $\chi_\lambda:H\to{\Bbb C}^*$ be the character of $H$ corresponding to a weight $\lambda\in X$. Let $H_\lambda\subset H$ be the kernel of the morphism $\chi_\lambda$. Consider the induced ${\Bbb C}^*$-torsor $\Phi_\lambda=\Phi/H_\lambda$ over $C$. The map $\tilde\phi$ has degree $\alpha\in{\Bbb N}[I]$ if $$ \text{for any }\lambda\in X\quad\text{we have}\quad \deg(\Phi_\lambda)=\langle\lambda,\alpha\rangle. $$ \end{remark} \begin{defn}\label{MMD1} The space $\MMD\alpha$ is the space of maps $\tilde\phi:C\to\tilde{\cal B}$ of degree $\alpha$ such that the generic point of $C$ maps into ${\cal B}\subset\tilde{\cal B}$. \end{defn} The equivalence of \ref{MMD} and \ref{MMD1} follows immediately from the Pl\"ucker embedding of ${\frak A}$ into $\prod\limits_{\lambda\in X^+}V_\lambda$. \begin{pro} $\MMD\alpha$ is a projective variety. \end{pro} \begin{proof} The space $\MMD\alpha$ is naturally embedded into the space $$ \prod_{k=1}^l {\Bbb P}(\oper{Hom}({\cal O}_C(-\langle\omega_k,\alpha\rangle),{\cal V}_{\omega_k})) $$ and is closed in it. \end{proof} \subsection{The stratification of the Drinfeld's compactification} In this subsection we will introduce the stratification of the space of quasimaps. {\bf Configurations of $I$-colored divisors.} Let us fix $\alpha\in{\Bbb N}[I]\subset Y,\ \alpha=\sum\limits_{k=1}^N a_ki_k$. Consider the configuration space $C^\alpha$ of colored effective divisors of multidegree $\alpha$ (the set of colors is $I$). The dimension of $C^\alpha$ is equal to the length $|\alpha|=\sum\limits_{k=1}^N a_k$. Multisubsets of a set $S$ are defined as elements of some symmetric power $S^{(m)}$ and we denote the image of $(s_1,\dots,s_m)\in S^m$ by $\{\{s_1,\dots,s_m\}\}$. We denote by $\Gamma(\alpha)$ the set of all partitions of $\alpha$, i.e.\ multisubsets $\Gamma=\{\{\gamma_1,\dots,\gamma_m\}\}$ of ${\Bbb N}[I]$ with $\sum\limits_{r=1}^m\gamma_r=\alpha$, $\gamma_r>0$. For $\Gamma\in\Gamma(\alpha)$ the corresponding stratum $C^\alpha_\Gamma$ is defined as follows. It is formed by configurations which can be subdivided into $m$ groups of points, the $r$-th group containing $\gamma_r$ points; all the points in one group equal to each other, the different groups being disjoint. For example, the main diagonal in $C^\alpha$ is the closed stratum given by partition $\alpha=\alpha$, while the complement to all diagonals in $C^\alpha$ is the open stratum given by partition $$ \alpha=\sum\limits_{k=1}^N(\underbrace{i_k+i_k+\ldots+i_k}_{a_k\operatorname{ times}}) $$ Evidently, $C^\alpha=\bigsqcup\limits_{\Gamma\in\Gamma(\alpha)}C^\alpha_\Gamma$. {\bf Normalization and defect of subsheaves.}\nopagebreak Let $F$ be a vector bundle on the curve $C$ and let $E$ be a subsheaf in $F$. Let $F/E=T(E)\oplus L$ be the decomposition of the quotient sheaf $F/E$ into the sum of its torsion subsheaf and a locally free sheaf, and let $\tilde E=\oper{Ker}(F\to L)$ be the kernel of the natural map $F\to L$. Then $\tilde E$ is a vector subbundle in $F$ which contains $E$ and has the following universal property: $$ \text{for any subbundle ${\cal E}'\subset F$ if ${\cal E}'$ contains $E$ then ${\cal E}'$ contains also $\tilde E$}. $$ Moreover, $\oper{rank}\tilde E=\oper{rank} E$, $\tilde E/E\cong T(E)$ and $c_1(\tilde E)=c_1(E)+\ell(T(E))$ (for any torsion sheaf on $C$ we denote by $\ell(T)$ its length). \begin{defn} We will call $\tilde E$ the {\em normalization} of $E$ in $F$ and $T(E)$ the {\em defect} of $E$. \end{defn} \begin{remark}\label{detnorm} If $\tilde E$ is the normalization of $E$ in $F$ then $\Lambda^k(\tilde E)$ is the normalization of $\Lambda^kE$ in $\Lambda^k F$. \end{remark} For any $x\in C$ and torsion sheaf $T$ on $C$ we will denote by $\ell_x(T)$ the length of the localization of $T$ in the point $x$. \begin{defn} For any quasimap $\varphi=({\cal L}_\lambda\subset{\cal V}_\lambda)_{\lambda\in X^+}\in\MMD\alpha$ we define the {\em normalization} of $\varphi$ as follows: $$ \tilde\varphi=(\tilde{\cal L}_\lambda\subset{\cal V}_\lambda), $$ and the {\em defect} of $\varphi$ as follows: $$ \oper{def}(\varphi)=(T({\cal L}_\lambda)) $$ (the defect of $\varphi$ is a collection of torsion sheaves). \end{defn} \begin{pro}\label{normalization} For any $\varphi\in\MMD\alpha$ there exists $\beta\le\alpha\in{\Bbb N}[I]$, partition $\Gamma=(\gamma_1,\dots,\gamma_m)\in\Gamma(\alpha-\beta)$ and a divisor $D=\sum\limits_{r=1}^m\gamma_rx_r\in C^{\alpha-\beta}_\Gamma$ such that $$ \tilde\varphi\in\MM\beta \qquad \ell_x(\oper{def}(\varphi)_\lambda)=\begin{cases} \displaystyle\langle\lambda,\gamma_r\rangle,&\text{if $x=x_r$}\\ \ds0,&\text{otherwise}\end{cases} $$ \end{pro} \begin{proof} Clear. \end{proof} \begin{defn} The pair $(\beta,\Gamma)$ will be called the {\em type of degeneration} of $\varphi$. We denote by $\DD\beta\Gamma$ the subspace of $\MMD\alpha$ consisting of all quasimaps $\varphi$ with the given type of degeneration. \end{defn} \begin{remark} Note that $\DD0\emptyset=\MM\alpha$. \end{remark} We have \begin{equation} \MMD\alpha=\bigsqcup\begin{Sb} \beta\le\alpha\\ \Gamma\in\Gamma(\alpha-\beta) \end{Sb}\DD\beta\Gamma \end{equation} The map $d_{\beta,\Gamma}: \DD\beta\Gamma\to\MM\beta\times C^{\beta-\alpha}_\Gamma$ which sends $\varphi$ to $(\tilde\varphi,D)$ (see \ref{normalization}) is an isomorphism. The inverse map $\sigma_{\beta,\Gamma}$ can be constructed as follows. Let $\varphi=({\cal L}_\lambda)\in\MM\beta$. Then $$ \sigma_{\beta,\Gamma}(\varphi,D)\stackrel{\text{\rm def}}{=}({\cal L}'_\lambda)\qquad {\cal L}'_\lambda\stackrel{\text{\rm def}}{=} \bigcap_{r=1}^m{\frak m}_{x_r}^{\langle\lambda,\gamma_r\rangle}\cdot{\cal L}_\lambda, $$ where ${\frak m}_x$ denotes the sheaf of ideals of the point $x\in C$. \subsection{Laumon's compactification} Let $V$ be an $n$-dimensional vector space.\linebreak >From now on we will assume that $G=SL(V)$ (in this case certainly $l=n-1$). In this case there is the Grassmann embedding of the flag variety, namely $$ {\cal B}=\{(U_1,U_2,\dots,U_{n-1})\in G_1(V)\times G_2(V)\times\dots\times G_{n-1}(V)\ |\ U_1\subset U_2\subset\dots\subset U_{n-1}\}, $$ where $G_k(V)$ is the Grassmann variety of $k$-dimensional subspaces in $V$. This embedding gives rise to another interpretation of $\MM\alpha$. We will denote by ${\cal V}$ the trivial vector bundle $V\otimes{\cal O}_C$ over $C$. Let $\alpha=\sum\limits_{k=1}^{n-1}a_ki_k$, where $i_k$ is the simple coroot dual to the highest weight $\omega_k$ of representation $G$ in $\Lambda^kV$. Then $\MM\alpha$ is the space of complete flags of vector subbundles $$ 0\subset{\cal E}_1\subset{\cal E}_2\subset\dots\subset{\cal E}_{n-1}\subset{\cal V} \quad\text{ such that }\quad c_1({\cal E}_k)=-\langle\omega_k,\alpha\rangle=-a_k. $$ \begin{defn}[{Laumon, \cite[4.2]{Laumon}}] The space $\MML\alpha$ of {\em quasiflags} of degree $\alpha$ is the space of complete flags of locally free subsheaves $$ 0\subset E_1\subset E_2\subset\dots\subset E_{n-1}\subset{\cal V} \quad\text{ such that }\quad c_1(E_k)=-\langle\omega_k,\alpha\rangle=-a_k. $$ \end{defn} It is known that $\MML\alpha$ is a smooth projective variety of dimension $2|\alpha|+\dim{\cal B}$ (see loc.\ cit., Lemma 4.2.3). \subsection{The stratification of the Laumon's compactification} There is a stratification of the space $\MML\alpha$ similar to the above stratification of $\MMD\alpha$. \begin{defn} For any quasiflag $E_\bullet=(E_1,\dots,E_{n-1})$ we define its {\em normalization} as $$ \tilde E_\bullet=(\tilde E_1,\dots,\tilde E_{n-1}),\text{ where }\tilde E_k\text{ is the normalization of $E_k$ in ${\cal V}$} $$ and {\em defect} $$ \oper{def}(E_\bullet)=(\tilde E_1/E_1,\dots,\tilde E_{n-1}/E_{n-1}) $$ Thus, the defect of $E_\bullet$ is a collection of torsion sheaves. \end{defn} \begin{pro}\label{Lnormalization} For any $E_\bullet\in\MML\alpha$ there exist $\beta\le\alpha\in{\Bbb N}[I]$, partition $\Gamma=(\gamma_1,\dots,\gamma_m)\in\Gamma(\alpha-\beta)$ and a divisor $D=\sum\limits_{r=1}^m\gamma_rx_r\in C^{\alpha-\beta}_\Gamma$ such that $$ \tilde E_\bullet\in\MM\beta, \qquad \ell_x(\oper{def}(E_k))=\begin{cases} \langle\omega_k,\gamma_r\rangle,&\text{if $x=x_r$}\\ 0,&\text{otherwise}\end{cases} $$ \end{pro} \begin{defn} The pair $(\beta,\Gamma)$ will be called the {\em type of degeneration} of $E_\bullet$. We denote by $\fL\beta\Gamma$ the subspace in $\MML\alpha$ consisting of all quasiflags $E_\bullet$ with the given type of degeneration. \end{defn} \begin{remark} Note that $\fL0\emptyset=\MM\alpha$. \end{remark} We have \begin{equation} \MML\alpha=\bigsqcup\begin{Sb} \beta\le\alpha\\ \Gamma\in\Gamma(\alpha-\beta) \end{Sb}\fL\beta\Gamma \end{equation} \subsection{The map from $\MML\alpha$ to $\MMD\alpha$} Consider the map $\pi:\MML\alpha\to\MMD\alpha$ which sends a quasiflag of degree $\alpha$ $E_\bullet\in\MML\alpha$ to a quasimap given by the collection $({\cal L}_{\omega_k})_{k=1}^{n-1}$ (see Remark \ref{rem1}) where ${\cal L}_{\omega_k}=\Lambda^k E_k\subset\Lambda^k{\cal V}={\cal V}_{\omega_k}$. \begin{pro} Let $E_\bullet$ be a quasiflag of degree $\alpha$ and let $(\beta,\Gamma)$ be its type of degeneration. Then $\pi(E_\bullet)$ is a quasimap of degree $\alpha$ and its type of degeneration is $(\beta,\Gamma)$. \end{pro} \begin{proof} Obviously we have $\deg{\cal L}_{\omega_k}=\deg\Lambda^k E_k=c_1(E_k)=-\langle\omega_k,\alpha\rangle$ which means that $\pi(E_\bullet)\in\MMD\alpha$. According to the Remark \ref{detnorm}, $\tilde{\cal L}_{\omega_k}=\Lambda^k\tilde E_k$ (i.e. $({\cal L}_{\omega_k})\in\MM\beta$), hence \begin{equation}\label{lx} \ell_x(\tilde{\cal L}_{\omega_k}/{\cal L}_{\omega_k})=\ell_x(\tilde E_k/E_k). \end{equation} This proves the Proposition. \end{proof} \begin{remark} Note that \refeq{lx} implies that $\pi$ preserves not only $\beta$ and $\Gamma$ but also $D$ (see \ref{normalization}, \ref{Lnormalization}). \end{remark} Recall that a proper birational map $f:{\cal X}\to{\cal Y}$ is called {\em small} if the following condition holds: let ${\cal Y}_m$ be the set of all points $y\in{\cal Y}$ such that $\dim f^{-1}(y)\ge m$. Then for $m>0$ we have \begin{equation}\label{small} \oper{codim}{\cal Y}_m>2m. \end{equation} {\bf Main Theorem. } The map $\pi$ is a small resolution of singularities. \section{The fibers of $\pi$} \subsection{} We fix ${\cal E}_\bullet\in\MM\beta$, a partition $\Gamma\in\Gamma(\alpha-\beta)$, and a divisor $D\in C_\Gamma^{\alpha-\beta}$. Then $({\cal E}_\bullet,D)\in\DD\beta\Gamma$. We define $F({\cal E}_\bullet,D)$ as $\pi^{-1}({\cal E}_\bullet,D)$. Let $D=\sum_{r=1}^m\gamma_rx_r$. We define the space ${\cal F}({\cal E}_\bullet,D)$ of commutative diagrams $$ \begin{CD} {\cal E}_1 @>>> {\cal E}_2 @>>> \dots @>>> {\cal E}_{n-1} \\ @V\varepsilon_1VV @V\varepsilon_2VV @. @V\varepsilon_{n-1}VV \\ T_1 @>\tau_1>> T_2 @>\tau_2>> \dots @>\tau_{n-2}>>T_{n-1} \end{CD} $$ such that a) $\varepsilon_k$ is surjective, b) $T_k$ is torsion, c) $\displaystyle \ell_x(T_k)=\begin{cases} \displaystyle\langle\omega_k,\gamma_r\rangle,&\text{if $\displaystyle x=x_r$}\\ \ds0,&\text{otherwise}\end{cases}$ \begin{lem}\label{torsion} We have an isomorphism $$ F({\cal E}_\bullet,D)\cong{\cal F}({\cal E}_\bullet,D). $$ \end{lem} \begin{proof} If $E_\bullet\in F({\cal E}_\bullet,D)$ then by the \ref{Lnormalization} the collection $(T_1,\dots,T_{n-1})=\oper{def}(E_\bullet)$ satisfies the above conditions. Vice versa, if the collection $(T_1,\dots,T_k)$ satisfies the above conditions, then consider $$ E_k=\oper{Ker}({\cal E}_k @>\varepsilon_k>> T_k). $$ Since the square $$ \begin{CD} {\cal E}_k @>\varepsilon_k>> T_k \\ @VVV @V\tau_kVV \\ {\cal E}_{k+1} @>\varepsilon_{k+1}>> T_{k+1} \end{CD} $$ commutes, we can extend it to the commutative diagram $$ \begin{CD} 0 @>>> E_k @>>> {\cal E}_k @>\varepsilon_k>> T_k @>>> 0 \\ @. @VVV @VVV @V\tau_kVV @.\\ 0 @>>> E_{k+1} @>>> {\cal E}_{k+1}@>\varepsilon_{k+1}>> T_{k+1} @>>> 0 \end{CD} $$ The induced morphism $E_k\to E_{k+1}$ is injective because ${\cal E}_k\to{\cal E}_{k+1}$ is, and \begin{multline*} \qquad c_1(E_k)=c_1({\cal E}_k)-\ell(T_k)= -\langle\omega_k,\beta\rangle-\sum_{x\in C}\ell_x(T_k)=\\= -\langle\omega_k,\beta\rangle-\sum_{r=1}^m\langle\omega_k,\gamma_r\rangle= -\langle\omega_k,\beta+(\alpha-\beta)\rangle=-\langle\omega_k,\alpha\rangle\qquad \end{multline*} This means that $E_\bullet\in F({\cal E}_\bullet,D)$. \end{proof} \begin{pro}\label{gfibre} If $D=\sum\limits_{r=1}^m\gamma_rx_r$ is a decomposition into disjoint divisors then \begin{equation} F({\cal E}_\bullet,D)\cong\prod_{r=1}^mF({\cal E}_\bullet,\gamma_rx_r). \end{equation} \end{pro} \begin{proof} Recall that if $T$ is a torsion sheaf on the curve $C$ then $$ T=\bigoplus_{x\in C}T_x, $$ where $T_x$ is the localization of $T$ in the point $x$. This remark together with Lemma \ref{torsion} proves the Proposition. \end{proof} The above Proposition implies, that in order to describe general fiber $F({\cal E}_\bullet,D)$ it is enough to have a description of the fibers $F({\cal E}_\bullet,\gamma x)$, which we will call {\em simple fibers}. \subsection{The stratification of a simple fiber} We will need the following obvious Lemma. \begin{lem}\label{coin} Let ${\cal E}$ be a vector bundle on $C$. Let ${\cal E}'\subset{\cal E}$ be a vector subbundle, and let $E\subset{\cal E}$ be a (necessarily locally free) subsheaf. Then $E'={\cal E}'\cap E$ is a vector subbundle in $E$. Moreover, the commutative square $$ \begin{CD} E' @>>> E \\ @VVV @VVV \\ {\cal E}' @>>> {\cal E} \end{CD} $$ can be extended to the commutative diagram $$ \begin{CD} E' @>>> E @>>> E/E' \\ @VVV @VVV @VVV \\ {\cal E}' @>>> {\cal E} @>>> {\cal E}/{\cal E}'\\ @VVV @VVV @VVV \\ {\cal E}'/E' @>>> {\cal E}/E @>>> \dfrac{{\cal E}/E}{{\cal E}'/E'}\cong\dfrac{{\cal E}/{\cal E}'}{E/E'} \end{CD} $$ in which both the rows and the columns form the short exact sequences. \end{lem} The sheaf in the lower-right corner of the diagram will be called {\em cointersection} of $E$ and ${\cal E}'$ inside ${\cal E}$ and denoted by $\coin{\cal E} E{{\cal E}'}$. Let $$ \gamma=\sum_{k=1}^{n-1}c_ki_k. $$ For every $E_\bullet\in F({\cal E}_\bullet,\gamma x)$ we define \begin{equation} \mu_{pq}(E_\bullet)\stackrel{\text{\rm def}}{=} l\left(\frac{{\cal E}_q}{E_p\cap{\cal E}_q}\right) \qquad(1\le q\le p\le n-1), \end{equation} \begin{equation}\label{nu} \nu_{pq}(E_\bullet)=\begin{cases} \mu_{pq}(E_\bullet)-\mu_{p+1,q}(E_\bullet),&\text{if }1\le q\le p<n-1\\ \mu_{pq}(E_\bullet),&\text{if }1\le q\le p=n-1 \end{cases} \end{equation} \begin{equation}\label{ka} \kappa_{pq}(E_\bullet)=\begin{cases} \nu_{pq}(E_\bullet)-\nu_{p,q-1}(E_\bullet),&\text{if }1<q\le p\le n-1\\ \nu_{pq}(E_\bullet),&\text{if }1=q\le p\le n-1 \end{cases} \end{equation} \begin{remark} The transformations \refeq{nu} and \refeq{ka} are invertible, so the numbers $\mu_{pq}$ can be uniquely reconstructed from $\nu_{pq}$ or $\kappa_{pq}$. Namely, \begin{equation}\label{inv} \displaystyle\nu_{pq}=\sum_{r=1}^q\kappa_{pr};\qquad \displaystyle\mu_{pq}=\sum_{s=p}^{n-1}\nu_{sq}=\sum_{r\le q\le p\le s}\kappa_{sr}. \end{equation} \end{remark} \begin{lem} We have \begin{equation}\label{pnu} \nu_{pq}(E_\bullet)=l\left(\dfrac{{\cal E}_q\cap E_{p+1}}{{\cal E}_q\cap E_p}\right). \end{equation} \begin{equation}\label{pka} \kappa_{pq}(E_\bullet)=l\left( \coin{{\cal E}_q\cap E_{p+1}}{{\cal E}_q\cap E_p}{{\cal E}_{q-1}\cap E_{p+1}}\right). \end{equation} \end{lem} \begin{proof} The commutative diagram with exact rows $$ \begin{CD} 0@>>> {\cal E}_q\cap E_p @>>> {\cal E}_q @>>> \dfrac{{\cal E}_q}{{\cal E}_q\cap E_p} @>>>0\\ @. @VVV @| @VVV\\ 0@>>> {\cal E}_q\cap E_{p+1} @>>> {\cal E}_q @>>> \dfrac{{\cal E}_q}{{\cal E}_q\cap E_{p+1}} @>>>0 \end{CD} $$ implies \refeq{pnu}. In order to prove \refeq{pka} note that $$ {\cal E}_{q-1}\cap E_p=({\cal E}_q\cap E_p)\cap({\cal E}_{q-1}\cap E_{p+1}) $$ and apply Lemma \ref{coin} and \refeq{pnu}. \end{proof} \begin{cor} Numbers $\mu_{pq}$, $\nu_{pq}$ and $\kappa_{pq}$ satisfy the following inequalities: \begin{eqnarray} 0\le\kappa_{pq}\\ 0\le\nu_{p1}\le\nu_{p2}\le\dots\le\nu_{pp}\\ 0\le\mu_{n-1,q}\le\mu_{n-2,q}\le\dots\le\mu_{qq}=c_q.\label{muin} \end{eqnarray} \end{cor} \begin{proof} See \refeq{pka},\refeq{ka},\refeq{nu} and compare the definition of $\mu_{qq}$ with \ref{torsion}. \end{proof} We will denote by $[p,q]$ the positive coroot \begin{equation}\label{pq} [p,q]\stackrel{\text{\rm def}}{=}\sum_{k=q}^pi_k\in R^+ \end{equation} \begin{lem}\label{part} For any $E_\bullet\in F({\cal E}_\bullet,\gamma x)$ we have $$ \sum_{1\le q\le p\le n-1}\kappa_{pq}(E_\bullet)[p,q]=\gamma. $$ \end{lem} \begin{proof} Applying \refeq{ka}, \refeq{pq} and \refeq{inv} we get \begin{multline*} \sum_{1\le q\le p\le n-1}\kappa_{pq}[p,q]= \sum_{1\le q\le p\le n-1}(\nu_{pq}-\nu_{p,q-1})[p,q]=\\=\!\!\! \sum_{1\le q\le p\le n-1}\!\!\!\nu_{pq}([p,q]-[p,q+1])=\!\!\! \sum_{1\le q\le p\le n-1}\!\!\!\nu_{pq}i_q= \sum_{q=1}^{n-1}\left(\sum_{p=q}^{n-1}\nu_{pq}\right)i_q= \sum_{q=1}^{n-1}\mu_{qq}i_q. \end{multline*} Now Lemma follows from \refeq{muin}. \end{proof} Let ${\frak K}(\gamma)$ be the set of all partitions of $\gamma\in{\Bbb N}[I]$ into the sum of positive coroots: $\gamma=\sum\limits_{s=1}^t\delta_s$, where $\delta_s\in R^+$ (note that ${\frak K}(\gamma)\ne\Gamma(\gamma)$). In other words, since every positive coroot for $G=SL(V)$ is equal to $[p,q]$ for some $p,q$, $$ {\frak K}(\gamma)=\{ (\kappa_{pq})_{1\le q\le p\le n-1}\ |\ \kappa_{pq}\ge0\quad\text{ and }\sum_{1\le q\le p\le n-1}\kappa_{pq}[p,q]=\gamma\}. $$ Let ${\frak M}(\gamma)$ denote the set of all collections $(\mu_{pq})$ which can be produced from some $(\kappa_{pq})\in{\frak K}(\gamma)$ as in \refeq{inv}. The Lemma \ref{part} implies that for any $E_\bullet\in F({\cal E}_\bullet,\gamma x)$ we have $(\mu_{pq}(E_\bullet))\in{\frak M}(\gamma)$. Define the stratum ${\frak S}((\mu_{pq})_\pq{n-1},({\cal E}_k)_{k=1}^{n-1})$ as follows: $$ {\frak S}((\mu_{pq})_\pq{n-1},({\cal E}_k)_{k=1}^{n-1})= \{ E_\bullet\in F({\cal E}_\bullet,\gamma x)\ |\ \mu_{pq}(E_\bullet)=\mu_{pq}\}. $$ To unburden the notations in the cases when it is clear which flag ${\cal E}_\bullet$ is used we will write just ${\frak S}_\mu$. We have obviously \begin{equation}\label{stratification} F({\cal E}_\bullet,\gamma x)=\bigsqcup_{\mu\in{\frak M}(\gamma)}{\frak S}_\mu. \end{equation} \begin{remark}\label{shortflag} We will also use the similar varieties ${\frak S}((\mu_{pq})_\pq{N},({\cal E}_k)_{k=1}^N)$ that can be defined in the same way for any short flag $({\cal E}_k)_{k=1}^N$ (that is the flag of subbundles ${\cal E}_1\subset\dots\subset{\cal E}_N\subset{\cal V}$ with $\oper{rank}{\cal E}_k=k$). \end{remark} \subsection{The strata ${\frak S}_\mu$} In order to study ${\frak S}_\mu$ we will introduce some more varieties. For every $1\le N\le n-1$, a short flag of subbundles $({\cal E}_k)_{k=1}^N$ (see Remark \ref{shortflag}) and a collection of numbers $(\nu_k)_{k=1}^N$ such that $0\le\nu_1\le\dots\le\nu_N$, we define the space ${\frak T}((\nu_k)_{k=1}^N,({\cal E}_k)_{k=1}^N)$ as follows: \begin{equation}\label{TT} {\frak T}((\nu_k)_{k=1}^N,({\cal E}_k)_{k=1}^N)=\{E\subset{\cal E}_N\ |\ \oper{rank}(E)=N\quad\text{ and }\quad l{\left(\frac{{\cal E}_k}{{\cal E}_k\cap E}\right)}=\nu_k\} \end{equation} We define {\em pseudoaffine} spaces by induction in dimension. First, the affine line ${\Bbb A}^1$ is a pseudoaffine space. Now a space $A$ is called pseudoaffine if it admits a fibration $A\to B$ with pseudoaffine fibers and pseudoaffine $B$. \begin{thm} The space ${\frak T}((\nu_k)_{k=1}^N,\!({\cal E}_k)_{k=1}^N)$ is pseudoaffine of dimension $\!\sum\limits_{k=1}^{N-1}\!\!\nu_k$. \end{thm} \begin{proof} We use induction in $N$. The case $N=1$ is trivial. There is only one subsheaf $E$ in the line bundle ${\cal E}_1$ with $\ell({\cal E}_1/E)=\nu_1$, namely $E={\frak m}_x^{\nu_1}\cdot{\cal E}_1$. This means that ${\frak T}(\nu_1,{\cal E}_1)$ is a point and the base of induction follows. If $N>1$ then consider the map $$ \tau:{\frak T}((\nu_k)_{k=1}^N,({\cal E}_k)_{k=1}^N)\to {\frak T}((\nu_k)_{k=1}^{N-1},({\cal E}_k)_{k=1}^{N-1}), $$ which sends $E\in{\frak T}((\nu_k)_{k=1}^N,({\cal E}_k)_{k=1}^N)$ to $E'=E\cap{\cal E}_{N-1}\in{\frak T}((\nu_k)_{k=1}^{N-1},({\cal E}_k)_{k=1}^{N-1})$. \begin{lem}\label{fibre} Let $L={\frak m}_x^{\nu_N-\nu_{N-1}}\cdot\left(\dfrac{{\cal E}_N}{{\cal E}_{N-1}}\right)$. For any $E'\in{\frak T}((\nu_k)_{k=1}^{N-1},({\cal E}_k)_{k=1}^{N-1})$ there is an isomorphism $$ \tau^{-1}(E')\cong\oper{Hom}(L,{\cal E}_{N-1}/E')\cong {\Bbb A}^{\ell({\cal E}_{N-1}/E')}={\Bbb A}^{\nu_{N-1}}. $$ \end{lem} Thus, the space ${\frak T}((\nu_k)_{k=1}^N,({\cal E}_k)_{k=1}^N)$ is affine fibration over a pseudoaffine space, hence it is pseudoaffine and its dimension is equal to $$ \dim\left({\frak T}((\nu_k)_{k=1}^{N-1},({\cal E}_k)_{k=1}^{N-1})\right)+\nu_{N-1}= \sum_{k=1}^{N-2}\nu_k+\nu_{N-1}=\sum_{k=1}^{N-1}\nu_k. $$ The Theorem is proved. \end{proof} \begin{proofof}{Lemma \ref{fibre}} Let $E\in\tau^{-1}(E')$. Since $E'=E\cap{\cal E}_{N-1}$ we can apply Lemma \ref{coin} which gives the following commutative diagram: $$ \begin{CD} E' @>>> E @>>> L \\ @VVV @VVV @VVV \\ {\cal E}_{N-1} @>j>> {\cal E}_N @>\psi>> {\cal E}_N/{\cal E}_{N-1} \\ @VVV @VVV @VVV \\ T_{N-1} @>>> T_N @>>> T_N/T_{N-1} \end{CD} $$ (Note that since ${\cal E}_N/{\cal E}_{N-1}$ is a line bundle and $\ell(T_N/T_{N-1})=\ell(T_N)-\ell(T_{N-1})=\nu_N-\nu_{N-1}$ the kernel of natural map ${\cal E}_N/{\cal E}_{N-1}\to T_N/T_{N-1}$ is isomorphic to $L$.) Let $\tilde{\cal E}_N=\psi^{-1}(L)$. Then $E$ is contained in $\tilde{\cal E}_N$ and we have the following commutative diagram: $$ \begin{CD} E' @>>> E @>>> L \\ @VVV @VVV @| \\ {\cal E}_{N-1} @>j>> \tilde{\cal E}_N @>\psi>> L \\ @VVV @V\varepsilon VV \\ T_{N-1} @= T_{N-1} \end{CD} $$ This means that the points of $\tau^{-1}(E')$ are in one-to-one correspondence with maps $\varepsilon:\tilde{\cal E}_N\to T_{N-1}$ such that $\varepsilon\cdot j$ is equal to the canonical projection from ${\cal E}_{N-1}$ to $T_{N-1}$. Applying the functor $\oper{Hom}(\bullet,T_{N-1})$ to the middle row of the above diagram we get an exact sequence: $$ 0\to\oper{Hom}(L,T_{N-1})\to\oper{Hom}(\tilde{\cal E}_N,T_{N-1})@>j^*>>\oper{Hom}({\cal E}_{N-1},T_{N-1})\to \oper{Ext}\nolimits^1(L,T_{N-1}). $$ The last term in this sequence is zero because $L$ is locally free and $T_{N-1}$ is torsion. This means that the space of maps $\varepsilon$ which we need to describe is a torsor over the group $\oper{Hom}(L,T_{N-1})$. Hence this space can be identified with the group. Thus, we have proved that $\tau^{-1}(E')\cong\oper{Hom}(L,T_{N-1})$ is an affine space. Now, $$ \dim(\tau^{-1}(E'))=\dim\oper{Hom}(L,T_{N-1})=\dim\oper{H}\nolimits^0(T_{N-1})=\ell(T_{N-1})= \nu_{N-1}. $$ The Lemma is proved. \end{proofof} \begin{thm}\label{saff} The space ${\frak S}((\mu_{pq})_\pq{N},({\cal E}_k)_{k=1}^N)$ is a pseudoaffine space of dimension $\mu_{21}+\mu_{32}+\dots+\mu_{N,N-1}$. \end{thm} \begin{proof} We use induction in $N$. If $N=1$ then ${\frak S}_\mu$ is a point and the base of induction follows. If $N>1$ consider the map $$ \sigma:{\frak S}((\mu_{pq})_\pq{N},({\cal E}_k)_{k=1}^N)\to {\frak T}((\mu_{N,k})_{k=1}^N,({\cal E}_k)_{k=1}^N), $$ which sends $(E_k)_{k=1}^N$ to $E_N\subset{\cal E}_N$. \begin{lem}\label{sfibre} Let $E\in{\frak T}((\nu_{N,k})_{k=1}^N,({\cal E}_k)_{k=1}^N)$. Consider $\tilde{\cal E}_k={\cal E}_k\cap E\quad(1\le k\le N-1)$ and set $\tilde\mu_{pq}=\mu_{pq}-\mu_{Nq}\quad(\pq{N-1})$. Then $(\tilde{\cal E}_k)_{k=1}^{N-1}$ is a short flag of subbundles and for any $E\in{\frak T}((\mu_{N,k})_{k=1}^N,({\cal E}_k)_{k=1}^N$ \begin{equation} \sigma^{-1}(E)\cong{\frak S}((\tilde\mu_{pq})_{\pq{N-1}}),(\tilde{\cal E}_k)_{k=1}^{N-1}). \end{equation} \end{lem} Thus ${\frak S}((\mu_{pq})_\pq{N},({\cal E}_k)_{k=1}^N)$ is a fiber space with pseudoaffine base and fiber, therefore it is pseudoaffine. Now, the calculation of the dimension \begin{multline*} \dim\left({\frak S}((\mu_{pq})_\pq{N},({\cal E}_k)_{k=1}^N)\right)= \sum_{k=1}^{N-1}\mu_{N,k}+\sum_{k=1}^{N-2}\tilde\mu_{k+1,k}=\\= \sum_{k=1}^{N-1}\mu_{N,k}+\sum_{k=1}^{N-2}(\mu_{k+1,k}-\mu_{N,k})= \mu_{N,N-1}+\sum_{k=1}^{N-2}\mu_{k+1,k}= \sum_{k=1}^{N-1}\mu_{k+1,k}, \end{multline*} finishes the proof of the Theorem. \end{proof} \begin{proofof}{Lemma \ref{sfibre}} Assume that $(E_k)_{k=1}^N\in{\frak S}_\mu$ and $E_N=E$. The commutative diagram $$ \begin{CD} 0 @>>> {\cal E}_q\cap E_p @>>> {\cal E}_q @>>> \dfrac{{\cal E}_q}{{\cal E}_q\cap E_p} @>>> 0\\ @. @VVV @| @VVV @.\\ 0 @>>> {\cal E}_q\cap E @>>> {\cal E}_q @>>> \dfrac{{\cal E}_q}{{\cal E}_q\cap E} @>>> 0 \end{CD} $$ implies that \begin{equation}\label{lengths} l\left(\dfrac{{\cal E}_q\cap E}{{\cal E}_q\cap E_p}\right)= l\left(\dfrac{{\cal E}_q}{{\cal E}_q\cap E_p}\right)- l\left(\dfrac{{\cal E}_q}{{\cal E}_q\cap E}\right)= \mu_{pq}-\mu_{Nq}=\tilde\mu_{pq}, \end{equation} hence $(E_k)_{k=1}^{N-1}\in{\frak S}_{\tilde\mu}$. Vice versa, if $(E_k)_{k=1}^{N-1}\in{\frak S}_{\tilde\mu}$ then the above commutative diagram along with \refeq{lengths} implies that $(E_k)_{k=1}^N\in{\frak S}_\mu$, where we have put $E_N=E$. \end{proofof} \subsection{The cohomology of the simple fiber} Now we will compute the dimension of the strata ${\frak S}_\mu$ in terms of the partition $\kappa$. \begin{defn} A space ${\cal X}$ is called {\em cellular} if it admits a stratification with pseudoaffine strata. \end{defn} Suppose ${\cal X}=\bigsqcup\limits_{\xi\in\Xi} S_\xi$ is a pseudoaffine stratification of a cellular space ${\cal X}$. For a positive integer $j$ we define $\chi(j)\stackrel{\text{\rm def}}{=}\{\xi\in\Xi\ |\ \dim S_\xi=j\}.$ \begin{lem}\label{hodge} The Hodge structure $\oper{H}\nolimits^\bullet({\cal X},{\Bbb Q})$ is a direct sum of Tate structures, and ${\Bbb Q}(j)$ appears with multiplicity $\chi(j)$. In other words, $$ \oper{H}\nolimits^\bullet({\cal X},{\Bbb Q})=\oplus_{j\in{\Bbb N}}{\Bbb Q}(j)^{\chi(j)}. $$ \end{lem} \begin{proof} Evident. \end{proof} Given a Tate structure ${\cal H}=\oplus_{j\in{\Bbb N}}{\Bbb Q}(j)^{\chi(j)}$ we consider a {\em generating function} $$ P({\cal H},t)=\sum_{j\in{\Bbb N}}\chi(j)t^j\in{\Bbb N}[t]. $$ For $\kappa\in{\frak K}(\gamma)$ we define $K(\kappa)\stackrel{\text{\rm def}}{=}\displaystyle\sum\limits_{\pq{n-1}}\kappa_{pq}$ as the number of summands in the partition $\kappa$. For $\gamma\in{\Bbb N}[I]$ the following $q$-analog of the Kostant's partition function was was introduced in \cite{Lusztig}: \begin{equation}\label{cke} {\cal K}_\gamma(t)=t^{|\gamma|}\sum_{\kappa\in{\frak K}(\gamma)}t^{-K(\kappa)}. \end{equation} \begin{lem}\label{dims} Let $\kappa\in{\frak K}(\gamma)$ and $\mu\in{\frak M}(\gamma)$ be defined as in \refeq{inv}. Then \begin{equation} \dim{\frak S}_\mu=\sum_{k=1}^{n-2}\mu_{k+1,k}=|\gamma|-K(\kappa), \end{equation} \end{lem} \begin{proof} Applying \refeq{inv} we get \begin{multline*} \sum_{k=1}^{n-2}\mu_{k+1,k}= \sum_{k=1}^{n-2} \left(\sum\begin{Sb}1\le q\le k\\k+1\le p\le n-1\end{Sb}\kappa_{pq}\right)= \sum_{\pq{n-1}}(p-q)\kappa_{pq}=\\= \sum_{\pq{n-1}}(|[p,q]|-1)\kappa_{pq}= |\gamma|-\sum_{\pq{n-1}}\kappa_{pq}=|\gamma|-K(\kappa). \end{multline*} {}\end{proof} \begin{cor}\label{ck} For any $\gamma\in{\Bbb N}[I]$, $x\in C$, the simple fiber $F({\cal E}_\bullet,\gamma x)$ is a cellular space, and the generating function of its cohomology is equal to the Lusztig--Kostant polynomial $$ P(\oper{H}\nolimits^\bullet(F({\cal E}_\bullet,\gamma x)),t)={\cal K}_\gamma(t). $$ \end{cor} \begin{proof} Apply \refeq{stratification}, \ref{saff}, \ref{hodge} and \ref{dims}. \end{proof} \begin{cor}\label{cck} Let $D=\sum\limits_{r=1}^m\gamma_rx_r\in C^{\alpha-\beta}_\Gamma$. The fiber $F({\cal E}_\bullet,D)$ is a cellular space and \begin{equation}\label{ccke} P(\oper{H}\nolimits^\bullet(F({\cal E}_\bullet,D)),t)={\cal K}_\Gamma \stackrel{\text{\rm def}}{=}\prod_{r=1}^m{\cal K}_{\gamma_r}(t). \end{equation} \end{cor} \begin{proof} Apply \ref{gfibre}, \ref{hodge} and \ref{ck}. \end{proof} \begin{lem}\label{est} Let $D=\sum\limits_{r=1}^m\gamma_r x_r$. We have $$ \dim F({\cal E}_\bullet,D)\le\left|\sum_{r=1}^m\gamma_r\right|-m. $$ \end{lem} \begin{proof} Note that for any $\kappa\in{\frak K}(\gamma_r)$ we have $K(\kappa)\ge1$, hence $\deg{\cal K}_{\gamma_r}\le|\gamma_r|-1$. Now, the Lemma follows from \ref{cck}. \end{proof} \begin{proofof}{Main Theorem} Consider the stratum $\DD\beta\Gamma$ of $\MMD\alpha$. Its dimension is $2|\beta|+\dim{\cal B}+m$ and codimension is $2|\alpha-\beta|-m$. The Lemma \ref{est} implies that the dimension of the fiber of $\pi$ over the stratum $\DD\beta\Gamma$ is less than or equal to $|\alpha-\beta|-m$, which is strictly less then the half codimension of the stratum. \end{proofof} \subsection{Applications} Let $\underline{\Bbb Q}$ denote the smooth constant Hodge irreducible module on $\MML\alpha$ (as a constructible complex it lives in cohomological degree $-2|\alpha|-\dim{\cal B}$). Let $IC$ denote the minimal extension of a smooth constant irreducible Hodge module from $\MM\alpha$ to $\MMD\alpha$. It is well known that the smallness of $\pi$ implies the following corollary. \begin{cor} $$ IC=\pi_*\underline{\Bbb Q}. $$ \end{cor} Now we can compute the stalks of $IC$ as cohomology of fibers of $\pi$: for $\varphi\in\MMD\alpha$ we have $$ IC_{(\varphi)}^\bullet=\oper{H}\nolimits^\bullet(\pi^{-1}(\varphi),\underline{\Bbb Q}) $$ as graded Hodge structures. \begin{cor}[Parity vanishing] $$ IC_{(\varphi)}^{j}=0\quad\text{if $j-\dim{\cal B}$ is odd.} $$ \end{cor} \begin{proof} Use \ref{cck}. \end{proof} \begin{cor} For $\varphi\in\DD\beta\Gamma$ we have $$ IC_{(\varphi)}^{-2|\alpha|-\dim{\cal B}+2j}={\Bbb Q}(j)^{{\frak k}_\Gamma(j)}, $$ where ${\frak k}_\Gamma(j)$ is the coefficient of $t^j$ in ${\frak K}_\Gamma(t)$. \end{cor}
1996-10-15T18:02:52
9610
alg-geom/9610017
en
https://arxiv.org/abs/alg-geom/9610017
[ "alg-geom", "math.AG", "math.QA", "q-alg" ]
alg-geom/9610017
Rodriguez Romo Suemi-FESC
V.K. Kharchenko, J. Keller, S. Rodriguez-Romo
Prime rings with PI rings of constants
20 pages, LaTex2e, to appear in Israel Journal of Mathematics, volume 96, part B, 1996 (357-377)
null
null
null
null
It is shown that if the ring of constants of a restricted differential Lie algebra with a quasi-Frobenius inner part satisfies a polynomial identity (PI) then the original prime ring has a generalized polynomial identitiy (GPI). If additionally the ring of constants is semiprime then the original ring is PI. The case of a non-quasi-Frobenius inner part is also considered.
[ { "version": "v1", "created": "Tue, 15 Oct 1996 16:59:22 GMT" } ]
2008-02-03T00:00:00
[ [ "Kharchenko", "V. K.", "" ], [ "Keller", "J.", "" ], [ "Rodriguez-Romo", "S.", "" ] ]
alg-geom
\section{Introduction} Rings of constants of restricted differential Lie algebras with an outer action on prime and semiprime rings were investigated in detail in papers [Kh82], [Po83], [Pi86] (see also [Kh91,Ch.4, Ch6(6.4)]). In the present paper we are going to consider actions with a nontrivial inner part. In the papers [Ko91] and [Kh81] it is shown that the minimal restriction required is that the inner part should be quasi-Frobenius (selfinjective). We are interested in the structure of a prime ring $R$ provided it is known that its ring of constants satisfies a polynomial identity. I.V.L'vov's example [Lv93] shows that in this case the ring $R$ does not need to be a PI-ring. We will show that in this case $R$ satisfies a generalized polynomial identity. The notion of a generalized polynomial identity was introduced by S.A.Amitsur in [Am65]. In his paper S.A.Amitsur proved a structure theorem for primitive rings with generalized polynomial identities. Later W.S.Martindale [Ma69] generalized this result to arbitrary prime rings. Using this theorem we will prove that if the ring of constants is a semiprime PI-ring and the inner part is quasi-Frobenius, then the ring $R$ is a PI-ring. \section{Preliminaries} Recall that a {\it derivation} of a ring $R$ is an additive mapping $d:R \rightarrow R$ satisfying the condition $(xy)^d=x^dy+xy^d.$ If $d_1, d_2$ are derivations then it is easy to see that the commutator $[d_1,d_2]=d_1d_2-d_2d_1$ is also a derivation. Therefore the set $Der R$ of all derivations of $R$ is a Lie subring in the ring of endomorphisms of the abelian group $(R,+)$. Moreover, if $z$ is a central element, then the composition of $d$ with the multiplication by $z$ is a derivation $$(xy)^{dz}=z(xy)^d=(zx^d)y+x(zy^d)$$ In this case the operators of multiplication may not commute with derivations: $x^{zd}\stackrel{\rm def}{=} (zx)^d=z^dx+zx^d$ or $$zd=dz+z^d. \eqno(1)$$ Thus the set $Der R$ is a right module over the center $Z.$ The module structure of $Der R$ is connected with the commutator operation by the formula $$[dz,d_1]=[d,d_1]z+dz^{d_1}. \eqno(2)$$ Note that $z^{d_1}$ is again a central element: $[z^{d_1},x]=[z^{d_1},x]+[z,x^{d_1}]=[z,x]^{d_1}$=0. Finally, if the characteristic $p$ of the ring $R$ is nonzero, $pR=0,$ then the $p$th power of any derivation will be a derivation by the Leibniz formula $$(xy)^{d^p}=\sum_{k=0}^{k=p} C_p^k x^{d^k}y^{d^{p-k}}=x^{d^p}y+xy^{d^p}.$$ Now it is natural to formulate the following definition. {\bf 2.1. Definition.} A set of derivations is called a {\it differential restricted Lie $Z$-algebra}, or shortly a {\it Lie $\partial $-algebra}, if it is a right $Z$-submodule of $DerR$ closed with respect to the operations $[d_1,d_2]=d_1d_2-d_2d_1$ and $d^{[p]}=d^p.$ Note that the notion of a Lie $\partial $-algebra can be formalized abstractly as a restricted Lie ring with a structure of right $Z$-module connected with the main operations by formula (2) and the following formula $$(dz)^{[p]}=d^{[p]}z^p+d\cdot (\ldots ((\overbrace{z^dz)^dz)^d\ldots )^d}^{p-1}z \eqno(3)$$ which follows from (1) (see details in [Kh91, pp. 6-11]; for a slightly more general approach see in [Pa87]). Now let $R$ be a prime ring. Denote by $R_{\cal F}$ its left Martindale ring of quotients (see, for example, [Kh91 pp.19-24]), by $Q$ the symmetric Martindale ring of quotients. Recall that the center $C$ of $R_{\cal F}$ is called the {\it extended} (or {\it generalized}) centroid of $R$ and it is a field (see [Ma69]). All derivations of $R$ can be uniquely extended to derivations of $Q$ and of $R_{\cal F}.$ The extended derivations are characterized in $Der Q$ by the property $R^d\subseteq R$ but the linear combinations over $C$ of extended derivations do not satisfy this property. Therefore we have to consider more general objects. {\bf 2.2. Definition.} A derivation $d$ of $Q$ is called $R$-{\it continuous} if there exists a nonzero two-sided ideal $I$ of $R$ such that $I^d\subseteq R.$ It is easy to see that the set ${\cal D}(R)$ of all $R$-continuous derivations is a differential restricted Lie $C$-subalgebra of $Der Q.$ In the present paper we consider Lie $\partial $-algebras of $R$-continuous derivations which are finite dimensional over $C.$ Let us fix the notations $R, C, Q, R_{\cal F}, {\cal D} (R)$ for a prime ring, its extended centroid, the symmetric Martindale ring of quotients, the left Martindale ring of quotients and the Lie $\partial $-algebra of $R$-continuous derivations, respectively. Throughout the paper $L$ denotes a restricted differential Lie $C$-algebra of $R$-continuous derivations, $L\subseteq {\cal D} (R),$ finite dimensional over $C,$ and $R^L=\{ r\in R:\forall \mu \in L \ \ \ r^{\mu }=0\} $ is its ring of constants. \section{The inner part of a Lie $\partial $-algebra} If $a$ is an element of $Q$ then the map $a^{-}:x\rightarrow xa-ax$ is an $R$-continuous derivation, i.e. $Q^{-}\subseteq {\cal D}(R).$ {\bf 3.1. Definition.} The space $K(L)$ generated over $C$ by all $q\in Q$ such that $q^{-}\in L$ is called the {\it inner linear part of $L.$} It is clear that $C^{-}=0,$ therefore $K(L)$ contains $C$ and in particular it contains the unit of $Q.$ {\bf 3.2. Lemma.} {\it The space $K(L)$ is a restricted Lie subalgebra of the adjoint restricted Lie algebra $Q^{(-)}.$} Recall that $Q^{(-)}$ is a restricted Lie algebra defined on the $C$-space $Q$ with the operations $[q_1,q_2]=q_1q_2-q_2q_1,\ \ q^{[p]}=q^p.$ For the proof of the lemma it is enough to show that $K(L)$ is closed with respect to these operations. This fact immediately follows from the formulae $$[a,b]^{-}=[a^{-},b^{-}] \eqno(4)$$ $$(a^p)^{-}=(a^{-})^{[p]}. \eqno(5)$$ {\bf 3.3. Lemma.} {\it $K(L)^{-}$ is equal to the subalgebra $L_{int}$ of all inner derivations of $L.$} The proof is evident. {\bf 3.4. Definition.} The associative subalgebra ${\cal B}(L)$ generated in $Q$ by $K(L)$ is called the {\it inner associative part of} $L.$ {\bf 3.5. Lemma.} {\it The algebra ${\cal B}(L)$ is of finite dimension over $C.$} {\bf Proof.} By the definition of operations in $K(L),$ the identity map $id$ is a homomorphism of restricted Lie algebras $id: K(L)\rightarrow {\cal B}(L)^{(-)}.$ Therefore ${\cal B}(L)$ as an associative envelope of ${\cal B}(L)^{(-)}$ is a homomorphic image of the universal restricted associative envelope $U_p(K(L)).$ The latter has dimension $(\dim K(L))^p.$ The lemma is proved. {\bf 3.6. Lemma.} {\it The algebra ${\cal B}(L)$ is stable under the action of $L,$ i.e. ${\cal B}(L)^{\mu } \subseteq {\cal B}(L)$ for all} $\mu \in L.$ The proof follows from the formula $$(q^{\mu })^{-}=[q^{-},\mu ]. \eqno(6)$$ \section{Differential operators} Denote by $\Phi (L)$ the associative subring generated in the endomorphism ring of the abelian group $(Q,+)$ by $L$ and by the operators of left and right multiplications by elements from ${\cal B}(L).$ By formula (1) the ring $\Phi (L)$ may not be an algebra over $C.$ Of course $\Phi (L)$ is an algebra over the subfield of central constants $$F=C^L\stackrel{\rm def}{=} \{ c\in C:\forall l\in L \ \ c^l=0 \} .$$ Nevertheless $\Phi (L)$ is a left and a right space over $C$ while the subring of left multiplications, ${\cal B}(L)^l,$ and that of right multiplications, ${\cal B}(L)^r,$ are algebras over $C.$ {\bf 4.1.} Let us fix derivations $\mu _1, \ldots ,\mu _m \in L$ such that $\mu _1+K(L)^{-}, \ldots ,\mu _m+K(L)^{-}$ form a basis for the right $C$-space $L/K(L)^{-}.$ An operator $\Delta $ is called {\it correct} if it is of the form: $$\Delta =\mu ^{s_1}_1\mu ^{s_2}_2 \ldots \mu ^{s_m}_m,$$ where $0\leq s_i<p$ and we suppose that $\mu ^0=1$ is the identity operator. Let $U$ be a right linear space generated by all correct operators. By formula (1) this set will be a left space over $C,$ also. {\bf 4.2. Proposition.} {\it The ring $\Phi (L)$ of differential operators is isomorphic as a left and a right space over $C$ to a tensor product over $C:$ $$\Phi (L)\simeq {\cal B}(L)^r\otimes {\cal B}(L)^l\otimes U\simeq U\otimes {\cal B}(L)^l\otimes {\cal B}(L)^r, \eqno(7)$$ where $U$ is the linear space generated by correct operators over $C$.} {\bf Proof.} It is enough to show that each differential operator $d\in \Phi (L)$ has a unique representation in the form $$d=\sum _{i,j,k} \alpha ^{(k)}_{ij}a^r_{ik}a^l_{jk}\Delta _k \eqno(8)$$ and a unique representation in the form $$d=\sum _{i,j,k} \Delta _ka^l_{ik}a^r_{jk}\alpha ^{(k)}_{ij}, \eqno(9)$$ where $a_{ik}, a_{jk} \in A$ and $A$ is some fixed basis of ${\cal B}(L)$ over $C$ (recall that by associativity, $a^r_{ik}a^l_{jk}=a^l_{jk}a^r_{ik}$) and the $\Delta _k$'s are correct words in $\{ \mu _1,\ldots ,\mu _m \} .$ The existence of this presentation follows from the relations $$\mu a^r=a^r\mu -(a^{\mu })^r \eqno(10)$$ $$\mu a^l=a^l\mu -(a^{\mu })^l \eqno(11)$$ $$\mu ^p=\mu _1c_1+ \ldots +\mu _mc_m+b^r-b^l \eqno(12)$$ $$\mu _i\mu _j=\mu _j\mu _i+\mu _1c_1+ \ldots +\mu_mc_m +b^r-b^l, \eqno(13)$$ where in formula (12) \ $\mu _1c_1+ \ldots +\mu _mc_m+b^{-}$ \ is a representation of $\mu ^p\in L$ as a linear combination of $\mu _i$'s modulo $K(L)^{-}$ and in (13) \ $\mu _1c_1+ \ldots +\mu _mc_m+b^{-}$ \ is the corresponding representation of $[\mu _i,\mu _j]\in L.$ The transformations of the left hand sides to the right hand sides (in the last formula only if $i>j$) allow us to reduce the operator to the form (8). If we write formulae (10), (11) in the form $$a^r\mu =\mu a^r+(a^{\mu })^r \eqno(14)$$ $$a^l\mu = \mu a^l+(a^{\mu })^l \eqno(15)$$ then in the same way the operator is reduced to the form (9). For the proof of the uniqueness it is possible to use the following results on differential identities (see [Kh91, theorem 2.2.2, corollary 2.5.8] or [Kh78]). {\bf 4.3. Proposition.} {\it If the derivations $\mu _1, \ldots ,\mu _m \in {\cal D}(R)$ are linearly independent modulo $Q^{-},$ and if the ring $R$ satisfies an identity of the type $$\sum ^{p^n}_{k=1} \sum _i a_{ki} x^{\Delta _k}b_{ki}=0,$$ where $\Delta _1, \ldots , \Delta _{p^n}$ --- are all correct operators and the coefficients $a_{ki},b_{ki}$ belong to $R_{\cal F},$ then $\sum _ia_{ki} \otimes b_{ki}=0$ in $R_{\cal F} \otimes _C R_{\cal F}$ for all $k, 1\leq k \leq p^n.$ In the same way if the identity $$\sum ^{p^n}_{k=1} (\sum _i a_{ki}xb_{ki})^{\Delta _k}=0$$ is valid then $\sum _i a_{ki}\otimes b_{ki}=0, 1\leq k\leq p^n.$} Since ${\cal D}(I)={\cal D}(R)$ and $Q(I)=Q(R)$ for each nonzero ideal $I$ of $R$ (see [Kh91, Lemma 1.8.4]), then proposition 4.3 shows that the restriction of a nonzero differential operator $d\in \Phi (L)$ to $I$ is nonzero. This note is important due to the following lemma: {\bf 4.4. Lemma.} {\it For each differential operator $d\in \Phi (L)$ there exists a nonzero two sided ideal $I$ of $R$ such that $I^d\subseteq R.$} The proof is easily obtained by induction from the formula $(I^2)^{\mu }=I^{\mu }I+II^{\mu }\subseteq I$ which is valid for the ideal $I$ such that $I^{\mu }\subseteq R.$ \section{Quasi-Frobenius algebras} Recall that a finite dimensional algebra $B$ over a field $C$ is called quasi-Frobenius if one of the following equivalent conditions is valid (see [CR62]). (Q1) {\it For each left ideal $\lambda $ and right ideal $\rho $ of $B$ the following equalities hold: $$l(r(\lambda ))=\lambda ,\ \ \ r(l(\rho ))=\rho ,$$ where $l(A)=\{ b\in B: bA=0\} $ is the left annihilator, $r(A)=\{ b\in B: Ab=0\} $ is the right annihilator.} (Q2){\it The left regular module $_BB$ is injective.} (Q3) {\it Modules $_BB$ and $(B_B)^*=Hom (B,C)$ have the same indecomposable components.} Recall that for any left (right) module $M$ the set of all linear functionals $M^*$ has a structure of right (left) module defined by the formula $(m^*b)(m)=m^*(bm)$ (respectively $m(bm^*)=(mb)m^*$). The modules $M$ and $N$ for $N\simeq M^*$ are called {\it conjugated modules}. If the module $M$ is of finite dimension then $(M^*)^*\simeq M$ and the conjugacy of modules (left and right), $M$ and $N,$ can be characterized by the existence of a nondegenerate associative bilinear form $(\ ,\ ):N\times M\rightarrow C.$ In this case for every basis $a_1,\ldots ,a_n$ of $M$ there exists a {\it dual} basis $a_1^*,\ldots ,a_n^*$ of $N$ which is characterized by the following properties $(a_i^*,a_i)=1, (a_i^*,a_j)=0, i\neq j.$ Condition (Q3) implies the following condition which is important for us: (Q4) {\it The sum of all right ideals $\rho $ of $B$ conjugated to left ideals of $B$ is equal to $B.$} It can be proved that this condition is also equivalent to $B$ being quasi-Frobenius. Moreover, as (Q1) is left-right symmetric then the left analog of (Q4) is also valid. (Q5) {\it The sum of all left ideals ${\lambda }$ of $B$ conjugated to right ideals of $B$ is equal to $B.$} The most important subclass of the class of quasi-Frobenius algebras is the class of Frobenius algebras. These algebras are defined by one of the following equivalent conditions ([CR62]). (F1) {\it For each left ideal $\lambda $ and right ideal $\rho $ of $B$ the following equalities hold: $$l(r(\lambda ))=\lambda ,\ \ \dim r(\lambda )+\dim \lambda =\dim B$$ \ $$ r(l(\rho ))=\rho , \ \ \dim l(\rho )+\dim \rho=\dim B.$$ } (F2) {\it There exists an element $\varepsilon \in B^*$ whose kernel contains no nonzero onesided ideals of $B.$} (F3) {\it There exists a nondegenerate associative bilinear form $B\times B\rightarrow C.$ } (F4) {\it The modules $_BB$ and $(B_B)^*$ are isomorphic.} Classical examples of Frobenius algebras are: group algebras of finite groups over a field of arbitrary characteristic, universal restricted enveloping algebras of finite dimensional Lie $p$-algebras, finite dimensional Hopf algebras, Clifford algebras. Finite dimensional semisimple algebras evidently satisfy (F1) therefore they are Frobenius. \section{Universal constants} Let $\lambda $ and $\rho $ be left and right conjugated ideals of ${\cal B}(L).$ Let us choose a basis $a_1, \ldots ,a_n$ of $\lambda $ and let $a_1^*, \ldots ,a_n^*$ be the dual basis of $\rho .$ It is well-known that the element $c=\sum a_i\otimes a_i^*$ of the tensor product ${\cal B}\otimes _C{\cal B}$ commutes with the elements of ${\cal B}, \ \ bc=cb$ for all $b\in {\cal B}.$ This implies that the set of values of the operator $c_{\lambda ,\rho }=\sum a_i^l(a_i^*)^r$ is contained in the centralizer of ${\cal B}.$ In particular for any $\mu \in K(L)^-$ we have $$c_{\lambda ,\rho }(x)^{\mu }=0. \eqno(16) $$ Let $U(L)$ be the associative subring of $\Phi (L)$ generated by $L$ and by the operators of multiplication by central elements. It is clear that $U(L)$ is both a left and a right space over $C$ and an algebra over the field of central constants $F=C^L.$ Consider the right ideal $I=K(L)^-\cdot U(L)$ of $U(L).$ First of all the formula $\mu a^-=a^-\mu -(a^{\mu })^-$ shows that $I$ is a two sided ideal of $U(L).$ The same formula and formulae (12), (13) show that the identity operator and operators of the form $a_1^-a_2^-\ldots a_s^-\Delta ,$ where $\Delta $ is a correct operator, $a_i\in K(L), \ s\geq 0,$ generate $U(L)$ as a left space over $C.$ {\bf 6.1. Proposition.} {\it The factor-algebra $U(L)/I=\overline{U}$ is Frobenius as an algebra over $F=C^L.$} {\bf Proof.} By the well-known R.Baer theorem [Ba27] the dimension of $C$ over $F$ is finite and therefore $\overline{U} =U(L)/I$ has a finite dimension over $F.$ Since $K(L)^-\subseteq I,$ the elements $\bar{\mu }_1=\mu _1+I, \ldots , \bar{\mu }_m=\mu _m+I$ generate $\overline{U} $ as a ring over $C.$ Moreover the relations $\bar{\mu } _i\bar{\mu } _j=\overline{[\mu _i,\mu _j]}+\bar{\mu } _j\bar{\mu } _i$ show that the images of correct words $\bar{\Delta } _k$ generate $\overline{U}$ as a left vector space over $C.$ The main note is that the elements $\bar{\Delta } _k$ are linearly independent over $C.$ If $$\sum _kc_k\Delta _k=\sum _k d_k\Delta _k\in I,$$ where $d_k$ are linear combinations of products of the type $a_1^-\cdots a_s^-,$ then taking into account that $a^-=a^r-a^l$ and using Proposition 4.3, we have $c^r_k=d_k$ for all $k,$ which is impossible since $c_k^r(1)=c_k,\ d_k(1)=0.$ Thus $\bar{\Delta }_k$ are linearly independent. Now let us define Berkson's linear map (see [Be64]) $\varphi :\overline{U} \rightarrow C$ which corresponds to the element $\sum c_k\bar{\Delta }_k$ the coefficient of $\bar{\Delta }_{p^m}=\bar{\mu }^{p-1}_1\ldots \bar{\mu }^{p-1}_m.$ The kernel of this linear map contains neither left nor right nonzero ideals, since the product $$(\bar{\mu }_1^{s_1}\ldots \bar{\mu }_m^{s_m})(\bar{\mu }^{p-s_1-1}_1\ldots \bar{\mu }^{p-s_m-1}_m)$$ written as a linear combination of correct words contains a unique member $\bar{\Delta }_{p^m}$ with a coefficient equal to 1. If $\psi :C\rightarrow F$ is any projection, then the linear functional $\varepsilon :d\mapsto \psi (\varphi (d))$ satisfies (F2) and therefore $\overline{U}$ is a Frobenius algebra. The proposition is proved. Let us consider the right subspace $\hat{U}$ of $\overline{U}$ over $C$ generated by all nonempty words $\bar{\Delta }_k.$ This space does not contain the unit (the identity operator) and it is a right (but, possibly, not a left) ideal because by formula (14) one has $$\bar{\Delta }c\bar{\mu }=\bar{\Delta }\bar{\mu }c+\bar{\Delta }c^{\mu }.$$ By formula (13), the product $\bar{\Delta }\bar{\mu }$ can be written as a linear combination $\sum \bar{\Delta }_kc_k,$ where $\bar{\Delta }_k$ are nonidentity correct operators. Thus, the left annihilator $A=l(\hat{U})$ in the algebra $\overline{U}$ is not equal to zero. Moreover, by (F1) its dimension over $F$ is connected with the dimension of $\hat{U}$ by the formula $ \dim _F\overline{U}=\dim _F\hat{U}+\dim _FA.$ On the other hand $\dim _F\overline{U}=\dim _F\hat{U}+\dim _FC$ i.e. the dimensions of $A$ and $C$ over $F$ coincide. It means that $A$ is one dimensional over $C$ i.e. $A=C\bar{f}$ (but possibly $A\neq \bar{f}C$ as $A$ may not be a right $C$-space), where $\bar{f}=\sum \bar{\Delta }_kc_k=\sum c_k^{\prime }\bar{\Delta }_k$ is a nonzero element of $\overline{U}.$ Thus, we have obtained that $\bar{f}\bar{\mu }_i=\bar{0}$ in $\overline{U}.$ In the ring of differential operators this means that $f\mu _i\in K(L)^{-}\cdot U(L).$ We have also that $fK(L)^{-}\subseteq K(L)^{-}\cdot U(L)$ as $I=K(L)^{-}\cdot U(L)$ is a two sided ideal. Thus $$fL\subseteq f(\sum (\mu _iC+K(L)^{-}))\subseteq K(L)^{-}\cdot U(L)$$ which, using formula (16), implies $$((c_{\lambda ,\rho } (x))^f)^{\mu }=0 \eqno(17)$$ for all ${\mu \in L}.$ Let us formulate the obtained result as a lemma (see also Lemma 4.6, [Kh95]). {\bf 6.2. Lemma.} {\it There exists a differential operator $f$ of the type $\sum \Delta _kc_k=\sum c^{\prime }_k\Delta _k,$ such that for each conjugated left ideal $\lambda $ and right ideal $\rho $ of ${\cal B}$ with dual bases $a_1, \ldots , a_n$ and $a_1^*,\ldots ,a_n^*,$ the operator $$u_{\lambda , \rho }=\sum_i a_i^l(a_i^*)^rf \eqno(18)$$ has values only in the ring of constants $Q^L.$ There exists a nonzero ideal $I$ of $R$ such that $$0\neq u_{\lambda ,\rho }(I)\subseteq R^L \eqno(19)$$} {\bf Proof.} The representation of $f$ in the form $\sum c^{\prime }_k\Delta _k$ follows from (10). Formula (19) follows from formula (17), proposition 4.3 and lemma 4.4. \section{PI rings of constants} In this secton we will prove the theorem about a generalized polynomial identity and discuss its generalization to the case when the inner part is not quasi-Frobenius. {\bf 7.1. Theorem.} {\it Let $L$ be a finite dimensional restricted differential Lie $C$-algebra of $R$-continuous derivations of a prime ring $R$ of positive characteristic $p>0.$ Suppose that the inner associative part ${\cal B}(L)$ of $L$ is quasi-Frobenius. If the ring of constants $R^L$ is PI then $R$ is GPI.} {\bf Proof.} Let $f(x_1,\ldots ,x_n)=0$ be a multilinear identity of $R^L.$ Let us choose arbitrary left ideals $\lambda _1,\ldots ,\lambda _n$ of ${\cal B}(L)$ having conjugated right ones $\rho _1, \ldots ,\rho _n.$ By Lemma 6.2 for every $j, 1\leq j\leq n$ there exists an operator $$u_j=u_{\lambda _j,\rho _j}=\sum_i a^l_{ij}(a_{ij}^*)^rf_j=\sum_{i,k} a_{ij}^l(a_{ij}^*)^rc_k^{\prime }\Delta _k$$ and a nonzero ideal $I_j$ of $R,$ such that $0\neq u_j(I_j)\subseteq R^L.$ If $I=\cap I_j$ then $u_j(x)\in R^L$ for all $x\in I$ and therefore the following differential identity holds in $I$ \ $$f(u_1(x_1),\ u_2(x_2),\ldots ,\ u_n(x_n))=0.$$ Let us fix some values of $x_2=b_2, \ldots ,x_n=b_n$ in $I.$ We have $$f(\sum_{i,k} (c_k^{\prime }a_{i1}x_1a_{i1}^*)^{\Delta _k},\ u_2(b_2),\ldots ,\ u_n(b_n))=0. \eqno(20)$$ By Leibnitz formula any expression of the type $(axb)^{\Delta }$ can be written in the form $$(axb)^{\Delta }=ax^{\Delta }b+\sum_s a_sx^{\Delta _s}b_s,$$ where $\Delta _s$ are subwords of $\Delta .$ In particular $$(c_k^{\prime }a_{i1}x_1a_{i1}^*)^{\Delta _k}= c_k^{\prime }a_{i1}x_1^{\Delta _k}a_{i1}^*+\sum_s a_sx_1^{\Delta _s}b_s. \eqno(21)$$ If $\Delta _{k_0}$ is the greatest operator such that $c^{\prime }_{k_0}$ is not zero, then this formula allows us to represent (20) in the form $$\sum^{k_0}_{k=1} \sum_i v_{ki}x_1^{\Delta _k}w_{ki}=0,$$ here we suppose that $\Delta _1<\Delta _2<\ldots <\Delta _{p^m}$ is the lexicographic ordering of all correct operators. By Proposition 4.3 applied to the prime ring $I$ we have $$\sum_i v_{k_0i}\otimes w_{k_0i}=0$$ in the tensor product $I_{\cal F}\otimes _{C(I)}I_{\cal F},$ where $C(I)$ is the generalized centroid of $I$ and $I_{\cal F}$ is the left Martindale ring of quotients of $I.$ It is well-known and it is easy to see that $I_{\cal F}=R_{\cal F}$ and $C(I)=C(R).$ Therefore for any $x_1\in R_{\cal F}$ we have the identity $$\sum_i v_{k_0i}x_1w_{k_0i}=0.$$ This identity with (21) and (20) implies that the identity $$c_{k_0}^{\prime }f(\sum_i a_{i1}x_1a_{i1}^*, u_2(b_2),\ldots ,u_n(b_n))=0 \eqno(22)$$ is valid for each $x_1\in R_{\cal F}.$ Since the values $b_2,\ldots ,b_n$ are arbitrary from $I$, we have an identity of the form $$f(\sum_i a_{i1}x_1a_{i1}^*,u_2(x_2),\ldots ,u_n(x_n))=0, \eqno(23)$$ where $x_1\in R_{\cal F}, x_2\in I, \ldots ,x_n\in I.$ Now let us fix values $x_1\in R_{\cal F}, x_3=b_3\in I, \ldots ,x_n=b_n \in I.$ Then in the same way we obtain $$f(\sum_i a_{i1}x_1a_{i1}^*,\sum_i a_{i2}x_2a_{i2}^*,\ldots ,u_n(x_n))=0,$$ where $x_1,x_2\in R_{\cal F}, x_3,\ldots x_n \in I.$ Continuing this process we will obtain the following identity on $R_{\cal F}:$ $$f(\sum_i a_{i1}x_1a_{i1}^*,\sum_i a_{i2}x_2a_{i2}^*,\ldots ,\sum_i a_{in}x_na_{in}^*)=0, \eqno(24)$$ This is a generalized identity valid in $R_{\cal F}\supseteq R.$ All we need is to prove that for some $\lambda _1,\ldots, \lambda_n;\rho _1, \ldots ,\rho _n$ this is not a trivial identity. It means that the left hand side of (24) is not zero in the free product $R_{\cal F} *_CC\langle x_1,\ldots ,x_n\rangle $ or, in other words, this identity does not follow from the trivial generalized identities $xc=cx,$ where $c\in C.$ Otherwise assume all these identities are trivial. Any application of a trivial identity does not change the order of the indeterminates, therefore all the generalized monomials (i.e. sums of all monomials with fixed order of sequence of the indeterminates) in the identities (24) should be (trivial) identities. These generalized monomials have the form $$\alpha _{\pi }(\sum_i a_{i\pi (1)}x_{\pi (1)}a^*_{i\pi (1)}) (\sum_i a_{i\pi (2)}x_{\pi (2)}a^*_{i\pi (2)})\cdots (\sum_i a_{i\pi (n)}x_{\pi (n)}a^*_{i\pi (n)}),$$ where $\pi $ is a permutation and $$f(x_1,\ldots x_n)=\sum_{\pi }\alpha _{\pi}x_{\pi (1)}\cdots x_{\pi (n)}.$$ Since one of the coefficients $\alpha _{\pi }$ is equal to one (let $\alpha _1=1$), $$(\sum_i a_{i1}x_1a^*_{i1})(\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0 \eqno(25)$$ Let us fix some values of $x_2,\ldots ,x_n$ in $R$ and apply Proposition 4.3 to (25), where we suppose $x=x_1,$ and all coefficients $a_{ki}, \ \ k=2,3,\ldots p^m$ are zero. We have $$(\sum_ia_{i1}\otimes a_{i1}^*)(\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0.$$ The set $\{ a_{i1}\} $ is a basis of the ideal $\lambda _1,$ i.e. this is a linearly independent set, therefore $$a_{i1}^*(\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0$$ for all $a_{i1}^*$ from the dual basis $\{ a_{i1}^*\} $ of the conjugated ideal $\rho _1.$ This implies that $$\rho _1(\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0.$$ Since the pair $(\lambda _1,\rho _1)$ was chosen in an arbitrary way, $$(\sum_{\rho ^*\simeq\ a\ left\ ideal\ of\ {\cal B}}\rho ) (\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0. \eqno(26)$$ By Property (Q5) of quasi-Frobenius algebras $$1\in {\cal B}= (\sum_{\rho ^*\simeq\ a\ left\ ideal\ of\ {\cal B}}\rho ) $$ and therefore $$(\sum_i a_{i2}x_2a^*_{i2})\cdots (\sum_i a_{in}x_na^*_{in})=0.$$ Now the evident induction works. The theorem is proved.\ The same proof can be applied also for some cases when the inner part is not quasi-Frobenius but has enough pairs of conjugated one-sided ideals. Indeed, let us denote by ${\cal B}_r$ the sum of all right ideals of a finite dimensional algebra ${\cal B}$ conjugated to left ones. {\bf 7.2. Lemma.} ${\cal B}_r$ {\it is a two-sided ideal of} ${\cal B}.$ {\bf Proof.} Let $\rho $ be a right ideal such that the dual left module $\rho ^*={\rm Hom}(\rho ,C)$ is isomorphic to a left ideal $\lambda .$ If $b\in {\cal B}$ then we have an exact sequence of homomorphisms of right ideals $\rho \rightarrow b\rho \rightarrow 0.$ The conjugated sequence has the form $\rho ^*\leftarrow (b\rho )^*\leftarrow 0,$ therefore the right ideal $b\rho $ has a conjugated module $(b\rho )^*$ which is isomorphic to a left subideal of $\lambda \simeq \rho ^*.$ Thus $b\rho \subseteq {\cal B}_r$ and ${\cal B}_r$ is a two-sided ideal. The lemma is proved. In the same way one can define an ideal ${\cal B}_l$ --- the sum of all left ideals conjugated to right ones. {\bf 7.3. Theorem.} {\it Let $L$ be a finite dimensional restricted differential Lie $C$-algebra of $R$-continuous derivations of a prime ring $R$ of positive characteristic $p>0.$ If the algebra of constants $R^L$ satisfies a multilinear polynomial identity of degree $n$ and ${\cal B}(L)^n_r\neq 0,$ then $R$ is a GPI-ring.} {\bf Proof.} In the same way as in the proof of Theorem 7.1 we have identities (24). If all of these identities are trivial then we also have the identities (26) which can be written in the form $${\cal B}(L)_r(\sum a_{i2}x_2a_{i2}^*)\cdots (\sum a_{in}x_na_{in}^*)=0. \eqno(27)$$ If $b$ is an arbitrary element of ${\cal B}(L),$ then $b(\sum_i a_{ik}x_ka_{ik}^*)=(\sum _i a_{ik}x_ka_{ik}^*)b.$ Therefore for $b\in {\cal B}(L)_r,$ identity (27) implies $$(\sum a_{i2}x_2a_{i2}^*)\cdots (\sum a_{in}x_na_{in}^*)b=0. \eqno(28)$$ By Proposition 4.3 we have $$(\sum a_{i2}\otimes a_{i2}^*)\cdots (\sum a_{in}x_na_{in}^*)b=0,$$ as in the proof of Theorem 7.1 we have $${\cal B}(L)_r(\sum a_{i3}x_3 a_{i3}^*)\cdots (\sum a_{in}x_na_{in}^*)b=0,$$ thus $$(\sum a_{i3}x_3 a_{i3}^*)\cdots (\sum a_{in}x_na_{in}^*){\cal B}(L)_r^2=0.$$ Now the evident induction implies ${\cal B}(L)_r^n=0.$ Hence one of the GPI's (24) is not trivial. The theorem is proved. In a symmetrical way one can prove that the condition ${\cal B}(L)^n_l\neq 0$ also implies that one of the identities (24) is not trivial. It can be proved that ${\cal B}^n_r=0$ iff ${\cal B}^n_l=0:$ {\bf 7.4. Proposition.} {\it Let ${\cal B}$ be a finite dimensional algebra. Then all $n+1$ conditions ${\cal B}_r^k{\cal B}_l^{n-k}=0, \ \ k=0,\ldots ,n$ are equivalent to each other.} {\bf Proof.} It is enough to show that the conditions for $k$ and $k+1$ are equivalent. The condition ${\cal B}_r^k{\cal B}_l^{n-k}=0$ is equivalent to ${\cal B}_r^k{\cal B}_l^{n-k-1}\lambda =0$ for all pairs of conjugated ideals $\rho , \lambda .$ Since the form $(\ ,\ ):\rho \times \lambda \rightarrow C$ is nondegenerate the last condition for given $\lambda , \rho $ is equivalent to $(\rho ,{\cal B}_r^k{\cal B}_l^{n-k-1}\lambda )=0.$ By associativity of the form this is equivalent to $(\rho {\cal B}_r^k{\cal B}_l^{n-k-1},\lambda )=0$ and since the form is nondegenerate this is equivalent to $\rho {\cal B}_r^k{\cal B}_l^{n-k-1}=0.$ The last conditions for all pairs of conjugated ideals $\lambda , \rho $ are equivalent to ${\cal B}_r^{k+1}{\cal B}_l^{n-k-1}=0.$ The proposition is proved. Now it is a question of interest whether the condition ${\cal B}(L)^n_r=0$ implies that all identities (24) are trivial generalized polynomial identities. The answer is yes: {\bf 7.5. Proposition.} {\it If under the conditions of theorem {\rm 7.3} \ ${\cal B}(L)_r^n=0,$ then all identities {\rm (24)} are trivial.} {\bf Proof.} It is enough to show that all the generalized monomials (25) are trivial identities. We will prove by inverse induction on $k$ that for arbitrary $b_1, \ldots ,b_k\in {\cal B}(L)_r$ the generalized polynomial $$(\sum _ia_{i\ k+1}x_{k+1}a_{i\ k+1}^*)\cdots (\sum _ia_{in}x_na_{in}^*)b_kb_{k-1}\cdots b_1=0 \eqno(29)$$ is a trivial generalized identity. If $k=n$ then (29) has the form $b_nb_{n-1}\cdots b_1=0$ that is a trivial identity as ${\cal B}(L)^n_r=0.$ Assume that (29) is a trivial identity. The identities $$b(\sum _i a_{is}xa_{is}^*)=(\sum _i a_{is}xa_{is}^*)b, \ \ b\in {\cal B}(L) \eqno(30)$$ are trivial generalized polynomial identities (as well as any linear generalized identity). Let $b_k=a_{ik}^*,$ then from (29) and (30) we have the following trivial identity $$a_{ik}^*(\sum _ia_{i\ k+1}x_{k+1}a_{i\ k+1}^*)\cdots (\sum _ia_{in}x_na_{in}^*)b_{k-1}\cdots b_1=0.$$ Multiplication of this equality on the left by $a_{ik}x_k$ and summation over $i$ gives the equality (29) with a smaller $k.$ The proposition is proved.\ \section{Semiprime PI-rings of constants} In this secton we will prove under the conditions of Theorem 7.1, that if the ring of constants $R^L$ is a semiprime PI-ring, then $R$ is also PI. {\bf 8.1. Theorem.} {\it Let $L$ be a finite dimensional restricted differential Lie $C$-algebra of $R$-continuous derivations of a prime ring $R$ of positive characteristic $p>0.$ Suppose that the inner associative part ${\cal B}(L)$ of $L$ is quasi-Frobenius. If the ring of constants $R^L$ is a semiprime PI-ring, then $R$ is PI.} {\bf Proof.} By Theorem 7.1 the ring $R$ satisfies a generalized polynomial identity. Moreover all generalized polynomial identities (24) hold in its left Martindale ring of quotients $R_{\cal F}.$ In particular they hold in the central closure $RC\subseteq R_{\cal F}$ of the ring $R.$ By the Martindale structure theorem [Ma69] this central closure has an idempotent $e,$ such that $D=eRCe$ is a skew field of finite dimension over $C.$ (Note that formally Martindale theorem can be applied only if the coefficients of the identity belong to $R.$ In our case they belong to $R_{\cal F}$ but may not belong to $R.$ Nevetheless Martindale's original proof is correct for our case too; see, for instance, [Kh91, Theorem 1.13.4] or the special investigation in [La86].) Thus, by the Martindale theorem, $RC$ is a primitive ring with a nonzero socle. The N. Jacobson structure theorem [Ja64] shows that $RC$ is a dense subring in the finite topology in the complete ring ${\cal E}$ of linear transformations of the left space $V=eRCe$ over the skew field $D.$ Moreover, the left Martindale quotient ring $(RC)_{\cal F}$ is equal to ${\cal E}$ (see, [Ha82, Lemma 1.1] and [Ha87, Remark 4.9] or [Kh91, Theorem 1.15.1]). It is easy to see that $R_{\cal F} \subseteq (RC)_{\cal F}={\cal E}.$ (Indeed, if $q\in R_{\cal F}$ and $Iq\subseteq R$ for a nonzero ideal $I$ of $R,$ then we can extend $q$ to the ideal $IC$ of $RC$ by the obvious formula $(\sum i_{\alpha }c_{\alpha })q=\sum (i_{\alpha }q)c_{\alpha }.$ This is well-defined. Indeed, if $\sum i_{\alpha }c_{\alpha }=0$ and $J$ is a nonzero ideal of $R$ such that $Jc_{\alpha }\subseteq R$ then $\sum (jc_{\alpha })i_{\alpha }=0$ for all $j\in J.$ Therefore $\sum (jc_{\alpha })(i_{\alpha }q)=0;$ i.e. $J(\sum c_{\alpha }(i_{\alpha }q))=0$ and $\sum (i_{\alpha }q)c_{\alpha }=0.$) Now all the coefficients of (24) belong to ${\cal E}$ and since addition and multiplication are continuous in the finite topology, the identities (24) hold in ${\cal E}.$ (Here one can use also Corollary 2.3.2 from [Kh91] which allows us to extend identities from $RC$ to $(RC)_{\cal F}.$) Now we are going to prove that the space $V$ is finite dimensional over $D.$ In that case the dimension of ${\cal E}$ over $C$ will also be finite: $d=\dim _C{\cal E}= (\dim _DV)^2 \cdot \dim _CD$ and ${\cal E}$ (and therefore $R$), like any $d$-dimensional algebra, will satisfy the standard polynomial identity: $$S_d(x_1, \ldots , x_{d+1})\equiv \sum (-1)^{\pi }x_{\pi (1)}\cdots x_{\pi (d+1)}=0.$$ On the contrary, suppose that $V$ has infinite dimension $\dim V=\beta .$ Let $M$ be the set of all linear transformations whose rank is less then $\beta .$ (Recall that the {\it rank } of a transformation $l$ is the dimension over $D$ of its image.) It is well-known that $M$ is a maximal ideal of ${\cal E}.$ So the factor ring $\bar{\cal E}={\cal E}/M$ is a simple ring with a unit. {\bf 8.2. Lemma.} { \it The ring $\bar{\cal E}$ is not Artinian.} {\bf Proof.} Let $\{ e_i, i\in I\} $ be a basis of $V$ over $D.$ and $$I_1\supset I_2\supset \ldots \supset I_n\supset \ldots $$ be a chain of subsets such that $|I_k \setminus I_{k+1}|=\beta ,$ and let $$A_n=\{ l\in {\cal E}:e_il=0 \ \ \forall i \in I\setminus I_n\} .$$ Then $$ (A_1+M)/M\supset A_2+M/M\supset \ldots \supset A_n+M/M\supset \ldots $$ is an infinite descending chain of right ideals of $\bar{\cal E}.$ Indeed, if $A_n+M=A_{n+1}+M,$ then for the transformation $w,$ defined by \[ e_iw=\left\{ \begin{array}{ll} e_i & \mbox{if $i\in I_n\setminus I_{n+1}$} \\ 0 & \mbox{otherwise} \end{array} \right. , \] we should get a presentation $w=a+m,$ where $a\in A_{n+1}, \ m\in M.$ Let $V_1$ be a subspace generated by $\{e_i:i\in I_n\setminus I_{n+1} \} .$ Then $V_1=V_1w\subseteq V_1a+V_1m=V_1m.$ However, $\dim _DV_1=\beta ,$ while $\dim _DV_1m\leq \dim Vm<\beta ,$ which is a contradiction. The lemma is proved.\ {\bf 8.3. Lemma.} {\it The ring $\bar{\cal E}$ does not satisfy a non trivial generalized polynomial identity.} {\bf Proof.} Like any simple ring with a unit, the ring $\bar{\cal E}$ is primitive. If it satisfies a GPI, then by the S.A. Amitsur structure theorem [Am65] it has a nonzero socle $S,$ which is a two-sided ideal and therefore $S=\bar{\cal E}.$ In N. Jacobson presentation of $\bar{\cal E}$ as a dense ring of linear transformations, the socle consists of all transformations of finite rank. This means that the unit has finite rank and therefore the space has finite dimension. Thus $\bar{\cal E}$ is the ring of all linear transformations of a finite dimensional space over a skew field. In particular $\bar{\cal E}$ is Artinian; this is a contradiction to Lemma 8.2. The lemma is proved. Let us consider now identities (24). We have seen that all these identities hold in ${\cal E}.$ If we apply the natural homomorphism $\varphi :{\cal E}\rightarrow \bar{\cal E}={\cal E}/M$ we obtain the following identities of the ring $\bar{\cal E}$ $$f(\sum _i\bar{a}_{i1}x_1\bar{a}^*_{i1},\ldots ,\sum _i\bar{a}_{in}x_n\bar{a}^*_{in})=0, \eqno(31)$$ where $\bar{a}=\varphi (a)=a+M.$ By Lemma 8.3 all we need is to prove that one of the identities (31) is a nontrivial GPI of $\bar{\cal E}.$ First of all we have to calculate the generalized centroid of $\bar{\cal E}.$ As $\bar{\cal E}$ is a simple ring with a unit, it equals its left Martindale quotient ring and therefore the generalized centroid is equal to the center. {\bf 8.4. Lemma.} {\it The center of $\bar{\cal E}$ is canonically isomorphic to $C, \ \ C(\bar{\cal E})=\varphi (C).$} {\bf Proof.} See [Ro58, Corollary 3.3]. We will need the following result which gives a criterium for determining when the ring of constants is semiprime (see Theorem 5.1 [Kh95]). {\bf 8.5. Theorem.} {\it Under the conditions of theorem 8.1, the ring of constants is semiprime if and only if ${\cal B}(L)$ is differentially semisimple, i.e. it has no nonzero differential (with respect to action of $L$) ideals with zero multiplication or, equivalently, it is a sum of a finite number of differentially simple algebras.} By this theorem we have that in our situation the algebra ${\cal B}(L)$ is differentially semisimple. {\bf 8.6. Lemma.} {\it The ideal $M$ is a differential ideal with respect to $L,$ i.e. $M^{\mu }\subseteq M$ for each $\mu \in L.$} {\bf Proof.} Note that $M$ is a differential ideal with respect to each derivarion of ${\cal E}.$ Indeed, if $l\in M$ than $l$ is a transformation of rank less then $\beta $ and the projection $e:V\rightarrow {\rm im} l$ also has rank less than $\beta ,$ in which case $l=le.$ We have $l^{\mu }=l^{\mu }e+le^{\mu }\in M$ for each derivation $\mu \in {\rm Der}({\cal E}).$ By proposition 1.8.1 [Kh91] any $R$-continuous derivation has a unique extension to $R_{\cal F}.$ In particular each derivation from $L$ is defined on $RC.$ Again by the same proposition we have that the elements of $L$ have a unique extensions to $(RC)_{\cal F}={\cal E}.$ Thus we have obtained that the ideal $M$ is differential with respect to $L.$ The lemma is proved. As a consequence we have that the intersection $M_0=M\cap {\cal B}(L)$ is a differential ideal of ${\cal B}(L),$ which is not equal to ${\cal B}(L)$ (it does not contain 1). The left annihilator $l(M_0)$ of $M_0$ in ${\cal B}(L)$ is also a differential ideal, therefore $l(M_0)\cap M_0 $ is a differential ideal with zero mulitiplication. By theorem 8.5, $l(M_0)\cap M_0=0.$ In the same way the left annihilator of the sum $l(M_0)+M_0$ is zero (it is contained in $l(M_0)$ and, therefore, has a zero multiplication). Now property (Q1) of quasi-Frobenius algebras implies that $l(M_0)+M_0=r(l(l(M_0)+M_0))=r(0)={\cal B}(L)$ and, finally $${\cal B}(L)=l(M_0)\oplus M_0=e{\cal B}(L) \oplus (1-e){\cal B}(L), \eqno(37)$$ where $e$ is a central idempotent defined by the corresponding decomposition of the unit $1=e\oplus (1-e).$ Let us return to identities (31). Suppose that in these identities $\{ a_{ij}\} $ and $\{ a_{ij}^* \} $ are bases of conjugated ideals $\lambda _j, \rho _j$ contained in $l(M_0).$ In that case the sets $A_j=\{ \bar{a}_{ij}, i=1, \ldots m \} $ are linearly independent over the center of $\bar{\cal E}$ (see lemma 8.4). Moreover, the $C$-space generated by all possible $a_{ij}^*$'s contains the unit $e$ of $l(M_0)$ because for each conjugated pair of ideals $\lambda ,\rho $ the one-sided ideals $e\lambda ,e\rho $ are also conjugated with respect to the same form (note that $e$ is a central idempotent of ${\cal B}(L)$). This implies that the linear space over the center of $\bar{\cal E}$ generated by all $\bar{a}_{ij}^*$'s contains the unit $\bar{e}$ of $\bar{\cal E}.$ This fact allows us to prove that one of the identities (31) is nontrivial in the same manner as it was done in the end of the proof of Theorem 7.1. By Lemma 8.3, Theorem 8.1 is proved. In this proof we used the fact that the inner part ${\cal B}(L)$ is differentially semisimple and that it has enough pairs of conjugated ideals. Therefore in the way analogous to Theorem 7.3 we can formulate a slightly more general result. {\bf 8.7. Theorem.} {\it Let $L$ be a finite dimensional restricted differential Lie $C$-algebra of $R$-continuous derivations of a prime ring $R$ of positive characteristic $p>0.$ Suppose that the inner part ${\cal B}(L)$ is a direct sum of differentially simple ideals $${\cal B}(L)=B_1\oplus B_2\oplus \ldots \oplus B_m.$$ If the algebra of constants $R^L$ satisfies a multilinear polynomial identity of degree $n$ and $(B_i)^n_r\neq 0, i=1,\ldots ,m,$ then $R$ is a PI-ring.} The only place where we have used that ${\cal B}(L)$ is quasi-Frobenius is decomposition (37). Therefore it is enough to show that each differential ideal of the direct sum of differentially simple algebras with units is a direct summand. If $B=B_1\oplus B_2\oplus \ldots \oplus B_m$ is a direct sum of differentially simple algebras then for any differential ideal $A$ we have that $A B_i$ is a differential ideal of $B_i.$ This implies that either $B_i\subseteq A$ or $A B_i=0.$ In the same way either $B_i\subseteq A$ or $B_iA=0.$ Let $l(A)$ be the left annihilator of $A,$ then $l(A)\cap A$ is a differential ideal with zero multiplication, so its product with each $B_i$ is zero. This is possible only if the intersection is zero. In the same way the left annihilator of the sum $l(A)+A$ has a zero multiplication and therefore it is equal to zero. It means that $l(A)+A$ contains all the components $B_i$ and $l(A)\oplus A=B.$ \ ACKNOWLEDGMENT The authors are grateful to Professor Dalit Baum for her help. \ {\bf REFERENCES} \ [Am65]. S.A.Amitsur, {\it Generalized polynomial identities and pivotal monomials}, Trans. Amer. Math. Soc., v.114(1965), 210--216. [Ba27]. R.Baer, {\it Algebraiche theorie der differentierbaren funktionen koper}, I. -- Sitzungsberichte. Heidelberger Academia, 1927, 15--32. [Be64]. A.J.Berkson, {\it The u-algebra of a restricted Lie algebra is Frobenius}, PAMS, v.15(1964), 14--15. [CR62]. C.W.Curtis and I.Reiner, {\it Representation Theory of Finite Groups and Associative Algebras}, Interscience Publishers, New York -- London, 1962. [Ha82]. M.Hacque, {\it Anneaux fidelement repr\'esent\'es sur leur socle droit}, Communications in Algebra, v.10, no.10(1982), 1027--1072. [Ha87]. M.Hacque, {\it Th\'eorie de Galois des anneaux presque-simples}, Journal of Algebra, v.108, no.2(1987), 534--577. [Ja64]. N.Jacobson, {\it Structure of Rings}, Amer. Math. Soc. Colloquium. Publ., Providence, 1964. [Kh78]. V.K.Kharchenko, {\it Differential identities of prime rings}, Algebra i logika, v.17, no.2(1978), 220--238. [Kh81]. V.K.Kharchenko, {\it On centralizers of finite dimensional algebras}, Algebra i logika, v.20, no.2(1981), 231--247. [Kh82]. V.K.Kharchenko, {\it Constants of derivations of prime rings}, Math. USSR Izvestija, v.18, no.2(1982), 381--401. [Kh91]. V.K.Kharchenko, {\it Automorphisms and Derivations of Associative Rings}, Kluwer Academic Publishers, v.69(1991). [Kh95]. V.K.Kharchenko, {\it On derivations of prime rings of positive characteristic}, Algebra i logika, 1995, to appear. [Ko91]. A.N.Korjukin, { \it To a question of bicentralizers in prime rings}, Sib. Mat. Journal, v.32, no.6(1991), 81--86. [La86]. C.Lanski, {\it A note on GPIs and their coefficients}, Proc. Amer. Math. Soc. v.98(1986), 17--19. [Lv93]. I.V.Lvov, {\it On centralizers of finite dimensional subulgebras in the algebra of linear transformations}, Third International Algebraic Conference, Krasnojarsk, 1993, 213--214. [Ma69]. W.S.Martindale, {\it Prime rings satisfying a generalized polynomial identity}, Journal of algebra, v.12, no.4(1969), 576--584. [Pa87]. D.S.Passman, {\it Prime ideals in enveloping rings}, Trans. Amer. Math. Soc., v.302(1987), 535--560. [Pi86]. Piers Dos Santos, {\it Derivationes des anneaux semi-premiers I}, Comm. in algebra, v.14, no.8(1986), 1523--1559. [Po83]. A.Z.Popov, {\it On derivations of prime rings}, Algebra i Logika, v.22, no.1(1983), 79--92. [Ro58]. A.Rosenberg, {\it The structure of the infinite general linear group}, Ann. Math. ser.2, 68(1958), 278--294 \end{document}
1996-10-31T13:12:06
9610
alg-geom/9610021
en
https://arxiv.org/abs/alg-geom/9610021
[ "alg-geom", "hep-th", "math.AG", "math.QA", "nlin.SI", "q-alg", "solv-int" ]
alg-geom/9610021
Nakajima Hiraku
Hiraku Nakajima
Jack polynomials and Hilbert schemes of points on surfaces
AMSLaTeXv1.2 + epic.sty + eepic.sty + youngtab.sty, 20pages
null
null
null
null
The Jack symmetric polynomials $P_\lambda^{(\alpha)}$ form a class of symmetric polynomials which are indexed by a partition $\lambda$ and depend rationally on a parameter $\alpha$. They reduced to the Schur polynomials when $\alpha=1$, and to other classical families of symmetric polynomials for several specific parameters. It is well-known that Schur polynomials can be realized as certain elements of homology groups of Grassmann manifolds. The purpose of this paper is to give a similar geometric realization for Jack polynomials. However, spaces which we use are totally different. Our spaces are Hilbert schemes of points on a surface X which is the total space of a line bundle L over the projective line. The parameter $\alpha$ in Jack polynomials relates to our surface X by $\alpha = -<C,C>$, where C is the zero section, and <C,C> is the self-intersection number of C.
[ { "version": "v1", "created": "Thu, 31 Oct 1996 11:25:57 GMT" } ]
2008-02-03T00:00:00
[ [ "Nakajima", "Hiraku", "" ] ]
alg-geom
\section{Introduction} The Jack (symmetric) polynomials $P_\lambda^{(\alpha)}(x)$ form a class of symmetric polynomials which are indexed by a partition $\lambda$ and depend rationally on a parameter $\alpha$. They reduced to the Schur polynomials when $\alpha = 1$, and to other classical families of symmetric polynomials for several specific parameters. Recently they attracts attention from various points of view, for example the integrable systems and combinatorics. They are simultaneous eigenfunctions of certain commuting families of differential operators, appearing in the integrable system called the Calogero-Sutherland system (see e.g., \cite{AMOS} and the reference therein). On the other hand, Macdonald studied their combinatorial properties. In fact, he introduced an even more general class of symmetric functions (a two parameter family), which have many common combinatorial features as Jack polynomials (see \cite[Chapter~6]{Symmetric}). It is well-known that Schur polynomials can be realized as certain elements of homology groups of Grassmann manifolds (see e.g., \cite[Chapter~14]{Fulton}). The purpose of this paper is to give a similar geometric realization for Jack polynomials. However, spaces which we use are totally different. Our spaces are Hilbert schemes of points on a surface $X$ which is the total space of a line bundle $L$ over the projective line $\operatorname{\C P}^1$. The parameter $\alpha$ in Jack polynomials relates to our surface $X$ by \begin{equation} \alpha = -\langle{C},{C}\rangle = -c_1(L)[{C}], \label{eq:selfinter}\end{equation} where $C$ is the zero section, and $\langle{C},{C}\rangle$ is the self-intersection number of $C$. It seems difficult to realize Jack polynomials in homology groups of Grassmann manifolds since they have no parameter. The Hilbert scheme $\HilbX{n}$ parametrizing $0$-dimensional subschemes of length $n$ on $X$ has been studied by various peoples (see \cite{Go-book} and the reference therein). The author and Grojnowski independently showed that the direct sum of the homology groups of $\HilbX{n}$ (the summation is over $n$) is a representation space of the Heisenberg algebra (boson Fock space) \cite{Gr,Na-hilb}. It is also well-known that the space of symmetric polynomials is the same representation space. This is the relationship between symmetric functions and Hilbert schemes. Now we explain our realization in more detail. There are two characterizations of Jack polynomials: \begin{aenume} \item an orthogonal basis such that the transition matrix to monomial symmetric functions is strictly upper triangular (see \thmref{thm:defJack}), or \item simultaneous eigenfunctions of a family of commuting differential operators (see above). \end{aenume} We use the characterization~(a). We first identify the complexified ring of symmetric functions with the direct sum of the middle degree homology groups of Hilbert schemes as above. We then identify the inner product with the intersection pairing (\thmref{thm:ident}). This result was essentially proved in \cite{Gr,Na-hilb} combined with \cite{ES2}. Then monomial symmetric functions are identified with fundamental classes of certain middle dimensional subvarieties in the Hilbert schemes (\thmref{thm:monom}), which were first introduced by Grojnowski~\cite{Gr}. Finally, in order to get mutually orthogonal elements, we use the {\it localization\/}. We define an $S^1$-action on the surface $X$ which induces an action on the Hilbert scheme. The localization, which goes back to Bott residue formula \cite{Bott}, enables computations of the intersection product to be reduced to the fixed point set of the $S^1$-action. Then cohomology classes which localize to different fixed point components are mutually orthogonal. This is the mechanism to construct a orthogonal basis. (In this paper, we shall use the equivariant cohomology to formulate the localization following \cite{AtBo}.) The relation to the subvarieties corresponding to monomial symmetric functions can be studied as follows. For each fixed point component, we associate a locally closed submanifold (``stratum'') consisting of points which converge to the fixed point components when they moved by the ${\mathbb C}^*$-actions which extends the $S^1$-action (see \eqref{eq:limit}). The dominance order is identified with the closure relation of the strata \eqref{eq:closure}. The subvarieties corresponding monomial symmetric functions are the closures of stratum. Then it is easy to check that the transition matrix is strictly upper triangular. In fact, almost all arguments work for the total space of a line bundle over any compact Riemann surface. We use the middle degree homology group of Hilbert schemes on which only $H^2(C)$ contributes. ($H^0(C)$ and $H^1(C)$ contribute only to lower degrees.) The motivation of this work comes from Wilson's observation \cite{Wilson} that ``completed phase space'' for the complex Calogero-Moser system (a cousin of the Calogero-Sutherland system) is diffeomorphic to the Hilbert scheme of points on the affine plane. (The author learned this observation from Segal's talk at Warwick, 1996 March.) This made the author to look for the connection between Hilbert schemes and Jack polynomials. However, the connection found in this paper does not follow the Wilson's route. Clarifying relation between these two connections should be an important problem. For example, it is desirable to have a geometric realization of commuting differential operators. It seems natural to conjecture that we get similar geometric realization of Macdonald polynomials (two parameter family explained above) if we replace the homology by the equivariant K-theory. Analogous phenomenon was found for the affine Weyl group and the affine Hecke algebra. The group ring of the affine Weyl group is realized on the homology group of the cotangent bundle of the flag manifold, while the Hecke algebra is realized on the equivariant K-theory. (See \cite{Gi-book}.) We hope to return back in near future. \subsection*{Acknowledgment} The author would like to thank K.~Hasegawa who told him the relation between Jack polynomials and the Calogero-Sutherland system just after Segal's talk at Warwick. It is also a pleasure to acknowledge discussions with T.~Gocho, A.~Matsuo and H.~Ochiai during the seminar on Hilbert schemes of points held in 1996 spring. In particular, Matsuo's talk on Macdonald polynomials was very helpful. \section{Preliminaries}\label{sec:pre} In this section we review the theory of symmetric functions and the equivariant cohomology for later use. \subsection{Symmetric Functions}\label{subsec:symmetric} First we briefly recall the theory of symmetric functions. See \cite{Symmetric} for detail. A {\it partition\/} $\lambda = (\lambda_1,\lambda_2,\lambda_3,\dots)$ is a nonincreasing sequence of nonnegative integers such that $\lambda_i = 0$ for all but finitely many $i$. Let $|\lambda| = \sum_i \lambda_i$. We say $\lambda$ is a partition of $n$ if $|\lambda| = n$. We also use another presentation $\lambda = (1^{m_1}2^{m_2}\dots)$ where $m_k = \# \{i\mid \lambda_i = k\}$. Number of nonzero entries in $\lambda$ is called {\it length\/} of $\lambda$ and denoted by $l(\lambda)$. If $\lambda$ and $\mu$ are partitions, we define $\lambda\cup\mu$ be the partition whose entries are those of $\lambda$ and $\mu$ arranged in the descending order. For a partition $\lambda = (\lambda_1,\lambda_2,\dots)$, we give a Young diagram such that the number of boxes in the $i$th column is $\lambda_i$. Remark that our convention differs from one used in \cite{Symmetric}. Our diagram is rotated by $\pi/2$ from one used in \cite{Symmetric}. The conjugate of a partition $\lambda$ is the partition $\lambda'$ whose diagram is the transpose of the diagram of $\lambda$, i.e., \begin{equation} \label{eq:conj} \lambda_i' = \# \{j\mid \lambda_j\ge i\}. \end{equation} We define $\lambda\ge\mu$ if $|\lambda| = |\mu|$ and \begin{equation} \lambda_1+\lambda_2+\cdots+\lambda_i \ge \mu_1+\mu_2+\cdots+\mu_i \qquad\text{for all $i$}. \label{eq:dom}\end{equation} This defines a partial order on the set of partitions and is called {\it dominance order}. Note that $\lambda\ge \mu$ if and only if $\mu'\ge \lambda'$. Let $\Lambda_N$ be the ring of symmetric functions \begin{equation*} \Lambda_N = {\mathbb Z}[x_1,\dots,x_N]^{{\mathfrak S}_N}, \end{equation*} where the symmetric group ${\mathfrak S}_N$ acts by the permutation of the variables. It is a graded ring: \begin{equation*} \Lambda_N = \bigoplus_{n \ge 0} \Lambda_N^n, \end{equation*} where $\Lambda_N^n$ consists of the homogeneous symmetric functions of degree $n$. It is more relevant for us to consider symmetric functions in ``infinitely many variables'' formulated as follows: let $M\ge N$ and consider the homomorphism \begin{equation*} {\mathbb Z}[x_1,\dots,x_M] \to {\mathbb Z}[x_1,\dots,x_N] \end{equation*} which sends $x_{N+1}$, \dots, $x_M$ to $0$. We have induced homomorphisms \begin{equation*} \rho^n_{M,N}\colon \Lambda_M^n \to \Lambda_N^n, \end{equation*} which is surjective for any $M\ge N$, and bijective for $M\ge N\ge n$. Let \begin{equation*} \Lambda^n \overset{\operatorname{\scriptstyle def.}}{=} \varprojlim \Lambda_N^n. \end{equation*} Then the ring of symmetric functions in infinitely many variables is defined by \begin{equation*} \Lambda \overset{\operatorname{\scriptstyle def.}}{=} \bigoplus_n \Lambda^n. \end{equation*} In the relationship between symmetric functions and Hilbert schemes, the degree $n$ corresponds the number of points, while the number of variables $N$ are irrelevant. This is the only reason why we use the different notation from \cite{Symmetric}. There are several distinguished classes of symmetric functions. The first class is the {\it monomial symmetric function\/} $m_\lambda$. Let $\lambda$ be a partition with $l(\lambda)\le N$. Let \begin{equation*} m_\lambda(x_1,\dots,x_N) \overset{\operatorname{\scriptstyle def.}}{=} \sum_{\alpha\in{\mathfrak S}_N\cdot\lambda} x_1^{\alpha_1}\cdots x_N^{\alpha_N} = \frac{1}{\#\{\sigma\in{\mathfrak S}_N \mid \sigma\cdot\lambda = \lambda\}} \sum_{\sigma\in{\mathfrak S}_N} x_1^{\lambda_{\sigma(1)}}\cdots x_N^{\lambda_{\sigma(N)}}, \end{equation*} where $\alpha = (\alpha_1,\dots,\alpha_N)$ runs over all distinct permutation of $(\lambda_1,\lambda_2, \dots, \lambda_N)$. If $l(\lambda) \le N$, we have \begin{equation*} \rho_{M,N}^n m_\lambda(x_1,\dots,x_M) = m_\lambda(x_1,\dots,x_N). \end{equation*} Hence $m_\lambda$ defines an element in $\Lambda$, which is also denoted by $m_\lambda$. Then $\{ m_\lambda\}_{\lambda}$ is a basis for $\Lambda$. The {\it $n$th power sum\/} is \begin{equation*} p_n \overset{\operatorname{\scriptstyle def.}}{=} \sum x_i^n = m_{(n)}. \end{equation*} For a partition $\lambda = (\lambda_1,\lambda_2,\dots)$, let $p_\lambda = p_{\lambda_1}p_{\lambda_2}\cdots$. Then $\{ p_{\lambda} \}_{\lambda}$ is a basis for $\Lambda\otimes{\mathbb Q}$. (It is {\it not\/} a ${\mathbb Z}$-basis for $\Lambda$.) For a positive real number $\alpha$, we define an inner product $\langle\cdot,\cdot\rangle$ on $\Lambda\otimes{\mathbb Q}$ by \begin{equation*} \langle p_\lambda, p_\mu\rangle \overset{\operatorname{\scriptstyle def.}}{=} \alpha^{l(\lambda)} z_\lambda\delta_{\lambda\mu}, \end{equation*} where $z_\lambda = \prod k^{m_k} m_k!$ for $\lambda = (1^{m_1}2^{m_2}\cdots)$. The Jack polynomials are defined by \begin{Theorem}[\protect{\cite[10.13]{Symmetric}}] For each partition $\lambda$, there is a unique symmetric polynomial $P_\lambda^{(\alpha)}$ satisfying \begin{enumerate} \def(\theequation.\arabic{enumi}){(\theequation.\arabic{enumi})} \def\arabic{enumi}{\arabic{enumi}} \item\label{tri} $P_\lambda^{(\alpha)} = m_\lambda + \sum_{\mu < \lambda} u_{\lambda\mu}^{(\alpha)}m_\mu$ for suitable coefficients $u_{\lambda\mu}^{(\alpha)}$, \item $\langleP_\lambda^{(\alpha)}, P_\mu^{(\alpha)}\rangle = 0$ if $\lambda\ne\mu$. \end{enumerate} \label{thm:defJack}\end{Theorem} The uniqueness is clear since the basis $\{P_\lambda^{(\alpha)}\}_{\lambda}$ is obtained by the Gram-Schmidt orthogonalization from $\{m_\lambda\}_{\lambda}$ with respect to {\it any\/} total order compatible with the dominance order. The nontrivial point lies in (\ref{thm:defJack}.\ref{tri}) where the summation is over $\mu$ which is smaller than $\lambda$ with respect to the dominance order. Our geometric construction gives a new proof of this theorem. The ``integral form'' $J_\lambda^{(\alpha)}$ of $P_\lambda^{(\alpha)}$ is defined by the normalization \begin{equation} J_\lambda^{(\alpha)} \overset{\operatorname{\scriptstyle def.}}{=} c_\lambda(\alpha)P_\lambda^{(\alpha)}, \qquad c_\lambda(\alpha) \overset{\operatorname{\scriptstyle def.}}{=} \prod_{s\in\lambda} (\alpha a(s) + l(s)+1), \label{eq:normal}\end{equation} where ``$s\in\lambda$'' means that $s$ is a box of the Young diagram corresponds to $\lambda$, and $a(s)$ (resp.\ $l(s)$) is the {\it arm\/} (resp.\ {\it leg\/}) length defined by \begin{equation} \label{fig:hooklength} \newcommand{\hfil}{\hfil} \newcommand{\heartsuit}{\heartsuit} \newcommand{\spadesuit}{\spadesuit} \Yvcentermath1 \young(\hfil,\hfil\heartsuit,\hfil\heartsuit\hfil,\hfil s\spadesuit\sps,\hfil\hf\hfil\hf)\qquad\qquad \begin{matrix} a(s) &= \text{number of $\heartsuit$} \\ l(s) &= \text{number of $\spadesuit$} \end{matrix} \end{equation} Since our diagram is rotated, there is no reason to call them arm and leg. But we follow the traditional convention. The {\it augmented\/} monomial symmetric function is defined by \begin{equation*} \tilde{m}_\lambda \overset{\operatorname{\scriptstyle def.}}{=} u_\lambda m_\lambda, \qquad u_\lambda\overset{\operatorname{\scriptstyle def.}}{=} \prod_k m_k \quad\text{for $\lambda = (1^{m_1}2^{m_2}\cdots)$}. \end{equation*} Macdonald conjectured \cite[10.26?]{Symmetric} that $J^{(\alpha)}_\lambda$ is expressed as a linear combination of augmented monomial symmetric function $\tilde{m}_\lambda$ with coefficients in ${\mathbb Z}_{\ge 0}[\alpha]$. This conjecture was proved affirmatively by Knop and Sahi \cite{KS} by a combinatorial method. \subsection{Equivariant Cohomology} We define the equivariant cohomology using the Borel construction. We assume the group is $S^1$, though the adaptation to the general compact Lie group is straightforward. Let $ES^1\to BS^1$ be the universal $S^1$-bundle which is given by the inductive limit of the Hopf fibration $S^{2n+1}\to \operatorname{\C P}^n$. For a topological space $M$ with a circle action, let $M_{S^1} = ES^1\times_{S^1} M$. We have a projection $M_{S^1} \to BS^1$, which is a fibration with fiber $M$. Then, the equivariant cohomology $\HT^*(M)$ is, by definition, the cohomology of $M_{S^1}$. Similarly, the equivariant cohomology with compact support, denoted by $\HTc^*(M)$ is defined by the cohomology of $M_{S^1}$ with compact support in the fiber direction of $M_{S^1}\to BS^1$. We assume coefficients are complex numbers. If $M$ is the space $pt$ consisting of a single point with a trivial $S^1$-action, we have $\HT^*(pt) = H^*(BS^1)$. Since $BS^1$ is the infinite dimensional projective space $\operatorname{\C P}^\infty$, $H^*(BS^1)$ is the polynomial ring with a generator $u$ in $H^2(BS^1)$. We normalize $u$ to be the first Chern class of the tautological line bundle, i.e., the dual of the hyperplane bundle. By the projection $M_{S^1}\to BS^1$, we have $\HT^*(pt)\cong{\mathbb C}[u]$-module structures on $\HT^*(M)$ and $\HTc^*(M)$. In order to relate the equivariant cohomology $\HT^*(M)$ to the ordinary cohomology $H^*(M)$, we consider the Leray-Serre spectral sequence associated with the fibration $M_{S^1}\to BS^1$. It is a spectral sequence converging to $H^*(M_{S^1}) = \HT^*(M)$ with $E_2$-term $E_2^{p,q} = H^p(BS^1)\otimes H^q(M)$. (Note $\pi_1(BS^1) = 0$.) The (decreasing) filtration is given by \begin{equation*} F^p \HT^*(M) = \{ u^p\varphi \mid \varphi\in \HT^*(M) \}. \end{equation*} Then Kirwan proved \begin{Theorem}\cite[5.8]{Kirwan} Let $M$ be a compact symplectic manifold with a Hamiltonian $S^1$-action. Then the Leray-Serre spectral sequence associated with $M_{S^1}\to BS^1$ degenerates at the $E_2$-term. Thus we have \begin{equation*} F^p \HT^{p+q}(M)/ F^{p+1}\HT^{p+q}(M) \cong H^p(BS^1)\otimes H^q(M). \end{equation*} \label{thm:degenerate}\end{Theorem} Kirwan's proof works for a noncompact symplectic manifold provided $f^{-1}((-\infty,c])$ is compact for all $c\in {\mathbb R}$, where $f$ is the moment map associated with the Hamiltonian $S^1$-action (see also \cite[Chapter~5]{Lecture}). Unfortunately, we do not have a natural K\"ahler metric on the Hilbert scheme of points, so we could not check the moment map satisfies the above condition. However, we can take another route to save Kirwan's argument. In stead of the gradient flow of the moment map, we use the ${\mathbb C}^*$-action which extends the $S^1$-action. (See \cite[Chapter~7]{Lecture}.) We have another short-cut when we assume $X$ is the total space of a line bundle over $\operatorname{\C P}^1$. Since the ordinary cohomology groups vanish in odd degree, it is obvious that the spectral sequence degenerates at $E_2$-term. \section{Homology Group of the Hilbert Scheme}\label{sec:homology} The purpose of this section is to identify the complexified ring of symmetric functions $\Lambda\otimes{\mathbb C}$ with the homology group of the Hilbert scheme of $X$. In fact, the result of this section holds when $X$ is the total space of a line bundle $L$ over {\it any\/} compact Riemann surface ${C}$. Let $X$ be as above. It is a nonsingular $2$-dimensional quasi-projective surface containing ${C}$ as the $0$-section. Let $\HilbX{n}$ be the Hilbert scheme parameterizing $0$-dimensional subschemes of length $n$. By a result of Fogarty~\cite{Fog}, it is a nonsingular $2n$-dimensional variety. Let us consider the homology group $H_*(\HilbX{n})$ of the Hilbert scheme. In \cite{Na-hilb}, we have constructed an action of the Heisenberg algebra on the direct sum $\bigoplus_n H_*(\HilbX{n})$. Let us briefly recall the construction. For each $i\in\mathbb Z\setminus\{0\}$, let $P_{C}[i]$ be a subvariety consisting $(\idl_1,\idl_2)\in \coprod_n \HilbX{n-i}\times\HilbX{n}$ such that \begin{equation} \begin{cases} \mathcal J_1\supset \mathcal J_2,\; \text{$\Supp(\mathcal J_1/\mathcal J_2) = \{ x\}$ for some $x\in {C}$} & (\text{when $i > 0$}) \\ \mathcal J_1\subset \mathcal J_2,\; \text{$\Supp(\mathcal J_2/\mathcal J_1) = \{ x\}$ for some $x\in {C}$} & (\text{when $i < 0$}) \end{cases}, \label{eq:corr}\end{equation} where we consider points in $\HilbX{n}$ as ideals of $\shfO_X$. Let $p_a$ be the projection to the $a$th factor in the product $\HilbX{n-i}\times\HilbX{n}$. Note that the restriction of the projection $p_2\colon P_{C}[i]\to \HilbX{n}$ is proper. We define a homomorphism $H_*(\HilbX{n})\to H_*(\HilbX{n-i})$ by \begin{equation*} \varphi\longmapsto p_{1*}\left(p_2^*\varphi\cap \big[P_{C}[i]\big]\right), \end{equation*} where $p_2^*\varphi\cap$ means the cap product of the pull-back of the Poincar\'e dual of $\varphi$ by $p_2^*$. Since $p_2\colon P_{C}[i]\to \HilbX{n}$ is proper, the support of $p_2^*\varphi\cap \big[P_{C}[i]\big]$ is compact, and hence $p_{1*}\left(p_2^*\varphi\cap \big[P_{C}[i]\big]\right)$ can be defined. Moving $n$, we get an endomorphism on $\bigoplus_n H_*(\HilbX{n})$, which we denote by the same symbol $P_{C}[i]$ for brevity. Since $P_{C}[i]\cap (\HilbX{n-i}\times\HilbX{n})$ is a $(2n-i)$-dimensional subvariety (see \cite[\S3]{Na-hilb} or \cite[\S8.3]{Lecture} for the proof), $P_{C}[i]$ maps $H_{2n+k}(\HilbX{n})$ to $H_{2(n-i)+k}(\HilbX{n-i})$. In particular, the middle degree part ($k=0$) is preserved. Then the main result of \cite{Na-hilb} is the following commutator relation \begin{equation} \big[P_{C}[i], P_{C}[j]\big] = (-1)^{i-1}i\delta_{i+j,0}\langle {C},{C}\rangle \operatorname{id}. \label{eq:comrel}\end{equation} (The factor $(-1)^{i-1}i$, which was not determined in \cite{Na-hilb}, was given by Ellingsrud-Str\o mme~\cite{ES2}. See also \cite[Chapter~9]{Lecture} for a proof in the spirit of this paper.) Let $1$ be the generator of $H_0(\HilbX{0}) = {\mathbb C}$. Applying $P_{C}[i]$ successively on $1$, we get a subspace in $\bigoplus_n H_{2n}(\HilbX{n})$ which is the irreducible representation of the Heisenberg algebra. For each partition $\lambda=(1^{m_1}2^{m_2}\cdots)$ define \begin{equation*} P^\lambda{C} \overset{\operatorname{\scriptstyle def.}}{=} P_{C}[-1]^{m_1} P_{C}[-2]^{m_2}\cdots 1 \in H_{2|\lambda|}(\HilbX{|\lambda|}). \end{equation*} Since the representation of the Heisenberg algebra generated by $P_{C}[i]$'s is irreducible, $P^\lambda{C}$'s are linearly independent. On the other hand, by the formula of G\"ottsche~\cite{Got,GS} for the Poincar\'e polynomial $P_t(\HilbX{n})$ of $\HilbX{n}$ (see also \cite[Chapter~7]{Lecture} for the proof based on Morse theory) we have \begin{equation*} \sum_{n=0}^\infty q^n P_t(\HilbX{n}) = \prod_{m=1}^\infty \frac{(1 + t^{2m-1}q^m)^{b_1(X)}} {(1 - t^{2m-2}q^m)(1 - t^{2m}q^m)}\, . \end{equation*} Hence \begin{equation} \sum_{n=0}^\infty q^n \mathop{\text{\rm dim}}\nolimits H_{2n}(\HilbX{n}) = \prod_{m=1}^\infty \frac{1}{1 - q^m}\, . \label{eq:char}\end{equation} The right hand side of \eqref{eq:char} is the same as the character of the irreducible representation of the Heisenberg algebra, hence $P^\lambda{C}$'s span $\bigoplus_n H_{2n}(\HilbX{n})$. In other words, $\{ P^\lambda{C}\,\}_{\lambda}$ is a basis for $\bigoplus_n H_{2n}(\HilbX{n})$. We identify $\bigoplus_n H_{2n}(\HilbX{n})$ with the polynomial ring ${\mathbb C}[p_1,p_2,\cdots]$ (and hence with the complexified ring of symmetric functions $\Lambda\otimes{\mathbb C}$) by \begin{equation} p_\lambda = p_1^{m_1}p_2^{m_2}\cdots \longmapsto P^\lambda{C} = P_{C}[-1]^{m_1} P_{C}[-2]^{m_2}\cdots 1\qquad \text{for $\lambda = (1^{m_1}2^{m_2}\cdots)$.} \label{eq:ident}\end{equation} Then the operator $P_{C}[-i]$ corresponds to the multiplication by $p_i$ when $i > 0$. The main result of this section is \begin{Theorem} \textup{(1)} The direct sum $\bigoplus_n H_{2n}(\HilbX{n})$ of the middle degree homology group is isomorphic to the complexified ring of symmetric functions $\Lambda\otimes{\mathbb C}$ under the identification~\eqref{eq:ident}. \textup{(2)} The intersection pairing $\langle\cdot,\cdot\rangle$ on $H_{2n}(\HilbX{n})$ is given by \begin{equation*} \langle P^\lambda{C}, P^\mu{C}\rangle = (-1)^n \delta_{\lambda\mu}z_\lambda (-\langle{C},{C}\rangle)^{l(\lambda)}, \end{equation*} where $\lambda$, $\mu$ is a partition of $n$. \label{thm:ident}\end{Theorem} If we define a new inner product by $\langle\cdot | \cdot\rangle \overset{\operatorname{\scriptstyle def.}}{=} (-1)^n\langle\cdot,\cdot\rangle$ on $H_{2n}(\HilbX{n})$, it is equal to the one used for the definition of Jack's symmetric functions, where the parameter $\alpha$ is $-\langle{C},{C}\rangle$. \begin{proof}[Proof of \thmref{thm:ident}] The only remaining is to prove the statement~(2). We identify $\bigoplus_n H_{2n}(\HilbX{n})$ with ${\mathbb C}[p_1,p_2,\cdots]$ by (1). Then the commutation relation \eqref{eq:comrel} means that the operator $P_{C}[i]$ for $i > 0$ corresponds to \begin{equation*} (-1)^{i-1} i\, \langle{C},{C}\rangle \pd{}{p_i}. \end{equation*} Hence for $i, j > 0$, we have \begin{equation} \begin{split} & \big[P_{C}[i]^m, P_{C}[-j]^n\big] = \left((-1)^{i-1} i\,\langle{C},{C}\rangle \pd{}{p_i}\right)^m\; p_j^n \\ =\;& \begin{cases} n(n-1)\cdots (n-m+1)\delta_{ij} \left\{(-1)^{i-1}i\,\langle{C},{C}\rangle\right\}^m P_{C}[-j]^{n-m}, &\text{for $n\ge m$}\\ 0, &\text{for $n < m$.} \end{cases} \end{split} \label{eq:PPcom}\end{equation} By construction, $P_{C}[i]$ is the adjoint of $P_{C}[-i]$ with respect to the intersection form. For $\lambda = (1^{m_1}2^{m_2}\cdots)$, $\mu = (1^{n_1}2^{n_2}\cdots)$ let ${\bar\lambda} = (2^{m_2}3^{m_3}\cdots)$, ${\bar\mu} = (2^{n_2}3^{n_3}\cdots)$. Then we have \begin{equation*} \begin{split} \langle P^\lambda{C}, P^\mu{C}\rangle =& \langle P_{C}[-1]^{m_1} P^{\bar\lambda}{C}, P_{C}[-1]^{n_1} P^{\bar\mu}{C}\rangle \\ =& \langle P_{C}[1]^{n_1}P_{C}[-1]^{m_1}P^{\bar\lambda}{C}, P^{\bar\mu}{C}\rangle. \end{split} \end{equation*} Since $P_{C}[-1]$ commutes with $P_{C}[i]$ for $i\ne -1$, we have \begin{equation*} \begin{split} &P_{C}[1]^{n_1}P_{C}[-1]^{m_1}P^{\bar\lambda}{C} = \Big[P_{C}[1]^{n_1}, P_{C}[-1]^{m_1}\Big] P^{\bar\lambda}{C}\\ =\;& \begin{cases} m_1!\; \langle{C},{C}\rangle^{m_1} P^{\bar\lambda}{C} &\text{for $m_1 = n_1$}\\ 0, &\text{otherwise,} \end{cases} \end{split} \end{equation*} where we have used \eqref{eq:PPcom}. Hence we have \begin{equation*} \langle P^\lambda{C}, P^\mu{C}\rangle = \delta_{m_1,n_1} m_1! \langle{C},{C}\rangle^{m_1} \langle P^{\bar\lambda}{C}, P^{\bar\mu}{C}\rangle. \end{equation*} Inductively, we get \begin{equation*} \begin{split} \langle P^\lambda{C}, P^\mu{C}\rangle &= \delta_{\lambda\mu} \prod_i \left((-1)^{i-1}i \langle{C},{C}\rangle\right)^{m_i}\, m_i!\\ &= \delta_{\lambda\mu} z_\lambda (-1)^{\sum i m_i} \left(-\langle{C},{C}\rangle\right)^{\sum m_i}. \end{split} \end{equation*} Since $\sum i m_i = n$, $\sum m_i = l(\lambda)$, we get the assertion. \end{proof} \section{Certain Subvarieties and Monomial Symmetric Functions}\label{sec:subvar} In this section we define certain middle dimensional subvarieties parametrized by partitions and identify them with monomial symmetric functions under the isomorphism given in \thmref{thm:ident}. The subvarieties were first introduced by Grojnowski~\cite{Gr}, and their identification with monomial symmetric functions was proved in \cite[Chapter~9]{Lecture}. We reproduce them for the sake of the reader. Let $X$ be as in \secref{sec:homology}, i.e., the total space of a line bundle over a compact Riemann surface $C$. Let us consider the ${\mathbb C}^*$-action on $X$ given by the multiplication on fibers. It induces an action on the Hilbert scheme $\HilbX{n}$. Let us consider a subvariety of the Hilbert scheme $\HilbX{n}$ defined by \begin{equation} \{ Z\in\HilbX{n} \mid \Supp(\shfO_Z)\subset{C} \}. \label{eq:LSigma}\end{equation} Since the ${\mathbb C}^*$-action retracts $\HilbX{n}$ to a neighborhood of the above subvariety, $\HilbX{n}$ is homotopically equivalent to the subvariety. It was pointed out by Grojnowski \cite{Gr} (see \cite[Chapter~7]{Lecture} for the proof) that its irreducible components $L^\lambda{C}$ are indexed by a partition $\lambda$ as follows: first let $\pi\colon\HilbX{n}\to S^n X$ be the Hilbert-Chow morphism which assigns to a closed subscheme $Z$ of $X$, the $0$-cycle consisting of the points of $Z$ with multiplicities given by the length of the local rings on $Z$. Then \eqref{eq:LSigma} is given by $\pi^{-1}(S^n{C})$, where $S^n{C}$ is considered as a subvariety of $S^n X$. We have a stratification of $S^n{C}$ given by \begin{equation*} S^n {C} = \bigcup_{\lambda} S^n_\lambda {C}, \quad \text{where }S^n_\lambda {C} \overset{\operatorname{\scriptstyle def.}}{=} \left\{ \left.\sum_{i=1}^k \lambda_i [x_i] \in S^n {C} \right| \text{$x_i \neq x_j$ for $i \neq j$} \right\}, \end{equation*} where $\lambda = (\lambda_1,\lambda_2,\dots)$ runs over partitions of $n$. Now consider a locally closed subvariety $\pi^{-1}(S^n_\lambda{C})$ of \eqref{eq:LSigma}. By a result of Brian\c{c}on~\cite{Bri}, the Hilbert-Chow morphism $\pi$ is semismall (see also \cite{GS} or \cite[Chapter~6]{Lecture}), the dimension of the fiber of $\pi$ over a point in $S^n_\lambda{C}$ is $n-l(\lambda)$. Hence the dimension of $\pi^{-1}(S^n_\lambda{C})$ is equal to $n$, which is independent of the partition $\lambda$. Moreover, the fiber of $\pi$ is irreducible again by a result of Brian\c{c}on~\cite{Bri}. Thus the irreducible components of \eqref{eq:LSigma} are given by \begin{equation} L^\lambda{C} \overset{\operatorname{\scriptstyle def.}}{=} \text{Closure of }\pi^{-1}(S^n_\lambda{C}). \label{eq:Llambda}\end{equation} We need another definition of $L^\lambda{C}$. Let us consider a fixed point $\idl$ of the ${\mathbb C}^*$-action in $\HilbX{n}$. Since the fixed point in $X$ is the zero section $C$, we have $\Supp(\shfO_X/\idl)\subset C$. Let us decompose $\idl$ as $\idl_1\cap\cdots\cap\idl_m$ according to the support, i.e., $\Supp(\shfO_X/\idl_k) = \{x_k\}$ and $x_k\ne x_l$ for $k\ne l$. For each $k$, we take a coordinate system $(z,\xi)$ around $x_k$ where $z$ is a coordinate of ${C}$ around $x_k$, and $\xi$ is a fiber coordinate of $L_{x_k}$. Then $\idl_k$ is generated by monomials in $z$ and $\xi$: \begin{equation*} \idl_k = (\xi^{\lambda^{(k)}_1}, z\xi^{\lambda^{(k)}_2}, \dots, z^{N-1}\xi^{\lambda^{(k)}_N}, z^N), \end{equation*} for some partition $\lambda^{(k)} = (\lambda^{(k)}_1,\lambda^{(k)}_2,\dots)$ with $N = l(\lambda^{(k)})$. For a partition $\lambda$ and a point $x\in C$, let \begin{equation} \idl_{\lambda,x} \overset{\operatorname{\scriptstyle def.}}{=} (\xi^{\lambda_1}, z\xi^{\lambda_2}, \dots, z^{N-1}\xi^{\lambda_N}, z^N), \label{eq:idlLx}\end{equation} where $N = l(\lambda)$ and $(z,\xi)$ is the coordinate system around $x$. (See Figure~\ref{fig:Young} for the visualization of $\idl_{\lambda,x}$.) Thus the fixed point can be written as $\idl = \bigcap \idl_{\lambda^{(k)},x_k}$ for some distinct points $x_k$'s in $C$ and partitions $\lambda^{(k)}$. If $x_k$ approaches to $x_l$, $\idl_{\lambda^{(k)},x_k}\cap\idl_{\lambda^{(l)},x_l}$ converges to $\idl_{\lambda^{(k)}\cup\lambda^{(l)}, x_l}$. This shows that the fixed point components are parametrized by partitions $\lambda = \bigcup \lambda^{(k)}$. Let us denote by $S^\lambda{C}$ the corresponding component. By the above discussion, we have \begin{equation*} S^\lambda{C} \overset{\operatorname{\scriptstyle def.}}{=} \left\{ \idl = \bigcap \idl_{\lambda^{(k)},x_k} \mid x_i\in {C},\; x_i \neq x_j \text{ for $i\neq j$},\; \lambda = \bigcup \lambda^{(k)} \right\}. \end{equation*} As a complex manifold, $S^\lambda{C}$ is isomorphic to a product of symmetric product \[ S^\lambda{C}\cong S^{m_1}{C}\times S^{m_2}{C}\times\cdots, \] where $\lambda = (1^{m_1}2^{m_2}\dots)$. Now another description of $L^\lambda{C}$ is \begin{equation} L^\lambda{C} = \text{Closure of } \{ \idl\in\HilbX{n} \mid \lim_{t\to\infty} t\cdot\idl \in S^\lambda{C} \}, \label{eq:limit}\end{equation} where $t\cdot$ denotes the ${\mathbb C}^*$-action. In fact, since the Hilbert-Chow morphism $\pi\colon\HilbX{n}\to S^nX$ is ${\mathbb C}^*$-equivariant, $t\cdot\idl$ stays a compact set, or equivalently converges to a fixed point if and only if $\Supp(\shfO_X/\idl)\subset{C}$. Moreover, $\lim_{t\to\infty} t\cdot\idl$ is contained in the open stratum \[ \left\{ \idl_{(\lambda_1),x_1}\cap \dots \cap \idl_{(\lambda_N),x_N} \mid x_i\in {C},\; x_i \neq x_j \text{ for $i\neq j$}\right\} \] of $S^\lambda{C}$ if and only if $\idl\in\pi^{-1}(S^n_\lambda{C})$. This shows the identification \eqref{eq:limit}. \begin{Remark} As is presented in \cite[Chapter~7]{Lecture}, we can start from \eqref{eq:limit} as a definition of $L^\lambda{C}$. Studying the weight decomposition of the tangent space at a point in $S^\lambda{C}$, we can prove that \eqref{eq:limit} is $n$-dimensional (cf.\ discussion for \propref{prop:pullback}). The irreducibility of \eqref{eq:limit} follows from that of $S^\lambda{C}$. Hence \eqref{eq:limit} is an irreducible component of \eqref{eq:LSigma}. \end{Remark} \begin{Theorem} Under the isomorphism $\Lambda\otimes{\mathbb C} \cong\bigoplus H_{2n}(\HilbX{n})$ given in \thmref{thm:ident}, the class $[L^\lambda{C}]$ corresponds to the monomial symmetric function $m_\lambda$. \label{thm:monom}\end{Theorem} \begin{proof} We shall show \begin{equation} P_{C}[-i][L^\lambda{C}] = \sum_\mu a_{\lambda\mu} [L^\mu{C}], \qquad\text{for any $i \in\mathbb Z_{>0}$}, \label{eq:Ll-ind}\end{equation} where the summation is over partitions $\mu$ of $|\lambda|+i$ which is obtained as follows: \begin{enumerate} \refstepcounter{equation}\label{eq:coeff} \def(\theequation.\arabic{enumi}){(\theequation.\arabic{enumi})} \def\arabic{enumi}{\arabic{enumi}} \item Add $i$ to a term in $\lambda$, say $\lambda_k$ \textup(possibly $0$\textup). \item Then arrange it in descending order. \item Define the coefficient $a_{\lambda\mu}$ by $\#\{ l\mid \mu_l = \lambda_k + i\}$. \end{enumerate} This is the same as the relation between the power sum and the monomial symmetric function: \begin{equation*} p_i m_\lambda = \sum_\mu a_{\lambda\mu} m_\mu. \end{equation*} By induction, we have $[L^\lambda{C}] = m_\lambda$ under the identification $P_{C}[-i] = p_i$. The equation~\eqref{eq:Ll-ind} is proved by studying the intersection product. Let $p_a\colon \HilbX{n}\times\HilbX{n}\to \HilbX{n}$ be the projection onto the $a$th factor ($a = 1,2$). By the definition~\eqref{eq:corr}, we represent $P_{C}[-i]$ as a subvariety. Then its set theoretical intersection with $p_2^{-1}(L^\lambda{C})$ is \begin{equation} \{ (\idl_1,\idl_2)\mid \idl_2\in L^\lambda{C},\; \idl_1\subset\idl_2,\; \Supp (\idl_2/\idl_1) = \{x\} \text{ for some } x\in{C}\}. \label{eq:intersection}\end{equation} Let $\mu = (\mu_1,\mu_2,\dots)$ be a partition of $|\lambda|+i$ which does not necessarily satisfy the condition in \eqref{eq:coeff}. Let $N = l(\mu)$. Since $\{[L^\mu{C}]\}_{\mu}$ is a basis for $H_{2|\mu|}(\HilbX{|\mu|})$, the left hand side of \eqref{eq:Ll-ind} can be written as a linear combination of $[L^\mu{C}]$'s. In order to determine the coefficients of $[L^\mu{C}]$ in $P_{{C}}[-i][L^\lambda{C}]$, it is enough to take arbitrary point $\idl_1$ in $L^\mu{C}$ and restrict cycles to a neighborhood of $\idl_1$. We choose the point $\idl_1 = \idl_{(\mu_1),x_1}\cap\dots\cap\idl_{(\mu_N),x_N}\in L^\mu{C}$ where $x_k$'s are distinct points in $C$. Here $\idl_{(\mu_k),x_k} = (\xi^{\mu_k}, z)$ for the coordinate $(z, \xi)$ around $x_k$. (See \eqref{eq:idlLx} for the definition of $\idl_{\lambda,x}$.) Suppose this point $\idl_1$ is contained in the image of \eqref{eq:intersection} under the projection $p_1$, i.e., there exists $\idl_2$ such that $(\idl_1,\idl_2)$ is a point in \eqref{eq:intersection}. Then the point $x$ must be one of $x_k$'s, and \begin{equation*} \idl_2 = \idl_{(\mu_1),x_1}\cap\dots\cap\idl_{(\mu_{k-1}),x_{k-1}} \cap\idl_{(\mu_k-i), x_k} \cap\idl_{(\mu_{k+1}),x_{k+1}}\cap\dots\cap\idl_{(\mu_N),x_N}, \end{equation*} i.e., $\idl_2$ is obtained from $\idl_1$ by replacing $\idl_{(\mu_k),x_k}$ by $\idl_{(\mu_k-i), x_k}$. Since $\idl_2$ must be a point in $L^\lambda{C}$, $\mu$ is obtained by (a) adding $i$ to $\lambda_k$, and then (b) arranging in descending order. Moreover, if $a_{\lambda\mu}$ is as in \eqref{eq:coeff}, there are exactly $a_{\lambda\mu}$ choices of $x_k$'s. This explains the coefficient $a_{\lambda\mu}$ in the formula. Thus the only remaining thing to check is that each choice of $(\idl_1, \idl_2)$ contributes to $P_{C}[-i][L^\lambda{C}]$ by $[L^\mu{C}]$. This will be shown by checking \begin{enumerate} \refstepcounter{equation}\label{eq:wanttoshow} \def(\theequation.\arabic{enumi}){(\theequation.\arabic{enumi})} \def\arabic{enumi}{\arabic{enumi}} \item $P_{C}[-i]$ and $p_2^{-1}(L^\lambda{C})$ intersect transversally, \item the intersection \eqref{eq:intersection} is isomorphic to $L^\mu{C}$ under the first projection $p_1$, \end{enumerate} in a neighborhood of $(\idl_1,\idl_2)$. Since $\idl_1$ and $\idl_2$ are isomorphic outside $x_k$, we can restrict our concern to $\idl_{(\mu_k),x_k}$ and $\idl_{(\mu_k-i),x_k}$. We take the following coordinate neighborhood around $(\idl_{(\mu_k),x_k}, \idl_{(\mu_k-i), x_k})$ in $\HilbX{\mu_k}\times\HilbX{\mu_k-i}$: \begin{align*} &\left\{ ( (\xi^{\mu_k}+f_1(\xi), z+ g_1(\xi)), (\xi^{\mu_k-i} + f_2(\xi), z+g_2(\xi))) \mid \text{$f_1$, $g_1$, $f_2$, $g_2$ as follows}\right\} \\ &\qquad\qquad f_1(\xi) = a_1\xi^{\mu_k-1} + a_2\xi^{\mu_k-2} + \dots + a_{\mu_k}, \\ &\qquad\qquad g_1(\xi) = b_1 + b_2\xi + \dots + b_{\mu_k-i}\xi^{\mu_k-i-1} \\ &\qquad\qquad\qquad\qquad\qquad + (b_{\mu_k-i+1} + b_{\mu_k-i+2}\xi + \dots + b_{\mu_k}\xi^{i-1})(\xi^{\mu_k-i} + f_2(\xi))\\ &\qquad\qquad f_2(\xi) = a'_1\xi^{\mu_k-i-1} + a'_2\xi^{\mu_k-i-2} + \dots + a'_{\mu_k-i}, \\ &\qquad\qquad g_2(\xi) = b'_1+ b'_2\xi + \dots + b'_{\mu_k-i}\xi^{\mu_k-i-1} \end{align*} where $(a_1,\dots, a_{\mu_k}, b_1,\dots, b_{\mu_k})$ (resp.\ $(a'_1,\dots, a'_{\mu_k-i}, b'_1,\dots, b'_{\mu_k-i})$) is in a neighborhood of $0$ in ${\mathbb C}^{2\mu_k}$ (resp.\ ${\mathbb C}^{2(\mu_k-i)}$). Then the above ideal is contained in $P_{C}[-i]$ if and only if the followings hold \begin{enumerate} \refstepcounter{equation} \def(\theequation.\arabic{enumi}){(\theequation.\arabic{enumi})} \def\arabic{enumi}{\arabic{enumi}} \item $\xi^{\mu_k} + f_1(\xi) = \xi^i(\xi^{\mu_k-i} + f_2(\xi))$, \item $g_1(\xi) - g_2(\xi)$ is divisible by $\xi^{\mu_k-i} + f_2(\xi)$. \end{enumerate} Namely, the defining equation for $P_{{C}}[-i]$ is \begin{equation*} \begin{split} & a_1 = a'_1, a_2 = a'_2, \dots, a_{\mu_k-i} = a'_{\mu_k-i}, \\ & a_{\mu_k-i+1} = \cdots = a_{\mu_k} = 0, \\ & b_1 = b'_1, b_2 = b'_2, \dots, b_{\mu_k-i} = b'_{\mu_k-i}. \end{split} \end{equation*} On the other hand, the defining equation for $p_2^{-1}(L^\lambda{C})$ is \begin{equation*} a'_1 = a'_2 = \cdots = a'_{\mu_k-i} = 0. \end{equation*} Now our assertions \eqref{eq:wanttoshow} are immediate. \end{proof} Our next task is to explain a geometric meaning of the dominance order \eqref{eq:dom}. It is given by modifying the stratification introduced in \cite{Bri,Iar}. For $i\ge 0$, let $\shfO_X(-iC)$ be the sheaf of functions vanishing to order $\ge i$ along the zero section $C$. Let $\idl\subset\shfO_X$ be an ideal of colength $n$ such that the support of $\shfO_X/\idl$ is contained in $C$. We consider the sequence $(\lambda'_1,\lambda'_2,\dots)$ of nonnegative integers given by \begin{equation*} \lambda'_i(\idl)\overset{\operatorname{\scriptstyle def.}}{=} \operatorname{length}\left( \frac{\shfO_X(-(i-1)C)}{\idl\cap\shfO_X(-(i-1)C)+\shfO_X(-iC)} \right). \end{equation*} The reason why we put the prime become clear later. The sequence in \cite{Bri,Iar} was defined by replacing $\shfO_X(-iC)$ by $\mathfrak m_x^i$ where $\mathfrak m_x$ is the maximal ideal corresponding to a point $x$. As in \cite[Lemma~1.1]{Iar}, we have $\idl\supset\shfO_X(-nC)$, hence $\lambda'_i(\idl) = 0$ for $i\ge n+1$. From the exact sequence \begin{equation*} 0 \to \frac{\shfO_X(-iC)}{\idl\cap\shfO_X(-iC)} \to \frac{\shfO_X(-(i-1)C)}{\idl\cap\shfO_X(-(i-1)C)} \to \frac{\shfO_X(-(i-1)C)}{\idl\cap\shfO_X(-(i-1)C)+\shfO_X(-iC)} \to 0, \end{equation*} we have \begin{equation*} \sum_{i=1}^n \lambda'_i(\idl) = n. \end{equation*} Let us decompose the ideal $\idl$ by its support, i.e., $\idl = \idl_1\cap\dots\cap\idl_N$ such that $\{ \Supp(\shfO_X/\idl_k) \}_k$ are $N$ distinct points. By definition, we have \begin{equation} \lambda_i'(\idl) = \sum_{k=1}^N \lambda_i'(\idl_k). \label{eq:sum}\end{equation} Suppose that $\idl$ satisfies $\Supp(\shfO_X/\idl) = \{x\}$ for some $x\in C$. We take a coordinate system $(z,\xi)$ around $x$ where $\xi$ is a coordinate for the fiber. If $\xi^i f_1(z),\dots, \xi^i f_d(z)$ form a basis of \begin{equation*} \frac{\shfO_X(-iC)}{\idl\cap\shfO_X(-iC)+\shfO_X(-(i+1)C)}, \end{equation*} Then $\xi^{i-1} f_1(z),\dots, \xi^{i-1} f_d(z)$ are linearly independent in \begin{equation*} \frac{\shfO_X(-(i-1)C)}{\idl\cap\shfO_X(-(i-1)C)+\shfO_X(-iC)}. \end{equation*} Hence we have $\lambda'_i(\idl) \ge \lambda'_{i+1}(\idl)$. Thus $(\lambda'_1(\idl), \lambda'_2(\idl), \dots)$ is a partition of $n$. By \eqref{eq:sum}, the same is true for general $\idl$ which do not necessarily satisfy $\Supp(\shfO_X/\idl)=\{x\}$. Let us denote the partition by $\lambda'(\idl)$. For a partition $\lambda' = (\lambda'_1,\lambda'_2,\dots)$ of $n$, let $W^{\lambda'}$ be the set of ideals $\idl\subset\shfO_X$ with colength $n = |\lambda'|$ such that $\shfO_X/\idl$ is supported on $C$ and $\lambda'(\idl) = \lambda'$. Since \begin{equation*} \operatorname{length}(\shfO_X(-iC)/\idl\cap\shfO_X(-iC)) \leq \sum_{j=i+1}^n \lambda'_j = n - \sum_{j=1}^i \lambda'_j \end{equation*} is a closed condition on $\idl$, the union \begin{equation*} \bigcup_{\mu'\ge \lambda'} W^{\mu'} \end{equation*} is a closed subset of $\{\idl\in \HilbX{n}\mid \Supp(\shfO_X/\idl)\subset C\}$. Thus we have \begin{equation} \label{eq:closure} \text{Closure of }W^{\lambda'} \subset\bigcup_{\mu'\ge \lambda'} W^{\mu'}. \end{equation} Suppose that $\lambda'$ is the conjugate of $\lambda$ as in \eqref{eq:conj}. We get the following third description of $L^\lambda{C}$. \begin{Proposition} $L^\lambda{C} = \text{Closure of }W^{\lambda'}$. \label{prop:third}\end{Proposition} \begin{proof} Let us write $\lambda = (\lambda_1,\dots,\lambda_N)$ with $N = l(\lambda)$. By \eqref{eq:limit}, a generic point $\idl$ in $L^\lambda{C}$ satisfies $\lim_{t\to\infty} t\cdot\idl = \idl_{(\lambda_1),x_1}\cap \dots \cap \idl_{(\lambda_N),x_N}$ such that $x_i$'s are distinct points in $C$. Since the support of $\idl$ cannot move as $t\to\infty$, we can decompose $\idl = \idl_1\cap\dots\cap\idl_N$ such that $\Supp(\shfO_X/\idl_k) = \{x_k\}$. Take a coordinate system $(z,\xi)$ around $x_k$ as before. Then $\idl_{(\lambda_k),x_k}$ was defined by $(z,\xi^{\lambda_k})$. Since $\lim_{t\to\infty} t\cdot\idl_k = \idl_{(\lambda_k),x_k}$, we have \begin{equation*} \idl_k = (\xi^{\lambda_k}, z + a_1 \xi + a_2 \xi^2 + \cdots + a_{\lambda_k-1}\xi^{\lambda_k-1}) \end{equation*} for some $a_1,\dots, a_{\lambda_k-1}$ (cf.\ the proof for \thmref{thm:monom}). Then $\lambda'(\idl_k) = (1^{\lambda_k})$. By \eqref{eq:sum}, we have $$ \lambda'(\idl) = \lambda'. $$ This shows $L^{\lambda}{C}\subset\text{Closure of }W^{\lambda'}$. Conversely, suppose $\idl$ is a point in $W^{\lambda'}$. Let \begin{equation*} \idl_i \overset{\operatorname{\scriptstyle def.}}{=} \frac{\idl\cap\shfO_X(-(i-1)C)}{\idl\cap\shfO_X(-iC)}. \end{equation*} Then $\idl' \overset{\operatorname{\scriptstyle def.}}{=} \bigoplus \idl_i$ satisfies $\lim_{t\to\infty} t\cdot\idl = \idl'$ and $\idl'\in W^{\lambda'}$. This shows $W^{\lambda'}\subset L^\lambda{C}$. \end{proof} \section{Equivariant Cohomology of Hilbert Schemes} Let $X$ be the total space of a line bundle over $\operatorname{\C P}^1$ and $C$ the zero section. $X$ is the quotient space of $({\mathbb C}^2\setminus\{0\})\times{\mathbb C}$ by the ${\mathbb C}^*$-action given by \begin{equation*} (z_0, z_1, \xi) \mapsto (\lambda z_0, \lambda z_1, \lambda^{-\alpha}\xi) \qquad \lambda\in{\mathbb C}^*, \end{equation*} where $\alpha = -c_1(L)[{C}] = -\langle{C},{C}\rangle$. We denote by $[(z_0, z_1, \xi)]$ the equivalence class containing $(z_0, z_1, \xi)$. The projection $X \to \operatorname{\C P}^1$ is given by $[(z_0, z_1, \xi)]\mapsto [z_0:z_1]$, hence $\xi$ is a coordinate for the fiber. We consider the $S^1$-action on $\HilbX{n}$ induced by the action on $X$ defined by \begin{equation} [(z_0, z_1, \xi)] \mapsto [(z_0, t^{-1} z_1, t^{\alpha}\xi)]\qquad \text{for $t\in S^1$}. \label{eq:action}\end{equation} Note that this differs from the $S^1$-action studied in the previous section given by $[(z_0, z_1, \xi)] \mapsto [(z_0, z_1, t\xi)]$. We shall study the equivariant cohomology of $\HilbX{n}$ with respect to the above $S^1$-action. The reason for using this $S^1$-action is to identify the normalization factor~\eqref{eq:normal} with the equivariant Euler class (see \propref{prop:pullback}). Let $pt$ be the space consisting of a single point with a trivial $S^1$-action. Let us denote the obvious $S^1$-equivariant morphism by $p\colon \HilbX{n}\to pt$. We have an equivariant push-forward $p_*$. Define a bilinear form $\langle \cdot, \cdot\rangle_{S^1}\colon \HTc^k(\HilbX{n})\otimes \HTc^k(\HilbX{n})\to\HT^{2(k-2n)}(pt)$ by \begin{equation} \langle \varphi,\psi\rangle_{S^1} \overset{\operatorname{\scriptstyle def.}}{=} p_*(\varphi\cup \psi), \label{eq:inner}\end{equation} where $\cup$ is the cup product \begin{equation*} \cup\colon\HTc^k(\HilbX{n})\otimes\HTc^k(\HilbX{n})\to \HTc^{2k}(\HilbX{n}). \end{equation*} This is trivial unless $k$ is even and $k\ge 2n$. We assume this condition henceforth. In particular, $\langle\cdot,\cdot\rangle_{S^1}$ is symmetric. Let $j^*\colon \HTc^k(\HilbX{n})\to \HT^k(\HilbX{n})$ be the natural homomorphism. Then the cup product $\cup$ factor through $j^*$: \begin{equation*} \varphi\cup \psi = \varphi\cup j^*\psi, \end{equation*} where $\cup$ in the right hand side is the cup product \begin{equation*} \cup \colon \HTc^k(\HilbX{n})\otimes\HT^k(\HilbX{n})\to \HTc^{2k}(\HilbX{n}). \end{equation*} In particular, $\langle\cdot,\cdot\rangle_{S^1}$ is well-defined on \begin{equation*} \HTc^k(\HilbX{n})/\operatorname{Ker} j^* \cong \operatorname{Im} j^*. \end{equation*} Let $\HilbX{n}_{S^1} \to BS^1$ be the fibration used in the definition of the equivariant cohomology. By \thmref{thm:degenerate} (see also the discussion after \thmref{thm:degenerate}), the spectral sequence of the fibration in cohomology degenerates. Let us denote by $F^*\HT^k(\HilbX{n})$ the corresponding decreasing filtration. Since \(H^q(\HilbX{n}) = 0\) for $q > 2n$, we have \begin{equation*} \begin{split} & F^p \HT^k(\HilbX{n}) = \HT^k(\HilbX{n}) \qquad\text{for $p\le k-2n$}, \\ & \HT^k(\HilbX{n})/F^{k-2n+1}\HT^k(\HilbX{n}) \cong \HT^{k-2n}(pt)\otimes H^{2n}(\HilbX{n}). \end{split} \end{equation*} If $j^*\psi \in F^{k-2n+1}\HT^k(\HilbX{n})$, then $\varphi\cup j^* \psi\in F^{2(k-2n)+1}\HTc^{2k}(\HilbX{n})$. Hence $p_*(\varphi\cup j^*\psi)\in F^{2(k-2n)+1}\HT^{2(k-2n)}(pt)\cong 0$. In particular, $\langle\cdot,\cdot\rangle_{S^1}$ is well-defined on $\operatorname{Im} j^*/\operatorname{Im} j^*\cap F^{k-2n+1}\HT^k(\HilbX{n})$. Consider the composition \begin{equation} \begin{gathered} \HTc^{k}(\HilbX{n})/\operatorname{Rad}\langle\ ,\ \rangle_{S^1} \xrightarrow{j^*} \operatorname{Im} j^*/\operatorname{Im} j^*\cap F^{k-2n+1}\HT^k(\HilbX{n})\\ \to \HT^k(\HilbX{n})/F^{k-2n+1}\HT^k(\HilbX{n}) \cong \HT^{k-2n}(pt)\otimes H^{2n}(\HilbX{n}), \end{gathered} \label{eq:compo}\end{equation} which is injective by the above discussion. We have an analogous injective homomorphism between {\it ordinary\/} cohomology groups \begin{equation*} H^{2n}_c(\HilbX{n})/\operatorname{Rad}\langle\ ,\ \rangle \to H^{2n}(\HilbX{n}). \end{equation*} However, the bilinear form $\langle\ ,\ \rangle$ on $H^{2n}_c(\HilbX{n})$ is nondegenerate by \thmref{thm:ident}(2). Hence $\operatorname{Rad}\langle\ ,\ \rangle = 0$. Moreover, the Poincar\'e duality implies $\mathop{\text{\rm dim}}\nolimits H^{2n}_c(\HilbX{n}) = \mathop{\text{\rm dim}}\nolimits H^{2n}(\HilbX{n})$. Therefore the natural homomorphism $ H^{2n}_c(\HilbX{n})\to H^{2n}(\HilbX{n})$ is an isomorphism. \begin{Theorem} Suppose $k$ is even and $k\ge 2n$. Then the composition~\eqref{eq:compo} \begin{equation} \HTc^{k}(\HilbX{n})/\operatorname{Rad}\langle\ ,\ \rangle_{S^1} \to \HT^{k-2n}(pt)\otimes H^{2n}(\HilbX{n}) \label{eq:isom}\end{equation} is an isomorphism. Moreover, the bilinear form $\langle\cdot ,\cdot \rangle_{S^1}$ on the left hand side is equal to one on the right hand side induced by the intersection pairing on $H^{2n}(\HilbX{n})\cong H^{2n}_c(\HilbX{n})$. \label{thm:isom}\end{Theorem} \begin{proof} The subvariety $L^\lambda{C}$ defined in \eqref{eq:Llambda} is invariant under the $S^1$-action. Hence its Poincar\'e dual defines an element in the equivariant cohomology $\HTc^{2n}(\HilbX{n})$, which we denote by $[L^\lambda{C}]_{S^1}$. Under \eqref{eq:compo} with $k=2n$, $[L^\lambda{C}]_{S^1}\bmod\operatorname{Rad}\langle\ ,\ \rangle_{S^1}$ is mapped to $1\otimes [L^\lambda{C}]$. More generally, if $u$ is the generator of $\HT^*(pt)$ which lives in $\HT^{2}(pt)$, $u^{k/2 - n}[L^\lambda{C}]_{S^1}$ is mapped to $u^{k/2 - n}\otimes [L^\lambda{C}]$. This shows that \eqref{eq:isom} is surjective. As we have already seen the injectivity, it is an isomorphism. Next consider the bilinear forms on the both hand sides of \eqref{eq:isom}. The doubt arises from that $\HT^k(\HilbX{n})\cong \bigoplus_{p+q=k}\HT^p(pt)\otimes H^q(\HilbX{n})$ is not an {\it algebra\/} isomorphism, it is only an isomorphism between {\it vector spaces}. If we consider the graded space $G\HT^k(\HilbX{n})$ instead of $\HT^k(\HilbX{n})$, it becomes an algebra isomorphism. However, as we pointed out above, $F^{k-2n+1}\HT^k(\HilbX{n})$ does not contribute to $\langle\cdot ,\cdot \rangle_{S^1}$. Hence we have the assertion. \end{proof} \section{Localization} We assume $X$ is the total space of a line bundle over $\operatorname{\C P}^1$ with $\alpha = -c_1(L)[\operatorname{\C P}^1] > 0$. Let us consider the $S^1$-action given in \eqref{eq:action}. The fixed point set consists of two components \begin{equation*} \{ [(1,0,0)] \} \text{ and } \{ [(0, 1,\xi)] \mid \xi\in{\mathbb C} \}. \end{equation*} The first is an isolated point and the latter is isomorphic to ${\mathbb C}$. If $Z\in \HilbX{n}$ is fixed by the induced action, its support must be contained in the above fixed point set. There are a family of fixed points $Z_\lambda$ with support $[(1,0,0)]$ indexed by a partition $\lambda = (\lambda_1,\lambda_2,\dots)$ of $n$ as follows: let $(z,\xi)$ be a coordinate system around $[(1,0,0)]$ given by $(z,\xi) \leftrightarrow [(1,z,\xi)]$. Then the corresponding ideal $\idl_\lambda$ is given by \begin{equation*} \idl_\lambda \overset{\operatorname{\scriptstyle def.}}{=} (\xi^{\lambda_1}, z\xi^{\lambda_2}, \dots, z^{N-1}\xi^{\lambda_N}, z^N), \end{equation*} where $N = l(\lambda)$. Namely, $\idl_\lambda = \idl_{\lambda,[(1,0,0)]}$ in the notation~\eqref{eq:idlLx}. The ideal can be visualized by the Young diagram as follows. (Remind that our diagram is rotated by $\pi/2$ from one used in \cite{Symmetric}.) If we write the monomial $z^{p-1}\xi^{q-1}$ at the position $(p,q)$, the ideal is generated by monomials outside the Young diagram. For example, the Young diagram in figure~\ref{fig:Young} corresponds to the ideal $(\xi^4, z\xi^3, z^2\xi, z^3)$. Although it is not necessary for our purpose, one can show that these ideals are a complete list of fixed points in $\HilbX{n}$ whose support are $[(1,0,0)]$ when $\alpha > 0$. \begin{figure}[htbp] \begin{center} \leavevmode \setlength{\unitlength}{0.01250000in} \begin{picture}(73,114)(0,-10) \path(22,62)(22,82)(2,82) (2,62)(22,62) \path(22,42)(22,62)(2,62) (2,42)(22,42) \path(42,42)(42,62)(22,62) (22,42)(42,42) \path(22,22)(22,42)(2,42) (2,22)(22,22) \path(42,22)(42,42)(22,42) (22,22)(42,22) \path(22,2)(22,22)(2,22) (2,2)(22,2) \path(42,2)(42,22)(22,22) (22,2)(42,2) \path(62,2)(62,22)(42,22) (42,2)(62,2) \put(7,87){\makebox(0,0)[lb]{\smash{$\xi^4$}}} \put(25,67){\makebox(0,0)[lb]{\smash{$z\xi^3$}}} \put(45,27){\makebox(0,0)[lb]{\smash{$z^2\xi$}}} \put(67,7){\makebox(0,0)[lb]{\smash{$z^3$}}} \end{picture} \caption{Young diagram and an ideal} \label{fig:Young} \end{center} \end{figure} Let $i_\lambda\colon Z_\lambda\to \HilbX{n}$ be the inclusion. We have a homomorphism between equivariant cohomology groups: \begin{equation*} \HT^*(Z_\lambda) \ni \varphi \longmapsto i_{\lambda*}\varphi \in \HTc^*(\HilbX{n}). \end{equation*} Let $e(T_{Z_\lambda}\HilbX{n})$ be the $S^1$-equivariant Euler class of the tangent space at $Z_\lambda$, considered as an element of $\HT^{4n}(Z_\lambda)$. Then we have \begin{Theorem} Let $\langle\cdot,\cdot\rangle_{S^1}$ be the bilinear pairing defined in \eqref{eq:inner}. Then we have \begin{equation*} \langle i_{\lambda*}\varphi, i_{\mu*}\psi\rangle_{S^1} = \begin{cases} 0 & \text{if $\lambda\ne\mu$}, \\ p_{\lambda*} \left(\varphi\cup \psi\cup e(T_{Z_\lambda}\HilbX{n})\right) &\text{if $\lambda=\mu$}, \end{cases} \end{equation*} where $p_\lambda\colon Z_\lambda\to pt$ is the obvious map. \label{thm:push}\end{Theorem} \begin{proof} By the projection formula, we have \begin{equation*} i_{\lambda*}\varphi\cup i_{\mu*}\psi = i_{\mu*}\left( i_{\mu}^*i_{\lambda*}\varphi \cup \psi\right). \end{equation*} The homomorphism $i_{\mu}^*i_{\lambda*}$ is zero when $\lambda\neq\mu$, and is given by the multiplication by the equivariant Euler class $e(T_{Z_\lambda}\HilbX{n})$ if $\lambda=\mu$. Therefore, \makeatletter\tagsleft@false\begin{align} & \langle i_{\lambda*}\varphi, i_{\lambda*}\psi\rangle_{S^1} = p_*\left(i_{\lambda*}\varphi\cup i_{\lambda*}\psi\right) = p_*i_{\lambda*}\left(\varphi \cup \psi \cup e(T_{Z_\lambda}\HilbX{n})\right) \notag \\ =\;& p_{\lambda*}\left(\varphi \cup \psi \cup e(T_{Z_\lambda}\HilbX{n})\right) \tag*{\qed} \end{align}\makeatother \renewcommand{\qed}{} \end{proof} Taking a class from $\HT^*(Z_\lambda)$ for each $\lambda$, we can give a set of mutually orthogonal elements in $\HTc^*(\HilbX{n})$. These are candidates for Jack polynomials. Our remaining task is to choose classes so that two conditions in \thmref{thm:defJack} will be satisfied. Thus we need to know monomial symmetric functions, or their geometric counterparts $[L^\lambda{C}]_{S^1}$. In order to study $[L^\lambda{C}]_{S^1}$, we need to know the $S^1$-module structure of its normal bundle at $Z_\lambda$. In fact, the formula will become clearer if we consider the two dimensional torus action given by \begin{equation*} [(z_0, z_1, \xi)] \mapsto [(z_0, t_1 z_1, t_2\xi)]\qquad \text{for $(t_1, t_2)\in T = S^1\times S^1$}. \end{equation*} The $S^1$-action studied in \secref{sec:subvar} is given by the restriction to $(1,t)\in T$, while the $S^1$-action given by \eqref{eq:action} is the restriction to $(t^{-1}, t^{\alpha})$. The fixed point $Z_\lambda$ defined above is also fixed by this torus action. The tangent space $T_{Z_\lambda}\HilbX{n}$ at $Z_\lambda$ is a $T$-module and have a weight decomposition \begin{equation*} T_{Z_\lambda}\HilbX{n} = \bigoplus_{p,q\in\mathbb Z} H(p,q), \end{equation*} where $H(p,q) = \{ v\in T_{Z_\lambda}\HilbX{n} \mid (t_1,t_2)\cdot v = t_1^p t_2^q v\;\text{for $(t_1,t_2)\in T$}\}$. \begin{Lemma} The character of the tangent space $T_{Z_\lambda}\HilbX{n}$ is given by \begin{equation*} \sum_{p,q} \mathop{\text{\rm dim}}\nolimits_{\mathbb C} H(p,q) T_1^p T_2^q = \sum_{s\in\lambda} \left(T_1^{l(s)+1}T_2^{-a(s)} + T_1^{-l(s)}T_2^{a(s)+1}\right). \end{equation*} \label{lem:weight}\end{Lemma} \begin{proof} The corresponding formula was proved in \cite{ES} (see also \cite[2.2.5]{Go-book} or \cite[Proposition~5.5]{Lecture}) when $X$ is replaced by ${\mathbb C}^2$ with the torus action given by \begin{equation*} (z_1,z_2) \mapsto (t_1 z_1, t_2 z_2)\qquad \text{for $(z_1,z_2)\in{\mathbb C}^2$, $(t_1,t_2)\in T$.} \end{equation*} (The presentation in \cite{ES} or \cite{Go-book} is different from the above. After substituting the arm/leg-length to their formula, we get the above formula.) Since the exponential map gives a $T$-equivariant isomorphism between a neighborhood of $0\in{\mathbb C}^2$ and a neighborhood of $[(1,0,0)]\in X$, we get the assertion. \end{proof} Since the $S^1$-equivariant Euler class of an $S^1$-module is given by the product of weights, we get the following. \begin{Corollary} The $S^1$-equivariant Euler class of the tangent space at $Z_\lambda$ is given by \begin{equation*} e(T_{Z_\lambda}\HilbX{n}) = u^{2n}\prod_{s\in\lambda} \left(-\alpha a(s) - l(s) - 1 \right)\cdot \prod_{s\in\lambda} \left(\alpha(a(s) + 1) + l(s)\right). \end{equation*} \label{cor:tangent}\end{Corollary} In particular, $Z_\lambda$ is an isolated fixed point of $S^1$-action and $e(T_{Z_\lambda}\HilbX{n})$ is nonzero since $\alpha > 0$. In particular, we have \begin{Corollary} Let $\varphi_\lambda$ be a nonzero element in $\HT^k(Z_\lambda)$. Then $\{ i_{\lambda*}\varphi_\lambda\pmod{\operatorname{Rad}\langle\ ,\ \rangle} \}_{\lambda}$ forms an orthogonal basis of $\HTc^{4n+k}(\HilbX{n})/\operatorname{Rad}\langle\ ,\ \rangle$. \label{cor:orth}\end{Corollary} Let us decompose the tangent space $T_{Z_\lambda}\HilbX{n}$ as \begin{gather*} N_{Z_\lambda}^{>0} \overset{\operatorname{\scriptstyle def.}}{=} \bigoplus_{p\in\mathbb Z} \bigoplus_{q > 0}H(p,q), \qquad N_{Z_\lambda}^{\le 0} \overset{\operatorname{\scriptstyle def.}}{=} \bigoplus_{p\in\mathbb Z} \bigoplus_{q \le 0}H(p,q), \\ T_{Z_\lambda}\HilbX{n} = N_{Z_\lambda}^{>0}\oplus N_{Z_\lambda}^{\le 0} \end{gather*} \begin{Proposition} The pull-back of the equivariant cohomology class $[L^\lambda{C}]_{S^1}$ by the inclusion $i_{\lambda}\colon Z_\lambda\to \HilbX{n}$ is equal to the $S^1$-equivariant Euler class of $N_{Z_\lambda}^{>0}$. It is given by \begin{equation*} i_\lambda^*[L^\lambda{C}]_{S^1} = e(N_{Z_\lambda}^{>0}) = u^{n}\prod_{s\in\lambda} \left(\alpha(a(s) + 1) + l(s)\right). \end{equation*} \label{prop:pullback}\end{Proposition} \begin{proof} By the description~\eqref{eq:limit} of $L^\lambda{C}$, it is nonsingular at $Z_\lambda$ and its tangent space is equal to $N_{Z_\lambda}^{\le 0}$. Hence the normal bundle is $N_{Z_\lambda}^{>0}$. Thus we have $i_\lambda^*[L^\lambda{C}]_{S^1} = e(N_{Z_\lambda}^{>0})$. The above formula follows from \lemref{lem:weight}. \end{proof} \begin{Theorem} For each partition $\lambda$, let $F_\lambda$ be the cohomology class \begin{equation} i_{\lambda*}\left(\frac{u^n}{e(N_{Z_\lambda}^{\le 0})}\right) = i_{\lambda*}\left( \frac{1}{ \prod_{s\in\lambda} \left(-\alpha a(s) - l(s) - 1 \right)}\right) \label{eq:Flambda}\end{equation} considered as an element in \begin{equation*} \HTc^{4n}(\HilbX{n})/\operatorname{Rad}\langle\ ,\ \rangle \cong \HT^{2n}(pt)\otimes H^{2n}_c(\HilbX{n}) \xrightarrow[u^{-n}\otimes\text{P.D.}]{\cong} H_{2n}(\HilbX{n}). \end{equation*} Then they satisfy \begin{align} & F_{\lambda} = [L^\lambda{C}] + \sum_{\mu < \lambda} u_{\lambda\mu}^{(\alpha)}[L^\mu{C}] \qquad\text{for some $u_{\lambda\mu}^{(\alpha)}$,} \label{eq:FL}\\ &\langle F_{\lambda}, F_{\mu}\rangle = 0 \qquad\text{for $\lambda\neq\mu$}. \label{eq:orth} \end{align} \label{thm:Jack}\end{Theorem} \begin{Corollary} Under the identification $\bigoplus H_{2n}(\HilbX{n})\cong\Lambda\otimes{\mathbb C}$ given in \thmref{thm:ident}, $F_{\lambda}$ corresponds to the Jack's symmetric function $P_{\lambda}^{(\alpha)}$, where the parameter $\alpha$ is given by $-\langle{C},{C}\rangle$ \textup(see \eqref{eq:selfinter}\textup). \end{Corollary} \begin{proof}[Proof of \protect{\thmref{thm:Jack}}] The equation \eqref{eq:orth} follows from \thmref{thm:push}. Let us take the generator $\varphi_\lambda\in\HT^0(Z_\lambda)$. By \corref{cor:orth}, we can write $[L^\lambda{C}]_{S^1}$ as a linear combination of $i_{\mu*}\varphi_\lambda$ modulo $\operatorname{Rad}\langle\ ,\ \rangle$. One can check that $Z_\mu$ is contained in $W^{\mu'}$. Hence $L^\lambda{C}=\text{Closure of } W^{\lambda'}$ (see \propref{prop:third}) contains $Z_\mu$ only if $\mu\le\lambda$ by \eqref{eq:closure}. In particular, the pull-back $i_\mu^*[L^\lambda{C}]_{S^1} = 0$ unless $\mu\le\lambda$. Hence we can write \begin{equation*} [L^\lambda{C}]_{S^1} = \sum_{\mu\le\lambda} v_{\lambda\mu}^{(\alpha)} i_{\mu*}\varphi_\lambda\mod{\operatorname{Rad}\langle\ ,\ \rangle} \end{equation*} for some $v_{\lambda\mu}^{(\alpha)}$. In order to compute $v_{\lambda\lambda}^{(\alpha)}$, we consider the pull-back by $i_\lambda^*$. We have \begin{align*} i_\lambda^*[L^\lambda{C}]_{S^1} &= e(N_{Z_\lambda}^{>0}) &&\text{(by \propref{prop:pullback})} \\ &= \frac{e(T_{Z_\lambda}\HilbX{n})} {e(N_{Z_\lambda}^{\le 0})} &&\text{(since $T_{Z_\lambda}\HilbX{n} = N_{Z_\lambda}^{>0}\oplus N_{Z_\lambda}^{\le 0}$)} \\ &=\frac{1}{u^n}i_\lambda^* i_{\lambda*}\left( \frac{u^n}{e(N_{Z_\lambda}^{\le 0})}\right) &&\text{(by the self-intersection formula)}. \end{align*} Thus we get the assertion. \end{proof} Finally note that the denominator in \eqref{eq:Flambda} coincides with the normalization factor~\eqref{eq:normal} up to sign. Hence the integral form $J_\lambda^{(\alpha)}$ is an {\it integral\/} class. This shows that $J_\lambda^{(\alpha)}$ is expressed as a linear combination of $m_\lambda$ with coefficients in ${\mathbb Z}$ for any positive integer $\alpha$. Thus we have $J_\lambda^{(\alpha)} = \sum_{\mu<\lambda} \tilde{u}_{\lambda\mu}^{(\alpha)} m_\mu$ with $\tilde{u}_{\lambda\mu}^{(\alpha)}\in{\mathbb Z}[\alpha]$. This proves a part of Macdonald conjecture \cite[10.26?]{Symmetric} mentioned in \subsecref{subsec:symmetric}.
1996-10-04T08:53:31
9610
alg-geom/9610004
en
https://arxiv.org/abs/alg-geom/9610004
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9610004
Sacha Sardo Infirri
Alexander V Sardo Infirri
Partial Resolutions of Orbifold Singularities via Moduli Spaces of HYM-type Bundles
LaTex2e, 30 pages with 1 table
null
null
null
null
Let $\Gamma$ be a finite group acting linearly on $\C^n$, freely outside the origin, and let $N$ be the number of conjugacy classes of $\Gamma$ minus one. A construction of Kronheimer of moduli spaces $X_\zeta$ of translation-invariant $\Gamma$-equivariant instantons on $\C^2$ is generalised to $\C^n$. The moduli spaces $X_\zeta$ depend on a parameter $\zeta\in\Q^N$. The following results are proved: for $\zeta=0$, $X_0$ is isomorphic to $\C^n/\Gamma$; if $\zeta\neq 0$, the natural maps $X_\zeta\to X_0$ are partial resolutions. The moduli $X_\zeta$ are furthermore shown to admit K\"ahler metrics which are Asymptotically Locally Euclidean (ALE). A description of the singularities of $X_\zeta$ using deformation complexes is given, and is applied in particular to the case $\Gamma\subset\SU(3)$. It is conjectured that for general $\Gamma$ and generic $\zeta$ that the singularities of $X_\zeta$ are at most quadratic. When $\Gamma\subset\SU(3)$ a natural holomorphic 3-form is constructed on the smooth locus of $X_\zeta$, which is conjectured to be non-vanishing. The morphims $X_\zeta\to X_0$ are expected to be crepant resolutions and $X_\zeta$ to be smooth for generic choices of the parameter $\zeta$. Related open problems in higher-dimensional complex geometry are also mentioned. The paper has a companion paper which identifies the moduli $X_\zeta$ with representation moduli of McKay quivers, and describes them completely in the case of abelian groups.
[ { "version": "v1", "created": "Fri, 4 Oct 1996 06:49:50 GMT" } ]
2008-02-03T00:00:00
[ [ "Infirri", "Alexander V Sardo", "" ] ]
alg-geom
\part{BlackBoard Bold} \newcommand{{\mathbb C}}{{\mathbb C}} \newcommand{{\C^*}}{{{\mathbb C}^*}} \newcommand{{\mathbb R}}{{\mathbb R}} \newcommand{\Z}{{\mathbb Z}} \newcommand{\Q}{{\mathbb Q}} \newcommand{{\mathbb N}}{{\mathbb N}} \newcommand{{\mathbb F}}{{\mathbb F}} \newcommand{{\mathbb H}}{{\mathbb H}} \newcommand{{\mathbb P}}{{\mathbb P}} \newcommand{_{\ge 0}}{_{\ge 0}} \newcommand{{\Q_+}}{{\Q_+}} \newcommand{\R\nneg}{{\mathbb R}_{\ge 0}} \newcommand{\protect\overline{T}{}}{\protect\overline{T}{}} \newcommand{\widetilde{M}}{\widetilde{M}} \newcommand{\widetilde{L}}{\widetilde{L}} \newcommand{\widetilde{P}}{\widetilde{P}} \newcommand{{\mathcal A}}{{\mathcal A}} \newcommand{{\mathcal B}}{{\mathcal B}} \newcommand{{\mathcal C}}{{\mathcal C}} \newcommand{{\mathcal D}}{{\mathcal D}} \newcommand{{\mathcal E}}{{\mathcal E}} \newcommand{{\mathcal F}}{{\mathcal F}} \newcommand{{\mathcal G}}{{\mathcal G}} \newcommand{{\mathcal H}}{{\mathcal H}} \newcommand{{\mathcal I}}{{\mathcal I}} \newcommand{{\mathcal J}}{{\mathcal J}} \newcommand{{\mathcal K}}{{\mathcal K}} \newcommand{{\mathcal L}}{{\mathcal L}} \newcommand{{\mathcal M}}{{\mathcal M}} \newcommand{{\mathcal N}}{{\mathcal N}} \newcommand{{\mathcal O}}{{\mathcal O}} \newcommand{{\mathcal P}}{{\mathcal P}} \newcommand{{\mathcal Q}}{{\mathcal Q}} \newcommand{{\mathcal R}}{{\mathcal R}} \newcommand{{\mathcal S}}{{\mathcal S}} \newcommand{{\mathcal T}}{{\mathcal T}} \newcommand{{\mathcal U}}{{\mathcal U}} \newcommand{{\mathcal V}}{{\mathcal V}} \newcommand{{\mathcal W}}{{\mathcal W}} \newcommand{{\mathcal X}}{{\mathcal X}} \newcommand{{\mathcal Y}}{{\mathcal Y}} \newcommand{{\mathcal Z}}{{\mathcal Z}} \newcommand{\boldsymbol{\alpha}}{\boldsymbol{\alpha}} \newcommand{\boldsymbol{\gamma}}{\boldsymbol{\gamma}} \newcommand{\boldsymbol{\omega}}{\boldsymbol{\omega}} \newcommand{{\mathbf a}}{{\mathbf a}} \newcommand{\bfa}{{\mathbf a}} \newcommand{{\mathbf e}}{{\mathbf e}} \newcommand{\bde}{{\mathbf e}} \newcommand{{\mathbf f}}{{\mathbf f}} \newcommand{{\mathbf g}}{{\mathbf g}} \newcommand{\bdh}{{\mathbf h}} \newcommand{{\mathbf A}}{{\mathbf A}} \newcommand{{\mathbf B}}{{\mathbf B}} \newcommand{{\mathbf C}}{{\mathbf C}} \newcommand{{\mathbf I}}{{\mathbf I}} \newcommand{{\mathbf J}}{{\mathbf J}} \newcommand{{\mathbf K}}{{\mathbf K}} \newcommand{{\mathbf Q}}{{\mathbf Q}} \newcommand{{\mathbf R}}{{\mathbf R}} \newcommand{{\mathbf V}}{{\mathbf V}} \newcommand{{\mathbf W}}{{\mathbf W}} \newcommand{{\mathfrak g}}{{\mathfrak g}} \newcommand{{\mathfrak u}}{{\mathfrak u}} \newcommand{{\mathbf g}}{{\mathbf g}} \newcommand{\gothk}{{\mathbf k}} \newcommand{{\mathbf t}}{{\mathbf t}} \newcommand{{\mathbf u}}{{\mathbf u}} \newcommand{\pi\mbox{-}\operatorname{\rm ext}}{\pi\mbox{-}\operatorname{\rm ext}} \newcommand{\circlearrowleft}{\circlearrowleft} \newcommand{\circlearrowright}{\circlearrowright} \newcommand{\upharpoonleft\downharpoonright}{\upharpoonleft\downharpoonright} \newcommand{\downharpoonleft\upharpoonright}{\downharpoonleft\upharpoonright} \newcommand{\centerdot}{\centerdot} \newcommand{M^{\Gamma }}{M^{\Gamma }} \newcommand{\GL^\Gamma\!}{\GL^\Gamma\!} \newcommand{\U^\Gamma\!}{\U^\Gamma\!} \newcommand{\cN^{\Gamma }}{{\mathcal N}^{\Gamma }} \newcommand{\cN^{\Gamma }}{{\mathcal N}^{\Gamma }} \newcommand{\cN^{\Gamma }\cap\mu^{-1}(0)}{{\mathcal N}^{\Gamma }\cap\mu^{-1}(0)} \newcommand{\cT_\zeta}{{\mathcal T}_\zeta} \newcommand{\cT^{\text{IC}}}{{\mathcal T}^{\text{IC}}} \newcommand{\cT^{\text{sing-IC}}}{{\mathcal T}^{\text{sing-IC}}} \newcommand{\cT_\zeta^{\text{IC}}}{{\mathcal T}_\zeta^{\text{IC}}} \newcommand{\cT_\zeta^{\text{sing-IC}}}{{\mathcal T}_\zeta^{\text{sing-IC}}} \newcommand{{\mathcal S}^{\text{IC}}}{{\mathcal S}^{\text{IC}}} \newcommand{\cT_{\zeta ,x}^{\text{IC}}}{{\mathcal T}_{\zeta ,x}^{\text{IC}}} \newcommand{C_{\zeta,x}}{C_{\zeta,x}} \newcommand{F_{\zeta,x}}{F_{\zeta,x}} \newcommand{T^\Pi}{T^\Pi} \newcommand{\overline}{\overline} \newcommand{\clos{S}}{\overline{S}} \newcommand{\clos{T}}{\overline{T}} \newcommand{\tilde \alpha}{\tilde \alpha} \newcommand{\widetilde X}{\widetilde X} \newcommand{\widetilde{X/\Gamma}}{\widetilde{X/\Gamma}} \newcommand{\Gamma}{\Gamma} \newcommand{\mbox{$\Gamma$}}{\mbox{$\Gamma$}} \newcommand{\widehat\Gamma}{\widehat\Gamma} \newcommand{\Lambda^2\to\Lambda^1}{\Lambda^2\to\Lambda^1} \newcommand{\Irr(\ga)}{\Irr(\Gamma)} \newcommand{\hom_\Gamma}{\hom_\Gamma} \newcommand{{\Rep_\cQ}}{{\Rep_{\mathcal Q}}} \newcommand{\contract}{\dashv} \newcommand{\bar{\partial}}{\bar{\partial}} \newcommand{\overline{D}}{\overline{D}} \newcommand{{\bar{q}}}{{\bar{q}}} \newcommand{{\bar{z}}}{{\bar{z}}} \newcommand{{\bar{\imath}}}{{\bar{\imath}}} \newcommand{{\bar{\jmath}}}{{\bar{\jmath}}} \newcommand{{\bar{k}}}{{\bar{k}}} \newcommand{{\bar{m}}}{{\bar{m}}} \newcommand{{\overline{\Pi}}}{{\overline{\Pi}}} \newcommand{{\overline w}}{{\overline w}} \newcommand{\dashrightarrow}{\dashrightarrow} \newcommand{\birationalto}{\dashrightarrow} \newcommand{\gitquot}[1]{/\!\!/_{\!#1}\,} \newcommand{\stackrel\pi\to}{\stackrel\pi\to} \newcommand{\stackrel\rho\to}{\stackrel\rho\to} \newcommand{>\!\!>}{>\!\!>} \newcommand{\squared}[1]{#1^2} \newcommand{\stackrel M\sim}{\stackrel M\sim} \newcommand{i.e.\xspace}{i.e.\xspace} \newcommand{c.f.\xspace}{c.f.\xspace} \newcommand{resp.\xspace}{resp.\xspace} \newcommand{neighbourhood}{neighbourhood} \newcommand{\text{s-group}}{\text{s-group}} \newcommand{geometric invariant theory}{geometric invariant theory} \newcommand{vector space}{vector space} \newcommand{McKay quiver}{McKay quiver} \newcommand{positive definite Hermitian inner product}{positive definite Hermitian inner product} \newcommand{asymptotically locally Euclidean}{asymptotically locally Euclidean} \newcommand{K\"ahler}{K\"ahler} \newcommand{\kah\ reduction}{K\"ahler\ reduction} \newcommand{singularity}{singularity} \newcommand{singularities}{singularities} \newcommand{Hermitian-Yang-Mills}{Hermitian-Yang-Mills} \DeclareMathOperator{\sign}{sign} \DeclareMathOperator{\mult}{mult} \DeclareMathOperator{\Map}{Map} \DeclareMathOperator{\conv}{conv} \DeclareMathOperator{\ord}{ord} \DeclareMathOperator{\Ind}{Ind} \DeclareMathOperator{\id}{id} \DeclareMathOperator{\pr}{pr} \DeclareMathOperator{\grad}{grad} \DeclareMathOperator{\Diff}{Diff} \DeclareMathOperator{\Vect}{Vect} \DeclareMathOperator{\pow}{pow} \DeclareMathOperator{\Image}{Im} \DeclareMathOperator{\rk}{rank} \DeclareMathOperator{\spn}{span} \DeclareMathOperator{\codim}{codim} \DeclareMathOperator{\Irr}{Irr} \DeclareMathOperator{\Sym}{Sym} \DeclareMathOperator{\trace}{trace} \DeclareMathOperator{\Proj}{Proj} \DeclareMathOperator{\Spec}{Spec} \DeclareMathOperator{\Exc}{Exc} \DeclareMathOperator{\Hilb}{Hilb} \DeclareMathOperator{\ch}{ch} \DeclareMathOperator{\td}{td} \DeclareMathOperator{\Pic}{Pic} \DeclareMathOperator{\adj}{ad} \DeclareMathOperator{\Hom}{Hom} \DeclareMathOperator{\End}{End} \DeclareMathOperator{\Aut}{Aut} \DeclareMathOperator{\Rep}{Rep} \DeclareMathOperator{\rad}{rad} \DeclareMathOperator*{\In}{In} \DeclareMathOperator*{\Out}{Out} \DeclareMathOperator{\supp}{supp} \DeclareMathOperator{\interior}{int} \DeclareMathOperator{\Star}{Star} \DeclareMathOperator{\Ext}{Ext} \DeclareMathOperator{\ver}{vert} \DeclareMathOperator{\ext}{ext} \DeclareMathOperator{\orb}{orb} \DeclareMathOperator{\Lie}{Lie} \DeclareMathOperator{\Cl}{Cl} \DeclareMathOperator{\Stab}{Stab} \DeclareMathOperator{\Fix}{Fix} \DeclareMathOperator{\Mat}{Mat} \DeclareMathOperator{\GL}{GL} \DeclareMathOperator{\gl}{\mathfrak{gl}} \DeclareMathOperator{\PGL}{PGL} \DeclareMathOperator{\PSL}{PSL} \DeclareMathOperator{\SL}{SL} \DeclareMathOperator{\PL}{PL} \DeclareMathOperator{\sll}{\mathfrak{sl}} \DeclareMathOperator{\U}{U} \DeclareMathOperator{\gu}{\mathfrak{u}} \newcommand{\gu}{\gu} \DeclareMathOperator{\SU}{SU} \DeclareMathOperator{\su}{\mathfrak{su}} \DeclareMathOperator{\PU}{PU} \DeclareMathOperator{\Vol}{Vol} \def\qsing#1/#2(#3){\frac{#1}{#2}(#3)} \def\<#1>{\langle#1\rangle} \newcommand{\ip}[2]{\langle#1,#2\rangle} \newcommand{\gen}[2]{\langle#1,\dots,#2\rangle} \newcommand{{}^{-1}}{{}^{-1}} \newcommand{{}^t\kern -0.2em}{{}^t\kern -0.2em} \newcommand{{}^{\vee}}{{}^{\vee}} \newcommand{{}^{\bot}}{{}^{\bot}} \newcommand{{}^{\ast}}{{}^{\ast}} \newcommand{{}_{\ast}}{{}_{\ast}} \newcommand{\sst}[1]{#1^{\protect\text{ss}}} \newcommand{\st}[1]{#1^{\protect\text{s}}} \newcommand{\nopagebreak}{\nopagebreak} \hyphenation{ortho-nor-mal} \newcommand{\cvec}[2]{{#1\choose#2}} \newcommand{\ntp}[3]{{{#1}_{#2},\dots,{#1}_{#3}}} \newcommand{\vect}[3]{{({#1}_{#2},\dots,{#1}_{#3})}} \newcommand{\nn}[3]{{{#1}^{#2},\dots,{#1}^{#3}}} \newcommand{\vv}[3]{{({#1}^{#2},\dots,{#1}^{#3})}} \newcommand{\ntup}[2]{#1,\dots,#2} \newcommand{\abs}[1]{\lvert#1\rvert} \newcommand{\norm}[1]{\lVert#1\rVert} \newcommand{\card}[1]{\sharp\,#1} \newcommand{\map}[5]{$$\begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array}$$} \newcommand{\map}{\map} \newcommand{\mapeq}[6]{\begin{equation} \begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array} \protect\label{#6} \end{equation}} \newcommand{\corresp}[4]{$$\begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array}$$} \newcommand{\correspeq}[5]{\begin{equation} \begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array} \protect\label{#5} \end{equation}} \newenvironment{point}[1]{\smallbreak\noindent{\bf #1\/}\nopagebreak}{\smallbreak} \newenvironment{empoint}[1]{\smallskip\noindent{\em #1\/}:}{\smallskip} \numberwithin{equation}{section} \def\*#1:#2*{\S\ref{sec:#1}.\ref{sec:#1:#2}} \newenvironment{Ventry}[1]% {\begin{list}{}{\renewcommand{\makelabel}[1]{\emph{##1:}\hfil}% \settowidth{\labelwidth}{{\emph{#1:}}}% \setlength{\leftmargin}{\labelwidth+\labelsep}}}% {\end{list}} \newcommand{\entrylabel}[1]{\mbox{\emph{#1:}}\hfil} \newenvironment{entry} {\begin{list}{}% {\renewcommand{\makelabel}{\entrylabel}% \setlength{\labelwidth}{12mm}% \setlength{\leftmargin}{14mm}% }% }% {\end{list}} \newlength{\Mylen} \newcommand{\Lentrylabel}[1]{% \settowidth{\Mylen}{\emph{#1}}% \ifthenelse{\lengthtest{\Mylen > \labelwidth}}% {\parbox[b]{\labelwidth {\makebox[0pt][l]{\emph{#1}}\\}}% {\emph{#1} \hfil\relax} \newenvironment{Lentry} {\renewcommand{\entrylabel}{\Lentrylabel}% \begin{entry}} {\end{entry}} \newcommand{\Pentrylabel}[1]{% {\emph{#1}} \hfil\relax} \newenvironment{Pentry} {\renewcommand{\entrylabel}{\Pentrylabel}% \begin{entry}} {\end{entry}} \newcommand{\Mentrylabel}[1]% {\raisebox{0pt}[1ex][0pt]{\makebox[\labelwidth][l]% {\parbox[t]{\labelwidth}{\hspace{0pt}\emph{#1:}}}}} \newenvironment{Mentry}% {\renewcommand{\entrylabel}{\Mentrylabel}\begin{entry}}% {\end{entry}} \begin{document} \title[\runningheadstring]{\titlestring\footnote{Maths Subject Classification \msc}} \author{Alexander V.\ Sardo Infirri} \email{[email protected]} \address{Research Institute for Mathematical Sciences\\ Ky\=oto University\\ Oiwake-ch\protect\=o\\ Kitashirakawa\\ Saky\protect\=o-ku\\ Ky\=oto 606-01\\ Japan} \date{2 October 1996} \begin{abstract} \hyphenation{trans-la-tion} Let~$\Gamma$ be a finite group acting linearly on~${\mathbb C}^n$, freely outside the origin, and let $N$ be the number of conjugacy classes of $\Gamma$ minus one. A construction of Kronheimer~\cite{kron:ale} of moduli spaces $X_\zeta$ of translation-invariant $\Gamma$-equivariant instantons on ${\mathbb C}^2$ is generalised to ${\mathbb C}^n$. The moduli spaces $X_\zeta$ depend on a parameter~$\zeta\in\Q^N$. The following results are proved: for $\zeta=0$, $X_0$ is isomorphic to ${\mathbb C}^n/\Gamma$; if $\zeta\neq 0$, the natural maps $X_\zeta\to X_0$ are partial resolutions. The moduli $X_\zeta$ are furthermore shown to admit K\"ahler metrics which are Asymptotically Locally Euclidean (ALE). A description of the singularities of $X_\zeta$ using deformation complexes is given, and is applied in particular to the case $\Gamma\subset\SU(3)$. It is conjectured that for general $\Gamma$ and generic $\zeta$ that the singularities of $X_\zeta$ are at most quadratic. When $\Gamma\subset\SU(3)$ a natural holomorphic 3-form is constructed on the smooth locus of $X_\zeta$, which is conjectured to be non-vanishing. The morphims $X_\zeta\to X_0$ are expected to be crepant resolutions and $X_\zeta$ to be smooth for generic choices of the parameter $\zeta$. Related open problems in higher-dimensional complex geometry are also mentioned. The paper has a companion paper~\cite{sacha:flows} which identifies the moduli $X_\zeta$ with representation moduli of McKay quivers, and describes them completely in the case of abelian groups. \end{abstract} \maketitle \tableofcontents \setcounter{section}{-1} \section{Introduction} \label{sec:intro} This paper is concerned with affine {\em orbifold singularities}, namely with singularities of the type $X={\mathbb C}^n/\Gamma$ for $\Gamma$ a finite group acting linearly on ${\mathbb C}^n$. More precisely, this paper gives a method for constructing {\em partial resolutions\/} of $X$, namely birational morphisms $Y\to X$ which are isomorphisms over the smooth locus of $X$. \subsection{Background} \label{sec:intro:back} The method in question was first introduced by Kronheimer~\cite{kron:ale}. It can be described in various ways, depending on one's point of view. One description (although maybe not the most straight-forward) is to construct moduli spaces $X_\zeta$ of instantons on the trivial bundle ${\mathbb C}^n\times R\to{\mathbb C}^n$. Here $R$ denotes the regular representation space for the group $\Gamma$, and $\zeta$ is a linearisation of the bundle action. The instantons are required to satisfy Hermitian-Yang-Mills-type equations, as well as additional $\Gamma$-equivariance and translation-invariance properties. In Kronheimer's case, $\Gamma\subset\SU(2)$, and the moduli spaces $X_\zeta$ can in fact be viewed as hyper-K\"ahler\ quotients. Kronheimer shows that $X_0$ is isomorphic to ${\mathbb C}^2/\Gamma$, that the natural maps $X_\zeta\to X_0$ are partial resolutions for $\zeta\neq 0$, and that indeed $X_\zeta$ coincides with the minimal resolution of ${\mathbb C}^2/\Gamma$ for generic choices of $\zeta$. Furthermore, $X_\zeta$ inherit natural hyper-K\"ahler\ metrics on their non-singular locus, which are shown to be Asymptotically Locally Euclidean (ALE): they asymptotically approximate the Euclidean metric at infinity (up to terms vanishing with the inverse of the fourth power of the radial coordinate). \subsection{Main Results} \label{sec:intro:main} In the present case $n$ is any integer greater than or equal to $2$, and $\Gamma\subset\U(n)$ is assumed to act on ${\mathbb C}^n$ freely outside the origin for any $n\geq 2$, which means that $X={\mathbb C}^n/\Gamma$ has an isolated singularity.\footnote{This is for the purpose of simplicity --- the method would seem to be applicable to the general case with some modifications} As a result, the moduli $X_\zeta$ are only {\em K\"ahler \/} rather than hyper-K\"ahler\ quotients (in actual fact they are more conveniently described in term of geometric invariant theory). The main result is \begin{nonumberthm}[c.f.\xspace Thms.~\ref{thm:X0free}, \ref{thm:partial_resolutions} and \ref{thm:ale} in the text] Let $\Gamma$ act linearly on ${\mathbb C}^n$ and freely outside the origin and let $X_\zeta$ be the moduli spaces constructed in Section~{\rm \ref{sec:setup:moment}}. Then $X_0$ is isomorphic to $X={\mathbb C}^n/\Gamma$, and for $\zeta\neq 0$, the natural morphisms $X_\zeta\to X_0$ are partial resolutions. Furthermore, the inherited K\"ahler\ metrics on the smooth loci of $X_\zeta$ are Asymptotically Locally Euclidean in the sense of~\cite{kron:ale}. \end{nonumberthm} \subsection{Two Conjectures} \label{sec:intro:discussion} The final sections of the paper discuss and develop two conjectures. \begin{nonumberconj}[c.f.\xspace Conj. \ref{conj:formal}] The singularities of $X_\zeta$ (for generic $\zeta$, say) are at most quadratic algebraic. \end{nonumberconj} This is a common occurrence for moduli spaces of this kind~\cite{nadel:quadratic,gold_mill:flat,gold_mill:invariance}. Its proof can be reduced to proving the formality of a certain differential graded Lie algebra (DGLA) by the methods of~\cite{gold_mill:invariance}. This is done in Section~\ref{sec:defcplx}, where the singularities of $X_\zeta$ are described in terms of deformation complexes~\cite{ahs,don_kron:4mfds}. The concept of formality is explained, and it is suggested that the complex relevant to $X_\zeta$ may be formal, in a way similar to Tian~\cite{tian:smoothness} and Todorov's~\cite{todo:weil-peterson} work. This would imply that the singularities of $X_\zeta$ at most quadratic algebraic. This conjecture is also checked by computer for low order abelian groups in~$\U(3)$ using the methods in the companion paper~\cite{sacha:flows}. \begin{nonumberconj}[c.f.\xspace Conj.\ \ref{conj:su3} in the main text] If $\Gamma\subset\SU(3)$, the morphisms $X_\zeta\to X_0$ are crepant, and if $\zeta$ is generic, $X_\zeta$ is smooth and its Euler number is equal to the orbifold Euler number of $X_0$ as defined in~\cite{dhvw:i}. \end{nonumberconj} The fact that $X_\zeta$ has at most quadratic singularities has been verified for the abelian subgroups of order less than~11. The smoothness of $X_\zeta$ has been verified in the abelian cases $\qsing 1/3(1,1,1)$, $\qsing 1/6(1,2,3)$, $\qsing 1/7(1,2,4)$, $\qsing 1/8(1,2,5)$, $\qsing 1/9(1,2,6)$, $\qsing 1/10(1,2,7)$ and $\qsing 1/11(1,2,8)$. Both these verifications were done by a brute-force listing of singularities of $X_\zeta$ for all possible $\zeta$, using the methods given in the companion paper~\cite{sacha:flows}. The cases $\Gamma\subset\SU(n)$ present a particular interest. The problem of constructing a crepant resolution of $C^3/\Gamma$ with the same orbifold Euler number was only recently completed~\cite{mar_ols_per,roan:mirror_cy,mark:res_168,roan:res_a5,ito:trihedral,roan:calabi-yau}. For the case ${\mathbb C}^4/\Gamma$, one can obtain some interesting analogous results if one considers terminalisations rather than resolutions~\cite{sacha:sl4}. In Section~\ref{sec:su3}, a natural holomorphic 3-form is constructed on the smooth locus of $X_\zeta$: this is conjectured to be non-vanishing. Its norm is shown to be constant if and only if the induced metric on $X_\zeta$ is Ricci-flat (which does not usually turn out to be the case, however). \subsection{Related Questions} \label{sec:intro:related} This paper has a companion paper~\cite{sacha:flows} in which the moduli $X_\zeta$ are identified with representation moduli of McKay quivers. This allows one to explicitly describe the case of Abelian~$\Gamma$ in terms of ``flows'' on the McKay quiver. Explicit computations are carried out for groups of low order, and the conjectures about the smoothness and the triviality of the canonical bundle are checked (by brute-force computer calculations) for abelian subgroups of $\SU(3)$ of order less than or equal to 11. Many questions are left open by the present work, besides the conjectures already mentioned. For instance, do the birational models $X_\zeta$ of ${\mathbb C}^n/\Gamma$ possess any special properties with regards to their singularities (are they terminal?) Are all terminal models for a given 3-fold singularity obtained by this construction? What is the relationship between the different $X_\zeta$? Are they related by flips/flops? Is it possible, by choosing very special values of $\zeta$, to produce blowups $X_\zeta\to X_0$ which are interesting from the point of view of higher dimensional geometry, for instance, for the construction of flips in dimensions~4 and greater? \subsection{Methods} \label{sec:intro:methods} The methods used include geometric invariant theory, K\"ahler\ quotients, and elementary theory of the moduli of bundles, the necessary aspects of which are reviewed in Sections~\ref{sec:git}, \ref{sec:hym} and~\ref{sec:setup}. Furthermore, the same construction is presented under different angles with the intention that the reader who is familiar with one of them (or with Kronheimer's work~\cite{kron:ale}) will be able to follow the discussion easily. The later sections devoted to the various conjectures raised by the main results touch on the theory of deformation complexes, and concepts of Kuranishi germs, formality, and so on. Some background is also provided, although not as extensive as to be able to describe it as ``self-contained'', given the conjectural nature of the material. \subsection{Outline} \label{sec:intro:outline} The outline of this paper is as follows. Section~\ref{sec:git} reviews material regarding geometric invariant theory\ quotients which is necessary to define the moduli $X_\zeta$. No essentially new material is involved, although the formulation of some of the results may be un-familiar to non-specialists. Section~\ref{sec:hym} review material concerning moduli of Hermitian-Yang-Mills connections. This is not essential to the understanding of the main results, although some familiarity is desirable for the understanding of Section~\ref{sec:defcplx}. Section~\ref{sec:setup} deals with the definition and construction of the moduli $X_\zeta$. Section~\ref{sec:partial} gives the proof that $X_0$ is isomorphic to $X$ and that $X_\zeta\to X_0$ are partial resolutions. Section~\ref{sec:ale} proves that the induced metrics on $X_\zeta$ are ALE. Section~\ref{sec:defcplx} contains the discussion of the singularities of $X_\zeta$ in the language of deformation complexes. This includes the conjecture that the singularities of $X_\zeta$ are at most quadratic algebraic. Section~\ref{sec:su3} deals with the case $\Gamma\subset\SU(3)$, the construction of the holomorphic three-form on the non-singular locus of $X_\zeta$ and conjecture~\ref{conj:su3}. \subsection{Acknowledgments} \label{sec:intro:ack} The present paper and its companion paper~\cite{sacha:flows} consist mostly\footnote{Minor portions have been rewritten to include references to advances in the field made since then.} of excerpts of my D.Phil.\ thesis~\cite{sacha:thesis}, and I wish to acknowledge the University of Oxford and Wolfson College for their hospitality during its preparation. I am grateful to the Rhodes Trust for financial support during my first three years, and to Wolfson College for a loan in my final year. The conversion from thesis to article format was done while I was a Research Assistant in RIMS, Kyoto. I also take the opportunity to thank my supervisors Peter Kronheimer and Sir~Michael Atiyah who provided me with constant advice, encouragement and support and whose mathematical insight has been an inspiration. I also wish to thank William Crawley-Boevey, Michel Brion, Gavin Brown, Jack Evans, Partha Guha, Katrina Hicks, Frances Kirwan, Alistair Mees, Alvise Munari, Martyn Quick, David Reed, Miles Reid, Michael Thaddeus, and, last but not least, my parents and family. \section{Geometric Invariant Theory} \label{sec:git} This section recalls the geometric invariant theory of affine varieties and proves some results which shall be needed in the sequel. These results have been included here, because, although they are well-known to the experts, no elementary treatment exists.\footnote{For the generalisation of these results to arbitrary quasi-projective varieties, see~\cite{thaddeus:git_flips,dolg_hu}.} \subsection{Linearisations and GIT quotients} \label{sec:git:lin} Let~$G$ be a reductive group acting linearly on a complex affine variety~$X$. In this situation,\footnote{When $X$ is only quasi-projective the definition of a linearisation involves specifying an ample bundle over $X$ as well as a lift of the action to the bundle.}\/ a ($G$-)\emph{linearisation} is a lifting of the $G$-action to the trivial line bundle $L\to X$. Such a linearisation is determined completely by the action of $G$ on the fibres of $L$, namely by a character $\zeta\colon G\to {\C^*}$. For every character $\zeta$, denote by $L_\zeta$ the trivial bundle endowed with the corresponding linearisation. The space of $G$-invariant sections of $L_\zeta$ is denoted by~$H^0(L_\zeta)^G$. The geometric invariant theory\ (GIT) quotient of $X$ by $G$ with respect to $\zeta$ is defined by $$X\gitquot\zeta G :=\Proj \bigoplus_{k\in{\mathbb N}}H^0(kL_\zeta)^G.$$ \begin{example} \label{example:zeta_zero} If $\zeta=0$ is the trivial character, then the corresponding quotient $X\gitquot 0G$ coincides with the usual affine GIT quotient $X\gitquot{}G$. In fact, suppose that $X=\Spec R$ for a finitely generated ring $R$ and let $z_0$ be a coordinate in the fibre of ~$L_0$. Then $\oplus_{k\in{\mathbb N}}H^0(kL_0)^G=R[z_0]^G=R^G[z_0]$, so taking Proj gives: $$X\gitquot 0 G =\Proj R^G[z_0]=\Spec R^G=X\gitquot{} G,$$ which is the usual affine GIT quotient. \end{example} \subsection{Stability and Extended $G$-equivalence} \label{sec:git:stab} The GIT quotient can be obtained by first restricting attention to the open set $\sst X(\zeta)\subseteq X$ of so called \emph{semi-stable} points. A point $x$ in $X$ is called \emph{semi-stable} (with respect to $\zeta$) if there exists a $G$-invariant section of $kL_\zeta$ (for some $k$ in ${\mathbb N}$) which is non-vanishing at~$x$. As a set, $X\gitquot\zeta G$ is the quotient of $\sst X(\zeta)$ by the \emph{extended $G$-equivalence} relation induced by the closure of the $G$-orbits: $$x\sim y \iff \overline{Gx}\cap\overline{Gy}\neq \emptyset.$$ Thus the $G$-invariant quotient map $\sst X(\zeta)\to X\gitquot\zeta G$ for the equivalence relation can map several $G$-orbits to the same point. For most of the points, this does not happen, however. This is because the closure of an open orbit is obtained by adding orbits of smaller dimension, so since the dimension of the orbit is a lower semi-continuous function on $X$, it follows that there is an open subset $\st X(\zeta)\subseteq\sst X(\zeta)$ of points which have full-dimensional closed $G$-orbits --- the so-called \emph{stable} points --- and there is a one-one correspondence between the orbits of $G$ in $\st X(\zeta)$ and their images in the GIT quotient. In other words, the GIT quotient contains, as an open set, the geometric quotient $\st X(\zeta)/G$. \subsection{Quotients for non-trivial linearisations} \label{sec:git:gen} Example~\ref{example:zeta_zero} showed that the GIT quotient for the trivial linearisation coincides with the affine GIT quotient. The following theorem uses the notion of stability to show that the quotients for non-trivial $\zeta$ are closely related to the affine quotient. \begin{thm} \label{thm:partial_res} The GIT quotients admit projective morphisms $$\rho_\zeta\colon X\gitquot\zeta G\to X\gitquot 0G$$ which are isomorphisms over $$\rho_\zeta^{-1}(\st X(0))\to\st X(0).$$ \end{thm} \begin{proof} Let $X=\Spec R$ for some finitely generated ring $R$, and let $z_\zeta$ be a complex coordinate in the fibre of~$L_\zeta$. The GIT quotient $X\gitquot\zeta G$ is given by taking Proj of the $G$-invariant part of $R[z_\zeta]$, where $R[z_\zeta]$ is to be considered as an algebra graded by the powers of~$z_\zeta$. The previous example showed that when $\zeta$ is zero, $\Proj R^G[z_0]=\Spec R^G$. For a general non-zero $\zeta$, $R[z_\zeta]^G\neq R^G[z_\zeta]$. However, the {\em degree-zero\/} part of $R[z_\zeta]^G$ is always~$R^G$, and this shows~\cite[Example~II.4.8.1 and Cor.~II.5.16]{hart:ag} that $X\gitquot\chi G$ is projective over $X\gitquot{}G=\Spec R^G$. Finally, the map $X\gitquot\zeta G\to X\gitquot 0G$ comes from the descent of the composition $\sst X(\zeta)\hookrightarrow \sst X(0) \to X\gitquot 0G$ and this is one-one whenever $\sst X(0) \to X\gitquot 0G$ is, so the last statement of the theorem follows. \end{proof} \begin{rmk} \label{rmk:partial_res} If $X$ contains a $0$-stable point, then $\st X(0)$ is open and non-empty, so dense in~$X$. Its image $\st X(0)\gitquot 0G$ is open and dense in $X\gitquot 0G$, and therefore, $\rho_\zeta\colon X\gitquot\zeta G\to X\gitquot 0G$ is an isomorphism on a dense open subset, i.e.\ $\rho_\zeta$ is birational. \end{rmk} \section{Moduli of Hermitian-Yang-Mills\ Connections} \label{sec:hym} This section recalls basic background concerning Hermitian-Yang-Mills\ connections and the construction of their moduli, following~\cite{don_kron:4mfds}. The construction of $X_\zeta$ will appear as a specialisation of the material in this section. However, the reader wishing to go straight to the point can skip this section and find a self-contained construction of $X_\zeta$ in section~\ref{sec:setup}. \subsection{Connections over Symplectic Manifolds} \label{sec:hym:symplectic} Suppose $X$ is a compact symplectic manifold $(X,\omega)$ of dimension $2n$ and let $E$ be a complex vector bundle over~$X$. The bundle of infinitesimal automorphisms of $E$ will be denoted $\gl(E)$ or~$\End E$. \subsubsection{Connections} Consider connections on $E$, namely linear maps $\nabla\colon \Omega^0_{X}(E) \to \Omega^1_{X}(E)$ which satisfy the Leibniz condition. Any connection on $ E$ can be expressed in a local neighbourhood\ $U\subset X$ as \begin{equation} d_\alpha= d+\alpha, \label{eq:dalpha} \end{equation} for $\alpha\in\Omega^1_U(\End E)$. On the other hand, if one considers the difference of two connections, one gets a {\em global\/} one-form with values in~$\End E$. \subsubsection{Hermitian Structure} Let $\bdh$ be a positive definite Hermitian inner product\ on the fibres of $E$ and denote by ${\mathfrak u}( E)\subset\gl( E)$ the real sub-bundle of unitary automorphisms determined by~$\bdh$. The connection $\nabla$ is said to be \emph{compatible} with the Hermitian structure if $$d \bdh(s,t) =\bdh(\nabla s,t)+\bdh(s,\nabla t).$$ Such a connection will have local one-form representatives $\alpha\in\Omega_U^1({\mathfrak u}( E))$; the space ${\mathcal A}$ of all such connections is an infinite-dimensional affine space modeled on~$\Omega_X^1({\mathfrak u}( E))$. \begin{rmk} Why does one fix a Hermitian structure on $E$? One reason is because fixing a Hermitian structure on $E$ amounts, by Chern-Weil theory, to fixing the topological invariants of the connections: for any Hermitian connection $d_\alpha$ on $E$, the Chern polynomials $c_1(E)$ and $c_2(E)-c_1(E)^2$ are represented respectively by $\frac{i}{2\pi}\trace F_\alpha$ and $\trace F_\alpha^2$. \end{rmk} \subsubsection{Symplectic Structure and Gauge Group} The space ${\mathcal A}$ has a symplectic structure defined by $$\boldsymbol{\omega}(a,b):=\int_X\trace(a\wedge b)\wedge\omega^{n-1},$$ for $a,b\in\Omega_X^1({\mathfrak u}( E))$ tangent vectors to~${\mathcal A}$. The \emph{gauge group} ${\mathcal G}$ is the group of automorphisms of $ E$ which respect the Hermitian structure in the fibres and cover the identity map of $X$. It acts on ${\mathcal A}$ by $$g\cdot\nabla:= g\nabla g^{-1}$$ (the condition that $g$ is unitary ensures that the new connection is compatible with $\bdh$) and preserves the symplectic form~$\boldsymbol{\omega}$. \subsubsection{Moment Map} The Lie algebra of the gauge group is~$\Lie {\mathcal G}=\Omega^0_X({\mathfrak u}( E))$. The moment map for the action of the gauge group is given by~\cite[Prop.\ 6.5.8]{don_kron:4mfds} \mapeq\mu{\mathcal A}{\Omega^0_X({\mathfrak u}( E))^*}\alpha{s\mapsto \int_X \trace s F_\alpha\wedge\omega^{n-1}}{eq:mucA} Note that the ${\mathcal G}$-equivariance follows because the curvature transforms as a tensor under gauge transformations. \subsection{Connections over K\"ahler\ manifolds, Holomorphic Structures and the Hermitian-Yang-Mills\ condition} \label{sec:hym:kahler} Now suppose that $X$ is in fact a K\"ahler\ manifold. \subsubsection{Complex Structure and the Decomposition of Curvature} There is a natural decomposition $\Omega^1_{X}(\End E) = (\Omega^{1,0}_{X}\oplus \Omega^{0,1}_{X})\otimes \Omega^0(\End E)$, and if a connection is expressed in a local holomorphic frame $\{z_i\}$ according to~\eqref{eq:dalpha} it takes the form $$\alpha = \sum_i \alpha_i dz_i - \alpha_i^* d \bar z_i,$$ for $\alpha_i$ smooth sections of~$\End E$. The local connection $d_\alpha$ splits into a sum of $(1,0)$ and $(0,1)$ parts $\partial_\alpha+\bar{\partial}_\alpha$ given by: \begin{align} \partial_\alpha &= \partial + \sum_i \alpha_i dz_i\\ \bar{\partial}_\alpha &= \bar{\partial} - \sum_i \alpha_i^* d{\bar{z}}_i. \end{align} The space ${\mathcal A}$ becomes a (flat) K\"ahler\ manifold when tangent vectors are identified with their $(0,1)$ parts, or in other words, when connections are represented by their $\bar{\partial}_\alpha$ operators. The holomorphic tangent space to ${\mathcal A}$ is of course isomorphic to $\Omega^{0,1}_{X}(\End E)$. The $(1,1)$ and $(2,0)$ parts of the curvature ${F_\alpha}$ of $d_\alpha$ are given by \begin{align}\label{eq:F11} F_\alpha^{1,1} &= \sum_{i, j} \left(\frac{\partial\alpha_j}{\partial{\bar{z}}_i} - \frac{\partial\alpha_j^*}{\partial z_i} -[\alpha_i,\alpha^*_j]\right)dz_i\wedge d{\bar{z}}_j,\\ \intertext{and} \label{eq:F20} F_\alpha^{2,0} &= \sum_{i,j} \left(-\frac{\partial\alpha_i}{\partial {z_j}} +\frac{1}{ 2}[\alpha_i,\alpha_j]\right)dz_i\wedge dz_j, \end{align} with $F_\alpha^{0,2}$ equal to minus the Hermitian adjoint of $F_\alpha^{2,0}$. \subsubsection{The Hermitian-Yang-Mills condition} A connection on $E$ is called \emph{Hermitian-Yang-Mills} (\emph{HYM}) if the inner product of its curvature with the K\"ahler\ form $\omega$ is a central element of $\Omega^1_X(\gu(E))$. This is in fact a moment map condition: using the identity $$F_\alpha\wedge\omega^{n-1}=\frac{1}{ n}\ip{F_\alpha}\omega \omega^n =: \frac{1}{ n}(\Lambda F_\alpha)\omega^n,$$ the map in equation~\eqref{eq:mucA} becomes $$\mu(\alpha)(s)=\frac{\Vol(X)}{(n-1)!} \trace(s\Lambda F_\alpha),$$ so, embedding the Lie algebra of ${\mathcal G}$ its dual in the usual way, we see that the moment map becomes a constant multiple of \map{\mu^*}{\mathcal A}{\Omega^0_X(\gu(E))}\alpha{\Lambda F_\alpha.} The moduli space of HYM connections is the K\"ahler\ quotient $\mu^{*-1}(0)/{\mathcal G}$. \subsubsection{Holomorphic Bundles} Suppose instead that the Hermitian connection $d_\alpha$ is required to induce a holomorphic structure on $E$. By the Newlander-Nirenberg theorem, prescribing such a structure is exactly equivalent to specifying a connection $d_\alpha$ which is \emph{integrable}, i.e.\xspace whose $(0,1)$-part is such that $\bar{\partial}_\alpha\circ \bar{\partial}_\alpha=0$. This is equivalent to the condition that $F_\alpha$ is of type $(1,1)$ and gives a K\"ahler\ subvariety ${{\mathcal A}^{1,1}}\subset{\mathcal A}$. This variety parametrises all the possible holomorphic structures which can be put on the $C^\infty$ bundle $E\to X$. The action of ${\mathcal G}$ on ${\mathcal A}$ extends to an action of its complexification ${\mathcal G}^{\mathbb C}$, which can be thought of naturally as the group of all general linear automorphisms of $E$ covering the identity map on $X$. Put $\tilde g:=(g^*)^{-1}$ and let \begin{align} g\cdot\bar{\partial}_\alpha &:= g\bar{\partial}_\alpha g^{-1},\\ g\cdot\partial_\alpha &:= \tilde g\partial_\alpha \tilde g^{-1}. \end{align} This action of ${\mathcal G}^{\mathbb C}$ preserves the space ${\mathcal A}^{1,1}$, and its orbits are equivalence classes of holomorphic bundles. To get a nice moduli space (a quasi-projective variety), one must restrict to the so-called semi-stable bundles, or in other words, consider a GIT quotient ${\mathcal A}^{1,1}\gitquot{}{\mathcal G}^{\mathbb C}$. A theorem of Uhlenbeck and Yau~\cite{uhl_yau:hym} states that the moduli space of stable bundles with the same topological type as $E\to X$ coincides with the moduli space of Hermitian-Yang-Mills connections on $E$. This is an infinite-dimensional version of the correspondence between symplectic and algebro-geometric quotients. \begin{rmk} One should really use the quotient of ${\mathcal G}^{\mathbb C}$ by the scalar automorphisms, since they act trivially on ${\mathcal A}$. The resulting group then has a trivial centre so there is only one linearisation of the action; it essentially determined by the degree and rank of $E$. \end{rmk} \section{Construction of $X_\zeta$} \label{sec:setup} A construction of $X_\zeta$ from scratch will be given in this section. Let $Q$ be an $n$-dimensional complex representation of a finite group \mbox{$\Gamma$}. Average over the group elements to get a positive definite Hermitian inner product\ on $Q$ such that $\Gamma\subset \U(Q)$. Let ${R}$ be the regular representation of \mbox{$\Gamma$}, i.e.\ the free \mbox{$\Gamma$}-module which is generated over ${\mathbb C}$ by a basis $\{e_\gamma | \gamma\in \Gamma\}$, and on which \mbox{$\Gamma$}\ acts via the morphism $\varphi\colon \Gamma\to\Aut_{\mathbb C} R$ defined by: $$\gamma\cdot e_\delta := \varphi(\gamma) e_\delta := e_{\gamma\delta}.$$ \subsection{Invariant HYM Connections} \label{sec:setup:hym} \begin{note} The reader who has not read Section~\ref{sec:hym} or is not interested in the ``moduli of bundles'' point of view can jump directly to~\ref{sec:setup:M}. \end{note} The construction of $X_\zeta$ is based on a variation on the construction in the previous section. It consists, roughly speaking, in applying the construction to the case where the compact K\"ahler\ manifold $X$ is replaced by the germ of the singularity $Q/\Gamma$. More precisely, start with $Q^*$, the dual vector-space to $Q$, and $E=Q^*\times R\to Q^*$ the trivial vector bundle with fibre $R$. Consider the connections on $E$ which are invariant under all translations in $Q^*$. These connections are determined by their value at one point, and they form a finite-dimensional vectorspace which can be identified with $M=Q\otimes_{\mathbb C} \End_{\mathbb C} R$ by choosing an isomorphism $\Omega^1_{Q^*}\cong Q$. The constructions in the previous section are now valid, because translation invariance eliminates any problems one might have with the non-compactness of base space $Q^*$. Most aspects are indeed a lot simpler: it suffices to set all the derivatives of the $\alpha_i$ equal to zero, and to ignore any integrals over the base space and all the formulas remain valid. An unusual feature of these invariant connections is that there are {\em several\/} moduli spaces: the usual Hermitian-Yang-Mills\ condition for a connection states that the contraction of the curvature with the K\"ahler\ form should be a central element of the Lie algebra of the gauge group. In case of general connections, there is only one gauge-invariant momentum level set because the centre of the gauge group ${\mathcal G}$ is trivial. In the case of invariant connections the gauge group that is relevant is a much smaller group $K^\Gamma$, consisting of gauge transformations which are invariant with respect to all translations and the action of $\Gamma$. This group consists of unitary endomorphisms of $R$ which commute with the action of $\Gamma$, and has a non-trivial centre (consisting of the traceless \mbox{$\Gamma$}-endomorphisms $\zeta\colon R\to R$). The non-triviality of the centre means that there are several gauge invariant momentum level sets, and hence several possible moduli $X_\zeta$. For the sake of the readers unfamiliar with the material in section~\ref{sec:hym}, the construction of $X_\zeta$ is given in detail without making any reference to the bundle construction. \subsection{The Vector Space $M^{\Gamma }$} \label{sec:setup:M} Let $M=Q\otimes_{\mathbb C} \End_{\mathbb C} R$ and let \mbox{$\Gamma$}\ act on $\End_{\mathbb C} R$ by conjugation via $\Gamma\stackrel{\varphi}\hookrightarrow \Aut R$: \begin{equation}\label{eq:gamma_action_endR} \gamma \cdot T := \varphi(\gamma)\,T\, \varphi(\gamma)^{-1} \end{equation} This makes $M$ into a \mbox{$\Gamma$}-module. Its \mbox{$\Gamma$}-invariant part is denoted ${M^\Gamma}$: \begin{equation}\label{eq:definition_M} M^\Gamma := \left( Q\otimes\End R \right)^\Gamma. \end{equation} The spaces $M^\Gamma$ and $M$ can be described explicitly by choosing a basis $\{q_l\}_{l=1}^n$ for $Q$, and defining the \emph{components} of $\alpha\in M$ by: \begin{equation}\label{eq:components_alpha} \alpha=\sum_{l=1}^n q_l\otimes \alpha_{l}. \end{equation} In this way the elements $\alpha\in M$ are identified with $n$-tuples of linear maps $\alpha_{i}:R\to R$. The elements of $M^\Gamma$ correspond to those $n$-tuples which satisfy the following equivariance condition \begin{equation}\label{eq:alpha_equivariance_condition} \sum_l \gamma_{kl} \alpha_{l} = \varphi(\gamma)\alpha_{k}\varphi(\gamma)^{-1},\rlap{$\qquad \forall k,\gamma$,} \end{equation} where $\boldsymbol{\gamma}=(\gamma_{kl})$ is the matrix corresponding to the action of the element $\gamma$ on $Q$ with respect to the basis $\{q_l\}_{l=1}^n$. \subsection{Symplectic and K\"ahler\ structure.} \label{sec:setup:symp} Endow $R$ with a fixed positive definite Hermitian inner product\ $\<\ ,\ >$ which makes the standard basis $e_\gamma$ orthonormal. The inner product on $R$ also defines a real structure on the space $\End_{\mathbb C} R$ of ${\mathbb C}$-linear endomorphisms of $R$ by the Hermitian adjoint operation in the usual way: $$\ip{T^*x}y := \ip{x}{Ty}. \qquad x,y\in R. $$ Define a positive definite Hermitian inner product\ ${\bdh}$ on $M$ by \mapeq{\bdh}{M\times M}{{\mathbb C}}{(\alpha,\beta)}{\sum_i \trace (\alpha_i\beta^*_i)}{eq:dfn_omega} The definition of $\bdh$ is independent of the basis of $Q$ up to a unitary transformations, and restricts to an inner product on $M^\Gamma$. As usual, $\bdh$ induces two forms on the underlying real vector-space to $M$: \begin{itemize} \item a non-degenerate symmetric bilinear form ${{\mathbf g}=\Re(\bdh)}$ called the \emph{Riemannian metric} associated to $\bdh$ \map{{\mathbf g}}{M\times M}{{\mathbb R}}{(\alpha,\beta)}{\frac{1}{2}\sum_i \trace (\alpha_i\beta^*_i+\beta_i\alpha^*_i)} \item a non-degenerate skew-symmetric bilinear form ${\omega=\Im(\bdh)}$ called the \emph{K\"ahler} form associated to $\bdh$ \map{\omega}{M\times M}{{\mathbb R}}{(\alpha,\beta)}{\frac{1}{2\sqrt{-1}} \sum_i \trace(\alpha_i\beta^*_i-\beta_i\alpha^*_i)} \end{itemize} This gives a Riemannian metric ${\mathbf g}$ and a K\"ahler form $\omega$ of type $(1,1)$ on $M$ and $M^\Gamma$, related as usual by \begin{equation} \label{eq:omega_g} \omega(\alpha, \beta )={\mathbf g}( \alpha , i\beta ). \end{equation} This makes $M^\Gamma, M$ and all their complex subvarieties into K\"ahler\ varieties. The group $\GL (R)$ of automorphisms of $R$ acts on $M$ by conjugation on $\End R$: \begin{equation} \alpha_i\mapsto g\alpha_i g^{-1},\qquad g\in\GL (R). \label{eq:action_glr} \end{equation} In fact, the scalars act trivially, and the action descends to an action of $G:=\PGL(R):=\GL(R)/\GL(1)$. The subgroup $\GL^\Gamma\! R$ of endomorphisms which commute with the action of $\Gamma$ acts on $M^\Gamma$ and, in the same way, there is a free action of $G^\Gamma=\PGL(R)^\Gamma$ on $M^\Gamma$. A maximal compact subgroup of $G$ (resp.\xspace $G^\Gamma$) is given by $K=\PU(R)$ (resp.\xspace $K^\Gamma:= \PU^\Gamma(R)$). The compact group $K$ (resp.\xspace $K^\Gamma$) leaves the K\"ahler\ structure on $M$ (resp.\xspace $M^\Gamma$) invariant. \subsection{The Variety $\protect\cN^{\Gamma }$ of Commuting Matrices} \label{sec:setup:cnga} Define the following natural map: \map{\psi}{Q\otimes\End R}{\Lambda^2 Q\otimes \End R}{\sum_k q_k \otimes \alpha_{k}}{\sum_{k,l} q_k\wedge q_l[\alpha_{k},\alpha_{l}]} This definition is independent of the basis of $Q$ (in terms of connections, it corresponds to calculating the $(0,2)$ part of the curvature). Denote the restriction of $\psi$ to $M^\Gamma$ by the same letter. Define \begin{equation}\label{eq:definition_cN} {\mathcal N} :=\psi{}^{-1}(0)\subset M; \end{equation} it is a cone (i.e.\ it is invariant under multiplication by non-zero scalars) which is an intersection of quadrics in $M$ given by the coordinate functions of $\psi$. In the representation of equation~ \eqref{eq:components_alpha}, its points consist of $n$-tuples of commuting $r\times r$ matrices: $${\mathcal N} = \{(\alpha_1,\dots,\alpha_n): \alpha_i\in \Mat_r({\mathbb C}), [\alpha_i,\alpha_j]=0.\}$$ Its $\Gamma$-invariant part ${{\mathcal N}^\Gamma}$ consists of those commuting matrices satisfying the equivariance condition~\eqref{eq:alpha_equivariance_condition}. \subsection{Moment Map} \label{sec:setup:moment} \nopagebreak Consider the vector space $M$ with the hermitian inner product $h$ and the action of $K$. The Lie algebra of $K$ is isomorphic to $\su R$, which consists of traceless skew-Hermitian endomorphisms of $R$. Using the invariant inner product $$\ip ab := \trace (a b^*)=-\trace ab,$$ identify $\su R$ with its dual in the usual way. Then the moment map is for the action of $K$ is \mapeq{\mu}M{\su R}{\alpha}{\sum_{k}[\alpha^*_{k},\alpha_{k}].}{eq:form_mu} (In the language of of connections, the map $\mu$ corresponds to contracting the curvature of the connection with the K\"ahler\ form $\omega=\sum_i dq_i\wedge d\bar{q_i}\in\Omega_{Q^*}^{1,1}$.) The moment map for the action of $K^\Gamma$ on $M^\Gamma$ is obtained simply by restriction. The K\"ahler\ quotients are defined by \begin{equation}\label{eq:kaehler_reductions} X_\zeta := \frac{\mu{}^{-1}(\zeta)\cap {\mathcal N}^\Gamma}{K^\Gamma}, \qquad \zeta\in \text{Centre}(\su^\Gamma R). \end{equation} As was remarked in section~\ref{sec:git}, to make this definition rigorous, one needs to make sense of the K\"ahler\ structure on $X_\zeta$. One way to do this is by restricting $\zeta$ to take on integral values. Then, by the correspondence between K\"ahler\ and GIT quotients, one has \begin{equation} X_\zeta \cong {\mathcal N}^\Gamma\gitquot\zeta G^\Gamma, \label{eq:git_xzeta} \end{equation} where $\zeta$ on the right-hand side specifies the linearisation of the action of $G^\Gamma$ on the trivial line bundle ${\mathcal N}^\Gamma\times{\mathbb C}$. \begin{rmk} In the case $\Gamma\subset SU(2)$, one has $\Lambda^2 Q\cong R_0$ --- the trivial representation. Identifying ${\mathbb C}^2$ with the quaternions, the vector space $M^\Gamma$ becomes a hyper-K\"ahler manifold with 3 distinct complex structures $I,J,K$ and corresponding associated K\"ahler forms $\omega_I, \omega_J, \omega_K$. The map $\psi$ is then a moment map for the complex symplectic form $\omega_{\mathbb C}= \omega_J+i\omega_K$ which is itself holomorphic with respect to $\omega_I$ and the quotients $X_\zeta$ are quotients of $M^\Gamma$ with respect to the hyper-K\"ahler\ moment map given by $(\mu,\psi)$, where the second (complex) variable is set to zero. Kronheimer~\cite{kron:thesis,kron:ale} exploits this fact to show that $X_\zeta$ are the minimal resolutions of ${\mathbb C}^2/\Gamma$ for generic values of $\zeta$. Furthermore, by varying the level set of $\psi$, he obtains universal deformations. If $\dim M>2$ however, $\psi^{-1}(\zeta)$ is not $G^\Gamma$-invariant for non-zero $\zeta$. \end{rmk} \begin{rmk} \label{rmk:GQaction} In fact, there is a further action on $M$ by $\GL Q$ (acting on $Q$ on the left). If $\rho\in\GL^\Gamma\! Q$, then it is easy to see that this preserves $M^\Gamma$, and indeed ${\mathcal N}^\Gamma$. Furthermore, the centre $Z(\Gamma)$ of $\Gamma$ is a subgroup of $\GL^\Gamma\! Q$ which acts trivially because of~\eqref{eq:alpha_equivariance_condition}, so there is an action of $G':=\GL^\Gamma\! Q/Z(\Gamma)$ which the quotients $X_\zeta $ inherit. The compact subgroup $K':=\U^\Gamma\! (Q)/Z(\Gamma)$ acts in a Hamiltonian fashion, and the moment map for this is given by \mapeq{\mu'}{X_\zeta}{(\Lie \U^{\Gamma} Q)^*}{[\alpha]}{b\mapsto \sum_{ij}b_{ij}\trace \alpha_i\alpha^*_j.}{eq:form_muQ} In general this does not provide one with very much information: for instance, if $Q$ is irreducible, $G'={\mathbb C}^*$ and $\mu'(\alpha)$ is the identity endomorphism of $Q$ times the sum of the norm squared of the $\alpha_i$'s. In the case where $\Gamma$ is abelian, however, $G'$ contains an algebraic torus of dimension $n$ acting freely. The components of $\mu'$ are the norm squared of the matrices $\alpha_i$. This is exploited in the companion paper~\cite{sacha:flows} to obtain a complete description of $X_\zeta$ by the methods of toric geometry. \end{rmk} \section{Variation of Quotients and Partial Resolutions} \label{sec:partial} The zero momentum quotient $X_0$ is better understood if viewed as a two-stage construction: first construct a ``universal quotient" ${\mathcal N}_0$ by ignoring the $\Gamma$-equivariance condition (i.e. perform the same construction with $\Gamma$ replaced by the trivial group) and then obtain $X_0$ as its \mbox{$\Gamma$}-invariant part. \subsection{The Universal Quotient} \label{sec:partial:uni} \nopagebreak Consider taking symplectic quotients of ${\mathcal N}$ with respect to the action of $K$. Since the centre of its Lie algebra is trivial, there is only {\em one\/} quotient, with momentum zero: $${\mathcal N}_0= \frac{ {\mathcal N}\cap \mu{}^{-1}(0)}{K}={\mathcal N}\gitquot{} G.$$ \begin{lemma} \label{lemma:N0} The reduction ${\mathcal N}_0$ is isomorphic to configuration space of $r=|\Gamma|$ points in $Q$: $${\mathcal N}_0 \cong \Sym^r(Q) := {Q^r/ \Sigma_r},$$ where $\Sigma_r$ denotes the permutation group on $r$ letters acting component-wise on the Cartesian product $Q^r$. \end{lemma} \begin{proof} The proof simply adapts Kronheimer's \cite[Lemma 5.2.1]{kron:thesis}. It is shown that the $K$-orbits in ${\mathcal N}\cap\mu^{-1}(0)$ can be identified in a one-one way with the $\Sigma_r$-orbits in $Q^r$. Let $\alpha\in{\mathcal N}\cap \mu{}^{-1}(0)$ have components $\alpha_i$ with respect to a basis $q_i$. The conditions $\psi(\alpha)=\mu(\alpha)=0$ give \begin{align*} [\alpha_i,\alpha_j] &=0,\qquad \text{for all }i,j\\ \sum_i [\alpha^*_i,\alpha_i] &=0. \end{align*} If one denotes by $A_i$ the operator $\adj(\alpha_i)$, one has, using the Jacobi identity and the above equations: \[ \sum_i A^*_i A_i(\alpha^*_j) = \sum_i [[\alpha^*_i,\alpha_i],\alpha^*_j]+ [[\alpha^*_j,\alpha^*_i],\alpha_i]= 0. \] The positivity of $A^*_iA_i$ implies $A^*_iA_i(\alpha^*_j)=0$ for all $i$ and $j$, and hence $A_j(\alpha^*_j)=[\alpha^*_j,\alpha_j]=0$. Thus ${\mathcal N}_0$ is the variety of $n$-tuples of normal commuting endomorphisms of $R$ modulo simultaneous conjugation. Any such $n$-tuple can simultaneously diagonalised by conjugation by a unitary matrix. This means that the orbit of an $n$-tuple is determined by the eigenvalues of its components; more precisely, there are orthonormal vectors $v_\gamma\in R$ indexed by the elements $\gamma\in\Gamma$ and corresponding eigenvalues $\lambda^i_\gamma\in{\mathbb C}$ such that: \begin{equation} \label{eq:alpha_eigenvectors} \alpha_i(v_\gamma)=\lambda^i_\gamma v_\gamma, \qquad \text{for all } i,\gamma. \end{equation} This gives $r$ elements $$\lambda_\gamma :=\sum_i \lambda^i_\gamma q_i \in Q,$$ which could be called the {\em eigenvalues\/} of $\alpha$. These are defined up to a permutation, because one can always conjugate by an elementary matrix which permutes the rows of the $\alpha_i$'s. In geometrical language, denote by~$\Delta\subset M$ the subspace of $n$-tuples of matrices which are diagonal with respect to the standard basis~$e_\gamma$ of~$R$. The unitary automorphism of $R$ which maps $e_\gamma$ to $v_\gamma$ moves $\alpha$ into $\Delta$. The slice $\Delta$ can be identified with $Q^r$ by mapping $\alpha\in \Delta$ to its $r$ eigenvalues (listed in some specified order). In this way $\Delta$ inherits an action of $\Sigma_r$, and the $\U(R)$-orbit of $\alpha\in\mu^{-1}(0)\cap {\mathcal N}$ intersects $\Delta$ in a single $\Sigma_r$-orbit. \end{proof} \subsection{The Zero-Momentum Quotient} \label{sec:partial:zero} \nopagebreak \begin{thm} \label{thm:X0free} If $\Gamma$ acts freely outside the origin, then $X_0\cong Q/\Gamma$ as varieties. \end{thm} \begin{proof} The proof consists in showing that the $K^\Gamma$-orbits in ${\mathcal N}^\Gamma\cap\mu^{-1}(0)$ can be identified in a one-one way with the $\Gamma$-orbits in $Q$. Let $\alpha\in{\mathcal N}^\Gamma\cap\mu^{-1}(0)$ and let $v_1$ be an eigenvector of $\alpha$ with eigenvalue $\lambda_1:=\sum_i\lambda^i_1 q_i\in Q$: $$\alpha_i(v_1)=\lambda^i_1 v_1,\quad i=1,\dots,n.$$ By equivariance of $\alpha$, $$\alpha_i (R(\gamma)v)=(Q(\gamma)\lambda_1)^i(R(\gamma)v), \text { for all } \gamma,\text{ and all }i,$$ so the eigenvectors and eigenvalues of $\alpha$ are given by $$\lambda_\gamma:=Q(\gamma)\lambda_1\text{ and }v_\gamma:=R(\gamma)v_1,\quad\text{for all }\gamma$$ i.e. they lie in orbits of $\Gamma$. Since $\Gamma$ acts freely outside the origin, the eigenvalues are either all zero or all distinct and non-zero. In the latter case, the eigenvectors therefore form a basis of $R$. The unitary automorphism of $R$ defined by $e_\gamma \mapsto v_\gamma$ commutes with the action of $\Gamma$, so defines an element of $K^\Gamma$. If $\Delta^\Gamma\subset M^\Gamma$ denotes the $n$-tuples of endomorphisms of $R$ which are diagonal with respect to the standard basis $\{e_\gamma\}$ then the automorphism carries $\alpha$ into $\Delta^\Gamma$. The map $\alpha\mapsto \sum_i\lambda^i_1 q_i$ identifies $\Delta^\Gamma$ with $Q$ in a manner that is compatible with the $\Gamma$-action on both sides. Furthermore, the $K^\Gamma$-orbit of $\alpha$ intersects $\Delta^\Gamma$ in precisely one $\Gamma$-orbit. Thus $X_0\cong \Delta^\Gamma/\Gamma\cong Q/\Gamma$. \end{proof} The following lemma will be useful in the section on ALE metrics (cf.~\cite{kron:thesis}). \begin{lemma} \label{lemma:iso_flat} If $\Gamma$ acts freely outside the origin the map $$\mu^{-1}(0)\cap{\mathcal N}^\Gamma/K^\Gamma\to \Delta^\Gamma/\Gamma$$ is an isometry when $\Delta^\Gamma$ is given the metric it inherits as a subspace of $M^\Gamma$, namely the Euclidean metric. Furthermore, the bundle $\mu^{-1}(0)\cap{\mathcal N}^\Gamma\to X_0$ is flat. \end{lemma} \begin{proof} The key point is that the subspace $\Delta^\Gamma$ is everywhere orthogonal to the orbits of $K^\Gamma$: a tangent vector to the orbits consists of an $n$-tuple of matrices of the form $[\xi,\alpha_i]$ for some $\xi\in\su^\Gamma(R)$, and these matrices are always zero on the diagonal, so orthogonal to $\Delta^\Gamma$. This shows that the bundle $\mu^{-1}(0)\cap{\mathcal N}^\Gamma \to X_0$ is flat, and the definition of the quotient metric on $X_0$ implies that the map $X_0=\mu^{-1}(0)\cap{\mathcal N}^\Gamma/K^\Gamma\to\Delta^\Gamma/\Gamma$ is an isometry. \end{proof} \subsubsection{Case when $\Gamma$ does not act freely outside the origin} Let $\alpha\in\cN^{\Gamma }\cap\mu^{-1}(0)$ have an eigenvalue $\lambda\in Q$. If the stabiliser $\Gamma_\lambda$ of $\lambda$ is trivial, then $\alpha$ has $r$ distinct eigenvalues, corresponding to the elements of the orbit $\Gamma\lambda$. This determines the components $\alpha_i$ completely on the whole of $R$. On the other hand, if $\lambda$ has a non-trivial stabiliser $\Gamma_\lambda$, then this determines $\alpha$ on the sub-representation $W_\lambda := \spn \Gamma\cdot E_\lambda\subset R$, where $E_\lambda$ is the eigenspace corresponding to $\lambda$. In fact, $E_\lambda$ is a representation of the stabiliser subgroup $\Gamma_\lambda$ and $W_\lambda$ is simply the representation of $\Gamma$ induced by $E_\lambda$: $$W_\lambda=\Ind_{\Gamma_\lambda}^\Gamma E_\lambda.$$ If $\dim E_\lambda<|\Gamma_\lambda|$, then $W_\lambda\neq R$ and $\alpha$ restricts to an endomorphism of $W^\perp_\lambda$. Let $\lambda'$ be an eigenvalue of the restriction; the equivariance condition then determines $\alpha$ on the factor $W_{\lambda'}$. Continuing in this way, one obtains a decomposition of $R$: $$R=W_\lambda\oplus W_{\lambda'}\oplus\dots.$$ From this discussion, one obtains the following description of the quotient $X_0$: \begin{thm} \label{thm:X0} There is an inclusion $Q/\Gamma \hookrightarrow X_0$; this inclusion is an isomorphism if and only if $\Gamma$ acts freely on $Q$ outside the origin. \end{thm} \begin{proof} The first statement follows because, for any orbit $\Gamma\lambda$ in $Q$, one can construct an $n$-tuple $\alpha$ of diagonal matrices whose $\lambda$-eigenspace has dimension equal to the stabiliser $\Gamma_\lambda$. The orbit of such an $\alpha$ under $K^\Gamma$ consists of commuting matrices with eigenvalue $\gamma\lambda$ with multiplicity $|\Gamma_\lambda|$, for all $\gamma\in\Gamma$. For the second statement, note that the {\em if\/} direction is theorem~\ref{thm:X0free}. The {\em only if\/} direction follows because if $\lambda$ is an eigenvalue of $\alpha$ with non-trivial stabiliser and with multiplicity one, one can set $\alpha$ to be zero on $W^\perp_\lambda$ (since $0$ is a fixed point of $\Gamma$, the equivariance condition~\eqref{eq:alpha_equivariance_condition} does not imply the existence of other eigenvalues). In general, $X_0$ corresponds configurations of $r=|\Gamma|$ points of $Q$ which are unions of orbits of $\Gamma$, and hence give rise to a decomposition of $R$ into induced representations $$R=\bigoplus_i \Ind_{\Gamma_{\lambda_i}}^\Gamma E_{\lambda_i},$$ where $E_{\lambda_i}$ denote the $\lambda$-eigenspace of an element $\alpha$. \end{proof} \begin{rmk} When $\Gamma$ doesn't act freely outside the origin, the quotient $X_0$ may end up containing all sorts of things. For instance, for the group action $\qsing 1/5(0,1,-1)$, the quotient $X_0$ contains a copy of ${\mathbb C}^3/\Z_5$, a copy of ${\mathbb C}^5$, eight copies of ${\mathbb C}^2$, etc... \end{rmk} \subsection{Non-zero Momentum and Partial Resolutions} \label{sec:partial:non-z} By theorem~\ref{sec:git}.\ref{thm:partial_res}, in the case where $\zeta$ is integral, there are projective morphisms $\rho_\zeta\colon X_\zeta\to X_0$ which are isomorphisms over the set of points which have finite $K^\Gamma$-stabilisers. \begin{prop} If $\Gamma$ acts freely outside the origin, The stabilisers of $K^\Gamma$ on ${\mathcal N}^\Gamma\cap\mu^{-1}(0)$ are trivial everywhere except at $\alpha=0$. \end{prop} \begin{proof} An automorphism $T$ of $R$ which fixes $\alpha\in {\mathcal N}^\Gamma\cap\mu^{-1}(0)$ must preserve the (simultaneous) eigenspaces of $\alpha$. If $T$ also commutes with the action of $\Gamma$, its action on an eigenvector $v\in R$ determines its action on the linear span of the $\Gamma$-orbit of $v$. In the case where $\Gamma$ acts freely and $\alpha$ is non-zero this means that $T$ is only allowed to multiply each eigenvector by the same non-zero constant --- and this constant must be of modulus one if $T$ is unitary. Such a $T$ thus corresponds to the identity element in the quotient group $K^\Gamma=\PU^\Gamma(R)$. \end{proof} Applying the theorem about K\"ahler\ quotients, one gets the following theorem: \begin{thm} \label{thm:partial_resolutions} If\/ $\Gamma$ acts on $Q$ freely outside the origin, and $\zeta$ is integral, there are projective morphisms $\rho_\zeta\colon X_\zeta\to X_0=Q/\Gamma$ which are isomorphisms outside the set $\rho_\zeta^{-1}(0)$. \end{thm} \begin{rmk} Even in the case that $\Gamma$ does not act freely outside the origin, it is likely that there are still birational maps from $X_\zeta$ to the component of $X_0$ which is isomorphic to $Q/\Gamma$ and which are isomorphisms outside the singular set. \end{rmk} \section{ALE Metrics} \label{sec:ale} The quotients $X_\zeta$ inherit a metric ${\mathbf g}_\zeta$ from the metric ${\mathbf g}$ on the ambient space $M^\Gamma$. This section shows that these are ALE metrics. A metric ${\mathbf g}$ on a real $m$-dimensional Riemannian manifold $X$ is called \emph{asymptotically locally Euclidean} (\emph{ALE}) if there exists a compact subset $C\subset X$ whose complement $X\setminus C$ has a finite covering $\widetilde{X\setminus C}$ which is diffeomorphic to the complement of a ball in ${\mathbb R}^m$, and such that, in the pulled-back coordinates $x_1,\dots,x_m$ on $\widetilde{X\setminus C}$, ${\mathbf g}$ takes the form \begin{equation} {\mathbf g}_{ij}=\delta_{ij} + a_{ij}, \label{eq:ale} \end{equation} where $|\partial^p a_{ij}|=O(r^{-4-p})$ for $p\geq 0$, where $r=\sqrt{\sum_i x_i^2}$ denotes the radial distance in ${\mathbb R}^m$ and $\partial$ denotes the differentiation with respect to the coordinates $x_1,\dots,x_m$. \begin{thm} \label{thm:ale} The metrics on $X_\zeta$ are ALE: for any $\zeta$, there is an expansion in powers of $r$ \begin{equation} {\mathbf g}_\zeta = \delta + \sum_{k\geq 2} h_k(\theta)r^{-2k}, \label{eq:expansion} \end{equation} where $(r,\theta)$ denote polar coordinates in ${\mathbb R}^{2n}\cong Q$. This expansion is analytic and may be differentiated term by term. \end{thm} \begin{proof} Kronheimer's proof~\cite[Prop.5.5.1]{kron:thesis} goes through with the appropriate modifications. The metric ${\mathbf g}_\zeta$ restricted to the unit ball $r=1$ is an analytic function of $\zeta$, so admits an expansion $${\mathbf g}_{\zeta| r=1} = \sum_\nu f_\nu \zeta^\nu$$ where $\nu$ are multi-indices in the coordinates of $\zeta$. The moment map being quadratic homogeneous implies that $${\mathbf g}_\zeta(r,\theta)={\mathbf g}_{r^{-2}\zeta}(1,\theta).$$ Hence the expansion for ${\mathbf g}_\zeta$ takes the form $$\sum_{k\geq 0} h_k(\theta)r^{-2k},$$ where the $h_k=\sum_{|\nu|=k}f_\nu\zeta^\nu$ are analytic functions of the radial coordinates. It remains to show that $h_0=\delta$ and that $h_1=0$. The first statement is equivalent to showing that the identification $X_0\to Q/\Gamma$ is an isometry. This was done in Lemma~\ref{lemma:iso_flat}. For the second statement, one must show that the variation of ${\mathbf g}_\zeta$ with $\zeta$ is zero at $\zeta=0$ in every direction $\lambda\in ((\Lie K^\Gamma)^*)^{K^\Gamma}$. The metric ${\mathbf g}_\zeta$ is determined entirely by the K\"ahler\ form $\omega_\zeta$ and the induced complex structure $J_\zeta$. Since the latter is the same for all $\zeta$, it is sufficient to prove that $$\partial_\lambda \omega_{\zeta|\zeta=0}=0$$ for all $\lambda$. A general formula for the variation of the induced symplectic form is given by Duistermaat and Heckman in~\cite{dui_hec:variation}. Away from the singularities, the projection $\mu^{-1}(\zeta)\cap{\mathcal N}^\Gamma\to X_\zeta$ is a principal $K^\Gamma$-bundle whose connection is given by the Levi-Civita connection for the induced metric on $X_\zeta$. If $\Omega_\zeta$ denotes the curvature, regarded locally as an element of $\Omega^2_{X_\zeta}\otimes\su^\Gamma(R)$, then the formula for the variation of $\omega_\zeta$ is given by $$\partial_\lambda \omega_\zeta = \<\lambda,\Omega_\zeta>.$$ In the present case, lemma~\ref{lemma:iso_flat} tells us that $\Omega_0=0$, so the variation is zero for $\zeta=0$, and this concludes the proof. \end{proof} \section{Deformation Complexes} \label{sec:defcplx} The question of the local geometry of the moduli spaces $X_\zeta$ can be studied using the tools of deformation complexes. \subsection{Differential Forms and Graded Lie Algebras} \label{sec:defcplx:diff} Define the vectorspaces $$M^{p,q} := \Omega^{p,q}_{Q^*}\otimes \End R,$$ for $p,q \in{\mathbb N}$, whose typical element $\beta$ is of the form $$\beta = \beta^I_{\bar J} dq_I\wedge d\bar q^{J},$$ where the summation convention is used for the multi-indices $I,J$ and where, as usual, $dq_I= \wedge_{i\in I} dq_i$, and $d\bar q^J=\wedge_{j\in J} d\bar q^j$. Define the \emph{degree} of $\beta$ to be $\deg \beta:=|I|+|J|$ and write $(-1)^\beta$ for $(-1)^{\deg\beta}$. The product of two elements $\alpha,\beta$ is defined to be $$\alpha\beta:=\alpha^I_{\bar J}\beta^{I'}_{\bar J'}dq_I\wedge d{\bar{q}}^{J}\wedge dq_{I'}\wedge d{\bar{q}}^{J'}, $$ and the bracket of any two elements $\alpha\in M^{p,q}$ and $\beta\in M^{p',q'}$ is defined by \begin{equation} [\alpha,\beta] := [\alpha^I_{\bar J},\beta^{I'}_{\bar J'}]dq_I\wedge d\bar q^{J}\wedge dq_{I'}\wedge d\bar q^{J'} = \alpha\beta-(-1)^{\alpha\beta}\beta\alpha. \label{eq:bracket} \end{equation} In the equation above and elsewhere, $(-1)^{\alpha\beta}$ means $(-1)^{\deg\alpha\deg\beta}$ and not $(-1)^{\deg\alpha}(-1)^{\deg\beta}$. Writing $M^r:=\sum_{p+q=r}M^{p,q}$, the algebra $M^*$ inherits the structure of a graded Lie algebra, namely a graded algebra with a bracket satisfying $$[M^r,M^s]\subset M^{r+s},$$ (graded) skew-commutativity $$[\alpha,\beta]=-(-1)^{\alpha\beta}[\beta,\alpha],$$ and the (graded) Jacobi identity: $$(-1)^{\alpha\gamma}[\alpha,[\beta,\gamma]]+(-1)^{\beta\alpha}[\beta,[\gamma,\alpha]]+(-1)^{\gamma\beta}[\gamma,[\alpha,\beta]]=0.$$ There are two sub-algebras $M^{*,0}$ and $M^{0,*}$ of $M^*$. Defining the adjoint of $\alpha=\alpha^I_{\bar J}dq_I\wedge d{\bar{q}}^{J}$ to be $$\alpha^* := (\alpha^I_{\bar J})^*d{\bar{q}}^I\wedge dq_{J},$$ then $(\alpha\beta)^* = (-1)^{\alpha\beta}\beta^*\alpha^*$ and $[\alpha,\beta]^*=(-1)^{\alpha\beta}[\beta^*,\alpha^*]$. The Jacobi identity implies that if $\alpha$ or $\beta$ has odd degree then \begin{equation} [\alpha,[\alpha,\beta]]=\frac{1}{2}[[\alpha,\alpha],\beta]. \label{eq:aab} \end{equation} If $\alpha=\alpha_i d{\bar{q}}^i \in M^{0,1}$, and one defines \map{\bar{\partial}_\alpha}{M^{p,q}}{M^{p,q+1}}{\beta}{[\alpha,\beta],} then, using~\eqref{eq:aab}, one sees that the sequences \begin{equation} \label{eq:Mpx_complex} M^{p,*}: M^{p,0}\xrightarrow{\bar{\partial}^0_\alpha}M^{p,1}\xrightarrow{\bar{\partial}^1_\alpha}M^{p,2}\to\dots \end{equation} of vectorspaces are complexes precisely when $[\alpha,\alpha]=0$, i.e.\ when $\alpha\in {\mathcal N}$. Write $H^{p,q}_\alpha$ for the cohomology groups $H^q(M^{p,*},\bar{\partial}_\alpha)$. If one introduces a metric on $M^{p,q}$ by using the standard inner product on $\End R$ and making $$\frac{1}{\sqrt{2^{p+q}}}\{dq_I\wedge d{\bar{q}}^{J}\}_{\substack{|I|=p,\\ |J|=q}}$$ orthonormal~\cite[p.80]{gri_har:ag}, one can define the adjoint operator $\bar{\partial}^*_\alpha$, and the Laplacian $\bar\Box_\alpha:=\bar{\partial}_\alpha\bar{\partial}^*_\alpha+\bar{\partial}^*_\alpha\bar{\partial}_\alpha$. Their kernels give harmonic representatives for the cohomology groups in the usual way $${\mathcal H}^{p,q}_\alpha:= \ker \bar\Box^{p,q}_\alpha \subset M^{p,q}.$$ If $\Lambda\colon M^{p,q}\to M^{p-1,q-1}$ denotes the operation of contraction with the K\"ahler\ form $\omega= dq_i\wedge d{\bar{q}}^i$, then the definition of the adjoint and the invariance of the trace under cyclic permutations give \begin{equation} \bar{\partial}^*_\alpha \beta= -\Lambda[\alpha^*,\beta], \label{eq:dstar} \end{equation} or, in coordinates, \begin{equation}(\bar{\partial}^*_\alpha \beta)^I_{\bar J} = [\alpha^*_j,\beta^I_{j\bar J}].\label{eq:dstar_coord} \end{equation} Writing $\kappa:= dq_1\wedge \dots \wedge dq_n$, the $n$-th power of $\omega$ is \begin{equation} \omega^n=n!(-1)^{(n-1)(n-2)/2}\frac{i^n}{2^n}\kappa\wedge\kappa^*. \label{eq:omegan} \end{equation} \subsection{Local Description of $X_\zeta$} \label{sec:defcplx:local} Under the identification $M=M^{0,1}$, the derivative of the action~\eqref{eq:action_glr} of $\GL(R)$ on $M$ is given by $$\partial_\alpha^0\colon M^{0,0}\to M^{0,1}.$$ On the other hand, the derivative of $\alpha\mapsto F^{0,2}_\alpha$ is (twice) $$\bar{\partial}^1_\alpha \colon M^{0,1}\to M^{0,2},$$ so the Zariski tangent space to ${\mathcal N}\gitquot{}\GL(R)$ at an element $\alpha$ is given by the first cohomology group of the Atiyah-Hitchin-Singer~\cite{ahs} deformation complex $$ M^{0,0}\xrightarrow{\bar{\partial}^0_\alpha}M^{0,1}\xrightarrow{\bar{\partial}^1_\alpha}M^{0,2},$$ i.e.\ by $H^{0,1}_\alpha$. From the point of view of the K\"ahler\ quotient, one can see this as follows: the Zariski tangent space to $X_\zeta$ at $[\alpha]$ is given by $\ker d\mu(\alpha)\cap\ker d\psi(\alpha)$. By definition, the derivative of $\mu$ is dual to the action of $K^\Gamma$, so $d\mu(\alpha)=-2\bar{\partial}^*_\alpha= 2\Lambda[\alpha^*,\ ]$, as can be verified by remarking that $\mu(\alpha)=\Lambda[\alpha^*,\alpha]$. Hence $\ker d\mu(\alpha)=\Image \bar{\partial}^*_\alpha$, so the tangent spaces indeed coincide. A local model for ${\mathcal N}\gitquot{}\GL(R)$ in a neighbourhood of a point $[\alpha]$ where $\alpha\in{\mathcal N}$ has trivial stabiliser is given by solving the equation $F^{0,2}_{a+\beta}=0$ for $\beta$ in the slice $$\{\beta\in M^{0,1}: \bar{\partial}_\alpha^*\beta=0,\|\beta\|\text{ small}\}.$$ This comes down to solving the system of equations \begin{align} &\bar{\partial}^*_\alpha \beta=0\\ &\bar{\partial}_\alpha \beta + \frac{1}{2}[\beta,\beta]=0, \end{align} in a neighbourhood of the origin. Kuranishi's argument~\cite{kura:complex,kura:newproof} shows that the solution set is given by the zero set of a map $\Phi\colon {\mathcal H}^{0,1}_\alpha\to {\mathcal H}^{0,2}_\alpha$ whose two-jet at the origin is given by \map{\Phi_{(2)}}{{\mathcal H}^{0,1}_\alpha}{{\mathcal H}^{0,2}_\alpha}{\beta}{{\mathcal H}_\alpha([\beta,\beta]).} where $${\mathcal H}_\alpha \colon M^{0,2} \to {\mathcal H}^{0,2}_\alpha$$ denotes the orthogonal projection to the harmonic subspace. Similar statements hold for $X_\zeta$ and $M^{0,*,\Gamma}$. \subsection{Kuranishi Germs and Formality} \label{sec:defcplx:kura} This whole discussion can be phrased in more abstract language of deformation functors and differential graded Lie algebras. Additional details and background can be found in~\cite{gold_mill:invariance,gold_mill:fundamental} and~\cite{dgms}. The algebra $M^{0,*}$ is actually a \emph{differential graded Lie algebra} (DGLA) when endowed with the differential $\bar{\partial}_\alpha$. The metric on $M^{p,q}$ makes it into an \emph{analytic DGLA}, namely a DGLA which possesses a norm compatible with its differential and bracket, and which induces what is essentially a Hodge decomposition of its graded pieces with finite dimensional topological summands ${\mathcal H}^i$ which are the analogues of the harmonic forms. When $\alpha$ has trivial stabiliser, $X_\zeta$ is locally analytically isomorphic in the neighbourhood of $[\alpha]$ to the \emph{Kuranishi germ} ${{\mathbf K}_M}$ associated to $M^{0,*,\Gamma}$. The results so far are stated in the following theorem. \begin{thm} \label{thm:formal} The sequence of vector spaces $(M^{0,*},\bar{\partial}_\alpha)$ is a complex (and therefore a differential graded Lie algebra) if and only if $\alpha\in{\mathcal N}$. Furthermore, if $\alpha\in{\mathcal N}^\Gamma$, the Zariski tangent space to $X_\zeta$ at $[\alpha]$ is isomorphic to the first cohomology group of its $\Gamma$-invariant part $$H^{0,1,\Gamma}_\alpha:=H^1(M^{0,*,\Gamma},\bar{\partial}_\alpha)$$ and if $\alpha$ has trivial $K^\Gamma$-stabilisers, then $X_\zeta$ is locally isomorphic to its Kuranishi germ $${\mathbf K}_{M^{0,*,\Gamma}}=\{\beta\in M^{0,1,\Gamma} | \bar{\partial}^*_\alpha\beta=\bar{\partial}_\alpha \beta + \frac{1}{2}[\beta,\beta]=0\}.$$ \end{thm} In general the Kuranishi germ ${\mathbf K}_L$ of an analytic DGLA $(L,d)$ is (Banach analytically) isomorphic to the germ at $0$ of $$\{\beta\in (\Image d)^\perp\subset L^1 | d\beta+\frac{1}{2}[\beta,\beta]=0\},$$ where $(\Image d)^\perp$ is a fixed complement of the image of $d$ in $L^1$. Goldman and~Millson~\cite{gold_mill:invariance} prove that ${\mathbf K}_L$ is invariant under \emph{quasi-isomorphisms}, namely chains of homomorphisms of DGLAs\footnote{An important point is that the intermediate $L',L'',\dots$ need {\em not\/} have any analytic structure and that the intermediate arrows need not preserve any splittings.} $$L\to L' \leftarrow L'' \to \dots \leftarrow L'''$$ which induce isomorphisms in cohomology. When $L$ is quasi-isomorphic to its cohomology (which is a DGLA when endowed with the zero differential), $L$ is called \emph{formal} and it follows that ${\mathbf K}_L$ is analytically isomorphic to the quadratic cone $${\mathbf Q}_L:=\{\beta\in{\mathcal H}^1 : [\beta,\beta]=0\}.$$ One way in which this can happen is if the bracket of two harmonic elements of degree $1$ is harmonic. This is the case, for instance for the moduli space of flat Hermitian-Yang-Mills connections over a compact K\"ahler\ manifold~\cite{nadel:quadratic,gold_mill:flat,gold_mill:invariance}. If in addition, the cup-product on ${\mathcal H}^1$ is zero, then ${\mathbf K}_L\cong {\mathcal H}^1$ and the deformation space is a smooth manifold (even if ${\mathcal H}^2\neq 0$). This is the case, for instance for the moduli space of complex structures over a \emph{Calabi-Yau $n$-fold}, namely a compact K\"ahler\ manifold with a nowhere vanishing holomorphic $(n,0)$-form~\cite{gold_mill:invariance}. These moduli were studied by F.~Bogomolov. The key fact which implies the formality of the DGLA and the vanishing of the cup-product in this case was proved by Tian~\cite{tian:smoothness} and Todorov~\cite{todo:weil-peterson}. In the case of the algebra $M^{0,*}$, formula~\eqref{eq:aab} with $\alpha$ and $\beta$ interchanged shows that the bracket of two harmonic elements in ${\mathcal H}^{0,1}_\alpha$ is $\bar{\partial}_\alpha$-closed. However, it does not follow that $\bar{\partial}^*_\alpha([\beta,\beta])=0$; indeed this is easily seen to be false, since $[\beta,\beta]=2\beta\beta$. Nevertheless, it does not seem unreasonable to expect that $M^{0,*,\Gamma}$ can also be proved to be formal for generic $\zeta$, maybe by imitating Tian and Todorov's method. \begin{conj} \label{conj:formal} The differential graded Lie algebra $(M^{0,*,\Gamma},\bar{\partial}_\alpha)$ is formal for all $\alpha\in{\mathcal N}^\Gamma\cap\mu^{-1}(\zeta)$ and generic $\zeta$, and therefore $X_\zeta$ has, for these $\zeta$, at worst quadratic algebraic singularities. \end{conj} Another conjecture is the following: \begin{conj} \label{conj:su3} If $\Gamma\subset\SU(3)$, can one imitate the Tian-Todorov proof and show that the Kuranishi germ of $(M^{0,*,\Gamma},\bar{\partial}_\alpha)$ is isomorphic to ${\mathcal H}^{0,1,\Gamma}_\alpha$ for generic $\zeta$, i.e.~that $X_\zeta$ is smooth? \end{conj} The fact that $X_\zeta$ has at most quadratic singularities has been verified for the abelian subgroups of order less than~11. The smoothness of $X_\zeta$ has been verified in the abelian cases $\qsing 1/3(1,1,1)$, $\qsing 1/6(1,2,3)$, $\qsing 1/7(1,2,4)$, $\qsing 1/8(1,2,5)$, $\qsing 1/9(1,2,6)$, $\qsing 1/10(1,2,7)$ and $\qsing 1/11(1,2,8)$. Both these verifications were done by exhaustive listing of singularities of $X_\zeta$ for all possible $\zeta$, using the methods given in the companion paper~\cite{sacha:flows}. A different approach is available in the specific case of $\SU(3)$; this is presented next. \section{Subgroups of $\protect\SU(3)$ and Cubic Forms} \label{sec:su3} Suppose that $\Gamma\subset\SU(3)$. If $\alpha\in\mu^{-1}(\zeta)$ and $\beta,\delta\in{\mathcal H}^{0,1,\Gamma}_\alpha$, then, as remarked in the previous section, $$\bar{\partial}_\alpha[\beta,\delta] =0,$$ but $[\beta,\delta]$ is not in ${\mathcal H}^{0,2,\Gamma}_\alpha$. However, considerations of type show that it differs from its harmonic projection by a term $\bar{\partial}_\alpha\epsilon$, for some $\epsilon\in M^{0,1,\Gamma}$. For $\eta\in{\mathcal H}^{0,1,\Gamma}_\alpha$ \begin{align} \trace(\eta[\beta,\delta])- \trace(\eta{\mathcal H}_\alpha([\beta,\delta])) &= \trace(\eta[\alpha,\epsilon]) \notag\\ &= \trace(\epsilon[\eta,\alpha])\notag\\ &= 0,\qquad \text{since } \eta\in{\mathcal H}^{0,1,\Gamma}_\alpha. \label{eq:trace_harmonic} \end{align} This shows that the tensor \corresp{H^{0,1,\Gamma}_\alpha\otimes H^{0,1,\Gamma}_\alpha\otimes H^{0,1,\Gamma}_\alpha}{{\mathbb C}}{(\eta,\beta,\delta)\phantom{{}^{,1,\Gamma}}}{\kappa\trace(\eta{\mathcal H}_\alpha([\beta,\delta])),} is totally symmetric on $H^{0,1,\Gamma}_\alpha$ (the isomorphism $\Omega^{3,3}_{Q^*}\cong{\mathbb C}$ has been used). An easy polarisation argument shows that it is completely determined by the corresponding cubic form \map{{\mathbf C}}{H^{0,1,\Gamma}_\alpha}{\mathbb C}\beta{\kappa\trace(\beta([\beta,\beta])).} \begin{prop} \label{prop:cubic} The singularity of $X_\zeta$ has no quadratic part if and only if ${\mathbf C}(\beta)=0$ for all $\beta\in{\mathcal H}^{0,1,\Gamma}_\alpha$ and all $\alpha\in\mu^{-1}(\zeta)$. \end{prop} \begin{proof} Suppose $X_\zeta$ has no quadratic part at $[\alpha]$. Then $\Phi_{(2)}(\beta)={\mathcal H}_\alpha([\beta,\beta])=0$. But this implies that ${\mathbf C}(\beta)=0$ by equation~\eqref{eq:trace_harmonic}. Conversely, if ${\mathbf C}(\beta)=0$ for all $\beta\in{\mathcal H}^{0,1,\Gamma}$ then the corresponding totally symmetric tensor vanishes on all triples $(\eta,\beta,\beta)$ for all $\eta,\beta\in {\mathcal H}^{0,1,\Gamma}_\alpha$. Since this is true for all $\eta$, it must be that ${\mathcal H}_\alpha([\beta,\beta])\in\Image\bar{\partial}_\alpha$, i.e.\ $\Phi_{(2)}(\beta)=0$ in ${\mathcal H}^{0,2,\Gamma}_\alpha$. \end{proof} There is a natural $3$-vector ${\Omega}$ whose value on three elements of $H^{0,1,\Gamma}_\alpha$ of is given by \begin{equation} \Omega(\eta,\beta,\delta):=\kappa\trace( \eta\beta\delta). \label{eq:Omega} \end{equation} This is symmetric under cyclic permutations of the entries, so decomposes into a totally skew-symmetric part $\Omega_{\text{skew}}$ and a totally symmetric part, which is nothing but the totally symmetric tensor corresponding to ${\mathbf C}$. The proposition above implies the \begin{cor} \label{cor:3form} If $X_\zeta$ is smooth, then $\Omega$ defines an element of $\Omega^{3,0}(X_\zeta)$. \end{cor} \begin{conj} The canonical sheaf ${\mathcal O}_{X_\zeta}(K_{X_\zeta})$ is locally free, and is generated by the non-vanishing $(3,0)$-form $\Omega$ when $X_\zeta$ is smooth. \end{conj} Taking the wedge of $\Omega$ with its complex conjugate gives \begin{align} \label{eq:owo1} \Omega\wedge\Omega^*(\eta,\beta,\delta,\eta^*,\beta^*,\delta^*) &= (\epsilon_{ijk}\trace \eta_i\beta_j\delta_k)\overline{(\epsilon_{{\bar{\imath}}{\bar{\jmath}}{\bar{k}}}\trace \eta_{\bar{\imath}}\beta_{\bar{\jmath}}\delta_{\bar{k}})}\kappa\wedge\kappa^*,\\ \label{eq:owo2} &=\left|\sum^\circ_{ijk}\trace \eta_i\beta_j\delta_k\right|^2\kappa\wedge\kappa^*, \end{align} where $\sum^\circ_{ijk}$ denotes the sum over distinct $i,j$ and $k$. On the other hand, the symplectic form $\omega_\zeta$ on $X_\zeta$ is simply the restriction of the symplectic form $\omega$ defined in~\eqref{eq:dfn_omega}, and equation~\eqref{eq:omegan} gives \begin{equation} \omega_\zeta\wedge\omega_\zeta\wedge\omega_\zeta = \frac{3i}{4}\kappa\wedge\kappa^*. \end{equation} Suppose that $(\eta,\beta,\delta)$ are an orthonormal triple in $T^{1,0}_\alpha X_\zeta$. Then the value of the coefficient of $\kappa\wedge\kappa^*$ in~\eqref{eq:owo2} is equal to $\|\Omega\|^2 \|\kappa\wedge\kappa^*\|^{-2}$. Hence $X_\zeta$ has trivial canonical bundle if this coefficient is never zero for all $\alpha\in\mu^{-1}(\zeta)$. \begin{lemma} \label{lemma:bochner} The K\"ahler\ manifold $X_\zeta$ is Ricci-flat if and only if the coefficient of $\kappa\wedge \kappa^*$ in~\eqref{eq:owo2} is constant for all orthonormal triples $(\eta,\beta,\delta)$ in $H^{0,1,\Gamma}_\alpha$ and all $\alpha\in\mu^{-1}(\zeta)\cap{\mathcal N}^\Gamma$. \end{lemma} \begin{proof} This follows because if $X_\zeta$ is Ricci-flat, there exists a holomorphic $(3,0)$-form $\Omega'$ which is covariant constant on $X_\zeta$. Hence $\Omega$ will differ from $\Omega'$ by a holomorphic function $f$. Now Liouville's theorem implies that $f$ is either constant or unbounded. Since $\Omega$ is clearly bounded (by $6\kappa\wedge\kappa^*$), $f$ must be constant. \end{proof} \subsection{Example} \label{sec:su3:ex} Let us work out a specific example. Consider the group $\Gamma=\mu_3$ of order $3$ acting on ${\mathbb C}^3$ with weights $(1,1,1)$. The following configuration of matrices is easily seen to define a point of $\mu^{-1}(\zeta)\cap{\mathcal N}^\Gamma$, where $\zeta=(-|A|^2,|A|^2-|B|^2,|B|^2)$, $(A,B\in{\mathbb R})$: \begin{equation} \label{eq:point1} \alpha_1=\begin{pmatrix} 0 &A &0 \\ 0 &0 &B \\ 0 &0 &0 \end{pmatrix}d{\bar{q}}^1, \quad \alpha_2=\alpha_3=0. \end{equation} The tangent space is three-dimensional and is generated by the following orthonormal elements (recall that $\|d{\bar{q}}^i\|^2=2$) \begin{equation} \beta_1=\frac{1}{\sqrt 2}\begin{pmatrix} 0 &0 &0 \\ 0 &0 &0 \\ 1 &0 &0 \end{pmatrix}d{\bar{q}}^1, \quad \beta_i= \frac{1}{\sqrt{2(A^2+B^2)}}\begin{pmatrix} 0 &A &0 \\ 0 &0 &B \\ 0 &0 &0 \end{pmatrix}d{\bar{q}}^i, \end{equation} for $i=2,3$, so this defines a smooth point of $X_\zeta$. The value of $\|\Omega\|^2$ at this point is \begin{equation} \left|\frac{1.A.B+1.B.A}{2\sqrt{2}(A^2+B^2)}\right|^2\kappa\wedge\kappa^* = \frac{1}{2}\left(\frac{AB}{A^2+B^2}\right)^2 \kappa\wedge\kappa^*, \label{eq:norm1} \end{equation} and so this is non-zero away from $AB=0$ (which correspond to non-generic values of $\zeta$). At the point \begin{equation}\label{eq:point2} \alpha_1=\begin{pmatrix} 0 &A+C &0 \\ 0 &0 &B+C \\ C &0 &0 \end{pmatrix}d{\bar{q}}^1, \quad \alpha_2= \alpha_3=0, \end{equation} in $\mu^{-1}(\zeta)\cap {\mathcal N}^\Gamma$, the tangent space is still three-dimensional, with orthonormal generators \begin{equation} \beta_i = \frac{1}{\sqrt{6}}\begin{pmatrix} 0 &1 &0 \\ 0 &0 &1 \\ 1 &0 &0 \end{pmatrix}d{\bar{q}}^i, \quad i=1,2,3, \end{equation} The value of $\|\Omega\|^2$ however is now \begin{equation} \left|6\left(\frac{1}{\sqrt 6}\right)^3\right|^2 \kappa\wedge\kappa^*=\frac{1}{6}\kappa\wedge\kappa^*. \label{eq:norm2} \end{equation} In fact, all the points of $\mu^{-1}(\zeta)\cap{\mathcal N}^\Gamma$ are of the form~ \eqref{eq:point1} or ~\eqref{eq:point2} (modulo permutations of the indices $1,2,3$)\footnote{See~\cite{sacha:flows} for an explanation of why this is so.}. Thus it has been shown, in a rather laborious way, that away from certain degenerate values of $\zeta$, $\Omega$ is non-vanishing on $X_\zeta$ and $K_{X_\zeta}$ is therefore trivial. In fact, $X_\zeta={\mathcal O}_{{\mathbb P}^2}(-3)$. Since the coefficient of $\kappa\wedge\kappa^*$ in~\eqref{eq:norm1} is always smaller than ${1/ 8}$, one also deduces that $\Omega\wedge\Omega^*$ is not a constant multiple of $\omega_\zeta\wedge\omega_\zeta\wedge\omega_\zeta$ on any of the quotients $X_\zeta$, and therefore by lemma~\ref{lemma:bochner} that the induced metric is never Ricci-flat. \begin{rmk} The space ${\mathcal O}_{{\mathbb P}^2}(-3)$ does have a standard Ricci-flat metric, as was first noted by Calabi~\cite{calabi}. \end{rmk} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1996-10-12T01:24:37
9610
alg-geom/9610013
en
https://arxiv.org/abs/alg-geom/9610013
[ "alg-geom", "math.AG" ]
alg-geom/9610013
Hans Boden
H. U. Boden and K. Yokogawa
Rationality of Moduli Spaces of Parabolic Bundles
latex2e
J. London Math. Soc. 59 (1999) 461-478
10.1112/S0024610799007061
null
null
The moduli space of parabolic bundles with fixed determinant over a smooth curve of genus greater than one is proved to be rational whenever one of the multiplicities associated to the quasi-parabolic structure is equal to one. It follows that if rank and degree are coprime, the moduli space of vector bundles is stably rational, and the bound obtained on the level is strong enough to conclude rationality in many cases.
[ { "version": "v1", "created": "Fri, 11 Oct 1996 23:16:19 GMT" } ]
2021-09-29T00:00:00
[ [ "Boden", "H. U.", "" ], [ "Yokogawa", "K.", "" ] ]
alg-geom
\section{Introduction} Let $X$ be a smooth complex curve of genus $g \geq 2$, $L$ a line bundle of degree $d$ over $X,$ and ${\cal M}_{r,L}$ the moduli space of semistable bundles $E$ of rank $r$ with determinant $L$. \begin{conj} \label{conj} ${\cal M}_{r,L}$ is {\it rational}, i.e. it is birational to a projective space. \end{conj} \noindent Despite many positive results \cite{newstead1}, this is still an open problem, even for $(r,d)=1.$ In this paper, we study a closely related problem, namely the birational classification of moduli spaces of parabolic bundles over $X.$ These moduli spaces occur naturally as \begin{enumerate} \item[(i)] unitary representation spaces of Fuchsian groups \cite{mehta-seshadri}, \item[(ii)] moduli spaces of Yang-Mills connections on X with an orbifold metric \cite{boden1}, and \item[(iii)] moduli spaces of certain semistable bundles over an elliptic surface \cite{bauer}. \end{enumerate} The theory developed in \cite{bh} and extended here shows that their birational type depends only on the {\em quasi-parabolic} structure (see Proposition \ref{prop:qp}). The methods of \cite{newstead1} then prove, in many cases, that these moduli spaces are rational. The weaker result, Theorem \ref{thm:rat1}, uses only Newstead's theorem, while the stronger one, Theorem \ref{thm:rat2}, requires an adaptation of his inductive argument. Using the theory developed in \S 4, it then follows from Theorem \ref{thm:rat2} that ${\cal M}_{r,L} \times {\Bbb P}^{r-1}$ is rational if $(r,d)=1$ (see Corollary \ref{cor}). Stable rationality of the moduli spaces had been proved in this case by Ballico \cite{ballico}, and our result is a strengthening of his. For instance, a consequence is that one can conclude Conjecture \ref{conj} under the assumption that $(r,d)=1$ for most values of the genus\footnote{Choosing $d' \equiv d \mod(r)$ with $0<d'<r,$ the hypothesis is that either $(d',g)=1$ or $(r-d',g)=1.$} (see Corollary \ref{cor:rat}). A number of useful facts are established along the way. One key point is Proposition \ref{prop:fine}, which gives a simple criterion for the existence of a universal bundle of stable parabolic bundles. We also extend the theory developed in \cite{bh} in several important ways (Theorems \ref{thm:bh1}, \ref{thm:bh2}, and \ref{thm:bh3}); the first two are standard but necessary for our purposes and the third is completely new. Its proof requires the idea of shifting a parabolic sheaf (Definition \ref{defn:shift}), which also provides a framework for the Hecke correspondence (equation (\ref{eqn:hecke})). All of these results play a crucial role in the proofs of Theorems \ref{thm:rat1} and \ref{thm:rat2}. A brief word about the organization of this paper: \S 2 introduces the notation used in the following sections, \S 3 discusses the existence of universal families, \S 4 summarizes and extends the theory of \cite{bh}, \S 5 describes shifting and the Hecke correspondence, and \S 6 contains the proofs of the main results and their corollaries. Before we begin, we would like to acknowledge a certain debt to the work of Newstead, upon which a number of our arguments depend, and without which this paper would be inconceivable. \section{Notation} Let $X$ be a smooth curve of genus $g \geq 2$ and $D$ a reduced divisor in $X.$ If $E$ is a ${\Bbb C}^r$ bundle over $X,$ then a {\it parabolic structure} on $E$ with respect to $D$ is just a collection of weighted flags in the fibers of $E$ over each $p \in D$ of the form \begin{eqnarray} \label{defn:parbun1} &E_p = F_1(p) \supset F_2(p) \supset \cdots \supset F_{\kappa_p}(p) \supset 0,&\\ &\, \;\; 0 \leq \alpha_1(p) < \alpha_2(p)< \cdots < \alpha_{\kappa_p}(p) < 1.& \label{defn:parbun2} \end{eqnarray} Holomorphic bundles $E$ with parabolic structures are called {\it parabolic bundles}, and we use the notation $E_*$ to indicate the bundle (or, equivalently, locally-free sheaf) $E$ together with a choice of parabolic structure. A morphism $\phi:E_* \longrightarrow E'_*$ of parabolic bundles is a bundle map satisfying $\phi(F_i(p)) \subset F'_{j+1}(p)$ whenever $\alpha_i(p) > \alpha'_j(p)$ for all $p \in D.$ We use the tensor product notation $H^0(E_*^\vee \otimes E'_*)$ for these morphisms, where $E^\vee_*$ denotes the dual parabolic bundle (cf. \cite{yoko}). A {\it quasi-parabolic} structure on $E$ is what is left after the weights are forgotten, it is determined topologically by its flag type $m,$ which specifies {\it multiplicities} $m(p) = (m_1(p), \ldots, m_{\kappa_p}(p))$ for each $p \in D$ defined by $m_i(p) = \dim F_i(p) - \dim F_{i+1}(p).$ A subbundle $E'$ inherits a parabolic structure from one on $E$ in a canonical way: The flag in $E'_p$ is gotten by intersecting with the flag in $E_p$ and the weights are determined by choosing maximal weights such that the inclusion map from $E'$ to $E$ is parabolic (p.\ 213, \cite{mehta-seshadri}). Parabolic structures on quotients have a similar description (loc.\ cit.). A parabolic bundle $E_*$ is called {\it stable} if every proper holomorphic subbundle $E'$ satisfies $\mu (E'_*) < \mu (E_*),$ where $$\mu (E_*) = \operatorname{pardeg} E_* / r = \deg E / r + \sum_{p \in D} \sum_{i=1}^{\kappa_p} m_i(p) \alpha_i (p)/r.$$ The parabolic bundle $E_*$ is called {\it semistable} if $\mu (E'_*) \leq \mu (E_*)$ for each subbundle $E'_*.$ The construction of the moduli space ${\cal M}_\alpha$ of semistable parabolic bundles, as a normal, projective variety, is given in \cite{mehta-seshadri}. The subspace ${\cal M}^s_\alpha$ of stable bundles is smooth and Zariski-open, in particular, if every semistable bundle is stable, then ${\cal M}_\alpha$ is smooth. Let $\Delta^r = \{ (a_1,\ldots,a_r) \mid 0 \leq a_1 \leq \cdots \leq a_r <1\}$ and define $W = \{ \alpha:D\longrightarrow \Delta^r \}.$ Points in $W$ determine both the weights and the multiplicities. Conversely, given a weight $\alpha$ in the sense of (\ref{defn:parbun2}), the associated point in $W$ is gotten by repeating each $\alpha_i(p)$ according to its multiplicity $m_i(p)$. We abuse notation slightly by referring to points in $W$ as weights. This gives an obvious notion of when a weight is {\it compatible} with a choice of multiplicities, and for a given $m,$ we define the open face of weights compatible with $m$ to be $$V_m = \{ \alpha \in W \mid \alpha_{i-1}(p)=\alpha_i(p) \Leftrightarrow \sum_{k=1}^j m_k(p) < i \leq \sum_{k=1}^{j+1} m_k(p) \}.$$ A weight in the interior of $W$ specifies full flags at each $p \in D.$ For every other choice of $m$, $V_m$ is contained in the boundary of $W.$ Now $W$ is a simplicial set, and the face relations give a natural ordering on $\{V_m\}$ and we write $V_m > V_{m'}$ if $V_{m'}$ is a proper face contained in the closure of $V_m.$ This agrees with the natural ordering on $m$ gotten by successive refinement. Weights for which ${\cal M}_\alpha$ is not necessarily smooth satisfy $\mu(E'_*) = \mu(E_*)$ for some proper subbundle $E'.$ Letting $E''$ be the quotient, then the short exact sequence of parabolic bundles $E'_* \stackrel{\iota}{\longrightarrow} E_* \stackrel{\pi}{\longrightarrow} E''_*$ determines a partition of $(d,r,m)$ in the obvious way: $(d',d''), (r',r'')$ and $(m',m'')$ are the degrees, ranks, and multiplicities of $(E',E'').$ (We define $m'$ and $m''$ here slightly unconventionally, namely \begin{eqnarray*} m'_i(p) &=& \dim (F_i(p) \cap \iota (E'_p)) - \dim (F_{i+1}(p) \cap \iota (E'_p)),\\ m''_i(p) &=& \dim (\pi (F_i(p)) \cap E''_p) - \dim (\pi (F_{i+1}(p)) \cap E''_p), \end{eqnarray*} for $p \in D$ and $1 \leq i \leq \kappa_p$.) Notice that $r',r'' > 0$ and $m'_i(p), m''_i(p) \geq 0.$ Write $\xi = (d',r',m').$ For fixed $\xi,$ the set of weights compatible with $m$ for which $\mu(E'_*) = \mu(E_*)$ is the intersection of a hyperplane $H_\xi$ in $W$ with $V_m$ given by the equation \begin{equation}\label{eqn:hyper} \sum_{i=1}^{\kappa_p} m_i(p) \alpha_i(p)) \sum_{p\in D} \sum_{i=1}^{\kappa_p} (r' m_i(p)-r m'_i(p)) \alpha_i(p) = r d' - r'd. \end{equation} There are only finitely many hyperplanes; the above equation puts a bound on $d'$ and all other quantities are already bounded. We shall refer to $H_\xi \cap V_m$ as a {\it wall} in $V_m.$ These walls induce a chamber structure on $V_m,$ a {\it chamber} being a connected component of $V_m \setminus \cup_\xi H_\xi$ (it is possible that $V_m \subset H_\xi$). Weights $\alpha \in W \setminus \cup H_\xi$ are called {\it generic}, and for these weights, ${\cal M}_\alpha = {\cal M}^s_\alpha.$ In the next section, we shall see that $V_m$ contains a generic weight if and only if the degree $d$ and the set of multiplicities $\{m_i(p)\}$ have greatest common divisor equal to one. \section{Families of parabolic bundles} In this section, we present Proposition \ref{prop:fine}, which establishes the existence of a universal family of stable parabolic bundles parametrized by ${\cal M}_\alpha^s$ whenever $V_m$ contains a generic weight. Although results of this type are well-known to experts, the proposition, as well as the proof, are original (cf. Th\'eor\`eme 32, \cite{seshadri}). It is important because, in the case of ordinary bundles, the non-existence of the universal family (\cite{ramanan}) is the obstruction to proving Corollary \ref{conj} by induction, and as shown in \S 6, the analogous argument works for parabolic bundles precisely because the necessary conditions for the vanishing of this obstruction given by Proposition \ref{prop:fine} are often satisfied. Given positive integers $m_1,\ldots,m_\kappa$ such that $m_1+\cdots m_\kappa = r,$ define ${\cal F}_m$ to be the variety of flags of type $m.$ These are simply flags ${\Bbb C}^r=F_1 \supset \cdots \supset F_s \supset 0$ with $\dim F_i -\dim F_{i+1} = m_i.$ Furthermore, for any bundle $E \longrightarrow S$ of rank $r,$ let ${\cal F}_m(E) \longrightarrow S$ be the bundle of flags of type $m.$ Given a bundle $U \rightarrow S \times X,$ we adopt the notation $U_s = U|_{\{s\} \times X}.$ We also use $\pi_S$ for the projection map $S\times X \rightarrow S.$ \begin{defn} \label{defn:family} Fix multiplicities $m(p)$ for each $p \in D.$ \begin{enumerate} \item[(i)] A family of quasi-parabolic bundles (of type $m$) parametrized by a variety $S$ is a bundle $U$ over $S \times X$ together with a section $\phi_p$ of the flag bundle ${\cal F}_{m(p)}(U|_{S \times \{p\}}) \longrightarrow S$ for each $p \in D.$ \item[(ii)] Two families $(U,\phi)$ and $(U',\phi')$ parametrized by $S$ are equivalent, written $(U,\phi) \sim (U',\phi'),$ if there exists a line bundle $L$ over $S$ and an isomorphism $U \cong U' \otimes \pi_S^* L$ under which $\phi \mapsto \phi'.$ \end{enumerate} \end{defn} Note that the section $\phi_p$ in (i) above is just a choice of a nested chain of subbundles of $U|_{S \times \{p\}}$ whose relative coranks are given by the multiplicities $m(p).$ A family of parabolic bundles is gotten by associating a fixed set of weights to each chain of subbundles. Let $U_*=(U,\phi,\alpha)$ be the resulting family of parabolic bundles and $U_{s,*}=(U_s, \phi(s),\alpha)$ be the parabolic bundle above $s \in S.$ Then $U_*$ is called a family of (semi)stable parabolic bundles if $U_{s,*}$ is (semi)stable for each $s \in S.$ It follows from the construction of Mehta and Seshadri that ${\cal M}_\alpha$ is a coarse moduli space. Proposition 1.8 of \cite{newstead2} then gives two conditions which are necessary and sufficient for a coarse moduli space to be fine, i.e. to admit a universal family. The second condition is not difficult to verify using an argument similar to that given in Lemma 5.10 of \cite{newstead2}. The first condition requires that we construct a family ${\cal U}^\alpha_*$ parametrized by ${\cal M}^s_\alpha$ with the property that ${\cal U}^\alpha_{e,*}$ is a parabolic stable bundle isomorphic to $E_*$ for all $[E_*] = e \in {\cal M}^s_\alpha.$ To construct this family, we need to review the construction of ${\cal M}_\alpha$ (\cite{bhosle}, \cite{mehta-seshadri}). Let $Q$ be the Hilbert scheme of coherent sheaves over $X$ which are quotients of ${\cal O}_X^{\oplus N}$ with fixed Hilbert polynomial (that of $E(k)$ for $k \gg g$), where $N= h^0(E).$ Let $U$ be the universal family on $Q \times X.$ Define $R$ to be the subscheme of $Q$ of points $r \in Q$ so that $U_r$ is a locally free sheaf which is generated by its global sections and $h^1(U_r)=0.$ Let $\widetilde{R}$ be the total space of the universal flag bundle over $R$ with flag type $\prod_{p\in D} {\cal F}_{m(p)},$ and let $\widetilde{U}$ be the pullback of $U$ to $\widetilde{R}.$ Then $\widetilde{U}$ is canonically a family of parabolic bundles parametrized by $\widetilde{R}$ by letting, for each $p \in D,$ $\phi_p$ be the tautological section and $\alpha(p)$ be the fixed weights. It follows that $\widetilde{R}$ has the local universal property for parabolic bundles (p.\ 16, \cite{bhosle}). The subsets $\widetilde{R}^{s}$ ($\widetilde{R}^{ss}$) corresponding to the stable (semistable) parabolic bundles are invariant under the natural action of $\operatorname{GL}(N) = \operatorname{Aut}({\cal O}_X^{\oplus N}),$ and ${\cal M}_\alpha$ is a good quotient of $\widetilde{R}^{ss}$ (with linearization induced by the weights $\alpha$), and ${\cal M}^s_\alpha$ is the geometric quotient of $\widetilde{R}^{s}.$ The center of $\operatorname{GL}(N)$ acts trivially on $R$ and $\widetilde{R},$ but nontrivially on the locally universal bundle $\widetilde{U}$. In fact, $\lambda(\hbox{id})$ acts on $\widetilde{U}$ by scalar multiplication by $\lambda$ in the fibers (this follows from p.\ 138, \cite{newstead2}). Given a line bundle $L$ over $\widetilde{R}^{s}$ with a natural lift of the $\operatorname{GL}(N)$ action such that $\lambda(\hbox{id})$ acts by multiplication by $\lambda,$ then using $\widetilde{U}^{s}$ to denote $\widetilde{U}|_{\widetilde{R}^{s} \times X},$ the quotient of $\widetilde{U}^{s} \otimes \pi^*_{\widetilde{R}^{s}} L^{-1},$ together with the tautological sections and weights $\{\phi_p, \alpha(p) \mid p \in D \}$ mentioned above, gives the desired family. \begin{prop} \label{prop:fine} Such a line bundle $L$ exists if either \begin{enumerate} \item[(i)] the elements of the set $\{d, m_i(p) \mid p \in D, 1 \leq i \leq \kappa_p \}$ have greatest common divisor equal to one, or \item[(ii)] the face $V_m$ containing $\alpha$ contains a generic weight. \end{enumerate} Moreover, these two conditions are equivalent, and when they are satisfied, the moduli space ${\cal M}^s_\alpha$ is fine. \end{prop} The idea of the proof is to find line bundles $L_k$ for each $k \in \{d, m_i(p) \}$ over $\widetilde{R}^{s}$ with natural actions of $\operatorname{GL}(N)$ such that $\lambda(\hbox{id})$ acts by scalar multiplication by $\lambda^k.$ Then (i) gives the existence of $k_1, \ldots, k_\ell \in \{d, m_i(p) \}$ and integers $a_1, \ldots, a_\ell$ so that $a_1 k_1 + \cdots a_\ell k_\ell = 1.$ The required line bundle is then the tensor product $L = L^{a_1}_{k_1} \otimes \cdots \otimes L^{a_\ell}_{k_\ell}.$ At the end of the proof, we will show that (i) and (ii) are equivalent. We start with a lemma. \begin{lem} Suppose $E_*$ is parabolic semistable or degree $d$ and rank $r$ and $H_*$ is a parabolic line bundle of degree $h$, then \begin{equation} \label{lem:ineq} h^1(H^\vee_* \otimes E_*) \neq 0 \quad \Rightarrow \quad d \leq r(2g-2+h) + r^2 n. \end{equation} \end{lem} \begin{pf} Serre duality for parabolic bundles (Proposition 3.7 of \cite{yoko}) implies that $$h^1(H^\vee_* \otimes E_*) \leq h^0(E^\vee_* \otimes {H}_* \otimes K(D)).$$ (If we had used $h^0(E^\vee_* \otimes \widehat{H}_* \otimes K(D)),$ the circumflex over $H_*$ indicating {\it strongly} parabolic morphisms, we would get the usual statement of Serre duality with equality, cf. \cite{yoko,by}.) Suppose that $\phi : E \longrightarrow H \otimes K(D)$ is a non-zero map and let $E'$ be the subbundle generated by $\operatorname{Ker} \phi.$ Then $$\deg E' \geq \deg E - \deg H\otimes K(D) = d - h - (2g -2 +n).$$ Considering $E'_*$ with its canonical parabolic structure as a subbundle of rank $r-1,$ the inequality (\ref{lem:ineq}) follows easily from this, semistability of $E_*,$ and the inequalities $\operatorname{pardeg} E'_* \geq \deg E'$ and $\operatorname{pardeg} E_* \geq \deg E + rn.$ \end{pf} \noindent {\it Proof of Proposition.} Write the weights $\alpha$ without repetition. Choose $\ell: D \longrightarrow {\Bbb Z}$ with $ 1 \leq \ell_p \leq \kappa_p+1$ and set $\beta(p) = \alpha_{\ell_p}(p).$ (Take $\beta(p) > \alpha_{\kappa_p}$ if $\ell_p = \kappa_p+1.$) For $h \in {\Bbb Z},$ define $$\chi(\ell,h) = d + r(1-g-h) - \sum_{p \in D} \sum_{i=1}^{\ell_p-1} m_i(p).$$ Let $H_*$ be the parabolic line bundle with $\deg H = h < d/r - r n -(2g-2)$ and with weights $\beta(p)$ at $p \in D.$ It follows from the lemma that if $E_*$ is semistable, then $h^1(H^\vee_* \otimes E_*)=0.$ Thus $h^0(H^\vee_* \otimes E_*) = \chi(\ell,h)$ by Riemann-Roch. Hence $(R^0 \pi_{\widetilde{R}^{s}}) (\widetilde{U}^{s} \otimes \pi_X^* H_*)$ is a locally free sheaf of rank $\chi(\ell,h)$ over $\widetilde{R}^{s}.$ Let $L(\ell,h)$ be the determinant of the corresponding bundle. By construction, the $\operatorname{GL}(N)$ action on $\widetilde{U}$ induces one on this bundle (and hence on $L(\ell,h)$); $\lambda(\hbox{id})$ acts by scalar multiplication by $\lambda$ on the bundle and by $\lambda^{\chi(\ell,h)}$ on $L(\ell,h).$ It is now a simple exercise in high school algebra to see that we can choose $h,h'$ and $\ell,\ell'$ so that $\lambda(\hbox{id})$ acts on $L(\ell,h) \otimes L(\ell',h')$ by $\lambda^k$ for any $k \in \{ d, m_i(p) \}.$ This proves the conclusion of the proposition assuming (i), and now we show that conditions (i) and (ii) are equivalent. Suppose first that (i) does not hold. Consider $E_*$ as a quasi-parabolic bundle without holomorphic structure, which will be specified later. Since the set $\{ d, m_i(p)\}$ is not relatively prime, there exists a prime number $q$ evenly dividing each element of the set. Clearly $q$ also divides $r.$ Set $d' =d/q, r' = r/q$ and $m'_i(p) =m_i(p)/q.$ Consider the quasi-parabolic bundle $E'_*$ with degree $d'$, rank $r',$ and multiplicities $m'.$ Any choice of weights $\alpha$ on $E_*$ induces (the same!) weights on $E'_*,$ and it follows that since $g \geq 2,$ there is some holomorphic structure for which $E'_*$ is semistable. Define the holomorphic structure on $E_*$ by $$E_* = E'_* \oplus \stackrel{q}{\cdots} \oplus E'_*.$$ It follows that $E_*$ is semistable but not stable for {\it any} choice of compatible weights. This implies that $V_m$ does not contain a generic weight. Suppose conversely that $V_m$ does not contain a generic weight. Since $V_m$ is affine, $V_m \subset H_\xi$ for some $\xi=(r',d',m')$ Using (\ref{eqn:hyper}), we conclude that for all $\alpha \in V_m,$ $$ \sum_{p \in D} \sum_{i=1}^{\kappa_p} (r m_i'(p) - r' m_i(p)) \alpha_i(p) = rd'-r'd .$$ (Here, we are still thinking of $\alpha$ without repetition.) We can vary each $\alpha_i(p)$ continuously by some small amount, and it follows that $$rm'_i(p) - r' m_i(p)=0= rd'-r'd$$ for all $i$ and $p.$ Since $r' < r,$ there exists a prime $q$ such that $q^k$ divides $r$ but not $r'.$ Hence $q$ divides $d$ and each element of the set $\{m_i(p) \mid p \in D, 1 \leq i \leq \kappa_p\}.$ $\quad \Box$ \section{The variation and degeneration theorems} In this section, we describe and extend the theory of \cite{bh}. This allows us to compare the moduli spaces of parabolic bundles ${\cal M}_\alpha$ and ${\cal M}_\beta$ when \begin{enumerate} \item[(i)] $\alpha, \beta \in V_m$ are generic weights in adjacent chambers, \item[(ii)] $\alpha \in V_\ell$ and $\beta \in V_m$ are generic weights not separated by any hyperplanes and $V_\ell > V_m.$ \end{enumerate} Cases (i) and (ii) correspond to Theorem 3.1 and Proposition 3.4 of \cite{bh}. We present slightly stronger versions of those results tailored for our purposes here. Starting with (i), suppose that $\alpha, \beta \in V_m$ are generic weights separated by a single hyperplane $H_\xi.$ Choose $\gamma \in H_\xi$ on the straight line connecting $\alpha$ to $\beta.$ Then ${\cal M}_\gamma$ is stratified by the Jordan-H\"older type of the underlying bundle, and since $\gamma$ lies on only one hyperplane, there are exactly two strata: the stable bundles ${\cal M}^s_\gamma$ and the strictly semistable bundles $\Sigma_\gamma.$ Writing $\xi=(r',d',m')$ for the partition, then it is not hard to see that $\Sigma_\gamma \cong {\cal M}_{\gamma'} \times {\cal M}_{\gamma''},$ with the obvious definitions for $\gamma'$ and $\gamma''$ coming from the partition $\xi.$ \begin{thm} \label{thm:bh1} There are natural algebraic maps $\phi_\alpha$ and $\phi_\beta$ $$\begin{array}{rcl}{\cal M}_\alpha& & {\cal M}_\beta\\ \;\; \phi_\alpha \!\! \searrow \!\!\!\!\!\!\! && \!\!\!\!\!\!\! \swarrow \!\! \phi_\beta\\& {\cal M}_\gamma \end{array}$$ which are generized blow-downs along projectivizations of vector bundles over $\Sigma_\gamma,$ where the projective fiber dimensions $e_\alpha$ and $e_\beta$ satisfy $e_\alpha + e_\beta + 1 = \operatorname{codim} \Sigma_\gamma.$ \end{thm} \begin{pf} The proof is the same as in \cite{bh}, the only difference being the actual computation of the numbers $e_\alpha$ and $e_\beta,$ which we discuss now. We assume that $E_* \sim_S E'_* \oplus E''_*,$ where $[E_*] \in \Sigma_\gamma$ and $\sim_S$ denotes Seshadri equivalence (i.e.\ isomorphic Jordan-H\"older form). The topological type of the parabolic bundles $E'_*$ and $E''_*$ does not change as $[E_*]$ varies within $\Sigma_\gamma.$ We use $(r',r''), (d',d'')$ and $(m',m'')$ to denote the ranks, degrees, and multiplicities of $(E'_*, E''_*),$ written as in \S 2. The moduli spaces ${\cal M}_\alpha, {\cal M}_\beta,$ and ${\cal M}_\gamma$ have dimension $$(g-1) r^2 +1 + \frac{1}{2}\sum_{p \in D} \left(r^2 - \sum_{i=1}^{\kappa_p} m_i(p)^2\right).$$ Using a similar formula for $\Sigma_\gamma={\cal M}^{\gamma'} \times {\cal M}^{\gamma''},$ we find that $$\operatorname{codim} \Sigma_\gamma = r' r''(2g - 1) -1+\sum_{p\in D}\sum_{i=1}^{\kappa_p} m'_i(p) m''_i(p).$$ Now we claim that $$h^0({E''_*}^\vee \otimes E'_*)=0=h^0({E'_*}^\vee \otimes E''_*).$$ This is true for any $\alpha' \in V_m,$ as one of these equations is true for $\alpha,$ the other for $\beta,$ but $H^0$ is constant as the weights are varied within $V_m.$ Let ${\cal U}'$ and ${\cal U}''$ be the families parametrized by $\Sigma_\gamma$ gotten by pulling back the universal families ${\cal U}^{\gamma'}$ and ${\cal U}^{\gamma''},$ whose existence follows from Proposition \ref{prop:fine}. Then the vector bundles referred to in the theorem are $$(R^1 \pi_{\Sigma_\gamma})({{\cal U}''}^\vee \otimes {\cal U}') \hbox{ and } (R^1 \pi_{\Sigma_\gamma})({{\cal U}'}^\vee \otimes {\cal U}'').$$ The projectivizations of these bundles have dimensions \begin{eqnarray} e_\alpha &=& h^1({E''_*}^\vee \otimes E'_*)-1 = r'' d' - r' d'' + r' r'' (g-1) + \chi ({\cal Q}) -1, \label{formula:ealpha}\\ e_\beta &=& h^1({E'_*}^\vee \otimes E''_*)-1 = r' d'' - r'' d' + r' r'' (g-1) + \chi ({\cal Q}') -1, \label{formula:ebeta} \end{eqnarray} where ${\cal Q}$ and ${\cal Q}'$ are skyscraper sheaves supported on $D$ obtained as the quotients \begin{eqnarray*} &\parhom (E''_*,E'_*) \longrightarrow \operatorname{{\frak H}{\frak o}{\frak m}} (E'',E') \longrightarrow {\cal Q},&\\ &\parhom (E'_*,E''_*) \longrightarrow \operatorname{{\frak H}{\frak o}{\frak m}} (E',E'') \longrightarrow {\cal Q}'.& \end{eqnarray*} It is a nice exercise to see $$ \chi ({\cal Q}) + \chi ({\cal Q}') = \sum_{p \in D} \left( r' r'' -\sum_{(i,j) \in S_e(p)} m'_i(p)m''_j(p) \right),$$ where $S_e(p) = \{ (i,j) \mid \gamma'_i(p) = \gamma''_j(p)\}.$ This shows $e_\alpha + e_\beta +1 = \operatorname{codim} \Sigma_\gamma.$ \end{pf} \begin{thm} \label{thm:bh2} Suppose that $\alpha \in V_\ell, \, \beta \in V_m, \, V_\ell > V_m,$ and that $\alpha$ and $\beta$ are generic and are not separated by any hyperplanes. Then there exists a fibration $\psi:{\cal M}_\alpha \longrightarrow {\cal M}_\beta$ with fiber a (possibly twisted) product of flag varieties and this fibration is locally trivial in the Zariski topology. In particular, ${\cal M}_\alpha$ is birational to the product of ${\cal M}_\beta$ with a product of flag varieties. \end{thm} \begin{pf} The hypothesis $V_\ell > V_m$ just means that the flag structure degenerates as we pass from $\alpha$ to $\beta.$ By induction, it is enough to prove the above statement when the degeneration of the flag structure is taking place at only one parabolic point. Given $E_*$ a parabolic bundle with multiplicities $m$ and weights $\alpha,$ let $E'_*$ be the parabolic bundle with multiplicities $\ell$ and weights $\beta$ resulting from forgetting part of the flag structure and interchanging the weights. One easily verifies that if $E_*$ is $\alpha$-stable, then $E_*'$ is $\beta$-stable, and the existence of the morphism $\psi$ then follows from the coarseness property of ${\cal M}_\beta.$ The remaining issue is to identify the fiber and to prove local triviality. For the first issue, notice that there is an inverse procedure to the forgetful map described above. Given a parabolic bundle $E'_*$ with multiplicities $\ell$ and weights $\beta,$ consider all parabolic bundles $E_*$ with weights $\alpha$ obtained from $E'_*$ by refining the flag stucture to one with multiplicities $m$ and exchanging the weights. For a given $E'_*,$ the set of all such possible refinements $E_*$ is parametrized by a flag variety. A straightforward numerical verification shows that applying this procedure to a $\beta$-stable parabolic bundle $E_*'$ yields an $\alpha$-stable $E_*$ for every possible refinement. It is not hard to see that the same procedure, when applied to the universal family ${\cal U}_*^\beta,$ identifies ${\cal M}_\alpha$ with the total space of the flag bundle of ${\cal U}^\beta$ restricted to ${\cal M}_\beta \times \{p\}$ and the map $\psi$ with the bundle projection. \end{pf} One might expect from Theorem \ref{thm:bh1} that the birational type of ${\cal M}_\alpha$ depends only on the underlying quasi-parabolic structure. This is the content of the following proposition. \begin{prop} \label{prop:qp} Suppose that $g \geq 2.$ Then the birational type of ${\cal M}_\alpha$ is independent of the choice of $\alpha \in V_m.$ \end{prop} \begin{pf} We prove the proposition by showing that ${\cal M}_\alpha$ and ${\cal M}_\beta$ are birational whenever $\alpha,\beta \in V_m$ are not separated by any walls (although one may lie on a wall which does not contain the other). So assume that $\alpha \in \cap_{i=1}^n H_{\xi_i}$ and $\beta \in \cap_{i=1}^m H_{\xi_i},$ where $m \geq n.$ By Theorem 4.1 \cite{mehta-seshadri}, ${\cal M}_\alpha$ and ${\cal M}_\beta$ are normal, projective varieties and $\dim {\cal M}_\alpha = \dim {\cal M}_\beta,$ hence we only need to construct an injective morphism $\phi:{\cal M}^s_\beta \longrightarrow {\cal M}^s_\alpha$ to conclude ${\cal M}_\alpha$ is birational to ${\cal M}_\beta.$ One easily verifies that every $\beta$-stable bundle is $\alpha$-stable, and the existence of $\phi$ follows from the coarseness of ${\cal M}_\alpha.$ \end{pf} \section{Shifting and the Hecke correspondence} In this section, we introduce the notion of a shifted parabolic bundle, which is the result of changing the weights, multiplicities, and degree of $E_*$ in a prescribed way. In some sense, shifting is a symmetry of a larger weight space, one which includes bundles of different degrees. Two applications of shifting are discussed at the end. Shifting is most naturally described in terms of parabolic sheaves. If ${\cal E}$ is a locally free sheaf on $X,$ then a {\it parabolic structure} on ${\cal E}$ consists of a weighted filtration of the form \begin{eqnarray} \label{eqn:filtration1} &{\cal E}={\cal E}_{\alpha_1} \supset{\cal E}_{\alpha_2} \supset \cdots \supset {\cal E}_{\alpha_l} \supset {\cal E}_{\alpha_{l+1}} = {\cal E}(-D),&\\ &0 \leq \alpha_1 < \alpha_2 < \cdots < \alpha_l < \alpha_{l+1}=1.\quad \quad& \label{eqn:filtration2} \end{eqnarray} We can define ${\cal E}_x$ for $x \in [0,1]$ by setting ${\cal E}_x = {\cal E}_{\alpha_i}$ if $\alpha_{i-1} < x \leq \alpha_i,$ and then extend to $x \in {\Bbb R}$ by setting ${\cal E}_{x+1} ={\cal E}_{x}(-D).$ We call the resulting filtered sheaf ${\cal E}_*$ a parabolic sheaf and ${\cal E}={\cal E}_0$ the underlying sheaf. We can define parabolic subsheaves, degree, and stability for these objects, and there is a categorical equivalence between locally free parabolic sheaves and parabolic bundles. We describe this in case $D = p,$ the general case being quite similar (\cite{yoko}, \cite{by}). Suppose that $E_*$ is a parabolic bundle given by flags and weights in the fibers as in (\ref{defn:parbun1}) and (\ref{defn:parbun2}). Define ${\cal E}_*$ by setting $${\cal E}_x = \ker(E\rightarrow E_{p}/F_i),$$ for $ \alpha_{i-1} < x < \alpha_i.$ Thus ${\cal E}_*$ is a parabolic sheaf. Conversely, given a parabolic sheaf ${\cal E}_*,$ the quotient ${\cal E}_0/{\cal E}_1 = {\cal E}/{\cal E}(-p)$ is a skyscraper sheaf with support $p$ and fiber that of ${\cal E}.$ Defining a flag in this fiber by setting $F_i = ({\cal E}_{\alpha_{i}}/{\cal E}_1)_p$ and associating the weight $\alpha_i,$ we obtain a parabolic bundle in the sense of (\ref{defn:parbun1}) and (\ref{defn:parbun2}). The category of parabolic sheaves is developed in \cite{yoko}, where one finds for example the definitions of tensor products ${\cal E}_* \otimes {\cal E}'_*$ and duals ${\cal E}^\vee_*.$ We use this notation freely in the calculations of \S 6 involving sheaf cohomology and point out that $H^i({\cal E}_*) = H^i({\cal E}).$ \begin{defn} \label{defn:shift} Given a parabolic sheaf ${\cal E}_*$ and $\eta \in {\Bbb R},$ define the shifted parabolic sheaf ${\cal E}_*[\eta]_*$ by setting ${\cal E}_*[\eta]_x = {\cal E}_{x+\eta}.$ \end{defn} \noindent {\it Remark.} The above operation can be refined in case $D = p_1 + \cdots + p_n.$ If $\eta=(\eta_1, \ldots, \eta_n),$ then one can shift ${\cal E}_*$ by $\eta_i$ at each $p_i \in D$ (\cite{yoko}, \cite{bh}). \medskip \noindent It is not difficult to verify that ${\cal E}_*[\eta]_*$ is (semi)stable if and only if ${\cal E}_*$ is (semi)stable, and it follows that this defines an isomorphism between the associated moduli spaces of parabolic bundles. We can easily describe the parabolic structure on the shifted bundle ${\cal E}'_* = {\cal E}_*[\eta]_*$ in case $0 < \eta \leq 1$ and $D = p.$ Let $E'_*$ denote the parabolic bundle associated to ${\cal E}'_*.$ If $i$ is the integer with $\alpha_i < \eta \leq \alpha_{i+1},$ then the weights of $E'_*$ are given by \begin{equation} \label{eqn:sft_wts} \alpha'_j = \begin{cases} \alpha_{j+i} - \eta & \hbox{ for } j=1, \ldots, r-i, \\ 1+\alpha_{j-r+i}-\eta & \hbox{ for } j = r-i+1,\ldots, r. \end{cases} \end{equation} The quasi-parabolic structure of $E'_*$ has multiplicities $m'$ given by a cyclic permutation of $m,$ i.e.\ $m' = (m_{i+1}, \ldots, m_\kappa, m_1, \ldots, m_i).$ Although ${\cal E}'$ is a subsheaf of ${\cal E},$ $E'$ is {\it not} a subbundle of $E,$ so one must appeal to sheaf theory in order to define the flag in $E'_p.$ This is a simple exercise in tracing through the equivalence between locally free parabolic sheafs and parabolic bundles given above. \begin{figure}[b] \begin{picture}(100,210)(-110,-10) \put(-70,150){${\cal E}_*$} \put(-70,100){${\Bbb R}$} \put(-40,110){\line(1,0){290}} \put(5,110){\line(0,1){70}} \put(3,100){$0$} \put(170,110){\line(0,1){20}} \put(168,100){$1$} \put(-12,185){${\cal E}={\cal E}_{\alpha_1}$} \put(-40,180){\line(1,0){78}} \put(40,180){\circle*{4}} \put(34,100){$\alpha_1$} \multiput(40,110)(0,4){12}{\line(0,1){2}} \multiput(40,178)(0,-4){4}{\line(0,-1){2}} \put(55,165){${\cal E}_{\alpha_2}$} \put(42,160){\line(1,0){48}} \put(40,160){\circle{4}} \put(90,160){\circle*{4}} \put(88,100){$\alpha_2$} \multiput(90,110)(0,4){7}{\line(0,1){2}} \multiput(90,158)(0,-4){4}{\line(0,-1){2}} \put(92,140){\line(1,0){48}} \put(109,145){${\cal E}_{\alpha_3}$} \put(90,140){\circle{4}} \put(140,140){\circle*{4}} \put(137,100){$\alpha_3$} \multiput(140,110)(0,4){5}{\line(0,1){2}} \multiput(140,138)(0,-4){2}{\line(0,-1){2}} \put(142,130){\line(1,0){48}} \put(140,130){\circle{4}} \put(190,130){\circle*{4}} \put(152,135){${\cal E}(-p)$} \put(185,100){$1\!+\!\alpha_1$} \multiput(190,110)(0,4){2}{\line(0,1){2}} \multiput(190,128)(0,-4){2}{\line(0,-1){2}} \put(192,120){\line(1,0){58}} \put(190,120){\circle{4}} \put(200,125){${\cal E}_{\alpha_2}(-p)$} \put(-80,50){${\cal E}_*[\eta]_*$} \put(-70,0){${\Bbb R}$} \put(-40,10){\line(1,0){250}} \put(5,10){\line(0,1){50}} \put(3,0){$0$} \put(170,10){\line(0,1){10}} \put(168,0){$1$} \put(-40,85){${\cal E}_{\alpha_1}$} \put(-40,80){\line(1,0){18}} \put(-20,80){\circle*{4}} \put(-40,0){$\alpha_1\!-\!\eta$} \multiput(-20,10)(0,4){12}{\line(0,1){2}} \multiput(-20,78)(0,-4){4}{\line(0,-1){2}} \put(-5,65){${\cal E}_{\alpha_2}$} \put(-18,60){\line(1,0){48}} \put(-20,60){\circle{4}} \put(30,60){\circle*{4}} \put(20,0){$\alpha_2\!-\!\eta$} \multiput(30,10)(0,4){7}{\line(0,1){2}} \multiput(30,58)(0,-4){4}{\line(0,-1){2}} \put(32,40){\line(1,0){48}} \put(49,45){${\cal E}_{\alpha_3}$} \put(30,40){\circle{4}} \put(80,40){\circle*{4}} \put(67,0){$\alpha_3\!-\!\eta$} \multiput(80,10)(0,4){5}{\line(0,1){2}} \multiput(80,38)(0,-4){2}{\line(0,-1){2}} \put(82,30){\line(1,0){48}} \put(80,30){\circle{4}} \put(130,30){\circle*{4}} \put(92,35){${\cal E}(-p)$} \put(110,0){$1\!+\!\alpha_1\!-\!\eta$} \multiput(130,10)(0,4){2}{\line(0,1){2}} \multiput(130,28)(0,-4){2}{\line(0,-1){2}} \put(132,20){\line(1,0){78}} \put(130,20){\circle{4}} \put(140,25){${\cal E}_{\alpha_2}(-p)$} \end{picture} \caption{The parabolic sheaf ${\cal E}_*$ shifted by $\eta$ with $\alpha_1 < \eta < \alpha_2.$} \end{figure} There are two interesting applications of shifting we discuss now. The first is the Hecke correspondence. Using ${\cal M}_{r,d}$ to denote the moduli space of semistable bundles of rank $r$ and degree $d,$ the Hecke correspondence gives a means of comparing ${\cal M}_{r,d}$ and ${\cal M}_{r,d'}$ through the use of parabolic bundles. For $r=2,$ this was observed in a remark at the end of \cite{mehta-seshadri}. To start, define $\epsilon_+(d,r), \epsilon_-(d,r),$ and $\epsilon(d,r)$ for $d,r \in {\Bbb Z}$ with $r >0$ by \begin{eqnarray*} \epsilon_\pm(d,r) &=& \inf \{ \pm({\textstyle \frac{d}{r}-\frac{d'}{r'}}) \mid d',r' \in {\Bbb Z}, \; 1 \leq r' < r, \hbox{ and } \pm ({\textstyle \frac{d}{r}-\frac{d'}{r'}}) > 0 \} \\ \epsilon(d,r) \;\; &=& \min \{ \epsilon_\pm(d,k) \mid k=1, \ldots, r \}. \end{eqnarray*} It is easy to see that $\epsilon_\pm(d,k) > 0$ for all $k,$ thus $\epsilon(d,r)>0$ as well. Suppose that $E$ is a bundle over $X$ of degree $d$ and rank $r$ and suppose further that $E'$ is a proper subbundle. If $\mu(E') < \mu(E),$ then $\mu(E)-\mu(E') \geq \epsilon_+(d,r).$ Similarly, if $\mu(E') > \mu(E),$ then $\mu(E')-\mu(E) \geq \epsilon_-(d,r).$ \begin{prop} \label{prop:num} Suppose that $E_*$ satisfies $ {\displaystyle \sum_{p \in D} \sum_{i=1}^{\kappa_p} m_i(p)} \alpha_i(p) < \epsilon(d,r)/2.$ \begin{enumerate} \item[(i)] If $E$ is stable as a regular bundle, then $E_*$ is parabolic stable. \item[(ii)] If $E_*$ is parabolic stable, then $E$ is semistable as a regular bundle. \end{enumerate} \end{prop} \begin{pf} (i) If $E'_*$ is a proper parabolic subbundle of $E_*$, then $$\mu(E'_*) \leq \mu(E') + \epsilon(d,r)/2 < \mu(E') + \epsilon_+(d,r) \leq \mu(E) < \mu(E_*),$$ thus $E_*$ is parabolic stable. (ii) If $E'$ is a subbundle of $E,$ then $$\mu(E') \leq \mu(E'_*) < \mu(E_*) < \mu(E) + \epsilon(d,r)/2 < \mu(E) + \epsilon_-(d,r),$$ hence $\mu(E') \leq \mu(E)$ and $E$ is semistable. \end{pf} We thus get a morphism ${\cal M}_\alpha \longrightarrow {\cal M}_{r,d}$ which is the map of Theorem \ref{thm:bh2} in case $(r,d)=1.$ By choosing the weights and quasi-parabolic structure correctly, we can fit ${\cal M}_{r,d}$ and ${\cal M}_{r,d-1}$ into a chain diagram of maps as follows. Let $D=p$ and $m=(1,\ldots, 1),$ and choose weights $\alpha=(\alpha_1,\ldots,\alpha_r)$ with $\alpha_1 + \cdots + \alpha_r < \epsilon(r,d)/2.$ Suppose $\alpha_1 < \eta < \alpha_2$ and set $E'_*$ to be the parabolic bundle $E_*$ shifted by $\eta.$ Notice that $E'_*$ has degree $d-1,$ multiplicities $m'=(1,\ldots,1),$ and weights $\alpha'=(\alpha_2 - \eta, \ldots, \alpha_r-\eta, 1-\eta+\alpha_1).$ If $\beta' \in V_{m'}$ is generic with $\beta'_1 + \cdots + \beta'_r < \epsilon(r,d)/2,$ then we can connect $\alpha'$ to $\beta'$ in $V_{m'}$ by a line passing through a finite number of hyperplanes $H_{\xi^1}, \ldots, H_{\xi^n},$ all of the form to which Theorem \ref{thm:bh1} applies. Choose weights $\alpha^i$ in the intermediate chambers and $\gamma^i \in H_{\xi^i}$ for $i=1,\ldots, n$ with $\alpha^n = \beta'.$ Applying Theorem \ref{thm:bh1} each time we cross a hyperplane, we get the following diagram: \begin{equation} \label{eqn:hecke} \begin{array}{lllllllll}{\cal M}_\alpha \cong {\cal M}_{\alpha'} & &&&{\cal M}_{\alpha^1} & & && \!\!\!{\cal M}_{\beta'}\\ \psi \downarrow & \searrow && \swarrow && \searrow && \swarrow & \downarrow \psi' \!\!\!\! \\ {\cal M}_{r,d} && {\cal M}_{\gamma^1} &&& & \cdots & & \!\!\!{\cal M}_{r,d-1} \end{array} \end{equation} \medskip \noindent where, by the above proposition, the vertical maps $\psi$ and $\psi'$ have fibers the (full) flag variety over ${\cal M}^s_{r,d}$ and ${\cal M}^s_{r,d-1},$ respectively. By Theorem \ref{thm:bh2}, $\psi$ is a fibration which is locally trivial in the Zariski topology provided $(r,d)=1,$ and the same follows for $\psi'$ if $(r,d-1)=1.$ The second application of shifting is to extend the results of \cite{bh} to a case which is natural from the point of view of representations of Fuchsian groups but less natural from the point of view of parabolic bundles. Assume for simplicity that $\mu(E_*) = 0$ and $D=p.$ Thus, $\deg E = -k$ for some $0 \leq k < r,$ and the relevant weight space is $$W_k = \{ (\alpha_1,\ldots,\alpha_r) \in \Delta^r \mid \alpha_1 + \cdots + \alpha_r = k \}.$$ Consider the union ${\displaystyle \widetilde{W} = \bigcup_{k=0}^{r-1} W_k},$ where we identify $$\partial_0 W_k = \{ \gamma \in W_k \mid \gamma_1=0 \}$$ with its companion set $$\partial_1 \overline{W}_{k+1} = \{ \overline{\gamma} \in \overline{W}_{k+1} \mid \overline{\gamma}_r=1 \}$$ via the identification \begin{equation} \label{eqn:glue} \partial_0 W_k \ni \gamma = (0, \gamma_2, \ldots, \gamma_n) \sim (\gamma_2, \ldots, \gamma_n, 1) = \overline{\gamma} \in \partial_1 \overline{W}_{k+1}. \end{equation} One can think of this set $\widetilde{W}$ as the space of all weights modulo shifting\footnote{Because every bundle can be shifted so that $\mu(E_*)=0.$}, which in this case is just the quotient $\operatorname{SU}(r)/Ad$ and which can be naturally identified with the standard $r-1$ simplex. {}From this point of view $\partial_0 W_k$ is an interior hyperplane of $\widetilde{W}$ because it satisfies condition (\ref{eqn:hyper}). However, Theorem \ref{thm:bh1} does not obviously carry over to this case because points in $W_k$ and $W_{k+1}$ are weights on parabolic bundles of different degrees. Given a parabolic bundle of degree $-k,$ what is needed is a canonical procedure to construct a parabolic bundle of degree $-(k+1).$ This is precisely what is provided by the shifting operation. Thought of in terms of $ \widetilde{W},$ the following theorem extends Theorem \ref{thm:bh1} to the case where $H_\xi= \partial_0 W.$ We use the notation ${\cal M}_\alpha(k,m)$ for the moduli space when $E_*$ has degree $-k,$ multiplicities $m,$ and weights $\alpha.$ \begin{thm} \label{thm:bh3} Suppose that $\gamma \in \partial_0 W_k \cap V_m$ does not lie on any other hyperplanes and that $\alpha \in W_k \cap V_m$ is a generic weight near to $\gamma.$ Choose $\eta \in {\Bbb R}$ with $0<\eta < \gamma_{m_1+1}.$ Define $\overline{\gamma}\in \partial_1 \overline{W}_{k+1}$ as in \rom(\ref{eqn:glue}\rom). Let $E'_*$ be $E_*$ shifted by $\eta,$ and denote the multiplicities of $E'_*$ by $m'.$ Set $k'= - \deg E' = k+m_1.$ Let $\beta \in W_{k'} \cap V_{m'}$ be generic near $\overline{\gamma}.$ Then there are projective algebraic maps $\phi_\alpha$ and $\phi_\beta$ $$\begin{array}{rcl}{\cal M}_\alpha(m,k)& & {\cal M}_\beta(m',k')\\ \;\; \phi_\alpha \!\! \searrow \!\!\!\!\!\! && \!\!\!\!\! \swarrow \!\! \phi_\beta\\& {\cal M}_\gamma(m,k) \end{array}$$ satisfying the conclusion of Theorem \ref{thm:bh1}. \end{thm} \begin{pf} By the choice of $\alpha, \beta$ and $\eta,$ we see that $\alpha_{m_1} < \eta < \alpha_{m_1+1}, \; \eta < \beta_1$ and $\eta < \gamma_{m_1+1}.$ Consequently, the shifting operation defines the following isomorphisms: \begin{eqnarray*} {\cal M}_\alpha(m,k) &\cong& {\cal M}_{\alpha'}(m',k'),\\ {\cal M}_\beta(m',k') &\cong& {\cal M}_{\beta'}(m',k'),\\ {\cal M}_\gamma(m,k) &\cong& {\cal M}_{\gamma'}(m',k'), \end{eqnarray*} where $\alpha', \beta', \gamma' \in V_{m'}$ are defined from $\alpha, \beta, \gamma$ as in (\ref{eqn:sft_wts}). Now Theorem \ref{thm:bh1} applies to the shifted moduli spaces to prove the theorem. One can calculate $e_\alpha$ and $e_\beta$ by applying formulas (\ref{formula:ealpha}) and (\ref{formula:ebeta}) to $ \alpha', \beta'$ and $\gamma'.$ \end{pf} \medskip \noindent {\it Remark.} Theorem \ref{thm:bh3} solves a problem mentioned at the end of \cite{bh} and extends the wall-crossing formula for knot invariants introduced in \cite{boden3}. \section{Rationality of moduli spaces of parabolic bundles} Let $L$ be a holomorphic line bundle over a curve $X$ of genus $g \geq 2.$ Denote by \begin{enumerate} \item[(i)] ${\cal M}_{r,L}$ the moduli space of semistable bundles $E$ of rank $r$ with $\det E = L,$ and by \item[(ii)] ${\cal M}_{\alpha, L}$ the moduli space of parabolic bundles $E_*$ with weights $\alpha$ and $\det E = L.$ \end{enumerate} The main results of \S 4 hold for the moduli spaces with fixed determinant with no essential difference. In view of Theorem \ref{thm:bh2}, the goal is therefore to prove rationality with the coarsest possible choice of flag structure. At one extreme, we have the trivial flag, whose moduli space is exactly ${\cal M}_{r,L}.$ Proposition 2 of \cite{newstead1} implies that ${\cal M}_{r,L}$ is rational if $\deg L = \pm 1 \mod (r),$ and then Theorem \ref{thm:bh2} and Proposition \ref{prop:qp} imply that ${\cal M}_{\alpha,L}$ is also rational for any $\alpha \in V_m$ provided $\deg L = \pm 1 \mod (r).$ \begin{thm} \label{thm:rat1} If $m(p)=(1,\ldots,1)$ for some $p \in D,$ then ${\cal M}_{\alpha, L}$ is rational for all $\alpha \in V_m.$ \end{thm} \begin{pf} First, use Theorem \ref{thm:bh2} to reduce to the case $D=p$ by forgetting all the other flag structures. If $E'_*$ denotes the bundle obtained by shifting $E_*$ by some $\eta$ with $\alpha_1 < \eta < \alpha_2,$ then $\det E' = L' = L(-p).$ It follows that shifting by $\eta$ defines an isomorphism from ${\cal M}_{\alpha,L}$ to ${\cal M}_{\alpha',L'}.$ Repeated application of shifting puts us in the case $\deg L = 1 \mod(r),$ and then Newstead's theorem and Theorem \ref{thm:bh2} imply that ${\cal M}_{\alpha,L}$ is rational. \end{pf} The above argument works in slightly more generality. We can always shift our bundle to be any of the ${\cal E}_x$ appearing in the filtration (\ref{eqn:filtration1}) and illustrated in Figure 1. Thus, whenever one of these terms in the filtration is of a degree to which Newstead's theorem applies, the corresponding moduli space of parabolic bundles is rational. The next theorem is a considerable strengthening of the previous one. \begin{thm} \label{thm:rat2} If $m_i(p)=1$ for some $p \in D$ and some $1 \leq i \leq \kappa_p,$ then ${\cal M}_{\alpha,L}$ is rational for all $\alpha \in V_m.$ \end{thm} Before delving into the proof of this theorem, we mention some interesting consequences. Recall first the following definition. \begin{defn} \label{defn:stablerat} A variety $V$ is stably rational of level $k$ if $V \times {\Bbb P}^k$ is rational. The level is the smallest integer $k$ with this property. \end{defn} The following result, with a weaker bound on the level, was proved in \cite{ballico}. \begin{cor} \label{cor} For $(r,d)=1, \; {\cal M}_{r,L}$ is stably rational with level $k \leq r-1.$ \end{cor} \begin{pf} Theorem \ref{thm:rat2} implies that ${\cal M}_{\alpha,L}$ is rational, where $m(p)=(r-1,1),$ and Theorem \ref{thm:bh2} shows that ${\cal M}_{\alpha,L}$ is birational to ${\cal M}_{r,L} \times {\Bbb P}^{r-1},$ which proves the corollary. \end{pf} We now apply this last result to Conjecture \ref{conj}. \begin{cor} \label{cor:rat} Suppose $(r,d)=1.$ By tensoring with a line bundle, we can assume that $0 < d < r.$ If either $(g,d) =1$ or $(g,r-d)=1,$ then ${\cal M}_{r,L}$ is rational. \end{cor} \begin{pf} Suppose first that $(g,r-d)=1.$ Let $L$ be a line bundle of degree $r(g-1)+d.$ Then Newstead's construction applies and proves that ${\cal M}_{r,L}$ is birational to ${\cal M}_{r-d,L} \times {\Bbb P}^\chi,$ where $\chi = (g-1)(r^2-(r-d)^2).$ But the above corollary implies that ${\cal M}_{r-d,L}$ is stably rational with level $k \leq r-d-1 \leq \chi,$ hence ${\cal M}_{r,L}$ is rational. The case $(g,d)=1$ follows by the same argument after applying duality, which interchanges $(r,d)$ and $(r,r-d).$ \end{pf} {\it Remark.} Conjecture \ref{conj} was previously known \cite{newstead1} in the following three cases: \begin{enumerate} \item[(i)] $d = \pm 1 \mod (r),$ \item[(ii)] $(r,d)=1$ and $g$ a prime power, and \item[(iii)] $(r,d)=1$ and the two smallest distinct primes factors of $g$ have sum greater than $r.$ \end{enumerate} Corollary \ref{cor:rat} applies in each case. More importantly, it applies in many cases not covered by (i), (ii) or (iii). In fact, for a given $r$ and $d$ with $(r,d)=1,$ one can easily list those $g$ for which the conjecture remains open. For example, if $r= 110$ and $d=43,$ then Corollary \ref{cor:rat} applies as long as $g$ is not a multiple of $d \cdot (r-d)= 43 \cdot 67=2881.$ \medskip \noindent {\it Proof of Theorem.} Set $d= \deg L.$ The theorem is clearly true for $r=1$ and follows from Theorem \ref{thm:rat1} for $r=2,$ so assume $r>2.$ Notice that by tensoring with a line bundle, we can suppose $$r(g-1) < d \leq rg.$$ By Theorem \ref{thm:bh2}, we can again assume that $D=p,$ and by shifting and another application of Theorem \ref{thm:bh2}, if necessary, we can arrange it so that $m(p)=(r-1,1).$ Write $$\alpha = \alpha(p) = (\overbrace{\alpha_1, \cdots, \alpha_1}^{r-1}, \alpha_2).$$ Proposition \ref{prop:fine} implies that $V_m$ contains a generic weight and that ${\cal M}_{\alpha,L}$ parametrizes a universal family ${\cal U}_*^\alpha.$ By Proposition \ref{prop:qp}, the birational type of ${\cal M}_{\alpha,L}$ is independent of choice of compatible weights, so we can assume that the weights are small enough to satisfy the hypothesis of Proposition \ref{prop:num} (this comes up at various technical points in the argument, e.g. the proof of Claim \ref{claim}). \medskip\noindent Consider the following two cases. \medskip\noindent {\sc Case I:} \quad $d=rg.$ \quad Choose $\eta$ with $\alpha_1 < \eta < \alpha_2,$ and let $E'_* = E_*[\eta]_*.$ Denote the weights of $E'_*$ by $\alpha'$ as in (\ref{eqn:sft_wts}). If $\det E = L,$ then $\det E' = L' = L(-(r-1)p)$ has degree $d'= d-(r-1).$ Since $d' = 1 \mod(r),$ Proposition 2 of \cite{newstead1} implies that ${\cal M}_{r,L'}$ is rational, and Theorem \ref{thm:bh2} then implies that ${\cal M}_{\alpha'\!,L'}$ is also rational. Rationality of ${\cal M}_{\alpha,L}$ now follows from the isomorphism of the moduli spaces ${\cal M}_{\alpha,L} \cong {\cal M}_{\alpha'\!,L'}$ defined by shifting by $\eta.$ \medskip\noindent {\sc Case II:} \quad $r(g-1) < d < rg.$ \quad The idea is to use induction to construct a nonempty, Zariski-open subset ${\cal M}$ of affine space of dimension $(r^2-1)(g-1)+r-1 \; (= \dim {\cal M}_{\alpha,L})$ and a family of stable parabolic bundles ${\cal U}_*$ parametrized by ${\cal M}$ with $\det {\cal U}_{\xi,*} = L$ for all $\xi \in {\cal M}.$ The universal property of ${\cal U}_*^\alpha$ then gives a map $\psi_{{\cal U}_*} : {\cal M} \longrightarrow {\cal M}_{\alpha,L}.$ If, in addition, we have ${\cal U}_{\xi_1,*} \cong {\cal U}_{\xi_2,*}\; \Leftrightarrow \; \xi_1 = \xi_2,$ then $\psi_{{\cal U}_*}$ is injective and rationality of ${\cal M}_{\alpha,L}$ follows from that of ${\cal M}$ and the dimension condition. Set $r' = rg-d, \, r'' = r-r'$ and $\alpha'= (\overbrace{\alpha_1, \cdots, \alpha_1}^{r'-1}, \alpha_2).$ Assume that both $\alpha$ and $\alpha'$ are generic. Let ${\cal U}_*^{\alpha'}$ be the universal family parametrized by ${\cal M}_{\alpha'\!,L}$ and $I_*= {\cal O}_X[\alpha_1]_*$ be the trivial parabolic line bundle with weight $\alpha_1.$ If $e' = [E'_*] \in {\cal M}_{\alpha'\!,L},$ then because ${E'_*}^\vee\otimes I_*$ is a stable parabolic bundle of negative parabolic degree, $h^0({E'_*}^\vee \otimes I_*)=0$ and \begin{equation} \label{eq:rank} n \stackrel{\hbox{\scriptsize\rm def}}{=} h^1({E'_*}^\vee \otimes I_*) = (2r'+r'')(g-1) + r'' + 1 \end{equation} is independent of $e'.$ Since ${\cal U}^{\alpha'}_{e',*} \cong E'_*,$ it follows that $$(R^1 \pi_{{\cal M}_{\alpha'\!,L}}) (({{\cal U}_*^{\alpha'}})^\vee \otimes \pi_X^*(I_*))$$ is locally free. The associated vector bundle $V \stackrel{\pi}{\longrightarrow} {\cal M}_{\alpha'\!,L}$ has rank $n$ and fiber over $e'$ naturally isomorphic to $H^1({E'_*}^\vee \otimes I_*).$ Let ${\cal U}_*'=(\pi^{r''} \times 1_X)^* ({\cal U}_*^{\alpha'})$ be the pullback family and ${\cal I}^{\oplus r''}_* =\pi_X^* I_*^{\oplus r''}$ the trivial family, where $\pi^{r''}:V^{\oplus r''} \longrightarrow {\cal M}_{al'\!,L}$. There is an extension \begin{equation} \label{eqn:universal_extension} 0 \longrightarrow {\cal I}_*^{\oplus r''} \longrightarrow {\cal U}_* \longrightarrow {\cal U}_*' \longrightarrow 0 \end{equation} of families over $V^{\oplus r''} \times X,$ such that, for $\xi \in V^{\oplus r''}_{e'},$ ${\cal U}_{\xi,*}$ is the parabolic bundle $E^\xi_*$ described as the short exact sequence \begin{equation} \label{eq:ses} 0 \longrightarrow I^{\oplus r''}_* \longrightarrow E^\xi_* \longrightarrow E'_* \longrightarrow 0 \end{equation} corresponding to the extension class $\xi \in H^1({E'_*}^\vee \otimes I^{\oplus r''}_*).$ Using stability of $E'_*$ and triviality of $I^{\oplus r''}_*,$ it follows that $$\operatorname{Aut}(E'_*) \times \operatorname{Aut}(I^{\oplus r''}_*) \cong {\Bbb C}^* \times \operatorname{GL}(r'', {\Bbb C}).$$ This group acts naturally as fiber-preserving maps on the bundle $V^{\oplus r''}$ since $$V^{\oplus r''}_{e'} \cong H^1({E'_*}^\vee \otimes I^{\oplus r''}_*) = H^1({E'_*}^\vee \otimes I_*)^{\oplus r''},$$ and two extension classes $\xi_1$ and $\xi_2$ in the same orbit have associated bundles $E^{\xi_1}$ and $E^{\xi_2}$ which are isomorphic. We can ignore the ${\Bbb C}^*$ action here because $(z,1)\cdot \xi = (1,z) \cdot \xi$ for $z \in {\Bbb C}^*$ and $\xi \in V^{\oplus r''}.$ Using the inductive hypothesis and local triviality of $V,$ we can choose a nonempty Zariski-open subset ${\cal M}'$ of ${\cal M}_{\alpha'\!,L}$ isomorphic to a Zariski-open subset of affine space of dimension $({r'}^2-1)(g-1)+r'-1$ such that $V|_{{\cal M}'} \cong {\cal M}' \times H^1({E'_*}^\vee \otimes I_*)$ ($E'_*$ is fixed). Lemma 2 of \cite{newstead1} applies here and produces a Zariski-open subspace ${\cal M}'\times W$ of $V^{\oplus r''}|_{{\cal M}'}$ invariant under the group action, and an affine subspace $A \subset W$ so that every orbit in $W$ intersects $A$ precisely once. In fact, $A$ can be chosen as a Zariski-open subset of the Grassmannian $G(r'',n).$ In any case, it should be clear that $A$ has dimension $r''(n-r'').$ Using equation (\ref{eq:rank}) and the fact that $r'+r'' = r,$ we see that ${\cal M}' \times A$ is a Zariski-open subset of affine space of dimension \begin{eqnarray*} \dim {\cal M}' \times A &=& ({r'}^2-1)(g-1)+r'-1 + r''(n - r'')\\ &=& ({r'}^2-1)(g-1)+r'-1 + r''((2r'+r'')(g-1) + 1) \\ &=& (r^2-1)(g-1)+r-1. \end{eqnarray*} Let ${\cal M}$ be the subset of $V^{\oplus r''}$ defined by $${\cal M} = \{ \xi \in {\cal M}' \times A \mid H^1({\cal U}_{\xi,*}) = 0 \},$$ and consider the bundle ${\cal U}_*$ restricted to ${\cal M},$ which we continue to denote ${\cal U}_*.$ For $\xi \in V^{\oplus r''},$ let $E^\xi_* = {\cal U}_{\xi,*}.$ Clearly $E^\xi_*$ is a parabolic bundle with weights $\alpha$ and determinant $L,$ thus ${\cal M}$ parametrizes a family of parabolic bundles. By the upper semi-continuity theorem, ${\cal M}$ is Zariski-open in ${\cal M}' \times A.$ We claim that ${\cal M}$ is nonempty. Fix $e'=[E'_*] \in {\cal M}'$ and consider the set $$N= \{ \xi \in H^1({E'_*}^\vee \otimes I^{\oplus r''}_*) \mid h^1(E^\xi_*)=0 \}.$$ If $N \cap A \neq \emptyset,$ then ${\cal M}$ is nonempty. Clearly, $N$ is invariant under the action of $\operatorname{GL}(r'',{\Bbb C}),$ so it is enough to show $N \cap W \neq \emptyset.$ There is a natural map $$\delta : H^1({E'_*}^\vee \otimes I^{\oplus r''}_*) \times H^0(E'_*) \longrightarrow H^1(I^{\oplus r''}_*)$$ with $\delta_\xi = \delta(\xi,\cdot) : H^0(E'_*) \longrightarrow H^1(I^{\oplus r''}_*)$ the coboundary map of the long exact sequence in homology of (\ref{eq:ses}). Now $H^0(E'_*)=H^0(E'),$ and since $\alpha_1+(r'-1) \alpha_2 < \epsilon(r,d)/2,$ by Proposition \ref{prop:num}, $E'$ is semistable as a non-parabolic bundle. Serre duality implies that $h^1(E')=h^0({E'}^\vee \otimes K),$ and we compute \begin{eqnarray*} \deg({E'}^\vee \otimes K) &=& -d +r'(1-g) \\ &\leq& (r+r')(1-g) - r'', \end{eqnarray*} which is negative since $r'' \geq 1$ and $g \geq 2.$ This implies that $h^1(E'_*) =0,$ and Riemann-Roch implies that $h^0(E'_*) = r''g.$ Because $h^1(I^{\oplus r''}_*)=r''g,$ we see that $$\xi \in N \Longleftrightarrow H^1(E^\xi_*) =0 \Longleftrightarrow \delta_\xi \hbox{ is an isomorphism.}$$ But $\delta$ is obviously onto and $\dim (\ker \delta) = r''n.$ The set $N$ has complement $$N^c = \{ \xi \in H^1({E'_*}^\vee \otimes I^{\oplus r''}_*) \mid \delta(\xi,s)=0 \hbox{ for some }0 \neq s \in H^0(I^{\oplus r''}_*) \}.$$ But $\delta(\xi,s) = 0 \Rightarrow \delta(\xi, zs)=0$ for all $z \in {\Bbb C},$ which shows that the map $\ker \delta \longrightarrow N^c$ has fibers of dimension $\geq 1.$ Hence $\dim N^c \leq \dim(\ker \delta) -1 < r''n,$ and we see that $N$ is nonempty and Zariski-open. Thus $N \cap W \neq \emptyset$ and it follows that ${\cal M}$ is nonempty. We now prove that ${\cal M}$ parametrizes a family of stable parabolic bundles, using again the inequality $(r-1) \alpha_1 + \alpha_2 < \epsilon(r,d)/2$ and Proposition \ref{prop:num}. \begin{claim} \label{claim} \begin{enumerate} \item[(i)] $E^\xi_*$ is stable for all $\xi \in {\cal M}.$ \item[(ii)] $E^{\xi_1}_* \cong E^{\xi_2}_* \Longleftrightarrow \operatorname{GL}(r'', {\Bbb C}) \cdot \xi_1 = \operatorname{GL}(r'', {\Bbb C}) \cdot \xi_2$ for all $\xi_1, \xi_2 \in {\cal M}.$ \end{enumerate} \end{claim} \begin{pf} (i) Suppose to the contrary that $E^\xi_*$ is not parabolic stable for some $\xi \in {\cal M}.$ Let $G_*$ be a rank $s$ parabolic subbundle of $E^\xi_*$ with $\mu(G_*) > \mu(E^\xi_*).$ Then $\mu(G) \geq \mu(E^\xi),$ since otherwise $$\mu(G_*) < \mu(G) + \epsilon(d,r)/2 < \mu(E^\xi) < \mu(E^\xi_*).$$ As in the argument of Lemma 6 of Newstead, the map $G \longrightarrow E'$ has a factorization as $G \rightarrow G^1 \rightarrow G^2 \rightarrow E'$ and the arguments there give the following inequalities: \begin{eqnarray} \label{eq:1} \deg(G^2) &\geq& \deg(G) \geq \frac{sd}{r}, \\ \label{eq:2} \operatorname{rank}(G^2) &\leq& \operatorname{rank}(G) - h^0(G) \leq \frac{sr'}{r}. \end{eqnarray} These imply that $\mu(G^2) - \mu(E') \geq 0.$ But $E'_*$ is parabolic stable, so by Proposition \ref{prop:num}, $E'$ is semistable and $\mu(G^2)=\mu(E').$ Thus, we must have equalities in equations (\ref{eq:1}) and (\ref{eq:2}), in particular $\mu(G) = \mu(E^\xi)$. But since $\mu(G_*) > \mu(E^\xi_*),$ we see that $G_*$ must inherit the weight $\alpha_2,$ which implies that $G^2_*$ also inherits $\alpha_2,$ and it now follows that $$\mu(G^2_*) - \mu(E'_*) = \frac{(s_2 -1) \alpha_1 + \alpha_2}{s_2} - \frac{(r'-1) \alpha_1 + \alpha_2}{r'} > 0,$$ where $s_2 = \operatorname{rank} G^2 < r'.$ This contradicts the parabolic stability of $E'_*$ and completes the proof of part (i). \medskip\noindent (ii) Since $\Leftarrow$ is true independent of the vanishing of $H^1,$ we only prove $\Rightarrow.$ Suppose $E^{\xi_1}_* \cong E^{\xi_2}_*$ and set $\pi_X(E^{\xi_i}_*) = e_i'=[{E^i_*}'] \in {\cal M}_{\alpha'\!,L}.$ Notice that $h^1(E^{\xi_i}_*) = 0,$ and so $h^0(E^{\xi_i}_*) = \chi(E^{\xi_i}_*) = r''.$ It follows that every holomorphic section of $E^{\xi_i}_*$ has its image contained in $I^{\oplus r''}_*.$ Hence any isomorphism $\varphi : E^{\xi_1}_* \longrightarrow E^{\xi_2}_*$ defines a commutative diagram $$\CD 0 @>>> I^{\oplus r''}_* @>>> E^{\xi_1}_* @>>> {E^1_*}' @>>> 0 \\ && @VV{\varphi''}V @VV{\varphi}V @VV{\varphi'}V \\ 0 @>>> I^{\oplus r''}_* @>>> E^{\xi_2}_* @>>> {E^2_*}' @>>> 0 \\ \endCD $$ where both $\varphi'$ and $\varphi''$ are isomorphisms, and so $\xi_2 = (\varphi' \times \varphi'') \cdot \xi_1.$ \end{pf} Part (i) of the claim and the universal property of ${\cal U}_*^\alpha$ gives a map ${\cal M} \stackrel{\psi}{\longrightarrow} {\cal M}_{\alpha,L},$ which is injective by part (ii). Since ${\cal M}$ is nonempty, $\dim {\cal M} = \dim {\cal M}_{\alpha,L},$ so rationality of ${\cal M}_{\alpha,L}$ follows from that of ${\cal M}.$ This concludes the proof in Case II. $\quad \Box$ \medskip \noindent {\it Remark.} We had originally hoped to prove rationality of ${\cal M}_{\alpha,L}$ with the weaker hypothesis that $\alpha$ is generic, but the argument does not hold in this generality. For consider the case $D=p.$ By tensoring with a line bundle and shifting, we can assume that $$r(g-1) < d \leq r(g-1) + m_1.$$ Hence, the subbundle split off in the induction is again a sum of parabolic line bundles with the same weights. The difficulty is in proving that the quotient $E'_*$ has {\it generic} weights $\alpha'.$ Proposition \ref{prop:fine} implies that $E'_*$ admits a generic weight if and only if the elements of the set $\{d, m'_i(p) \}$ greatest common divisor equal to one. The statement $$(d,m_1,\ldots,m_\kappa)=1 \Rightarrow (d,m'_1,\ldots,m'_\kappa)=1,$$ which is what we would need to prove here, is unfortunately false (notice that $m'_1=m_1-d+r(g-1)$ and $m'_i=m_i$ otherwise). \medskip\noindent {\it Acknowledgements.} Both authors would like to express their gratitude to the Max-Planck-Institut f\"ur Mathematik for providing financial support. The first author is also grateful to the Institut des Hautes \'Etudies Scientifique for partial support. We would also like to thank I.\ Dolgachev and L.\ G\"ottsche for helpful discussions.
1996-10-04T09:04:22
9610
alg-geom/9610005
en
https://arxiv.org/abs/alg-geom/9610005
[ "alg-geom", "dg-ga", "math.AG", "math.DG" ]
alg-geom/9610005
Sacha Sardo Infirri
Alexander V Sardo-Infirri
Resolutions of Orbifold Singularities and Flows on the McKay Quiver
LaTex2e, 57 pages with 1 table and 19 figures
null
null
null
null
Let $\Gamma$ be a finite group acting linearly on $\C^n$, freely outside the origin. In previous work a generalisation of Kronheimer's construction of moduli of Hermitian-Yang-Mills bundles with certain invariance properties was given. This produced varieties $X_\zeta$ (parameterised by $\zeta\in\Q^N$) which are partial resolutions of $\C^n/\Gamma$. In this article, it is shown the same $X_\zeta$ can be described as moduli spaces of representations of the McKay quiver associated to the action of $\Gamma$. It it shown that, for abelian groups, $X_\zeta$ are toric varieties defined by convex polyhedra which are the solution sets for a generalisation of the transportation problem on the McKay quiver. The generalised transportation problem is solved for an arbitrary quiver to give a description of the extreme points, faces, and tangent cones to the solution polyhedra in terms of certain distinguished trees in the quiver. Applied the McKay quiver, this gives an explicit procedure for calculating $X_\zeta $, its Euler number, and its singularities for any $\zeta$. The $\zeta$-parameter-space is thus partitioned into a finite disjoint union of cones inside which the biregular type of $X_\zeta$ remains constant. Finally, the example $\C^3/\Z_5$ (weights $1,2,3$) is worked out in detail, and figures of smooth and singular $X_\zeta$ and their corresponding flows are drawn. A further example of smooth crepant resolution $X_\zeta$ is drawn for the singularity ${1/11}(1,4,6)$.
[ { "version": "v1", "created": "Fri, 4 Oct 1996 06:59:13 GMT" } ]
2008-02-03T00:00:00
[ [ "Sardo-Infirri", "Alexander V", "" ] ]
alg-geom
\part{BlackBoard Bold} \newcommand{{\mathbb C}}{{\mathbb C}} \newcommand{{\C^*}}{{{\mathbb C}^*}} \newcommand{{\mathbb R}}{{\mathbb R}} \newcommand{\Z}{{\mathbb Z}} \newcommand{\Q}{{\mathbb Q}} \newcommand{{\mathbb N}}{{\mathbb N}} \newcommand{{\mathbb F}}{{\mathbb F}} \newcommand{{\mathbb H}}{{\mathbb H}} \newcommand{{\mathbb P}}{{\mathbb P}} \newcommand{_{\ge 0}}{_{\ge 0}} \newcommand{{\Q_+}}{{\Q_+}} \newcommand{\R\nneg}{{\mathbb R}_{\ge 0}} \newcommand{\protect\overline{T}{}}{\protect\overline{T}{}} \newcommand{\widetilde{M}}{\widetilde{M}} \newcommand{\widetilde{L}}{\widetilde{L}} \newcommand{\widetilde{P}}{\widetilde{P}} \newcommand{{\mathcal A}}{{\mathcal A}} \newcommand{{\mathcal B}}{{\mathcal B}} \newcommand{{\mathcal C}}{{\mathcal C}} \newcommand{{\mathcal D}}{{\mathcal D}} \newcommand{{\mathcal E}}{{\mathcal E}} \newcommand{{\mathcal F}}{{\mathcal F}} \newcommand{{\mathcal G}}{{\mathcal G}} \newcommand{{\mathcal H}}{{\mathcal H}} \newcommand{{\mathcal I}}{{\mathcal I}} \newcommand{{\mathcal J}}{{\mathcal J}} \newcommand{{\mathcal K}}{{\mathcal K}} \newcommand{{\mathcal L}}{{\mathcal L}} \newcommand{{\mathcal M}}{{\mathcal M}} \newcommand{{\mathcal N}}{{\mathcal N}} \newcommand{{\mathcal O}}{{\mathcal O}} \newcommand{{\mathcal P}}{{\mathcal P}} \newcommand{{\mathcal Q}}{{\mathcal Q}} \newcommand{{\mathcal R}}{{\mathcal R}} \newcommand{{\mathcal S}}{{\mathcal S}} \newcommand{{\mathcal T}}{{\mathcal T}} \newcommand{{\mathcal U}}{{\mathcal U}} \newcommand{{\mathcal V}}{{\mathcal V}} \newcommand{{\mathcal W}}{{\mathcal W}} \newcommand{{\mathcal X}}{{\mathcal X}} \newcommand{{\mathcal Y}}{{\mathcal Y}} \newcommand{{\mathcal Z}}{{\mathcal Z}} \newcommand{\boldsymbol{\alpha}}{\boldsymbol{\alpha}} \newcommand{\boldsymbol{\gamma}}{\boldsymbol{\gamma}} \newcommand{\boldsymbol{\omega}}{\boldsymbol{\omega}} \newcommand{{\mathbf a}}{{\mathbf a}} \newcommand{\bfa}{{\mathbf a}} \newcommand{{\mathbf e}}{{\mathbf e}} \newcommand{\bde}{{\mathbf e}} \newcommand{{\mathbf f}}{{\mathbf f}} \newcommand{{\mathbf g}}{{\mathbf g}} \newcommand{\bdh}{{\mathbf h}} \newcommand{{\mathbf A}}{{\mathbf A}} \newcommand{{\mathbf B}}{{\mathbf B}} \newcommand{{\mathbf C}}{{\mathbf C}} \newcommand{{\mathbf I}}{{\mathbf I}} \newcommand{{\mathbf J}}{{\mathbf J}} \newcommand{{\mathbf K}}{{\mathbf K}} \newcommand{{\mathbf Q}}{{\mathbf Q}} \newcommand{{\mathbf R}}{{\mathbf R}} \newcommand{{\mathbf V}}{{\mathbf V}} \newcommand{{\mathbf W}}{{\mathbf W}} \newcommand{{\mathfrak g}}{{\mathfrak g}} \newcommand{{\mathfrak u}}{{\mathfrak u}} \newcommand{{\mathbf g}}{{\mathbf g}} \newcommand{\gothk}{{\mathbf k}} \newcommand{{\mathbf t}}{{\mathbf t}} \newcommand{{\mathbf u}}{{\mathbf u}} \newcommand{\pi\mbox{-}\operatorname{\rm ext}}{\pi\mbox{-}\operatorname{\rm ext}} \newcommand{\circlearrowleft}{\circlearrowleft} \newcommand{\circlearrowright}{\circlearrowright} \newcommand{\upharpoonleft\downharpoonright}{\upharpoonleft\downharpoonright} \newcommand{\downharpoonleft\upharpoonright}{\downharpoonleft\upharpoonright} \newcommand{\centerdot}{\centerdot} \newcommand{M^{\Gamma }}{M^{\Gamma }} \newcommand{\GL^\Gamma\!}{\GL^\Gamma\!} \newcommand{\U^\Gamma\!}{\U^\Gamma\!} \newcommand{\cN^{\Gamma }}{{\mathcal N}^{\Gamma }} \newcommand{\cN^{\Gamma }}{{\mathcal N}^{\Gamma }} \newcommand{\cN^{\Gamma }\cap\mu^{-1}(0)}{{\mathcal N}^{\Gamma }\cap\mu^{-1}(0)} \newcommand{\cT_\zeta}{{\mathcal T}_\zeta} \newcommand{\cT^{\text{IC}}}{{\mathcal T}^{\text{IC}}} \newcommand{\cT^{\text{sing-IC}}}{{\mathcal T}^{\text{sing-IC}}} \newcommand{\cT_\zeta^{\text{IC}}}{{\mathcal T}_\zeta^{\text{IC}}} \newcommand{\cT_\zeta^{\text{sing-IC}}}{{\mathcal T}_\zeta^{\text{sing-IC}}} \newcommand{{\mathcal S}^{\text{IC}}}{{\mathcal S}^{\text{IC}}} \newcommand{\cT_{\zeta ,x}^{\text{IC}}}{{\mathcal T}_{\zeta ,x}^{\text{IC}}} \newcommand{C_{\zeta,x}}{C_{\zeta,x}} \newcommand{F_{\zeta,x}}{F_{\zeta,x}} \newcommand{T^\Pi}{T^\Pi} \newcommand{\overline}{\overline} \newcommand{\clos{S}}{\overline{S}} \newcommand{\clos{T}}{\overline{T}} \newcommand{\tilde \alpha}{\tilde \alpha} \newcommand{\widetilde X}{\widetilde X} \newcommand{\widetilde{X/\Gamma}}{\widetilde{X/\Gamma}} \newcommand{\Gamma}{\Gamma} \newcommand{\mbox{$\Gamma$}}{\mbox{$\Gamma$}} \newcommand{\widehat\Gamma}{\widehat\Gamma} \newcommand{\Lambda^2\to\Lambda^1}{\Lambda^2\to\Lambda^1} \newcommand{\Irr(\ga)}{\Irr(\Gamma)} \newcommand{\hom_\Gamma}{\hom_\Gamma} \newcommand{{\Rep_\cQ}}{{\Rep_{\mathcal Q}}} \newcommand{\contract}{\dashv} \newcommand{\bar{\partial}}{\bar{\partial}} \newcommand{\overline{D}}{\overline{D}} \newcommand{{\bar{q}}}{{\bar{q}}} \newcommand{{\bar{z}}}{{\bar{z}}} \newcommand{{\bar{\imath}}}{{\bar{\imath}}} \newcommand{{\bar{\jmath}}}{{\bar{\jmath}}} \newcommand{{\bar{k}}}{{\bar{k}}} \newcommand{{\bar{m}}}{{\bar{m}}} \newcommand{{\overline{\Pi}}}{{\overline{\Pi}}} \newcommand{{\overline w}}{{\overline w}} \newcommand{\dashrightarrow}{\dashrightarrow} \newcommand{\birationalto}{\dashrightarrow} \newcommand{\gitquot}[1]{/\!\!/_{\!#1}\,} \newcommand{\stackrel\pi\to}{\stackrel\pi\to} \newcommand{\stackrel\rho\to}{\stackrel\rho\to} \newcommand{>\!\!>}{>\!\!>} \newcommand{\squared}[1]{#1^2} \newcommand{\stackrel M\sim}{\stackrel M\sim} \newcommand{i.e.\xspace}{i.e.\xspace} \newcommand{c.f.\xspace}{c.f.\xspace} \newcommand{resp.\xspace}{resp.\xspace} \newcommand{neighbourhood}{neighbourhood} \newcommand{\text{s-group}}{\text{s-group}} \newcommand{geometric invariant theory}{geometric invariant theory} \newcommand{vector space}{vector space} \newcommand{McKay quiver}{McKay quiver} \newcommand{positive definite Hermitian inner product}{positive definite Hermitian inner product} \newcommand{asymptotically locally Euclidean}{asymptotically locally Euclidean} \newcommand{K\"ahler}{K\"ahler} \newcommand{\kah\ reduction}{K\"ahler\ reduction} \newcommand{singularity}{singularity} \newcommand{singularities}{singularities} \newcommand{Hermitian-Yang-Mills}{Hermitian-Yang-Mills} \DeclareMathOperator{\sign}{sign} \DeclareMathOperator{\mult}{mult} \DeclareMathOperator{\Map}{Map} \DeclareMathOperator{\conv}{conv} \DeclareMathOperator{\ord}{ord} \DeclareMathOperator{\Ind}{Ind} \DeclareMathOperator{\id}{id} \DeclareMathOperator{\pr}{pr} \DeclareMathOperator{\grad}{grad} \DeclareMathOperator{\Diff}{Diff} \DeclareMathOperator{\Vect}{Vect} \DeclareMathOperator{\pow}{pow} \DeclareMathOperator{\Image}{Im} \DeclareMathOperator{\rk}{rank} \DeclareMathOperator{\spn}{span} \DeclareMathOperator{\codim}{codim} \DeclareMathOperator{\Irr}{Irr} \DeclareMathOperator{\Sym}{Sym} \DeclareMathOperator{\trace}{trace} \DeclareMathOperator{\Proj}{Proj} \DeclareMathOperator{\Spec}{Spec} \DeclareMathOperator{\Exc}{Exc} \DeclareMathOperator{\Hilb}{Hilb} \DeclareMathOperator{\ch}{ch} \DeclareMathOperator{\td}{td} \DeclareMathOperator{\Pic}{Pic} \DeclareMathOperator{\adj}{ad} \DeclareMathOperator{\Hom}{Hom} \DeclareMathOperator{\End}{End} \DeclareMathOperator{\Aut}{Aut} \DeclareMathOperator{\Rep}{Rep} \DeclareMathOperator{\rad}{rad} \DeclareMathOperator*{\In}{In} \DeclareMathOperator*{\Out}{Out} \DeclareMathOperator{\supp}{supp} \DeclareMathOperator{\interior}{int} \DeclareMathOperator{\Star}{Star} \DeclareMathOperator{\Ext}{Ext} \DeclareMathOperator{\ver}{vert} \DeclareMathOperator{\ext}{ext} \DeclareMathOperator{\orb}{orb} \DeclareMathOperator{\Lie}{Lie} \DeclareMathOperator{\Cl}{Cl} \DeclareMathOperator{\Stab}{Stab} \DeclareMathOperator{\Fix}{Fix} \DeclareMathOperator{\Mat}{Mat} \DeclareMathOperator{\GL}{GL} \DeclareMathOperator{\gl}{\mathfrak{gl}} \DeclareMathOperator{\PGL}{PGL} \DeclareMathOperator{\PSL}{PSL} \DeclareMathOperator{\SL}{SL} \DeclareMathOperator{\PL}{PL} \DeclareMathOperator{\sll}{\mathfrak{sl}} \DeclareMathOperator{\U}{U} \DeclareMathOperator{\gu}{\mathfrak{u}} \newcommand{\gu}{\gu} \DeclareMathOperator{\SU}{SU} \DeclareMathOperator{\su}{\mathfrak{su}} \DeclareMathOperator{\PU}{PU} \DeclareMathOperator{\Vol}{Vol} \def\qsing#1/#2(#3){\frac{#1}{#2}(#3)} \def\<#1>{\langle#1\rangle} \newcommand{\ip}[2]{\langle#1,#2\rangle} \newcommand{\gen}[2]{\langle#1,\dots,#2\rangle} \newcommand{{}^{-1}}{{}^{-1}} \newcommand{{}^t\kern -0.2em}{{}^t\kern -0.2em} \newcommand{{}^{\vee}}{{}^{\vee}} \newcommand{{}^{\bot}}{{}^{\bot}} \newcommand{{}^{\ast}}{{}^{\ast}} \newcommand{{}_{\ast}}{{}_{\ast}} \newcommand{\sst}[1]{#1^{\protect\text{ss}}} \newcommand{\st}[1]{#1^{\protect\text{s}}} \newcommand{\nopagebreak}{\nopagebreak} \hyphenation{ortho-nor-mal} \newcommand{\cvec}[2]{{#1\choose#2}} \newcommand{\ntp}[3]{{{#1}_{#2},\dots,{#1}_{#3}}} \newcommand{\vect}[3]{{({#1}_{#2},\dots,{#1}_{#3})}} \newcommand{\nn}[3]{{{#1}^{#2},\dots,{#1}^{#3}}} \newcommand{\vv}[3]{{({#1}^{#2},\dots,{#1}^{#3})}} \newcommand{\ntup}[2]{#1,\dots,#2} \newcommand{\abs}[1]{\lvert#1\rvert} \newcommand{\norm}[1]{\lVert#1\rVert} \newcommand{\card}[1]{\sharp\,#1} \newcommand{\map}[5]{$$\begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array}$$} \newcommand{\map}{\map} \newcommand{\mapeq}[6]{\begin{equation} \begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array} \protect\label{#6} \end{equation}} \newcommand{\corresp}[4]{$$\begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array}$$} \newcommand{\correspeq}[5]{\begin{equation} \begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array} \protect\label{#5} \end{equation}} \newenvironment{point}[1]{\smallbreak\noindent{\bf #1\/}\nopagebreak}{\smallbreak} \newenvironment{empoint}[1]{\smallskip\noindent{\em #1\/}:}{\smallskip} \numberwithin{equation}{section} \def\*#1:#2*{\S\ref{sec:#1}.\ref{sec:#1:#2}} \newenvironment{Ventry}[1]% {\begin{list}{}{\renewcommand{\makelabel}[1]{\emph{##1:}\hfil}% \settowidth{\labelwidth}{{\emph{#1:}}}% \setlength{\leftmargin}{\labelwidth+\labelsep}}}% {\end{list}} \newcommand{\entrylabel}[1]{\mbox{\emph{#1:}}\hfil} \newenvironment{entry} {\begin{list}{}% {\renewcommand{\makelabel}{\entrylabel}% \setlength{\labelwidth}{12mm}% \setlength{\leftmargin}{14mm}% }% }% {\end{list}} \newlength{\Mylen} \newcommand{\Lentrylabel}[1]{% \settowidth{\Mylen}{\emph{#1}}% \ifthenelse{\lengthtest{\Mylen > \labelwidth}}% {\parbox[b]{\labelwidth {\makebox[0pt][l]{\emph{#1}}\\}}% {\emph{#1} \hfil\relax} \newenvironment{Lentry} {\renewcommand{\entrylabel}{\Lentrylabel}% \begin{entry}} {\end{entry}} \newcommand{\Pentrylabel}[1]{% {\emph{#1}} \hfil\relax} \newenvironment{Pentry} {\renewcommand{\entrylabel}{\Pentrylabel}% \begin{entry}} {\end{entry}} \newcommand{\Mentrylabel}[1]% {\raisebox{0pt}[1ex][0pt]{\makebox[\labelwidth][l]% {\parbox[t]{\labelwidth}{\hspace{0pt}\emph{#1:}}}}} \newenvironment{Mentry}% {\renewcommand{\entrylabel}{\Mentrylabel}\begin{entry}}% {\end{entry}} \begin{document} \title[Orbifold Singularities and McKay Flows]{\titlestring\footnote{Maths Subject Classification (1991): 14M25 (Primary) 05C35, 90C08, 32S45 (Secondary)}} \author{Alexander V.\ Sardo Infirri} \email{[email protected]} \address{Research Institute for Mathematical Sciences\\ Ky\=oto University\\ Oiwake-ch\protect\=o\\ Kitashirakawa\\ Saky\protect\=o-ku\\ Ky\=oto 606-01\\ Japan} \date{2 October 1996} \begin{abstract} Let~$\Gamma$ be a finite group acting linearly on~${\mathbb C}^n$, freely outside the origin, and let $N$ be the number of conjugacy classes of $\Gamma$ minus one. In~\cite{sacha:thesis,sacha:ale} a generalisation of Kronheimer's construction \cite{kron:ale} of moduli of Hermitian-Yang-Mills bundles with certain invariance properties was given. This produced varieties $X_\zeta$ (parameterised by $\zeta\in\Q^N$) which are partial resolutions of ${\mathbb C}^n/\Gamma$. In this article, it is shown that $X_\zeta$ can be described as moduli spaces of representations of the McKay quiver associated to the action of $\Gamma$, subject to certain natural commutation relations. This allows a complete description of these varieties in the case when $\Gamma$ is abelian. They are shown to be toric varieties corresponding to convex polyhedra which are the solution sets for a generalisation of the transportation problem on the McKay quiver. The generalised transportation problem is solved for a general quiver to give a description of the extreme points, faces, and tangent cones to the solution polyhedra in terms of certain distinguished trees in the underlying graph to the quiver. Applied to the case of the McKay quiver, this gives an explicit computational procedure for calculating $X_\zeta $, its Euler number, and giving a complete list of the singularities which can occur for all $\zeta$. The $\zeta$-parameter-space $\Q^N$ is thus partitioned into a finite disjoint union of cones inside which the biregular type of $X_\zeta$ remains constant. Passing from one cone to the other correponds to a birational transformation. The example ${\mathbb C}^3/\Z_5$ (weights $1,2,3$) is worked out in detail: there are two types of $\zeta$-cones: ones for which $X_\zeta $ is a smooth resolution, and others where it has a singularity isomorphic to a cone over a quadric in ${\mathbb C}^4$. This gives some evidence for the conjecture expressed in~\cite{sacha:ale} acording to which the singularities of $X_\zeta$ are at most quadratic for a generic $\zeta$. Computer calculations also show that the cases where $\Gamma\subset SU(3)$, and $|\Gamma|\leq 11$ yield crepant $X_\zeta$. For generic $\zeta$, the Euler number of $X_\zeta$ is also equal to $|\Gamma|$, and the $X_\zeta$ give smooth crepant resolutions fo the singularity. A further example of a picture of the polyhedron corresponding to $X_\zeta$ is drawn for the singularity $\frac{1}{11}(1,4,6)$. \end{abstract} \maketitle \tableofcontents \setcounter{section}{-1} \section{Introduction} \label{sec:intro} This paper is concerned with affine \emph{orbifold singularities}, namely with singularities of the type $X={\mathbb C}^n/\Gamma$ for $\Gamma$ a finite group acting linearly on ${\mathbb C}^n$. In~\cite{sacha:thesis,sacha:ale} a method using moduli of invariant Hermitian-Yang-Mills bundles was given for constructing partial resolutions of $X$ carrying natural asymptotically locally Euclidean (ALE) metrics. In this article, the same construction is described in terms of representation moduli of quivers. This allows a complete description of the moduli to be given for the case of abelian groups using the language of toric varieties. It turns out that the convex polyhedra describing the moduli are the solutions of a generalisation of a well-known network optimization problem (the transportation problem) on the McKay quiver, in which the quiver plays the role of a network where commodities are transported, and the parameter $\zeta$ specifies the supplies and demands at each vertex. \subsection{Background} \label{sec:intro:back} The resolution of the Kleinian singularities ${\mathbb C}^2/\Gamma$ for $\Gamma$ a finite subgroup of $\SL(2)$ is a classical subject; their minimal resolution $\widetilde X$ was first constructed by Du Val, and Brieskorn~\cite{briesk} showed that the components of the exceptional divisors form graphs which are dual to the homogeneous Dynkin diagrams for the Lie algebras $A,D,E$. In 1980, McKay~\cite{mckay:graphs} remarked that this establishes a correspondence between the extended homogeneous Dynkin diagrams $\overline A,\overline D,\overline E$ and the irreducible representations of $\Gamma$. More recently~\cite{reid_ito} another correspondence was constructed between conjugacy classes ``of weight 1'' and crepant divisors for any quotient singularity generated by a finite subgroup of $\SL(n)$. In the case where $\Gamma\subset\SL(3)$, Dixon, Harvey, Vafa and Witten proposed a definition of the orbifold Euler characteristic~\cite{dhvw:i,dhvw:ii} and conjectured the existence of smooth resolutions with trivial canonical bundle (i.e.\xspace which are crepant) and whose Euler number is given by the DHVW orbifold Euler number. The final cases in the proof of this conjecture were only recently completed \cite{mar_ols_per,roan:mirror_cy,mark:res_168,roan:res_a5,ito:trihedral,roan:crepant}. For the case $\Gamma\subset\SL(4)$, the author has obtained some interesting analogous results if one considers terminalisations rather than resolutions~\cite{sacha:sl4}. A promising link between the representation theory of $\Gamma$ and the construction of resolutions was established in 1986; Kronheimer~\cite{kron:thesis,kron:ale} constructed a family of hyper-K\"ahler\ quotients $X_\zeta $ by looking at a dimensional reduction of the Hermitian Yang-Mills (HYM) equations on a $\Gamma$-equivariant bundle over ${\mathbb C}^2$. He used the representation theory of $\Gamma$ and McKay's observation to prove that, for generic $\zeta$, these quotients are isomorphic to the minimal resolution $\widetilde X$. The author's thesis~\cite{sacha:thesis} (from which the present paper and its companion~\cite{sacha:ale} are in large part extracted) grew out of the ambition to generalise the results as far as possible to dimensions higher than two. Despite the fact that there is little hope for the results for $\GL(n)$ and general $n$ to be as strong as those for $\SL(2)$ groups, some of the nice properties of the $\SL(2)$ case generalise to the $\GL(n)$ case.\footnote{Some interesting new aspects are also present for the $\SL(3)$ case~\cite{sacha:thesis}.} In~\cite{sacha:thesis} several descriptions were given of the generalised construction. Probably the most concise is the one given in~\cite{sacha:ale}: construct moduli spaces $X_\zeta$ of instantons on the trivial bundle ${\mathbb C}^n\times R\to{\mathbb C}^n$. Here $R$ denotes the regular representation space for the group $\Gamma$, and $\zeta$ is a linearisation of the bundle action. The instantons are required to satisfy Hermitian-Yang-Mills-type equations, as well as additional $\Gamma$-equivariance and translation-invariance properties. In the present paper, an entirely different view point is adopted, namely that of considering $X_\zeta$ are representation moduli of the McKay quiver. More precisely, one considers representations $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ of the McKay quiver ${\mathcal Q}$ into the multiplicity space ${\mathbf R}$ of the regular representation of $\Gamma$, subject to certain commutation relations ${\mathcal K}$. A reductive group $\PGL({\mathbf R})$ acts on the above space, and one obtains GIT quotients $X_\zeta$ which depend on a rational parameter $\zeta$. \subsection{Main Results} \label{sec:intro:main} This paper assumes that $\Gamma\subset\GL(n)$ acts on ${\mathbb C}^n$ freely outside the origin for any $n\geq 2$, which means that $X={\mathbb C}^n/\Gamma$ has an isolated singularity.\footnote{This is for the purpose of simplicity --- the method would seem to be applicable to the general case with some modifications.} The main results are as follows. Firstly, $X_\zeta$ are identified with representation moduli of the McKay quiver. This is fairly straight-forward and involves mostly translating quiver concepts over to the language used in~\cite{sacha:ale}. There are two further sets of results. \subsubsection{Toric description of $X_\zeta$} The first is a toric description of the representation variety $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ and its moduli $X_\zeta$. The convex polyhedron $C_\zeta$ corresponding to $X_\zeta$ is identified with the projection to ${\mathbb R}^n$ of the solution polyhedron for the transportation problem on the McKay quiver. The transportation problem is a well-known linear network optimization problem~\cite{kenn_helg,gond_mino:graphs}: given an assignment of real numbers to the vertices of a quiver\footnote{The term \emph{network} is usually used in this context instead of quiver.} ${\mathcal Q}$ (thought of as representing the demand and supply of certain commodities), the aim is to find an assignment of non-negative real numbers to the arrows (a \emph{flow} on ${\mathcal Q}$), in such a way that the demands and supplies at each vertex are satisfied (i.e.\xspace the equation $\partial f=\zeta$ holds). For any given assignment of weights $\zeta$, the solution set is a convex polyhedron inside ${\mathbb R}^{{\mathcal Q}_1}$ denoted by $F_\zeta$. The notation in use is as follows: \begin{notation} Let $\Gamma$ be the cyclic group of order $r$ acting freely outside the origin in $Q$ with weights $w_1,\dots,w_n\in\Z_r=\Z/r\Z$ i.e.\xspace via \map{\rho}{\mu_r\subset{\C^*}}{{\C^*}^n}{\lambda}{\begin{pmatrix} \lambda^{w^1} \\ & \lambda^{w^2} \\ &\\ & & & \lambda^{w^n} \end{pmatrix}.} The McKay quiver ${\mathcal Q}$ of $\Gamma$ has vertices ${\mathcal Q}_0=\Z_r$, and arrows $a_v^i:=v\to v-w_i \pmod r$ for each $v\in\Z_r$ and $i=1,\dots,n$. Let $\pi\colon{\mathcal Q}_1\to\{1,\dots,n\}$ be the map defined by $\pi(a_v^i):=i$ and let $\pi\colon{\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^n$ be the induced linear map defined by mapping the standard bases. (Here and elsewhere we use the exponential notation ${\mathbb R}^{{\mathcal Q}_1}:=\Map({\mathcal Q}_1,{\mathbb R})$. The standard basis of ${\mathbb R}^{{\mathcal Q}_1}$ is given by the ``indicator'' functions $\chi_a$ defined by $\chi_a(a')=1$ if $a=a'$ and $\chi_a(a')=0$ otherwise). Let $\Pi$ be the sub-lattice of $\Z^n$ of index $r$ defined by $$\Pi := \ker\hat\rho=\{ x\in \Z^n | \sum x_iw_i\equiv 0\pmod r\}.$$ For each element $\zeta\in\Z^{{\mathcal Q}_0}$ such that $\sum_{v\in{\mathcal Q}_0} \zeta(v)=0$, consider the corresponding solution polyhedron $F_\zeta$ to the transportation problem on ${\mathcal Q}$, and let $C_\zeta=\pi F_\zeta$ be its projection to ${\mathbb R}^n$. \end{notation} Adopting the above notation, one can state the first main theorem. \begin{nonumberthm}[c.f.\xspace Thms.\ \ref{thm:toric-d-moduli} and \ref{thm:slices}] The moduli $X_\zeta$ are isomorphic to the toric varieties $$X_\zeta\cong\protect\overline{T}{}^{\Pi,C_\zeta},$$ where $C_\zeta$ correspond to the projection to ${\mathbb R}^n$ of the solution polyhedra to the transportation problem on ${\mathcal Q}$. \end{nonumberthm} \subsubsection{Generalised Transportation Problem} \label{sec:intro:ab:trans} The problem of determining $C_\zeta$ can be considered as a generalised transportation problem. The solution to the classical transportation problem is well known: \begin{nonumberthm}[Classical Transportation Problem (c.f.\xspace Theorem~\ref{thm:classical})] The extreme points of $F_\zeta$ are precisely those flows in $F_\zeta$ whose supports\footnote{The support of a flow is the set of arrows on which it is non-zero.} are \emph{trees}, i.e.\xspace contain no cycles. \end{nonumberthm} The basic idea behind the proof of this theorem is to associate to each cycle $c$ a corresponding flow $\tilde\chi_c$ as follows. A \emph{cycle} in ${\mathcal Q}$ is a sequence $c_1,\dots,c_m$ of arrows such that they form a cycle when their orientation is disregarded. The disjoint union of the arrows $c_i$ whose direction agrees (resp.\xspace disagrees) with the ordering $c_1,\dots,c_k$ is called the \emph{positive\/} (resp.\xspace \emph{negative\/}) part of $c$ and is denoted it by $c^+$ (resp.\xspace $c^-$). For each arrow $a\in{\mathcal Q}_1$, let $\chi_a$ denote the flow which takes the value $1$ on the arrow $a$ and zero elsewhere. To each cycle $c$, one can define the \emph{basic flow} associated to $c$ by $$\tilde\chi_c:=\sum_{c_i\in c^+}\chi_{c_i} - \sum_{c_i\in c^-}\chi_{c_i}\in\Z^n.$$ Note that for any cycle $c$, $\partial \tilde\chi_c=0$, i.e.\xspace the associated flow does not contribute anything at any vertex. If $f\in F_\zeta$ contains $c$ in its support, then one can add $\pm\epsilon\tilde\chi_c$ to $f$ for some small $\epsilon$ and still remain in $F_\zeta$. Hence $f$ cannot be an extreme point. Vice-versa, if $f$ is not extreme, then one can reconstruct a cycle in its support. The generalisation of this theorem to describe the projection $C_\zeta=\pi(F_\zeta)$ is quite straight-forward. One begins by defining the \emph{type} of any given cycle as the element $\pi(\tilde\chi_c)\in\Z^n$. Cycles of type $0\in\Z^n$ play a special role, much as ordinary cycles do in the classical case. \begin{dfn} The \emph{closure} of $S\subset{\mathcal Q}_1$ is defined to be the smallest over-set $\overline{S}\supseteq S$ such that $$c^-\subseteq \overline{S}\iff c^+\subseteq\overline{S},$$ for all cycles $c$ of type zero. Two configurations $S,S'$ will be called \emph{equivalent} (written $S\sim S'$) if $\overline{S}=\overline{S'}$. \end{dfn} Th same methods as in the classical case allows one to prove the following generalisation. \begin{nonumberthm}[Extreme Points of $C_\zeta$] The extreme points of $C_\zeta$ are the images under $\pi$ of precisely those flows in $F_\zeta$ whose support contains no cycles of non-zero type. \end{nonumberthm} In fact, by taking into account all cycles in the support of a given flow, it is possible to determine the dimension of the face which $\pi(f)$ belongs to. Some additional terminology will be convenient to state the results. Let ${\mathcal C}$ denote the set of all \emph{configurations}, i.e.\xspace the set of non-empty subsets of ${\mathcal Q}_1$. For a configuration $S\subset{\mathcal Q}_1$, let $F_0(S)$ be the cone generated by the flows $\tilde\chi_c$ for the cycles $c$ whose negative part is included in $S$ and let $Z_0(S)$ be its maximal vector subspace (i.e.\xspace the subspace generated by the flows $\tilde\chi_c$ for the cycles $c\subseteq S$). The \emph{rank} of $S$ is defined to be the dimension of $\pi Z_0({\overline S})$. The set of configurations (resp.\xspace trees) of rank $k$ is denoted ${\mathcal C}^k$ (resp.\xspace ${\mathcal T}^k$). Also, write ${\mathcal C}_\zeta$ (resp.\xspace ${\mathcal T}_\zeta$) for the subset of configurations (resp.\xspace trees) which are \emph{admissible for $\zeta$},\/ namely configurations (resp.\xspace trees) $S\in{\mathcal C}$ which arise as the support of some element in $F_\zeta$. \begin{nonumberthm}[c.f.\xspace Theorem \ref{thm:faces}] \label{thm:intro:faces} For all $\zeta$, the map \map{\text{Face}_\zeta}{{\mathcal C}_\zeta}{\text{Faces of }\pi F_\zeta}{S}{\pi F_\zeta \cap(\pi f+\pi Z_0(\overline{S})))} is independent of the choice of $f\in F_\zeta\cap\supp^{-1}(S)$, and induces a bijection $$\text{Face}_\zeta\colon {\mathcal C}^k_\zeta/\!\!\sim \xrightarrow{\;\cong\;} \text{$k$-faces of }\pi F_\zeta.$$ Furthermore, for all $[S]\in{\mathcal C}^k_\zeta/\!\!\sim$, $$T_{\text{Face}_\zeta(S)}\pi F_\zeta =\pi F_0(\overline{S}),$$ where the left-hand side denotes the tangent cone to the polyhedron $\pi F_\zeta$ at the face $\text{Face}_\zeta(S)$. In other words, $\pi F_0(\overline{S})$ gives the tangent cone corresponding to the configuration $S$ (which is independent of the value of $\zeta$) and $\text{Face}_\zeta(S)$ gives the corresponding face of $C_\zeta$ (whose direction is also independent of $\zeta$). \end{nonumberthm} This theorem has a number of important corollaries. \begin{nonumbercor}[c.f.\xspace Cor.\ \ref{cor:iso_C_zeta}] \label{cor:intro:iso_C_zeta} If $\zeta$ and $\zeta'$ have the same admissible configurations (${\mathcal C}_\zeta={\mathcal C}_{\zeta'}$) or even just the same admissible trees (${\mathcal T}_\zeta={\mathcal T}_{\zeta'}$) then the corresponding polyhedra $C_{\zeta}$ and $C_{\zeta'}$ are geometrically isomorphic. Two polyhedra are said to be \emph{geometrically isomorphic} if they are combinatorially isomorphic and their tangent cones at the corresponding faces are identical. In particular, their associated fans and toric varieties must be identical. \end{nonumbercor} \begin{nonumbercor}[c.f.\xspace Cor. \ref{cor:number_extreme_pts}] \label{cor:intro:number_extreme_pts} Let ${\mathcal T}\subset{\mathcal C}$ denote the configurations which are {\em trees,\/} and let $\ext C_\zeta$ denote the extreme points of $C_\zeta$. Then $$\card{\ext C_\zeta}=\card{{\mathcal C}^0_\zeta/\!\!\sim}=\card{{\mathcal T}^0_\zeta/\!\!\sim}.$$ \end{nonumbercor} \begin{rmk} Since any tree admits a unique flow such that $\partial f = \zeta$, it is very easy to determine ${\mathcal T}_\zeta$. Furthermore, if $\zeta$ is generic, then ${\mathcal C}_\zeta\subset{\mathcal C}_{\text{span}}$, the subset of \emph{spanning configurations},\/ i.e.\xspace configurations whose arrows join any two vertices of the quiver (not necessarily in an oriented way). In this case, $T\in{\mathcal T}^0_\zeta$ if and only if the tree $T$ admits an assignment of elements of $\Z^n$ to the vertices which satisfies particular properties (a so-called \emph{$n$-weighting} in the terminology of Section~\ref{sec:2:examples:weight:fixed}). This condition can again be checked by a very simple algorithm. To sum up: the above corollary allows one to determine the extreme points of $C_\zeta$ for any $\zeta$ very easily. \end{rmk} \begin{nonumbercor}[c.f.\xspace Cor.\ \ref{cor:fan} and Lemma \ref{lemma:generic_flow}] \label{cor:intro:fan} The extreme points of the polyhedron $\pi F_\zeta$ correspond to the trees $T$ in ${\mathcal T}^0_\zeta$ and the tangent cone to $\pi F_\zeta$ at the point corresponding to $T$ is $\pi F_0(\clos{T})=\pi F_0(T)$. Thus the fan associated to the polyhedron $\pi F_\zeta $ is given by the dual cones $\pi F_0(T)^\vee$ for the trees $T\in{\mathcal T}^0_\zeta$ and all their faces. \end{nonumbercor} The theorem and corollaries above allow a complete understanding of the extreme points, faces, and tangent cones to the solution polyhedra of a generalised transportation problem. Applied to the case of the McKay quiver, they allow one to determine the Euler number and the singularities of $X_\zeta$ as well as giving a complete list of the singularities which can occur for all $\zeta$. Several interesting questions regarding these moduli nevertheless remain, such as whether they produce smooth (resp.\xspace terminal) resolutions in the $\SL(3)$ (resp.\xspace $\SL(4)$) case. A conjecture is given in the companion paper~\cite{sacha:ale} and several examples are computed in Section~\ref{sec:2:examples} of this paper. Depending on the point of view, isomorphic objects are denoted by different notations. For ease of reference, Table~\ref{tab:notation} gives a correspondence table between the point of view in~\cite{sacha:ale} and in Parts~\ref{sec:1} and~\ref{sec:2} of this paper. \renewcommand{\arraystretch}{1.4} \begin{table}[htbp] \begin{center} \leavevmode \setlength{\fboxrule}{0pt} \begin{tabular}{|l|l|l|} \hline Moduli of Bundles & Moduli of Representations& Toric Varieties\\ \cite{sacha:ale} & Part 1 of this paper & Part 2 of this paper\\ \hline $M^\Gamma $&$ \Rep_{\mathcal Q}({\mathbf R}) $&$ \protect\overline{T}{}^{\Lambda^1_{\mathbb R},\Lambda^1_{{\mathbb R}+}}={\mathbb C}^{{\mathcal Q}_1}$\\ \hline ${\mathcal N}^\Gamma $&$ \Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R}) $&$ \protect\overline{T}{}^{\Lambda,C}$\\ \hline $G^\Gamma=\PGL^\Gamma(R) $&$ \PGL({\mathbf R}) $&$ \protect\overline{T}{}^{\Lambda^{0,0}}=\protect\overline{T}{}^{0,0}$\\ \hline $X_\zeta $&$ {\mathcal M}_\zeta $& $\protect\overline{T}{}^{\Pi,C_\zeta}$\\ \hline \end{tabular} \vspace{1em} \caption{Correspondence between the notations in use in this paper and in the companion paper~\cite{sacha:ale}.} \label{tab:notation} \end{center} \end{table} \subsection{Methods and Outline} \label{sec:intro:outline} The methods used to prove the main results are notationally cumbersome due to the quiver notation, but on the whole straightforward. A familiarity with toric geometry will make reading easier, although all the required facts and notation are explained when needed. The paper is divided into two parts; the first dealing with general groups, and the second specializing to the case of abelian ones. Each part begins with a section summarizing the notation in use. \subsubsection{Part \ref{sec:1}} Section~\ref{sec:1:notation} summarises the notation. Section~\ref{sec:1:quivers} outlines the basic facts and notation concerning quivers, their representations and their moduli. The necessary details regarding geometric invariant theory (GIT) quotients are also given. Section~\ref{sec:1:mckay} explains the case of the McKay quiver associated to the representation of a finite group. The McKay correspondence is briefly mentioned, and the ``canonical'' commutation relations ${\mathcal K}$ are defined. \subsubsection{Part \ref{sec:2}} Section~\ref{sec:2:notation} summarises the notation. Section~\ref{sec:2:abel} specialises further the discussion from Part~\ref{sec:1} to the case of abelian groups. The McKay quiver and commutation relations for these can be described quite simply and some specialised notation is introduced for this purpose. The representation variety and its moduli are proved to be toric varieties corresponding to an $(n+\card{\Gamma}-1)$-dimensional convex cone $C$ and to its $n$-dimensional convex polyhedral slices $C_\zeta$ respectively. From Section~\ref{sec:2:flow} onwards, the focus is on the transportation problem; its solution polyhedra coincide with $C_\zeta$ in the case of the McKay quiver. Section~\ref{sec:2:flow} explains the ordinary transportation problem on a network. Section~\ref{sec:2:gtp} explains the generalised transportation problem and states the theorems describing its solution polyhedra. An example of an application to the McKay quiver and several important corollaries are given in Section~\ref{sec:2:mckay}. The proofs of the theorems are given in Section~\ref{sec:2:flow:proofs}, except for the proof of some technical lemmas regarding flows which are left until Section~\ref{sec:2:exact}. Section~\ref{sec:2:sing} contains a discussion of the singularities of $X_\zeta$. The paper concludes in Section~\ref{sec:2:examples} with some practical examples and computations. \subsection{Acknowledgments} \label{sec:intro:ack} The present paper and its companion paper~\cite{sacha:ale} consist mostly\footnote{Minor portions have been rewritten to include references to advances in the field made since then (notably~\cite{ito:trihedral,roan:crepant,reid_ito}).} of excerpts of my D.Phil.\ thesis~\cite{sacha:thesis}, and I wish to acknowledge the University of Oxford and Wolfson College for their hospitality during its preparation. I am grateful to the Rhodes Trust for financial support during my first three years, and to Wolfson College for a loan in my final year. The conversion from thesis to article format was done while I was a Research Assistant in RIMS, Kyoto. I also take the opportunity to thank my supervisors Peter Kronheimer and Sir~Michael Atiyah who provided me with constant advice, encouragement and support and whose mathematical insight has been an inspiration. I also wish to thank William Crawley-Boevey, Michel Brion, Gavin Brown, Jack Evans, Partha Guha, Katrina Hicks, Frances Kirwan, Alistair Mees, Alvise Munari, Martyn Quick, David Reed, Miles Reid, Michael Thaddeus, and, last but not least, my parents and family. \section{Summary of Notation for Part \ref{sec:1}} \label{sec:1:notation} \subsection{General} \begin{Pentry} \item[$k$] Algebraically closed field. \item[$\Z_+$] Non-negative integers. \item[${\mathbb R}_+$] Non-negative reals. \end{Pentry} \subsection{Quivers, Representations} \begin{Pentry} \item[$Q$] $n$-dimensional complex vector-space. \item[$\Gamma$] Finite sub-group of $\SL(Q)$ \item[${\mathcal Q}$] Generic quiver ${\mathcal Q}=({\mathcal Q}_0,{\mathcal Q}_1)$.\\ From Section~\ref{sec:1:mckay} onwards, ${\mathcal Q}={\mathcal Q}_{\Gamma,{\mathcal Q}}$, the McKay quiver associated to $(\Gamma,Q)$. \item[${\mathcal Q}_0$] Vertices of ${\mathcal Q}$; in the case of the McKay quiver, elements of ${\mathcal Q}_0$ index the irreducible representations $R_v$ of $\Gamma$. \item[${\mathcal Q}_1$] Arrows of ${\mathcal Q}$. \item[$R_v$] Irreducible representation of $\Gamma$ corresponding to vertex $v$ of the McKay quiver. \item[$R$] Regular representation of $\Gamma$; $R=\oplus_{v\in{\mathcal Q}_0} {{\mathbf R}_v}\otimes R_v$. \item[${\mathbf R}_v$] The trivial $\Gamma$-module giving the multiplicity of $R_v$ in $R$. \item[${\mathbf R}$] Multiplicity space for $R$; ${\mathbf R}=\oplus_{v\in{\mathcal Q}_0}{{\mathbf R}_v}$ \item[${\mathbf V}$] Generic representation space of ${\mathcal Q}$; ${\mathbf V}=\oplus_{v\in{\mathcal Q}_0}{\mathbf V}_v$. \item[$R_{\mathbf V}$] The $\Gamma$-module $\oplus_{v\in{\mathcal Q}_0} {\mathbf V}_v\otimes R_v$ corresponding to ${\mathbf V}$. \item[$\alpha$] $\Gamma$-invariant element of $(Q\otimes\End R)$. \item[$q_i$] Basis of $Q$ ($i=1,\dots,n$). \item[$\alpha_i$] Component of $\alpha$ with respect to $\{q_i\}$: $\alpha=\sum_{i=1}^n q_i\otimes\alpha_i$. \item[$\tilde\alpha$] Corresponding representation of the McKay quiver into ${\mathbf R}$. \item[$a$] Arrow of ${\mathcal Q}$; also written $t(a)\to h(a)$, where $t(a)$ denotes the tail and $h(a)$ the head of $a$. \item[$\tilde\alpha_a$] Value of $\tilde\alpha$ on the arrow $a$; $\tilde\alpha_a\colon {\mathbf R}_{t(a)}\to {\mathbf R}_{h(a)}$. \item[${\mathcal K}$] Relations on ${\mathcal Q}$. \item[$\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$] Space of all representations of ${\mathcal Q}$ into ${\mathbf R}$ satisfying the relations ${\mathcal K}$. \item[$\PGL({\mathbf R})$] The isomorphism group for representations of ${\mathcal Q}$ into ${\mathbf R}$ $$\PGL({\mathbf R}):=\times_{v\in{\mathcal Q}_0}\GL({{\mathbf R}_v})/{\C^*}.$$ \item[$\mathfrak{k}$] The Lie algebra of $\PU({\mathbf R})$ (a real form of $\PGL({\mathbf R})$). \item[${\mathcal M}_{{\mathcal Q},{\mathcal K},\zeta}$] Representation moduli of $({\mathcal Q},{\mathcal K})$ $${\mathcal M}_{{\mathcal Q},{\mathcal K},\zeta}:= \Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})\gitquot{\chi}\PGL({\mathbf R}).$$ The linearisation $\chi$ is related to $\zeta$ by $\zeta=d\chi(1)_{|\mathfrak{k}}$. \item[$X_{\zeta}$] The moduli above in the case when ${\mathcal Q}$ is the McKay quiver, and ${\mathcal K}$ are the commutation relations defined in Section~\ref{sec:mckay:mckay:ccr:relns}. \item[$\rho_{\zeta}$] The natural projective morphism $X_{\zeta}\to X_{0}$ (a partial resolution). \end{Pentry} \part{General Groups} \label{sec:1} \section{Quivers, Relations and Representation Moduli} \label{sec:1:quivers} \subsection{Quivers} \label{sec:1:quivers:quivers} A \emph{quiver} is an oriented graph, possibly with multiple arrows between the same vertices and with loops (arrows which begin and end at the same vertex). Formally, a quiver ${\mathcal Q}$ consists of a pair of finite sets ${\mathcal Q}=({\mathcal Q}_0,{\mathcal Q}_1)$ with two maps ${\mathcal Q}_1 \overset{h}{\underset{t}\rightrightarrows} {\mathcal Q}_0$; the elements of ${\mathcal Q}_0$ are called \emph{vertices} and those of ${\mathcal Q}_1$ are called \emph{arrows}.\/ The elements $t(a)$ and $h(a)$ are called the \emph{tail\/} and \emph{head\/} of the arrow $a\in{\mathcal Q}_1$ respectively. The arrow $a$ is also sometimes denoted $t(a)\to h(a)$. \subsection{Representations} \label{sec:1:quivers:reps} Let $k$ be an algebraically closed field. A \emph{representation of a quiver} is a realization of its diagram of vertices and arrows in some category: it corresponds to replacing the vertices by objects and the arrows by morphisms between the objects. From now on, only the category of $k$-vector-spaces is considered; a representation ${\mathbf V}$ of a quiver ${\mathcal Q}$ is thus taken to be a collection of finite dimensional vector-spaces ${\mathbf V}_v$, indexed by the vertices $v\in {\mathcal Q}_0$, and of linear maps ${\mathbf V}_{v\to v'}\colon {\mathbf V}_{v}\to {\mathbf V}_{v'},$ indexed by the arrows $v\to v' \,\in{\mathcal Q}_1$. The set of all representations of ${\mathcal Q}$ into a fixed ${\mathcal Q}_0$-graded vector-space ${\mathbf V}=\oplus_{v\in {\mathcal Q}_0} {\mathbf V}_v$ forms a vector-space, denoted $\Rep_{\mathcal Q} {\mathbf V}$. \begin{example} \label{ex:endo} Representations of the quiver ${\mathcal Q}$ with one vertex and one loop are just endomorphisms of a vector-space. \end{example} \begin{example} A \emph{primitive} representation of ${\mathcal Q}$ is an element of $$\Rep_{{\mathcal Q}}(k^{{\mathcal Q}_0}),$$ where $k^{{\mathcal Q}_0}$ denotes the free $k$-vector-space on the vertices. A primitive representation thus corresponds to an assignment of an element of $k$ to each arrow in ${\mathcal Q}$, i.e.\xspace to an element of the vector-space $k^{{\mathcal Q}_1}$. \label{ex:primitive} \end{example} \subsection{Relations} \label{sec:1:quivers:relations} If the morphisms ${\mathbf V}_{v\to v'}$ are required to satisfy relations between them, then one talks about a \emph{representation of a quiver with relations}. More formally, a \emph{relation} is defined as a formal sum of paths in a quiver. A (non-trivial) \emph{path} $p$ is a sequence $a_1\dots a_n$ of arrows which \emph{compose}, i.e.\xspace such that $t(a_i)=h(a_{i+1})$ for $1\leq i < n$: $$\stackrel{h(p)}\bullet \xleftarrow[{a_1}]{} \bullet \xleftarrow[{a_2}]{} \dots \xleftarrow[{a_n}]{}\stackrel{t(p)}\bullet.$$ The vertex $h(a_1)$ ($t(a_n)$) is called the \emph{head} (\emph{tail}) of the path $p$ and denote it by $t(p)$ ($h(p)$). If ${\mathbf V}$ is a representation of ${\mathcal Q}$, then there is an induced morphism $${\mathbf V}(p)={\mathbf V}_{a_1}{\mathbf V}_{a_{2}}\dots {\mathbf V}_{a_n}\colon {\mathbf V}_{t(p)} \to {\mathbf V}_{h(p)}$$ corresponding to each path. The \emph{trivial path} $e_v$ consists of a single vertex $v$ and no arrows; ${\mathbf V}_{e_v}$ is of course the identity endomorphism of ${\mathbf V}_v$. A \emph{representation \/} is said to \emph{satisfy} $r=\sum\lambda_ip^i$ if $\sum\lambda_i{\mathbf V}(p^i)=0$. If ${\mathcal K}$ denotes a set of relations, then the set of representations of ${\mathcal Q}$ into ${\mathbf V}$ satisfying the relations ${\mathcal K}$ will be denoted $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf V})$. It is an affine variety inside $\Rep_{\mathcal Q}({\mathbf V})$. \begin{rmk} There are in general many classes of relations one can consider, but in this paper, only \emph{commutation relations} will be considered, namely relations generated by differences of two paths, both paths having the same tail and head. \end{rmk} \subsubsection{Morphisms of Representations} \label{sec:1:quivers:morph} \nopagebreak Let ${\mathbf V},{\mathbf V}'$ be two representations of the same quiver ${\mathcal Q}$ (possibly with relations). A \emph{morphism} of representations $\theta\colon {\mathbf V}\to {\mathbf V}'$ is given by morphisms $\theta_v\colon {\mathbf V}_v\to {\mathbf V}'_v$ for each vertex $v$, which satisfy ${\mathbf V}'_{a}\theta_{t(a)}=\theta_{h(a)}{\mathbf V}_a$ for each arrow $a$. If ${\mathbf V}_v={\mathbf V}_v'$ and the linear maps $\theta_v$ are all isomorphisms, then $\theta$ is called an \emph{isomorphism of representations}. \subsubsection{Representation Moduli} \label{sec:1:quivers:moduli} \nopagebreak Given a quiver ${\mathcal Q}$, some relations ${\mathcal K}$ and a representation space ${\mathbf V}$, one is naturally interested in the \emph{moduli space} of representations. Loosely speaking, this is the set of isomorphism classes of representations of ${\mathcal Q}$ into ${\mathbf V}$. More precisely, the group $$\GL({\mathbf V}):=\times_{v\in{\mathcal Q}_0}\GL({\mathbf V}_v)$$ acts on $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf V})$ and its orbits consist of equivalence classes of representations. The scalar subgroup ${\C^*}\subset\GL({\mathbf V})$ acts trivially, and one is left with a free action of the quotient $\PGL({\mathbf V}):=\GL({\mathbf V})/{\C^*}$. Since $\PGL({\mathbf V})$ is not compact, one must resort to geometric invariant theory\ (GIT) in order to obtain quotients which are well-defined as quasi-projective varieties. The ``canonical'' quotient is the affine GIT quotient \begin{equation} \label{eq:zero-quot} {\mathcal M}_{{\mathcal Q},{\mathcal K},0}:= \Rep_{{\mathcal Q},{\mathcal K}}({\mathbf V})\gitquot{} \PGL({\mathbf V}). \end{equation} There are other possible quotients however and their existence is in fact the basis of this paper. In general, given a complex affine variety $X$ acted upon by a reductive group $G$, and given a character $\chi\colon G\to{\C^*}$, one defines $$X\gitquot\chi G := \Proj \bigoplus_{r\in{\mathbb N}} {\mathcal O}_X(rL_\chi)^G,$$ where $L_\chi$ is the trivial bundle over $X$, on which $G$ acts via $\chi$. The reader is refered to~\cite{sacha:ale,sacha:thesis} for a detailed treatment. When the complex reductive group $G$ has a real form $K$, rather than specifying the character $\chi$ of $G$ one can specify instead its derivative at the identity $\Lie G\to {\mathbb R}$, or even the restriction of the later to the Lie algebra of the real form, giving an element of $\zeta=d\chi(1)_{|\Lie K}\in(\Lie K)^*$. For the moduli of representations of ${\mathcal Q}$ into ${\mathbf V}$, a real form of $\PGL({\mathbf V})$ is given by $\PU({\mathbf V}):=\times_{v\in{\mathcal Q}_0}\U({\mathbf V}_v)/\U(1)$, and the dual of its Lie algebra is the subspace $${\mathfrak{k}}:=\{\zeta\in \times_{v\in{\mathcal Q}_0} \End_{\mathbb R}({\mathbf V}_v)^*| \sum_{v\in{\mathcal Q}_0}\trace \zeta_v=0\},$$ where $\zeta_v$ denotes the restriction of $\zeta$ to $\End_{\mathbb R}({\mathbf V}_v)$. For $\zeta$ an integral element of $\mathfrak{k}$, define \begin{equation} \label{eq:zeta-quot} {\mathcal M}_{{\mathcal Q},{\mathcal K},\zeta_\chi} := \Rep_{{\mathcal Q},{\mathcal K}}({\mathbf V})\gitquot\chi \PGL({\mathbf V}),\quad\text{where }\zeta_\chi=d\chi(1)_{|\Lie K} \end{equation} This is seen to coincide with Definition~\ref{eq:zero-quot} in the case $\zeta_\chi=0$ (i.e.\xspace $\chi=1$). \begin{rmk} In the above, ``$\zeta$ integral'' means integral with respect to the natural lattice in $\mathfrak{k}$ (i.e.\xspace the kernel of the exponential map). Strictly speaking, one is not restricted to integral values, any rational value will do, provided one makes sense of ``fractional linearisations'' in the obvious manner. We do not bother, as nothing substantially new is gained from this approach (the moduli for $k\zeta$ are isomorphic to those for $\zeta$). \end{rmk} \section{The McKay Quiver} \label{sec:1:mckay} The McKay quiver ${\mathcal Q}_{\Gamma,Q}$ is a quiver which is naturally associated to a representation $Q$ of a finite group $\Gamma$. As explained below, its vertices are the irreducible representations of $\Gamma$ and the arrows describe how the tensor product of $Q$ with each irreducible decomposes into a sum of irreducibles. It seems natural to expect a relation between ${\mathcal Q}_{\Gamma,Q}$ and the quotient singularity $X=Q/\Gamma$. In fact, for the case of a finite subgroup $\Gamma\subset \SU(2)$ there is a remarkable relation between ${\mathcal Q}_{\Gamma,Q}$ and the minimal resolution $\widetilde{X}\to X$. McKay remarked~\cite{mckay:graphs} that the quivers ${\mathcal Q}_{\Gamma,{\mathbb C}^2}$ are precisely the extended Dynkin diagrams of type $\overline A,\overline D$, and $\overline E$. The ordinary Dynkin diagrams of type $A,D$ and $E$ had previously been shown by Brieskorn~\cite{briesk} to be the dual graphs to the graphs of rational curves in the exceptional fibre of $\widetilde{X}\to X$. Kronheimer~\cite{kron:thesis,kron:crendus,kron:ale} showed that one can construct the minimal resolution $\widetilde{X}$ by considering what turn out to be, upon closer inspection, moduli spaces of representations of ${\mathcal Q}_{\Gamma,{\mathbb C}^2}$ into the regular representation space of $\Gamma$. The representations are also required to satisfy some commutation relations, denoted ${\mathcal K}$. In this section, the generalisation of these commutation relations to any McKay quiver is described. \subsection{The McKay Quiver} \label{sec:mckay:mckay:dfn} \nopagebreak Let $\Gamma$ be a finite group and let $\{R_i, i\in I\}$ be the set of irreducible representations of $\Gamma$. Any $\Gamma$-module $R_{\mathbf V}$ decomposes into a sum of irreducibles: \begin{equation} R_{\mathbf V}= \bigoplus_{i\in I}{\mathbf V}_i\otimes R_i, \label{eq:irred_decomp} \end{equation} and gives an $I$-graded vector-space (trivial as a $\Gamma$-module) ${\mathbf V}=\oplus_{i\in I}{\mathbf V}_i$, called the \emph{multiplicity space} for $R_{\mathbf V}$. The dimension of ${\mathbf V}_i$ is called the \emph{multiplicity} of $R_i$ in $R_{\mathbf V}$ and the vector $\dim {\mathbf V}=(\dim {\mathbf V}_i)_{i\in I}$ is called the \emph{dimension vector} of the $\Gamma$-module $R_{\mathbf V}$. Conversely, given any $I$-graded vector-space ${\mathbf V}$, one can construct a corresponding $\Gamma$-module $R_{\mathbf V}$ by formula~\eqref{eq:irred_decomp}. The \emph{McKay quiver} ${\mathcal Q}={\mathcal Q}_{\Gamma,Q}$ of a representation $Q$ is constructed as follows. The vertices ${\mathcal Q}_0=I$ are the irreducible representations of $\Gamma$, and there are $a_{ij}$ (possibly zero) arrows from vertex $i$ to vertex $j$. The non-negative integers $a_{ij}$ (which form what is called the \emph{adjacency matrix} of ${\mathcal Q}$) are defined by the following irreducible decompositions (for each $i\in I$) \begin{equation} \label{eq:aijdecomp} Q\otimes R_i=\bigoplus_j a_{ji}R_j. \end{equation} More invariantly, one may write \begin{align} Q\otimes R_i &= \bigoplus_j A^*_{ji}\otimes R_j\\ \intertext{where} A_{ji}:&=\Hom_\Gamma(Q\otimes R_i,R_j), \end{align} and think of the arrows from $i$ to $j$ as giving a basis for the $a_{ij}$-dimensional vector-space $A_{ij}$. If ${\mathbf V}$ is a ${\mathcal Q}_0$-graded vector-space, then the representations of ${\mathcal Q}$ into ${\mathbf V}$ correspond to the $\Gamma$-module endomorphisms $R_{\mathbf V}\to Q\otimes R_{\mathbf V}$. In fact, \begin{equation} \begin{split} \Hom_\Gamma(R_{\mathbf V},Q\otimes R_{\mathbf V}) &= \Hom_\Gamma(\oplus_i {\mathbf V}_i\otimes R_i, Q\otimes (\oplus_j{\mathbf V}_j\otimes R_j))\\ &= \bigoplus_{i,j} A_{ij}^* \otimes \Hom({\mathbf V}_i,{\mathbf V}_j), \end{split} \end{equation} so given $\alpha\in\Hom_\Gamma(R_{\mathbf V},Q\otimes R_{\mathbf V})$, one can pair it with an arrow $a=i\to j$ (considered as a basis element of $A_{ij}$) and obtain a map ${\mathbf V}_i\to {\mathbf V}_j$. Thus \begin{equation} \label{eq:hom-rep} \Hom_\Gamma(R_{\mathbf V},Q\otimes R_{\mathbf V}) = \Rep_{\mathcal Q}({\mathbf V}). \end{equation} The representation of ${\mathcal Q}$ which corresponds to $\alpha\in\Hom_\Gamma(R_{\mathbf V},Q\otimes R_{\mathbf V})$ will be denoted $\tilde \alpha$. \begin{rmk} Note that the orientation of ${\mathcal Q}$ is reversed when the representation $Q$ is replaced by its dual $Q^*$, so when $Q$ is self-dual all the arrows $i\to j$ have opposite arrows $j\to i$. The pair $i\rightleftarrows j$ is usually denoted $i\mbox{---} j$. \end{rmk} \begin{example}[The McKay Correspondence] \label{ex:dynkin} For a subgroup $\Gamma<\SU(2)$, the standard 2-dimensional representation $Q$ is always self-dual, since $\Lambda ^2 Q\cong {\mathbb C}$ as $\Gamma$-modules. McKay's observation~\cite{mckay:graphs} is that the quivers ${\mathcal Q}_{\Gamma,Q}$ coincide with the extended homogeneous\footnote{A Dynkin diagram is called \emph{homogeneous} if it has no multiple bonds} Dynkin diagrams $\overline A_k,\overline D_k, \overline E_6,\overline E_7,\overline E_8$ represented in Figure~\ref{fig:dynkin}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 5cm \epsfbox{fig/dynkin.eps} \end{center} \caption[The extended homogeneous Dynkin diagrams $\overline A_k,\overline D_k, \overline E_6,\overline E_7,\overline E_8$]{The extended homogeneous Dynkin diagrams $\overline A_k,\overline D_k, \overline E_6,\overline E_7,\overline E_8$ (The extra vertex is indicated marked $\bullet$.)} \label{fig:dynkin} \end{figure} \end{example} \subsection{The Commutation Relations} \label{sec:mckay:mckay:ccr:relns} \nopagebreak Recall from Section~\ref{sec:mckay:mckay:dfn} that if $R_{\mathbf V}$ is any $\Gamma$-module, the $\Gamma$-module homomorphisms $R_{\mathbf V} \to Q\otimes R_{\mathbf V}$ correspond to representations of the McKay quiver into the $I$-graded vector-space ${\mathbf V}$ which is the multiplicity space for $R_{\mathbf V}$. One natural $\Gamma$-module to consider as a candidate for $R_{\mathbf V}$ is the ``canonical one'' given by the regular representation space $R$ of $\Gamma$. Its multiplicity space will be denoted ${\mathbf R}=\oplus_{v\in {\mathcal Q}_0} {\mathbf R}_v$. Its components satisfy the well-known equalities $\dim {\mathbf R}_v=\dim R_v$. The simplest way to define the commutation relations ${\mathcal K}$ is to use the isomorphism~\eqref{eq:hom-rep}. Let $q_i$ be a basis of $Q$ which is orthonormal with respect to some positive definite $\Gamma$-invariant hermitian inner product. Let $\tilde \alpha\in\Rep_{\mathcal Q}({\mathbf R})$ be a representation of ${\mathcal Q}$ into ${\mathbf R}$, and let $\alpha\in Q\otimes\End_{\mathbb C} R$ be the $\Gamma$-invariant element which corresponds via the isomorphism~\eqref{eq:hom-rep}. The group $\Gamma$ acts on $\End_{\mathbb C} R$ in the natural way, i.e.\xspace by conjugation, and the element $\alpha $ decomposes with respect to the basis $q_i$ into endomorphisms $\alpha _i\in\End_{\mathbb C} R$ for $i=1,\dots, n$ which satisfy the following equivariance condition \begin{equation}\label{eq:alpha_equivariance_condition} \sum_l \gamma_{kl} \alpha_{l} = \varphi(\gamma)\alpha_{k}\varphi(\gamma)^{-1},\rlap{$\qquad \forall k,\gamma$,} \end{equation} where $\boldsymbol{\gamma}=(\gamma_{kl})$ is the matrix corresponding to the action of the element $\gamma$ on $Q$ with respect to the basis $\{q_l\}_{l=1}^n$. The commutation relations are defined by the condition \begin{equation} \label{eq:alpha_commute} [\alpha _i,\alpha _j]=0. \end{equation} Thus the variety ${\mathcal M}_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ of representations of ${\mathcal Q}$ into ${\mathbf R}$ satisfying these relations coincides with the variety ${\mathcal N}^\Gamma$ defined in~\cite{sacha:ale,sacha:thesis}. \subsection{Representation Moduli} \label{sec:1:mckay:moduli} Given $\zeta\in\mathfrak{k}$, the representation moduli of the McKay quiver of $(\Gamma,Q)$ in the multiplicity space ${\mathbf R}$ for the regular representation $R$ of $\Gamma$ are defined by the GIT quotients $${\mathcal M}_\zeta= \Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R}) \gitquot\zeta \PGL({\mathbf R}).$$ Since the group $\PGL({\mathbf R})$ coincides with the group $\PGL^\Gamma(R)$ (the projectivization of the group of $\Gamma$-invariant linear endomorphisms of $R$), one sees that the moduli ${\mathcal M}_\zeta$ coincide with the moduli $X_\zeta$ of Hermitian-Yang-Mills type bundles defined and studied in~\cite{sacha:ale}. \part{Abelian Groups} \label{sec:2} In Part~\ref{sec:1} the moduli $X_\zeta$ were shown to coincide with the moduli of Hermitian-Yang-Mills bundles defined in~\cite{sacha:ale} and hence ($\Gamma$ acts freely outside the origin) to provide partial resolutions $\rho_\zeta\colon X_\zeta \to X_0=Q/\Gamma$ of the isolated quotient singularity. The focus from now on will be the moduli $X_\zeta$ in the case where $\Gamma$ is an abelian group (of order $r$) acting linearly on $Q\cong{\mathbb C}^n$. (Later we shall specialise to the cyclic group of order $r$.) When $\Gamma$ is abelian, the singularity $X$ is toric and can be resolved within the toric category, in general in many ways. In fact, as shown in the first section below, the space of representations $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ and its quotients $X_\zeta$ are also toric varieties; in toric notation (reviewed in~\ref{sec:2:abel:toric}) $$\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R}) = \protect\overline{T}{}^{\Lambda,C}\text{ and }X_\zeta = \protect\overline{T}{}^{\Pi,C_\zeta},$$ where $\Lambda,\Pi$ are certain lattices, and where the $n$-dimensional convex polyhedra $C_\zeta$ are obtained by `slicing' the $n+r-1$-dimensional cone $C$ by affine $n$-planes. The rest of this paper is devoted to describing the slices $C_\zeta $, and in particular their extreme points and their tangent cones. The number of extreme points is the Euler characteristic of $X_\zeta $, and the geometry of the tangent cones of $C_\zeta $ at these points describes the singularities of $X_\zeta $. There are two main steps in the study of the slices $C_\zeta $. The first step is to reduce their description to a network flow problem on the McKay quiver. The particular problem turns out to be a generalisation of the classical transportation problem~\cite{kenn_helg,gond_mino:graphs}. This is done in Section~\ref{sec:2:abel:toric}. The second step is to describe the polyhedra which solve the generalised transportation problem. A solution for a general quiver is given in Section~\ref{sec:2:flow}. The main result is Theorem~\ref{thm:general}, which says that the flows which correspond to the extreme points of $C_\zeta$ are precisely the ones whose support contains no cycles of non-zero type in its ``closure'' (see \*2:flow:cycle* for the definition). This is then applied to the case of the McKay quiver in Sections~\ref{sec:2:exact} and~\ref{sec:2:sing} to describe the slices $C_\zeta $ for all $\zeta$. The last section in this part chapter concerns various conjectures, in particular for singularities of dimension~3. One hope is that, in general, the quotients $X_\zeta $ provide a class of partial desingularisations of the isolated quotient singularity $X$, which will be in some sense ``natural,'' and, in good cases, non-singular and minimal with respect to this property. This is motivated by the fact that Kronheimer~\cite{kron:thesis} has shown that for the case $\Gamma \subset\SU(2)$, the $X_\zeta $ coincide with the minimal smooth resolution of the singularity for generic values of $\zeta$. One candidate for a ``good case'' is the case when $\Gamma\subset\SU(3)$. \begin{conj} If $\Gamma\subset\SU(3)$ is abelian and acts on ${\mathbb C}^3$ freely outside the origin then $\rho_\zeta\colon X_\zeta \to X_0={\mathbb C}^3/\Gamma$ is crepant and is a smooth resolution for generic values of $\zeta$. \end{conj} Using the description of $C_\zeta$, this conjecture is reduced in Section~\ref{sec:2:crep} to to a statement concerning the existence and combinatorial properties of certain trees in the McKay quiver. Brute-force computer calculations show the conjecture to be true for the singularities $\qsing 1/6(1,2,3)$, $\qsing 1/7(1,2,4)$, $\qsing 1/8(1,2,5)$, $\qsing 1/9(1,2,6)$, $\qsing 1/10(1,2,7)$ and $\qsing 1/11(1,2,8)$, but there is as yet no general proof. Other conjectures regarding, for example, the Euler number of the resolutions also translate to graph theoretical and combinatorial statements on the McKay quiver. \section{Summary of Notation for Part \ref{sec:2}} \label{sec:2:notation} \begin{Pentry} \item[$\Gamma$] Finite abelian group $\Gamma$ of order $r$ (usually $\mu_r$). \item[$\mu_{r}$] Group of $r$-th roots of unity in ${\C^*}$. \item[$\widehat\Gamma$] Character group $\widehat\Gamma:=\Hom(\Gamma,{\C^*})$. \item[$\rho$] The morphism giving the action of $\Gamma$ on ${\C^*}^n$: $\rho\colon\Gamma\to{\C^*}^{n}$. \item[$w_{i}$] Weights ($w_i\in \Z_{r}; i=1,\dots,n$) for the action of $\Gamma$ on ${\mathbb C}^{n}$. \item[$\hat{\rho}$] Dual morphism $\hat\rho\colon\Z^{n}\to\widehat\Gamma$. \item[$\Pi$] A sub-lattice of $\Z^{n}$ of index $r$) given by $$\Pi := \ker\hat\rho= \{ x\in \Z^n | \sum x_iw_i\equiv 0\pmod r\}.$$ \item[$\qsing1/r(w_1,\dots,w_n)$] The quotient singularity ${\mathbb C}^n/\mu_r$, where $\mu_r$ acts on ${\mathbb C}^n$ with weights $(w_1,\dots,w_n)$. \item[$R_{v}$] One-dimensional representation on which $\Gamma$ acts by $\lambda\to\lambda^{v}$. \item[$\nu$] Natural morphism $\nu\colon\Gamma\to\PGL({\mathbf R})$. \item[$\hat{\nu}$] Dual morphism $\hat\nu\colon\Lambda^{0,0}\to\widehat\Gamma$. \item[$a_{v}^{i}$] Arrow of type $i$: $a_{v}^{i}:=v\to v-w_{i}$ ($v\in{\mathcal Q}_{0}$, $i\in\{1,\cdots,n\}$). \item[$\tilde{\alpha}(a_{v}^{i})$] Value of $\tilde{\alpha}$ on $a_{v}^{i}$; Equal to the $(v-w_{i},v)$-th entry of the matrix $\alpha$. \end{Pentry} \subsection{Toric Geometry} \begin{Pentry} \item[$M$] Generic lattice. \item[$M_{\Q}$] Associated rational vector-space. \item[$T^{M}$] Algebraic torus $T^{M}:=\Spec{\mathbb C}[M]$. \item[$P$] Generic polyhedron in $M_{\Q}$. \item[$\protect\overline{T}{}^{M,P}$] Quasi-projective toric variety $\Proj{\mathbb C}[\widetilde{P}\cap\widetilde{M}]$ defined by $M$ and $P$. \item[$\widetilde{M}$] Product lattice $\widetilde{M}:=\Z\times M$ \item[$\widetilde{P}$] (Closure of) the cone over $P$: $\widetilde{P} := \overline{\Q_{\ge 0}(\{1\}\times P)}\subset M_{\Q}$. \item[$T_{F}P$] Tangent cone of $P$ at face $F$. \end{Pentry} \subsection{Flows} \begin{Pentry} \item[$k^A$] Set (lattice, vectorspace, cone) of maps $A\to k$. Used for $k=\Z,\Z_+, {\mathbb R}, {\mathbb R}_+, {\mathbb C}$, and $A={\mathcal Q}_0, {\mathcal Q}_1, S$. \item[$f$] Generic flow (element of ${\mathbb R}^{{\mathcal Q}_{1}}$). \item[$f^{\pm}$] Positive/negative part of $f$. \item[$\Lambda^{1}$] Lattice $\Z^{{\mathcal Q}_{1}}$ of integer-valued flows on ${\mathcal Q}$. \item[$\Lambda^{1}_{{\mathbb R}}$] Vectorspace of real-valued flows on ${\mathcal Q}$. \item[$\Lambda^{1}_{{\mathbb R}_{+}}$] First quadrant in $\Lambda^{1}_{{\mathbb R}}$ (non-negative real-valued flows). \item[$\chi_{a}$] Flow taking the value 1 on $a$ and zero elsewhere. Basis element of $\Lambda^{1}$. \item[$\chi_{v}^{i}$] Alternative notation for $\chi_{a_{v}^{i}}$. \item[$\Lambda^{0}$] Lattice $\Z^{{\mathcal Q}_{0}}$ of assignments of integers to the vertices of ${\mathcal Q}$; $\Lambda^0_{\mathbb R}$ coincides with $\Lie \GL({\mathbf R})$. \item[$\Lambda^{0,0}$] Sub-lattice $\{\zeta\in\Z^{{\mathcal Q}_{0}}| \sum_{v\in{\mathcal Q}_{0}}\zeta(v)=0\}$ of $\Lambda^0$; $\Lambda^{0,0}_{\mathbb R}$ coincides with $\Lie \PGL({\mathbf R})$. \item[$\zeta$] Element of $\Lambda^{0,0}$. \item[$\In(v)$] Set of arrows $a$ such that $h(a)=v$. \item[$\Out(v)$] Set of arrows $a$ such that $t(a)=v$. \item[$\partial$] Natural map $\partial\colon\Lambda^{1}\to\Lambda^{0,0}$ which calculates the net contribution of a flow at each vertex: $$\partial f(v)=\sum_{a\in\In(v)} f(a) - \sum_{a\in\Out(v)} f(a).$$ \item[$\chi_{v}$] Function taking the value 1 on $v$ and zero elsewhere. Basis element of $\Lambda^{0}$. \item[$\Lambda^{2}$] Sub-lattice of $\Lambda^{1}$ generated by the elements corresponding to the commutation relations $$\chi_v^i +\chi_{v- w_i}^j-\chi_v^j-\chi_{v- w_j}^i,\qquad\text{for } i,j=1,\dots, n$$ \item[$\Lambda$] Quotient lattice $\Lambda^{1}/\Lambda^{2}$. \item[$C$] Image of cone $\Lambda^{1}_{{\mathbb R}_{+}}$ inside the quotient $\Lambda$. Also written $\Lambda_{{\mathbb R}_{+}}$. \item[$C_{\zeta}$] Convex polyhedron obtained by slicing $C$ with the hyper-plane $\partial f=\zeta$. Equal to the image of $F_{\zeta}$ under $\pi$. \item[$\pi$] Map $\pi\colon{\mathcal Q}_1\to\{1,\dots,n\}$ which assigns to each arrow of the McKay quiver its type: $\pi(a^i_v) := i$. Induces a map $\pi\colon\Lambda^{1}_{{\mathbb R}_{+}}\to{\mathbb R}^{n}$ on the flows: $$\pi(f)=\left(\sum_{v\in{\mathcal Q}_{0}} f(a_{v}^{1}),\dots,\sum_{v\in{\mathcal Q}_{0}} f(a_{v}^{n})\right).$$ \end{Pentry} \subsection{Configurations} \begin{Pentry} \item[$S$] Configuration of arrows (non-empty subset of ${\mathcal Q}_{1}$). \item[$\supp f$] Support of the flow $f$ (arrows on which $f$ takes a non-zero value). \item[${\mathcal C}$] Set of all configurations of arrows. \item[${\mathcal T}$] Set of all configurations of arrows which are trees (i.e.\xspace which contain no cycles). \item[$F_{\zeta}$] Admissible (i.e.\xspace non-negative) flows which satisfy $\partial f=\zeta$. \item[$F_{\zeta}(S)$] Flows satisfying $\partial f=\zeta$ which are non-negative outside $S$. (A convex cone if $\zeta=0$). \item[$Z_{\zeta}(S)$] Flows satisfying $\partial f=\zeta$ which are zero outside $S$. (Equal to the maximal subspace contained in $F_{0}(S)$ if $\zeta=0$). \item[$f_{\zeta}$] Map $f_{\zeta}\colon{\mathcal T}\to\Lambda^{1}_{{\mathbb R}}$ assigning to each tree $T$ the unique flow $f=f_{\zeta}(T)$ such that $\partial f=\zeta$. \item[${{\Lambda^{0,0}_{{\mathbb R}}}}(T)$] Open convex cone in $\Lambda^{0,0}_{{\mathbb R}}$ of values of $\zeta$ for which the tree $T$ is the support of some flow in $F_{\zeta}$. \item[${\mathcal D}$] Generic subset of ${\mathcal C}$. \item[${\mathcal D}_{\zeta}$] Elements of ${\mathcal D}$ which are the support of some $f\in F_{zeta}$. \item[${\mathcal D}_{\spn}$] Elements of ${\mathcal D}$ which span all vertices of ${\mathcal Q}$. \item[${\mathcal D}^{k}$] Elements of ${\mathcal D}$ of rank $k$ (i.e.\xspace elements $S$ such that $\rk \pi F_{0}(\overline{S})=k$. \end{Pentry} \subsection{Cycles and Paths} \begin{Pentry} \item[$p$] Generic path in ${\mathcal Q}$. \item[$p^{+}$] Positive part of $p$. \item[$p^{-}$] Negative part of $p$. \item[$\tilde{\chi}_{p}$] Basic flow associated to the path $p$. \item[$c$] Generic cycle in ${\mathcal Q}$. \item[$\{v\}(j_{0},\dots,j_{k})$] Notation for basic flows. \end{Pentry} \subsection{Miscellaneous} \begin{Pentry} \item[${\mathcal W}_S$] Weighting of the vertices of ${\mathcal Q}_{0}$ (a map ${\mathcal Q}_{0}\to\Z^{n}$ with special properties). \end{Pentry} \section{Representations and Relations in the Abelian Case} \label{sec:2:abel} \subsection{The McKay Quivers} \label{sec:2:abel:mckay} In the abelian case, all the irreducible representations are one-dimensional and are determined by an element of the dual group $\widehat\Gamma$, which is a finite abelian group isomorphic to $\Gamma$. Thus the vertex set of ${\mathcal Q}$ is nothing but $\widehat\Gamma$. The group $\Gamma$ will always be identified with a subgroup of ${\mathbb C}^{*n}\subset\GL(n)$, and thus $\widehat\Gamma$ will be thought of as a product of finite groups $\Z_{r_i}$ with the group operation denoted additively. \begin{example} \label{ex:5123} Let $\Gamma=\mu_5\subset{\C^*}$, the group of $5$-th roots of unity, acting on ${\mathbb C}^3$ with \emph{weights} $1,2$ and $3$, i.e.\xspace via $$\lambda\to \begin{pmatrix} \lambda^1 & 0 & 0 \\ 0 & \lambda^2 & 0 \\ 0 & 0 & \lambda^3 \end{pmatrix}. $$ This action is denoted symbolically by $\frac{1}{ 5}(1,2,3)$. The character group of $\mu_5$ is $\widehat{\mu_5}=\Z_5:=\Z/5\Z $, so ${\mathcal Q}$ has $5$ vertices. For each $i\in\Z_5$, let $\chi_i\colon \lambda\mapsto\lambda^i$ be the corresponding character, so that $\chi_i\chi_j=\chi_{i+j}$. Writing $R_i$ for the irreducible representations, one has $R_i\otimes R_j = R_{i+j}$. The representation $Q$ is just $R_1\oplus R_2\oplus R_3$, so that $$Q\otimes R_i=R_{i+1}\oplus R_{i+2}\oplus R_{i+3}.$$ Thus, the quiver has arrows from the vertex $i$ to the vertices $i-1, i-2$ and $i-3$ for each $i$ (see Figure~\ref{fig:all5123} below).\footnote{The additions are modulo 5.} \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/all5123.eps} \end{center} \caption{The McKay quiver for the action $\frac{1}{ 5}(1,2,3)$} \label{fig:all5123} \end{figure} \end{example} In fact, it is easy to see that the McKay quiver for the action of any cyclic group has a similar appearance. To see what the arrows are, decompose $Q$ into a sum of one-dimensional irreducibles \begin{equation} \label{eq:qdecomp} Q=\bigoplus_{i=1}^n R_{w_i}, \end{equation} where $w_i\in\widehat\Gamma$ for $i=1,\dots, n$ are the \emph{weights} of the action of $\Gamma$ on $Q$. Using ~\eqref{eq:irred_decomp}, one sees easily that \begin{equation} Q\otimes R_v = R_{v+w_1}\oplus \dots\oplus R_{v+w_n}, \label{eq:qtensorrv} \end{equation} so there is one arrow from $v$ to $v-w_i$ for each vertex $v$ and each weight $w_i$, giving a total of $nr$ arrows. The arrows corresponding to the weight $w_i$ are written $$a_v^i := v\to v -w_i,\quad \text{for } v\in{\mathcal Q}_0$$ and are said to be of \emph{type}~$i$. The McKay quiver for a general abelian group $\Gamma$ can be obtained by decomposing $\Gamma$ into products of cyclic groups. Define the \emph{product} of ${\mathcal Q}$ and ${{\mathcal Q}'}$ to be the quiver with vertices ${\mathcal Q}_0\times{{\mathcal Q}'}_0$ and with arrows \begin{multline} \{(v,t(a'))\to (v,h(a')) : v\in{\mathcal Q}_0,a'\in{{\mathcal Q}'}_1\}\\ \cup \{( t(a),v')\to (h(a),v') : v'\in{{\mathcal Q}'}_0,a\in{\mathcal Q}_1\}. \end{multline} Then the McKay quiver for $\Gamma$ is given by taking the product of the quivers for the cyclic factors. \subsection{Commutation Relations} \label{sec:2:abel:comm} Pick a basis $q_i$ of $Q$ such that the action of $\Gamma$ on $Q$ is diagonal, with $q_i$ corresponding to the irreducible component $R_{w_i}$. Then decomposing $R$ into irreducibles and using Shur's Lemma \begin{equation} \begin{split} \Hom_\Gamma(R,R_{w_i}\otimes R) &= \bigoplus_{v,v'\in{\mathcal Q}_0}\Hom_\Gamma({{\mathbf R}_v}\otimes R_v,{\mathbf R}_{v'}\otimes R_{w_i}\otimes R_{v'})\\ &= \bigoplus_{v\in{\mathcal Q}_0\phantom{v'}} \Hom({{\mathbf R}_v},{\mathbf R}_{v-w_i }) \end{split} \label{eq:homga} \end{equation} This means that the scalar map $\tilde \alpha(a_v^i)\colon {\mathbf R}_{v}\to {\mathbf R}_{v-w_i }$ is multiplication by the $(v-w_i,v)$-th entry of the $i$-th component matrix $\alpha_i\in\End R$. The condition $[\alpha_i,\alpha_j]=0$ therefore translates to the commutation relation \begin{equation} \label{eq:comrelations2} \tilde \alpha(a_v^i)\tilde \alpha(a_{v-w_i }^j)-\tilde \alpha(a_v^j)\tilde \alpha(a_{v-w_j }^i)=0 \end{equation} for all $v\in{\mathcal Q}_0$ and $i,j\in\{1,\dots, n\}$. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/comm7123.eps} \end{center} \caption[A picture of the commutation relation~\eqref{eq:comrelations2}]{A picture of the commutation relation~\eqref{eq:comrelations2} for $v=0$, $i=1$ and $j=3$ in the McKay quiver for $\frac{1}{ 7}(1,2,3)$. (The product of the representation along the continuous arrows must equal the product along the dotted arrows.)} \label{fig:comm7_123} \end{figure} \subsection{Toric Descriptions of the Representation Moduli} \label{sec:2:abel:toric} \subsubsection{Toric Varieties} Let $M$ be an integral lattice isomorphic to $\Z^n$, let $M_\Q:=M\otimes\Q$ be the associated rational vector space, and let $C\subset M_\Q$ be a convex $n$-dimensional cone. Let $T^M:=\Spec{\mathbb C}[M]$ be the algebraic torus whose character group is $M$. Then $$\protect\overline{T}{}^{M,C}:=\Spec{\mathbb C}[C\cap M]$$ is a compactification of $T^M$, called the \emph{toric variety associated to the cone} $C$ with respect to the lattice $M$. The covariant notation $\protect\overline{T}{}_{M^\vee, C^\vee}$ which uses the dual objects\footnote{The dual of a cone $C$ is the cone $C^\vee := \{n\in M^\vee | n(c) \geq 0, \forall c\in C\}$} is widely in use in the literature, but the contravariant version is more convenient for the purposes of this paper. This definition can be extended from cones to general convex polyhedra as follows. If $P$ is any convex polyhedron in $M_\Q$, denote by $\widetilde{P}$ the closure of the cone over $P$, namely \begin{equation} \label{eq:Ptw} \widetilde{P} := \overline{\Q_{\ge 0}(\{1\}\times P)} \subset \widetilde{M}_\Q=\Q\times M. \end{equation} Then one can define: \begin{equation} \label{eq:toric-defn} \protect\overline{T}{}^{M,P}:=\Proj{\mathbb C}[\widetilde{P}\cap\widetilde{M}], \end{equation} and this is easily seen to extend the previous definition for cones. \subsubsection{GIT quotients of Toric Varieties} Note that there is a natural line bundle $L^{M,P}:={\mathcal O}(1)$ on $\protect\overline{T}{}^{M,P}$ on which $T^M$ acts and whose sections are given simply by evaluation on the lattice points of $P$. If one translates $P$ by an element $\zeta\in M$, one gets the same toric variety ($\protect\overline{T}{}^{M,P+\zeta}=\protect\overline{T}{}^{M,P}$) and the same line bundle $L$, but with a \emph{different \/} linearisation of the action of $T^M$, differing from the old by the character $\zeta$. The consequence of this for GIT quotients is that if $M'$ is any sub-lattice of $M$, and if $\zeta$ an element of $M'$ (and hence a character of $T^{M'}$) one can consider either the quotient $$\protect\overline{T}{}^{M,P}\gitquot{\zeta}T^{M'}$$ with $T^{M'}$ taken to act on $L^{M,P}$ via multiplication by the section $\zeta$ or, equivalently, the quotient $$T^{M,\zeta+P}\gitquot{} T^{M'}$$ where $T^{M'}$ now acts on $L^{M,\zeta+P}$ in the ordinary way. Using this notation, one can state the following proposition which describes how toric varieties behave under GIT quotients (c.f.\xspace~\cite{thaddeus:flips}). \begin{prop} \label{prop:toric_quotients} Let $0\to M''\to M\to M'\to 0$ be an exact sequence of lattices and $P\subset M_\Q$ be a convex polyhedron. Write $$(M,P)\gitquot{}M' := (M'',P\cap M_\Q'').$$ Then $$\protect\overline{T}{}^{ M,P}\gitquot{} T^{ M'} = \protect\overline{T}{}^{(M,P)\gitquot{}M'}.$$ \end{prop} \begin{proof} An element $x^m\in{\mathbb C}[ M\cap P]$ is invariant under $T^{ M'}$ if $m$ belongs to the stabiliser of $T^{ M'}$ in $T^M$, which is nothing but $M''$. Thus ${\mathbb C}[\widetilde{P}\cap \widetilde{M}]^{T^ {M'}} ={\mathbb C}[\widetilde{P}\cap\widetilde{M''}]= {\mathbb C}[\widetilde{(P\cap M_\Q'')}\cap\widetilde{M''}]$ and the result follows by taking $\Proj$. \end{proof} \subsubsection{The Representation Variety} The description of the previous sections shows that $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ is a subvariety of ${\mathbb C}^{{\mathcal Q}_1}$ defined by the equations~\eqref{eq:comrelations2}. The goal is now to describe the quotients by the group $\PGL({\mathbf R})$. It turns out that $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ is a toric variety and that $\PGL({\mathbf R})$ is an algebraic sub-torus, so that one is able to use Proposition~\ref{prop:toric_quotients}. Consider the lattice $\Lambda ^1:= \Z^{{\mathcal Q}_1}$ and the first quadrant $\Lambda ^1_{{\mathbb R}+}:={\mathbb R}_+^{{\mathcal Q}_1}$ inside its associated real vector-space $\Lambda^1_{\mathbb R}$. Recall that the toric variety which corresponds to ${\mathbb C}^{{\mathcal Q}_1}$ is $$\protect\overline{T}{}^{\Lambda ^1,\Lambda ^1_{{\mathbb R}+}} := \Spec {\mathbb C}[\Lambda ^1_+],$$ where $\Lambda ^1_+$ is the semi-group $\Z_+^{{\mathcal Q}_1}$ and ${\mathbb C}[\Lambda ^1_+]$ denotes its group algebra. More precisely, the ${\mathbb C}$-points of the scheme $\Spec {\mathbb C}[\Lambda ^1_+]$ are into one-one correspondence with the points of ${\mathbb C}^{{\mathcal Q}_1}$ as follows. If $x\colon {\mathbb C}[\Lambda ^1_+]\to{\mathbb C}$ is an algebra homomorphism representing a point of $\Spec {\mathbb C}[\Lambda ^1_+]$ then, evaluating it on the generators of $\Lambda ^1_+$ gives a map ${\mathcal Q}_1\to {\mathbb C}$, i.e.\xspace a point of ${\mathbb C}^{{\mathcal Q}_1}$. Conversely, if $\tilde \alpha \in{\mathbb C}^{{\mathcal Q}_1}$, there is an induced morphism of semi-groups \corresp{(\Lambda ^1_+,+)}{({\mathbb C},\cdot)}{f}{\prod_a(\tilde \alpha (a))^{f(a)},} and this induces an algebra morphism ${\mathbb C}[\Lambda ^1_+]\to{\mathbb C}$. Under this identification, the points of $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ correspond to the semi-group morphisms $ \Lambda ^1_+ \to{\mathbb C}$ which are the identity when restricted to the sub-semi-group $\Lambda ^2\cap\Lambda ^1_+$ of $\Lambda ^1_+$. Here $\Lambda ^2$ is the sub-lattice of $\Lambda ^1$ generated by the elements $$\chi_v^i +\chi_{v- w_i}^j-\chi_v^j-\chi_{v- w_j}^i,\qquad\text{for } i,j=1,\dots, n$$ (and $\chi_v^i$ denotes the indicator function $\chi_{a_v^i}$ for the element $a_v^i$, i.e.\xspace the basis element of $\Lambda^1$ which takes the value $1$ on the arrow $a_v^i=v\to v-w_i\in {\mathcal Q}_1$ and zero elsewhere). \begin{prop} \label{prop:rep-variety-toric} The quotient $\Lambda:=\Lambda ^1/\Lambda ^2$ is a lattice (i.e.\xspace is torsion free). If one denotes by $\Lambda _+$ (resp.\xspace $C=\Lambda_{{\mathbb R}+}$) the lattice (resp.\xspace the cone) generated by the image of the elements of $\Lambda ^1_+$ in $\Lambda$, then the variety of representations of the McKay quiver with relations into ${\mathbf R}$ is given by $$\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R}) = \protect\overline{T}{}^{\Lambda,C} = \Spec {\mathbb C}[\Lambda _+].$$ \end{prop} \begin{proof} The only fact which has to be proved is that $\Lambda $ is a lattice. This will be postponed till Lemmas~\ref{lemma:kerc_lambda2}--\ref{lemma:imagec}. \end{proof} The group $\GL({\mathbf R})$ on the other hand coincides with ${\C^*}^{{\mathcal Q}_0}$. If $\Lambda^0$ denotes the lattice $\Z^{{\mathcal Q}_0}$, then one can write $\GL({\mathbf R})=T^{\Lambda^0}$ in the notation of Section~\ref{sec:2:abel:toric}. Similarly, writing $\Lambda^{0,0}$ for the sub-lattice of $\Lambda^0$ whose elements $\zeta$ satisfy $\sum_{v\in{\mathcal Q}_0}(v)=0$, one has $\PGL({\mathbf R})= T^{\Lambda^{0,0}}$. Applying Proposition~\ref{prop:toric_quotients}, one obtains the following toric description of the moduli $X_\zeta$. \begin{thm} \label{thm:toric-d-moduli} The moduli spaces $X_\zeta$ are quasi-projective toric varieties $$X_\zeta = \protect\overline{T}{}^{\Pi ,C_\zeta },$$ where $\Pi:=\ker\hat\rho=\pi\times\partial(\Lambda)/\Lambda^{0,0}$ is a sub-lattice of $\Z^n$ of index $\card{\Gamma}$ and $C_\zeta$ are convex polyhedra in $\Pi_{\mathbb R}$ in obtained by slicing the cone $C$ with the affine planes $\zeta +\Lambda ^{0,0}_{\mathbb R}$. \end{thm} \begin{proof} By definition $$X_\zeta=\protect\overline{T}{}^{\Lambda,C}\gitquot{\zeta}T^{\Lambda^{0,0}},$$ and this coincides with $$\protect\overline{T}{}^{\Lambda,\zeta+C}\gitquot{}T^{\Lambda^{0,0}}.$$ If $\Lambda':=\pi\times\partial(\Lambda)/\Lambda^{0,0}$ was a lattice then applying Proposition~\ref{prop:toric_quotients} would give $X_\zeta = \protect\overline{T}{}^{\Lambda' ,C_\zeta }$. It therefore remains to prove that $\Lambda'=\Pi$. This is done in Section~\ref{sec:2:exact}: Lemmas~\ref{lemma:kerc_lambda2}--\ref{lemma:imagec} show that there is an exact sequence of abelian groups \begin{equation} \label{eq:exact1} 0\to\Lambda\xrightarrow{\pi\times\partial} \Z^n\times\Lambda^{0,0} \xrightarrow{\hat\rho-\hat\nu} \Z^n/\Pi \cong\widehat\Gamma \to 0. \end{equation} \end{proof} Note that the morphisms appearing in the exact sequence in the above proof are: \begin{itemize} \item The projection $\pi\colon\Lambda^1\to\Z^n$ (or better, its descent to $\Lambda\to\Z^n$). \item The (descent to $\Lambda$) of the morphism of lattices $\partial\colon\Lambda^1\to\Lambda^{0,0}$ dual to the action of $\PGL({\mathbf R})=T^{0,0}$ on $\Rep_{{\mathcal Q}}({\mathbf R})={\mathbb C}^{{\mathcal Q}_1}$ given by \map{\Hat\partial}{T^{0,0}}{T^1}{\lambda}{a\mapsto\lambda_{t(a)}^{-1}\lambda_{h(a)}.} Here $\lambda_v$ denotes the component of $\lambda\in T^{0,0}$ corresponding to the vertex $v\in{\mathcal Q}_0$. \item The morphism of lattices \map{\hat\nu}{\Lambda^{0,0}}{\widehat\Gamma}{\zeta}{\sum_{v\in{\mathcal Q}_0}\zeta(v)v} dual to the action of \/$\Gamma$ on $\End R$ by conjugation. \item The morphism of lattices\map{\hat\rho}{\Z^n}{\widehat\Gamma}{\zeta}{\sum_{v\in{\mathcal Q}_0}x_iw_i} dual to the action $\rho\colon \Gamma\to {\C^*}^n$. \end{itemize} The fact that $\pi\times\partial$ descends to $\Lambda$ corresponds to the fact that $\GL^\Gamma(Q)={\C^*}^n$ and $\PGL({\mathbf R})=T^{0,0}$ act on $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ rather than simply on $\Rep_{{\mathcal Q}}({\mathbf R})={\mathbb C}^{{\mathcal Q}_1}$. On the other hand, the fact that their image maps into $\ker (\hat\rho-\hat\nu)$ corresponds to the fact that $\Rep_{{\mathcal Q},{\mathcal K}}({\mathbf R})$ can be identified with the $\Gamma$-invariant part of $\Hom(R,Q\otimes R)$. The action of $T^\Pi$ arises from the existence of the map $$\pi\colon{\mathcal Q}_1\to \{1,\dots, n\},$$ which assigns to each arrow its type. This induces a $\Z$-linear map $\pi \colon {\Lambda^1}\to \Z^n$, which is denoted by the same letter, and one obtains an action of ${\C^*}^n$ on ${\mathbb C}^{{\mathcal Q}_1}$ via the corresponding morphism of algebraic tori: $\tau\cdot\tilde \alpha:=\Hat\pi (\tau)\tilde \alpha$. Explicitly, for $\tau\in{\C^*}^n$, $\tilde \alpha \in{\mathbb C}^{{\mathcal Q}_1}$, and $a\in{\mathcal Q}_1$, $$ (\tau\cdot \tilde \alpha)(a) =\tau_{\pi(a)}\tilde \alpha(a), $$ where $\tau\in {\C^*}^n$ has components $\tau_i:=\tau(e_i)$ with respect to the standard basis $e_i$ of $\Z^n$, and $\chi _a$ denotes the basis element of $\Lambda ^1$ which is the indicator function of the singleton $\{a\}\subset{\mathcal Q}_1$. The action of ${\C^*}^n$ corresponds to multiplying the value of $\tilde \alpha $ on all arrows $a$ with the same $\pi(a)$ by the same factor $\tau_{\pi(a)}$. Since $\Gamma$ is abelian, $\rho(\Gamma)\subset{\C^*}^n$ acts trivially on the representations, and one has an action of $T^\Pi:={\C^*}^n/\rho(\Gamma)$. According to the toric formalism, one has $$T^\Pi = {\C^*}^n/\Image\rho = T^{\ker\hat\rho},$$ and thus $\Pi=\ker\hat\rho$. Note also that the exact sequence implies that the cone $C:=\Lambda_{{\mathbb R}+}$ is isomorphic to $({\pi\times\partial})(\Lambda^1_{{\mathbb R}+})$. Therefore, one has the following theorem. \begin{thm} \label{thm:slices} The slices $C_\zeta$ (which describe the toric moduli $X_\zeta$ when $\zeta$ is integral) are given by \begin{align} \label{eq:czeta} C_\zeta &\phantom{:}= \pi_{\mathbb R}(F_\zeta ),\\ \intertext{ where } \label{eq:fzeta} F_\zeta &:= {\mathbb R}_+^{{\mathcal Q}_1}\cap \partial^{-1}_{\mathbb R}(\zeta ). \end{align} \end{thm} The next section is devoted to the study of the polyhedra $F_\zeta $ and $C_\zeta $. \section{Flow Polyhedra} \label{sec:2:flow} The question of determining the polyhedra $C_\zeta$ is in fact a generalisation of a basic problem, well-known to network optimization specialists: \begin{problem}[The Transportation Problem ({\protect\cite{kenn_helg,gond_mino:graphs}})] \label{pb:transp} Given a quiver --- a ``network" in optimization parlance --- ${\mathcal Q}$ and an element $\zeta\in{\mathbb R}^{{\mathcal Q}_0}$ specifying given demands and supplies of some commodity at the vertices, find all\footnote{Actually, the interest in the transportation problem usually centers around finding the extreme flow which minimizes a certain cost function, not in finding \emph{all\/} extreme flows. Furthermore, there are usually capacity constraints for the maximal flow allowed along each arrow.} the non-negative elements $f\in{\mathbb R}_+^{{\mathcal Q}_1}$, representing flows of commodities along each arrow, whose net contribution at each vertex balances the demands and supplies specified by $\zeta$. \end{problem} Writing $\In(v)$ ($\Out(v)$) for the arrows which have their head (tail) at a given vertex $v$, then $\partial\colon{\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^{{\mathcal Q}_0}$ is given by $$\partial f(v)=\sum_{a\in\In(v)} f(a) - \sum_{a\in\Out(v)} f(a).$$ The element $\partial f$ is called the \emph{net contribution} of $f$. The flow $f\in{\mathbb R}^{{\mathcal Q}_1}$ is said to be a \emph{$\zeta$-flow} if \begin{equation} \partial f = \zeta. \label{eq:demand} \end{equation} If $f$ is non-negative, it is called an \emph{admissible $\zeta$-flow}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/flow7123.eps} \end{center} \caption[An admissible flow on the McKay quiver for $\frac{1}{ 7}(1,2,3)$ and its net contribution at each vertex]{An admissible flow on the McKay quiver for $\frac{1}{ 7}(1,2,3)$. (The numbers along the edges indicate the values of $f$ and the numbers at the vertices indicate the values of $\partial f=\zeta$.)} \label{fig:flow7123} \end{figure} The problem is therefore to determine the set $F_\zeta$ of all admissible flows for a given $\zeta$. Note that the solution set is empty unless $\zeta$ satisfies $\sum_{v\in{\mathcal Q}_0}\zeta(v)=0$ (i.e.\xspace supply = demand). Recall that the sub-space for which this holds was denoted $\Lambda^{0,0}_{\mathbb R}$. The set $F_\zeta$ is the intersection of the cone ${\mathbb R}_+^{{\mathcal Q}_1}$ (the first quadrant) with an affine translate of the vector-space $\ker\partial$, so $F_\zeta$ is a convex polyhedron; the interest lies in determining its extreme points. For instance, if one is trying to minimize a convex cost function of the flows, then the minimum will be attained at one of these extreme points. A solution to this problem is given by the easy: \begin{thm}\footnote{This theorem is the basis for the ``Simplex on a graph'' algorithm~\cite{luen}. See Proposition~3.16 in \cite{kenn_helg}, noting that trees correspond to what is called a ``basic solution'' or a ``basis'' for the linear programming problem~\cite[\S 2.3, Defn., p.17]{luen}. For the case of the permutation polytope, see~\cite[\S 5, Th.~1.1]{yem_kov_kra}} \label{thm:classical} The extreme points of $F_\zeta$ are precisely the admissible $\zeta$-flows whose support contains no cycles. \end{thm} Here, the \emph{support} of a flow is the set of arrows on which it takes non-zero values. A \emph{cycle} means a sequence of arrows in ${\mathcal Q}_1$ which form a cycle in the underlying graph to ${\mathcal Q}$, when their orientation is disregarded. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/uncycle7123.eps} \end{center} \caption{A cycle in the McKay quiver for $\frac{1}{7}(1,2,3)$} \label{fig:uncycle7123} \end{figure} \section{Generalised Transportation Problem} \label{sec:2:gtp} In the present paper, the interest is in the image $C_\zeta$ of $F_\zeta$ under the projection $\pi\colon {\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^n$, and one is therefore let to the following slightly more general problem. \begin{problem} \label{pb:labeledq} Given a quiver ${\mathcal Q}$, a natural integer $n$ and a projection $\pi\colon{\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^{n}$, characterize the (supports of the) flows which, under $\pi$, map to an extreme point of $\pi F_\zeta $ for a given $\zeta$. \end{problem} \begin{rmk} This problem can be thought of as a transportation problem with some extra structure: instead of associating a cost with \emph{each\/} arrow, one supposes that there are linear relations between the arrow costs which are parametrised by $n$ independent variables --- in other words, suppose that the cost function really is a (convex) function of these $n$ variables, so that the cost-minimizing flows will correspond to extreme points of $\pi F_\zeta$. \end{rmk} If $f$ is any flow with $\pi(f)=x\in \pi F_\zeta $, then the support $S$ of $f$ will be called a $\zeta$-\emph{configuration (of arrows)} for $x$, or simply a \emph{configuration} of $x$ (there will in general be \emph{many \/} $\zeta $-configurations for a given $x$). If $x$ is an extreme point of $\pi F_\zeta$, then any configuration $S$ of $x$ will be called an \emph{extreme configuration}. The problem is thus to determine all the possible extreme configurations. \subsection{Cycle Type and Closure} \label{sec:2:flow:cycle} In order to state the solution to the problem, one needs to introduce some concepts relating to cycles. Suppose that $c=(c_1,\dots, c_k)$ is a cycle, i.e.\xspace a sequence of arrows $c_i\in{\mathcal Q}_1$ such that they form a circuit, when suitably oriented. Consecutive arrows $c_i$'s are not allowed to be the same (although they can join the same vertices)\footnote{Note that it is necessary to state this explicitly for quivers, due to the possibility of several arrows joining the same two vertices.}. An arrow $c_i$ is said to \emph{agree} with the ordering $c_1,\dots c_k$ if the tail of $c_i$ is an extremity of $c_{i-1}$ and the head of $c_i$ is an extremity of $c_{i+1}$, and is said to \emph{disagree} otherwise. Call the disjoint union\footnote{The disjoint union is used in case some cycles contain the same arrow twice.} of the arrows $c_i$ whose direction agrees (disagrees) with the ordering $c_1,\dots,c_k$ the \emph{positive\/} (\emph{negative\/}) part of $c$ and denote it by $c^+$ ($c^-$). For each arrow $a\in{\mathcal Q}_1$, let $\chi_a$ denote the basis element of $\Z^{{\mathcal Q}_1}$ which takes the value $1$ on the arrow $a$ and zero elsewhere. To each cycle $c$, one can define the \emph{basic flow} associated to $c$ by $$\tilde\chi_c:=\sum_{a\in c^+}\chi_a - \sum_{a\in c^-}\chi_a\in\Z^n.$$ The \emph{type} of $c$ is defined to be element $\pi(\tilde\chi_c)\in\Z^n$. Cycles of \emph{type zero} are of special interest. They correspond to cycles having zero total number of arrows of any given type, the number being counted algebraically, according to whether the arrow agrees or disagrees with the orientation of $c$. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/nullcycle7123.eps} \end{center} \caption[A cycle of type zero in the McKay quiver for $\frac{1}{ 7}(1,2,3)$]{A cycle of type zero in the McKay quiver for $\frac{1}{ 7}(1,2,3)$. (The numbers next to the arrows indicate their type.)} \label{fig:nullcycle7123} \end{figure} The \emph{cycle closure} of a configuration $S$ is defined to be the smallest over-set $\clos{S}$ of $S$ such that, for any cycle $c$ in ${\mathcal Q}$ of type $0$, \begin{equation} c^+\subseteq \clos{S}\iff c^-\subseteq \clos{S}. \label{eq:cycle_closure} \end{equation} The closure of $S$ can be computed by searching for all the cycles $c$ of type $0$ satisfying $c^+\subseteq S$, adjoining $c^-$ to $S$, and repeating this procedure until~\eqref{eq:cycle_closure} is satisfied. Note that, even if $S$ contains no cycles of non-zero type (for instance, if $S$ is a tree), the arrows one adjoins may create such cycles in $\clos{S}$. Two configurations $S,S'$ will be called \emph{equivalent} (written $S\sim S'$) if $\overline{S}=\overline{S'}$. \subsection{Statements of the Generalised Theorems} \label{sec:2:flow:state} The generalised version of Theorem~\ref{thm:classical} can now be stated. \begin{thm}[Generalised Extreme Flows] \label{thm:general} The extreme points of the polyhedron $\pi F_\zeta$ are the images under $\pi$ of the admissible $\zeta$-flows whose supports have no cycles of non-zero type in their closures. If $S$ is a such a configuration and if there is a $\zeta$-flow with support $S$, then the image of that flow under $\pi$ is an extreme point of $\pi F_\zeta$. \end{thm} Note that if $\pi$ is the identity map, then \emph{any\/} cycle is of non-zero type and one recovers Theorem~\ref{thm:classical}. The above theorem can be generalised to get a description of the faces of $\pi F_\zeta $ of all dimensions. Recall that the \emph{tangent cone\/} of $P$ at one of its faces $F$ is the convex cone $$T_F P:={\mathbb R}_+(P-F):={\mathbb R}_+(P-f), \text{ for any }f\in \interior F,$$ where $\interior F$ denotes the relative interior of the face $F$. For any configuration of arrows $S$ define $$ F_\zeta(S) := \{f\in\partial^{-1}(\zeta) : a\not\in S \implies f(a) \geq 0\}$$ and $$Z_\zeta(S) := \{f\in\partial^{-1}(\zeta) : a\not\in S \implies f(a) = 0\}.$$ One has $F_\zeta(\emptyset)=F_\zeta$ and $Z_\zeta(S)=F_\zeta\cap\supp^{-1}(S)$. When $\zeta=0$, $F_0(S)$ is a cone and $Z_0(S)$ is its maximal vector subspace. The cone $F_0(S)$ (resp.\xspace the vector-space $Z_0(S)$) is generated over ${\mathbb R}_+$ by the flows $\tilde\chi_c$ for cycles $c$ such that $c^-\subseteq S$ (resp.\xspace $c\subseteq S$). Let ${\mathcal C}$ denote the set of all \emph{configurations},\/ i.e.\xspace the set of non-empty subsets of ${\mathcal Q}_1$. The \emph{rank} of $S$ is defined to be the rank of $\pi Z_0(\overline S)$. It is trivial to see that the rank function determines a partition of ${\mathcal C}$ into non-empty disjoint sets ${\mathcal C}={\mathcal C}^0\amalg{\mathcal C}^1\amalg\dots\amalg{\mathcal C}^n$ which respects the equivalence relation $\sim$ induced by $S\mapsto\overline{S}$. As a further piece of notation, for any subset of ${\mathcal D}\subseteq{\mathcal C}$, denote by ${\mathcal D}^k$ the subset of configurations in ${\mathcal D}$ which have rank $k$. Also, write ${\mathcal D}_\zeta$ for the subset of configurations which are \emph{admissible for $\zeta$},\/ namely configurations $S\in{\mathcal D}$ which arise as the support of some element in $F_\zeta$. Theorem~\ref{thm:general} says that the configurations corresponding to the extreme points of $F_\zeta$ are precisely those belonging to ${\mathcal C}^0_\zeta$. In general one has the following complete description of the extreme faces and tangent cones of $C_\zeta$: \begin{thm}[Extreme Faces and Tangent Cones] \label{thm:faces} For all $\zeta$, the map \map{\text{Face}_\zeta}{{\mathcal C}_\zeta}{\text{Faces of }\pi F_\zeta}{S}{\pi F_\zeta \cap(\pi f+\pi Z_0(\overline{S})))} is independent of the choice of $f\in F_\zeta\cap\supp^{-1}(S)$, and induces a bijection $$\text{Face}_\zeta\colon {\mathcal C}^k_\zeta/\!\!\sim \xrightarrow{\;\cong\;} \text{$k$-faces of }\pi F_\zeta.$$ Furthermore, for all $[S]\in{\mathcal C}^k_\zeta/\!\!\sim$, $$T_{\text{Face}_\zeta(S)}\pi F_\zeta =\pi F_0(\overline{S}),$$ where the left-hand side denotes the tangent cone to the polyhedron $\pi F_\zeta$ at the face $\text{Face}_\zeta(S)$. In other words, $\pi F_0(\overline{S})$ gives the tangent cone corresponding to the configuration $S$ (which is independent of the value of $\zeta$) and $\text{Face}_\zeta$ gives the corresponding face of $C_\zeta$ (whose direction is also independent of $\zeta$). \end{thm} \begin{rmk} Note that both the direction of the face corresponding to $S$ and the tangent cone of $\pi F_\zeta$ at this face are independent of the value of $\zeta$. \end{rmk} In fact, (see Lemma~\ref{lemma:0cycles}) $Z_0(S)$ is generated by the flows for the cycles supported in $S$, so that Theorem~\ref{thm:general} corresponds to the case when $S$ has rank zero. The concepts needed for the proof of Theorem~\ref{thm:faces} are developed in Section~\ref{sec:2:flow:basic}. The theorem is then proved in Section~\ref{sec:2:flow:classpf} for the classical case (where ${\mathbb R}^{{\mathcal Q}_1}={\mathbb R}^n$ and $\pi\colon{\mathcal Q}_1\to{\mathcal Q}_1$ is the identity). The general case follows easily from this and is treated in Section~\ref{sec:2:flow:genpf}. Before this, an example of an application and several important corollaries are given. \section{Application to the McKay Quiver} \label{sec:2:mckay} \subsection{Example} \label{sec:2:flow:examples} Consider the McKay quiver for the action $\frac{1}{ 5}(1,2,3)$. Recall that identifying ${\mathcal Q}_0$ with $\Z_5$, the arrows are $a_v^i := v\to v-i$ for $v\in \Z_5$ and $i\in\{1,2,3\}$. Let $\pi $ be the map ${\mathcal Q}_1 \to\{1,2,3\}$ which assigns to each arrow $a_v^i$ its type $i$. This induces a projection $\pi\colon {\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^3$ which assigns to the basis element $\chi_v^i := \chi_{a_v^i}$ the basis element $e_i$ of ${\mathbb R}^3$. With respect to this map, the cycles of type zero are those with total number of arrows of any given type equal to zero, where the number of arrows is counted algebraically according to the orientation of the cycle. Consider the configuration $T\subset{\mathcal Q}_1$ represented in Figure~\ref{fig:tree5123}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 1in \epsfbox{fig/tree5123.eps} \end{center} \caption{A configuration $T\in{\mathcal C}^0$ in the McKay quiver for $\frac{1}{ 5}(1,2,3)$.} \label{fig:tree5123} \end{figure} What is the closure of~$T$? A little thought shows\footnote{See Section~\ref{sec:2:examples:comm} for more details on this.} that $\clos{T}$ is simply $T$ itself. Since $\clos{T}$ has no cycles, $Z_0(\clos{T})=0$, so $T\in{\mathcal C}^0$. If $T$ is admissible for $\zeta $, and $f$ is any $\zeta $-flow with support $T$, then the theorem says that $\pi(f)$ is a $0$-face of $\pi F_\zeta $, i.e.\xspace an extreme point. Furthermore, the tangent cone to $\pi F_\zeta $ at $\pi(f)$ is the cone $\pi F_0(T)$; this is generated by the types of the cycles whose negative part is contained in $T$. These are listed in Figure~\ref{fig:cycles5123}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 7.5cm \epsfbox{fig/cycles5123.eps} \end{center} \caption{The different types of cycles which generate the cone $\pi F_0(T)$ for the configuration $T$ in Figure~\ref{fig:tree5123}.} \label{fig:cycles5123} \end{figure} There are four different types\nopagebreak \begin{align*} v_1 &= (1,1,-1)\\ v_2 &= (1,-2,1)\\ v_3 &= (-1,0,2)\\ v_4 &= (-1,3,0), \end{align*} and they generate $\pi F_0(T)$ as a cone. Any three of these form a basis for the lattice $\Pi\subset\Z^3$ of index $5$ given by $$\Pi:=\ker\hat\rho = \{(a,b,c): a+2b+3c \equiv 0\pmod 5\},$$ and they satisfy the single relation $v_1+v_3 = v_2+v_4$. This corresponds via the usual toric formalism to a singularity of $X_\zeta$ of the type $xw=yz \subset{\mathbb C}^4$. The variety $X_\zeta$ therefore has such a singularity whenever the tree $T$ is admissible for $\zeta$, i.e.\xspace whenever the flow $f_\zeta(T)$ is positive. Writing this out explicitly, one sees that the admissible cone for $T$, ${{\Lambda^{0,0}_{{\mathbb R}}}}(T)$ is given by the following inequalities\footnote{See \ref{sec:2:mckay:flow:trees} for the exact definition of ${{\Lambda^{0,0}_{{\mathbb R}}}}(T)$.} \begin{align*} -\zeta_0 &>0\\ -\zeta_2 &>0\\ -\zeta_4 &>0\\ -\zeta_4-\zeta_1 &>0 \end{align*} \subsection{Corollaries} \label{sec:2:flow:cor} \subsubsection{$k$-Adjacency} The ``simplex on a graph'' algorithm~\cite[Alg.~3.3]{kenn_helg} can be interpreted as moving from one extreme point of $F_\zeta$ to an adjacent one by varying the flow along a single cycle in the network\footnote{See~\cite[\S 5.4]{evans}, or, for the special case of the ``permutation polytope,''~\cite[\S 5, Th.~1.3]{yem_kov_kra}.}. Geometrically this says that two extreme flows are joined by an edge of $F_\zeta$ if and only if the union of their supports contains only \emph{one\/} cycle (up to multiples obtained by going around the cycle $k\in\Z$ times). In order to generalize the notion of being joined by an edge, the following terminology is introduced: two points of a polyhedron are called \emph{$k$-adjacent} if they are contained in a $k$-dimensional face, but in no face of smaller dimension. Note that by convexity, two points are contained in the same face if and only if their midpoint is also in that face. Since a configuration of the midpoint is obtained by taking the union of configurations of the two points, the answer to the question is deduced as a corollary to Theorem~\ref{thm:faces}: \begin{cor}[$k$-Adjacency] \label{cor:k_adjacent} Two points $x,x'\in\pi F_\zeta$ are $k$-adjacent if and only if for some (and hence for any) configurations $S$ of $x$ and $S'$ of $x'$, $${S\cup S'}\in{\mathcal C}^k,$$ i.e.\xspace their union has rank $k$. \end{cor} Note that the case~ $k=0$ gives the condition for two configurations to give rise to the same extreme point. The case $k=1$ says that two configurations are joined by an edge of~ $\pi F_\zeta $ if and only if their union only contains cycles whose types are multiples of a fixed element $v\in{\mathbb R}^n$. For the case when $\pi$ is the identity, this reduces to the classical statement that the union $S\cup S'$ contains only one cycle (up to multiples obtained by going around the cycle $k\in\Z$ times). \subsubsection{Trees} \label{sec:2:mckay:flow:trees} It can be shown easily (see Corollary~\ref{cor:spanning_trees}) that any extreme point of~ $\pi F_\zeta$ has at least one configuration~ $T$ which is a \emph{tree}, i.e.\xspace it has no cycles. Thus, to know the extreme points and the tangent cones of $\pi F_\zeta$ at these points, one need only answer the question of which trees occur as extreme configurations. Denoting the set of trees in ${\mathcal C}$ by ${\mathcal T}$, the generalised theorem implies that the extreme points of $C_\zeta$ are given by the images of the flows whose supports are members of~ ${\mathcal T}^0_\zeta$. This set is easy to determine: for a given $\zeta$, there is, on each tree $T$, a unique flow $f_\zeta(T)\in\Lambda^1$ which meets the demand $\zeta$ and is supported on $T$; $f_\zeta(T)$ is called the \emph{($\zeta$-)flow admitted by the tree\/} $T$. (See \ref{sec:2:examples} for an example). This defines a map $$f_\zeta\colon{\mathcal T}\to\Lambda^1.$$ The set $f_\zeta^{-1}(\Lambda^1_+)$ of trees $T$ which admit non-negative $\zeta$-flows is of course the set ${\mathcal T}_\zeta$ of \emph{admissible tree}\emph{s\/} for $\zeta $. The classical Theorem~\ref{thm:classical} says that $$\ext F_\zeta = f_\zeta({\mathcal T}_\zeta).$$ Similarly, the generalised Theorem~\ref{thm:general} gives the following: \begin{cor} \label{cor:number_extreme_pts} Let $f_\zeta$ be the map which assigns to each tree $T$ the unique $\zeta$-flow with support equal to $T$. Then the extreme points of $C_\zeta$ are given by $$\ext \pi F_\zeta = \pi f_\zeta ({\mathcal T}^0_\zeta ),$$ where ${\mathcal T}^0 := {\mathcal C}^0\cap {\mathcal T}$ is the set of trees whose closures have type zero. Furthermore, Theorem~\ref{thm:faces} implies that $$\card{\ext C_\zeta}=\card{{\mathcal C}^0_\zeta/\!\!\sim}=\card{{\mathcal T}^0_\zeta/\!\!\sim},$$ where $\sim$ is the equivalence relation induced by closure. \end{cor} Note (see Lemma~\ref{lemma:maximal_config} for a proof) that for the sets $S\in {\mathcal C}^0$ one has $\pi F_0(\clos{S})=\pi F_0(S)$. This implies the following corollary: \begin{cor} \label{cor:fan} The extreme points of the polyhedron $\pi F_\zeta$ correspond to the trees $T$ in ${\mathcal T}^0_\zeta$ and the tangent cone to $\pi F_\zeta$ at the point corresponding to $T$ is $\pi F_0(\clos{T})=\pi F_0(T)$. Thus the fan\footnote{Recall that the \emph{fan \/} associated to a convex polyhedron $P$ is the collection of dual cones to the tangent cones of $P$ at all its faces.} $\Sigma _\zeta$ associated to the polyhedron $\pi F_\zeta $ is given by the dual cones $\pi F_0(T)^\vee$ for the trees $T\in{\mathcal T}^0_\zeta$ and all their faces. In fact, its $k$-skeleton, i.e.\xspace the set of its $k$-dimensional cones, is $\Sigma _\zeta^{(k)} :=\{\pi F_0(\clos{S})^\vee : S\in{\mathcal C}^{n-k}_\zeta \}$. \end{cor} \begin{rmk} This corollary says that the singularities of $C_\zeta$ are precisely those given (with respect to the lattice $\Pi$) by the cones $\{\pi F_0(T)| T\in{\mathcal T}^0_\zeta\}$. \end{rmk} \subsubsection{Variation of the Flow Polyhedra with $\zeta $} \label{sec:2:flow:cor:var} The following corollary of Theorem~\ref{thm:faces} describes when two different values of $\zeta$ give isomorphic polyhedra: \begin{cor} \label{cor:iso_C_zeta} If $\zeta$ and $\zeta'$ have the same admissible configurations (${\mathcal C}_\zeta={\mathcal C}_{\zeta'}$) or even just the same admissible trees (${\mathcal T}_\zeta={\mathcal T}_{\zeta'}$) then the corresponding polyhedra $C_{\zeta}$ and $C_{\zeta'}$ are geometrically isomorphic. Two polyhedra are said to be geometrically isomorphic is they are combinatorially isomorphic and their tangent cones at the corresponding faces are identical. In particular, their associated fans are identical, and their corresponding toric varieties are isomorphic. \end{cor} Note that when one multiplies $\zeta$ by a non-zero number, the polyhedron $C_\zeta$ is simply scaled-up. Fixing a tree $T\in{\mathcal T}^0$ determines an open convex cone ${{\Lambda^{0,0}_{{\mathbb R}}}}(T)\subseteq \Lambda^{0,0}_{\mathbb R}$ of values of $\zeta$ for which $T$ is $\zeta$-admissible. The cone ${{\Lambda^{0,0}_{{\mathbb R}}}}(T)$ is called the \emph{admissible cone of $T$}.\/ If ${\mathcal E}$ is a fixed subset of ${\mathcal T}^0$, then the condition ${\mathcal T}^0_\zeta={\mathcal E}$ defines an open cone in the $\zeta$-parameter space $\Lambda^{0,0}_{\mathbb R}$. As ${\mathcal E}$ varies, one obtains a partition of $\Lambda ^{0,0}_{\mathbb R}$ into a union of open cones inside which the polyhedra $C_\zeta $ are geometrically isomorphic. \subsubsection{Degeneracies} Another fact which will be of interest to us is that, for {\em generic\/} values of $\zeta$, any $\zeta$-flow has a support which is a spanning subgraph of the quiver, i.e.\xspace which connects any two vertices. This is written ${\mathcal C}_\zeta\subset{\mathcal C}_{\spn }$. For instance, the faces of ${{\Lambda^{0,0}_{{\mathbb R}}}}(T)$ consist of degenerate values of $\zeta$ for which some extreme $\zeta$-flows have supports which are strict subsets of $T$ and so cannot be spanning subsets. In the next section, basic flows associated to paths and cycles are studied in more detail. They provide the key to the proofs of the other results. \section{Proofs} \label{sec:2:flow:proofs} \subsection{Basic Flows} \label{sec:2:flow:basic} Many proofs in the context of network flows use the basic technique of decomposing a flow into certain basic components associated to paths in ${\mathcal Q}$. A \emph{path} in ${\mathcal Q}$ means a sequence $p=(p_1,\dots,p_k)$ of arrows in ${\mathcal Q}_1$ which form a connected path in the underlying graph to ${\mathcal Q}$, once their orientation has been disregarded. Consecutive $p_i$'s are not allowed to be the same (although they can join the same vertices). As for cycles, the disjoint union of the arrows of the path $p=(p_1,\dots,p_k)$ which agree\footnote{See Section~\ref{sec:2:flow:cycle} for the definition of \emph{agree.\/}} (resp.\xspace disagree) with the sense of traversal specified by the sequence $p_1,\dots,p_k$ are called the \emph{positive} (resp.\xspace\emph{negative}) arrows of $p$ and denoted $p^+$ (resp.\xspace $p^-$). A path will sometimes be confused with its set of arrows, for instance in statements such as ``a path $p$ is {\em in\/} a set $S\subseteq{\mathcal Q}_1$ (written $p\subseteq S$)'' which means of course that all its arrows belong to the set $S$. To each path $p$ the \emph{basic}\footnote{Basic flows are also termed \emph{simple flows}~\cite{busacker_saaty:graphs} or \emph{elementary flows} by other authors.} flow $\tilde\chi_p\colon{\mathcal Q}_1\to\Z$ is defined by \begin{equation} \tilde\chi_p(a)=\sum_{a\in p^+}\chi_a - \sum_{a\in p^-}\chi_a\in\Z^n. \label{eq:basic_flow} \end{equation} Note that if $p$ is a path from $v$ to $v'$, then $\partial \tilde{\chi}_p=\chi _{v'}-\chi _v$. Conversely, one has the following lemma. \begin{lemma} \label{lemma:basicf_path} If $f$ is an integral flow with $\partial f=\chi _{v'}-\chi _v$ for two vertices $v,v'$ of ${\mathcal Q}$, then there exists a path $p$ from $v$ to $v'$ such that $p^{\pm}\subseteq\supp f^\pm$. \end{lemma} \begin{proof} By induction on the $1$-norm of $f$: $\norm{f}_1:=\sum_{a\in{\mathcal Q}_1}\abs{f(a)}$. If $\norm{f}_1=1$ then, obviously, $f=\chi _a$ or $-\chi_a$, for some arrow $a\in{\mathcal Q}_1$ which is either $v\to v'$ or $v'\to v$ respectively. Either way, $f=\tilde{\chi }_{p}$ for the corresponding one-arrow path $p$ from $v$ to $v'$. Now suppose that $\norm{f}_1>1$ and that $\partial f=\chi _{v'}-\chi _v$. There must be an arrow $a$ such that one of the following statements holds \begin{enumerate} \item[(1)] $t(a)=v$ and $f(a)>0$, or \item[(2)] $h(a)=v$ and $f(a)<0$. \end{enumerate} If (1) holds then the flow $f'=f-\chi _a$ satisfies the induction hypothesis with $\partial f'=\chi _{v'}-\chi _{h(a)}$, so there exists a path $p'\subseteq\supp f'$ from $h(a)$ to $v'$. But then $p=ap'$ is a path from $v$ to $v'$ with $p^+ = p^{\prime +}\cup \{a\}\subseteq\supp f^{\prime+}\cup\{a\} = \supp f^+$. Case (2) follows in a similar way setting $f'=f+\chi _a$ and $p=(-a)p$. \end{proof} Obviously, if $p$ is actually a cycle, then the basic flow associated to $p$ is a $0$-flow (it satisfies $\partial \tilde\chi_p = 0$), and is supported in $p$. Figure~\ref{fig:basicf7123} gives an example of such a flow. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/basicf7123.eps} \end{center} \caption{A basic $0$-flow associated to the cycle in Figure~\protect\ref{fig:uncycle7123}} \label{fig:basicf7123} \end{figure} The converse to this statement is given by the following lemma~\cite[Th.~7.2]{busacker_saaty:graphs}, \cite[{\S5.1.6,Th.~3}]{gond_mino:graphs}, . \begin{lemma}[Decomposition into basic flows] Any\/ $0$-flow $f$ can be decomposed into a positive linear combination of basic flows for finitely many cycles $c_i$ such that $c^\pm_i\subseteq \supp f^\pm$. \label{lemma:0cycles} \end{lemma} \begin{proof} Let $f\in\ker\partial$ and $S$ be its support. If $S$ contains a cycle $c$ then let $v$ be a vertex such that $c^+$ has an arrow $a$ with $t(a)=v$. Then $f':= f-f(a)\tilde\chi_c\in\ker\partial$ and $\supp f' = \supp f \setminus \{a\}$ is strictly smaller than $\supp f$. Repeating this reasoning finitely many times, one obtains a flow $f'\in\ker\partial$ such that $S'=\supp f'$ contains no cycles. Suppose $S'$ is non-empty, and let $p$ be a path in $S'$ which is \emph{maximal, \/} i.e.\xspace is not contained in any strictly larger path. Let $p$ go from vertex $v$ to vertex $v'$ and $a$ denotes the first arrow of $p$ (the one joined to $v$). Then this is the only arrow in the quiver which is joined to $v$ and on which $f'$ is non-zero, so $\partial f'(v)$ must be non-zero. Since this is impossible, $S'$ must be empty, and $f=\sum_i x_i \tilde\chi_{c_i}$ for some cycles $c_i\subseteq S$ and real numbers $x_i$. Replacing $c_i$ by $-c_i$ if necessary, one may assume $x_i>0$. (Note also that if $f$ is integral to start with, the $x_i$ are also integral.) I claim that these cycles may be chosen in such a way that their orientations are \emph{conformal}, i.e.\xspace such that any two cycles $c_i,c_j$ satisfy $c^+_i\cap c^-_j=\emptyset$. Indeed suppose that $c_i$ and $c_j$ are not conformal, and suppose, say, $x_i\geq x_j$. On the subset $U\subseteq{\mathcal Q}_1$ on which their orientations disagree, the sum of $x_ic_i$ and $x_jc_j$ will cancel each other out, and only $x_i-x_j$ units of flow will survive. The complement $c_i\cup c_j\setminus U$ will consist of one or more disjoint cycles $d_1,\dots,d_k$, all conformal to $c_i$. Thus upon adding the two basic flows $$x_i\tilde\chi_{c_i}+x_j\tilde\chi_{c_j} = (x_i-x_j)\tilde\chi_{c_i} + x_j(\tilde{\chi }_{d_1}+\dots+\tilde{\chi }_{d_k}),$$ giving $k+1$ conformal cycles with all coefficients positive or zero. Repeating this procedure for all the pairs of non-conformal cycles gives the desired conformal decomposition. Now writing $$f=\sum_i\tilde\chi_{c_i}=\sum_i (\chi_{c^+_i}-\chi_{c^-_i}),$$ and since $c^+_i\cap c^-_j=\emptyset$, one has $$f^\pm = \sum_i\chi_{c^\pm_i},$$ which is the desired result. \end{proof} \begin{rmk} The decomposition lemma above can be viewed as a consequence of the following homological argument. Regard ${\mathcal Q}$ as a CW-complex, and for each cycle $p$ in ${\mathcal Q}$, adjoin a 2-cell whose boundary is the element $\tilde\chi_{p}$. Then the resulting CW-complex $\widetilde{\mathcal Q}$ has $H^1(\widetilde{\mathcal Q})=0$. An elementary proof was given because it illustrates the basic technique of decomposing a flow into basic flows. \end{rmk} \subsection{Proof of Theorem~\protect\ref{thm:faces} --- Classical Case} \label{sec:2:flow:classpf} The classical analog of Theorem~\ref{thm:faces} (i.e.\xspace the case when $\pi$ is the identity map) can now be proved. The statement says that the $k$-dimensional faces of $F_\zeta$ correspond to the $\zeta $-configurations $S$ such that $\rk Z_0(S)=k$, i.e.\xspace whose cycles span a lattice of rank $k$. This theorem follows from the following four facts: \begin{fact} \label{fact:1} If $f\in F_\zeta$ then the cone ${\mathbb R}_+(F_\zeta-f)$ contains a subspace of dimension $k$ if and only if $f$ is contained in the (relative) interior of a face of dimension $k$ (and hence in no lower dimensional face). \end{fact} \begin{fact} \label{fact:2} For any $f\in F_\zeta$, one has ${\mathbb R}_+(F_\zeta-f)= F_0(\supp f)$. \end{fact} \begin{fact} \label{fact:3} The maximal vector subspace in the cone $ F_0(S)$ is $Z_0(S)$. \end{fact} \begin{fact} \label{fact:4} The subspace $Z_0(S)$ is generated by the basic flows for the cycles supported in $S$. \end{fact} Fact~\ref{fact:1} follows from the definition of a $k$-dimensional face, \ref{fact:3} is trivial and~\ref{fact:4} follows from Lemma~\ref{lemma:0cycles}. Fact~\ref{fact:2} is proved in the following lemma. \begin{lemma} \label{lemma:tangent_fzeta} If $f\in F_\zeta$, then ${\mathbb R}_+(F_\zeta -f) = F_0(\supp f)$. \end{lemma} \begin{proof} Let $f\in F_\zeta$, so that $F_\zeta=(f+\ker\partial)\cap{\mathbb R}_+^{{\mathcal Q}_1}$. The non-negative multiples of elements in $(f+\ker\partial)\cap{\mathbb R}_+^{{\mathcal Q}_1} -f$ belong to $\ker\partial$ and are non-negative outside $\supp f$, so belong to $ F_0(\supp f)$. Conversely, suppose $m\in F_0(\supp f)$. Then $\partial(f+\epsilon m)=\zeta $ for all $\epsilon\in{\mathbb R}$, and it suffices to show that there exists $\epsilon >0$ such that $f+\epsilon m\in {\mathbb R}_+^{{\mathcal Q}_1}$. But this follows because $f$ is bounded below by a positive number on $\supp f$, whereas $m\geq 0$ outside $\supp f$. \end{proof} \subsection{Proof of Theorem~\protect\ref{thm:faces} --- General Case} \label{sec:2:flow:genpf} In order to prove the general case of Theorem~\ref{thm:faces}, a bit more has to be said about the various configurations which can occur for a point $x\in\pi F_\zeta$. \begin{lemma} \label{lemma:maximal_config} If $S$ is a $\zeta $-configuration for $x$ then so is $\clos{S}$. Furthermore, all $\zeta $-configurations for $x$ have the same closure. \end{lemma} \begin{proof} Let $f\in F_\zeta$ be a flow such that $\pi f=x$ and let $S=\supp f$. The proof begins by constructing a flow $\overline f\in F_\zeta\cap \pi^{-1}(x)$ whose support is $\clos{S}$. If $c$ is a cycle of type $0$ in ${\mathcal Q}$ such that $c^-\subseteq S$ and $c^+\not\subseteq S$, one can add a small positive multiple of $\tilde\chi_c$ to $f$ and obtain a non-negative flow $f'$. Since $c$ is a cycle, $\partial\tilde\chi_c=0$, and since $c$ has type $0$, $\pi f=\pi f'$ and hence $f'\in F_\zeta$. Continuing in this way until all cycles of type $0$ satisfy~\eqref{eq:cycle_closure}, one obtains the required flow $\overline f$. For the second statement of the lemma, note that if $f,f'$ are two elements of $F_\zeta\cap\pi^{-1}(x) $, then $\partial(f-f')=0$, so by Lemma~\ref{lemma:0cycles}, has a decomposition into basic flows for cycles $c_i$: $$f-f' = \sum_i x_i\tilde\chi_{c_i},$$ with $c_i^+\subseteq f$, $c_i^-\subseteq f'$ and $x_i>0$. Now $c_i^+\subseteq S\implies c_i^-\subseteq \overline S$, so $$f'=f+\sum_i \tilde\chi_{c_i}$$ implies that $S'\subseteq \overline S$. By symmetry, one also has $S\subseteq \overline{S'}$, and so $\overline S = \overline{S'}$. \end{proof} A little corollary needed later is \begin{cor} \label{cor:spanning_trees} Any $S\in{\mathcal C}^0$ contains a tree $T$ such that $\clos{T}=\clos{S}$. \end{cor} \begin{proof} Let $S=\supp f$ be a configuration for $x$. Eliminate any cycles in the support of $f$ by adding basic flows corresponding to those cycles, as in the proof of Lemma~\ref{lemma:maximal_config}. The resulting flow $f'$ is supported in a tree $T$ and, since all the basic flows have type zero it satisfies $\pi(f')=x$. Lemma~\ref{lemma:maximal_config} shows that $\clos{T}=\clos{S}$. \end{proof} \begin{proof}[Proof of the general Theorem~\ref{thm:faces}] Consider a point $x\in\pi F_\zeta$. The cone ${\mathbb R}_+(\pi F_\zeta-x)$ gives the tangent cone to $\pi F_\zeta$ at the minimal face of $F_\zeta$ containing $x$. It is obtained by taking the union over all $f\in F_\zeta\cap\pi^{-1}(x)$ of $\pi {\mathbb R}_+(F_\zeta-f)$. By Lemma~\ref{lemma:tangent_fzeta} this gives $$\bigcup\{\pi F_0(S) : S \text{ a $\zeta $-configuration for }x\},$$ which by Lemma~\ref{lemma:maximal_config} is $\sigma(S)=\pi F_0(\clos{S})$, for any $S=\supp f$ and $f\in F_\zeta\cap\pi^{-1}(x)$. The minimal face of $\pi F_\zeta$ containing $x$ is given by intersecting $\pi Z_0(\clos{S})$, the largest subspace in the cone $\pi F_0(\clos{S})$, with the tangent cone of $\pi F_\zeta $ at $x$ and then translating by $x$. This gives $$x+(\pi F_\zeta )_x \cap Z_0(\clos{S}),$$ which is precisely the face $\text{Face}_\zeta(S)$ mentioned in the theorem and all faces of $\pi F_\zeta$ are obtained in this way. It is obvious that equivalent configurations have the same rank and give the same face, so the partition of ${\mathcal C}$ has the stated properties and $\text{Face}_\zeta$ is a bijection ${\mathcal C}^k_\zeta/\!\!\sim\to k\text{-faces}$. \end{proof} \section{Exactness Results} \label{sec:2:exact} The exactness of the sequence~\eqref{eq:exact1} is proven in this section. \subsection{Basic Flows: Sequential Notation} \label{sec:2:exact:sequ} \nopagebreak We introduce some notation which is convenient to describe basic flows. This is used in~ Section~\ref{sec:2:exact:exact} to give an elementary proof of the exactness of~\eqref{eq:exact1}. For $v\in{\mathcal Q}_0$ and $j\in\{1,\dots,n\}$, write $\chi_v^j$ for the basis element $\Lambda ^1$ which is the indicator function of the singleton $\{a_v^j\}\subset{\mathcal Q}_1$. Define the following symbols: \begin{align} \{v\}(j) &:= \chi_v^j\\ \{v\}(-j) &:= -\chi_{v+w_j}^j. \end{align} For $k>0$, and $j_0,\dots,j_k\in\pm\{1,\dots,n\}$, define $$ \{v\}(j_0,\dots,j_k):=\{v\}(j_0)+\{v-w_{j_0}\}(j_1,\dots,j_k).$$ Also define $(j)\{v\} := \{v+w_j\}(j)$ and $$(j_k,\dots,j_0)\{v\}=(j_k,\dots,j_1)\{v+w_{j_0}\} + (j_0)\{v\}.$$ This sequential notation is designed especially for representing basic flows and has the advantage of including only the relevant information. For instance, if $p=(p_1,\dots,p_k)$ is a path in ${\mathcal Q}$, then the basic flow associated to $p$ is, in this notation, $$\tilde\chi_p=\{t(p_1)\}(j_1,\dots,j_k),$$ where $$j_i:= \begin{cases} \phantom{-}\pi(p_i),\qquad & p_i\in p^+\\ -\pi(p_i),\qquad & p_i\in p^- \end{cases} $$ For example, the basic flow represented in Figure~\ref{fig:basicf7123} can be written as $$\{0\}(1,1,2,1,1,-3,-3)$$ in this notation. Figure~\ref{fig:bfl11124} shows yet another example. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/bfl11124.eps} \end{center} \caption{The basic flow $\{0\}(1,1,3,-2,1,3,-1,-1)$ in the McKay quiver for $\frac{1}{ 11}(1,2,4)$.} \label{fig:bfl11124} \end{figure} Often, when there is a need to specify both endpoints, the notation $\{v\}(j_0,\dots,j_k)=:\{v\}(j_0,\dots,j_k)\{v'\}$ with $v'=v-\sum w_{j_i}$ will be adopted. The following identities are easily checked for any $v\in{\mathcal Q}_0$ and $j,j_i\in\pm\{1,\dots,n\}$: \begin{gather} (j)\{v\}+\{v\}(-j) = 0 \label{eq:v-id1},\\ \{v\}(j_0,\dots,j_k,-j_k,\dots,-j_0)=0 \label{eq:v-id2},\\ \{v\}(j_0,\dots,j_k)\{v'\}+\{v'\}(j_{k+1},\dots,j_l)= \{v\}(j_0,\dots,j_l), \label{eq:v-id3} \end{gather} where the last identity holds of course for $v'=v-\sum_{i=0}^k w_{j_i}$. If $\sum_i w_{j_i}=0$, i.e.\xspace the corresponding flow $\{v\}(j_1,\dots, j_k)$ is a $\partial$-closed flow associated to a cycle, there is another identity which allows us to cyclically permute the indices: \begin{equation} \label{eq:v-id4} \{v\}(j_0,\dots,j_k)=\{v-w_{j_0}\}(j_1,\dots,j_k,j_0). \end{equation} Recall that $\Lambda^2$ is the sub-lattice of $\Lambda^1$ generated by the elements corresponding to the commutation relations. These can be written in various notations: $$r^{ij}_v=\chi_v^i +\chi_{v-w_i}^j-\chi_v^j-\chi_{v-w_j}^i= \{v\}(i,j,-i,-j).$$ Putting these identities together gives the following lemma. \begin{lemma} For any permutation $\sigma$ of $\{0,\dots,k\}$ one has $$\{v\}(j_{\sigma(0)},\dots,j_{\sigma(k)}) = \{v\}(j_0,\dots,j_k)\mod \Lambda^2.$$ \label{lemma:permutation} \end{lemma} \vspace*{-\belowdisplayskip} \vspace*{-\topsep} \vspace*{-\partopsep} \begin{proof} It is enough to show that the equality holds for any transposition $\sigma=(q,q+1)$ of consecutive elements. Let $j_q=i$ and $j_{q+1}=l$. By the identities \eqref{eq:v-id2}--\eqref{eq:v-id3} one has $$ \{v\}(j_0,\dots,j_k) = \{v\}(j_0,\dots,j_{q-1})+\{v_q\}(i,l)+ \{v_{q+2}\}(j_{q+2},\dots,j_k), $$ where $v_a:=v-\sum_{i=0}^{a-1} w_{j_i}$ for any $a\in\{0,\dots, k\}$. Using $\{v_q\}(l,i,-l,-i)+\{v_q\}(i,l)=\{v_q\}(l,i)$, one sees that $$\{v\}(j_{\sigma(0)},\dots,j_{\sigma(k)})-\{v\}(j_0,\dots,j_k) =\{v_q\}(l,i,-l,-i).$$ Now if $i,l\in\{1,\dots,n\} $ then the result follows because $\{v_q\}(l,i,-l,-i)=r^{il}_{v_q}$. If on the other hand $-i,l\in\{1,\dots,n\} $, then the result is true because \begin{align*} \{v_q\}(l,i,-l,-i) &=\{v_q +w_i\}(-i,l,i,-l),\qquad\text{ (by equation~\ref{eq:v-id4})}\\ &=r^{-i,l}_{v_q +w_i}. \end{align*} The other two possibilities ($(i,-l)$ and $(-i,-l)$) also follow in this way. \end{proof} Another useful result follows from these identities and Corollary~\ref{lemma:0cycles}: \begin{lemma} Any $\partial$-closed flow can be written as the basic flow associated to a single cycle in ${\mathcal Q}$ (not necessarily contained in its support). \label{lemma:null_single_cycle} \end{lemma} \begin{proof} Note that in the proof of Lemma~\ref{lemma:0cycles}, $f\in\ker\partial$ was decomposed into a sum of basic flows for cycles $c$. The sum of two basic flows corresponding to cycles can be written, using identities \eqref{eq:v-id2}--\eqref{eq:v-id4} as \begin{multline} \{v\}(j_1,\dots,j_k)+\{v'\}(j'_1,\dots,j'_{k'}) = \\ \{v\}(j_1,\dots,j_k,a_1,\dots,a_l,j'_1,\dots,j'_{k'},-a_l,\dots,-a_1), \label{eq:v-sum} \end{multline} where $a_i\in\pm\{1,\dots,n\}$ are such that $v-\sum_i w_{a_i}=v'$, and one has $\sum_i w_{j_i}=\sum_i w_{j'_i}=0$ since the basic flows on the left-hand side correspond to cycles. This proves the lemma. \end{proof} \subsection{Exactness Results} \label{sec:2:exact:exact} \nopagebreak The promised proof of the exactness of~\eqref{eq:exact1} can now be given. Recall the morphism of lattices $\pi\times\partial\colon\Lambda^1\to \Z^n\times\Lambda^{0,0}$ defined in section~\ref{sec:2:abel:toric}. \begin{lemma} One has $\ker \pi\times\partial =\Lambda^2$, i.e.\xspace $$\Lambda^2\to\Lambda^1\stackrel{\pi\times\partial}\to\Z^n\times\Lambda^{0,0}$$ is exact. \label{lemma:kerc_lambda2} \end{lemma} \begin{proof} Since $\pi\times\partial(r^{ij}_v)=0$ one only has to prove that $\ker \pi\times\partial\subseteq\Lambda^2$. Let $f\in\ker \pi\times\partial$, and suppose that $c'$ is a cycle such that $f=\tilde\chi_{c'}$, as described in Lemma~\ref{lemma:null_single_cycle}. Then $f$ is of the form $$f=\{v\}(j_0,\dots,j_k),$$ for some $v\in{\mathcal Q}_0$ and $j_i\in\pm\{1,\dots,n\}$. Since $\pi f=0$, one has $\#\{i: j_i=j\}=\#\{i:j_i=-j\}$ for any $j\in\{1,\dots,n\}$. Thus, up to permutations, $(j_0,\dots,j_k)$ can be rewritten as $(j_1,-j_1,j_2,-j_2,\dots,j_l,-j_l)$ for some elements $j_k\in\{1,\dots,n\}$. Since $\{v\}(j_1,-j_1,j_2,-j_2,\dots,j_l,-j_l)=0$, the result follows by Lemma~\ref{lemma:permutation}. \end{proof} From this lemma it follows that~\eqref{eq:exact1} is exact, and induces an inclusion $\Lambda\hookrightarrow \Z^n\times\Lambda^{0,0}$. Recall that $\Pi=\ker\hat\rho$, where \map{\hat\rho}{\Z^n}{\widehat\Gamma}{x}{\sum_i x_iw_i.} \begin{lemma} \label{lemma:Lambda_n_pi} One has $\Pi=\pi(\ker\partial)$. \end{lemma} \begin{proof} Suppose $x\in\Pi $. One wants to find $f\in\ker\partial$ such that, for all $i\in\{1,\dots,n\}$, $f$ satisfies $$\sum_{a:\pi(a)=i} f(a) = x_i.$$ This is easy: just take the basic flow given by \begin{equation} f=\{v\}(\underbrace{ 1,\dots, 1}_{x_1}, \underbrace{ 2,\dots, 2}_{x_2},\dots, \underbrace{ n,\dots, n}_{x_n}), \label{eq:cycle_lambda_n} \end{equation} for any $v\in{\mathcal Q}_0$. (If $x_i$ is negative under any brace, the notation is taken to mean $-x_i$ copies of $-i$.) Now $\sum_{i=1}^n w_i{x_i} = 0\text{ in }\widehat\Gamma$ is equivalent to the fact that the basic flow corresponds to a cycle, and so $\partial f=0$. The converse follows from Lemma~\ref{lemma:null_single_cycle} and the preceding sentence. \end{proof} \begin{lemma} \label{lemma:imagec} There is an exact sequence of abelian groups $$0\to\Lambda \xrightarrow{\pi\times\partial} \Z^n\times\Lambda^{0,0} \xrightarrow{\hat\rho-\hat\nu} \Z^n/\Pi \cong\widehat\Gamma \to 0,$$ where \map{\hat\nu}{\Lambda^{0,0}}{\widehat\Gamma}{\zeta}{\sum_{v\in{\mathcal Q}_0}\zeta(v)v} is the morphism of lattices dual to the action of \/$\Gamma$ on $\End R$ by conjugation. \end{lemma} \begin{proof} One needs to show that if $(x,\zeta)\in\Z^n\times\Lambda^{0,0}$ satisfies \begin{equation} \sum_{i=1}^n x_iw_i - \sum_v \zeta(v)v=0, \label{eq:zerosum} \end{equation} then there exists a flow $f\in\Lambda^1$ such that $\pi(f)=x$ and $\partial f=\zeta$. One can construct this flow in two steps. For convenience, suppose that the weights of the action of $\Gamma$ on $Q$ have been normalised so that $w_1=1$. First construct a flow $g$ such that $\partial g=\zeta$ with only arrows of type $1$: start at vertex $0$ and let $g(1\to 0)=\zeta(0)$. Next, let $g(2\to 1)=\zeta(1)+\zeta(0)$, and continue in this way, setting $$g(k+1\to k) = \sum_{j=0}^k\zeta(j),$$ and $g=0$ on all the other arrows. Then $\partial g =\zeta$ by construction, and $\pi(g)=X_1e_1$, where $X_1=\sum_j\zeta(j)j$. Now add on any flow of the form $$g'=\{v\}(\underbrace{ 1,\dots, 1}_{x_1-X_1}, \underbrace{ 2,\dots, 2}_{x_2},\dots, \underbrace{ n,\dots, n}_{x_n}),$$ using the same conventions as in the previous lemma for the case when the integers under the braces are negative. Equation~\eqref{eq:zerosum} means that $g'$ is the basic flow associated to a cycle, and so $\partial g'=0$. The flow $f=g+g'$ has the required properties. \end{proof} \section{Singular Configurations} \label{sec:2:sing} In this section some comments are made regarding the singularities of $C_\zeta$. Recall that the tangent cone to $C_\zeta =\pi F_\zeta $ at a point $x$ which has a configuration $S\in{\mathcal C}^k$, is the cone $\pi F_0(\clos{S})$. This is singular if its dual $(\pi F_0(\clos{S}))^\vee$ is generated by a part of a basis of ${\Pi}^*$. In this case, $S$ is called a \emph{singular configuration}.\/ The set of singular configurations is denoted ${\mathcal S}$. \begin{prop} The following statements are true. \begin{enumerate} \item $C_\zeta$ is singular at the faces corresponding to configurations $S\in {\mathcal S}_\zeta$. \item $C_\zeta$ is non-singular in co-dimension $k$ if and only if ${\mathcal S}_\zeta^{n-k}=\emptyset$. \item $C_\zeta$ is generically non-singular in co-dimension $k$ if and only if ${\mathcal S}_{\spn}^{n-k}=\emptyset$. \item $C_\zeta$ is generically non-singular if and only if ${\mathcal S}_{\spn}=\emptyset$. \end{enumerate} \end{prop} \begin{proof} Part (1) of the following proposition follows from Theorem~\ref{thm:faces} and Lemma~\ref{lemma:generic_flow}. Part~(2) follows because $S\in{\mathcal S}^{n-k} \iff \dim \sigma_S = k$. The last two statements follow because ${\mathcal S}_{\spn}=\cup_{\zeta\text{ generic }}{\mathcal S}_\zeta$. \end{proof} \begin{question} What is the lowest $k$ for which ${\mathcal S}^k_\zeta$ is empty (i.e.\xspace in what co-dimension is $C_\zeta$ smooth)? The cone $C_0$ is smooth in co-dimension $1$ (it has an isolated singularity), so it seems likely that ${\mathcal S}^k=\emptyset$ for $k\geq1$. Of course, this is equivalent to the statement that ${\mathcal S}^1=\emptyset$. \end{question} The statements about singular trees are translated here for the record. \begin{conj} The polyhedra $C_\zeta$ are non-singular in co-dimension $n-1$, i.e.\xspace ${\mathcal S}^1=\emptyset$. If\/ $\Gamma\subset\SU(3)$, then $C_\zeta$ are non-singular for generic values of $\zeta$, i.e.\xspace ${\mathcal S}^0_{\text{span}}=\emptyset$. Furthermore, the Euler number of $C_\zeta$ for generic $\zeta$ is equal to the order of $\Gamma$, i.e\ $\card{{\mathcal T}^0_\zeta/\!\!\sim}=\card{\Gamma}$. \end{conj} Let us look at the case of singular \emph{points.\/} Suppose that $S\in{\mathcal C}^0$ is a singular configuration. This means that the primitive generators $\rho(S)=(\rho^1_S,\dots,\rho^k_S)$ of $\pi F_0(S)$ do not form a basis of $\Pi $. If $k=n$, $\pi F_0(\clos{S})$ corresponds to a finite abelian quotient singularity, whereas, if $k>n$, the singularity is determined by the linear relations holding between the $\rho_S^i$. To determine whether a given set $S\in{\mathcal C}^0$ is singular or not, one must find all the cycles $c$ such that $c^-\subseteq S$, calculate their type, and see what primitive vectors of $\Pi$ one obtains. It is sufficient to restrict one's attention to the cycles which are not decomposable into a union of cycles. Let us look at some examples. \section{Examples and Computations} \label{sec:2:examples} \subsection{Commutators} \label{sec:2:examples:comm} In this section, the commutator of a configuration $S$ is defined; this is specific to the McKay quiver and its purpose is purely computational: it gives a necessary criterion for determining when $S\in{\mathcal C}^0$ which is extremely useful in practical calculations. Let us begin by working out what the closure of a subset $S\subseteq{\mathcal Q}_1$ corresponds to in the case of the McKay quiver for the action $\frac{1}{ r}(w_1,\dots,w_n)$. Let $\pi$ be the map ${\mathcal Q}_1\to\{1,\dots,n\}$ which assigns to each arrow $a_v^i = v\to v-w_i$ its type $i$. This induces a map $\pi\colon{\mathbb R}^{{\mathcal Q}_1}\to{\mathbb R}^n$ which assigns to the basis element $\chi_v^i=\chi_{a_v^i}$ the basis element $e_i$ of ${\mathbb R}^n$. Recall that to calculate the closure of $S$, one must find the smallest over-set of $S$ which contains the positive part of cycles of type zero if and only if it contains the negative part. The simplest cycles which have type zero are cycles with only four arrows: starting from vertex $v$, go forward along arrow $a_v^i$ to $v-w_i$, forward again along $a_{v-w_i}^j$ to $v-w_i-w_j$, back along $a_{v-w_j}^i$ to $v-w_j$ and back along $a_v^j$ to $v$. This cycle is denoted by $c_v^{ij}$. The basic flow corresponding to this $c^{ij}_v$ is $\{v\}(i,j,-i,-j)$ in the notation of Section~\ref{sec:2:exact:sequ}. If $S$ contains $c^{ij+}_v=\{a_v^i,a_{v-w_i}^j\}$ for some $v,i,j$ the closure of $S$ must contain $c^{ij-}_v=\{a_v^j,a_{v-w_j}^i\}$. The two pairs of arrows will be called \emph{complementary pairs}.\/ \begin{dfn} Denote by $p^{ij}_v$ the pair of arrows $\{a_v^i,a_{v-w_i}^j\}$. If $S$ is a subset of ${\mathcal Q}_1$, the smallest over-set $S^{\bowtie}\supseteq S$ satisfying the ``commutation condition'' \begin{equation} p^{ij}_v \subseteq S^{\bowtie} \iff p^{ji}_v\subseteq S^{\bowtie} \label{eq:S_commute} \end{equation} is called the \emph{commutator} of $S$. \end{dfn} \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/commutator.eps} \end{center} \caption[The commutator of the set in Figure~\protect\ref{fig:uncycle7123}]{The commutator of the set in Figure~\protect\ref{fig:uncycle7123}. Notice that this is not an invariant set: there is no way to assign elements of $\Z^n$ to the vertices in such a way that equation~\protect\ref{eq:S_inv} is satisfied.} \label{fig:commutator} \end{figure} The fact that $\clos{S}\supseteq S^{\bowtie}$ is extremely useful in practical computations, because it gives the following easily checked necessary condition for $S$ to be an IC-set. \begin{lemma} If $S$ is a spanning IC-set then the morphism ${\mathcal W}_S\colon {\mathcal Q}_0 \to\Z^n$ defined by equations~\ref{eq:lambda_def} satisfies the following conditions for all $v\in{\mathcal Q}_0$, and $i,j\in\{1,\dots,n\} $: \begin{equation} p^{ij}_v \subseteq S \implies \begin{cases} {\mathcal W}_S({v-w_j})-{\mathcal W}_S({v})&=e_j \\ {\mathcal W}_S({v-w_j-w_i})-{\mathcal W}_S({v-w_j})&=e_i, \end{cases} \label{eq:lambda_commute} \end{equation} where $\{e_i\}$ denotes the standard basis of $\Z^n$. \end{lemma} It is rather surprising that this condition is in fact \emph{not\/} sufficient. Figure~\ref{fig:notcc} gives an example of a set $S$ for which $S=S^{\bowtie}\neq \clos{S}$. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/notcc.eps} \end{center} \caption{A configuration $S$ for which $S=S^{\bowtie}\neq \clos{S}$. The dotted arrows indicate the arrows in $\clos{S}\setminus S^{\bowtie}$} \label{fig:notcc} \end{figure} \subsection{Weightings and Stabilisers} \label{sec:2:examples:weight} There is another way of understanding the condition $S\in {\mathcal C}^k$ for a set $S\subseteq{\mathcal Q}_1$ which relates directly to the toric geometry of $X_\zeta =\protect\overline{T}{}^{\Pi,C_\zeta }$. The basic idea is that a $k$-dimensional face of $C_\zeta $ corresponds to elements of $X_\zeta $ which are fixed by a torus of codimension $k$. For instance, the extreme points correspond to fixed points of $T^\Pi$. Let us begin with this case for simplicity. \subsubsection{Fixed Points and $n$-weightings} \label{sec:2:examples:weight:fixed} Recall that the lattice $\Z^{{\mathcal Q}_0}$ is denoted by $\Lambda ^0$. The sub-lattice of co-rank $1$ defined by the equation $\sum_{v\in{\mathcal Q}_0} \zeta(v) = 0$ is denoted $\Lambda ^{0,0}$. For any subset of arrows $S\subseteq{\mathcal Q}_1$ one attempts to find a morphism ${\mathcal W}_S\colon \Lambda^{0,0}\to\Z^n$ satisfying \begin{equation} {\mathcal W}_S(\partial \chi_a)=\pi(\chi_a),\quad\text{ for }a\in S. \label{eq:S_inv} \end{equation} If ${\mathcal W}_S$ exists, it satisfies \begin{equation} {\mathcal W}_S(\partial\tilde{\chi }_p)=\pi\tilde{\chi }_p \label{eq:lambda_def} \end{equation} for any path $p=(p^1,\dots,p^k)$ in $S$. In particular, this equation implies that $\pi\tilde{\chi }_c=0$ must hold for all cycles $c\subseteq S$. In other words, if ${\mathcal W}_S$ exists, then $\rk \pi Z_0(S)=0$. Conversely, if all cycles in $S$ have zero type then one can find a morphism ${\mathcal W}_S$ satisfying~\eqref{eq:S_inv}. Note that this is possible if and only if there is morphism ${\mathcal W}'_S\colon \Lambda^0\to\Pi$ which extends it. The latter corresponds to a labeling $${\mathcal W}'_S\colon{\mathcal Q}_0\to\Z^n$$ of the vertices of the quiver by elements of $\Z^n$ in such a way that equation~\ref{eq:S_inv} holds. This is called an \emph{$n$-weighting} of $S$, and $S$ is called~\emph{invariant} if it admits such a weighting. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 3cm \epsfbox{fig/lambdas7123.eps} \end{center} \caption{Example of an $n$-weighting of a configuration of arrows.} \label{fig:lambdas7123} \end{figure} In summary, $S$ is an extreme configuration if and only if $\clos{S}$ is invariant, if and only if $\pi Z_0(S)=0$. Such $S$ are called \emph{ invariant closure configurations\/} or \emph{IC-configuration}\emph{s\/} for short. Particularly important is the set ${\mathcal T}^0$ of \emph{IC-trees}; it is easy to determine whether $T\in {\mathcal T}$ is an IC-tree: just check whether the arrows $a\in \clos{T}$ all satisfy ${\mathcal W}_T\partial\chi _a= \pi \chi _a$. To see the relationship to the toric geometry of $X_\zeta $, recall that $X_\zeta$ is the GIT quotient of a certain affine variety in ${\mathbb C}^{{\mathcal Q}_1}$ by $T^{0,0}$. The extreme points of $C_\zeta $ therefore correspond to elements $\tilde \alpha \in{\mathbb C}^{{\mathcal Q}_1}$ which are mapped to the same $T^{0,0}$-orbit under the morphism of algebraic tori $$\widehat{{\mathcal W}_S}\colon {\C^*}^n\to T^1.$$ This is the case for any $\tilde \alpha $ such that $\supp\tilde \alpha \subseteq S^{\Box}$, where $S^{\Box}$ is the set of all arrows in ${\mathcal Q}_1$ which satisfy~\eqref{eq:S_inv} (this includes of course $\overline S$). This discussion can be extended to higher dimensional faces, although it is not as practical for concrete calculations. \subsubsection{Higher Dimensional Faces} \label{sec:2:examples:weight:higher} The sub-lattice of integral flows on ${\mathcal Q}$ which are zero outside $S$ will be denoted $\Z^S\subseteq\Z^{{\mathcal Q}_1}$. Recall that $Z_0(S)$ is the set of $0$-flows which are zero outside $S$: $Z_0(S)= \Z^S\cap \ker\partial$. Define, for each $S\subseteq{\mathcal Q}_1$, a morphism of lattices $$ {\mathcal W}_S\colon \partial \Z^S\to{\pi \Z^S/\pi Z_0(S)}$$ defined by \begin{equation} \label{eq:ws} {\partial\chi_a}:= \pi Z_0(S) + \pi\chi_a,\quad \rlap{$a\in S$.} \end{equation} This fits into a commutative diagram $$\begin{CD} \label{CD:ws} \partial\Z^S & @>{{\mathcal W}_S}>> & {\pi\Z^S/\pi Z_0(S)}\\ @A{\partial}AA & & @AA{\pr}A\\ \Z^S & @>>{\pi_{|\Z^S}}> & \pi\Z^S. \end{CD} $$ Taking the image of this under the functor $\ \widehat{\ } = \Hom(\ \cdot\ ,{\mathbb C}^*)$, one obtains a corresponding diagram of algebraic tori: $$ \begin{CD} \label{CD:tori} T^{0,0}&\supseteq T^{\partial\Z^S} & @<{\widehat {{\mathcal W}_S}}<< & T^{\pi\Z^S/\pi Z_0(S)}&\\ & @V{\widehat\partial}VV & & @VV{\widehat\pr}V &\\ T^1 & \supseteq T^{\Z^S} & @<<{\widehat {\pi_{|\Z^S}}}< & T^{\pi\Z^S}& \subseteq {\C^*}^n . \end{CD} $$ This diagram shows that the action of the sub-torus $ T^{\pi \Z^S/\pi Z_0(S)}$ of $T^{\pi \Z^S}$ on $\tilde \alpha \in {\mathbb C}^S$ via $\widehat{\pi_{|\Z^S}}\widehat{\pr}$ leaves $\tilde \alpha $ in the same orbit of $T^{0,0}$. In fact, since the torus corresponding to $\Z^{S^c}$ (where $S^c$ denotes the complement of $S$ in ${\mathcal Q}_1$) acts trivially on ${\mathbb C}^S$, one sees that $$ T^{\pi\Z^{{\mathcal Q}_1}/\pi Z_0(S)}$$ acts on ${\mathbb C}^S$ fixing the $T^{0,0}$-orbits and is a sub-torus of $T^{\pi \Z^{{\mathcal Q}_1}} = {\C^*}^n$ of codimension $\rk\pi Z_0(S)$. \subsubsection{Existence and Uniqueness of $n$-weightings} \label{sec:2:examples:weight:exist} \begin{dfn} Two vertices in $v,v'\in {\mathcal Q}$ are said to be \emph{connected by a subset\/} $S\subseteq{\mathcal Q}_1$ if there is a path $p$ in $S$ whose endpoints are $\{v,v'\}$. A subset $S\subseteq {\mathcal Q}_1$ is called a \emph{spanning set} if it connects any two vertices in ${\mathcal Q}_0$. \end{dfn} One sees easily that the condition that $S$ be a spanning subset is equivalent to the statement $\partial\Z^S = \Lambda ^{0,0}$. Thus if $S$ is a spanning subset, any $n$-weighting ${\mathcal W}_S$ of $S$ is completely determined by~\eqref{eq:lambda_def}, and so is unique. In fact, if $T\subseteq S$ is any spanning tree in $S$, then $T$ determines a unique morphism ${\mathcal W}_T\colon \Lambda ^{0,0}\to\Z^n$, and one sees that $S$ is invariant if and only \begin{equation} \label{eq:S_inv_wt} {\mathcal W}_T\partial\chi _a = \pi \chi _a, \quad\forall a\in S\setminus T. \end{equation} The following lemma shows that, for generic values of $\zeta $, one can effectively use the condition above to check the invariance of configurations. \begin{lemma} \label{lemma:generic_flow} If $\zeta$ is generic in $\Lambda ^{0,0}_{\mathbb R} = \Lambda ^{0,0}\otimes_\Z {\mathbb R}$ and $f$ is a $\zeta$-admissible flow then $\supp(f)$ is a spanning subset. \end{lemma} \begin{proof} If $\supp(f)$ is not a spanning subset then there exists a partition of ${\mathcal Q}_0$ into disjoint non-empty subsets $S_0$ and $S_1$ such that no element of $S_0$ is connected to any element of $S_1$. Consider the restriction of $f$ to the $S_i$; the previous statement implies that $\partial (f_{|S_i})=(\partial f)_{|S_i}$. Since, for any $\zeta$-flow $g$, ``the flow is conserved," i.e.\xspace $\sum_{v\in{\mathcal Q}_0}(\partial g)_v=0$, one has \begin{align*} \sum_{v\in S_i} \zeta_v &= \sum_{v\in S_i} (\partial f)_v \\ &= \sum_{v\in{\mathcal Q}_0} \left( \partial (f_{|S_i}) \right)_v \\ &=0, \end{align*} but this does not happen for a generic $\zeta$ in ${\mathbb R}^{{\mathcal Q}_0}_0$. \end{proof} First, an example which has some singular cones: the action of the group of fifth roots of unity on ${\mathbb C}^3$ with weights $1$, $2$ and $3$. In this case, there are whole cones of values of $\zeta$ for which $X_\zeta$ is smooth, and others where $X_\zeta $ is singular. \begin{rmk}[About the computations] The computations were done using several computer programs. A Pascal program was used to produce a list of all the IC-trees for any action of a cyclic group. Then, for each value of $\zeta$, another Pascal program was used to determine which trees were admissible and to work out the corresponding flows and extreme points. Then a Mathematica program was run to draw the pictures of the polyhedra and of the extreme flows. \end{rmk} \subsection{Example: the action $\qsing 1/5(1,2,3)$.} \label{sec:2:ex:5123} Recall the action $\frac{1}{ 5}(1,2,3)$ of the group $\mu_5$ of fifth roots of unity on ${\mathbb C}^3$ with weights $1,2$ and $3$ considered in example~\ref{ex:5123}. The McKay quiver in this case is the regular oriented graph with $5$ vertices and an arrow between each vertex $v$ and $v-1$, $v-2$ and $v-3 \pmod 5$. The set ${\mathcal T}^0$ of IC-trees contains a total of 55 trees, once one has factored out by the symmetry which consists in permuting the vertices cyclicly (the action of $\widehat\Gamma$ on itself). To draw $C_\zeta$, choose values of $\zeta$, calculate the $\zeta$-flows on all the IC-trees, discard those which are negative, and project the resulting points to ${\mathbb R}^3$ via the map $\pi$. The resulting convex polyhedra $C_\zeta$ are the intersection of the positive orthant with a finite number of half-spaces. The polyhedron $C_\zeta$ for the value $\zeta=(-1,-1,-1,-1,4)$ is shown in Figure~\ref{fig:poly5sing}. The trees and flows corresponding to the extreme points appear in Figure~\ref{fig:tree5sing}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 9cm \epsfbox{fig/poly5123sing.eps} \end{center} \caption[$C_\zeta$ for $\qsing 1/5(1,2,3)$, $\zeta=(-1,-1,-1,-1,4)$.]{$C_\zeta$ for $\qsing 1/5(1,2,3)$, $\zeta=(-1,-1,-1,-1,4)$. (The view is from ``behind'', from the point with coordinates $(-1.3,-1,-1)$. The vertices have been numbered (where possible) in order of increasing $z$ coordinate.)} \label{fig:poly5sing} \end{figure} \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123sing1.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing2.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing3.eps} \vspace*{-1.5cm} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123sing4.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing5.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing6.eps} \vspace*{-1.5cm} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123sing7.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing8.eps} \epsfysize= 4cm \epsfbox{fig/tree5123sing9.eps} \vspace*{-1.5cm} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123sing10.eps} \end{center} \vspace*{-1cm} \caption[Extreme flows for $\qsing 1/5(1,2,3)$, $\zeta=(-1,-1,-1,-1,4)$.]{Extreme flows for $\qsing 1/5(1,2,3)$, $\zeta=(-1,-1,-1,-1,4)$. (The values of the flows are indicated to the right of the arrows. The numbers in the top left-hand corners correspond to the vertex numbers in Figure~\ref{fig:poly5sing}.)} \label{fig:tree5sing} \end{figure} One sees immediately from the figure that $X_\zeta$ has a singularity at the point corresponding to the extreme point $(1,3,1)$: the tangent cone there has four generators. The other extreme points are the intersection of three faces: in order to determine whether they are in fact singular or not one must check whether the primitive generators in $\Pi\subset \Z^3$ of the tangent cone actually generate $\Pi$. In this case it turns out that they do, so they are smooth points. In fact, there are a total of 7 singular non-isomorphic IC-trees. These have been listed in Figure~\ref{fig:singtree}. The first tree (rotated by $\frac{2\pi}{ 5}$) corresponds to the singular point $(1,3,1)$ above. The corresponding tangent cone to $C_\zeta$ was described in Section~\ref{sec:2:flow:state} --- it has four generators $v_1,\dots,v_4$, any three of which generate $\Pi$, and satisfying a single relation of the form $v_1+v_3=v_2+v_4$. The corresponding singularity is therefore of the type $xw=yz \subset{\mathbb C}^4$: a cone over a quadric surface. \begin{figure}[htbp] \begin{center} \vspace*{-3cm} \leavevmode \epsfbox{fig/pstree5_123.eps} \end{center} \vspace*{-3cm} \caption{Non-isomorphic singular IC-trees in ${\mathcal T}^0$ for the action $\frac{1}{ 5}(1,2,3)$.} \label{fig:singtree} \end{figure} Checking the cases listed in Figure~\ref{fig:singtree}, one sees that it is possible to find generic values of $\zeta$ for which none of these trees (nor their rotations by elements of $\hat\Gamma$) are admissible: for instance, $\zeta=(9,8,-3,-2,-12)$ is such a value; this gives a smooth resolution of ${\mathbb C}^3/\Z_5$ which has Euler number $9$. The corresponding polyhedron and flows are shown in Figures~\ref{fig:poly5non} and~\ref{fig:tree5non}. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 9cm \epsfbox{fig/poly5123non.eps} \end{center} \caption[$C_\zeta$ for $\qsing 1/5(1,2,3)$, $\zeta=(9,8,-3,-2,-12)$.]{$C_\zeta$ for $\qsing 1/5(1,2,3)$, $\zeta=(9,8,-3,-2,-12)$. (The view is from ``behind'', from the point with coordinates $(-1.3,-1,-1)$. The vertices have been numbered (where possible) in order of increasing $z$ coordinate.)} \label{fig:poly5non} \end{figure} \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123non1.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non2.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non3.eps} \vspace*{-1.5cm} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123non4.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non5.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non6.eps} \vspace*{-1.5cm} \leavevmode \epsfysize= 4cm \epsfbox{fig/tree5123non7.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non8.eps} \epsfysize= 4cm \epsfbox{fig/tree5123non9.eps} \end{center} \vspace*{-1cm} \caption[Extreme flows for $\qsing 1/5(1,2,3)$, $\zeta=(9,8,-3,-2,-12)$.]{Extreme flows for $\qsing 1/5(1,2,3)$, $\zeta=(9,8,-3,-2,-12)$. (The values of the flows are indicated to the right of the arrows. The numbers in the top left-hand corners correspond to the vertex numbers in Figure~\ref{fig:poly5non}.)} \label{fig:tree5non} \end{figure} \begin{rmk} In fact, all the singular trees in Figure~\ref{fig:singtree} correspond to the same type of singularity, namely a cone over a quadric surface.\footnote{Note that this is consistent with the conjecture in~\cite{sacha:ale} regarding the quadratic nature of the singularities of $X_\zeta$.} All the IC-trees whose cones are simplicial correspond to non-singular points. Thus one can tell whether $X_\zeta$ is non-singular simply by checking whether all extreme points have three edges emanating from them. This is also the case for other singularities: for instance, $\qsing 1/6(1,2,4)$, $\qsing 1/7(1,2,5)$, $\qsing 1/8(1,2,6)$, $\qsing 1/9(1,2,7)$; it is not always the case however: for instance, for $\qsing 1/7(1,2,3)$ and $\qsing 1/10(1,2,8)$ where there exists elements of ${\mathcal T}^0$ which correspond to $\Z_2$-quotient singularities. \end{rmk} \subsection{Crepant Resolutions} \label{sec:2:crep} Let $\Sigma_\zeta $ denote the fan determined by the polyhedron $C_\zeta$. By Corollary~\ref{cor:fan}, the one-skeleton of $\Sigma_\zeta$ is given by $$\Sigma_\zeta^{(1)} = \{ \sigma(S)^\vee: S\in{\mathcal C}^{n-1}_\zeta\}, $$ so $X_\zeta$ has trivial canonical bundle if the primitive generators $v_S$ of the cones $\sigma(S)^\vee$ for $S\in{\mathcal C}^{n-1}$ all belong to the hyper-plane defined by the equation $\sum n_i = 1$ in $\Pi^* = \Z^{3}+\frac{\Z}{ r}(w_1,w_2,w_3)$. \begin{example} \label{ex:3111} Consider the action $\qsing 1/3(1,1,1)$. This has only three IC-trees up to isomorphism. They consist of trees with two arrows of the same type. For generic values of $\zeta$, a little thought shows that the polyhedron $C_\zeta$ is the positive quadrant with a small equilateral triangle chopped off. The dual fan is the barycentric subdivision of the positive quadrant by the ray passing through the point $v=\frac{1}{3}(1,1,1)\in \Pi^*$. This is a non-singular fan, and since the point $v$ belongs to the hyper-plane $\sum n_i = 1$, this has trivial canonical bundle. The variety $X_\zeta$ is the total space of the bundle ${\mathcal O}(-3)$ over ${\mathbb P}^2$. \end{example} To conclude this section, a picture of a more complicated example is draw. Note that $\Gamma\subset\SU(3)$ and that the variety $X_{\zeta}$ is a smooth crepant resolution with Euler number 11. \begin{figure}[htbp] \begin{center} \leavevmode \epsfysize= 9cm \epsfbox{fig/poly11_146.eps} \end{center} \caption{An example of $C_\zeta$ for the action $\qsing 1/11(1,4,6)$.} \label{fig:11_146} \end{figure} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1996-10-01T04:57:38
9610
alg-geom/9610001
en
https://arxiv.org/abs/alg-geom/9610001
[ "alg-geom", "math.AG" ]
alg-geom/9610001
Sacha Sardo Infirri
Alexander V. Sardo Infirri
Crepant Terminalisations and Orbifold Euler Numbers for SL(4) Singularities
LaTex2e, 22 pages with 1 table
null
null
null
null
Let $X$ and $Y$ be two analytic canonical Gorenstein orbifolds. A resolution of singularities $Y\to X$ is called an Euler resolution if $Y$ and $X$ have the same orbifold Euler number. If $Y$ is only terminal rather than smooth, it is called an Euler terminalisation. It is proved that Euler terminalisations exist for toric varieties in any dimension, for 4-dimensional toroidal varieties, and for singularities $\C^4/G$ where $G$ belongs to certain classes of $\SL(4)$ subgroups. The method of proof is expected to be applicable to a sizeable number of finite $\SL(4)$ subgroups and to lead to a generalisation of the Dixon-Harvey-Vafa-Witten orbifold Euler number conjecture to dimension~4.
[ { "version": "v1", "created": "Tue, 1 Oct 1996 02:48:18 GMT" } ]
2008-02-03T00:00:00
[ [ "Infirri", "Alexander V. Sardo", "" ] ]
alg-geom
\part{Maps} \newcommand{\map}[5]{$$\begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array}$$} \newcommand{\map}{\map} \newcommand{\mapeq}[6]{\begin{equation} \begin{array}{rccc} #1\colon & #2 & \longrightarrow & #3 \\ & #4 & \longmapsto & #5 \end{array} \protect\label{#6} \end{equation}} \newcommand{\corresp}[4]{$$\begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array}$$} \newcommand{\correspeq}[5]{\begin{equation} \begin{array}{ccc} #1 & \to & #2 \\ #3 & \mapsto & #4 \end{array} \protect\label{#5} \end{equation}} \newcommand{\operatorname{\rm Cl}}{\operatorname{\rm Cl}} \newcommand{\operatorname{\rm N}}{\operatorname{\rm N}} \newcommand{\operatorname{\rm Stab}}{\operatorname{\rm Stab}} \newcommand{\operatorname{\rm Fix}}{\operatorname{\rm Fix}} \newcommand{\operatorname{\rm Mat}}{\operatorname{\rm Mat}} \newcommand{\operatorname{\rm Hom}}{\operatorname{\rm Hom}} \newcommand{\operatorname{\rm GL}}{\operatorname{\rm GL}} \newcommand{\operatorname{\rm PGL}}{\operatorname{\rm PGL}} \newcommand{\operatorname{\rm PSL}}{\operatorname{\rm PSL}} \newcommand{\operatorname{\rm SL}}{\operatorname{\rm SL}} \newcommand{\operatorname{\rm PL}}{\operatorname{\rm PL}} \def\<<#1>>{\langle#1\rangle} \def\qsing#1/#2(#3){\frac{#1}{#2}(#3)} \newcommand{\PP}{{\mathbb P}} \newcommand{\Q}{{\mathbb Q}} \newcommand{\R}{{\mathbb R}} \newcommand{\F}{{\mathbb F}} \newcommand{\Z}{{\mathbb Z}} \newcommand{{\mathbb C}}{{\mathbb C}} \newcommand{{\C^*}}{{{\mathbb C}^*}} \newcommand{{\mathcal P}}{{\mathcal P}} \newcommand{{\mathcal O}}{{\mathcal O}} \newcommand{{\mathcal T}}{{\mathcal T}} \newcommand{\bar E}{\bar E} \newcommand{\bar F}{\bar F} \newcommand{\bar G}{\bar G} \newcommand{\bar H}{\bar H} \newcommand{\bar V}{\bar V} \newcommand{\bar p}{\bar p} \newcommand{\bar g}{\bar g} \newcommand{\bar f}{\bar f} \newcommand{\bar\Gamma}{\bar\Gamma} \newcommand{G_\eta}{G_\eta} \newcommand{H_\eta}{H_\eta} \newcommand{\bar\Ge}{\barG_\eta} \newcommand{\bar G_{\eta\xi}}{\bar G_{\eta\xi}} \newcommand{\bar G_{\xi}}{\bar G_{\xi}} \newcommand{\bar G_{\eta\eta'}}{\bar G_{\eta\eta'}} \newcommand{\bar h}{\bar h} \newcommand{\subpart}{\prec} \newcommand{canonical Gorenstein orbifold\xspace}{canonical Gorenstein orbifold\xspace} \newcommand{{\cGo}s\xspace}{{canonical Gorenstein orbifold\xspace}s\xspace} \newcommand{{\text{an.}}}{{\text{an.}}} \newcommand{{\text{alg.}}}{{\text{alg.}}} \newcommand{{\text{proj.}}}{{\text{proj.}}} \newcommand{{\widetilde Y}}{{\widetilde Y}} \newcommand{{\text{diag}}}{{\text{diag}}} \newcommand{H${}^*$}{H${}^*$} \newcommand{I${}^*$}{I${}^*$} \newcommand{\text{Sing}}{\text{Sing}} \newcommand{\text{Term}}{\text{Term}} \newcommand{\TermC}{\text{Term}\text{C}} \newcommand{\TermN}{\text{Term}\text{N}} \newcommand{\text{Res}}{\text{Res}} \newcommand{\text{Abel}}{\text{Abel}} \newcommand{\Abel\text{N}}{\text{Abel}\text{N}} \newcommand{\Abel\text{C}}{\text{Abel}\text{C}} \newcommand{\text{wt}}{\text{wt}} \newcommand{\text{Glob}}{\text{Glob}} \newcommand{\Glob\Term}{\text{Glob}\text{Term}} \newcommand{\Glob\Res}{\text{Glob}\text{Res}} \newcommand{\Glob\Abel}{\text{Glob}\text{Abel}} \newcommand{\text{--}}{\text{--}} \newcommand{\dashrightarrow}{\dashrightarrow} \newcommand{\chi_{\text{orb}}}{\chi_{\text{orb}}} \newtheorem{thm}{Theorem}[section] \newtheorem{nonumberthm}{Theorem} \renewcommand{\thenonumberthm}{} \newtheorem{nonumbercor}{Corollary} \renewcommand{\thenonumbercor}{} \newtheorem{prop}[thm]{Proposition} \newtheorem{cor}[thm]{Corollary} \newtheorem{lemma}[thm]{Lemma} \newtheorem{crit}[thm]{Criterion} \theoremstyle{definition} \newtheorem{dfn}[thm]{Definition} \newtheorem{problem}[thm]{Problem} \newtheorem{conj}{Conjecture} \newtheorem{nonumberconj}{Conjecture} \renewcommand{\thenonumberconj}{} \newtheorem{example}[thm]{Example} \newtheorem{assumption}[thm]{Assumption} \newtheorem{convention}[thm]{Convention} \newtheorem{notation}[thm]{Notation} \theoremstyle{remark} \newtheorem{rmk}[thm]{Remark} \newtheorem{question}{Question} \newtheorem{claim}[thm]{Claim} \newtheorem{fact}[thm]{Fact} \newtheorem{summ}[thm]{Summary} \newtheorem{case}{Case} \renewcommand{\thecase}{} \newtheorem{note}{Note} \renewcommand{\thenote}{} \newtheorem{ack}{Acknowledgment} \renewcommand{\theack}{} \begin{document} \title[\runningheadstring]{\titlestring\footnote{{Maths Subject Classification (1991): 32S45 (Primary) 14L30 14E30 (Secondary)}}} \author{Alexander V. Sardo Infirri} \email{[email protected]} \address{Research Institute for Mathematical Sciences\\ Ky\=oto University\\ Oiwake-ch\protect\=o\\ Kitashirakawa\\ Saky\protect\=o-ku\\ Ky\=oto 606-01\\ Japan} \date{1 October 1996} \begin{abstract} Let $X$ and $Y$ be two analytic canonical Gorenstein orbifolds. A resolution of singularities $Y\to X$ is called an \emph{Euler resolution} if $Y$ and $X$ have the same orbifold Euler number. If $Y$ is only terminal rather than smooth, it is called an \emph{Euler terminalisation}. It is proved that Euler terminalisations exist for toric varieties in any dimension, for 4-dimensional toroidal varieties, and for singularities ${\mathbb C}^4/G$ where $G$ belongs to certain classes of $\operatorname{\rm SL}(4)$ subgroups. The method of proof is expected to be applicable to a sizeable number of finite $\operatorname{\rm SL}(4)$ subgroups and to lead to a generalisation of the Dixon-Harvey-Vafa-Witten orbifold Euler number conjecture to dimension~4. \end{abstract} \maketitle \tableofcontents \setcounter{section}{-1} \section{Introduction} \label{sec:intro} An analytic $n$-fold $X$ will be called a {\em canonical Gorenstein orbifold\/} if it has at most canonical Gorenstein singularities and is such that, for each $x\in X$, there exists a finite group $\pi_x<\operatorname{\rm SL}(n)$ such that $$(X,x)\cong ({\mathbb C}^n/\pi_{x}, 0)$$ as germs of analytic spaces. \subsection{The Orbifold Euler Number} \label{sec:intro:orbi} The {\em orbifold Euler number\/} of $X$ is defined as the (finite) sum \begin{equation} \label{eq:orb-eul-def} \chi_{\text{orb}}(X) := \sum_{k\geq 1} k\chi(m^{-1}(k)), \end{equation} where $\chi$ is the ordinary Euler number and \map{m}{X}{\Z}{x}{|\operatorname{\rm Cl}(\pi_x)|} is the upper semi-continuous map assigning to each point $x$ the number of conjugacy classes of $\pi_x$. It is easy to show~\cite{roan:calabi-yau} that if $M$ is an $n$-fold admitting a $G$-action whose non-trivial elements' fixed-point loci have codimension at least two, and such that $M/G$ has only Gorenstein singularities, then \begin{align} \chi_{\text{orb}}(M/G) &\phantom{:}= \chi_{DHVW}(M;G),\\ &:= \sum_{[g]\in\operatorname{\rm Cl}(G)}\chi(M^g/\operatorname{\rm N}^G_g) \end{align} where the right-hand side denotes the Dixon-Harvey-Vafa-Witten Euler number proposed in~\cite{dhvw:i,dhvw:ii}. If $X$ and $Y$ are {\cGo}s\xspace and $Y\to X$ is a bi-meromorphic map such that $\chi_{\text{orb}}(Y)=\chi_{\text{orb}}(X)$, then $Y$ will be called an {\em Euler blow-up\/} of $X$. If $Y$ is in addition smooth, then $Y$ will be called an {\em Euler resolution}. Restated in the above terminology, the Dixon-Harvey-Vafa-Witten Euler conjecture~\cite{dhvw:i} is \begin{quotation} Every $3$-dimensional canonical Gorenstein orbifold\xspace has an Euler resolution. \end{quotation} It took ten years to give a positive answer to the conjecture \cite{mar_ols_per,roan:mirror_cy,mark:res_168,roan:res_a5,ito:trihedral,roan:crepant}. \subsection{Euler Terminalisations} \label{sec:intro:euler-term} Right from the start it was recognized that the analogous conjecture in dimension 4 (Do all 4-dimensional {\cGo}s\xspace have Euler resolutions?) is trivially false: the simplest non-smooth example ${\mathbb C}^4/\<<-1>>$ is already terminal. However, rephrased slightly, the conjecture can be made to look much more promising. For this, note that the existence of Euler resolutions in dimension 3 is equivalent to saying that the minimal models for these singularities are smooth. In other words ``smooth'' is equivalent to ``terminal'' for 3-dimensional Gorenstein finite quotient singularities. \begin{dfn} A Euler blow-up $Y\to X$ such that $Y$ has only terminal singularities is called an {\em Euler terminalisation\/} of $X$. The property $\text{Term}(X)$ is defined to be true if and only if such a $Y$ exists. \end{dfn} \begin{dfn} The property $\text{Term}(n)$ is defined to be true if and only if $\text{Term}(X)$ is true for all $n$-dimensional {\cGo}s\xspace $X$. \end{dfn} Thus, $\text{Term}(2)$ is true in virtue of classical work and $\text{Term}(3)$ is true by the recent work mentioned above. \begin{question} \label{conj} Is $\text{Term}(n)$ true? \end{question} The next open case is of course $\text{Term}(4)$. As in the case of dimension~3, the question reduces to the problem of constructing Euler terminalisations for the local singularities ${\mathbb C}^{4}/G$ for all the small subgroups $G$ of $\operatorname{\rm SL}(4)$. We shall use the following terminology. A subgroup $G < \operatorname{\rm SL}(V)$ will be called \emph{reducible} if $V$ is reducible as a $G$-module. The {\em type\/} of a group $G<\operatorname{\rm SL}(n)$ denotes the dimensions of the irreducible representations of $G$ appearing in the chosen special linear representation ${\mathbb C}^n$. For instance, irreducible groups have type $(n)$ and abelian groups have type $(1,1,\dots,1)$. For any $n$, denote by $Z_n$ the cyclic central subgroup of $\operatorname{\rm SL}(n)$. For any element $g\in G$, $\operatorname{\rm N}^G_g$ denotes the centralizer of $g$ in $G$, namely $\{h\in G | h^{-1}gh=g\}$ \subsection{Main Results} \label{sec:intro:results} \subsubsection{Toric and Toroidal cases} The first result is that $\text{Term}(X)$ is true in all dimensions for toric varieties. Further more, it holds also for toroidal varieties (analytic varieties which are locally isomorphic to toric varieties) in dimension~4 and in dimension~$n$ if termination of flips can be proved. \begin{thm} \label{thm:toric-toroidal} All simplicial toric {\cGo}s\xspace have Euler terminalisations. Furthermore, if flips terminate in dimension $n$, then all $n$-dimensional \emph{toroidal } {\cGo}s\xspace have Euler terminalisations. In particular, this is the case in dimension 4. \end{thm} The proof of the toric case is straightforward; crepant blow-ups of $X$ correspond to subdividing the first quadrant in $\R^n$ by rays whose generators all lie in the same hyper-plane; the orbifold Euler number of any cone is equal to its volume (meaning the volume of the simplex spanned by its generators), and since the sum of the volumes of the cones in the subdivision is equal to the total volume of the original quadrant, the orbifold Euler number is seen to remain invariant under crepant blow-ups. The fact that among all the crepant blow-ups there exists a terminal one is a consequence of the Toric Minimal Model Program~\cite{reid:toricmmp}. There is no minimal model program as yet for non-toric varieties in dimensions 4 and over. However, in the toroidal case, a well-known technique makes it is possible to construct flips by patching together local toric flips and using uniqueness. Thus if termination of flips is also known, (as it is in dimension 4~\cite{kmm:intro_mmp}), the existence of the terminal model follows and this again must have the same orbifold Euler number as the original variety. \subsubsection{$\operatorname{\rm SL}(4)$ subgroups of type $(3,1)$} The second set of results concerns 4-dimensional non-abelian singularities created by finite $\operatorname{\rm SL}(4)$ subgroups of type (3,1). Let $G<\operatorname{\rm SL}(4)$ be a finite subgroup which stabilises a line $V^2\subset V={\mathbb C}^{4}$ and let $V^1$ be a $G$-submodule such that $V=V^1\oplus V^2$. Denote by $\eta$ the generic point of the line $\{0\}\times V^2$, and by $G_\eta$ the stabiliser of $\eta$. Note that $G_\eta$ is a subgroup of $\operatorname{\rm SL}(V^1)\times\{1\}\cong \operatorname{\rm SL}(3)$. \begin{thm} If $G<\operatorname{\rm SL}(4)$ fixes a line and is such that the group $G_\eta$ does not contain $Z_3$ as a subgroup, then ${\mathbb C}^4/G$ has an Euler terminalisation with only toric singularities. \end{thm} The method of proof essentially consists in using the results of Roan \cite{roan:calabi-yau} to construct a 3-dimensional Euler resolution of ${\mathbb C}^3/G_\eta$ which is equivariant under the larger group $G$. As mentioned in the next section, the assumption that $G_\eta$ does not contain the group $Z_3$ is not essential to the method. However, so far the author has been unable to construct the equivariant resolutions without it. An attempt is made in Section~\ref{sec:3-1:centre:2}. \subsubsection{$\operatorname{\rm SL}(4)$ subgroups containing $Z_4$} The third set of results reduces the question $\text{Term}({\mathbb C}^4/G)$ for the irreducible $G$ which contain $Z_4$ to a conjecture regarding the existence of the equivariant resolutions of $\operatorname{\rm SL}(3)$ mentioned above. The conjecture is proved for irreducible subgroups of $\operatorname{\rm SL}(3)$ which do not contain $Z_3$, but remains open in the general case. \begin{rmk} The material presented here can no doubt be pushed further, but the author has so far been unable to do so. Nevertheless, it is hoped that the attentive reader will be able to perceive a direction in which to proceed. Is is also conceivable that the general strategy that emerges from this approach may be applicable to an understanding of quotient singularities in higher dimensions. \end{rmk} The idea here is as follows. Suppose that $G<\operatorname{\rm SL}(n)$ acts on ${\mathbb C}^{n}$ and contains the centre $Z_{n}$, and write $\bar G:=G/Z_n$ and $\bar V := \text{Bl}_0V/Z_n$. Then Lemma~\ref{lemma:bV} implies that $\bar V/\bar G$ is another canonical Gorenstein orbifold\xspace and $\bar V/\bar G \to V/G$ is an Euler blow-up. With the aid of another Lemma (The Patching Lemma~\ref{lemma:patching}), the problem is thus reduced to the construction and patching of Euler blow-ups of local neighbourhoods of $\bar V/\bar G$. The advantage of these is that the stabilisers of $\bar G$ in the tangent space to the blowup $\bar V$ are simpler than those of $G$ (because the stabiliser of $\bar G$ must fix a line in the tangent space to $\bar V$, so its type must be $(t,1)$, where $t$ is a partition of $n-1$). This approach is spelt out in Section~\ref{sec:4-centre} for irreducible subgroups of $\operatorname{\rm SL}(4)$ which contain $Z_{4}$; it reduces the problem to constructing and patching together equivariant $\operatorname{\rm SL}(3)$ resolutions. \subsubsection{Discussion} What if, on the other hand, the group $G$ is irreducible, but doesn't contain the centre of $\operatorname{\rm SL}(V)$? A complete answer to Question~\ref{conj} for all finite $\operatorname{\rm SL}(4)$ singularities seems out of reach of the methods suggested here, if only because the cases when $G$ does not contain $Z_4$ include simple groups, such as the alternating group $A_5$; if dimension~3 is any indication~\cite{mark:res_168,roan:res_a5}, it seems that ad-hoc methods will be necessary to construct a terminalisation. However, the possibility is open that, as a general rule, the non-simple finite subgroups $\operatorname{\rm SL}(n)$ not containing $Z_n$ are few in number and relatively amenable in form. For instance, if $\dim V=2$ no cases occur. For $\dim V=3$, one has --- apart from the simple groups (H) and (I) which require ad-hoc methods --- ``half'' the groups of type (C) and ``half'' of those of type (D). It turns out that under the assumption $Z_3\not < G$ these are semi-direct products of abelian groups with the alternating group $A_3$ and the symmetric group $S_3$ respectively. This allows one to construct their Euler resolutions from toric resolutions --- see~\cite{roan:calabi-yau}. Extension of this method to the $\operatorname{\rm SL}(4)$ case would seem to be feasible. Table~\ref{tab:sl4} outlines the state of the $\text{Term}({\mathbb C}^{4}/G)$ problem. \begin{table}[htbp] \begin{center} \leavevmode \setlength{\fboxrule}{0pt} \begin{tabular}{|l|l|l|l|l|} \hline \multicolumn{3}{|c|}{Type of Group} & Method & Status\\ \hline \multirow{7}{12mm}{Irred.\ \textbf{(4)}} & \multirow{2}{16mm}{$G\not > Z_{4}$} & \multirow{1}{22mm}{Simple} & \fbox{Ad hoc method?} &\\ \cline{3-5} & & \multirow{1}{22mm}{Non-Simple} & \fbox{\parbox{3.2cm}{Small \# of cases? Semi-direct products of abelian groups with $A_{4}$ and $S_{4}$?}} & \\ \cline{2-5} & \multicolumn{2}{l|}{$G > Z_{4}$} & \fbox{\parbox{3.2cm}{4-d Euler blow-up of origin. Patch together equivariant lower-d blowups.}} & \S\ref{sec:4-centre}\ \ \ ?\\ \hline \multirow{5}{12mm}{Red.} & \multirow{2}{16mm}{\textbf{(3,1)}} & \multirow{1}{22mm}{$ G_\eta\not >Z_{3}$} & \fbox{\parbox{3.2cm}{Construct equivariant 3-d Euler resolution. Reduces to the toric case.}} & \S\ref{sec:3-1:nocentre} OK\\ \cline{3-5} & & \multirow{1}{22mm}{$G_\eta >Z_{3}$} & \fbox{\parbox{3.2cm}{Euler blow-up of fixed line. Reduce to 2-d equivariant blow-up.}} & \S\ref{sec:3-1}\ \ \ ?\\ \cline{2-5} & \multicolumn{2}{l|}{\textbf{(2,2)}} & \fbox{Use 2-d results} & \S\ref{sec:2-2} ?\\ \cline{2-5} & \multicolumn{2}{l|}{\textbf{(2,1,1)}} & \fbox{Use 2\&3-d results} & \S\ref{sec:2-1-1} ?\\ \cline{2-5} & \multicolumn{2}{l|}{\textbf{(1,1,1,1)} (Abelian)} & \fbox{Toric MMP} & \S\ref{sec:toric}\ \ \ OK\\ \hline \end{tabular} \vspace{1em} \caption{Constructing Euler Terminalisations of ${\mathbb C}^4/G$ for $G<\operatorname{\rm SL}(4)$. (Bold numbers in parentheses indicate the {\em type\/} of the group. In the cases $(3,1)$, the group $G_\eta$ denotes the stabiliser in $G$ of a generic point of the line fixed by $G$. )} \label{tab:sl4} \end{center} \end{table} The {\em Status\/} column indicates the section of this paper which makes some contribution to the problem. Questions which are solved in this paper are indicated by the mention ``OK''. A question mark indicates that the question is still open and that methods of this paper are relevant. Finally, where no results are known, nothing is indicated in the Status column, but some guesses are given as to the likely situation, based on the present state of knowledge. \subsection{Open Problems} \label{sec:open} Interesting open problems arise in relation to recent work of Ito and Reid~\cite{reid_ito} which establishes a one-one correspondence between crepant divisors of $V/G$ and conjugacy classes ``of weight 1'' in (the dual group to) $G$. One interesting question is how this correspondence behaves under the Euler blow-ups which are constructed for the $\operatorname{\rm SL}(4)$ singularities mentioned above. A deeper understanding of this (apart from being of interest in itself) would also no doubt allow one to say more about the singularities of the terminalisation. For instance, for toric and toroidal varieties, the Euler blow-up can be made projective rather than just analytic. But in general, the gluing process in the Patching Lemma~\ref{lemma:patching} does not in itself guarantee that the blow-up $X$ will be projective, because a divisor which is reducible when considered locally in the neighbourhood of one of the points $[\xi]$ might have two of its components identified by the gluing process. Such troublesome cases might conceivably be ruled out by a deeper understanding of the correspondence between divisors and conjugacy classes. \subsection{Outline} \label{sec:intro:outline} Section~\ref{sec:toric} deals with the construction of Euler terminalisations for the toric and toroidal cases. Section~\ref{sec:3-1} deals with the finite $\operatorname{\rm SL}(4)$ groups of type (3,1). Section~\ref{sec:blowing-up} presents a blowing-up construction and some other technical lemmas which are then used in Section~\ref{sec:4-centre} in dealing with irreducible finite $\operatorname{\rm SL}(4)$ subgroups containing $Z_4$. Finally, Section~\ref{sec:2-} makes some comments regarding the finite $\operatorname{\rm SL}(4)$ groups of type (2,2) and (2,1,1). \subsection{Acknowledgments} \label{sec:intro:ack} I wish to acknowledge S.Mori and S-S.Roan for the many ideas and suggestions they contributed to this paper. This research was undertaken at the Research Institute for Mathematical Sciences of Kyoto University thanks to an European Commission Science and Technology Fellowship; I am grateful to my host institution and its staff for their hospitality and to the European Commission for their financial support. \section{Toric and Toroidal Cases} \label{sec:toric} \begin{thm}[Toric Minimal Model Program] \label{thm:toric-mmp} Let $Y$ be a toric variety and $X $ be a simplicial toric variety which admits a projective birational toric morphism $f\colon X \to Y$. Then there exists a sequence $X \stackrel h\dashrightarrow Z\xrightarrow{g}Y$ such that \begin{enumerate} \item $h$ is a composite of toric divisorial contractions or toric flips. \item $g$ is a projective morphism and $Z$ is a simplicial toric variety with terminal singularities such that $K_Z$ is relatively nef for $g$. \end{enumerate} Note that if $Y$ has canonical singularities, then $K_Z=g^*K_Y$ in the sense of $\Q$-Cartier divisors, i.e.\xspace $g$ is crepant. \end{thm} \begin{proof} See~\cite[Theorem 0.2]{reid:toricmmp}, where the result is proved under the assumption that $Y$ is projective. As remarked in~\cite{reid:toricmmp} this assumption is not essential and the result is valid for non-complete toric varieties also (the easiest way to see this is to reduce the non-projective case to the projective one by completing the fan in an appropriate way). \end{proof} \begin{cor} \label{cor:abelian-terminal} All toric {\cGo}s\xspace admit (toric) Euler terminalisations. \end{cor} \begin{proof} Note that a toric variety has at most orbifold singularities if and only if it is simplicial. It is therefore sufficient to prove that any crepant blow-up of a simplicial Gorenstein toric variety must have the same orbifold Euler number as the original. But this is true because the orbifold Euler number of a simplicial toric variety is just the volume\footnote{Also called the {\em multiplicity\/} of the cone.} of the cone, meaning the volume of the simplex defined by the cone's generators; a crepant blow-up corresponds to a fan subdivision by one-dimensional rays whose primitive generators all belong to the same plane, and therefore the sum of the volumes of the cones in such a subdivision is equal to the volume of the original cone. \end{proof} The minimal model program for general varieties is at present only proved in dimension~3. In dimension~4, although termination has been shown~\cite{kmm:intro_mmp} existence of flips remains a problem. Nevertheless, a technique well-known to minimal model program specialists allows one to use the Theorem above and the termination result to say something about toroidal varieties, i.e.\xspace varieties which are only locally isomorphic to toric varieties. The argument can phrased for general $n$, even though at present, termination has only been proved for $n\leq 4$. \begin{thm} \label{thm:toroidal-mmp} Assume that flips terminate in dimension $n$. Then all $n$-dimensional toroidal canonical Gorenstein orbifold\xspace admit Euler terminalisations (which are themselves toroidal). In particular, this is true in dimension~4. \end{thm} \begin{proof} Let $Y$ be a canonical Gorenstein orbifold\xspace locally isomorphic to a toric variety and let $p\colon X\to Y$ be any resolution obtained by toric blowups. Suppose that $K_X$ is not $p$-nef, and let $c\colon X \to W$ be an extremal contraction. If it is a divisorial contraction, then replace $X$ by $W$. If $c$ is a small contraction, consider its restriction $C_{U_X}\colon U_X \to U_W$ to the inverse image of a local toric neighbourhood $U_Y\subset Y$. Theorem~\ref{thm:toric-mmp} implies the existence of a local flip $c^+_{U_X}\colon U^+_X \to U_W$ over each neighbourhood $U_W$. A flip being unique if it exists, the local flips patch together on the overlaps to form a global flip $c^+\colon X^+\to W$. Thus, existence of flips is guaranteed in this case. Applying the same procedure repeatedly (and using the termination hypothesis) results in a projective morphism $p\colon Z\to Y$ such that $Z$ has $\Q$-factorial terminal singularities with $K_Z$ being $p$-nef, which means that $p$ is crepant, since $Y$ is canonical. Furthermore, any toric terminalisation which is crepant must have the same orbifold Euler number by the volume argument in Corollary~\ref{cor:abelian-terminal}. \end{proof} Corollary~\ref{cor:abelian-terminal} and Theorem~\ref{thm:toroidal-mmp} together give Theorem~\ref{thm:toric-toroidal}. \section{Groups $G$ of type (3,1)} \label{sec:3-1} \subsection{Notation} \label{sec:3-1:notation} Let $G<\operatorname{\rm SL}(n+1)$ be a finite subgroup such that $V={\mathbb C}^{n+1}$ decomposes into two irreducible $G$-modules: $V=V^1\oplus V^2$, with $V^1$ of dimension $n$, and $V^2$ of dimension 1. Denote by $\eta$ the generic point of the line $\{0\}\times V^2$, and by $G_\eta$ the stabiliser of $\eta$. Note that $G_\eta$ is a subgroup of $\operatorname{\rm SL}(n)\times\{1\}\cong \operatorname{\rm SL}(n)$. The quotient $C:= G/G_\eta$ is cyclic, being a naturally a subgroup of $\operatorname{\rm GL}(V^2)\cong{\C^*}$. The canonical quotient map is denoted $\pi\colon G\to C$ and the induced map on conjugacy classes is denoted $\pi_*\colon\operatorname{\rm Cl}(G)\to \operatorname{\rm Cl}(C)$. Note that $G$ can be considered as a subgroup of $\operatorname{\rm GL}(V^1)=\operatorname{\rm GL}(n)$ by forgetting about the last row and column of any matrix element. Let $h\in G$ be an element such that $G=\<<G_\eta,h>>$ --- i.e.\xspace a representative for a generator of $C$ --- and denote by $h_1$ the $n\times n$ sub-matrix consisting of the first $n$ rows and columns of $h$. Let $\lambda_h$ be a complex $n$-th root of $\det(h_1)^{-1}$. Then $h':= \lambda_h h_1\in\operatorname{\rm SL}(n)$ and normalises $G_\eta$. Hence $G_\eta$ is a normal subgroup of $G':=\<<G_\eta,h'>> < \operatorname{\rm SL}(n)$ with quotient a cyclic subgroup $C'$. Note that $G'$ and $C'$ are defined up to a choice of the $n$-th root $\lambda_h$. The canonical quotient map is denoted $\pi'\colon G'\to C'$ and the induced map on conjugacy classes is denoted $\pi^{\prime}_*\colon\operatorname{\rm Cl}(G')\to \operatorname{\rm Cl}(C')$. \subsection{Equivariant Resolution} \label{sec:3-1:resolution} From now on, let $n=3$, i.e.\xspace consider the case where $G<\operatorname{\rm SL}(4)$ fixes a line in ${\mathbb C}^4$ (and therefore $G_\eta<\operatorname{\rm SL}(3)$). \begin{conj} \label{conj:3-1-equiv-Euler} Let $G<\operatorname{\rm SL}(4)$ be a finite subgroup which stabilises a line $V^2\subset V={\mathbb C}^4$ and let $G_\eta$ denote the stabiliser in $G$ of the generic point of $V^{2}$ and let $G'$ and $C'$ be defined as in Section~\ref{sec:3-1:notation} above. Then there exists a $G'$-invariant Euler resolution $W^1\to V^1/G_\eta$, satisfying \begin{equation} \label{eq:chi-pi} \chi((W^1)^{c'}/{C'})=|\pi^{\prime -1}_*([c'])|, \end{equation} for all $c'\in C'$. \end{conj} If true, we shall see that this conjecture implies that $\text{Term}({\mathbb C}^{4}/G)$ is true for groups stabilizing a line (i.e.\xspace types $(3,1)$ and $(2,1,1)$) and (cf.\ Section~\ref{sec:4-centre}) for all irreducible $G$ which contain $Z_{4}$. \begin{prop} \label{prop:3-1-invariant-resolution-gives-terminalisation} Suppose that Conjecture~\ref {conj:3-1-equiv-Euler} is true for $G$. Then $W^1\times V^2/C\to V/G$ is an Euler blowup, and $V/G$ has an Euler terminalisation with only toric singularities. \end{prop} \begin{proof} First, note that since $V^1/G_\eta$ has dimension at most three, it has a minimal model. The fact that there are only a finite number of distinct minimal models implies that ${\C^*}$ acts on any of them: for if the action of some element $\lambda\in{\C^*}$ produced a different minimal model, then, by continuity, one could produce countably many distinct minimal models by acting with a countable family of distinct neighbours of $\lambda$. Second, note that the equality~\eqref{eq:chi-pi} would follow from the same equality with $C'$ replaced by the group $C$. For if $\phi$ is any element of $G$, one can find $\lambda\in{\C^*}$ such that $\phi'=\lambda \phi\in G'$. The invariant sets $(W^1)^\phi$ and $(W^1)^{\phi'}$ have the same homotopy type, by an easy application of Bialynicki-Birula's well-known decomposition theorem~\cite[Thm.~4.1]{bb:algebraic_groups} to the smooth variety $W^1$. Hence $\chi((W^1)^\phi)=\chi((W^1)^{\phi'})$ and so, averaging, \begin{equation} \label{eq:chi-equality} \chi((W^1)^\phi/C)=\chi((W^1)^{\phi'}/C'). \end{equation} On the other hand, if $c$ and $c'$ denote the images in $C$ and $C'$ of $\phi$ and $\phi'$ respectively, \begin{equation} \label{eq:pi-equality} |\pi_*^{-1}([c])|=|\pi^{\prime -1}_*([c'])|, \end{equation} where $\pi$ denotes the projection $\pi\colon G\to C$ and $\pi_*\colon\operatorname{\rm Cl}(G)\to\operatorname{\rm Cl}(C)$ the induced map on conjugacy classes, and similarly for the primed symbols. Hence one can assume formula~\eqref{eq:chi-pi} to be valid for the group $C$. The variety $(W^1\times V^2)/G$ is a blow-up of $V/G$ having only cyclic quotient singularities resulting from the residual action of $C=G/G_\eta$ on $W^1\times V^2$. Its orbifold Euler number can be expressed as a sum: \begin{align} \chi_{\text{orb}}(W^1\times V^2/C) &=\sum_{[c]\in\operatorname{\rm Cl}(C)} \chi((W^1\times V^2)^c/C),\quad\text{since $\operatorname{\rm N}^C_c=C$}\notag\\ &= \sum_{[c]\in\operatorname{\rm Cl}(C)} \chi((W^1)^c/C),\\ &\qquad\text{since $(V^2)^c$ is contractible}. \label{eq:orb-sum} \end{align} On the other hand, one has: \begin{equation} \label{eq:cclass-sum} |\operatorname{\rm Cl}(G)|= \sum_{[c]\in \operatorname{\rm Cl}(C)} |\pi_*^{-1}([c])|, \end{equation} which agrees with the previous sum term-by-term. Hence, $Y:=(W^1\times V^2)/G\to (V^1\times V^2)/G$ is an Euler blow-up with only toric (cyclic) singularities. Applying the minimal model program (Theorem~\ref{thm:toroidal-mmp}) to $Y$, one obtains a crepant terminalisation $t\colon Z\to Y$ which satisfies $\chi_{\text{orb}}(Z)=\chi_{\text{orb}}(Y)=\chi_{\text{orb}}(V/G)$, and has only toric singularities. \end{proof} Thus in the case where $G$ fixes a line in $V={\mathbb C}^4$ it suffices to prove the existence of a $G'$-equivariant Euler resolution $W^1\to V^1/G_\eta$ which satisfies equation~\eqref{eq:chi-pi}. \begin{rmk} The same method as above can be used to deal with the easier case when $n=2$ and $G<\operatorname{\rm SL}(3)$ and fixes a line in ${\mathbb C}^3$. In 2 dimensions, the minimal model is unique, so there is no need to check $G'$-stability of the Euler resolution of $V^1/G_\eta$. \end{rmk} \subsection{Case where $G_\eta$ doesn't contain $Z_3$} \label{sec:3-1:nocentre} In this section, Conjecture~\ref{conj:3-1-equiv-Euler} is proved in the case where $G_\eta$ is irreducible and does not contain $Z_{3}$. In order to construct a $G$-equivariant Euler resolution $W^1\to V^1/G_\eta$, the cases to be considered are first restricted using the following lemma (which makes use of the classification of small finite sub-groups of $\operatorname{\rm SL}(3)$ --- see~\cite{yau_yu}, although the notation adopted here is that of~\cite{roan:calabi-yau}, which is slightly different). \begin{lemma} \label{lemma:3-1-nocentre-exact} Suppose that \begin{equation} \label{eq:3-1-nocentre-exact} 1 \to G_\eta \to G' \xrightarrow{\pi'} C' \to 1, \end{equation} is an exact sequence of finite groups of $\operatorname{\rm SL}(3)$ such that $C'$ is cyclic and non-trivial, and $G_\eta$ is non-abelian and doesn't contain $Z_3=\<<\omega_3>>$. Then exactly one of the following is true. \begin{enumerate} \item $G_\eta$ is of type (B) and $G'$ is of type (B). \item $G_\eta$ is of type (C), (D), (H) or (I) and $G' = \<<G_\eta,\omega_3>>$. \item $G_\eta$ is of type (C) and $G'$ is of type (D), with $G'=\<<G_\eta,R>>$ or $G'=\<<G_\eta,\omega_3R>>$. \end{enumerate} \end{lemma} \begin{proof} The groups of type (E), (F), (G), (H${}^*$), (I${}^*$) all contain $Z_3$, so do not occur as the group $G_\eta$ by assumption. The only finite subgroup of $\operatorname{\rm SL}(3)$ containing the simple group (H) (resp.\ the simple group (I)) as a normal subgroup is (H${}^*$) (resp.\ (I${}^*$))~\cite[p.36]{yau_yu}. Thus for these, the result follows immediately. If $G_\eta$ has type (B), a simple argument~\cite[\S1.4, p.18]{yau_yu} shows that $G'$ must also have type (B). It remains to deal with the case where $G_\eta$ has type (C) or (D). Throughout the rest of this proof, write $$T:= \begin{pmatrix} 0& 1& 0\cr 0& 0& 1\cr 1& 0& 0 \end{pmatrix}, $$ for the element which, together with a diagonal group, generates a group of type (C). To get a group of type (D), recall that one must add to a group of type (C) an element of the form \begin{equation} \label{eq:phi} \phi= \begin{pmatrix} a& 0& 0\cr 0& 0& c\cr 0& b& 0\cr \end{pmatrix} \quad\text{with }abc=-1. \end{equation} \begin{claim} If $G_\eta$ has type (C) or (D), then $G'$ must also be of type (C) or (D) (though not necessarily the same type as $G_\eta$). \end{claim} \begin{proof} Denote by $x_1, x_2, x_3$ the standard coordinates on ${\mathbb C}^3$. If $G_\eta$ is of type (C) or (D) and does not contain the centre $Z_3$, then the monomial $x_1x_2x_3$ is invariant under $G_\eta$ up to scale~\cite[\S 1.3]{yau_yu}. It follows from an easy argument that $G'$ must also leave ${\mathbb C} x_1x_2x_3$ invariant. Thus $G'$ cannot be primitive. This means that $G'$ must be of type (C) or (D). \end{proof} The next step is to study the normal diagonal subgroups of $G_\eta$ and $G'$, which are denoted by $H_\eta$ and $H'$ respectively. \begin{notation} The standard toric notation for diagonal matrices will be used: $$\qsing1/d(r_1, r_2, \dots, r_n):\equiv [\exp({\frac{2\pi i r_1}{ d}}), \exp({\frac{2\pi i r_2}{d}}), \dots, \exp({\frac{2\pi i r_n}{d}})]$$ \end{notation} \begin{claim} \label{claim:order3} If $H'$ contains an element of order $3$ then that element must be $\omega_3$ or $\omega_3^2$. As a consequence, all the elements of $H_\eta$ have orders prime to $3$. \end{claim} \begin{proof} If $x\in H'$ has order $3$ and its does not belong to the centre $Z_3$ then it can be chosen to be of the form $x=\qsing1/3(i,i+1,i+2)$ for some $i\in\{0,1,2\}$. But then \begin{align*} (TxT^{-1})x^{-1} &= \qsing1/3(i+1,i+2,i) - \qsing1/3(i,i+1,i+2)\\ &= \qsing1/3(1,1,-2)\\ &= \qsing1/3(1,1,1)=\omega_3 \end{align*} On the other hand, since $x$ normalises $G_\eta$, one has $xT^{-1}x^{-1}\inG_\eta$ and so $\omega_3=TxT^{-1}x^{-1}\inG_\eta$, which contradicts the hypothesis of the lemma. Thus the only elements of order $3$ in $H'$ are $\omega_3$ and $\omega_3^2$. As a consequence, $H_\eta$ has no elements of order $3$, since $Z_3\not < H_\eta$. The claim for $H_\eta$ follows immediately from this. \end{proof} \begin{claim} \label{claim:hprime} $H' < \<< H_\eta,\omega_3 >>$, i.e. $H'$ is either equal to $H_\eta$ or equal to $\<< H_\eta,\omega_3>>$. \end{claim} \begin{proof} Let $\varphi\in H'$. I begin by showing that $\varphi^3\in H_\eta$. Since $H'$ is normal in $G'$, the element $T \varphi T^{-1}$ is diagonal, and so, therefore, is $f:=\varphi^{-1}T \varphi T^{-1}$. One has $fT = \varphi^{-1} T\varphi\in G_\eta$, since $G_\eta$ is normal in $G'$, so $f$ belongs to $G_\eta$. Since $f$ is also diagonal, it follows that $f\inH_\eta$. Writing $\varphi = \qsing 1/d(a,b,-a-b)$, one has \begin{align*} f &=\varphi^{-1}(T \varphi T^{-1})= \qsing1/d(-a,-b,a+b)+\qsing1/d(b,-a-b,a)\\ &=\qsing1/d(b-a,-a-2b,2a+b)\\ \intertext{and} T^{-1}f T &= \qsing1/d(2a+b,b-a,-a-2b). \end{align*} Dividing the second element by the first gives $$T^{-1} f T f^{-1} = \qsing1/d(3a,3b,-3(a+b))= \varphi^3,$$ so $\varphi^3\inH_\eta$. Let $x:=\varphi^{-3}\inH_\eta$. By Claim~\ref{claim:order3}, the order of $x$ is prime to 3, so there exists an integer $l$ such that $x^{3l}=x$. Writing $\alpha := x^l=(\varphi^{-3l})\inH_\eta$, one has $(\alpha\varphi)^3=1$, and so Claim~\ref{claim:order3} again implies that $\alpha \varphi =\omega_3$ or $\omega_3^2$. Thus $H'=\<<H_\eta,\omega_3>>$. \end{proof} Now one can deal with the groups $G_\eta$ and $G'$ themselves. To begin with, since the quotient $C'=G'/G_\eta$ is cyclic, $G'=\<< G_\eta,\phi>>$ for some $\phi\in G'\setminus G_\eta$. Now any element $\phi$ in a group of type (C) or (D) has associated to it a permutation $\sigma(\phi)\in S_3$, defined according to how it permutes the coordinates $x_1$, $x_2$, $x_3$. If $\sigma(\phi)$ is the identity, then $\phi$ is diagonal, whereas if $\sigma(\phi)$ is a permutation of order 3 then $\phi T$ or $\phi T^{-1}$ is diagonal. In these cases, since $T\inG_\eta$, it follows from the claim above that $G' = \<< G_\eta,\omega_3>>$. The only remaining possibility is that $\sigma(\phi)$ equals a transposition or order 2, which can be assumed to be the transposition $(12)$, by multiplying $\phi$ by a suitable power of $T$. Thus $\phi$ is of the form~\eqref{eq:phi}. \begin{claim} \label{claim:type_D} For any $\phi$ of the form~\eqref{eq:phi}, define $$\tilde\phi:= T^{-1}\phi^2T\phi= \begin{pmatrix} A& 0& 0\cr 0& 0& C\cr 0& B& 0\cr \end{pmatrix},$$ with $A=-1$, $B=b^2c$ and $C=-B^{-1}$. Suppose that $Z_3\not<\<<\tilde\phi,T>>$. Then there exists an element $t\in \<<T,\tilde\phi>>$ such that $t\tilde\phi=R$, where $$R:=\begin{pmatrix}-1& 0& 0\cr 0& 0& -1\cr 0& -1& 0 \end{pmatrix}.$$ \end{claim} \begin{proof} Define $f:=\tilde\phi T\tilde\phi^{-1}T=[-B,-B,B^{-2}]$ and $f':=fT^{-1}fT=[1,-B^3,-B^{-3}]$. If the order of $B$ is a multiple of $3$, say $m=3k$, then $f^{2k}=\omega_3$ or~$\omega_3^2$, so $Z_3<\<<\tilde\phi,T>>$. Thus if $Z_3\not<\<<\tilde\phi,T>>$ then the order of $B$ is prime to three, and a suitable power of $f'$ gives the required element $t=[-1,-B^{-1},-B]$, which satisfies $t\tilde\phi=R$. \end{proof} Now the element $\phi^2$ is diagonal, so belongs to $\<< H_\eta, \omega_3>>$, by claim~\ref{claim:hprime}. Note also that $\phi=\tilde\phi(T\phi^{-2}T^{-1})\in\tilde\phi H'$. {\bf Case 1: $\phi^2 \in H_\eta$} In this case, $C'$ has order 2, so $\omega_3\not\in H'$ --- i.e.\xspace $H'=H_\eta$. By Claim~\ref{claim:type_D}, $t\tilde\phi=R$ for some element $t\in \<<T,\tilde\phi>> < G_\eta$. Thus $G'=\<<G_\eta,\phi>>=\<<G_\eta,\tilde\phi,H'>> = \<<G_\eta,R>>$, since $H'=H_\eta$. {\bf Case 2: $\phi^2\not\inH_\eta$} Then $\phi^2=\varphi\omega_3^k$ for $k=1$ or~$2$ and $\varphi\inH_\eta$. Let $\phi':=\omega_3^k\phi$. Then $(\phi')^2=\varphi^2\inH_\eta$, so by the preceding case, one may assume $t\tilde\phi'=R$ for some $t\in\<<T,\tilde\phi'>>< G_\eta$. Now $\tilde\phi=\widetilde{\phi'}$, so $G'=\<<G_\eta,\phi>>=\<<G_\eta,\tilde\phi,H'>> = \<<G_\eta, Rt^{-1},H'>> = \<<G_\eta,R,\omega_3>>$. This completes the proof of the lemma. \end{proof} \begin{prop} \label{prop:3-1-nocentre-equiv-Euler} Conjecture~\ref{conj:3-1-equiv-Euler} is true when $G_\eta$ is irreducible and doesn't contain $Z_{3}$. \end{prop} \begin{proof} Recall from the proof of Prop.~\ref{prop:3-1-invariant-resolution-gives-terminalisation}, that ${\C^*}$ acts on any Euler resolution. Thus, in all cases where $G'=\<<G_\eta,\omega_3>>$, any smooth crepant resolution of $V^1/G_\eta$ admits a $G'$-action --- indeed a $G$-action, as remarked in the proof of Prop.~\ref{prop:3-1-invariant-resolution-gives-terminalisation}, since $G<\<<G',{\C^*}>>$. Hence, Lemma~\ref{lemma:3-1-nocentre-exact} implies that it suffices to deal with the cases where $G=\<<G_\eta,\omega_3>>$ or $G_\eta=\<<H_\eta,T>>$ is of type (C) with $G'$ of type (D), either equal to $\<<G_\eta,R>>$ or equal to $\<<G_\eta, R,\omega_3>>$. In these cases (see~\cite{roan:calabi-yau}) an Euler resolution of ${\mathbb C}^3/G_\eta$ is obtained by taking a toric Euler resolution of $\widetilde{{\mathbb C}^3/H_\eta} \to {\mathbb C}^3/H_\eta$ which is $\<<T>>$-stable and then resolving the singularities of $\widetilde{{\mathbb C}^3/H_\eta}/\<<T>>$. The existence of a $T$-stable toric Euler resolution follows from the fact that $H_\eta$ is normal in $G_\eta$. However, $H_\eta$ is also normal in $G'$, so $\widetilde{{\mathbb C}^3/H_\eta}$ can also be chosen to be $R$-stable. The singularities of $\widetilde{{\mathbb C}^3/H_\eta}/\<<T>>$ are fixed points of $R$, so resolving them gives the desired $G'$-invariant resolution of ${\mathbb C}^3/G_\eta$. The proof that $\chi((W^1)^{c'}/C')=|\pi_*^{\prime -1}([c'])|$ is done by treating case by case the three possibilities for $G'$ given by Lemma~\ref{lemma:3-1-nocentre-exact}. {\bf Case 1: $G'=\<<G_\eta,\omega_3>>$.} In this case, since $\omega_3\in{\C^*}$, the same argument as the second paragraph of the proof of Proposition~\ref{prop:3-1-invariant-resolution-gives-terminalisation} gives $\chi((W^1)^{\omega_3})=\chi(W^1)=|\operatorname{\rm Cl}(G_\eta)|$. On the other hand, $G'$ is just a direct product of $G_\eta$ and $Z_3$, so $\pi_*'$ is everywhere $|\operatorname{\rm Cl}(G_\eta)|:1$. {\bf Case 2: $G'=\<<G_\eta,R>>$.} If we denote by $H_\eta$ and $H'$ the normal diagonal subgroups of $G_\eta$ and $G'$ respectively, then they are equal. Since they do not contain $\omega_3$, their order is $d^2$ (for some $d$ prime to~3) and they are a semi-direct factor in $G'$~\cite[Lemma 10]{roan:crepant}. The inverse image of the trivial class in $C'$ is the number of $G'$-conjugate elements in $G_\eta$. Since $G_\eta$ is normal, this is the same as the number of $G_\eta$-conjugate elements in $G_\eta$, i.e.\xspace equal to $\operatorname{\rm Cl}(G_\eta)=\chi(W^1)$. For the non-trivial class $[R]$, \cite[Formula~(32)]{roan:crepant} implies that $\chi((W^1)^R/\<<R>>)=d$ and the proof of ~\cite[Lemma 10]{roan:crepant} again gives $\pi_*^{\prime-1}([R])=|Z_R\cap H'|=d$. {\bf Case 3: $G'=\<<G_\eta, \omega_3 R>>$.} As we remarked in Case~1, the Euler number does not depend on scalar factors, so $\chi((W^1)^{\omega_3 R})=\chi((W^1)^R)$. On the other hand, the discussion in the second paragraph of this proof implies that $|\pi^{\prime-1}_*([\omega_3 R])|=|\pi^{\prime-1}_*([R])|$. The result thus follows from Case~2. \end{proof} Hence, if $G_\eta$ is irreducible and does not contain $Z_3$, $V/G$ has an Euler terminalisation with only toric singularities. \begin{question} The proof of Propositions~\ref{prop:3-1-nocentre-equiv-Euler} depends on the classification of $\operatorname{\rm SL}(3)$ groups. Is it possible to find a proof which doesn't depend on the classification? \end{question} \subsection{Case where $G_\eta$ contains $Z_3$} \label{sec:3-1:centre} Unfortunately, the author was not able to prove the corresponding result to Conjecture~\ref{conj:3-1-equiv-Euler} in the case where $G_\eta$ is irreducible but contains $Z_3$. A method is suggested in Section~\ref{sec:3-1:centre:2}, but requires further work. If it the conjecture can be proved in all cases, the work above implies that $\text{Term}({\mathbb C}^{4}/G)$ is true for all groups of type $(3,1)$ and $(2,1,1)$. The results of the next section imply that in that case, $\text{Term}({\mathbb C}^{4}/G)$ is true for all irreducible $G$ which contain $Z_{4}$. \section{Blowing up in the presence of the centre $Z_n$} \label{sec:blowing-up} \subsection{Invariant sets in projective space} First, a lemma about the Euler number of a invariant sets in projective space. \begin{lemma} \label{lemma:euler-proj} Let $H$ be a finite abelian group acting linearly on $\PP^n$. Then $$\chi((\PP^n)^H) = \chi(\PP^n)=n+1.$$ \end{lemma} \begin{proof} Suppose $H$ has order $r$. Diagonalise the action of $H$, and order the weights of the action so that they form a non-decreasing sequence of elements of $\{0,\dots,r-1\}$. The sequence will consist of $d_1$ occurrences of the smallest weight $w_1$, followed by $d_2$ occurrences of the second smallest weight $w_2$, and so on, ending with $d_s$ occurrences of the greatest weight $w_s$. Since there are $n+1$ (not necessarily distinct) weights in the sequence, the multiplicities $d_i$ sum to $n+1$. Computing the invariant part of $\PP^n$ with respect to the $H$-action, one sees that it consists of a disjoint union over all $i\in\{1,\dots,s\}$ of projective spaces $\PP^{d_i-1}$. Taking the sum of the Euler numbers of the invariant components, and using the fact that the Euler number of $\PP^d$ is $d+1$, one obtains the value $\sum d_i$, which by the previous paragraph indeed coincides with the Euler number $n+1$ of $\PP^n$. \end{proof} \subsection{Blowing up the origin} \label{sec:blowing-up:blowing-up} Now let $V= {\mathbb C}^n$ and let $\text{Bl}_0V$ be the blow-up of $V$ at the origin. This has a natural $G$-action and one has the following commutative diagram $$\begin{CD} & \text{Bl}_0V @>\sigma_0>> & V \\ & \downarrow & & & \downarrow \\ & \text{Bl}_0V/G @>{\sigma'_0}>> & V/G . \end{CD}$$ A standard discrepancy calculation yields the following result. \begin{lemma} The morphism $\text{Bl}_0V/G \to V/G$ is crepant if and only if $G$ contains $Z_n=\<<\omega_n>>$. \end{lemma} The following lemma contains the basic idea to constructing Euler blow-ups. \begin{lemma} \label{lemma:bV} Assume that $G$ contains $Z_n$ and write $\bar G:=G/Z_n$. Then $\bar V := \text{Bl}_0V/Z_n$ is smooth, and $\bar V/\bar G \to V/G$ is a projective Euler blow-up. \end{lemma} \begin{proof} The only place where singularities of $\bar V$ could arise is on the image $\bar E=E/Z_n$ of the exceptional divisor $E$ of $\sigma_0$. Identifying $E$ with $\PP(V)$, a local chart for $\text{Bl}_0V$ at a point $\xi\in E$ is given in suitable local coordinates by $$(x_1,\frac{x_2}{x_1},\dots,\frac{x_n}{x_1}),$$ so therefore $\omega_n$ acts there as $(\omega_n, 1,1,1)$, i.e.\xspace as a pseudo-reflection. The Euler number computation goes as follows: \begin{align*} \chi_{\text{orb}}(\bar V/\bar G) &=\sum_{[\bar g]\in\operatorname{\rm Cl}(\bar G)}\chi(\PP(V)^{\bar g}/\operatorname{\rm N}^{\bar G}_{\bar g}), \\ & \qquad\text{since $\bar V\sim E\cong\PP(V)$\ \ (homotopy)}\\ &=\sum_{[\bar g]\in\operatorname{\rm Cl}(\bar G)} \chi(\PP(V)), \\ & \qquad\text{averaging and using Lemma~\ref{lemma:euler-proj}},\\ &=|\operatorname{\rm Cl}(\bar G)| n = |\operatorname{\rm Cl}(G)|\\ &=\chi_{\text{orb}}(V/G). \end{align*} \end{proof} Thus if $G$ contains $Z_n$ then $\text{Term}(V/G)\equiv\text{Term}(\bar V/\bar G)$. \subsection{Patching} \label{sec:blowing-up:patching} In many cases, an Euler blow-up is constructed by patching together local Euler blow-ups. The following lemma summarises the necessary conditions to carry this out this procedure. \begin{lemma}[Patching Lemma] \label{lemma:patching} Let $G$ be a finite subgroup of $\operatorname{\rm SL}(n)$ which contains $Z_n$, so that $Y:=\bar V/\bar G\to V/G$ is an Euler blow-up. Let $\bar E:\equiv E/Z_n$, where $E$ is the exceptional divisor of $\text{Bl}_0 V\to V$ and denote by $p\colon\bar V\to\bar E$ the projection. There exists a finite collection of points $y\in\bar E/\bar G$ and corresponding analytic neighbourhoods $\bar E_\xi\subset\bar E/\bar G$ such that $Y$ is covered by $\{Y_y\}$, where $Y_y:=p^{-1}(\bar E_y)$. Suppose that for each $y$, there exists an Euler blow-up $\varphi_y\colon X_y\to Y_y$, such that if $y\neq y'$, one nevertheless has \begin{equation} \label{eq:patching} X_{y|Y_y\cap Y_{y'}} = X_{y'|Y_y\cap Y_{y'}}. \end{equation} Then the analytic canonical Gorenstein orbifold\xspace\ $X$ obtained by gluing together all the $ X_y$ is an Euler blow-up of $Y$. \end{lemma} \begin{proof} The existence of the finite covering $\{\bar E_y\}$ follows because $\bar E$ is compact. Since, all the $X_y$ are birational to each other above the overlaps, equation~\eqref{eq:patching} implies that $X$ is well-defined. Furthermore, since no crepant divisors are introduced during the local blow-ups and since the orbifold Euler numbers of $X_y$ and $Y_y$ are the same, $X$ is crepant over $Y$ and has the same orbifold Euler number. \end{proof} \section{Irreducible $G$ which contain $Z_4$} \label{sec:4-centre} \subsection{Notation} \label{sec:4-centre:notation} When $G$ contains $Z_4$, $\bar V/\bar G\to V/G$ is an Euler blow-up. Let $\bar E$ be the exceptional divisor of $\bar V\to V/Z_4$, and let $p\colon \bar V\to \bar E$ be the projection. Let $\xi\in \bar E$ be a point in the base, and consider the tangent space of $\bar V$ at $\xi$. This decomposes into $\bar G$-modules $$V_\xi^1\oplus V_\xi^2,$$ where $V^1_\xi$ is the tangent space to $\bar E$ and $V^2_\xi$ is the line tangent to the fibre of $p$, and stabilized by $\bar G$. Let $\xi'\in V_\xi^2$ be the generic point, so that its stabiliser $\bar G_{\xi'}$ is a subgroup of $\operatorname{\rm SL}(3)$. \subsection{Local Blow-ups} \label{sec:4-centre:blowups} Let $\xi\in \bar V$ and let $\bar\xi$ denote its image in $Y=\bar V/\bar G$. A local analytic neighbourhood of $\bar\xi$ is isomorphic to $$Y_{\bar\xi} := (V_\xi^1\oplus V_\xi^2)/\bar G_\xi= (V_\xi^1/\bar G_{\xi'}\oplus V_\xi^2)/C_{\xi},$$ where $C_\xi$ denotes the quotient $C_\xi:= \bar G_{\xi}/\bar G_{\xi'}$, which is cyclic, being a naturally a subgroup of $\operatorname{\rm GL}(V_\xi^2)\cong{\C^*}$. Restricting attention to the ``base'' $\bar E$, one has, corresponding to each $\bar\xi$ in $\bar E $, a quotient singularity $\bar E_{\bar\xi} = V_\xi^1/\bar G_{\xi'}$ which is an $SL(3)$-singularity. Since $\bar E/\bar G$ is compact, the choice of a finite number of points $\bar\xi$ is sufficient for $\bigcup_{\bar\xi} Y_{\bar\xi}$ to cover the whole of $\bar V/\bar G$. \subsection{Gluing} \label{sec:4-centre:gluing} \begin{lemma} \label{lemma:4-centre-gluing} Let $W_\xi^1\to V_\xi^1/\bar G_{\xi'}$, be the $\bar G_\xi$-equivariant Euler resolution such as that in Conjecture~\ref{conj:3-1-equiv-Euler}. Define $$ X_\xi := (W_\xi^1\times V_\xi^2)/\bar G_{\xi}.$$ Then $\varphi_\xi\colon X_\xi\to Y_\xi$ are Euler blowups which have only cyclic quotient singularities, and for each $\xi,\xi'$, $ X_\xi$ and $ X_{\xi'}$ agree above the inverse image of $Y_\xi\cap Y_{\xi'}$. Thus by Lemma~\ref{lemma:patching}, they glue to form a complex analytic Euler blow-up $Y \to \bar V/\bar G$ which has only cyclic quotient singularities. \end{lemma} \begin{proof} One must check that $\chi_{\text{orb}}(X_\xi)=\chi_{\text{orb}}(Y_\xi)$ and that , on $\varphi_\xi^{-1}(Y_\xi\cap Y_{\xi'})$ the blow-ups corresponding to $\xi$ and $\xi'$ agree. The orbifold Euler number equality is checked in Proposition~\ref{prop:3-1-invariant-resolution-gives-terminalisation}. For the agreement of the blow-ups on the overlaps, knowledge of the $SL(3)$ singularities implies that these overlaps only occur over curves of 2-dimensional singularities (the components of the non-isolated singularities are all curves for $\operatorname{\rm SL}(3)$). Over these, the resolutions which are being glued-in are trivial families of minimal resolutions: they are therefore unique, and so resolutions coming from neighbourhoods corresponding to different $\xi$'s will agree. \end{proof} This gives a crepant analytic blow-up $Y$ which is locally analytically isomorphic to a cyclic quotient (and hence locally analytically $\Q$-factorial). \subsection{Terminalisation and the Orbifold Euler Number} \label{sec:4-centre:terminalisation} Since the orbifold Euler number can be calculated by summing the contributions of the various analytic neighbourhoods, the equality $$\chi_{\text{orb}}(X)=\chi_{\text{orb}}(\bar V/\bar G)$$ will follow by showing that the resolutions glued in above preserve the orbifold Euler number. This is proved in Proposition~\ref{prop:3-1-nocentre-equiv-Euler}. Thus $X$ is an Euler blow-up with only cyclic singularities. Applying the minimal model program to $X$ (Theorem~\ref{thm:toroidal-mmp}), one obtains a crepant terminalisation $t\colon Z\to X$ which satisfies $\chi_{\text{orb}}(Z)=\chi_{\text{orb}}(X)=\chi_{\text{orb}}(V/G)$, and has only toric singularities. \begin{rmk} By studying which toric flips can occur, one might be able to prove that the singularities of $T$ are in fact at most cyclic. They would then have to necessarily be isolated. For if a 4--dimensional Gorenstein cyclic singularity consisted of a curve of singularities, these would also have to be (3-dimensional) terminal Gorenstein cyclic quotients. But the classification of 3-dimensional terminal cyclic quotients~\cite{mor_stev:terminal} shows that they are all of the form $\frac{1}{r}(1,-1,a)$, and so can only be Gorenstein if they are smooth. \end{rmk} \subsection{Case where $G$ is of type $(3,1)$ revisited} \label{sec:3-1:centre:2} A method similar to the one above can be applied to the case treated in Section~\ref{sec:3-1:centre}, namely the case where $G$ is a group of type (3,1) and the stabiliser group $G_\eta<\operatorname{\rm SL}(3)$ is irreducible and contains $Z_3$. This goes some way towards a solution of Conjecture~\ref{conj:3-1-equiv-Euler}. Since $G_\eta >Z_3$, there exists an Euler blow-up $\bar V^1/\bar\Ge\to V^1/G_\eta$ and this is equivariant under the $G$ (and hence $G'$) action, since it is obtained from blowing up the origin of $V^1$, which is of course fixed by $G$. Its singularities are of the type $(2,1)$ and $(1,1,1)$ and the singular locus is invariant under $G$. An analytic resolution of $\bar V^1/\bar\Ge$ can be constructed by the same gluing procedure as in Lemma~\ref{lemma:4-centre-gluing}: in any local analytic neighbourhood of $\bar V^1/\bar\Ge$, construct an Euler resolution, doing this equivariantly under the $G'$-action. These glue together, since they can only intersect over smooth points. This gives a resolution $W^1\to\bar V^1/\bar\Ge$. It remains to show that it admits a $\bar G$-action and that it satisfies the Euler number property of equation~\eqref{eq:chi-pi}. \section{Groups $G$ of type (2,2) and (2,1,1)} \label{sec:2-} \subsection{Groups of type $(2,2)$} \label{sec:2-2} Let ${\mathbb C}^4=V^1\oplus V^2$ with $V^i$ 2-dimensional irreducible $G$-modules and denote by $\eta_i$ the generic point of $V^i$. If $G<\operatorname{\rm SL}(2)\times\operatorname{\rm SL}(2)$ then the following lemma constrains the stabilisers of $\eta_i$. \begin{lemma} \label{lemma:terminal} Suppose that $V=V^1\oplus V^2$ and that $G<\operatorname{\rm SL}(V^1)\times\operatorname{\rm SL}(V^2)<\operatorname{\rm SL}(V)$. Denote by $G_{i}$ the stabiliser of the generic point $\eta_i\in V^i$ for $i=1,2$. If both stabilisers $G_1$ and $G_2$ are trivial, then $V/G$ must be terminal. \end{lemma} \begin{proof} This can be proved by an easy discrepancy calculation, or equivalently, by using the concept of ``weights'' for the group action~\cite{reid_ito} as follows. The number of crepant divisors of $V/G$ is equal to the ``number of elements of $G(1)$ of weight one''. Here, $G(1):=\operatorname{\rm Hom}(\mu_r, G)$, where $r$ is the least common multiple of the orders of the elements of $G$. The {\em weight\/} $\text{wt}(\hat g)$ of $\hat g\in G(1)$ is defined by evaluating $\hat g$ on a primitive generator $\epsilon$ of $\mu_r$, diagonalising the resulting matrix $\hat g(\epsilon)$ and expressing the diagonal elements in terms of powers of $\epsilon$ ranging between $0$ and $r-1$. Because $G<\operatorname{\rm SL}(V)$, the sum of these powers divided by $r$ is a non-negative integer, called the {\em weight\/} of $\hat g$. Note that if we simply want to calculate the {\em number\/} of elements of $G(1)$ of a given weight, we can identify $G$ with $G(1)$ by fixing a primitive generator of $\mu_r$, and pretend to be calculating the weights of the elements of $G$. Suppose $V/G$ is not terminal, so that there exists and element $g\in G$ of weight one. For each $i=1,2$, denote by $g_i$ the part of the matrix of $g$ which represents its action on the module $V^i$. Since $\text{wt}(g)=\text{wt}(g_1)+\text{wt}(g_2)$, one of the $g_j$'s must be equal to the identity (whereas the other one must be a non-trivial matrix). But $G_j=1$ and $g_j=1$ for some $j$ would imply that $g=1$, so one of the two stabilisers must be non-trivial. \end{proof} Thus, $G<\operatorname{\rm SL}(2)\times\operatorname{\rm SL}(2)$ then $V/G$ not terminal implies (Lemma~\ref{lemma:terminal}) that one of $G_{\eta_i}$ must be non-trivial, say $G_\eta=G_{\eta_2}\neq 1$. Denoting by $W^1\to V^1/G_\eta$ the minimal resolution, one obtains a crepant blow-up $(W^1\times V^2)/(G/G_\eta)\to V/G$ with singularities of type $(2,1,1)$ or $(1,1,1,1)$. In order to prove that the orbifold Euler number remains unchanged under the blowup $(W^1\times V^2)/G\to V/G$, one must prove a formula similar to that of equation~\eqref{eq:chi-pi}, namely: $$\chi((W^{1})^{d}/Z_d)=|\pi_*^{-1}([d])|,$$ for all $d\in D:=G/G_{\eta}<\operatorname{\rm GL}(V^1)$. This could presumably be achieved by examining the finite $\operatorname{\rm SL}(2)$ and $\operatorname{\rm GL}(2)$ subgroups and determining how many fit into an exact sequence of the form given in equation~\eqref{eq:3-1-nocentre-exact}, with the cyclic group $C$ replaced by the $\operatorname{\rm GL}(2)$ subgroup $D$. \begin{question} What happens if $G$ is not a subgroup of $\operatorname{\rm SL}(2)\times\operatorname{\rm SL}(2)$? \end{question} \subsection{Groups of type $(2,1,1)$} \label{sec:2-1-1} Suppose the irreducible decomposition of ${\mathbb C}^4$ is $V^1\oplus V^2 \oplus V^3$ with $\dim V^1=2$ and $\dim V^i=2$ for $i=2,3$. Denote by $\eta_i$ the generic point of $V^i$, by $G_3$ the stabiliser of $\eta_3$ in $G$ and by $G_{32}$ the stabiliser of $\eta_2$ in $G_3$. Then we have exact sequences \begin{alignat}{3} \label{eq:2-1-1-exact} 1 &\to G_3 &\to G &\xrightarrow{\pi_3} C \to 1,\\ \intertext{and} \label{eq:2-1-1-exact2} 1 &\to G_{32} &\to G_3 &\xrightarrow{\pi_{32}} D \to 1, \end{alignat} with $C$ and $D$ cyclic subgroups, $G_{32}<\operatorname{\rm SL}(2)$ and $G_3<\operatorname{\rm SL}(3)$. The analytic germ of $V/G$ is isomorphic to $$(V^1\times V^2 \times V^3)/G \cong \bigl( (V^1/G_{32}\times V^2)/D\times V^3\bigr )/C.$$ The term $V^1/G_{32}$ has a minimal resolution $Z^1$ which is unique and therefore admits an action of $G_3$. The germ $(Z^1\times V^2)/D$ is a cyclic $\operatorname{\rm SL}(3)$ singularity and has an Euler resolution $W^1$. It remains to be shown that $W^1$ admits a $G$-action and that $\chi_{\text{orb}}((W^1\times V^3)/C)=\chi_{\text{orb}}(V/G)$. As in Section~\ref{sec:3-1}, the later statement would follow from the equalities $$\chi((W^1)^{c'}/C')=|\pi_{3*}^{\prime -1}([c'])|,$$ for $c'\in C'$, where the primed objects are defined similarly to those in Section~\ref{sec:4-centre:notation}. \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1996-10-09T23:07:16
9610
alg-geom/9610009
en
https://arxiv.org/abs/alg-geom/9610009
[ "alg-geom", "math.AC", "math.AG" ]
alg-geom/9610009
Qun Chen
Ragnar-Olaf Buchweitz and Qun Chen
Hilbert-Kunz functions of cubic curves and surfaces
LaTex 2e with Xy-pic v3.2 for commutative diagrams
null
null
null
null
We determine the Hilbert-Kunz function of plane elliptic curves in odd characteristic, as well as over arbitrary fields the generalized Hilbert-Kunz functions of nodal cubic curves. Together with results of K. Pardue and P. Monsky, this completes the list of Hilbert-Kunz functions of plane cubics. Combining these results with the calculation of the (generalized) Hilbert-Kunz function of Cayley's cubic surface, it follows that in each degree and over any field of positive characteristic there are curves resp. surfaces taking on the minimally possible Hilbert-Kunz multiplicity.
[ { "version": "v1", "created": "Wed, 9 Oct 1996 20:50:48 GMT" } ]
2008-02-03T00:00:00
[ [ "Buchweitz", "Ragnar-Olaf", "" ], [ "Chen", "Qun", "" ] ]
alg-geom
\section{Introduction} Let $S=k[x_0,\cdots,x_n]$ be the standard polynomial ring in $n+1$ variables over a field $k$ of prime characteristic $p$. Given a finite graded $S$-module $M$, the Hilbert-Kunz function of $M$ is defined on powers of the characteristic, $q=p^n$, $n\in {\Bbb{N}}$, through \[ HK_M (q):=\dim_k M/m^{[q]}M \] where $m^{[q]}=(x^q_0,\cdots,x^q_n)$ is the $q$-th Frobenius power of the maximal homogeneous ideal $m=(x_0, \cdots, x_n)$. If $I\subset S$ is a homogeneous ideal, and $R=S/I$ the homogeneous coordinate ring of the underlying projective scheme $X\subset {\Bbb{P}}^n_k$, the function $HK_R(q)$ is also called the Hilbert-Kunz function of $X$. Introduced by E. Kunz [5] in 1969, these functions were first studied in detail by P. Monsky [6], and he obtained the asymptotic formula \[ HK_M (q)=c q^m+ O(q^{m-1}) \] with $c\geq 1$ some real number and $m$ the Krull dimension of $M$. The number $c$ is called the Hilbert-Kunz multiplicity of $M$ and P. Monsky conjectures it to be rational. In general, it seems very difficult to determine these functions explicitly and a conceptual interpretation of the constant $c$ is missing (see [4] for some surprising examples). Here we exhibit the Hilbert-Kunz functions of plane elliptic curves in odd characteristic and of plane nodal cubics. Combining this work with results in [7] and [8] completes the explicit determination of Hilbert-Kunz functions of plane cubic curves. The Hilbert-Kunz functions of reducible cubics were already determined by K. Pardue in [8], and he also predicted the following list for the irreducible ones on the basis of computer experiments. Note that the Hilbert-Kunz function is invariant under extensions of the coefficient field $k$, so that one may assume $k$ algebraically closed. \newtheorem{theorem}{Theorem} \begin{theorem} Let $f$ be the equation of an irreducible cubic curve $C$ in ${\Bbb{P}}^2_k$ over an algebraically closed field $k$, and let $HK_R(q)$ be the Hilbert-Kunz function of the homogeneous coordinate ring $R=S/f$. \begin{description} \item[(1) (K. Pardue {[8]})] If $C$ is a cuspidal cubic, \[ HK_R (q)=\left\{ \begin{array}{ll} \frac{7}{3} q^2 &\text{for}\quad p=3\; , \\ \frac{7}{3} q^2-\frac{4}{3} & \text{for}\quad p\neq 3\; . \end{array} \right. \] \item[(2) (Thm. \ref{th:nodal} below)] If $C$ is a nodal cubic, \[ HK_R (q)=\left\{ \begin{array}{ll} \frac{7}{3}q^2-\frac{1}{3}q-1 & \text{for}\quad q\not\equiv 2\bmod 3\;, \\ \frac{7}{3}q^2-\frac{1}{3}q-\frac{5}{3} & \text{for}\quad q\equiv 2\bmod 3\;. \end{array} \right. \] \item[(3) (Thm. \ref{theorem:main} below)] If $C$ is an elliptic curve and $p\neq 2$, \[ HK_R(q)=\frac{9}{4} q^2-\frac{5}{4}\quad .\] \item[(4) (P. Monsky {[7]})] If $C$ is an elliptic curve and $p=2$, \[ HK_R (q)=\left\{ \begin{array}{ll} \frac{9}{4}q^2 & \text{if the $j$-invariant is $0$}\;, \\ \frac{9}{4}q^2-1 &\text{if the $j$-invariant is not $0$}\;. \end{array} \right. \] \end{description} \label{theorem:major} \end{theorem} At this stage, Hilbert-Kunz functions or multiplicities of plane curves of higher degree remain mysterious. However, a corollary of our work shows that for any $d\ge 2$ and for any field $k$ of prime characteristic there exists a plane curve of degree $d$ in ${\Bbb{P}}^2_k$ whose Hilbert-Kunz multiplicity is $\frac{3}{4} d$ --- and this is the minimal possible value for such curves. In particular, the minimal Hilbert-Kunz multiplicity in each degree is rational and independent of the characteristic. We then determine explicitly the Hilbert-Kunz function of Cayley's cubic surface in ${\Bbb{P}}^3_k$, and the result allows us to conclude as well that for any $d\ge 2$ and for any field $k$ of prime characteristic there exists a surface of degree $d$ in ${\Bbb{P}}^3_k$ whose Hilbert-Kunz multiplicity is $\frac{2}{3} d$ --- and this is again the minimal possible value, again rational and independent of the characteristic. \section{Minimal Values of Hilbert-Kunz Functions} Let $I=(f)$ be a principal ideal generated by a homogeneous form $f$ of degree $d > 0$ in $S$. The considerations in this section apply to the values of the {\em generalized\/} Hilbert-Kunz function of $R=S/I$, introduced by A. Conca in [1], and defined as \[ HK_{R,x}(q) =\dim_k S/(f,x_0^{q},\ldots,x_n^{q}) \quad, \] where $q$ is now {\em any\/} nonnegative integer, $k$ any field. Unless $k$ is of positive characteristic $p$, and $q$ is a power of $p$, this dimension will generally depend upon the choice of the coordinate system $x = (x_0,\ldots,x_n)$. For each $q\in{\Bbb N}$ and each choice of coordinates $x$, set $x^{[q]}=(x_0^{q},\ldots,x_n^{q})$ and consider the following graded $S$-modules of finite length, \begin{eqnarray*} \Theta & = & \frac{S}{x^{[q]}}=\bigoplus_i \Theta_i \quad , \\ \theta & = & \frac{S}{f+x^{[q]}}=\bigoplus_i \theta_i \quad , \\ \vartheta & = & \frac{(x^{[q]}:f)}{x^{[q]}}=\bigoplus_i \vartheta_i \quad . \end{eqnarray*} They are related by the exact sequence of graded $S$-modules \begin{equation} \begin{array}{cccccccccccc} 0 & \longrightarrow & \vartheta (-d)& \longrightarrow & \Theta (-d) & \stackrel{f}{\longrightarrow} & \Theta & \longrightarrow & \theta & \longrightarrow & 0 &, \end{array} \label{eq:exa} \end{equation} and $HK_{R,x}(q)=\dim_k \theta$. Evaluating dimensions yields universal bounds for the generalized Hilbert-Kunz function of $R=S/(f)$, when $f$ varies over polynomials of degree $d$ in $n+1$ variables, \begin{equation} q^{n+1}=\sum_i \dim_k \Theta_i \geq HK_{R,x}(q)\geq \sum_i \max \{ \dim_k \Theta_i-\dim_k \Theta_{i-d}, 0 \} \quad. \label{eq:ineq} \end{equation} The upper bound, $HK_{R,x}(q)=q^{n+1}$, is achieved iff $f\in x^{[q]}$; for example if $d > (n+1)(q-1)$, or if $q=p^n$ is a power of the characteristic, $d\geq q$ and $f=l^d$ for some linear form $l$. Here we are more concerned with the lower bound, that is taken on if and only if $f$ {\em is of maximal rank at\/} $q$, meaning that in each degree $i$ the $k$-linear map $f|\Theta_{i-d}$ is of maximal rank. Whether a given polynomial $f$ is of maximal rank at $q$, can be decided by looking at the socle degree of the artinian ring $\theta$, \begin{equation}a(q) =\max\{i:\theta_i\neq 0\}\quad, \label{eq:a} \end{equation} and at the initial degree of $\vartheta$, \begin{equation} \imath(q)=\min\{i:\vartheta_i \neq 0\}\quad. \label{eq:i} \end{equation} Indeed, as the socle degree of $\Theta$ is $(n+1)(q-1)$, outside the range $d\leq i\leq (n+1)(q-1)$ source or target of $f|\Theta_{i-d}$ is zero, whereas for a degree $i$ inside that range the map is not surjective iff $i \leq a(q)$, not injective iff $i-d \geq \imath(q)$. Accordingly, all the $k$-linear maps induced by $f$ are of maximal rank iff $a(q) < \imath(q) + d$. Moreover, the exact sequence (\ref{eq:exa}) is selfdual, whence it suffices to know either $a(q)$ or $\imath(q)$: \newtheorem{lemma}{Lemma} \begin{lemma} For each $q\in{\Bbb N}$, and independent of $f$, one has \begin{equation} a(q)+\imath(q)=(n+1)(q-1)\quad . \end{equation} Given $q$, all $k$-linear maps $f|\Theta_{i-d}$ are of maximal rank iff \begin{equation} a(q) < \frac{(n+1)(q-1)+d}{2} < \imath(q)+d\quad . \end{equation} Moreover, each of the inequalities implies the other. \label{lemma:mr} \end{lemma} {\noindent \it Proof:\quad } The ring $\Theta=S/x^{[q]}$ is a zerodimensional complete intersection with its socle in degree $(n+1)(q-1)$. Thus for any finite graded $\Theta$-module $M$, we have an isomorphism of graded $\Theta$-modules \[ {\rm Hom}_k (M,k) \cong {\rm Hom}_{\Theta} (M,\omega_{\Theta})\quad , \] where $\omega_{\Theta}=\Theta ((n+1)(q-1))$ is the canonical module of $\Theta$, and ${\rm Hom}_k (M,k)$ is the $\Theta$-module graded naturally through \[ ({\rm Hom}_k (M,k))_i = {\rm Hom}_k (M_{-i},k)\quad . \] As ${\rm Hom}_{\Theta}(\theta,\Theta) \cong (x^{[q]}:f)/x^{[q]}=\vartheta$, we get \[ {\rm Hom}_k (\theta,k)\cong \vartheta ((n+1)(q-1))\quad . \] For the dimension of the finite dimensional $k$-vector space $\theta_i$, this yields \[ \dim_k \theta_i =\dim_k {\rm Hom}_k (\theta_i,k) = \dim_k ({\rm Hom}_k (\theta,k))_{-i} = \dim_k \vartheta_{(n+1)(q-1)-i}\quad, \] and the equality follows from the definition of $a(q)$ and $\imath(q)$. As $f$ induces maps of maximal rank iff $a(q) < \imath(q) + d$, we can eliminate either one of the two invariants to obtain the last claim. $\hfill \square$ \[ \] If $d>(n+1)(q-1)$, the information is already complete: $f$, inducing the zero map in (\ref{eq:exa}), is trivially of maximal rank at $q$, and $HK_{R,x}(q) = q^{n+1}$. Also, if $f$ is a polynomial of a single variable, $n=0$, there are no secrets to discover. If $n>0$, the (usual) Hilbert series \[ H_{\Theta}(t) = \sum_i (\dim_k \Theta_i)t^i = (1+t+t^2+\cdots+t^{q-1})^{n+1}\quad,\] of the artinian $k$-algebra $\Theta$ is a {\em reciprocal\/} and {\em unimodal\/} polynomial of degree $l=(n+1)(q-1)$ in $t$, meaning that its coefficients, $\alpha_i=\dim_k \Theta_i$, satisfy \begin{eqnarray*} \alpha_i & = & \alpha_{l-i} \quad {\rm for\ every}\quad i \quad, \\ \alpha_i & < & \alpha_{i+1} \quad {\rm for} \quad 0\leq i <\left\lfloor \frac{l}{2} \right\rfloor\quad . \end{eqnarray*} In particular, $\dim_k\Theta_i - \dim_k\Theta_{i-d} > 0$ iff $0\leq i \leq m(q)$, where \[ m(q) =\left\lfloor \frac{(n+1)(q-1)+(d-1)}{2} \right\rfloor\quad,\] --- as for $a(q)$, we suppress the dependence upon $d$ from the notation. Thus the lower bound, $L(q)$, in inequality (\ref{eq:ineq}) evaluates to \begin{eqnarray*} L(q) & := & \sum_i \max\{ \dim_k \Theta_i-\dim_k \Theta_{i-d}, 0 \} \\ & = & \sum_{m(q)-d+1}^{m(q)} \dim_k \Theta_i \quad,\quad {\rm as}\ H_{\Theta}(t)\ {\rm is\ unimodal\ and\ reciprocal}\\ & = & {\rm\ coefficient\ of\ \;} t^{m(q)} {\rm\;\ in\ } \frac{(1-t^d)(1-t^q)^{n+1}}{(1-t)^{n+2}}\\ & = & \frac{1}{2\pi\sqrt{-1}}\int_{|z|=\epsilon} \frac{(1-z^d)(1-z^q)^{n+1}}{(1-z)^{n+2}z^{m(q)+1}}dz\quad. \end{eqnarray*} \noindent As $m(q)$ is the largest integer smaller than $\frac{1}{2}((n+1)(q-1)+d)$, we get the following result. \begin{theorem} If $n>0$, and if $f$ is a homogeneous polynomial of degree $d\leq (n+1)(q-1)$\; in \;$n+1$\; many variables, then the socle degree of the graded artinian $k$-algebra $\theta$ satisfies \begin{equation} a(q) \geq m(q)\quad . \label{eq:aq} \end{equation} Furthermore, the following statements are equivalent: \begin{description} \item[{\rm (i)}]\quad The polynomial $f$ is of maximal rank at $q$. \item[{\rm (ii)}]\quad The socle degree $a(q)$ is minimal, $a(q)=m(q)$. \item[{\rm (iii)}]\quad The initial degree $\imath(q)$ is maximal, $\imath(q)=(n+1)(q-1)-m(q)$. \item[{\rm (iv)}]\quad The Hilbert-Kunz function of $f$ at $q$ achieves the lower bound $L(q)$. \end{description} \label{theorem:minimal} \end{theorem} {\noindent \it Proof:\quad} The first statement follows from the exact sequence (\ref{eq:exa}), as \[ \dim_k \Theta_{m(q)} > \dim_k \Theta_{m(q)-d}\quad . \] As $a(q)$ is an integer and $m(q)$ is the largest integer smaller than $\frac{1}{2}((n+1)(q-1)+d)$, the just established lower bound for $a(q)$ implies the equivalences in view of Lemma \ref{lemma:mr}. $\hfill \square$ \[ \] {\noindent \it Example:\quad} For $d=2,3$ and $n=2,3$, we get the following table: \[ \begin{tabular}{|l|l|l|l|} \hline \emph{$d$} & \emph{$n$} & \emph{$m(q)$} & \emph{lower bound for $HK_{R,x} (q)$} \\ \hline 2 & 2 & $\left\lfloor \frac{3q}{2} \right\rfloor -1$ & $\left\lfloor \frac{3}{2} q^2 \right\rfloor =\left\{ \begin{array}{ll} \frac{3}{2} q^2 & \mbox{for $q$ even} \\ \frac{3}{2} q^2-\frac{1}{2} & \mbox{for $q$ odd} \end{array} \right.$ \\ \cline{2-4} & 3 & $2(q-1)$ & $\frac{4}{3} q^3-\frac{1}{3} q$ \\ \hline 3 & 2 & $\left\lfloor \frac{3q-1}{2} \right\rfloor$ & $\left\lfloor \frac{9}{4} q^2-\frac{5}{4} \right\rfloor =\left\{ \begin{array}{ll} \frac{9}{4} q^2-2 & \mbox{for $q$ even} \\ \frac{9}{4} q^2-\frac{5}{4} & \mbox{for $q$ odd} \end{array} \right.$ \\ \cline{2-4} & 3 & $2q-1$ & $2q^3-q$ \\ \hline \end{tabular} \] For $d=2$, it can be extracted from [1] that the quadric $x_0^2-x_1x_2$ for $n=2$, respectively the quadric $x_0x_1-x_2x_3$ for $n=3$, have generalized Hilbert-Kunz functions that take on the minimum value at each $q$. \[ \] {\noindent \it Remark 1:\quad} P. Monsky noted that expressing the minimal possible value $L(q)$ of $HK_{R,x}(q)$ as a residue leads to an intriguing lower bound for Hilbert-Kunz multiplicities in terms of the integrals \[ \beta_{n+1}=\frac{1}{\pi} \int^{+\infty}_{-\infty} (\frac{\sin \alpha}{\alpha})^{n+1} d\alpha = \frac{1}{2^{n}n!}\sum_{i=0}^{\left\lfloor n/2 \right\rfloor} (-1)^i(n+1-2i)^{n}\binom{n+1}{i}\quad , \] as \begin{equation} d\beta_{n+1} = \lim_{q\to\infty}\frac{1}{q^n{2\pi\sqrt{-1}}} \int_{|z|=\epsilon} \frac{(1-z^d)(1-z^q)^{n+1}}{(1-z)^{n+2}z^{m(q)+1}}dz\quad. \label{eq:beta} \end{equation} Thus, for a hypersurface of degree $d$ in ${\Bbb{P}}^n_k$ over a field $k$ of positive characteristic, the Hilbert-Kunz multiplicity satisfies \begin{equation} c \geq d\beta_{n+1}\quad. \label{eq:lower} \end{equation} A direct combinatorial proof is as follows. Expanding $(1-t^d)/(1-t)^{n+2}$ into its Taylor series at $t=0$, one can write \[ \frac{1-t^d}{(1-t)^{n+2}} = \sum_{\nu=0}^{d-n-1} R(\nu)t^{\nu} + \sum_{\nu \geq 0} P(\nu)t^{\nu}\quad,\] where $R(\nu)\in{\Bbb Z}$, and $P(\tau) = \frac{d}{n!}\tau^n + O(\tau^{n-1})$ is the corresponding (Hilbert) polynomial, univariate over ${\Bbb Q}$ of degree $n$ with leading coefficient $\frac{d}{n!}$. Now use that $m(q)/q = (n+1)/2 + O(1/q)$, that the coefficient of $t^{m(q)}$ in $\frac{1}{q^n}(1-t^q)^{n+1}\sum_{\nu=0}^{d-n-1} R(\nu)t^{\nu}$ tends to zero with $q$, and that \[ (1-t^q)^{n+1}\sum_{\nu \geq 0} P(\nu)t^{\nu} = \sum_{\nu\geq 0} \big(\sum_{i=0}^{\left\lfloor \nu/q \right\rfloor} (-1)^i P(\nu - iq)\binom{n+1}{i}\big)t^{\nu}\quad, \] to get \begin{eqnarray*} \lim_{q\to\infty} \frac{L(q)}{q^n} &=& \lim_{q\to\infty} \sum_{i=0}^{\left\lfloor m(q)/q\right\rfloor} (-1)^i\frac{P(m(q)-iq)}{q^n}\binom{n+1}{i}\\ &=&\lim_{q\to\infty} \sum_{i=0}^{\left\lfloor m(q)/q\right\rfloor} (-1)^i\frac{d}{n!}(\frac{m(q)}{q}-i)^n\binom{n+1}{i}\\ &=&\frac{d}{2^n n!} \sum_{i=0}^{\left\lfloor n/2\right\rfloor} (-1)^i(n+1-2i)^n\binom{n+1}{i}\\ &=&d\beta_{n+1}\quad. \end{eqnarray*} That this combinatorial expression equals the indicated integral can now be checked in any table of integrals, e.g. [3, $3.836.5^3$, p 458]. It follows that the sequence $\{ \beta_{n} \}$ of rational numbers decreases to zero. The first few values are \[ \beta_1=1, \; \;\beta_2=1, \; \; \beta_3=\frac{3}{4}, \; \; \beta_4=\frac{2}{3}, \; \; \beta_5=\frac{115}{192}, \; \; \beta_6=\frac{11}{20}\quad . \] {\noindent \it Remark 2:\quad} The same argument applies to $\liminf_{q\to\infty} HK_{R,x}(q)/q^n$, but it is not yet known whether a ``generalized'' Hilbert-Kunz multiplicity exists, i.e. whether $HK_{R,x}(q)/q^n$ tends to a limit for $q\to\infty$. \[ \] For elliptic curves in odd characteristic, we will prove that the corresponding cubic polynomial is of maximal rank at any power $q$ of the characteristic, whereas for the polynomial $x_0x_1x_2x_3( \frac{1}{x_0}+\frac{1}{x_1}+\frac{1}{x_2}+\frac{1}{x_3})$, representing Cayley's cubic surface, this will be even established at any $q\in{\Bbb N}$ over any field $k$. The proof is accomplished by showing that $a(q)$ equals the minimum value $m(q)$, and the respective Hilbert-Kunz function can then be read off from the table above. None of this applies to elliptic curves in characteristic $2$, (see [7] for details), nor does it hold for singular irreducible cubic curves in any characteristic. But in the latter case, the Hilbert-Kunz function can be determined completely from the rational parametrization of the curve as we show next. \section{Singular Irreducible Cubic Curves} The Hilbert-Kunz function of a cuspidal cubic is known from [8], see also [1], but our treatment here deals with the nodal and cuspidal case at the same time. Let $k$ be an algebraically closed field --- of any characteristic for now ---, and denote by $C$ a singular irreducible plane cubic curve in ${\Bbb{P}}^2_k$. In suitable coordinates, $C$ is given by a Weierstra\ss\ equation $$ f(x,y,z) = z(y^2 + a_1 xy - a_2 x^2) - x^3 = 0 \quad, $$ so that $o=[0,0,1]\in{\Bbb{P}}^2$ is its unique singular point. The curve has a node at $o$ iff the tangential quadric $Q(x,y) = y^2 + a_1 xy - a_2 x^2$ has distinct roots iff $a^2_1 + 4 a_2 \ne 0$, otherwise it is cuspidal. The curve $C$ is rational and a rational parametrization $\nu:{\Bbb{P}}^1\to C\subset {\Bbb{P}}^2$ normalizes the curve, pulling back ${\cal{O}}_{{\Bbb{P}}^2}(1)$ along $C\hookrightarrow{\Bbb{P}}^2$ and then $\nu$ to ${\cal{O}}_{{\Bbb{P}}^1}(3)$. Algebraically, such a parametrization is given by the monomorphism of $k$-algebras \[ \alpha(x,y,z)=(sQ(s,t),tQ(s,t),s^3)\quad , \] $$ \alpha: R ={ k[x,y,z]\over f(x,y,z)} \cong \bigoplus_{n\in {\Bbb{Z}}} H^0(C,{\cal{O}}_{C}(n))\hookrightarrow \bigoplus_{n\in {\Bbb{Z}}} H^0({\Bbb{P}}^1, {\cal{O}}_{{\Bbb{P}}^1}(3n)) \cong k[s,t]^{(3)} =:{\tilde R}\quad, $$ where ${\tilde R}=k[s,t]^{(3)}$ is the Veronese subring of the polynomial ring $k[s,t]$ spanned by all homogeneous polynomials whose degree is divisible by $3$. Notice that ${\tilde R}_n$ consists of all homogeneous polynomials of degree $3n$. The cokernel of $\alpha$ can be identified as follows. A section $p(s,t)\in H^0({\Bbb{P}}^1, {\cal{O}}_{{\Bbb{P}}^1}(3n))$ comes via $\alpha$ from a section in $H^0(C,{\cal{O}}_{C}(n))$ iff $p(s,t)$ takes on the same value at the two points $Q(s,t)=0$. Explicitly, write $p(s,t) = e_1(s) + e_2(s) t + e(s,t)Q(s,t)$ with uniquely determined polynomials $e_1,e_2\in k[s]$ and $e\in k[s,t]$. The component $e_2(s)t$ represents the class of $p(s,t)$ in $k[s,t]/(k[s] + Q k[s,t])$ and $(k[s,t]/(k[s] + Q k[s,t]))^{(3)}$ is the cokernel of $\alpha$. If $p(s,t)\in k[s,t]^{(3)}$, then $e_2(s) = b(s^3) s^2$ for some unique univariate polynomial $b$ that is necessarily of degree $\frac{1}{3}\deg p -1$. \begin{lemma} {\rm (i)}\quad The map $$ \beta : {\tilde R} = k[s,t]^{(3)}\to k[z](-1)\quad, $$ associating to $p(s,t)$ the polynomial $b(z)$, is a degree preserving epimorphism of $R$-modules, the $R$-module structure on ${\tilde R}$ given by $\alpha$, the one on $k[z]$ by the natural projection $R=k[x,y,z]/f\to k[z]$. \noindent {\rm (ii)}\quad The sequence of graded $R$-modules \begin{equation} \begin{array}{ccccccccc} 0 & \to & R & \stackrel{\alpha}{\longrightarrow} & {\tilde R} = k[s,t]^{(3)} & \stackrel{\beta}{\longrightarrow} & k[z](-1) & \to & 0 \end{array} \label{eq:exa2} \end{equation} is exact. \label{lemma:normalization} \end{lemma} {\noindent \it Proof:\quad } If $p(s,t) = e_1(s) + b(s^3)s^2 t + e(s,t)Q(s,t)$ is the unique representation of $p(s,t)\in {\tilde R}$, then \begin{eqnarray*} \alpha(x)p(s,t) &=& 0 + 0\cdot s^2t + (sp(s,t))Q(s,t) \quad, \\ \alpha(y)p(s,t) &=& 0 + 0\cdot s^2t + (tp(s,t))Q(s,t) \quad, \\ \alpha(z)p(s,t) &=& s^3e_1(s) + (s^3b(s^3))s^2 t + (s^3e(s,t))Q(s,t) \quad, \end{eqnarray*} are the corresponding unique representations of $\alpha(x)p(s,t), \alpha(y)p(s,t)$ and $\alpha(z)p(s,t)$ respectively. This shows that the image of $\beta$ is annihilated by $x,y$ and that $\beta(\alpha(z)p) = z\beta(p)$. Furthermore, $\beta(s^2t)$ generates the image of $\beta$ already as $k[z]$-module, thus a fortiori as $R$-module. As $s^2t$ is of degree one with respect to the grading on ${\tilde R} = k[s,t]^{(3)}$, (i) follows. For (ii), note first that $\beta\alpha(1) = 0$, whence $\beta\alpha = 0$. To prove that the kernel of $\beta$ is precisely the image of $\alpha$, consider Hilbert functions: In degree $i\in {\Bbb N}$, \[ \dim_k R_i =\left\{ \begin{array}{ll} 1 & for \quad i=0 \\ 3i & for \quad i>0 \end{array} \right. \quad, \quad {\rm whereas} \quad \dim_k {\tilde R}_i = 3i + 1\quad. \] Accordingly, the quotient ${\tilde R}_i/R_i$ is zero for $i=0$ and onedimensional for $i>0$. Thus the cokernel of $\alpha$ and $k[z](-1)$ have the same Hilbert function and (ii) follows. $\hfill\square$ \[ \] Multiplication with $x^q,y^q,z^q$ on (\ref{eq:exa2}) results in a commutative diagram of graded $R$-modules whose exact rows and columns define the modules $A$ through $G$, \[ \diagram & 0 \dto & 0 \dto & 0 \dto \\ 0 \rto & A \dto \rto^{\matrix \overline{\alpha} \endmatrix} & B \dto \rto^{\matrix \overline{\beta} \endmatrix} & C \dto \rto & D \rto & 0 \\ 0 \rto & R^{\oplus 3}(-q) \xto[ddd]_{\left( \matrix x^q \\ y^q \\ z^q \endmatrix \right)} \rto^-{\matrix \alpha^{\oplus 3} \endmatrix} & {\tilde R}^{\oplus 3}(-q) \xto[ddd]_{\left( \matrix \alpha(x)^q \\ \alpha(y)^q \\ \alpha(z)^q \endmatrix \right) } \rto^-{\matrix \beta^{\oplus 3} \endmatrix} & k[z]^{\oplus 3}(-1-q) \xto[ddd]_{\left( \matrix 0 \\ 0 \\ z^q \endmatrix \right)} \rto & 0 \\ &&&& \\ &&&& \\ 0 \rto & R \dto \rto^{\matrix \alpha \endmatrix} & {\tilde R} \dto \rto^{\matrix \beta \endmatrix} & k[z](-1) \dto \rto & 0 \\ & E \dto \rto & F \dto \rto & G \dto \rto & 0 \\ & 0 & 0 & 0 &&\quad. \enddiagram \] In this diagram, $E,F,G$, and then also $D$, are finite dimensional and one has \begin{equation} HK_{R,(x,y,z)}(q) = \dim_kE = \dim_k F - \dim_k G + \dim_k D\quad. \label{eq:sum} \end{equation} The dimension of $G\cong (k[z]/z^q)(-1)$ equals $q$, and the next Lemma determines the dimension of $F$, that is the value of the generalized Hilbert-Kunz function for ${\tilde R}$ with respect to $(x,y,z)$ at $q$. \begin{lemma} {\rm (i)}\quad Set $P=k[s,t]$, the polynomial ring in two variables with its natural grading. For any $q\in{\Bbb{N}}$, the $P$-module $M = k[s,t]/(\alpha(x)^q,\alpha(y)^q,\alpha(z)^q)$ has minimal graded resolution $$ \diagram 0 \rto & P(-4q)\oplus P(-5q) \xto[rrrr]^-{\left( \matrix t^q & -s^q & 0 \\ s^{2q} & 0 &-Q(s,t)^q \endmatrix \right) } & & & & P(-3q)^{\oplus3} \xto[rr]^-{\left( \matrix s^qQ(s,t)^q \\ t^qQ(s,t)^q \\ s^{3q} \endmatrix \right)} & & P \rto & M \rto & 0\ . \enddiagram $$ {\rm (ii)}\quad The middle column in the diagram above is obtained from that resolution by applying the functor $(\ \;)^{(3)}$, and in particular $F=M^{(3)}$.\\ \noindent {\rm (iii)}\quad \[ HK_{{\tilde R},(x,y,z)}(q) = \dim_k F = \left\{ \begin{array}{ll} {7\over 3} q^2 - {1\over 3} & \text{if}\quad q\not\equiv 0\bmod 3\quad, \\ {7\over 3} q^2 & \text{if} \quad q\equiv 0\bmod 3\quad. \end{array} \right. \] \label{lemma:F} \end{lemma} {\noindent \it Proof:\quad } (i)\quad As $s$ does not divide $Q(s,t)$, the module $M$ is artinian, and the result follows from the Hilbert-Burch theorem: the (signed) $(2\times 2)$-minors of the leftmost matrix are respectively $s^q Q(s,t)^q = \alpha(x)^q\; ,\; t^q Q(s,t)^q = \alpha(y)^q\;, \; s^{3q} =\alpha(z)^q\;$. \noindent (ii) is clear and (iii) follows then easily from \[ \dim_k F_i = \dim_k M_{3 i} = \dim_k P_{3 i} - 3 \dim_k P_{3 i - 3 q} + \dim_k P_{3 i - 4 q} + \dim_k P_{3 i - 5 q}\] and $\dim_k P_j = \max\{0,j+1\}$ for $j\in\Bbb Z$. $\hfill\square$ \[ \] In equation (\ref{eq:sum}) for $HK_{R,(x,y,z)}(q)$, it remains to determine the dimension of $D$. To this end, we exhibit the map ${\overline \beta}$ explicitly. As multiplication by $z^q$ is injective on $k[z]$, one has $C\cong k[z]^{\oplus 2}(-1-q)$. Furthermore, as $B \cong \big( k[s,t](-4q)\oplus k[s,t](-5q)\big)^{(3)}$ by Lemma \ref{lemma:F}, a homogeneous element in $B$ is represented by a pair $(p_1(s,t),p_2(s,t))$ of homogeneous polynomials satisfying \[ \deg p_1 = \deg p_2 + q \equiv -4q\bmod 3\quad,\] and such pair is mapped to \[ p_1(s,t)(t^q,-s^q,0) + p_2(s,t)(s^{2q},0,-Q(s,t)^q) \] in ${\tilde R}^{\oplus 3}(-q)$. Thus \[ {\overline \beta}(p_1,p_2) = (\beta(t^qp_1 + s^{2q}p_2),\beta(-s^q p_1)) \in k[z]^{\oplus 2}(-1-q)\cong C\quad. \] As the field $k$ is algebraically closed, the quadric $Q$ factors, $Q(s,t) = (t-us)(t-vs)\;;\; u,v\in k$; and the unique representation of $t^q \bmod Q(s,t)$ is \[ t^q = \tau_1 s^q + \tau_2 s^{q-1}t + \tau(s,t) Q(s,t)\quad, \] where \[ \tau_1 = -uv\sum_{i=0}^{q-2} u^{q-2-i}v^{i} \quad,\quad \tau_2 = \sum_{i=0}^{q-1} u^{q-1-i}v^{i}\quad. \] Writing now \[ p_1 = b_1(s) + b_2(s)t + b_3(s,t)Q(s,t)\quad,\quad p_2 = c_1(s) + c_2(s)t + c_3(s,t)Q(s,t)\quad, \] for suitable polynomials $b_i,c_i$, it follows that \[ {\overline \beta} (p_1,p_2) = \big(\tau_2 b_1 s^{q-3}+ (\tau_1 - a_1 \tau_2) b_2 s^{q-2} + c_2 s^{2q-2}, -b_2 s^{q-2}\big)\big|_{s^3=z}\quad. \] So the image of ${\overline \beta}$ in $C\cong k[z]^{\oplus 2}(-1-q)$ is generated as a $k[z]$-module by the three pairs \[ (\tau_2,0)z^{(q-3+\epsilon)/3}\quad,\quad (\tau_1 -a_1 \tau_2 ,-1)z^{(q-2+\eta)/3}\quad,\quad (1,0)z^{(2q-2+\zeta)/3}\quad, \] where $\epsilon,\eta,\zeta\in \{0,1,2\}$ are such that the exponents become integers. Accordingly, the dimension of $D$ is equal to $(2q-5 +\epsilon+\eta)/3$ if $\tau_2\ne 0$, whereas it equals $(3q-4+\eta+\zeta)/3$ if $\tau_2 = 0$. Thus, the decisive factor is whether or not $\tau_2$ vanishes. \begin{lemma} \begin{description} \item[(i)] If $q$ is a power of the characteristic of $k$, then $C$ is nodal over $k$ iff $\tau_2\ne 0$. \item[(ii)] If $\tau_2\ne 0$, then for any $q$ \[ \dim_k D =\left\{ \begin{array}{ll} 2\lfloor q/3\rfloor & \text{for} \quad q\not\equiv 0\bmod 3\;, \\ {2\over 3}q - 1 & \text{for} \quad q\equiv 0\bmod 3\;. \end{array} \right. \] \end{description} \end{lemma} {\noindent \it Proof:\quad } (i)\quad If $q$ is a power of the characteristic of $k$, then $\tau_2 = (u-v)^{q-1}$, whence $\tau_2 \ne 0$ iff $u\ne v$ iff $C$ is nodal. \smallskip\noindent (ii) just evaluates the formula for $\dim_k D$ found above in terms of $q\bmod 3$.$\hfill\square$ \[ \] Putting everything together yields the Hilbert-Kunz function in the nodal case. \begin{theorem} Let $C$ be a nodal cubic over a field $k$ of prime characteristic $p$. For a power $q$ of $p$, the Hilbert-Kunz function at $q$ is \[ HK_C(q) =\left\{ \begin{array}{ll} {7\over 3}q^2 - {1\over 3}q - 1 & \text{for} \quad q\not\equiv 2\bmod 3\; ,\\ {7\over 3}q^2 - {1\over 3}q - {5\over 3} & \text{for} \quad q\equiv 2\bmod 3\;. \end{array} \right. \] \label{th:nodal} \end{theorem} If $C$ is a cuspidal cubic, then $\tau_2=0$ for any $q$ and we get immediately the generalized Hilbert-Kunz function --- in accordance with [8] and [1]: \[ HK_{C,(x,y,z)}(q) =\left\{ \begin{array}{ll} {7\over 3}q^2 & \text{for} \quad q\equiv 0\bmod 3\;, \\ {7\over 3}q^2 - {4\over 3} & \text{for} \quad q\not\equiv 0\bmod 3\;. \end{array} \right. \] Note however that, if $q$ is not a power of the characteristic, this last result will in general depend upon the choice of the coordinate system made relative to the given Weierstra\ss\ form. The case of the generalized Hilbert-Kunz function for a nodal cubic can be extracted as well --- and the dependence upon the coordinate system becomes apparent: If the distinct roots $u,v$ satisfy $u^q-v^q = 0$ for some $q$, the generalized Hilbert-Kunz function ``jumps up'', it takes on the value from the cuspidal case. For any given $q$, we can avoid this situation by replacing $y$ with $y+\alpha x$, for a general $\alpha\in k$. The curve $C$ is then still in Weierstra\ss\ form, and with respect to $(x,y+\alpha x, z)$, the generalized Hilbert-Kunz function takes on the value predicted by Theorem \ref{th:nodal}. Unless the algebraically closed field $k$ is an algebraic closure of a finite field, one can even find an $\alpha\in k$ that works for all $q$ simultaneously. \section{Elliptic Curves in Odd Characteristic} In this section, we prove the announced result for elliptic curves in odd characteristic and deduce that the Hilbert-Kunz multiplicity of a generic plane curve equals $\frac{3}{4}d$ when $d\geq 2$. \begin{theorem} Let $f(x,y,z)\in S=k[x,y,z]$ be a cubic polynomial defining a plane elliptic curve over a field $k$ of odd characteristic $p$. For any $n\in {\Bbb{N}}$ and $q=p^n$, the socle degree $a(q)$ of $\theta = S/(f+m^{[q]})$ is minimal, \[a(q)=\frac {3}{2}q-\frac {1}{2}\quad,\] and the Hilbert-Kunz function of $R=S/(f)$ at $q$ is given by \[ HK_R (q)=\frac {9}{4}q^2 - \frac {5}{4}\quad .\] \label{theorem:main} \end{theorem} \newtheorem{corollary}{Corollary} \begin{corollary} For any field $k$ of prime characteristic $p$ and any integer $d\geq 2$, there is a curve $C\subset {\Bbb{P}}^2_k$ of degree $d$ whose Hilbert-Kunz multiplicity achieves the minimum $\frac{3}{4}d\;$. \end{corollary} {\noindent \it Proof:\quad } As shown in [8], the Hilbert-Kunz multiplicity of the quadric $g=x^2-yz$ equals $\frac{3}{2}$. For elliptic curves in any prime characteristic, Theorem \ref{theorem:major} shows that their Hilbert-Kunz multiplicities are minimal, equal to $\frac{9}{4}$. As any integer $d\geq 2$ can be written $d=2u+3v$ for some $u,v\in {\Bbb{N}}$, additivity of the Hilbert-Kunz multiplicity, see [6], implies that the curve of degree $d$, defined by $h=g^uf^v$, $f$ a nonsingular cubic, will achieve the minimum. $\hfill \square$ \[ \] {\noindent \it Remark 3. } Semi-continuity of the Hilbert-Kunz multiplicity yields that the Hilbert-Kunz multiplicity of a generic plane curve of degree $d\geq 2$ equals $\frac{3}{4}d$. Clearly $c=1$ if the degree $d=1$. So the Hilbert-Kunz multiplicity of a generic curve is rational and independent of the (positive) characteristic. \[ \] In section 4.1, we recall a classical result about determinants of Hankel matrices whose entries are Legendre polynomials, and in section 4.2, we use it to determine the invariant $a(q)$ and to establish Theorem \ref{theorem:main}. \subsection {Hankel Determinants of Legendre Polynomials} The Hankel matrices associated to a sequence $a=\{ a_i \}$ are \[ H^{(n)}_k (a) = \left( \begin{array}{llll} a_n & a_{n+1} & \cdots & a_{n+k-1} \\ a_{n+1} & a_{n+2} & \cdots & a_{n+k} \\ \cdots & \cdots & \cdots & \cdots \\ a_{n+k-1} & \cdots & \cdots & a_{n+2(k-1)} \end{array} \right)\quad, \] with corresponding Hankel determinants \[ D^{(n)}_k (a) = \det H^{(n)}_k (a) \quad . \] The generating function for the Legendre polynomials, $\{P_n (t)\}_{n\in\Bbb N},$ is \[ F(t,x)=\frac{1}{\sqrt{1-2tx+x^2}} = \sum P_n (t) x^n \in{\Bbb Z}[\tfrac{1}{2},t][\![ x]\!] \quad .\] For each $k$, consider the determinant of the following Hankel matrix whose entries are Legendre polynomials, \[ D^{(0)}_k (P_i (t)) =\det \left( \begin{array}{llll} P_0 (t) & P_1 (t) & \cdots & P_{k-1} (t) \\ P_1 (t) & P_2 (t) & \cdots & P_k (t) \\ \cdots & \cdots & \cdots & \cdots \\ P_{k-1} (t) & \cdots & \cdots & P_{2k-2} (t) \end{array} \right)\quad. \] In [2], J. Geronimus gave the following beautiful formula. \begin{theorem} \[ D^{(0)}_k (P_i (t)) = 2^{- (k-1)^2} (t^2-1)^{\frac{1}{2} (k-1)k} \quad . \] \label{theorem:G} \end{theorem} P. Monsky communicated a direct proof to us that we now present. \begin{lemma} If $t\in {\Bbb{R}}$ and $t>1$, then \[ P_n (t)=\frac{1}{\pi} \int^{\pi}_0 [t+\sqrt{t^2-1} \cos\alpha]^n d \alpha . \] \end{lemma} {\noindent \it Proof:\quad} If $x$ is small, \[ (1-2tx+x^2)^{-\frac{1}{2}} =[(1-tx)^2-(x\sqrt{t^2-1})^2]^{-\frac{1}{2}} \] \[=\frac{1}{\pi} \int^{\pi}_0 \frac{d\alpha}{(1-tx)-x\sqrt{t^2-1} \cos\alpha } =\frac{1}{\pi} \int^{\pi}_0 \frac{d\alpha} {1-x[t+\sqrt{t^2-1}\cos\alpha ]}\quad . \] The expansion of the integrand into a power series of $x$ yields the lemma. $\hfill \square$ \[ \] {\noindent \it Proof of Theorem \ref{theorem:G} (P. Monsky):\quad} As both sides of Geronimus' formula are polynomials in $t$, it suffices to prove it for $t\in {\Bbb{R}}$ and $t>1$. Let $V$ be the vector space of real valued continuous functions on $[0,\pi]$. Define a symmetric bilinear form $(\cdot,\cdot)$ on $V$ through \[ (f,g)=\frac{1}{\pi} \int^{\pi}_0 f(\alpha)g(\alpha) d\alpha \quad . \] If $h_1,\ldots,h_s \in V$, let $\Delta (h_1,\ldots,h_s)$ be the determinant of the matrix $((h_i,h_j))$. Set $g_m=[t+\sqrt{t^2-1} \cos\alpha]^m$. By the preceding lemma, $(g_i,g_j) = P_{i+j}$. So the required Hankel determinant is $\Delta (g_0,\ldots,g_{k-1})$. Let $V_m\subset V$ be the subspace spanned by $\{1,\cos\alpha,\ldots,(\cos\alpha)^m\}$. Then $f_m=(t^2-1)^{\frac{m}{2}} \cos (m\alpha) \in V_m$, and it is a linear combination of $g_0,\ldots,g_m$. Furthermore, modulo $V_{m-1}$, $\cos(m\alpha) \equiv 2^{m-1} (\cos\alpha)^m$, and consequently $f_m \equiv 2^{m-1} (t^2-1)^{\frac{m}{2}} (\cos \alpha)^m \equiv 2^{m-1} g_m$. We conclude that \[ \Delta (f_0,\cdots,f_{k-1})=(\prod^{k-1}_{m=1} 2^{m-1})^2 \Delta (g_0,\cdots,g_{k-1})=2^{(k-2)(k-1)}\Delta (g_0,\cdots,g_{k-1}).\] But using the orthogonality of the $f_i$ one finds that \[ \Delta (f_0,\cdots,f_{k-1})=2^{-(k-1)}(t^2-1)^{\frac{(k-1)k}{2}} \] and Theorem \ref{theorem:G} follows. $\hfill \square$ \[ \] Now consider \[ G(t,x)=\sqrt{1-2tx+x^2}=\sum \widetilde{P_n (t)} x^n\quad. \] As $F(t,x) G(t,x) =1$, Geronimus' formula yields also the following corollary. \begin{corollary} \[ D^{(2)}_k (\widetilde{P_i (t)} )= (-1)^k D^{(0)}_{k+1} (P_i (t)) = (-2)^{-k^2} (t^2-1)^{\frac{1}{2} k(k+1)} \quad . \] \end{corollary} {\noindent \it Remark 4. } The coefficients of Legendre polynomials are rational numbers whose denominators are powers of $2$. Thus Geronimus' identity and the above corollary hold over any ring in which $2$ is a unit, in particular over a field of odd characteristic. \subsection{The Invariant $a(q)$ } We first prove the following theorem showing that in odd characteristic there are no nontrivial syzygies of low degree between the equation of an elliptic curve and Frobenius powers of the variables. \begin{theorem} Let $k$ be a field of odd characteristic $p$, and let $f\in k[x,y]$ be a cubic polynomial defining an elliptic curve in ${\Bbb{A}}^2_k$. For any $q=p^n$, with $n\in {\Bbb{N}}$, if $f|u x^q+v y^q+w$ for $u,v,w\in k[x,y]$ of degree at most $\frac{1}{2}(q-1)$, then $f$ divides each of $u,v,w$. \label{theorem:second} \end{theorem} {\noindent \it Proof:\quad} We give the proof for $q\equiv 1\bmod 4$. The argument in the other case, $q\equiv 3\bmod 4$, is analogous and left to the reader. Without loss of generality assume that $k$ is algebraically closed. Since the result is invariant under the action of $GL(2, k)$, and the characteristic of $k$ is odd, we can put the cubic into the form $f=y^2-x(1-2tx+x^2)$ with $t^2\neq 1$. If $f | u x^q + v y^q + w $ for some $u , v , w $ of degree at most $\frac{1}{2} (q-1) $, then \begin{equation} u x^q + v y^q + w = f h \quad , \label{eq:main} \end{equation} where $h$ is a polynomial in $x,y$. Set \begin{eqnarray*} l & = & \frac{q-1}{2}\quad, \\ g & = & x(1-2tx+x^2)\quad . \end{eqnarray*} We can then write \[ u=\sum^l_{i=0} a_i y^i =A_0 + yA_1 + fu_1\quad, \] with $a_i\in k[x]$ of degree at most $l-i, u_1 \in k[x,y]$ and \[ A_0=\sum^{\frac{l}{2}}_{j=0} a_{2j}g^j\quad,\quad A_1=\sum^{\frac{l}{2}-1}_{j=0} a_{2j+1}g^j\quad, \] polynomials in $x$. Similarly, we write \begin{eqnarray*} v & = & B_0 +yB_1+fv_1 \\ w & = & C_0 +yC_1+fw_1 \\ y^q & = & y^{2l+1}=yg^l+ f\gamma \end{eqnarray*} for polynomials $v_1, w_1, \gamma\in k[x,y]; B_0,B_1,C_0,C_1\in k[x]$. Equation (\ref{eq:main}) then becomes \[ (x^qA_0+C_0+g^{l+1}B_1)+y(x^qA_1+C_1+g^lB_0)=fh_1 \quad .\] Viewing both sides as polynomials in $y$, we get $h_1=0$ and \begin{equation} x^qA_0+C_0+g^{l+1}B_1 = 0\quad , \label{eq:1} \end{equation} \begin{equation} x^qA_1+C_1+g^lB_0 = 0\quad . \label{eq:2} \end{equation} As $\deg C_0\leq \deg x^q A_0 \leq \frac{7}{2} l + 1$, it follows that \[ \deg B_1 \leq (\frac{7}{2} l + 1) - 3 (l + 1) = \frac{l}{2} -2\quad , \] and similarly $\deg B_0 \leq \frac{l}{2} -1$. Thus we can write \begin{eqnarray*} B_1 & = & \alpha_{\frac{l}{2}-2} x^{\frac{l}{2}-2} + \cdots + \alpha_0 \quad, \\ B_0 & = & \beta_{\frac{l}{2}-1} x^{\frac{l}{2}-1} + \cdots + \beta_0\quad, \end{eqnarray*} for tuples \[ \alpha = ( \alpha_i ) \in k^{\frac{l}{2}-1} \quad , \quad \beta= (\beta_i) \in k^{\frac{l}{2}} \quad . \] Since $\deg C_0 \leq \frac{3}{2} l $ and ${\rm ord}(x^q A_0) \geq q=2l+1 $, the intermediate powers of $x$ in $g^{l+1} B_1 $ have zero coefficients, whence we get a linear system of equations for $\alpha$, say $E \alpha =0$, where $E : k^{\frac{l}{2}-1} \longrightarrow k^{\frac{l}{2}}$ is represented by the matrix \[ E=\left( \begin{array}{llll} e_2 & e_3 & \cdots & e_{\frac{l}{2}} \\ e_3 & e_4 & \cdots & e_{\frac{l}{2}+1} \\ e_4 & \cdots & \cdots & e_{\frac{l+2}{2}} \\ \cdots & \cdots & \cdots & \cdots \\ e_{\frac{l}{2}+1} & \cdots & \cdots & e_{l-1} \end{array} \right)\quad, \] whose entries $e_i$ are the coefficients in the expansion \[ (1-2tx+x^2)^{\frac{q+1}{2}} = x^{q+1} + e_q x^q +\cdots + e_0 \quad . \] Analogously, the corresponding powers of $x$ in $g^l B_0$ yield a system of equations for $\beta$, say $H \beta = 0$, where $H : k^{\frac{l}{2}} \longrightarrow k^{\frac{l}{2}+2}$ is represented by the matrix \[ H=\left( \begin{array}{llll} h_0 & h_1 & \cdots & h_{\frac{l}{2}-1} \\ h_1 & h_2 & \cdots & h_{\frac{l}{2}} \\ \cdots & \cdots & \cdots & \cdots \\ h_{\frac{l}{2}+1} & \cdots & \cdots & h_l \end{array} \right)\quad, \] whose entries $h_i$ are the coefficients in the expansion \[ (1-2tx+x^2)^{\frac{q-1}{2}}= x^{q-1} + h_{q-2} x^{q-2} + \cdots + h_0\quad . \] As $q$ is a power of the characteristic $p$, one has \[(1-2tx+x^2)^{\frac{q-1}{2}} \equiv \frac{(1 - (2tx)^q + x^{2q})^{\frac{1}{2}}} {\sqrt{1-2tx+x^2}} \equiv (1 - (2tx)^q + x^{2q})^{\frac{1}{2}} \sum P_n (t) x^n \bmod p\quad,\] whence $h_i \equiv P_i (t)\bmod p$ and, analogously, $e_i \equiv \widetilde{P_i (t)}\bmod p$ for $i<q$, where $P_i (t)$ and $\widetilde{P_i (t)}$ are as in section 4.1. As $t^2\neq 1$ by assumption, Geronimus' Theorem and its Corollary imply \[ {\rm rank} \; E = \frac{l}{2} -1 \quad,\quad {\rm rank} \; H = \frac{l}{2}\quad , \] whence each $\alpha_i$ or $\beta_j$ equals zero, thus $B_0=B_1=0$, so that $v=fv_1$. As $\deg C_i <{\rm ord} (x^qA_i)$; for $i=0,1$; it follows further from equations (\ref{eq:1}) and (\ref{eq:2}) that $C_i=A_i=0$ and the theorem follows. $\hfill \square$ \[ \] Now we can finish the {\it Proof of Theorem \ref{theorem:main}:} As $d=3<3(q-1)$, for any power $q=p^n; n\in {\Bbb{N}};$ of an odd prime $p$, Theorem \ref{theorem:minimal} yields the lower bound $a(q) \geq \frac{3}{2}q-\frac{1}{2}$, and the upper bound $\imath (q)\leq \frac{3}{2} q-\frac{5}{2}$. It remains thus to show $\vartheta_{\frac{3}{2} q-\frac{7}{2} }=0$, or, equivalently, if $f | u x^q+v y^q + w z^q $ for $u ,v ,w \in k[x,y,z]_{\frac{1}{2} (q-1) }$, then $f | u ,v , w $. As it suffices to verify the above statement in the affine part $(z=1)$ of ${\Bbb{P}}^2_k $, the result in Theorem \ref{theorem:second} finishes the proof. $\hfill \square$ \section{Cayley's Cubic Surface} Let $S=k[x,y,z,w]$ be the polynomial ring in four variables over an arbitrary field $k$ and let $f=xyz+xyw+xzw+yzw$ be the Cayley cubic. We consider the generalized Hilbert-Kunz function of $R=S/f$, given at $q\in\Bbb N$ through \[ HK_{R,(x,y,z,w)} (q)=\dim_k S/(f,x^q,y^q,z^q,w^q) \quad. \] \begin{theorem} The socle degree of the artinian ring $\theta= S/(f,x^q,y^q,z^q,w^q)$ is \[ a(q)=\left\{ \begin{array}{ll} 0 & {\text if} \quad q=1\;, \\ 2n-1 & {\text if} \quad q>1\;, \end{array} \right. \] and the value of the generalized Hilbert-Kunz function of Cayley's cubic at $q\in\Bbb N$ is \[ HK_{R,(x,y,z,w)} (q)= 2q^3 - q\quad.\] \label{theorem:third} \end{theorem} {\noindent \it Proof:\quad } If $q=1$, then $\theta \cong k$ and $a(1)=0$. Now assume $q>1$. Since $d=3<4(q-1)$, Theorem \ref{theorem:minimal} yields the lower bound $a(q)\geq 2q-1$. Thus it remains to show $\theta_{2q}=0$, i.e, that any monomial $x^iy^jz^kw^l\in \theta_{2q}$ is equivalent to $0$. We prove this by ``descent'' on the {\it dominant exponent\/}, $e=\max\{i,j,k,l\}$, of a monomial $x^iy^jz^kw^l$ in $S_{2q}$. Any monomial with sufficiently large dominant exponent --- for example, when it exceeds $q$ --- will be equivalent to $0$. We may thus assume that for a fixed integer $e$, any monomial whose dominant exponent exceeds $e$ is equivalent to $0$, to consider then monomials with dominant exponent equal to $e$. Due to symmetry, it suffices to consider monomials $x^iy^jz^kw^l$ with $e=i\geq j \geq k \geq l$. {\it Case 1: \quad} If $l>0$, then \begin{eqnarray*} x^iy^jz^kw^l & \equiv & x^iy^{j-l}z^{k-l} \cdot (-1)^l x^l (yz+yw+zw)^l \bmod f\\ & \equiv & (-1)^lx^{i+l} y^{j-l}z^{k-l}(yz+yw+zw)^l \bmod f\\ & \equiv & 0\bmod (f,x^q,y^q,z^q,w^q) \end{eqnarray*} by assumption. {\it Case 2: \quad} Now suppose $l=0$. If $k\leq 1$, then $i+j+1\geq 2q$ with $i\geq j$, so that $i\geq q$ whence the monomial is obviously equivalent to $0$. If $k>1$, consider first the monomial $g=x^{i-k+1}y^jz^kw^{k-1}$ and set $h=\min\{ i-k+1,k-1 \}$. Since $j>i-k+1>0$ and $j\geq k>k-1>0$, the argument employed above shows that $g$ is equivalent to a linear combination of monomials whose dominant exponent is $j+h$. As $j+h>i$, our assumption insures that $g$ is equivalent to $0$. On the other hand, \[ g=x^{i-k+1}y^jz^kw^{k-1} \equiv (-1)^{k-1} x^i y^{j-k+1} z (yz+yw+zw)^{k-1} \bmod f\quad ,\] and in the expansion of the right hand side of this congruence, each term involving $w$ involves all four variables, has dominant exponent $e=i$, and is thus equivalent to $0$ by Case 1. So \[ g=x^{i-k+1}y^jz^kw^{k-1} \equiv (-1)^{k-1} x^i y^{j-k+1}z (yz)^{k-1} \equiv (-1)^{k-1} x^i y^j z^k\quad,\] whence $x^iy^jz^k\equiv 0$. The claimed result for the generalized Hilbert-Kunz function follows now from Theorem \ref{theorem:minimal}, as $d=3 < 4(q-1)$ whenever $q >1$, and its validity for $q=0,1$ is clear. $\hfill \square$ \[ \] \begin{corollary} For any field $k$ of prime characteristic $p$ and any integer $d\geq 2$, there is a surface $X\subset {\Bbb{P}}^3_k$ of degree $d$ whose Hilbert-Kunz multiplicity achieves $\frac{2}{3}d$, the minimum possible for such surfaces. \end{corollary} {\noindent \it Proof:\quad} The Hilbert-Kunz multiplicity of the quadric surface $g=xy-zw$ equals $\frac{4}{3}$ by [1]. As just established, the Hilbert-Kunz multiplicity of the Cayley cubic $f$ is equal to $2$. Since $d=2u+3v$ for some $u,v\in {\Bbb{N}}$, additivity of the Hilbert-Kunz multiplicity implies that the surface defined by $g^uf^v$ has Hilbert-Kunz multiplicity equal to $\frac{2}{3}d$. $\hfill \square$ \[ \] {\noindent \it Remark 5. } For any field $k$ of positive characteristic, by virtue of the above corollary and semi-continuity, a generic surface in ${\Bbb{P}}^3_k$ of degree $d\geq 2$ achieves the minimal Hilbert-Kunz multiplicity $\frac{2}{3}d$. Also, $c=1$ if $d=1$. So the Hilbert-Kunz multiplicity of a generic surface is rational and independent of the characteristic. Note that the Hilbert-Kunz multiplicity of Cayley's cubic is minimal although this surface is singular --- in contrast to the case of cubic curves. \[ \] {\noindent \it Acknowledgments.\quad} The authors would like to thank P. Monsky, whose influence on this paper should be clear to every reader. We are especially grateful for his permission to include the direct proof of Geronimus' theorem and for modifications that lead to a shorter proof of Theorem \ref{theorem:second}. We also thank K.~Pardue, who got us interested in the subject, and A.~Conca for useful conversations.
1996-10-10T01:44:46
9610
alg-geom/9610010
en
https://arxiv.org/abs/alg-geom/9610010
[ "alg-geom", "math.AG" ]
alg-geom/9610010
Misha S. Verbitsky
Misha Verbitsky
Deformations of trianalytic subvarieties of hyperk\"ahler manifolds
51 pages, LaTeX2e
Selecta Math. (N.S.) 4 (1998), no. 3, 447--490.
null
null
null
Let $M$ be a compact complex manifold equipped with a hyperk\"ahler metric, and $X$ be a closed complex analytic subvariety of $M$. In alg-geom/9403006, we proved that $X$ is trianalytic, i. e., complex analytic with respect to all complex structures induced by the hyperk\"ahler structure, provided that $M$ is generic in its deformation class. Here we study the complex analytic deformations of trianalytic subvarieties. We prove that all deformations of $X$ are trianalytic and naturally isomorphic to $X$ as complex analytic varieties. We show that this isomorphism is compatible with the metric induced from $M$. Also, we prove that the Douady space of complex analytic deformations of $X$ in $M$ is equipped with a natural hyperk\"ahler structure.
[ { "version": "v1", "created": "Wed, 9 Oct 1996 23:39:04 GMT" } ]
2008-02-03T00:00:00
[ [ "Verbitsky", "Misha", "" ] ]
alg-geom
\section{Introduction.} \label{_Intro_Section_} \subsection{An overview} This is the third article studying closed complex analytic subvarieties of compact holomorphically symplectic\footnote{Holomorphically symplectic means ``equipped with a holomorphic symplectic form''. See \ref{_holomorphi_symple_Definition_} for details.} K\"ahler manifolds. In the first article in series (\cite{Verbitsky:Symplectic_I_}), we proved that, when a holomorphically symplectic manifold $M$ is generic in its deformation class, all subvarieties $X\subset M$ are also holomorphically symplectic, i. e. restriction of holomorphic symplectic form to nonsingular strata of $X$ is non-degenerate. In the second article \cite{Verbitsky:Symplectic_II_}, we obtained a more precise result about the structure of such $X$, which is related to the hyperk\"ahler structure of $M$. By Yau's proof of Calabi conjecture, $M$ admits a natural hyperk\"ahler metric (\ref{_symplectic_=>_hyperkahler_Proposition_}; for a definition of hyperk\"ahler manifold, see \ref{_hyperkahler_manifold_Definition_}). A hyperk\"ahler structure (which is essentially a quaternion action in the tangent bundle to $M$) gives a rise to a whole family of complex structure on $M$, parametrized by ${\Bbb C} P^1$. These complex structures are called {\bf complex structures induced by the hyperk\"ahler structure} (\ref{_indu_comple_str_Definition_}). Denote the set of all induced complex structures by $\c R_M$. A closed subset $X\subset M$ is called {\bf trianalytic} if $X$ is complex analytic with respect to all induced complex structures $L \in \c R_M$ (\ref{_trianalytic_Definition_}). For an induced complex structure $L$, we denote by $(M, L)$ the $M$ considered as a complex manifold, with complex structure $L$. In \cite{Verbitsky:Symplectic_II_}, we proved that for all $L\in \c R_M$, with exception of may be a countable set, all complex analytic subsets of $(M, L)$ are trianalytic (\ref{_hyperkae_embeddings_Corollary_}). Unlike the second article \cite{Verbitsky:Symplectic_II_}, which supersedes results of the first \cite{Verbitsky:Symplectic_I_}, the present one (the third) elaborates on the results of the second. We start where \cite{Verbitsky:Symplectic_II_} left. Consider a complex subvariety $X \subset (M, L)$ which happens to be trianalytic. Such subvarieties are called {\bf complex analytic subvarieties of trianalytic type}. We describe the deformations of $X$ in $(M, L)$ and the Douady space of such deformations.\footnote{The Douady space \cite{_Douady_} is the same as Chow scheme, but in complex analytic situation.} In \cite{Verbitsky:Symplectic_II_}, we gave a simple cohomological criterion of trianaliticity, for arbitrary complex analytic subvariety $X \subset (M, L)$ (see \ref{_G_M_invariant_implies_trianalytic_Theorem_} of this article). \ref{_G_M_invariant_implies_trianalytic_Theorem_} immediately implies that a complex analytic deformation of a trianalytic subvariety is again trianalytic. The main result of this article is the following theorem. \hfill \theorem \label{_iso_intro:Theorem_} Let $M$ be a hyperk\"ahler manifold, $L$ an induced complex structure, and $X, X'\subset (M, L)$ be closed complex analytic subvarieties in the same deformation class. Assume that $X$ is trianalytic.\footnote{Then the subvariety $X'$ is also trianalytic, as implied by \ref{_G_M_invariant_implies_trianalytic_Theorem_}.} Consider $X, X'$ as K\"ahler subvarieties of $(M, L)$, with K\"ahler metric induced from $M$. Then \begin{description} \item[(i)] $X, X'$ are naturally isomorphic as complex varieties, and this isomorphism is compatible with the K\"ahler metric. \item[(ii)] The embedding $X\stackrel i \hookrightarrow M$ is compatible with metrics given by geodesics. \item[(iii)] Also, every map $X\stackrel \phi{\:\longrightarrow\:} (M, L)$ is complex analytic, provided that $\phi$ is a deformation of $i$ in the space of isometries $X{\:\longrightarrow\:} M$. \end{description} \hfill {\bf Caution} The varieties $X, X'$ need not to be non-singular. \hfill \ref{_iso_intro:Theorem_} (i) follows from \ref{_triana_subse_isome_Theorem_} and \ref{_triana_subse_comple_ana_Theorem_}. \ref{_iso_intro:Theorem_} (ii) follows from \ref{_hype_embe_comple_geode:Corollary_}, and (iii) from \ref{_isome_embe_Proposition_}. \hfill As a corollary, we obtain the following interesting result (\ref{_Doua_hyperka_Theorem_}). \hfill \theorem \label{_defo_intro:Theorem_} Let $M$ be a compact holomorphically symplectic K\"ahler manifold, and $X\subset M$ a closed complex analytic subvariety. Consider the Douady space $D_M(X)$ of deformations of $X$ inside $M$. Then $D_M(X)$ is compact and is equipped with a natural hyperk\"ahler structure, in the sense of \cite{_Verbitsky:Hyperholo_bundles_},\footnote{If $D_M(X)$ is non-singular, this means exactly that $D_M(X)$ is hyperk\"ahler. In singular case, there is no satisfactory definition of a hyperk\"ahler structure. The definition of \cite{_Verbitsky:Hyperholo_bundles_} is a palliative, which is probably much stronger that the correct definition. See \cite{_Simpson:hyperka-defi_}, \cite{_Deligne:defi_} for alternative definitions.} (see \ref{_singu_hype_Definition_}). \hfill \remark This gives a new set of examples of hyperk\"ahler varieties, in addition to those produced by \cite{_Verbitsky:Hyperholo_bundles_}. The varieties obtained through \cite{_Verbitsky:Hyperholo_bundles_} are usually not compact; on contrary, our new examples are compact. \hfill The proof of \ref{_iso_intro:Theorem_} is based on the following argument. First, we show that every trianalytic submanifold $X\subset M$ (not necessarily closed) is {\bf completely geodesic} in $M$ (\ref{_hype_embe_comple_geode:Corollary_}). \footnote{{\bf Completely geodesic} means that geodesics in $X$ are also geodesics in $M$.} A simple geometric argument shows that a family $\c X$ of completely geodesic submanifolds admits a natural {\bf connection} (\ref{_conne_Definition_}, \ref{_conne_in_fam_of_comple_geode_Proposition_}). This connection is compatible with the holomorphic structure if the family $\c X$ is holomorphic (\ref{_conne_in_fam_of_comple_geode_Proposition_}). \hfill For $M$ a compact hyperk\"ahler variety, $I$ an induced complex structure and $X$ a trianalytic subvariety of $M$, we study the Douady space $D(X,I)$ which classifies complex analytic deformations of $X$ in $(M,I)$, where $(M, I)$ is $M$ considered as a complex manifold. We prove that the real analytic subvariety $D(X)$ underlying $D(X,I)$ does not depend from the choice of induced complex structure. This endows $D(X)$ with a 2-dimensional sphere of complex structures, which induce quaternionic action in the Zariski tangent space. A real analytic variety with such a system of complex structures is called {\bf hypercomplex} (\ref{_hypercomplex_Definition_}). We return to the study of the families of trianalytic subvarieties. As we have shown, the base of the universal family is hypercomplex. The fibers are trianalytic, and therefore, hypercomplex as well. The natural connection in such a family is compatible with the quaternionic action, because this connection is {\bf holomorphic} with respect to each of induced complex structures. Thus, a curvature of this connection is ${\Bbb H}^*$-invariant (\ref{_curva_SU(2)_inva_Lemma_}). On the other hand, the curvature lies in the tensor product of three representations of ${\Bbb H}^*$ of weight one. A trivial linear algebra argument shows that such a tensor product does not contain non-trivial ${\Bbb H}^*$-invariant vectors. This proves that the natural connection in the family of trianalytic subvarieties is {\bf flat} (\ref{_conne_in_triana_flat_Theorem_}). Let $\c X\stackrel \pi {\:\longrightarrow\:} S$ be a family of closed trianalytic subvarieties in a compact hyperk\"ahler manifold $M$. Let $X_s^{ns}$ be the set of non-singular points in $X_s:= \pi^{-1}(s)$, $s\in S$. The submanifold $X_s^{ns}$ is completely geodesic. Thus, the family $\c X$ is equipped with a natural connection. For every two points $s_1, s_2 \in S$, we show that this connection might be integrated to an isomorphism of hyperk\"ahler manifolds $\Psi^{s_2}_{s_1}:\; X^{ns}_{s_1} {\:\longrightarrow\:} X^{ns}_{s_2}$ (\ref{_triana_subse_isome_Theorem_}). Since $X_s^{ns}$ is completely geodesic in $M$, its completion as a metric space is naturally isomorphic to a closure $X_s$ of $X_s^{ns}$ in $M$. Since the map $\Psi^{s_2}_{s_1}:\; X^{ns}_{s_1} {\:\longrightarrow\:} X^{ns}_{s_2}$ is an isometry, it is naturally extended to an isomorphism of metric spaces $\bar \Psi^{s_2}_{s_1}:\; X^{ns}_{s_1} {\:\longrightarrow\:} X^{ns}_{s_2}$ (\ref{_triana_subse_isome_Theorem_}). From construction it follows that $\Psi^{s_2}_{s_1}$ is compatible with each of induced holomorphic structures. We extend this assertion to $\bar \Psi^{s_2}_{s_1}$. This is done in two steps. A general type argument shows that a homeomorphism of complex varieties is bimeromorphic if it is holomorphic in a dense open set. The leap from bimeromorphic to holomorphic is done via a convoluted algebro-geometric argument involving normalization and finite unramified maps. We finish this paper with the definition of singular hyperk\"ahler varieties. We cite the previously known examples of singular hyperk\"ahler varieties and then show that constructed above maps $\bar \Psi^{s_2}_{s_1}:\; X^{ns}_{s_1} {\:\longrightarrow\:} X^{ns}_{s_2}$ are isomorphisms of hyperk\"ahler varieties. The Douady space of trianalytic subvarieties of a compact hyperk\"ahler manifold is proven to be hyperk\"ahler. \subsection{Contents} \begin{itemize} \item Section \ref{_Intro_Section_} is an introduction, independent from the rest of this paper. \item Section \ref{_basics_Section_} is a compendium of results pertaining to hyperk\"ahler geometry and Yang--Mills theory. We define hyperk\"ahler manifolds, trianalytic subvarieties, hyperholomorphic bundles, and cite their most basic properties. A reader acquainted with a literature may skip this section. \item Section \ref{_nonsingu_preli_Section_} illustrates our results with a simplicistic example of deformations of a smooth trianalytic subvariety. This section is also independent from the rest, and its results are further superseded by more general statements. It is safe to skip this section too. \item Section \ref{_comple_geode_defo_Section_} contains a study of completely geodesic submanifolds and their deformations. Using an easy geometric argument, we show that a family of completely geodesic submanifolds is equipped with a natural connection (\ref{_conne_in_fam_of_comple_geode_Proposition_}). We study this connection and show that, under certain additional assumptions, this connection induces isometry (metric equivalence) of the fibers of this family. This section does not use results of hyperk\"ahler geometry. \item In Section \ref{_comple_geode_hyperho_Section_}, we prove that hyperk\"ahler submanifolds are always completely geodesic. We use arguments from Yang--Mills theory and the theory of hyperholomorphic bundles (\cite{_Verbitsky:Hyperholo_bundles_}). \item Section \ref{_triholo_Section_} shows that the deformational results of Section \ref{_comple_geode_defo_Section_} can be applied to the deformations of hyperk\"ahler submanifolds, not necessary closed. Again, we use arguments from the theory of hyperholomorphic bundles. In Appendix to this section, we prove equivalence of real analytic structures induced on a hyperk\"ahler manifold by different induced complex structures. The proof is based on twistor geometry (\cite{_HKLR_}, \cite{_NHYM_}). \item Section \ref{_Douady_Section_} studies the Douady deformation space $D(X)$ of trianalytic subvariety $X$ of a compact hyperk\"ahler manifold $M$. A general argument of K\"ahler geometry shows that the Douady space is compact (\cite{_Lieberman_}). Using this, we prove that the underlying real analytic variety does not depend on the complex structure. As an application, we obtain that the Douady space is {\bf hypercomplex} (\ref{_hypercomplex_Definition_}), i. e., has an integrable quaternionic action in the tangent space. In the Appendix to this section, we give an independent proof of the compactness of the Douady space. We describe the Douady space in terms of the space of isometric embeddings from $X$ to $M$. The Appendix is based on Wirtinger's inequality (\ref{_Wirti_for_Kahle_Theorem_}). The body of this section depends only on the result of the Appendix to Section \ref{_triholo_Section_} (the equivalence of induced real analytic structures on a given hyperk\"ahler manifold). \item In Section \ref{_Conne_in_fami_Section_}, we define a curvature of a connection in a family of manifolds. We consider the natural connection on a family of trianalytic subvarieties of a compact hyperk\"ahler manifolds. Using the results of Section \ref{_Douady_Section_} (hypercomplex structure on the Douady space), we show that this connection is {\bf flat} i. e. its curvature is zero. This result is not used anywhere outside of Section \ref{_singu_hype_Section_} (and even there, we don't really need it). However, the flatness of the connection shows that the natural isomorphism $\Psi^{s_1}_{s_2}$ of the fibers of a family $\c X {\:\longrightarrow\:} S$ of trianalytic subvarieties is independent from the choice of a path $\gamma:\; [0,1] {\:\longrightarrow\:} S$ in its homotopy class. \item Section \ref{_isome=>holo_Section_} deals with the map $\Psi^{s_1}_{s_2}:\; X_{s_1} {\:\longrightarrow\:} X_{s_2}$ of fibers of a family of trianalytic subvarieties of $M$, obtained by integrating the natural connection. This map is by construction a homeomorphism and is holomorphic outside of singularities, with respect to each of induced complex structures. Using general argument, we show that $\Psi^{s_1}_{s_2}$ is bimeromorphic and induces an isomorphism of normalizations of $X_{s_1}$, $X_{s_2}$. Then, we apply the knowledge that, outside of singularities, $X_{s_i}$ are completely geodesic in $M$. This allows us to show that $\Psi^{s_1}_{s_2}$ induces an isomorphism on the Zariski tangent spaces. An algebro-geometric argument shows that these properties are sufficient to prove that $\Psi^{s_1}_{s_2}:\; X_{s_1} {\:\longrightarrow\:} X_{s_2}$ is a complex analytic isomorphism, with respect to each of induced complex structures. In Section \ref{_isome=>holo_Section_}, we use the results of Section \ref{_comple_geode_defo_Section_} (structure of deformations of completely geodesic submanifolds) and Section \ref{_triholo_Section_} (that these results may be applied to the deformations of trianalytic submanifolds). \item In Section \ref{_singu_hype_Section_}, we give a definition of singular hyperk\"ahler varieties. After a short discussion, we give examples of hyperk\"ahler varieties, stemming from the theory of hyperholomorphic bundles (\cite{_Verbitsky:Hyperholo_bundles_}). All trianalytic subvarieties of hyperk\"ahler manifolds are hyperk\"ahler varieties, which is clear from the definition. Also, the map $\Psi^{s_1}_{s_2}:\; X_{s_1} {\:\longrightarrow\:} X_{s_2}$ constructed in Section \ref{_isome=>holo_Section_}, is an isomorphism of hyperk\"ahler varieties. We show that the Douady space of deformations of a trianalytic subvariety is also a hyperk\"ahler variety. \end{itemize} \section{Basic definitions and results.} \label{_basics_Section_} This section used mainly for reference, and contains a compilation of results and definitions from literature. An impatient reader is advised to skip it and proceed to the next section. \subsection{Hyperk\"ahler manifolds} \label{_basics_hyperka_Section_} This subsection contains a compression of the basic and most known results and definitions from hyperk\"ahler geometry, found, for instance, in \cite{_Besse:Einst_Manifo_} or in \cite{_Beauville_}. \hfill \definition \label{_hyperkahler_manifold_Definition_} (\cite{_Besse:Einst_Manifo_}) A {\bf hyperk\"ahler manifold} is a Riemannian manifold $M$ endowed with three complex structures $I$, $J$ and $K$, such that the following holds. \begin{description} \item[(i)] the metric on $M$ is K\"ahler with respect to these complex structures and \item[(ii)] $I$, $J$ and $K$, considered as endomorphisms of a real tangent bundle, satisfy the relation $I\circ J=-J\circ I = K$. \end{description} \hfill The notion of a hyperk\"ahler manifold was introduced by E. Calabi (\cite{_Calabi_}). \hfill Clearly, hyperk\"ahler manifold has the natural action of quaternion algebra ${\Bbb H}$ in its real tangent bundle $TM$. Therefore its complex dimension is even. For each quaternion $L\in \Bbb H$, $L^2=-1$, the corresponding automorphism of $TM$ is an almost complex structure. It is easy to check that this almost complex structure is integrable (\cite{_Besse:Einst_Manifo_}). \hfill \definition \label{_indu_comple_str_Definition_} Let $M$ be a hyperk\"ahler manifold, $L$ a quaternion satisfying $L^2=-1$. The corresponding complex structure on $M$ is called {\bf an induced complex structure}. The $M$ considered as a complex manifold is denoted by $(M, L)$. \hfill \definition \label{_holomorphi_symple_Definition_} Let $M$ be a complex manifold and $\Omega$ a closed holomorphic 2-form over $M$ such that $\Omega^n=\Omega\wedge\Omega\wedge...$, is a nowhere degenerate section of a canonical class of $M$ ($2n=dim_{\Bbb C}(M)$). Then $M$ is called {\bf holomorphically symplectic}. \hfill Let $M$ be a hyperk\"ahler manifold; denote the Riemannian form on $M$ by $<\cdot,\cdot>$. Let the form $\omega_I := <I(\cdot),\cdot>$ be the usual K\"ahler form which is closed and parallel (with respect to the Levi-Civitta connection). Analogously defined forms $\omega_J$ and $\omega_K$ are also closed and parallel. A simple linear algebraic consideration (\cite{_Besse:Einst_Manifo_}) shows that the form $\Omega:=\omega_J+\sqrt{-1}\omega_K$ is of type $(2,0)$ and, being closed, this form is also holomorphic. Also, the form $\Omega$ is nowhere degenerate, as another linear algebraic argument shows. It is called {\bf the canonical holomorphic symplectic form of a manifold M}. Thus, for each hyperk\"ahler manifold $M$, and induced complex structure $L$, the underlying complex manifold $(M,L)$ is holomorphically symplectic. The converse assertion is also true: \hfill \proposition \label{_symplectic_=>_hyperkahler_Proposition_ (\cite{_Beauville_}, \cite{_Besse:Einst_Manifo_}) Let $M$ be a compact holomorphically symplectic K\"ahler manifold with the holomorphic symplectic form $\Omega$, a K\"ahler class $[\omega]\in H^{1,1}(M)$ and a complex structure $I$. There is a unique hyperk\"ahler structure $(I,J,K,(\cdot,\cdot))$ over $M$ such that the cohomology class of the symplectic form $\omega_I=(\cdot,I\cdot)$ is equal to $[\omega]$ and the canonical symplectic form $\omega_J+\sqrt{-1}\:\omega_K$ is equal to $\Omega$. \hfill \ref{_symplectic_=>_hyperkahler_Proposition_} immediately follows from the conjecture of Calabi, pro\-ven by Yau (\cite{_Yau:Calabi-Yau_}). \blacksquare \subsection{Hyperholomorphic bundles} \label{_hyperholo_Subsection_} This subsection contains several versions of a definition of hyperholomorphic connection in a complex vector bundle over a hyperk\"ahler manifold. We follow \cite{_Verbitsky:Hyperholo_bundles_}. \hfill Let $M$ be a hyperk\"ahler manifold. We identify the group $SU(2)$ with the group of unitary quaternions. This gives a canonical action of $SU(2)$ on the tangent bundle, and all its tensor powers. In particular, we obtain a natural action of $SU(2)$ on the bundle of differential forms. Of a special interest to us are those forms which are $SU(2)$-invariant. \hfill \lemma \label{_SU(2)_inva_type_p,p_Lemma_} Let $\omega$ be a differential form over a hyperk\"ahler manifold $M$. The form $\omega$ is $SU(2)$-invariant if and only if it is of Hodge type $(p,p)$ with respect to all induced complex structures on $M$. {\bf Proof:} This is \cite{_Verbitsky:Hyperholo_bundles_}, Proposition 1.2. \blacksquare \hfill Further in this article, we use the following statement. \lemma \label{_SU(2)_commu_Laplace_Lemma_} The action of $SU(2)$ on differential forms commutes with the Laplacian. {\bf Proof:} This is Proposition 1.1 of \cite{_Verbitsky:Hyperholo_bundles_}. \blacksquare Thus, for compact $M$, we may speak of the natural action of $SU(2)$ in cohomology. \hfill Let $B$ be a holomorphic vector bundle over a complex manifold $M$, $\nabla$ a connection in $B$ and $\Theta\in\Lambda^2\otimes End(B)$ be its curvature. This connection is called {\bf compatible with a holomorphic structure} if $\nabla_X(\zeta)=0$ for any holomorphic section $\zeta$ and any antiholomorphic tangent vector $X$. If there exist a holomorphic structure compatible with the given Hermitian connection then this connection is called {\bf integrable}. \hfill One can define a {\bf Hodge decomposition} in the space of differential forms with coefficients in any complex bundle, in particular, $End(B)$. \hfill \theorem \label{_Newle_Nie_for_bu_Theorem_} Let $\nabla$ be a Hermitian connection in a complex vector bundle $B$ over a complex manifold. Then $\nabla$ is integrable if and only if $\Theta\in\Lambda^{1,1}(M, \operatorname{End}(B))$, where $\Lambda^{1,1}(M, \operatorname{End}(B))$ denotes the forms of Hodge type (1,1). Also, the holomorphic structure compatible with $\nabla$ is unique. {\bf Proof:} This is Proposition 4.17 of \cite{_Kobayashi_}, Chapter I. $\hbox{\vrule width 4pt height 4pt depth 0pt}$ \hfill \definition \label{_hyperho_conne_Definition_} Let $B$ be a Hermitian vector bundle with a connection $\nabla$ over a hyperk\"ahler manifold $M$. Then $\nabla$ is called {\bf hyperholomorphic} if $\nabla$ is integrable with respect to each of the complex structures induced by the hyperk\"ahler structure. As follows from \ref{_Newle_Nie_for_bu_Theorem_}, $\nabla$ is hyperholomorphic if and only if its curvature $\Theta$ is of Hodge type (1,1) with respect to any of complex structures induced by a hyperk\"ahler structure. As follows from \ref{_SU(2)_inva_type_p,p_Lemma_}, $\nabla$ is hyperholomorphic if and only if $\Theta$ is a $SU(2)$-invariant differential form. \hfill \example \label{_tangent_hyperholo_Example_} (Examples of hyperholomorphic bundles) \begin{description} \item[(i)] Let $M$ be a hyperk\"ahler manifold, $TM$ its tangent bundle equipped with Levi--Civita connection $\nabla$. Then $\nabla$ is integrable with respect to each induced complex structure, and hence, Yang--Mills. \item[(ii)] For $B$ a hyperholomorphic bundle, all its tensor powers are also hyperholomorphic. \item[(iii)] Thus, the bundles of differential forms on a hyperk\"ahler manifold are also hyperholomorphic. \end{description} \subsection{Stable bundles and Yang--Mills connections.} This subsection is a compendium of the most basic results and definitions from the Yang--Mills theory over K\"ahler manifolds, concluding in the fundamental theorem of Uhlenbeck--Yau \cite{_Uhle_Yau_}. \hfill \definition\label{_degree,slope_destabilising_Definition_} Let $F$ be a coherent sheaf over an $n$-dimensional compact K\"ahler manifold $M$. We define $\deg(F)$ as \[ \deg(F)=\int_M\frac{ c_1(F)\wedge\omega^{n-1}}{vol(M)} \] and $\text{slope}(F)$ as \[ \text{slope}(F)=\frac{1}{\text{rank}(F)}\cdot \deg(F). \] The number $\text{slope}(F)$ depends only on a cohomology class of $c_1(F)$. Let $F$ be a coherent sheaf on $M$ and $F'\subset F$ its proper subsheaf. Then $F'$ is called {\bf destabilizing subsheaf} if $\text{slope}(F') \geq \text{slope}(F)$ A holomorphic vector bundle $B$ is called {\bf stable} if it has no destabilizing subsheaves. \hfill Later on, we usually consider the bundles $B$ with $deg(B)=0$. \hfill Let $M$ be a K\"ahler manifold with a K\"ahler form $\omega$. For differential forms with coefficients in any vector bundle there is a Hodge operator $L: \eta{\:\longrightarrow\:}\omega\wedge\eta$. There is also a fiberwise-adjoint Hodge operator $\Lambda$ (see \cite{_Griffi_Harri_}). \hfill \definition \label{Yang-Mills_Definition_} Let $B$ be a holomorphic bundle over a K\"ahler manifold $M$ with a holomorphic Hermitian connection $\nabla$ and a curvature $\Theta\in\Lambda^{1,1}\otimes End(B)$. The Hermitian metric on $B$ and the connection $\nabla$ defined by this metric are called {\bf Yang-Mills} if \[ \Lambda(\Theta)=constant\cdot \operatorname{Id}\restrict{B}, \] where $\Lambda$ is a Hodge operator and $\operatorname{Id}\restrict{B}$ is the identity endomorphism which is a section of $End(B)$. Further on, we consider only these Yang--Mills connections for which this constant is zero. \hfill A holomorphic bundle is called {\bf indecomposable} if it cannot be decomposed onto a direct sum of two or more holomorphic bundles. \hfill The following fundamental theorem provides examples of Yang-\--Mills \linebreak bundles. \theorem \label{_UY_Theorem_} (Uhlenbeck-Yau) Let B be an indecomposable holomorphic bundle over a compact K\"ahler manifold. Then $B$ admits a Hermitian Yang-Mills connection if and only if it is stable, and this connection is unique. {\bf Proof:} \cite{_Uhle_Yau_}. \blacksquare \hfill \proposition \label{_hyperholo_Yang--Mills_Proposition_} Let $M$ be a hyperk\"ahler manifold, $L$ an induced complex structure and $B$ be a complex vector bundle over $(M,L)$. Then every hyperholomorphic connection $\nabla$ in $B$ is Yang-Mills and satisfies $\Lambda(\Theta)=0$ where $\Theta$ is a curvature of $\nabla$. \hfill {\bf Proof:} We use the definition of a hyperholomorphic connection as one with $SU(2)$-invariant curvature. Then \ref{_hyperholo_Yang--Mills_Proposition_} follows from the \hfill \lemma \label{_Lambda_of_inva_forms_zero_Lemma_} Let $\Theta\in \Lambda^2(M)$ be a $SU(2)$-invariant differential 2-form on $M$. Then $\Lambda_L(\Theta)=0$ for each induced complex structure $L$.\footnote{By $\Lambda_L$ we understand the Hodge operator $\Lambda$ associated with the K\"ahler complex structure $L$.} {\bf Proof:} This is Lemma 2.1 of \cite{_Verbitsky:Hyperholo_bundles_}. \blacksquare \hfill Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure. For any stable holomorphic bundle on $(M, I)$ there exists a unique Hermitian Yang-Mills connection which, for some bundles, turns out to be hyperholomorphic. It is possible to tell when this happens (though in the present paper we never use this knowledge). \hfill \theorem Let $B$ be a stable holomorphic bundle over $(M,I)$, where $M$ is a hyperk\"ahler manifold and $I$ is an induced complex structure over $M$. Then $B$ admits a compatible hyperholomorphic connection if and only if the first two Chern classes $c_1(B)$ and $c_2(B)$ are $SU(2)$-invariant.\footnote{We use \ref{_SU(2)_commu_Laplace_Lemma_} to speak of action of $SU(2)$ in cohomology of $M$.} {\bf Proof:} This is Theorem 2.5 of \cite{_Verbitsky:Hyperholo_bundles_}. \blacksquare \subsection{Generic holomorphically symplectic manifolds} In this section, we define generic holomorphically symplectic manifolds. Such manifolds, as seen later (\ref{_hyperkae_embeddings_Corollary_}), admit a hyperk\"ahler structure $\c H$ such that every closed complex analytic subvariety is trianalytic with respect to $\c H$ (for the definition of trianalytic subvarieities, see \ref{_trianalytic_Definition_}). We follow \cite{Verbitsky:Symplectic_I_} (see also \cite{Verbitsky:Symplectic_II_}). \hfill Let $M$ be a compact holomorphically symplectic K\"ahler manifold. By \ref{_symplectic_=>_hyperkahler_Proposition_}, $M$ has a unique hyperk\"ahler structure with the same K\"ahler class and holomorphic symplectic form. Therefore one can without ambiguity speak about the action of $SU(2)$ on $H^*(M,{\Bbb R})$ (see \ref{_SU(2)_commu_Laplace_Lemma_}). Of course, this action essentially depends on the choice of K\"ahler class. \hfill \definition \label{_generic_manifolds_Definition_} Let $\omega\in H^{1,1}(M)$ be the K\"ahler class of a holomorphically symplectic manifold $M$. We say that $\omega$ {\bf induces the $SU(2)$-action of general type} when all elements of the group \[ \bigoplus\limits_p H^{p,p}(M)\cap H^{2p}(M,{\Bbb Z})\subset H^*(M)\] are $SU(2)$-invariant. A holomorphically symplectic manifold $M$ is called {\bf of general type} if there exists a K\"ahler class on $M$ which induces an $SU(2)$-action of general type. \hfill As \ref{_gene_type_co_div_by2_Remark_} below implies, holomorphically symplectic manifolds of general type have no Weil divisors. In particular, such manifolds are never algebraic. \hfill \proposition \label{_generic_are_dense_Proposition_} Let $M$ be a hyperk\"ahler manifold and $S$ be the set of induced complex structures over $M$. Denote by $S_0\subset S$ the set of $L\in S$ such that the natural K\"ahler metric on $(M,L)$ induces the $SU(2)$ action of general type. Then $S_0$ is dense in $S$. {\bf Proof:} This is Proposition 2.2 from \cite{Verbitsky:Symplectic_II_} \blacksquare \hfill One can easily deduce from the results in \cite{_Todorov:Moduli_I_II_} and \ref{_generic_are_dense_Proposition_} that the set of points associated with holomorphically symplectic manifolds of general type is dense in the classifying space of holomorphically symplectic manifolds of K\"ahler type. \subsection{Trianalytic subvarieties in compact hyperk\"ahler manifolds.} In this subsection, we give a definition and a few basic properties of trianalytic subvarieties of hyperk\"ahler manifolds. We follow \cite{Verbitsky:Symplectic_II_}. \hfill Let $M$ be a compact hyperk\"ahler manifold, $\dim_{\Bbb R} M =2m$. \hfill \definition\label{_trianalytic_Definition_} Let $N\subset M$ be a closed subset of $M$. Then $N$ is called {\bf trianalytic} if $N$ is a complex analytic subset of $(M,L)$ for any induced complex structure $L$. \hfill Throughout this paper, we implicitly assume that our trianalytic subvarieties are connected. Most results are trivially generalized to the general case. \hfill Let $I$ be an induced complex structure on $M$, and $N\subset(M,I)$ be a closed analytic subvariety of $(M,I)$, $dim_{\Bbb C} N= n$. Denote by $[N]\in H_{2n}(M)$ the homology class represented by $N$. Let $\inangles N\in H^{2m-2n}(M)$ denote the Poincare dual cohomology class. Recall that the hyperk\"ahler structure induces the action of the group $SU(2)$ on the space $H^{2m-2n}(M)$. \hfill \theorem\label{_G_M_invariant_implies_trianalytic_Theorem_} Assume that $\inangles N\in H^{2m-2n}(M)$ is invariant with respect to the action of $SU(2)$ on $H^{2m-2n}(M)$. Then $N$ is trianalytic. {\bf Proof:} This is Theorem 4.1 of \cite{Verbitsky:Symplectic_II_} (see Subsection \ref{_SU(2)-inv=>triana_Subsection_} for a sketch of a proof). \blacksquare \remark \label{_triana_dim_div_4_Remark_} Trianalytic subvarieties have an action of quaternion algebra in the tangent bundle. In particular, the real dimension of such subvarieties is divisible by 4. \hfill \ref{_G_M_invariant_implies_trianalytic_Theorem_} has the following immediate corollary: \corollary \label{_hyperkae_embeddings_Corollary_} Let $M$ be a compact holomorphically symplectic manifold of general type, $S\subset M$ be its complex analytic subvariety, and $\omega$ be a K\"ahler class which induces an $SU(2)$-action of general type. Consider the hyperk\"ahler structure associated with $\omega$ by \ref{_symplectic_=>_hyperkahler_Proposition_}. Then $S$ is trianalytic with respect to $\c H$. \blacksquare \remark \label{_gene_type_co_div_by2_Remark_} {}From \ref{_hyperkae_embeddings_Corollary_} and \ref{_triana_dim_div_4_Remark_}, it follows that a holomorphically symplectic manifold of general type has no closed complex analytic subvarieties of odd dimension; in particular, such a manifold has no divisors. \subsection{Wirtinger's inequality and its use in K\"ahler geometry} Let $M$ be a compact K\"ahler manifold, $X\subset M$ a closed real analytic subvariety. In this section, we give criteria for $X$ to be complex analytic, in terms of certain integrals. We follow \cite{Verbitsky:Symplectic_II_} and \cite{_Stolzenberg_}. \hfill \definition \label{_volume_Riema_Definition_} Let $H$ be an ${\Bbb R}$-linear space equipped with a positively defined scalar product, $\dim H =h$. The exterrior form $\operatorname{Vol}\in \Lambda^h(H)$ is called {\bf a volume form} if the the standard hypercube with the side 1 has the volume 1 in the measure defined by $\operatorname{Vol}$. \hfill Clearly, the volume form is defined up to a sign. This sign is determined by the choice of orientation on $H$. In the same manner we define the top degree differential form $\operatorname{Vol}$ called {\bf a volume form} on any oriented Riemannian manifold. \hfill Let $V$ be a Hermitian linear space, $W\subset V$ be a ${\Bbb R}$-linear subspace, $\dim_{\Bbb R} W = 2n$. Consider space $\Lambda^{2n}(W)$ of volume forms on $M$. Let $\omega$ be the imaginary part of the Hermitian form on $V$. Consider the vectors $\omega^n$, $\operatorname{Vol} \in \Lambda^{2n}(W)$. Since $\operatorname{Vol}$ is non-zero, and $\Lambda^{2n}(W)$ is one-dimensional, we can speak of a fraction $\frac{\omega^n}{\operatorname{Vol}}$, which is a real number, defined up to a sign (the form $\operatorname{Vol}$ is defined up to a sign). Denote by $\Xi_W$ the number $\left|\frac{\omega^n}{\operatorname{Vol}}\right|$. \hfill \lemma\label{_Wirtinger_Lemma_} (Wirtinger's inequality) In these assumptions, $\Xi_W\leq 2^n.$ Moreover, if $\Xi_W=2^n$, then $W$ is a complex subspace of $V$. {\bf Proof:} \cite{_Stolzenberg_} page 7. \blacksquare \hfill Let $M$ be a K\"ahler manifold, $N\subset (M,I)$ be a closed real analytic subvariety of even dimension, $N_{ns}\subset N$ the set of non-singular points of $N$. For each $x\in N_{ns}$, consider $T_xN$ as a subspace of $T_xM$. Denote the corresponding number $\Xi_{T_xN}$ by $\Xi_L (x)$. The following statement is a direct consequence of \ref{_Wirtinger_Lemma_}: \hfill \proposition\label{_N_is_analytic_if_eta_is_constant_Proposition_} Let $J$ be an induced complex structure. The subset $N\subset M$ is complex analytic if and only if \[ \forall x\in N_{ns} \;\;\;\; \Xi_L(x)=2^n. \] \blacksquare \hfill Let $\operatorname{Vol} N_{ns}$ be the volume form of $N_{ns}$, taken with respect to the Riemannian form (see \ref{_volume_Riema_Definition_}). \hfill \theorem \label{_Wirti_for_Kahle_Theorem_} Let $M$ be a K\"ahler manifold, $N\subset M$, $\dim N =2n$ a closed real analytic subvariety, $N_{ns}\subset M$ the set of its non-singular points. Assume that the improper integrals $\int_{N_{ns}} \operatorname{Vol} N_{ns}$, $\int_{N_{ns}} \omega^n$ exist. Then \[ 2^n\int_{N_{ns}} \operatorname{Vol} N_{ns} \geq \int_{N_{ns}} \omega^n \] and the equality is reached if and only if $N$ is complex analytic in $M$. {\bf Proof:} This is a direct consequence of \ref{_N_is_analytic_if_eta_is_constant_Proposition_} \blacksquare \hfill We use the term ``symplectic volume'' for the number $\frac{1}{2^n}\int_{N_{ns}} \omega^n $ and ``Riemannian volume'' for $\int_{N_ns} \operatorname{Vol} N_{ns}$. Then, \ref{_Wirti_for_Kahle_Theorem_} might be rephrased in the form ``a real analytic cycle is complex analytic if and only if its symplectic volume is equal to its Riemannian volume''. \subsection{$SU(2)$-invariant cycles in cohomology and trianalytic subvarieties} \label{_SU(2)-inv=>triana_Subsection_} This subsection consists of a sketch of a proof of \ref{_G_M_invariant_implies_trianalytic_Theorem_}. We follow \cite{Verbitsky:Symplectic_II_}. \hfill For a K\"ahler manifold $M$, $m=dim_{\Bbb C} M$ and a form $\alpha\in H^{2p}(M,{\Bbb C})$, define \[ \deg(\alpha):=\int_M \omega^{m-p}\wedge\alpha, \] where $\omega$ is the K\"ahler form. Let $M$ be a compact hyperk\"ahler manifold, $\alpha$ a differential form. We denote by $\deg_L\alpha$ the degree associated with an induced complex structure $L$. \hfill \proposition \label{_G_M_invariant_cycles_over_Proposition_} Let $M$ be a compact hyperk\"ahler manifold and $\alpha$ be an $SU(2)$-invariant form of non-zero degree. Then the dimension of $\alpha$ is divisible by 4. Moreover, \[ \deg_{I}\alpha = \deg_{I'}\alpha, \] for every pair of induced complex structures $I$, $I'$. {\bf Proof:} This is Proposition 4.5 of \cite{Verbitsky:Symplectic_II_}. \blacksquare \hfill \ref{_G_M_invariant_implies_trianalytic_Theorem_} follows immediately from \ref{_G_M_invariant_cycles_over_Proposition_} and \ref{_Wirti_for_Kahle_Theorem_}. By \ref{_Wirti_for_Kahle_Theorem_}, a real analytic subvariety $N \subset M$ is trianalytic if and only if its symplectic volume, taken with respect to any of induced complex structures, is defined and equal to its Riemannian volume. In notation of \ref{_G_M_invariant_implies_trianalytic_Theorem_}, the symplectic volume of $N$ taken with respect to $L$ is equal to $\frac {1}{2^n} \deg_L\inangles N$. {}From \ref{_G_M_invariant_cycles_over_Proposition_}, we obtain that symplectic volume of $N$ is the same for all induced complex structures. Since $N$ is complex analytic with respect to $I$, the symplectic volume of $N$ is equal to the Riemannian volume of $N$, again by \ref{_Wirti_for_Kahle_Theorem_}. This proves \ref{_G_M_invariant_implies_trianalytic_Theorem_}. \section[Deformations of non-singular trianalytic subvarieties.] {Deformations of non-singular trianalytic \\subvarieties.} \label{_nonsingu_preli_Section_} Let $M$ be a holomorphically symplectic K\"ahler manifold, $X\subset M$ a closed complex analytic submanifold of trianalytic type. We study the deformations of $X$ in $M$, in order to prove \ref{_iso_intro:Theorem_}. It turns out that \ref{_iso_intro:Theorem_} is almost trivial in the case $X$ non-singular. In this section, we give the proof of \ref{_iso_intro:Theorem_} (i) for non-singular $X$; the general case is proven independently. We hope that a simple argument will be insightful, even if we need to produce a separate proof for the general case. This section is perfectly safe to skip. \hfill \remark For a smooth trianalytic submanifold $X\subset M$, $X$ is obviously hyperk\"ahler, in a natural way. The quaternion action comes from quaternion action in $TM$, and the metric is induced from $M$ too. \hfill We start from the following simple, but important, lemma. \hfill \lemma \label{_TM-restrict_X_hyperholo:Lemma_} Let $M$ be a complex manifold equipped with a hyperk\"ahler metric, $X\subset M$ a closed complex analytic submanifld of trianalytic type. Consider the restriction $TM\restrict{X}$ of the tangent bundle to $M$ on $X$. We equip $TM_\restrict{X}$ with a connection $\nabla$ coming from Levi-Civita connection in $TM$. Then $(TM\restrict{X}, \nabla)$ is {\bf hyperholomorphic}.\footnote{See Subsection \ref{_hyperholo_Subsection_} for the definition of {\bf hyperholomorphic}. By \ref{_hyperholo_Yang--Mills_Proposition_}, a hyperholomorphic connection is Yang-Mills.} \hfill {\bf Proof:} Consider the natural action of $SU(2)$ on the space \[ \Lambda^2\left(X, \operatorname{End}\left(TM\restrict{X}\right)\right). \] We need to show that the curvature $\Theta$ of $\nabla$ is $SU(2)$-invariant. Let $\nabla_{LC}$ be the Levi--Civita connection on $TM$, and $(\nabla_{LC})^2\in \Lambda^2(M, \operatorname{End}(TM))$ be its curvature. Clearly, $\Theta$ is a pull-back of $(\nabla_{LC})^2$ to $X$. Therefore, it suffices to show that $(\nabla_{LC})^2$ is $SU(2)$-invariant. This is \ref{_tangent_hyperholo_Example_}. \blacksquare \hfill As a corollary, we obtain the following proposition. \hfill \proposition \label{_TM-restrict-X_decompo_for_compa_Proposition_} In assumptions of \ref{_TM-restrict_X_hyperholo:Lemma_}, let $M$, $X$ be compact. Then the following statements are true. \begin{description} \item[(i)] The bundle $TM\restrict{X}$ is naturally isomorphic to the direct sum $TM\restrict X \cong TX \oplus NX$, where $N$ is the normal bundle to $X$ inside $M$. This isomorphism is compatible with the natural connections and Hermitian metrics on $TM\restrict X$, $TX$, $NX$. \item[(ii)] $NX$ is hyperholomorphic. \item[(iii)] For each section $\gamma$ of $NX$, $\gamma$ is nowhere degenerate, and there is a natural decomposition \begin{equation} \label{_NX=O-gamma+rest:Equation_} NX = {\cal O}_\gamma \oplus N_\gamma X, \end{equation} where ${\cal O}_\gamma$ is a trivial sub-bundle of $NX$ generated by $\gamma$, and $N_\gamma X$ its orthogonal complement. The decomposition \eqref{_NX=O-gamma+rest:Equation_} is compatible with connection. \item[(iv)] In assumptions of (iii), consider the connection $\nabla_\gamma$ induced from $NX$ to ${\cal O}_\gamma X$. Then is $\nabla_\gamma$ flat. \end{description} \hfill \remark In fact, \ref{_TM-restrict-X_decompo_for_compa_Proposition_} (i)-(ii) holds true in assumptions of \ref{_TM-restrict_X_hyperholo:Lemma_} for general, not necessarily compact, $X$ and $M$ (see \ref{_NX_splits_for_hype_Proposition_}, \ref{_NX_hyperholo_Proposition_}). However, the proof is easier in compact case. \hfill {\bf Proof of \ref{_TM-restrict-X_decompo_for_compa_Proposition_}.} Consider the embedding \[ TX \hookrightarrow TM\restrict X \] of bundles with connection. By \ref{_tangent_hyperholo_Example_}, $TX$ is hyperholomorphic, and hence Yang--Mills. By \ref{_TM-restrict_X_hyperholo:Lemma_}, $TM\restrict X$ is hyperholomorphic as well. We obtain that $TX$ is a destabilizing subsheaf in $TX$. By \ref{_UY_Theorem_}, a Yang--Mills bundle is a direct sum of stable bundles.\footnote{Bundles which are direct sum of stable are called {\bf polystable}.} Thus, $TM$ is a direct sum of $TX$ and its orthogonal complement $NX$ (see \ref{_YM_exact_split_Proposition_} for details). This proves \ref{_TM-restrict-X_decompo_for_compa_Proposition_} (i). Since the curvature of $TM\restrict X$ is decomposed onto a direct sum of the curvature of $NX$ and the curvature of $TX$, the curvature of $NX$ is $SU(2)$-invariant. This proves \ref{_TM-restrict-X_decompo_for_compa_Proposition_} (ii). \hfill The condition (iii) follows from (ii) and \ref{_UY_Theorem_}, since every holomorphic section $\gamma$ of a Yang--Mills bundle $B$ of zero degree (such as hyperholomorphic bundles) spans a destabilizing subsheaf ${\cal O}_\gamma$. Since $B$ is Yang--Mills, it is polystable; thus, ${\cal O}_\gamma$ is its direct summand. This proves \ref{_TM-restrict-X_decompo_for_compa_Proposition_} (iii). To prove (iv), we notice that ${\cal O}_\gamma$ is a trivial holomorphic bundle with Yang--Mills connection. By \ref{_UY_Theorem_}, Yang--Mills connection is unique, and therefore, ${\cal O}_\gamma$ is flat. \ref{_TM-restrict-X_decompo_for_compa_Proposition_} is proven. \blacksquare \hfill We recall the following general results of the theory of deformations of complex subvarieties. Let $X\subset M$ be a closed complex analytic subvariety of a compact complex manifold. The {\bf Douady space} of deformations of $X$ inside of $M$ is defined (\cite{_Douady_}). We denote the Douady space by $D_M X$, or sometimes by $D(X)$. By definition, $D_M X$ is a finite-dimensional complex analytic variety. The points of $D_MX$ are identified with the subvarieties $X'\subset M$, where $X'$ is a deformation of $X$. Let $\gamma:\; {\Bbb R} {\:\longrightarrow\:} D_m X$ be a real analytic map. Assume that for $t=t_0$, $\gamma(t)$ is a smooth subvariety of $M$. In deformation theory, the differential $\frac{d\gamma}{dt}(t_0)$ is interpreted as a holomorphic section of the normal bundle $N \gamma (t_0)$. Thus, the Zariski tangent space $T_{X} D_M (X)$ is naturally embedded to the space of holomorphic sections of $N X$. The following proposition is an easy application of the Kodaira--Spencer theory. \hfill \proposition \label{_compa_subva_iso_holo:Proposition_} Let $X \subset M$ be a closed complex analytic submanifold of a compact complex manifold $M$, and $D_M X$ the corresponding Douady space. Let $\gamma:\; {\Bbb R} {\:\longrightarrow\:} D_M X$ be a real analytic map. Assume that for all $t_0\in {\Bbb R}$, the subvariety $\gamma(t_0)$ is smooth. Assume also that the section \[ \nu_{t_0} = \frac{d \gamma}{dt}(t_0) \in \Gamma_{\gamma(t_0)} (N \gamma(t_0)) \] is nowhere degenerate. Finally, assume that $\nu_{t_0}$ splits from $N \gamma(t_0)$, i. e, there exists a decomposition of a holomorphic vector bundle \[ N \gamma (t_0) = {\cal O}_{\nu_{t_0}} \oplus N' \] where ${\cal O}_{\nu_{t_0}}$ is the trivial subbundle of $N \gamma(t_0)$ generated by $\nu_{t_0}$. Then the subvarieties $\gamma(t)\subset M$ are naturally isomorphic for all $t$. {\bf Proof:} Well known; see, for instance, \cite{_Kodaira_Spencer_} \blacksquare \hfill \proposition \label{_compa_subva_iso_metric:Proposition_} In assumptions of \ref{_compa_subva_iso_holo:Proposition_}, let $M$ be K\"ahler. Consider the induced Hermitian structure on $N\gamma(t)$, for all $t\in {\Bbb R}$. Assume that for all $t\in {\Bbb R}$, the section \[ \nu_{t_0} = \frac{d \gamma}{dt}(t_0) \] has constant length. Assume also that the orthogonal decomposition \[ TM\restrict X = NX \oplus NX^\bot \] is compatible with the Levi-Civita connection in $TM\restrict X$. Consider the isomorphisms $\psi_{t_1,t_2}:\; \gamma(t_1){\:\longrightarrow\:} \gamma(t_2)$ constructed in \ref{_compa_subva_iso_holo:Proposition_}. Then the maps $\psi_{t_1,t_2}$ are compatible with the K\"ahler metric induced from $M$. {\bf Proof:} The proof follows from Kodaira--Spencer construction; for a complete argument, see \ref{_conne_in_fami_of_comple_geo_Proposition_} (ii). \blacksquare \hfill The following theorem is the main result of this section. \hfill \theorem \label{_iso_for_smooth_subva:Theorem_} Let $M$ be a compact holomorphically symplectic K\"ahler manifold, and $X\subset M$ a complex submanifold of trianalytic type. Let $X'$ be a deformation of $X$ in $M$. Then there exists a complex analytic isomorphism $\psi:\; X {\:\longrightarrow\:} X'$. Moreover, if the K\"ahler metric on $M$ is hyperk\"ahler, then $\psi:\; X {\:\longrightarrow\:} X'$ is compatible with the K\"ahler structure induced from $M$. \hfill {\bf Proof:} The deformations of $X$ are infinitesimally classified by the sections of $NX$. Let $\gamma$ be such a section. Applying \ref{_TM-restrict-X_decompo_for_compa_Proposition_}, we obtain that assumptions of \ref{_compa_subva_iso_holo:Proposition_} and \ref{_compa_subva_iso_metric:Proposition_} hold for the deformations of $X$ inside $M$. Now, \ref{_iso_for_smooth_subva:Theorem_} is directly implied by \ref{_compa_subva_iso_holo:Proposition_} and \ref{_compa_subva_iso_metric:Proposition_}. \blacksquare \hfill \remark \ref{_iso_for_smooth_subva:Theorem_} gives a proof of \ref{_iso_intro:Theorem_} (i), for the subvarieties $X\subset M$ which are smooth. We prove \ref{_iso_for_smooth_subva:Theorem_} for general subvarieties in \ref{_triana_subse_comple_ana_Theorem_}. \section[Completely geodesic embeddings of Riemannian manifolds.] {Completely geodesic embeddings \\of Riemannian manifolds.} \label{_comple_geode_defo_Section_} Before we proceed with the proof of \ref{_iso_intro:Theorem_}, we have to prove a serie of preliminary results from deformation theory. The arguments of deformation theory are greatly simplified in the hyperk\"ahler case, because hyperk\"ahler embeddings are completely geodesic\footnote{For a definition of completely geodesical embeddings, see \ref{_comple_geode:Definition_}.} (\ref{_hype_embe_comple_geode:Corollary_}). We prove a number of simple statements from the deformation theory of completely geodesical embeddings of K\"ahler manifolds. \subsection{Completely geodesic submanifolds} \nopagebreak \hspace{5mm} \proposition \label{_comple_geodesi_basi_Proposition_ Let $X \stackrel \phi\hookrightarrow M$ be an embedding of Riemannian manifolds (not necessarily compact) compatible with the Riemannian structure.\footnote{Such embeddings are called {\bf Riemannian embeddings}.} Then the following conditions are equivalent. \begin{description} \item[(i)] For every point $x\in X$, there exist a neighbourhod $U \ni x$ of $x$ in $X$ such that for all $x' \in U$ there is a geodesic in $M$ going from $\phi(x)$ to $\phi(x')$ which lies in $\phi(X)\subset M$. \item[(ii)] Consider the Levi-Civita connection $\nabla$ on $TM$, and restriction of $\nabla$ to $TM \restrict{X}$. Consider the orthogonal decomposition \begin{equation} \label{TM_decompo_Equation_} TM\restrict{X} = TX \oplus TX^\bot. \end{equation} Then, this decomposition is preserved by the connection $\nabla$. \end{description} {\bf Proof:} Well known; see, for instance, \cite{_Besse:Einst_Manifo_}. \hbox{\vrule width 4pt height 4pt depth 0pt} \hfill \definition\label{_comple_geode:Definition_} Let $X \stackrel i \hookrightarrow M$ be a Riemannian embedding satisfying either of the conditions of \ref{_comple_geodesi_basi_Proposition_}. Then $i$ is called {\bf a completely geodesic embedding}, and the image $i(X)\subset M$ is called {\bf a completely geodesic submanifold}. \hfill \lemma \label{_comple_geo_compa_holo_Lemma_} Let $M$ be a K\"ahler manifold, $X\subset M$ a complex submanifold. If $X$ is completely geodesic, then the decomposition \eqref{TM_decompo_Equation_} is compatible with the holomorphic structure on $TX$, $TM\restrict X$. {\bf Proof:} Clear. \blacksquare \subsection{Deformations of submanifolds} \label{_defo_subva_conven_Subsection_} Let $\c X \stackrel \pi {\:\longrightarrow\:} S$ be a family of complex manifolds equipped with a map $\c X \stackrel \phi {\:\longrightarrow\:} M$. Denote the pre-image $\pi^{-1}(s)\subset \c X$ by $X_s$. Assume that for all $s\in S$ the restriction $\phi_s:\; X_s{\:\longrightarrow\:} M$ of $\c X \stackrel \phi {\:\longrightarrow\:} M$ to $X_s \subset \c X$ is a smooth embedding. Assume that for $s_0\in S$, the image $\phi_{s_0}(X_{s_0})\subset M$ coinsides with $X$. \hfill \definition The collection of data \[ \left( \c X \stackrel \pi {\:\longrightarrow\:} S, \phi, \phi_{s_0}:\; X_{s_0} \oldtilde {\:\longrightarrow\:} X\right ) \] is called {\bf a family of submanifolds of $M$}, and {\bf a family of deformations of $X$}. The same definition can be formulated for $M$, $X$, $S$, $\c X$ real analytic; in this case, we speak of {\bf family of real analytic submanifolds of $M$}. Also, the submanifolds might be replaced by subvarieties in order to obtain the definition of a family of subvarieties. \subsection{Section of a normal bundle arising from a deformation} \label{_norma_vecto_Subsection_} Let $M$ be a complex or real analytic manifold and \[ \left( \c X \stackrel \pi {\:\longrightarrow\:} S, \phi:\; \c X {\:\longrightarrow\:} M \right) \] be a system of submanifolds. For each tangent vector $t \in T_{s_0} S$, deformation theory gives a canonical section $\eta$ of the normal bundle $N X_{s_0}$. This section is holomorphic if we work in complex situation. We recall how this section is obtained. For a sufficiently small neighbourhood of $x\in \c X$, we can always find coordinates in $X_s$ which analytically depend on $s$. For each point $x\in X_{s_0}$, denote by $x(s)$ the point of $X_s$ with the same coordinates as $x$. Consider the vector $\tilde \eta_x :=\frac{dx(s)}{ds}(t)\in T_xM$, which is a derivative of $x(s)$ along $t$. The vector $\tilde \eta_x$ obviously depends on the choice of coordinates in $X_s$. Let $\eta_x\in NX_{s_0}\restrict x$ be the image of $\tilde\eta_x$ under the natural map $T_x M {\:\longrightarrow\:} N X_{s_0}\restrict x$. Clearly, $\eta_x$ is independent from the choice of coordinates. Gluing $\eta_x$ together, we obtain the canonical section $\eta\in NX_{s_0}$. We need the following technical lemma in Section \ref{_triholo_Section_}. \hfill \lemma \label{_norma_sec_holom_Lemma_} Let $M$ be a complex analytic manifold and \[ \left( \c X \stackrel \pi {\:\longrightarrow\:} S, \phi:\; \c X {\:\longrightarrow\:} M \right) \] be a real analytic system of submanifolds. Assume that for all $s\in S$, the submanifold $\pi^{-1}(s) := X_s\subset M$ is complex analytic. Then, for all $t\in T_{s_0} S$, the corresponding section $\eta \in N X_{s_0}$ is holomorphic. \hfill {\bf Proof:} Shrinking $\c X$ if necessary, we can find a system of holomorphic coordinates in $X_s$, $s\in S$ which depends smoothly upon $s$. Consider the section $\tilde \eta\in TM\restrict{X_{s_0}}$ obtained by deriving $x(s)$ along $t$ as above. Since the map $x{\:\longrightarrow\:} x(s)$ is by construction holomorphic, the section $\tilde\eta$ is also holomorphic. Then, $\eta$ is holomorphic by construction. \blacksquare \subsection{Connections in families of manifolds} \label{_conne_Subsection_} \hfill \definition \label{_conne_Definition_} Let $\pi:\; \c X {\:\longrightarrow\:} S$ be a family of real analytic manifolds parametrized by $S$. Consider the bundles $T\c X$, $N_\pi X$, where $N_\pi X$ is a fiberwise normal bundle to the fibers of $\pi$. There is a natural projection $p:\; T\c X {\:\longrightarrow\:} N_\pi X$. The {\bf connection} in $\c S$ is a section of $p$, i. e. such an embedding $\nabla:\; N_\pi X {\:\longrightarrow\:} T\c X$ that $p \circ s = \operatorname{Id} \restrict {N_\pi X}$. This definition is naturally adopted to the case of complex analytic family of manifolds. If the section $\nabla:\; N_\pi X {\:\longrightarrow\:} T\c X$ is holomorphic, the connection is called {\bf a holomorphic connection}. This definition makes sense only when the base variety $S$ is smooth. However, the connection in a family gives a connection on a pullback of this family, under all maps $s:\; S' {\:\longrightarrow\:} S$. Connections on the pullback families are naturally compatible. We shall sometimes speak of {\bf connection} in a family where the base $S$ is not smooth. This means that for all maps $s:\; S' {\:\longrightarrow\:} S$, where $S'$ is smooth, the pullback family is equipped with a connection, and these connections are compatible. All the definitions and results (which we state and prove in the case of a smooth base) are naturally adopted to the case when the base $S$ is singular. \hfill \proposition \label{_conne_in_fam_of_comple_geode_Proposition_} Let \[ \left( \c X \stackrel \pi {\:\longrightarrow\:} S, \phi:\; \c X {\:\longrightarrow\:} M \right) \] be a family of submanifolds of a Riemannian manifold $M$. Assume that for all $s\in S$, the submanifold $X_s:= \psi(\pi^{-1}(s)) \subset M$ is completely geodesic. Then the family $\c X$ is equipped with a natural connection. Moreover, if $M$ is K\"ahler and the family $\c X$ is complex analytic, then the connection $\nabla$ is holomorphic. \hfill {\bf Proof:} Let $T_\pi X\subset T\c X$ be the bundle of vectors tangent to the fibers of $\pi$. To split the exact sequence \[ 0 {\:\longrightarrow\:} T_\pi X \stackrel i{\:\longrightarrow\:} T\c X {\:\longrightarrow\:} T_\pi X {\:\longrightarrow\:} 0, \] we have to construct a surjection \begin{equation} \label{_surje_secti_for_conne_Equation_} T{\c X} \stackrel p {\:\longrightarrow\:} T_\pi X \end{equation} satisfying $i\circ p = \operatorname{Id}_{T_\pi X}$. Consider the orthogonal decomposition \begin{equation} \label{_ortho_compo_for_norma_in_fami_Equation_} \phi^*TM=\phi^*N_M X_t \oplus TX_t, \end{equation} where $N_M X_t$ is the normal bundle to $\phi(X_t)$ in $M$. This gives a natural epimorphism \[ \phi^* TM \stackrel e{\:\longrightarrow\:} T_\pi X.\] Taking a composition of $e$ with \[ d\phi:\; T\c X {\:\longrightarrow\:} \phi^* TM,\] we obtain a surjection $p$ of \ref{_surje_secti_for_conne_Equation_}. In K\"ahler case, the decomposition \eqref{_ortho_compo_for_norma_in_fami_Equation_} is holomorphic by \ref{_comple_geo_compa_holo_Lemma_}. Thus, we constructed a connection which is holomorphic. \blacksquare \hfill For each real analytic path $\gamma:\; [0,1] {\:\longrightarrow\:} S$, $t, t'\in [0,1]$, we may {\bf integrate} the connection along $\gamma$. This notion is intuitively clear (and well known). We recall briefly the definition of the integral maps associated to connection, in order to fix the notation. Restricting the family $\c X$ to the image of $\gamma$ in $S$ does not change the result of integration. To simplify the exposition, we assume that $S$ coinsides with the image $D \cong [0,1]$ of $\gamma$. In this case, the bundle $N_\pi X$ is naturally isomorphic to a pullback $\pi^* T D$. Denote by $\nu$ a canonical unit section of $\pi^* TD$ corresponding to the unit tangent field to $[0,1]$. Then, $\nabla \nu$ is a vector field in $T\c X$. Integrating $\nu$ to a diffeomorphism, we obtain a map $\exp(t \nu)$ defined in an open subset of $\c X$, for $t\in {\Bbb R}$. Clearly, $\exp(t\nu)$ maps the points of $X_{t_1}$ to the points of $X_{t_1+t}$. The resulting diffeomorphism we denote by \[ \Psi^{t_1}_{t_1+t}:\; U_{t_1} {\:\longrightarrow\:} U_{t+t_1}, \] where $U_{t_1}$ is an intersection of $X_{t_1}$ with the domain of $\exp(t\nu)$. Clearly, the same definition applies to the arbitrary families of manifolds with connection. When $\c X$ is a complex analytic family with a holomorphic connection, the diffeomorphisms $\Psi^t_{t'}$ are complex analytic. \subsection{Deformations of completely geodesic submanifolds (the main statement)} Let \[ \left( \c X \stackrel \pi {\:\longrightarrow\:} S, \phi, \phi_{s_0}:\; X_{s_0} \oldtilde {\:\longrightarrow\:} X\right ) \] be a real analytic deformation of $X\subset M$, and $\gamma:\; [0,1] {\:\longrightarrow\:} S$ be a real analytic map. Slightly abusing the notation, we shall consider the fibers $X_s$ of $\pi$ as subvarieties in $M$. Assume that $X_{\gamma(t)}$ is a completely geodesic complex analytic submanifold of $M$, for all $t$. Let $\eta_t \in \Gamma \left(N X_{\gamma(t)}\right)$ be a section of the normal bundle to $X_{\gamma(t)}$ corresponding to the vector $\frac {d \gamma}{dt} \in T_{\gamma(t)} S$. Using the natural holomorphic embedding \begin{equation} \label{_N_embe_to_TM_Equation_} N X_{\gamma(t)} \hookrightarrow TM\restrict{X_{\gamma(t)}} \end{equation} associated with a decomposition $TM\restrict{X_{\gamma(t)}} = N X_{\gamma(t)}\oplus TX_{\gamma(t)}$, we may consider $\eta_t$ as a section of $TM\restrict{X_{\gamma(t)}}$. \hfill \proposition \label{_conne_in_fami_of_comple_geo_Proposition_} Let $t$, $t'\in [0,1]$. Let $\Psi^t_{t'}:\; U_t {\:\longrightarrow\:} U_{t'}$ be the map of Subsection \ref{_conne_Subsection_} integrating the connection $\nabla$ of \ref{_conne_in_fam_of_comple_geode_Proposition_}, where $U_t$, $U_{t'}$ are open subsets of $X_{\gamma(t)}$, $X_{\gamma(t')}$ defined in Subsection \ref{_conne_Subsection_}. Assume that for all $t$ the section $\eta_t\in TM\restrict{X_{\gamma(t)}}$ is parallel with respect to the natural connection on $TM\restrict{X_{\gamma(t)}}$ obtained from Levi--Civita on $TM$. Assume also that the Riemannian form on $M$ is real analytic.\footnote{This assumption is extraneous; we use it to simplify the exposition. For the case we are interested in ($M$ a hyperk\"ahler manifold) the Riemannian form is real analytic by \ref{_Riema_on_hype_real_ana_Corollary_}.} Then $\Psi^t_{t'}$ satisfies the following conditions. \begin{description} \item[(i)] $\Psi^t_{t'}$ is compatible with a Riemannian structure on $U_t$, $U_{t'}$. \item[(ii)] Assume also that $M$ admits a smooth Riemannian compactification $\bar M$ such that the family $\c X$ admits a compactification in $\bar M$ (such is the case when $M$ and $\c X$ are quasiprojective, and $\phi$ is algebraic). Then the maps $\Psi^t_{t'}:\; U_t {\:\longrightarrow\:} U_{t'}$ of (i) can be found in such a way that $U_t$ is the set of non-singular points of $X_{\gamma(t)}$. \item[(iii)] In assumptions of (ii), we can extend $\Psi^t_{t'}$ to an isomorphism of metric spaces \[ \bar \Psi^t_{t'}:\; \bar X_{\gamma(t)} {\:\longrightarrow\:} \bar X_{\gamma(t')}, \] where $\bar X_{\gamma(t)}$, $\bar X_{\gamma'(t)}$ are closures of $X_{\gamma(t)}$, $X_{\gamma'(t)}$ in $\bar M$, with induced metrics. \end{description} The next subsection is taken by the proof of \ref{_conne_in_fami_of_comple_geo_Proposition_}. \subsection{Deformations of completely geodesic submanifolds (the proofs)} The statement (i) is sufficient to prove in a small neighbourhood, by analytic continuation. Thus, we may pick an open subset $U\in {\c X}$ in such a way that the restriction of $\phi:\; \c X {\:\longrightarrow\:} M$ to \[ \c X':= \pi^{-1}(\gamma([0,1])\cap U \] is an embedding. Shrinking $\c X$ to $\c X'$ and pulling from $M$ the Riemannian metric, we see that \ref{_conne_in_fami_of_comple_geo_Proposition_} (i) is implied by the following lemma. \hfill \lemma \label{_integra_field_Killing_Lemma_} Let $\c X \stackrel \pi{\:\longrightarrow\:} [0,1]$ be a real analytic family of manifolds, equipped with a Riemannian metrics. Let $T _\pi X\subset T\c X$ be the relative tangent bundle consisiting of all vectors tangent to the fibers of $\pi$, and $N _\pi X= T \c X /T _\pi X$ be the normal bundle to the fibration $\pi$. Assume that the fibers $\pi^{-1} (t) \subset \c X$ are completely geodesic in $\c X$. Consider the connection $\nabla:\; N _\pi X {\:\longrightarrow\:} T\c X$ of \ref{_conne_in_fam_of_comple_geode_Proposition_}. Let $\Psi^t_{t'}:\; U_t {\:\longrightarrow\:} U_{t'}$ be the maps obtained by integrating the connection as in Subsection \ref{_conne_Subsection_} and $\eta_t \in N _\pi X$ be the normal fields arising from the deformation theory. Assume that $\eta_t$'s are parallel with respect to the natural connection on $N _\pi X$. Then the maps $\Psi^t_{t'}$ are compatible with Riemannian metrics. \hfill {\bf Proof:} Let $\eta\in T \c X$ be the tangent vector field obtained by gluing all $\eta_t$ together. Then $\Psi_t$ can be considered as integral map of this vector field. Thus, to prove that $\Psi_t$ is compatible with the Riemannian structure, we have to show that $\eta$ is {\bf a Killing vector field}.\footnote{A Killing field is a vector field which integrates to a diffeomorphism which is compatible with a Riemannian metric.} Denote by $\nabla_x:\; T\c X {\:\longrightarrow\:} T\c X$ the action of covariant derivative along the vector field $x$. By \cite{_Besse:Einst_Manifo_}, Theorem 1.81, to prove that $\eta$ is Killing it suffices to prove that for all fields $a, b\in T\c X$, we have \begin{equation}\label{_Killing_Equation_} (\nabla_a \eta, b) + (\nabla_b \eta,a) =0, \end{equation} where $(\cdot,\cdot)$ is the Riemannian form. Take coordinates $(x_0, x_1,...x_n)$ on $\c X$ in such a way that $x_0$ comes from a projection $\pi:\; \c X {\:\longrightarrow\:} [0,1]$ and $x_1, ... x_n$ are coordinates along the fibers. Let $\frac{d}{d x_i}$ be the corresponding vector fields. To prove that $\eta$ is Killing it suffices to check \eqref{_Killing_Equation_} for $a$, $b$ coordinate vector fields. Since $\eta$ is parallel along $X_t$, we have $\nabla_{\frac{d}{d x_i}} \eta=0$ for $i= 1,..., n$. Thus, it suffices to prove \eqref{_Killing_Equation_} in case $a = \frac{d}{dx_0}$. For appropriate choice of coordinates, $\frac{d}{d x_0}=\eta$; thus, \eqref{_Killing_Equation_} is implied by \begin{equation}\label{_nabla_eta_coo_Equation_} \left(\nabla_\eta \eta,\frac{d}{d x_i}\right) =0, \;\; i= 0,\dots n. \end{equation} Since Levi-Civita connection is compatible with the metrics, we have \begin{equation}\label{_Levi_Civi_compa_Equation_} D_\eta \left(\eta, \frac{d}{d x_i}\right) = \left(\nabla_\eta \eta,\frac{d}{d x_i}\right) + \left(\eta,\nabla_\eta\frac{d}{d x_i}\right), \end{equation} where $D_\eta$ is the usual (directional) derivative along $\eta$. For $i=0$, this gives \[ D_\eta (\eta, \eta) = (\nabla_\eta \eta,\eta) + (\eta,\nabla_\eta\eta)=0 \] ($\eta$ is parallel, and hence has constant length). This proves \eqref{_nabla_eta_coo_Equation_} for $i=0$. Since $\frac{d}{d x_i}$ are coordinate vector fields they commute, and therefore, $\nabla_\eta\frac{d}{d x_i}= -\nabla_{\frac{d}{d x_i}}\eta=0$ for $i>0$. Thus, the equation \eqref{_Levi_Civi_compa_Equation_} for $i>0$ is reduced to \[ D_\eta \left(\eta, \frac{d}{d x_i}\right) = \left(\nabla_\eta \eta,\frac{d}{d x_i}\right). \] Since the vector fields $\frac{d}{d x_i}$, $i>0$ are tangent to the fibration $\pi$, the function $(\eta, \frac{d}{d x_i})$ is identically zero. This proves \eqref{_nabla_eta_coo_Equation_} for $i>0$. \ref{_integra_field_Killing_Lemma_} and consequently \ref{_conne_in_fami_of_comple_geo_Proposition_} (i) is proven. \blacksquare \hfill To prove \ref{_conne_in_fami_of_comple_geo_Proposition_} (ii) we have to show that the connection of \ref{_conne_in_fam_of_comple_geode_Proposition_} can be integrated for all smooth points of $\c X$. We give a sketch of a simple geometric argument. By (i), the maps $\Psi_t$ are isometries. Thus, the distance from the given point to the singular set of $X_t$ is invariant under the maps $\Psi_t$. To integrate the connection, we write a tangent vector field which we subsequently integrate. Since the normal field $\eta_t$ is parallel, this tangent field is uniformly bounded. On a certain distance from the singular set $Sing(X_t)$, depending on this uniform bound, the connection might be always integrated. In more precise terms, for all $\epsilon>0$, all $x\in X_t$, with the distance between $x$ and $Sing(X_t)$ no less than $\epsilon$, there exist $\delta>0$ and a map $\Psi^t_{t+\delta}:\;X_t {\:\longrightarrow\:} X_{t+\delta}$ defined in a neighbourhood of $x$ which integrates the connection. {}From this statement and \ref{_conne_in_fami_of_comple_geo_Proposition_} (i), \ref{_conne_in_fami_of_comple_geo_Proposition_} (ii) follows directly. Finally, to prove \ref{_conne_in_fami_of_comple_geo_Proposition_} (iii) we notice that $X_t$ is completely geodesic in $M$ for all $t$. Thus, the completion of $X_t$ as a metric space coinsides with the closure of $X_t$ in $\bar M$. Every isometry of metric spaces extends to a completion, and thus, $\Psi^t_{t'}$ extends to a closure of $X_t$, $X_{t'}$ in $\bar M$. \blacksquare \section[Hyperholomorphic bundles and completely geodesical embeddings.]{Hyperholomorphic bundles \\and completely geodesical embeddings.} \label{_comple_geode_hyperho_Section_} \hfill In this section, we prove that hyperk\"ahler embeddings are completely geodesic. \subsection{Hyperholomorphic structure on the normal bundle} Let $M$ be a hyperk\"ahler manifold, not necessary compact, and $B$ a vector bundle. Recall that in Section \ref{_basics_Section_}, we defined {\bf hyperholomorphic connections} in $B$ (\ref{_hyperho_conne_Definition_}). These are connections $\nabla$ such that the curvature $\Theta:=\nabla^2\in \Lambda^2(M,End(B))$ is an $SU(2)$-invariant 2-form with respect to the natural $SU(2)$-action in $\Lambda^2(M)$. The hyperholomorphic connections are always Yang-Mills (\ref{_hyperholo_Yang--Mills_Proposition_}). The Levi--Civita connection in the tangent bundle $TM$ is a prime example of a hyperholomorphic connection (\ref{_tangent_hyperholo_Example_}). \hfill \proposition \label{_NX_hyperholo_Proposition_} Let $M$ be a hyperk\"ahler manifold, not necessary compact, and $X\subset M$ a trianalytic submanifold. Consider the normal bundle $NX$, equipped with a connection $\nabla$ induced from the Levi--Civita connection on $M$. Then $NX$ is hyperholomorphic. \hfill {\bf Proof:} Let $L$ be an induced complex structure $M$. Consider the manifold $(X, L)$ as a complex submanifold of $(M, L)$. The normal bundle $NX = N(X,L)$ has a natural holomorphic structure, which is compatible with the connection $\nabla$. Therefore, the curvature $\Theta\in \Lambda^2(X, End(NX))$ is of type $(1,1)$ with respect to the Hodge decomposition defined by $L$. Thus, $\Theta$ is of type $(1,1)$ with respect to any of the induced complex structures. By definition, this means that $\nabla$ is hyperholomorphic. \blacksquare \subsection{Hyperk\"ahler embeddings are completely geodesic} In assumptions of \ref{_NX_hyperholo_Proposition_}, consider the bundle $TM\restrict X$ with metrics and connection induced from $TM$. There is a natural exact sequence of holomorphic vector bundles over $(X, L)$: \[ 0{\:\longrightarrow\:} TX {\:\longrightarrow\:} TM\restrict X {\:\longrightarrow\:} NX {\:\longrightarrow\:} 0 . \] The natural connection in each of these bundles is hyperholomorphic (\ref{_NX_hyperholo_Proposition_}). \hfill \proposition \label{_NX_splits_for_hype_Proposition_} Let $X$ be a hyperk\"ahler manifold, not necessary compact, $L$ induced complex structure, and \begin{equation}\label{_holo_seque_Equation_} 0{\:\longrightarrow\:} E_1{\:\longrightarrow\:} E_2 {\:\longrightarrow\:} E_3{\:\longrightarrow\:} 0 \end{equation} be an exact sequence of holomorphic vector bundles. Let $g_2$ be a Hermitian structure on $E_2$. Consider the metrics $g_1$, $g_2$, $g_3$ and the connections $\nabla_1$, $\nabla_2$, $\nabla_3$ on $E_1$, $E_2$, $E_3$ induced by $g_2$. Assume that $\nabla_1$, $\nabla_2$, $\nabla_3$ are hyperholomorphic. Then the exact sequence \eqref{_holo_seque_Equation_} splits, and moreover, the orthogonal decomposition $E_2 = E_1 \oplus E_1^\bot$ is preserved by the connection $\nabla_2$. \hfill {\bf Proof:} The same statement is well known for the Yang--Mills connections: every exact sequence of holomorphic bundles with compatible Yang--Mills metrics splits (see Appendix to this section). By \ref{_hyperholo_Yang--Mills_Proposition_}, hyperholomorphic connections are always Yang--Mills. \blacksquare \hfill \definition Let $N \stackrel i \hookrightarrow M$ be an embedding of hyperk\"ahler manifolds. We say that $i$ is a {\bf hyperk\"ahler embedding} if $i$ is compatible with the quaternionic structure and Riemannian metric. \hfill \corollary\label{_hype_embe_comple_geode:Corollary_} Let $N \stackrel i \hookrightarrow M$ be a hyperk\"ahler embedding. Then $i$ is completely geodesic. {\bf Proof:} Follows directly from \ref{_comple_geodesi_basi_Proposition_} and \ref{_NX_splits_for_hype_Proposition_} \blacksquare \hfill \subsection{Appendix: every exact sequence of Yang--Mills bundles splits.} \hspace{6mm} \proposition \label{_YM_exact_split_Proposition_} Let $X$ be a K\"ahler manifold, not necessary compact, and \begin{equation}\label{_holo_seque2_Equation_} 0{\:\longrightarrow\:} E_1{\:\longrightarrow\:} E_2 {\:\longrightarrow\:} E_3{\:\longrightarrow\:} 0 \end{equation} be an exact sequence of holomorphic vector bundles. Let $g_2$ be a Hermitian structure on $E_2$. Consider the induced metrics $g_1$, $g_2$, $g_3$. Assume that either $g_1$ or $g_3$ is Yang--Mills. Then the exact sequence \eqref{_holo_seque_Equation_} splits, and moreover, the orthogonal complement $E_1^\bot\subset E_2$ is preserved by the connection $\nabla_2$. \hfill {\bf Proof:} Consider the second fundamental form \[ A\in \Lambda^{0,1}(X, \operatorname{Hom}(E_1,E_3)) \] of the exact sequence \eqref{_holo_seque2_Equation_}. The curvatures $\Theta_i$ of $E_i$ are expressed through $A$ as follows (\cite{_Griffi_Harri_}): \begin{equation}\label{_Theta_1_through_seco_Equation_} \Theta_1 = \Theta_2 \restrict{E_1} + {}^t\bar A\wedge A, \end{equation} \begin{equation}\label{_Theta_3_through_seco_Equation_} \Theta_3 = \Theta_2 \restrict{E_3} - A\wedge{}^t\bar A. \end{equation} The 2-forms ${}^t\bar A\wedge A$, $A\wedge{}^t\bar A$ are {\bf positive} unless $A=0$. Thus, the endomorphisms $\Lambda \left(A\wedge{}^t\bar A\right)\in End(E_1)$, $\Lambda \left({}^t\bar A\wedge A\right)\in End(E_3)$ have positive trace (again, unless $A=0$). By our assumption, $\Lambda (\Theta_2)=0$ and either $\Lambda(\Theta_1)=0$ or $\Lambda(\Theta_3)=0$. Applying $\Lambda$ to both sides of \eqref{_Theta_1_through_seco_Equation_} and \eqref{_Theta_3_through_seco_Equation_}, we obtain that either $\Lambda \left(A\wedge{}^t\bar A\right)=0$ or $\Lambda \left({}^t\bar A\wedge A\right)=0$. Therefore, $A=0$ and the exact sequence \eqref{_holo_seque2_Equation_} splits. \blacksquare \section[Triholomorphic sections of hyperholomorphic bundles and deformations of trianalytic submanifolds.] {Triholomorphic sections \\of hyperholomorphic bundles\\ and deformations of trianalytic submanifolds.} \label{_triholo_Section_} In the previous section, we proved that hyperk\"ahler embeddings are completely geodesic. In this section, we show that, furthermore, the deformational results of Section \ref{_comple_geode_defo_Section_} are fully applicable to the deformations of trianalytic submanifolds. \subsection{Triholomorphic sections of normal bundle} \hfill \definition Let $M$ be a hyperk\"ahler manifold, $B$ a vector bundle equipped with a hyperholomorphic connection, and $\alpha$ a section of $B$. Then $\alpha$ is called {\bf triholomorphic} if for each induced complex structure $L$ on $M$, $\alpha$ is a holomorphic section of $B$ considered as a holomorphic bundle over $(M, L)$. \hfill Let $M$ be a hyperk\"ahler manifold, not necessarily compact, and $X\subset M$ a trianalytic submanifold, not necessarily closed. Fix an induced complex structure $I$ on $M$. Let \[ \left( \pi:\; \c X {\:\longrightarrow\:} S, \phi:\; \c X {\:\longrightarrow\:} M, \phi(\pi^{-1}(s_0)) = X \right) \] be a family of deformations of $(X,I) \subset (M, I)$ (see Section \ref{_comple_geode_defo_Section_} for details). Let $\gamma:\; [0,1] {\:\longrightarrow\:} S$ be a real analytic path in $S$, such that $\gamma(0) = s_0$. Assume that for all $t\in [0,1]$, the submanifold $X_t = \phi(\gamma^{-1}(t))\subset M$ is trianalytic. Consider the normal bundle $NX$ to $X$, with the metric and connection induced from $M$. By \ref{_NX_hyperholo_Proposition_}, $NX$ is hyperholomorphic. Let $\eta$ be the section of $NX$ corresponding to $\frac{d\gamma}{dt}$ as in Subsection \ref{_norma_vecto_Subsection_}. \hfill \proposition \label{_norma_triholo_Proposition_} In the above assumptions, $\eta$ is a triholomorphic section of $NX$. \hfill {\bf Proof:} Let $L$ be an induced complex structure, and $(M, L)$, $(X, L)$ the manifolds $M$ and $X$ considered as complex manifolds with the complex structure $L$. The normal bundle $N(X,L)$ is naturally identified with $NX$ as a real vector bundle. Therefore $\eta$ can be considered as a section of $N(X,L)$. By \ref{_real_ana_indu_on_hype_equiva_Proposition_} (see Appendix to this section), $\c X$ can be considered as a real analytic deformation of $(X, L)$. {}From \ref{_norma_sec_holom_Lemma_} it is clear that $\eta$ is holomorphic as a section of $N(X,L)$. This proves \ref{_norma_triholo_Proposition_}. \blacksquare \subsection{Triholomorphic sections are parallel} \hfill \proposition\label{_triholo_parallel_Proposition_} Let $M$ be a hyperk\"ahler manifold, not necessary compact, and $B$ a vector bundle with a hyperholomorphic connection $\nabla$. Let $\nu$ be a trianalytic section of $B$. Then, $\nu$ is parallel: \begin{equation}\label{_nu_para_Equation_} \nabla\nu=0 \end{equation} {\bf Proof:} Let $L$ be an induced complex structure. Since $\nu$ is triholomorphic, $\bar \partial_L \nu =0$, where $\bar \partial_L:\; B {\:\longrightarrow\:} B\times \Lambda^{0,1}_L(M)$ is the $(0,1)$-part of the connection, taken with respect to $L$. Taking $L=I, -I$, we obtain \begin{equation} \label{_bar6_I+bar6_-I_Equation_} \bar \partial_I + \bar\partial_{-I}(\nu)=0. \end{equation} On the other hand, $\bar\partial_{-I}= \partial_I$, where $\partial_I$ is the $(1,0)$-part of $\nabla$ taken with respect to $I$. Thus, \[ \bar \partial_I + \bar\partial_{-I} =\bar \partial_I + \partial_{I} =\nabla. \] {}From \eqref{_bar6_I+bar6_-I_Equation_} we obtain that $\nabla(\nu)=0$. \blacksquare \hfill {}From \ref{_norma_triholo_Proposition_} and \ref{_triholo_parallel_Proposition_}, we obtain that the section $\eta \in NX$ is parallel with respect to the connection. Then, \ref{_conne_in_fami_of_comple_geo_Proposition_} (i) can be applied to the following effect. \hfill \corollary \label{_norma_sec_para_for_hype_from_it_Corollary_} Let $M$ be a hyperk\"ahler manifold, not necessarily compact, and $X\subset M$ a trianalytic submanifold, not necessarily closed, and \[ \left( \pi:\; \c X {\:\longrightarrow\:} S, \phi:\; \c X {\:\longrightarrow\:} M, \phi(\pi^{-1}(s_0)) = X \right) \] be a real analytic family of deformations of $X\subset M$ (see Section \ref{_comple_geode_defo_Section_} for details). Let $\gamma:\; [0,1] {\:\longrightarrow\:} S$ be a real analytic path in $S$, such that $\gamma(0) = s_0$. Assume that for all $t\in [0,1]$, the submanifolds $X_t = \phi(\gamma^{-1}(t))\subset M$ are trianalytic. Fix an induced complex structure $I$ on $M$. Let $U_t\subset (X_t, I)$, $t\in [0,1]$ be the subsets constructed in Subsection \ref{_conne_Subsection_}\footnote{By \ref{_hype_embe_comple_geode:Corollary_}, trianalytic submanifolds are completely geodesic, and thus, \ref{_conne_in_fam_of_comple_geode_Proposition_} can be applied.} with the corresponding holomorphic isomorphisms $\Psi_t:\; U_0{\:\longrightarrow\:} U_t$. Then, assumptions of \ref{_conne_in_fami_of_comple_geo_Proposition_} (i) hold. Thus, for all $t, t'\in [0,1]$, the maps $\Psi^t_{t'}$ are isometries. Moreover, $\Psi^t_{t'}$ are compatible with the hyperk\"ahler structure. \hfill {\bf Proof:} Assumptions of \ref{_conne_in_fami_of_comple_geo_Proposition_} (i) hold by \ref{_norma_triholo_Proposition_} and \ref{_triholo_parallel_Proposition_}. The maps $\Psi^t_{t'}$ are isometries by \ref{_conne_in_fami_of_comple_geo_Proposition_} (i). The maps $\Psi^t_{t'}$ are compatible with the hyperk\"ahler structure because the they are obtained by integrating a certain connection in the family $\c X$. This connection is constructed from the Hermitian metric, and thus, does not depend from the choice of induced complex structure. Taking different induced complex structures, we obtain the same maps $\Psi^t_{t'}$. Thus, $\Psi^t_{t'}$ is holomorphic with respect to each of induced complex structures (see also \ref{_norma_sec_holom_Lemma_}). \blacksquare \subsection{Appendix: real analytic structures on hyperk\"ahler manifolds} Consider the real analytic structures on a given hyperk\"ahler manifold arising from the different induced complex structures. We prove that these real analytic structures are equivalent. \hfill \proposition \label{_real_ana_indu_on_hype_equiva_Proposition_} Let $M$ be a hyperk\"ahler manifold, $I_1$, $I_2$ induced complex structures. Let $(M, I_1)$ and $(M, I_2)$ be the corresponding complex manifolds, and $(M, I_1)_{\Bbb R}$, $(M, I_2)_{\Bbb R}$ be the real analytic manifolds underlying $(M, I_1)$, $(M, I_2)$. Consider the tautological map $(M, I_1)_{\Bbb R}\stackrel \phi {\:\longrightarrow\:} (M, I_2)_{\Bbb R}$. Then $\phi$ is compatible with the real analytic structure. \hfill {\bf Proof:} Consider the {\bf twistor space} for $M$ (see \cite{_Besse:Einst_Manifo_}), $\operatorname{Tw}(M)$, with the natural holomorphic map $\pi:\; \operatorname{Tw}(M) {\:\longrightarrow\:} {\Bbb C} P^1$. Let $Sec(M)$ be the space of holomorphic sections of the map $\pi$. Then $Sec(M)$ is identified naturally with an open subspace of a Douady space for $\operatorname{Tw}(M)$, and thus, has a natural complex structure. According to \cite{_HKLR_} (see \cite{_NHYM_} for details), $Sec(M)$ is a complexification of $(M,I)$, in the sense of Grauert. In other words, the complex valued real analytic functions on $(M,I)$ are naturally identified with the germs of complex analytic functions on $Sec(M)$. Since $Sec(M)$ is defined independently from the choice of an induced complex structure, the tautological map $\phi$ is an equivalence. This proves \ref{_real_ana_indu_on_hype_equiva_Proposition_}. \blacksquare \hfill \ref{_real_ana_indu_on_hype_equiva_Proposition_} implies that we may speak of a real analytic manifold underlying a given hyperk\"ahler manifold. \hfill \corollary \label{_Riema_on_hype_real_ana_Corollary_} Let $M$ be a hyperk\"ahler manifold. Consider the Riemannian form $g$ on $M$ as a section of the real analytic bundle of symmetric 2-forms. Then $g$ is real analytic. \hfill {\bf Proof:} Let $I, J, K$ be the induced complex structures which form the standard basis in quaternions, and $\omega_I$, $\omega_J$, $\omega_K$ be the corresponding K\"ahler forms. Then $\Omega:=\omega_J + \sqrt{-1}\:\omega_K$ is the natural holomorphically symplectic form on $(M,I)$, and as such, $\Omega$ is real analytic. Then, its real part $\omega_J$ is also real analytic. Since the complex structure operator $J$ is real analytic, we obtain that the form $g(\cdot,\cdot):= - \omega(\cdot, J\cdot)$ is also real analytic. \blacksquare \section{Douady spaces for trianalytic cycles and real analytic structure} \label{_Douady_Section_} \subsection{Real analytic structure on the Douady space.} In this section, we consider the Douady space $D_M(X)$ for a subvariety $X$ of a compact complex manifold $M$ equipped with a hyperk\"ahler structure. We prove that, when $X$ is trianalytic, $D_M(X)$ is {\bf hypercomplex} (\ref{_hypercomplex_Definition_}). We also show that the real analytic variety underlying $D_M(X)$ does not change if we replace a complex structure on $M$ by another induced complex structure. These results are technical and we use them mainly to simplify the exposition. \hfill Let $M$ be a compact hyperk\"ahler manifold, and $X\subset M$ a closed trianalytic subvariety. For each induced complex structure $L$, consider $(X, L)$ as a complex subvariety of $(M, L)$. Let $D_L(X)$ be the Douady deformation space of $(X, L)$ in $(M, L)$. The points $[X']$ of $D_L(X)$ correspond to the subvarieties $X'\subset (M, L)$ which are deformation of $X$. By \ref{_G_M_invariant_implies_trianalytic_Theorem_} (see also \cite{Verbitsky:Symplectic_II_}), every such $X'$ is trianalytic. Let $I_1$, $I_2$ be induced complex structures on $M$. The following proposition shows that a trianalytic subvariety $X'$ can be obtained as a deformation of $(X, I_1)$ in $(M, I_1)$ if and only if $X'$ can be obtained as a deformation of $(X, I_2)$ in $(M, I_2)$. \hfill \proposition \label{_X_in_Douady_indep_from_I_1_Proposition_} Let $I_1$, $I_2$ be induced complex structures on a compact hyperk\"ahler manifold $M$, and $X\subset M$ a closed trianalytic subvariety. Consider the corresponding Douady spaces $D_{I_1}(X)$, $D_{I_2}(X)$. Then, for a trianalytic subvariety $X'\subset M$, $[X'] \in D_{I_1}(X)$ if and only if $[X']\in D_{I_2}(X)$. \hfill {\bf Proof:} Let us recall the notion of {\bf degree} of a trianalytic subvariety. Let $M_1$ be a K\"ahler manifold. By {\bf degree} of a subvariety $X \subset M_1$ we understand a number \[ \deg X := \int_X \omega^{\dim_{\Bbb C} X}, \] where $\omega$ is the K\"ahler form. If $X\subset M$ is a trianalytic subvariety of a hyperk\"ahler manifold, we can associate a number $\deg X$ to each of induced complex structures. In \cite{Verbitsky:Symplectic_I_}, we prove that $\deg X$ is in fact independent from the choice of induced complex structure. This enables us to speak of {\bf degree} of a closed trianalytic subvariety of a compact complex manifold. We return to the proof of \ref{_X_in_Douady_indep_from_I_1_Proposition_}. Let $D_1$ be the union of all components of the Douady space for $(M,I_1)$. By \cite{_Lieberman_}, \cite{_Fukjiki_Kahler_}, $D_1$ is compact; in particular, $D_1$ has a finite number of connected components. Let $D_1 = \cup D^i_1$, $i\in \Upsilon$ be the decomposition of $D$ unto a union of its connected components. Let $X'\subset M$ be a closed trianalytic subvariety of $M$ such that $[X'] \in D_{I_2}(X)$. There exists a real analytic path $\gamma:\; [0,1] {\:\longrightarrow\:} D_{I_2}(X)$ joining $[X]$ and $[X']$. For all $t\in [0,1]$, the point $\gamma(t)$ lies in $D_1^i$ for some $i\in \Upsilon$. Let $S^i\subset [0,1]$ be the set of all $t\in [0,1]$ such that $\gamma(t)\in D_1^i$. We are going to show that $S^i$ are compact for all $i$. This will clearly imply that for all $i$ except one, $S^i$ is empty, thus proving that $[X]$ and $[X']$ lie in the same component of $D_1$. Let $D_{\Bbb R}$ be the set of all real analytic deformations of $X$ in $M$ with natural topology. The forgetful map $\psi:\; D_1 {\:\longrightarrow\:} D_{\Bbb R}$ is continous and injective. Since $D_1$ is compact, $\psi$ is a closed embedding. The composition $\gamma\circ \psi:\; [0,1] {\:\longrightarrow\:} D_{\Bbb R}$ is obviously continous, and thus, $\gamma:\; [0,1] {\:\longrightarrow\:} D_1$ is also continous. This implies that all $S_i$, $i\in \Upsilon$ are closed, thus proving that all $S_i$ except one are empty. \ref{_X_in_Douady_indep_from_I_1_Proposition_} is proven. \blacksquare \hfill Let $D_{I_1}(X)_{\Bbb R}$, $D_{I_2}(X)_{\Bbb R}$ be the real analytic varieties underlying $D_{I_1}(X)$, $D_{I_2}(X)$. \ref{_X_in_Douady_indep_from_I_1_Proposition_} gives a tautological bijection \[ \psi:\; D_{I_1}(X)_{\Bbb R}{\:\longrightarrow\:} D_{I_2}(X)_{\Bbb R}.\] \hfill \proposition\label{_Douady_real_ana_inde_from_indu_Proposition_} The map $\psi:\; D_{I_1}(X)_{\Bbb R}{\:\longrightarrow\:} D_{I_2}(X)_{\Bbb R}$ is an isomorphism of real analytic varieties. \hfill {\bf Proof:} Let $I$ be an induced complex structure on $M$. For a real analytic function $f$ on $M$, consider the function $\hat f:\; D_I(X) {\:\longrightarrow\:} {\Bbb R}$, \[ [X'] \stackrel{\hat f}{\:\longrightarrow\:} \int_{X'} f \cdot \operatorname{Vol} X', \] with $\operatorname{Vol} X'$ the volume form on $X'$. For all $f$, the function $\hat f$ is real analytic. From the definition of Douady spaces it might be seen that converging power series of different $\hat f$ generate the sheaf $\c A_I$ of real analytic functions on $D_I(X)$. For an induced complex structure $I'$, the same set $\{ \hat f\}$ generates the sheaf $\c A_{I'}$ of real analytic functions on $D_{I'}(X)$ (\ref{_real_ana_indu_on_hype_equiva_Proposition_}). Thus, the sheaves $\c A_I$ and $\c A_{I'}$ coinside. \blacksquare \hfill \ref{_Douady_real_ana_inde_from_indu_Proposition_} shows that we may speak of the real analytic variety underlying $D_M(X)$ without specifying an induced complex structure $I$. \hfill \definition \label{_hypercomplex_Definition_} Let $Y$ be a real analytic variety equipped with three complex structures $I$, $J$ and $K$. Assume that for every point $y\in Y$, the action of $I$, $J$, $K$ on the Zariski tangent space $T_yY$ satisfies $I\circ J = - J\circ I =K$. Then $Y$ is called {\bf a hypercomplex variety}. \hfill \remark \label{_Douady_hyperc_Remark_} \ref{_Douady_real_ana_inde_from_indu_Proposition_} immediately implies that the induced complex structures on $M$ equip the Douady space $D_M(X)$ with a hypercomplex structure. \subsection{Appendix: isometric embeddings are hyperk\"ahler.} Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a closed trianalytic subvariety. There is an alternative way to describe the Douady space $D(X)$. The \ref{_isome_embe_Proposition_} below describes $D(X)$ in terms of of the space of isometric embeddings $\nu:\; X {\:\longrightarrow\:} M$. \hfill Let $M$ be a compact K\"ahler manifold, and $X\subset M$ be a closed analytic subvariety. One can make sense of the {\bf fundamental class} of $X$, $[X] \in H^{\dim_{\Bbb R}(X)}(M)$, lying in homology of $M$. Consider $X$ with an induced structure of a metric space, and let $X \stackrel \nu {\:\longrightarrow\:} M'$ be an isometric embedding, where $M'$ is a compact Riemannian manifold. The same argument which allows us to define the fundamental class of $X$, allows us to define the fundamental class of $\nu(X)$. \hfill \proposition \label{_isome_embe_Proposition_} Let $M$ be a closed analytic subvariety, $X\subset M$ a closed trianalytic subvariety, and $\nu:\; X {\:\longrightarrow\:} M$ an isometric embedding. Assume that the fundamental class of $\nu(X)$ is equal to the fundamental class of $X$. Then $\nu(X)$ is a trianalytic subvariety of $M$. \hfill {\bf Proof:} We use notation introduced in Subsection \ref{_SU(2)-inv=>triana_Subsection_}. For each induced complex structure $I$, we have \begin{equation}\label{_deg_same_appl_nu_Equation_} \deg_I (X) - \deg_I (\nu(X)), \end{equation} because $\nu(X)$ has the same fundamental class as $X$. Since $X$ is trianalytic, applying Wirtinger's inequality (\ref{_Wirti_for_Kahle_Theorem_}), we obtain that \begin{equation} 2^n \int_{X_{ns}} \operatorname{Vol} X_{ns} = \int_{X_{ns}} \omega^n, \end{equation} where $\omega$ is the K\"ahler class of $(M,I)$, $n=\frac{1}{2}\dim_{\Bbb R}(X)$ and $X_{ns}$ is the nonsingular part of $(X,I)$. By definition, we have \begin{equation} \int_{X_{ns}} \omega^n = \deg_I (X), \ \ \int_{\nu(X_{ns})} \omega^n = \deg_I (\nu(X)). \end{equation} Since $\nu$ is an isometry, and $\operatorname{Vol}(X)$ is an invariant of a metric, we have \begin{equation} \label{_Vol_same_appl_nu_Equation_} \int_{X_{ns}} \operatorname{Vol} X_{ns} = \int_{\nu(X_{ns})} \operatorname{Vol} \left(\nu(X_{ns}) \right). \end{equation} Together the equations \eqref{_deg_same_appl_nu_Equation_}--\eqref{_Vol_same_appl_nu_Equation_} give \[ 2^n\int_{X_{ns}} \operatorname{Vol} X_{ns} = \int_{\nu(X_{ns})} \omega^n. \] Applying Wirtinger's inequality (\ref{_Wirti_for_Kahle_Theorem_}) once again, we obtain that $\nu(X)$ is complex analytic with respect to $I$. This proves \ref{_isome_embe_Proposition_}. \blacksquare \hfill We just proved \ref{_iso_intro:Theorem_} (iii). As another application of \ref{_isome_embe_Proposition_}, we give a direct proof that $D(X)$ is compact. \hfill \corollary Let $M$ be a compact hyperk\"ahler manifold, $X\stackrel {\bar i}\hookrightarrow M$ be a closed trianalytic subvariety and $D(X)$ its Douady space. Then $D(X)$ is compact. \hfill {\bf Proof:} Let $X_{ns}$ be the non-singular part of $X$, and $i:\; X_{ns} \hookrightarrow M$ the natural embedding. Then $i$ is an isometry. Consider a deformation $\nu$ of $i$ in class of isometries. The arument proving \ref{_isome_embe_Proposition_} shows that the closure of $\nu(X_{ns})$ is trianalytic in $M$. Consider the space $\underline{Is(X_{ns})}$ of isometries $X_{ns}\stackrel \nu\hookrightarrow M$. Let $Is(X_{ns})$ be a connected component of this space, containing $i$. There is a natural continuous surjection $p:\; Is(X_{ns}) {\:\longrightarrow\:} D(X)$, which maps $\nu \in Is(X_{ns})$ to a closure of $\nu(X_{ns})$ in $M$. To prove that $D(X)$ is compact it suffices to show that $Is(X_{ns})$ is compact. This is an implication of the following general statement, which is clear. \hfill \claim Let $X$, $M$ be Riemannian manifolds, $M$ compact. Let $Is(X, M)$ be the space of isometries (maps which preserve the geodesic distance) from $X$ to $M$. Then $Is(X, M)$ is compact. \blacksquare \section[Connections in the families of trianalytic subvarieties] {Connections in the families of trianalytic \\subvarieties} \label{_Conne_in_fami_Section_} \subsection{Introduction} \label{_intro_conne_Subsection_} Let $M$ be a compact hyperk\"ahler manifold, and \begin{equation} \label{_fami_Equation_} \left( \pi:\; {\c X} {\:\longrightarrow\:} S, \phi:\; {\c X} {\:\longrightarrow\:} M \right ) \end{equation} a real analytic family of subvarieties, not necessarily closed. Assume that for all $s\in S$, the fiber $X_s = \phi(\pi^{-1} (s))$ is trianalytic in $M$. By \ref{_conne_in_fam_of_comple_geode_Proposition_} and \ref{_hype_embe_comple_geode:Corollary_}, the family ${\c X}$ is then equipped with a connection $\nabla:\; N_\pi X {\:\longrightarrow\:} T {\c X}$. It is natural to ask whether this connection is {\bf flat} (see \ref{_flat_conne_in_fami_Definition_}). The answer is affirmative, under certain additional assumptions. \hfill \definition \label{_admits_co_Definition_} We say that the family \eqref{_fami_Equation_} {\bf admits a compactification} if the following conditions hold. \begin{description} \item[(i)] For each $s \in S$, the closure $\bar X_s$ of the fiber $X_s$ in $M$ is a trianalytic subvariety of $M$. \item [(ii)] For all $s\in S$, $\bar X_s$ lie in the same component of the Douady space. \end{description} \hfill We show that in assumptions of \ref{_admits_co_Definition_}, the connection \[ \nabla:\; N_\pi X {\:\longrightarrow\:} T {\c X}\] is indeed flat (\ref{_conne_in_triana_flat_Theorem_}). However, we don't know whether it is flat for general families of trianalytic submanifolds. \hfill \subsection{Flat connections and curvature.} \hfill \definition (curvature of a connection) Let $\pi: {\c X} {\:\longrightarrow\:} S$ be a family of manifolds, $T_\pi X \subset T{\c X}$ be the fiberwise tangent bundle and $N_\pi X = T{\c X} / T_\pi X$ be a fiberwise normal bundle. Let $\nabla:\; N_\pi X {\:\longrightarrow\:} T{\c X}$ be a connection in ${\c X}$. Then the {\bf curvature} of $\nabla$ is the following tensor $\Theta\in \operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \left( N_\pi X\right), T_\pi{\c X}\right)$. For two sections $a,b \in N_\pi X$, consider the corresponding vector fields $\nabla a$, $\nabla b\in T{\c X}$. Consider the projection $T {\c X} \stackrel {pr}{{\:\longrightarrow\:}} T{\c X} / \nabla N_\pi X = T_\pi X$. Let $\Theta(a,b):= pr([\nabla a, \nabla b])$. Clearly, $\Theta$ is a tensor. \hfill \proposition Let $\left( \pi:\; {\c X} {\:\longrightarrow\:} S \right)$ be a family of manifolds equipped with a connection $\nabla$. Then the following conditions are equivalent. \begin{description} \item[(i)] The connection $\nabla$ is flat. \item[(ii)] Let $s_1, s_2\in S$ be a pair of points, and $\gamma, \gamma':\; [0,1] {\:\longrightarrow\:} S$ real analytic paths. Let $U_0 \subset X_{s_0}$, $U_1 \subset X_{s_1}$ be the open subsets such that the connection $\nabla$ might be integrated to a map $\Psi:\; U_{s_0} {\:\longrightarrow\:} U_{s_1}$ along $\gamma$, and $U_0' \subset X_{s_0}$, $U_1' \subset X_{s_1}$ be such that $\nabla$ might be integrated to a map $\Psi':\; U_{s_0}' {\:\longrightarrow\:} U_{s_1}'$ along $\gamma'$. Then $\Psi$ coinsides with $\Psi'$ in the intersection $U_{s_0} \cap U'_{s_0}$. \end{description} {\bf Proof:} Well known. \blacksquare \hfill \definition \label{_flat_conne_in_fami_Definition_} Let ${\c X} {\:\longrightarrow\:} S$ be a family of manifolds, equipped with a connection $\nabla$. Then $\nabla$ is called {\bf flat} if its curvature is zero. \subsection{Curvature of holomorphic connections} Let $\pi:\; {\c X} {\:\longrightarrow\:} S$ be a complex analytic family of manifolds. Then the vector bundles $T{\c X}$ and $N_\pi X$ are equipped with a natural holomorphic structure. The connection $\nabla:\; N_\pi X {\:\longrightarrow\:} T {\c X}$ is called {\bf holomorphic} if the map $\nabla:\; N_\pi X {\:\longrightarrow\:} T{\c X}$ is holomorphic. \hfill \claim \label{_holo_conne_curva_C-line_Claim_} Let $\pi:\; {\c X} {\:\longrightarrow\:} S$ be a family of manifolds equipped with with a holomorphic connection $\nabla$. Consider the curvature \[ \Theta\in \operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \left(N_\pi X\right), T_\pi{\c X}\right). \] Then $\Theta$ is ${\Bbb C}$-linear; in other words, $\Theta$ belongs to the subspace \[ \operatorname{Hom}_{\Bbb C}\left ( \Lambda^2_{\Bbb C} \left(N_\pi X\right), T_\pi{\c X}\right) \subset \operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \left(N_\pi X\right), T_\pi{\c X}\right). \] {\bf Proof:} Clear. \blacksquare \subsection{Flat connections in a family of trianalytic varieties. } Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a closed trianalytic subvariety, $D(X)$ its Douady space and \[ \left( \underline \pi: \; \underline{{\c X}} {\:\longrightarrow\:} D(X), \underline \phi:\; \underline{{\c X}} {\:\longrightarrow\:} M \right ) \] a universal family of subvarieties corresponding to $D(X)$. Let ${\c X}$ be the union of all smooth points in all fibers of $\underline \pi$, and $\pi$, $\phi$ be the restrictions of $\underline \pi$ and $\underline \phi$ to ${\c X}$. Consider the corresponding family of manifolds \[ \left( \pi: {{\c X}} {\:\longrightarrow\:} D(X), \phi {{\c X}} {\:\longrightarrow\:} M \right ). \] We are in the same situation as described in Subsection \ref{_intro_conne_Subsection_}. Thus, the family ${\c X}$ is equipped with a natural connection $\nabla$. \hfill \theorem \label{_conne_in_triana_flat_Theorem_} In these assumptions, $\nabla$ is flat. \hfill {\bf Proof:} The proof of \ref{_conne_in_triana_flat_Theorem_} takes the rest of this section. \hfill Clearly, $N_\pi X = \pi^* TS$. By \ref{_Douady_hyperc_Remark_}, $TS$ has a natural quaternionic action, and thus, an action of $SU(2) \subset {\Bbb H}^*$. The bundle $T_\pi X$ is also equipped with a natural action of quaternions. This endows the bundle $\operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \right(N_\pi X\left), T_\pi{\c X}\right)$ with an action of $SU(2)$. \hfill \lemma \label{_curva_SU(2)_inva_Lemma_} Let $\Theta$ be a curvature of $\nabla$, $\Theta \in\operatorname{Hom}_{\Bbb R} \left ( \Lambda^2_{\Bbb R} \left( N_\pi X\right), T_\pi{\c X}\right)$. Then $\Theta$ is $SU(2)$-invariant. \hfill {\bf Proof:} Let $U(1) \stackrel {\tilde i}\hookrightarrow SU(2)$ be an embedding corresponding to an algebra embedding ${\Bbb C} \stackrel i\hookrightarrow {\Bbb H}$. Let $I$ be an induced complex structure associated with $i$. The connection $\nabla$ is holomorphic with respect to $I$. Therefore $\Theta$ is ${\Bbb C}$-linear with respect to the action of ${\Bbb C}$ given by ${\Bbb C} \stackrel i \hookrightarrow {\Bbb H}$ (\ref{_holo_conne_curva_C-line_Claim_}). Thus, $\Theta$ is $U(1)$-invariant with respect to the action of $U(1)$ given by $\tilde i:\; U(1)\hookrightarrow SU(2)$. The group $SU(2)$ is generated by the images of $\tilde i$ for all algebra embeddings $i:\; {\Bbb C} \hookrightarrow {\Bbb H}$. Thus, $\Theta$ is $SU(2)$-invariant. \blacksquare \hfill Consider $N_\pi X$, $T_\pi X$ as representation of $SU(2)$. These representations are of weight 1. The bundle \[ \operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \left(N_\pi X\right), T_\pi{\c X}\right) \] is again a representation of $SU(2)$. Since $N_\pi X$, $T_\pi X$ are of weight one, $\operatorname{Hom}_{\Bbb R}\left ( \Lambda^2_{\Bbb R} \left(N_\pi X\right), T_\pi{\c X}\right)$ is a direct sum of representations of weight 3 and 1. Thus, this bundle cannot have non-zero $SU(2)$-invariant sections. Therefore, ($\Theta$ is $SU(2)$-invariant) implies ($\Theta =0$). \ref{_conne_in_triana_flat_Theorem_} is proven. \blacksquare \section[Isometries of trianalytic subvarieties of a compact hy\-per\-k\"ah\-ler manifold]{Isometries of trianalytic subvarieties \\ of a compact hyperk\"ahler manifold} \label{_isome=>holo_Section_} \subsection{Premises} Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a closed trianalytic subvariety. {}From \ref{_conne_in_fami_of_comple_geo_Proposition_}, \ref{_conne_in_triana_flat_Theorem_} and \ref{_norma_sec_para_for_hype_from_it_Corollary_}, we immediately obtain the following theorem. \hfill \theorem \label{_triana_subse_isome_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a closed trianalytic subvariety, $D(X)$ the corresponding Douady space (\ref{_Douady_hyperc_Remark_}). Let $\gamma:\; [0,1] {\:\longrightarrow\:} D(X)$ be a real analytic path. Denote by $X_t$ the subvarieties corresponding to points $\gamma(t)\in D(X)$. Consider $X_t$ as metric spaces with the metric induced from $M$. Then, for each $t_1$, $t_2\in [0,1]$, there exist a natural isometry $\Psi^{t_1}_{t_2}:\; X_{t_1} {\:\longrightarrow\:} X_{t_2}$, mapping non-singular points to non-singular points, and acting compatible with the hyperk\"ahler structure on $X_{t_1}^{ns}$.\footnote{The non-singular part $X_t^{ns}$ is naturally equipped with a structure of a hyperk\"ahler manifold.} This isometry depends only on the homotopy class of $\gamma$. \blacksquare \hfill This almost finishes the proof \ref{_iso_intro:Theorem_} (i). It remains to prove the following theorem. \hfill \theorem \label{_triana_subse_comple_ana_Theorem_} In assumptions of \ref{_triana_subse_isome_Theorem_}, let $I$ be an induced complex structure on $M$. Then the isometry $\Psi^{t_1}_{t_2}:\; X_{t_1} {\:\longrightarrow\:} X_{t_2}$ is compatible with the complex analytic structure induced by $I$. \hfill This section is taken fully with the proof of \ref{_triana_subse_comple_ana_Theorem_} \subsection{Homeomorphisms of complex varieties and normalization.} \hfill \claim \label{_conti_fu_extend_to_holo_Claim_} (\cite{_Grauert_}) Let $X$ be a normal complex analytic variety, $U$ a dense open subset in $X$ and $f:\; U {\:\longrightarrow\:} {\Bbb C}$ a bounded holomorphic function. Then $f$ can be extended to a holomorphic function on $X$. \blacksquare \hfill This statement has an immediate corollary. \hfill \corollary \label{_conti_map_norma_holo_Corollary_} Let $\phi:\; X {\:\longrightarrow\:} Y$ be a continous map of complex analytic varieties. Assume that $\phi$ is holomorphic in an open dense subset $U\subset X$. Then, if $X$ is normal, $\phi$ is holomorphic. \hfill {\bf Proof:} Let $x\in X$ be an aritrary point, $V$ its neighbourhood, sufficiently small. Taking coordinates in a neighbourhood $W$ of $\phi(V)$ and applying \ref{_conti_fu_extend_to_holo_Claim_}, we find that $\phi\restrict {U\cap V}$ extends continously to a holomorphic map $\tilde \phi$ from $U\cap V$ to $W$. Since $U\cap V$ is dense in $V$, $\tilde \phi$ coinsides with $\phi\restrict V$. Thus, $\phi$ is holomorphic. \blacksquare \hfill Thus, in the situation of \ref{_triana_subse_comple_ana_Theorem_}, were $X_t$ normal, $\Psi^{t}_{t'}$ would have been holomorphic and \ref{_triana_subse_comple_ana_Theorem_} would have been proven. Unfortunately, we have no means to show that $X_t$ is normal. However, from \ref{_conti_map_norma_holo_Corollary_} we obtain some information about maps of arbitrary varieties too. \hfill \corollary \label{_conti_map_mero_Corollary_} Let $\phi:\; X {\:\longrightarrow\:} Y$ be a continous map of complex analytic varieties, which is holomorphic on an open dense subset $U\subset X$. Then $\phi$ is meromorphic. \hfill {\bf Proof:} Take a normalization $\tilde X \stackrel n {\:\longrightarrow\:} X$. Applying \ref{_conti_map_norma_holo_Corollary_} to the composition $n\circ \phi:\; \tilde X {\:\longrightarrow\:} Y$, we obtain that $n\circ \phi$ is holomorphic. Thus, $\phi$ is meromorphic as a composition of holomorphic $n\circ \phi$ and a meromorphic map $n^{-1}$. \blacksquare \hfill Applying \ref{_conti_map_mero_Corollary_} to the situation of \ref{_triana_subse_comple_ana_Theorem_}, we obtain the following corollary. \hfill \corollary \label{_Psi_bimero_Corollary_} In the situation of \ref{_triana_subse_isome_Theorem_}, the map \[ \Psi^{t_1}_{t_2}:\; X_{t_1} {\:\longrightarrow\:} X_{t_2} \] is bimeromorphic for each induced complex structures. \blacksquare \hfill It remains to make a leap from ``bimeromorphic'' to ``holomorphic''. This is done in two steps. We prove a number of algebro-geometric statements about the behaviour of $\Psi^{t_1}_{t_2}$, concluding with \ref{_pi_i_properties_Proposition_}. In \ref{_alge_geo_suffi_pi_i_iso_Proposition_}, we show that these statements are strong enough to show that the $\Psi^{t_1}_{t_2}$ is holomorphic. The premise of \ref{_alge_geo_suffi_pi_i_iso_Proposition_} is purely algebro-geometric and its proof is independent from the rest of this section. We finish this Subsection with the following statement, which we use in \ref{_pi_i_properties_Proposition_}. \hfill \proposition \label{_Psi_iso_on_norma_Proposition_} The map $\Psi^{t_1}_{t_2}:\; X_{t_1} {\:\longrightarrow\:} X_{t_2}$ induces an isomorphism of normalizations \[ \tilde \Psi^{t_1}_{t_2}:\;\tilde X_{t_1} {\:\longrightarrow\:} \tilde X_{t_2}. \] {\bf Proof:} From the definition of normalization \cite{_Grauert_}, the following lemma is evident. \hfill \lemma Let $\Psi:\; X {\:\longrightarrow\:} Y$ be a homeomorphism of complex varieties. Assume that in an open dense subset $U\subset X$, $\Psi$ is holomorphic. Then $\Psi$ induces an isomorphism $\tilde \Psi:\; \tilde X {\:\longrightarrow\:} \tilde Y$ of normalizations, if the following statement holds. \begin{description} \item[(*)] There exist a Stein covering $\{ U_i\}$ of $X$ such that $\{ \Psi(U_i)\}$ is a Stein covering for $Y$. \end{description} \blacksquare To prove \ref{_Psi_iso_on_norma_Proposition_}, it remains to show that the property {\bf (*)} holds for the map $\Psi^{t_1}_{t_2}:\; X_{t_1} {\:\longrightarrow\:} X_{t_2}$. For each point $x \in X_{t_1}$, it suffices to construct a Stein neighbourhood $U$ of $X_{t_1}$, such that $\Psi^{t_1}_{t_2}(U)$ is Stein. For a K\"ahler variety, consider an open ball $B$ of radius $r$, taken with respect to the metric defined by geodesics. Then, for $r$ sufficiently small, $B$ is Stein (\cite{_Greene_}). Since $\Psi^{t_1}_{t_2}$ is an isometry (\ref{_triana_subse_isome_Theorem_}), an image of an open ball of radius $r$ is again an open ball of radius $r$. This gives a system of Stein neighbourhoods satisfying {\bf (*)}. \ref{_Psi_iso_on_norma_Proposition_} is proven. \blacksquare \subsection{Homeomorphisms of completely geodesic subvarieties induce isomorphisms of Zariski tangent spaces.} \hfill \proposition\label{_homeo_indu_iso_Zariski_Proposition_} Let $M_1$, $M_2$ be a K\"ahler manifolds, $X_1^{ns}\subset M_1$, $X_2^{ns}\subset M_2$ be completely geodesic complex submanifolds, not necessarily closed, and $X_1$, $X_2$ be the closures of $X_1^{ns}$, $X_2^{ns}$ in $M_1$, $M_2$. Assume that $X_1$, $X_2$ are complex analytic subvarieties of $M_1$, $M_2$. Let $\phi:\; X_1{\:\longrightarrow\:} X_2$ be a morphism of complex varieties, such that $\phi\restrict{X_1^{ns}}$ is an isometry. Then $\phi$ induces an isomorphism of Zariski tangent spaces. \hfill {\bf Proof:} To prove \ref{_homeo_indu_iso_Zariski_Proposition_}, we interpret the Zariski tangent space $T_x X_1$ in terms of the metric structure on $X_i^{ns}$, where $x$ is a point of $X_1$. Let $\gamma:\; [0,1] {\:\longrightarrow\:} X_1$ be a path satisfying $\gamma\left([0,1] \backslash\{0.5\}\right) \subset X_1^{ns}$, $\gamma(0.5) =x$. There is a natural topology on the total space $Tot(T X_1^{ns})$ , which comes from the embedding $Tot(TX_1) \stackrel i\hookrightarrow Tot(TM_1)$. Since $X_1$ is completely geodesic, $i$ is an isometry. This topology is compatible with the map $d\phi:\; TX_1 {\:\longrightarrow\:} TX_2$, because $d\phi$ is also an isometry. Assume that $\gamma$ is differentiable outside of $\{0.5\}$ and \[ \lim\limits_{t\to +0.5} \frac{d\gamma}{dt} = \lim\limits_{t\to -0.5} \frac{d\gamma}{dt}. \] (the limits are taken in the metric completion of $TX_1$, which might be considered as a subset of $Tot(TM_1)$). The Zariski tangent space $T_x X_1$ can be identified with equivalence classes of such paths as follows. Two paths $\gamma, \gamma'$ are equivalent if \[ \lim\limits_{t\to 0.5} \frac{\rho(\gamma_1(t), \gamma_2(t))}{(t-0.5)^2} =0, \] where $\rho$ is the distance function in $M_1$. Since the distance in $M_1$ coinsides with the distance in $X_1$, and $\phi$ is isometry, this equivalence relation is compatible with $\phi$. \blacksquare \subsection{Algebro-geometric properties of the map $\Psi^{s_1}_{s_2}$.} Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a closed trianalytic subvariety. Fix a choice of induced complex structure. Let $D(x)$ be the Douady space of $X$, and $s_1, s_2$ points on $D(X)$, and $\gamma$ a real analytic path in $D(x)$ connecting $s_1$ and $s_2$. Let $X_1 = X_{s_1}, X_2 = X_{s_2}$ be the subvarieties of $M$ corresponding to $s_1$, $s_2$, and $\Psi:\; X_1 {\:\longrightarrow\:} X_2$ be the bimeromorphic map of \ref{_Psi_bimero_Corollary_}. Let $X_1^{ns}$, $X_2^{ns}$ be the non-singular part of $X_1$, and $\Psi^{ns}$ be the restriction of $\Psi$ to $X_1^{ns}$. Let $\Gamma^{ns}\subset M\times M$ be a graph of $\Phi^{ns}$, and $\Gamma$ its closure. Clearly, $\Gamma$ is a complex analytic subvariety of $M\times M$. Let $\pi_i:\; \Gamma {\:\longrightarrow\:} X_i$ be the projections of $\Gamma$ to $X_i$, which are obviously morphisms of complex varieties. \hfill \proposition\label{_pi_i_properties_Proposition_} The maps $\pi_i$, $i= 1,2$, have the following properties. \begin{description} \item[(i)] $\pi_i$ is finite. \item[(ii)] $\pi_i$ is dominant and induces isomorphism of normalizations \item[(iii)] For every point $x\in \Gamma$, the differential $d \pi_i :\; T_x \Gamma {\:\longrightarrow\:} T_{\pi_i(x)} X_i$ is an isomorphism. \end{description} {\bf Proof:} The statement (ii) follows from \ref{_Psi_iso_on_norma_Proposition_} and (iii) from \ref{_homeo_indu_iso_Zariski_Proposition_}. To prove \ref{_pi_i_properties_Proposition_} (i), consider the normalization map $n:\; \tilde \Gamma {\:\longrightarrow\:} \Gamma$. On the level of rings of functions, we have an embedding \[ {\cal O}(X_i) \hookrightarrow {\cal O}(\Gamma) \hookrightarrow {\cal O}(\tilde \Gamma). \] Since $\tilde \Gamma$ is a normalization of $X_i$ by (ii), the ring ${\cal O}(\tilde \Gamma)$ is finitely generated as a ${\cal O}(X_i)$-module. Since ${\cal O}(X_i)$ is Noetherian (\cite{_Grauert_}), this implies that ${\cal O}(\Gamma)$ is also finitely generated as a ${\cal O}(X_i)$-module. This proves \ref{_pi_i_properties_Proposition_} (i). \blacksquare \hfill To prove that the map $\Psi:\; X_1 {\:\longrightarrow\:} X_2$, a.k.a. $\Psi^{s_1}_{s_2}:\; X_{s_1} {\:\longrightarrow\:} X_{s_2}$ is holomorphic, it suffices to show that $\pi_i:\; \Gamma {\:\longrightarrow\:} X_i$ is an isomorphism. In the following Subsection, we show that conditions (i) -- (iii) of \ref{_pi_i_properties_Proposition_} are {\it a priori} sufficient to establish that $\pi$ is an isomorphism. This will finish the proof of \ref{_triana_subse_comple_ana_Theorem_}. \subsection{Finite dominant unramified morphisms of complex varieties.} \hfill \proposition \label{_alge_geo_suffi_pi_i_iso_Proposition_} Let $\phi:\; X {\:\longrightarrow\:} Y$ be a map of complex varieties satisfying conditions (i)--(iii) of \ref{_pi_i_properties_Proposition_}. Then $\phi$ is an isomorphism. \hfill {\bf Proof:} The map $\phi$ is one-to-one in general point (by (ii)). Thus, to prove that $\phi$ is an isomorphism it suffices to show that $\phi$ is etale. On the other hand, $\phi$ is unramified by (iii). By definition of etale morphisms, to prove that $\phi$ is etale it remains to show that $\phi$ is flat. To conclude the proof of \ref{_alge_geo_suffi_pi_i_iso_Proposition_}, we use the following lemma. \hfill \lemma \label{_unrami_domi_flat_Lemma_} Let $\phi:\; X {\:\longrightarrow\:} Y$ be a dominant morphism of complex varieties. Assume that for every point $x\in X$, the map $\phi$ induces an isomorphism $d\phi:\; T_x X {\:\longrightarrow\:} T_{\phi(x)} Y$ of Zariski tangent spaces. Then $\phi$ is flat. \hfill {\bf Proof:} Let $y=\phi (x)$. Conside the associated morphism of local rings ${\cal O}_y Y \stackrel {\phi_x} \hookrightarrow {\cal O}_x X$. To prove that $\phi$ is flat, it suffices to show that $\phi$ is an isomorphism. Let ${\mathfrak m}_x$, ${\mathfrak m}_y$ be the maximal ideals in ${\cal O}_x X$, ${\cal O}_y Y$. Then ${\mathfrak m}_x/{\mathfrak m}_x^2$ is generated by $\phi_x({\mathfrak m}_y)$. Thus, by Nakayama, $\phi_x({\mathfrak m}_y)$ generate ${\mathfrak m}_x$, and we obtain ${\mathfrak m}_x= \phi_x({\mathfrak m}_y)\otimes {\cal O}_x X$. Consider ${\cal O}_x X$ as ${\cal O}_y Y$-module. Then ${\cal O}_x X/{\mathfrak m}_y {\cal O}_x X$ is one-dimensional. Applying Nakayama once more, we obtain that ${\cal O}_x X$ is an ${\cal O}_y Y$-module generated by $1\in{\cal O}_x X$. Thus, the map $\phi_x$ is surjective. Since $\phi$ is dominant, $\phi_x$ is also injective. Thus, ${\cal O}_x X$ is a free ${\cal O}_{\phi(x)} Y$-module for every $x\in X$. By the local criterion of flatness, this implies that $\phi$ is flat. \ref{_unrami_domi_flat_Lemma_} is proven. This finishes the proof of \ref{_alge_geo_suffi_pi_i_iso_Proposition_} and \ref{_triana_subse_comple_ana_Theorem_} \blacksquare \section{Singular hyperk\"ahler varieties.} \label{_singu_hype_Section_} It is an intriguing question, what is the ``correct''\footnote{``Correct'' in Platonic sense: some mathematicians presume that the unique ``correct'' definition of each and every significant mathematical object exists in itself and independently of human perception.} definition of a singular hyperk\"ahler variety. We don't pretend to answer this question. Instead, we give an {\it ad hoc} set of axioms which describe some known examples (deformation spaces of stable bundles and trianalytic subvarieties). It is likely that this {\it ad hoc} definition is stronger than the ``correct'' one. A more elegant approach was suggested by Deligne and Simpson (\cite{_Deligne:defi_}, \cite{_Simpson:hyperka-defi_}). \hfill \definition\label{_singu_hype_Definition_} (\cite{_Verbitsky:Hyperholo_bundles_}, Definition 6.5) Let $M$ be a hypercomplex variety (\ref{_hypercomplex_Definition_}). The following data define a structure of {\bf hyperk\"ahler variety} on $M$. \begin{description} \item[(i)] For every $x\in M$, we have an ${\Bbb R}$-linear symmetric positively defined bilinear form $s_x:\; T_x M \times T_x M {\:\longrightarrow\:} {\Bbb R}$ on the corresponding real Zariski tangent space. \item[(ii)] For each triple of induced complex structures $I$, $J$, $K$, such that $I\circ J = K$, we have a holomorphic differential 2-form $\Omega\in \Omega^2(M, I)$. \item[(iii)] Fix a triple of induced complex structure $I$, $J$, $K$, such that $I\circ J = K$. Consider the corresponding differential 2-form $\Omega$ of (ii). Let $J:\; T_x M {\:\longrightarrow\:} T_x M$ be an endomorphism of the real Zariski tangent spaces defined by $J$, and $Re\Omega\restrict x$ the real part of $\Omega$, considered as a bilinear form on $T_x M$. Let $r_x$ be a bilinear form $r_x:\; T_x M \times T_x M {\:\longrightarrow\:} {\Bbb R}$ defined by $r_x(a,b) = - Re\Omega\restrict x (a, J(b))$. Then $r_x$ is equal to the form $s_x$ of (i). In particular, $r_x$ is independent from the choice of $I$, $J$, $K$. \end{description} \noindent \remark \nopagebreak \begin{description} \item[(a)] It is clear how to define a morphism of hyperk\"ahler varieties. \item[(b)] For $M$ non-singular, \ref{_singu_hype_Definition_} is equivalent to the usual one (\ref{_hyperkahler_manifold_Definition_}). If $M$ is non-singular, the form $s_x$ becomes the usual Riemann form, and $\Omega$ becomes the standard holomorphically symplectic form. \item[(c)] It is easy to check the following. Let $X$ be a hypercomplex subvariety of a hyperk\"ahler variety $M$. Then, restricting the forms $s_x$ and $\Omega$ to $X$, we obtain a hyperk\"ahler structure on $X$. In particular, trianalytic subvarieties of hyperk\"ahler manifolds are always hyperk\"ahler, in the sense of \ref{_singu_hype_Definition_}. \end{description} \hfill {\bf Caution:} Not everything which is seemingly hyperk\"ahler satisfies the conditions of \ref{_singu_hype_Definition_}. Take a quotient $M/G$ os a hyperk\"ahler manifold by an action of finite group $G$, acting in accordance with hyperk\"ahler structure. Then $M/G$ is, generally speaking, {\it not} hyperk\"ahler (in fact, $M/G$ is {\it never} hyperk\"ahler). For instance, take a quotient of a 2-dimensional torus $T$ by $G=\{\pm 1\}$ acting as an involution $t{\:\longrightarrow\:} -t$. This is a beautiful and well known example of a hyperk\"ahler automorphism; the quotient space has 16 isolated singular points, which, if blown up, give a K3 surface. For $x$ a singular point of $T/\{\pm 1\}$, its Zariski tangent space has real dimension 6. On the other hand, for a hypercomplex variety, there is a quaternion action in every Zariski tangent space, and thus, the dimension real dimension of Zariski tangent space must be divisible by 4. We obtain that the space $T/\{\pm 1\}$ is not even hypercomplex. How this happens? We take a twistor space $\operatorname{Tw}(T)$ of $T$ and take a quotient of $\operatorname{Tw}(T)$ by the natural action of $G=\{\pm 1\}$, which is holomorphic. The quotient is a complex variety fibered over ${\Bbb C} P^1$ For $\operatorname{Tw}(T)/\{\pm 1\}$ to be hypercomplex, this fibration must be trivial, in a real analytic category. But the functor of forgetting the complex structure does not commute with taking finite quotients! Thus, even if $Tw(T)$ is (as a real analytic space) trivially fibered over ${\Bbb C} P^1$, there is no way to push down this trivialization to $\operatorname{Tw}(T)/\{\pm 1\}$. \hfill The following theorem, proven in \cite{_Verbitsky:Hyperholo_bundles_} (Theorem 6.3), gives a convenient way to construct examples of hyperk\"ahler varieties. \hfill \theorem \label{_hyperho_defo_hyperka_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure and $B$ a stable holomorphic bundle over $(M, I)$. Let $D(B)$ be a deformation space of stable holomorphic structures on $B$. Assume that $c_1(B)$, $c_2(B)$ are $SU(2)$-invariant, with respect to the standard action of $SU(2)$ on $H^*(M)$. Then $D(B)$ has a natural structure of a hyperk\"ahler variety. \nopagebreak \blacksquare \hfill The following theorem is implicit in \cite{_Verbitsky:Hyperholo_bundles_}. \hfill \theorem \label{_hyperholo_functo_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $I$ an induced complex structure and $B_1, B_2, ..., B_n$ stable holomorphic bundles over $(M, I)$. Let $D(B_i)$ be a deformation space of stable holomorphic structures on $B_i$. Assume that $c_1(B_i)$, $c_2(B_i)$, $i=1, 2, ..., n$ are $SU(2)$-invariant, with respect to the standard action of $SU(2)$ on $H^*(M)$. Let $\Pi$ be a natural tensor operation on the vector bundles, such that, e. g., \[ B_1, ... B_n {\:\longrightarrow\:} B_1 \otimes B_2 \otimes \Lambda^2 B_3 \otimes S^7 B_4 \otimes ... \otimes B_n ^*. \] Assume that $\Pi(B_1, ... B_n)$ cannot be decomposed to a direct sum of holomorphic bundles. Then $\Pi(B_1, ... B_n)$ is stable, and the associated map \[ D(B_1) \times D(B_2) \times ..., \times D(B_n){\:\longrightarrow\:} D\left(\Pi(B_1, ..., B_n)\right) \] (defined in a certain neighbourhood of $[B_1]\times [B_2]\times ..., \times [B_n] \in D(B_1) \times D(B_2) \times ..., \times D(B_n)$ is a morphism of hyperk\"ahler varieties. \blacksquare \hfill \ref{_hyperholo_functo_Theorem_} gives a natural way to construct trianalytic subvarieites of hyperk\"ahler varieties. \hfill The following theorem is almost trivial; the reader is advised to invent his or her own proof instead of reading ours (which is by necessity sketchy). \hfill \theorem \label{_Doua_hyperka_Theorem_} Let $M$ be a compact hyperk\"ahler manifold, $X\subset M$ a trianalytic subvariety and $D(X)$ its Douady space. Then $D(X)$ is a compact hyperk\"ahler variety. \hfill {\bf Proof:} Clearly, $D(X)$ is hypercomplex, in a natural way (\ref{_Douady_hyperc_Remark_}). It remains to construct the forms $s_x$ and $\Omega$. Let $Y$ be a trianalytic subvariety of $M$ which is a deformation of $X$, and $U$ be a sufficiently small neighbourhood of $[Y]\in D(X)$. Let $\left(\pi:\; {\c X}_U {\:\longrightarrow\:} U,\phi:\; {\c X}_U {\:\longrightarrow\:} M\right)$ be the universal family of subvarieties of $M$, attached to $U\subset D(X)$. The space ${\c X}_U$ is hypercomplex and the map $\pi$ is compatible with the hypercomplex structure. Moreover, the map $\phi:\; {\c X}_U {\:\longrightarrow\:} M$ is an immersion, so ${\c X_U}$ is hyperk\"ahler (hyperk\"ahler structure is obtained as a pullback from $M$). \ref{_conne_in_triana_flat_Theorem_} provides a natural trivialization of ${\c X}_U {\:\longrightarrow\:} U$. Thus, for each $y\in Y$, there exists a natural section $\sigma_y:\; U\hookrightarrow {\c X}_U$. By \ref{_triana_subse_comple_ana_Theorem_}, this section is compatible with the hypercomplex structure. Restricting the hyperk\"ahler structure from ${\c X_U}$ to $\sigma_y(U)$, we obtain a hyperk\"ahler structure on $U$. It is easy to check that this hyperk\"ahler structure is independent from the choice of section $\sigma_y$. Gluing the hyperk\"ahler structures from different $U$, we obtain the proof of \ref{_Doua_hyperka_Theorem_}. \blacksquare \hfill {\bf Acknowledgements:} I am indebted to D. Kaledin for enlightening discussions; he also suggested some of the proofs. I am thankful to Tony Pantev for the frutiful discussions. Mohan Ramachandran communicated me the reference to \cite{_Greene_}. I am grateful to A. Beilinson and J. Bernstein for the help with algebraic geometry. My gratitude to M. Finkelberg, D. Kazhdan and S.-T. Yau for their interest. \hfill